Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Download as pdf or txt
Download as pdf or txt
You are on page 1of 629

PHYSICAL CHEMISTRY

FOR THE BIOSCIENCES

LibreTexts Textmap
Physical Chemistry for the Biosciences

LibreTexts Textmap
This text is disseminated via the Open Education Resource (OER) LibreTexts Project (https://LibreTexts.org) and like the hundreds
of other texts available within this powerful platform, it is freely available for reading, printing and "consuming." Most, but not all,
pages in the library have licenses that may allow individuals to make changes, save, and print this book. Carefully
consult the applicable license(s) before pursuing such effects.
Instructors can adopt existing LibreTexts texts or Remix them to quickly build course-specific resources to meet the needs of their
students. Unlike traditional textbooks, LibreTexts’ web based origins allow powerful integration of advanced features and new
technologies to support learning.

The LibreTexts mission is to unite students, faculty and scholars in a cooperative effort to develop an easy-to-use online platform
for the construction, customization, and dissemination of OER content to reduce the burdens of unreasonable textbook costs to our
students and society. The LibreTexts project is a multi-institutional collaborative venture to develop the next generation of open-
access texts to improve postsecondary education at all levels of higher learning by developing an Open Access Resource
environment. The project currently consists of 14 independently operating and interconnected libraries that are constantly being
optimized by students, faculty, and outside experts to supplant conventional paper-based books. These free textbook alternatives are
organized within a central environment that is both vertically (from advance to basic level) and horizontally (across different fields)
integrated.
The LibreTexts libraries are Powered by NICE CXOne and are supported by the Department of Education Open Textbook Pilot
Project, the UC Davis Office of the Provost, the UC Davis Library, the California State University Affordable Learning Solutions
Program, and Merlot. This material is based upon work supported by the National Science Foundation under Grant No. 1246120,
1525057, and 1413739.
Any opinions, findings, and conclusions or recommendations expressed in this material are those of the author(s) and do not
necessarily reflect the views of the National Science Foundation nor the US Department of Education.
Have questions or comments? For information about adoptions or adaptions contact info@LibreTexts.org. More information on our
activities can be found via Facebook (https://facebook.com/Libretexts), Twitter (https://twitter.com/libretexts), or our blog
(http://Blog.Libretexts.org).
This text was compiled on 07/09/2023
TABLE OF CONTENTS
Licensing

1: Introduction to Physical Chemistry


1.1: Nature of Physical Chemistry
1.2: Units
1.3: Atomic Mass, Molecular Mass, and the Chemical Mole

2: Properties of Gases
2.1: Some Definitions
2.2: An Operational Definition of Temperature
2.3: Ideal Gases
2.4: Real Gases
2.5: Condensation of Gases and the Critical State
2.6: Kinetic Theory of Gases
2.7: The Maxwell Distribution Laws
2.8: Molecular Collisions and the Mean Free Path
2.9: Graham's Laws of Diffusion and Effusion
2.E: Properties of Gases (Exercises)

3: The First Law of Thermodynamics


3.1: Work and Heat
3.2: The First Law of Thermodynamics
3.3: Heat Capacities
3.4: Gas Expansion
3.5: Calorimetry
3.6: Thermochemistry
3.7: Bond Energies and Enthalpies
3.E: Exercises

4: The Second Law of Thermodynamics


4.1: Spontaneous Processes
4.2: Entropy
4.3: The Second Law of Thermodynamics
4.4: The Third Law of Thermodynamics
4.5: Evaluating Entropy and Entropy Changes
4.6: Gibbs Energy
4.7: Standard Molar Gibbs Energy of Formation
4.8: Dependence of Gibbs Energy on Temperature and Pressure
4.9: Phase Equilibria
4.10: Thermodynamics of Rubber Elasticity
4.E: Exercises

5: Solutions
5.1: Concentration Units
5.2: Partial Molar Quantities
5.3: The Thermodynamics of Mixing

1 https://chem.libretexts.org/@go/page/182383
5.4: Binary Mixtures of Volatile Liquids
5.5: Real Solutions
5.6: Colligative Properties
5.7: Electrolyte Solutions
5.8: Ionic Activity
5.9: Colligative Properties of Electrolyte Solutions
5.10: Biological Membranes
5.E: Solutions (Exercises)

6: Chemical Equilibrium
6.1: Chemical Equilibrium in Gaseous Systems
6.2: Reactions in Solutions
6.3: Heterogeneous Equilibria
6.4: The Influence of Temperature, Pressure, and Catalysts on the Equilibrium Constant
6.5: Binding of Ligands and Metal Ions to Macromolecules
6.6: Bioenergetics
6.E: Chemical Equilibrium (Exercises)

7: Electrochemistry
7.1: Electrochemical Cells
7.2: Single Electrode Potentials
7.3: Thermodynamics of Electrochemical Cells
7.4: Types of Electrochemical Cells
7.5: Applications of EMF Measurements
7.6: Biological Oxidation
7.7: Membrane Potential
7.E: Electrochemistry (Exercises)

8: Acids and Bases


8.1: Definitions of Acids and Bases
8.2: Acid-Base Properties of Water
8.3: Dissociation of Acids and Bases
8.4: Diprotic and Polyprotic Acids
8.5: Buffer Solutions
8.6: Acid-Base Titrations
8.7: Amino Acids
8.8: Maintaining the pH of Blood

9: Chemical Kinetics
9.1: Reaction Rates
9.2: Reaction Order
9.3: Molecularity of a Reaction
9.4: More Complex Reactions
9.5: The Effect of Temperature on Reaction Rates
9.6: Potential Energy Surfaces
9.7: Theories of Reaction Rates
9.8: Isotope Effects in Chemical Reactions
9.9: Reactions in Solution
9.10: Fast Reactions in Solution
9.11: Oscillating Reactions

2 https://chem.libretexts.org/@go/page/182383
9.E: Chemical Kinetics (Exercises)

10: Enzyme Kinetics


10.1: General Principles of Catalysis
10.2: The Equations of Enzyme Kinetics
10.3: Chymotrypsin- A Case Study
10.4: Multisubstrate Systems
10.5: Enzyme Inhibition
10.6: Allosteric Interactions
10.7: The Effect of pH on Enzyme Kinetics
10.8: The Effect of Temperature on Enzyme Kinetics
10.E: Exercises

11: Quantum Mechanics and Atomic Structure


11.1: The Wave Theory of Light
11.2: Planck's Quantum Theory
11.3: The Photoelectric Effect
11.4: Bohr's Theory of the Hydrogen Emission Spectrum
11.5: de Broglie's Postulate
11.6: The Heisenberg Uncertainty Principle
11.7: The Schrödinger Wave Equation
11.8: Particle in a One-Dimensional Box
11.9: Quantum-Mechanical Tunneling
11.10: The Schrödinger Wave Equation for the Hydrogen Atom
11.11: Many-Electron Atoms and the Periodic Table
11.E: Quantum Mechanics and Atomic Structure (Exercises)

12: The Chemical Bond


12.1: Lewis Structures
12.2: Valence Bond Theory
12.3: Hybridization of Atomic Orbitals
12.4: Electronegativity and Dipole Moment
12.5: Molecular Orbital Theory
12.6: Diatomic Molecules
12.7: Resonance and Electron Delocalization
12.8: Coordination Compounds
12.9: Coordination Compounds in Biological Systems
12.E: The Chemical Bond (Exercises)

13: Intermolecular Forces


13.1: Intermolecular Interactions
13.2: The Ionic Bond
13.3: Types of Intermolecular Forces
13.4: Hydrogen Bonding
13.5: The Structure and Properties of Water
13.6: Hydrophobic Interaction
13.E: Intermolecular Forces (Exercises)

3 https://chem.libretexts.org/@go/page/182383
14: Spectroscopy
14.1: Vocabulary
14.2: Microwave Spectroscopy
14.3: Infrared Spectroscopy
14.4: Electronic Spectroscopy
14.5: Nuclear Magnetic Resonance
14.6: Electron Spin Resonance
14.7: Fluorescence and Phosphorescence
14.8: Lasers
14.9: Optical Rotatory Dispersion and Circular Dichroism
14.E: Spectroscopy (Exercises)

15: Photochemistry and Photobiology


15.1: Introduction to Photochemistry
15.2: Photosynthesis
15.3: Vision
15.4: Biological Effects of Radiation

16: Macromolecules
16.1: Size, Shape, and Molar Mass of Macromolecules
16.2: Structure of Synthetic Polymers
16.3: Structure of Proteins and DNA
16.4: Protein Stability

Index

Glossary
Detailed Licensing

4 https://chem.libretexts.org/@go/page/182383
Licensing
A detailed breakdown of this resource's licensing can be found in Back Matter/Detailed Licensing.

1 https://chem.libretexts.org/@go/page/417257
CHAPTER OVERVIEW

1: Introduction to Physical Chemistry


1.1: Nature of Physical Chemistry
1.2: Units
1.3: Atomic Mass, Molecular Mass, and the Chemical Mole

1: Introduction to Physical Chemistry is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1
1.1: Nature of Physical Chemistry
1.1: Nature of Physical Chemistry is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1.1.1 https://chem.libretexts.org/@go/page/41460
1.2: Units
There is international agreement that the units used for physical quantities in science and technology should be those of the
International System of Units, or SI (standing for the French Système International d’Unités). The Physical Chemistry Division
of the International Union of Pure and Applied Chemistry, or IUPAC, produces a manual of recommended symbols and
terminology for physical quantities and units based on the SI. The manual has become known as the Green Book (from the color of
its cover) and is referred to here as the IUPAC Green Book. This e-book will, with a few exceptions, use symbols recommended in
the third edition (2007) of the IUPAC Green Book (E. Richard Cohen et al, Quantities, Units and Symbols in Physical Chemistry,
3rd edition. RSC Publishing, Cambridge, 2007). These symbols are listed for convenient reference in Appendices C and D.
Any of the symbols for units listed in Tables 1.1–1.3, except kg and C, may be preceded by one of the prefix symbols of Table 1.4

to construct a decimal fraction or multiple of the unit. (The symbol g may be preceded by a prefix symbol to construct a fraction or
multiple of the gram.) The combination of prefix symbol and unit symbol is taken as a new symbol that can be raised to a power
without using parentheses, as in the following examples:
The physical quantity formally called amount of substance is a counting quantity for particles, such as atoms or molecules, or
for other chemical entities. The counting unit is invariably the mole, defined as the amount of substance containing as many
particles as the number of atoms in exactly 12 grams of pure carbon-12 nuclide, C. See Appendix A for the wording of the
12

official IUPAC definition. This definition is such that one mole of H O molecules, for example, has a mass of 18.0153 grams
2

(where 18.0153 is the relative molecular mass of H O) and contains 6.02214 × 10 molecules (where 6.02214 × 10 mol
2
23 23 −1

is the Avogadro constant to six significant digits). The same statement can be made for any other substance if 18.0153 is
replaced by the appropriate atomic mass or molecular mass value.
The symbol for amount of substance is n . It is admittedly awkward to refer to n (H O) as “the amount of substance of water.”
2

This e-book simply shortens “amount of substance” to amount. An alternative name suggested for n is “chemical amount.”
Thus, “the amount of water in the system” refers not to the mass or volume of water, but to the number of H O molecules in the
2

system expressed in a counting unit such as the mole.

1.2: Units is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
1.1: Units by Howard DeVoe is licensed CC BY 4.0. Original source: https://www2.chem.umd.edu/thermobook.

1.2.1 https://chem.libretexts.org/@go/page/41461
1.3: Atomic Mass, Molecular Mass, and the Chemical Mole
1.3: Atomic Mass, Molecular Mass, and the Chemical Mole is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by LibreTexts.

1.3.1 https://chem.libretexts.org/@go/page/41462
CHAPTER OVERVIEW

2: Properties of Gases
The study of gases allows us to understand the behavior of matter at its simplest: individual particles, acting independently, almost
completely uncomplicated by interactions and interferences between each other. This knowledge of gases will serve as the pathway
to our understanding of the far more complicated condensed phases (liquids and solids) in which the theory of gases will no longer
give us correct answers, but it will still provide us with a useful model that will at least help us to rationalize the behavior of these
more complicated states of matter.
2.1: Some Definitions
2.2: An Operational Definition of Temperature
2.3: Ideal Gases
2.4: Real Gases
2.5: Condensation of Gases and the Critical State
2.6: Kinetic Theory of Gases
2.7: The Maxwell Distribution Laws
2.8: Molecular Collisions and the Mean Free Path
2.9: Graham's Laws of Diffusion and Effusion
2.E: Properties of Gases (Exercises)

2: Properties of Gases is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1
2.1: Some Definitions
2.1: Some Definitions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

2.1.1 https://chem.libretexts.org/@go/page/41403
2.2: An Operational Definition of Temperature
2.2: An Operational Definition of Temperature is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.

2.2.1 https://chem.libretexts.org/@go/page/41404
2.3: Ideal Gases
In an ideal gas, there are no interactions between the particles, hence, the particles do not exert forces on each other. However,
particles do experience a force when they collide with the walls of the container. Let us assume that each collision with a wall is
elastic. Let us assume that the gas is in a cubic box of length a and that two of the walls are located at x = 0 and at x = a . Thus, a
particle moving along the x direction will eventually collide with one of these walls and will exert a force on the wall when it
strikes it, which we will denote as F . Since every action has an equal and opposite reaction, the wall exerts a force −F on the
x x

particle. According to Newton’s second law, the force −F on the particle in this direction gives rise to an acceleration via
x

Δvx
−Fx = m ax = m (2.3.1)
Δt

Here, t represents the time interval between collisions with the same wall of the box. In an elastic collision, all that happens to the
velocity is that it changes sign. Thus, if v is the velocity in the x direction before the collision, then −v is the velocity after, and
x x

Δv = −v − v = −2 v
x x x x, so that
vx
−Fx = −2m (2.3.2)
Δt

In order to find Δt, we recall that the particles move at constant speed. Thus, a collision between a particle and, say, the wall at
x = 0 will not change the particle’s speed. Before it strikes this wall again, it will proceed to the wall at x = a first, bounce off that

wall, and then return to the wall at x = 0 . The total distance in the x direction traversed is 2a, and since the speed in the x
2a
direction is always v , the interval Δt =
x . Consequently, the force is
vx

2
mvx
−Fx = − (2.3.3)
a

Thus, the force that the particle exerts on the wall is


2
mvx
Fx = (2.3.4)
a

The mechanical definition of pressure is


⟨F ⟩
P = (2.3.5)
A

where ⟨F ⟩ is the average force exerted by all N particles on a wall of the box of area A . Here A = a . If we use the wall at x = 0
2

we have been considering, then


N ⟨Fx ⟩
P = (2.3.6)
2
a

because we have N particles hitting the wall. Hence,


2
N m⟨vx ⟩
P = (2.3.7)
3
a

from our study of the Maxwell-Boltzmann distribution, we found that


kB T
2
⟨vx ⟩ = (2.3.8)
m

Hence, since a 3
=V ,
N kB T nRT
P = = (2.3.9)
V V

which is the ideal gas law.

2.3.1 https://chem.libretexts.org/@go/page/41405
Figure 2.3.1: (Left) Pressure vs. volume for different temperatures (isotherms of the ideal-gas equation of state. (Right) Pressure vs.
temperature for different densities ρ = N /V .
The ideal gas law is an example of an equation of state, which was introduced in Lecture 1. One way to visualize any equation of
state is to plot the so-called isotherms, which are graphs of P vs. V at fixed values of T . For the ideal-gas equation of state
P = nRT /V , some of the isotherms are shown in the figure below (left panel): If we plot P vs. T at fixed volume (called the

isochores), we obtain the plot in the right panel. What is important to note, here, is that an ideal gas can exist only as a gas. It is not
possible for an ideal gas to condense into some kind of “ideal liquid”. In other words, a phase transition from gas to liquid can be
modeled only if interparticle interactions are properly accounted for.
Note that the ideal-gas equation of state can be written in the form
¯
PV PV P
= = =1 (2.3.10)
nRT RT ρRT

where V¯ = V /n is called the molar volume. Unlike V , which increases as the number of moles increases (an example of what is
called an extensive quantity in thermodynamics), \(\bar{V}\) does not exhibit this dependence and, therefore, is called intensive.
The quantity
¯
PV PV P
Z = = = (2.3.11)
nRT RT ρRT

is called the compressibility of the gas. In an ideal gas, if we “compress” the gas by increasing P , the density ρ must increase as
well so as to keep Z = 1 . For a real gas, Z , therefore, gives us a measure of how much the gas deviates from ideal-gas behavior.

Figure 2.3.2: Plot of the compressibility Z vs. P for several gases, together with the ideal gas, where Z = 1 .
Figure 2.2 shows a plot of Z vs. P for several real gases and for an ideal gas. The plot shows that for sufficiently low pressures
(hence, low densities), each gas approaches ideal-gas behavior, as expected.

Contributors and Attributions


Mark Tuckerman (New York University)

2.3: Ideal Gases is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

2.3.2 https://chem.libretexts.org/@go/page/41405
2.4: Real Gases
According to Boyle's law, the product PV is a constant at any given temperature, so a plot of PV as a function of the pressure of an ideal gas yields a horizontal straight line. This implies that any
increase in the pressure of the gas is exactly counteracted by a decrease in the volume as the molecules are crowded closer together. But we know that the molecules themselves are finite objects
having volumes of their own, and this must place a lower limit on the volume into which they can be squeezed. So we must reformulate the ideal gas equation of state as a relation that is true only in
the limiting case of zero pressure:
lim P V = nRT (2.4.1)
P →0

So what happens when a real gas is subjected to a very high pressure? The outcome varies with both the molar mass of the gas and its temperature, but in general we can see the the effects of both
repulsive and attractive intermolecular forces:
Repulsive forces: As a gas is compressed, the individual molecules begin to get in each other's way, giving rise to a very strong repulsive force acts to oppose any further volume decrease. We
would therefore expect the PV vs P line to curve upward at high pressures, a nd this is in fact what is observed for all gases at sufficiently high pressures.
Attractive forces: At very close distances, all molecules repel each other as their electron clouds come into contact. At greater distances, however, brief statistical fluctuations in the distribution
these electron clouds give rise to a universal attractive force between all molecules. The more electrons in the molecule (and thus the greater the molecular weight), the greater is this attractive
force. As long as the energy of thermal motion dominates this attractive force, the substance remains in the gaseous state, but at sufficiently low temperatures the attractions dominate and the
substance condenses to a liquid or solid.
The universal attractive force described above is known as the dispersion, or London force. There may also be additional (and usually stronger) attractive forces related to charge imbalance in the
molecule or to hydrogen bonding. These various attractive forces are often referred to collectively as van der Waals forces. A plot of PV/RT as a function of pressure is a very sensitive indicator of
deviations from ideal behavior, since such a plot is just a horizontal line for an ideal gas. The two illustrations below show how these plots vary with the nature of the gas, and with temperature.

Intermolecular attractions, which generally increase with molecular weight, cause the PV product to decrease as higher pressures bring the molecules closer together and thus within the range of these
attractive forces; the effect is to cause the volume to decrease more rapidly than it otherwise would. The repulsive forces always eventually win out. But as the molecules begin to intrude on each
others' territory, the stronger repulsive forces cause the curve to bend upward.

The temperature makes a big difference! At higher temperatures, increased thermal motions overcome the effects of intermolecular attractions which normally dominate at lower pressures. So all
gases behave more ideally at higher temperatures. For any gas, there is a special temperature (the Boyle temperature) at which attractive and repulsive forces exactly balance each other at zero
pressure. As you can see in this plot for methane, this some of this balance does remain as the pressure is increased.

The van der Waals Equation of State


How might we modify the ideal gas equation of state to take into account the effects of intermolecular interactions? The first and most well known answer to this question was offered by the Dutch
scientist J.D. van der Waals (1837-1923) in 1873. The ideal gas model assumes that the gas molecules are merely points that occupy no volume; the "V" term in the equation is the volume of the
container and is independent of the nature of the gas.

van der Waals recognized that the molecules themselves take up space that subtracts from the volume of the container, so that the “volume of the gas” V in the ideal gas equation should be replaced by
the term (V–b), in which b relates to the excluded volume, typically of the order of 20-100 cm3 mol–1. The excluded volume surrounding any molecule defines the closest possible approach of any two
molecules during collision. Note that the excluded volume is greater then the volume of the molecule, its radius being half again as great as that of a spherical molecule.
The other effect that van der Waals needed to correct for are the intermolecular attractive forces. These are ignored in the ideal gas model, but in real gases they exert a small cohesive force between
the molecules, thus helping to hold the gas together and reducing the pressure it exerts on the walls of the container.
Because this pressure depends on both the frequency and the intensity of collisions with the walls, the reduction in pressure is proportional to the square of the number of molecules per volume of
space, and thus for a fixed number of molecules such as one mole, the reduction in pressure is inversely proportional to the square of the volume of the gas. The smaller the volume, the closer are the
molecules and the greater will be the effect. The van der Waals equation replaces the P term in the ideal gas equation with P + (a/V ) in which the magnitude of the constant a increases with the
2

strength of the intermolecular attractive forces.


The complete van der Waals equation of state can be written as

Although most students are not required to memorize this equation, you are expected to understand it and to explain the significance of the terms it contains. You should also understand that the van
der Waals constants a and b must be determined empirically for every gas. This can be done by plotting the P-V behavior of the gas and adjusting the values of a and b until the van der Waals

2.4.1 https://chem.libretexts.org/@go/page/41406
equation results in an identical plot. The constant a is related in a simple way to the molecular radius; thus the determination of a constitutes an indirect measurement of an important microscopic
quantity.
Table 2.4.1 : van der Waals constants for some gases
Substance molar mass, g a (L2-atm mole–2)

hydrogen H2 2 0.244

helium He 4 0.034

methane CH4 16 2.25

water H2O 18 5.46

nitrogen N2 28 1.39

carbon dioxide CO2 44 3.59

carbon tetrachloride CCl4 154 20.4

The van der Waals equation is only one of many equations of state for real gases. More elaborate equations are required to describe the behavior of gases over wider pressure ranges. These generally
take account of higher-order nonlinear attractive forces, and require the use of more empirical constants. Although we will make no use of them in this course, they are widely employed in chemical
engineering work in which the behavior of gases at high pressures must be accurately predicted.

Contributors and Attributions


Stephen Lower, Professor Emeritus (Simon Fraser U.) Chem1 Virtual Textbook

2.4: Real Gases is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

2.4.2 https://chem.libretexts.org/@go/page/41406
2.5: Condensation of Gases and the Critical State
The most striking feature of real gases is that they cease to remain gases as the temperature is lowered and the pressure is increased.

The Van der Waals Equation of State


In the last lecture, we saw that for the pair potential

⎧∞ r ≤σ

u(r) = ⎨ C6 (2.5.1)
⎩− r >σ
6
r

we could write the second virial coefficient as


2 3
C6
B2 (T ) = π N0 σ [1 − ] (2.5.2)
6
3 3 kB T σ

Let us introduce to simplifying variables


2
3
b = π N0 σ (2.5.3)
3
2
2πN C6
0
a = (2.5.4)
3

in terms of which
a
B2 (T ) = b − (2.5.5)
RT

With these definitions, the virial equation of state becomes


2
nRT n a
P = + RT (b − ) (2.5.6)
2
V V RT
2
nRT nb an
= (1 + )− (2.5.7)
2
V V V

If we assume nb/V is small, then we can also write


nb 1
1+ ≈ (2.5.8)
V nb
1−
V

so that
2
nRT an
P = − (2.5.9)
2
V − nb V

which is known as the van der Waals equation of state.

Figure 2.5.1 : Two hard spheres of diameter σ at closest contact. The distance between their centers is also σ . A sphere of radius σ
just containing the two particles is shown in cross-section.
The first term in this equation is easy to motivate. In fact, it looks very much like the equation of state for an ideal gas having
volume V − nb rather than V . This part of the van der Waals equation is due entirely to the hard-wall potential u (r). Essentially, o

this potential energy term describes a system of “billiard balls” of diameter σ. The figure below shows two of these billiard ball
type particles at the point of contact (also called the distance of closest approach). At this point, they undergo a collision and

2.5.1 https://chem.libretexts.org/@go/page/41407
separate, so that cannot be closer than the distance shown in the figure. At this point, the distance between their centers is also σ, as
the figure indicates. Because of this distance of closest approach, the total volume available to the particles is not V but some
volume less than V . This reduction in volume can be calculated as follows: Figure 2.5.1 shows a shaded sphere that just contains
the pair of billiard ball particles. The volume of this sphere is the volume excluded from any two particles. The radius of the sphere
is σ as the figure shows. Hence, the excluded volume for the two particles is 4π σ /3, which is the volume of the shaded sphere.
3

From this, we see that the excluded volume for any one particle is just half of this or 2π σ /3. The excluded volume for a mole of
3

such particles is just 2π σ N /3, which is the parameter b :


3
0

2 3
b = π σ N0 (2.5.10)
3

Given n moles of gas, the total excluded volume is then nb, so that the total available volume is simply V − nb .
Isotherms of the van der Waals equation are shown in the figure below (left panel). In the figure the volume axis is the molar
volume denoted V¯ = V /n . At sufficiently high temperature, the isotherms approach those of an ideal gas. However, we also see
something strange in some of the isotherms. Specifically, we see a region in which P and V increase together, and we know that
this cannot actually happen in a real gas. It should be clear that many approximations and assumptions go into the derivation of the
van der Waals equation so that some of the important physics is missing from the model. Hence, we should not be surprised if the
van der Waals equation has some unphysical behavior buried in it.

Figure 2.5.2 : Isotherms of the van der Waals equation of state using parameters a and b computed for carbon dioxide.
In fact, we know that at sufficiently low temperatures, any real gas, when compressed, must undergo a transition from gas to liquid.
The signature of such a transition is a discontinuous change in the volume, signifying the condensation of the gas into a liquid that
occupies a significantly lower volume. Unfortunately, the van der Waals equation does not correctly predict this behavior, and
hence, it must be added in ad hoc. This is done by drawing a horizontal line through the isotherm (Figure 2.5.2) and in the figure
below. The vertical position of the line is

Figure 2.5.3 : Illustration of the tie line in the van der Waals equation.
chosen so that the area above the line (between the line and the isotherm) and below the line (again between the line and the
isotherm) is exactly the same. In this way, we entirely remove the artifact of the unphysical increase of P with V when we compute
the compressional work on the gas from ∫ P (V ) dV , to be discussed in our section on thermodynamics. This horizontal line is
called the tie line.

2.5.2 https://chem.libretexts.org/@go/page/41407
As it happens, there is exactly one isotherm along which the van der Waals equation correctly predicts the gas-to-liquid phase
transition. Along this isotherm, the volume discontinuity captured by the tie line is shrunken down to a single point (so that there is
no possibility of an increase of P with V !). This isotherm, in fact, corresponds to the highest possible temperature at which such a
transition can occur. As we approach this isotherm from higher temperatures, this isotherm is a kind of dividing line between the
system’s remaining a gas at all value of P and V and the system’s actually undergoing a gas-to-liquid transition. Hence, this
isotherm is called the critical isotherm (Figure 2.5.4): The temperature of this isotherm is called the critical temperature, denoted
T . The point at which the curve flattens out, signifying the phase transition, is called the critical point. If we draw a curve through
c

the isotherms joining all points of these isotherms at which the tie lines begin, continue the curve up to the critical isotherm, and
down the other side where the tie lines end, this curve reaches a maximum at the critical point. This is illustrated below:

Figure 2.5.4 : Isotherms of the van der Waals equation of state for four different temperatures with explicit illustration of the critical
isotherm.
The shape of the critical isotherm at the critical point allows us to determine the exact temperature, pressure, and volume at which
the phase transition from gas to liquid will occur. At this point, the isotherm is both horizontal and flat. This means that both the
first and second derivatives of P with respect to V must vanish:
2
∂P ∂ P
= 0, =0 (2.5.11)
2
∂V ∂V

Substituting the van der Waals equation into these two conditions, we find the following:
2
nRT 2an
− + =0 (2.5.12)
2 3
(V − nb) V

2
2nRT 6an
− =0 (2.5.13)
3 4
(V − nb) V

Hence, we have two equations in two unknowns V and T for the critical temperature and critical volume. Once these are
determined, the van der Waals equation, itself, allows us to determine the critical pressure.
To solve the equations, first divide one by the other. This gives us a simple condition for the volume:
V − nb V
= (2.5.14)
2 3

3V − 3nb = 2V (2.5.15)

V = 3nb ≡ Vc (2.5.16)

This is the critical volume. Now use either of the two conditions to obtain the critical temperature Tc . If we use the first one, we
find

2.5.3 https://chem.libretexts.org/@go/page/41407
2
nRTc 2an
= (2.5.17)
2 3
(Vc − nb) Vc

2
nRTc 2an
= (2.5.18)
2 3
(3nb − nb) (3nb)
2
nRTc 2an
= (2.5.19)
2 2 3 3
4n b 27n b
8a
RTc = (2.5.20)
27b

Finally, plugging the critical temperature and volume into the van der Waals equation, we obtain the critical pressure

Figure 2.5.5 : Illustration of the curve (dashed line) joining the beginning and ends of the tie lines. This curve reaches a maximum
at the critical point.
as
2
nRTc an
P = − (2.5.21)
2
Vc − nb Vc
2
8an/27b an
= − (2.5.22)
2
3nb − nb
(3nb)
a
= (2.5.23)
2
27b

PVT Surface of a real gas


The plot illustrates this behavior; as the volume is decreased, the lower-temperature isotherms suddenly change into straight lines.
Under these conditions, the pressure remains constant as the volume is reduced. This can only mean that the gas is “disappearing"
as we squeeze the system down to a smaller volume. In its place, we obtain a new state of matter, the liquid. In the green-shaded
region, two phases, liquid, and gas, are simultaneously present. Finally, at very small volume all the gas has disappeared and only
the liquid phase remains. At this point the isotherms bend strongly upward, reflecting our common experience that a liquid is
practically incompressible.

2.5.4 https://chem.libretexts.org/@go/page/41407
To better understand this plot, look at the isotherm labeled . As the gas is compressed from to , the pressure rises in much
the same way as Boyle's law predicts. Compression beyond , however, does not cause any rise in the pressure. What happens
instead is that some of the gas condenses to a liquid. At , the substance is entirely in its liquid state. The very steep rise to
corresponds to our ordinary experience that liquids have very low compressibilities. The range of volumes possible for the liquid
diminishes as the critical temperature is approached.

The critical point


Liquid and gas can coexist only within the regions indicated by the green-shaded area in the diagram above. As the temperature and
pressure rise, this region becomes more narrow, finally reaching zero width at the critical point. The values of P, T, and V at this
juncture are known as the critical constants Pc , Tc , and Vc . The isotherm that passes through the critical point is called the critical
isotherm. Beyond this isotherm, the gas and liquids become indistinguishable; there is only a single fluid phase, sometimes referred
to as a supercritical liquid.

Critical Behaviour of carbon dioxide

At temperatures below 31°C (the critical temperature), CO2 acts somewhat like an ideal gas even at a rather high pressure ( ).
Below 31°, an attempt to compress the gas to a smaller volume eventually causes condensation to begin. Thus at 21°C, at a
pressure of about 62 atm ( ), the volume can be reduced from 200 cm3 to about 55 cm3 without any further rise in the pressure.
Instead of the gas being compressed, it is replaced with the far more compact liquid as the gas is essentially being "squeezed" into
its liquid phase. After all of the gas has disappeared ( ), the pressure rises very rapidly because now all that remains is an almost
incompressible liquid. Above this isotherm ( ), CO2 exists only as a supercritical fluid.

What happens if you have some liquid carbon dioxide in a transparent cylinder at just under its Pc of 62 atm, and you then
compress it slightly? Nothing very dramatic until you notice that the meniscus has disappeared. By successively reducing and
increasing the pressure, you can "turn the meniscus on and off".
One intriguing consequence of the very limited bounds of the liquid state is that you could start with a gas at large volume and low
temperature, raise the temperature, reduce the volume, and then reduce the temperature so as to arrive at the liquid region at the
lower left, without ever passing through the two-phase region, and thus without undergoing condensation!

Supercritical fluids
The supercritical state of matter, as the fluid above the critical point is often called, possesses the flow properties of a gas and the
solvent properties of a liquid. The density of a supercritical fluid can be changed over a wide range by adjusting the pressure; this,
in turn, changes its solubility, which can thus be optimized for a particular application. The picture at the right shows a commercial
laboratory device used for carrying out chemical reactions under supercritical conditions.

2.5.5 https://chem.libretexts.org/@go/page/41407
Supercritical carbon dioxide is widely used to dissolve the caffeine out of coffee beans and as a dry-cleaning solvent.
Supercritical water has recently attracted interest as a medium for chemically decomposing dangerous environmental pollutants
such as PCBs. Supercritical fluids are being increasingly employed as as substitutes for organic solvents (so-called "green
chemistry") in a range of industrial and laboratory processes. Applications that involve supercritical fluids include extractions, nano
particle and nano structured film formation, supercritical drying, carbon capture and storage, as well as enhanced oil recovery
studies.

Contributors and Attributions


Stephen Lower, Professor Emeritus (Simon Fraser U.) Chem1 Virtual Textbook

2.5: Condensation of Gases and the Critical State is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.

2.5.6 https://chem.libretexts.org/@go/page/41407
2.6: Kinetic Theory of Gases
The kinetic theory describes a gas as a large number of submicroscopic particles (atoms or molecules), all of which are in constant,
random motion. The rapidly moving particles constantly collide with each other and with the walls of the container. Kinetic theory
explains macroscopic properties of gases, such as pressure, temperature, viscosity, thermal conductivity, and volume, by
considering their molecular composition and motion. The theory posits that gas pressure is due to the impacts, on the walls of a
container, of molecules or atoms moving at different velocities.

The Model
The five basic tenets of the kinetic-molecular theory are as follows:
1. A gas is composed of molecules that are separated by average distances that are much greater than the sizes of the molecules
themselves. The volume occupied by the molecules of the gas is negligible compared to the volume of the gas itself.
2. The molecules of an ideal gas exert no attractive forces on each other, or on the walls of the container.
3. The molecules are in constant random motion, and as material bodies, they obey Newton's laws of motion. This means that
the molecules move in straight lines (see demo illustration at the left) until they collide with each other or with the walls of the
container.
4. Collisions are perfectly elastic; when two molecules collide, they change their directions and kinetic energies, but the total
kinetic energy is conserved. Collisions are not “sticky".
5. The average kinetic energy of the gas molecules is directly proportional to the absolute temperature. Notice that the term
“average” is very important here; the velocities and kinetic energies of individual molecules will span a wide range of values,
and some will even have zero velocity at a given instant. This implies that all molecular motion would cease if the temperature
were reduced to absolute zero.
According to this model, most of the volume occupied by a gas is empty space; this is the main feature that distinguishes gases
from condensed states of matter (liquids and solids) in which neighboring molecules are constantly in contact. Gas molecules are in
rapid and continuous motion; at ordinary temperatures and pressures their velocities are of the order of 0.1-1 km/sec and each
molecule experiences approximately 1010collisions with other molecules every second.
If gases do in fact consist of widely-separated particles, then the observable properties of gases must be explainable in terms of the
simple mechanics that govern the motions of the individual molecules. The kinetic molecular theory makes it easy to see why a gas
should exert a pressure on the walls of a container. Any surface in contact with the gas is constantly bombarded by the molecules.

Figure 2.6.1: When a molecule collides with a rigid wall, the component of its momentum perpendicular to the wall is reversed. A
force is thus exerted on the wall, creating pressure. Image used with permisison from OpenSTAX

2.6.1 https://chem.libretexts.org/@go/page/41408
At each collision, a molecule moving with momentum mv strikes the surface. Since the collisions are elastic, the molecule bounces
back with the same velocity in the opposite direction. This change in velocity ΔV is equivalent to an acceleration a; according to
Newton's second law, a force f = ma is thus exerted on the surface of area A exerting a pressure P = f/A.

Kinetic Interpretation of Temperature


According to the kinetic molecular theory, the average kinetic energy of an ideal gas is directly proportional to the absolute
temperature. Kinetic energy is the energy a body has by virtue of its motion:
2
mv
KE = (2.6.1)
2

As the temperature of a gas rises, the average velocity of the molecules will increase; a doubling of the temperature will increase
this velocity by a factor of four. Collisions with the walls of the container will transfer more momentum, and thus more kinetic
energy, to the walls. If the walls are cooler than the gas, they will get warmer, returning less kinetic energy to the gas, and causing it
to cool until thermal equilibrium is reached. Because temperature depends on the average kinetic energy, the concept of
temperature only applies to a statistically meaningful sample of molecules. We will have more to say about molecular velocities
and kinetic energies farther on.
Kinetic explanation of Boyle's law: Boyle's law is easily explained by the kinetic molecular theory. The pressure of a gas
depends on the number of times per second that the molecules strike the surface of the container. If we compress the gas to a
smaller volume, the same number of molecules are now acting against a smaller surface area, so the number striking per unit of
area, and thus the pressure, is now greater.
Kinetic explanation of Charles' law: Kinetic molecular theory states that an increase in temperature raises the average kinetic
energy of the molecules. If the molecules are moving more rapidly but the pressure remains the same, then the molecules must
stay farther apart, so that the increase in the rate at which molecules collide with the surface of the container is compensated for
by a corresponding increase in the area of this surface as the gas expands.
Kinetic explanation of Avogadro's law: If we increase the number of gas molecules in a closed container, more of them will
collide with the walls per unit time. If the pressure is to remain constant, the volume must increase in proportion, so that the
molecules strike the walls less frequently, and over a larger surface area.
Kinetic explanation of Dalton's law: "Every gas is a vacuum to every other gas". This is the way Dalton stated what we now
know as his law of partial pressures. It simply means that each gas present in a mixture of gases acts independently of the
others. This makes sense because of one of the fundamental tenets of KMT theory that gas molecules have negligible volumes.
So Gas A in mixture of A and B acts as if Gas B were not there at all. Each contributes its own pressure to the total pressure
within the container, in proportion to the fraction of the molecules it represents.

Derivation of the Ideal Gas Law


One of the triumphs of the kinetic molecular theory was the derivation of the ideal gas law from simple mechanics in the late
nineteenth century. This is a beautiful example of how the principles of elementary mechanics can be applied to a simple model to
develop a useful description of the behavior of macroscopic matter. We begin by recalling that the pressure of a gas arises from the
force exerted when molecules collide with the walls of the container. This force can be found from Newton's law
dv
f = ma = m (2.6.2)
dt

in which v is the velocity component of the molecule in the direction perpendicular to the wall and m is its mass.

To evaluate the derivative, which is the velocity change per unit time, consider a single molecule of a gas contained in a cubic box
of length l. For simplicity, assume that the molecule is moving along the x-axis which is perpendicular to a pair of walls, so that it is
continually bouncing back and forth between the same pair of walls. When the molecule of mass m strikes the wall at velocity +v

2.6.2 https://chem.libretexts.org/@go/page/41408
(and thus with a momentum mv ) it will rebound elastically and end up moving in the opposite direction with –v. The total change
in velocity per collision is thus 2v and the change in momentum is 2mv.
After the collision the molecule must travel a distance l to the opposite wall, and then back across this same distance before
colliding again with the wall in question. This determines the time between successive collisions with a given wall; the number of
collisions per second will be v/2l. The force F exerted on the wall is the rate of change of the momentum, given by the product of
the momentum change per collision and the collision frequency:
2
d(m vx ) vx mvx
F = = (2m vx ) × ( ) = (2.6.3)
dt 2l l

Pressure is force per unit area, so the pressure P exerted by the molecule on the wall of cross-section l becomes 2

2 2
mv mv
P = = (2.6.4)
3
l V

in which V is the volume of the box.


As noted near the beginning of this unit, any given molecule will make about the same number of moves in the positive and
negative directions, so taking a simple average would yield zero. To avoid this embarrassment, we square the velocities before
averaging them
2 2 2 2 2 2
v +v +v +v . . . v ∑ v
¯2 1 2 3 4 N i i
v = = (2.6.5)
N N

and then take the square root of the average. This result is known as the root mean square (rms) velocity.
−−
¯2
vrms = √v (2.6.6)

We have calculated the pressure due to a single molecule moving at a constant velocity in a direction perpendicular to a wall. If we
now introduce more molecules, we must interpret v as an average value which we will denote by v¯ . Also, since the molecules are
2 2

moving randomly in all directions, only one-third of their total velocity will be directed along any one Cartesian axis, so the total
pressure exerted by N molecules becomes
2
N mν̄
P = (2.6.7)
3 V

Recalling that mv¯ /2 is the average translational kinetic energy ϵ, we can rewrite the above expression as
2

1 ¯2 2
PV = N mv = Nϵ (2.6.8)
3 3

The 2/3 factor in the proportionality reflects the fact that velocity components in each of the three directions contributes ½ kT to the
kinetic energy of the particle. The average translational kinetic energy is directly proportional to temperature:
3
ϵ= kT (2.6.9)
2

in which the proportionality constant k is known as the Boltzmann constant. Substituting Equation 2.6.9 into Equation 2.6.8 yields
2 3
PV = ( N) ( kT ) = N kT (2.6.10)
3 2

The Boltzmann constant k is just the gas constant per molecule. For n moles of particles, the Equation 2.6.10 becomes
P V = nRT (2.6.11)

which is the Ideal Gas law.

Contributors and Attributions


Stephen Lower, Professor Emeritus (Simon Fraser U.) Chem1 Virtual Textbook

2.6: Kinetic Theory of Gases is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

2.6.3 https://chem.libretexts.org/@go/page/41408
2.7: The Maxwell Distribution Laws
In the context of the Kinetic Molecular Theory of Gases, a gas contains a large number of particles in rapid motions. Each particle
has a different speed, and each collision between particles changes the speeds of the particles. An understanding of the properties of
the gas requires an understanding of the distribution of particle speeds.

Many molecules, many velocities


At temperatures above absolute zero, all molecules are in motion. In the case of a gas, this motion consists of straight-line jumps
whose lengths are quite great compared to the dimensions of the molecule. Although we can never predict the velocity of a
particular individual molecule, the fact that we are usually dealing with a huge number of them allows us to know what fraction of
the molecules have kinetic energies (and hence velocities) that lie within any given range.
The trajectory of an individual gas molecule consists of a series of straight-line paths interrupted by collisions. What happens when
two molecules collide depends on their relative kinetic energies; in general, a faster or heavier molecule will impart some of its
kinetic energy to a slower or lighter one. Two molecules having identical masses and moving in opposite directions at the same
speed will momentarily remain motionless after their collision.
If we could measure the instantaneous velocities of all the molecules in a sample of a gas at some fixed temperature, we would
obtain a wide range of values. A few would be zero, and a few would be very high velocities, but the majority would fall into a
more or less well defined range. We might be tempted to define an average velocity for a collection of molecules, but here we
would need to be careful: molecules moving in opposite directions have velocities of opposite signs. Because the molecules are in a
gas are in random thermal motion, there will be just about as many molecules moving in one direction as in the opposite direction,
so the velocity vectors of opposite signs would all cancel and the average velocity would come out to zero. Since this answer is not
very useful, we need to do our averaging in a slightly different way.
The proper treatment is to average the squares of the velocities, and then take the square root of this value. The resulting quantity is
known as the root mean square (RMS) velocity
−−−−−
2
∑ν
vrms =√ (2.7.1)
n

where n is the number of molecules in the system. The formula relating the RMS velocity to the temperature and molar mass is
surprisingly simple (derived below), considering the great complexity of the events it represents:
−−−−−
3RT
vrms = √ (2.7.2)
M

where
M is the molar mass in kg mol–1, and
R is gas constant.

Equation 2.7.2 can also be expressed as


−−−− −
3 kb T
vrms = √ (2.7.3)
m

where
m is the molecular mass in kg
kb is Boltzmann constant and is just the “gas constant per molecule"
R R
kb = = (2.7.4)
23
Na 6.02 × 10

Equation 2.7.3 is just the per atom version of Equation 2.7.2 which is expressed in terms of per mol. Either equation will work.

2.7.1 https://chem.libretexts.org/@go/page/41409
Example 2.7.1

What is the vrms of a nitrogen molecule at 300 K?


Solution
The molar mass of N is 28.01 g/mol. Substituting in the above equation and expressing R in energy units, we obtain
2

−1 −1
(3)(8.31 J mol K )(300 K) J
2 5
v = = 2.67 × 10
−3 −1
28.01 × 10 Kg mol Kg

Recalling the definition of the joule (1 J = 1 kg m2 s–2) and taking the square root,
−−−−−−−−−−−−−−−−−−−−−−−−−−−−− −
 2 –2
⎛ J ⎞ ⎛ 1 kg m s ⎞
 5
v̄ = 2.67 × 10 = 517 m/s
⎷⎝
Kg ⎠ ⎝ 1 J ⎠

or

m 1 km 3600 s
517 ( )( ) = 1860 km/hr
3
s 10 m 1 hr

Comment: this is fast! The velocity of a rifle bullet is typically 300-500 m s–1; convert to common units to see the comparison
for yourself. A simpler formula for estimating average molecular velocities than 2.7.2 is
−−

T
vrms = 157 √
M

in which v is in units of meters/sec, T is the absolute temperature and M the molar mass in grams.

The Boltzmann Distribution


If we were to plot the number of molecules whose velocities fall within a series of narrow ranges, we would obtain a slightly
asymmetric curve known as a velocity distribution. The peak of this curve would correspond to the most probable velocity. This
velocity distribution curve is known as the Maxwell-Boltzmann distribution, but is frequently referred to only by Boltzmann's
name. The Maxwell-Boltzmann distribution law was first worked out around 1850 by the great Scottish physicist, James Clerk
Maxwell (left, 1831-1879), who is better known for discovering the laws of electromagnetic radiation. Later, the Austrian physicist
Ludwig Boltzmann (1844-1906) put the relation on a sounder theoretical basis and simplified the mathematics somewhat.
Boltzmann pioneered the application of statistics to the physics and thermodynamics of matter, and was an ardent supporter of the
atomic theory of matter at a time when it was still not accepted by many of his contemporaries.

Figure 2.7.1 : Maxwell (left) and Boltzman (right) are responsible for the velocity distrubtion of gas molecules

2.7.2 https://chem.libretexts.org/@go/page/41409
The Maxwell-Boltzmann distribution is used to determine how many molecules are moving between velocities v and v + dv .
Assuming that the one-dimensional distributions are independent of one another, that the velocity in the y and z directions does not
affect the x velocity, for example, the Maxwell-Boltzmann distribution is given by
1/2 2
dN m −mv
= ( ) exp [ ] dv (2.7.5)
N 2πkb T 2kb T

where
dN /N is the fraction of molecules moving at velocity v to v + dv ,
m is the mass of the molecule,
kb is the Boltzmann constant, and
T is the absolute temperature.1
Additionally, the function can be written in terms of the scalar quantity speed v instead of the vector quantity velocity. This form of
the function defines the distribution of the gas molecules moving at different speeds, between v and v , thus
1 2

3/2 2
m −mv
2
f (v) = 4π v ( ) exp [ ] (2.7.6)
2π kb T 2 kb T

Finally, the Maxwell-Boltzmann distribution can be used to determine the distribution of the kinetic energy of for a set of
molecules. The distribution of the kinetic energy is identical to the distribution of the speeds for a certain gas at any temperature.2
The Maxwell-Boltzmann distribution is a probability distribution and just like any such distribution, can be characterized in a
variety of ways including.
Average Speed: The average speed is the sum of the speeds of all of the particles divided by the number of particles.
Most Probable Speed: The most probable speed is the speed associated with the highest point in the Maxwell distribution.
Only a small fraction of particles might have this speed, but it is more likely than any other speed.
Width of the Distribution: The width of the distribution characterizes the most likely range of speeds for the particles. One
measure of the width is the Full Width at Half Maximum (FWHM). To determine this value, find the height of the distribution
at the most probable speed (this is the maximum height of the distribution). Divide the maximum height by two to obtain the
half height, and locate the two speeds in the distribution that have this half-height value. On speed will be greater than the most
probably speed and the other speed will be smaller. The full width is the difference between the two speeds at the half-
maximum value.

Velocity distributions depend on temperature and mass


Higher temperatures allow a larger fraction of molecules to acquire greater amounts of kinetic energy, causing the Boltzmann plots
to spread out. Figure 2.7.2 shows how the Maxwell-Boltzmann distribution is affected by temperature. At lower temperatures, the
molecules have less energy. Therefore, the speeds of the molecules are lower and the distribution has a smaller range. As the
temperature of the molecules increases, the distribution flattens out. Because the molecules have greater energy at higher
temperature, the molecules are moving faster.

Figure 2.7.2 : The Maxwell Distribution as a function of temperature for nitrogen molecules
Notice how the left ends of the plots are anchored at zero velocity (there will always be a few molecules that happen to be at rest.)
As a consequence, the curves flatten out as the higher temperatures make additional higher-velocity states of motion more

2.7.3 https://chem.libretexts.org/@go/page/41409
accessible. The area under each plot is the same for a constant number of molecules.
All molecules have the same kinetic energy (mv2/2) at the same temperature, so the fraction of molecules with higher velocities will
increase as m, and thus the molecular weight, decreases. Figure 2.7.3 shows the dependence of the Maxwell-Boltzmann
distribution on molecule mass. On average, heavier molecules move more slowly than lighter molecules. Therefore, heavier
molecules will have a smaller speed distribution, while lighter molecules will have a speed distribution that is more spread out.

Figure 2.7.3 : The speed probability density functions of the speeds of a few gases at a temperature of 298.15 K ). The y-axis is in
s/m so that the area under any section of the curve (which represents the probability of the speed being in that range) is
dimensionless.
The Maxwell-Boltzmann equation, which forms the basis of the kinetic theory of gases, defines the distribution of speeds for a gas
at a certain temperature. From this distribution function, the most probable speed, the average speed, and the root-mean-square
speed can be derived.

Related Speed Expressions


Usually, we are more interested in the speeds of molecules rather than their component velocities. The Maxwell–Boltzmann
distribution for the speed follows immediately from the distribution of the velocity vector, above. Note that the speed of an
individual gas particle is:
−−−−−−−−−−
2 2 2
v = √ vx + vy = vz (2.7.7)

Three speed expressions can be derived from the Maxwell-Boltzmann distribution:


the most probable speed,
the average speed, and
the root-mean-square speed.
The most probable speed is the maximum value on the distribution plot (Figure 2.7.4}. This is established by finding the velocity
when the derivative of Equation 2.7.6 is zero
df (v)
=0 (2.7.8)
dv

which is
−−−−−
2RT
vmp = √ (2.7.9)
M

2.7.4 https://chem.libretexts.org/@go/page/41409
Figure 2.7.4 : The Maxwell-Boltzmann distribution is shifted to higher speeds and is broadened at higher temperatures. from
OpenStax. The speed at the top of the curve is called the most probable speed because the largest number of molecules have that
speed.
The average speed is the sum of the speeds of all the molecules divided by the number of molecules.

−−−−−
8RT
vavg = v̄ = ∫ vf (v)dv = √ (2.7.10)
0
πM

The root-mean-square speed is square root of the average speed-squared.


−−−−−
¯2 3RT
vrms =v = √ (2.7.11)
M

where
Ris the gas constant,
Tis the absolute temperature and
M is the molar mass of the gas.

It always follows that for gases that follow the Maxwell-Boltzmann distribution:
vmp < vavg < vrms (2.7.12)

Problems
1. Using the Maxwell-Boltzman function, calculate the fraction of argon gas molecules with a speed of 305 m/s at 500 K.
2. If the system in problem 1 has 0.46 moles of argon gas, how many molecules have the speed of 305 m/s?
3. Calculate the values of C , C , and C
mp avg rms for xenon gas at 298 K.
4. From the values calculated above, label the Boltzmann distribution plot (Figure 1) with the approximate locations of
(C_{mp}\), C , and C .
avg rms

5. What will have a larger speed distribution, helium at 500 K or argon at 300 K? Helium at 300 K or argon at 500 K? Argon at
400 K or argon at 1000 K?

Answers
1. 0.00141
2. 3.92 × 10 argon molecules
20

3. cmp = 194.27 m/s, cavg = 219.21 m/s, crms = 237.93 m/s


4. As stated above, Cmp is the most probable speed, thus it will be at the top of the distribution curve. To the right of the most
probable speed will be the average speed, followed by the root-mean-square speed.
5. Hint: Use the related speed expressions to determine the distribution of the gas molecules: helium at 500 K. helium at at 300 K.
argon at 1000 K.

References
1. Dunbar, R.C. Deriving the Maxwell Distribution J. Chem. Ed. 1982, 59, 22-23.
2. Peckham, G.D.; McNaught, I.J.; Applications of the Maxwell-Boltzmann Distribution J. Chem. Ed. 1992, 69, 554-558.
3. Chang, R. Physical Chemistry for the Biosciences, 25-27.

2.7.5 https://chem.libretexts.org/@go/page/41409
Contributors and Attributions
Prof. David Blauch (Davidson College)
Stephen Lower, Professor Emeritus (Simon Fraser U.) Chem1 Virtual Textbook
Adam Maley (Hope College)

2.7: The Maxwell Distribution Laws is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

2.7.6 https://chem.libretexts.org/@go/page/41409
2.8: Molecular Collisions and the Mean Free Path
Collision theory is a theory proposed independently by Max Trautz in 1916 and William Lewis in 1918, that qualitatively explains
how chemical reactions occur and why reaction rates differ for different reactions. The collision theory states that when suitable
particles of the reactant hit each other, only a certain percentage of the collisions cause any noticeable or significant chemical
change; these successful changes are called successful collisions. The successful collisions have enough energy, also known as
activation energy, at the moment of impact to break the preexisting bonds and form all new bonds. This results in the products of
the reaction. Increasing the concentration of the reactant particles or raising the temperature, thus bringing about more collisions
and therefore many more successful collisions, increases the rate of reaction.

Collision Energy
Consider two particles A and B in a system. The kinetic energy of these two particles is
2 2
p p
A B
KAB = + (2.8.1)
2mA 2mB

Let us change to center-of-mass (P) and relative (p) momenta, which are given by
mB pa − mA pB
P = pA + pB , p = (2.8.2)
M

where M = mA + mB is the total mass of the two particles. Substituting this into the kinetic energy, we find
2 2 2 2
p p P p
A B
KAB = + = + (2.8.3)
2mA 2mB 2M 2μ

where
mA mB
μ = (2.8.4)
M

is called the reduced mass of the two particles. Note that the kinetic energy separates into a sum of a center-of-mass term and a
relative term.
Now the relative position is r = r − r
A B so that the relative velocity is ṙ = ṙ − ṙ or v = v − v . Thus, if the two particles
A B A B

are approaching each other such that v A = −vB , then v = 2v . However, by equipartitioning the relative kinetic energy, being
A

mass independent, is
2
p 3
⟨ ⟩ = kB T (2.8.5)
2μ 2

which is called the collision energy.

Collision cross section


Consider two molecules in a system. The probability that they will collide increases with the effective “size” of each particle.
However, the size measure that is relevant is the apparent cross-section area of each particle. For simplicity, suppose the particles
are spherical, which is not a bad approximation for small molecules. If we are looking at a sphere, what we perceive as the size of
the sphere is the cross section area of a great circle. Recall that each spherical particle has an associated “collision sphere” that just
encloses two particles at closest contact, i.e., at the moment of a collision, and that this sphere is a radius d , where d is the diameter
of each spherical particle (see lecture 5). The cross-section of this collision sphere represents an effective cross section for each
particle inside which a collision is imminent. The cross-section of the collision sphere is the area of a great circle, which is πd . We 2

denote this apparent cross section area σ. Thus, for spherical particles A and B with diameters d and d , the individual cross
A B

sections are
2 2
σA = π d , σB = π d (2.8.6)
A B

The collision cross section, σ AB is determined by an effective diameter d characteristic of both particles. The collision
AB

probability increases of both particles have large diameters and decreases if one of them has a smaller diameter than the other.
Hence, a simple measure sensitive to this is the arithmetic average

2.8.1 https://chem.libretexts.org/@go/page/41410
1
dAB = (dA + dB ) (2.8.7)
2

and the resulting collision cross section becomes


2
σAB = πd (2.8.8)
AB

2
dA + dB
= π( ) (2.8.9)
2
π
2 2
= (d + 2 dA dB + d ) (2.8.10)
A B
4
1
−−−−−
= (σA + 2 √σA σB + σB ) (2.8.11)
4

1 σA + σB
−−−−−
= [( ) + √σA σB ] (2.8.12)
2 2

which, interestingly, is an average of the two types of averages of the two individual cross sections, the arithmetic and geometric
averages!

Average collision Frequency


Consider a system of particles with individual cross sections σ. A particle of cross section σ that moves a distance l in a time Δt
will sweep out a cylindrical volume (ignoring the spherical caps) of volume σl (Figure 2.8.1). If the system has a number density
ρ, then the number of collisions that will occur is

Ncoll = ρσl (2.8.13)

Figure 2.8.1 : Collision cylinder. Any particle that partially overlaps with this volume will experience a collision with a test particle
tracing out this volume.
We define the average collision rate as N coll /Δt, i.e.,
Ncoll ρσl
γ = = = ρσ⟨|v|⟩ (2.8.14)
Δt Δt

where ⟨|v|⟩ is the average relative speed. If all of the particles are of the same type (say, type A ), then performing the average over
a Maxwell-Boltzmann speed distribution gives
−−−−−
8 kB T
⟨|v|⟩ = √ (2.8.15)
πμ

where μ = m A /2 is the reduced mass. The average speed of a particle is


−−−−−
8 kB T
⟨| vA |⟩ = √ (2.8.16)
πmA

so that

⟨|v|⟩ = √2⟨| vA |⟩ (2.8.17)

Mean Free Path


The mean free path is defined as the distance a particle will travel, on average, before experiencing a collision event. This is

defined as the product of the speed of a particle and the time between collisions. The former is ⟨|v|⟩/√2, while the latter is 1/γ.

2.8.2 https://chem.libretexts.org/@go/page/41410
Hence, we have
⟨|v|⟩ 1
λ = = (2.8.18)
– –
√2ρσ⟨|v|⟩ √2ρσ

Random Walks
In any system, a particle undergoing frequent collisions will have the direction of its motion changed with each collision and will
trace out a path that appears to be random. In fact, if we treat the process as statistical, then, we are, in fact, treating each collision
event as a random event, and the particle will change its direction at random times in random ways! Such a path might appear as
shown in Figure \(\PageIndex{2. Such a path is often referred to as a random walk path.

Figure 2.8.2 : Random walk path. The Path of a Single Particle in a Gas Sample. The frequent changes in direction are the result of
collisions with other gas molecules and with the walls of the container.
In order to analyze such paths, let us consider a random walk in one dimension. We’ll assume that the particle move a mean-free
path length λ between collisions and that each collision changes the direction of the particles motion, which in one dimension,
means that the particle moves either to the right or to the left after each event. This can be mapped onto a metaphoric “coin toss”
that can come up heads “H” or tails “T”, with “H” causing motion to the right, and “T” causing motion to the left.
In three dimensions, we consider the three spatial directions to be independent, hence, the probability distribution for a particle to
diffuse to a location r = (x, y, z) is just a product of the three one-dimensional distributions:
1 2 2 2
−(x +y +z )/4Dt
P(r) = P (x) P (y) P (z) = e (2.8.19)
3/2
(4πDt)

where D is the diffusion constant. If we are only interested in diffusion over a distance r, we can introduce spherical coordinates,
integrate over the angles, and we find that
4π 2
−r /4Dt
P (r, t) = E (2.8.20)
3/2
(4πDt)

Rates of Diffusion or Effusion


Graham’s law is an empirical relationship that states that the ratio of the rates of diffusion or effusion of two gases is the square
root of the inverse ratio of their molar masses. The relationship is based on the postulate that all gases at the same temperature have
the same average kinetic energy. We can write the expression for the average kinetic energy of two gases with different molar
masses:
1 MA 1 MB
2 2
KE = v = v (2.8.21)
rms,A rms,B
2 NA 2 NA

Multiplying both sides by 2 and rearranging give


2
v
rms,B MA
= (2.8.22)
2
v MB
rms,A

Taking the square root of both sides gives

2.8.3 https://chem.libretexts.org/@go/page/41410
−−−−
vrms,B MA
=√ (2.8.23)
vrms,A MB

Thus the rate at which a molecule, or a mole of molecules, diffuses or effuses is directly related to the speed at which it moves.
Equation 2.8.23 shows that Graham’s law is a direct consequence of the fact that gaseous molecules at the same temperature have
the same average kinetic energy.

Figure 2.8.3 : The Wide Variation in Molecular Speeds Observed at 298 K for Gases with Different Molar Masses
The lightest gases have a wider distribution of speeds and the highest average speeds.

Molecules with lower masses have a wider distribution of speeds and a higher average speed.

Gas molecules do not diffuse nearly as rapidly as their very high speeds might suggest. If molecules actually moved through a
room at hundreds of miles per hour, we would detect odors faster than we hear sound. Instead, it can take several minutes for us to
detect an aroma because molecules are traveling in a medium with other gas molecules. Because gas molecules collide as often as
1010 times per second, changing direction and speed with each collision, they do not diffuse across a room in a straight line.

The denser the gas, the shorter the mean free path.

Example 2.8.2

Calculate the rms speed of a sample of cis-2-butene (C4H8) at 20°C.


Given: compound and temperature
Asked for: rms speed
Strategy:
Calculate the molar mass of cis-2-butene. Be certain that all quantities are expressed in the appropriate units and then use
Equation ??? to calculate the rms speed of the gas.
Solution:
To use Equation 2.8.23, we need to calculate the molar mass of cis-2-butene and make sure that each quantity is expressed in
the appropriate units. Butene is C4H8, so its molar mass is 56.11 g/mol. Thus

2.8.4 https://chem.libretexts.org/@go/page/41410
−−−−−−−−−−−−−−−−−−−−−−−−−−− −

 J
−−−−−  3 × 8.3145 × (20 + 273) K
3RT K ⋅ mol

urms =√ = = 361 m/s (2.8.24)
M ⎷ −3
56.11 × 10 kg

or approximately 810 mi/h.

Exercise 2.8.2

Calculate the rms speed of a sample of radon gas at 23°C.


Answer: 1.82 × 102 m/s (about 410 mi/h)

Contributors and Attributions


Mark Tuckerman (New York University)
Wikipedia

2.8: Molecular Collisions and the Mean Free Path is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.

2.8.5 https://chem.libretexts.org/@go/page/41410
2.9: Graham's Laws of Diffusion and Effusion
Learning Objectives
Understand the difference between effusion and diffusion
To derive and apply Graham's Law of Effusion

Diffusion is the gradual mixing of gases due to the motion of their component particles even in the absence of mechanical agitation
such as stirring. The result is a gas mixture with uniform composition. Diffusion is also a property of the particles in liquids and
liquid solutions and, to a lesser extent, of solids and solid solutions. The related process, effusion, is the escape of gaseous
molecules through a small (usually microscopic) hole, such as a hole in a balloon, into an evacuated space.
The phenomenon of effusion had been known for thousands of years, but it was not until the early 19th century that quantitative
experiments related the rate of effusion to molecular properties. The rate of effusion of a gaseous substance is inversely
proportional to the square root of its molar mass. This relationship is referred to as Graham’s law, after the Scottish chemist
Thomas Graham (1805–1869). The ratio of the effusion rates of two gases is the square root of the inverse ratio of their molar
masses:
−−−−
rate of effusion A MB
=√ (2.9.1)
rate of effusion B MA

Graham Law
The rate of effusion of a gaseous substance is inversely proportional to the square root of its molar mass.

Graham’s law is an empirical relationship that states that the ratio of the rates of diffusion or effusion of two gases is the square
root of the inverse ratio of their molar masses. The relationship is based on the postulate that all gases at the same temperature have
the same average kinetic energy (recall that a result of the Kinetic Theory of Gases is that the temperature, in degrees Kelvin, is
directly proportional to the average kinetic energy of the molecules.) . We can write the expression for the average kinetic energy of
two gases with different molar masses:
1 MA 1 MB
2 2
KE = v = v (2.9.2)
rms,A rms,B
2 NA 2 NA

Multiplying both sides by 2 and rearranging give


2
v MA
rms,B
= (2.9.3)
2
v MB
rms,A

Taking the square root of both sides gives


−−−−
vrms,B MA
=√ (2.9.4)
vrms,A MB

Thus the rate at which a molecule, or a mole of molecules, diffuses or effuses is directly related to the speed at which it moves.
Equation 2.9.4 shows that Graham’s law is a direct consequence of the fact that gaseous molecules at the same temperature have
the same average kinetic energy.

2.9.1 https://chem.libretexts.org/@go/page/41411
Figure 2.9.1 : The Relative Rates of Effusion of Two Gases with Different Masses. The lighter He atoms ( M = 4.00 g/mol) effuse
through the small hole more rapidly than the heavier ethylene oxide (C2H4O) molecules (M = 44.0 g/mol), as predicted by
Graham’s law.
Heavy molecules effuse through a porous material more slowly than light molecules, as illustrated schematically in Figure 2.9.1 for
ethylene oxide C H O and helium H e. Helium (M = 4.00 g/mol) effuses much more rapidly than ethylene oxide (M = 44.0
2 4

g/mol). Because helium is less dense than air, helium-filled balloons “float” at the end of a tethering string. Unfortunately, rubber
balloons filled with helium soon lose their buoyancy along with much of their volume. In contrast, rubber balloons filled with air
tend to retain their shape and volume for a much longer time. Because helium has a molar mass of 4.00 g/mol, whereas air has an
−−−
average molar mass of about 29 g/mol, pure helium effuses through the microscopic pores in the rubber balloon √
29

4.00
= 2.7

times faster than air. For this reason, high-quality helium-filled balloons are usually made of Mylar, a dense, strong, opaque
material with a high molecular mass that forms films that have many fewer pores than rubber. Hence, mylar balloons can retain
their helium for days.

Example 2.9.1

During World War II, scientists working on the first atomic bomb were faced with the challenge of finding a way to obtain
large amounts of U . Naturally occurring uranium is only 0.720% U , whereas most of the rest (99.275%) is U , which
235 235 238

is not fissionable (i.e., it will not break apart to release nuclear energy) and also actually poisons the fission process. Because
both isotopes of uranium have the same reactivity, they cannot be separated chemically. Instead, a process of gaseous effusion
was developed using the volatile compound U F (boiling point = 56°C).
6

1. Calculate the ratio of the rates of effusion of 235UF6 and 238UF6 for a single step in which UF6 is allowed to pass through a
porous barrier. (The atomic mass of 235U is 235.04, and the atomic mass of 238U is 238.05.)
2. If n identical successive separation steps are used, the overall separation is given by the separation in a single step (in this
case, the ratio of effusion rates) raised to the nth power. How many effusion steps are needed to obtain 99.0% pure 235UF6?
Given: isotopic content of naturally occurring uranium and atomic masses of 235U and 238U
Asked for: ratio of rates of effusion and number of effusion steps needed to obtain 99.0% pure 235UF6
Strategy:
A. Calculate the molar masses of 235UF6 and 238UF6, and then use Graham’s law to determine the ratio of the effusion rates.
Use this value to determine the isotopic content of 235UF6 after a single effusion step.
B. Divide the final purity by the initial purity to obtain a value for the number of separation steps needed to achieve the
desired purity. Use a logarithmic expression to compute the number of separation steps required.
Solution:
1. A The first step is to calculate the molar mass of UF6 containing 235U and 238U. Luckily for the success of the separation
method, fluorine consists of a single isotope of atomic mass 18.998. The molar mass of 235UF6 is 234.04 + (6)(18.998) =
349.03 g/mol
The molar mass of 238UF6 is
238.05 + (6)(18.998) = 352.04 g/mol
The difference is only 3.01 g/mol (less than 1%). The ratio of the effusion rates can be calculated from Graham’s law using
Equation 2.9.1:

2.9.2 https://chem.libretexts.org/@go/page/41411
−−−−−−−−−−−
235
rate UF6 352.04 g/mol
=√ = 1.0043 (2.9.5)
238
rate UF6 349.03 g/mol

2.

Thus passing UF6 containing a mixture of the two isotopes through a single porous barrier gives an enrichment of 1.0043,
so after one step the isotopic content is (0.720%)(1.0043) = 0.723% 235UF6.
3. B To obtain 99.0% pure 235UF6 requires many steps. We can set up an equation that relates the initial and final purity to the
number of times the separation process is repeated: final purity = (initial purity)(separation)n
In this case, 0.990 = (0.00720)(1.0043)n, which can be rearranged to give
0.990
n
1.0043 = = 137.50 (2.9.6)
0.00720

4. Taking the logarithm of both sides gives


n ln(1.0043) = ln(137.50) (2.9.7)

ln(137.50)
n = = 1148 (2.9.8)
ln(1.0043)

Thus at least a thousand effusion steps are necessary to obtain highly enriched 235U. Figure 2.9.2 shows a small part of a
system that is used to prepare enriched uranium on a large scale.

Figure 2.9.2 : A Portion of a Plant for Separating Uranium Isotopes by Effusion of UF6. The large cylindrical objects (note the
human for scale) are so-called diffuser (actually effuser) units, in which gaseous UF6 is pumped through a porous barrier to
partially separate the isotopes. The UF6 must be passed through multiple units to become substantially enriched in 235U.

Exercise 2.9.1

Helium consists of two isotopes: 3He (natural abundance = 0.000134%) and 4He (natural abundance = 99.999866%). Their
atomic masses are 3.01603 and 4.00260, respectively. Helium-3 has unique physical properties and is used in the study of
ultralow temperatures. It is separated from the more abundant 4He by a process of gaseous effusion.
a. Calculate the ratio of the effusion rates of 3He and 4He and thus the enrichment possible in a single effusion step.
b. How many effusion steps are necessary to yield 99.0% pure 3He?
Answer: a. ratio of effusion rates = 1.15200; one step gives 0.000154% 3He; b. 96 steps

Gas molecules do not diffuse nearly as rapidly as their very high speeds might suggest. If molecules actually moved through a
room at hundreds of miles per hour, we would detect odors faster than we hear sound. Instead, it can take several minutes for us to
detect an aroma because molecules are traveling in a medium with other gas molecules. Because gas molecules collide as often as
1010 times per second, changing direction and speed with each collision, they do not diffuse across a room in a straight line.

2.9.3 https://chem.libretexts.org/@go/page/41411
Summary
Gaseous particles are in constant random motion. Gaseous particles tend to undergo diffusion because they have kinetic energy.
Diffusion is faster at higher temperatures because the gas molecules have greater kinetic energy. Effusion refers to the movement of
gas particles through a small hole. Graham's Law states that the effusion rate of a gas is inversely proportional to the square root of
the mass of its particles.

2.9: Graham's Laws of Diffusion and Effusion is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.

2.9.4 https://chem.libretexts.org/@go/page/41411
2.E: Properties of Gases (Exercises)
2.1: Some Definitions
2.2: An Operational Definition of Temperature

2.3: Ideal Gases


2.4: Real Gases
2.5: Condensation of Gases and the Critical State
Q2.2
There are some gases in existence, such as NS and NCl , which do not follow Boyle's Law, regardless of the pressure conditions
2 2

they're under. Explain why this is the case.

S2.2
Boyle's Law assumes that the only factors changing are pressure (P ) and volume (V ), therefore, in order to apply Boyle's Law,
both number of molecules (n ) and temperature (T ) must remain constant. However, both N S and N C l do not have constant n
2 2

because they undergo association reactions when in contact with like molecules.
2N H2 ⇌ N2 S2 (2.E.1)

2N C l2 ⇌ N2 C l2 (2.E.2)

Therefore, Boyle's Law may not be applied for these gases.

Q2.3
An ideal gas originally at 0.85 atm and 66ºC was allowed to expand until its final volume, pressure, and temperature were 94 mL,
0.60 atm, and 45ºC, respectively. What was its initial volume?

S2.3
P1 = 0.85 atm

T1 = 339 K

V2 = 94 mL

P2 = 0.60 atm

T2 = 318 K

(P2 )(V2 )(T1 )


V1 = (2.E.3)
(T2 )(P1 )

(0.60 atm)(94 mL)(339 K)


= = 71 mL (2.E.4)
(318 K)(0.85 atm)

Q2.4
Some ballpoint pens have a small hole in the main body of the pen. What is the purpose of the hole?

S2.4
The hole is there to allow the pressure to equalize and allow the ink to flow out of the pen.

Q2.5
Starting with the ideal-gas equation, show how you ca calculate the molar mass of a gas from a knowledge of its density

S2.5

2.E.1 https://chem.libretexts.org/@go/page/41328
Q2.6
At STP (standard temperature and pressure), 0.280 /;L of a gas weighs 0.400/;g. Calculate the molar mass of the gas.

S2.6
First, we need to determine the moles of gas in the sample:
(P )(V )
n = (2.E.5)
(R)(T )

(1.00 atm)(0.280 L)
= = 0.0125 moL (2.E.6)
(0.08206 L − atm/mol − K)(273 K)

(0.400 g)
= = 32.0 g/moL (2.E.7)
(0.0125 mol)

Q2.7
Ozone molecules in the stratosphere absorb much of the harmful radiation from the sun. Typically, the temperature and partial
pressure of ozone in the stratosphere are 250 K and 0.0010 atm, respectively. How many ozone molecules are present in 1.0 L of
air under these conditions? Assume ideal-gas behavior.

Q2.9
Dissolving 3.00 g of an impure sample of CaCO3 in an excess of HCl acid produced 0.656 L of CO2 (measured at 20ºC and 792
mmHg). Calculate the percent by mass of CaCO3 in the sample.

Q2.14
a. What volume of air at 1.0 atm and 22ºC is needed to fill a 0.98 L bicycle tire to a pressure of 5.0 atm at the same temperature?
(Note that the 5.0 atm is the gauge pressure, which is the difference between the pressure in the tire and the atmospheric
pressure. Initially, the gauge pressure in the tire was 0 atm.)
b. What is the total pressure in the tire when the gauge reads 5.0 atm?
c. The tire is pumped with a hand pump full of air at 1.0 atm; compressing the gas in the cylinder adds all of the air in the pump to
the air in the tire. If the volume of the pump is 33% of the tire's volume, what is the gauge pressure in the tire after 2 full strokes
of the pump?

Q2.15
A student breaks a thermometer and spills most of the mercury (Hg) onto the floor of a laboratory that measures 15.2 m long, 6.6 m
wide, and 2.4 m high. The vapor pressure of mercury at 20ºC is 1.7 x 10-6 atm.
a. Calculate the mass of mercury vapor (in grams)in the room at 20ºC.
b. Does the concentration of mercury vapor exceed the air quality regulation of 0.050 mm Hg m-3 of air?
c. One way to treat small quantities of spilled mercury is to spray powdered sulfur over the metal. Suggest a physical and a
chemical reason for this treatment.

S2.15
a) The volume of the room is 15.2 m x 6.6 m x 2.4 m = 240 m3.
The pressure is 1.7 x 10-6 atm x 101325 Pa/atm = 0.17 Pa
Ideal gas equation:
P V = nRT (2.E.8)

PV
n = (2.E.9)
RT

3
0.17 P a ∗ (240) m
n = (2.E.10)
3
8.314 P a − m
∗ (293K)
K − mol

= 0.017 mole (2.E.11)

2.E.2 https://chem.libretexts.org/@go/page/41328
Mass of mercury vapor : mHg = nHg * MHg =(0.017 mole)(200.6 g/mole)(1000 mg/gram)} = 3400 mg
b) The mass density of mercury vapor :
[Hg] = mHg / VHg(l) =
3400 mg
(2.E.12)
−3
240 m

=14 mg/m3
This value is indeed larger than the 0.050 mg Hg m-3 allowed.
c) Using powdered sulfur to absorb the spilled mercury works because the sulfur molecules will strongly bind to the mercury, and
the mercury sulfide that forms is a solid. This makes removing spilled mercury (a liquid) easier.

Q2.16
An oxide has the density of 1.75 g L-1 at 1.12 atm and 100ºC. Is this oxide O2(g) or O3(g)?

S2.16
ρRT
M = (2.E.13)
P

−1 −1 −1
(1.75gL )(0.08206Latm K mol )(100 + 273)K) g
= = 47.8 (2.E.14)
1.12atm mol

The molar mass of O3(g) is 48 g mol-1, so the oxide must be O3(g)

Q2.21
You are given a gas at 750 torr with a volume of 0.62 Liters at 25°C. A while later, you measure the temperature and notice a 5°C
temperature increase in the system. Assuming there is no pressure change and that this is an ideal gas, what is the new volume?

S2.21
First: List out what you have

Vi=0.62 L
Ti=25ºC +273.15= 298.15
We do not know the amount of moles so we should find that.
PV=nRT

So ni=2.47x10-4 moles
Pf=0.987 atm
Vf=?
Tf=298.15 K + 5 K= 303.15 K
Vf=(nRT)/P

=0.631 L

2.E.3 https://chem.libretexts.org/@go/page/41328
Q2.23a
The composition of air in your alveoli is 75% N2, 14% O2, 5% CO2, and 6% H2O, by volume. (a) Calculate the average molar mass
of this air sample. (b) Calculate the partial pressures of N2, O2, CO2, and H2O in your alveoli. assuming that alveolar pressure is
760 torr (At constant pressure and temperature, the volume of a gas is directly proportional to the number of moles of gas).

Q2.23b
The composition of air in your alveoli is 75% N2, 14% O2, 5% CO2, and 6% H2O, by volume. (a) Calculate the average molar mass
of this air sample. (b) Calculate the partial pressures of N2, O2, CO2, and H2O in your alveoli. (At constant pressure and
temperature, the volume of a gas is directly proportional to the number of moles of gas.)

Q2.24
In an expandable balloon we find a mixture of two gases only. The mass percent of the gases are: 65% 02 and 35 % HQ2. The mass
of the gases combined is 4.6 grams. Calculate the total volume these two gases occupy at 298 K and 1.00 atm.

S2.24
You can find the Volume by using the Ideal Gas Law equation:
nRT
V = (2.E.15)
P

The mass of the two gases is given by:


mH2 = 4.6 × 0.35 = 1.6g

mO = 4.6 × 0.65 = 3.0g


2

The number of moles of the mixture= n total = nH2 + nO2

1.60g 3.0g
ntotal = + = 0.884mol
−1 −1
Q2.0g. mol 32g. mol

L. atm
(0.884mol)(0.08206 )(298K)
K. mol
The volume the two gases occupy = V = = 21.6L
1.0atm

Q2.25
The temperature in a small room in summer time is 31oC. The room (dimension 4mx4mx3m) is filled with 0.88 kg of water vapor.
What is the partial pressure of water vapor in the air?

Q2.27
A flask contains a mixture of two ideal gases, A and B. Show graphically how the total pressure of the system depends on the
amount of A present. That is, plot the total pressure versus the mole fraction of A. Do the same for B on the same graph. The total
number of moles of A and B is remained as constant.

S2.27

Q2.29
When you drink something from a straw, how does the liquid get from the glass up through the straw into your mouth?

S2.29
When you close your mouth around the straw, you close off the atmosphere from outside to inside. During suction, you increase the
volume inside your mouth, thus lowering the internal pressure and the atmospheric pressure exerted onto the liquid over comes the
internal pressure and so the liquid flows into your mouth.

Q2.31
Consider the reaction of ferrous sulfide with a dilute acid:
F eS(s) + 2H C l(aq) → H2 S(g) + F eC l2(l) (2.E.16)

2.E.4 https://chem.libretexts.org/@go/page/41328
An amount of 5.10 grams of ferrous sulfide completely reacted in this reaction. The gas collected over the water at 27ºC has a
volume 1.78 L. Find the pressure of the collected gas in atm. The vapor pressure of water at 27°C is 26.7 mmHg.
Hint: Look at Dalton's law of partial pressure

S2.31
Base on the amount of reactant, use stoichiometry to find the number of moles of H2S gas produced and get partial pressure of
hydrogen sulfide gas. Then use Dalton's law of partial pressure to find the total pressure of collected gas.
1moleF eS 1moleH2 S
5.10gF eS × × = 0.058moleH2 S (2.E.17)
87.92gF eS 1moleF eS

L ⋅ atm
0.058moleH2 S × 0.08206 × (27 + 273)K
nRT mole ⋅ K
PH2 S = = = 0.802atm (2.E.18)
V 1.78L

1atm
Ptotal = PH2 S + PH2 O = 0.802atm + (26.7mmH g × ) = 0.837atm (2.E.19)
760mmH g

Q2.32a
Tanks containing mixtures of oxygen and nitrogen are routinely lowered to the Aquarius underwater research lab in Florida. What
is the percent composition of oxygen gas in the tank if the total pressure in the tank once it has reached Aquarius is 6.5 atm, and the
partial pressure is 0.55 atm?
S2.32a
For any ideal gas,
P V = nRT (2.E.20)

At constant pressure and temperature, V is proportional to n. Thus:


P 2 =x 2 Ptotal (2.E.21)
O O

P 2
0.55atm
O
x 2 = = = .085 (2.E.22)
O
Ptotal 6.5atm

The percent composition is 8.5%

Q2.32b
A mixture of helium and oxygen gas is used to fill up an apparatus. If the total pressure within the apparatus is 10 atm and the
partial pressure of helium is 2 atm within that apparatus, calculate the percent by volume of diatomic oxygen gas.

S2.32b
PH e
χH e = , where χ = M ole F raction (2.E.23)
Ptotal

2atm
χH e =
10atm

χH e = 0.2

% by volume H e = χH e ⋅ 100

= 0.2 ⋅ 100 = 20%

100% by volume total gas − 20% by volume H e = 80% by volume O2 (2.E.24)


––––––––––––––––––

Click here for more on mole fractions.

Q2.33
An unknown mass of NO2 in a closed chamber decomposes completely to oxygen and nitrogen gases. The final pressure in the
chamber is 0.687 atm. What are the partial pressures of each component of the gas mixture in the chamber?

2.E.5 https://chem.libretexts.org/@go/page/41328
S2.33
Every two moles of NO2 produce one mole of N2 and O2. By the ideal gas equation, since we know that the gases are at the same
temperature and occupy the same volume, the pressure of each gas is proportional to the number of moles of that gas present, and
thus the pressure of each gas is proportional to the mole fractions of that gas. Because NO2 represents 1/3 of the total moles of gas
present, and O2 represents 2/3 of the total gas present, NO2 represents 1/3 of the total pressure, and O2 represents 2/3 of the total
pressure.
PN2 = 1/3 * 0.687 atm = 0.229 atm
PO2 = 2/3 * 0.687 atm = 0.458 atm

Q2.41
Choose the more ideal system
a. H2O at 110 °C and 1 atm vs H2O at 100°C and I atm
b. Cl2 at 10 atm vs Cl2 at 0.5 atm
c. N2 at 60 °C vs N2 at 100°C

2.6: Kinetic Theory of Gases


Q2.41
List the ideal conditions (the conditions that help allow a gas to behave ideally).

S2.41
Ideal conditions come from the kinetic theory of gases. This theory relies on three primary assumptions (as follows):
a. The gas is made up of a large amount of molecules. These molecules obey Newton's laws pertaining to motion and are always
moving around in random motion.
b. In comparison to the volume of the gas, the volume of the molecules are so negligible that they can essentially be ignored.
c. No forces are capable of acting on the molecules except for extremely short durations of elastic collisions.
Ideal conditions are at:
high temperatures, and
low pressures
This is because at these conditions is when there are relatively large distance between molecules.

Q2.42
1.00 mole of an unknown gas occupies a 0.865 L container at 0.90 atm and 25ºC. If the value of van der Waals constant a= 0.550,
what is the value of b for the gas? What does the value of b indicate about the gas?

S2.42
P = 4.0atm (2.E.25)

V = 0.70L

T = 77.2K
2
atm ⋅ L
a = 2.50
mol2

2
an nRT 1
(P + )(V − nb) = nRT → b = −V ( ) (2.E.26)
2 2
V an −1
P +
2
V

L ⋅ atm
(1mol)(.08206 )(77.2K)
mol ⋅ K 1
b = − .70L( ) (2.E.27)
(Q2.5)1mol2 −1mol
4.0atm +
2
.70L


L
= 4 × 10 3
mol

2.E.6 https://chem.libretexts.org/@go/page/41328
The small value of b indicates that the gas molecules have a small molecular mass.

Q2.48
The compressibility for compound 1 is 0.84 and 0.56 for compound. Which compound will occupy a smaller volume?

S2.48
Z is a prediction of the behavior of gases. A smaller value means that there are stronger intermolecular forces amongst the
molecules, meaning a smaller volume occupied. Therefore, compound 2 will occupy a smaller volume than compound 1.

Q2.50
True or false: The concept of temperature is a macroscopic. Explain your reasoning.

S2.50
True. The kinetic theory of gases hypothesizes that it deals with an extremely large amount of molecules. Therefore, temperature is
a macroscopic concept since the average kinetic energy of the molecules within a given system is proportional to the temperature.
A large amount of molecules must be included into this average in order for it to be properly utilized.

Q2.51
What assumptions are made when applying the kinetic molecular theory to gases? Are all of these assumptions necessary? Why or
why not?

S2.51
There are five assumptions made when applying the kinetic molecular theory to gases are:
1. Gases consist of a large amount of atoms or molecules and these atoms/ molecules have spaces of separation between them that
are much larger than their size.
2. Although the molecules have mass, we treat them as if they do not since their volume is negligibly small.
3. The molecules are always moving around randomly.
4. All collisions, both molecule-molecule collisions and molecule-wall collisions, are considered to be elastic rather than inelastic.
5. The molecules are not considered to be attracted nor repulsed to/ by one another.
Most of these assumptions are in fact necessary to consider when applying the kinetic molecular theory to gases, however,
assumption number 4 is not always crucial. This is due to the fact that it does not matter whether or not the collisions between the
molecules and the walls are elastic if the walls of the container are at the same temperature as the molecules inside of the the
container (i.e. the gas) because although kinetic energy may be transferred from molecule to molecule, it will not be converted into
any other forms of energy.

Q2.52
A 3.50 L cube-shaped container container 3.5x1023 molecules of N at 30ºC. What is pressure exerted on one wall?
2

S2.52
23
N = 3.5 × 10 (2.E.28)

V = 3.50L

T = 303K
−23
kB = 1.38 × 10

2N ¯
¯¯¯
2N 3
P = E t rans → P = ( kB T ) (2.E.29)
3V 3V 2

23
2(3.5 × 10 ) 3
−23
= [ (1.38 × 10 )(303K)] (2.E.30)
3(3.50L) 2

2
= 4.18 × 10 atm (2.E.31)

2.E.7 https://chem.libretexts.org/@go/page/41328
Q2.53
Imagine a container with walls 123 cm³. Argon (with a speed of 700 m/s) atoms at 300 K are colliding at right angles with the
container's walls at a rate of 5.0 × 10 m/s . What would the force and pressure exerted on the container's walls be?
20

2.7: The Maxwell Distribution Laws


Q2.54
Calculate the root-mean-square speed and the molar kinetic energy of N2 at 25oC

S2.54
The root-mean-square speed of N is: 2

−−−− −
3 kB T
vrms = √ (2.E.32)
m

or in terms of molar mass M instead of molecular mass (m)


−−−−−
3RT
vrms = √ (2.E.33)
M

−−−−−−−−−−−−−−−−−−−−−−−−−−−− −

 (3)(8.3145 J K mol−1 )(298.15 K )

= = 5.15 m/s (2.E.34)
 1 kg

⎷ 28.02 g mol−1
1000 ; g

The molar kinetic energy of N at 25oC is:


2

3 8.3145J 1kJ
= ∗ ∗ 298.15K ∗
2 Kmol 1000J

= 3.72 kJ/mol

Q2.55
Find the v rms of N_{2(g)} at 25ºC. What temperature must C l 2(g)
be to have the same v rms ?

S2.55
−−−−−
3RT
vrms = √ (2.E.35)
M

−−−−−−−−−−−−−−−−−−−−−−−−−−−− −
−1 −1
3(8.314J ⋅ K ⋅ mol )(25 + 273)K
−1
vrmsN2 = √ = 515m s (2.E.36)
−3 −1
28.02x 10 kg ⋅ mol

−1
since vrmsN2 = vrmsC l2 , then vrmsC l2 = 515m s (2.E.37)

−−−−−−−−−−−−−−−−−−−−− −
−1 −1
3(8.314J ⋅ K ⋅ mol )(T )
−1
vrmsC l2 = √ = 515m s (2.E.38)
−3 −1
70.9x 10 g ⋅ mol

T = 754 K
Therefore, Cl2(g) must be 754 K to have the same vrms as N2(g) at 298 K

Q2.60
a. Plot the distribution of speeds for Cl2 at 100 K, 300 K, 600 K, and 1000 K.
b. Plot the distribution of speeds for three gases: N2, Cl2, and CH4 at the same temperature.

2.E.8 https://chem.libretexts.org/@go/page/41328
S2.60
a) The plot of the distribution of speeds for Cl2 at 100 K, 300 K, 600 K, and 1000 K

b) The plot of the distribution of N2,Cl2 and CH4 at the same temperature (300K)

Q2.60
Which gas would you expect to move faster at 255K, O2 or I2? And why?

S2.60
We'd expect O2 to move faster. This is because it is lighter. Think of the formula for average speed because it is divided by kg/mol
the smaller the mass, the faster it goes.

Q2.61
According to the Maxwell speed distribution, would O 2(g) have a wider speed distribution at 200 K or 1000 K?

S2.61
At lower temperatures, the speed distribution will be narrower and there will will be a smaller most probable speed while the higher
temperature will have a wider speed distribution and a higher most probable speed. Therefore, O will have a wider speed
2(g)

distribution at 1000 K.

Q2.62a
A CO2 molecule at unknown temperature at sea level is released to travel upward. Assuming that the temperature is constant and
that the molecule doesn’t collide with another molecules and reaches a terminal height of 8.5 meters above sea level, what is the
temperature? Do the same calculation for a H2 atom. [Hint: To calculate the altitude, h, the molecule will travel, equate its kinetic
energy with the potential energy, mgh, where m is the mass and g is the acceleration due to gravity (9.81 m s-2).]

Q2.62b
A CO2 molecule at unknown temperature at sea level is released to travel upward. Assuming that the temperature is constant and
that the molecule doesn’t collide with another molecules and reaches a terminal height of 5000 meters above sea level, what is the
temperature? Do the same calculation for a H2 atom. [Hint: To calculate the altitude, h, the molecule will travel, equate its kinetic
energy with the potential energy, mgh, where m is the mass and g is the acceleration due to gravity (9.81 m s-2).]

2.E.9 https://chem.libretexts.org/@go/page/41328
Q2.63
Calculate the root-mean-square speed in cm/s for the following particles (hint, what is the average speed?): 0.010, 0.043, 0.027,
0.012, 0.041, 0.023, 0.011, 0.004, 0.007, and 0.009 (all in m/s).

Q2.63
The root-mean-square speed of 6 particles is 2.47 ms-1. The speed of 5 of the particles are 1.0, 2, 1.5, 3.0 and Q2.5. Calculate the
unknown speed of the 6th particle. Find the average speed of the 6 particles.

S2.63
The root-mean-square of the ensemble of particles is
−−−−−−−
6 2
∑ c
i=1 i
vrms = √ (2.E.39)
m

−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−
2 2 2 2 2
1.0 + Q 2.0 + 1.5 + 3.0 + 2.5 + c6
2.47 = √ (2.E.40)
6

We solve for c and we get


6

2 −1 2 2 2 2 2 2 2 −2
c = (Q2.47m s ) (6) − (1.0 + 2.0 + 1.5 + 3.0 + 2.5 ) m s (2.E.41)
6

2 2 −2 2 −2
c = 36.6 m s − 22.5 m s (2.E.42)
6

2 2 −2
c = 14.1 m s (2.E.43)
6

−1
c6 = 3.75m s (2.E.44)

The average speed of the 6 particles is


6
∑ ci
i=1
c̄ = (2.E.45)
N

−1
(1.0 + Q2.0 + 1.5 + 3.0 + Q2.5 + 3.75)ms
c̄ = (2.E.46)
6

−1
= 2.30ms (2.E.47)

Q2.64a
Compare the velocities of Oxygen and Hydrogen at room temperature (25oC) by computing their root mean square speed?

Q2.64b
The speeds of six gaseous molecules at a given temperature are 1.0m/s , 1.2m/s, 1.5m/s, 1.7m/s, 2.0m/s, and 2.3m/s.
a. Compare the root-mean-square (rms) speeds and the average speed of those molecules.
b. Is always the average speed of the molecules the smaller of the two? Explain.

S2.64b
a) The root-mean-square-speed of those gases:
\[v_{rms} = \sqrt{\dfrac{1}{N}*(v_1^{2}+v2^{2}+v_3^{2}+v_4^{2}+v_5^{2}+v_6^{2})}\]
−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−− −
1 2 2 2 2 2 2
=√ ∗ (1 + 1.2 + 1.5 + 1.7 + 2 + 2.3 ) (2.E.48)
N

= 1.68 m/s (2.E.49)

The average speed is :


\[v_{avg} = \dfrac{1+1.2+1.5+1.7+2+2.3}{6}\]
= 1.62 m/s (2.E.50)

2.E.10 https://chem.libretexts.org/@go/page/41328
b) The average speed is always smaller then the root-mean-square speed because the average speed depends on the temperature,
rather than the molecular sizes.

Q2.65
Calculate cmp for a sample of N2(g) at 500K.

S2.65
−−−−−
2RT
c mp = √ (2.E.51)
M

−−−−−−−−−−−−−−−−−−−−−−−−
−1 −1
2 (8.314J K mol )(500K)
−1
c mp = √ = 545m s (2.E.52)
−1
.02802kg mol

Q2.66
2
mc
3/2 −
m
Using the initial equation f (c) = 4πc 2
e 2 KB T derive an expression for c mp
π KB T

S2.66
Step 1: $$f(c)=4\pi c^2 \dfrac{m}{\pi K_BT} ^{3/2} e^{-\dfrac{mc^2}{2K_BT}}\]
$$=c^2 \dfrac{m}{\pi K_BT} ^{3/2} e^{-\dfrac{mc^2}{2K_BT}}\]
2
mc

Step 2: (Hint: Apply the chain rule to c and e 2 2 KB T )


2 2
mc mc
3/2 − −
df (c) m mc
2 KB T 2 2 KB T
= 4π [2c e +c e − ] (2.E.53)
dc 2π KB T 2 KB T

Step 3:
2
mc
− 2
m mc
3/2 2 KB T (2 −
4π( ) ce ) (2.E.54)
2π KB T KB T

df (c)
Step 4: (Hint: Set =0 ,c=cmp )
dc

2
mcmp
− 2
m mcmp
3/2 2 KB T (2 −
π( ) cmp e ) =0 (2.E.55)
2π KB T KB T

Step 5:
2
mcmp
2− =0 (2.E.56)
KB T

Step 6:
2 KB T
2
cmp = (2.E.57)
m

Step 7: (Hint: Use Avogadro's number)


2 KB T NA 2RT
2
cmp = = (2.E.58)
m NA M

Step 8:
−−−−−
2RT
cmp = √ (2.E.59)
M

2.E.11 https://chem.libretexts.org/@go/page/41328
Q2.68a
Calculate the value of cmp for CH3OH at 30°C. What is the ratio of the number of molecules of speed 400 m s-1 to the number of
molecules of speed cmp?

Q2.68b
Calculate the value of the most probable speed, average speed, and the root-mean-square speed for C 5 H12 at 25°C.

S2.68b
−−−−− −−−−−−−−−−−−−−−−−−−−−−− −
2RT 2(8.314J/K ∗ mol)(298.15K)
cmp = √ =√ = 26Q2.13m/s (2.E.60)
−3
M 72.15 × 10 kg/mol

−−−−−− −
− −−−−− −

8RT 4 2RT 4
c̄ =√ =√ ∗√ =√ ∗ cmp
π∗M π M π
−− −−−−−
4
=√ ∗ 262.13 m/s = 295.78m/s
3.14159

−−−−− −
− −−−−− −

3RT 3 2RT 3
crmp =√ =√ ∗√ =√ ∗ cmp
M 2 M 2


3
=√ ∗ 262.13 m/s = 321.04 m/s
2

Q2.68
Calculate the value of cmp for CH3OH at 30°C. What is the ratio of the number of molecules of speed 400 m s-1 to the number of
molecules of speed cmp?

2.8: Molecular Collisions and the Mean Free Path


Q2.69
You open a container that has 1 mole of Neon gas and 1 mole of hydrogen gas. Hydrogen gas moves at faster speeds than Neon.
a. Compare the mean free path between the two gases in a closed room.
b. Assume travels across the room and there are no collision between molecules. Which gas reaches the other end first?

S2.69
a. The Neon gas molecules are larger than the Hydrogen gas molecules. This decreases the distance the molecules have to travel
before they run into each other. For this reason, the mean free path for Neon will be lower than the mean free path for
Hydrogen.
b. Since there are no collisions between molecules the mean free path does not play a role in this case. Therefore, Hydrogen will
reach the other end first because it travels faster than Neon.

Q2.70a
How many times is the average distance between collision for a gas molecule larger than its molecular diameter, given that
Pressure
Po= 780mm Hg, and temperature T=300K, molecular diameter is 2x10-10 meters.

Q2.70b
The mean free path of gas molecules is given as:
1
γ = – (2.E.61)
√2π d 2 (N /V )

a. rewrite the equation so that the mean free path is expressed in terms of gas pressure
b. Name some that factors affect this property directly and indirectly?
c. Consider a gas piston containing only hydrogen gas. While not changing the temperature, there observed to be an increase in
the mean free path of the gas molecules. What action could have caused this to happen?

2.E.12 https://chem.libretexts.org/@go/page/41328
Hint: Look at concept of mean free path

S2.70b
Mean free path is the average distance molecules traveled between each successive collision. Think of what could have affect this
distance
a)
nRT
P V = nRT ⇒ V = (2.E.62)
P

N P
N /V = (2.E.63)
n RT

but(N /n) = NA (2.E.64)

N P NA RT
⇒ = ⇒ γ = – (2.E.65)
V RT √2π d 2 P NA

b) direct factors: density and molecule's dameter, indirect factors: temperature, pressure, volume
c) An increase in mean free path means that the collisions are happen less frequently, therefore the density of the gas decreased. An
action that could have caused this to happen is an increase in volume. A larger volume while molecules while not changing the
number of molecules and speed lead to molecules are further apart and have less collisions.

Q2.70c
Gas A has a density of 0.835 g/cm3 and Gas B has a density of 14.67 g/cm3. From just this information, which do you expect to
have a larger mean free path? Explain why.

S2.70c
We would expect gas A to have a larger mean free path.

Q2.71
A cylinder contains 25 molecules of an unknown ideal gas. What is the mean free path of the gas molecules? The cylinder has a
volume of 20.0 cm3 and the gas has an atomic radius of 1.24 Å. (1 cm = 108 Å)
S2.71

For an ideal gas, the mean free path is calculated by:


c̄ 1
λ = = (2.E.66)
– N
Z1 √2π d 2
V

Thus, the mean free path of the unknown gas is:


1
−17 −19
λ = = 1.17 × 10 cm = 1.17 × 10 m (2.E.67)
– 8 25
√2π(1.24 × 10 cm )2
2
20.0cm

Q2.71
A rattle containing 50 spherical beads is shaken around by a toddler. What is the mean free path of the beads if the volume of the
rattle is 500 cm3 and the diameter of each individual bead is 0.2 cm?

S2.71
Use the appropriate equation and plug in given values.
1
λ = – (2.E.68)
√2π d 2 (N /V )

1
λ = (2.E.69)
– 0.2 500
2 3
√2π( m ) (50beads/ m )
100 100

−4
λ = 1.8 ⋅ 10 m (2.E.70)
–––––––––––––––––

2.E.13 https://chem.libretexts.org/@go/page/41328
Click here for more details about mean free path.

Q2.72a
Calculate the mean free path and the binary number of collisions per liter per second between Ar molecules at 298 K and 1.00 atm.
Use 3.62 Å as the collision diameter of the Ar molecules. Assume ideal gas behavior.

Q2.72b
Calculate the mean free path and the binary number of collisions per liter per second between Ar molecules at 298 K and 1.00 atm.
Use 3.62 Å as the collision diameter of the Ar molecules. Assume ideal gas behavior.

Q2.72c
C O2 has a collision diameter of 0.4 nanometers. At 25 oC and 1.2 atm pressure, what is the mean free path between binary
collisions, and what is the rate of binary collisions per liter assuming the volume contains only pure carbon dioxide gas behaving
ideally?

Q2.72d
For the molecules HCl at 350K and 1.21 atm, calculate the mean free path and the binary number of collisions per liter per second.
The collision diameter of HCl molecules is 5.50 Å. Assume the HCl acts as an ideal gas.

S2.72d
Step 1:
N
Calculate to calculate the mean free path
V

$$PV=NRT=\dfrac{N}{N_A}RT\]
Step 2:
$$\dfrac{N}{V}=\dfrac{PN_A}{RT}=\dfrac{(1.21atm)(6.022\times 10^{23}mol^{-1})}{(0.08206L\,atm\,K^{-1}mol^{-1})
(350K)}=Q2.537\]
Step 3:
Turn L into m −3

$$Q2.537\times 10^{22}L^{-1}(\dfrac{1000L}{1m^{3}})=Q2.537\times 10^{22}m^{-3}\]


Step 4:
$$\lambda=\dfrac{1}{\sqrt{2}\pi d^{2}(\dfrac{N}{V})}=\dfrac{1}{\sqrt{2}\pi (5.50\times 10^{-10}m)^{2}(Q2.537\times
10^{22}m^{-3})}=Q2.93\times 10^{-5}m\]
Calculating the binary number of collisions
Step 1: (Hint: Use the equation for average molecular speed.)
$$\bar{C}=\sqrt{\dfrac{8RT}{\pi M}}=\sqrt{\dfrac{8(8.3145JK^{-1}mol^{-1}(300K)}{\pi (36.46\times 10^{-3}\dfrac{kg}
{mol})}}\]
Step 2:
$$z_{11}=\sqrt{\dfrac{2}{2}}\pi d^2\bar{C}\dfrac{N}{V}^{2}\]
$$=1.80526\times 10^{29}m^{-3}s^{-1}\]
Step 3: (Hint: Turn into L −1 −1
s )
$$z_{11}=\sqrt{\dfrac{2}{2}}\pi (5.50\times 10^{-10}m)^{2}(417.39ms^{-1})(\dfrac{1m^{3}}{1000L})=1.80527\times^{26}
L^{-1}s^{-1}\]

Q2.73
In a 4L container you have 5 grams of 0Q2. Assume the collision diameter of 02 is 3.55 Angstrongs . Calculate the mean free path
of the gas.

2.E.14 https://chem.libretexts.org/@go/page/41328
S2.73
The mean free path is given by
1
λ = (2.E.71)
– N
√2π d 2 ( )
V

N
V is given and we can find ( ) by calculating N.
V

23 −1
1mol
N = NA . n = (6.022 × 10 mol )(5gO2 )( ) (2.E.72)
32gO2

22
N = 9.40 × 10 (2.E.73)

22
N 9.41 × 10 22 −1 −3
( ) = = Q2.35 × 10 L (1000Lm ) (2.E.74)
V 4L

N 25 3
( ) = Q2.35 × 10 m (2.E.75)
V

The mean free path is


1
λ = (2.E.76)
– N
√2π d 2 ( )
V

1 −8
λ = = 7.60 × 10 m (2.E.77)
– −10 2 25 −3
√2π(3.55 × 10 m ) (Q2.35 × 10 m )

Q2.74a
If a sealed container has of xenon molecules traveling at a speed of 24.6 m/s at 108 °C. How many moles of xenon gas do we have
in the container?

S2.74a
1 3
¯ 2
E trans = mv = kB T
2 2
−23 2 2
3 kB T 3(1.381 × 10 J/K)(381.15K) 1kg ∗ m / s
−23
m = = ∗ = Q2.609 × 10 kg
v2 (24.6m/s)2 1J

m −23
1000g mol −22
n = = Q2.609 × 10 kg ∗ ∗ = 1.987 × 10 mol
molarmass 1kg 131.3g

Q2.74b
At room temperature (25oC) there are two flasks A and B containing gaseous Oxygen and Hydrogen, respectively, at their
equilibrium states. Assuming there is no exchange of heat or work between the systems inside the flasks and the surroundings.
Compare the initial velocities between an Oxygen and a Hydrogen molecule in each flask.

Q2.76a
In a insulated container of volume 2.50 L filled with nitrogen gas, the Z1 value is 6.75e19 collisions/second. The temperature
measured is 55ºC. Assuming that the collision diameter of N2 is 3.75 Å,
a. how many moles of nitrogen gas are present the container?
b. find the Z11 value of the gas.
c. How does the number of moles affect Z1 and Z11 values of the gas?
Hint: look at collision frequency

2.E.15 https://chem.libretexts.org/@go/page/41328
S2.76a
a)

– N
2
Z1 = √2π d c̄ ( ) (2.E.78)
V

V Z1
⇒ N = (2.E.79)
– 2
√2π d c̄

−−−−−−−−−−−−−−−−−−−−−−− −

 J
−−−−−  8(8.314 )(55 + 273)K
8RT  mol ⋅ K
c̄ = √ = ) = 704.3m/s (2.E.80)

πM π ⋅ 0.014kg

3
1m
19
(2.50L)( )(6.75 × 10 collision/s)
1000L 23
N = – = 1.017 × 10 molecules (2.E.81)
−10 2
√2π(3.75 × 10 m ) (704.3m/s)

23
N 1.017 × 10 molecules
n = = = 0.169moleN2 (2.E.82)
NA molecules
23
6.022 × 10
mole

b)
1 N
Z11 = Z1 ( ) (2.E.83)
2 V

23
1 1.017 × 10 molecules collisions
19 45
Z11 = (6.75 × 10 ) = 1.37 × 10 (2.E.84)
3 3
2 1m m ⋅s
2.50L ×
1000L

c) Since N = nNA and Z1 is proportional to N, therefore


Z1 ∝ n (2.E.85)

2
similarly, Z11 is proportional to N , therefore:
2
Z11 ∝ n (2.E.86)

Q2.76b
For molecular oxygen at 56 C , calculate the number of collisions a single molecule makes in 1 second and the total number of
o

binary collisions at P=1.0 atm and P=0.25 atm. How does pressure relate to these two quantities? The collision diameter of oxygen
is 3.02 A.

S2.76b
Step 1: (Hint: Calculate the density and average molecular speed)
At P=1.0atm
$$\dfrac{N}{V}=\dfrac{PN_A}{RT}=\dfrac{(1.0atm)(6.022\times 10^{23}mol^{-1})}{0.08206\,L\,atm\,K^{-1}mol^{-1})
(273+56)}=Q2.23\times 10^{22}L^{-1}(\dfrac{1000L}{1m^{3}})=2.23\times 10^{22}m^{-3}\]
Step 2:
$$\bar{c}=\sqrt{\dfrac{8RT}{\pi M}}=\sqrt{\dfrac{8(8.3145JK^{-1}mol^{-1})(273+56K)}{\pi (31.998\times
10^{-3}kg\,mol^{-1})}}=466.578ms^{-1}\]
Step 3:
$$z_1=\sqrt{2}\pi d^2\bar{c}\dfrac{N}{V}=\sqrt{2}\pi(3.02\times 10^{-10}m)^{2}(466.578ms^{-1})(2.23\times
10^{22}m^{-3})=4.683\times 10^{28} collisions m^{-3}s^{-1}\]
Step 4:

2.E.16 https://chem.libretexts.org/@go/page/41328
$$z_{11}=\dfrac{\sqrt{2}}{2}\pi d\bar{c}(\dfrac{N}{V})^{2}=\dfrac{1}{2}(4.2\times 10^{6})(2.23\times
10^{22}m^{-3})=4.683\times 10^{28} collisionsm^{-3}s^{-1}\]
For an ideal gas, z is proportional to P and z is proportional to P . Reducing P to a fourth of its original value (from 1.0 atm to
1 11
2

0.25 atm) will also reduce 2, to one fourth its value at P=1.0atm.
Therefore at P=0.25 atm,
5 −1 26 −3
z1 = 4.2 × 10 collisions s $$$$ z11 = 4.683 × 10 collisions m s−1 (2.E.87)

Q2.77a
The relationship between the molar mass and root mean square speed of an ideal gas can be described as:
−−−−−
3RT
υrms = √ (2.E.88)
M

Use this equation to derive Grahm's Law of Effusion:


−−−

r1 M2
=√ (2.E.89)
r2 M1

S2.77a

For two gasses at standard temperature and pressure,


−−−−− −−−−−
3RT 3RT
υ1rms = √ υ2rms = √ (2.E.90)
M1 M2

For any of the two gases,


3RT 3RT
2 2
υ = υ = (2.E.91)
1rms 2rms
M1 M2

2 2
υ M1 υ M2
1rms 2rms
R = = (2.E.92)
3T 3T

2 2
υ M1 = υ M2 (2.E.93)
1rms 2rms

2
υ M2
1rms
= (2.E.94)
2
υ M1
2rms

−−−

υ1rms M2
=√ (2.E.95)
υ2rms M1

Which can be rewritten as:


−−−

r1 M2
=√ (2.E.96)
r2 M1

2.9: Graham's Laws of Diffusion and Effusion


Q2.77b
Derive Graham's Law of Effusion (1) from the equation for calculating Kinetic Energy (2).
−−−
v1 m2
(1) =√ (2.E.97)
v2 m1

1 2
(2) KE = mv (2.E.98)
2

2.E.17 https://chem.libretexts.org/@go/page/41328
S2.77b
−−−
v1 m2 1 2
=√ KE = mv (2.E.99)
v2 m1 2

K E1 = K E2 , f or two substances (2.E.100)

1 1
2 2
m1 v = m2 v , multiply both sides by two (2.E.101)
1 2
2 2
2 2
m1 v = m2 v (2.E.102)
1 2

2
v m2
1
= , square root both sides of the equation (2.E.103)
2
v m1
2
−−−
v1 m2
=√ (2.E.104)
v2 m1
––––––––––––

Click here and here for more information about effusion and Graham's law.

Q2.78a
How long would it take for oxygen to diffuse in and carbon dioxide to diffuse out of a cell that is 0.1µm2 thick given that the rate of
diffusion of carbon dioxide is 0.0016 mm2/s?

S2.78a
−−−−−
tC O2 rO2 MC O2
= =√
tO2 rC O2 MO2

−−−−−−−−−−
44.01g/mol
2 2
rO2 = (0.0016m m /s)√ = 0.0019m m /s
3Q2.0g/mol

2
(1mm) s
2 −3
tC O2 = 0.1μm ∗ ∗ = 6.25 × 10 s
2 2
(100μm) 0.0016mm

2
(1mm) s
2 −3
tO = 0.1μm ∗ ∗ = 5.26 × 10 s
2
2 2
(100μm) 0.0019mm

Q2.78b
A sample of red gas collected from the emissions of a paper mill effuses through an orifice of unknown size at standard
temperature and pressure in 14.1 minutes. Pure diatomic nitrogen gas at standard temperature and pressure effuses through the hole
in 11.0 minutes. What is the red gas? How do you know?

S2.78b
mN2 = 28.01 g/mol
tN2/tunk = (munk/mN2)1/2
munk = (tN2/tunk)2 * mN2 = (14.1 min/11.0 min)2 * 28.01 g mol-1
munk = 46.02 g mol-1
The red gas is nitrogen dioxide, a common emission from pulp mills. We know because the diffusion rate corresponds with a molar
mass of 46.01g, the molar mass of NO2

Q2.80
An amount of CO2 gas is effused through a small opening at the rate of 0.596 L/min. A homogenous mixture of O2 and N2 in the
atmosphere is put under the same conditions for effusion and after 5 minutes, 5.03 L of the mixture was effused. Find the percent
composition of the homogenous mixture (molar fraction).
Hint: Look at concept of effusion and its rate

2.E.18 https://chem.libretexts.org/@go/page/41328
S2.80
−−−−−−−−
rC O2 Mmixture
=√ (2.E.105)
rmixture MC O
2

L
0.596
r 2 g g
CO 2 min 2
Mmixture = ( ) (MC O ) = ( ) (44.01 ) = 15.447 (2.E.106)
2
5.03L
rmixture mol mol
5min

Mmixture = MN XN + MO XO (2.E.107)
2 2 2 2

Mmixture = MN XN + MO (1 − XN ) (2.E.108)
2 2 2 2

g g g
15.447 = (14.01 )(XN ) + (16.00 )(1 − XN ) (2.E.109)
2 2
mol mol mol

15.447 = 14.01 XN + 16.00 − 16.00 XN = −1.99 XN + 16.00 (2.E.110)


2 2 2

XN2 = 0.278 (2.E.111)

XO2 = 1 − 0.278 = 0.722 (2.E.112)

%of N2 = 27.8% (2.E.113)

%of O2 = 72.2% (2.E.114)

Q2.81a
You are trying to isolate gaseous Nitrogen-14 via effusion in a 50:50 mixture of Nitrogen-14 and Nitrogen-15 respectively. What
will be the percentage of enrichment of the mixture after a single stage of separation?

S2.81a
Use Graham's Law of Diffusion and Effusion to solve this problem.

−−−
r1 M2
=√ (2.E.115)
r2 M1

−−−−
14 15
rate of N N
=√ (2.E.116)
15 14
rate of N N

−−−−−−−−−−−−−−
14 −1
rate of N 15.0001 g mol
=√ (2.E.117)
15 −1
rate of N 14.00307 g mol

14
rate of N
= 1.035 (2.E.118)
15
rate of N

Since the ratio of rates of effusion is 1.035, in a single step of separation 3.5% more nitrogen-14 will effuse out of the mixture than
nitrogen-15. Thus the percent of enrichment would be 3.5% since there is 3.5% more of N-14 now.
Click here and bhere for more information about effusion and Graham's law.

Q2.81b
A mixture of Neon and Helium gas is effused at standard temperature and pressure. Assuming the mixture is equimolar, what is the
composition of the gas mixture after effusion?

S2.81b
For the effusion of a binary gas mixture, we can use Graham's Law:
−−−

r1 M2
=√ (2.E.119)
r2 M1

The ratio of the two gasses following effusion given r1 corresponds with neon and r2 corresponds with helium is:

2.E.19 https://chem.libretexts.org/@go/page/41328
−−− −−−−− −
 g

r1  4.003
mol
= = 0.4454 (2.E.120)
g
r2 ⎷ 20.180
mol

Therefore, to calculate the mole fraction of each gas,


0.4454
xN e = = 0.3081 (2.E.121)
0.4454 + 1

xH e = 1 − xN e = 1 − 0.3081 = 0.6919 (2.E.122)

Q2.82
An equimolar mixture of diatomic nitrogen gas and an unknown monatomic gas diffuse through an aperture of unknown size. It is
determined that 1.18 times as many moles of the monatomic gas pass through the opening as moles of nitrogen gas. What is the
gas?

Q2.82
mN2 = 28.01 g mol-1
nunk/nN2 = (mN2/munk)1/2
munk = mN2/(nunk/nN2)2 = 28.01 g mol-1/(1.18)2 = 20.11 g mol-1
The gas is monatomic and has a mass of roughly 20.11 g/mol, so Neon is the most likely gas.

2.E: Properties of Gases (Exercises) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

2.E.20 https://chem.libretexts.org/@go/page/41328
CHAPTER OVERVIEW

3: The First Law of Thermodynamics


3.1: Work and Heat
3.2: The First Law of Thermodynamics
3.3: Heat Capacities
3.4: Gas Expansion
3.5: Calorimetry
3.6: Thermochemistry
3.7: Bond Energies and Enthalpies
3.E: Exercises

3: The First Law of Thermodynamics is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1
3.1: Work and Heat
One of the pioneers in the field of modern thermodynamics was James P. Joule (1818 - 1889). Among the experiments Joule
carried out, was an attempt to measure the effect on the temperature of a sample of water that was caused by doing work on the
water. Using a clever apparatus to perform work on water by using a falling weight to turn paddles within an insulated canister
filled with water, Joule was able to measure a temperature increase in the water.

Figure 3.1.1 : (left) James Prescott Joule (1818 - 1889) (right) Joule's apparatus for measuring the work equivalent of heat. (CC BY-
SA 3.0; Dr. Mirko Junge)
Thus, Joule was able to show that work and heat can have the same effect on matter – a change in temperature! It would then be
reasonable to conclude that heating, as well as doing work on a system will increase its energy content, and thus it’s ability to
perform work in the surroundings. This leads to an important construct of the First Law of Thermodynamics:

The capacity of a system to do work is increased by heating the system or doing work on
it.
The internal energy (U) of a system is a measure of its capacity to supply energy that can do work within the surroundings,
making U the ideal variable to keep track of the flow of heat and work energy into and out of a system. Changes in the internal
energy of a system (ΔU ) can be calculated by
ΔU = Uf − Ui (3.1.1)

where the subscripts i and f indicate initial and final states of the system. U as it turns out, is a state variable. In other words, the
amount of energy available in a system to be supplied to the surroundings is independent on how that energy came to be available.
That’s important because the manner in which energy is transferred is path dependent.
There are two main methods energy can be transferred to or from a system. These are suggested in the previous statement of the
first law of thermodynamics. Mathematically, we can restate the first law as

ΔU = q + w

or

dU = dq + dw

where q is defined as the amount of energy that flows into a system in the form of heat and w is the amount of energy lost due to
the system doing work on the surroundings.

Heat
Heat is the kind of energy that in the absence of other changes would have the effect of changing the temperature of the system. A
process in which heat flows into a system is endothermic from the standpoint of the system (q > 0, q
system < 0 ).
surroundings

Likewise, a process in which heat flows out of the system (into the surroundings) is called exothermic (q < 0, system

qsurroundings> 0 ). In the absence of any energy flow in the form or work, the flow of heat into or out of a system can be measured

by a change in temperature. In cases where it is difficult to measure temperature changes of the system directly, the amount of heat

3.1.1 https://chem.libretexts.org/@go/page/41412
energy transferred in a process can be measured using a change in temperature of the soundings. (This concept will be used later in
the discussion of calorimetry).
An infinitesimal amount of heat flow into or out of a system can be related to a change in temperature by

dq = C dT

where C is the heat capacity and has the definition


dq
C =
∂T

Heat capacities generally have units of (J mol-1 K-1) and magnitudes equal to the number of J needed to raise the temperature of 1
mol of substance by 1 K. Similar to a heat capacity is a specific heat which is defined per unit mass rather than per mol. The
specific heat of water, for example, has a value of 4.184 J g-1 K-1 (at constant pressure – a pathway distinction that will be
discussed later.)

 Example 3.1.1: Heat required to Raise Temperature

How much energy is needed to raise the temperature of 5.0 g of water from 21.0 °C to 25.0 °C?

Solution
q = mC ΔT

J
= (5.0 g )(4.184 )(25.0 °C − 21.0 °C )
g °C

= 84 J

 What is a partial derivative?

A partial derivative, like a total derivative, is a slope. It gives a magnitude as to how quickly a function changes value when
one of the dependent variables changes. Mathematically, a partial derivative is defined for a function f (x , x , … x ) by
1 2 n

∂f f (x1 + Δx1 , x2 + Δx2 , … , xi + Δxi , … xn + Δxn ) − f (x1 , x2 , … xi , … xn )


( ) = lim ( )
∂xi Δi →0 Δxi
xj ≠i

Because it measures how much a function changes for a change in a given dependent variable, infinitesimal changes in the in
the function can be described by
∂f
df = ∑ ( )
∂xi
i xj ≠i

So that each contribution to the total change in the function f can be considered separately.
For simplicity, consider an ideal gas. The pressure can be calculated for the gas using the ideal gas law. In this expression,
pressure is a function of temperature and molar volume.
RT
p(V , T ) =
V

The partial derivatives of p can be expressed in terms of T and V as well.


∂p RT
( ) =− (3.1.2)
2
∂V V
T

and
∂p R
( ) = (3.1.3)
∂T V
V

3.1.2 https://chem.libretexts.org/@go/page/41412
So that the change in pressure can be expressed
∂p ∂p
dp = ( ) dV + ( ) dT (3.1.4)
∂V ∂T
T V

or by substituting Equations 3.1.2 and 3.1.3


RT R
dp = (− ) dV + ( ) dT
2
V V

Macroscopic changes can be expressed by integrating the individual pieces of Equation 3.1.4 over appropriate intervals.
V2 T2
∂p ∂p
Δp = ∫ ( ) dV + ∫ ( ) dT
V1
∂V T1
∂T
T V

This can be thought of as two consecutive changes. The first is an isothermal (constant temperature) expansion from V to V 1 2

at T and the second is an isochoric (constant volume) temperature change from T to T at V . For example, suppose one
1 1 2 2

needs to calculate the change in pressure for an ideal gas expanding from 1.0 L/mol at 200 K to 3.0 L/mol at 400 K. The set up
might look as follows.
V2 T2
RT R
Δp = ∫ (− ) dV + ∫ ( ) dT
2
V1 V T1
V
 
isothermal expansion isochoric heating

or
3.0 L/mol 400, K
R(400 K) R
Δp = ∫ (− ) dV + ∫ ( ) dT
2
1.0 L/mol V 200 K 1.0 L/mol

3.0 L/mol 400 K


R(200 K) RT
=[ ] +[ ]
V 3.0 L/mol
1.0 L/mol 200 K

200 K 200 K 400 K 200 K


= R [( − ) +( − )]
3.0 L/mol 1.0 L/mol 3.0 L/mol 3.0 L/mol

= −5.47 atm

Alternatively, one could calculate the change as an isochoric temperature change from T1 to T2 at V1 followed by an
isothermal expansion from V to V at T :
1 2 2

T2 V2
R RT
Δp = ∫ ( ) dT + ∫ (− ) dV
2
T1
V V1 V

or

3.1.3 https://chem.libretexts.org/@go/page/41412
400, K 3.0 L/mol
R R(400 K)
Δp = ∫ ( ) dT + ∫ (− ) dV
2
200 K 1.0 L/mol 1.0 L/mol V

400 K 3.0 L/mol


RT R(400 K)
=[ ] +[ ]
1.0 L/mol V
200 K 1.0 L/mol

400 K 200 K 400 K 400 K


= R [( − ) +( − )]
1.0 L/mol 1.0 L/mol 3.0 L/mol 1.0 L/mol

= −5.47 atm

This results demonstrates an important property of pressure in that pressure is a state variable, and so the calculation of
changes in pressure do not depend on the pathway!

Work
Work can take several forms, such as expansion against a resisting pressure, extending length against a resisting tension (like
stretching a rubber band), stretching a surface against a surface tension (like stretching a balloon as it inflates) or pushing electrons
through a circuit against a resistance. The key to defining the work that flows in a process is to start with an infinitesimal amount of
work defined by what is changing in the system.
Table 3.1.1: Changes to the System
Type of work Displacement Resistance dw

Expansion dV (volume) -pext (pressure) -pextdV

Electrical dQ (charge) W (resistence) -W dQ

Extension dL (length) -t (tension) t dL

Stretching dA -s (surf. tens.) sdA

The pattern followed is always an infinitesimal displacement multiplied by a resisting force. The total work can then be determined
by integrating along the pathway the change follows.

 Example 3.1.2: Work from a Gas Expansion

What is the work done by 1.00 mol an ideal gas expanding from a volume of 22.4 L to a volume of 44.8 L against a constant
external pressure of 0.500 atm?

Solution
dw = −pext dV

since the pressure is constant, we can integrate easily to get total work
V2

w = −pexp ∫ dV
V1

= −pexp (V2 − V1 )

8.314 J
= −(0.500 am)(44.8 L − 22.4 L) ( )
0.08206 atm L

= −1130 J = −1.14 kJ

Note: The ratio of gas law constants can be used to convert between atm·L and J quite conveniently!

3.1: Work and Heat is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
3.2: Work and Heat by Patrick Fleming is licensed CC BY-NC-SA 4.0.

3.1.4 https://chem.libretexts.org/@go/page/41412
3.2: The First Law of Thermodynamics
3.2: The First Law of Thermodynamics is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

3.2.1 https://chem.libretexts.org/@go/page/41413
3.3: Heat Capacities
Let us turn our attention from the macroscopic to the microscopic level. According to the first law of thermodynamics, the heat
energy absorbed as we raise the temperature of a substance cannot be destroyed. But where does it go? In the case of a monatomic
gas, like neon, this question is easy to answer. All the energy absorbed is converted into the kinetic energy of the neon molecules
(atoms). In other sections, we found that the kinetic energy of the molecules in a sample of gas is given by the expression
3
Ek = nRT
2

Thus if the temperature of a sample of neon gas is raised from T1 to T2, the kinetic energy of the molecules increases from 3/2 nRT1
to 3/2 nRT2, a total change of
3 3
nR (T2 − T1 ) = ( R) n (T2 − T1 )
2 2

Inserting the value of R in appropriate units, we obtain


3 J J
(8.314 ) n (T2 − T1 ) = (12.47 ) n (T2 − T1 ) (3.3.1)
2 K mol K mol

This is the same quantity that is obtained by substituting the experimental value of CV for neon (calculated in Example 2 from Heat
Capacities) into Eq. (4) from Heat Capacities. In other words the quantity of heat found experimentally exactly matches the
increase in kinetic energy of the molecules required by the kinetic theory of gases.
Table 3.3.1 lists the CV values not only for neon but for some other gases as well. We immediately notice that only the noble gases
and other mon-atomic gases such as Hg and Na have molar heat capacities equal to 3/2R, or 12.47 J K–1 mol–1. All other gases have
higher molar heat capacities than this. Moreover, as the table shows, the more complex the molecule, the higher the molar heat
capacity of the gas. There is a simple reason for this behavior.
Table 3.3.1 : Molar Heat Capacities at Constant Volume (CV) for Various Gases (Values at 298 K Unless Otherwise Stated).
Gas Cv /J K-1 mol-1 Gas Cv /J K-1 mol-1

Monoatomic Gases Triatomic Gases

Ne 12.47 CO2 28.81

Ar 12.47 N2O 30.50

Hg 12.47 (700K) SO2 31.56

Na 12.47 (1200K)

Diatomic Gases Alkanes

N2 20.81 CH4 27.42

O2 21.06 C2H6 44.32

Cl2 25.62 C3H8 65.20

C4H10 89.94

A molecule which has two or more atoms is not only capable of moving from one place to another (translational motion), it can
also rotate about itself, and it can change its shape by vibrating. When we heat a mole of Cl2 molecules, for example, we not only
need to supply them with enough energy to make them move around faster (increase their translational kinetic energy), we must
also supply an additional quantity of energy to make them rotate and vibrate more strongly than before. For a mole of more
complex molecules like n-butane even more energy is required since the molecule is capable of changing its shape in all sorts of
ways. In the butane molecule there are three C—C bonds around which segments of the molecule can rotate freely. All the bonds
can bend or stretch, and the whole molecule can rotate as well. Such a molecule is constantly flexing and writhing at room
temperature. As we raise the temperature, this kind of movement occurs more rapidly and extra energy must be absorbed in order
to make this possible.

3.3.1 https://chem.libretexts.org/@go/page/41414
When we heat solids and liquids, the situation is somewhat different than for gases. The rapid increase of vapor pressure with
temperature makes it virtually impossible to heat a solid or liquid in a closed container, and so heat capacities are always measured
at constant pressure rather than at constant volume. Some Cp values for selected simple liquids and solids at the melting point are
shown in Table 3.3.2. In general the heat capacities of solids and liquids are higher than those of gases. This is because of the
intermolecular forces operating in solids and liquids. When we heat solids and liquids, we need to supply them with potential
energy as well as kinetic energy. Among the solids, the heat capacities of the metals are easiest to explain since the solid consists of
individual atoms. Each atom can only vibrate in three dimensions. According to a theory first suggested by Einstein, this
vibrational energy has the value 3RT, while the heat capacity is given by 3R = 24.9 J K–1 mol–1.
Table 3.3.2 : Molar Heat Capacities at Constant Pressure Cp for Various Solids and Liquids at the Melting Point.
Substance Cp (solid)/J K-1 mol-1 Cp (liquid)/J K-1 mol-1

Monoatomic Substances

Hg 27.28 27.98

Pb 29.40 30.33

Na 28.20 31.51

Diatomic Substances

Br2 53.8 75.7

I2 54.5 80.7

HCl 50.5 62.2

HI 47.5 68.6

Polyatomic Substances

H2O 37.9 76.0

NH3 49.0 77.0

Benzene 129.0 131.0

n-Heptane 146.0 203.1

As can be seen from the table, most monatomic solids have Cp values slightly larger than this. This is because solids expand
slightly on heating. The atoms get farther apart and thus increase in potential as well as vibrational energy.
Solids which contain molecules rather than atoms have much higher heat capacities than 3R. In addition to the vibration of the
whole molecule about its site in the crystal lattice, the individual atoms can also vibrate with respect to each other. Occasionally
molecules can rotate in the crystal, but usually rotation is only possible when the solid melts. As can be seen from the values for
molecular liquids in Table 3.3.2, this sudden ability to rotate causes a sharp increase in the heat capacity. For monatomic
substances, where there is no motion corresponding to the rotation of atoms around each other, the heat capacity of the liquid is
only very slightly higher than that of the solid.

3.3: Heat Capacities is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
15.3: Heat Capacity and Microscopic Changes by Ed Vitz, John W. Moore, Justin Shorb, Xavier Prat-Resina, Tim Wendorff, & Adam Hahn
is licensed CC BY-NC-SA 4.0.

3.3.2 https://chem.libretexts.org/@go/page/41414
3.4: Gas Expansion
In Gas Expansion, we assume Ideal behavior for the two types of expansions:

Isothermal Expansion

This shows the expansion of gas at constant temperature against weight of an object's mass (m) on the piston. Temperature is held
constant, therefore the change in energy is zero (U=0). So, the heat absorbed by the gas equals the work done by the ideal gas on its
surroundings. Enthalpy change is also equal to zero because the change in energy zero and the pressure and volume is constant.

Isothermal Irreversible/Reversible process

The graphs clearly show work done (area under the curve) is greater in a reversible process.

Adiabatic Expansions
Adiabatic means when no heat exchange occurs during expansion between system and surrounding and the temperature is no
longer held constant.

Reversible Adiabatic Expansion


This equation shows the relationship between PV and is useful only when it applies to ideal gas and reversible adiabatic change.
The equation is very similar to Boyle's law except it has exponent (gamma) due to change in temperature. The work done by an
adiabatic reversible process is given by the following equation:
where T2 is less than T1. The internal energy of the system decreases as the gas expands. The work can be calculated in two ways
because the Internal energy (U) does not depend on path. The graph shows that less work is done in an adiabatic reversible process
than an Isothermal reversible process.

3.4.1 https://chem.libretexts.org/@go/page/41415
References
1. Chang, Raymond. Physical Chemistry for the Biosciences. Sausalito, CA: University Science, 2005.

3.4: Gas Expansion is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
Gas Expansion is licensed CC BY-NC-SA 4.0.

3.4.2 https://chem.libretexts.org/@go/page/41415
3.5: Calorimetry
As chemists, we are concerned with chemical changes and reactions. The thermodynamics of chemical reactions can be very
important in terms of controlling the production of desired products and preventing safety hazards such as explosions. As such,
measuring and understanding the thermochemistry of chemical reactions is not only useful, but essential!

Calorimetry
The techniques of calorimetry can be used to measure q for a chemical reaction directly. The enthalpy change for a chemical
reaction is of significant interest to chemists. An exothermic reaction will release heat (q < 0, q
reaction > 0 ) causing the
surroundings

temperature of the surrounding to increase. Conversely, an endothermic reaction (q > 0, q


reaction < 0 ) will draw heat
surroundings

from the surroundings, causing the temperature of the surrounding to drop. Measuring the temperature change in the surroundings
allows for the determination of how much heat was released or absorbed in the reaction.

Bomb Calorimetry
Bomb calorimetry is used predominantly to measure the heat evolved in combustion reactions, but can be used for a wide variety of
reactions. A typical bomb calorimetry set up is shown here. The reaction is contained in a heavy metallic container (the bomb)
forcing the reaction to occur at constant volume. As such, the heat evolved (or absorbed) by the reaction is equal to the change in
internal energy (DUrxn). The bomb is then submerged in a reproducible quantity of water, the temperature of which is monitored
with a high-precision thermometer.
For combustion reactions, the bomb will be loaded with a small sample of the compound to be combusted, and then the bomb is
filled with a high pressure (typically about 10 atm) of O2. The reaction is initiated by supplying heat using a short piece of resistive
wire carrying an electrical current.

3.5.1 https://chem.libretexts.org/@go/page/41416
Figure 3.5.1 : A Bomb Calorimeter. After the temperature of the water in the insulated container has reached a constant value, the
combustion reaction is initiated by passing an electric current through a wire embedded in the sample. Because this calorimeter
operates at constant volume, the heat released is not precisely the same as the enthalpy change for the reaction. (CC BY-SA-NC;
Anonymous by request).
The calorimeter must be calibrated by carrying out a reaction for which ΔU is well known, so that the resulting temperature
rxn

change can be related to the amount of heat released or absorbed. A commonly used reaction is the combustion of benzoic acid.
This makes a good choice since benzoic acid reacts reliably and reproducibly under normal bomb calorimetry conditions. The
“water equivalent” of the calorimeter can then be calculated from the temperature change using the following relationship:
nΔUc + ewrire + eother
W =
ΔT

where n is the number of moles of benzoic acid used, ΔU is the internal energy of combustion for benzoic acid (3225.7 kJ mol-1 at
c

25 oC), ewire accounts for the energy released in the combustion of the fuse wire, eother account for any other corrections (such as
heat released due to the combustion of residual nitrogen in the bomb), and DT is the measured temperature change in the
surrounding water bath.
Once the “water equivalent” is determined for a calorimeter, the temperature change can be used to find ΔUc for an unknown
compound from the temperature change created upon combustion of a known quantity of the substance.
W ΔT − ewire − eother
ΔUc =
nsample

The experiment above is known as “isothermal bomb calorimetry” as the entire assembly sits in a constant temperature laboratory.
Another approach is to employ “adiabatic bomb calorimetry” in which the assembly sits inside of a water jacket, the temperature of
which is controlled to match the temperature of the water inside the insulated container. By matching this temperature, there is no
thermal gradient, and thus no heat leaks into or out of the assembly during an experiment (and hence the experiment is effectively
“adiabatic”).

Finding ΔU c

The enthalpy of combustion can be calculated from the internal energy change if the balanced chemical reaction is known. Recall
from the definition of enthalpy

ΔH = ΔU + Δ(pV )

and if the gas-phase reactants and products can be treated as ideal gases (pV = nRT )

3.5.2 https://chem.libretexts.org/@go/page/41416
ΔH = ΔU + RT Δngas

at constant temperature. For the combustion of benzoic acid at 25 oC


15
C H COOH(s) + O (g) ⟶ 7 CO (g) + 3 H O(l)
6 5 2 2 2 2

it can be seen that Δn gas is -0.5 mol of gas for every mole of benzoic acid reacted.

 Example 3.5.1: Combustion of Naphthalene

A student burned a 0.7842 g sample of benzoic acid (C H O ) in a bomb calorimeter initially at 25.0 oC and saw a
7 6 2

temperature increase of 2.02 oC. She then burned a 0.5348 g sample of naphthalene (C H ) (again from an initial temperature
10 8

of 25 oC) and saw a temperature increase of 2.24 oC. From this data, calculate ΔH for naphthalene (assuming ewire and eother
c

are unimportant.)

Solution
First, the water equivalent:
1 mol
[(0.7841 g) ( )] (3225.7 kJ/mol)
122.124 g

W = = 10.254 kJ/°C
2.02 °C

Then ΔU for the sample:


c

(10.254 kJ/ °C )(2.24 °C )


ΔUc = = 5546.4 kJ/°C
1 mol
(0.5308 g) ( )
128.174 g

ΔHc is then given by

ΔHc = ΔUc + RT Δngas

The reaction for the combustion of naphthalene at 25 oC is:

C H (s) + 12 O (g) ⟶ 10 CO (g) + 4 H O(l)


10 8 2 2 2

with Δn gas = −2 .
So
8.314
ΔHc = 5546.4 kJ/mol + ( kJ/(mol K)) (298 L)(−2) = 5541 kJ/mol
1000

The literature value (Balcan, Arzik, & Altunata, 1996) is 5150.09 kJ/mol. So that’s not too far off!

3.5: Calorimetry is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
3.4: Calorimetry by Patrick Fleming is licensed CC BY-NC-SA 4.0.

3.5.3 https://chem.libretexts.org/@go/page/41416
3.6: Thermochemistry
The standard enthalpy of formation is defined as the change in enthalpy when one mole of a substance in the standard state (1 atm
of pressure and 298.15 K) is formed from its pure elements under the same conditions.

Introduction
The standard enthalpy of formation is a measure of the energy released or consumed when one mole of a substance is created under
standard conditions from its pure elements. The symbol of the standard enthalpy of formation is ΔHf.
Δ = A change in enthalpy
o = A degree signifies that it's a standard enthalpy change.

f = The f indicates that the substance is formed from its elements


The equation for the standard enthalpy change of formation (originating from Enthalpy's being a State Function), shown below, is
commonly used:
o o o
ΔH = ∑ ΔH (products) − ∑ ΔH (Reactants) (3.6.1)
reaction f f

This equation essentially states that the standard enthalpy change of formation is equal to the sum of the standard enthalpies of
formation of the products minus the sum of the standard enthalpies of formation of the reactants.

Example 3.6.1

Given a simple chemical equation with the variables A, B and C representing different compounds:
A+B ⇋ C

and the standard enthalpy of formation values:


ΔHfo[A] = 433 KJ/mol
ΔHfo[B] = -256 KJ/mol
ΔHfo[C] = 523 KJ/mol
the equation for the standard enthalpy change of formation is as follows:
ΔHreactiono = ΔHfo[C] - (ΔHfo[A] + ΔHfo[B])
ΔHreactiono = (1 mol)(523 kJ/mol) - ((1 mol)(433 kJ/mol) + (1 mol)(-256 kJ/mol)\)
Because there is one mole each of A, B and C, the standard enthalpy of formation of each reactant and product is multiplied by
1 mole, which eliminates the mol denominator:
ΔHreactiono = 346 kJ
The result is 346 kJ, which is the standard enthalpy change of formation for the creation of variable "C".

The standard enthalpy of formation of a pure element is in its reference form its standard enthalpy formation is zero.

Carbon naturally exists as graphite and diamond. The enthalpy difference between graphite and diamond is too large for both to
have a standard enthalpy of formation of zero. To determine which form is zero, the more stable form of carbon is chosen. This is
also the form with the lowest enthalpy, so graphite has a standard enthalpy of formation equal to zero. Table 1 provides sample
values of standard enthalpies of formation of various compounds.
Table 1: Sample Table of Standard Enthalpy of Formation Values. Table T1 is a more comprehensive table.
Compound ΔHfo

O2(g) 0 kJ/mol

C(graphite) 0 kJ/mol

CO(g) -110.5 kJ/mol

CO2(g) -393.5 kJ/mol

3.6.1 https://chem.libretexts.org/@go/page/41417
Compound ΔHfo

H2(g) 0 kJ/mol

H2O(g) -241.8 kJ/mol

HF(g) -271.1 kJ/mol

NO(g) 90.25 kJ/mol

NO2(g) 33.18 kJ/mol

N2O4(g) 9.16 kJ/mol

SO2(g) -296.8 kJ/mol

SO3(g) -395.7 kJ/mol

All values have units of kJ/mol and physical conditions of 298.15 K and 1 atm, referred to as the "standard state." These are the
conditions under which values of standard enthalpies of formation are typically given. Note that while the majority of the values of
standard enthalpies of formation are exothermic, or negative, there are a few compounds such as NO(g) and N2O4(g) that actually
require energy from its surroundings during its formation; these endothermic compounds are generally unstable.

Example 3.6.2

Between Br2(l) and Br2(g) at 298.15 K, which substance has a nonzero standard enthalpy of formation?
Solution
Br2(l) is the more stable form, which means it has the lower enthalpy; thus, Br2(l) has ΔHf = 0. Consequently, Br2(g) has a
nonzero standard enthalpy of formation.
Note: that the element phosphorus is a unique case. The reference form in phosphorus is not the most stable form, red
phosphorus, but the less stable form, white phosphorus.
Recall that standard enthalpies of formation can be either positive or negative.

Example 3.6.3

The enthalpy of formation of carbon dioxide at 298.15K is ΔHf = -393.5 kJ/mol CO2(g). Write the chemical equation for the
formation of CO2.
Solution
This equation must be written for one mole of CO2(g). In this case, the reference forms of the constituent elements are O2(g)
and graphite for carbon.
O2 (g) + C (graphite) ⇌ C O2 (g) (3.6.2)

The general equation for the standard enthalpy change of formation is given below:
o o o
ΔH = ∑ ΔH (products) − ∑ ΔH (Reactants) (3.6.3)
reaction f f

Plugging in the equation for the formation of CO2 gives the following:
ΔHreactiono= ΔHfo[CO2(g)] - (ΔHfo[O2(g)] + ΔHfo[C(graphite)]
Because O2(g) and C(graphite) are in their most elementally stable forms, they each have a standard enthalpy of formation
equal to 0:
ΔHreactiono= -393.5 kJ = ΔHfo[CO2(g)] - ((1 mol)(0 kJ/mol) + (1 mol)(0 kJ/mol))
ΔHfo[CO2(g)]= -393.5 kJ

3.6.2 https://chem.libretexts.org/@go/page/41417
Example 3.6.4

Using the values in the above table of standard enthalpies of formation, calculate the ΔHreactiono for the formation of NO2(g).
Solution
N O2(g) is formed from the combination of N O (g)
and O 2(g)
in the following reaction:
2N O(g) + O2 (g) ⇋ 2N O2 (g)

To find the ΔHreactiono, use the formula for the standard enthalpy change of formation:
o o o
ΔH = ∑ ΔH (products) − ∑ ΔH (Reactants) (3.6.4)
reaction f f

The relevant standard enthalpy of formation values from Table 1 are:


O2(g): 0 kJ/mol
NO(g): 90.25 kJ/mol
NO2(g): 33.18 kJ/mol
Plugging these values into the formula above gives the following:

o
ΔH = (2 mol )(33.18 kJ/ mol ) − [(2 mol )(90.25 kJ/ mol ) + (1 mol )(0 kJ/ mol )] (3.6.5)
reaction

o
ΔH = −114.1 kJ (3.6.6)
reaction

Kirchhoff's Law describes the enthalpy of a reaction's variation with temperature changes. In general, enthalpy of any substance
increases with temperature, which means both the products and the reactants' enthalpies increase. The overall enthalpy of the
reaction will change if the increase in the enthalpy of products and reactants is different.

Kirchoff's Law - Enthalpy is Temperature Dependent


At constant pressure, the heat capacity is equal to change in enthalpy divided by the change in temperature.
ΔH
cp = (3.6.7)
ΔT

Therefore, if the heat capacities do not vary with temperature then the change in enthalpy is a function of the difference in
temperature and heat capacities. The amount that the enthalpy changes by is proportional to the product of temperature change and
change in heat capacities of products and reactants. A weighted sum is used to calculate the change in heat capacity to incorporate
the ratio of the molecules involved since all molecules have different heat capacities at different states.
Tf

HT = HT + ∫ cp dT (3.6.8)
f i

Ti

If the heat capacity is temperature independent over the temperature range, then Equation 3.6.7 can be approximated as

HTf = HTi + cp (Tf − Ti ) (3.6.9)

with
cp is the (assumed constant) heat capacity and
HTi and H are the enthalpy at the respective temperatures.
Tf

Equation 3.6.9 can only be applied to small temperature changes, (<100 K) because over a larger temperature change, the heat
capacity is not constant. There are many biochemical applications because it allows us to predict enthalpy changes at other
temperatures by using standard enthalpy data.

Contributors and Attributions


Janki Patel (UCD), Kostia Malley (UCD), Jonathan Nguyen (UCD), Garrett Larimer (UCD)

3.6: Thermochemistry is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

3.6.3 https://chem.libretexts.org/@go/page/41417
3.7: Bond Energies and Enthalpies
In the absence of standard formation enthalpies, reaction enthalpies can be estimated using average bond enthalpies. This method is
not perfect, but it can be used to get ball-park estimates when more detailed data is not available. A bond dissociation energy D is
defined by
XY (g) → X(g) + Y (g) (3.7.1)

with ΔH ≡ D(X − Y )
In this process, one adds energy to the reaction to break bonds, and extracts energy for the bonds that are formed.

ΔHrxn = ∑(bonds broken) − ∑(bonds formed) (3.7.2)

As an example, consider the combustion of ethanol:

In this reaction, five C-H bonds, one C-C bond, and one C-O bond, and one O=O bond must be broken. Also, four C=O bonds, and
one O-H bond are formed.

Bond Average Bond Energy (kJ/mol)

C-H 413

C-C 348

C-O 358

O=O 495

C=O 799

O-H 463

The reaction enthalpy is then given by


ΔHc = 5(413 kJ/mol) + 1(348 kJ/mol) + 1(358 kJ/mol)

+ 1(495 kJ/mol) − 4(799 kJ/mol)– 2(463 kJ/mol)

= − 856 kJ/mol (3.7.3)

Because the bond energies are defined for gas-phase reactants and products, this method does not account for the enthalpy change
of condensation to form liquids or solids, and so the result may be off systematically due to these differences. Also, since the bond
enthalpies are averaged over a large number of molecules containing the particular type of bond, the results may deviate due to the
variance in the actual bond enthalpy in the specific molecule under consideration. Typically, reaction enthalpies derived by this
method are only reliable to within ± 5-10%.

Contributors and Attributions


Patrick E. Fleming (Department of Chemistry and Biochemistry; California State University, East Bay)

3.7: Bond Energies and Enthalpies is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

3.7.1 https://chem.libretexts.org/@go/page/41418
3.E: Exercises
Q3.6
Calculate the work done by the reaction
Zn(s) + H SO (aq) → ZnSO (aq) + H (g) (3.E.1)
2 4 4 2

when 1 mole of hydrogen gas is collected at 273 K and 1.0 atm. (Neglect volume changes other than the change in gas volume, and assume
that the has behaves ideally.)

Q3.11a
If the internal energy of 3 moles of Cl2 is 3.96 KJ, what is the temperature of the system?

Q3.11b
The internal energy of 3 moles of neon gas (Ne) is 4.87 kJ. Assuming conditions are ideal, what is the temperature of the system?

S3.11b
Since we can assume that conditions are ideal, we know that the temperature is 25°C or 298 K. This assumption allows us to ignore the small
change in volume of the gas.

Q3.12a
Some ice cubes are placed inside a sealed plastic bag on a hot day, what is the sign of w , q, ΔH , ΔU ? What does this mean about the
system?

Q3.12b
A sports bottle containing Gatorade is held in a man's hand while running on a treadmill. Suppose the Gatorade is the system in question.
Answer the subsequent questions.
a. Will the temperature rise due to the movement caused by the man's running?
b. Was any heat actually added to the system? If so, what was the source?
c. Was work done onto the system?
d. In what ways has the internal energy of the system been affected, if affected at all?

S3.12b
a. The man's running motion causes the Gatorade to be shaken. The shaking motion adds energy to the system and increases the random
motion of the molecules. Since the motion of the molecules rises, so does the temperature of the system.
b. Yes, heat exchange did occur between the system and its surroundings (the heat given off from the man's hand).
c. Yes, work was done onto the system via the shaking motion.
d. The internal energy has been increased.

Q3.12c
Consider an ice cube melting.
a. Is the change of enthalpy positive or negative for this process?
b. Is the change of entropy positive or negative for this process?

S3.12c
a. Enthalpy is positive because the ice must gain heat in order to melt.
b. Entropy is also positive because the water molecules are gaining kinetic energy and moving to a more disordered state.

Q3.13
2.5 moles of CO2 expands against an external pressure of 1.5 atm. It was initially at 9.0 atm and 30ºC. The final volume is 35.0 L.
a. What is the final temperature?
b. Calculate q, w, and ΔU for the process.

S3.13
PV (1.5atm)(35L)
a) T = = = 255K
Rn L ⋅ atm
(.08206 )(2.5mol)
mol ⋅ K

3.E.1 https://chem.libretexts.org/@go/page/41329
P1 J 9.0atm kJ
b) w = −(nRT )ln = −(2.5mol)(8.314 )(255K)ln = −9.5
P2 mol ⋅ K 1.5atm mol

3 3 J kJ
qp = nC pT = n RT = (2.5mol) (8.314 )(255K) = 7.95
2 2 mol ⋅ K mol

kJ kJ kJ
ΔU = q + w = 7.95 + −9.5 = −1.55
mol mol mol

- For calculating work, expression for Isothermal reversible process has been used instead of correct -Pext*dV as it is a irreversible expansion
against external pressure. -Wrong method for calculating q. Instead Internal energy would be calculated using nCv*dT.

Q3.17
Determine the at 343.15 K for . (Note: The heat of vaporization for water is known to be 40.79 kJ mol-1 at 273.15 K, the of
was determined to be 85.6 J K mol-1 and the
-1
of was determined to be 54.35 J K-1 mol-1). Use the reaction below and give
your final answer in kJ mol-1.

S3.17

Q3.18
,Is (q = −w true for a cyclic process?

S3.18
Yes, q = −w is a true statement. It is true because ΔU for any cyclic process is equal to zero.

Q3.22
A 10.0 g sheet of gold with a temperature of 18.0°C is laid flat on a sheet of iron that weights 20.0 g and has a temperature of 55.6°C. Given
that the specific heats of Au and Fe are 0.129 J g-1 °C-1 and 0.444 J g-1 °C-1, respectively, what is the final temperature of the combined
metals? (Hint: The heat gained by the gold must be equal to the heat lost by the iron.)

Q3.23
You are cooking 100 grams of chicken soup at 300K. You know that it will take 500 J to raise the temperature of your soup to the desired
350K. Assuming constant pressure, calculate the molar heat capacity of this chicken soup.

Q3.24
Why does the molar heat of vaporization for water decrease as temperature increases?

Q3.24
The value of ∆H for heating 1 mole of hexane is 28.9kJ/mole at 342 K and 25.6 kJ/mole at 298 K. Explain why those values are different.

Q3.25
What is ΔH if of 3.00 moles of O2(g) are cooled from 175ºC to 35.0ºC and
¯ −3 −6 2 −1 −1
Cp = (7.50 × 10 T − 2.30 × 10 T Jmol K (3.E.2)

S3.25
¯
dH = nCp dT (3.E.3)

3.E.2 https://chem.libretexts.org/@go/page/41329
308K
−3 −6 2 −1 −1
ΔH = (3mol) ∫ (7.50 × 10 T − 2.30 × 10 T )Jmol K dT (3.E.4)
448K

−3 −6
7.50 × 10 2
2.30 × 10 3 −1 −1 308K
ΔH = (3mol)( T − T Jmol K |448K ) (3.E.5)
2 3

−3 2 −7 3
ΔH = (3mol)[(3.75 × 10 (308K ) − 7.67 × 10 (308K ) ) (3.E.6)

−3 2 −7 3 −1 −1
− (3.75 × 10 (448K ) − 7.67 × 10 (448K ) )]Jmol K

−1
ΔH = (3mol)[(333.3) − (683.7)]Jmol (3.E.7)

ΔH = −1050J (3.E.8)

Q3.25
For a constant-pressure process, the heat capacity per mole is given by the expression:
Cp per mole = (16.0 + 4.35 x 10-3 T + 1.15 T2) J K-1 mol-1
Find the molar heat capacity of oxygen gas from 38°C to 115oC

S3.25
115
−3 2
ΔH =∫ (16 + 4.35 × 10 T + 1.15 T )dT (3.E.9)
38

−3 2 3 −3 2 3
= (16 ∗ 115 + (4.35 × 10 )/2 ∗ 115 + 1.15/3 ∗ 115 ) − (16 ∗ 38 + (4.35 × 10 )/2 ∗ 38 + 1.15/3 ∗ 38 ) (3.E.10)

= 5.63J/mol (3.E.11)

Q3.26
What is the heat capacity ratio (γ) if of an ideal gas if its C V = 42 J K
−1
mol
−1
?

S3.26
¯
¯¯¯¯¯
¯ ¯
¯¯¯¯¯
¯
since C P = CV + R

¯
¯¯¯¯¯
¯ ¯
¯¯¯¯¯
¯
CP CV + R
γ = = (3.E.12)
¯
¯¯¯¯¯
¯ ¯
¯¯¯¯¯
¯
CV CV
−1 −1
(42 + 8.314)J K mol
γ = (3.E.13)
−1 −1
42J K mol

γ = 1.2 (3.E.14)

Q3.29
Explain (in terms of thermo chemistry) why Emperor Penguins huddle together in the Antarctic.

Q3.30a
Through laboratory experiment, a group of students found out that molar mass of a solid element can be calculated in term of the specific heat
capacity :
Molar mass x the specific heat of metals= 25 (Eq.1)
They were also obtaining the values of the three metals, Aluminum (0.900 J/g oC), Zinc (0.381 J/g oC), and Arsenic (0.329 J/g oC).
Determine which metal violate the above equation (Eq.1)

S3.30a
Aluminum (Al) can not be applied in this equation because the specific heat capacity of Al is approximately equal 1.0 J g-1 (oC-1)

Q3.30b
What is the molar heat capacity of lead given its specific heat is 0.128J/(gºC) at 25ºC?

S3.30b
The first thing we do is find the molar mass of lead which is 207.2 g/mol.
We multiply our molar mass by the specific heat to get the molar heat capacity.

3.E.3 https://chem.libretexts.org/@go/page/41329
Q3.31
What is the total work done on 2.00 moles of an ideal monatomic gas at 3.00 atm and 4.00 L when it adiabatically expands to 7.00 L?

S3.31
The equation for work is
¯
w = nCV (T2 − T1 ) (3.E.15)

For a monatomic gas,


3 5
¯ ¯
CV = R and CP = R (3.E.16)
2 2

¯
CP
γ = = 1.67 (3.E.17)
¯
CV

Solve for T1:


P V = nRT (3.E.18)

PV (3atm)(4L)
T1 = = (3.E.19)
nR L atm
(2moles)(0.08206 )
mol K

T1 = 73.12K (3.E.20)

Solve for T2:


γ γ
P1 V = P2 V (3.E.21)
1 2

1.67
(3atm)(4L)
P2 = = 1.178atm (3.E.22)
1.67
(7L)

PV (1.178atm)(7L)
T2 = = (3.E.23)
nR L atm
(2moles)(0.08206 )
mol K

T2 = 50.24K (3.E.24)

Solve for work:


¯
w = nCV (T2 − T1 ) (3.E.25)

3 J
w = (2mol)( )(8.314 )(50.24K − 73.12K) (3.E.26)
2 mol K

w = −571J (3.E.27)

Q3.31
You notice your system changes volume from 10 L to 30 L with a constant pressure at 770 torr. Then, there is a decrease in pressure to 740
torr with the volume staying constant at 30 L. Calculate the total amount of work done.

S3.31
We split up the solution into two parts.
Work done during the volume change and the work done during the pressure change.
First part:

Second part:
Convert pressure to atm:

3.E.4 https://chem.libretexts.org/@go/page/41329
=107.4 J
Then add up the two works
wtot= -2046.3+107.4=-1938.9 J

Q3.33
Calculate the values of q, w, ΔU, and ΔH for the reversible adiabatic expansion of 1 mole of a diatomic ideal gas from 10.00 m3 to 35.0 m3
with an initial temperature of 298 K.

Q3.33
Calculate the values of q, w, ΔU, and ΔH for the reversible adiabatic expansion of 1 mole of a diatomic ideal gas from 10.00 m3 to 40.0 m3
with an initial temperature of 298K.

Q3.34
A quantity of 0.27 mole of argon is confined in a container at 3.0 atm and 298 K and then allowed to expand adiabatically under two different
conditions: (a) reversibly to 0.80 atm and (b) against a constant pressure of 0.80 atm. Calculate the final temperature in each case.

Q3.34
In an adiabatic process 0.5 moles of He2 in a container are allowed to expand against a constant pressure of 1.00 atm and 298 K. The final
temperature after the gas has expanded is 205 K. Calculate the initial pressure inside the container.

S3.34
In an adiabatic process
CV = (T2 − T1 ) = −Pex (V2 − V1 )

we get
3 T2 T1
(T2 − T1 ) = −Pex ( − )
2 P2 P1

we rearrange and get


3 Pex T1
(T2 − T1 ) = − T2
2 P1

we solve for
P2
P1 and get P
1 =
T2 5 3
( )−
T1 2 2

1.00atm
P1 =
205K 5 3
( )−
298K 2 2

P1 = 4.5atm

Q3.34
A quantity of 0.27 mole of argon is confined in a container at 3.0 atm and 298 K and then allowed to expand adiabatically under two different
conditions:
1. reversibly to 0.80 atm and
2. against a constant pressure of 0.80 atm.
Calculate the final temperature in each case.

3.E.5 https://chem.libretexts.org/@go/page/41329
Q3.35a
Two moles of ideal diatomic gas H expand from 2.00 atm to 1.30 atm while being cooled from 550 K to 250 K. Compute the Change of
2

Enthalpy (∆H) and Change of Internal Energy (∆U) for 2 different processes: a. reversible and b. irreversible. Explain and calculate the
difference a and b?

Q3.35b
0.5 mole of Helium at 298 K are allowed to expand reversibly from 5 atm to 1 atm. Find q, w and ΔU if
a. the process is adiabatic and
b. if the process is isothermal.

Q3.36a
808 kJ/mol of heat is released by burning x (gram) amount of Iron in a constant volume bomb calorimeter. The calorimeter containing 100 g
water inside has a heat capacity of 2100 JoC-1. Evaluate the amount of Iron needed to increase the water by 5.20C?

Q3.36b
In your constant-volume bomb calorimeter you have a 0.3677 g sample of Potassium. The water surrounding the sample increase from 18.08
ºC to 20.667 ºC. The calorimeter has a heat capacity of 1515 J/(gºC). What is the change in internal energy?

S3.36b
Since the bomb calorimeter has a constant volume ΔV , there is no (PV) work so the first law of thermodynamics
ΔU = q + w (3.E.28)

can be simplified to
ΔU = q (3.E.29)

Then the relationship between heat and temperature change


q = cv ΔT (3.E.30)

can be changed
ΔU = m cv ΔT (3.E.31)

which for this case is


ΔU = (0.3677 g)(1, 515 J/(g C ))(20.667 º ºC − 18.08 ºC ) (3.E.32)

ΔU = 1411.1 (3.E.33)

Q3.37
The enthalpy of combustion for benzoic acid (C H C OOH ) is commonly used as the standard
6 5 ΔHcombustion = −3226.7 kJ mol
−1
, for
calibrating bomb calorimeters at a constant-volume process.
a. When 1.567 g of benzoic acid burned in a calorimeter the temperature increased by 3.678 C . Calculate the heat capacity of the
o

calorimeter.
b. For a different experiment 0.6792g of ethanol was burned in the calorimeter which rose from 16.42 C to 26.58 C. Calculate the enthalpy
o o

of combustion for ethanol, the value of ΔrU for the combustion, and the molar enthalpy of formation.

S3.37
(a)

qcalorimeter = −qbenzoic acid (3.E.34)

1.567g
o −1
qcalorimeter (3.678 C ) = −( )(−3226.7kj mol ) (3.E.35)
122.1g mol−1

41.41kj
o −1
ccalorimeter = = 11.26kj C (3.E.36)
o
3.678 C

(b) (Hint: Calorimeter absorbs heat transfered from the combustion of maltose)

qcalorimeter = −qmaltose (3.E.37)

Step 1: Calculate ΔrU

3.E.6 https://chem.libretexts.org/@go/page/41329
o −1 o o
.6792g o
(11.26kj C )(26.58 C − 16.42 C ) = −( )(ΔrU ) (3.E.38)
−1
46.06844gmol

114.4016kj
o 3 −1
Δ rU =− = −7.759 × 10 jk mol (3.E.39)
.014743282mol

Step 2: Calculation of enthalpy of combustion


C2 H 5OH(s) + 3 O2 (g)
→ 2C O2(g) + 3 H2 O(l) (3.E.40)

Δn = −1

Therefore,
o o 3 − −3 −1 −1
ΔrH = ΔrU + RT Δn = (−7.759 × 10 kj mol 1) + (8.3146 × 10 kj mol K )(298.15K)(−1) = −7.761 (3.E.41)

3 −1
× 10 kj mol

Step 3: Calculate the molar enthalpy of formation.


o o o o
¯ ¯ ¯ ¯
ΔrH = 2 Δf H [C O2(g) ] + 3 Δf H [ H2 O(l) ] − Δf H [ C2 H5 OH(s) ] − 3 Δf H [ O2(g) ] (3.E.42)

−1 −1 −1 3 −1
= 2(−393.5kj mol ) + 3(−285.8kj mol ) − 3(0kj mol ) − (−7.761 × 10 kj mol ) (3.E.43)

3 −1
= 6.116 × 10 kj mol (3.E.44)

Q3.38a
A quantity of 2.50 x 102 mL of 0.468 M HBr is mixed with 2.50 x 102 mL of 0.425 M KOH in a constant-pressure calorimeter that has a heat
capacity of 437 J °C-1. The initial temperature if the HBr and KOH solutions is the same at 21.35°C. For the process
+ −
H (aq) + OH (aq) ⇌ H2 O(l) (3.E.45)

the heat of neutralization is -56.2 kJ mol-1. What is the final temperature of the mixed solution?

Q3.38b
In a constant-pressure calorimeter with a heat capacity of 395J C , you mix 1.5 × 10 mL of HBr at 0.788M and 1.5 × 10 mL
o −1 3 3

C a(OH ) 2at .394M. For this process to occur the heat of neutralization is −56.2kj mol . The initial temperature of both solutions is −1

23.08 C .
o

+ −
H + OH → H2 O(l) (3.E.46)
(aq) (aq)

Find the final temperature of the mixed solution.

S3.38b
Step 1: (Hint: Determine the number of moles of the reactants)
−1
nH + = (1.5L)(0.788 mol L ) = 1.182 mol (3.E.47)

−1
nOH − = 2(1.5L)(0.394 mol L ) = 1.182 mol (3.E.48)

Step 2: (Hint: Calculate the the thermal energy released by the reaction)
−1
qP = (−52.2 kJ mol )(1.182 mol) = −66.42 kJ (3.E.49)

Step 3: (Hint: Assume the the density of the solutions are the same as water and specific heat is same as water too.)

s_{solution}=4.184\,J\, g^{-1}^{o}C^{-1}

msolution = mH Br + mC a(OH ) (3.E.50)


2

3 −1 3 −1 3
= (1.5 × 10 mL)(1.00gm L ) + (1.5 × 10 mL)(1.00gm L ) = 3.00 × 10 g (3.E.51)

Step 4:

o −1 3 −1 o −1
1kJ
66.4284 kJ = 395 J C + (3.00 × 10 g)(4.184 J g C )( )ΔT (3.E.52)
1000 J

o −1
66.4284 = (12.947 C )ΔT (3.E.53)

o
ΔT = 5.13 C (3.E.54)

3.E.7 https://chem.libretexts.org/@go/page/41329
Step 5:
ΔT = Tf − Ti (3.E.55)

o
Tf = 28.21 C (3.E.56)

Q3.38
A quantity of 2.50 × 10 mL of 0.468; M H Br is mixed with 2.50 × 10 mL of 0.425 M KOH in a constant-pressure calorimeter that
2 2

has a heat capacity of 437 J °C-1. The initial temperatures of the H Br and KOH solutions are the same at 21.35 °C. The heat of
neutralization for the below reaction is \(-56.2\; kJ/mol\)
+ −
H + OH → H2 O(l) (3.E.57)
(aq) (aq)

What is the final temperature of the mixed solution?

Q3.39
1.52 g of toluene (C H ) is combusted in a constant volume bomb calorimeter at 25°C. Calculate the Δ
7 8
¯
rU and Δ
rH
¯
for this reaction given
that 39.183 kJ of heat was released.

S3.39
First we need to construct the balanced reaction that occurs
C7 H8 (l) + 9 O2 (g) → 7C O2 (g) + 4 H2 O(l) (3.E.58)

−39.183kJ
¯
Δr U = qr = −qcal = = −2375 kJ/mol
1.52g

92.14(g/mol)

Δr H̄ = Δr Ū + Δ(P V ) = Δr Ū + Δn(RT )

−3
= −2375 kJ/mol + (−2)(8.314 × 10 kJ/K ∗ mol)(298.15 K) = −2380 kJ/mol

Q3.39
The of ΔU a of the combustion of naphthalene (C10H8) are -480 kj mol-1. The heat evolved from the combustion in a constant-volume
calorimeter is 7.5 KJ. Calculate the mass of naphthalene in grams in the calorimeter.

Q3.40
For the reaction

C2 H5 OH(l) + O2(g) → C O2(g) + H2 O(l) (3.E.59)

a. Balance the reaction


b. Calculate the Δ H of reaction using the Δ
r f H data

Q3.40B
Consider the following reaction:

7
C2 H6 (g) + O2 (g) → 2C O2 (g) + 3 H2 O(l)
2

Δr H = −1560kJ/mol

What is the value of Δ rH



if
a. the equation is multiplied throughout by 4;
b. the direction of the reaction is reversed;
c. C H was a liquid instead of gas?
2 6

S3.40B
a. Δ H = 4(−1560kJ/mol) = −6240kJ/mol;
r

b. Δ H = 1560kJ/mol;
r

c. Ethane exist as a liquid at temperatures of below 184.6 K. If we used ethane liquid instead of ethane gas, then according to

ΔH = ΔU + Δn(RT ) (3.E.60)

3.E.8 https://chem.libretexts.org/@go/page/41329
the number of moles would change and the temperature of the reaction would also change.

Q3.41a
On what basis is the 0 value for standard enthalpy of formation assigned to a substance at a specific temperature? Base on that, identify which
substances have ΔfH° = 0 and which do not at 1 atm and 25°C (give explanation for those that do not).
H2(g), O3(g), Al(s), Br2(g), NaCl(s), Cl2(g)
Hint: Look at standard enthalpy of formation

S3.41a
0 value for standard enthalpy of formation is assigned to elemental substances that are in their natural/standard physical state at that specific
condition of temperature and pressure.
a. H2(g), Al(s), Cl2(g) are substances with ΔfH° = 0 kJ/mol
b. O3(g) is an elemental substance but does not occur naturally at 25°C, O2(g) is.
c. Br2(g) is also not the standard state of bromine. At STP, the physical state of bromine is liquid
d. NaCl does not have a 0 value because it is not an elemental substance

Q3.41b
Which of the following substances has a standard enthalpy of formation, at 298 K, of 0 K?
O2(g), Br2(g), Hg(s), CH4(g)

Q3.41c
Why is the standard enthalpy of formation of O2 zero at a temperature of 298 K?

S3.41c
The standard enthalpy of formation of O2 is zero at 298 K since O2 is the most stable allotropic form of oxygen at that particular temperature.

Q3.42a
a) The ionization of a weak acetic acid is written as:
− +
C H3 C OOH (aq) ⇌ C H3 C OO +H (3.E.61)

Suppose the change of reaction enthalpy for this reaction is +1.0 kJ/mol and the ΔfH° of CH3COO- is -486.01 kJ/mol. What would be the
molar enthalpy of formation of aqueous acetic acid?
b) The acetic acid is reacted with a strong base NaOH. Calculate the enthalpy of neutralization for this reaction of weak acid strong base.
ΔfH°[OH-(aq)] = -229.6 kJ/mol and ΔfH°[H2O(l)] = -285.8 kJ/mol
Hint: Look at enthalpy of neutralization

S3.42a
a)
∘ ∘ ∘
Δr H̄ = ∑ [ Δf H̄ ]reactants − ∑ [ Δf H̄ ]products (3.E.62)

∘ ∘ − + ∘ ∘
¯ ¯ ¯ ¯
Δr H = Δf H [C H3 C OO ] + Δf H [ H ] − Δf H [C H3 C OOH ] (3.E.63)

kJ kJ kJ ∘
¯
+1.0 = −486.01 +0 − Δf H [C H3 C OOH ] (3.E.64)
mol mol mol

∘ kJ
¯
Δf H [C H3 C OOH ] = −487.01 (3.E.65)
mol

b) Note that acetic acid is a weak acid, therefore only dissociates partially or does not dissociate at all. Therefore the complete ionic reaction
of reaction is:
+ − − + 2
C H3 C OOH (aq) + N a (aq) + OH (aq) = C H3 C OO (aq) + N a (aq) + H O(l) (3.E.66)

+
Since Na cancels out, the net equation is
− − 2
C H3 C OOH (aq) + OH (aq) = C H3 C OO (aq) + H O(l) (3.E.67)

∘ ∘ ∘ ∘ ∘
− −
Δr H̄ = Δf H̄ [C H3 C OO ] + Δf H̄ [ H2 O] − Δf H̄ [C H3 C OOH ] − Δf H̄ [OH ] (3.E.68)

3.E.9 https://chem.libretexts.org/@go/page/41329
∘ kJ
¯
Δr H = (−486.01 − 285.5 + 486.01 + 229.6) (3.E.69)
mol

∘ kJ
¯
Δr H = −55.9 (3.E.70)
mol

Q3.42b
Calculate the enthalpy of formation for Br- ions in the following reactions:
a. H Br(g) → C l
− +
(aq) + Br (aq)

kJ
b. Δr H

= −105.7
mol

S3.42b
∘ ∘ ∘
∘ ¯
¯¯¯
¯ + ¯
¯¯¯
¯ − ¯
¯¯¯
¯
Δr H = Δf H (H ) + Δf H (Br ) − Δf H (H Br) (3.E.71)

∘ ∘ ∘
¯
¯¯¯
¯ − ∘ ¯
¯¯¯
¯ + ¯
¯¯¯
¯
Δf H (Br ) = Δr H − Δf H (H ) + Δf H (H Br) (3.E.72)

¯
¯¯¯
¯
∘ kJ kJ

Δf H (Br ) = 36.4 − 0 + (−.105.7 ) (3.E.73)
mol mol

¯
¯¯¯
¯
∘ kJ

Δf H (Br ) = 69.3 (3.E.74)
mol

Q3.42c
o
When determining standard enthalpies of formation of ions in aqueous solutions, set the value of H +
ion to 0 for Δ f
¯
H [H
+

(aq)
] = 0.

o
a. For the following reaction H Br (g) → H
+

(aq)
+ Br

(aq)
, ΔrH o
= −98.4kj mol
−1
. Calculate the value of Δ f
¯
H for the Br ions.
b. The standard enthalpy of neutralization between HBr and NaOH solutions is −53.4kj mol −1
. At 25
)
o
C , calculate the standard enthalpy
of formation of the hydroxide ion.

S3.42c
(a)
$$\Delta rH^{o}=-98.4kj\,mol^{-1}=\Delta _f\bar{H}^{o}[H^+_{(aq)}]+\Delta _f\bar{H}^{o}[Br^-_{(aq)}]-\Delta _f\bar{H}^{o}
[HBr_{(g)}]\]
$$\Delta _f\bar{H}^{o}[Br^-_{(aq)}]=-98.4kj\,mol^{-1}-\Delta _f\bar{H}^{o}[H^+_{(aq)}]+\Delta _f\bar{H}^{o}[HBr_{(g)}]\]
$$=-98.4kj\,mol^{-1}-0kj\,mol^{-1}+(-36.29kj\,mol^{-1})\]
$$=-134.69kj\,mol^{-1}\]
(b) The neutralization for 1 mole of H 2O is ΔrH o
= −53.4kj mol
−1

$$\Delta rH^{o}=\Delta_f\bar{H}^{o}[H_2O_{(l)}]-\Delta_f\bar{H}^{o}[H^+_{(aq)}]-\Delta_f\bar{H}^{o}[OH^-_{(aq)}]\]
$$\Delta_f\bar{H}^{o}[OH^-_{(aq)}]=\Delta_f\bar{H}^{o}[H_2O_{(l)}]-\Delta_f\bar{H}^{o}[H^+_{(aq)}]+53.4kj\,mol^{-1}\]
$$=(-285.8kj\,mol^{-1})-0kj\,mol^{-1}+53.4kj\,mol^{-1}\]
$$=-232.4kj\,mol^{-1}\]

Q3.42d
For the following chemical reaction, calculate the ΔrH .
º
2− +
H3 P O4 (aq) → H2 P O (aq) + H (aq) (3.E.75)
4

∘ −1
△f H = −1277 kJ mol (3.E.76)
H3 P O4 (aq)

∘ −1
△f H +
= 0 kJ mol (3.E.77)
H (aq)

∘ −1
△f H 2−
= −1302.5 kJ mol (3.E.78)
H2 P O (aq)
4

S3.42d
Use the following equation:
o o o
ΔH = ∑ ΔH (products) − ∑ ΔH (Reactants) (3.E.79)
reaction f f

Solve

3.E.10 https://chem.libretexts.org/@go/page/41329
o o 2− o + o
ΔH = ΔH (H2 P O ) + ΔH (H ) − ΔH (H3 P O4 ) (3.E.80)
reaction f 4 f f

−1 −1
= −1302.5 kJ mol − (−1277 kJ mol ) (3.E.81)

−1
= −25.5 kJ mol (3.E.82)
––––––––––––––––––

Click here for more information about standard enthalpies of reaction.

Q3.42e
The reaction for the dissociation of hydrobromic acid in water is as follows:
KJ
+ − ∘
H Br(g) ⇌ H + Br Δr H = −85.2 (3.E.83)
(aq) (aq)
mol

KJ

Δf H (H Br) = −36.4 (3.E.84)
mol

a.) Given the enthalpy of formation for HBr, what is the enthalpy of formation for the Bromine ion? Assume that the enthalpy of formation
for the Hydrogen ion is 0.
b.) What is the enthalpy of reaction for the neutralization of HBr with KOH? Assume that the potassium and bromine ions remain in solution
following the neutralization.

∘ −
KJ ∘
KJ
Δf H (OH ) = −36.4 Δf H (H2 O(l) ) = −285.8 (3.E.85)
(aq)
mol mol

S3.42e
a.) For any reaction,
∘ ∘
∘ ¯ ¯
Δr H = ∑ ν Δf H (products) − ∑ ν Δf H (reactants) (3.E.86)

For this reaction,


∘ ∘ ∘
∘ − +
Δr H = Δf H̄ (Br ) + Δf H̄ (H ) − Δf H̄ (H Br(aq) ) (3.E.87)
(aq) (aq)

∘ ∘
− ∘
Δf H̄ (Br ) = Δr H + Δf H̄ (H Br(aq) ) (3.E.88)
(aq)

KJ KJ KJ
= (−85.2 ) + (−36.4 ) = −121.6 (3.E.89)
mol mol mol

b.) For any neutraliztion reaction in solution,


− +
OH +H ⇌ H2 O(l) (3.E.90)
(aq) (aq)

∘ ∘
∘ ¯ ¯ −
Δr H = Δf H (H2 O(l) ) − Δf H (OH ) (3.E.91)
(aq)

KJ KJ KJ
= −285.8 − (−36.4 ) = −249.4 (3.E.92)
mol mol mol

Q3.43a
How much heat is given off when one explodes 24.65 grams of nitroglycerin? The equation for the combustion of nitroglycerin is given
below:
4 C3 H5 N3 O9 (l) → 12C O2(g) + 10 H2 O(g) + 6 N2 (g) + O2 (g) (3.E.93)

∘ −1
△r H = −1.529 M J mol (3.E.94)
C3 H5 N3 O9 (l)

S3.43a
The molar mass of nitroglycerin is 227.0865 g/mol.
To solve this question simply multiple the enthalpy of reaction by the amount moles of nitroglycerin.
−1
24.65 g ÷ 227.0865 g mol = 0.109 moles of nitroglycerin (3.E.95)
−1
1.529 M J mol ⋅ 0.109 mol = 0.166M J of heat (3.E.96)
–––––––––––––––––

Q3.43b
What is the enthalpy change from the formation of 5 grams of H2CO3 via CO2(g) + H2O(g) àH2CO3(g) H2CO3 (g) via CO2 (g) + H2O (g) →
H2CO3 (g)? The change in enthalpy associated with this reaction is +360.19 kJ/mol.

3.E.11 https://chem.libretexts.org/@go/page/41329
S3.43b
The enthalpy change from this reaction is given by the molar enthalpy change times the number of moles of product formed. Thus:
mco2 = 44.01 g mol-1 ;
nco2 = 5g /44.01 g mol-1 = 0.1136 mol
ΔH =0.1136 mol * 360.19 kJ/mol = 40.9kJ

Q3.43c
The following reaction was carried out at standard temperature and pressure:
KJ

2 C(graphite) + 3 H2(g) → C2 H6(g) Δr H = −84.7 (3.E.97)
mol

If -148.16 KJ of heat was released during this reaction, how many grams of Ethane were formed?

S3.43c
For any reaction:

m ∘
q = nΔf H = Δf H (3.E.98)
M

This can be rearranged as


qM
m = (3.E.99)

Δf H

The molar mass of ethane is


g g g
M = 2(12.01 ) + 6(1.008 ) = 30.07 (3.E.100)
mol mol mol

The mass of ethane produced, then, is


g

qM (−148.16KJ)(30.07 )
mol
m = = = 52.6g (3.E.101)
∘ KJ
Δf H −84.7
mol

Q3.44a
When 2.3 g of calcium carbonate is decomposed under constant pressure, 178.3 kJ of heat were released. What is the Δ rH

for this reaction?

C aC O3 (s) → C aO(s) + C O2 (g) (3.E.102)

S3.44a

178.3kJ
Δr H = qp = = 7759.15kJ/mol (3.E.103)
2.3g

100.09g/mol

Q3.44b
In a high school chemistry experiment, 10 grams of methane were combusted in the presence of excess oxygen under standard conditions. A
calorimeter measured that 554.9 kJ of heat were released by the reaction to the surroundings. Assuming the calorimeter’s measurement
accounts for all heat changes associated with the reaction, write the reaction for the combustion of methane and calculate the change in
enthalpy associated with it.

Q3.46
The isomerization process from 1-butene to 2-butene has a reaction enthalpy of -7.1 kJ/mol.

1-butene 2-butene
If the hydrogenation of 1 butene has a reaction enthalpy of -126.8 kJ/mol. How much energy would the hydrogenation of 2-butene takes?
Keep in mind that both processes yield the same product.
Hint: Look at example problems of change of standard reaction enthalpy

3.E.12 https://chem.libretexts.org/@go/page/41329
S3.46
1-butene + 2H2 → butane ΔfH° = -126.8 kJ/mol
1-butene → 2-butene ΔfH° = -7.1 kJ/mol
Reverse the 2nd reaction and combine the two reactions
1-butene + 2H2 → butane ΔfH° = -126.8 kJ/mol
2-butene → 1-butene ΔfH° = +7.1 kJ/mol
ΔrH° = -126.8 kJ/mol +7.1 kJ/mol
ΔrH° = -119.7 kJ/mol

Q3.47a
From the reaction below, calculate the enthalpy of formation of methane gas.
C H4 (g) + 2 O2 (g) → C O2 (g) + 2 H2 O(l) (3.E.104)
∘ −1
△r H = −891kJ mol (3.E.105)

∘ −1
△f H = −393.5kJ mol (3.E.106)
C O2 (g)

∘ −1
△f H = −285.8kJ mol (3.E.107)
H2 O(l)

S3.47a
Use the following equation:
o o o
ΔH = ∑ ΔH (products) − ∑ ΔH (Reactants) (3.E.108)
reaction f f

Plug in appropriate values and solve. Change in enthalpy of oxygen is 0 because it is the reference form of a pure element.
o o o o o
ΔHr = ΔH (C O2 ) + (2 ⋅ ΔH (H2 O)) − (2 ⋅ ΔH (O2 )) − ΔH (C H4 ) (3.E.109)
f f f f

−1 −1 −1 o
−891 kJ mol = (−393.5 kJ mol ) + (−285.8 kJ mol ) − (ΔH (C H4 )) (3.E.110)
f

o −1
ΔH (C H4 ) = 211.7 kJ mol (3.E.111)
f –––––––––––––––

Click here for more information on standard enthalpies.

Q3.47b
For the following combustion reaction:


KJ
C3 H7 OH(l) + 3 O2(g) → 4 H2 O(l) + 3C O2(g) Δr H = −2005.8 (3.E.112)
mol

Calculate the enthalpy of formation for 2-propanol, given the following enthalpies of formation
∘ KJ ∘ KJ
Δf H̄ (H2 O) = −285.8 Δf H̄ (C O2 ) = −393.5 (3.E.113)
mol mol

S3.47b
In general:
∘ ∘ ∘
¯ ¯
Δr H = ∑ ν Δf H (products) − ∑ ν Δf H (reactants) (3.E.114)

In general,
∘ ∘ ∘
∘ ¯ ¯ ¯
Δr H = 3 Δf H (C O2 ) + 4 Δf H (H2 O) − Δf H (C3 H7 OH ) (3.E.115)

∘ ∘ ∘
¯ ¯ ¯ ∘
Δf H (C3 H7 OH ) = 3 Δf H (C O2 ) + 4 Δf H (H2 O) − Δr H (3.E.116)

KJ KJ KJ KJ
= 3 (−393.5 ) + 4 (−285.8 ) − (−2005.8 ) = −317.9 (3.E.117)
mol mol mol mol

Q3.48
Diatomic fluorine gas has a Cp of 31.3 J/K*mol. What is its molar enthalpy of formation on the surface of the planet Venus (at standard
pressure), where temperature is commonly 465 C?

3.E.13 https://chem.libretexts.org/@go/page/41329
S3.48
At constant pressure, the change in enthalpy with respect to temperature is given by Cp. By determining the change in temperature and
multiplying it by Cp we can find the change in enthalpy.
ΔT = 465 C - 25 C = 440K
ΔH = 31.3 J/K*mol * 440K = 13.772 kJ/mol

Q3.50
The combustion of graphite is
C(gr) + O2(g) → CO2(g)
Calculate the change in the enthalpy of combustion from 298 K to 398 K. The molar Cp values are (in J K-1 mol-1): C(gr):8.52, O2(g):29.4,
and CO2(g):37.1

Q3.53
What values are needed to calculate the change in standard enthalpy of a reaction?

Q3.54
Consider the following reactions:
C(graphite) + O2 (g) → CO (s) ∆rH = -110.5 kJ/mole
CO2 (g) → Co (s) + O2 (g) ∆rH = 283.0 kJ/mole
Find the enthalpy of formation of carbon dioxide (CO2)

S3.54
The ΔH of the reaction is the sum of ΔH and ΔH1 2

\[2C + O_2 \rightarrow 2CO \,\,, \Delta H_1 = -110.5 kJ/mol


\[2CO + O2 \rightarrow 2CO_2 \Delta H_2 = -283 kJ/mol
thus the ΔH of reaction is (-110.5)+(-283.0) = -393.5 kJ/mol

Q3.55
What is ΔrHº of the following reaction? ΔfHº for O(g) is 249.4 kJ mol-1 and ΔfHº for O3(g) is 142.7 kJ mol-1

2 O2(g) → O(g) + O3(g) (3.E.118)

S3.55
∘ ∘ ∘
ΔHrxn = ∑ ΔH − ∑ ΔH (3.E.119)
prod reactants

∘ ∘ ∘ ∘
ΔHrxn = [ΔH (O(g)) + ΔH (O3 (g))] − 2ΔH (O2 (g)) (3.E.120)

kJ

ΔHrxn = (249.4 + 142.7) (3.E.121)
mol

kJ

ΔHrxn = 392.1 (3.E.122)
mol

Q3.60
Methane combusts, what is the enthalpy of combustion for this reaction?

Substance Standard Molar Enthalpies of Formation at 25ºC

CH4(g) -74.85 kj/mol

CO2(g) -393.5 kj/mol

H2O(g) -241.8 kj/mol

H2O(l) -285.8 kj/mol

3.E.14 https://chem.libretexts.org/@go/page/41329
S3.60
Write the balanced formula for this reaction to identify the coefficients

C H4(g) + 2 O2(g) → 2 H2 O(l) + C O2(g) (3.E.123)

ΔH = ∑ Hproducts − sum Hreactants (3.E.124)

= 2xH (H2 O) + H (C O2 ) − H (C H4 ) (3.E.125)

We do not include O because it is a standard state


2

= [(2x − 285.8 + −393.5) − (−74.85)]kJ/mol (3.E.126)

= −890.25 kJ/mol (3.E.127)

Q3.62
Predict whether the values of q, w, ΔU, and ΔH are positive, zero, or negative for each of the following processes: (a) freezing of ice at 1 atm
and 273 K, (b) melting of solid butanol at 1 atm and the normal melting point, (c) reversible isothermal compression of an ideal gas, and (d)
reversible adiabatic compression of an ideal gas.

Q3.62
Predict whether the values of q, w, ΔU, and ΔH are positive, zero, or negative for each of the following processes: (a) freezing of ice at 1 atm
and 273 K, (b) melting of solid butanol at 1 atm and the normal melting point, (c) reversible isothermal compression of an ideal gas, and (d)
reversible adiabatic compression of an ideal gas.

Q3.63
The first law of thermodynamics states that energy is conserved. It cannot be destroyed nor created. Using Einstein’s equation E= mc2 explain
how large amounts of energy are necessary to fuse atoms together. If energy is conserved where did all this energy go?

S3.63
When large amounts of energy are used to fuse atoms together (eg: fusion) we can observe that the mass of the atoms is very small compared
to the amount of energy used. However, using Einstein's equation E= mc2 we see that the mass is multiplied times the speed of light squared
which will result in a very large number. The energy was not destroyed but rather stored in the chemical bonds of the atoms.

Q3.64
For the nuclear process we cannot apply the assumption that Δ f H =0 for most stable elements because there are different components in the
reactant and product sides. Explain this difference?

Q3.66
You eat 2 pounds of fries for lunch and the fuel value is approximately 2.9kcal g . If you don't store any of the energy, calculate the amount
−1

of water (evaporated perspiration) needed to keep your body temperature constant.

S3.66
Step1: (Hint: Calculate the fuel value for 2 pound of fries)
454g f ries 2.9kcal f ries 4.184kJ 4
(2 pounds of f ries)( )( )( ) = 1.1017 × 10 kj (3.E.128)
1 pound f ries g f ries 1kcal

Step 2: (Hint: Use Δ vap H to calculate grams of H 2O )

H2 O(l) → H2 O(g) (3.E.129)

−1
Δvap H = 44.01kjmol at 298K (3.E.130)

4
1.1017 × 10 jk 18.02 H2 O 3
( )( ) = 4.51 × 10 g H2 O (3.E.131)
−1
44.01kj mol 1mol H2 O

Q3.68
What is the final temperature of the following reaction:
C3 H8(g) + 5 O2(g) → 3C O2(g) + 4 H2 O(l) (3.E.132)

Hint: first find Δ


rH

, then look up the standard molar heat capacities of the products.

3.E.15 https://chem.libretexts.org/@go/page/41329
S3.68
∘ ∘ ∘
Δr H = ∑ nΔf H − ∑ nΔf Hreactant
product

= 3(−393.5kJ/mol) + 4(−241.8kJ/mol) − (−104.6kJ/mol) − 5(0kJ/mol)

= −2043.1kJ/mol

∘ ¯ ¯
Δr H = qp,rxn = −qp,absorbedbyproduct = (nC p,C O2 + nC p, H2 O )ΔT

−qp,absbyprod
ΔT = Tf − Ti =
¯ ¯
nC p,C O2 + nC p, H2 O

−qp,absbyprod
Tf = Ti +
¯ ¯
nC p,C O2 + nC p, H2 O

2043.1kJ/mol
= 298.15K + = 8613.6K
−3 −3
3(37.1 ∗ 10 kJ/K ∗ mol) + 4(33.6 ∗ 10 kJ/K ∗ mol)

Q3.70
The reaction below is a oxidation of 2-propanol to make acetone

a) Calculate the change of enthalpy of reaction by using bond enthalpies


b) Compare the result with the reaction of the combustion 2-propanol that has an enthalpy of -1825.5 kJ/mol. What does this say about the
two reactions. (Hint: Look at bond enthalpies).

S3.70
a)

Types of bonds formed #of bonds formed Bond Enthalpy (kJ/mol) Enthalpy change (kJ/mol)

C-H 6 414 2484

C-C 2 347 694

C=O 1 724 724

O-H 1 460 460


Δr H = ∑ BE(reactants) − ∑ BE(products) (3.E.133)

kJ kJ
= [(2898 + 694 + 351 + 460) − (2484 + 694 + 724 + 460)] = +41
mol mol

b) According to the enthalpy values, the enthalpy of the oxidation of 2-propanol into ketone is positive while its combustion is largely
negative. This tells that although oxygen is involved, the two reactions are different. The combustion releases heat where as the other process
needs an absorption of energy in order for the reaction to take place. In this case is usually the help of another catalyst.

Q3.71a
The molar enthalpy of sublimation for an unknown organic substance is found to be 58.05 KJ/mol, and the molar enthalpy of vaporization is
37.12 KJ/mol. Using these values, provide an estimate for the molar enthalpy of fusion for this substance.
Hint: Sublimination is the transition from solid to gas phase. Fusion and Vaporization, respectively, are the the transition from solid
to liquid, then liquid to gas.

S3.71a
In general,
∘ ∘ ∘
¯ ¯ ¯
Δsub H = Δvap H + Δf us H (3.E.134)

∘ ∘ ∘ KJ KJ KJ
¯ ¯ ¯
Δf us H = Δsub H − Δvap H = 58.05 − 37.12 = 20.93 (3.E.135)
mol mol mol

Q3.71b
º º
Calculate the standard enthalpy of sublimation of ethanol. ΔfusH (ethanol) = 4.9 kJ mol-1 and ΔvapH (ethanol) = 38.56 kJ mol-1.

3.E.16 https://chem.libretexts.org/@go/page/41329
S3.71b
You can estimate the standard enthalpy of sublimation by adding up the standard enthalpies of fusion and vaporization since sublimation
consists of two simultaneous phase changes. See here for more information on it.
º
ΔsubH (ethanol) = 4.9+38.56 = 43.46 kJ mol-1

Q3.72
Nitric acid is neutralized by hydroxide via the reaction HNO3 (aq) + OH-(aq) → H2O(l) + NO3-(aq). The dissociation of water has an enthalpy
change of 55.8 kJ/mol. What is the enthalpy change for the dissociation of nitric acid?

Chemical HNO3 OH- H2O NO3-

ΔfH (kJ/mol) -207.6 -229.6 -285.8 -206.6

S3.72
The dissociation of water is H2O(l) -> H+(aq) + OH-(aq), which when added to the neutralization reaction above, gives the overall reaction
+ −
H N O3 (aq) → H (aq) + N O (aq) (3.E.136)
3

which is the dissociation reaction for nitric acid. The enthalpy change of the first reaction is given by the difference in the total enthalpies of
formation of the products and reactants, i.e.
ΔrH=(-206.6 kJ/mol -285.8 kJ/mol) - (-229.6 kJ/mol -207.6 kJ/mol) = -55.2 kJ/mol.
When added to the enthalpy change for the dissociation of water to obtain the enthalpy change for the dissociation of nitric acid, we have:
ΔnetH=-55.2 kJ/mol + 55.8 kJ/mol = 0.6 kJ/mol

3.E: Exercises is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

3.E.17 https://chem.libretexts.org/@go/page/41329
CHAPTER OVERVIEW

4: The Second Law of Thermodynamics


4.1: Spontaneous Processes
4.2: Entropy
4.3: The Second Law of Thermodynamics
4.4: The Third Law of Thermodynamics
4.5: Evaluating Entropy and Entropy Changes
4.6: Gibbs Energy
4.7: Standard Molar Gibbs Energy of Formation
4.8: Dependence of Gibbs Energy on Temperature and Pressure
4.9: Phase Equilibria
4.10: Thermodynamics of Rubber Elasticity
4.E: Exercises

4: The Second Law of Thermodynamics is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1
4.1: Spontaneous Processes
4.1: Spontaneous Processes is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

4.1.1 https://chem.libretexts.org/@go/page/41419
4.2: Entropy
Entropy is a state function that is often erroneously referred to as the 'state of disorder' of a system. Qualitatively, entropy is simply
a measure how much the energy of atoms and molecules become more spread out in a process and can be defined in terms of
statistical probabilities of a system or in terms of the other thermodynamic quantities. Entropy is also the subject of the Second and
Third laws of thermodynamics, which describe the changes in entropy of the universe with respect to the system and surroundings,
and the entropy of substances, respectively.

Statistical Definition of Entropy


Entropy is a thermodynamic quantity that is generally used to describe the course of a process, that is, whether it is a spontaneous
process and has a probability of occurring in a defined direction, or a non-spontaneous process and will not proceed in the defined
direction, but in the reverse direction. To define entropy in a statistical manner, it helps to consider a simple system such as in
Figure 4.2.1. Two atoms of hydrogen gas are contained in a volume of V . 1

Figure 4.2.1 . Two hydrogen atoms in a volume V 1

Since all the hydrogen atoms are contained within this volume, the probability of finding any one hydrogen atom in V is 1. 1

However, if we consider half the volume of this box and call it V ,the probability of finding any one atom in this new volume is ,
2
1

since it could either be in V or outside. If we consider the two atoms, finding both in V , using the multiplication rule of
2 2

probabilities, is
1 1 1
× = . (4.2.1)
2 2 4

For finding four atoms in V would be


2

1 1 1 1 1
× × × = . (4.2.2)
2 2 2 2 16

N
Therefore, the probability of finding N number of atoms in this volume is 1

2
. Notice that the probability decreases as we increase
the number of atoms.
If we started with volume V and expanded the box to volume V , the atoms would eventually distribute themselves evenly
2 1

because this is the most probable state. In this way, we can define our direction of spontaneous change from the lowest to the
highest state of probability. Therefore, entropy S can be expressed as
S = kB ln Ω (4.2.3)

where Ω is the probability and k is a proportionality constant. This makes sense because entropy is an extensive property and
B

relies on the number of molecules, when Ω increases to W , S should increase to 2S. Doubling the number of molecules doubles
2

the entropy.
So far, we have been considering one system for which to calculate the entropy. If we have a process, however, we wish to
calculate the change in entropy of that process from an initial state to a final state. If our initial state 1 is S = K ln Ω and the 1 B 1

final state 2 is S = K ln Ω ,
2 B 2

Ω2
ΔS = S2 − S1 = kB ln (4.2.4)
Ω1

using the rule for subtracting logarithms. However, we wish to define Ω in terms of a measurable quantity. Considering the system
of expanding a volume of gas molecules from above, we know that the probability is proportional to the volume raised to the
number of atoms (or molecules), αV . Therefore,
N

4.2.1 https://chem.libretexts.org/@go/page/41420
N
V2 V2
ΔS = S2 − S1 = kB ln ( ) = N kB ln (4.2.5)
V1 V1

We can define this in terms of moles of gas and not molecules by setting the k or Boltzmann constant equal to
B , where R is the
R

NA

gas constant and N is Avogadro's number. So for a expansion of an ideal gas and holding the temperature constant,
A

N
N V2 V2
ΔS = R ln ( ) = nR ln (4.2.6)
NA V1 V1

because N

NA
= n , the number of moles. This is only defined for constant temperature because entropy can change with

temperature. Furthermore, since S is a state function, we do not need to specify whether this process is reversible or irreversible.

Thermodynamic Definition of Entropy


Using the statistical definition of entropy is very helpful to visualize how processes occur. However, calculating probabilities like
Ω can be very difficult. Fortunately, entropy can also be derived from thermodynamic quantities that are easier to measure.

Recalling the concept of work from the first law of thermodynamics, the heat (q) absorbed by an ideal gas in a reversible,
isothermal expansion is
V2
qrev = nRT ln . (4.2.7)
V1

If we divide by T, we can obtain the same equation we derived above for ΔS :


qrev V2
ΔS = = nR ln . (4.2.8)
T V1

We must restrict this to a reversible process because entropy is a state function, however the heat absorbed is path dependent. An
irreversible expansion would result in less heat being absorbed, but the entropy change would stay the same. Then, we are left with
qirrev
ΔS > (4.2.9)
T

for an irreversible process because

ΔS = ΔSrev = ΔSirrev . (4.2.10)

This apparent discrepancy in the entropy change between an irreversible and a reversible process becomes clear when considering
the changes in entropy of the surrounding and system, as described in the second law of thermodynamics.
It is evident from our experience that ice melts, iron rusts, and gases mix together. However, the entropic quantity we have defined
is very useful in defining whether a given reaction will occur. Remember that the rate of a reaction is independent of spontaneity. A
reaction can be spontaneous but the rate so slow that we effectively will not see that reaction happen, such as diamond converting
to graphite, which is a spontaneous process.

The Second Law as Energy Dispersion


Energy of all types -- in chemistry, most frequently the kinetic energy of molecules (but also including the phase change/potential
energy of molecules in fusion and vaporization, as well as radiation) changes from being localized to becoming more dispersed in
space if that energy is not constrained from doing so. The simplest example stereotypical is the expansion illustrated in Figure 1.
The initial motional/kinetic energy (and potential energy) of the molecules in the first bulb is unchanged in such an isothermal
process, but it becomes more widely distributed in the final larger volume. Further, this concept of energy dispersal equally
applies to heating a system: a spreading of molecular energy from the volume of greater-motional energy (“warmer”) molecules in
the surroundings to include the additional volume of a system that initially had “cooler” molecules. It is not obvious, but true, that
this distribution of energy in greater space is implicit in the Gibbs free energy equation and thus in chemical reactions.
“Entropy change is the measure of how more widely a specific quantity of molecular energy is dispersed in a process, whether
isothermal gas expansion, gas or liquid mixing, reversible heating and phase change, or chemical reactions.” There are two
requisites for entropy change.
1. It is enabled by the above-described increased distribution of molecular energy.

4.2.2 https://chem.libretexts.org/@go/page/41420
2. It is actualized if the process makes available a larger number of arrangements for the system’s energy, i.e., a final state that
involves the most probable distribution of that energy under the new constraints.
Thus, “information probability” is only one of the two requisites for entropy change. Some current approaches regarding
“information entropy” are either misleading or truly fallacious, if they do not include explicit statements about the essential
inclusion of molecular kinetic energy in their treatment of chemical reactions.

References
1. This literal greater spreading of molecular energy in 3-D space in an isothermal process is accompanied by occupancy of more
quantum states (“energy levels”) within each microstate and thus more microstates for the final macrostate (i.e., a larger Ω in
R\ln{W}). Similarly, in any thermal process higher energy quantum states can be significantly occupied – thereby increasing
the number of microstates in the product macrostate as measured by the Boltzmann relationship.
2. I. N. Levine, Physical Chemistry, 6th ed. 2009, p. 101, toward the end of “What Is Entropy?”
3. Chang, Raymond. Physical Chemistry for the Biosciences. Sausalito, California: University Science Books, 2005.

Contributors and Attributions


Frank L. Lambert, Professor Emeritus, Occidental College
Konstantin Malley (UCD)

Carnot Cycle
In the early 19th century, steam engines came to play an increasingly important role in industry and transportation. However, a
systematic set of theories of the conversion of thermal energy to motive power by steam engines had not yet been developed.
Nicolas Léonard Sadi Carnot (1796-1832), a French military engineer, published Reflections on the Motive Power of Fire in 1824.
The book proposed a generalized theory of heat engines, as well as an idealized model of a thermodynamic system for a heat
engine that is now known as the Carnot cycle. Carnot developed the foundation of the second law of thermodynamics, and is often
described as the "Father of thermodynamics."

The Carnot Cycle


The Carnot cycle consists of the following four processes:
I. A reversible isothermal gas expansion process. In this process, the ideal gas in the system absorbs q amount heat from a heat
in

source at a high temperature Thigh, expands and does work on surroundings.


II. A reversible adiabatic gas expansion process. In this process, the system is thermally insulated. The gas continues to expand
and do work on surroundings, which causes the system to cool to a lower temperature, T . low

III. A reversible isothermal gas compression process. In this process, surroundings do work to the gas at T , and causes a loss of
low

heat, q .
out

IV. A reversible adiabatic gas compression process. In this process, the system is thermally insulated. Surroundings continue to do
work to the gas, which causes the temperature to rise back to T .
high

Figure 4.2.1 : An ideal gas-piston model of the Carnot cycle. (CC BY 4.0; XiSen Hou via Hope College)

4.2.3 https://chem.libretexts.org/@go/page/41420
P-V Diagram
The P-V diagram of the Carnot cycle is shown in Figure 4.2.2. In isothermal processes I and III, ∆U=0 because ∆T=0. In adiabatic
processes II and IV, q=0. Work, heat, ∆U, and ∆H of each process in the Carnot cycle are summarized in Table 4.2.1.

Figure 4.2.2 : A P-V diagram of the Carnot Cycle.


Table 4.2.1 : Work, heat, ∆U, and ∆H in the P-V diagram of the Carnot Cycle.
Process w q ΔU ΔH

V2 V2
I −nR Thigh ln( ) nR Thigh ln( ) 0 0
V1 V1

II ¯
n Cv (Tlow − Thigh) 0 ¯
n Cv (Tlow − Thigh)
¯
n Cp (Tlow − Thigh)

V4 V4
III −nR Tlow ln( ) nR Tlow ln( ) 0 0
V3 V3

IV ¯
n Cv (Thigh − Tlow ) 0 ¯
n Cv (Thight − Tlow )
¯
n Cp (Thigh − Tlow )

V2 V4 V2 V4
Full Cycle −nR Thigh ln( ) − nR Tlow
nRln(
Thigh ln(
) ) + nR Tlow ln( ) 0 0
V1 V3 V1 V3

T-S Diagram
The T-S diagram of the Carnot cycle is shown in Figure 4.2.3. In isothermal processes I and III, ∆T=0. In adiabatic processes II and
IV, ∆S=0 because dq=0. ∆T and ∆S of each process in the Carnot cycle are shown in Table 4.2.2.

Figure 4.2.3 : A T-S diagram of the Carnot Cycle. (CC BY 4.0; XiSen Hou via Hope College)
Table 4.2.1 : Work, heat, and ∆U in the T-S diagram of the Carnot Cycle.
Process ΔT ΔS
V2
I 0 −nR ln( )
V1

4.2.4 https://chem.libretexts.org/@go/page/41420
Process ΔT ΔS

II Tlow − Thigh 0
V4
III 0 −nR ln( )
V3

IV Thigh − Tlow 0

Full Cycle 0 0

Efficiency
The Carnot cycle is the most efficient engine possible based on the assumption of the absence of incidental wasteful processes such
as friction, and the assumption of no conduction of heat between different parts of the engine at different temperatures. The
efficiency of the carnot engine is defined as the ratio of the energy output to the energy input.
net work done by heat engine −wsys
efficiency = =
heat absorbed by heat engine qhigh

V2 V4
nRThigh ln( ) + nRTlow ln( )
V1 V3
=
V2
nRThigh ln( )
V1

Since processes II (2-3) and IV (4-1) are adiabatic,


CV /R
T2 V3
( ) = (4.2.11)
T3 V2

and
CV /R
T1 V4
( ) = (4.2.12)
T4 V1

And since T1 = T2 and T3 = T4,


V3 V2
= (4.2.13)
V4 V1

Therefore,
V2 V2
nRThigh ln( ) − nRTlow ln( )
V1 V1
efficiency = (4.2.14)
V2
nRThigh ln( )
V1

Thigh − Tlow
efficiency = (4.2.15)
Thigh

Summary
The Carnot cycle has the greatest efficiency possible of an engine (although other cycles have the same efficiency) based on the
assumption of the absence of incidental wasteful processes such as friction, and the assumption of no conduction of heat between
different parts of the engine at different temperatures.

Problems
1. You are now operating a Carnot engine at 40% efficiency, which exhausts heat into a heat sink at 298 K. If you want to increase
the efficiency of the engine to 65%, to what temperature would you have to raise the heat reservoir?
2. A Carnot engine absorbed 1.0 kJ of heat at 300 K, and exhausted 400 J of heat at the end of the cycle. What is the temperature
at the end of the cycle?

4.2.5 https://chem.libretexts.org/@go/page/41420
3. An indoor heater operating on the Carnot cycle is warming the house up at a rate of 30 kJ/s to maintain the indoor temperature
at 72 ºF. What is the power operating the heater if the outdoor temperature is 30 ºF?

References
1. Goldstein, M. J. Chem. Educ., 1980, 57, 114-116
2. Bader, M. J. Chem. Educ., 1973, 50, 834
3. W. F. Luder. J. Chem. Educ., 1944, 21, 600-601
4. Salter, C. J. Chem. Educ., 2000, 77, 1027-1030

4.2: Entropy is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
Statistical Entropy is licensed CC BY-NC-SA 4.0.
Carnot Cycle by XiSen Hou is licensed CC BY 4.0.

4.2.6 https://chem.libretexts.org/@go/page/41420
4.3: The Second Law of Thermodynamics
The Second Law of Thermodynamics states that the state of entropy of the entire universe, as an isolated system, will always
increase over time. The second law also states that the changes in the entropy in the universe can never be negative.

Introduction
Why is it that when you leave an ice cube at room temperature, it begins to melt? Why do we get older and never younger? And,
why is it whenever rooms are cleaned, they become messy again in the future? Certain things happen in one direction and not the
other, this is called the "arrow of time" and it encompasses every area of science. The thermodynamic arrow of time (entropy) is the
measurement of disorder within a system. Denoted as ΔS , the change of entropy suggests that time itself is asymmetric with
respect to order of an isolated system, meaning: a system will become more disordered, as time increases.

Major players in developing the Second Law


Nicolas Léonard Sadi Carnot was a French physicist, who is considered to be the "father of thermodynamics," for he is
responsible for the origins of the Second Law of Thermodynamics, as well as various other concepts. The current form of
the second law uses entropy rather than caloric, which is what Sadi Carnot used to describe the law. Caloric relates to heat
and Sadi Carnot came to realize that some caloric is always lost in the motion cycle. Thus, the thermodynamic reversibility
concept was proven wrong, proving that irreversibility is the result of every system involving work.
Rudolf Clausius was a German physicist, and he developed the Clausius statement, which says "Heat generally cannot
flow spontaneously from a material at a lower temperature to a material at a higher temperature."
William Thompson, also known as Lord Kelvin, formulated the Kelvin statement, which states "It is impossible to convert
heat completely in a cyclic process." This means that there is no way for one to convert all the energy of a system into
work, without losing energy.
Constantin Carathéodory, a Greek mathematician, created his own statement of the second law arguing that "In the
neighborhood of any initial state, there are states which cannot be approached arbitrarily close through adiabatic changes
of state."

Figure 4.3.1 : Sade Carnot(Louis-LéopoldBoilly public Figure 4.3.1 : Rudolf Clausiius (Theo Schafgans, public
domain) domain)

4.3.1 https://chem.libretexts.org/@go/page/41421
Figure 4.3.1 : Lord Kelvin (Smithsonian Institution, public Figure 4.3.1 : Constantin Carathéodory (Public
domain) Domain

To understand why entropy increases and decreases, it is important to recognize that two changes in entropy have to considered at
all times. The entropy change of the surroundings and the entropy change of the system itself. Given the entropy change of the
universe is equivalent to the sums of the changes in entropy of the system and surroundings:
qsys qsurr
ΔSuniv = ΔSsys + ΔSsurr = + (4.3.1)
T T

In an isothermal reversible expansion, the heat q absorbed by the system from the surroundings is
V2
qrev = nRT ln (4.3.2)
V1

Since the heat absorbed by the system is the amount lost by the surroundings, qsys = −qsurr .Therefore, for a truly reversible
process, the entropy change is
V2 V2
nRT ln −nRT ln
V1 V1
ΔSuniv = + =0 (4.3.3)
T T

If the process is irreversible however, the entropy change is


V2
nRT ln
V1
ΔSuniv = >0 (4.3.4)
T

If we put the two equations for ΔS univ together for both types of processes, we are left with the second law of thermodynamics,

ΔSuniv = ΔSsys + ΔSsurr ≥ 0 (4.3.5)

where ΔS univ equals zero for a truly reversible process and is greater than zero for an irreversible process. In reality, however,
truly reversible processes never happen (or will take an infinitely long time to happen), so it is safe to say all thermodynamic
processes we encounter everyday are irreversible in the direction they occur.

The second law of thermodynamics can also be stated that "all spontaneous processes produce an increase in the entropy of
the universe".

4.3: The Second Law of Thermodynamics is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

4.3.2 https://chem.libretexts.org/@go/page/41421
4.4: The Third Law of Thermodynamics
 Learning Objectives
The absolute entropy of a pure substance at a given temperature is the sum of all the entropy it would acquire on warming
from absolute zero (where S = 0 ) to the particular temperature.
Calculate entropy changes for phase transitions and chemical reactions under standard conditions

The atoms, molecules, or ions that compose a chemical system can undergo several types of molecular motion, including
translation, rotation, and vibration (Figure 4.4.1). The greater the molecular motion of a system, the greater the number of possible
microstates and the higher the entropy. A perfectly ordered system with only a single microstate available to it would have an
entropy of zero. The only system that meets this criterion is a perfect crystal at a temperature of absolute zero (0 K), in which each
component atom, molecule, or ion is fixed in place within a crystal lattice and exhibits no motion (ignoring quantum zero point
motion).

Figure 4.4.1 : Molecular Motions. Vibrational, rotational, and translational motions of a carbon dioxide molecule are illustrated
here. Only a perfectly ordered, crystalline substance at absolute zero would exhibit no molecular motion (classically; there will
always be motion quantum mechanically) and have zero entropy. In practice, this is an unattainable ideal. (CC BY-SA-NC;
Anonymous by request)
This system may be described by a single microstate, as its purity, perfect crystallinity and complete lack of motion (at least
classically, quantum mechanics argues for constant motion) means there is but one possible location for each identical atom or
molecule comprising the crystal (Ω = 1 ). According to the Boltzmann equation, the entropy of this system is zero.

S = k ln Ω

= k ln(1)

=0

In practice, absolute zero is an ideal temperature that is unobtainable, and a perfect single crystal is also an ideal that cannot be
achieved. Nonetheless, the combination of these two ideals constitutes the basis for the third law of thermodynamics: the entropy of
any perfectly ordered, crystalline substance at absolute zero is zero.

 Definition: Third Law of Thermodynamics

The entropy of a pure, perfect crystalline substance at 0 K is zero.

The third law of thermodynamics has two important consequences: it defines the sign of the entropy of any substance at
temperatures above absolute zero as positive, and it provides a fixed reference point that allows us to measure the absolute entropy
of any substance at any temperature. In this section, we examine two different ways to calculate ΔS for a reaction or a physical
change. The first, based on the definition of absolute entropy provided by the third law of thermodynamics, uses tabulated values of
absolute entropies of substances. The second, based on the fact that entropy is a state function, uses a thermodynamic cycle similar
to those discussed previously.

4.4.1 https://chem.libretexts.org/@go/page/432051
Standard-State Entropies
One way of calculating ΔS for a reaction is to use tabulated values of the standard molar entropy (S ), which is the entropy of 1
o

mol of a substance under standard pressure (1 bar). Often the standard molar entropy is given at 298 K and is often demarked as
ΔS
o
298
. The units of S are J/(mol•K). Unlike enthalpy or internal energy, it is possible to obtain absolute entropy values by
o

measuring the entropy change that occurs between the reference point of 0 K (corresponding to S = 0 ) and 298 K (Tables T1 and
T2).
As shown in Table 4.4.1, for substances with approximately the same molar mass and number of atoms, S values fall in the order
o

o o o
S (gas) ≫ S (liquid) > S (solid). (4.4.1)

For instance, S for liquid water is 70.0 J/(mol•K), whereas S for water vapor is 188.8 J/(mol•K). Likewise, S is 260.7
o o o

J/(mol•K) for gaseous I and 116.1 J/(mol•K) for solid I . This order makes qualitative sense based on the kinds and extents of
2 2

motion available to atoms and molecules in the three phases (Figure 4.4.1). The correlation between physical state and absolute
entropy is illustrated in Figure 4.4.2, which is a generalized plot of the entropy of a substance versus temperature.

Figure 4.4.2 : A Generalized Plot of Entropy versus Temperature for a Single Substance. Absolute entropy increases steadily with
increasing temperature until the melting point is reached, where it jumps suddenly as the substance undergoes a phase change from
a highly ordered solid to a disordered liquid (ΔSfus). The entropy again increases steadily with increasing temperature until the
boiling point is reached, where it jumps suddenly as the liquid undergoes a phase change to a highly disordered gas (ΔSvap). (CC
BY-SA-NC; anonymous).

The Third Law Lets us Calculate Absolute Entropies


The absolute entropy of a substance at any temperature above 0 K must be determined by calculating the increments of heat q
required to bring the substance from 0 K to the temperature of interest, and then summing the ratios q/T . Two kinds of
experimental measurements are needed:
1. The enthalpies associated with any phase changes the substance may undergo within the temperature range of interest. Melting
of a solid and vaporization of a liquid correspond to sizeable increases in the number of microstates available to accept thermal
energy, so as these processes occur, energy will flow into a system, filling these new microstates to the extent required to
maintain a constant temperature (the freezing or boiling point); these inflows of thermal energy correspond to the heats of
fusion and vaporization. The entropy increase associated with transition at temperature T is
ΔHf usion
. (4.4.2)
T

2. The heat capacity C of a phase expresses the quantity of heat required to change the temperature by a small amount ΔT , or
more precisely, by an infinitesimal amount dT . Thus the entropy increase brought about by warming a substance over a range

4.4.2 https://chem.libretexts.org/@go/page/432051
of temperatures that does not encompass a phase transition is given by the sum of the quantities C dT

T
for each increment of
temperature dT . This is of course just the integral
T
Cp
S0→T = ∫ dt (4.4.3)
0
T

Because the heat capacity is itself slightly temperature dependent, the most precise determinations of absolute entropies require that
the functional dependence of C on T be used in the integral in Equation 4.4.3, i.e.,:
T
Cp (T )
S0→T = ∫ dt. (4.4.4)
0
T

When this is not known, one can take a series of heat capacity measurements over narrow temperature increments ΔT and measure
the area under each section of the curve. The area under each section of the plot represents the entropy change associated with
heating the substance through an interval ΔT . To this must be added the enthalpies of melting, vaporization, and of any solid-solid
phase changes.

Figure 4.4.3 : Heat capacity/temperature as a function of temperature. (CC BY; Stephan Lower)
Values of C for temperatures near zero are not measured directly, but can be estimated from quantum theory. The cumulative areas
p

from 0 K to any given temperature (Figure 4.4.3) are then plotted as a function of T , and any phase-change entropies such as
Hvap
Svap = (4.4.5)
Tb

are added to obtain the absolute entropy at temperature T . As shown in Figure 4.4.2 above, the entropy of a substance increases
with temperature, and it does so for two reasons:
As the temperature rises, more microstates become accessible, allowing thermal energy to be more widely dispersed. This is
reflected in the gradual increase of entropy with temperature.
The molecules of solids, liquids, and gases have increasingly greater freedom to move around, facilitating the spreading and
sharing of thermal energy. Phase changes are therefore accompanied by massive and discontinuous increase in the entropy.

Calculating ΔS sys

We can make careful calorimetric measurements to determine the temperature dependence of a substance’s entropy and to derive
absolute entropy values under specific conditions. Standard entropies are given the label S for values determined for one mole
o
298

of substance at a pressure of 1 bar and a temperature of 298 K. The standard entropy change (ΔS ) for any process may be o

computed from the standard entropies of its reactant and product species like the following:
o o o
ΔS = ∑ νS (products) − ∑ ν S (reactants) (4.4.6)
298 298

Here, ν represents stoichiometric coefficients in the balanced equation representing the process. For example, ΔS
o
for the
following reaction at room temperature

mA + nB ⟶ xC + yD (4.4.7)

is computed as the following:


o o o o o
ΔS = [x S (C) + y S (D)] − [m S (A) + nS (B)] (4.4.8)
298 298 298 298

4.4.3 https://chem.libretexts.org/@go/page/432051
Table 4.4.1 lists some standard entropies at 298.15 K. You can find additional standard entropies in Tables T1 and T2
Table 4.4.1 : Standard Molar Entropy Values of Selected Substances at 25°C
Gases Liquids Solids

Substance S
o
[J/(mol•K)] Substance S
o
[J/(mol•K)] Substance S
o
[J/(mol•K)]

He 126.2 H2O 70.0 C (diamond) 2.4

H2 130.7 CH3OH 126.8 C (graphite) 5.7

Ne 146.3 Br2 152.2 LiF 35.7

Ar 154.8 CH3CH2OH 160.7 SiO2 (quartz) 41.5

Kr 164.1 C6H6 173.4 Ca 41.6

Xe 169.7 CH3COCl 200.8 Na 51.3

H2O 188.8 C6H12 (cyclohexane) 204.4 MgF2 57.2

N2 191.6 C8H18 (isooctane) 329.3 K 64.7

O2 205.2 NaCl 72.1

CO2 213.8 KCl 82.6

I2 260.7 I2 116.1

A closer examination of Table 4.4.1 also reveals that substances with similar molecular structures tend to have similar S values. o

Among crystalline materials, those with the lowest entropies tend to be rigid crystals composed of small atoms linked by strong,
highly directional bonds, such as diamond (S = 2.4 J/(mol ∙ K) ). In contrast, graphite, the softer, less rigid allotrope of carbon,
o

has a higher S (5.7 J/(mol•K)) due to more disorder (microstates) in the crystal. Soft crystalline substances and those with larger
o

atoms tend to have higher entropies because of increased molecular motion and disorder. Similarly, the absolute entropy of a
substance tends to increase with increasing molecular complexity because the number of available microstates increases with
molecular complexity. For example, compare the S values for CH3OH(l) and CH3CH2OH(l). Finally, substances with strong
o

hydrogen bonds have lower values of S , which reflects a more ordered structure.
o

Entropy increases with softer, less rigid solids, solids that contain larger atoms, and solids
with complex molecular structures.
To calculate ΔS for a chemical reaction from standard molar entropies, we use the familiar “products minus reactants” rule, in
o

which the absolute entropy of each reactant and product is multiplied by its stoichiometric coefficient in the balanced chemical
equation. Example 4.4.1 illustrates this procedure for the combustion of the liquid hydrocarbon isooctane (C H ; 2,2,4- 8 18

trimethylpentane).

 Example 4.4.1

Use the data in Table 4.4.1 to calculate ΔS


o
for the reaction of liquid isooctane with O (g)
2
to give CO (g)
2
and H O(g)
2
at
298 K.
Given: standard molar entropies, reactants, and products
Asked for: ΔS°
Strategy:
Write the balanced chemical equation for the reaction and identify the appropriate quantities in Table 4.4.1. Subtract the sum
of the absolute entropies of the reactants from the sum of the absolute entropies of the products, each multiplied by their
appropriate stoichiometric coefficients, to obtain ΔS for the reaction. o

Solution:
The balanced chemical equation for the complete combustion of isooctane (C 8
H
18
) is as follows:
25
C H (l) + O (g) ⟶ 8 CO (g) + 9 H O(g)
8 18 2 2 2 2

4.4.4 https://chem.libretexts.org/@go/page/432051
We calculate ΔS for the reaction using the “products minus reactants” rule, where m and n are the stoichiometric coefficients
o

of each product and each reactant:


o o o
ΔSrxn = ∑ m S (products) − ∑ nS (reactants)

25
o o o o
= [8 S (C O2 ) + 9 S (H2 O)] − [ S (C8 H18 ) + S (O2 )]
2

= {[8 mol C O2 × 213.8 J/(mol ⋅ K)] + [9 mol H2 O × 188.8 J/(mol ⋅ K)]}

25
− {[1 mol C8 H18 × 329.3 J/(mol ⋅ K)] + [ mol O2 × 205.2 J/(mol ⋅ K)]}
2

= 515.3 J/K

ΔS is positive, as expected for a combustion reaction in which one large hydrocarbon molecule is converted to many
o

molecules of gaseous products.

 Exercise 4.4.1
Use the data in Table 4.4.1 to calculate ΔS
o
for the reaction of H (g)
2
with liquid benzene (C 6
H
6
) to give cyclohexane (
C H
6
) at 298 K.
12

Answer
-361.1 J/K

 Example 4.4.2: Determination of ΔS°

Calculate the standard entropy change for the following process at 298 K:

H O(g) ⟶ H O(l)
2 2

Solution
The value of the standard entropy change at room temperature, ΔS o
298
, is the difference between the standard entropy of the
product, H2O(l), and the standard entropy of the reactant, H2O(g).
o o o
ΔS =S (H O(l)) − S (H O(g))
298 298 2 298 2

−1 −1 −1 −1
= (70.0 J mol K ) − (188.8 Jmol K )

−1 −1
= −118.8 J mol K

The value for ΔS o


298
is negative, as expected for this phase transition (condensation), which the previous section discussed.

 Exercise 4.4.2

Calculate the standard entropy change for the following process at 298 K:

H (g) + C H (g) ⟶ C H (g)


2 2 4 2 6

Answer
−120.6 J mol−1 K−1

 Example 4.4.3: Determination of ΔS°


Calculate the standard entropy change for the combustion of methanol, CH3OH at 298 K:

2 CH OH(l) + 3 O (g) ⟶ 2 CO (g) + 4 H O(l)


3 2 2 2

Solution

4.4.5 https://chem.libretexts.org/@go/page/432051
The value of the standard entropy change is equal to the difference between the standard entropies of the products and the
entropies of the reactants scaled by their stoichiometric coefficients. The standard entropy of formations are found in Table
4.4.1.

o o
ΔS = ΔS
298

o o
= ∑ νS (products) − ∑ ν S (reactants)
298 298

o o o o
= 2S (CO (g)) + 4 S (H O(l))] − [2 S (CH OH(l)) + 3 S (O (g))]
298 2 298 2 298 3 298 2

= [(2 × 213.8) + (4 × 70.0)] − [(2 × 126.8) + (3 × 205.03)]

= −161.6 J/mol ⋅ K

 Exercise 4.4.3

Calculate the standard entropy change for the following reaction at 298 K:

Ca (OH) (s) ⟶ CaO(s) + H O(l)


2 2

Answer
24.7 J/mol•K

Summary
Energy values, as you know, are all relative, and must be defined on a scale that is completely arbitrary; there is no such thing as
the absolute energy of a substance, so we can arbitrarily define the enthalpy or internal energy of an element in its most stable form
at 298 K and 1 atm pressure as zero. The same is not true of the entropy; since entropy is a measure of the “dilution” of thermal
energy, it follows that the less thermal energy available to spread through a system (that is, the lower the temperature), the smaller
will be its entropy. In other words, as the absolute temperature of a substance approaches zero, so does its entropy. This principle is
the basis of the Third law of thermodynamics, which states that the entropy of a perfectly-ordered solid at 0 K is zero.
In practice, chemists determine the absolute entropy of a substance by measuring the molar heat capacity (C ) as a function of p

temperature and then plotting the quantity C /T versus T . The area under the curve between 0 K and any temperature T is the
p

absolute entropy of the substance at T . In contrast, other thermodynamic properties, such as internal energy and enthalpy, can be
evaluated in only relative terms, not absolute terms.
The second law of thermodynamics states that a spontaneous process increases the entropy of the universe, Suniv > 0. If ΔSuniv < 0,
the process is nonspontaneous, and if ΔSuniv = 0, the system is at equilibrium. The third law of thermodynamics establishes the zero
for entropy as that of a perfect, pure crystalline solid at 0 K. With only one possible microstate, the entropy is zero. We may
compute the standard entropy change for a process by using standard entropy values for the reactants and products involved in the
process.

Contributors and Attributions


Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
2bd...a7ac8df6@9.110).
Stephen Lower, Professor Emeritus (Simon Fraser U.) Chem1 Virtual Textbook

4.4: The Third Law of Thermodynamics is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.

4.4.6 https://chem.libretexts.org/@go/page/432051
4.5: Evaluating Entropy and Entropy Changes
Learning Objectives
To use thermodynamic cycles to calculate changes in entropy.

The atoms, molecules, or ions that compose a chemical system can undergo several types of molecular motion, including
translation, rotation, and vibration (Figure 4.5.1). The greater the molecular motion of a system, the greater the number of possible
microstates and the higher the entropy. A perfectly ordered system with only a single microstate available to it would have an
entropy of zero. The only system that meets this criterion is a perfect crystal at a temperature of absolute zero (0 K), in which each
component atom, molecule, or ion is fixed in place within a crystal lattice and exhibits no motion (ignoring quantum effects). Such
a state of perfect order (or, conversely, zero disorder) corresponds to zero entropy. In practice, absolute zero is an ideal temperature
that is unobtainable, and a perfect single crystal is also an ideal that cannot be achieved. Nonetheless, the combination of these two
ideals constitutes the basis for the third law of thermodynamics: the entropy of any perfectly ordered, crystalline substance at
absolute zero is zero.

Figure 4.5.1 : Molecular Motions. Vibrational, rotational, and translational motions of a carbon dioxide molecule are illustrated
here. Only a perfectly ordered, crystalline substance at absolute zero would exhibit no molecular motion and have zero entropy. In
practice, this is an unattainable ideal.

Third Law of Thermodynamics


The entropy of any perfectly ordered, crystalline substance at absolute zero is zero.

The third law of thermodynamics has two important consequences: it defines the sign of the entropy of any substance at
temperatures above absolute zero as positive, and it provides a fixed reference point that allows us to measure the absolute entropy
of any substance at any temperature.In practice, chemists determine the absolute entropy of a substance by measuring the molar
heat capacity (Cp) as a function of temperature and then plotting the quantity Cp/T versus T. The area under the curve between 0 K
and any temperature T is the absolute entropy of the substance at T. In contrast, other thermodynamic properties, such as internal
energy and enthalpy, can be evaluated in only relative terms, not absolute terms. In this section, we examine two different ways to
calculate ΔS for a reaction or a physical change. The first, based on the definition of absolute entropy provided by the third law of
thermodynamics, uses tabulated values of absolute entropies of substances. The second, based on the fact that entropy is a state
function, uses a thermodynamic cycle similar to those discussed previously.

Calculating ΔS from Standard Molar Entropy Values


One way of calculating ΔS for a reaction is to use tabulated values of the standard molar entropy (S°), which is the entropy of 1
mol of a substance at a standard temperature of 298 K; the units of S° are J/(mol•K). Unlike enthalpy or internal energy, it is
possible to obtain absolute entropy values by measuring the entropy change that occurs between the reference point of 0 K
[corresponding to S = 0 J/(mol•K)] and 298 K.

4.5.1 https://chem.libretexts.org/@go/page/432052
Figure 4.5.2 : A Generalized Plot of Entropy versus Temperature for a Single Substance. Absolute entropy increases steadily with
increasing temperature until the melting point is reached, where it jumps suddenly as the substance undergoes a phase change from
a highly ordered solid to a disordered liquid (ΔSfus). The entropy again increases steadily with increasing temperature until the
boiling point is reached, where it jumps suddenly as the liquid undergoes a phase change to a highly disordered gas (ΔSvap).
As shown in Table 4.5.1, for substances with approximately the same molar mass and number of atoms, S° values fall in the order
S°(gas) > S°(liquid) > S°(solid). For instance, S° for liquid water is 70.0 J/(mol•K), whereas S° for water vapor is 188.8 J/(mol•K).
Likewise, S° is 260.7 J/(mol•K) for gaseous I2 and 116.1 J/(mol•K) for solid I2. This order makes qualitative sense based on the
kinds and extents of motion available to atoms and molecules in the three phases. The correlation between physical state and
absolute entropy is illustrated in Figure 4.5.2, which is a generalized plot of the entropy of a substance versus temperature.
Table 4.5.1 : Standard Molar Entropy Values of Selected Substances at 25°C
Gases Liquids Solids

Substance S° [J/(mol•K)] Substance S° [J/(mol•K)] Substance S° [J/(mol•K)]

He 126.2 H2O 70.0 C (diamond) 2.4

H2 130.7 CH3OH 126.8 C (graphite) 5.7

Ne 146.3 Br2 152.2 LiF 35.7

Ar 154.8 CH3CH2OH 160.7 SiO2 (quartz) 41.5

Kr 164.1 C6H6 173.4 Ca 41.6

Xe 169.7 CH3COCl 200.8 Na 51.3

H2O 188.8 C6H12 (cyclohexane) 204.4 MgF2 57.2

N2 191.6 C8H18 (isooctane) 329.3 K 64.7

O2 205.2 NaCl 72.1

CO2 213.8 KCl 82.6

I2 260.7 I2 116.1

Note

Entropy increases with softer, less rigid solids, solids that contain larger atoms, and solids with complex molecular structures.

A closer examination of Table 4.5.1 also reveals that substances with similar molecular structures tend to have similar S° values.
Among crystalline materials, those with the lowest entropies tend to be rigid crystals composed of small atoms linked by strong,
highly directional bonds, such as diamond [S° = 2.4 J/(mol•K)]. In contrast, graphite, the softer, less rigid allotrope of carbon, has a
higher S° [5.7 J/(mol•K)] due to more disorder in the crystal. Soft crystalline substances and those with larger atoms tend to have
higher entropies because of increased molecular motion and disorder. Similarly, the absolute entropy of a substance tends to

4.5.2 https://chem.libretexts.org/@go/page/432052
increase with increasing molecular complexity because the number of available microstates increases with molecular complexity.
For example, compare the S° values for CH3OH(l) and CH3CH2OH(l). Finally, substances with strong hydrogen bonds have lower
values of S°, which reflects a more ordered structure.

Note

ΔS° for a reaction can be calculated from absolute entropy values using the same “products minus reactants” rule used to
calculate ΔH°.

To calculate ΔS° for a chemical reaction from standard molar entropies, we use the familiar “products minus reactants” rule, in
which the absolute entropy of each reactant and product is multiplied by its stoichiometric coefficient in the balanced chemical
equation. Example 4.5.1 illustrates this procedure for the combustion of the liquid hydrocarbon isooctane (C8H18; 2,2,4-
trimethylpentane).

Example 4.5.1

Use the data in Table 4.5.1 to calculate ΔS° for the reaction of liquid isooctane with O2(g) to give CO2(g) and H2O(g) at 298
K.
Given: standard molar entropies, reactants, and products
Asked for: ΔS°
Strategy:
Write the balanced chemical equation for the reaction and identify the appropriate quantities in Table 4.5.1. Subtract the sum
of the absolute entropies of the reactants from the sum of the absolute entropies of the products, each multiplied by their
appropriate stoichiometric coefficients, to obtain ΔS° for the reaction.
Solution:
The balanced chemical equation for the complete combustion of isooctane (C8H18) is as follows:
25
C8 H18 (l) + O2 (g) → 8C O2 (g) + 9 H2 O(g)
2

We calculate ΔS° for the reaction using the “products minus reactants” rule, where m and n are the stoichiometric coefficients
of each product and each reactant:
∘ ∘ ∘
ΔSrxn = ∑ m S (products) − ∑ nS (reactants) (4.5.1)

∘ ∘ ∘
25 ∘
= [8 S (C O2 ) + 9 S (H2 O)] − [ S (C8 H18 ) + S (O2 )] (4.5.2)
2

= {[8 mol C O2 × 213.8 J/(mol ⋅ K)] + [9 mol H2 O × 188.8 J/(mol ⋅ K)]} (4.5.3)

25
− {[1 mol C8 H18 × 329.3 J/(mol ⋅ K)] + [ mol O2 × 205.2 J/(mol ⋅ K)]} (4.5.4)
2

= 515.3 J/K (4.5.5)

ΔS° is positive, as expected for a combustion reaction in which one large hydrocarbon molecule is converted to many
molecules of gaseous products.

Exercise 4.5.1

Use the data in Table 4.5.1 to calculate ΔS° for the reaction of H2(g) with liquid benzene (C6H6) to give cyclohexane (C6H12).
Answer: −361.1 J/K

Calculating ΔS from Thermodynamic Cycles


We can also calculate a change in entropy using a thermodynamic cycle. As you learned previously, the molar heat capacity (Cp) is
the amount of heat needed to raise the temperature of 1 mol of a substance by 1°C at constant pressure. Similarly, Cv is the amount

4.5.3 https://chem.libretexts.org/@go/page/432052
of heat needed to raise the temperature of 1 mol of a substance by 1°C at constant volume. The increase in entropy with increasing
temperature in Figure 4.5.2 is approximately proportional to the heat capacity of the substance.
Recall that the entropy change (ΔS) is related to heat flow (qrev) by ΔS = qrev/T. Because qrev = nCpΔT at constant pressure or
nCvΔT at constant volume, where n is the number of moles of substance present, the change in entropy for a substance whose
temperature changes from T1 to T2 is as follows:
qrev ΔT
ΔS = = nCp (constant pressure) (4.5.6)
T T

As you will discover in more advanced math courses than is required here, it can be shown that this is equal to the following:For a
review of natural logarithms, see Essential Skills 6 in Chapter 11 "Liquids".
T2
ΔS = nCp ln (constant pressure) (4.5.7)
T1

Similarly,
T2
ΔS = nCv ln (constant volume) (4.5.8)
T1

Thus we can use a combination of heat capacity measurements (Equation 18.20 or Equation 18.21) and experimentally measured
values of enthalpies of fusion or vaporization if a phase change is involved (Equation 18.18) to calculate the entropy change
corresponding to a change in the temperature of a sample.
We can use a thermodynamic cycle to calculate the entropy change when the phase change for a substance such as sulfur cannot be
measured directly. As noted in the exercise in Example 6, elemental sulfur exists in two forms (part (a) in Figure 4.5.3): an
orthorhombic form with a highly ordered structure (Sα) and a less-ordered monoclinic form (Sβ). The orthorhombic (α) form is
more stable at room temperature but undergoes a phase transition to the monoclinic (β) form at temperatures greater than 95.3°C
(368.5 K). The transition from Sα to Sβ can be described by the thermodynamic cycle shown in part (b) in Figure 4.5.3, in which
liquid sulfur is an intermediate. The change in entropy that accompanies the conversion of liquid sulfur to Sβ (−ΔSfus(β) = ΔS3 in
the cycle) cannot be measured directly. Because entropy is a state function, however, ΔS3 can be calculated from the overall
entropy change (ΔSt) for the Sα–Sβ transition, which equals the sum of the ΔS values for the steps in the thermodynamic cycle,
using Equation 18.20 and tabulated thermodynamic parameters (the heat capacities of Sα and Sβ, ΔHfus(α), and the melting point of
Sα.)

Figure 4.5.3 : Two Forms of Elemental Sulfur and a Thermodynamic Cycle Showing the Transition from One to the Other(a)
Orthorhombic sulfur (Sα) has a highly ordered structure in which the S8 rings are stacked in a “crankshaft” arrangement.
Monoclinic sulfur (Sβ) is also composed of S8 rings but has a less-ordered structure. (b) At 368.5 K, Sα undergoes a phase transition
to Sβ. Although ΔS3 cannot be measured directly, it can be calculated using the values shown in this thermodynamic cycle.
If we know the melting point of Sα (Tm = 115.2°C = 388.4 K) and ΔSt for the overall phase transition [calculated to be 1.09
J/(mol•K) in the exercise in Example 6], we can calculate ΔS3 from the values given in part (b) in Figure 4.5.3 where Cp(α) = 22.70
J/mol•K and Cp(β) = 24.77 J/mol•K (subscripts on ΔS refer to steps in the cycle):

4.5.4 https://chem.libretexts.org/@go/page/432052
ΔSt = ΔS1 + ΔS2 + ΔS3 + ΔS4 (4.5.9)

T2 ΔHfus T4
1.09 J/(mol ⋅ K) = Cp(α) ln( )+ + ΔS3 + Cp(β) ln( ) (4.5.10)
T1 Tm T3

388.4 1.722 kJ/mol


= 22.70 J/(mol ⋅ K) ln( ) +( × 1000 J/kJ) (4.5.11)
368.5 388.4 K

368.5
+ ΔS3 + 24.77 J/(mol ⋅ K) ln( ) (4.5.12)
388.4

= [1.194 J/(mol ⋅ K)] + [4.434 J/(mol ⋅ K)] + ΔS3 + [−1.303 J/(mol ⋅ K)] (4.5.13)

Solving for ΔS3 gives a value of −3.24 J/(mol•K). As expected for the conversion of a less ordered state (a liquid) to a more
ordered one (a crystal), ΔS3 is negative.

How are Entropies Measured


The absolute entropy of a substance at any temperature above 0 K must be determined by calculating the increments of heat q
required to bring the substance from 0 K to the temperature of interest, and then summing the ratios q/T. Two kinds of
experimental measurements are needed:
1. The enthalpies associated with any phase changes the substance may undergo within the temperature range of interest.
Melting of a solid and vaporization of a liquid correspond to sizeable increases in the number of microstates available to
accept thermal energy, so as these processes occur, energy will flow into a system, filling these new microstates to the
extent required to maintain a constant temperature (the freezing or boiling point); these inflows of thermal energy
correspond to the heats of fusion and vaporization. The entropy increase associated with melting, for example, is just
ΔHfusion/Tm.
2. The heat capacity C of a phase expresses the quantity of heat required to change the temperature by a small amount ΔT , or
more precisely, by an infinitesimal amount dT . Thus the entropy increase brought about by warming a substance over a
range of temperatures that does not encompass a phase transition is given by the sum of the quantities C dT/T for each
increment of temperature dT . This is of course just the integral
o
T
Cp
S0→T o =∫ dt (4.5.14)
0 T

Because the heat capacity is itself slightly temperature dependent, the most precise determinations of absolute entropies require
that the functional dependence of C on T be used in the above integral in place of a constant C.
o
T
Cp (T )
S0→T o =∫ dt (4.5.15)
0 T

When this is not known, one can take a series of heat capacity measurements over narrow temperature increments ΔT and
measure the area under each section of the curve.

Figure 4.5.4 : Heat capitity/temperature as a function of temperature


The area under each section of the plot represents the entropy change associated with heating the substance through an interval
ΔT. To this must be added the enthalpies of melting, vaporization, and of any solid-solid phase changes. Values of Cp for
temperatures near zero are not measured directly, but can be estimated from quantum theory.

4.5.5 https://chem.libretexts.org/@go/page/432052
Figure 4.5.5 : Molar entropy as a function of temperature
The cumulative areas from 0 K to any given temperature (taken from the experimental plot on the left) are then plotted as a
function of T, and any phase-change entropies such as Svap = Hvap / Tb are added to obtain the absolute entropy at temperature
T.

Summary
Entropy changes can be calculated using the “products minus reactants” rule or from a combination of heat capacity
measurements and measured values of enthalpies of fusion or vaporization.
The third law of thermodynamics states that the entropy of any perfectly ordered, crystalline substance at absolute zero is zero. At
temperatures greater than absolute zero, entropy has a positive value, which allows us to measure the absolute entropy of a
substance. Measurements of the heat capacity of a substance and the enthalpies of fusion or vaporization can be used to calculate
the changes in entropy that accompany a physical change. The entropy of 1 mol of a substance at a standard temperature of 298 K
is its standard molar entropy (S°). We can use the “products minus reactants” rule to calculate the standard entropy change (ΔS°)
for a reaction using tabulated values of S° for the reactants and the products.

4.5: Evaluating Entropy and Entropy Changes is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.

4.5.6 https://chem.libretexts.org/@go/page/432052
4.6: Gibbs Energy
Gibbs energy is the energy of a chemical reaction that can be used to do non-mechanical work. Gibbs Energy is described as
G = H −TS (4.6.1)

Where H is enthalpy, T is temperature, and S is entropy. ΔG is used to predict spontaneity within a system by

ΔGsys = ΔHsys – Δ(T S )sys (4.6.2)

Gibbs energy was developed in the 1870’s by Josiah Willard Gibbs. He originally termed this energy as the “available energy” in a
system. His paper published in 1873, “Graphical Methods in the Thermodynamics of Fluids,” outlined how his equation could
predict the behavior of systems when they are combined.
At a constant temperature and pressure, the Gibbs Energy of a system can be described as
ΔGsys = ΔHsys – T ΔSsys (4.6.3)

This equation can be used to determine the spontaneity of the process.


If ΔGsys ≤ 0, the process is spontaneous.
If ΔGsys = 0, the system is at equilibrium.
If ΔGsys > 0, the process is not spontaneous.
Furthermore,
If ΔGsys < 0, the process is exergonic
If ΔGsys > 0, the process is endergonic.
Gibbs Energy is a useful tool to describe in what manner the reaction is conducted. If ΔH >> TΔS, the reaction is enthalpy driven.
However, if TΔS >> ΔH, the reaction is driven by entropy. The Clausius-Clapeyron Equation is an application derived from Gibb's
energy:
P2 Δvap H T2 − T1
ln( ) = ( ) (4.6.4)
P1 R T2 T1

Another important application of Gibb's energy is the Maxwell relations (also available in a link at the end of the wiki page.)

Temperature & Pressure Dependence


When the pressure & temperature of a reaction are not held constant,
G = H −TS (4.6.5)

For an infinitesimal process,


ΔG = ΔH – Δ(T S) (4.6.6)

ΔG = ΔH – T ΔS– SΔT (4.6.7)

For a reaction where temperature is held constant,


ΔG = ΔH – T ΔS (4.6.8)

From the First Law of Thermodynamics, we know


H = U +PV (4.6.9)

ΔH = ΔU + P ΔV + V ΔP (4.6.10)

Since
ΔU = T ΔS– P ΔV (4.6.11)

We find that

ΔG = V ΔP − SΔT (4.6.12)

4.6.1 https://chem.libretexts.org/@go/page/41424
showing the obvious dependence of ΔG on temperature and pressure. To observe the change in Gibbs energy due to temperature
change alone (pressure held constant) the equation becomes
ΔG = −SΔT (4.6.13)

Solving for S and plugging it into Eq. (1), the Gibbs-Helmholtz Equation is found:
∂(ΔG/T ) ΔH
( ) =− (4.6.14)
2
∂T T
p

The Gibbs-Helmholtz Equation is very important because it relates the change in Gibbs energy to its temperature dependence, and
the position of equilibrium to change in enthalpy.
To observe the change in Gibbs energy due to pressure change alone (temperature held constant) the equation becomes

ΔG = V ΔP (4.6.15)

If the gas is assumed to be ideal then


P2
ΔG = nRT ln( ) (4.6.16)
P1

for an initial and final pressure (P1 and P2) at a constant T.

Standard-State Free Energy of Formation


Gibbs Energy is defined as a state function (a property that depends only on conditions describing the system, not how the change
occurs as in a path function.) This is because each component of the equation (H,T, and S) are all state functions. Therefore, we can
know the change in Gibbs energy without knowing every detail of the process. In a process that takes place at constant temperature
and pressure (298 K, 1 atm) the standard molar free energy of formation can be determined by the change in free energy from the
reactants and products. Using predetermined values, Eqn. (17) can be used
o ∘ ∘
ΔG )f = ∑(C oef f icientproducts G ) − ∑(C oef f icientreactants Greactants ) (4.6.17)
products

Standard-State Free Energy of Reaction


Gibbs Energy can be found at standard-state conditions using
ΔG° = ΔH ° − T ΔS° (4.6.18)

ΔH° and ΔS° values can be found in the appendix of any general chemistry textbook, or using this link.

Free Energy of Reaction


Gibbs energy can be found at any conditions by relating it to the standard-state free energy of reaction, using
ΔG = ΔG° + RT ln Q (4.6.19)

Where Q is the reaction quotient. Very rarely does chemistry actually occur at the given "standard-state" conditions. Using the
above equation and standard-state values, chemists can determine the overall Gibb's energy for the system, regardless of the
conditions.

References
1. Olmsted, J. Chemistry, 4th ed.; John Wiley & Sons, Inc: Hoboken 2006.
2. Chang, R. Physical Chemistry for the Biosciences; University Science Books: Sausalito, 2005.
3. Christian, S.D. Gibbs-Duhem equation and free energy calculations. J. Chem. Educ. 1962, 39 (10) 521-524.
4. Porter, S.K. The volume-entropy-energy surface of J.W. Gibbs. J. Chem Educ. 1971, 48(4) 231-234.

Problems
1. Calculate the standard-state free energy of formation for the H2O2(l) from H2 and O2, given the following values:
ΔfG (H2): 0 kJ/mol
ΔfG (O2): 0 kJ/mol

4.6.2 https://chem.libretexts.org/@go/page/41424
ΔfG (H2O2(l)): -120.4 kJ/mol

Hint: (product*coefficient) - [(Reactant1*coefficient)+(Reactant2*coefficient)]


2. Consider the following reaction:
C aC O3 C aO + C O2 (4.6.20)

At what temperature will this reaction become favorable? Note: Assume ΔHr°and ΔSr° are temperature independent.
Given values:

Substance ΔHf° Sf°

CaCO3 -1206.9 kJ/mol 92.9 J/K mol

CaO -635.6 kJ/mol 39.8 J/K mol

CO2 -393.5 kJ/mol 213.6 J/K mol

Hint: At what value of ΔG does a reaction become spontaneous, and therefore favorable? Next, find the ΔH value for the
reaction (products-reactants) and the ΔS values for the reaction (products-reactants.) Solve for T.
3. Is the above decomposition of calcium carbonate enthalpy driven or entropy driven?

Hint: Is ΔH >> TΔS, or TΔS >> ΔH?

Anwers to Problems
1. -120.4 kJ/mol
2. 1107.8 K
3. Enthalpy

Contributor
Christine Gobrogge (Hope College)

4.6: Gibbs Energy is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

4.6.3 https://chem.libretexts.org/@go/page/41424
4.7: Standard Molar Gibbs Energy of Formation
4.7: Standard Molar Gibbs Energy of Formation is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.

4.7.1 https://chem.libretexts.org/@go/page/41425
4.8: Dependence of Gibbs Energy on Temperature and Pressure
 Learning Objectives
To get an overview of Gibbs energy and its general uses in chemistry.
Understand how Gibbs energy pertains to reactions properties
Understand how Gibbs energy pertains to equilibria properties
Understand how Gibbs energy pertains to electrochemical properties

Gibbs free energy, denoted G, combines enthalpy and entropy into a single value. The change in free energy, ΔG, is equal to the
sum of the enthalpy plus the product of the temperature and entropy of the system. ΔG can predict the direction of the chemical
reaction under two conditions:
1. constant temperature and
2. constant pressure.
If ΔG is positive, then the reaction is nonspontaneous (i.e., an the input of external energy is necessary for the reaction to occur)
and if it is negative, then it is spontaneous (occurs without external energy input).

Introduction
Gibbs energy was developed in the 1870’s by Josiah Willard Gibbs. He originally termed this energy as the “available energy” in a
system. His paper published in 1873, “Graphical Methods in the Thermodynamics of Fluids,” outlined how his equation could
predict the behavior of systems when they are combined. This quantity is the energy associated with a chemical reaction that can be
used to do work, and is the sum of its enthalpy (H) and the product of the temperature and the entropy (S) of the system. This
quantity is defined as follows:

G = H −TS (4.8.1)

or more completely as
G = U +PV −TS (4.8.2)

where
U is internal energy (SI unit: joule)
P is pressure (SI unit: pascal)

V is volume (SI unit: m )


3

T is temperature (SI unit: kelvin)

S is entropy (SI unit: joule/kelvin)

H is the enthalpy (SI unit: joule)

Gibbs Energy in Reactions


Spontaneous - is a reaction that is consider to be natural because it is a reaction that occurs by itself without any external action
towards it. Non spontaneous - needs constant external energy applied to it in order for the process to continue and once you stop the
external action the process will cease. When solving for the equation, if change of G is negative, then it's spontaneous. If change of
G if positive, then it's non spontaneous. The symbol that is commonly used for FREE ENERGY is G. can be more properly
consider as "standard free energy change"
In chemical reactions involving the changes in thermodynamic quantities, a variation on this equation is often encountered:

ΔG = ΔH − T ΔS (4.8.3)
change in free energy change in enthalpy (temperature) change in entropy

 Example 1.1

Calculate ∆G at 290 K for the following reaction:

2 NO(g) + O (g) → 2 NO (g)


2 2

4.8.1 https://chem.libretexts.org/@go/page/41426
Given
∆H = -120 kJ
∆S = -150 JK -1

Solution
now all you have to do is plug in all the given numbers into Equation 3 above. Remember to divide ΔS by 1000 J/kJ so that
after you multiply by temperature, T , it will have the same units, kJ , as ΔH .

1 kJ
ΔS = −150 J /K ( ) = −0.15 kJ/K
1000 J

and substituting into Equation 3:

ΔG = −120 kJ − (290 K )(−0.150 kJ/ K )

= −120 kJ + 43 kJ

= −77 kJ

 Exercise 1.1: The Haber Process

What is the ΔG for this formation of ammonia from nitrogen and hydrogen gas.

N +3 H ⇌ 2 NH
2 2 3

The Standard free energy formations: NH3 =-16.45 H2=0 N2=0

Answer
−1
ΔG = −32.90 kJ mol

Since the changes of entropy of chemical reaction are not measured readily, thus, entropy is not typically used as a criterion. To
obviate this difficulty, we can use G. The sign of ΔG indicates the direction of a chemical reaction and determine if a reaction is
spontaneous or not.
ΔG < 0 : reaction is spontaneous in the direction written (i.e., the reaciton is exergonic)
ΔG = 0 : the system is at equilibrium and there is no net change either in forward or reverse direction.
ΔG > 0 : reaction is not spontaneous and the process proceeds spontaneously in the reserve direction. To drive such a reaction,

we need to have input of free energy (i.e., the reaction is endergonic)


The factors affect ΔG of a reaction (assume ΔH and ΔS are independent of temperature):

ΔH ΔS ΔG Example

at low temperature: + , at high


+ + 2HgO(s) -> 2Hg (l) + O2 (g)
temperature: -

+ - at all temperature: + 3O2 (g) ->2O3 (g)

- + at all temperature: - 2H2O2 (l) -> 2H2O (l) + O2 (g)

at low temperature: - , at high


- - NH3 (g) + HCl (g) -> NH4Cl (s)
temperature: +

Note:
1. ΔG depends only on the difference in free energy of products and reactants (or final state and initial state). ΔG is independent
of the path of the transformation and is unaffected by the mechanism of a reaction.
2. ΔG cannot tell us anything about the rate of a reaction.

4.8.2 https://chem.libretexts.org/@go/page/41426
The standard Gibbs energy change ΔG (at which reactants are converted to products at 1 bar) for:
o

aA + bB → cC + dD (4.8.4)

o o o o o
ΔrG = c Δf G (C ) + dΔf G (D) − aΔf G (A) − b Δf G (B) (4.8.5)

0 0 0
Δf G = ∑ vΔf G (products) − ∑ vΔf G (reactants) (4.8.6)

The standard-state free energy of reaction ( ΔG ) is defined as the free energy of reaction at standard state conditions:
o

o o o
ΔG = ΔH − T ΔS (4.8.7)

 Note
If |ΔH | >> |T ΔS|: the reaction is enthalpy-driven
If ΔH << T ΔS : the reaction is entropy-driven

Standard-State Free Energy of Formation


The partial pressure of any gas involved in the reaction is 0.1 MPa.
The concentrations of all aqueous solutions are 1 M.
Measurements are generally taken at a temperature of 25° C (298 K).
The standard-state free energy of formation is the change in free energy that occurs when a compound is formed from its elements
in their most thermodynamically stable states at standard-state conditions. In other words, it is the difference between the free
energy of a substance and the free energies of its constituent elements at standard-state conditions:
o o o
ΔG = ∑ ΔG − ∑ ΔG (4.8.8)
fproducts frea cta nts

 Example 1.2

Used the below information to determine if N H 4N O3(s) will dissolve in water at room temperature.

Compound ΔH
o
f
ΔS
f
o

N H4 N O3(s) -365.56 151.08

NH
+

4(aq )
-132.51 113.4

NO

3(aq )
205.0 146.4

Solution
This question is essentially asking if the following reaction is spontaneous at room temperature.
H O
2
+ −
NH NO (s) ⟶ NH (aq) + NO (aq)
4 3 4 3

This would normally only require calculating ΔG and evaluating its sign. However, the ΔG values are not tabulated, so they
o o

must be calculated manually from calculated ΔH and ΔS values for the reaction.
o o

Calculate ΔH : o

o o o
ΔH = ∑ nΔH − ∑ mΔH
f f
products rea cta nts

o
kJ −
kJ
ΔH = [(1 mol N H3 ) (−132.51 ) + (1 mol N O ) (−205.0 )]
3
mol mol

kJ
− [(1 mol N H4 N O3 ) (−365.56 )]
mol

o
ΔH = −337.51 kJ + 365.56 kJ = 28.05 kJ

4.8.3 https://chem.libretexts.org/@go/page/41426
Calculate ΔS : o

o o o
ΔS = ∑ nΔS − ∑ SΔH
fproducts frea cta nts

J J
o −
ΔS = [(1 mol N H3 ) (113.4 ) + (1 mol N O ) (146.6 )]
3
mol K mol K

J
− [(1 mol N H4 N O3 ) (151.08 )]
mol K

o
ΔS = 259.8 J/K − 151.08 J/K = 108.7 J/K

Calculate ΔG : o

These values can be substituted into the free energy equation


o
TK = 25 C + 273.15K = 298.15 K

o
1 kJ
ΔS = 108.7 J /K ( ) = 0.1087 kJ/K
1000 J

o
ΔH = 28.05 kJ

Plug in ΔH , ΔS and T into Equation 1.7


o o

o o o
ΔG = ΔH − T ΔS

o
ΔG = 28.05 kJ − (298.15 K )(0.1087 kJ/ K )

o
ΔG = 28.05 kJ − 32.41 kJ

o
ΔG = −4.4 kJ

This reaction is spontaneous at room temperature since ΔG


o
is negative. Therefore N H4 N O3(s) will dissolve in water at
room temperature.

 Example 1.3

Calculate ΔG for the following reaction at 25 o


C . Will the reaction occur spontaneously?

N H3(g) + H C l(g) → N H4 C l(s)

given for the reaction


ΔH = −176.0 kJ

ΔS = −284.8 J/K

Solution
calculate ΔG from the formula

ΔG = ΔH − T ΔS

but first we need to convert the units for ΔS into kJ/K (or convert ΔH into J) and temperature into Kelvin

1 kJ
ΔS = −284.8 J /K ( ) = −0.284.8 kJ/K
1000 J
o
T = 273.15 K + 25 C = 298 K

The definition of Gibbs energy can then be used directly

ΔG = ΔH − T ΔS

4.8.4 https://chem.libretexts.org/@go/page/41426
ΔG = −176.0 kJ − (298 K )(−0.284.8 kJ/ K )

ΔG = −176.0 kJ − (−84.9 kJ)

ΔG = −91.1 kJ

Yes, this reaction is spontaneous at room temperature since ΔG is negative.

Gibbs Energy in Equilibria


Let's consider the following reversible reaction:
A+B ⇋ C +D (4.8.9)

The following equation relates the standard-state free energy of reaction with the free energy at any point in a given reaction (not
necessarily at standard-state conditions):
o
ΔG = ΔG + RT ln Q (4.8.10)

ΔG = free energy at any moment


ΔG = standard-state free energy
o

R is the ideal gas constant = 8.314 J/mol-K


T is the absolute temperature (Kelvin)
ln Q is natural logarithm of the reaction quotient

At equilibrium, ΔG = 0 and Q=K. Thus the equation can be arranged into:


[C ][D]
o
ΔG = ΔG + RT ln (4.8.11)
[A][B]

with
ΔG
o
= standard free energy change
R = gas constant = 1.98 * 10-3 kcal mol-1 deg-10
T = is usually room temperature = 298 K
[C ][D]
K =
[A][B]

The Gibbs free energy ΔG depends primarily on the reactants' nature and concentrations (expressed in the ΔG
o
term and the
logarithmic term of Equation 1.11, respectively).
At equilibrium, ΔG = 0 : no driving force remains
′ [C ][D]
o
0 = ΔG + RT ln (4.8.12)
[A][B]

[C ][D]
o
ΔG = −RT ln (4.8.13)
[A][B]

The equilibrium constant is defined as


[C ][D]
Keq = (4.8.14)
[A][B]

When K eq is large, almost all reactants are converted to products. Substituting K eq into Equation 1.14, we have:
o
ΔG = −RT ln Keq (4.8.15)

or
o
ΔG = −2.303RT log10 Keq (4.8.16)

Rearrange,

4.8.5 https://chem.libretexts.org/@go/page/41426
o
−Δ G /(2.303RT )
Keq = 10 (4.8.17)

This equation is particularly interesting as it relates the free energy difference under standard conditions to the properties of a
system at equilibrium (which is rarely at standard conditions).
Table 1.1: Converting K ea to ΔG
Keq ΔGo (kcal/mole)

10
−5
6.82

10
−4
5.46

10
−3
4.09

10
−2
2.73

10
−1
1.36

1 0

10
1
-1.36

10
2
-2.73

10
3
-4.09

10
4
-5.46

10
5
-6.82

 Example 1.4

What is ΔG for isomerization of dihydroxyacetone phosphate to glyceraldehyde 3-phosphate?


o

If at equilibrium, we have K eq = 0.0475 at 298 K and pH 7. We can calculate:


o −3
ΔG = −2.303 RT log10 Keq = (−2.303) ∗ (1.98 ∗ 10 ) ∗ 298 ∗ (log10 0.0475) = 1.8 kcal/mol

Given:
The initial concentration of dihydroxyacetone phosphate = 2 × 10 −4
M

The initial concentration of glyceraldehyde 3-phosphate = 3 × 10 −6


M

Solution
From equation 2:
ΔG = 1.8 kcal/mol + 2.303 RT log10(3*10-6 M/2*10-4 M) = -0.7 kcal/mol

 Note

Under non-standard conditions (which is essential all reactions), the spontaneity of reaction is determined by ΔG, not ΔG . o

Gibbs Energy in Electrochemistry


The Nernst equation relates the standard-state cell potential with the cell potential of the cell at any moment in time:

o
RT
E =E − ln Q (4.8.18)
nF

with
E = cell potential in volts (joules per coulomb)
n = moles of electrons
F = Faraday's constant: 96,485 coulombs per mole of electrons

4.8.6 https://chem.libretexts.org/@go/page/41426
By rearranging this equation we obtain:

o
RT
E =E − ln Q (4.8.19)
nF

multiply the entire equation by nF


o
nF E = nF E − RT ln Q (4.8.20)

which is similar to:


o
ΔG = ΔG + RT ln Q (4.8.21)

By juxtaposing these two equations:


o
nF E = nF E − RT ln Q (4.8.22)

o
ΔG = ΔG + RT ln Q (4.8.23)

it can be concluded that:


ΔG = −nF E (4.8.24)

Therefore,
o o
ΔG = −nF E (4.8.25)

 Some remarks on the Gibbs "Free" Energy


Free Energy is not necessarily "free": The appellation “free energy” for G has led to so much confusion that many
scientists now refer to it simply as the Gibbs energy. The “free” part of the older name reflects the steam-engine origins of
thermodynamics with its interest in converting heat into work: ΔG is the maximum amount of energy which can be “freed”
from the system to perform useful work. By "useful", we mean work other than that which is associated with the expansion
of the system. This is most commonly in the form of electrical work (moving electric charge through a potential difference),
but other forms of work (osmotic work, increase in surface area) are also possible.
Free Energy is not energy: A much more serious difficulty with the Gibbs function, particularly in the context of
chemistry, is that although G has the units of energy (joules, or in its intensive form, J mol–1), it lacks one of the most
important attributes of energy in that it is not conserved. Thus although the free energy always falls when a gas expands or
a chemical reaction takes place spontaneously, there need be no compensating increase in energy anywhere else. Referring
to G as an energy also reinforces the false but widespread notion that a fall in energy must accompany any change. But if
we accept that energy is conserved, it is apparent that the only necessary condition for change (whether the dropping of a
weight, expansion of a gas, or a chemical reaction) is the redistribution of energy.The quantity –ΔG associated with a
process represents the quantity of energy that is “shared and spread”, which as we have already explained is the meaning of
the increase in the entropy. The quotient –ΔG/T is in fact identical with ΔStotal, the entropy change of the world, whose
increase is the primary criterion for any kind of change.
Free Energy is not even "real": G differs from the thermodynamic quantities H and S in another significant way: it has no
physical reality as a property of matter, whereas H and S can be related to the quantity and distribution of energy in a
collection of molecules (e.g., the third law of thermodynamics). The free energy is simply a useful construct that serves as a
criterion for change and makes calculations easier.

References
1. Chang, Raymond. Physical Chemistry for the Biosciences. Sansalito, CA: University Sciences, 2005.
2. Atkins, Peter and de Paula, Julio. Physical Chemistry for the Life Sciences. New York, NY: W. H. Freeman and Company, 2006.
Page 153-163, 286.
3. Stryer, Lubert. Biochemistry (Third Edition). New York, NY: W.H. Freeman and Company, 1988. Page 181-184.

4.8: Dependence of Gibbs Energy on Temperature and Pressure is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by LibreTexts.
Gibbs (Free) Energy by Cathy Doan, Han Le, Stephen Lower is licensed CC BY-NC-SA 4.0.

4.8.7 https://chem.libretexts.org/@go/page/41426
4.9: Phase Equilibria
 Learning Objectives
To know how and why the vapor pressure of a liquid varies with temperature.
To understand that the equilibrium vapor pressure of a liquid depends on the temperature and the intermolecular forces
present.
To understand that the relationship between pressure, enthalpy of vaporization, and temperature is given by the Clausius-
Clapeyron equation.

Nearly all of us have heated a pan of water with the lid in place and shortly thereafter heard the sounds of the lid rattling and hot
water spilling onto the stovetop. When a liquid is heated, its molecules obtain sufficient kinetic energy to overcome the forces
holding them in the liquid and they escape into the gaseous phase. By doing so, they generate a population of molecules in the
vapor phase above the liquid that produces a pressure—the vapor pressure of the liquid. In the situation we described, enough
pressure was generated to move the lid, which allowed the vapor to escape. If the vapor is contained in a sealed vessel, however,
such as an unvented flask, and the vapor pressure becomes too high, the flask will explode (as many students have unfortunately
discovered). In this section, we describe vapor pressure in more detail and explain how to quantitatively determine the vapor
pressure of a liquid.

Evaporation and Condensation


Because the molecules of a liquid are in constant motion, we can plot the fraction of molecules with a given kinetic energy (KE)
against their kinetic energy to obtain the kinetic energy distribution of the molecules in the liquid (Figure 4.9.1), just as we did for
a gas. As for gases, increasing the temperature increases both the average kinetic energy of the particles in a liquid and the range of
kinetic energy of the individual molecules. If we assume that a minimum amount of energy (E ) is needed to overcome the
0

intermolecular attractive forces that hold a liquid together, then some fraction of molecules in the liquid always has a kinetic energy
greater than E . The fraction of molecules with a kinetic energy greater than this minimum value increases with increasing
0

temperature. Any molecule with a kinetic energy greater than E has enough energy to overcome the forces holding it in the liquid
0

and escape into the vapor phase. Before it can do so, however, a molecule must also be at the surface of the liquid, where it is
physically possible for it to leave the liquid surface; that is, only molecules at the surface can undergo evaporation (or
vaporization), where molecules gain sufficient energy to enter a gaseous state above a liquid’s surface, thereby creating a vapor
pressure.

Figure 4.9.1 : The Distribution of the Kinetic Energies of the Molecules of a Liquid at Two Temperatures. Just as with gases,
increasing the temperature shifts the peak to a higher energy and broadens the curve. Only molecules with a kinetic energy greater
than E0 can escape from the liquid to enter the vapor phase, and the proportion of molecules with KE > E0 is greater at the higher
temperature. (CC BY-SA-NC; Anonymous by request)
Graph of fraction of molecules with a particular kinetic energy against kinetic energy. Green line is temperature at 400 kelvin,
purple line is temperature at 300 kelvin.

4.9.1 https://chem.libretexts.org/@go/page/41427
To understand the causes of vapor pressure, consider the apparatus shown in Figure 4.9.2. When a liquid is introduced into an
evacuated chamber (part (a) in Figure 4.9.2), the initial pressure above the liquid is approximately zero because there are as yet no
molecules in the vapor phase. Some molecules at the surface, however, will have sufficient kinetic energy to escape from the liquid
and form a vapor, thus increasing the pressure inside the container. As long as the temperature of the liquid is held constant, the
fraction of molecules with KE > E will not change, and the rate at which molecules escape from the liquid into the vapor phase
0

will depend only on the surface area of the liquid phase.

Figure 4.9.2 : Vapor Pressure. (a) When a liquid is introduced into an evacuated chamber, molecules with sufficient kinetic energy
escape from the surface and enter the vapor phase, causing the pressure in the chamber to increase. (b) When sufficient molecules
are in the vapor phase for a given temperature, the rate of condensation equals the rate of evaporation (a steady state is reached),
and the pressure in the container becomes constant. (CC BY-SA-NC; Anonymous by request)
As soon as some vapor has formed, a fraction of the molecules in the vapor phase will collide with the surface of the liquid and
reenter the liquid phase in a process known as condensation (part (b) in Figure 4.9.2). As the number of molecules in the vapor
phase increases, the number of collisions between vapor-phase molecules and the surface will also increase. Eventually, a steady
state will be reached in which exactly as many molecules per unit time leave the surface of the liquid (vaporize) as collide with it
(condense). At this point, the pressure over the liquid stops increasing and remains constant at a particular value that is
characteristic of the liquid at a given temperature. The rates of evaporation and condensation over time for a system such as this are
shown graphically in Figure 4.9.3.

Figure 4.9.3 : The Relative Rates of Evaporation and Condensation as a Function of Time after a Liquid Is Introduced into a Sealed
Chamber. The rate of evaporation depends only on the surface area of the liquid and is essentially constant. The rate of
condensation depends on the number of molecules in the vapor phase and increases steadily until it equals the rate of evaporation.
(CC BY-SA-NC; Anonymous by request)
Graph of rate against time. The green line is evaporation while the pruple line is condensation. Dynamic equilibrium is established
when the evaporation and condensation rates are equal.

Equilibrium Vapor Pressure


Two opposing processes (such as evaporation and condensation) that occur at the same rate and thus produce no net change in a
system, constitute a dynamic equilibrium. In the case of a liquid enclosed in a chamber, the molecules continuously evaporate and

4.9.2 https://chem.libretexts.org/@go/page/41427
condense, but the amounts of liquid and vapor do not change with time. The pressure exerted by a vapor in dynamic equilibrium
with a liquid is the equilibrium vapor pressure of the liquid.
If a liquid is in an open container, however, most of the molecules that escape into the vapor phase will not collide with the surface
of the liquid and return to the liquid phase. Instead, they will diffuse through the gas phase away from the container, and an
equilibrium will never be established. Under these conditions, the liquid will continue to evaporate until it has “disappeared.” The
speed with which this occurs depends on the vapor pressure of the liquid and the temperature. Volatile liquids have relatively high
vapor pressures and tend to evaporate readily; nonvolatile liquids have low vapor pressures and evaporate more slowly. Although
the dividing line between volatile and nonvolatile liquids is not clear-cut, as a general guideline, we can say that substances with
vapor pressures greater than that of water (Figure 4.9.4) are relatively volatile, whereas those with vapor pressures less than that of
water are relatively nonvolatile. Thus diethyl ether (ethyl ether), acetone, and gasoline are volatile, but mercury, ethylene glycol,
and motor oil are nonvolatile.

Figure 4.9.4 : The Vapor Pressures of Several Liquids as a Function of Temperature. The point at which the vapor pressure curve
crosses the P = 1 atm line (dashed) is the normal boiling point of the liquid. (CC BY-SA-NC; Anonymous by request)
The equilibrium vapor pressure of a substance at a particular temperature is a characteristic of the material, like its molecular mass,
melting point, and boiling point. It does not depend on the amount of liquid as long as at least a tiny amount of liquid is present in
equilibrium with the vapor. The equilibrium vapor pressure does, however, depend very strongly on the temperature and the
intermolecular forces present, as shown for several substances in Figure 4.9.4. Molecules that can hydrogen bond, such as ethylene
glycol, have a much lower equilibrium vapor pressure than those that cannot, such as octane. The nonlinear increase in vapor
pressure with increasing temperature is much steeper than the increase in pressure expected for an ideal gas over the corresponding
temperature range. The temperature dependence is so strong because the vapor pressure depends on the fraction of molecules that
have a kinetic energy greater than that needed to escape from the liquid, and this fraction increases exponentially with temperature.
As a result, sealed containers of volatile liquids are potential bombs if subjected to large increases in temperature. The gas tanks on
automobiles are vented, for example, so that a car won’t explode when parked in the sun. Similarly, the small cans (1–5 gallons)
used to transport gasoline are required by law to have a pop-off pressure release.

Volatile substances have low boiling points and relatively weak intermolecular
interactions; nonvolatile substances have high boiling points and relatively strong
intermolecular interactions.

4.9.3 https://chem.libretexts.org/@go/page/41427
Vapor Pressure & Boiling Point

A Video Discussing Vapor Pressure and Boiling Points. Video Source: Vapor Pressure & Boiling Point(opens in new window)
[youtu.be]
The exponential rise in vapor pressure with increasing temperature in Figure 4.9.4 allows us to use natural logarithms to express
the nonlinear relationship as a linear one.

−ΔHvap 1
ln P = ( ) +C (4.9.1)
R T

where
ln P is the natural logarithm of the vapor pressure,
ΔHvap is the enthalpy of vaporization,
R is the universal gas constant [8.314 J/(mol•K)],

T is the temperature in kelvins, and

C is the y-intercept, which is a constant for any given line.

Plotting ln P versus the inverse of the absolute temperature (1/T ) is a straight line with a slope of −ΔHvap/R. Equation 4.9.1,
called the Clausius–Clapeyron Equation, can be used to calculate the ΔH of a liquid from its measured vapor pressure at two or
vap

more temperatures. The simplest way to determine ΔH is to measure the vapor pressure of a liquid at two temperatures and
vap

insert the values of P and T for these points into Equation 4.9.2, which is derived from the Clausius–Clapeyron equation:

P1 −ΔHvap 1 1
ln( ) = ( − ) (4.9.2)
P2 R T1 T2

Conversely, if we know ΔHvap and the vapor pressure P at any temperature T , we can use Equation 4.9.2 to calculate the vapor
1 1

pressure P at any other temperature T , as shown in Example 4.9.1.


2 2

The Clausius-Clapeyron Equation

4.9.4 https://chem.libretexts.org/@go/page/41427
A Video Discussing the Clausius-Clapeyron Equation. Video Link: The Clausius-Clapeyron Equation(opens in new window)
[youtu.be]

 Example 4.9.1: Vapor Pressure of Mercury

The experimentally measured vapor pressures of liquid Hg at four temperatures are listed in the following table:
experimentally measured vapor pressures of liquid Hg at four temperatures
T (°C) 80.0 100 120 140

P (torr) 0.0888 0.2729 0.7457 1.845

From these data, calculate the enthalpy of vaporization (ΔHvap) of mercury and predict the vapor pressure of the liquid at
160°C. (Safety note: mercury is highly toxic; when it is spilled, its vapor pressure generates hazardous levels of mercury
vapor.)
Given: vapor pressures at four temperatures
Asked for: ΔHvap of mercury and vapor pressure at 160°C

Strategy:
A. Use Equation 4.9.2 to obtain ΔHvap directly from two pairs of values in the table, making sure to convert all values to the
appropriate units.
B. Substitute the calculated value of ΔHvap into Equation 4.9.2 to obtain the unknown pressure (P2).

Solution:
A The table gives the measured vapor pressures of liquid Hg for four temperatures. Although one way to proceed would be to
plot the data using Equation 4.9.1 and find the value of ΔHvap from the slope of the line, an alternative approach is to use
Equation 4.9.2 to obtain ΔHvap directly from two pairs of values listed in the table, assuming no errors in our measurement. We
therefore select two sets of values from the table and convert the temperatures from degrees Celsius to kelvin because the
equation requires absolute temperatures. Substituting the values measured at 80.0°C (T1) and 120.0°C (T2) into Equation 4.9.2
gives

0.7457 T orr −ΔHvap 1 1


ln( ) = ( − )
0.0888 T orr 8.314 J/mol ⋅ K (120 + 273) K (80.0 + 273) K

−ΔHvap
−4 −1
ln(8.398) = (−2.88 × 10 K )
8.314 J/mol ⋅ K

−4 −1
2.13 = −ΔHvap (−3.46 × 10 )J ⋅ mol

ΔHvap = 61, 400 J/mol = 61.4 kJ/mol

B We can now use this value of ΔHvap to calculate the vapor pressure of the liquid (P2) at 160.0°C (T2):

−61, 400 J/mol


P2 1 1
ln( ) = ( − )
0.0888 torr −1 (160 + 273) K (80.0 + 273) K
8.314 J/mol K

Using the relationship e ln x


=x , we have
P2
ln( ) = 3.86
0.0888 T orr

P2
3.86
=e = 47.5
0.0888 T orr

P2 = 4.21T orr

4.9.5 https://chem.libretexts.org/@go/page/41427
At 160°C, liquid Hg has a vapor pressure of 4.21 torr, substantially greater than the pressure at 80.0°C, as we would expect.

 Exercise 4.9.1: Vapor Pressure of Nickel

The vapor pressure of liquid nickel at 1606°C is 0.100 torr, whereas at 1805°C, its vapor pressure is 1.000 torr. At what
temperature does the liquid have a vapor pressure of 2.500 torr?

Answer
1896°C

Boiling Points
As the temperature of a liquid increases, the vapor pressure of the liquid increases until it equals the external pressure, or the
atmospheric pressure in the case of an open container. Bubbles of vapor begin to form throughout the liquid, and the liquid begins
to boil. The temperature at which a liquid boils at exactly 1 atm pressure is the normal boiling point of the liquid. For water, the
normal boiling point is exactly 100°C. The normal boiling points of the other liquids in Figure 4.9.4 are represented by the points
at which the vapor pressure curves cross the line corresponding to a pressure of 1 atm. Although we usually cite the normal boiling
point of a liquid, the actual boiling point depends on the pressure. At a pressure greater than 1 atm, water boils at a temperature
greater than 100°C because the increased pressure forces vapor molecules above the surface to condense. Hence the molecules
must have greater kinetic energy to escape from the surface. Conversely, at pressures less than 1 atm, water boils below 100°C.
Table 4.9.1 : The Boiling Points of Water at Various Locations on Earth
Place Altitude above Sea Level (ft) Atmospheric Pressure (mmHg) Boiling Point of Water (°C)

Mt. Everest, Nepal/Tibet 29,028 240 70

Bogota, Colombia 11,490 495 88

Denver, Colorado 5280 633 95

Washington, DC 25 759 100

Dead Sea, Israel/Jordan −1312 799 101.4

Typical variations in atmospheric pressure at sea level are relatively small, causing only minor changes in the boiling point of
water. For example, the highest recorded atmospheric pressure at sea level is 813 mmHg, recorded during a Siberian winter; the
lowest sea-level pressure ever measured was 658 mmHg in a Pacific typhoon. At these pressures, the boiling point of water
changes minimally, to 102°C and 96°C, respectively. At high altitudes, on the other hand, the dependence of the boiling point of
water on pressure becomes significant. Table 4.9.1 lists the boiling points of water at several locations with different altitudes. At
an elevation of only 5000 ft, for example, the boiling point of water is already lower than the lowest ever recorded at sea level. The
lower boiling point of water has major consequences for cooking everything from soft-boiled eggs (a “three-minute egg” may well
take four or more minutes in the Rockies and even longer in the Himalayas) to cakes (cake mixes are often sold with separate high-
altitude instructions). Conversely, pressure cookers, which have a seal that allows the pressure inside them to exceed 1 atm, are
used to cook food more rapidly by raising the boiling point of water and thus the temperature at which the food is being cooked.

As pressure increases, the boiling point of a liquid increases and vice versa.

 Example 4.9.2: Boiling Mercury

Use Figure 4.9.4 to estimate the following.


a. the boiling point of water in a pressure cooker operating at 1000 mmHg
b. the pressure required for mercury to boil at 250°C

4.9.6 https://chem.libretexts.org/@go/page/41427
Mercury boils at 356 °C at room pressure. To see video go to www.youtube.com/watch?v=0iizsbXWYoo
Given: Data in Figure 4.9.4, pressure, and boiling point
Asked for: corresponding boiling point and pressure

Strategy:
A. To estimate the boiling point of water at 1000 mmHg, refer to Figure 4.9.4 and find the point where the vapor pressure
curve of water intersects the line corresponding to a pressure of 1000 mmHg.
B. To estimate the pressure required for mercury to boil at 250°C, find the point where the vapor pressure curve of mercury
intersects the line corresponding to a temperature of 250°C.

Solution:
a. A The vapor pressure curve of water intersects the P = 1000 mmHg line at about 110°C; this is therefore the boiling point
of water at 1000 mmHg.
b. B The vertical line corresponding to 250°C intersects the vapor pressure curve of mercury at P ≈ 75 mmHg. Hence this is
the pressure required for mercury to boil at 250°C.

 Exercise 4.9.2: Boiling Ethlyene Glycol


Ethylene glycol is an organic compound primarily used as a raw material in the manufacture of polyester fibers and fabric
industry, and polyethylene terephthalate resins (PET) used in bottling. Use the data in Figure 4.9.4 to estimate the following.
a. the normal boiling point of ethylene glycol
b. the pressure required for diethyl ether to boil at 20°C.

Answer a
200°C
Answer b
450 mmHg

Summary
Because the molecules of a liquid are in constant motion and possess a wide range of kinetic energies, at any moment some fraction
of them has enough energy to escape from the surface of the liquid to enter the gas or vapor phase. This process, called
vaporization or evaporation, generates a vapor pressure above the liquid. Molecules in the gas phase can collide with the liquid
surface and reenter the liquid via condensation. Eventually, a steady state is reached in which the number of molecules
evaporating and condensing per unit time is the same, and the system is in a state of dynamic equilibrium. Under these conditions,
a liquid exhibits a characteristic equilibrium vapor pressure that depends only on the temperature. We can express the nonlinear
relationship between vapor pressure and temperature as a linear relationship using the Clausius–Clapeyron equation. This
equation can be used to calculate the enthalpy of vaporization of a liquid from its measured vapor pressure at two or more
temperatures. Volatile liquids are liquids with high vapor pressures, which tend to evaporate readily from an open container;
nonvolatile liquids have low vapor pressures. When the vapor pressure equals the external pressure, bubbles of vapor form within
the liquid, and it boils. The temperature at which a substance boils at a pressure of 1 atm is its normal boiling point.

4.9.7 https://chem.libretexts.org/@go/page/41427
Contributors and Attributions
Modified by Joshua Halpern (Howard University)

4.9: Phase Equilibria is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
11.5: Vapor Pressure is licensed CC BY-NC-SA 3.0.

4.9.8 https://chem.libretexts.org/@go/page/41427
4.10: Thermodynamics of Rubber Elasticity
4.10: Thermodynamics of Rubber Elasticity is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.

4.10.1 https://chem.libretexts.org/@go/page/41428
4.E: Exercises
Q4.9
What is the entropy changed when 10.2 moles of diatomic gas is cooled from a temperature of 25 °C to 10 °C at constant pressure?

Q4.9
Determine the value of while reversibly heating 5 moles of an ideal gas from 25 °C to 73 °C at constant volume.

S4.9
At constant volume, dqrev = CV dT.

Q4.10
If 1.50 moles of an ideal gas were compressed isothermally from 6.6 L to 2.5 L:
a. What is Ssys when compressed irreversibly?
b. Ssys when done reversibly?
c. What do the values mean for the overall process?

S4.10
V2 L ⋅ atm 2.5L kJ
a. △ Ssys = nR ln →△ Ssys = (1.5mol)(8.314 )(ln ) = −12.1
V1 mol ⋅ K 6.6L mol

b. The change in entropy of the system is the same whether an irreversible or reversible process is being carried out.
c. The negative value means that the reaction that was carried out was unfavorable. Entropy increases as molecules become more
disordered.

Q4.13
Nitrogen gas is reversibly heated at constant pressure of 101.3 kPa. The molar heat capacity of nitrogen gas is:
Cp = A + BT + CT2, where A = 27.4 J/mol-K, B = 5.23 x 10-3 J/mol-K2, and C = -0.0300 x 10-7 J/mol-K3.
Assuming N2 is an ideal gas, calculate the change in entropy when 42 g of N2(g) is heated from 250 K to 350 K.

Q4.21
Imagine you are to carry out the isothermal and reversible gas expansion with an ideal gas. You have 0.5 moles of the gas at 280K.
You wish to expand this gas from its initial volume of 2 liters to a new volume of 8 liters. What would your values for heat, work,
change in entropy, and change in internal energy have to be? What if you wanted to instead carry his experiment out as isothermal
and irreversible with 2.0 atm of external pressure?

Q4.22
Consider 0.50 moles of an ideal gas at 30oC. It expands from a pressure of 4.5 atm to a pressure of 2.0 atm at the same temperature.
a. For a reversible process, find the entropy change for the surrounding.
b. For an irreversible process, calculate the entropy change for the universe which against a constant external pressure of 2.0 atm.

S4.22
a) Reversible process:
delta S of system =(\dfrac{q}{T}\)=(\dfrac{-W}{T}\)=(\dfrac{-1}{T}*(-nRT\ln\dfrac{V2}{V1})\)=(\nR\ln\dfrac{P1}{P}\)

4.E.1 https://chem.libretexts.org/@go/page/41330
qsys/T = - Wsys/T = -1/T *(-nRT\ln (V2/V1) = nR\ln (P1/P2)
= 0.5 moles * 8.3145 Jmol-1K-1 * \ln(4.5/2)
= 3.37 J/K
delta S of surrounding = -3.37 J/K
b) Irreversible process:
delta S of system is 3.37 J/K
Delta S(surrouding)=(\dfrac{-1}{T}*P2*V2*(1-\dfrac{V1}{V2})\)= (\-nR(1-\dfrac{P2}{P1})\)
delta S of surrounding = -1/T * P2*V2*(1 - V1/V2) = -nR(1 - P2/P1)
=-0.5 * 8.3145 *(1-2/4.5)=2.31 J/K
delta of the universe is : 3.37 J/K +2.31J/K=5.68 J/K

Q4.23
Calculate ΔSsys for the heating of 2.00 moles of nitrogen from 25ºC to 200ºC. The heat capacity of oxygen is:
Cp = (3.268 + 0.00325T) J K-1 mol-1

S4.23
¯
CP
ΔS = ∫ n dT (4.E.1)
T

(200+273)K −1 −1
(3.268 + 0.00325T )J mol K
ΔS = ∫ (2mol) dT (4.E.2)
(25+273)K T

473K
6.536 −1
ΔS = ∫ ( + 0.0065)J mol dT (4.E.3)
298K
T

473 −1
ΔS = (6.536 ln(T ) + 0.0065T ) J mol (4.E.4)
298

−1
ΔS = 4.16J mol (4.E.5)

Q4.27
The value of (CO2(aq)) is less than that of (CO2(g)) at 298 K. (Note: Refer to table below). Would this be the case at a
temperature of 0 K as well?

Substance / J K-1 mol-1

CO2(g) 213.6

CO2(aq) 117.6

S4.27
Since the gaseous form of a compound always has a greater enthalpy than the more ordered liquid or aqueous form, this would still
be the case at a temperature of 0 K as well.

Q4.27
Is the standard entropy of CO2(g) higher than the standard entropy of CO2(s) at 298 K? What about at 0 K?

Q4.28
You have an ideal gas that expands 3.56 L with the initial volume at 22 L. All you know is that you have 0.45 moles and that the
temperature is 22 ºC. What is the change in Gibbs free energy for this system?

4.E.2 https://chem.libretexts.org/@go/page/41330
S4.28

=167.3

Q4.28
0.70 moles of an ideal gas expands adiabatically from 1.0 atm to 2.5 atm at a temperature of 30ºC. Calculate the values of q, w, U,
S, and G.

S4.28
P1 J 1.0atm 1kJ kJ
q = 0 w = −nRT ln = −(.70mol)(8.314 )(303K)(ln )( = 1.62
P2 mol ⋅ K 2.5atm 1000kJ mol

kJ
△ U = w = 1.62
mol

P1 J 1.0atm J
△ Suniv =△ Ssys = nR ln = (.70mol)(8.314 ) ln = −5.33
P2 mol ⋅ K 2.5atm mol

dH P2 J 2.5atm 1kJ
dH = V dP → = nRT ln = (.70mol)(8.314 (303K)(ln (
dP P1 mol ⋅ K 1.0atm 1000J

kJ
= 1.62
mol

3
J J J
△ G =△ H − T △ S = 1.62 × 10 − (303K)(−5.33 = 3234
mol mol/cdotK mol

kJ
= 3.23
mol

Q4.30
Use the values listed listed below to calculate the value of ΔrG° for the following reaction:
pyruvate(aq) → 2 lactate ion(aq) + 2H+(aq)
(ΔfG°[pyruvate(aq)] = -472.4 kJ mol-1, ΔfG°[lactate ion] = -516.7 kJ mol-1). Is pyruvate or the lactate ion favored under standard
conditions? Explain.

Q4.30
Use the values listed listed below to calculate the value of ΔrG° for the following reaction:
pyruvate(aq) → 2 lactate ion(aq) + 2H+(aq)
(ΔfG°[pyruvate(aq)] = -472.4 kJ mol-1, ΔfG°[lactate ion] = -516.7 kJ mol-1)

Q4.31
A scientist measures studies the thermodynamics of a protein of interest. He interested in finding the Free Energy of unfolding in
the quaternary structure of protein at a concentration of 5µM. He finds that the Gibbs free energy of the reaction -0.84 Kcal/mol.
Why does the scientist want to know the value of (\Delta G\) Explain how the reaction proceeds.

S4.31
The value of (\Delta G\) of the reaction will let the scientist know how likely the reaction of the unfolding of the protein is to occur.
If (\Delta G\) is negative the reaction will go forward, towards the unfolded state. In this case (\Delta G\) is negative so the reaction
goes forward to the unfolding of the protein.

4.E.3 https://chem.libretexts.org/@go/page/41330
Q4.32
For the reaction

H2(g) + C l2(g) → 2H C l(g) (4.E.6)

Using the formula ∆rG = ∆fG (products)-∆fG (reactants) to compute the Gibbs energy of the reaction?

Q4.34
Consider the conversion of graphite carbon into diamond
C(graphite) → C(diamond) (4.E.7)

a. Determine ΔrH and ΔrS o) for the reaction. Is the conversion spontaneous at any temperature?
o (

b. The molar volume of graphite is 2.1cm greater than diamond in density.


3

c. If you apply pressure on graphite at 25 C , can you convert graphite to diamond? If so, determine the required pressure to make
o

the process spontaneous.

S4.34
(a)
$$\Delta rH^{o}=\Delta _f\bar{H}^{o}[C_{(diamond)}]-\Delta _f\bar{H}^{o}[C_{(graphite)}]\]
$$=1.90Kjmol^{-1}-0kjmol^{-1}\]
$$=1.90kj\,mol^{-1}\]
$$\Delta rS^{o}=\Delta S^{o}[C_{(diamond)}]-\Delta S^{o}[C_{(graphite)}]\]
$$=2.4J\,K^{-1}mol^{-1}-5.7JK^{-1}mol^{-1}\]
$$=-3.3J\,K^{-1}mol^{-1}\]
At 0 o
C ,
$$\Delta rG^{o}=\Delta rH^o-(273K)\Delta rS^{o}\]
$$=(1.90kj\,mol^{-1})-(273K)(-3.3\times10^{-3}kj\,K^{-1}mol^{-1})\]
$$=2.8834kj\,mol^{-1}\]
(b)
Step 1:
$$G_2=G_1+V(P_2-P_1)\]
Step 2: (Hint: Apply molar quantities)
$$\bar{G}_2=\bar{G}_1+\bar{V}\Delta P\]
Step 3: (Hint: Apply the equation to diamond and graphite)
$$\bar{G}_2[Graphite]=\bar{G}_1[graphite]+\bar{V}[graphite]\Delta P\]
$$\bar{G}_2[diamond]=\bar{G}_1[diamond]+\bar{V}[diamond]\Delta P\]
Step 4: (Hint: Combine the equations)
$$\Delta rG_2=\Delta rG_1+[\bar{V}[diamond]-\bar{V}[graphite]]\Delta P\]
Step 5: (Hint: At 25 o
C , P1 = 1bar so ΔrG1 = ΔrG
o
= 2.883kj mol
−1
)
$$\Delta rG_2=2.883kj\,mol^{-1}+(-2.1cm^3mol^{-1})\Delta P(\dfrac{1L}{1000cm^{3}})(\dfrac{1atm}{1.01 bar})
(\dfrac{101.3J}{1L\,atm})(\dfrac{1kj}{1000J})\]
$$\Delta rG_2=2.883kj\,mol^{-1}-2.1\times10^{-4}\Delta P\,kj\,bar^{-1}mol^{-1}\]
Step 6: (Hint: To make the process spontaneous)

4.E.4 https://chem.libretexts.org/@go/page/41330
$$\Delta rG_2=2.883kj\,mol^{-1}-2.1\times10^{-4}\Delta P\,kj\,bar^{-1}mol^{-1}\lt0\]
$$2.883kj\,mol^{-1}\lt2.1\times10^{-4}\Delta P\,kj\,bar^{-1}mol^{-1}\]
$$\Delta P\lt\dfrac{2.883kj\,mol^{-1}}{2.1\times10^{-4}}=1.4\times10^{4} bar\]

Q4.36
Predict the signs of ΔH , ΔS and ΔG of the system for the following processes at 1atm:
a. Ice melt at 200 K
b. Ice melt at 273.15 K
c. Ice melt at 300 K

S4.36
ΔH is positive for all because melting is endothermic. ΔS is positive for all because going from solid to liquid is becoming more
disordered. ΔG for (a) is positive because reaction is not spontaneous when melting at below the melting point, ΔG for (b) is 0
because reaction is in equilibrium at the melting point, ΔG for (c) is negative because reaction is spontaneous when melting at
above the melting point.

(a) (b) (c)

ΔH + + +

ΔS + + +

ΔG + 0 -

Q4.37
Crystals of AgCl(s) form spontaneously when aqueous solutions of silver(I)nitrate and sodium chloride are combined. What can
you deduce about the signs of the changes in enthalpy and entropy?

Q4.38
The formation reaction to form magnesium oxide is as follow:
1
Mg(s) + O (g) → MgO(s) (4.E.8)
2 2

Verify the equation ΔG° = ΔH° + TS° using values from Appendix
Hint: Gibbs free definition above is free energy of the whole reaction, not free energy of formation

S4.38
∘ ∘ ∘ kJ kJ
¯ ¯ ¯
Δr G = ∑ Δf G [products] − ∑ Δf G [reactants] = −569.5 − 0 − 0 = −569.5 (4.E.9)
mol mol

∘ ∘ ∘ kJ kJ
Δr H̄ = ∑ Δf H̄ [products] − ∑ Δf H̄ [reactants] = −601.8 − 0 − 0 = −601.8 (4.E.10)
mol mol

∘ ∘ ∘ 205.0 J
¯ ¯ ¯
Δr S = ∑ Δf S [products] − ∑ Δf S [reactants] = (26.78 − 32.68 − ) = −108.4 (4.E.11)
2 mol ⋅ K

∘ ∘ kJ kJ kJ
¯
Δr H̄ − T Δr S = −601.8 − (298K)(−0.1084 ) = −569.5 (4.E.12)
mol mol ⋅ K mol

The answers agree therefore the equation is true.

Q4.39a
At 49ºC, a particular reaction is spontaneous. What would be the smallest possible value of ΔrS for the reaction if ΔrH was 10 kJ?

4.E.5 https://chem.libretexts.org/@go/page/41330
S4.39a
Use the following equation

ΔG = ΔH − T ΔS (4.E.13)

Since the process is spontaneous:


Note: Don't forget to flip the inequality when dividing by a negative number.
ΔG < 0 (4.E.14)

ΔH − T ΔS < 0 (4.E.15)

(10 kJ) − (322 K)ΔS < 0 (4.E.16)

−(322 K)ΔS < 10 kJ (4.E.17)

ΔS > 0.031 (4.E.18)

Thus the minimum value is the closest number above 0.031 kJ K-1.
Click here for more information on Gibb's free energy.

Q4.39b
Calculate the minimum change in entropy of reaction for a spontaneous process at 340K with an enthalpy change of reaction equal
to 800 J. Your answer must be in KJ/K and have 2 significant figures.

Q4.39c
The entropy change for a reaction is found to be 47.50 J/K. The enthalpy change for the reaction is 21.25 KJ. At what temperature
is this reaction spontaneous?

S4.39c
Δr G = Δr H − T Δr S (4.E.19)

Because we want to find the temperature at equilibrium we set the Gibbs Free Energy equal to 0. Thus:
Δr H = T Δr S (4.E.20)

3
Δr H 21.25 × 10 J
T = = = 447.4K (4.E.21)
J
Δr S 47.50
K

Q4.40a
The standard Gibbs energy change, ∆rGo, for the reaction is +108 kJ. What can you deduce about the state of reaction when the
reactants are mixed together?

S4.40a
When the two gases are mixed, there is no interaction between them. So, the products would not form.

Q4.40b
A scientist mixes together a set of reagents which are known to undergo a reaction with a negative change in Gibbs’ free energy in
a beaker. When the scientist comes back in a few hours, they find by looking at the beaker that the reaction does not appear to have
proceeded. What are two reasons that the scientist would not observe a change after two hours?

S4.40b
The Gibbs' free energy is an indicator of a reaction's favorability, but does not indicate that that reaction will proceed quickly. If the
rate constant for the reaction is low, or the concentration of one or more of the reagents is low, the reaction might not proceed at a
pace that the scientist would observe over a few hours or even longer.

4.E.6 https://chem.libretexts.org/@go/page/41330
Q4.41
If water changes from a liquid to a gas at 50ºC and 0.123 bar, what must the pressure be to convert water from a liquid to a gas at
75ºC? The molar enthalpy of vaporization for water (ΔvapH) is 42.3 kJ mol-1

S4.41
¯
P2 ΔH vap 1 1
ln( ) = ( − ) (4.E.22)
P1 R T1 T2

T1 = (50 + 273) K = 323 K (4.E.23)

T2 = (75 + 273) K = 348 K (4.E.24)

3 −1
P2 42.3 × 10 Jmol 1 1
ln( ) = ( − ) (4.E.25)
−1 −1
0.123 bar 8.314J mol K 323 K 348 K

P2
ln( ) = 1.132 (4.E.26)
0.123 bar

P2
= 3.10 (4.E.27)
0.123 bar

P2 = 0.381 bar (4.E.28)

Q4.42
What is the change in pressure when ice melts ast room temperature if the density of ice is 0.919 g/ml and the ΔfusH = 6.01
KJ/mol?

Q4.42
Calculate the freezing point depression of ice if a pressure of 450 atm is exerted on it by a car that weighs 1500-kg. Note: The
molar volumes are as follows: molar volume of liquid water is 18.03 mL mol-1 and the molar volume of ice is 19.65 mL mol-1.

S4.42
Use the slope of the S-L curve from the phase diagram:

This means that the freezing point depression determined for when 450 atm of pressure is applied to ice by a car is 3.30 K.

Q4.43
Here is a phase diagram for carbon.
a. What phase will carbon be if temperature is increased at constant pressure from the triple point between the vapor, liquid, and
graphite?
b. At 1.0 GPa and 5000 K, what phase will carbon be in?
File:Carbon basic phase diagram.png

4.E.7 https://chem.libretexts.org/@go/page/41330
S4.43
a. Carbon will be in its vapor phase if the temperature is increased from the lower triple point.
b. At 1.0 GPa and 5000 K, carbon will be in the liquid phase.

Q4.46
Look at the phase diagram of water. Which phase is more dense, liquid or solid? How can you tell by just looking at the graph?
The link below contains a phase diagram of water:
http://chemwiki.ucdavis.edu/Wikitext...Phase_Diagrams

S4.46
Through looking at the phase diagram, we can see that the liquid phase is more dense. You can see this because the line separating
the two phases is negative.

Q4.48
Explain what happens to ΔvapH as the temperature increases significantly.

Q4.49
You go on an adventure and find yourself stranded on the top of a mountain where the air pressure is 0.68 atm. You find water but
you want to make sure is safe to drink so you want to boil the water. At what temperature would you expect the water to start
boiling? Hint: The molar heat of vaporization of H2O is 41.0 kJ mol-1

Q4.50
At higher altitude, water is boiled faster than at lower altitude. The reason is because the pressure is smaller. Given the heat of
vaporization of water at 100oC is 42.0 KJ/mol. Calculate the temperature to boil the water at a place where air pressure is 0.7 atm?

Q4.52
The equations are as follows
C pdT
a. dS =
T
ΔH
b. ΔS =
T
c. S0 = 0

dq
d. dS =
T

Determine the conditions needed for each equation.

S4.52
a. Constant pressure
b. Constant pressure and temperature
c. Absolute 0, no residual entropy, pure crystalline
d. reversible process

Q4.54
Calculate the entropy change when nitrogen gas is heated from 25°C to 75°C while simultaneously allowed to expand from 0.562 L
to 1.245 L. Assume ideal gas behavior.

S4.54
1. Find how many moles of gas we have: P V = nRT

PV 1atm ∗ 0.562L
n = = = 0.02297mol (4.E.29)
RT (0.08206L ∗ atm/K ∗ mol)(298.15K)

4.E.8 https://chem.libretexts.org/@go/page/41330
2. entropy change during expansion:
V2 1.245L
DeltaS = nR ln = (0.02297mol)(8.314J/K ∗ mol) ln = 0.1519J/K (4.E.30)
V1 0.562L

3. entropy change during heating at constant volume:


T2 5 T2
¯
ΔS = nC v ln = n( R) ln
T1 2 T1

5 348.15K
= (0.02297mol)( )(8.314J/K ∗ mol) ln = 0.0987J/K
2 283.15K

4. Combine 2&3:
ΔStot = ΔSexpansion + ΔSheating = 0.1519J/K + 0.0987J/K = 0.2506J/K (4.E.31)

Q4.54
A 4 × 10 moles sample of Xenon (Xe) is heated from 300 K to 400 K in a 1 L container at 1.5 atm. What will be the change in
−2

entropy if Xe is able to expand to twice the volume of the container?

Q4.55
Photosynthesis depends upon the absorption of visible light. However, the infrared radiation can not be used for photosynthesis.
Explain why.

S4.55
The infrared light : Cells do absorb the photon of infrared radiation, but it could not provide enough energy to do work in a
biological system.
Whereas, for the visible light, each photon of visible light contains just enough energy to excite the electrons of molecules.
Photosynthesis is carried out by using the visible light.

Q4.56a
An amount of ideal diatomic gas is expanded from 4.8 atm to 1.7 atm while experiencing the change of temperature from 35°C to
128ºC. The change in entropy of the gas expansion process is 15.0 J/K. Determine the ΔU, ΔH, and ΔS of the entire process.
Assume the temperature is not enough to unleash the rotational and vibrational energy of the gas.
Hint: Break the problem into 2 processes (isothermic and isobaric)

S4.56a
Step 1: Consider the isothermal process of the ideal gas (expanding from 4.8 atm to 1.7atm at constant temperature of 35°C
ΔU = 0, ΔH = 0, ΔS = 15.0 J/K
V2 P1
ΔS = nRln = nRln (4.E.32)
V1 P2

J
15.0
ΔS K
n = = = 1.738moles (4.E.33)
P1 J 4.8atm
Rln (8.314 )ln
P2 mol ⋅ K 1.7atm

Step 2: Consider the change in temperature at constant pressure of 1.7 atm


7 7 J
3
ΔH = Cp ΔT = nRΔT = (1.738moles)(8.314 )(128 − 35)K = 1.88 × 10 J (4.E.34)
5 5 K ⋅ mol

J
3
ΔU = ΔH − nRΔT = (1.88 × 10 J) − (1.738moles)(8.314 )(128 − 35)K = 537J (4.E.35)
K ⋅ mol

4.E.9 https://chem.libretexts.org/@go/page/41330
T2 7 T2 7 J (128 + 273)K J
ΔS = Cp ln = nRln = (1.738moles)(8.314 )ln = 5.34 (4.E.36)
T1 5 T1 5 K ⋅ mol (35 + 273)K K

For the entire process


3 3
ΔH = 0J + 1.88 × 10 J = 1.88 × 10 J (4.E.37)

ΔU = 0J + 537J = 537J (4.E.38)

J J J
ΔS = 15.0 + 5.34 = 20.34 (4.E.39)
K K K

Q4.56b
Calculate the ΔU, ΔH, and ΔS of 3.00 moles of an ideal monatomic gas that are compressed from 7.00 L to 1.00 L while being
heated from 25 ºC to 100 ºC.

S4.56b
Break the process down into two steps: isothermal compression from 7 L to 1 L at 298 K and heating at a constant pressure from
298 K to 373 K.
a. Compression:
Because the compressions is isothermal, ΔU=0 & ΔH=0 for this step
V2 1
−1 −1
ΔS = nR ln( ) = (3mol)(8.314J mol K ) ln( ) (4.E.40)
V1 7

−1
ΔS = −48.5J K (4.E.41)

b. Heating:
5 5
−1 −1
ΔH = CP ΔT = nRΔT = (3mol)(8.314J mol K )(373 − 298)K (4.E.42)
2 2

ΔH = 4677J (4.E.43)

−1 −1
ΔU = ΔH − nRΔT = 4677J − (3mol)(8.314J mol K )(373 − 298)K (4.E.44)

ΔU = 2806J (4.E.45)

T2 5 T2
ΔS = CP ln( ) = nR ln( ) (4.E.46)
T1 2 T1

5 373
−1 −1 −1
ΔS = (3mol)(8.314J mol K ) ln( ) = 14.0J K (4.E.47)
2 298

c. For entire process

ΔH = 0J + 4677J = 4.68kJ (4.E.48)

ΔU = 0J + 2806J = 2.81kJ (4.E.49)

−1 −1 −1
ΔS = −48.5J K + 14.0J K = −34.5J K (4.E.50)

Q4.57a
Briefly provide examples and descriptions for each of the three laws of thermodynamics.

S4.57a
The first law is the conservation of mass and energy. Basically you can't create anything out of thin air and perpetual motion
machines are impossible.
The second law states how entropy is always increasing and the entropy of the universe can never become negative. In other
words things will always attempt to go to equilibrium. This is seen commonly when batteries run out. You never see a battery
charge itself spontaneously because that would violate the second law.

4.E.10 https://chem.libretexts.org/@go/page/41330
The third law states how a pure crystalline structure at 0 K will essentially have 0 entropy. The third law is visualized by
knowing how frozen and solid materials have less entropy because the molecules within can only vibrate.

Q4.57b
The three laws of thermodynamics are often generally stated to be:
1. Energy and matter cannot be created or destroyed.
2. The entropy of the universe is always increasing.
3. At absolute zero, the entropy of a system is zero.
With this knowledge, which of the three laws of thermodynamics apply to the examples below? There may be more than one.
a.) A graduate student in UC davis claims to have created the world's first perpetual motion engine. This machine produces no heat
and can be used, (he claims), to power the campus for the next decade! This is not possible.

b.)
If TF becomes 0, entropy becomes undefined. Hence, it is not possible to reach absolute zero.
c.) In an adiabatic gas expansion where:
ΔU = Q + w (4.E.51)

No thermal energy or matter is exchanged. Hence:


ΔU = w (4.E.52)

d.) In an imperfect crystal of CO, carbon monoxide, there are multiple energy states at 0K. Thus, it experiences residual entropy
and, quite unusually, follows:
lim S >0 (4.E.53)
T →0K

S4.57b
a.) first and second laws; The existence of the perpetual motion machine violates the first and second laws of thermodynamics.
Because
ΔU = Q + w (4.E.54)

and heat is a form of entropy, a perpetual motion machine does not increase the entropy of the universe. By definition, a machine
which does work must do less work than it has internal energy.
b.) second and third law; In order for absolute zero to be reached, the system would have to be removed from the universe. Because
T2
¯
ΔS = nC p ln ( ) (4.E.55)
T1

and entropy must always increase, a system within the universe would always be at a temperature lower than its surroundings and
hence could experience heat transfer which would mitigate the absolute zero condition.
c.) first law; Energy which is "lost" from the system due to lack of heat transfer is compensated by the system doing more work.
d.) third law; In a perfect crystal, where there is only one energy state, there is no residual entropy at 0K

Q4.58
The standard enthalpy of formation of ice is -291.83 kJ/mol and its standard molar entropy is 41.0 J/(mol K). The standard enthalpy
of formation of liquid water is -285.83 kJ/mol and its standard molar entropy is 69.95 J/(mol K). Estimate the melting point of
water based on this information. What would you expect the value be? Do your values match? What assumptions made during this
calculation could be responsible for the difference?

4.E.11 https://chem.libretexts.org/@go/page/41330
Q4.61
In a reversible adiabatic expansion, what do you expect to happen to the volume? What about the temperature?

S4.61
We expect both volume and temperature to increase.

Q4.63
Aniline can hydrogen bond, whereas benzene is nonpolar. Explain why, contrary to our expectation, benzene melts at a higher
temperature than aniline. Why is the boiling point of aniline higher than that of benzene?

Q4.64
Molar Gibbs energies of solid, liquid and vapor depend on temperature and pressure. (a) Briefly explain how temperature affects
molar Gibbs energy. (b) Explain how pressure changes affect molar Gibbs energy. Why is water an exemption?

S4.64
a. At a constant pressure the phase with the lowest molar Gibbs energy is the most stable phase at that temperature. At high
temperatures the vapor phase has the lowest molar Gibbs energy, therefore it is the most stable. As temperature decreases, liquid
becomes the stable phase, and at even lower temperatures, solid becomes the most stable phase.
∂ΔG
b. From the equation ( ¯
) =V we observe that the molar volume affects the molar Gibbs energy depending on pressure. In
∂P

this case the vapor phase has the greatest increase in molar Gibbs energy because the molar volume is much larger than that of
liquids and solids. In general, an increase in external pressure will raise both the melting point and the boiling point of a
substance. The exception to this, is water. This is because the molar volume of ice is actually greater than that of liquid water.
Thus, for water an increase in pressure will lower the melting point.

Q4.65
When ice melts to liquid water, entropy is increasing and the the surrounding becomes cooler since ice absorbs heat. Explain the
relationship between those two facts.

Q4.66
For any spontaneous endothermic process, what must be true about the entropy of the process?

Q4.67
For the following situation take a rubber band and stretch it quickly. It will feel warm. Next stretch it out for a few seconds and
then release it. Consequently, it will feel cool. Why do you think this happens? Explain thermodynamically.

S4.67
ΔG = ΔH − T ΔS (4.E.56)

The warming effect: Stretching the rubber band creates a nonspontaneous process (ΔG > 0) and −(ΔG < 0) . The warming
effect means that ΔH < 0) . This makes ΔS negative. T must be positive and ΔS must be negative. This tells us that the rubber
band under tension is more disordered in its natural state.
The cooling effect: If theres no tension and the rubber band snaps back to its natural state, ΔG is negative and −ΔG is positive.
Since it cools the ΔH > 0 so T ΔS is positive. Entropy increases when the rubber band goes from stretched state to natural state.

Q4.69
Cold packs utilizes the chemical reaction of water and ammonium nitrate (or similar substances) to turn cold. Predict the signs of
ΔH , ΔS and ΔG for the reaction when the membrane that separates the two substances in a cold pack breaks.

4.E.12 https://chem.libretexts.org/@go/page/41330
S4.69
The reaction is spontaneous so ΔG is negative; the cold pack feels cold because it's taking in heat so ΔH is positive; ΔS is
positive because disorder increases as ammonium nitrate breaks down to ammonium ion and nitrate ion.

Q4.71
At what temperature will toluene have a vapor pressure of 517 torr. The normal boiling point of toluene is 110.6°C with the molar
enthalpy of vaporization being 35.2 kJ/mol. (Hint: Use Clausius-Clapeyron equation).

S4.71
¯
P2 Δv ap H 1 1
ln = ( − ) (4.E.57)
P1 R T2 T1

J
3
35.2 × 10
517torr mol 1 1
ln = ( − ) (4.E.58)
760torr J T2 (110.6 + 273)K
8.314
K ⋅ mol


⇒ T2 = 397.5K = 124 C (4.E.59)

Q4.72a
You have discovered a way to create a new organic compound of the formula C x Hy Nz Oα . What measurements would you need to
take to find out values for the following:properties:
∘ ∘ ∘
¯
¯¯¯
¯ ¯¯
¯ ¯
¯¯¯
△f H = ? △f S = ? △f G = ? (4.E.60)

S4.72a
For enthalpy, you need to take measurements of the change of enthalpy of one mole of the unknown compound. Thus you would
need to take measurements for the molar heat capacity and how much temperature changes in for formation of one mole. See here
and here.
For entropy of formation you would need to divide enthalpy by the temperature at which the formation takes place because the
change in entropy can equate to enthalpy divided by temperature. See here.
For Gibb's free energy of formation, you simply need the enthalpy and entropy of formation along with the temperature where one
mole of the substance is formed. See here.
ΔG = ΔH − T ΔS (4.E.61)

Q4.72b
You are given hexane, C6H14, and a constant pressure calorimeter. What steps would you take to determine standard gibbs free
energy of formation, for this substance?
Hint: of hexane can be found using a measurement obtained by the calorimeter

S4.72b
An equation for the complete combustion of hexane can be written as follows:

C6 H14(l) + O2(g) → C O2(l) + H2 O(l) (4.E.62)

The calorimeter would then be used to find the heat released by this reaction by the following rule:

Δr H = −qP ,calorimeter = − (Ccalorimeter ΔT ) (4.E.63)

after which the standard heat of formation for hexane could be calculated using the known heat of formation for H2O, O2, and CO2.
Under the 3rd law of dynamics, the standard entropy would then be calculated by
∘ qrxn
¯
ΔS =− (4.E.64)
T

4.E.13 https://chem.libretexts.org/@go/page/41330
where T=298K
From this, standard entropy for the formation of hexane should be obtained from the known standard entropies of H2O, O2, and
CO2, through the following reaction for the formation of hexane from elemental reactants:

6 C(graphite) + 7 H2(g) → C6 H14(l) (4.E.65)

The standard gibbs free energy of formation, for this substance is then determined through the relationship:
∘ ∘
¯ ∘
Δf G = Δf H̄ − T Δr S (4.E.66)

Q4.73
A 93.5 Liter drum at 65 C houses 20g of water. How many moles of water are in the vapor phase, and what mole fraction is in the
vapor phase? The vapor pressure of water at 65C is 187.54 mmhg.

S4.73
PV
=n (4.E.67)
/RT

ntot = 20g/18.02 g mol-1= 1.11 mol


nvap = (187.54 mmhg 93.5 L)/(8.314 L atm mol-1 K-1* 338 K * 710 mmhg atm-1) = 8.79 *10-3 mol
xvap= 8.79 *10-3 mol / 1.11 mol = 7.92*10-3

4.E: Exercises is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

4.E.14 https://chem.libretexts.org/@go/page/41330
CHAPTER OVERVIEW

5: Solutions
5.1: Concentration Units
5.2: Partial Molar Quantities
5.3: The Thermodynamics of Mixing
5.4: Binary Mixtures of Volatile Liquids
5.5: Real Solutions
5.6: Colligative Properties
5.7: Electrolyte Solutions
5.8: Ionic Activity
5.9: Colligative Properties of Electrolyte Solutions
5.10: Biological Membranes
5.E: Solutions (Exercises)

5: Solutions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1
5.1: Concentration Units
5.1: Concentration Units is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

5.1.1 https://chem.libretexts.org/@go/page/41429
5.2: Partial Molar Quantities
5.2: Partial Molar Quantities is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

5.2.1 https://chem.libretexts.org/@go/page/41430
5.3: The Thermodynamics of Mixing
When solids, liquids or gases are combined, the thermodynamic quantities of the system experience a change as a result of the
mixing. This module will discuss the effect that mixing has on a solution’s Gibbs energy, enthalpy, and entropy, with a specific
focus on the mixing of two gases.

Introduction
A solution is created when two or more components mix homogeneously to form a single phase. Studying solutions is important
because most chemical and biological life processes occur in systems with multiple components. Understanding the
thermodynamic behavior of mixtures is integral to the study of any system involving either ideal or non-ideal solutions because it
provides valuable information on the molecular properties of the system.
Most real gases behave like ideal gases at standard temperature and pressure. This allows us to combine our knowledge of ideal
systems and solutions with standard state thermodynamics in order to derive a set of equations that quantitatively describe the
effect that mixing has on a given gas-phase solution’s thermodynamic quantities.

Gibbs Free Energy of Mixing


Unlike the extensive properties of a one-component system, which rely only on the amount of the system present, the extensive
properties of a solution depend on its temperature, pressure and composition. This means that a mixture must be described in terms
of the partial molar quantities of its components. The total Gibbs free energy of a two-component solution is given by the
expression
¯
¯¯¯ ¯
¯¯¯
G = n1 G1 + n2 G2 (5.3.1)

where
G is the total Gibbs energy of the system,
ni is the number of moles of component i,and
¯
¯¯¯
Gi is the partial molar Gibbs energy of component i.
The molar Gibbs energy of an ideal gas can be found using the equation

¯
¯¯¯ ¯
¯¯¯
∘ P
G =G + RT ln (5.3.2)
1bar


¯
¯¯¯
where G is the standard molar Gibbs energy of the gas at 1 bar, and P is the pressure of the system. In a mixture of ideal gases, we
find that the system’s partial molar Gibbs energy is equivalent to its chemical potential, or that
¯
¯¯¯
Gi = μi (5.3.3)

This means that for a solution of ideal gases, Equation 5.3.2 can become

¯
¯¯¯ Pi

Gi = μi = μ + RT ln (5.3.4)
i
1bar

where
µi is the chemical potential of the ith component,
µi° is the standard chemical potential of component i at 1 bar, and
Pi is the partial pressure of component i.
Now pretend we have two gases at the same temperature and pressure, gas 1 and gas 2. The Gibbs energy of the system before the
gases are mixed is given by Equation 5.3.1, which can be combined with Equation 5.3.4 to give the expression
∘ ∘
Ginitial = n1 (μ + RT ln P ) + n2 (μ + RT ln P ) (5.3.5)
1 2

If gas 1 and gas 2 are then mixed together, they will each exert a partial pressure on the total system, P1 and P2 , so that
P + P = P . This means that the final Gibbs energy of the final solution can be found using the equation
1 2

∘ ∘
Gf inal = n1 (μ + RT ln P1 ) + n2 (μ + RT ln P2 ) (5.3.6)
1 2

5.3.1 https://chem.libretexts.org/@go/page/41431
The Gibbs energy of mixing, Δ mix , can then be found by subtracting G
G initial from G f inal .
Δmix G = Gf inal − Ginitial (5.3.7)

P1 P2
= n1 RT ln + n2 RT ln (5.3.8)
P P

= n1 RT ln χ1 + n2 RT ln χ2 (5.3.9)

where
Pi = χi P (5.3.10)

and χ is the mole fraction of gas i. This equation can be simplified further by knowing that the mole fraction of a component is
i

equal to the number of moles of that component over the total moles of the system, or
ni
χi = . (5.3.11)
n

Equation 5.3.9 then becomes


Δmix G = nRT (χ1 ln χ1 + χ2 ln χ2 ) (5.3.12)

This expression gives us the effect that mixing has on the Gibbs free energy of a solution. Since χ and χ are mole fractions that
1 2

range from 0 to 1, we can conclude that Δ G will be a negative number. This is consistent with the idea that gases mix
mix

spontaneously at constant pressure and temperature.

Entropy of mixing
Figure 5.3.1 shows that when two gases mix, it can really be seen as two gases expanding into twice their original volume. This
greatly increases the number of available microstates, and so we would therefore expect the entropy of the system to increase as
well.

Figure 5.3.1 : The mixing of two gases can be seen as two expansions. (a) Expansion of gas 1 alone when teh barrier is removed.
The molecules have twice as many microstates in the open box. (b) Expansion of gas 2 along. (c) the simultaneous expansion of
gases 1 and 2 is equivalent to mixing
Thermodynamic studies of an ideal gas’s dependence of Gibbs free energy of temperature have shown that
dG
( ) = −S (5.3.13)
dT
P

This means that differentiating Equation 5.3.12 at constant pressure with respect to temperature will give an expression for the
effect that mixing has on the entropy of a solution. We see that
dGmix
( ) = nR(x1 ln x1 + x2 ln x2 ) (5.3.14)
dT
P

= −Δmix S (5.3.15)

5.3.2 https://chem.libretexts.org/@go/page/41431
Δmix S = −nR(x1 ln x1 + x2 ln x2 ) (5.3.16)

Since the mole fractions again lead to negative values for ln x1 and ln x2, the negative sign in front of the equation makes ΔmixS
positive, as expected. This agrees with the idea that mixing is a spontaneous process.

Enthalpy of mixing
We know that in an ideal system ΔG = ΔH − T ΔS , but this equation can also be applied to the thermodynamics of mixing and
solved for the enthalpy of mixing so that it reads
Δmix H = Δmix G + T Δmix S (5.3.17)

Plugging in our expressions for Δ mix G (Equation 5.3.12) and Δ mix S (Equation 5.3.16) , we get
Δmix H = nRT (x1 ln x1 + x2 ln x2 ) + T [−nR(x1 ln x1 + x2 ln x2 )] = 0 (5.3.18)

This result makes sense when considering the system. The molecules of ideal gas are spread out enough that they do not interact
with one another when mixed, which implies that no heat is absorbed or produced and results in a Δ H of zero. Figure 5.3.2mix

illustrates how T Δ S and Δ G change as a function of the mole fraction so that Δ H of a solution will always be equal to
mix mix mix

zero (this is for the mixing of two ideal gasses).

Figure 5.3.2 : A graph of T Δ mix S , TΔ mix H , and T Δ


mix G as a function of x for the mixing of two ideal gases.
1

References
1. Chang, R. Physical Chemistry for the Biosciences, 1st Herdon, VA: University Science Books, 132-133. Print.
2. Meyer, E.F. (1987). Theromodynamics of “Mixing” of Ideal Gases. J. Chem. Educ. 64, 676-677.

Outside Links
Satter, S. (2000). Thermodynamics of Mixing Real Gases. J. Chem. Educ. 77, 1361-1365.
Brandani, V., Evangelista, F. (1987). Correlation and prediction of enthalpies of mixing for systems containing alcohols with
UNIQUAC associated-solution theory. Ind. Eng. Chem. Res. 26 (12), 2423–2430.

Problems
1. Use Figure 2 to find the x1 that has the largest impact on the thermodynamic quantities of the final solution. Explain why this is
true.
2. Calculate the effect that mixing 2 moles of nitrogen and 3 moles of oxygen has on the entropy of the final solution.
3. Another way to find the entropy of a system is using the equation ΔS = nRln(V2/V1). Use this equation and the fact that volume
is directly proportional to the number of moles of gas at constant temperature and pressure to derive the final expression for
TΔ mix S . (Hint: Use the derivation of T Δ G as a guide).
mix

Answers
1. x1= 0.5
2. Increases the entropy of the system by 27.98 J/molK

5.3.3 https://chem.libretexts.org/@go/page/41431
Contributors and Attributions
Elizabeth Billquist (Hope College)

5.3: The Thermodynamics of Mixing is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
Thermodynamics of Mixing is licensed CC BY-NC-SA 4.0.

5.3.4 https://chem.libretexts.org/@go/page/41431
5.4: Binary Mixtures of Volatile Liquids
5.4: Binary Mixtures of Volatile Liquids is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

5.4.1 https://chem.libretexts.org/@go/page/41432
5.5: Real Solutions
A non-ideal solution is a solution that does not abide to the rules of an ideal solution where the interactions between the molecules
are identical (or very close) to the interactions between molecules of different components. That is, there is no forces acting
between the components: no Van-der-Waals nor any Coulomb forces. We assume ideal properties for dilute solutions.

5.5: Real Solutions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

5.5.1 https://chem.libretexts.org/@go/page/41433
5.6: Colligative Properties
 Learning Objectives
Express concentrations of solution components using mole fraction and molality
Describe the effect of solute concentration on various solution properties (vapor pressure, boiling point, freezing point, and
osmotic pressure)
Perform calculations using the mathematical equations that describe these various colligative effects
Describe the process of distillation and its practical applications
Explain the process of osmosis and describe how it is applied industrially and in nature

The properties of a solution are different from those of either the pure solute(s) or solvent. Many solution properties are dependent
upon the chemical identity of the solute. Compared to pure water, a solution of hydrogen chloride is more acidic, a solution of
ammonia is more basic, a solution of sodium chloride is more dense, and a solution of sucrose is more viscous. There are a few
solution properties, however, that depend only upon the total concentration of solute species, regardless of their identities. These
colligative properties include vapor pressure lowering, boiling point elevation, freezing point depression, and osmotic pressure.
This small set of properties is of central importance to many natural phenomena and technological applications, as will be
described in this module.

Mole Fraction and Molality


Several units commonly used to express the concentrations of solution components were introduced in an earlier chapter of this
text, each providing certain benefits for use in different applications. For example, molarity (M) is a convenient unit for use in
stoichiometric calculations, since it is defined in terms of the molar amounts of solute species:
mol solute
M = (5.6.1)
L solution

Because solution volumes vary with temperature, molar concentrations will likewise vary. When expressed as molarity, the
concentration of a solution with identical numbers of solute and solvent species will be different at different temperatures, due to
the contraction/expansion of the solution. More appropriate for calculations involving many colligative properties are mole-based
concentration units whose values are not dependent on temperature. Two such units are mole fraction (introduced in the previous
chapter on gases) and molality.
The mole fraction, χ, of a component is the ratio of its molar amount to the total number of moles of all solution components:
mol A
χA = (5.6.2)
total mol of all components

Molality is a concentration unit defined as the ratio of the numbers of moles of solute to the mass of the solvent in kilograms:
mol solute
m = (5.6.3)
kg solvent

Since these units are computed using only masses and molar amounts, they do not vary with temperature and, thus, are better suited
for applications requiring temperature-independent concentrations, including several colligative properties, as will be described in
this chapter module.

 Example 5.6.1: Calculating Mole Fraction and Molality


The antifreeze in most automobile radiators is a mixture of equal volumes of ethylene glycol and water, with minor amounts of
other additives that prevent corrosion. What are the (a) mole fraction and (b) molality of ethylene glycol, C2H4(OH)2, in a
solution prepared from 2.22 × 10 g of ethylene glycol and 2.00 × 10 g of water (approximately 2 L of glycol and 2 L of
3 3

water)?

Solution

5.6.1 https://chem.libretexts.org/@go/page/41434
(a) The mole fraction of ethylene glycol may be computed by first deriving molar amounts of both solution components and
then substituting these amounts into the unit definition.
1 mol C2 H4 (OH)2
mol C2 H4 (OH)2 = 2220 g × = 35.8 mol C2 H4 (OH)2
62.07 g C2 H4 (OH)2

1 mol H2 O
mol H2 O = 2000 g × = 111 mol H2 O
18.02 g H2 O

35.8 mol C2 H4 (OH)2


χethylene glycol = = 0.245
(35.8 + 111) mol total

Notice that mole fraction is a dimensionless property, being the ratio of properties with identical units (moles).
(b) To find molality, we need to know the moles of the solute and the mass of the solvent (in kg).
First, use the given mass of ethylene glycol and its molar mass to find the moles of solute:
mol C2 H2 (OH)2
2220 g C2 H4 (OH)2 ( ) = 35.8 mol C2 H4 (OH)2
62.07 g

Then, convert the mass of the water from grams to kilograms:


1 kg
2000 g H2 O ( ) = 2 kg H2 O
1000 g

Finally, calculate molarity per its definition:


mol solute
molality =
kg solvent

35.8 mol C2 H4 (OH)2


molality =
2 kg H2 O

molality = 17.9 m

 Exercise 5.6.1

What are the mole fraction and molality of a solution that contains 0.850 g of ammonia, NH3, dissolved in 125 g of water?

Answer
7.14 × 10−3; 0.399 m

 Example 5.6.2: Converting Mole Fraction and Molal Concentrations

Calculate the mole fraction of solute and solvent in a 3.0 m solution of sodium chloride.

Solution
Converting from one concentration unit to another is accomplished by first comparing the two unit definitions. In this case,
both units have the same numerator (moles of solute) but different denominators. The provided molal concentration may be
written as:
3.0 mol NaCl

1.0 kg H2 O

The numerator for this solution’s mole fraction is, therefore, 3.0 mol NaCl. The denominator may be computed by deriving the
molar amount of water corresponding to 1.0 kg
1000 g mol H2 O
1.0 kg H2 O ( )( ) = 55 mol H2 O
1 kg 18.02 g

5.6.2 https://chem.libretexts.org/@go/page/41434
and then substituting these molar amounts into the definition for mole fraction.
mol H2 O
XH O =
2
mol NaCl + mol H2 O

55 mol H2 O
XH2 O =
3.0 mol NaCl + 55 mol H2 O

XH O = 0.95
2

mol NaCl
XNaCl =
mol NaCl + mol H2 O

3.0 mol NaCl


XNaCl =
3.0 mol NaCl + 55 mol H2 O

XNaCl = 0.052

 Exercise 5.6.2

The mole fraction of iodine, I , dissolved in dichloromethane, CH


2 2
Cl
2
, is 0.115. What is the molal concentration, m, of iodine
in this solution?

Answer
1.50 m

Vapor Pressure Lowering


As described in the chapter on liquids and solids, the equilibrium vapor pressure of a liquid is the pressure exerted by its gaseous
phase when vaporization and condensation are occurring at equal rates:
liquid⇌ gas (5.6.4)

Dissolving a nonvolatile substance in a volatile liquid results in a lowering of the liquid’s vapor pressure. This phenomenon can be
rationalized by considering the effect of added solute molecules on the liquid's vaporization and condensation processes. To
vaporize, solvent molecules must be present at the surface of the solution. The presence of solute decreases the surface area
available to solvent molecules and thereby reduces the rate of solvent vaporization. Since the rate of condensation is unaffected by
the presence of solute, the net result is that the vaporization-condensation equilibrium is achieved with fewer solvent molecules in
the vapor phase (i.e., at a lower vapor pressure) (Figure 5.6.1). While this kinetic interpretation is useful, it does not account for
several important aspects of the colligative nature of vapor pressure lowering. A more rigorous explanation involves the property of
entropy, a topic of discussion in a later text chapter on thermodynamics. For purposes of understanding the lowering of a liquid's
vapor pressure, it is adequate to note that the greater entropy of a solution in comparison to its separate solvent and solute serves to
effectively stabilize the solvent molecules and hinder their vaporization. A lower vapor pressure results, and a correspondingly
higher boiling point as described in the next section of this module.

5.6.3 https://chem.libretexts.org/@go/page/41434
Figure 5.6.1 : The presence of nonvolatile solutes lowers the vapor pressure of a solution by impeding the evaporation of solvent
molecules.
This figure contains two images. Figure a is labeled “pure water.” It shows a beaker half-filled with liquid. In the liquid, eleven
molecules are evenly dispersed in the liquid each consisting of one central red sphere and two slightly smaller white spheres are
shown. Four molecules near the surface of the liquid have curved arrows drawn from them pointing to the space above the liquid in
the beaker. Above the liquid, twelve molecules are shown, with arrows pointing from three of them into the liquid below. Figure b
is labeled “Aqueous solution.” It is similar to figure a except that eleven blue spheres, slightly larger in size than the molecules, are
dispersed evenly in the liquid. Only four curved arrows appear in this diagram with two from the molecules in the liquid pointing to
the space above and two from molecules in the space above the liquid pointing into the liquid below.
The relationship between the vapor pressures of solution components and the concentrations of those components is described by
Raoult’s law: The partial pressure exerted by any component of an ideal solution is equal to the vapor pressure of the pure
component multiplied by its mole fraction in the solution.

PA = XA P (5.6.5)
A

where PA is the partial pressure exerted by component A in the solution, P is the vapor pressure of pure A, and XA is the mole

A

fraction of A in the solution. (Mole fraction is a concentration unit introduced in the chapter on gases.)
Recalling that the total pressure of a gaseous mixture is equal to the sum of partial pressures for all its components (Dalton’s law of
partial pressures), the total vapor pressure exerted by a solution containing i components is

Psolution = ∑ Pi = ∑ Xi P (5.6.6)
i

i i

A nonvolatile substance is one whose vapor pressure is negligible (P° ≈ 0), and so the vapor pressure above a solution containing
only nonvolatile solutes is due only to the solvent:

Psolution = Xsolvent P (5.6.7)
solvent

 Example 5.6.3: Calculation of a Vapor Pressure

Compute the vapor pressure of an ideal solution containing 92.1 g of glycerin, C3H5(OH)3, and 184.4 g of ethanol, C2H5OH, at
40 °C. The vapor pressure of pure ethanol is 0.178 atm at 40 °C. Glycerin is essentially nonvolatile at this temperature.

Solution
Since the solvent is the only volatile component of this solution, its vapor pressure may be computed per Raoult’s law as:

Psolution = Xsolvent P
solvent

First, calculate the molar amounts of each solution component using the provided mass data.
1 mol C3 H5 (OH)3
92.1 g C3 H5 (OH)3 × = 1.00 mol C3 H5 (OH)3
92.094 g C3 H5 (OH)3

1 mol C2 H5 OH
184.4 g C2 H5 OH × = 4.000 mol C2 H5 OH
46.069 g C2 H5 OH

Next, calculate the mole fraction of the solvent (ethanol) and use Raoult’s law to compute the solution’s vapor pressure.

5.6.4 https://chem.libretexts.org/@go/page/41434
4.000 mol
XC2 H5 OH = = 0.800
(1.00 mol + 4.000 mol)


Psolv = XsolvP = 0.800 × 0.178 atm = 0.142 atm
solv

 Exercise 5.6.3

A solution contains 5.00 g of urea, CO(NH2)2 (a nonvolatile solute) and 0.100 kg of water. If the vapor pressure of pure water
at 25 °C is 23.7 torr, what is the vapor pressure of the solution?

Answer
23.4 torr

Elevation of the Boiling Point of a Solvent


As described in the chapter on liquids and solids, the boiling point of a liquid is the temperature at which its vapor pressure is equal
to ambient atmospheric pressure. Since the vapor pressure of a solution is lowered due to the presence of nonvolatile solutes, it
stands to reason that the solution’s boiling point will subsequently be increased. Compared to pure solvent, a solution, therefore,
will require a higher temperature to achieve any given vapor pressure, including one equivalent to that of the surrounding
atmosphere. The increase in boiling point observed when nonvolatile solute is dissolved in a solvent, ΔT , is called boiling point
b

elevation and is directly proportional to the molal concentration of solute species:

ΔTb = Kb m (5.6.8)

where
Kb is the boiling point elevation constant, or the ebullioscopic constant and
m is the molal concentration (molality) of all solute species.
Boiling point elevation constants are characteristic properties that depend on the identity of the solvent. Values of Kb for several
solvents are listed in Table 5.6.1.
Table 5.6.1 : Boiling Point Elevation and Freezing Point Depression Constants for Several Solvents
Freezing Point (°C at 1
Solvent Boiling Point (°C at 1 atm) Kb (Cm−1) Kf (Cm−1)
atm)

water 100.0 0.512 0.0 1.86

hydrogen acetate 118.1 3.07 16.6 3.9

benzene 80.1 2.53 5.5 5.12

chloroform 61.26 3.63 −63.5 4.68

nitrobenzene 210.9 5.24 5.67 8.1

The extent to which the vapor pressure of a solvent is lowered and the boiling point is elevated depends on the total number of
solute particles present in a given amount of solvent, not on the mass or size or chemical identities of the particles. A 1 m aqueous
solution of sucrose (342 g/mol) and a 1 m aqueous solution of ethylene glycol (62 g/mol) will exhibit the same boiling point
because each solution has one mole of solute particles (molecules) per kilogram of solvent.

 Example 5.6.4: Calculating the Boiling Point of a Solution

What is the boiling point of a 0.33 m solution of a nonvolatile solute in benzene?

Solution
Use the equation relating boiling point elevation to solute molality to solve this problem in two steps.

5.6.5 https://chem.libretexts.org/@go/page/41434
1. Calculate the change in boiling point.
−1
ΔTb = Kb m = 2.53 °C m × 0.33 m = 0.83 °C

Add the boiling point elevation to the pure solvent’s boiling point.
Boiling temperature = 80.1 °C + 0.83 °C = 80.9 °C

 Exercise 5.6.4

What is the boiling point of the antifreeze described in Example 5.6.4?

Answer
109.2 °C

 Example 5.6.5: The Boiling Point of an Iodine Solution

Find the boiling point of a solution of 92.1 g of iodine, I


2
, in 800.0 g of chloroform, CHCl
3
, assuming that the iodine is
nonvolatile and that the solution is ideal.

Solution
We can solve this problem using four steps.

1. Convert from grams to moles of I using the molar mass of I in the unit conversion factor.
2 2

Result: 0.363 mol


Determine the molality of the solution from the number of moles of solute and the mass of solvent, in kilograms.
Result: 0.454 m
Use the direct proportionality between the change in boiling point and molal concentration to determine how much the boiling
point changes.
Result: 1.65 °C
Determine the new boiling point from the boiling point of the pure solvent and the change.
Result: 62.91 °C
Check each result as a self-assessment.

 Exercise 5.6.5: glycerin:Water Solution


What is the boiling point of a solution of 1.0 g of glycerin, C 3
H (OH)
5 3
, in 47.8 g of water? Assume an ideal solution.

Answer

5.6.6 https://chem.libretexts.org/@go/page/41434
100.12 °C

Distillation of Solutions
Distillation is a technique for separating the components of mixtures that is widely applied in both in the laboratory and in
industrial settings. It is used to refine petroleum, to isolate fermentation products, and to purify water. This separation technique
involves the controlled heating of a sample mixture to selectively vaporize, condense, and collect one or more components of
interest. A typical apparatus for laboratory-scale distillations is shown in Figure 5.6.2.

Figure 5.6.2 : A typical laboratory distillation unit is shown in (a) a photograph and (b) a schematic diagram of the components.
(credit a: modification of work by “Rifleman82”/Wikimedia commons; credit b: modification of work by “Slashme”/Wikimedia
Commons)
Figure a contains a photograph of a common laboratory distillation unit. Figure b provides a diagram labeling typical components
of a laboratory distillation unit, including a stirrer/heat plate with heat and stirrer speed control, a heating bath of oil or sand,
stirring means such as boiling chips, a still pot, a still head, a thermometer for boiling point temperature reading, a condenser with a
cool water inlet and outlet, a still receiver with a vacuum or gas inlet, a receiving flask for holding distillate, and a cooling bath.
Oil refineries use large-scale fractional distillation to separate the components of crude oil. The crude oil is heated to high
temperatures at the base of a tall fractionating column, vaporizing many of the components that rise within the column. As
vaporized components reach adequately cool zones during their ascent, they condense and are collected. The collected liquids are
simpler mixtures of hydrocarbons and other petroleum compounds that are of appropriate composition for various applications
(e.g., diesel fuel, kerosene, gasoline), as depicted in Figure 5.6.3.

5.6.7 https://chem.libretexts.org/@go/page/41434
Figure 5.6.3 : Crude oil is a complex mixture that is separated by large-scale fractional distillation to isolate various simpler
mixtures.
This figure contains a photo of a refinery, showing large columnar structures. A diagram of a fractional distillation column used in
separating crude oil is also shown. Near the bottom of the column, an arrow pointing into the column shows a point of entry for
heated crude oil. The column contains several layers at which different components are removed. At the very bottom, residue
materials are removed as indicated by an arrow out of the column. At each successive level, different materials are removed
proceeding from the bottom to the top of the column. The materials are fuel oil, followed by diesel oil, kerosene, naptha, gasoline,
and refinery gas at the very top. To the right of the column diagram, a double sided arrow is shown that is blue at the top and
gradually changes color to red moving downward. The blue top of the arrow is labeled, “small molecules: low boiling point, very
volatile, flows easily, ignites easily.” The red bottom of the arrow is labeled, “large molecules: high boiling point, not very volatile,
does not flow easily, does not ignite easily.”

Depression of the Freezing Point of a Solvent


Solutions freeze at lower temperatures than pure liquids. This phenomenon is exploited in “de-icing” schemes that use salt (Figure
5.6.4), calcium chloride, or urea to melt ice on roads and sidewalks, and in the use of ethylene glycol as an “antifreeze” in

automobile radiators. Seawater freezes at a lower temperature than fresh water, and so the Arctic and Antarctic oceans remain
unfrozen even at temperatures below 0 °C (as do the body fluids of fish and other cold-blooded sea animals that live in these
oceans).

Figure 5.6.4 : Rock salt (NaCl), calcium chloride (CaCl2), or a mixture of the two are used to melt ice. (credit: modification of work
by Eddie Welker)
The decrease in freezing point of a dilute solution compared to that of the pure solvent, ΔTf, is called the freezing point depression
and is directly proportional to the molal concentration of the solute

ΔTf = Kf m (5.6.9)

where
m is the molal concentration of the solute in the solvent and
Kf is called the freezing point depression constant (or cryoscopic constant).
Just as for boiling point elevation constants, these are characteristic properties whose values depend on the chemical identity of the
solvent. Values of Kf for several solvents are listed in Table 5.6.1.

5.6.8 https://chem.libretexts.org/@go/page/41434
 Example 5.6.5: Calculation of the Freezing Point of a Solution
What is the freezing point of the 0.33 m solution of a nonvolatile nonelectrolyte solute in benzene described in Example 5.6.4?

Solution
Use the equation relating freezing point depression to solute molality to solve this problem in two steps.

1. Calculate the change in freezing point.


−1
ΔTf = Kf m = 5.12 °C m × 0.33 m = 1.7 °C

2. Subtract the freezing point change observed from the pure solvent’s freezing point.

Freezing Temperature = 5.5 °C − 1.7 °C = 3.8 °C

 Exercise 5.6.6

What is the freezing point of a 1.85 m solution of a nonvolatile nonelectrolyte solute in nitrobenzene?

Answer
−9.3 °C

 Colligative Properties and De-Icing

Sodium chloride and its group 2 analogs calcium and magnesium chloride are often used to de-ice roadways and sidewalks,
due to the fact that a solution of any one of these salts will have a freezing point lower than 0 °C, the freezing point of pure
water. The group 2 metal salts are frequently mixed with the cheaper and more readily available sodium chloride (“rock salt”)
for use on roads, since they tend to be somewhat less corrosive than the NaCl, and they provide a larger depression of the
freezing point, since they dissociate to yield three particles per formula unit, rather than two particles like the sodium chloride.
Because these ionic compounds tend to hasten the corrosion of metal, they would not be a wise choice to use in antifreeze for
the radiator in your car or to de-ice a plane prior to takeoff. For these applications, covalent compounds, such as ethylene or
propylene glycol, are often used. The glycols used in radiator fluid not only lower the freezing point of the liquid, but they
elevate the boiling point, making the fluid useful in both winter and summer. Heated glycols are often sprayed onto the surface
of airplanes prior to takeoff in inclement weather in the winter to remove ice that has already formed and prevent the formation
of more ice, which would be particularly dangerous if formed on the control surfaces of the aircraft (Video 5.6.1).

5.6.9 https://chem.libretexts.org/@go/page/41434
Aircraft De-Icing - Close Up, Details [HD]

Video 5.6.1 : Freezing point depression is exploited to remove ice from the control surfaces of aircraft.

Phase Diagram for an Aqueous Solution of a Nonelectrolyte


The colligative effects on vapor pressure, boiling point, and freezing point described in the previous section are conveniently
summarized by comparing the phase diagrams for a pure liquid and a solution derived from that liquid. Phase diagrams for water
and an aqueous solution are shown in Figure 5.6.5.

Figure 5.6.5 : These phase diagrams show water (solid curves) and an aqueous solution of nonelectrolyte (dashed curves).
This phase diagram indicates the pressure in atmospheres of water and a solution at various temperatures. The graph shows the
freezing point of water and the freezing point of the solution, with the difference between these two values identified as delta T
subscript f. The graph shows the boiling point of water and the boiling point of the solution, with the difference between these two
values identified as delta T subscript b. Similarly, the difference in the pressure of water and the solution at the boiling point of
water is shown and identified as delta P. This difference in pressure is labeled vapor pressure lowering. The lower level of the vapor
pressure curve for the solution as opposed to that of pure water shows vapor pressure lowering in the solution. Background colors
on the diagram indicate the presence of water and the solution in the solid state to the left, liquid state in the central upper region,
and gas to the right.
The liquid-vapor curve for the solution is located beneath the corresponding curve for the solvent, depicting the vapor pressure
lowering, ΔP, that results from the dissolution of nonvolatile solute. Consequently, at any given pressure, the solution’s boiling
point is observed at a higher temperature than that for the pure solvent, reflecting the boiling point elevation, ΔTb, associated with
the presence of nonvolatile solute. The solid-liquid curve for the solution is displaced left of that for the pure solvent, representing
the freezing point depression, ΔTf, that accompanies solution formation. Finally, notice that the solid-gas curves for the solvent and
its solution are identical. This is the case for many solutions comprising liquid solvents and nonvolatile solutes. Just as for
vaporization, when a solution of this sort is frozen, it is actually just the solvent molecules that undergo the liquid-to-solid
transition, forming pure solid solvent that excludes solute species. The solid and gaseous phases, therefore, are composed solvent
only, and so transitions between these phases are not subject to colligative effects.

5.6.10 https://chem.libretexts.org/@go/page/41434
Osmosis and Osmotic Pressure of Solutions
A number of natural and synthetic materials exhibit selective permeation, meaning that only molecules or ions of a certain size,
shape, polarity, charge, and so forth, are capable of passing through (permeating) the material. Biological cell membranes provide
elegant examples of selective permeation in nature, while dialysis tubing used to remove metabolic wastes from blood is a more
simplistic technological example. Regardless of how they may be fabricated, these materials are generally referred to as
semipermeable membranes.
Consider the apparatus illustrated in Figure 5.6.6, in which samples of pure solvent and a solution are separated by a membrane
that only solvent molecules may permeate. Solvent molecules will diffuse across the membrane in both directions. Since the
concentration of solvent is greater in the pure solvent than the solution, these molecules will diffuse from the solvent side of the
membrane to the solution side at a faster rate than they will in the reverse direction. The result is a net transfer of solvent molecules
from the pure solvent to the solution. Diffusion-driven transfer of solvent molecules through a semipermeable membrane is a
process known as osmosis.

Figure 5.6.6 : Osmosis results in the transfer of solvent molecules from a sample of low (or zero) solute concentration to a sample
of higher solute concentration.
The figure shows two U shaped tubes with a semi permeable membrane placed at the base of the U. In figure a, pure solvent is
present and indicated by small yellow spheres to the left of the membrane. To the right, a solution exists with larger blue spheres
intermingled with some small yellow spheres. At the membrane, arrows pointing from three small yellow spheres on both sides of
the membrane cross over the membrane. An arrow drawn from one of the large blue spheres does not cross the membrane, but
rather is reflected back from the surface of the membrane. The levels of liquid in both sides of the U shaped tube are equal. In
figure b, arrows again point from small yellow spheres across the semipermeable membrane from both sides. This diagram shows
the level of liquid in the left, pure solvent, side to be significantly lower than the liquid level on the right. Dashed lines are drawn
from these two liquid levels into the middle of the U-shaped tube and between them is the term osmotic pressure.
When osmosis is carried out in an apparatus like that shown in Figure 5.6.6, the volume of the solution increases as it becomes
diluted by accumulation of solvent. This causes the level of the solution to rise, increasing its hydrostatic pressure (due to the
weight of the column of solution in the tube) and resulting in a faster transfer of solvent molecules back to the pure solvent side.
When the pressure reaches a value that yields a reverse solvent transfer rate equal to the osmosis rate, bulk transfer of solvent
ceases. This pressure is called the osmotic pressure (Π ) of the solution. The osmotic pressure of a dilute solution is related to its
solute molarity, M, and absolute temperature, T, according to the equation
Π = M RT (5.6.10)

where R is the universal gas constant.

 Example 5.6.7: Calculation of Osmotic Pressure

What is the osmotic pressure (atm) of a 0.30 M solution of glucose in water that is used for intravenous infusion at body
temperature, 37 °C?

5.6.11 https://chem.libretexts.org/@go/page/41434
Solution
We can find the osmotic pressure, \(Π\), using Equation 5.6.10 , where T is on the Kelvin scale (310 K) and the value of R is
expressed in appropriate units (0.08206 L atm/mol K).
Π = M RT

= 0.03 mol/L × 0.08206 L atm/mol K × 310 K

= 7.6 atm

 Exercise 5.6.7

What is the osmotic pressure (atm) a solution with a volume of 0.750 L that contains 5.0 g of methanol, CH3OH, in water at 37
°C?

Answer
5.3 atm

If a solution is placed in an apparatus like the one shown in Figure 5.6.7, applying pressure greater than the osmotic pressure of the
solution reverses the osmosis and pushes solvent molecules from the solution into the pure solvent. This technique of reverse
osmosis is used for large-scale desalination of seawater and on smaller scales to produce high-purity tap water for drinking.

Figure 5.6.7 : Applying a pressure greater than the osmotic pressure of a solution will reverse osmosis. Solvent molecules from the
solution are pushed into the pure solvent.
The figure shows a U shaped tube with a semi permeable membrane placed at the base of the U. Pure solvent is present and
indicated by small yellow spheres to the left of the membrane. To the right, a solution exists with larger blue spheres intermingled
with some small yellow spheres. At the membrane, arrows point from four small yellow spheres to the left of the membrane. On
the right side of the U, there is a disk that is the same width of the tube and appears to block it. The disk is at the same level as the
solution. An arrow points down from the top of the tube to the disk and is labeled, “Pressure greater than Π subscript solution.”
Examples of osmosis are evident in many biological systems because cells are surrounded by semipermeable membranes. Carrots
and celery that have become limp because they have lost water can be made crisp again by placing them in water. Water moves into
the carrot or celery cells by osmosis. A cucumber placed in a concentrated salt solution loses water by osmosis and absorbs some
salt to become a pickle. Osmosis can also affect animal cells. Solute concentrations are particularly important when solutions are
injected into the body. Solutes in body cell fluids and blood serum give these solutions an osmotic pressure of approximately 7.7
atm. Solutions injected into the body must have the same osmotic pressure as blood serum; that is, they should be isotonic with
blood serum. If a less concentrated solution, a hypotonic solution, is injected in sufficient quantity to dilute the blood serum, water
from the diluted serum passes into the blood cells by osmosis, causing the cells to expand and rupture. This process is called

5.6.12 https://chem.libretexts.org/@go/page/41434
hemolysis. When a more concentrated solution, a hypertonic solution, is injected, the cells lose water to the more concentrated
solution, shrivel, and possibly die in a process called crenation (Figure 11.5.8).

Figure 5.6.8 : Red blood cell membranes are water permeable and will (a) swell and possibly rupture in a hypotonic solution; (b)
maintain normal volume and shape in an isotonic solution; and (c) shrivel and possibly die in a hypertonic solution. (credit a/b/c:
modifications of work by “LadyofHats”/Wikimedia commons)

This figure shows three scenarios relating to red blood cell membranes. In a, H subscript 2 O has two arrows drawn from it
pointing into a red disk. Beneath it in a circle are eleven similar disks with a bulging appearance, one of which appears to have
burst with blue liquid erupting from it. In b, the image is similar except that rather than having two arrows pointing into the red
disk, one points in and a second points out toward the H subscript 2 O. In the circle beneath, twelve of the red disks are present. In
c, both arrows are drawn from a red shriveled disk toward the H subscript 2 O. In the circle below, twelve shriveled disks are
shown.

Determination of Molar Masses


Osmotic pressure and changes in freezing point, boiling point, and vapor pressure are directly proportional to the concentration of
solute present. Consequently, we can use a measurement of one of these properties to determine the molar mass of the solute from
the measurements.

 Example 5.6.8: Determining Molar Mass from Freezing Point Depression

A solution of 4.00 g of a nonelectrolyte dissolved in 55.0 g of benzene is found to freeze at 2.32 °C. What is the molar mass of
this compound?

Solution
We can solve this problem using the following steps.

1. Determine the change in freezing point from the observed freezing point and the freezing point of pure benzene (Table
11.5.1).

5.6.13 https://chem.libretexts.org/@go/page/41434
ΔTf = 5.5 °C − 2.32 °C = 3.2 °C

1. Determine the molal concentration from Kf, the freezing point depression constant for benzene (Table 11.5.1), and ΔTf.
\(ΔT_\ce{f}=K_\ce{f}m\)
ΔTf 3.2 °C
m = = = 0.63 m
−1
Kf 5.12 °Cm

1. Determine the number of moles of compound in the solution from the molal concentration and the mass of solvent used to make
the solution.
0.62 mol solute
Moles of solute = × 0.0550 kg solvent = 0.035 mol
1.00 kg solvent

2. Determine the molar mass from the mass of the solute and the number of moles in that mass.
\(\mathrm{Molar\: mass=\dfrac{4.00\:g}{0.034\:mol}=1.2×10^2\:g/mol}\)

 Exercise 5.6.8

A solution of 35.7 g of a nonelectrolyte in 220.0 g of chloroform has a boiling point of 64.5 °C. What is the molar mass of this
compound?

Answer
1.8 × 102 g/mol

 Example 5.6.9: Determination of a Molar Mass from Osmotic Pressure

A 0.500 L sample of an aqueous solution containing 10.0 g of hemoglobin has an osmotic pressure of 5.9 torr at 22 °C. What is
the molar mass of hemoglobin?

Solution
Here is one set of steps that can be used to solve the problem:

5.9 torr × 1 atm


1. Π = = 7.8 × 10
−3
atm
760 torr

Π = M RT

−3
Π 7.8 × 10 atm
−4
M = = = 3.2 × 10 M
RT (0.08206 L atm/mol K)(295 K)

−4
3.2 × 10 mol
1. moles of hemoglobin = × 0.500 L solution = 1.6 × 10
−4
mol
1 L solution

2. Determine the molar mass from the mass of hemoglobin and the number of moles in that mass.
10.0 g
4
molar mass = = 6.2 × 10 g/mol
−4
1.6 × 10 mol

5.6.14 https://chem.libretexts.org/@go/page/41434
 Exercise 5.6.9

What is the molar mass of a protein if a solution of 0.02 g of the protein in 25.0 mL of solution has an osmotic pressure of 0.56
torr at 25 °C?

Answer
2.7 × 104 g/mol

Colligative Properties of Electrolytes


As noted previously in this module, the colligative properties of a solution depend only on the number, not on the kind, of solute
species dissolved. For example, 1 mole of any nonelectrolyte dissolved in 1 kilogram of solvent produces the same lowering of the
freezing point as does 1 mole of any other nonelectrolyte. However, 1 mole of sodium chloride (an electrolyte) forms 2 moles of
ions when dissolved in solution. Each individual ion produces the same effect on the freezing point as a single molecule does.

 Example 5.6.10: The Freezing Point of a Solution of an Electrolyte

The concentration of ions in seawater is approximately the same as that in a solution containing 4.2 g of NaCl dissolved in 125
g of water. Assume that each of the ions in the NaCl solution has the same effect on the freezing point of water as a
nonelectrolyte molecule, and determine the freezing temperature the solution (which is approximately equal to the freezing
temperature of seawater).

Solution
We can solve this problem using the following series of steps.
Convert from grams to moles of NaCl using the molar mass of NaCl in the unit conversion factor. Result: 0.072 mol NaCl
Determine the number of moles of ions present in the solution using the number of moles of ions in 1 mole of NaCl as the
conversion factor (2 mol ions/1 mol NaCl). Result: 0.14 mol ions
Determine the molality of the ions in the solution from the number of moles of ions and the mass of solvent, in kilograms.
Result: 1.1 m
Use the direct proportionality between the change in freezing point and molal concentration to determine how much the
freezing point changes. Result: 2.0 °C
Determine the new freezing point from the freezing point of the pure solvent and the change. Result: −2.0 °C
Check each result as a self-assessment.

 Exercise 5.6.10

Assume that each of the ions in calcium chloride, CaCl2, has the same effect on the freezing point of water as a nonelectrolyte
molecule. Calculate the freezing point of a solution of 0.724 g of CaCl2 in 175 g of water.

Answer
−0.208 °C

Assuming complete dissociation, a 1.0 m aqueous solution of NaCl contains 2.0 mole of ions (1.0 mol Na+ and 1.0 mol Cl−) per
each kilogram of water, and its freezing point depression is expected to be
ΔTf = 2.0 mol ions/kg water × 1.86 °C kg water/mol ion = 3.7 °C. (5.6.11)

When this solution is actually prepared and its freezing point depression measured, however, a value of 3.4 °C is obtained. Similar
discrepancies are observed for other ionic compounds, and the differences between the measured and expected colligative property

5.6.15 https://chem.libretexts.org/@go/page/41434
values typically become more significant as solute concentrations increase. These observations suggest that the ions of sodium
chloride (and other strong electrolytes) are not completely dissociated in solution.
To account for this and avoid the errors accompanying the assumption of total dissociation, an experimentally measured parameter
named in honor of Nobel Prize-winning German chemist Jacobus Henricus van’t Hoff is used. The van’t Hoff factor (i) is defined
as the ratio of solute particles in solution to the number of formula units dissolved:
moles of particles in solution
i = (5.6.12)
moles of formula units dissolved

Values for measured van’t Hoff factors for several solutes, along with predicted values assuming complete dissociation, are shown
in Table 5.6.2.
Table 5.6.2 : Expected and Observed van’t Hoff Factors for Several 0.050 m Aqueous Electrolyte Solutions
Electrolyte Particles in Solution i (Predicted) i (Measured)

HCl H+, Cl− 2 1.9

NaCl Na+, Cl− 2 1.9

MgSO4 Mg2+, SO
2−
4
2 1.3

MgCl2 Mg2+, 2Cl− 3 2.7

FeCl3 Fe3+, 3Cl− 4 3.4

glucose (a non-electrolyte) C12H22O11 1 1.0

In 1923, the chemists Peter Debye and Erich Hückel proposed a theory to explain the apparent incomplete ionization of strong
electrolytes. They suggested that although interionic attraction in an aqueous solution is very greatly reduced by solvation of the
ions and the insulating action of the polar solvent, it is not completely nullified. The residual attractions prevent the ions from
behaving as totally independent particles (Figure 5.6.9). In some cases, a positive and negative ion may actually touch, giving a
solvated unit called an ion pair. Thus, the activity, or the effective concentration, of any particular kind of ion is less than that
indicated by the actual concentration. Ions become more and more widely separated the more dilute the solution, and the residual
interionic attractions become less and less. Thus, in extremely dilute solutions, the effective concentrations of the ions (their
activities) are essentially equal to the actual concentrations. Note that the van’t Hoff factors for the electrolytes in Table 5.6.2 are
for 0.05 m solutions, at which concentration the value of i for NaCl is 1.9, as opposed to an ideal value of 2.

Figure 5.6.9 : Ions become more and more widely separated the more dilute the solution, and the residual interionic attractions
become less.
The diagram shows four purple spheres labeled K superscript plus and four green spheres labeled C l superscript minus dispersed in
H subscript 2 O as shown by clusters of single red spheres with two white spheres attached. Red spheres represent oxygen and
white represent hydrogen. In two locations, the purple and green spheres are touching. In these two locations, the diagram is
labeled ion pair. All red and green spheres are surrounded by the white and red H subscript 2 O clusters. The white spheres are
attracted to the purple spheres and the red spheres are attracted to the green spheres.

5.6.16 https://chem.libretexts.org/@go/page/41434
Summary
Properties of a solution that depend only on the concentration of solute particles are called colligative properties. They include
changes in the vapor pressure, boiling point, and freezing point of the solvent in the solution. The magnitudes of these properties
depend only on the total concentration of solute particles in solution, not on the type of particles. The total concentration of solute
particles in a solution also determines its osmotic pressure. This is the pressure that must be applied to the solution to prevent
diffusion of molecules of pure solvent through a semipermeable membrane into the solution. Ionic compounds may not completely
dissociate in solution due to activity effects, in which case observed colligative effects may be less than predicted.

Key Equations

(PA = XA P )
A

Psolution = ∑i Pi = ∑i Xi P
i

Psolution = Xsolvent P
solvent

ΔTb = Kbm
ΔTf = Kfm
Π = MRT

Footnotes
1. A nonelectrolyte shown for comparison.

Glossary
boiling point elevation
elevation of the boiling point of a liquid by addition of a solute

boiling point elevation constant


the proportionality constant in the equation relating boiling point elevation to solute molality; also known as the ebullioscopic
constant

colligative property
property of a solution that depends only on the concentration of a solute species

crenation
process whereby biological cells become shriveled due to loss of water by osmosis

freezing point depression


lowering of the freezing point of a liquid by addition of a solute

freezing point depression constant


(also, cryoscopic constant) proportionality constant in the equation relating freezing point depression to solute molality

hemolysis
rupture of red blood cells due to the accumulation of excess water by osmosis

hypertonic
of greater osmotic pressure

hypotonic
of less osmotic pressure

ion pair
solvated anion/cation pair held together by moderate electrostatic attraction

isotonic
of equal osmotic pressure

5.6.17 https://chem.libretexts.org/@go/page/41434
molality (m)
a concentration unit defined as the ratio of the numbers of moles of solute to the mass of the solvent in kilograms

osmosis
diffusion of solvent molecules through a semipermeable membrane

osmotic pressure (Π)


opposing pressure required to prevent bulk transfer of solvent molecules through a semipermeable membrane

Raoult’s law
the partial pressure exerted by a solution component is equal to the product of the component’s mole fraction in the solution and
its equilibrium vapor pressure in the pure state

semipermeable membrane
a membrane that selectively permits passage of certain ions or molecules

van’t Hoff factor (i)


the ratio of the number of moles of particles in a solution to the number of moles of formula units dissolved in the solution

5.6: Colligative Properties is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
11.4: Colligative Properties by OpenStax is licensed CC BY 4.0. Original source: https://openstax.org/details/books/chemistry-2e.

5.6.18 https://chem.libretexts.org/@go/page/41434
5.7: Electrolyte Solutions
An electrolyte solution is a solution that generally contains ions, atoms or molecules that have lost or gained electrons, and is
electrically conductive. For this reason they are often called ionic solutions, however there are some cases where the electrolytes
are not ions. For this discussion we will only consider solutions of ions. A basic principle of electrostatics is that opposite charges
attract and like charges repel. It also takes a great deal of force to overcome this electrostatic attraction.

Introduction
The general form of Coulomb's law describes the force of attraction between charges:
q1 m q2
F =k (5.7.1)
2
r

However, we must make some changes to this physics formula to be able to use it for a solution of oppositely charged ions. In
Coulomb's Law, the constant
1
k = (5.7.2)
4πε0

, where ε is the permittivity of free space, such as in a vacuum. However, since we are looking at a solution, we must consider the
0

effect that the medium (the solvent in this case) has on the electrostatic force, which is represented by the dielectric constant ε :
q1 q2
F = (5.7.3)
2
4π ε0 εr

Polar substances such as water have a relatively high dielectric constant.

Standard Definitions of Enthalpy, Entropy, and Gibbs Energy for Ions


Ions are not stable on their own, and thus no ions can ever be studied separately. Particularly in biology, all ions in a certain cell or
tissue have a counterion that balances this charge. Therefore, we cannot measure the enthalpy or entropy of a single ion as we can

¯
¯¯¯
¯
atoms of a pure element. So we define a reference point. The Δf H of a hydrogen ion H
+
is equal to zero, as are the other
thermodynamic quantities.

¯
¯¯¯
¯ +
Δf H [H (aq)] = 0 (5.7.4)


¯
¯¯¯ +
Δf G [ H (aq)] = 0 (5.7.5)


¯¯
¯ +
S [H (aq)] = 0 (5.7.6)

When studying the formation of ionic solutions, the most useful quantity to describe is chemical potential μ , defined as the partial
molar Gibbs energy of the ith component in a substance:

¯
¯¯¯ ∂G

μi = Gi = ( ) =μ + RT ln xi (5.7.7)
i
∂ni
T ,P ,nj

where x can be any unit of concentration of the component: mole fraction, molality, or for gases, the partial pressure divided by
i

the pressure of pure component.

Ionic Solutions
To express the chemical potential of an electrolyte in solution in terms of molality, let us use the example of a dissolved salt such as
magnesium chloride, M gC l . 2

2+ −
M gC l2 ⇌ M g + 2C l (5.7.8)

We can now write a more general equation for a dissociated salt:


z+ z−
Mν + Xν − ⇌ ν+ M + ν− X (5.7.9)

where ν represents the stoichiometric coefficient of the cation or anion and z represents the charge, and M and X are the metal
± ±

and halide, respectively.

5.7.1 https://chem.libretexts.org/@go/page/41435
The total chemical potential for these anion-cation pair would be the sum of their individual potentials multiplied by their
stoichiometric coefficients:
μ = ν+ μ+ + ν− μ− (5.7.10)

The chemical potentials of the individual ions are:



μ+ = μ + RT ln m+ (5.7.11)
+


μ− = μ− + RT ln m− (5.7.12)

And the molalities of the individual ions are related to the original molality of the salt m by their stoichiometric coefficients
m+ = ν+ m (5.7.13)

Substituting Equations 5.7.11 and 5.7.12 into Equation 5.7.10,


∘ ∘ ν+ ν−
μ = (ν+ μ+ + ν− m u− ) + RT ln(m+ m− ) (5.7.14)

since the total number of moles ν = ν+ + ν− , we can define the mean ionic molality as the geometric average of the molarity of
the two ions:
1
ν+ ν−
m± = (m+ m− ) ν
(5.7.15)

then Equation 5.7.14 becomes


∘ ∘
μ = (ν+ μ+ + ν− μ− ) + ν RT ln m± (5.7.16)

We have derived this equation for a ideal solution, but ions in solution exert electrostatic forces on one another to deviate from
ideal behavior, so instead of molarities we must use the activity a to represent how the ion is behaving in solution. Therefore the
mean ionic activity is defined as
1
ν+ ν−
a± = (a+ + a− ) ν
(5.7.17)

where

a± = γ m± (5.7.18)

and γ is the mean ionic activity coefficient, which is dependent on the substance.
±

Substituting the mean ionic activity of \Equation 5.7.18 into Equation 5.7.16,
∘ ∘ ∘ ∘ ν ∘ ∘
μ = (ν+ μ+ + ν− μ− ) + ν RT ln a± = (ν+ μ+ + ν− μ− ) + RT ln a = (ν+ μ+ + ν− μ− ) + RT ln a (5.7.19)
±

when a = a . Equation 5.7.19 then represents the chemical potential of a nonideal electrolyte solutions. To calculate the mean
ν
±

ionic activity coefficient requires the use of the Debye-Hückel limiting law, part of the Debye-Hückel theory of electrolytes .

 Example 5.7.1

Let us now write out the chemical potential in terms of molality of the salt in our first example, M gC l . First from Equation 2

5.7.8, the stoichiometric coefficients of the ions are:

ν+ = 1, ν− = 2, ν =3

The mean ionic molality is


1
1 2
m± = (m+ m− ) 3

= (ν+ m × ν− m ) 3

1
1 2
= m(1 2 ) 3

= 1.6 m

The expression for the chemical potential of

5.7.2 https://chem.libretexts.org/@go/page/41435
M gC l2 (5.7.20)

is

μMgC l =μ + 3RT ln 1.6\mm
2 MgC l2

References
1. Chang, Raymond. Physical Chemistry for the Biosciences. Sausalito, California: University Science Books, 2005.

Contributors and Attributions


Konstantin Malley

5.7: Electrolyte Solutions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
Electrolyte Solutions is licensed CC BY-NC-SA 4.0.

5.7.3 https://chem.libretexts.org/@go/page/41435
5.8: Ionic Activity
The properties of electrolyte solutions can significantly deviate from the laws used to derive chemical potential of solutions. In
nonelectrolyte solutions, the intermolecular forces are mostly comprised of weak Van der Waals interactions, which have a r −7

dependence, and for practical purposes this can be considered ideal. In ionic solutions, however, there are significant electrostatic
interactions between solute-solvent as well as solute-solute molecules. These electrostatic forces are governed by Coulomb's law,
which has a r dependence. Consequently, the behavior of an electrolyte solution deviates considerably from that an ideal
−2

solution. Indeed, this is why we utilize the activity of the individual components and not the concentration to calculate deviations
from ideal behavior. In 1923, Peter Debye and Erich Hückel developed a theory that would allow us to calculate the mean ionic
activity coefficient of the solution, γ , and could explain how the behavior of ions in solution contribute to this constant.
±

Assumptions of Debye-Hückel Theory


The Debye-Hückel theory is based on three assumptions of how ions act in solution:
1. Electrolytes completely dissociate into ions in solution.
2. Solutions of Electrolytes are very dilute, on the order of 0.01 M.
3. Each ion is surrounded by ions of the opposite charge, on average.

Debye and Hückel developed the following equation to calculate the mean ionic activity coefficient γ : ±

6
1.824 × 10 –
log γ± = − ∣ z+ z− ∣ √I (5.8.1)
3/2
(εT )

where
ε is the dielectric constant,
z+ and z are the charges of the cation and anion, respectively, and

I is the ionic strength of the solution.


The Equation 5.8.1 is known as the Debye-Hückel Limiting Law. The ionic strength is calculated by the following relation:
1
2
I = ∑ mi z (5.8.2)
i
2
i

where m and z are the molality and the charge of the ith ion in the electrolyte. Since most of the electrolyte solutions we study
i i

are aqueous (ε = 78.54) and have a temperature of 298 K, the Limiting Law in Equation 5.8.1 reduces to

log γ± = −0.509 ∣ z+ z− ∣ √I (5.8.3)

Example 5.8.1

Calculate ionic strength, mean ionic activity coefficient γ , and the mean ionic molality m for a 0.02 molal aqueous solution
± ±

of zinc chloride, ZnCl .2

Solution
Zinc chloride will dissolve as
2 + −
ZnCl ⟶ Zn (aq) + 2 Cl (aq)
2

The concentrations of the zinc and chloride ions will then be 0.02 and 0.04 molal, respectively. First calculate the mean ionic
molality. The mean ionic molality is defined as the average molality of the two ions (see Electrolyte Solutions):
1
ν+ ν−
m± = (m+ m− ) ν

where ν is the stoichiometric coefficient of the ions, and the total of the coefficients in the exponent. In our case, the mean
ionic molality is

5.8.1 https://chem.libretexts.org/@go/page/41436
1
ν (Zn) ν (C l)
ν( Z n) +ν( C l)
m± = (m m )
Zn Cl

1
1 2
= [(0.02 ) (0.04 ) ] 3

= [(0.02)(0.0016)]3

−2
= 3.17 × 10

To calculate the mean ionic activity coefficient, we first need the ionic strength of the solution from Equation 5.8.2:
1 2 2
I = [(0.02)(+2 ) + (0.04)(−1 ) ]
2

1
= (0.08 + 0.04) = 0.06
2

Now we can use Equation 5.8.3 to calculate the activity coefficient:


−−−

log γ± = −0.509 ∣ (+2)(−1) ∣ √0.06

= (−0.509)(2)(0.245)

= −0.250

−0.25
γ± = 10

= 0.1627

Exercise 5.8.1

Determine the mean ionic activity coefficient and mean activity of a 0.004 molal of Ba(HCO 3
)
2
?

Answer
−0.10188
γ± = 10 ≈ 0.7909

a± = 0.7909 × 0.2785 m = 0.2203 m

Example activity coe cients

Figure 5.8.1 : Example activity coefficients. https://www.youtube.com/watch?v=MZCNooIEzQQ


The kinetic salt effect is the effect of salts preset in solution on the rate of a reaction.

Kinetic Salt Effect


In biological systems, salts influence how well proteins and DNA function. Salts are formed by ionic bonds, between a metal and
an electromagnetic atom(s). Some examples of salts include NaCl, KCl, and Na2SO4. Salt molecules are able to disassociate,
forming cations and anions. An increase in the charge (- or +) of a transition state or an activated complex results in an increase in

5.8.2 https://chem.libretexts.org/@go/page/41436
solvation (creating more order in the system), and causes a decrease in the change of entropy (ΔS). In contrast, a decrease in the
charge of the transition state causes an increase in ΔS.
− + o
I +C ⇌ E → P roduct (5.8.4)

The kinetic salt effect describes the way salts stabilize reactants. For example, in the above reaction, each reactant has a charge.
The negatively-charged reactant is stabilized by the positive charges from the salt, and the positively-charged reactant is stabilized
by the negative charges from the salt. As a result, the rate at which the reactants come together decreases, thus decreasing the rate
at which E forms. Because a charged intermediate is also stabilized in the solution, the half life of the intermediate at equilibrium
increases, shifting the reaction toward product formation. Because the rate of the product formation is higher due to increased
amounts of the intermediate present on the solution, first order kinetics is used to derive the rate constant equation:

log KT S = log KT S o + 2 ZA ZB √I (5.8.5)

where
Z is the charge on the cation and ion from the salt. ZAZB is a product value.
I is the ionic strength. I is also dependent on the solubility of the salt in the reaction mixture. Ionic strength is directly
proportional to the solubility of the salt. Changing the ionic strength manipulates the solvation of the reactants and
intermediates, thus changing ΔS, and affecting the reaction rate.
log K TS
is the rate constant without the salt in the reaction mixture.
o

A is also a constant for the solvent the solution is in. The A value for water is 0.509 at 298 K.
The relationship between ZAZB, I, and the rate of the reaction is presented in tabular form below:

ZAZB Rate of Reaction

+ Increases (salt present in the reaction mixture)

- Decreases (the ionic strength increases)

Note: I=0 at very dilute salt concentrations or if the salt is inert.

References
1. Atkins, P.W. Physical Chemistry. 5th Ed. New York: WH Freeman, 1994.
2. Chang, Raymond. Physical Chemistry for the Biosciences. Sausalito, California: University Science Books, 2005.

Contributors and Attributions


Konstantin Malley (UCD)
Artika Singh (UCD)

5.8: Ionic Activity is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

5.8.3 https://chem.libretexts.org/@go/page/41436
5.9: Colligative Properties of Electrolyte Solutions
An electrolyte solution is a solution that generally contains ions, atoms or molecules that have lost or gained electrons, and is
electrically conductive. For this reason they are often called ionic solutions, however there are some cases where the electrolytes
are not ions. For this discussion we will only consider solutions of ions. A basic principle of electrostatics is that opposite charges
attract and like charges repel. It also takes a great deal of force to overcome this electrostatic attraction.

Introduction
The general form of Coulomb's law describes the force of attraction between charges:
q1 m q2
F =k (5.9.1)
2
r

However, we must make some changes to this physics formula to be able to use it for a solution of oppositely charged ions. In
Coulomb's Law, the constant
1
k = (5.9.2)
4πε0

, where ε is the permittivity of free space, such as in a vacuum. However, since we are looking at a solution, we must consider the
0

effect that the medium (the solvent in this case) has on the electrostatic force, which is represented by the dielectric constant ε :
q1 q2
F = (5.9.3)
2
4π ε0 εr

Polar substances such as water have a relatively high dielectric constant.

Standard Definitions of Enthalpy, Entropy, and Gibbs Energy for Ions


Ions are not stable on their own, and thus no ions can ever be studied separately. Particularly in biology, all ions in a certain cell or
tissue have a counterion that balances this charge. Therefore, we cannot measure the enthalpy or entropy of a single ion as we can

¯
¯¯¯
¯
atoms of a pure element. So we define a reference point. The Δf H of a hydrogen ion H
+
is equal to zero, as are the other
thermodynamic quantities.

¯
¯¯¯
¯ +
Δf H [H (aq)] = 0 (5.9.4)


¯
¯¯¯ +
Δf G [ H (aq)] = 0 (5.9.5)


¯¯
¯ +
S [H (aq)] = 0 (5.9.6)

When studying the formation of ionic solutions, the most useful quantity to describe is chemical potential μ , defined as the partial
molar Gibbs energy of the ith component in a substance:

¯
¯¯¯ ∂G

μi = Gi = ( ) =μ + RT ln xi (5.9.7)
i
∂ni
T ,P ,nj

where x can be any unit of concentration of the component: mole fraction, molality, or for gases, the partial pressure divided by
i

the pressure of pure component.

Ionic Solutions
To express the chemical potential of an electrolyte in solution in terms of molality, let us use the example of a dissolved salt such as
magnesium chloride, M gC l . 2

2+ −
M gC l2 ⇌ M g + 2C l (5.9.8)

We can now write a more general equation for a dissociated salt:


z+ z−
Mν + Xν − ⇌ ν+ M + ν− X (5.9.9)

where ν represents the stoichiometric coefficient of the cation or anion and z represents the charge, and M and X are the metal
± ±

and halide, respectively.

5.9.1 https://chem.libretexts.org/@go/page/41437
The total chemical potential for these anion-cation pair would be the sum of their individual potentials multiplied by their
stoichiometric coefficients:
μ = ν+ μ+ + ν− μ− (5.9.10)

The chemical potentials of the individual ions are:



μ+ = μ + RT ln m+ (5.9.11)
+


μ− = μ− + RT ln m− (5.9.12)

And the molalities of the individual ions are related to the original molality of the salt m by their stoichiometric coefficients
m+ = ν+ m (5.9.13)

Substituting Equations 5.9.11 and 5.9.12 into Equation 5.9.10,


∘ ∘ ν+ ν−
μ = (ν+ μ+ + ν− m u− ) + RT ln(m+ m− ) (5.9.14)

since the total number of moles ν = ν+ + ν− , we can define the mean ionic molality as the geometric average of the molarity of
the two ions:
1
ν+ ν−
m± = (m+ m− ) ν
(5.9.15)

then Equation 5.9.14 becomes


∘ ∘
μ = (ν+ μ+ + ν− μ− ) + ν RT ln m± (5.9.16)

We have derived this equation for a ideal solution, but ions in solution exert electrostatic forces on one another to deviate from
ideal behavior, so instead of molarities we must use the activity a to represent how the ion is behaving in solution. Therefore the
mean ionic activity is defined as
1
ν+ ν−
a± = (a+ + a− ) ν
(5.9.17)

where

a± = γ m± (5.9.18)

and γ is the mean ionic activity coefficient, which is dependent on the substance.
±

Substituting the mean ionic activity of \Equation 5.9.18 into Equation 5.9.16,
∘ ∘ ∘ ∘ ν ∘ ∘
μ = (ν+ μ+ + ν− μ− ) + ν RT ln a± = (ν+ μ+ + ν− μ− ) + RT ln a = (ν+ μ+ + ν− μ− ) + RT ln a (5.9.19)
±

when a = a . Equation 5.9.19 then represents the chemical potential of a nonideal electrolyte solutions. To calculate the mean
ν
±

ionic activity coefficient requires the use of the Debye-Hückel limiting law, part of the Debye-Hückel theory of electrolytes .

 Example 5.9.1

Let us now write out the chemical potential in terms of molality of the salt in our first example, M gC l . First from Equation 2

5.9.8, the stoichiometric coefficients of the ions are:

ν+ = 1, ν− = 2, ν =3

The mean ionic molality is


1
1 2
m± = (m+ m− ) 3

= (ν+ m × ν− m ) 3

1
1 2
= m(1 2 ) 3

= 1.6 m

The expression for the chemical potential of

5.9.2 https://chem.libretexts.org/@go/page/41437
M gC l2 (5.9.20)

is

μMgC l =μ + 3RT ln 1.6\mm
2 MgC l2

References
1. Chang, Raymond. Physical Chemistry for the Biosciences. Sausalito, California: University Science Books, 2005.

Contributors and Attributions


Konstantin Malley

5.9: Colligative Properties of Electrolyte Solutions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.
Electrolyte Solutions is licensed CC BY-NC-SA 4.0.

5.9.3 https://chem.libretexts.org/@go/page/41437
5.10: Biological Membranes
5.10: Biological Membranes is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

5.10.1 https://chem.libretexts.org/@go/page/41438
5.E: Solutions (Exercises)
5.1: Concentration Units

5.5: Real Solutions


Q5.6
Calculate the molality for a 0.005M NH4SO4. What is the difference between the molarity and molality of the solution?

Q5.6
What is the difference between the molarity and the molality of a solution? Calculate the molality for a 0.05M ammonium sulfite
[NH4SO3] solution.

S5.6
The molarity of a solution is the ratio between the number of moles of solute and the solution.
The molality of a solution is the ratio between the number of moles of solute and the mass of solvent:

Q5.7
After taking tylenol for stomach pains, a man has 0.300 mg of acetaminophen in his 5.0 L blood. After 4 hours, there is 0.010 mg left
in his blood. Calculate the number of moles of acetaminophen per ml of blood and the total number of moles and grams before and
after. (Molar Mass of Acetaminophen: 151.2 g/mol)

S5.7
Before:
1mol
−3 −6
n = .300 × 10 g× = 1.98 × 10 mol
151.2g

−6
1.98 × 10 mol 1000ml mol
−4
M = ( ) = 3.96 × 10
5.0L 1L ml

After:
1mol
−3 −8
n = .010 × 10 g× = 6.61 × 10 mol
151.2g

−8
6.61 × 10 mol 1000ml mol
−5
M = ( ) = 1.32 × 10
5.0L 1L ml

5.2: Partial Molar Quantities


Q5.16
If a diver swims to 50 ft below sea level, did the oxygen in his blood increase or decrease? What about a hiker who is 50 ft above sea
level?

Q5.16
Charlie goes on a submarine ride with his family. On their submarine ride, they have dinner which Charlie washes down with a Coke.
Charlie notes that upon the removal of the cap from the Coke bottle, it does not make much of a noise that he has grown accustomed
to hearing when opening a soft drink bottle. Soon after their dinner, Charlie and his family make their way back up to the surface to
disembark from their submarine journey. As they were ascending, Charlie started to burp. What has just occurred?

S5.16
When on a submarine, you submerge into the water and are below sea level. When below sea level, the total and partial pressures of
CO2 increases in comparison to when you are at or above sea level. Since Charlie's family has brought the Coke down from land (at
sea level), the solubility of the CO2 is increased and the concentration of the CO2 in solution is raised. When Charlie returns to sea

5.E.1 https://chem.libretexts.org/@go/page/41331
level, the conditions go back to normal which causes both the pressure and the solubility to decrease and the CO2 inside of Charlie
comes out and causes him to burp in order to release the CO2 gas.

Q5.17
Calculate the Henry's Law constant for O2 in blood at a partial pressure of 0.25. There's about 2.5 kg of blood in the individual and 15
g of O present.
2

S5.17
P02
PO2 = k ⋅ m → k = (5.E.1)
m

.25
k =
mol
.188
kg

1mol
M = (15.0g)( )
32.0g

Q5.18
Imagine the ideal gas F (at 25°C) has a solubility of 123 × 10 mol/L at a partial pressure of 0.5 atm, and
2
−4
321 × 10
−4
at 5 atm.
Use Henry’s law to determine the number of moles of F at the given pressures.
2

5.6: Colligative Properties


Q5.19
Write the equation showing the relationship between the molality and the change of temperature. Show possible reasons that leading
to the derivation of this equation.

S5.19
ΔT = Kb × m2 (5.E.2)

Molality is independent of temperature.

Q5.20
There are two liquids X and Y. The boiling point of liquid X is 125ºC while the boiling point of liquid Y is 250ºC. If liquids X and Y
were mixed to form an ideal solution, would the boiling point of the solution be less than, greater than, or equal to that of (a) liquid X?
(b) liquid Y?

S5.20
a. Boiling point of solution > Boiling point of liquid X
b. Boiling point of solution < Boiling point of liquid Y

Q5.25
Why don't you keep your honey in the fridge? Explain using the terms hypertonic, and hypotonic.

S5.25
Honey is a hypertonic solution and bacteria prefer a hypotonic environment.

Q5.27
The freezing point depression measurement of Compound X in Reagent A yields a molar mass of 144 g; the same measurement in
Reagent Y gives a value of 72 g. What is the reason for this discrepancy? (Hint: Consider solvent-solute interactions).

Q5.28
What would be the volume of 20 moles of sugar (C12H22O11) dissolved in 0.5 g/cm^-3 of water at 10K? Given the freezing point
depression constant for water is 1.86 K/mol kg, what would be the molality of the sugar?

5.E.2 https://chem.libretexts.org/@go/page/41331
Q5.28
You have a boiling pot with 4 L of water. You add 1000 grams of NaCl. Water boils at 100 o C and freezes at 0 o C. To what
temperature is the boiling point of water elevated to with the addition of NaCl? To what temperature is the freezing point depressed
to?

Q5.29a
Blood plasma, an enriched with platelets, contains several growth factors that help healing of tissues and bones. There are a lot of
medical treatments which inject the comparable concentration of a certain solution into human blood. Explain why ?

S5.29a
The injected solution should be comparable to that of blood plasma. There is a certain amount of platelets which keep the state the
blood plasma in equilibrium. If we inject too much, it significantly affects to human blood by changing the concentration of platelets.

Q5.29b
Immerse a human cell in hypertonic or hypotonic solution can cause harmful consequences. Explain what would happen in each
situation and the case when we got injected with a high concentrated solutions in our vein?

Q5.30
The Davis water tower is 250 ft tall (about 76 meters). What is the osmotic pressure required to push water from the ground to the top
of the tower?

S5.30
π = hρg (5.E.3)

−3 −2
π = (76m)(1000kg m )(9.81m s ) (5.E.4)

1atm
5 −2
π = (7.456 ∗ 10 N m )( ) (5.E.5)
5 −2
1.013 ∗ 10 N m

π = 7.36atm (5.E.6)

Q5.31
A solution of liquid A is mixed with liquid B and exerts ideal behavior. At 103 C the total vapor pressure of a solution is 413mmHg
o

containing 1.4 moles of A and 2.0 moles of B. After you add another mole of A to the solution, the vapor pressure increases to
486mmHg. At 103 C calculate the vapor pressure of pure A and B.
o

S5.31
Step1: (Hint: Calculate the mole fraction of A and B)
molesA 1.4molesA
xA = ( ) =( ) = .42 (5.E.7)
(molesA + molesB) (1.4molesA + 2.0molesB)

molesB 2.0molesB
xB = ( ) =( ) = .58 (5.E.8)
(molesA + molesB) (1.4molesA + 2.0molesB)

Step 2: (Hint: Set up total pressure equation)


∗ ∗
P = xA P + xB P (5.E.9)
A B

∗ ∗
413 = (.42)P + (.58)P (5.E.10)
A B

Step 3: (Hint: Calculate the mole fraction of A and B with the addition of 1 mole of A)
molesA 2.4molesA
xA = ( ) =( ) = .55 (5.E.11)
(molesA + molesB) (2.4molesA + 2.0molesB)

molesB 2.0molesB
xB = ( ) =( ) = .45 (5.E.12)
(molesA + molesB) (2.4molesA + 2.0molesB)

Step 4: (Hint: Set up the second total pressure equation)

5.E.3 https://chem.libretexts.org/@go/page/41331
∗ ∗
486 = (.55)P + (.45)P (5.E.13)
A B

Step 5: (Hint: Solve for P and P by substitution of the two total pressure equations)
A

B

∗ ∗
413 = (.42)P + (.58)P (5.E.14)
A B

∗ ∗
486 = (.55)P + (.45)P (5.E.15)
A B

Rearrange equation 1
∗ ∗
P = 983.33 − 1.38095 P (5.E.16)
A B

Substitute into any equation and solve for P B


∗ ∗
486 = (.55)(983.33 − 1.38095 P ) + (.45)P (5.E.17)
B B


P = 177.149mmH g (5.E.18)
B

Substitute P into any equation to get P



B

A

∗ ∗
P = 983.33 − 1.38095 P (5.E.19)
A B

= 738.696mmH g (5.E.20)

Q5.33
An ideal solution is composed of equal amount of component A with molar mass of 28.05g/mol and component B with molar mass of
80.91g/mol. At 35°C, component A has a vapor pressure of 0.2 atm and component B has a vapor pressure of 0.65 atm.
a. Calculate the mole fraction of each component in the solution
b. Calculate the partial pressure of each component at 35°C
c. If some of the vapor condenses back into a liquid, what will the mole fraction of each component in this liquid and the vapor
pressure above this liquid be?

S5.33
(a) remember mole fraction = (mole of solute) / (total mole of solution) ;
weightA
nA = (5.E.21)
28.05g/mol

weightB
nB =
80.91g/mol

weightA

nA 28.05g/mol 1/28.05
xA = = = = 0.7426
nA + nB weightA weightB (1/28.05) + (1/80.91)
+
28.05g/mol 80.91g/mol

weightB

nB 80.91g/mol 1/80.91
xB = = = = 0.2574
nA + nB weightA weightB (1/28.05) + (1/80.91)
+
28.05g/mol 80.91g/mol

(b)P A = xA ∗ P

A
= 0.7426 ∗ 0.2atm = 0.1485atm


PB = xB ∗ P = 0.2574 ∗ 0.62atm = 0.1596atm
B

PA 0.1485atm
(c)xA = = = 0.4820
PA + PB 0.1485atm + 0.1596atm

PB 0.1596atm
xB = = = 0.5180
PA + PB 0.1485atm + 0.1596atm


PA = xA ∗ P = (0.4820)(0.2atm) = 0.0964atm
A


PB = xB ∗ P = (0.5180)(0.62atm) = 0.3212atm
B

5.E.4 https://chem.libretexts.org/@go/page/41331
5.3: The Thermodynamics of Mixing
Q5.35a
Give a brief explanation for the lower vapor pressure of a solvent in presence of a solute from the entropy point of view. (Hint: Look
at colligative property of lower vapor pressure).

S5.35a
The action of mixing a solvent and a solute leads is associated with an increase in entropy. The final solution therefore has a greater
entropy than that of solvent, giving it less driving force for evaporation. This leads to the solvent's decrease in evaporation which
results in a lower vapor pressure.

5.4: Binary Mixtures of Volatile Liquids


Q5.35b
Describe the process of vapor-pressure lowering from the colligative properties of solutions.

S5.35b
Suppose you have an ideal solution to which you can apply Raoult's law:

(Note: This solution has both a solvent and a solute which is nonvolatile).
Next, since we know that , we can ultimately rearrange the equation to be:

In the equation above, is directly proportional to the mole fraction of the solute.

Q5.35c
Sugar is added to water. Would there be more water molecules in the vapor above the sugar solution or in the vapor above the pure
water?

S5.35c
We would expect more water molecules in the vapor above the pure water.

Q5.36
4.0 moles of a compound was added to 500 g of HCl. What was the freezing point of the pure solvent if the freezing point of the
K ⋅ kg
solution is 188.6 K. (Kf=.55 )
mol

S5.36
Tf = 188.6K

4mol mol
m = =8
.500kg kg

K ⋅ kg mol
△ T = −(Kf ⋅ m) → Ti = Tf + [(kf )(m)] = 188.6K + (.55 )(8 = 193K (5.E.22)
mol kg

Q5.36a
After dissolving a substance weighing 0.246 grams in 100.0 g of water, the freezing point of the solution went down 1K below that of
the pure solvent. What would be the molar mass of this substance?

S5.36a
The following information is needed.
△T = Kf m2 (5.E.23)

Kf H2 O = 1.86

Now solve for the molar mass of the solute.

5.E.5 https://chem.libretexts.org/@go/page/41331
1 K = 1.86m2 (5.E.24)

m2 = 0.538

m2 = mol ÷ kg of solvent

−1
mol2 = 0.538 mol kg ⋅ 0.1 kg of water

mol2 = 0.0538 moles of substance

M olar M ass = 0.246 grams of substance ÷ 0.0538 moles of substance

M olar M ass = 4.572 g/mol


–––––––––––––––––––––––––––

Click here for more information on freezing point depression.

Q5.36b
5.525 grams of toluene is mixed with 20.0 grams of acetone. Calculate the freezing point of acetone. The freezing point of the
solution is found to be -102.6o C. The Kf of acetone is 2.40 oC kg/mol, and the molar mass of toluene is 92.14 g/mol

S5.36b
We can relate the molality of the solution to the freezing point depression:
ΔT
m = (5.E.25)
Kf

Rearranged:

ΔT = Kf m (5.E.26)

Tacetone − Tsolution = Kf m (5.E.27)

mols of toluene
Tacetone = Tsolution + Kf (5.E.28)
kg of solvent

1mol
(5.252g × )

C 92.14g

= −102.6 C + (2.40 ) (5.E.29)
−3
kg ⋅ mol 20.0 × 10 kg


= −95.8 C (5.E.30)

Q5.37a
Two aqueous urea solutions have osmotic pressures of 1.4 atm and 3.0 atm, respectively, at a certain temperature. What is the osmotic
pressure of a solution prepared by mixing equal volumes of these two solutions at the same temperature?

Q5.37b
A lab technician has 2 solutions of ammonia. Solution A exerts an osmotic pressure of 8 atm. When it is mixed with solution B in
equal volume to produce solution C, it is determined that solution C exerts an osmotic pressure of 6.5 atm. What is the osmotic
pressure of solution B?

Q5.38
A research scientist used 50 mL of sterilized water to conduct experiments and obtained odd results. She suspected that it had been
contaminated with another compound and determined that the boiling point of the solution had increased to 105oC. How many moles
of the compound (assume it dissolves as a molecule) had been added to the water?

Q5.39
Time-released pills are designed to have longer effect at a constant rate. The active ingredient is concealed by insoluble substances so
that the active ingredient has to be dissolved by getting through the holes of insoluble layers covering outside. Explain how the
insoluble layers regulate the absorption of the active ingredient.

Q5.41
Acetic Acid can make hydrogen bonds with water molecules and also with benzene. A solution of 4.0 g of acetic acid in 95g C6 H6

has a freezing point of 3.5 C . What is the molar mass of the solute?
o

5.E.6 https://chem.libretexts.org/@go/page/41331
S5.41
ΔT = Kf m (5.E.31)

\(m=molality\)
o o −1
3.5 C = (5.12 C kg mol )(m) (5.E.32)

−1
m = .68359mol kg (5.E.33)

moles solvent
m = (5.E.34)
kg solvent

moles solvent = .71757moles (5.E.35)

g acetic acid
mol acetic acid = (5.E.36)
molar mass, acetic acid

4.0g acetic acid


7.1957 moles = (5.E.37)
.555887g mol−1

Q5.43
Determine on whether each of the following statements is true or false and briefly explain your answers. (a) Colligative properties
depends on both the number of solute molecules and the size of the solute molecules. (b) The addition of a nonvolatile solute to a
solvent only changes the chemical potential of liquid (μ (l)). (c) The mole fraction of a component is never the same as the activity
1

of the same component.

S5.43
a. False; the definition of colligative properties are properties that ONLY depend on the number of solute molecules.
b. True; μ (s) and μ (g) are not affected because solute is nonvolatile and insoluble to the solid solvent. (c) False; in the case of
1 1

ideal solutions, the activity coefficient (γ ) is 1 and a = (1) ∗ x .


i i i

Q5.45
Provide an example of a nonideal solution explain why it is nonideal.(Hint: Look over the concept of nonideal solutions)

S5.45
An example of a non-ideal solution is a mixture of diethyl ether and water. Diethyl ether is a non-polar solvent used for dissolving
organic molecules whereas water is polar. Intermolecular forces between these two components are weak; hence they will form two
layers upon mixing; making the mixture a non-ideal solution.

Q5.46a
An unknown substance has a molal boiling-point elevation constant, Kb of 1.23 K/m. Find the molar mass of this substance. The
enthalpy of vaporization for this substance is 38.56 kJ/mol at 78.37 oC.
S5.46a
For any substance:
2
RT M1
0
Kb = (5.E.38)
¯
Δvap H

Rearranging this equation, we get:

Kb Δvap H̄
M1 = (5.E.39)
2
RT
0

K J
3
(1.23 )(38.56 × 10 )
mol mol
= (5.E.40)
J
2
(8.3145 )(78.37 + 276.15 ) K
K ⋅ mol

kg g
= 0.04616 = 46.16 (5.E.41)
mol mol

5.E.7 https://chem.libretexts.org/@go/page/41331
Q5.46b
Find Kb (the molal boiling-point elevation constant) for acetone solution at 56ºC, given that:

¯
¯¯¯
¯ −1
△vap H Acetone = 31.3 kJ mol (5.E.42)

S5.46b
You need to use the following equation:
2
RT Mmolar
b
Kb = (5.E.43)
¯
¯¯¯
¯
△vap H

Plug in values and solve for the Kb.


8.3145 58.08
−1 −1 2 −1
( kJ mol K )(298K ) ( kg mol )
1000 1000
Kb = (5.E.44)
−1
31.3 kJ mol
−1
Kb = 1.37K mol kg (5.E.45)
–––––––––––––––––––––

Click here for more information on boiling point elevation.

Q5.47a
A home gardener and amateur scientist decides to prepare her olives by brining them. She begins the process by soaking the olives in
fresh water. During this time she notices that the olives absorb moisture and swell. She then soaks the olives in a strong salt brine.
During this time, she notices that the olives lose moisture and shrink. To what physical process can she attribute the shrinking and
swelling? Why do the olives absorb the fresh water and lose moisture to the brine?

S5.47a
The movement of water in and out of the olives is due to diffusion. The high concentration of solutes in the brine relative to the
concentration in the olives causes water to diffuse from olives into the brine, and the low concentration of solutes in the fresh water
surrounding the olives causes the fresh water to diffuse into the olives, which have a comparatively high concentration of solutes.

Q5.47b
Describe the difference between a hypertonic, and hypotonic. Why does a cell shrink in a hypertonic solution, and expand in a
hypotonic solution.

Q5.48
The following data give the pressures for water-methanol solution at 39.9oC. Find the activity coefficients of both components based
on:
a. Raoult's law
b. Henry's law

xwater 0 0.0490 0.3120 0.4750 0.6535 0.7904

Pwater/torr 0 4.51 23.03 31.53 39.78 44.81

Pmethanol/torr 255.6 253.39 188.27 152.87 116.22 80.89

Q5.49
A solution is made up of some amount of sucrose (C12H22O11) and 1.03 kg water. Calculate the amount of sucrose (in grams) in
solution if the solution freezes at -2ºC and the activity coefficient of sucrose is 0.789.

S5.49
ΔT = Kb m (5.E.46)

ΔT 2K
m = = (5.E.47)
−1
Kb 1.86 K mol kg

−1
m = 1.075molkg (5.E.48)

5.E.8 https://chem.libretexts.org/@go/page/41331
a is effective molar concentration so a = m in:
a = γx (5.E.49)

x is the actual amount of sucrose in solution, so


a 1.075
x = = (5.E.50)
γ 0.789

−1
x = 1.36mol kg (5.E.51)

1.36mol 342.3g
mass sucrose = ( )(1.03kg H2 O)( ) (5.E.52)
kg H2 O mol

mass sucrose = 479g (5.E.53)

5.7: Electrolyte Solutions

5.8: Ionic Activity


Q5.54
You know that the mean activity coefficient is 0.42. From this information find what the ionic strength of the compound LiNO2 is.

I=0.548

Q5.56
A 0.15 m Ca(NO3)2 solution has a mean ionic activity coefficient of 0.17 at 25°C. Calculate the mean molality, the mean ionic
activity, and the activity of the compound.

Q5.58
In 0.03m aqueous NaNO3 solution at temperature 300K., the size of the ionic atmosphere is 1/k, also known as Debye radius, is
27.8x10-10 m. Using the Debye radius formula to calculate the ionic strength I.

Q5.60
For a 0.0030m aqueous solution of C aC l at 298K, calculate the ionic strength. (a) Calculate the activity coefficients of C a and
2
2+

Cl ions in the solution and (b) calculate the mean ionic activity coefficients of these ions. (Hint: Use the Debye-Huckel limiting

law)

S5.60
1
2
I = Σi mi z (5.E.54)
i
2

1
2 2
= [(0.0030m)(2 ) + (0.0060m)(−1 ) ] (5.E.55)
2

= −.009m (5.E.56)

(a)
2

logγi = −0.509 z √I (5.E.57)
i

For C a , 2+

2 −−−−
logγ+ = −0.5092 √0.009 (5.E.58)

logγ+ = −0.193 (5.E.59)

γ+ = .640985 (5.E.60)

5.E.9 https://chem.libretexts.org/@go/page/41331
For C l ,

2 −−−

logγ− = −0.509(−1 ) √.009 (5.E.61)

logγ− = −0.048288 (5.E.62)

γ− = .89477 (5.E.63)

(b)
1 1

v− v− 1 2
γ± = (γ γ ) v = [(.640985 ) (.89477 ) ] 3 = .8006 (5.E.64)
+ −

Q5.65
For an ideal mixture, what mole fraction of each gas maximizes the entropy of mixing:
Δmix S = −nR(x1 lnx1 + x2 lnx2 ) (5.E.65)

Hint: Rewrite the above function in terms of only either x1 or x2, then derive.

S5.65
If we take the partial derivative of the entropy of mixing:

Δmix S = −nR(x1 lnx1 + x2 lnx2 ) (5.E.66)

δΔmix S δ
[ ] = [− nR(x1 lnx1 + x2 lnx2 )] (5.E.67)
δx1 δx1
x
,n,R x2 ,n,R
2

If we rewrite the function in terms of x1


δ
= [− nR(x1 lnx1 + (1 − x1 )ln(1 − x1 )] (5.E.68)
δx1
x2 ,n,R

and derive using the chain rule:


1 1
= − ( x1 + lnx1 + (1 − x1 ) (− ) + ln(1 − x1 )(−1)) (5.E.69)
x1 1 − x1

= 1 − lnx1 + 1 + ln(1 − x1 ) (5.E.70)

= −lnx1 + ln(1 − x1 ) (5.E.71)

x1
= −ln (5.E.72)
1 − x1

To find the local maximum, we set this to 0 and solve for x1


x1
0 = −ln (5.E.73)
1 − x1

x1
0
e = (5.E.74)
1 − x1

x1
1 = (5.E.75)
1 − x1

1 − x1 = x1 (5.E.76)

x1 = 0.5 (5.E.77)

To check that this is a maximum and not a minimum, we plug in values to the left and right of the given x1 value (ex. 0.4 and 0.6). We
then find that at x1<0.5 the derivative, or slope, is positive and at x1>0.5m the slope is negative. This is further confirmed by the
following diagram, where is maximized at mole fraction 0.5.

5.E.10 https://chem.libretexts.org/@go/page/41331
Q5.64
The osmotic pressure arbitrary non-ideal diluted solution is measured at 2 different concentrations at 300K. At concentration of 5.20
g/L, osmotic pressure is 10.20 x 10-3 atm and at concentration of 8.90 g/L the osmotic pressure is 2.30 x 10-2 atm. Use this
information to estimate the molar mass of the solute. (Hint: For dilute solutions, only the second virial coefficient is concerned. Look
at equation for osmotic pressure and make a graph of π/c vs. c and look for the y-intercept of graph.)

S5.64

π (atm) 1.06 x 10-5 2.30 x 10-5

c (g/L) 5.20 8.90

π/c ( atm•L/g) 2.038 x 10-6 2.584 x 10-6

slope of the graph of π/c vs. c is generated as:


atm ⋅ L
−6 −6
(2.584 × 10 − 2.038 × 10 )
2
g atm ⋅ L
−7
slope = = 1.476 × 10 (5.E.78)
g 2
g
(8.90 − 5.20)
L

π RT RT RT B
= (1 + Bc) = +( )c (5.E.79)
c M M M

2
RBT atm ⋅ L
−7
slope = = 1.476 × 10 (5.E.80)
2
M g

Plug back into equation to solve for "y-intercept"


RT
−6 −7
2.584 × 10 = + (1.476 × 10 )(8.90) (5.E.81)
M

RT atm ⋅ L
−6
⇒ = 1.271 × 10 (5.E.82)
M g

−6
atm ⋅ L L⋅ −6
atm ⋅ L −5
g
M = RT (1.271 × 10 ) = (0.08206 )(300K)(1.271 × 10 ) = 3.13 × 10 (5.E.83)
g K ⋅ mol g mol

Q5.65
Assume that you have made an ideal solution of an unknown compound made of two components. Why does the mole fraction 0.5
achieve the maximum Δ S? mix

S5.65
Use the equation.
Δmix S = −nR(x1 ln x1 + x2 ln x2 ) (5.E.84)

Find the partial derivative since we want to find out the max value for this equation for x1 . We find x1 since it will show us x2.

5.E.11 https://chem.libretexts.org/@go/page/41331
∂ ∂
[ Δmix S ]x ,n,R =[ − nR (x1 ln x1 + x2 ln x2 )]x ,n,R (5.E.85)
2 2
∂x1 ∂x1

since x1 + x2 = 1 (5.E.86)


= −[ nR (x1 ln x1 + (1 − x1 ) ln(1 − x1 ))]x2 ,n,R (5.E.87)
∂x1

derive (5.E.88)

1 −1
= − [(x1 ⋅ ) + (ln x1 ) + ((1 − x1 ) ⋅ ) + ((−1) ⋅ ln(1 − x1 ))] (5.E.89)
x1 1 − x1

= −1 − ln x1 + 1 + ln(1 − x1 ) (5.E.90)

= ln(1 − x1 ) − ln x1 (5.E.91)

1 − x1
= ln (5.E.92)
x1

Now that we have the partial derivative, we set it to 0 and solve to attempt to find the maximum value.
1 − x1
0 = ln (5.E.93)
x1

1 − x1
ln
0 x1
e =e (5.E.94)

1 − x1
1 = (5.E.95)
x1

x1 = 1 − x1 (5.E.96)

2 x1 = 1 (5.E.97)

x1 = 0.5 (5.E.98)
––––––––

Now we know that x1 = 0.5 is either a min or max. If we plug in a number smaller then 0.5 and bigger than 0.5 into the partial
derivative we found earlier, we can conclude that the slope of the equation is positive to the left of 0.5 and negative to the right of the
equation. This is characteristic of maximum values. Thus x1 = 0.5 and x2 = 0.5 attain the maximum values of ΔmixS since
x1 = 0.5 is a max.
See here for more details about ΔmixS. See figure ( 11) on that page for more details about the max.

Q5.66
Producing a certain gas mixture involves mixing high-pressure nitrogen and low-pressure hydrogen together. In one such production,
3.22 moles of N2 at 4.2 atm are mixed with 1.25 moles of H2 at 0.8 atm. Determine the change in Gibbs’ free energy.

5.9: Colligative Properties of Electrolyte Solutions


Q5.72
You are trying to produce a passive reverse osmosis system to obtain pure water from the ocean by using a semi-permeable membrane
on one end of the intake pipe. At what depth would the bottom of the pipe need to be in order for pure water to begin to fill the pipe?
Assume the ocean temperature is 17ºC, and the seawater is a 0.70 M NaCl solution. The density of seawater is 1.03 g/cm3.

5.10: Biological Membranes

5.E: Solutions (Exercises) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

5.E.12 https://chem.libretexts.org/@go/page/41331
CHAPTER OVERVIEW

6: Chemical Equilibrium
6.1: Chemical Equilibrium in Gaseous Systems
6.2: Reactions in Solutions
6.3: Heterogeneous Equilibria
6.4: The Influence of Temperature, Pressure, and Catalysts on the Equilibrium Constant
6.5: Binding of Ligands and Metal Ions to Macromolecules
6.6: Bioenergetics
6.E: Chemical Equilibrium (Exercises)

6: Chemical Equilibrium is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1
6.1: Chemical Equilibrium in Gaseous Systems
6.1: Chemical Equilibrium in Gaseous Systems is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.

6.1.1 https://chem.libretexts.org/@go/page/41439
6.2: Reactions in Solutions
6.2: Reactions in Solutions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

6.2.1 https://chem.libretexts.org/@go/page/41440
6.3: Heterogeneous Equilibria
6.3: Heterogeneous Equilibria is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

6.3.1 https://chem.libretexts.org/@go/page/41441
6.4: The Influence of Temperature, Pressure, and Catalysts on the Equilibrium
Constant
6.4: The Influence of Temperature, Pressure, and Catalysts on the Equilibrium Constant is shared under a CC BY-NC-SA 4.0 license and was
authored, remixed, and/or curated by LibreTexts.

6.4.1 https://chem.libretexts.org/@go/page/41442
6.5: Binding of Ligands and Metal Ions to Macromolecules
6.5: Binding of Ligands and Metal Ions to Macromolecules is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by LibreTexts.

6.5.1 https://chem.libretexts.org/@go/page/41443
6.6: Bioenergetics
6.6: Bioenergetics is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

6.6.1 https://chem.libretexts.org/@go/page/41444
6.E: Chemical Equilibrium (Exercises)
Q6.1
Imagine a catalytic reaction that changes a reagent between two states, i.e.:
yY + cC ⇌ cC + zZ (6.E.1)

The equilibrium constants K , K , and K with respect to concentration, mole fraction, and pressure are given by
c χ p

z
[Z]
Kc = (6.E.2)
[Y ]y

z
χ
Z
Kχ = y
(6.E.3)
χ
Y

z
P
Z
Kp = (6.E.4)
y
P
Y

Given this information, express K in terms of K and K .


p c χ

Q6.2a
C aC O3(s) ⇌ C aO(s) + C O2(s) (6.E.5)

a. The partial pressure of CO2 is 0.1 atm, calculate K p

b. If the reaction occurred in a 1L flask, what is the amount of CaO formed when 0.12 moles CaCO3 was set to react?
c. What is the minimum about of CaO needed to cause the formation of CaCO3?

Q6.2b
The decomposition of sodium chlorate is:
2N aC lO3(s) ⇌ 2N aC l(s) + 3 O2(g) (6.E.6)

Suppose 0.760 mole of sodium chlorate was to be decomposed in a 3.75 L container. At the temperature 336°C, the percent
decomposition of NaClO3 is 3.5%. Find the
a. pressure of O gas in the container in atm and
2

b. equilibrium constant K of the reaction. Assume O gas is an ideal gas.


p 2

Hint: Use partial pressure and look at chemical equilibrium of gaseous systems

S6.2b
a) number of moles of NaClO3 decomposed is:
#moles decomposed
= 0.035 ⇒ #mole decomposed = 0.0266mole N aC lO3 (6.E.7)
.760moles

number of moles of O2 formed is:


3molO3
0.0266mol N aC lO3 × = 0.0399mol O2 (6.E.8)
2mol N aC lO3

nRT
PO V = nRT ⇒ PO = (6.E.9)
2 2
V

L⋅K
(0.0399mol)(0.08206 )(336 + 273)K
K ⋅ mol
PO2 = = 0.532atm (6.E.10)
3.75L

b)

6.E.1 https://chem.libretexts.org/@go/page/41332
1.013bar
0.532atm ×
PO2 1atm
Kp = = = 0.539 (6.E.11)

P 1bar

Q6.2c
Using the decomposition reaction of hydrogen peroxide (H2O2), assume that at 1048 °C and the pressure of the oxygen gas (O2) is
1.5 bar.
2 H O (aq) ⇌ 2 H O(l) + O (g) (6.E.12)
2 2 2 2

a. Determine the K for the reaction above.


P

b. If 0.92 mole of H O is placed in a 3.0-L beaker at 1048°C, determine the fraction of H O that will decompose.
2 2 2 2

c. Determine the same as (b) if 1.3 moles of H O were placed inside the beaker instead.
2 2

d. What is the minimum amount of H O (in moles) necessary in order to reach equilibrium?
2 2

S6.2c

a.
b. In order to calculate the number of moles of H2O2 that are decomposed, you first need to determine the moles of O2 formed by
the reaction. Treat O2 as an ideal gas.

c. If 1.3 moles of H2O2 were used instead of 0.92 mole, the pressure of the O2 would not be affected and would remain at 1.5 bar.
The number of moles of H2O2 decomposed would still be 0.0802 mole, therefore, the fraction of H2O2 decomposed would be as
follows:

d. The pressure of O2 must be greater than or equal to 1.5 bar in order for equilibrium to take place, therefore, the number of moles
of H2O2 cannot be less than 0.0802 mol.

Q6.3
Find the value of Kp for cellular respiration if P O2 = 350torr , with a 3:2 ratio to the pressure of CO2

S6.3
C6 H12 O6 (s) + O2 (g) ⇌ C O2 (g) + H2 O(l) (6.E.13)

2
PC O2 = (350torr) = 233.3torr (6.E.14)
3

233.3torr

750torr
Kp = = 0.666 (6.E.15)
350torr

750torr

Q6.5
For the reaction of
N O → 2 NO (6.E.16)
2 4 2

6.E.2 https://chem.libretexts.org/@go/page/41332
with Kp = 0.167 at 300 K. If 1.0 g of N2O4 is placed into a 250.0 mL container:
a. What would be its pressure if none dissociates?
b. What is its partial pressure at equilibrium (with dissociation)?
c. What is the total pressure in the container?

Q6.9a
Calculate Δ vap H for the evaporation of methanol when the temperature is raised from 20 °C to 100 °C and if the K 2 / K1 ratio is
22.14?
C H3 OH(l) → C H3 OH(g) (6.E.17)

Q6.9b
Using the decomposition reaction of coppr (II) oxide below:
4C u O(s) ⇌ 2C u2 O(s) + O2(g) (6.E.18)

determine the standard enthalpy of the reaction. (Note: The equilibrium vapor pressures of O2 are 15.4 mmHg at 600°C and 927
mmHg at 850°C).

S6.9b
In this case, we must use the van't Hoff equation:
o
K2 Δ 1 1
H
ln = ( − ) (6.E.19)
K1 R T1 T2

in order to solve this particular problem. We can conclude that is proportional to due to the defined relationship of:

.
Therefore,

Q6.10
A container of water at 20ºC was placed in a freezer that was at a temperature of -5.0ºC. The vapor pressure of water in the
container went from 0.60 bar to 0.38 bar. Calculate the enthalpy of fusion for the reaction that took place.

S6.10
k2 = 0.38 bar

k1 = 0.60 bar

T2 = 268 K

T1 = 293 K


k2 Δr H T2 − T1
ln = (6.E.20)
k1 R T2 T1


k2 T2 T1
→ Δr H = ln (R) (6.E.21)
k1 T2 − T1

0.38bar J (293 K)(268 K) 1 kJ kJ


ln 8.314 ( ) = 11.9 (6.E.22)
0.60bar mol ⋅ K 268 K − 293 K 1000 J mol

6.E.3 https://chem.libretexts.org/@go/page/41332
Q6.11
The chemical responsible for the brown air throughout the Los Angeles area is NO2(g). To learn more about NO2(g), you decide to
study this pollutant spectroscopically (by light absorption). You fill a gas cell with N2O4, equilibrate the temperature to 298.0 K
and then open the stopcock on the cell to equilibrate the pressure to the barometric pressure (723.4 mm Hg) that day. You then
reequilibrate the cell at 323.1 K, 348.0 K, and 372.9 K. The following data is obtained:

T/K 298.0 323.1 348.0 372.9

[NO2]/M 0.01262 0.02140 0.02756 0.02920

Find KP, Δ rxn G, Δ rxn H, and Δ rxn S at each temperature for the reaction

N O (g) ⇌ 2 NO (g). (6.E.23)


2 4 2

Q6.14
What is the equilibrium constant and standard Gibbs energy change for:
C O(g) ⇌ C O(g) + ½O2(g) (6.E.24)

Assume that when C O(g) dissociates into C O(g) and O2(g) at 300 K and 1.5 atm, the overall dissociation is one two fifths
complete.

Q6.15
The standard Gibbs energies of formation of proban-1-ol and proban-2-ol are -171.3 kJ mol-1 and -180.3 kJ mol-1, respectively.
Find the ratio of equilibrium vapor pressures of each isomer at 300 K.

S6.15
The ratio of equilibrium pressure is : 171.3/180.3 = 0.95

Q6.16
At what temperature does a particular reaction favor the formation of products at equilibrium if ΔrHº = 215.7 kJ mol-1 and ΔrSº =
348.8 J mol-1 K-1?

S6.16
The reaction favors the formation of products at equilibrium when
∘ ∘ ∘
Δr G = Δr H − T Δr S <0 (6.E.25)

∘ ∘ 3 −1 −1 −1
Δr H − T Δr S = 215.7 ∗ 10 J mol − T (348.8J mol K ) <0 (6.E.26)

3 −1
215.7 ∗ 10 J mol
T > (6.E.27)
−1 −1
348.8J mol K

T > 616K (6.E.28)

Q6.17
Using a table of thermodynamic data, find Ksp at 298.15K for:
C O(g) → C O(g) + ½O2(g) (6.E.29)

(Hint, this problem can be solved using the van’t Hoff equation in its integrated form)

Q6.18
Consider the the Haber process:
N2(g) + 3 H2(g) ⇌ 2N H3(g) (6.E.30)

6.E.4 https://chem.libretexts.org/@go/page/41332
How does the equilibrium shift under the following changes
a. the volume of nitrogen gas is increased
b. the temperature is decreased
c. the pressure of the system is decreased

S6.18
a. The reaction shifts to the right : more products formed.
b. The reaction shifts to the right : favor exothermic reaction
c. The reaction shifts to the left : more pressure applied.

Q6.19
Given the general reaction A(g) ⇔2B(g), calculate the degree of dissociation of A at 25ºC and 7.00 bar if ΔrGº = 6.76 kJ mol-1.
According to Le Chatelier's Principal, in what direction should this reaction proceed?

S6.19

Δr G = −RT lnKP (6.E.31)

3 −1
∘ 6.76 ∗ 10 J mol
Δr G
−1 −1
−(8.314J mol K )298K
KP = e −RT =e (6.E.32)

KP = 0.0653 (6.E.33)

2

KP = P (6.E.34)
2
1 −α

2
4α 0.0653
= (6.E.35)
2
1 −α 7

α = .048 (6.E.36)

Degree of dissociation α = 0.048, the reaction proceeds 4.8%. Le Chatelier's Principal says that reactions will move towards the
side with fewer moles of gas at high pressures.

Q6.19
How does Le Chatelier's Principle relate to the following equation?

S6.19
Le Chatelier's Principle basically states that a system will adjust itself in efforts to re-establish equilibrium when an outside stress is
placed on it. Using the equation provided, we can conclude that raising the temperature causes the equilibrium to shift from left to
right in an endothermic reaction. This suggests that it is favoring the formation of products. We can also conclude that the opposite
is true. This conclusion reinforces Le Chatelier's Principle because the temperature acts as the external stress placed on the system
in this case.

Q6.20
H2 and H molecules are at equilibrium pressures of .35 bar and .30 bar, respectively. If the size of the container they are in is
reduced by a factor of two, what will be the new partial pressures?

S6.20
H2 (g) ⇌ 2H (g) (6.E.37)

2 2
P .30
H
Kp = = = .675 (6.E.38)
PH2 .35

6.E.5 https://chem.libretexts.org/@go/page/41332
At new volume:
PH = .60 (6.E.39)

PH = .70 (6.E.40)
2

Pressures will increase with decreased volume. Less molecules of gas will be favored.

PH PH
2

0.60 bar - 2x 0.70 bar + x

2
.60 − (2 x )
2
.675 = → 4x + .675x + .113 → x = .272 (6.E.41)
.70 + x

PH = .70bar + .272bar = .972bar (6.E.42)


2

PH = .60bar − (2).272bar = .056bar (6.E.43)

Q6.21

With Le Chatlier's principle in mind, does raising the temperature favor the forward reaction or the reverse?

S6.21
This reaction favors the reverse reaction.

Q6.23a
When a gas was heated at atmospheric pressure and 25°C, its color deepened. Heating above 150°C caused the color to fade, and at
550°C the color was barely detectable. At 550°C, however, the color was partially restored by increasing the pressure of the
system. Which of the following scenarios best fits the above description? Justify your answer.
a. A mixture of Gas A and Gas B
b. Pure Gas B
c. A mixture of Gas X and Gas Y
(Hint: Gas B is bluish, and Gas Y is yellow. The other gases are colorless. Gas A and B are in their natural state, Gas AB has a
ΔfH° = -43.2 kJ / mol. The reaction of Gas X into Gas Y is endothermic.)

Q6.23b
Hydrogen gas and iodine react at equilibrium in a glass canister to form Hydrogen Iodide, a strong acid:

H2(g) + I2(g) ⇌ H I(g) (6.E.44)

Iodine gas is a deep purple color. Both Hydrogen Iodide and hydrogen gas are colorless. Assume that iodine sublimates readily at
37o C, and that the reaction is endothermic in the direction written. What color is the gas mixture in the canister in the following
scenarios? Provide an explanation for each.
a. The canister is heated to 40o C
b. The canister is heated to 500o C
c. Negative pressure is applied to the cannister at 500o C
Hint: At room temperature, the gas is colorless. Solid iodine must be in gas phase to react.

S6.23b
a.) The color of the gas will be a deep purple. The iodine will have sublimated, but the temperature is not high enough to drive the
reaction forward.
b.) The gas will be colorless, or almost completely so. This is because an increase in the temperature in an endothermic reaction
drives the equilibrium constant higher. This can be justified using Le Chatelier's principle, which states that added stress to an

6.E.6 https://chem.libretexts.org/@go/page/41332
equilibrium will be offset by the system.
H2(g) + I2(g) + heat ⇌ H I(g) (6.E.45)

Thus, when heat is added, the system will compensate by driving the reactants of the reaction forward into products. Another way
to assess this is by the altered form of the van't Hoff equation:
∘ ∘
Δr H Δr S
lnK = − + (6.E.46)
RT R

Assuming that neither change in entropy nor enthalpy changes due to change in temperature, and that in an endothermic reaction
the enthalpy is positive, as the temperature increases, so does K.
c.) The gas will become slightly purple. This is because a decrease in pressure will alter K and drive the reaction in the direction
which produces more moles of gas. By this, Le Chatelier's principle once again holds.

Q6.24
2+ 2+
Mg + P b → Mg +Pb (6.E.47)

Write the two separate reactions happening here. Label which one is happening at the cathode end and which is happening at the
anode end.

S6.24
Cathode:
2+ −
Pb + 2e → Pb (6.E.48)

Anode:
2+ −
Mg → Mg + 2e (6.E.49)

Q6.24
The reaction of N2 (g) and H2 (g) gas produces NH3 (g). This reaction is exothermic explain what happens when you increase the
temperature of the reaction. What happens when you increase the pressure?

S6.24
Since this reaction is exothermic heat is produced. Using Le Chatelier's principle we see that an increase in the temperature of the
reaction will drive the reaction backward since heat is already in the product side. If we increase pressure we decrease the volume
of the reaction therefore by Le Chatelier's principle an increase in the pressure should drive the reaction forward.

Q6.25
At places such as high mountain, the air pressure is lower than 1 atm, resulting lower partial pressure of Oxygen. What would you
expect for the concentration of hemoglobin for the people living at such places?

Q6.26
Show the steps to get from
[P L]
Y = (6.E.50)
[L] + [P ]

to
1 Kd
=1+ (6.E.51)
Y L

Q6.27
After an experiment of protein-binding you find the data respectively

6.E.7 https://chem.libretexts.org/@go/page/41332
2+
Mg
Total : 60 120 180 240 480
μM

2+
Mg boundtoprotein
: 33.8 120 180 240 480
μM

Determine the dissociation constant of C a 2+


graphically. The protein concentration was kept at 96 μM for each run.
Solution:

54.2μM pH = pKa (6.E.52)

6
1 × 10 M −5
54.2μM ( ) = 5.42 × 10 (6.E.53)
1μM

−5.42E−5
Ka = 10 (6.E.54)

Ka = .9998 (6.E.55)

Q6.28
The dissociation constant for the following reaction is 3.2x10-4. Dissolve 0.02M of C9H8O4 in water, calculate the molarity of
reactants at equilibrium.
C9H8O4 + H2O <---> H3O+ + C9H7O4-

Q6.29
The reaction
+ 2− +
Glyceraldehyde 3 − phosphate + N AD + HP O → 1, 3 − Biphosphoglycerate + N ADH + H (6.E.56)
4

∘ −1
△r G = 6.3 kJ mol (6.E.57)

is catalyzed by GAPDH (Glyceraldehyde 3-phosphate Dehydrogenase). At 298 K, predict whether or not the reaction will be
spontaneous with the following information:

[G3P] = 1.5x10-5 M; [BPG] = 3.0x10-3 M


[NAD+] = 1.2x10-5 M; [NADH]=1.0x10-4 M
[HPO42-]= 1.2x10-5 M; pH = 7.5

S6.29
Use the following equation.

Calculate [H+]

6.E.8 https://chem.libretexts.org/@go/page/41332
+ −7.5 −8
[H ] = 10 = 3.162X 10 M (6.E.58)

+
since pH = −log([ H ])

Plug values into equation and solve. Note: Since there is no coefficient in front of H+ in the reaction, x=1.
8.3145
−1 −1 −1
△r G = 6.3 kJmol +( kJ/mol K )(298K) ln (6.E.59)
1000
−3 −8 −7 1 −4
(3 ⋅ 10 M )(3.162 ⋅ 10 M /(1 ⋅ 10 )] (1.0 ⋅ 10 M)

−5 −5 −5
(1.5 ⋅ 10 M ) (1.2 ⋅ 10 M ) (1.2 ⋅ 10 M)

−1
△r G = 49.90 kJ mol (6.E.60)
–––––––––––––––––––––––

Click here for some more information on bioenergetics and free energy.

Q6.29
The reaction:
Glucose + AT P ⇌ Glucose6 − phosphate + ADP (6.E.61)

At 298K, the equilibrium constant for the reaction is 3.7 × 10


−3
.Will the reaction occur spontaneously if the reaction is at the
following concentrations:
[Glucose]=3.2 × 10 −4
M , [ATP]=2.5 × 10 −3
M , [G-6-P]=1.2 × 10 −5
M , [ADP]=1.0 × 10 −5
M .

S6.29
∘ −3
ΔG = −RT ln(Keq ) = −(8.314J/K ∗ mol)(298K)ln(3.7 × 10 )

= 13872.97J/mol = 13.87297kJ/mol

[P roduct]

ΔG = ΔG + RT ln( )
[Reactant]

−5 −5
[1.2 × 10 ][1.0 × 10 ]
−3
= 13.87297kJ/mol + (8.314 × 10 kJ/K ∗ mol)(298K)ln( )
−4 −3
[3.2 × 10 ][2.5 × 10 ]

= −7.942kJ/mol

ΔG is negative, so the reaction is spontaneous at the given concentrations.

Q6.30
The established standard Gibbs energy for hydrolysis of ATP to ADP at 310K is −30.5kj mol
−1
. At −4.6 C
o
, determine the
in the process of the muscle of a hippo. (Hint: ΔrH = −20.1kjmol
′ ′
o o −1
ΔrG

S6.30
Step 1:
′ ′ ′
o o o
ΔrG = ΔrH − T ΔrS (6.E.62)

Step 2:
′ ′
o o −1 −1
′ ΔrH − ΔrG (−20.1kj mol ) − (−3.0kj mol )
o −2 −1 −1
ΔrS = $$$$ = = 3.355 × 10 kj K mol (6.E.63)
T 310K

Step 3:
′ ′ ′
o o o
ΔrG = ΔrH − T ΔrS (6.E.64)

−1 −2 −1 −1
= (−20.1kj mol ) − (298.15K)(3.355 × 10 kj K mol ) (6.E.65)

$$=-29.1kj\,mol^{-1}$$

6.E.9 https://chem.libretexts.org/@go/page/41332
Q6.31a
Which step of glycolysis would not occur spontaneously at standard-state conditions and why?

Q6.31b
Consider a hydrolysis of PEP, a phosphate compound.
PEP + H2O → pyruvate + Pi ΔrG°' = -61.9 kJ/mol
At the temperature of 288K, the following reaction took place and has a reaction Gibbs free energy of -49.5 kJ/mole. Find the
concentration of PEP for the reaction if the other concentrations are: [Pi] = 3.54 x 10-1 M, [Pyruvate] = 1.85 x 10-2 M.
Hint: Use equation ΔrG° = -RTlnKp

S6.31b
′ [pyruvate][ Pi ]

Δr G = Δr G + RT ln (6.E.66)
[P EP ]

−2 −1
J J J (1.85 × 10 M )(3.54 × 10 M)
−49500 = −61900 + (8.314 )(288K)ln (6.E.67)
mol mol K ⋅ mol [P EP ]

−5
[P EP ] = 3.69 × 10 M (6.E.68)

Q6.32
For the following reaction:

F ructose1, 6 − bisphosphate ⇌ dihydroxyacetonephosphate + glyceraldehyde3 − phosphate (6.E.69)

The ΔG° ′
= 5.7kcal/mol . Calculate the equilibrium constant and determine if the reaction is spontaneous or not at 310K.

S6.32

ΔG° = 5.7kcal/mol = 23.8488kJ/mol

′ ′
ΔG° = −RT ln(keq )

′ 23848.8J/mol
ΔG°

′ (−8.314J/K ∗ mol)(310K) −5
keq = e −RT =e = 9.58 × 10

Small k ′
eq suggest that this reaction is not a spontaneous process under the given conditions.

Q6.32a
Consider the following reaction:
Glucose + Fructose → Sucrose + H2O
Find:
a. the value of the standard free energy at 300 K.
b. the ratio between ∆rGo' at 300 K and ∆rGo' at 333 K.

S6.32a
a) The standard free energy = (-1544.3)- (-908.9 + -875.9) = 240.2 kJ/mol
b)delta Go at 300 K = delta G1 + RTln Q = -34.6 kJ/mol
delta GO at 333 K = delta G2 + RTln Q = -26.8 kJ/mol

Q6.32b
Consider the formation of the dipeptide glycylclycine. Using the following information, calculate the ΔrG∘ .

6.E.10 https://chem.libretexts.org/@go/page/41332
2 Glycine → Glycylglycine + H2 O (6.E.70)

∘ −1
△r G = 29.5 kJ mol (6.E.71)

[Glycine] = 1.4M (6.E.72)

[Glycylglycine] = 0.7M (6.E.73)

[ H2 O] = 1.0M (6.E.74)

S6.32b
Use the following equation and plug in values.

∘ ∘
△r G = △r G + RT ln(K) (6.E.75)

′ [Glycylglycine] ⋅ [ H2 O]
∘ ∘
△r G = △r G + RT ln (6.E.76)
2
[Glycine]

8.3145 0.7M ⋅ 1.0M


∘ −1 −1 −1
△r G = 29.5 kJ mol +( kJ mol K )(298K) ln (6.E.77)
2
1000 1.4 M
∘ −1
△r G = 26.7 kJ mol (6.E.78)
––––––––––––––––––––––

Click here for some more information on bioenergetics and free energy.

Q6.32c
Consider the following reaction:
+
2N H3(g) + H2 O(l) ⇌ 2N H (6.E.79)
4(aq)

∘ KJ
¯
Δf G (N H3(g) ) = −16.6 (6.E.80)
mol

∘ KJ
¯
Δf G (H2 O(l) ) = −237.2 (6.E.81)
mol

∘+
KJ
¯
Δf G (N H ) = −79.3 (6.E.82)
4(aq)
mol

What is the equilibrium constant of this process? Is this process typically spontaneous process under standard conditions?

S6.32c
To calculate change in Gibbs free energy for the equation, we use the standard molar Gibbs energies of formation for the reactants
and products:
∘+ ∘ ∘
¯ ¯ ¯
Δr G = 2 Δf G (N H ) − Δf G (H2 O(l) ) − 2 Δf G (N H3(g) ) (6.E.83)
4(aq)

kJ kJ kJ
= 2 (−79.3 ) − (−267.2 ) − 2 (−16.6 ) (6.E.84)
mol mol mol

kJ
= 141.8 (6.E.85)
mol

To calculate the equilibrium constant:


J
141800
Δr G mol
ln K = − =− = −57.23 (6.E.86)
RT J
(8.3145 )(298K)
K ⋅ mol

−57.23 −25
K =e = 1.40 × 10 (6.E.87)

Because the K value is so low for this process, the reaction does not occur spontaneously at 25o C

6.E.11 https://chem.libretexts.org/@go/page/41332
Q6.33a
Calculate the ΔrGº and equilibrium constant for the following reaction at 298K:
3-Phosphoglycerate → Phosphoenolpyruvate +H2O
Given that:
2-Phosphoglycerate → 3-Phosphoglycerate ΔrGº = -4.2 kJ mol-1
2-Phosphoglycerate → Phosphoenolpyruvate +H2O ΔrGº = -16.4 kJ mol-1

S6.33a
ΔrGº = 4.2 kJ mol-1 - 16.4 kJ mol-1 = 12.2 kJ mol-1

Δr G = −RT lnK (6.E.88)

3 −1
12.2 ∗ 10 J mol
lnK = (6.E.89)
−1 −1
−(8.314J mol K )298K

.00492
K =e (6.E.90)

K = 1.00 (6.E.91)

Q6.33b
Two acid dissociation reactions for carbonic and acetic acid and their corresponding pKa values are displayed below.
CH3CO2H ⇌ CH3CO2- + H+ ; pKa = 4.75
H2CO3 ⇌ HCO3- + H+; pKa = 6.37
Assuming the reactions take place at 298K, what is the equilibrium constant and change in gibbs’ free energy associated with the
following reaction?
CH3CO2- + H2CO3 ⇌ HCO3- + CH3CO2H

Q6.34
Suppose the isomerization of DHAP to GAP in glycolysis has an enthalpy of -1.20 kJ/mol. At 25°C the Gibbs free energy of the
reaction is 1.98 kJ/mol. Determine the equilibrium constant K of the isomerization at 25°C and at 35°C
Hint: Use van't Hoff equation

S6.34
For K at 25°C

Δr G = −RT lnK25∘ (6.E.92)

J J
1980 = −(8.314 )(25 + 273)K(lnK25∘ ) ⇒ K25∘ = 0.450 (6.E.93)
mol K⋅

For K at 35°C

K2 Δr H 1 1
ln = ( − ) (6.E.94)
K1 R T1 T2

K35∘ −1200J/mol 1 1
ln = ( − ) (6.E.95)
0.45 J (25 + 273)K (35 + 273)K
8.314
K ⋅ mol

K35∘ = 0.152 (6.E.96)

Q6.35a
Steady state and equilibrium state have an important role in understanding enzyme kinetics. What are some significant differences
between the two?

6.E.12 https://chem.libretexts.org/@go/page/41332
S6.35a
In a steady state, there is no net change over time in concentrations of reactants and products of a reaction since they are being
produced and consumed at constant rates. In this sense a steady state is a dynamic equilibrium. A steady state can be going in either
the forward reaction or the backward reaction.
A chemical equilibrium on the other hand is when a reaction goes in the forward and backward reaction at the same rate so there is
no net change in the system.
This is important because cells maintain steady states so that they are able to use particular reactions continuously. If a cell were at
chemical equilibrium it would be dead because it would be at the point where all reactions are not going anymore, among other
reasons.
Also, it is important to note that in a steady state, reactions are reversible compared to chemical equilibrium where the reaction rate
is zero.
Click here to see more information about steady states.

Q6.35b
Is the following an example of a steady state, equilibrium state, or neither:
a. Glycolysis, ingestion, and respiration
b. NaCl_{(s)}\rightleftharpoons Na^{+}_{(aq)}+Cl^{-}_{(aq)}
c. Sodium potassium ATPase
d. Citric Acid Cycle
e. oxidation of gold

S6.35b
a. steady state
b. equilibrium state
c. steady state
d. steady state
e. neither

Q6.36
Diatomic hydrogen gas and diatomic iodine gas are in equilibrium with hydrogen Iodide gas in a closed environment of unknown
volume at an unknown temperature.
Find the pressure equilibrium constant if the partial pressures of the gases are as follows: PH2 = 812 mmhg PI2= 587 mmhg PHI =
980 mmhg
Find the activity equilibrium constant if the activity constants of the gases are as follows γH2 = 1.45 γI2=0.844 γHI = 1.23
Use your results to find the thermodynamic equilibrium constant.

S6.36
The reaction in question is H2 + I2 ⇌ 2HI
The pressure equilibrium constant is given by PHI2/(PH2 PI2), the activity equilibrium constant is given by γHI2/(γH2 γI2), and the
thermodynamic equilibrium constant is given by the product of those two.
Kp = (980 mmhg)2/(812 mmhg * 587 mmhg) = 2.01
Kγ = 1.232/(1.45 * 0.844) = 1.24
Kthermo = 1.24 * 2.01 = 2.49

Q6.38
S O2 + C l2 → S O2 C l2 (6.E.97)

6.E.13 https://chem.libretexts.org/@go/page/41332
This reaction happens at 273K. You are given a Kp of 0.683. The pressure for SO2 is 0.58 bar for Cl2 is 0.93 bar, and SO2Cl2 is
0.776 bar. From this information, determine (delta)rG.

=(-8.314Jmol-1K-1)(273K)ln(0.683)
=865.4 J mol-1

=1.69 kj mol-1

Q6.40
Assuming oxygen binding to hemoglobin can be represented by the following reaction:
H b(aq) + O2(g) → H b O2(aq) (6.E.98)

If the value of ΔrG° for the reaction is -11.2 kJ mol-1 at 37°C, calculate the value of ΔrG° for the reaction.

Q6.42
At T=300K, given the mole ratio between 2 isomers Cis-2-butene and Trans-2-butene in an equilibrium mixture is 1:4. Evaluate
∆rG of the reversible reaction:
Cis-2-butene <------> Trans-2-butene

Q6.44
When discussing the reaction in biological cells, why would you use concentrations instead of activities?

S6.44
Concentrations are generally smaller and it's easier to compare the concentrations between two parts of a cell or between different
molecules in a reaction whereas activity describes behavior.

Q6.46
The following data shows the oxygen binding concentration in snails. The protein concentration is 15mM. Find n and K by using d

the Scatchard plot

[O2 ]total 10 mM 14 mM 18 mM 22 mM 26 mM 30 mM 34 mM 38 mM 42 mM 46 mM

[O2 ]bound 9 mM 12 mM 16 mM 19 mM 22 mM 25 mM 28 mM 32 mM 35 mM 39 mM

S6.46
y = mx + b

Y 1 n
= (− )Y +
[L] Kd Kd

[O2 ]bound
Y =
[P ]

. [L] = [O2 ]f ree

6.E.14 https://chem.libretexts.org/@go/page/41332
[O2]tot 10 14 18 22 26 30 34 38 42 46

[O2]bound 9 12 16 19 22 25 28 32 35 39

[O2]free 1 2 2 3 4 5 6 6 7 7

Y 0.6 0.8 1.067 1.267 1.467 1.667 1.867 2.133 2.333 2.6

Y/[L] 0.6 0.4 0.833 0.422 0.367 0.333 0.311 0.356 0.333 0.371

1
− = −0.102
Kd

Kd = 9.804

n
= 0.564
Kd

n = Kd ∗ 0.564 = 5.529

6.E: Chemical Equilibrium (Exercises) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

6.E.15 https://chem.libretexts.org/@go/page/41332
CHAPTER OVERVIEW

7: Electrochemistry
Electrochemistry is the study of chemical processes that cause electrons to move. This movement of electrons is called electricity,
which can be generated by movements of electrons from one element to another in a reaction known as an oxidation-reduction
("redox") reaction.
7.1: Electrochemical Cells
7.2: Single Electrode Potentials
7.3: Thermodynamics of Electrochemical Cells
7.4: Types of Electrochemical Cells
7.5: Applications of EMF Measurements
7.6: Biological Oxidation
7.7: Membrane Potential
7.E: Electrochemistry (Exercises)

7: Electrochemistry is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1
7.1: Electrochemical Cells
Electrochemistry is the study of chemical processes that cause electrons to move. This movement of electrons is called electricity,
which can be generated by movements of electrons from one element to another in a reaction known as an oxidation-reduction
("redox") reaction. A redox reaction is a reaction that involves a change in oxidation state of one or more elements. When a
substance loses an electron, its oxidation state increases; thus, it is oxidized. When a substance gains an electron, its oxidation state
decreases, thus being reduced. For example, for the redox reaction

H +F → 2 HF
2 2

can be rewritten as follows:


Oxidation reaction
+ −
H → 2H +2 e
2

Reduction reaction:
− −
F +2 e → 2F
2

Overall Reaction
+ −
H +F → 2H +2 F
2 2

Oxidation is the loss of electrons, whereas reduction refers to the acquisition of electrons, as illustrated in the respective reactions
above. The species being oxidized is also known as the reducing agent or reductant, and the species being reduced is called the
oxidizing agent or oxidant. In this case, H2 is being oxidized (and is the reducing agent), while F2 is being reduced (and is the
oxidizing agent). The following acronym is useful in remembering this concept:

"OIL RIG"
"OIL RIG" is a useful mnemonic for remembering the definitions of oxidation nd reduction.
Oxidation Is Losing electrons; Reduction Is Gaining electrons

Example 7.1.1: IRon-Vanadium Couple

Given the redox reaction


3 + 2 + 2 + 3 +
Fe +V → Fe +V

which species is oxidized? Which is reduced? Identify the reducing agent and the oxidizing agent.
Solution
Fe
3 +
is reduced into Fe and
2 +
V
2 +
is oxidized into V
3 +
. This is because the oxidized species loses electrons, and the
reduced species gains electrons.
Iron gains an electron
3 + 2 +
Fe → Fe

and vanadium loses an electron


2 + 3 +
V → V .

Thus, Fe 3 +
is the oxidizing agent and V 2 +
is the reducing agent.

Oxidation States

7.1.1 https://chem.libretexts.org/@go/page/41445
Rules for Assigning Oxidation States
1. Free elements have an oxidation state of 0. (e.g., He, N , O has an oxidation state of 0)
2 2

2. The oxidation state of one ion must equal the net charge. (e.g., F oxidation state is -1, K oxidation state is +1)
− +

3. The sum of the oxidation state has to equal the total net charge for a compound. (e.g., MnO has a net charge of -1 with

Mn(+7)O4(-8)= -1)
4. The alkali metals (Group I elements) have an oxidation state of +1. (e.g., Li O, Li= +1)
2

5. The alkaline earth metals (Group II elements) always have an oxidation state of +2. (e.g., CaO, Ca=+2)
6. Oxygen has an oxidation state of -2 in a compound
7. Fluorine has an oxidation state of -1 in a compound
8. Hydrogen has an oxidation state of +1 in a compound.
9. Transition metals and other metals may have more than one common ionic charge. (e.g., Chromium's common ionic
charges are Cr+2 and Cr+3)

Exercise 7.1.1

What is the oxidation state of magnesium in MgF ? 2

Answer
Using rule 5 and 7.
MgF
2
total charge=0 Total Charge=(+2)+(-1*2)=0

Exercise 7.1.2

What is the oxidation state of hydrogen in H2


O ?

Answer
Using rule 6 and 8
H O
2
total charge=0 Total Charge=(+1*2)+(-2)=0

Balancing Redox Reactions


Method 1: Oxidation Number Method
Step 1: Assign oxidation numbers to each atom.
Step 2: Determine the net change in charge to determine the ratio of atoms
Step 3: Use the ratio to eliminate the net charge change
Step 4: Use the ratio as coefficients for the elements
Step 5: Add H+ (under acidic conditions), OH- (under basic conditions), and H2O to balance charges.
Method 2: Half-Reaction Method

Step 1: Determine oxidation numbers for each atom


Step 2: Use oxidation numbers to determine what is oxidized and what is reduced.
Step 3: Write a half-reaction for reduction
Step 4: Write a half-reaction for oxidation
Step 5: Balance all elements except H and O
if have acid redox reaction: Balance the O using H 2O , balance the H using protons
if have base redox reaction: Balance O using OH −

Step 6: Add up the charge on each side


Step 7: Balance the charges by adding electrons
Step 8: Multiply the half-reactions by factors that cancel out electrons
Step 9: Add the two half-reactions back together to eliminate out intermediates

7.1.2 https://chem.libretexts.org/@go/page/41445
Example 7.1.2: Manganese

Balance the following reaction in an acidic aqueous Solution


− 2 +
MnO +H C O → Mn + CO
4 2 2 4 2

Solution

Reduction half-reaction:
− + − 2 +
2 × (5 e +8 H + MnO 4 → Mn + 4 H O)
2

Oxidation half-reaction:
+ −
5 × (H C O → 2 CO +2 H +2 e )
2 2 4 2

Combining and canceling gives the following:


− + 1 − 2 + + −
10 e + 16 H + 2 MnO 4 +5 H C O → 2 Mn + 8 H O + 10 CO + 10 H + 10 e
2 2 4 2 2

Answer
+ 1 − 2 +
6H + 2 MnO +5 H C O → 2 Mn + 8 H O + 10 CO
4 2 2 4 2 2

Voltaic (Galvanic) Cells


In 1793, Alessandro Volta discovered that electricity could be produced by placing different metals on the opposite sides of a wet
paper or cloth. He made his first battery by placing Ag and Zn on the opposite sides of a moistened cloth with salt or weak acid
Solution. Therefore, these batteries acquired the name voltaic cells. Voltaic (galvanic) cells are electrochemical cells that contain a
spontaneous reaction, and always have a positive voltage. The electrical energy released during the reaction can be used to do
work. A voltaic cell consists of two compartments called half-cells. The half-cell where oxidation occurs is called the anode. The
other half-cell, where reduction occurs, is called the cathode. The electrons in voltaic cells flow from the negative electrode to the
positive electrode—from anode to cathode (see figure below). (Note: the electrodes are the sites of the oxidation and reduction
reactions). The following acronym is useful in keeping this information straight:
Red Cat and An Ox
Reduction Cathode and Anode Oxidation
For an oxidation-reduction reaction to occur, the two substances in each respective half-cell are connected by a closed circuit such
that electrons can flow from the reducing agent to the oxidizing agent. A salt bridge is also required to maintain electrical neutrality
and allow the reaction to continue.

The figure above shows that Zn(s) is continuously oxidized, producing aqueous Zn 2 +
:
2 + −
Zn(s) → Zn (aq) + 2 e (7.1.1)

Conversely, in the cathode, Cu 2 +


is reduced and continuously deposits onto the copper bar:

7.1.3 https://chem.libretexts.org/@go/page/41445
2 + −
Cu (aq) + 2 e → Cu(s) (7.1.2)

As a result, the Solution containing Zn(s) becomes more positively charged as the Solution containing Cu(s) becomes more
negatively charged. For the voltaic cell to work, the Solutions in the two half-cells must remain electrically neutral. Therefore, a
salt bridge containing KNO3 is added to keep the Solutions neutral by adding NO3-, an anion, into the anode Solution and K , a +

cation, into the cathode Solution. As oxidation and reduction proceed, ions from the salt bridge migrate to prevent charge buildup
in the cell compartments.
The cell diagram (or cell notation) is a shorthand notation to represent the redox reactions of an electrical cell. For the cell
described, the cell diagram is as follows:
2 + 2 +
Zn(s)| Zn (aq)|| Cu (aq)|Cu(s) (7.1.3)

A double vertical line (||) is used to separate the anode half reaction from the cathode half reaction. This represents the salt
bridge.
The anode (where oxidation occurs) is placed on the left side of the (||).
The cathode (where reduction occurs) is placed on the right side of the (||).
A single vertical line (|) is used to separate different states of matter on the same side, and a comma is used to separate like
states of matter on the same side. For example:
2 + 3 + +
Fe (aq), Fe (aq)|| Ag (aq)|Ag(s) (7.1.4)

Figure 7.1.2 : A voltaic cell works by the different reactivity of metal ions, and not require external battery source. Image taken at
Hope College as part of their General Chemistry Lab curriculum.
The figure above shows a set of electrochemical half-cells that can be used to measure various voltages within galvanic cells. The
cells shown are made of agar saturated with KCl Solution so as to act as a salt bridge. The zinc electrode in the middle can be used
as a reference while the various concentrations of copper (labeled 1, 2, 3, 4 and 5) can be tested to form a calibration curve. The
potential of the unknown can be used to determine the concentration of an unknown copper Solution. This application of the Nernst
equation allows for rapid data collection without the need for a complicated salt bridge apparatus.

Example 7.1.3: Copper-silver Reaction


Write the cell diagram for this reaction:

2 +
Cu(s) + 2 Ag+(aq) → Cu (aq) + 2 Ag(s)

Solution
2 + +
Cu(s)| Cu (aq)|| Ag (aq)|Ag(s)

Example 7.1.4: Aluminum-Tin Reaction


Write cell reactions for this cell diagram:

3 + 2 +
Al(s)| Al (aq)|| Sn (aq)|Sn(s)

Answer

7.1.4 https://chem.libretexts.org/@go/page/41445
Oxidation: {Al(s) → Al3+(aq) +3e-} x 2
Reduction: {Sn2+(aq) +2e- → Sn(s)} x 3

Net: 2Al(s) + 3Sn2+(aq) → 2Al3+(aq) + 3Sn(s)

Cell Potentials
The oxidation of Zn(s) into Zn2+ and the reduction of Cu2+ to Cu(s) occur spontaneously. In other words, the redox reaction
between Zn and Cu2+ is spontaneous. This is due to the difference in potential energy between the two substances. The difference
in potential energy between the anode and cathode dictates the direction of electronic movement. Electrons move from areas of
higher potential energy to areas of lower potential energy. In this case, the anode has a higher potential energy; electrons therefore
move from anode to cathode. The potential difference between the two electrodes is measured in units of volts. One volt (V) is the
potential difference necessary to generate a charge of 1 coulomb (C) from 1 joule (J) of energy.
For a voltaic cell, this potential difference is called the cell potential (or EMF for electromotive force, although it is not really a
force), which is denoted Ecell. For a spontaneous reaction, Ecell is positive and ΔG (Gibbs free energy, used to determine if a
reaction occurs spontaneously) is negative. Thus, when ΔG is negative the reaction is spontaneous. Merging electrochemistry with
thermodynamics gives this formula:
ΔG = −nF Ecell (7.1.5)

Cell potential is different for each voltaic cell; its value depends upon the concentrations of specific reactants and products as well
as temperature of the reaction. For standard cell potential, temperature of the reaction is assumed to be 25o Celsius, the
concentration of the reactants and products is 1 M, and reaction occurs at 1 atm pressure. The standard cell potential is denoted
Eocell, and can be written as oxidation potential + reduction potential. For voltaic cells:
o o o
E =E −E (7.1.6)
cell cathode Anode

Warning
To use Equation 7.1.6, the cell potentials must be in reduction form.

Since standard potentials are given in the form of standard reduction potential for each half-reaction, to calculate the standard cell
potential E , the substance is being oxidized must be identified; then the standard reduction potential of the oxidation reaction is
o
cell

subtracted from the standard reduction potential of the reducing reaction.

Example 7.1.5

What is the cell potential for the following reaction?


2 + 2 +
Zn(s) + Cu (aq) → Zn (aq) + Cu(s)

Solution
Zn(s) is being oxidized, and Cu(s) is being reduced. The potentials for the two half reaction are given in the reduction form:
2 + −
Zn(s) → Zn (aq) + 2 e

2 + −
Cu (aq) + 2 e → Cu(s)

The cell potentials indicate which reaction takes place at the anode and which at the cathode. The cathode has a more positive
potential energy, and thus:
Cu(s) is the cathode
Zn(s) is the anode.
To calculate E o
cell
, subtract the E of the oxidized half reaction from the E
o o
cell
of the reduction half reaction, we use Equation
7.1.6}:

o o o
E =E −E
cell cathode anode

7.1.5 https://chem.libretexts.org/@go/page/41445
Oxidation half reaction: Eo= -0.763V
2 + −
Zn(s) → Zn (aq) + 2 e

Reduction half reaction: Eo= +0.342V


2 + −
Cu (aq) + 2 e → Cu(s)

Net reaction:
2 + 2 +
Zn(s) + Cu (aq) → Zn (aq) + Cu(s)

Therefore:
o o o
E =E −E = 0.340 V − (−0.763 V ) = +1.103 V
cell cathode anode

Exercise 7.1.6

Calculate Eocell for the following redox reaction under standard conditions:
2 + 3 +
2 Al(s) + 3 Sn (aq) → 2 Al (aq) + 3 Sn(s)

Answer
Oxidation:
3 + − o
{Al(s) → Al (aq) + 3 e } ×2 −E = +1.676 V

Reduction:
2 + − o
{ Sn (aq) + 2 e → Sn(s)} × 3 E = −0.137 V

Net reaction:
2 + 3 +
2 Al(s) + 3 Sn (aq) → 2 Al (aq) + 3 Sn(s)

Then using Equation 7.1.6


o
E = −0.137 V − (−1.676V )
cell

= +1.539 V .

Voltage is an Intensive Property


Standard reduction potential is an intensive property, meaning that changing the stoichiometric coefficient in a half reaction does
not affect the value of the standard potential. For example,
Oxidation:{Al(s) → Al3+(aq) +3e-} x 2 is still Eo= -1.676
Reduction:{Sn2+(aq) +2e- → Sn(s)} x 3 is still Eo= -0.137
If the stoichiometric coefficient is multiplied by 2, the standard potential does not change:

Example 7.1.7: Iron/Vanadium Chemistry

Calculate the cell potential in the following redox reaction under standard conditions:
3 + 2 + 2 + 3 +
Fe (aq) + V (aq) → Fe (aq) + V (aq)

Solution
Consult the table of standard reduction potentials (Table P1) for each half reaction:
3+ − 2+ o
Fe +e → Fe with E = 0.771 V
(aq) (aq)

7.1.6 https://chem.libretexts.org/@go/page/41445
2+ 3+ − o
V → V +e with E = −0.255 V
(aq) (aq)

The cell potential is


o o o
E =E −E = 0.771 V − (−0.255 V ) = 1.026 V
cell cathode anode

Glossary
Anode: Electrode in an electrochemical cell on which the oxidation reaction occurs.
Cathode: Electrode in an electrochemical cell on which the reduction reaction occurs
Electrochemistry: A field of chemistry that focuses on the interchange between electrical and chemical energy
Electricity: Flow of electrons over a wire that is affected by the presence and flow of electric charge.
Electrolysis: The decomposition of a substance by means of electric current. This method pushes a redox reaction toward the
non-spontaneous side.
Electrolytic cell: Electrochemical cell that is being pushed toward the non-spontaneous direction by electrolysis.
Electromotive force, EMF (or cell potential): Difference of potential energy of electrons between the two electrodes.
Oxidation number: Charge on an atom if shared electrons where assigned to the more electronegative atom.
Oxidation: Lose of electrons, can occur only in combination with reduction. [remember: Oxidation Is Loss, Reduction Is Gain =
OIL RIG]
Reduction: Gain of electrons, can occur only in combination with oxidation. [remember: OIL RIG]
Redox reaction: Shorthand for reduction-oxidation reaction.
Voltaic cell or galvanic cell: An electrochemical cell that uses redox reaction to produce electricity spontaneously.

References
1. Atkins, P. de Paula, J. Physical Chemistry for the Life Sciences. pg 209-225. 2006. Oxford Univeristy Press. New York.
2. Zumdahl, S. Zumdahl, S. Chemistry. pg 215-220. 2007. Houghton Mifflin Company. New Jersey.
3. Petrucci, Ralph H. General Chemistry Principles & Modern Applications. Pearson Prentice Hall. New Jersey
4. Jones, Atkins. Chemistry Molecules, Matter, and Change. W. Freeman and Company, New York.
5. Smela M, Currier S, Bailey E, Essignmann J. Journal of the electrochemical society. pg 392-394. 2001Academic Press, New
York.
6. Van R, S. J. 1977. The open electrochemistry journal. pg 561-566. 1994. Bentham Open,
7. Huddle, Penelope Ann; White, Margaret Dawn; Rogers, Fiona. "Using a Teaching Model to Correct Known Misconceptions in
Electrochemistry." J. Chem. Educ.2000 77 104.
8. Sanger, Michael J.; Greenbowe, Thomas J. "An Analysis of College Chemistry Textbooks As Sources of Misconceptions and
Errors in Electrochemistry." J. Chem. Educ. 1999, 76, 853.

Contributors and Attributions


Matthew Bui (UCD), Wen Chung Chou (UCD)
Justin Shorb (Hope College), Yong Chul Yoon (Hope College)

7.1: Electrochemical Cells is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

7.1.7 https://chem.libretexts.org/@go/page/41445
7.2: Single Electrode Potentials
7.2: Single Electrode Potentials is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

7.2.1 https://chem.libretexts.org/@go/page/41446
7.3: Thermodynamics of Electrochemical Cells
7.3: Thermodynamics of Electrochemical Cells is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.

7.3.1 https://chem.libretexts.org/@go/page/41447
7.4: Types of Electrochemical Cells
Concentration Cells
A concentration cell is an electrolytic cell that is comprised of two half-cells with the same electrodes, but differing in
concentrations. A concentration cell acts to dilute the more concentrated solution and concentrate the more dilute solution, creating
a voltage as the cell reaches an equilibrium. This is achieved by transferring the electrons from the cell with the lower
concentration to the cell with the higher concentration.
The standard electrode potential, commonly written as Eocell, of a concentration cell is equal to zero because the electrodes are
identical. But, because the ion concentrations are different, there is a potential difference between the two half-cells. One can find
this potential difference via the Nernst Equation,
0.0592
o
Ecell = E − logQ (7.4.1)
cell
n

at 25oC. The E stands for the voltage that can be measured using a voltmeter (make sure if the voltmeter measures it in millivolts
that you convert the number before using it in the equation). Note that the Nernst Equation indicates that cell potential is dependent
on concentration, which results directly from the dependence of free energy on concentration. Remember that to find Q you use this
equation:
aA + bB ⇌ cC + dD (7.4.2)

c d
(C ) ∗ (D)
Q = (7.4.3)
a b
(A) ∗ (B)

When Q=1, meaning that the concentrations for the products and reactants are the same, then taking the log of this equals zero.
When this occurs, the Ecell is equal to the Eocell.
Another way to use the Eocell , or to find it, is using the equation below.
o o o
E =E −E (7.4.4)
cell cathode anode

A voltage can also be generated by constructing an electrochemical cell in which each compartment contains the same redox active
solution but at different concentrations. The voltage is produced as the concentrations equilibrate. Suppose, for example, we have a
cell with 0.010 M AgNO3 in one compartment and 1.0 M AgNO3 in the other. The cell diagram and corresponding half-reactions
are as follows:
+
Ag(s) | Ag (aq, 0.010 M ) || Ag + (aq, 1.0 M ) | Ag(s) (7.4.5)

cathode:
+ −
Ag (aq, 1.0 M ) + e → Ag(s) (7.4.6)

anode:
+ −
Ag(s) → Ag (aq, 0.010 M ) + e (7.4.7)

Overall
+ +
Ag (aq, 1.0 M ) → Ag (aq, 0.010 M ) (7.4.8)

As the reaction progresses, the concentration of Ag will increase in the left (oxidation) compartment as the silver electrode
+

dissolves, while the Ag concentration in the right (reduction) compartment decreases as the electrode in that compartment gains
+

mass. The total mass of Ag(s) in the cell will remain constant, however. We can calculate the potential of the cell using the Nernst
equation, inserting 0 for E°cell because E°cathode = −E°anode:
0.0591 V 0.0591 V 0.010

Ecell = E −( ) log Q = 0 − ( ) log( ) = 0.12 V (7.4.9)
cell
n 1 1.0

An electrochemical cell of this type, in which the anode and cathode compartments are identical except for the concentration of a
reactant, is called a concentration cell. As the reaction proceeds, the difference between the concentrations of Ag+ in the two

7.4.1 https://chem.libretexts.org/@go/page/41448
compartments will decrease, as will Ecell. Finally, when the concentration of Ag+ is the same in both compartments, equilibrium
will have been reached, and the measured potential difference between the two compartments will be zero (Ecell = 0).

Example 7.4.2

Calculate the voltage in a galvanic cell that contains a manganese electrode immersed in a 2.0 M solution of MnCl2 as the
cathode, and a manganese electrode immersed in a 5.2 × 10−2 M solution of MnSO4 as the anode (T = 25°C).
Given: galvanic cell, identities of the electrodes, and solution concentrations
Asked for: voltage
Strategy:
A. Write the overall reaction that occurs in the cell.
B. Determine the number of electrons transferred. Substitute this value into the Nernst equation to calculate the voltage.
Solution:
A This is a concentration cell, in which the electrode compartments contain the same redox active substance but at different
concentrations. The anions (Cl− and SO42−) do not participate in the reaction, so their identity is not important. The overall
reaction is as follows:
2+ 2+
Mn → Mn −2
(7.4.10)
(aq,2.0M) (aq,5.2×10 M)

B For the reduction of Mn2+(aq) to Mn(s), n = 2. We substitute this value and the given Mn2+ concentrations into Equation
??? :

−2
0.0591 V 0.0591 V 5.2 × 10

Ecell = E −( ) log Q = 0 V − ( ) log( ) = 0.047 V
cell
n 2 2.0

Thus manganese will dissolve from the electrode in the compartment that contains the more dilute solution and will be
deposited on the electrode in the compartment that contains the more concentrated solution.

Exercise 7.4.2

Suppose we construct a galvanic cell by placing two identical platinum electrodes in two beakers that are connected by a salt
bridge. One beaker contains 1.0 M HCl, and the other a 0.010 M solution of Na2SO4 at pH 7.00. Both cells are in contact with
the atmosphere, with P = 0.20 atm. If the relevant electrochemical reaction in both compartments is the four-electron
O2

reduction of oxygen to water:


+ −
O2(g) + 4 H + 4e → 2 H2 O(l) (7.4.11)
(aq)

What will be the potential when the circuit is closed?


Answer 0.41 V

Fuel Cells
References
1. Petrucci, Ralph H. General Chemistry: Principles and Modern Applications 9th Ed. New Jersey: Pearson Education Inc. 2007.
2. Zumdahl, steven S. Chemistry, 9th Ed. New York: Houghton Mifflin Co. 2007.
3. Ciparick, Joseph D. "Half cell reactions: Do students ever see them? (TD)." J. Chem. Educ. 1991, 68, 247.
4. Tanis, David O. "Galvanic cells and the standard reduction potential table (F&R)." J. Chem. Educ. 1990, 67, 602.

Contributors and Attributions


SSReno, Rainna Lim

7.4: Types of Electrochemical Cells is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

7.4.2 https://chem.libretexts.org/@go/page/41448
7.5: Applications of EMF Measurements
7.5: Applications of EMF Measurements is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

7.5.1 https://chem.libretexts.org/@go/page/41449
7.6: Biological Oxidation
Before the concept of biological oxidation can be understood and explored, the fundamental chemical process through which
oxidation and reduction take place should be first established.

Foundation
All reactions which involve electron flow are considered oxidation-reduction reactions. The basic definition can be defined as: One
reactant is oxidized (loses electrons), while another is reduced (gains electrons). A couple of basic oxidation-reduction or "redox"
example's are given here.

 Example 1

The reaction of magnesium metal with oxygen, involves the oxidation of magnesium
2M g(s) + O2 (g) → 2M gO(s) (7.6.1)

Since the magnesium solid is oxidized, we expect to see a loss of electrons. Similarly, since oxygen must therefore be reduced,
we should see a gain of electrons.

As the magnesium is oxidized there is a loss of 2 electrons while simultaneously, oxygen gains those two electrons. Another
example of a redox reaction is with the two gasses CO2 and H2. This redox reaction also demonstrates the importance of
implementing "oxidation numbers" in the methodology of redox reactions, allowing for the determination of which reactant is
being reduced and which reactant is being oxidized.

 Example 2
The reaction of carbon dioxide gas with hydrogen gas, involving the oxidation of hydrogen

C O2 (g) + H2 (g) → 2C O(g) + H2 O(g) (7.6.2)

Since the hydrogen gas is being oxidized (reductant), we expect to see an overall loss of electrons for the resulting molecule.
Similarly, we expect to see a gain in the overall number of electrons for the resulting molecule of the oxidant (CO2).

Here it is possible to infer that the carbon of CO2 is being reduced by review of its unique oxidation number. Such that, C (of
CO2) goes from an oxidation number of +4 to C (of CO) having an oxidation number of +2, representing a loss of two
electrons. Similarly, H2 is noted as going from an oxidation number of 0 to +1, or gaining one electron in a reduction process.
For more information on oxidation numbers, review the following link: Oxidation-Reduction Reactions

A Basic Biological Model


The flow of electrons is a vital process that provides the necessary energy for the survival of all organisms. The primary source of
energy that drives the electron flow in nearly all of these organisms is the radiant energy of the sun, in the form of electromagnetic
radiation or Light. Through a series of nuclear reactions, the sun is able to generate thermal energy (which we can feel as warmth)
from electromagnetic radiation (which we perceive as light). However, the particular wavelength of the electromagnetic spectrum
we are able to detect with the human eye is only between 400 and 700 nm in wavelength. It should therefore be noted that the
visible part of the electromagnetic spectrum is actually a small percentage of the whole; where a much greater percentage remains
undetectable for the human eye.

7.6.1 https://chem.libretexts.org/@go/page/41450
Figure 1: The electromagnetic spectrum with an emphasis on the visible light region
In physics, the use of the term "light" refers to electromagnetic radiation of any wavelength, independent of its detectability for the
human eye. For plants, the upper and lower ends of the visible spectrum are the wavelengths that help drive the process of splitting
water (H2O) during photosynthesis, to release its electrons for the biological reduction of carbon dioxide (CO2) and the release of
diatomic oxygen (O2) to the atmosphere. It is through the process of photosynthesis that plants are able to use the energy from light
to convert carbon dioxide and water into the chemical energy storage form called glucose.
Plants represent one of the most basic examples of biological oxidation and reduction. The chemical conversion of carbon dioxide
and water into sugar (glucose) and oxygen is a light-driven reduction process:

6C O2 + 6 H2 O → C6 H12 O6 + 6 O2 (7.6.3)

The process by which non photosynthetic organisms and cells obtain energy, is through the consumption of the energy rich
products of photosynthesis. By oxidizing these products, electrons are passed along to make the products carbon dioxide, and
water, in an environmental recycling process. The process of oxidizing glucose and atmospheric oxygen allowed energy to be
captured for use by the organism that consumes these products of the plant. The following reaction represents this process:

C6 H12 O6 + O2 → 6C O2 + 6 H2 O + Energy (7.6.4)

It is therefore through this process that heterotrophs (most generally "animals" which consume other organisms obtain energy) and
autotrophs (plants which are able to produce their own energy) participate in an environmental cycle of exchanging carbon dioxide
and water to produce energy containing glucose for organismal oxidation and energy production, and subsequently allowing the
regeneration of the byproducts carbon dioxide and water, to begin the cycle again. Therefore, these two groups of organisms have
been allowed to diverge interdependently through this natural life cycle.

Physical Chemistry's Understanding


Biological oxidation-reduction reactions, or simply biological oxidations utilize multiple stages or processes of oxidation to
produce large amounts of Gibbs energy, which is used to synthesize the energy unit called adenosine triphosphate or ATP. To
efficiently produce ATP, the process of glycolysis must be near an abundance of oxygen. Since glycolysis by nature is not an
efficient process, if it lacks sufficient amounts of oxygen the end product pyruvate, is reduced to lactate with NADH as the
reducing agent. However, in a more favorable aerobic process, the degradation of glucose through glycolysis proceeds with two
additional processes known as the citric acid cycle and the terminal respiratory chain; yielding the end products carbon dioxide
and water, which we exhale with each breath.

Figure 2: The three main processes for the breakdown of glucose into carbon dioxide and water
The products NADH and FADH2 formed during glycolysis and the citric acid cycle are able to reduce molecular oxygen (O2)
thereby releasing large amounts of Gibbs energy used to make ATP. The process by which electrons are transferred from NADH or
FADH2 to O2 by a series of electron transfer carriers, is known as oxidative phosphorylation. It is through this process that ATP is
able to form as a result of the transfer of electrons.
Thee specific examples of redox reactions that are used in biological processes, involving the transfer of electrons and hydrogen
ions as follows. During some biological oxidation reactions, there is a simultaneous transfer of hydrogen ions with electrons (1). In
other instances, hydrogen ions may be lost by the substance being oxidized while transferring only its electrons to the substance
being reduced (2). A third type of biological oxidation might involve only a transfer of electrons (3). It should be noted that
biological oxidation rarely proceeds in a direct manner, and generally involves complex mechanisms of several enzymes. The
outline below recaps the three processes of biological oxidation stated above, in descending order.

7.6.2 https://chem.libretexts.org/@go/page/41450
Table 1: Transfer of hydrogen ions and electrons for the general reaction scheme of A + B with intermediate stage shown
Reactants Intermediate Stage Products

AH2 + B [A + 2H+ + 2e- + B] A + BH2

AH2 + B [A + 2H+ + 2e- + B] A + B2- + 2H+

A2- + B [A + 2e- + B] A + B2-

In the last stage of the metabolic process (the terminal respiratory chain), the sequence by which electrons are carried is determined
by relative redox potentials. The carrier molecules used to transfer electrons in this stage are called cytochromes, which are an
electron-carrying protein containing a heme group. The iron atom of each cytochrome molecule can exist either in the oxidized
(Fe3+) or reduced (Fe2+) form. Within the terminal respiratory chain, each carrier molecule alternates between the reduced state and
the oxidized state, with molecular oxygen as the final electron acceptor at the end.
TRC.JPG

Figure 3. The terminal respitory chain showing electron transport and phosphorylation. Electrons from the citric acid cycle are
transferred from one carrier to another, where each carrier alternates between the reduced and oxidized state. Molecular oxygen
represents the final electron acceptor.
It is through the knowledge of redox potentials, that the knowledge of biological processes can be further expanded. The standard
reduction potential is denoted as Eo' and is often based on the hydrogen electrode scale of pH 7, rather than pH 0, a common
reference point for listed values. Moreover, the superscript symbol ( o ) denotes standard-state conditions, while the adjacent
superscript symbol ( ' ) denotes the pH scale of 7 for biochemical processes.
It therefore becomes possible to trace the energy transfer in cells back to the fundamental flow of electrons from one particular
molecule to another. Where this electron flow occurs via the physics principle of higher potential to lower potential; similar to a
ball rolling down a hill, as opposed to the opposite direction. All of these reactions involving electron flow can be attributed to the
basic definition of the oxidation-reduction pathway stated above.

References
1. Chang, Raymond. PHYSICAL CHEMISTRY for the Chemical and Biological Sciences . 3rd. Sausalito, CA: University Science
Books , 373-389. Print.
2. Nelson, David, and Michael Cox. LEHNINGER PRINCIPLES OF BIOCHEMISTRY . 5th . New York, NY: Freeman and
Company, 22. Print.

Contributors and Attributions


Brent Younglove (Hope)

7.6: Biological Oxidation is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
Biological Oxidation is licensed CC BY-NC-SA 4.0.

7.6.3 https://chem.libretexts.org/@go/page/41450
7.7: Membrane Potential
Membrane potential is what we use to describe the difference in voltage (or electrical potential) between the inside and outside of a
cell.

Introduction
Without membrane potentials human life would not be possible. All living cells maintain a potential difference across their
membrane. Simply stated, membrane potential is due to disparities in concentration and permeability of important ions across a
membrane. Because of the unequal concentrations of ions across a membrane, the membrane has an electrical charge. Changes in
membrane potential elicit action potentials and give cells the ability to send messages around the body. More specifically, the
action potentials are electrical signals; these signals carry efferent messages to the central nervous system for processing and
afferent messages away from the brain to elicit a specific reaction or movement. Numerous active transports embedded within the
cellular membrane contribute to the creation of membrane potentials, as well as the universal cellular structure of the lipid bilayer.
The chemistry involved in membrane potentials reaches to many scientific disciplines. Chemically it involves molarity,
concentration, electrochemistry and the Nernst equation. From a physiological standpoint, membrane potential is responsible for
sending messages to and from the central nervous system. It is also very important in cellular biology and shows how cell biology
is fundamentally connected with electrochemistry and physiology. The bottom line is that membrane potentials are at work in your
body right now and always will be as long as you live.

History
The subject of membrane potential stretches across multiple scientific disciplines; Membrane Potential plays a role in the studies of
Chemistry, Physiology and Biology. The culmination of the study of membrane potential came in the 19th and early 20th centuries.
Early in the 20th century, a man named professor Bernstein hypothesized that there were three contributing factors to membrane
potential; the permeability of the membrane and the fact that [K+] was higher inside and lower on the outside of the cell. He was
very close to being correct, but his proposal had some flaws. Walther H. Nernst, notable for the development of the Nernst equation
and winner of 1920 Nobel Prize in chemistry, was a major contributor to the study of membrane potential. He developed the Nernst
equation to solve for the equilibrium potential for a specific ion. Goldman, Hodgkin and Katz furthered the study of membrane
potential by developing the Goldman-Hodgkin-Katz equation to account for any ion that might permeate the membrane and affect
its potential. The study of membrane potential utilizes electrochemistry and physiology to formulate a conclusive idea of how
charges are separated across a membrane.

Figure 1. Differences in concentration of ions on opposite sides of a cellular membrane produce a voltage difference called the
membrane potential. The largest contributions usually come from sodium (Na+) and chloride (Cl–) ions which have high
concentrations in the extracellular region, and potassium (K+) ions, which along with large protein anions have high concentrations
in the intracellular region. Calcium ions, which sometimes play an important role, are not shown.

Membrane Potential and Cellular Biology


In discussing the concept of membrane potentials and how they function, the creation of a membrane potential is essential. The
lipid bilayer structure of the cellular membrane, with its lipid-phosphorous head and fatty acid tail, provides a perfect building
material that creates both a hydrophobic and hydrophilic side to the cellular membrane. The membrane is often referred to as a
mosaic model because of its semi-permeability and its ability to keep certain substances from entering the cell. Molecules such as
water can diffuse through the cell based on concentration gradients; however, larger molecules such as glucose or nucleotides
require channels. The lipid bilayer also houses the Na+/K+ pump, ATPase pump, ion transporters, and voltage gated channels, and it

7.7.1 https://chem.libretexts.org/@go/page/41451
is the site of vesicular transport. The structure regulates which ions enter and exit to determine the concentration of specific ions
inside of the cell.
Why is membrane potential essential to the survival of all living creatures?
Animals and plants require the breakdown of organic substances through cellular respiration to generate energy. This process,
which produces ATP, is dependent on the electron transport chain. Electrons travel down this path to be accepted by oxygen or
other electron acceptors. The initial electrons are obtained from the breakdown of water molecules. The hydrogen build up in the
extracellular fluid leaving a gradient. As per membrane potentials, when there a gradient, the molecules flow in the opposite
direction. In this case, hydrogen flows back into the cell through a protein known as ATP synthase which creates ATP in the
process. This action is essential to life because the number of ATP created from each glucose increases drastically. Chemical
disequilibrium and membrane potentials allow bodily functions to take place.

Figure 2
Transport proteins, more specifically the 'active' transport proteins, can pump ions and molecules against their concentration
gradient. This is the main source of charge difference across the cellular membrane.

Physiology of Membrane Potential


Understanding Membrane Potential
The following points should help you to understand how membrane potential works
The difference between the electrical and chemical gradient is important.
Electrical Gradient
Opposes the chemical gradient.
Represents the difference in electrical charge across the membrane
Chemical Gradient
Opposes the electrical gradient
Represents the difference in the concentration of a specific ion across the membrane.
A good example is K+. The membrane is very permeable to K+ and the [K+] inside the cell is great, therefore a positive
charge is flowing out of the cell along with K+. The [K+] inside the cell decreases causing the concentration gradient to flow
towards the outside of the cell. This also causes the inside of the cell to become more electronegative increasing its electrical
gradient.
The Nernst equation can help us relate the numerical values of concentration to the electrical gradient.
Leak Channels
Channels that are always open
Permit unregulated flow of ions down an electrochemical gradient.
Na+/K+ ATPase Pump
Actively transports Na+ out of the cell and K+ into the cell.

7.7.2 https://chem.libretexts.org/@go/page/41451
Helps to maintain the concentration gradient and to counteract the leak channels.

Membrane Potential and Physiology of Human Nerve Cells


Human nerve cells work mainly on the concept of membrane potentials. They transmit chemicals known as serotonin or dopamine
through gradients. The brain receives these neurotransmitters and uses it to perform functions.

Na+ has a much higher concentration outside of the cell and the cell membrane is very impermeable to Na+
K+ has a high concentration inside the cell due to the fact that the cell membrane is very permeable to K+
A- is used to refer to large ions that are found completely inside of the cell and cannot penetrate the cell membrane.

Concentration (in Millimoles/ Liter) and permeability of Ions Responsible for Membrane Potential in a Resting Nerve
Cell

ION Extracellular Intracellular Relative Permeability

Na+ 150 15 1

K+ 5 150 25-30

A- 0 65 0

Check out this YouTube video if you want to know more about how the Na+/K+ pump and how the membrane potential works.
www.youtube.com/watch?v=iA-Gdkje6pg

How To Calculate A Membrane Potential


The calculation for the charge of an ion across a membrane, The Nernst Potential, is relatively easy to calculate. The equation is as
follows: (RT/zF) log([X]out/[X]in). RT/F is approximately 61, therefore the equation can be written as
(61/z) ln([X]out/[X]in)
R is the universal gas constant (8.314 J.K-1.mol-1).
T is the temperature in Kelvin (°K = °C + 273.15).
z is the ionic charge for an ion. For example, z is +1 for K+, +2 for Mg2+, -1 for F-, -1 for Cl-, etc. Remember, z does not have a
unit.
F is the Faraday's constant (96485 C.mol-1).
[X]out is the concentration of the ion outside of the species. For example the molarity outside of a neuron.
[X]in is the concentration of the ion inside of the species. For example, the molarity inside of a neuron.
The only difference in the Goldman-Hodgkin-Katz equation is that is adds together the concentrations of all permeable ions as
follows
(RT/zF) log([K+]o+[Na+]o+[Cl-]o /[K+]i+[Na+]i+[Cl-]i)

7.7.3 https://chem.libretexts.org/@go/page/41451
Figure 3. (Clockwise From Upper Left) 1) The charges are equal on both sides; therefore the membrane has no potential. 2)There is
an unbalance of charges, giving the membrane a potential. 3) The charges line up on opposite sides of the membrane to give the
membrane its potential. 4) A hypothetical neuron in the human body; a large concentration of potassium on the inside and sodium
on the outside.

References
1. Kaiser, Chris A., et al. Molecular Cell Biology. 6th ed. New York. W. H. Freeman, 2007.
2. Orians, Gordon H., et al. Life: The Science of Biology. 8th ed. Gordonville, VA. Sinaver Associates, Inc., 2008.,
3. Petrucci, Ralph H., et al. General Chemistry: Principles and Modern Applications. 9th ed. New Jersey. Pearson Education
International, 2007.Fuel Cells#
4. Sherwood, Lauralee. Human Physiology: From Cells to Systems (International Edition). International ed ed. New York: Brooks
Cole, 2009. Print.
5. Hietler, W.J.. "Membrane Potential Tutorial." St. Andrews Biology Dept.. St. Andrews University., 13 Aug. 2007. Web. 24 May
2010. <http:/http://www.st-andrews.ac.uk/~wjh/neurotut/

Problems
1. List the following in order from highest to lowest permeability. A-, K+, Na+
2. Which of the following statement is NOT true?
a. The membrane potential usually requires a minimal difference of electrocharges across the membrane
b. Membrane impermeability plays a role in membrane potentials.
c. Membrane potential exists in all cellular structures, except for neurons.
d. The active transports play a vital role in membrane potentials.
3. What would be the equilibrium potential for the ion K+ be if [K+]in= 5mM and [K+]in=150mM?
4. True or false: At resting membrane potential, the inside of the membrane is slightly negatively charged while the outside is
slightly positively charged.

Answers:
1. K+ > Na+ > A-
2. Answer c.)is not true; membrane potential exists in neurons and is responsible for action potential propagation in neurons.
3. Ek+ = (61/z) log([K+]out/[K+]in) = (61/1) log([5mM]/[150mM]) = -90mV

7.7.4 https://chem.libretexts.org/@go/page/41451
z=1
4. True. The resting membrane potential is negative as a result of this disparity in concentration of charges.

Contributors
Dan Chong, Matt Klingler (UCD)

7.7: Membrane Potential is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

7.7.5 https://chem.libretexts.org/@go/page/41451
7.E: Electrochemistry (Exercises)
Q7.1
For an electrochemical cell driven by the reaction:
− +
Cl + 2 Tl → 2 Cl + 2 Tl (7.E.1)
2

Determine the electromotive force of the cell under standard conditions. Standard reduction potentials for the half cells are given below
− −
Cl (g) + 2 e → 2 Cl (7.E.2)
2

with Eo= +1.36 V


+ −
Tl +e → Tl (7.E.3)

with Eo= -0.336 V

S7.1
The change in gibbs free energy is an extrinsic quality that can be used to determine the EMF of the cell based on the SRPs of the half-
cells. Because the number of electrons exchanged in each reaction are known, we can cancel out the Faraday constants in the equation to
relate Gibbs' free energy directly to the EMF of the cell.
ΔGTot = ΔG1 + ΔG2 → EototntotF = Eo1n1F + Eo2n2F → Eotot = (Eo1n1 + Eo2n2)/ntot = (1.36V*(2) -0.336V(1))/2 = 2.38 V

Q7.8
For the following electrochemical cells, the value of K is given. Determine the change in Gibbs free energy for the reaction in the cell
described, and determine the standard potential of the cell.
Cu2+ + Sn -> Sn2+ + Cu; K = e38.8
2Li + Hg2+ -> 2Li+ + Hg; K = e381
Au3+ + 3Zn -> 3Zn+ + Au; K = e264

Q7.9
Refer to the reaction below for the following question.
4+ 2+ 2+
Zn + S n ⇌ Zn + Sn (7.E.4)

Calculate the E° for the specified reaction at 298 K.

S7.9
First, identify the anode and the cathode:
Anode:
Cathode:
With this information established, you can then calculate the E° using the half-reactions.

Q7.9
2+ 2+
Cu + 2T l → C u + 2T l (7.E.5)

If the equilibrium constant is 8.66 × 10 , what is the electric potential of the cell?
22

Q7.10a
In an experiment using standard conditions, you find the cell emf to be 0.0428 V for the following cell:
+ +
Ag(s) |Ag ||Ag |Ag(s) (7.E.6)

If the concentration of Ag+ is 4.5 M at the cathode, calculate the concentration of Ag+ at the anode.

7.E.1 https://chem.libretexts.org/@go/page/41333
S7.10a
Firstly, note the half reaction.
+ −
Ag +e → Ag (7.E.7)

Then solve using the Nernst equation, noting that E∘ is 0 for concentration cells (same electrodes with different concentrations).

RT [Anode]
o
E =E − ln (7.E.8)
vF [C athode]

−1 ∘
(8.3145 J mol )(298 K) X
0.0428 = 0 − ⋅ ln (7.E.9)
(1)(96, 485J) 4.5M

X
0.0428 = −0.0257 ⋅ ln (7.E.10)
4.5M

−1.665 = ln(x) − ln(4.5) (7.E.11)

ln(4.5) + (−1.665) = ln(x) (7.E.12)

−0.161 ln(x)
e =e (7.E.13)
+
X = 0.85M → [Ag ] = 0.85M (7.E.14)
–––––––––––––––

Click here for more information on concentration cells.

Q7.10b
A concentration cell contains I and I . What is the pressure of I if the cell emf is -.05 V and the pressure of \(I_2\)is 3.75 bar?
2

S7.10b
2 I(g) ⇌ I2(g) (7.E.15)

−.0257V 3.75bar
E = ln → x = 13.0bar (7.E.16)
2 x

Q7.14
Cu(s)|Cu2+ (0.5M) || Cu2+(aq)|Cu(s)
Calculate the electric potential difference for the above concentration cell

Q7.14

Using the above concentration cell, determine the emf. Assume the temperature is at 25°C.

S7.14

The overall reaction is:

The emf of the cell is dependent upon the concentrations of Co2+ at the anode and at the cathode.

Q7.16
Referring to the reaction:

and taking into account that at 298 K, the E°' is -0.219 V. Use this information to determine the E' at pH = 3. Note: Both FAD and FADH2
have equal moles of concentration.

7.E.2 https://chem.libretexts.org/@go/page/41333
S7.16

Q7.15
Calculate E for the following reaction:
[(Zn(s)+Cu^{2+}(aq) \rightleftharpoons Cu(s) + Zn^{2+}(aq)\]

S7.15
Zn(s) → 2 e

+ Zn
2+
(aq) -Anode [Zn2+]=.25M
Cu
2+
(aq) + 2 e

→ C u(s) -Cathode [Cu2+]=.35M

E = Ecathode − Eanode = .342V − (−.762V ) = 1.104V (7.E.17)

2+
.0257V Cu 0.0257V .35M

E =E − ln = 1.10V − ln = 1.08V (7.E.18)
2+
2 Zn 2 .25M

Q7.14
Co(s)|Co2+ (0.05M) ||Co2+(aq) (0.1M)|Co(s)
Calculate the emf for the above concentration cell.

Q7.21
To breakdown consumed alcohol, our body converted ethanol into an acetyldehyde by forming a redox reaction with N AD . +

+ +
ethanol + N AD → acetaldehyde + N ADH + H (7.E.19)

Determine ΔrG°' of reaction at 298 K and pH = 7 (Hint: Set up half reactions and find E°' of the whole reaction.)

S7.21
+ − ∘′
ethanol → acetaldehyde + 2 H + 2e E = +0.197V (7.E.20)

+ + − ∘′
N AD +H + 2e → N ADH E = −0.320V (7.E.21)

+ + ∘′
Overall : ethanol + N AD → acetaldehyde + N ADH + H E = −0.123 (7.E.22)

C
∘′ ∘′
ΔrG = −vF ΔE = −(2)(96500 )(−0.123V ) = 23739J = 23.7kJ/mol (7.E.23)
mol

Q7.21
The hydrolysis of ATP with the following concentrations:

ATP4- + H2O ADP3- + H+ + HPO4²-


10mM 0.5mM pH 6 5mM

Calculate the change in biochemical standard Gibbs energy of reaction.

Q7.22
Calculate the equilibrium constant and ∆rGo' for hydrolysis of ATP at 25.0oC:
Glucose + ATP → ADP + glucose-6-phosphate (7.E.24)

[Glucose]initial = 0.05M
[Glucose]equilibrium = 0.01M

7.E.3 https://chem.libretexts.org/@go/page/41333
S7.22
KC = (0.04M/1M)/(0.01M/1M)=4.0
delta GO = -8.3145J/Kmole *(25+273.15)K*ln 4.0=-3.44 *103kJ/mol

Q7.23
Calculate the ΔrGº/ of a certain biochemical reaction that involves the exchange of 2 electrons and where Eº/ = 3.25 V

S7.23
∘′ ∘′
Δr G = −vF E (7.E.25)

∘′ −1
Δr G = −(2)(96500C mol )(3.25V ) (7.E.26)

∘′ −1
Δr G = −627kJ mol (7.E.27)

Q7.25
In intestinal epithelial cells, transport of lactose across a membrane is coupled with the transport of 2 H+ ions from a high to low
concentration.


Δr G = 16.2kJ ψ = −0.15V (7.E.28)

How much will the pH change across the membrane at 298 K and 1 atm be if one lactose molecule is transported?
Hint: What is the change in Gibbs Free Energy for the movement of 2H+ across this membrane? Write this as an expression as a function of
change in pH.

S7.25
When we consider the chemical potential of a single H+
∘ +
μH + = μ +
+ RT ln[ H ] + F zψ (7.E.29)
H

z = the ion charge


F = Faraday constant
\psi = electrical potential
+
pH = −log[ H ] (7.E.30)

+ +
ln[ H ] = −2.3log[ H ] = −2.3pH (7.E.31)


μH + = μ + − 2.3RT pH + F ψ (7.E.32)
H

Change in Gibbs free energy is defined as the initial and final chemical potentials, or:

Δr G = μH + (low) − μH + (high) (7.E.33)

= −2.3RT ΔpH + F Δψ (7.E.34)

As a function of change in pH, the equation can be written as


Δr G − F Δψ
ΔpH = (7.E.35)
−2.3RT

Since this equation describes chemical potential across a membrane for a H+ ions, the sign for change in Gibb's free energy should be
opposite that of the transport lactose, or -16.2 KJ. Setting change in electric potential as -0.15 V and converting 12.7 KJ to J
C
−16200J − (96500 )(−0.15V )
mol
ΔpH = (7.E.36)
J
(298K)(−2.3)(8.3145 )
K⋅mol

= 0.30 (7.E.37)

Q7.28
You are studying muscle cells. You notice inside the muscle there is 255 mM of Li+ and 16 mM of Rb+ and there is 58 Mm of Li+ and
249 mM of Rb+ outside the cell. From this information, decide if the membrane is more permeable to Li+ or Rb+.

7.E.4 https://chem.libretexts.org/@go/page/41333
S7.28

=-0.075

=0.0407
So Rb+ has a greater magnitude so the membrane is more permeable to it

Q7.29
Calculate the membrane potential of a cell that has [K+]out = 8 and [K+]in = 200. Assume standard temperature of 298 K.

Q7.30
Find E° for:
Cu²+ + e- -> Cu+
K+ + e- -> K
How can E° be used to find Ksp?

Q7.30a
Given the values of E° for the following reactions:
Ag+ + e- → Ag E° = +0.851 V
AgCl + e- → Ag + Cl- E° = +0.222 V
Determine the solubility product (Ksp) of AgCl at 23.5°C.

Q7.30b
Given the half-reduction reaction:
3+ −
Fe + 3e → Fe (7.E.38)

− −
F eC l3 + e → F e + 3C l (7.E.39)

Calculate the Ksp of FeCl3 at 298 K

Q7.31
Given the following reaction at pH 7.0:
+ −
C H3 C OOH + C O2(g) + 2 H + 2e → C H3 C OC OOH + H2 O (7.E.40)

a. Write down the reaction that produces H ion in term of Eo


+

b. Calculate the concentration of C H C OC OOH at 25oC. Given the emf of the cell is -0.70 V.
3

Q7.31
A certain reaction in solution has provides H+ ions and has an Eº of .48V. If a calomel electrode is placed in solution (as part of a pH
meter) and has an Eref of .11 V. Find the pH of the solution.

S7.31
E − Eref
pH = (7.E.41)
0.0591V

7.E.5 https://chem.libretexts.org/@go/page/41333
.37V
pH = (7.E.42)
0.0591V

pH = 6.26 (7.E.43)

Q7.32a
If a since and iron block were to be exposed for an extended period of time, which would rust first?

S7.32a
Zinc is a more electropositive metal with a reduction potential of -0.762 V while iron has a reduction potential of -0.447 V. Since Zinc is
more electropositive, it will be more likely to be oxidized so it will rust first.

Q7.32b
Aluminum is used widely to be a better protective metal than lead because it forms a corrosion resistance layer in most of environments.
Explain this using Standard Reduction Potentials values?

Q7.34
The ΔrS of the Daniell cell is found to be
o
−26.8J K
−1
mol
−1
. Given this value at o
95 C calculate the emf of the cell and the
temperature coefficient.

S7.34
Step 1: (Hint: ΔrS is directly related to temperature coefficient.)
o

o
o
∂E
ΔrS = vF ( )P (7.E.44)

o o −1 −1
∂E ΔrS −26.8J K mol
o −4 −1
ΔrS = vF ( )P = = = −1.38860 × 10 vK (7.E.45)
−1
∂T vF 2(96500C mol )

Step 2:
o o o
o E
∂E E −E 368.15K−1.104V
369.15K 298.15K −4 −1
)P = = = 1.39 × 10 V K (7.E.46)
∂ 368.15 − 298.15K 70K

o
E = 1.094V (7.E.47)
268.15K

Q7.36
Calculate the equilibrium constant for the following reaction:
2+ −
F eC l2 ⇌ F e + 2C l (7.E.48)

(Hint: look up the reduction potential for the half reactions)

S7.36
Half reactions are
2+ −
F e →F e + 2e

ε = 0.44V (this − is − oxidation)
− −
C l2 + 2 e → 2C l

ε = 1.36V (this − is − reduction)

∘ ∘ ∘
ε =ε −ε = 1.36V − 0.44V = 0.92V
cathode anode

∘ (96500C /mol)(0.92V )
vF ε

RT (8.314J/K ∗ mol)(298.15K) 15
K =e =e = 3.58 ∗ 10

S7.36

7.E.6 https://chem.libretexts.org/@go/page/41333
+ −
Anode : 2K → 2 K + 2e

ε = 5.84V
+ −
C athode : 2 H 2e → H2

ε = 0V

∘ ∘ ∘
ε =ε −ε = −5.84V
cathode anode

∘ ∘ ∘ ∘ ∘
¯ ∘ ¯ + ¯ ¯ ¯ +
Δr G = −vF ε = 2 Δf G [ K ] + Δf G [ H2 ] − 2 Δf G [K] − 2 Δf G [ H ]

+∘ 1 ∘
1 ∘ ∘ ∘
+
¯ ¯ ¯ ¯
Δf G [ K ] = − vF ε − Δf G [ H2 ] + Δf G [K] + Δf G [ H ]
2 2
1 1 5
=− (2)(96500C /mol)(−5.84V ) − (0kJ/mol) + (0kJ/mol) + (0kJ/mol) = 5.6 ∗ 10 J/mol
2 2

Q7.36
Do you expect a high Kw value at 276 K or 373 K? Explain.

S7.36
We expect a higher Kw value for 373 K.

Q7.38a
What is the the reaction for a solution with Au3++3e- →Au and Ag++e- →Ag

S7.38a
The first thing we do is look at the standard reduction potentials table. We see that the reduction potential is higher for Au3+ so that will
be the one that gets reduced. This means we reverse the reaction for Ag+.
3+ −
Au + 3e → Au (7.E.49)

+ -
3x(Ag→Ag +e ) we need the charges to add up to zero so we multiply it by 3 so there will be 3 electrons here as well
We end up with:
3+ +
Au + 3Ag → Au + 3Ag (7.E.50)

Q7.38b
The magnitudes of the standard electrode potentials of two metals, A and B, are
A2+ + 2e- → A |E°| = 0.63 V
B2+ + 2e- →B |E°| = 1.24 V
Where the | | notation denotes that only the magnitude (but not the sign) of the E° value is shown. When the half-cells of A and B are
connected, electrons flow from A to B. When A is connected to a SHE, electrons flow from A to SHE. (a) What are the signs for each of
the E°? (b) What is the standard emf of a cell made up of A and B?

Q7.38c
The electrochemical properties of a pair of newly discovered metals, Q and G, are being studied in a lab. Their associated reduction
reactions are shown below. A lab technician collecting the data makes a mistake, and does not record the sign of the standard potential for
the half-cells in their notes. However, he remembers that when he connected the cell containing Q to the standard hydrogen electrode, the
Q electrode was oxidized, and when he connected G to the standard hydrogen electrode, it was reduced. What are the signs of the standard
potentials for the half-cells? When the two half-cells are connected, what is the electromotive force of the complete Q/G cell?
\[Q+ + e- -> Q= 0.38 V \]
\[G+ + e- -> G = 0.72 V \]

Q7.39
Consider a galvanic (voltaic) cell with silver wire suspended in 0.5M Ag+ solution and a cobalt wire in 1.0M Co2+ solution. Write out the
half cell reactions and overall reaction to find the cell's equilibrium constant and observed emf.
Hint: Look at examples of setting up half reactions

7.E.7 https://chem.libretexts.org/@go/page/41333
S7.39
Half cell reactions:
+ −
C athode : 2[Ag (aq) + e → Ag(s)] (7.E.51)

2+ −
Anode : C o(s) → C o (aq) + 2 e (7.E.52)

+ 2+
Overall : 2Ag (aq) + C o(s) → 2Ag(s) + C o (aq) (7.E.53)


E = 0.800 V − (−.277 V ) = +1.077 V (7.E.54)


RT
E = lnK (7.E.55)
vF

J
8.314 (298K)
K⋅mol
+1.077 V = lnK (7.E.56)
C
(2)(26485)
mol

36
K = 2.51 × 10 (7.E.57)

2+
RT [C o ] 0.0257V 1.0M

E =E − ln = +1.077V − ln (7.E.58)
+
v [Ag ] 2 0.5M

E = +1.068 V (7.E.59)

Q7.40a
Answer the following questions about aluminum:
a. Given that the standard electrical potential of Al3+ is -1.662, why does aluminum tarnish in air?
b. Would the the following reaction be spontaneous?
2Al + 3Br2 → Al2 Br6 (7.E.60)

c. Predict the reaction between aluminum and chloride gas.


Hint: look up values for E∘and appropriate half reactions.

S7.40a
(a) Silver tarnishes in air because the reduction potential for oxygen is high enough that it oxidizes aluminum.
∘ ∘ ∘
E =E −E (7.E.61)
cell cathode anode

∘ ∘ ∘
E = Eoxygen −E (7.E.62)
cell aluminum


E = 1.23 V − (−1.662 V ) > 0 (7.E.63)
cell

Since E∘ > 0, then △rG is less than zero (since △rG = -vFE). Thus the reaction is spontaneous and aluminum will be oxidized (tarnish) in
contact with air.

(b) Find the appropriate half reactions to fulfill the equation.


2Al + 3Br2 → Al2 Br6 (1) (7.E.64)
+
3 − ∘
Al + 3e → Al, where E = −1.662 (7.E.65)
Al

− − ∘
Br2 + 2 e → 2Br , where E = +1.087 (7.E.66)
Br

Multiply both sides by the appropriate integer to fulfill equation (1).


+
3 − ∘
−2 times Al + 3e → Al, where E = −1.662 (7.E.67)
Al

− − ∘
3 times Br2 + 2 e → 2Br , where E = +1.087 (7.E.68)
Br

Switching around the half reactions you should achieve the following, which matches up to the equation (1).
+
− 3 − −
2Al + 3Br2 + 6 e → 2Al + 6Br + 6e (7.E.69)


Use the equation to find the E . Keep in mind cathodes are reduced in the half reaction (you want the electrons to be on the left hand side
of the half reaction after balancing) and anodes are oxidized in the half reaction (electrons on the right). Also note how standard reduction
potential is an intrinsic property so there is no need to multiply by molar coefficients represented in the balanced equation.

7.E.8 https://chem.libretexts.org/@go/page/41333
∘ ∘ ∘
E =E −E (7.E.70)
cell cathode anode

∘ ∘ ∘
E =E −E (7.E.71)
cell bromide aluminum


E = (1.087 V ) − (−1.662 V ) > 0 (7.E.72)
cell


E = 2.749 V (7.E.73)
cell

Since E∘ > 0, then △rG is less than zero (since △rG = -vFE). Thus the reaction is spontaneous.

(c)2Al + 3Cl2 → 2AlCl3 since Al3+ is aluminum's most stable form and Cl-. Then it's a matter of balancing the equation.
Check here for additional information on Redox Chemistry.

Q7.40b
Using the electrochemical principle to explain how the reduction potential for oxygen is not adequate to oxidize the gold metal.

Q7.40c
Answer the following questions for the half-reaction below and E°[Pt|O2,H2O] = 1.229 V.
Fe2+(aq) + e- → Fe(s) E° = -0.447 V
(a) Predict whether or not iron will rust in air.
(b) Will the following half-reaction have a positive E°? (E°[Pt|Fe3+,Fe2+] = 0.771 V)
Fe(s) → Fe3+(aq) + 3e-
(c) Predict the reaction between iron and bromine gas (E°[Br2] = 1.087 V)

Q7.40d
The standard reduction potential for Platinum, Pt, is as follows:
2+ − ∘
Pt + 2e → P t(s) E = +1.2V (7.E.74)

a. Platinum is often used as a metal in fine jewelry because it does not tarnish noticeably. Why is this the case?
b. Based on the above standard reduction potential, will the following reaction be spontaneous,
2+ 2− 2−
Pt + Cl → P tC l (7.E.75)
4

given that
2− − − ∘
P tC l + 2e → P t(s) + 4C l E = +0.73V (7.E.76)
4

c.) Gold is also used as a metal in fine jewelry. Based on your answer to part (a.) and the standard reduction potential for gold Au3+, Eo=
1.69, does gold tarnish?

S7.40d
a.) Platinum does not tarnish, or oxidize, very quickly in air since its reduction potential is only slightly lower than that of oxygen. Thus,
any oxidation which does occur does so very slowly.
+ − ∘
O2 + 4 H + 4e → H2 O E = +1.229V (7.E.77)

b.) If the reaction has a positive standard reduction potential, then it is spontaneous. The reaction can be broken up into two half reactions
whose values are provided.
− 2− − ∘
P t(s) + 4C l → P tC l + 2e E = −0.73V (7.E.78)
4

2+ − ∘
Pt + 2e → P t(s) E = +1.2V (7.E.79)

2+ 2− 2−
Pt + Cl → P tC l (7.E.80)
4

so the standard reduction potential is:


o
E = −0.73V + 1.2V = +0.47V (7.E.81)

Thus, the reaction is spontaneous.

7.E.9 https://chem.libretexts.org/@go/page/41333
c.) No. Since the standard reduction potential of gold is greater than that of oxygen, oxygen cannot spontaneously react with gold and
oxidize it.

Q7.41
In figure 7.1 we observe a voltaic cell. Here the anode is oxidized (loss of electrons) and the cathode is reduced (gains electrons). The
Zinc metal loses electron and Zn (s) becomes Zn 2+ ions. Explain how the Cu2+ ions become solid Copper.

Q7.42a
In the following reaction at 298K,
+
C a(s) + 2 H →← C a(aq) + H2(g) (7.E.82)

Calculate the pressure of H2 in bar to maintain equilibrium. The concentration of \([Ca^{2+}]=0.026M and the pH of the solution is
buffered at 1.60 pH.

S7.42a
Step1:
+2 +2 −
Anode : C a → Ca + 2e (7.E.83)

+ −
C athode : 2 H + 2e → H2 (7.E.84)

o
E = 0V − (−0.34) = 0.34 (7.E.85)

Step 2: EMF calculation


+2
0.0257V [C a ] PH2
o
E =E − ln (7.E.86)
+ 2
v [H ]

+ −1.60 −2
F rom the pH , [ H ] = 10 = 2.51 × 10 M

At E=0,

0.0257V (0.026)PH 2
0 = 0.34 − ln (7.E.87)
−2 2
2 (2.51 × 10 )

9
PH2 = 7.5 × 10 bar (7.E.88)

Q7.42b
For the reaction at 300K:
+ 2+
M g(s) + 2 H (aq) ⇌ M g (aq) + H2 (g) (7.E.89)

2+.
Given pressure following the reaction is 296.07 atm, and molarity of acid used is 0.025M. Compute the molarity of the Mg

Q7.44a
o
For the ion Sn2+
, use the electric potential table to determine the value of Δf G
¯
.

S7.44a
2+ −
Sn + 2e → S n(s) (7.E.90)
(aq)

o
E = −0.137 (7.E.91)

2+ −
Anode : S n(s) → S n + 2e

+ −
C athode : 2 H + 2e → H2(g)
(aq)

Overall Equation,
+ 2+ o
S n(s) + 2 H → Sn + H2 E = 0.137V (7.E.92)
(aq) (aq)

o o o o
o o ¯ 2+ ¯ ¯ ¯ +
ΔrG = −vF E = Δf G [S n ] + Δf G [ H2(g) ] − Δf G [S n(s) ] − 2 Δf G [ H ] (7.E.93)
(aq) (aq)

7.E.10 https://chem.libretexts.org/@go/page/41333
o
−1 ¯ 2+ −1 −1 −1
−(2)(96500C mol )(0.137V ) = Δf G [S n ] + 0kj mol − 0kj mol − 2(0kj mol ) (7.E.94)
(aq)

o
¯ 2+ 4 −1
Δf G [S n ] = −2.6 × 10 J mol (7.E.95)
(aq)

Q7.44b

Calculate the Δ f
Ḡ for K +
in the following reaction:
+ +
2K + 2 H → 2K + H2 (7.E.96)

Q7.46a
Determine the standard emf Eo for the following reaction (at 298 K and pH 0)
I-(aq) + H+(aq) + Cr2O72- (aq) → I2(s) + Cr3+(aq) + H2O(l)
Hint: Use standard change in free energy of formation and use the relationship between Go and Eo

S7.46a
− −
oxidation : 6 I (aq) → 3 I2 + 6 e (7.E.97)

− + −2 3+
reduction : 6 e + 14 H (aq) + C r2 O7 (aq) → 2C r (aq) + 7 H2 O(l) (7.E.98)

− + −2 3+
Overall : 6 I (aq) + 14 H (aq) + C r2 O7 (aq) → 2C r (aq) + 7 H2 O + 3 I2 (s) (7.E.99)

∘ ∘ 3+ ∘ ∘ ∘ − ∘ + ∘ 2−
Δr G = 2 Δf G (C r ) + 3 Δf G (I2 ) + 7 Δf G (H2 O) − 6 Δf G (I ) − 14 Δf G (H ) − Δf G (C r2 O7 ) (7.E.100)

kJ kJ

Δr G = [2(−214.2) + 3(0) + 7(−237.2) − 6(51.57) − 14(0) − (−1301.1)] = −1097.12 (7.E.101)
mol mol

kJ 1000J
∘ −1097.12 ×
Δr G mol 1kJ

E =− =− = +1.89V (7.E.102)
vF C
(6)(96500 )
mol

Q7.46b

Calculate the ε for the combustion of ethanol given that Δ

f
¯
Gethanol = −174.8kJ/mol .

S7.46b
Overall : C2 H5 OH + 3 O2 → 2C O2 + 3 H2 O

+ −
Anode : C2 H5 OH + 3 H2 O → 2C O2 + 12 H + 12 e

+ −
C athode : 3 O2 + 12 H + 12 e → 6 H2 O

∘ ∘ ∘ ∘ ∘
¯ ¯ ¯ ¯
Δr G = 2 Δf G [C O2 ] + 3 Δf G [ H2 O] − Δf G [ C2 H5 OH ] − 3 Δf G [ O2 ]

= 2(−394.4kJ/mol) + 3(−237.14kJ/mol) − (−174.8kJ/mol) − 3(0kJ/mol)

= −1325.42kJ/mol

∘ 3
Δr G −1325.42 ∗ 10 J/mol

ε =− =− = 1.145V
vF (12)(96500C /mol)

Q7.47a
Explain the biological significance of cytochromes by comparing the following half reaction in cytochrome c, a, or f vs. an electrode:
+ +
3 − 2
Fe +e → Fe (7.E.103)

S7.47a
For cytochrome c, E∘' = +0.254

For Pt|Fe3+ , Fe2+, E∘ = +0.771


Since cytochrome c has a smaller reduction potential, it makes it easier for for Iron to exist in both the oxidized form Fe3+ or the reduced
form Fe2+. This allows cytochromes to transfer electrons easily between other cytochromes in the electron transport chain. keep in mind
that different cytochromes have slightly different reduction potentials.

7.E.11 https://chem.libretexts.org/@go/page/41333
See here more details about Electron Transfer Proteins.

Q7.47b
Practice
The standard reduction potential for the reduction of copper in aqueous solution is as follows:
2+ − + ∘
Cu +e → Cu E = +0.153V (7.E.104)

In hemocyanin, the primary oxygen carrier in invertebrate blood, however, the standard reduction potential is:

2+ − + ∘
Cu +e → Cu E = +0.540V (7.E.105)

Why do these reduction potentials differ?

S7.47b
The standard reduction potential is used to measure the readiness with which an atom may be reduced at standard conditions (25 oC, 1
atm, reactants at 1M). In biological half-reactions, standard reduction potentials differ because the process cannot occur at standard
concentrations. Eo' occurs at pH 7, as opposed to pH 0, the reference point for electrode half reactions. To obtain a more intuitive
understanding, we may use the Nernst equation:
+
′ RT [C u ]
o o
E =E − ln (7.E.106)
2+
γF [C u ]

At a higher pH, more electrons are found in solution, due to the formation of Hydrogen ions. By le chatelier's principle, this pushes the
above half reaction forward so that the concentration of Cu2+ is greater than that of Cu+. This makes the value produced by the natural log
negative. Thus. Eo' is greater than Eo at higher pH.

7.E: Electrochemistry (Exercises) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

7.E.12 https://chem.libretexts.org/@go/page/41333
CHAPTER OVERVIEW

8: Acids and Bases


8.1: Definitions of Acids and Bases
8.2: Acid-Base Properties of Water
8.3: Dissociation of Acids and Bases
8.4: Diprotic and Polyprotic Acids
8.5: Buffer Solutions
8.6: Acid-Base Titrations
8.7: Amino Acids
8.8: Maintaining the pH of Blood

8: Acids and Bases is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1
8.1: Definitions of Acids and Bases
8.1: Definitions of Acids and Bases is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

8.1.1 https://chem.libretexts.org/@go/page/41452
8.2: Acid-Base Properties of Water
8.2: Acid-Base Properties of Water is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

8.2.1 https://chem.libretexts.org/@go/page/41453
8.3: Dissociation of Acids and Bases
8.3: Dissociation of Acids and Bases is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

8.3.1 https://chem.libretexts.org/@go/page/41454
8.4: Diprotic and Polyprotic Acids
8.4: Diprotic and Polyprotic Acids is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

8.4.1 https://chem.libretexts.org/@go/page/41455
8.5: Buffer Solutions
8.5: Buffer Solutions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

8.5.1 https://chem.libretexts.org/@go/page/41456
8.6: Acid-Base Titrations
8.6: Acid-Base Titrations is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

8.6.1 https://chem.libretexts.org/@go/page/41457
8.7: Amino Acids
8.7: Amino Acids is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

8.7.1 https://chem.libretexts.org/@go/page/41458
8.8: Maintaining the pH of Blood
8.8: Maintaining the pH of Blood is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

8.8.1 https://chem.libretexts.org/@go/page/41459
CHAPTER OVERVIEW

9: Chemical Kinetics
Chemical kinetics is the study of the speed with which a chemical reaction occurs and the factors that affect this speed. This
information is especially useful for determining how a reaction occurs.

Topic hierarchy
9.1: Reaction Rates
9.2: Reaction Order
9.3: Molecularity of a Reaction
9.4: More Complex Reactions
9.5: The Effect of Temperature on Reaction Rates
9.6: Potential Energy Surfaces
9.7: Theories of Reaction Rates
9.8: Isotope Effects in Chemical Reactions
9.9: Reactions in Solution
9.10: Fast Reactions in Solution
9.11: Oscillating Reactions
9.E: Chemical Kinetics (Exercises)

9: Chemical Kinetics is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1
9.1: Reaction Rates
In this Module, the quantitative determination of a reaction rate is demonstrated. Reaction rates can be determined over particular
time intervals or at a given point in time. A rate law describes the relationship between reactant rates and reactant concentrations.
Reaction rates are usually expressed as the concentration of reactant consumed or the concentration of product formed per unit
time. The units are thus moles per liter per unit time, written as M/s, M/min, or M/h. To measure reaction rates, chemists initiate the
reaction, measure the concentration of the reactant or product at different times as the reaction progresses, perhaps plot the
concentration as a function of time on a graph, and then calculate the change in the concentration per unit time.

Figure 9.1.1 The Progress of a Simple Reaction (A → B). The mixture initially contains only A molecules (purple). Over time, the
number of A molecules decreases and more B molecules (green) are formed (top). The graph shows the change in the number of A
and B molecules in the reaction as a function of time over a 1 min period (bottom).
The progress of a simple reaction (A → B) is shown in Figure 9.1.1; the beakers are snapshots of the composition of the solution at
10 s intervals. The number of molecules of reactant (A) and product (B) are plotted as a function of time in the graph. Each point in
the graph corresponds to one beaker in Figure 9.1.1. The reaction rate is the change in the concentration of either the reactant or the
product over a period of time. The concentration of A decreases with time, while the concentration of B increases with time.
Δ[B] Δ[A]
rate = =− (9.1.1)
Δt Δt

Square brackets indicate molar concentrations, and the capital Greek delta (Δ) means “change in.” Because chemists follow the
convention of expressing all reaction rates as positive numbers, however, a negative sign is inserted in front of Δ[A]/Δt to convert
that expression to a positive number. The reaction rate calculated for the reaction A → B using Equation 9.1.1 is different for each
interval (this is not true for every reaction, as shown below). A greater change occurs in [A] and [B] during the first 10 s interval,
for example, than during the last, meaning that the reaction rate is greatest at first.

Reaction rates generally decrease with time as reactant concentrations decrease.

Study 1: The Hydrolysis of Aspirin


We can use Equation 9.1.1 to determine the reaction rate of hydrolysis of aspirin, probably the most commonly used drug in the
world (more than 25,000,000 kg are produced annually worldwide). Aspirin (acetylsalicylic acid) reacts with water (such as water
in body fluids) to give salicylic acid and acetic acid, as shown in Figure 9.1.2.

9.1.1 https://chem.libretexts.org/@go/page/41341
Figure 9.1.2
Because salicylic acid is the actual substance that relieves pain and reduces fever and inflammation, a great deal of research has
focused on understanding this reaction and the factors that affect its rate. Data for the hydrolysis of a sample of aspirin are in Table
9.1.1 and are shown in the graph in Figure 9.1.3. These data were obtained by removing samples of the reaction mixture at the

indicated times and analyzing them for the concentrations of the reactant (aspirin) and one of the products (salicylic acid).
Table 9.1.1 : Data for Aspirin Hydrolysis in Aqueous Solution at pH 7.0 and 37°C*
Time (h) [Aspirin] (M) [Salicylic Acid] (M)

0 5.55 × 10−3 0

2.0 5.51 × 10−3 0.040 × 10−3

5.0 5.45 × 10−3 0.10 × 10−3

10 5.35 × 10−3 0.20 × 10−3

20 5.15 × 10−3 0.40 × 10−3

30 4.96 × 10−3 0.59 × 10−3

40 4.78 × 10−3 0.77 × 10−3

50 4.61 × 10−3 0.94 × 10−3

100 3.83 × 10−3 1.72 × 10−3

200 2.64 × 10−3 2.91 × 10−3

300 1.82 × 10−3 3.73 × 10−3

*The reaction at pH 7.0 is very slow. It is much faster under acidic conditions, such as those found in the stomach.

The average reaction rate for a given time interval can be calculated from the concentrations of either the reactant or one of the
products at the beginning of the interval (time = t0) and at the end of the interval (t1). Using salicylic acid, the reaction rate for the
interval between t = 0 h and t = 2.0 h (recall that change is always calculated as final minus initial) is calculated as follows:
[salicyclic acid]2 − [salicyclic acid]0
rate(t=0−2.0 h)
= (9.1.2)
2.0 h − 0 h
−3
0.040 × 10 M−0 M
−5
= = 2.0 × 10 M/h (9.1.3)
2.0 h

The reaction rate can also be calculated from the concentrations of aspirin at the beginning and the end of the same interval,
remembering to insert a negative sign, because its concentration decreases:
[aspirin]2 − [aspirin]0
rate(t=0−2.0 h)
=− (9.1.4)
2.0 h − 0 h
−3 −3
(5.51 × 10 M) − (5.55 × 10 M)
=− (9.1.5)
2.0 h
−5
= 2 × 10 M/h (9.1.6)

If the reaction rate is calculated during the last interval given in Table 9.1.1(the interval between 200 h and 300 h after the start of
the reaction), the reaction rate is significantly slower than it was during the first interval (t = 0–2.0 h):
[salicyclic acid]300 − [salicyclic acid]200
rate(t=200−300h) = (9.1.7)
300 h − 200 h
−3 −3
(3.73 × 10 M) − (2.91 × 10 M)
=− (9.1.8)
100 h
−6
= 8.2 × 10 M/h (9.1.9)

9.1.2 https://chem.libretexts.org/@go/page/41341
Study 2: The Fermentation of Sucrose
In the preceding example, the stoichiometric coefficients in the balanced chemical equation are the same for all reactants and
products; that is, the reactants and products all have the coefficient 1. Consider a reaction in which the coefficients are not all the
same, the fermentation of sucrose to ethanol and carbon dioxide:

C12 H22 O11 (aq) + H2 O(l) → 4 C2 H5 OH(aq) + 4C O2 (g) (9.1.10)


sucrose

The coefficients indicate that the reaction produces four molecules of ethanol and four molecules of carbon dioxide for every one
molecule of sucrose consumed. As before, the reaction rate can be found from the change in the concentration of any reactant or
product. In this particular case, however, a chemist would probably use the concentration of either sucrose or ethanol because gases
are usually measured as volumes and the volume of CO2 gas formed depends on the total volume of the solution being studied and
the solubility of the gas in the solution, not just the concentration of sucrose. The coefficients in the balanced chemical equation tell
us that the reaction rate at which ethanol is formed is always four times faster than the reaction rate at which sucrose is consumed:
Δ[ C2 H5 OH] 4Δ[sucrose]
=− (9.1.11)
Δt Δt

The concentration of the reactant—in this case sucrose—decreases with time, so the value of Δ[sucrose] is negative. Consequently,
a minus sign is inserted in front of Δ[sucrose] in Equation 9.1.11 so the rate of change of the sucrose concentration is expressed as
a positive value. Conversely, the ethanol concentration increases with time, so its rate of change is automatically expressed as a
positive value.
Often the reaction rate is expressed in terms of the reactant or product with the smallest coefficient in the balanced chemical
equation. The smallest coefficient in the sucrose fermentation reaction (Equation 9.1.10) corresponds to sucrose, so the reaction
rate is generally defined as follows:
Δ[sucrose] 1 Δ[ C2 H5 OH]
rate = − = ( ) (9.1.12)
Δt 4 Δt

Example 9.1.1

Consider the thermal decomposition of gaseous N2O5 to NO2 and O2 via the following equation:
Δ

2 N2 O5 (g) −→ 4NO2 (g) + O2 (g)

Write expressions for the reaction rate in terms of the rates of change in the concentrations of the reactant and each product
with time.
Given: balanced chemical equation
Asked for: reaction rate expressions
Strategy:
A. Choose the species in the equation that has the smallest coefficient. Then write an expression for the rate of change of that
species with time.
B. For the remaining species in the equation, use molar ratios to obtain equivalent expressions for the reaction rate.
Solution
A Because O2 has the smallest coefficient in the balanced chemical equation for the reaction, define the reaction rate as the rate
of change in the concentration of O2 and write that expression.
B The balanced chemical equation shows that 2 mol of N2O5 must decompose for each 1 mol of O2 produced and that 4 mol of
NO2 are produced for every 1 mol of O2 produced. The molar ratios of O2 to N2O5 and to NO2 are thus 1:2 and 1:4,
respectively. This means that the rate of change of [N2O5] and [NO2] must be divided by its stoichiometric coefficient to obtain
equivalent expressions for the reaction rate. For example, because NO2 is produced at four times the rate of O2, the rate of
production of NO2 is divided by 4. The reaction rate expressions are as follows:
Δ[ O2 ] Δ[NO2 ] Δ[ N2 O5 ]
rate = = =−
Δt 4Δt 2Δt

9.1.3 https://chem.libretexts.org/@go/page/41341
Exercise 9.1.1

The key step in the industrial production of sulfuric acid is the reaction of SO2 with O2 to produce SO3.

2S O2(g) + O2(g) → 2S O3(g) (9.1.13)

Write expressions for the reaction rate in terms of the rate of change of the concentration of each species.
Answer
Δ[ O2 ] Δ[SO2 ] Δ[SO3 ]
rate = − =− =
Δt 2Δt 2Δt

Example 9.1.2

Using the reaction shown in Example 9.1.1, calculate the reaction rate from the following data taken at 56°C:

2 N2 O5(g) → 4N O2(g) + O2(g) (9.1.14)

Time (s) [N2O5] (M) [NO2] (M) [O2] (M)

240 0.0388 0.0314 0.00792

600 0.0197 0.0699 0.0175

Given: balanced chemical equation and concentrations at specific times


Asked for: reaction rate
Strategy:
A. Using the equations in Example 9.1.1, subtract the initial concentration of a species from its final concentration and
substitute that value into the equation for that species.
B. Substitute the value for the time interval into the equation. Make sure your units are consistent.
Solution
A Calculate the reaction rate in the interval between t1 = 240 s and t2 = 600 s. From Example 9.1.1 , the reaction rate can be
evaluated using any of three expressions:
Δ[ O2 ] Δ[NO2 ] Δ[ N2 O5 ]
rate = = =−
Δt 4Δt 2Δt

Subtracting the initial concentration from the final concentration of N2O5 and inserting the corresponding time interval into the
rate expression for N2O5,
Δ[ N2 O5 ] [ N2 O5 ]600 − [ N2 O5 ]240
rate = − =−
2Δt 2(600 s − 240 s)

B Substituting actual values into the expression,


0.0197 M − 0.0388 M
−5
rate = − = 2.65 × 10 M/s
2(360 s)

Similarly, NO2 can be used to calculate the reaction rate:


Δ[NO2 ] [NO2 ]600 − [NO2 ]240 0.0699 M − 0.0314 M
−5
rate = = = = 2.67 × 10 M/s (9.1.15)
4Δt 4(600 s − 240 s) 4(360 s)

Allowing for experimental error, this is the same rate obtained using the data for N2O5. The data for O2 can also be used:
Δ[ O2 ] [ O2 ]600 − [ O2 ]240 0.0175 M − 0.00792 M −5
rate = = = = 2.66 × 10 M/s (9.1.16)
Δt 600 s − 240 s 360 s

Again, this is the same value obtained from the N2O5 and NO2 data. Thus, the reaction rate does not depend on which reactant
or product is used to measure it.

9.1.4 https://chem.libretexts.org/@go/page/41341
Exercise 9.1.2

Using the data in the following table, calculate the reaction rate of SO2(g) with O2(g) to give SO3(g).

2S O2(g) + O2(g) → 2S O3(g) (9.1.17)

Time (s) [SO2] (M) [O2] (M) [SO3] (M)

300 0.0270 0.0500 0.0072

720 0.0194 0.0462 0.0148

Answer 9.0 × 10−6 M/s

Instantaneous Rates of Reaction


The instantaneous rate of a reaction is the reaction rate at any given point in time. As the period of time used to calculate an
average rate of a reaction becomes shorter and shorter, the average rate approaches the instantaneous rate. Comparing this to
calculus, the instantaneous rate of a reaction at a given time corresponds to the slope of a line tangent to the concentration-versus-
time curve at that point—that is, the derivative of concentration with respect to time.
−Δ[R] d[R]
rate = lim =− (9.1.18)
Δt→0 Δt dt

The distinction between the instantaneous and average rates of a reaction is similar to the distinction between the actual speed of a
car at any given time on a trip and the average speed of the car for the entire trip. Although the car may travel for an extended
period at 65 mph on an interstate highway during a long trip, there may be times when it travels only 25 mph in construction zones
or 0 mph if you stop for meals or gas. The average speed on the trip may be only only 50 mph, whereas the instantaneous speed on
the interstate at a given moment may be 65 mph. Whether the car can be stopped in time to avoid an accident depends on its
instantaneous speed, not its average speed. There are important differences between the speed of a car during a trip and the speed of
a chemical reaction, however. The speed of a car may vary unpredictably over the length of a trip, and the initial part of a trip is
often one of the slowest. In a chemical reaction, the initial interval typically has the fastest rate (though this is not always the case),
and the reaction rate generally changes smoothly over time.

Chemical kinetics generally focuses on one particular instantaneous rate, which is the
initial reaction rate, t = 0. Initial rates are determined by measuring the reaction rate at
various times and then extrapolating a plot of rate versus time to t = 0.

9.1: Reaction Rates is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

9.1.5 https://chem.libretexts.org/@go/page/41341
9.2: Reaction Order
The kinetic theory of gases can be used to model the frequency of collisions between hard-sphere molecules, which is proportional
to the reaction rate. Most systems undergoing a chemical reaction, however, are much more complex. The reaction rates may be
dependent on specific interactions between reactant molecules, the phase(s) in which the reaction takes place, etc. The field of
chemical kinetics is thus by-and-large based on empirical observations. From experimental observations, scientists have established
that reaction rates almost always have a power-law dependence on the concentrations of one or more of the reactants. In the
following sections, we will discuss different power laws that are commonly observed in chemical reactions.

0
th
Order Reaction Kinetics
Consider a closed container initially filled with chemical species A . At t = 0 , a stimulus, such as a change in temperature, the
addition of a catalyst, or irradiation, causes an irreversible chemical reaction to occur in which A transforms into product B :

aA ⟶ bB (9.2.1)

The rate that the reaction proceeds, r, can be described as the change in the concentrations of the chemical species with respect to
time:
1 d [A] 1 d [B]
r =− = (9.2.2)
a dt b dt

where [X] denotes the molar concentration of chemical species X with units of mol
3
.
L

Let us first examine a reaction A ⟶ B in which the reaction rate, r, is constant with time:
d [A]
r =− =k (9.2.3)
dt

mol
where k is a constant, also known as the rate constant with units of 3
. Such reactions are called zeroth order reactions because
m s
the reaction rate depends on the concentrations of species A and B to the 0 th
power. Integrating [A] with respect to T , we find that
[A] = −kt + c1 (9.2.4)

At t = 0 , [A] (0) = [A] . Plugging these values into the equation, we find that c
0 1 = [A]
0
. The final form of the equation is:
[A] = [A] − kt (9.2.5)
0

A plot of the concentration of species A with time for a 0 order reaction is shown in Figure 9.2.1, where the slope of the line is
th

−k and the y -intercept is [A] . Such reactions in which the reaction rates are independent of the concentrations of products and
0

reactants are rare in nature. An example of a system displaying 0 order kinetics would be one in which a reaction is mediated by
th

a catalyst present in small amounts.

1
st
Order Reaction Kinetics
Experimentally, it is observed than when a chemical reaction is of the form

∑ νi Ai = 0 (9.2.6)

the reaction rate can be expressed as


νi
r =k ∏ [ Ai ] (9.2.7)

reactants

where it is assumed that the stoichiometric coefficients νi of the reactants are all positive. Thus, or a reaction A ⟶ B , the
reaction rate depends on [A] raised to the first power:
d [A]
r = = −k [A] (9.2.8)
dt

9.2.1 https://chem.libretexts.org/@go/page/41342
Figure 9.2.1 : Plots of [A] (solid line) and [B] (dashed line) over time for a 0
th
order reaction.
1
For first order reactions, k has the units of . Integrating and applying the condition that at t =0 s , [A] = [A]
0
, we arrive at the
s
following equation:
−kt
[A] = [A] e (9.2.9)

Figure 9.2.2 displays the concentration profiles for species A and B for a first order reaction. To determine the value of k from

Figure 9.2.2 : Plots of [A] (solid line) and [B] (dashed line) over time for a 1 order reaction.
st

experimental data, it is convenient to take the natural log of Equation 9.2.9:


ln ([A]) = ln ([A] ) − kt (9.2.10)
0

For a first order irreversible reaction, a plot of ln ([A]) vs. t is straight line with a slope of −k and a y -intercept of ln ([A] ) . 0

2
nd
Order Reaction Kinetics
Another type of reaction depends on the square of the concentration of species A - these are known as second order reactions. For a
second order reaction in which 2A ⟶ B , we can write the reaction rate to be

1 d [A] 2
r =− = k[A] (9.2.11)
2 dt

3
m
For second order reactions, k has the units of . Integrating and applying the condition that at t = 0 s , [A] = [A] , we arrive0
mol ⋅ s
at the following equation for the concentration of A over time:
1
[A] = (9.2.12)
1
2kt +
[A]
0

Figure 9.2.3 shows concentration profiles of A and B for a second order reaction. To determine k from experimental data for

9.2.2 https://chem.libretexts.org/@go/page/41342
Figure 9.2.3 : Plots of [A] (solid line) and [B] (dashed line) over time for a 2 nd
order reaction.
second-order reactions, it is convenient to invert Equation 9.2.12:
1 1
= + 2kt (9.2.13)
[A] [A]
0

A plot of 1/ [A] vs. t will give rise to a straight line with slope k and intercept 1/[A] . 0

Second order reaction rates can also apply to reactions in which two species react with each other to form a product:
k
A+B ⟶ C (9.2.14)

In this scenario, the reaction rate will depend on the concentrations of both A and B to the first order:
d [A] d [B]
r =− =− = k [A] [B] (9.2.15)
dt dt

To integrate the above equation, we need to write it in terms of one variable. Since the concentrations of A and B are related to
each other via the chemical reaction equation, we can write:
[B] = [B] − ([A] − [A]) = [A] + [B] − [A] (9.2.16)
0 0 0 0

d [A]
= −k [A] ([A] + [B] − [A] ) (9.2.17)
0 0
dt

We can then use partial fractions to integrate:


d [A] 1 d [A] d [A]
kdt = = ( − ) (9.2.18)
[A] ([A] + [B] − [A] ) [B] − [A] [A] [B] − [A] + [A]
0 0 0 0 0 0

1 [A] [B]0
kt = ln (9.2.19)
[A]0 − [B]0 [B] [A]0

Figure 9.2.4 displays the concentration profiles of species A , B , and C for a second order reaction in which the initial
concentrations of A and B are not equal.

Figure 9.2.4 : Plots of [A] (solid line), [B] (dashed line) and [C] (dotted line) over time for a 2 nd
order reaction in which the initial
concentrations of the reactants, [A] and [B] , are not equal.
0 0

9.2.3 https://chem.libretexts.org/@go/page/41342
Contributors and Attributions
Mark Tuckerman (New York University)

9.2: Reaction Order is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

9.2.4 https://chem.libretexts.org/@go/page/41342
9.3: Molecularity of a Reaction
In general, it is necessary to experimentally measure the concentrations of species over time in order to determine the apparent rate
law governing the reaction. If the reactions are elementary reactions, (i.e. they cannot be expressed as a series of simpler reactions),
then we can directly define the rate law based on the chemical equation. For example, an elementary reaction in which a single
reactant transforms into a single product, is unimolecular reaction. These reactions follow 1 order rate kinetics. An example of
st

this type of reaction would be the isomerization of butane:


nC4 H10 ⟶ i C4 H10 (9.3.1)

From the chemical reaction equation, we can directly write the rate law as
d [nC4 H10 ]
= −k [nC4 H10 ] (9.3.2)
dt

without the need to carry out experiments.


Elementary bimolecular reactions that involve two molecules interacting to form one or more products follow second order rate
kinetics. An example would be the following reaction between a nitrate molecule and carbon monoxide to form nitrogen dioxide
and carbon dioxide:
N O3 + C O ⟶ N O2 + C O2 (9.3.3)

For the above elementary reaction, we can directly write the rate law as:
d [N O3 ]
= −k [N O3 ] [C O] (9.3.4)
dt

Trimolecular elementary reactions involving three reactant molecules to form one or more products are rare due to the low
probability of three molecules simultaneously colliding with one another.

9.3: Molecularity of a Reaction is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

9.3.1 https://chem.libretexts.org/@go/page/41343
9.4: More Complex Reactions
A major goal in chemical kinetics is to determine the sequence of elementary reactions, or the reaction mechanism, that comprise
complex reactions. For example, Sherwood Rowland and Mario Molina won the Nobel Prize in Chemistry in 1995 for proposing
the elementary reactions involving chlorine radicals that contribute to the overall reaction of O → O in the troposphere. In the
3 2

following sections, we will derive rate laws for complex reaction mechanisms, including reversible, parallel and consecutive
reactions.

Parallel Reactions
Consider the reaction in which chemical species A undergoes one of two irreversible first order reactions to form either species B
or species C :
k1
A → B (9.4.1)

k2

A → C (9.4.2)

The overall reaction rate for the consumption of A can be written as:
d [A]
= −k1 [A] − k2 [A] = − (k1 + k2 ) [A] (9.4.3)
dt

Integrating [A] with respect to t , we obtain the following equation:


−( k1 +k2 )t
[A] = [A] e (9.4.4)
0

d [B]
Plugging this expression into the equation for , we obtain:
dt

d [B]
−( k1 +k2 )t
= k1 [A] = k1 [A] e (9.4.5)
0
dt

Integrating [B] with respect to t , we obtain:


k1 [A]0
−( k1 +k2 )t
[B] = − (e ) + c1 (9.4.6)
k1 + k2

At t = 0 , [B] = 0 . Therefore,
k1 [A]0
c1 = (9.4.7)
k1 + k2

k1 [A]
0 −( k1 +k2 )t
[B] = (1 − e ) (9.4.8)
k1 + k2

Likewise,
k2 [A]
0 −( k1 +k2 )t
[C] = (1 − e ) (9.4.9)
k1 + k2

The ratio of [B] to [C] is simply:


[B] k1
= (9.4.10)
[C] k2

An important parallel reaction in industry occurs in the production of ethylene oxide, a reagent in many chemical processes and
also a major component in explosives. Ethylene oxide is formed through the partial oxidation of ethylene:
k1

2 C2 H4 + O2 ⟶ 2 C2 H4 O (9.4.11)

However, ethylene can also undergo a combustion reaction:


k2

C2 H4 + 3 O2 ⟶ 2 C O2 + 2 H2 O (9.4.12)

9.4.1 https://chem.libretexts.org/@go/page/41344
To select for the first reaction, the oxidation of ethylene takes place in the presence of a silver catalyst, which significantly
increases k compared to k . Figure 9.4.1 displays the concentration profiles for species A , B , and C in a parallel reaction in
1 2

which k > k .
1 2

Figure 9.4.1 : Plots of [A] (solid line), [B] (dashed line) and [C] (dotted line) over time for a parallel reaction.

Consecutive Reactions
Consider the following series of first-order irreversible reactions, where species A reacts to form an intermediate species, I , which
then reacts to form the product, P:
k1 k2

A ⟶ I ⟶ P (9.4.13)

We can write the reaction rates of species A , I and P as follows:


d [A]
= −k1 [A] (9.4.14)
dt

d [I]
= k1 [A] − k2 [I] (9.4.15)
dt

d [P]
= k2 [I] (9.4.16)
dt

As before, integrating [A] with respect to t leads to:


−k1 t
[A] = [A] e (9.4.17)
0

The concentration of species I can be written as


k1 [A]0
−k1 t −k2 t
[I] = (e −e ) (9.4.18)
k2 − k1

Then, solving for [P], we find that:


1 −k1 t −k2 t
[P] = [A] [1 + (k2 e − k1 e )] (9.4.19)
0
k1 − k2

Figure 9.4.2 displays the concentration profiles for species A , I , and P in a consecutive reaction in which k = k . As can be seen 1 2

from the figure, the concentration of species I reaches a maximum at some time, t . Oftentimes, species I is the desired product.
max

Returning to the oxidation of ethylene into ethylene oxide, it is important to note another reaction in which ethylene oxide can
decompose into carbon dioxide and water through the following reaction
5 k3

C2 H4 O + O2 ⟶ 2 C O2 + 2 H2 O (9.4.20)
2

Thus, to maximize the concentration of ethylene oxide, the oxidation of ethylene is only allowed proceed to partial completion
before the reaction is stopped.
Finally, in the limiting case when k 2 ≫ k1 , we can write the concentration of P as
1
−k1 t −k1 t
[P] ≈ [A] {1 + k2 e } = [A] (1 − e ) (9.4.21)
0 0
−k2

Thus, when k 2 ≫ k1 , the reaction can be approximated as A → P and the apparent rate law follows 1 order kinetics. st

9.4.2 https://chem.libretexts.org/@go/page/41344
Figure 9.4.2 : Plots of [A] (solid line), [I] (dashed line) and [P] (dotted line) over time for consecutive first order reactions.

Consecutive Reactions With an Equilibrium


Consider the reactions
k1 k2

A ⇌ I → P (9.4.22)
k−1

We can write the reaction rates as:


d [A]
= −k1 [A] + k−1 [I] (9.4.23)
dt

d [I]
= k1 [A] − k−1 [I] − k2 [I] (9.4.24)
dt

d [P]
= k2 [I] (9.4.25)
dt

The exact solutions of these is straightforward, in principle, but rather involved, so we will just state the exact solutions, which are
[A]
0 −( k1 +K−λ)t/2 −( k1 +K+λ)t/2
[A] (t) = [(λ − k1 + K) e + (λ + k1 − K) e ] (9.4.26)

k1 [A]0
−( k1 +K−λ)t/2 −( k1 +K+λ)t/2
[I] (t) = [e −e ] (9.4.27)
λ

−( k1 +K−λ)t/2 −( k1 +K+λ)t/2
2 1 e e
[P] (t) = 2 k1 k2 [A] [ − ( − )] (9.4.28)
0 2
(k1 + K) −λ
2 λ k1 + K − λ k1 + K + λ

where

K = k2 + k−1 (9.4.29)
−−−−−−−−−−−−−−−
2
λ = √ (k1 − K) − 4 k1 k−1 (9.4.30)

Steady-State Approximations
Consider the following consecutive reaction in which the first step is reversible:
k1 k2

A ⇌ I → P (9.4.31)
k−1

We can write the reaction rates as:


d [A]
= −k1 [A] + k−1 [I] (9.4.32)
dt

d [I]
= k1 [A] − k−1 [I] − k2 [I] (9.4.33)
dt

d [P]
= k2 [I] (9.4.34)
dt

These equations can be solved explicitly in terms of [A], [I], and [P], but the math becomes very complicated quickly. If, however,
k +k
2 ≫ k
−1 (in other words, the rate of consumption of I is much faster than the rate of production of I ), we can make the
1

approximation that the concentration of the intermediate species, I , is small and constant with time:

9.4.3 https://chem.libretexts.org/@go/page/41344
d [I]
≈0 (9.4.35)
dt

Equation 21.22 can now be written as


d [I]
= k1 [A] − k−1 [I] − k2 [I] ≈0 (9.4.36)
ss ss
dt

where [I] is a constant represents the steady state concentration of intermediate species, [I]. Solving for [I] ,
ss ss

k1
[I] = [A] (9.4.37)
ss
k−1 + k2

We can then write the rate equation for species A as


d [A] k1 k1 k2
= −k1 [A] + k−1 [I] = −k1 [A] + k−1 [A] = − [A] (9.4.38)
ss
dt k−1 + k2 k−1 + k2

Integrating,
k1 k2
− t
k−1 + k2
[A] = [A] e (9.4.39)
0

Equation 21.28 is the same equation we would obtain for apparent 1 order kinetics of the following reaction:
st


k

A⟶ P (9.4.40)

where
k1 k2

k = (9.4.41)
k−1 + k2

Figure 9.4.3 displays the concentration profiles for species, A , I , and P with the condition that k2 + k−1 ≫ k1 . These types of
reaction kinetics appear when the intermediate species, I , is highly reactive.

Figure 9.4.3 : Plots of [A] (solid line), [I] (dashed line) and [P] (dotted line) over time for k 2 + k−1 ≫ k1 .

Lindemann Mechanism
Consider the isomerization of methylisonitrile gas, C H 3N C , to acetonitrile gas, C H 3CN :
k
C H3 N C ⟶ C H3 C N (9.4.42)

If the isomerization is a unimolecular elementary reaction, we should expect to see 1 order rate kinetics. Experimentally, st

however, 1 order rate kinetics are only observed at high pressures. At low pressures, the reaction kinetics follow a 2 order rate
st nd

law:
d [C H3 N C ]
2
= −k[C H3 N C ] (9.4.43)
dt

To explain this observation, J.A. Christiansen and F.A. Lindemann proposed that gas molecules first need to be energized via
intermolecular collisions before undergoing an isomerization reaction. The reaction mechanism can be expressed as the following
two elementary reactions

9.4.4 https://chem.libretexts.org/@go/page/41344
k1

A+ M ⇌ A +M (9.4.44)
k−1

k2

A → B (9.4.45)

where M can be a reactant molecule, a product molecule or another inert molecule present in the reactor. Assuming that the
concentration of A is small, or k ≪ k + k , we can use a steady-state approximation to solve for the concentration profile of

1 2 −1

species B with time:



d [A ]
∗ ∗
= k1 [A] [M] − k−1 [ A ] [M] − k2 [ A ] ≈0 (9.4.46)
ss ss
dt

Solving for [A ] ,

k1 [M] [A]

[A ] = (9.4.47)
k2 + k−1 [M]

The reaction rates of species A and B can be written as


d [A] d [B] k1 k2 [M] [A]

− = = k2 [ A ] = = kobs [A] (9.4.48)
dt dt k2 + k−1 [M]

where
k1 k2 [M]
kobs = (9.4.49)
k2 + k−1 [M]

At high pressures, we can expect collisions to occur frequently, such that k −1 [M] ≫ k2 . Equation 21.33 then becomes
d [A] k1 k2
− = [A] (9.4.50)
dt k−1

which follows 1 order rate kinetics.


st

At low pressures, we can expect collisions to occurs infrequently, such that k−1 [M] ≪ k2 . In this scenario, equation 21.33
becomes
d [A]
− = k1 [A] [M] (9.4.51)
dt

which follows second order rate kinetics, consistent with experimental observations.

Equilibrium Approximations
Consider again the following consecutive reaction in which the first step is reversible:
k1 k2

A ⇌ I → P (9.4.52)
k−1

Now let us consider the situation in which k ≪ k and k . In other words, the conversion of I to P is slow and is the rate-
2 1 −1

limiting step. In this situation, we can assume that [A] and [I] are in equilibrium with each other. As we derived before for a
reversible reaction in equilibrium,

k1 [I]
Keq = ≈ (9.4.53)
k−1 [A]

or, in terms of [I],

[I] = Keq [A] (9.4.54)

These conditions also result from the exact solution when we set k 2 ≈0 . When this is done, we have the approximate expressions
from the exact solution:

9.4.5 https://chem.libretexts.org/@go/page/41344
K ≈ k−1 (9.4.55)
−−−−−−−−−−−−−−−−− −−−−−−−−−−−−−−
2 2 2
λ ≈ √ (k1 − k−1 ) + 4 k1 k−1 = √ k + 2 k1 k−1 + k = k1 + k−1 (9.4.56)
1 −1

λ − k1 + K ≈ k1 + k−1 + k1 − k−1 = 2 k1 (9.4.57)

λ + k1 − K ≈ k1 + k−1 + k1 − k−1 = 2 k1 (9.4.58)

k1 + K − λ ≈ k1 + k−1 − k1 − k−1 = 0 (9.4.59)

k1 + K + λ ≈ k1 + k−1 + k1 + k−1 = 2 (k1 + k−1 ) (9.4.60)

and the approximate solutions become


[A]
0 −( k1 +k−1 )t
[A] (t) = [2 k−1 + 2 k1 e ] (9.4.61)
2 (k1 + k−1 )

k1 [A]
0 −( k1 +k−1 )t
[I] (t) = [1 − e ] (9.4.62)
(k1 + k−1 )

In the long-time limit, when equilibrium is reached and transient behavior has decayed away, we find
[I] k1
≡ Keq → (9.4.63)
[A] k−1

Plugging the above equation into the expression for d [P] /dt,
d [P] k1 k2
= k2 [I] = k2 Keq [A] = [A] (9.4.64)
dt k−1

The reaction can thus be approximated as a 1 order reaction


st


k
A⟶ P (9.4.65)

with


k1 k2
k = (9.4.66)
k−1

Figure 9.4.4 displays the concentration profiles for species, A , I , and P with the condition that k ≪ k = k . When k = k , 2 1 −1 1 −1

we expect [A] = [I] . As can be seen from the figure, after a short initial startup time, the concentrations of species A and I are
approximately equal during the reaction.

Figure 9.4.4 : Plots of [A] (solid line), [I] (dashed line) and [P] (dotted line) over time for k 2 ≪ k1 = k−1 .

This page titled 9.4: More Complex Reactions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Mark E.
Tuckerman.
1.20: Complex reaction mechanisms by Mark Tuckerman is licensed CC BY-NC-SA 4.0.

9.4.6 https://chem.libretexts.org/@go/page/41344
9.5: The Effect of Temperature on Reaction Rates
When molecules collide, the kinetic energy of the molecules can be used to stretch, bend, and ultimately break bonds, leading to
chemical reactions. If molecules move too slowly with little kinetic energy, or collide with improper orientation, they do not react
and simply bounce off each other. However, if the molecules are moving fast enough with a proper collision orientation, such that
the kinetic energy upon collision is greater than the minimum energy barrier, then a reaction occurs. The minimum energy
requirement that must be met for a chemical reaction to occur is called the activation energy, E . a

Figure 9.5.1 : In Greek mythology Sisyphus was punished by being forced roll an immense boulder up a hill, only to watch it roll
back down, and to repeat this action forever. If this were a chemical reaction, then it would never be observed, since the reactants
must overcome the energy barrier to get to the other side (products).
The reaction pathway is similar to what happens in Figure 9.5.1. To get to the other end of the road, an object must roll with
enough speed to completely roll over the hill of a certain height. The faster the object moves, the more kinetic energy it has. If the
object moves too slowly, it does not have enough kinetic energy necessary to overcome the barrier; as a result, it eventually rolls
back down. In the same way, there is a minimum amount of energy needed in order for molecules to break existing bonds during a
chemical reaction. If the kinetic energy of the molecules upon collision is greater than this minimum energy, then bond breaking
and forming occur, forming a new product (provided that the molecules collide with the proper orientation).

Figure 9.5.2 : Reaction coordinate diagram for the bimolecular nucleophilic substitution (S N ) reaction between bromomethane
2

and the hydroxide anion. from Wikipedia.


The activation energy (E ), labeled ΔG in Figure 9.5.2, is the energy difference between the reactants and the activated complex,
a

also known as transition state. In a chemical reaction, the transition state is defined as the highest-energy state of the system. If the
molecules in the reactants collide with enough kinetic energy and this energy is higher than the transition state energy, then the
reaction occurs and products form. In other words, the higher the activation energy, the harder it is for a reaction to occur and vice
versa.
Overcoming the energy barrier from thermal energy involves addressing the fraction of the molecules that possess enough kinetic
energy to react at a given temperature. According to kinetic molecular theory, a population of molecules at a given temperature is
distributed over a variety of kinetic energies that is described by the Maxwell-Boltzman distribution law.

9.5.1 https://chem.libretexts.org/@go/page/41345
Figure 9.5.3 : Kinetic energy distributions (similar to Maxwell-Boltzman distributions for velocity) for a gas at two temperatures
and critical energies for overcoming an activation barrier.
The two distribution plots shown here are for a lower temperature T1 and a higher temperature T2. The area under each curve
represents the total number of molecules whose energies fall within particular range. The shaded regions indicate the number of
molecules which are sufficiently energetic to meet the requirements dictated by the two values of Ea that are shown. It is clear from
these plots that the fraction of molecules whose kinetic energy exceeds the activation energy increases quite rapidly as the
temperature is raised. This the reason that virtually all chemical reactions (and all elementary reactions) proceed more rapidly at
higher temperatures.

Arrhenius Equation
By 1890 it was common knowledge that higher temperatures speed up reactions, often doubling the rate for a 10-degree rise, but
the reasons for this were not clear. Finally, in 1899, the Swedish chemist Svante Arrhenius (1859-1927) combined the concepts of
activation energy and the Boltzmann distribution law into one of the most important relationships in physical chemistry:

Take a moment to focus on the meaning of this equation, neglecting the A factor for the time being. First, note that this is another
form of the exponential decay law discussed in the previous section of this series. What is "decaying" here is not the
concentration of a reactant as a function of time, but the magnitude of the rate constant as a function of the exponent –Ea /RT.
And what is the significance of this quantity? Recalling that RT is the average kinetic energy, it becomes apparent that the
exponent is just the ratio of the activation energy Ea to the average kinetic energy. The larger this ratio, the smaller the rate
(hence the negative sign). This means that high temperature and low activation energy favor larger rate constants, and thus speed
up the reaction. Because these terms occur in an exponent, their effects on the rate are quite substantial.

Svante August Arrhenius

Svante August Arrhenius (19 February 1859 – 2 October 1927) was a Swedish scientist, originally a physicist, but often
referred to as a chemist, and one of the founders of the science of physical chemistry. He received the Nobel Prize for
Chemistry in 1903, becoming the first Swedish Nobel laureate, and in 1905 became director of the Nobel Institute where he
remained until his death. The Arrhenius equation, Arrhenius definition of an acid, lunar crater Arrhenius, the mountain of
Arrheniusfjellet and the Arrhenius Labs at Stockholm University are named after him. Today, Arrhenius is best known for his
study published in 1896, on the greenhouse effect.

9.5.2 https://chem.libretexts.org/@go/page/41345
The two plots in Figure 9.5.4 show the effects of the activation energy (denoted here by E ‡ ) on the rate constant. Even a modest
activation energy of 50 kJ/mol reduces the rate by a factor of 108.

Figure 9.5.4 : Arrhenius plots. The logarithmic scale in the right-hand plot leads to nice straight lines.
Looking at the role of temperature, a similar effect is observed. (If the x-axis were in "kilodegrees" the slopes would be more
comparable in magnitude with those of the kilojoule plot at the above right.)

Determining the activation energy


The Arrhenius equation
−Ea /RT
k = Ae (9.5.1)

can be written in a non-exponential form that is often more convenient to use and to interpret graphically (Figure 9.5.4). Taking the
logarithms of both sides and separating the exponential and pre-exponential terms yields
−Ea /RT −Ea /RT
ln k = ln(Ae ) = ln A + ln(e ) (9.5.2)

−Ea −Ea 1
ln k = ln A + =( )( ) + ln A (9.5.3)
RT R T

which is the equation of a straight line whose slope is – E /R. This affords a simple way of determining the activation energy
a

from values of k observed at different temperatures, by plotting ln k as a function of 1/T .

Example 9.5.1: Isomerization of Cyclopropane

For the isomerization of cyclopropane to propene,

the following data were obtained (calculated values shaded in pink):

T, °C 477 523 577 623

1/T, K–1 × 103 1.33 1.25 1.18 1.11

k, s–1 0.00018 0.0027 0.030 0.26

ln k –8.62 –5.92 –3.51 –1.35

9.5.3 https://chem.libretexts.org/@go/page/41345
From the calculated slope, we have
– (Ea/R) = –3.27 × 104 K
Ea=– (8.314 J mol–1 K–1) (–3.27 × 104 K) = 273 kJ mol–1
This activation energy is high, which is not surprising because a carbon-carbon bond must be
broken in order to open the cyclopropane ring. (C–C bond energies are typically around 350
kJ/mol.) This is why the reaction must be carried out at high temperature.

Calculating E without a plot


a

Because the ln k vs.-1/T plot yields a straight line, it is often convenient to estimate the activation
energy from experiments at only two temperatures. The ln A term is eliminated by subtracting the
expressions for the two ln-k terms via the following steps:
Ea
ln k1 = ln A − (9.5.4)
kB T1

at T and
1

Ea
ln k2 = ln A − (9.5.5)
kB T2

at T . By rewriting the second equation:


2

Ea
ln A = ln k2 + (9.5.6)
kB T2

and substitute for ln A into the first equation:


Ea Ea
ln k1 = ln k2 + − (9.5.7)
kB T2 kB T1

This simplifies to:


Ea Ea
ln k1 − ln k2 = − + (9.5.8)
kB T1 kB T2

k1 Ea 1 1
ln =− ( − ) (9.5.9)
k2 kB T1 T2

Example 9.5.2

A widely used rule-of-thumb for the temperature dependence of a reaction rate is that a ten degree rise in the temperature
approximately doubles the rate. This is not generally true, especially when a strong covalent bond must be broken. For a
reaction that does show this behavior, what would the activation energy be?
Solution
Center the ten degree interval at 300 K. Substituting into the above expression yields
(8.314)(ln 2/1) (8.314)(0.693)
Ea = = (9.5.10)
1 1 0.00339 K −1 – 0.00328 K −1

295 305

= (5.76 J mol K ) / (0.00011 K ) = 52400 J mol = 52.4 kJ mol–1


–1 –1 –1 –1

Example 9.5.3

It takes about 3.0 minutes to cook a hard-boiled egg in Los Angeles, but at the higher altitude of Denver, where water boils at
92°C, the cooking time is 4.5 minutes. Use this information to estimate the activation energy for the coagulation of egg
albumin protein.

9.5.4 https://chem.libretexts.org/@go/page/41345
Solution
The ratio of the rate constants at the elevations of Los Angeles and Denver is 4.5/3.0 = 1.5, and the respective temperatures are
373 K and 365 K. With the subscripts 2 and 1 referring to Los Angeles and Denver respectively:

(8.314)(ln 1.5) (8.314)(0.405)


Ea = = (9.5.11)
1 1 −1 −1
0.00274 K – 0.00268 K

365 K 373 K

–1 –1
(3.37 J mo l K )
–1 –1
= = 5740 J mo l = 5.73 kJ mo l (9.5.12)
−5 –1
5.87 × 10 K

Comment: This low value seems reasonable because thermal denaturation of proteins primarily involves the disruption of
relatively weak hydrogen bonds; no covalent bonds are broken (although disulfide bonds can interfere with this interpretation).

The pre-exponential factor


Up to this point, the pre-exponential term, A in the Arrhenius equation, has been ignored because it is not directly involved in
relating temperature and activation energy, which is the main practical use of the equation.

However, because A multiplies the exponential term, its value clearly contributes to the value of the rate constant and thus of the
rate. Recall that the exponential part of the Arrhenius equation expresses the fraction of reactant molecules that possess enough
kinetic energy to react, as governed by the Maxwell-Boltzmann law. This fraction can run from zero to nearly unity, depending on
the magnitudes of E and of the temperature.
a

If this fraction were 0, the Arrhenius law would reduce to


k =A (9.5.13)

In other words, A is the fraction of molecules that would react if either the activation energy were zero, or if the kinetic energy of
all molecules exceeded E — admittedly, an uncommon scenario (although barrierless reactions have been characterized).
a

The Role of Collisions


What would limit the rate constant if there were no activation energy requirements? The most obvious factor would be the rate at
which reactant molecules come into contact. This can be calculated from kinetic molecular theory and is known as the frequency-
or collision factor, Z .

In some reactions, the relative orientation of the molecules at the point of collision is important, so a geometrical or steric factor
(commonly denoted by ρ (Greek lower case rho) can be defined. In general, we can express A as the product of these two factors:
A = Zρ (9.5.14)

Values of ρ are generally very difficult to assess; they are sometime estimated by comparing the observed rate constant with the one
in which A is assumed to be the same as Z . Usually, the more complex the reactant molecules, the lower the steric factors. The
deviation from unity can have different causes: the molecules are not spherical, so different geometries are possible; not all the
kinetic energy is delivered into the right spot; the presence of a solvent (when applied to solutions) and other factors (Figure 9.5.4).

9.5.5 https://chem.libretexts.org/@go/page/41345
Figure 9.5.5 : The Effect of Molecular Orientation on the Reaction of NO and O3. Most collisions of NO and O3 molecules occur
with an incorrect orientation for a reaction to occur. Only those collisions in which the N atom of NO collides with one of the
terminal O atoms of O3 are likely to produce NO2 and O2, even if the molecules collide with E > Ea.

Contributors and Attributions


Stephen Lower, Professor Emeritus (Simon Fraser U.) Chem1 Virtual Textbook
Wikipedia

9.5: The Effect of Temperature on Reaction Rates is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.

9.5.6 https://chem.libretexts.org/@go/page/41345
9.6: Potential Energy Surfaces
A potential energy surface (PES) describes the potential energy of a system, especially a collection of atoms, in terms of certain
parameters, normally the positions of the atoms. The surface might define the energy as a function of one or more coordinates; if
there is only one coordinate, the surface is called a potential energy curve or energy profile. It is helpful to use the analogy of a
landscape: for a system with two degrees of freedom (e.g. two bond lengths), the value of the energy (analogy: the height of the
land) is a function of two bond lengths (analogy: the coordinates of the position on the ground). The Potential Energy Surface
represents the concepts that each geometry (both external and internal) of the atoms of the molecules in a chemical reaction is
associated with it a unique potential energy. This creates a smooth energy “landscape” and chemistry can be viewed from a
topology perspective (of particles evolving over "valleys""and passes").

Potential Energy Curves (1-D Potential Energy Surfaces)


The PES is the energy of a molecule as a function of the positions of its nuclei r. This energy of a system of two atoms depends on
the distance between them. At large distances the energy is zero, meaning “no interaction”. At distances of several atomic
diameters attractive forces dominate, whereas at very close approaches the force is repulsive, causing the energy to rise. The
attractive and repulsive effects are balanced at the minimum point in the curve. Plots that illustrate this relationship are quite useful
in defining certain properties of a chemical bond.

Figure 9.6.1 : A potential Energy Curve for a covalent bond.


The internuclear distance at which the potential energy minimum occurs defines the bond length. This is more correctly known as
the equilibrium bond length, because thermal motion causes the two atoms to vibrate about this distance. In general, the stronger
the bond, the smaller will be the bond length.
Attractive forces operate between all atoms, but unless the potential energy minimum is at least of the order of RT, the two atoms
will not be able to withstand the disruptive influence of thermal energy long enough to result in an identifiable molecule. Thus we
can say that a chemical bond exists between the two atoms in H2. The weak attraction between argon atoms does not allow Ar2 to
exist as a molecule, but it does give rise to the van Der Waals force that holds argon atoms together in its liquid and solid forms.

Potential, Kinetic, and Total Energy for a System


Potential energy and kinetic energy Quantum theory tells us that an electron in an atom possesses kinetic energy K as well as
potential energy V , so the total energy E is always the sum of the two: E = V + K . The relation between them is surprisingly
simple: K =– 0.5V . This means that when a chemical bond forms (an exothermic process with ΔE < 0 ), the decrease in
potential energy is accompanied by an increase in the kinetic energy (embodied in the momentum of the bonding electrons),
but the magnitude of the latter change is only half as much, so the change in potential energy always dominates. The bond
energy – ΔE has half the magnitude of the fall in potential energy.

Mathematical definition and computation


The geometry of a set of atoms can be described by a vector, r, whose elements represent the atom positions. The vector r could be
the set of the Cartesian coordinates of the atoms, or could also be a set of inter-atomic distances and angles. Given r, the energy as
a function of the positions, V (r), is the value of V (r) for all values of r of interest. Using the landscape analogy from the
introduction, V (r) gives the height on the "energy landscape" so that the concept of a potential energy surface arises. An example
is the PES for water molecule (Figure 9.6.1) that show the energy minimum corresponding to optimized molecular structure for
water- O-H bond length of 0.0958 nm and H-O-H bond angle of 104.5°

9.6.1 https://chem.libretexts.org/@go/page/41346
Figure 9.6.2 : PES for water molecule: Shows the energy minimum corresponding to optimized molecular structure for water- O-
H bond length of 0.0958nm and H-O-H bond angle of 104.5°. of Wikipedia (Credit: Aimnature).

The Dimensionality of a Potential Energy Surface


To define an atom’s location in 3-dimensional space requires three coordinates (e.g., x, y ,and z or r, θ and phi in Cartesian
and Spherical coordinates) or degrees of freedom. However, a reaction and hence the corresponding PESs do not depend of the
absolute position of the reaction, only the relative positions (internal degrees). Hence both translation and rotation of the entire
system can be removed (each with 3 degree of freedom, assuming non-linear geometries). So the dimensionality of a PES is
3N − 6 (9.6.1)

where N is the number of atoms involves in the reaction, i.e., the number of atoms in each reactants). The PES is a
hypersurface with many degrees of freedom and typically only a few are plotted at any one time for understanding. See
Calculate Number of Vibrational Modes to get a more details picture of how this applies to calculating the number of
vibrations in a molecule

To study a chemical reaction using the PES as a function of atomic positions, it is necessary to calculate the energy for every
atomic arrangement of interest. Methods of calculating the energy of a particular atomic arrangement of atoms are well described
in the computational chemistry article, and the emphasis here will be on finding approximations of (V (r) to yield fine-grained
energy-position information.
For very simple chemical systems or when simplifying approximations are made about inter-atomic interactions, it is sometimes
possible to use an analytically derived expression for the energy as a function of the atomic positions. An example is
H + H2 → H2 + H (9.6.2)

system as a function of the three H-H distances. For more complicated systems, calculation of the energy of a particular
arrangement of atoms is often too computationally expensive for large scale representations of the surface to be feasible.

Application of Potential Energy Surfaces


A PES is a conceptual tool for aiding the analysis of molecular geometry and chemical reaction dynamics. Once the necessary
points are evaluated on a PES, the points can be classified according to the first and second derivatives of the energy with respect to
position, which respectively are the gradient and the curvature. Stationary points (or points with a zero gradient) have physical
meaning: energy minima correspond to physically stable chemical species and saddle points correspond to transition states, the
highest energy point on the reaction coordinate (which is the lowest energy pathway connecting a chemical reactant to a chemical
product). Three
PES do not show kinetic energy, only potential energy.
At T = 0 K (no KE), species will want to be at the lowest possible potential energy, (i.e., at a minimum on the PES).
Between any two minima (valley bottoms) the lowest energy path will pass through a maximum at a saddle point, which we
call that saddle point a transition-state structure.
The PES concept finds application in fields such as chemistry and physics, especially in the theoretical sub-branches of these
subjects. It can be used to theoretically explore properties of structures composed of atoms, for example, finding the minimum
energy shape of a molecule or computing the rates of a chemical reaction.

9.6.2 https://chem.libretexts.org/@go/page/41346
Contributors and Attributions
Wikipedia
Stephen Lower, Professor Emeritus (Simon Fraser U.) Chem1 Virtual Textbook

9.6: Potential Energy Surfaces is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

9.6.3 https://chem.libretexts.org/@go/page/41346
9.7: Theories of Reaction Rates
The macroscopic discussion of kinetics discussed in previous sections can be now expanded into a more microscopic picture in
terms of molecular level properties (e..g, mass and velocities) involving two important theories: (1) collision theory and (2)
transition-state theory.

Collision Theory
If two molecules need to collide in order for a reaction to take place, then factors that influence the ease of collisions will be
important. The more energy there is available to the molecules, the faster they will move around, and the more likely they are to
bump into each other. Higher temperatures ought to lead to more collisions and a greater frequency of reactions between molecules.
In the drawing below, the cold, sluggish molecules on the left are not likely to collide, but the energetic molecules on the right are
due to collide at any time.

The rate at which molecules collide which is the frequency of collisions is called the collision frequency, Z , which has units of
collisions per unit of time. Given a container of molecules A and B , the collision frequency between A and B is defined by:
−−−−−
8 kB T
Z = NA NB σAB √ (9.7.1)
πμAB

where:
NA and NB are the numbers of molecules A and B, and is directly related to the concentrations of A and B.
The mean speed of molecules obtained from the Maxwell-Boltzmann distribution for thermalized gases
−−−−−
8 kB T
√ (9.7.2)
πμAB

σAB is the averaged sum of the collision cross-sections of molecules A and B. The collision cross section represents the
collision region presented by one molecule to another.
μ is the reduced mass and is given by

mA mB
μ = (9.7.3)
mA + mB

The concepts of collision frequency can be applied in the laboratory: (1) The temperature of the environment affects the average
speed of molecules. Thus, reactions are heated to increase the reaction rate. (2) The initial concentration of reactants is directly
proportional to the collision frequency; increasing the initial concentration will speed up the reaction.
For a successful collision to occur, the reactant molecules must collide with enough kinetic energy to break original bonds and form
new bonds to become the product molecules. This energy is called the activation energy for the reaction; it is also often referred to
as the energy barrier.

9.7.1 https://chem.libretexts.org/@go/page/41347
The fraction of collisions with enough energy to overcome the activation barrier is given by:
−Ea

f =e RT (9.7.4)

where:
f is the fraction of collisions with enough energy to react
Ea is the activation energy
The fraction of successful collisions is directly proportional to the temperature and inversely proportional to the activation energy.

The fraction of successful collisions is directly proportional to the temperature and inversely proportional to the activation
energy.

The more complicated the structures of the reactants, the more likely that the value of the rate constant will depend on the
trajectories at which the reactants approach each other. This kind of electrophilic addition reaction is well-known to all students of
organic chemistry. Consider the addition of a hydrogen halide such as HCl to the double bond of an alkene, converting it to a
chloroalkane.

Experiments have shown that the reaction only takes place when the HCl molecule approaches the alkene with its hydrogen-end,
and in a direction that is approximately perpendicular to the double bond, as shown at below.

The reason for this becomes apparent when we recall that HCl is highly polar owing to the high electronegativity of chlorine, so
that the hydrogen end of the molecule is slightly positive. The steric factor, ρ is then introduced to represent is the probability of the
reactant molecules colliding with the right orientation and positioning to achieve a product with the desirable geometry and
stereospecificity. Values of ρ are generally very difficult to assess and range from 0 to 1, but are sometime estimated by comparing
the observed rate constant with the one in which the preexponential constant A is assumed to be the same as Z .
The lesson you should take from this example is that once you start combining a variety of chemical principles, you gradually
develop what might be called "chemical intuition" which you can apply to a wide variety of problems. This is far more important
than memorizing specific examples.

9.7.2 https://chem.libretexts.org/@go/page/41347
All Three Factors Combined
The rate constant of the gas-phase reaction is proportional to the product of the collision frequency and the fraction of successful
reactions. As stated above, sufficient kinetic energy is required for a successful reaction; however, they must also collide properly.
Compare the following equation to the Arrhenius equation:
−Ea

k = Zρe RT (9.7.5)

where
k is the rate constant for the reaction
ρ is the steric factor.
Zρ is the pre-exponential factor, A, of the Arrhenius equation. It is the frequency of total collisions that collide with the right
orientation. In practice, it is the pre-exponential factor that is directly determined by experiment and then used to calculate the
steric factor.
Ea is activation energy
T is absolute temperature
R is gas constant.
Although the collision theory deals with gas-phase reactions, its concepts can also be applied to reactions that take place in
solvents; however, the properties of the solvents (for example: solvent cage) will affect the rate of reactions. Ultimately, collision
theory illustrates how reactions occur; it can be used to approximate the rate constants of reactions, and its concepts can be directly
applied in the laboratory. Read this for a more detailed discussion of Collision Theory.

Transition-State Theory
Transition state theory (TST) provides a more accurate alternative to the previously used Arrhenius equation and the collision
theory. The transition state theory attempts to provide a greater understanding of activation energy, E , and the thermodynamic
a

properties involving the transition state. Collision theory of reaction rate, although intuitive, lacks an accurate method to predict the
probability factor for the reaction. The theory assumes that reactants are hard spheres rather than molecules with specific structures.
In 1935, Henry Eyring helped develop a new theory called the transition state theory to provide a more accurate alternative to the
previously used Arrhenius equation and the collision theory. The Eyring equation involves the statistical frequency factory, v,
which is fundamental to the theory.
According to TST, between the state where molecules are reactants and the state where molecules are products, there is a state
known as the transition state. In the transition state, the reactants are combined in a species called the activated complex. The
theory suggests that there are three major factors that determine whether a reaction will occur:
1. The concentration of the activated complex
2. The rate at which the activated complex breaks apart
3. The way in which the activated complex breaks apart: whether it breaks apart to reform the reactants or whether it breaks apart
to form a new complex, the products.
Collision theory proposes that not all reactants that combine undergo a reaction. However, assuming the stipulations of the collision
theory are met and a successful collision occurs between the molecules, transition state theory allows one of two outcomes: a return
to the reactants, or a rearranging of bonds to form the products.
Consider a bimolecular reaction:
A +B → C (9.7.6)

[C ]
K = (9.7.7)
[A][B]

where K is the equilibrium constant. In the transition state model, the activated complex AB is formed:

A + B ⇌ AB → C (9.7.8)


[AB]

K = (9.7.9)
[A][B]

9.7.3 https://chem.libretexts.org/@go/page/41347
There is an energy barrier, called activation energy, in the reaction pathway. A certain amount of energy is required for the reaction
to occur. The transition state, AB , is formed at maximum energy. This high-energy complex represents an unstable intermediate.

Once the energy barrier is overcome, the reaction is able to proceed and product formation occurs.

Figure 9.7.1 : Reaction coordinate diagram for the bimolecular nucleophilic substitution (S N ) reaction between bromomethane
2

and the hydroxide anion. form Wikipedia.


The rate of a reaction is equal to the number of activated complexes decomposing to form products. Hence, it is the concentration
of the high-energy complex multiplied by the frequency of it surmounting the barrier.

rate = v[AB ] (9.7.10)


= v[A][B]K (9.7.11)

The rate can be rewritten:


rate = k[A][B] (9.7.12)

Combining Equations 9.7.12 and 9.7.11 gives:



k[A][B] = v[A][B]K (9.7.13)

k = vK (9.7.14)

where
vis the frequency of vibration,
kis the rate constant and
K is the thermodynamic equilibrium constant.

Statistical mechanics (not shown) provides that the frequency, v, is equivalent to the thermal energy, kBT, divided by Planck's
constant, h.
kB T
v = (9.7.15)
h

where
kB is the Boltzmann's constant (1.381 x 10-23 J/K),
T is the absolute temperature in Kelvin (K) and
h is Planck's constant (6.626 x 10-34 Js).
Substituting Equation 9.7.15 into Equation 9.7.14 :
kB T

k = K (9.7.16)
h

Equation ref is often tagged with another term (M


1−m
) that makes the units equal with M is the molarity and m is the
molecularly of the reaction.
kB T
‡ 1−m
k = K (M ) (9.7.17)
h

9.7.4 https://chem.libretexts.org/@go/page/41347
It is important to note here that the equilibrium constant K can be calculated by absolute, fundamental properties such as bond

length, atomic mass, and vibration frequency. This gives the transition rate theory the alternative name absolute rate theory, because
the rate constant, k, can be calculated from fundamental properties.

Thermodynamics of Transition State Theory


To reveal the thermodynamics of the theory, K must be expressed in terms of ΔG . ΔG is simply,
‡ ‡ ‡

o‡ o o
ΔG = G (transitionstate) − G (reactants) (9.7.18)

By definition, at equilibrium, ΔG can be expressed as:


‡ ‡
ΔG = −RT ln K (9.7.19)

Rearrangement gives:

ΔG
‡ −
[K ] =e RT (9.7.20)

From Equation 9.7.17



ΔG
− 1−m
k = ve RT
(M ) (9.7.21)

It is also possible to obtain terms for the change in enthalpy and entropy for the transition state. Because
‡ ‡ ‡
ΔG = ΔH − T ΔS (9.7.22)

it follows that the derived equation becomes,


kB T ‡ ‡
ΔS /R −Δ H /RT 1−m
k = e e M (9.7.23)
h

Equation 9.7.23 is known as the Eyring Equation and was developed by Henry Eyring in 1935, is based on transition state theory
and is used to describe the relationship between reaction rate and temperature. It is similar to the Arrhenius Equation, which also
describes the temperature dependence of reaction rates.
The linear form of the Eyring Equation is given below:
† ‡
k −ΔH 1 kB ΔS
ln = + ln + (9.7.24)
T R T h R

k 1
The values for ΔH

and ΔS can be determined from kinetic data obtained from a

ln vs. plot. The Equation is a straight
T T
‡ ‡
−ΔH ΔS kB
line with negative slope, , and a y-intercept, + ln .
R R h

Figure 9.7.2 : Linearized TST theory

Conclusion
In this article, the complete thermodynamic formulation of the transition state theory was derived. This equation is more reliable
than either the Arrhenius equation and the equation for the Collision Theory. However, it has its limitations, especially when
considering the concepts of quantum mechanics. Quantum mechanics implies that tunneling can occur, such that particles can
bypass the energy barrier created by the transition state. This can especially occur with low activation energies, because the
probability of tunneling increases when the barrier height is lowered.

9.7.5 https://chem.libretexts.org/@go/page/41347
In addition, transition state theory assumes that an equilibrium exists between the reactants and the transition state phase. However,
in solution non-equilibrium situations can arise, upsetting the theory. Several more complex theories have been presented to correct
for these and other discrepancies. This theory still remains largely useful in calculating the thermodynamic properties of the
transition state from the overall reaction rate. This presents immense usefulness in medicinal chemistry, in which the study of
transition state analogs is widely implemented.

References
1. Chang, Raymond. Physical Chemistry for the Biosciences. University Science Books, California 2005
2. Truhlar, D. G.; Garrett, B. C.; Klippenstein, S. J., Current Status of Transition-State Theory. The Journal of physical chemistry
1996, 100, (31), 12771-12800
3. Laidler, K.; King, C., Development of transition-state theory. The Journal of physical chemistry 1983, 87, (15), 2657
4. Lienhard, Gustav, Enzyme Catalysis and Transition -State Theory: Transition State Analogs. Cold Spring Harb Symp Quant
Biol 1972. 36:45-51

Contributors and Attributions


Phong Dao, Bilal Latif, Lu Zhao (UC Davis)

9.7: Theories of Reaction Rates is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

9.7.6 https://chem.libretexts.org/@go/page/41347
9.8: Isotope Effects in Chemical Reactions
The kinetic isotope effect (KIE) is a phenomenon associated with isotopically substituted molecules exhibiting different reaction
rates. Isotope effects such as KIEs are invaluable tools in both physical and biological sciences and are used to aid in the
understanding of reaction kinetics, mechanisms, and solvent effects.

Introduction
Research was first introduced on this topic over 50 years ago and has grown into an enormous field. The scientists behind much of
the understanding and development of kinetic isotope effects were Jacob Bigeleisen and Maria Goeppert Mayer who published the
first paper on isotope effects [J. Chem. Phys., 15, 261 (1947)]. Kinetic isotope effects specifically explore the change in rate of a
reaction due to isotopic substitution.
An element is identified by its symbol, mass number, and atomic number. The atomic number is the number of protons in the
nucleus while the mass number is the total number of protons and neutrons in the nucleus. Isotopes are two atoms of the same
element that have the same number of protons but different numbers of neutrons. Isotopes are specified by the mass number.

As an example consider the two isotopes of chlorine, you can see that their mass numbers vary, with 35Cl being the most abundant
isotope, while their atomic numbers remain the same at 17.
35 37
C l and Cl (9.8.1)

The most common isotope used in light atom isotope effects is hydrogen ( H ) commonly replaced by its isotope deuterium ( H ).
1 2

Note: Hydrogen also has a third isotope, tritium ( H ). Isotopes commonly used in heavy atom isotope effects include carbon ( C ,
2 12

C , nitrogen ( N, N ), oxygen, sulfur, and bromine. Not all elements exhibit reasonably stable isotopes (i.e. Fluorine, F ), but
13 14 15 19

those that due serve as powerful tools in isotope effects.

Potential Energy Surfaces


Understanding potential energy surfaces is important in order to be able to understand why and how isotope effects occur as they
do. The harmonic oscillator approximation is used to explain the vibrations of a diatomic molecule. The energies resulting from the
quantum mechanic solution for the harmonic oscillator help to define the internuclear potential energy of a diatomic molecule and
are
1
En = (n + ) hν (9.8.2)
2

where
n is a positive integer (n=1,2,3...),
h is Planck's constant and
ν is the frequency of vibration.

The Morse potential is an analytic expression that is used as an approximation to the intermolecular potential energy curves:
2
−β(l−lo )
V (l) = De (1 − e ) (9.8.3)

where
V (l) is the potential energy,
D is the dissociation energy of the molecule,
e

β is the measure of the curvature of the potential at its minimum,

l is displacement, and

l is the equilibrium bond length.


o

The D , β, and l variables can be looked up in a textbook or CRC handbook.


e o

9.8.1 https://chem.libretexts.org/@go/page/41348
Below is an example of a Morse potential curve with the zero point vibrational energies of two isotopic molecules (for example R-
H and R-D where R is a group/atom that is much heavier than H or D). The y-axis is potential energy and the x axis is internuclear
distance. In this figure EDo and EHo correspond to the zero point energies of deuterium and hydrogen. The zero point energy is the
lowest possible energy of a system and equates to the ground state energy. Zero point energy is dependent upon the reduced mass
of the molecule as will be shown in the next section. The heavier the molecule or atom, the lower the frequency of vibration and the
smaller the zero point energy. Lighter molecules or atoms have a greater frequency of vibration and a higher zero point energy. We
see this is the figure below where deuterium is heavier than hydrogen and therefore has the lower zero point energy.

This results in different bond dissociation energies for R-D and R-H. The bond dissociation energy for R-D (ED) is greater than the
bond dissociation energy of R-H (EH). This difference in energy due to isotopic replacement results in differing rates of reaction,
the effect that is measured in kinetic isotope effects. The reaction rate for the conversion of R-D is slower than the reaction rate for
the conversion of R-H.

p>
It is important to note that isotope replacement does not change the electronic structure of the molecule or the potential energy
surfaces of the reactions the molecule may undergo. Only the rate of the reaction is affected.

Activation Energies
The energy of the vibrational levels of a vibration (i.e., a bond) in a molecule is given by
1
En = (n + ) hν (9.8.4)
2

where we assume that the molecule is in its ground state and we can compare zero-point vibrational energies,
1
Eo = ( ) hv (9.8.5)
2

Using the harmonic oscillator approximation the fundamental vibrational frequency is

9.8.2 https://chem.libretexts.org/@go/page/41348


1 k
ν = √ (9.8.6)
2π μ

where
k is the force constant of the bond and
μ is the reduced mass
m1 m2
μ = (9.8.7)
m1 + m2

The Arrhenius equation is used to determine reaction rates and activation energies and since we are interested in the change in rate
of reactions with different isotopes, this equation is very important,
Ea

k = Ae kT
(9.8.8)

where
kis the reaction rate,
Ea is the activation energy, and
A is the Arrhenius constant.

The Arrhenius equation can be used to compare the rates of a reaction with R-H and R-D,
H
E
a

kH = AH e kT
(9.8.9)

D
E
a

kD = AD e kT (9.8.10)

where kH and kD are the rates of reaction associated with R-H and the isotope substituted R-D. We will then assume the Arrhenius
constants are equal (A = A ). The ratio of the rates of reaction gives an approximation for the isotope effect resulting in:
H D

H D
kH −
E
a
−E
a

=e kT (9.8.11)
kD

By using the relationship that for both R-H and R-D


1
Eo = ( ) hν (9.8.12)
2

a substitution can be made resulting in


h( ν −ν )
kH H D

=e 2kT (9.8.13)
kD

The vibrational frequency (Equation 5) can then be substituted for R-H and R-D and the value of the expected isotope effect can be
calculated.
kRH kRD
h( − )
μRH μRD
kH
=e 4πkT (9.8.14)
kD

The same general procedure can be followed for any isotope substitution.
In summary, the greater the mass the more energy is needed to break bonds. A heavier isotope forms a stronger bond. The resulting
molecule has less of a tendency to dissociate. The increase in energy needed to break the bond results in a slower reaction rate and
the observed isotope effect.

Kinetic Isotope Effects


Kinetic Isotope Effects (KIEs) are used to determine reaction mechanisms by determining rate limiting steps and transition states
and are commonly measured using NMR to detect isotope location or GC/MS to detect mass changes. In a KIE experiment an atom
is replaced by its isotope and the change in rate of the reaction is observed. A very common isotope substitution is when hydrogen

9.8.3 https://chem.libretexts.org/@go/page/41348
is replaced by deuterium. This is known as a deuterium effect and is expressed by the ratio kH/kD (as explained above). Normal
KIEs for the deuterium effect are around 1 to 7 or 8. Large effects are seen because the percentage mass change between hydrogen
and deuterium is great. Heavy atom isotope effects involve the substitution of carbon, oxygen, nitrogen, sulfur, and bromine, with
effects that are much smaller and are usually between 1.02 and 1.10. The difference in KIE magnitude is directly related to the
percentage change in mass. Large effects are seen when hydrogen is replaced with deuterium because the percentage mass change
is very large (mass is being doubled) while smaller percent mass changes are present when an atom like sulfur is replaced with its
isotope (increased by two mass units).

Primary KIEs
Primary kinetic isotope effects are rate changes due to isotopic substitution at a site of bond breaking in the rate determining step of
a reaction.

 Example

Consider the bromination of acetone: kinetic studies have been performed that show the rate of this reaction is independent of
the concentration of bromine. To determine the rate determining step and mechanism of this reaction the substitution of a
deuterium for a hydrogen can be made.

kH
When hydrogen was replaced with deuterium in this reaction a of 7 was found. Therefore the rate determining step is the
kD

tautomerization of acetone and involves the breaking of a C-H bond. Since the breaking of a C-H bond is involved, a
substantial isotope effect is expected.

Heavy Atom Isotope Effects


A rule of thumb for heavy atom isotope effects is that the maximum isotopic rate ratio is proportional to the square root of the
inverse ratio of isotopic masses.

Expected:
−−

k32 34
=√ = 1.031 (9.8.15)
k34 32

Experimental:
k32
= 1.072 (9.8.16)
k34

Secondary KIEs
Secondary kinetic isotope effects are rate changes due to isotopic substitutions at a site other than the bond breaking site in the rate
determining step of the reaction. These come in three forms: α , β, and γ effects.

β secondary isotope effects occur when the isotope is substituted at a position next to the bond being broken.

9.8.4 https://chem.libretexts.org/@go/page/41348
kH

(CH ) CHBr + H O −→ (CH ) CHOH (9.8.17)


3 2 2 3 2

kD

(CD ) CHBr + H O −→ (CD ) CHOH (9.8.18)


3 2 2 3 2

This is thought to be due to hyperconjugation in the transition state. Hyperconjugation involves a transfer of electron density from a
sigma bond to an empty p orbital (for more on hyperconjugation see outside links).

Solvent Effects in Reactions


Reactions may be affected by the type of solvent used (for example H2O to D2O or ROH to ROD). There are three main ways
solvents effect reactions:
1. The solvent can act as a reactant resulting in a primary isotope effect.
2. Rapid hydrogen exchange can occur between substrate molecules labeled with deuterium and hydrogen atoms in the solvent.
Deuterium may change positions in the molecule resulting in a new molecule that is then reacted in the rate determining step of
the reaction.
3. The nature of solvent and solute interactions may also change with differing solvents. This could change the energy of the
transition state and result in a secondary isotope effects.

References
1. Baldwin, J.E., Gallagher, S.S., Leber, P.A., Raghavan, A.S., Shukla, R.; J. Org. Chem. 2004, 69, 7212-7219 (This is a great
paper using kinetic isotope effects to determine a reaction mechanism. It will interest the organic chemistry oriented reader.)
2. Bigeleisen, J., Goeppert, M., J. Chem. Phys. 1947, 15, 261.
3. Chang, R.; Physical Chemistry for the Chemical and Biological Sciences; University Science Books: Sausalito, CA, 2000, pp
480-483.
4. Isaacs, N.; Physical Organic Chemistry; John Wiley & Sons Inc.: New York, NY; 1995, 2nd ed, pp 287-313.
5. March, J., Smith, M.B.; March’s Advanced Organic Chemistry; John Wiley & Sons, Inc.: Hoboken, NJ, 2007; 6th ed.
6. McMurry, J.; Organic Chemistry; Brooks & Cole: Belmont, CA; 2004, 6th ed.
7. McQuarrie, D.; Quantum Chemistry; University Science Books: Sausalito, CA, 2008, 2nd ed.
8. Rouhi, A.; C&EN. 1997, 38-42.

Problems
1. Describe the difference between primary and secondary kinetic isotope effects.
2. Estimate the kN-H/kN-D for a deuterium substitution on nitrogen given that vH=9.3x1013 Hz and the activation energy is equal to
5.31 kJ/mol.
3. Using the 'rule of thumb' for heavy isotope effects, calculate the expected effect for a bromine isotope substitution, 79Br and
81Br.

4. Explain some of the main ways kinetic isotope effects are used.
5. As discussed, the rate-limiting step in the bromination of acetone is the breaking of a carbon-hydrogen bond. Estimate kC-
H/KC-D for this reaction at 285 K. (Given: vtildeC-H=3000 cm-1 and vtildeC-D=2100 cm-1)

Solutions
1. Primary isotope effects involve isotopic substitution at the bond being broken in a reaction, while secondary isotope effects
involve isotopic substituion on bonds adjacent to the bond being broken.
2. 8.5
3. 1.0126
4. To determine reaction mechanisms, to determine rate limiting steps in reactions, to determine transition states in reactions.
5. 9.685

9.8: Isotope Effects in Chemical Reactions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
Kinetic Isotope Effects is licensed CC BY-NC-SA 4.0.

9.8.5 https://chem.libretexts.org/@go/page/41348
9.9: Reactions in Solution
 Learning Objectives

Make sure you thoroughly understand the following essential ideas:


Describe some of the major differences between the kinetics of reactions in the gas phase, compared with those in liquid
solutions.
What role do solvent cages play in solution kinetics?
Explain the distinction between diffusion-control and activation-control of reaction rates in solutions.
How can the polarity of a solvent affect the energetics of a reaction mechanism?

The kinetics fundamentals we covered in the earlier sections of this lesson group relate to processes that take place in the gas phase.
But chemists and biochemists are generally much more concerned with solutions. This lesson will take you through some of the
extensions of basic kinetics that you need in order to understand the major changes that occur when reactions take place in liquid
solutions.

What's different about kinetics in liquid solutions?


Most of the added complications of kinetics and rate processes in liquid solutions arise from the much higher density of the liquid
phase. In a typical gas at atmospheric pressure, the molecules occupy only about 0.2 per cent of the volume; the other 99.8 percent
is empty space. In a liquid, molecules may take up more than half the volume, and the "empty" spaces are irregular and ever-
changing as the solvent molecules undergo thermal motions of their own.

Figure 9.9.1 : Liquid vs. Gases.


In a typical liquid solution, the solvent molecules massively outnumber the reactant solute molecules, which tend to find
themselves momentarily (~10–11 sec) confined to a "hole" within the liquid. This trapping becomes especially important when the
solvent is strongly hydrogen-bonded as is the case with water or alcohol.

Figure 9.9.2 : Brownian motion of a particle in solution


When thermal motions occasionally release a solute molecule from this trap, it will jump to a new location. The jumps are very fast
(10–12 - 10–13 sec) and short (usually a few solvent-molecule diameters), and follow an entirely random pattern, very much as in
Brownian motion. Consider a simple bimolecular process A + B → products. The reactant molecules will generally be jumping
from hole to hole in the solvent matrix, only occasionally finding themselves in the same solvent cage where thermal motions are
likely to bring them into contact.
Table 9.9.1 : Solvent cages and encounter pairs

9.9.1 https://chem.libretexts.org/@go/page/3537
A pair of reactants end up in the Eventually the two reactants form
same solvent cage, where they an encounter pair. If they fail to Finally, after about 10–11 sec, the
The products form and begin to
bounce around randomly and react the first time, they have solvent cage breaks up and the
move away from each other.
exchange kinetic energy with the many more opportunities during products diffuse away.
solvent molecules. the lifetime of the cage.

The process can be represented as

A + B → {AB} → products (9.9.1)

in which the {AB} term represents the caged reactants including the encounter pair and the activated complex.
Contrast this scenario with a similar reaction taking place in the gas phase; the molecules involved in the reaction will often be the
only ones present, so a significant proportion of the collisions will be A -B encounters. However, if the collision should fail to be
energetically or geometrically viable, the reactant molecules fly apart and are unlikely to meet again anytime soon. In a liquid,
however, the solute molecules are effectively in a constant state of collision — if not with other reactants, then with solvent
molecules which can exchange kinetic energy with the reactants. So once an A-B encounter pair forms, the two reactants get
multiple whacks at each other, greatly increasing the probability that they will obtain the kinetic energy needed to kick them over
the activation hump before the encounter pair disintegrates.

Limiting Cases: Diffusion-Controlled and Activation-Controlled Reactions


The encounter pair model introduces some new rate parameters:
k1

A+B ⇌ AB ⟶ products (9.9.2)


k−1

The first step is an equilibrium between the reactants outside and inside the solvent cage. The rate constants k and k reflect those
1 2

relating to diffusion of molecules through the solvent; their values are strongly dependent on the viscosity (and thus the
temperature) of the solvent. (Note that k is a second-order rate constant, while k is first-order.)
1 2

Diffusion is the transport of a substance through a concentration gradient; that is, from a region of higher concentration to one of
lower concentration. Think of the way the color of tea spreads out when a tea bag is immersed in hot water. Diffusion occurs
because random thermal motions are statistically more likely to move molecules out of a region of higher concentration than in the
reverse direction, simply because in the latter case fewer molecules are available to make the reverse trip. Eventually the
concentrations become uniform and equilibrium is attained.
As molecules diffuse through a liquid, they must nudge neighboring molecules out of the way. The work required to do this
amounts to an activation energy, so diffusion can be thought of as a kinetic process with its own rate constant kd and activation
energy. These parameters depend on the sizes of the solute and solvent molecules and on how strongly the latter interact with each
other. This suggests two important limiting cases for reactions in solution.
For water at room temperature, k1 is typically 109-1010 dm–3 mol–1 s–1 and k2 is around 10–9-10–10 dm–3 mol–1 s–1. Given these
values, k3 > 1012 s–1 implies diffusion control, while values < 109 s–1 are indicative of activation control.
Diffusion Controlled (k ≫ k ): If the activation energy of the A+B reaction is very small or if escape of molecules from the
3 2

{AB} cage is difficult, the kinetics will be dominated by k1, and thus by the activation energy of diffusion. Such a process is
said to be diffusion controlled. Reactions in aqueous solution in which Ea > 20 kJ/mol are likely to fall into this category.
Activation Controlled (k ≪ k ): Alternatively, if the activation energy of the A+B reaction dominates the kinetics, and the
3 2

reaction is activation-controlled.
Several general kinds of reactions are consistently very "fast" and thus are commonly found to be diffusion-controlled in most
solvents:
Units Matter

Gas-phase rate constants are normally expressed in units of mol s–1, but rate constants of reactions in solution are conventionally
given in mol/L units, or dm3 mol–1 s–1. Conversion between them depends on a number of assumptions and is non-trivial.
Recombination of atoms and radicals
For example the formation of I2 from I atoms in hexane at 298 K has k3 = 1.3×1012 dm3 mol–1 s–1.

9.9.2 https://chem.libretexts.org/@go/page/3537
Acid-base reactions which involve the transport of H+ and OH– ions tend to be very fast.
The most famous of these is one of the fastest reactions known:
+ –
H + OH → H2 O

for which k3 = 1.4×1011 dm3 mol–1 s–1 at 298 K.


Polar solvents such as water and alcohols interact with ions and polar molecules through attractive dipole-dipole and ion-dipole
interactions, leading to lower-energy solvated forms which stabilize these species. In this way, a polar solvent can alter both the
thermodynamics and kinetics (rate) of a reaction.

Solvent Thermodynamic Effect


If the products of the reaction are markedly more or less polar than the reactants, solvent polarity can change the overal
thermodynamics (equilibrium constant) of the reaction. Nowhere is this more apparent than when an ionic solid such as salt
dissolves in water. The Na+ and Cl– ions are bound together in the solid through strong coulombic forces; pulling the solid apart in
a vacuum or in a nonpolar solvent is a highly endothermic process. In contrast, dissolution of NaCl in water is slightly exothermic
and proceeds spontaneously.

Figure 9.9.3 : The solvent thermodynamics effect for table salt in water.
The water facilitates this process in two important ways. First, its high dielectric constant of 80 reduces the force between the
separated ions to 1/80 of its normal value. Secondly, the water molecules form a solvation shell around the ions (lower left),
rendering them energetically (thermodynamically) more stable than they were in the NaCl solid.

Solvent Kinetic Effect


In the same way, a reaction whose mechanism involves the formation of an intermediate or activated complex having a polar or
ionic character will have its activation energy, and thus its rate, subject to change as the solvent polarity is altered. As an example
we will consider an important class of reactions that you will hear much about if you take a course in organic chemistry. When an
aqueous solution of a strong base such as KOH is added to a solution of tertiary-butyl chloride in ethanol, the chlorine is replaced
by a hydroxyl group, leaving t-butyl alcohol as a product:

This reaction is one of a large and important class known as SN1 nucleophilic substitution processes that are discussed in most
organic chemistry courses. In these reactions, a species that possesses a pair of non-bonding electrons (also called a nucleophile or
Lewis base) uses them to form a new bond with an electrophile — a compound in which a carbon atom has a partial positive charge
owing to its bonds to electron-withdrawing groups. In the example here, other nucleophiles such as NH3 or even H2O would serve
as well.
In order to reflect the generality of this process and to focus on the major changes that take place, we will represent this reaction as

Extensive studies of this class of reactions in the 1930's revealed that it proceeds in two activation energy-controlled steps,
followed by a simple dissociation into the products:

9.9.3 https://chem.libretexts.org/@go/page/3537
In step , which is rate-determining, the chlorine leaves the alkyl chloride which becomes an intermediate known as a carbocation
("cat-ion"). These ions, in which the central carbon atom lacks a complete octet, are highly reactive, and in step the carbocation
is attacked by the hydroxide ion which supplies the missing electron. The immediate product is another cation in which the positive
charge is on the oxygen atom. This oxonium ion is unstable and rapidly dissociates ( )into the alcohol and a hydrogen ion.

Figure 9.9.4 : reaction Coordinate


The reaction coordinate diagram helps us understand the effect of solvent polarity on this reaction. Polar solvent molecules interact
most strongly with species in which the electric charge is concentrated in one spot. Thus the carbocation is stabilized to a greater
degree than are the activated complexes in which the charge is spread out between the positive and negative ends. As the heavy
green arrows indicate, a more polar solvent will stabilize the carbocation more than it will either of the activated complexes; the
effect is to materially reduce the activation energy of the rate-determining step, and thus speed up the reaction. Because neither the
alkyl chloride nor the alcohol is charged, the change in solvent polarity has no effect on the equilibrium constant of the reaction.
This is dramatically illustrated by observing the rate of the reaction in solvents composed of ethanol and water in varying amounts:
Table 9.9.2 : Data
% water 10 20 30 40 50 60
6
k1 × 10 1.7 9.1 40.3 126 367 1294

9.9: Reactions in Solution is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
17.5: Kinetics of Reactions in Solution by Stephen Lower is licensed CC BY 3.0. Original source:
http://www.chem1.com/acad/webtext/virtualtextbook.html.

9.9.4 https://chem.libretexts.org/@go/page/3537
9.10: Fast Reactions in Solution
The traditional experimental methods described above all assume the possibility of following the reaction after its components have
combined into a homogeneous mixture of known concentrations. But what can be done if the time required to complete the mixing
process is comparable to or greater than the time needed for the reaction to run to completion?

Flow methods
Flow instruments are a rapid mixing devices used to study the chemical kinetics of fast reactions in solution. There are different
flavors that can be implement depending on the nature of the reaction as discussed below.

Continuous Flow Approach


For reactions that take place in milliseconds, the standard approach since the 1950s has been to employ a flow technique of some
kind. An early example was used to study fast gas-phase reactions in which one of the reactants is a free radical such as OH that
can be produced by an intense microwave discharge acting on a suitable source gas mixture. This gas, along with the other reactant
being investigated, is made to flow through a narrow tube at a known velocity.

Figure 9.10.1 : A continuous flow fast kinetic system.


If the distance between the point at which the reaction is initiated and the product detector is known, then the time interval can be
found from the flow rate. By varying this distance, the time required to obtain the maximum yield can then be determined.
Although this method is very simple in principle, it can be complicated in practice.

Stopped Flow Approach


Owing to the rather large volumes required, continuous flow method is more practical for the study of gas-phase reactions than for
solutions, for which the stopped-flow method described below is generally preferred. These are by far the most common means of
studying fast solution-phase reactions over time intervals of down to a fraction of a millisecond. The use of reasonably simple
devices is now practical even in student laboratory experiments. These techniques make it possible to follow not only changes in
the concentrations of reactants and products, but also the buildup and decay of reaction intermediates.

Figure 9.10.2 : A stop flow fast kinetic system.


The basic stopped-flow apparatus consists of two or more coupled syringes that rapidly inject the reactants into a small mixing
chamber and then through an observation cell that can be coupled to instruments that measure absorption, fluorescence, light
scattering, or other optical or electrical properties of the solution. As the solution flows through the cell, it empties into a stopping
syringe that, when filled, strikes a backstop that abruptly stops the flow. The volume that the stopping syringe can accept is
adjusted so that the mixture in the cell has just become uniform and has reached a steady state; at this point, recording of the cell
measurement begins and its change is followed.

9.10.1 https://chem.libretexts.org/@go/page/41350
Figure 9.10.3 : Stop-flow equipment at a biochemistry research laboratory for measuring rapid reactions and properties such as
enzyme kinetics. from Wladimir Labeikovsky.

Quenched Flow Approach


In a quenched-flow instrument, the reaction is stopped after a certain amount of time has passed after mixing. The stopping of the
reaction is called quenching and it can be achieved by various means, for example by mixing with another solution, which stops the
reaction (chemical quenching), quickly lowering the temperature (freeze quenching) or even by exposing the sample to light of a
certain wavelength (optical quenching).
Of course, there are many reactions that cannot be followed by changes in light absorption or other physical properties that are
conveniently monitored. In such cases, it is often practical to quench (stop) the reaction after a desired interval by adding an
appropriate quenching agent. For example, an enzyme-catalyzed reaction can be stopped by adding an acid, base, or salt solution
that denatures (destroys the activity of) the protein enzyme. Once the reaction has been stopped, the mixture is withdrawn and
analyzed in an appropriate manner.

Figure 9.10.4 : A quench flow fast kinetic system.


The quenched-flow technique works something like the stopped-flow method described above, with a slightly altered plumbing
arrangement. The reactants A and B are mixed and fed directly through the diverter valve to the measuring cell, which is not shown
in this diagram. After a set interval that can vary from a few milliseconds to 200 sec or more, the controller activates the quenching
syringe and diverter valve, flooding the cell with the quenching solution.

Relaxation Methods
To investigate reactions that are complete in less than a millisecond, one can start with a pre-mixed sample in which one of active
reactants is generated in situ. Alternatively, a rapid change in pressure or temperature can alter the composition of a reaction that
has already achieved equilibrium.

Flash Photolysis
Many reactions are known which do not take place without light of wavelength sufficiently short to supply the activation energy
needed to break a bond, often leading to the creation of a highly reactive radical. A good example is the combination of gaseous
Cl2 with H2, which proceeds explosively when the system is illuminated with visible light. In flash photolysis, a short pulse of light
is used to initiate a reaction whose progress can be observed by optical or other means.

9.10.2 https://chem.libretexts.org/@go/page/41350
Photolysis refers to the use of light to decompose a molecule into simpler units, often ions or free radicals. In contrast to
thermolysis (decomposition induced by high temperature), photolysis is able to inject energy into a molecule almost
instantaneously and can be much "cleaner," meaning that there are fewer side reactions that often lead to complex mixtures of
products. Photolysis can also be highly specific; the wavelength of the light that triggers the reaction can often be adjusted to
activate one particular kind of molecule without affecting others that might be present.

Norrish and Porter

All this had been known for a very long time, but until the mid-1940's there was no practical way of studying the kinetics of
the reactions involving the highly reactive species produced by photolysis. In 1945, Ronald Norrish of Cambridge University
and his graduate student George Porter conceived the idea of using a short-duration flash lamp to generate gas-phase CH2
radicals, and then following the progress of the reaction of these radicals with other species by means of absorption
spectroscopy.

Figure 9.10.5 : Basic principle of a flash-photolysis relaxation experiment where an excitation pulse purturbed a system at
equilibrium and the subsequent dynamics are resolved in time.
In a flash photolysis experiment, recording of the absorbance of the sample cell contents is timed to follow the flash by an interval
that can be varied in order to capture the effects produced by the product or intermediate as it is formed or decays. Norrish and
Porter shared the 1967 Nobel Prize in Chemistry for this work.

Figure 9.10.6 : A flash-photolysis relaxation experiment


Many reactions, especially those that take place in solution, occur too rapidly to follow by flow techniques, and can therefore only
be observed when they are already at equilibrium. The classical examples of such reactions are two of the fastest ones ever
observed, the dissociation of water
+ −
2 H2 O → H3 O + OH (9.10.1)

and the formation of the triiodide ion in aqueous solution



I + I2 → I – (9.10.2)
3

Reactions of these kinds could not be studied until the mid-1950s when techniques were developed to shift the equilibrium by
imposing an abrupt physical change on the system.

Temperature Jumps
The rate constants of reversible reactions can be measured using a relaxation method. In this method, the concentrations of
reactants and products are allowed to achieve equilibrium at a specific temperature. Once equilibrium has been achieved, the
temperature is rapidly changed, and then the time needed to achieve the new equilibrium concentrations of reactants and products is
measured. For example, if the reaction
k1
A⇌ B (9.10.3)
k−1

9.10.3 https://chem.libretexts.org/@go/page/41350
is endothermic, then according to the Le Chatelier principle, subjecting the system to a rapid jump in temperature will shift the
equilibrium state to one in which the product B has a higher concentration. The composition of the system will than begin to shift
toward the new equilibrium composition at a rate determined by the kinetics of the process.

For the general case illustrated here, the quantity "x" being plotted is a measurable quantity such as light absorption or electrical
conductivity that varies linearly with the composition of the system. In a first-order process, x will vary with time according to
−kt
xt = xo e (9.10.4)

After the abrupt perturbation at time to, the relaxation time t* is defined as the half-time for the return to equilibrium — that is, as
the time required for x to decrease by \(Δx/e = Δx/2.718\). The derivation of t* and the relations highlighted in yellow can be
o

found in most standard kinetics textbooks. Temperature jumps are likely most commonly used.
The rate law for the reversible reaction in Equation 9.10.3 can be written as
d [B]
= k1 [A] − k−1 [B] (9.10.5)
dt

Consider a system comprising A and B that is allowed to achieve equilibrium concentrations at a temperature, T . After 1

equilibrium is achieved, the temperature of the system is instantaneously lowered to T and the system is allowed to achieve new
2

equilibrium concentrations of A and B , [A] and [B] . During the transition time from the first equilibrium state to the second
eq,2 eq,2

equilibrium state, we can write the instantaneous concentration of A as


[A] = [B] − [B] (9.10.6)
eq,1

The rate of change of species B can then be written as


d [B]
= k1 ([B] − [B]) − k−1 [B] = k1 [B] − (k1 + k−1 ) [B] (9.10.7)
eq,1 eq,1
dt

At equilibrium, d [B] /dt = 0 and [B] = [B] eq,2


, allowing us to write

k1 [B] = (k1 + k−1 ) [B] (9.10.8)


eq,1 eq,2

Using the above equation, we can rewrite the rate equation as


dB
= (k1 + k−1 ) dt (9.10.9)

([B] − [B])
eq,2

Integrating yields

−ln ([B] − [B] ) = − (k1 + k−1 ) t + C (9.10.10)


eq,2

We can rearrange the above equation in terms of B


−( k1 +k−1 )t
[B] = C e + [B] (9.10.11)
eq,2

At t = 0 , [B] = [B] eq,1


, so C = [B]eq,1 − [B]eq,2 . Plugging the the value of C , we arrive at

9.10.4 https://chem.libretexts.org/@go/page/41350
−( k1 +k−1 )t
[B] − [B] = ([B] − [B] )e (9.10.12)
eq,2 eq,1 eq,2

which can also be expressed as


−( k1 +k−1 )t −t/τ
Δ [B] = Δ[B] e = Δ[B] e (9.10.13)
0 0

where Δ [B] is the difference in the concentration of B from the final equilibrium concentration after the perturbation, and τ is the
relaxation time. A plot of ln (Δ [B] /Δ[B] ) versus t will be linear with a slope of − (k + k ) , where k and k are the rate
0 1 −1 1 −1

constants at temperature, T .2

Contributors and Attributions


Mark Tuckerman (New York University)
Stephen Lower, Professor Emeritus (Simon Fraser U.) Chem1 Virtual Textbook

This page titled 9.10: Fast Reactions in Solution is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Stephen Lower via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

9.10.5 https://chem.libretexts.org/@go/page/41350
9.11: Oscillating Reactions
It should be clear by now that chemical kinetics is governed by the mathematics of systems of differential equations. Thus far, we have only looked
at reaction systems that give rise to purely linear differential equations, however, in many instances the rate equations are nonlinear. When the
differential equations are nonlinear, the behavior is considerably more complex. In particular, nonlinear equations can lead to oscillatory solutions
and can also exhibit the phenomenon of chaos. Chaotic systems are systems that are highly sensitive to small changes in the parameters of the
equations or in the initial conditions. Basically, this means that the behavior of a chaotic system can be unpredictable, since such small changes can
occur in the form of small errors in determining the parameters (rounding to the nearest tenth or hundredth) or in specifying the initial conditions,
and these small changes can cause the system to evolve in time in a very different way.

The Iodine Clock Reaction


The iodine clock reaction is a popular chemistry experiment in which one can visualize how different rate constants in consecutive reactions affect
the concentration of species during the reaction. Iodine anions (I ) are colorless. When I is reacted with hydrogen peroxide and protons, triiodide
− −

is formed, which has a dark blue color. Consider the following series of irreversible reactions:
k1
− + −
H2 O2 + 3 I +2 H ⟶ I + 2 H2 O (9.11.1)
3

k2
− 2− − 2−
I + 2 S2 O ⟶ 3 I + S4 O (9.11.2)
3 3 6

The rate laws for this system are



d [I ] 3 2
3 − − 2−
= k1 [ I ] − k2 [ I ] [ S2 O ] (9.11.3)
3 3
dt

d [I ] 3 2
− − 2−
= −k1 [ I ] + 3 k2 [ I ] [ S2 O ] (9.11.4)
3 3
dt
2−
d [ S2 O ] 2
3 − 2−
= −k2 [ I ] [ S2 O ] (9.11.5)
3 3
dt

In order to make the equations look a little simpler, let us introduce the variables:
− − 2−
x = [I ], y = [I ], z = [ S2 O ] (9.11.6)
3 3

In terms of these, the rate equations are


dx 3 2
= −k1 x + 3 k2 y z (9.11.7)
dt
dy
3 2
= k1 x − k2 y z (9.11.8)
dt

dz 2
= −k2 y z (9.11.9)
dt

If we solve these numerically, we find the following time dependence of the three concentrations: This is a clear example of nonlinearity. Note how
the concentration of I remains close to 0 for a period of time and then suddenly starts to increase. In a sense, think of the “straw that broke the

camel’s back”. As we pile straws on the back of the camel, the camel remains upright until that last straw, which suddenly breaks the back of the
camel, and the camel suddenly falls to the ground. This is also an illustration of nonlinearity.

9.11.1 https://chem.libretexts.org/@go/page/41351
Chemistry experiment 28 - Iodine clock reaction

Video 9.11.1 : Famous iodine clock reaction: oxidation of potassium iodide by hydrogen peroxide (https://www.youtube.com/watch?
v=_qhYDuJt8fI).
Despite the complexity of the rate equations, we can still analyze the approximately and predict the behavior seen in Figure 9.11.1. In this reaction
mechanism, k ≫ k . Given the rate law for I ,
2 1 3


d [I ] 3 2
3 − − 2−
= k1 [ I ] − k2 [ I ] [ S2 O ] (9.11.10)
3 3
dt


d [I ]
3
if we use the steady-state approximation, we can set the equal to 0, yielding
dt

− 3
k1 [I ]

[I ] = (9.11.11)
3 2
k2 2−
[ S2 O ]
3

Since k ≫ k , the concentration of [ I ] is approximately 0 as long as there are S O ions present. As soon as all of the S O is consumed, the
2 1 3

2
2−
3 2
2−
3

concentration of I can build up in the solution, changing the solution to a dark blue color. Figure 9.11.1 displays the concentration profiles for I ,

3

I

3
, and S O . As can be seen from the figure, the concentration of I (red line) remains at approximately 0 mol/L until all of the S O (blue
3
2−

3

3 2
2−

line) has been depleted.

Figure 9.11.1 : Concentrations as functions of time of the three species in the iodine clock reaction.

Oscillating Reactions
In all of the examples we have seen thus far, the concentration of intermediate species displays a single maximum during the course of the reaction.
There is another class of reactions called oscillating reactions in which the concentration of intermediate species oscillates with time. Consider the
following series of reactions
k1

A+Y ⟶ X (9.11.12)

9.11.2 https://chem.libretexts.org/@go/page/41351
k2
X+Y ⟶ P (9.11.13)

k3

B+X ⟶ 2 X+Z (9.11.14)

k4

2 X ⟶ Q (9.11.15)

k5

Z⟶ Y (9.11.16)

In the above reaction mechanism, A and B are reactants; X , Y , and Z are intermediates; and P and Q are products. The third reaction in which B
and X react to form X and Z is known as an ”autocatalytic reaction” in which at least one of the reactants is also a product. Such reactions are a key
feature of oscillating reactions, as will be discussed below.

The Belousov Zhabotinsky reaction 8 x normal …

Video 9.11.2 : The famous Belousov Zhabotinsky chemical reaction in a petri dish. The action is speeded up 8 x from real life. Pacemaker
nucleation sites emit circular waves. Breaking the wavefront with a wire triggers pairs of spiral defects which emit more closely spaced waves which
eventually fill the container.
Let us assume that the concentrations of A and B are large, such that we can approximate them to be constant with time. The rate equation for
species X can be written as
d [X] 2
= k1 [A] [Y] − k2 [X] [Y] + k3 [B] [X] − 2 k4 [X] (9.11.17)
dt

Using the steady-state approximation, we can set dX/dt = 0 and rewrite Equation 9.11.17 as
2
(−2 k4 ) [X] + (k2 [Y] − k3 [B]) [X] + k1 [A] [Y] = 0 (9.11.18)

We can then use the quadratic formula to solve for X :


−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−
2
(k2 [Y] − k3 [B]) ± √ (k2 [Y] − k3 [B]) − 4 (−2 k4 ) (k1 [A] [Y])
[X] = − (9.11.19)
2 (−2 k4 )

Thus, there are two solutions for the concentration of X accessible to the reaction system. To examine solutions for [X], let us first assume that [Y] is
large. Under these conditions, the first two reactions in the reaction mechanism largely determine the concentration of [X]. We can thus approximate
Equation 9.11.17 as
0 ≈ k1 [A] [Y] − k2 [X] [Y] (9.11.20)

Solving for [X] yields


k1 [A]
[X] ≈ (9.11.21)
k2

As the reaction continues, species Y is depleted and the assumption that [Y] is large becomes invalid. Instead the 3
rd
and 4
th
steps of the reaction
mechanism determine the concentration of X . In this limit, we can approximate Equation 9.11.17 as
2
0 ≈ k3 [B] [X] − 2 k4 [X] (9.11.22)

Solving for [X] yields


k3 [B]
[X] ≈ (9.11.23)
2k4

9.11.3 https://chem.libretexts.org/@go/page/41351
In the second mechanism, the autocatalytic reaction step leads to an increase in the concentration of X and Z, which in turn leads to an increase in
the concentration of Y . The feedback loop between the production of species X and Y leads to oscillatory behavior in the system. This reaction
mechanism is known as the Belousov-Zhabotinksii reaction first discovered in the 1950s.
The autocatalytic reaction of X in which X reacts with B to form more X in reaction 3
The regeneration of species Y in reaction 5
The competition between reaction 2 and 3 for the consumption of X and the involvement of Y in reaction 2
The actual Belousov-Zhabotinskii reaction is complex, involving many individual steps, and involves the oscillation between the concentration of
H BrO and Br . The reaction equations are

2

− − +
BrO + Br +2 H ⟶ H BrO2 + H OBr (9.11.24)
3
− +
H BrO2 + Br +H ⟶ 2 H OBr (9.11.25)
− +
H OBr + Br +H ⟶ Br2 + H2 O (9.11.26)
− +
2 H BrO2 ⟶ BrO + H OBr + H (9.11.27)
3

− + .
BrO + H BrO2 + H ⟶ 2 BrO + H2 O (9.11.28)
3 2

. 3+ + 4+
BrO + Ce +H ⟶ H BrO2 + C e (9.11.29)
2

. 4+ − 3+ +
BrO + Ce + H2 O ⟶ BrO + Ce +2 H (9.11.30)
2 3

The essential steps in this mechanism can be reduced to the following set of reactions. Note that we leave this unbalanced and only include the
species whose concentrations as functions of time we seek.
k1
− −
BrO + Br ⟶ H BrO2 + H OBr (9.11.31)
3

k2

H BrO2 + Br ⟶ 2 H OBr (9.11.32)

k3
− 4+
BrO + H BrO2 ⟶ 2 H BrO2 + 2 C e (9.11.33)
3

k4

2 H BrO2 ⟶ BrO + H OBr (9.11.34)
3

k5
4+ −
Ce ⟶ f Br (9.11.35)

Setting the variables as follows:


− 4+
x = [H BrO2 ] , y = [Br ], z = [C e ] (9.11.36)

We make the approximation that [BrO −

3
] to be a constant a . In this case, the rate equations become
dx 2
= k1 ay − k2 xy + k3 ax − k4 x (9.11.37)
dt
dy
= −k1 ay − k2 xy + f k5 z (9.11.38)
dt
dz
= 2 k3 ax − k5 z (9.11.39)
dt

Solving these equations numerically, we obtain the trajectory of two of the species show in the Figure 9.11.2.

Figure 9.11.2 : Oscillating pattern of the concentrations in the Belousov-Zhabotinskii reaction.


On the other hand, we can drive this system to become chaotic by changing the parameters a little. When this is done, we find the follow plot of the
concentration of x:

9.11.4 https://chem.libretexts.org/@go/page/41351
Figure 9.11.3 : Chaotic behavior in the Belousov-Zhabotinskii reaction.

Contributors and Attributions


Mark Tuckerman (New York University)

9.11: Oscillating Reactions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

9.11.5 https://chem.libretexts.org/@go/page/41351
9.E: Chemical Kinetics (Exercises)
9.1: Reaction Rates
Q9.1
Write the rate of reaction in terms of the rate of disappearance of reactant and the rate of formation of products:
a. N O + O (g)
→ NO +O3(g) 2(g) 2(g)

b. 2C H2 + 7O → 4C O
6(g) + 6H 2(g) 2(g) 2 O(aq)

c. H +I
2(g)
→ 2H I 2(g) (g)

d. 4OH + H S → SO
(g) + 2H O
2 (g) 2(g) 2 (aq) + H2(g)

S9.1
− Δ [N O] − Δ [ O3 ] Δ[N O2 ] Δ[ O2 ]
1. rate of reaction = = = =
Δt Δt Δt Δt
− Δ [ C2 H6 ] − Δ [ O2 ] Δ[C O2 ] Δ[ H2 O]
2. rate of reaction = = = =
2Δt 7Δt 4Δt 6Δt
− Δ [ H2 ] − Δ [ I2 ] Δ[H I ]
3. rate of reaction = = =
Δt Δt 2Δt
− Δ [OH ] − Δ [ H2 S] Δ[S O2 ] Δ[ H2 O] Δ[ H2 ]
4. rate of reaction = = = = =
4Δt Δt Δt Δt Δt

9.2: Reaction Order


Q9.2
Determine the value of the rate constant for the elementary reaction:
I2(g) + H2(g) → 2H I(aq) (9.E.1)

given that when [Br2] is 0.15 M and [H2] is 0.2M, the rate of reaction is 0.005 M s-1 at 298 K.

S9.2
rate of reaction = k[Br2][H2]
0.005 Ms-1 =k ( 0.15M)2(0.2M)
k= 1.11 M-1s-1

Q9.3
Given the rate of the third order reaction:
A+B+C → P (9.E.2)

-1
is 0.05 Ms
If the [A] = 0.05 M, [B] = 0.01M, and [C] = 0.25M. What is the third order rate constant?

S9.3
rate of reaction = k[A][B][C]
0.05Ms-1 =k (0.05M)(0,01M)(0.25M)
k= 0.05Ms-1/( 0.05) (0.01)( 0.25) M3
k= 400M-2s-1

Q9.3
What are the units of the rate constant for a second-order reaction?

9.E.1 https://chem.libretexts.org/@go/page/41334
S9.3
The reaction rate:
d[A]
2
Rate = = −k[A] (9.E.3)
dt

The rate constant of a second-order reaction, k, is M-1min-1 or M-1s-1

Q9.4
Calculate the fraction of the starting quantity of A that will be used up after 60 s. Given the reaction below which is found to be the
first order in A and t = 40 s
1/2

A → B+C (9.E.4)

S9.4
with \(t_{1/2} = 40 \; s\]

The remain of fraction A after 60 s:

The fraction will be used up after 60s:

Q9.5
Given the first order reaction is completed 90% in 30 mins at 298 K. Calculate the rate constant.

S9.5
90%= 0.9

Q9.6
Assume the half life of the first order decay of radioactive isotope takes about 1 year (365 days). How long will it take the
radioactivity of that isotope to decay by 60%?

S9.6

60% is lost
100%- 60%= 40% is remained

9.E.2 https://chem.libretexts.org/@go/page/41334
40%= 0.4

Q9.7a
The decomposition of dinitrogen peroxide (N 2 O5 ) is a first-order reaction with a rate constant of 0.045 min-1 at 300 K.
2 N2 O5(g) → 4N O2 (g) + O2 (9.E.5)

If there were initially 0.040 mol of N 2 O5 , calculate the moles of N 2 O5 remaining after 5 minutes.

S9.7a
The integrated rate equation of a first-order reaction is:
−kt
[A] = [A]0 e (9.E.6)

Substituting concentration for moles of reactant and plugging in the known values:
−0.045∗5
nA = 0.0400 e (9.E.7)

nA = 0.032 (9.E.8)

Q9.7b
You forgot to do your lab assignment the day before and realize you only have 5 hours to reduce your initial concentration of your
reactant by 90%. You know the half life is 2 hours and before you get out of bed, you wonder if you can even finish the lab on time.
Assuming you have a first order reactant, will 5 hours be enough time for your compound to reduce by 90%?

S9.7b
First, you need to figure the half life of your compound. We do this by solving for k:
0.693
t1/2 = (9.E.9)
k

0.693
k = (9.E.10)
1
t
2

0.693
k = = 0.3465 (9.E.11)
2hrs

with this half life, we can find the time it will take by solving for t:
[A]
ln = −kt (9.E.12)
[A]o

We do not have the initial and final concentration, but that is okay.
Assuming the initial concentration is 100 g, we can assume that the final concentration is 10 g because that would be reduced 90%
as stated in the problem.
[10]
ln = −kt (9.E.13)
[100]

9.E.3 https://chem.libretexts.org/@go/page/41334
ln[0.1]
= t = 6.65 hours (9.E.14)
−0.3465

So 5 hours will be definitely not enough time to reduce the compound

Q9.8a
Assume you are dealing with a first order reactant. You know your rate constant is 1.5x10-4/min, but your employer wants to the
half-life of the reactant?

S9.8a
0.693
t1/2 = (9.E.15)
k

0.693
t1/2 = (9.E.16)
−4

1.5x10

3
= 4.62 × 10 mins (9.E.17)

Q9.8b
The half-life of a second order reaction

2A → P (9.E.18)

is given by:
1
t1/2 = (9.E.19)
k[A]o

Calculate the half-life of a reaction with initial reactant concentration [A] = 0.90 M and a rate constant of 0.20 M-1 min-1.

S9.8b
1
−1
t1/2 = = 0.18mi n (9.E.20)
0.20 × 0.90

Q9.9a
A reaction
A → B (9.E.21)

is observed. You do not know the stoichiometry of the reactants or products, but observe that when you increase the initial
concentration of A from 0.4 M to 0.8 M, the half-life decreased from 10 minutes to 5 minutes. Determine the reaction order and the
rate constant.

S9.9a
A reaction whose half-life changes when the reactant concentration is changed is a second-order reaction.
1
t1/2 = (9.E.22)
k[A]o

1
k = (9.E.23)
[A]0 t1/2

1 −1 −1
k = = 0.5 M mi n (9.E.24)
(0.4M )(5min)

Q9.9b
You notice the initial concentration has no affect on the half-life. From this information alone, what can you determine about the
the order of the reactant?

S9.9b
You can determine from this information that it would be a first order reactant. The concentration of first order reactants has not
affect on the rate or half-life of the reactant. This can be seen in the formula. t_{\frac{1}{2}}=\frac{0.693}{k} This formula does
not account for the initial concentration of the reactant.

9.E.4 https://chem.libretexts.org/@go/page/41334
Q9.10a
Calculate the order of the reaction and the rate constant of Cytobutane decompose to ethylene based on equation
\[C_4H_{8 g)} \rightarrow 2C_2H_{4(g)}\]
0
and temperature at 215 C, constant volume, the pressure 200, 158, 124, 98, 77.5, 61 mmHg.

S9.10a

Q9.10b
Given the following data of concentration [A] over a period of time, decide if the data represents first order or second order. Solve
for K. Show graphs.

S9.10b
The data best fits a Second order graph.

The equation for a second order reaction is:


1/[A]= kt+ 1/[A0]
When solving for K, the equation is rearranged to:
1/[A]- 1/[A0] =k
t
Plug in the numbers given:
1/[.3]- 1/[.5]= k

9.E.5 https://chem.libretexts.org/@go/page/41334
(0-54)
.025=k
One, can also obtain the slop of the graph and k=.025

Q9.11a
If a compound’s ½ life is 15.6 days. What is the value of k? How long will it take to decompose to 10%. Use first order reaction.

S9.11a
Equation for half-life of a First-Order Reaction is:
ln(2)
t½ = (9.E.25)
k

If we plug in the information given:


ln(2)
15.6days = (9.E.26)
k

−2 −1
k = 4.4 × 10 day (9.E.27)

The second equation needed is the first order reaction, which is:
ln([A]/[A0])= -kt1/2
so, ln ([A]/[A0])=.1
t=-1/k x ln([A]/[A0])
t= -1/(4.4 x 10-2 day-1) x (.1)
t=52 days
12) In a second order reaction 2A--> products, the final concentration is .28M. What is the initial concentration if k=.32M-1s-1 and
the times is 5 seconds.
Solution:
Second order reaction equation is:
1/[A]= kt+ 1/[A0]
Filling the information given:
1/[.28M]=.32M-1s-1(5)+ 1/[A0]
1/[A0]= 1.97
[A0]= .50M

Q9.11b
Calculate the half life of a compound if 90% of a given sample of the compound decomposed in 30min

S9.11b
[A]=[A]0 e^-kt
[A]0 = 90%
[A] = 10%
t = 30min * 60s/1min = 1800s
[10%]/[90%]=e^ -k*1800
ln[10%/90%]/1800 = k = 3.95E-4
t1/2= ln(2)/k
t1/2 = .693/3.95E-4= 1754s

9.E.6 https://chem.libretexts.org/@go/page/41334
1754s*1min/60s = 29 min

Q9.12
Given rate constant for second order reaction
2N O2(g) → 2N O(g) + O2(g) (9.E.28)

is 1.08 M-1s-1 at 600 0 C. Find the time that would take for the concentration of NO2 from 1.24 M to 0.56 M?

S9.12
1/ [A] – 1/[A]0 = kt
t= 1/k ( 1/[A] – 1/[A]0)
= (1/ 1.08 M-1s-1 ) (1/0.56 – 1/1.24) = 7.2 s

Q9.13a
Is which order of reaction half-life is independent of initial concentration?

S9.13a
First order because the half-life equation for first order is
t1/2= ln(2)/k, it does not have [A0]

Q9.13b
The decomposition of N2) is the first order. At 3650C, t1/2 is 1.79 x 103 min. given intial pressure of 1.05 atm.. Calculate total
pressure.

S9.13b
P = PN2 O + PN2 + PO2 (9.E.29)

= 0.525 + 0.525 + 0.2625 = 1.31 atm (9.E.30)

Q9.14a
The integrated rate law for the zero-order reaction A → B is [A]t = [A]0 - kt
a) skektch the following plots:
(i) rate vs. [A]
(ii) [A] vs. t

S9.14a
rate = k
rate is independent of [A]

(ii) [A] vs. t


[A]t = [A]0 - kt

9.E.7 https://chem.libretexts.org/@go/page/41334
b) Derive an expression for the half-life of the reaction.
At t = t1/2 , [A] = [A]0/2 so, [A]0/2 = [A]0 - kt1/2
t1/2 = 1/2k [A}0
c) Calculate the time in half-lives when the integrated rate law is no longer valid (that is, when [A] = 0)
[A]t = 0 = [A]0 - kt
t1/2 = 1/2k [A}0 ⇒ k = 1/2t1/2 [A}0
Therefore, to consume all of the reactants it takes

t = [A}0/k = [A]0/2 so, [A]0 / (1/(2t1/2)) [A]0 = 2t1/2


integrated rate law is no longer valid after 2 half-lives

Q9.14b
Jack, Jill, and you are in a physical chemistry class. The professor writes the following equations on the black board.
A → B (9.E.31)

−kt
[A] = [ A0 ] e (9.E.32)

a. The professor assigns you to derive the first-order reaction in front of the class.
b. Jack was assigned to find the rate constant if the reaction half-life is 10 hours. Since you are a good friend of Jack, you decide
to help him double check his answer by solving for the rate constant.

S9.14b
a)

[A] t
d[A]
∫ =∫ −kdt (9.E.33)
[A]0 [A] 0

[A]
ln[A] ∣ = −kt (9.E.34)
[A]
0

ln[A] − ln[A]0 = −kt (9.E.35)

[A]
ln = −kt (9.E.36)
[A]0

[A]
−kt
=e (9.E.37)
[A]0

−kt
[A] = [ A0 ] e (9.E.38)

b)

ln 2
t1/2 = (9.E.39)
k

9.E.8 https://chem.libretexts.org/@go/page/41334
ln 2
10hours = (9.E.40)
k

ln 2 1hour 1min
k = × × (9.E.41)
10hours 60min 60secs

k = (9.E.42)

Q9.15a
In the nuclear industry, workers use a rule of thumb that the readioactivity from any sample will be relatively harmless after 10
half-lives. Calculate the fraction of a radioactive sample that remains after this time period (hint: Radioactive decay obeys first-
order kinetics)

S9.15a
[A] = [A]oe-kt
t1/2 = ln2 / k
[A] = (100)e-(1)(10) =.00454
% remaining after 10 half-lives

Q9.15b
As a school field trip to apply what you learned in physical chemistry class, you and your class went to a nuclear power plant. The
plant's manager, Mr. Setsaw gives the class a tour of the facilities. When the class reaches the simulation room, a meltdown
simulation is offered to the class, but only if the class can answer his questions can the class experience it.
a. Radioactive decay occurs in what order?
b. The gas used in the simulation is harmless because the fraction of the original remaining is 0.004 and the time that passed 30
years. What is the rate constant?

S9.15b
a) Radioactive decay occurs in first order.
b)

[A]
−kt
=e (9.E.43)
[A]0

−k 30years
0.004 = e (9.E.44)

ln[0.004] = −k 30years (9.E.45)

−30
k = (9.E.46)
ln[0.004]

k = (9.E.47)

Q9.16a
Many reactions involving heterogeneous catalysis are zero order; that is, rate = k. An example is the decomposition of phosphine (
PH ) over tungsten (W):
3

4 PH (g) → P (g) + 6 H (g) (9.E.48)


3 4 2

The rate for this reaction is independent of [PH3] as long as phosphine's pressure is sufficiently high (>= 1 atm). Explain.

9.E.9 https://chem.libretexts.org/@go/page/41334
S9.16a
With sufficient PH3, all of the catalytic sites on the tungsten surface are occupied. Further increases in the amount of phosphine
cannot affect the reaction, and the rate is independent of [PH3].

Q9.16b
Jill and you just went to an enzyme kinetic workshop. However, Jill is confused about why at high concentrations of substrates the
reaction order is zero-order. Please explain to Jill in a way that anyone can understand.

S9.16b
At high concentrations of substrates, all the enzymes are working at their fastest rate. As soon as the enzyme is done converting one
substrate to its product, another substrate comes into the active site. Therefore, the rate of the enzyme working is constant and this
is the definition for zero-order reaction.

9.3: Molecularity of a Reaction


Q9.41a
A mixture of compounds M and N whose half-lives are 40 minutes and 17 minutes, respectively. They decompose by first-order
kinetics. If their concentrations are equal initially, how long does concentration of N to be half that of M?

S9.41a
[I] is the initial concentration of M and N

-> t = 30.1 min

Q9.41b
Compounds A and B both decay by first-order kinetics. The half-life of A is 20 minutes and the half-life of B is 48 minutes. If a
container initially contains equal concentrations of compounds A and B, after how long will the concentration of B be twice that of
A?

S9.41b
1. Write, in mathematical terms, the information given in the problem and what the problem is asking for.

t1/2,A = 20.0min (9.E.49)

t1/2,B = 48.0min (9.E.50)

[A]0 = [B]0 (9.E.51)

A and B decay by first-order, so


d[A]
− = kA [A] (9.E.52)
dt

d[B]
− = kB [B] (9.E.53)
dt

(Note: the rate constants for A and B are not equal, so indicate which is which with subscripts.)
Want find t at which the following is true:

[B] = 2[A] (9.E.54)

9.E.10 https://chem.libretexts.org/@go/page/41334
2. Substitute the integrated rate equations for [A] and [B]
−kB t −kA t
[B]0 e = 2[A]0 e (9.E.55)

3. Write expressions for the rate constants in terms of half-lives, and substitute into the equation.
ln2 ln2
t1/2,A = ⇒ kA = (9.E.56)
kA t1/2,A

ln2 ln2
t1/2,B = ⇒ kB = (9.E.57)
kB t1/2,B

ln2 ln2
− t − t
t1/2,B t1/2,B
[B]0 e = 2[A]0 e (9.E.58)

4. Solve for t
Since initial concentrations of A and B are equal:
ln2 ln2
− t − t
t1/2,B t1/2,B
e = 2e (9.E.59)

Take natural log of both sides:


t t
− ln2 = (1 − ) ln2 (9.E.60)
t1/2,B t1/2,A

1
t = (9.E.61)
1 1

t1/2,A t1/2,B

5. Plug in values for half-lives


\[ t = \dfrac{1}{\dfrac{1}{20.0 min} - \dfrac{1}{48.0 min}} = 34.2\ min \]
Answer: 34.2 minutes

Q9.42a
In Q3 thermodynamics and in Q9 chemical kinetics, the term “reversible “ is used. How do you understand this term? (it has a
same meaning in these two chapters)

S9.42a
Actually, this word is used to describe a “reversible” reaction in which both forward and backward reactions can happen in kinetics.
In thermodynamics, “reversible” is used to describe a process that is in equilibrium along the pathway from the initial to final
states.

Q9.42b
If a reaction has come to thermodynamic equilibrium, can we say anything in particular about the system's kinetics?

S9.42b
Equilibrium occurs when all reactants and products are being consumed at the same rate that they are created. Take the simple
example:
k1

A ⇌ B (9.E.62)
k−1

When the system is at equilibrium,


d[A] d[B]
= =0 (9.E.63)
dt dt

9.E.11 https://chem.libretexts.org/@go/page/41334
k−1 [A] = k1 [B] (9.E.64)

k−1 [B]
= (9.E.65)
k1 [A]

k−1
Keq = (9.E.66)
k1

Q9.43a
The recombination of bromine atoms in an organic solvent, like carbon tetrafloride, is considered as a diffusion-controlled process
Br + Br → Br2 (9.E.67)

we have the viscosity of CF4 is 9.80 x 10-4 Nsm-2 at 30oC, what is the rate of recombination at 30oC?

S9.43a

Q9.43b
Calculate the rate constant of the diffusion-controlled reaction
2I → I2 (9.E.68)

in dichloromethane at 15°C, which has a viscosity of 0.449 mPa·s at 15°C.

S9.43b
1. Use equation 9.50 to calculate the rate constant.
8 RT
kD = (9.E.69)
3 η

J N m
8.314 × 288 K ×
8 mol K J 1000 L
kD = × (9.E.70)
2 3
3 Pa N /m m
0.449 mP a s × ×
1000 mP a Pa

Answer:
10 −1 −1
kD = 1.42 × 10 M s (9.E.71)

Q9.52
Japanese survivors have been exposed to the risk of radiation after the atomic bomb. One man was measured to have iodine-131
levels at 9.7 mC. Calculate the number of atoms of I-131 to which this radioactivity corresponds.

S9.52
First, convert the rate mCi s-1
1 mCi=1.10X10-3 Ci
1 Ci=3.7 X 1010 s-1
The rate (r can be derived as such
10 −1
1 Ci 3.7 × 10 s 8 −1
r = (9.7 C i) ( )( ) = 3.59 × 10 s (9.E.72)
1000 mC i 1 Ci

9.E.12 https://chem.libretexts.org/@go/page/41334
The accepted value for the half life of I-131 is 8.02 days. Using this information, the number of I-131 atoms can be calculated using
the nuclear decay equation. Use the
Nuclear Decay Equation:
ln 2
λN = N (9.E.73)
t1/2

or
ln 2
N = r (9.E.74)
t1/2

The half-life for the radioactive beta decay of iodine is 8.02 days
131 131
I → ¯e
Xe + β + ν (9.E.75)
53 54

therefore, the rate is


8 −1
r = 3.59 × 10 s (9.E.76)

Plug into the equation and convert days to seconds.


(8.02 days/ln2)(24 hours/1 day)(3600 seconds/1 hour)(3.59X 10-8) = 2.49 X 1014 I-131 atoms.
N= (

Q9.54
Calculate the rate law for the following acid-catalyzed reaction:
+
H
+ −
C H3 C OC H3 + Br2 ⟶ C H3 C OC H2 Br + H + Br (9.E.77)

Rate of Disappearance Br2/


Expt. # [C H3 COC H3 ]o /M [Br2 ]o /M [H
+
]o /M
M*s-1

1 0.3 0.05 0.05 5.7 × 10


−5

2 0.3 0.1 0.05 5.7 × 10


−5

3 0.3 0.05 0.05 1.2 × 10


−4

4 0.4 0.05 0.2 3.1 × 10


−4

5 0.4 0.05 0.05 7.6 × 10


−5

S9.54
Find the rate law.
rate = k[CH3COCH3]^x [Br2]^y [H+]^z

[exp 1] / [exp 5]:


5.7e-5 / 7.6e-5 = (0.30/0.40)^x (0.050/0.050)^y (0.050/0.050)^z
3/4 = (3/4)^x
x=1

[exp 1] / [exp 2]:


5.7e-5 / 5.7e-5 = (0.30/0.30) (0.050/0.10)^y (0.050/0.050)^z
1 = (1/2)^y
y=0

[exp 1] / [exp 3]:


5.7e-5 / 1.2e-4 = (0.30/0.30) (0.050/0.10)^z
19/40 = (1/2)^z
z≈1

9.E.13 https://chem.libretexts.org/@go/page/41334
Use [exp 1] to find rate constant.
5.7e-5 M/s = k(0.30 M)(0.050 M)
k = 3.8e-3 M^(-1)-s^(-1)

"calculate the rate of disappearance of bromine if the initial concentration are .600mol/L, 0.200 mol/L, and .10 mol/L for
propanone, bromine and H+."
rate = (3.8e-3 M^(-1)-s^(-1))(0.600 M)(0.10 M)
rate = 2.28e-4 M/s

Q9.55
Determine the rate law for the following reaction:
N2 O2 + H2 → H2 O + N2 O (9.E.78)

In addition, determine which of the following actions would alter the value of k ?
a. Increase in pressure of N O
2 2

b. Increase in volume size of container


c. Increase in temperature
d. Addition of catalyst to the container.
e. None of the above ;a rate constant is always constant.

Q9.56
Consider the mechanism for the association of iodine atoms to create molecular iodine.

2 I(g) ⇌ I (9.E.79)
2(g)


I (g) + M (g) → I2(g) + M(g) (9.E.80)
2

With the respect of the first step is at equilibrium, determine the expected rate law (d/dt)[I2(g)] in terms of k1, k-1, k2, [I], and [M].

Q9.57
Consider the following reaction:
C3 H8(g) + 5 O2(g) − > 3C O2(g) + 4 H2 O(g) (9.E.81)

If propane (C3H8) is burning at a rate of 0.15 M/s^-1, calculate the rate of formation of CO2.

9.4: More Complex Reactions


S9.57
First, express the reaction with the differential rate equation for the reactants and products involved.
-(d/dt)[C3H8] = (1/3)(d/dt)[CO2]
Then, use the given burning rate of propane and plug it into the differential equation.
(d/dt)[CO2] = 3(0.15M/s^-1)
(d/dt)[CO2] = 0.45 M/s^-1

9.5: The Effect of Temperature on Reaction Rates


Q9.26
If kept in a refrigerator, fresh fish will last for 3 days. If kept in a freezer, it will last for 6 months. Assuming that the temperature in
the refrigerator is 5°C, and the temperature in the freezer is -10°C, calculate the activation energy for the bacterial spoiling of fish.
Assume that the spoiling time is the 1/e lifetime instead of the half-life.

9.E.14 https://chem.libretexts.org/@go/page/41334
Q9.27
Find the activation energy of a reaction whose rate constant is multiplied by 6.50 when T is increased from 300.0 K to 310.0 K. For
a reaction with Ea = 19 kJ/mol, by what factor is k multiplied when T increases from 300.0 K to 310.0 K?

Q9.28
The kinetics of the browning of juice from Golden Delicious apples was studied; at 20°C k=7.87×10-3/week, and at 37°C
k=0.139/week. What is the activation energy for the browning of Golden Delicious apple juice?

9.6: Potential Energy Surfaces

9.7: Theories of Reaction Rates

9.8: Isotope Effects in Chemical Reactions

9.9: Reactions in Solution

9.10: Fast Reactions in Solution

9.11: Oscillating Reactions

9.E: Chemical Kinetics (Exercises) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

9.E.15 https://chem.libretexts.org/@go/page/41334
CHAPTER OVERVIEW

10: Enzyme Kinetics


Enzyme kinetics is the study of the chemical reactions that are catalyzed by enzymes. In enzyme kinetics, the reaction rate is
measured and the effects of varying the conditions of the reaction are investigated.
10.1: General Principles of Catalysis
10.2: The Equations of Enzyme Kinetics
10.3: Chymotrypsin- A Case Study
10.4: Multisubstrate Systems
10.5: Enzyme Inhibition
10.6: Allosteric Interactions
10.7: The Effect of pH on Enzyme Kinetics
10.8: The Effect of Temperature on Enzyme Kinetics
10.E: Exercises

Thumbnail: Enzyme changes shape by induced fit upon substrate binding to form enzyme-substrate complex. Hexokinase has a
large induced fit motion that closes over the substrates adenosine triphosphate and xylose. (PDB: 2E2N, 2E2Q). (CC BY
4.0; Thomas Shafee).

10: Enzyme Kinetics is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1
10.1: General Principles of Catalysis
As can be seen from the Arrhenius equation, the magnitude of the activation energy, E , determines the value of the rate constant,
a

k , at a given temperature and thus the overall reaction rate. Catalysts provide a means of reducing E and increasing the reaction
a

rate. Catalysts are defined as substances that participate in a chemical reaction but are not changed or consumed. Instead they
provide a new mechanism for a reaction to occur which has a lower activation energy than that of the reaction without the catalyst.
Homogeneous catalysis refers to reactions in which the catalyst is in solution with at least one of the reactants whereas
heterogeneous catalysis refers to reactions in which the catalyst is present in a different phase, usually as a solid, than the reactants.
Figure 10.1.1 shows a comparison of energy profiles of a reaction in the absence and presence of a catalyst.

Figure 10.1.1 : Comparison of energy profiles with and without catalyst present.
Consider a non-catalyzed elementary reaction
k

A⟶ P (10.1.1)

which proceeds at rate k at a certain temperature. The reaction rate can be expressed as
d [A]
= −k [A] (10.1.2)
dt

In the presence of a catalyst C, we can write the reaction as


kcat

A+C ⟶ P+C (10.1.3)

and the reaction rate as


d [A]
= −k [A] − kcat [A] [C] (10.1.4)
dt

where the first term represents the uncatalyzed reaction and the second term represents the catalyzed reaction. Because the reaction
rate of the catalyzed reaction is often magnitudes larger than that of the uncatalyzed reaction (i.e. k ≫ k ), the first term can often
cat

be ignored.

Acid Catalysis
A common example of homogeneous catalysts are acids and bases. For example, given an overall reaction is S→ P . If k is the
rate, then
d [P]
= k [S] (10.1.5)
dt

The purpose of an enzyme is to enhance the rate of production of the product P. The equations of the acid-catalyzed reaction are
k1
+ −
S + AH ⇌ SH +A (10.1.6)
k−1

k2
+ +
SH + H2 O → P + H3 O (10.1.7)

k3
+ −
H3 O +A ⇌ AH + H2 O (10.1.8)
k−3

The full set of kinetic equations is

10.1.1 https://chem.libretexts.org/@go/page/41361
d [S] −
+
= −k1 [S] [AH ] + k−1 [SH ] [A ] (10.1.9)
dt

d [AH ] − −
+ +
= −k1 [S] [AH ] + k−1 [SH ] [A ] − k−3 [AH ] + k3 [ H3 O ] [A ] (10.1.10)
dt
+
d [SH ]
+ − +
= k1 [S] [AH ] − k−1 [SH ] [A ] − k2 [SH ] (10.1.11)
dt

d [A ]
+ − − +
= k1 [S] [AH ] − k−1 [SH ] [A ] − k2 [ A ] [ H3 O ] + k−3 [AH ] (10.1.12)
dt

d [P]
+
= k2 [SH ] (10.1.13)
dt
+
d [ H3 O ]
+ + −
= −k2 [SH ] − k3 [ H3 O ] [A ] + k−3 [AH ] (10.1.14)
dt

We cannot easily solve these, as they are nonlinear. However, let us consider two cases k2 ≫ k−1 [ A

] and −
k2 ≪ k−1 [ A ] . In
both cases, SH is consumed quickly, and we can apply a steady-state approximation:
+

+
d [SH ]
− + +
= k1 [S] [AH ] − k−1 [ A ] [SH ] − k2 [SH ] =0 (10.1.15)
dt

Rearranging in terms of SH +
yields
k1 [S] [AH ]
+
[SH ] = (10.1.16)

k−1 [ A ] + k2

and the rate of production of P can be written as


d [P] k1 k2 [S] [AH ]
+
= k2 [SH ] = (10.1.17)

dt k−1 [ A ] + k2

In the case where k 2 ≫ k−1 [ A



] , Equation 10.1.17 can be written as
d [P]
= k1 [S] [AH ] (10.1.18)
dt

which is known as a general acid-catalyzed reaction. On the other hand, if k2 ≪ k−1 [ A



] , we can use an equilibrium
approximation to write the rate of production of P as
d [P] k1 k2 [S] [AH ] k1 k2 +
= = [S] [ H ] (10.1.19)

dt k−1 [ A ] k−1 K

where K is the acid dissociation constant:


− +
[A ] [H ]
K = (10.1.20)
[AH ]

In this case, the reaction is hydrogen ion-catalyzed.

Enzyme Catalysis
To live, grow, and reproduce, microorganisms undergo a variety of chemical changes. They alter nutrients so they can
enter the cell and they change them once they enter in order to synthesize cell parts and obtain energy.
Metabolism refers to all of the organized chemical reactions in a cell. Reactions in which chemical compounds are
broken down are called catabolic reactions while reactions in which chemical compounds are synthesized are termed
anabolic reactions. All of these reactions are under the control of enzymes.
Enzymes are substances present in the cell in small amounts that function to speed up or catalyze chemical
reactions. On the surface of the enzyme is usually a small crevice that functions as an active site or catalytic site to
which one or two specific substrates are able to bind. (Anything that an enzyme normally combines with is called a

10.1.2 https://chem.libretexts.org/@go/page/41361
substrate.) The binding of the substrate to the enzyme causes the flexible enzyme to change its shape slightly through
a process called induced fit to form a tempore intermediate called an enzyme-substrate complex (Figure 1).

Figure 10.1.2 : Enzymes are substances present in the cell in small amounts which speed up or catalyze chemical
reactions. Enzymes speed up the rate of chemical reactions because they lower the energy of activation, the energy
that must be supplied in order for molecules to react with one another. Enzymes lower the energy of activation by
forming an enzyme-substrate complex.
Enzymes speed up the rate of chemical reactions because they lower the energy of activation, the energy that must be supplied in
order for molecules to react with one another (Figure 10.1.3). Like homogeneous catalysts discussed above, enzymes lower the
energy of activation by forming an enzyme-substrate complex allowing products of the enzyme reaction to be formed and released
(Figure 10.1.2).

Figure 10.1.3 : An enzyme speeds up a chemical reaction by lowering its energy of activation, the energy that must be
supplied in order for molecules to react with one another.

Enzyme-Substrate Reactions
Enzymes are substances present in the cell in small amounts which speed up or catalyze chemical reactions. Enzymes speed up
the rate of chemical reactions because they lower the energy of activation, the energy that must be supplied in order for
molecules to react with one another. Enzymes lower the energy of activation by forming an enzyme-substrate complex.

Many enzymes require a nonprotein cofactor to assist them in their reaction. In this case, the protein portion of the
enzyme, called an apoenzyme, combines with the cofactor to form the whole enzyme or haloenzyme (Figure 10.1.4).
Some cofactors are ions such as Ca++, Mg++, and K+; other cofactors are organic molecules called coenzymes which
serve as carriers for chemical groups or electrons. NAD+, NADP+, FAD, and coenzyme A (CoA) are examples of
coenzymes.

10.1.3 https://chem.libretexts.org/@go/page/41361
Figure 10.1.4 : An apoenzyme and cofactor combine to form a haloenzyme. If the cofactor is an organic molecule, it is
called a coenzyme.
Enzymes are generally globular proteins. (Some RNA molecules called ribozymes can also be enzymes. These are usually found in
the nuclear region of cells and catalyze the splitting of RNA molecules.). Enzymes are catalysts that breakdown or synthesize more
complex chemical compounds. They allow chemical reactions to occur fast enough to support life. Enzymes speed up the rate of
chemical reactions because they lower the energy of activation, the energy that must be supplied in order for molecules to react
with one another. Anything that an enzyme normally combines with is called a substrate. Enzymes are very efficient with a
typically enzyme generally able to catalyze between 1 and 10,000 molecules of substrate per second. The means that enzymes are
only have to be present in small amounts in the cell since. They are not altered during their reaction and are highly specific for their
substrate, with generally one specific enzyme dedicated for each specific chemical reaction.

Factors that affect the rate of enzyme reactions


Enzyme activity is affected by a number of factors including:
a. The concentration of enzyme: Assuming a sufficient concentration of substrate is available, increasing enzyme
concentration will increase the enzyme reaction rate.
b. The concentration of substrate: At a constant enzyme concentration and at lower concentrations of substrates, the
substrate concentration is the limiting factor. As the substrate concentration increases, the enzyme reaction rate increases.
However, at very high substrate concentrations, the enzymes become saturated with substrate and a higher concentration of
substrate does not increase the reaction rate.
c. Inhibitors: inhibitors will inhibit the activity of enzyme and decrease the rate of reaction. Enzyme inhibitors will bind to
enzyme active sites and could modify the chemistry of an active site which can stop a substrate from entering.
d. The temperature: Each enzyme has an optimum temperature at which it works best. A higher temperature generally
results in an increase in enzyme activity (Arrhenius kinetics). As the temperature increases, molecular motion increases
resulting in more molecular collisions. If, however, the temperature rises above a certain point, the heat will denature the
enzyme, causing it to lose its three-dimensional functional shape by denaturing its hydrogen bonds. Cold temperature, on
the other hand, slows down enzyme activity by decreasing molecular motion.
e. The pH: Each enzyme has an optimal pH that helps maintain its three-dimensional shape. Changes in pH may denature
enzymes by altering the enzyme's charge. This alters the ionic bonds of the enzyme that contribute to its functional shape.
f. The salt concentration: Each enzyme has an optimal salt concentration. Changes in the salt concentration may also
denature enzymes.

Applications of Enzymes
Enzymes are essential to maintain homeostasis because any malfunction of an enzyme could lead to diseases. Therefore,
pharmaceutical companies study enzyme to manipulate and synthesis new medicine. Besides their medicinal applications, enzymes
in industry are important because enzymes help breaking down cellulose, wastes, etc. Enzymes are essential in the process of
making new products in many industries such as pharmaceutical, food, paper, wine, etc.

References
1. "Chapter 4 Enzyme."Copyrighted IHW, Oct. 2005. Web. 5 Mar. 2011. biologymad.com/resources/Ch%2...0-20enzyme.pdf
2. Chang, Raymond. Physical Chemistry for the Biosciences. Sausalito, CA: University Science Books, 2005. Pages 363-384.

Contributors and Attributions


Mark Tuckerman (New York University)

10.1.4 https://chem.libretexts.org/@go/page/41361
Dr. Gary Kaiser (COMMUNITY COLLEGE OF BALTIMORE COUNTY, CATONSVILLE CAMPUS)

10.1: General Principles of Catalysis is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

10.1.5 https://chem.libretexts.org/@go/page/41361
10.2: The Equations of Enzyme Kinetics
In biological systems, enzymes act as catalysts and play a critical role in accelerating reactions, anywhere from 10 to 10 times 3 17

faster than the reaction would normally proceed. Enzymes are high-molecular weight proteins that act on a substrate, or reactant
molecule, to form one or more products.

Michaelis-Menten Enzyme Kinetics


Enzymes are highly specific catalysts for biochemical reactions, with each enzyme showing a selectivity for a single reactant, or
substrate. For example, the enzyme acetylcholinesterase catalyzes the decomposition of the neurotransmitter acetylcholine to
choline and acetic acid. Many enzyme–substrate reactions follow a simple mechanism that consists of the initial formation of an
enzyme–substrate complex, ES , which subsequently decomposes to form product, releasing the enzyme to react again.

Figure 10.2.1 : An enzyme catalyzes the reaction of two substrates and to form one product. from Wikipedia.
This is described within the following multi-step mechanism
k1 k2

E + S ⇌ ES ⇌ E + P (10.2.1)
k−1 k−2

where k , k , k , and k
1 –1 2 –2 are rate constants. The reaction’s rate law for generating the product [P ] is
d[P ]
rate = = k2 [ES] − k−2 [E][P ] (10.2.2)
dt

However, if we make measurement early in the reaction, the concentration of products is negligible, i.e.,
[P ] ≈ 0 (10.2.3)

and we can ignore the back reaction (second term in right side of Equation 10.2.2). Then under these conditions, the reaction’s rate
is
d[P ]
rate = = k2 [ES] (10.2.4)
dt

To be analytically useful we need to write Equation 10.2.4 in terms of the reactants (e.g., the concentrations of enzyme and
substrate). To do this we use the steady-state approximation, in which we assume that the concentration of ES remains
essentially constant. Following an initial period, during which the enzyme–substrate complex first forms, the rate at which ES
forms
d[ES]
= k1 [E][S] = k1 ([E ]0 − [ES])[S] (10.2.5)
dt

is equal to the rate at which it disappears


d[ES]
− = k−1 [ES] + k2 [ES] (10.2.6)
dt

where [E] is the enzyme’s original concentration. Combining Equations 10.2.5 and 10.2.6 gives
0

k1 ([E ]0 − [ES])[S] = k−1 [ES] + k2 [ES] (10.2.7)

which we solve for the concentration of the enzyme–substrate complex

10.2.1 https://chem.libretexts.org/@go/page/41362
[E ]0 [S] [E ]0 [S]
[ES] = = (10.2.8)
k−1 + k2 Km + [S]
+ [S]
k1

where K is the Michaelis constant. Substituting Equation 10.2.8 into Equation 10.2.4 leaves us with our final rate equation.
m

d[P ] k2 [E ]0 [S]
= (10.2.9)
dt Km + [S]

A plot of Equation 10.2.9, as shown in Figure 10.2.1, is instructive for defining conditions where we can use the rate of an
enzymatic reaction for the quantitative analysis of an enzyme or substrate.

Figure 10.2.1 : Plot of Equation 10.2.9 showing limits for the analysis of substrates and enzymes in an enzyme-catalyzed chemical
kinetic method of analysis. The curve in the region highlighted in red obeys equation 10.2.11 and the curve in the area highlighted
in green follows Equation 10.2.10.
For high substrate concentrations, where [S] ≫ K , Equation 10.2.9 simplifies to
m

d[P ] k2 [E ]0 [S] k2 [E ]0 [S]


= ≈ = k2 [E ]0 = Vmax (10.2.10)
dt Km + [S] [S]

where V maxis the maximum rate for the catalyzed reaction. Under these conditions the reaction is zero-order in substrate and we
can use Vmax to calculate the enzyme’s concentration, typically using a variable-time method. At lower substrate concentrations,
where [S] ≪ K , Equation 10.2.9 becomes
m

d[P ] k2 [E ]0 [S] k2 [E ]0 [S] Vmax [S]


= ≈ = (10.2.11)
dt Km + [S] Km Km

The reaction is now first-order in substrate, and we can use the rate of the reaction to determine the substrate’s concentration by a
fixed-time method.
The Michaelis constant K is the substrate concentration at which the reaction rate is at half-maximum, and is an inverse measure
m

of the substrate's affinity for the enzyme—as a small K indicates high affinity, meaning that the rate will approach V
m more max

quickly. The value of K is dependent on both the enzyme and the substrate, as well as conditions such as temperature and pH.
m

The Michaelis constant K is the substrate concentration at which the reaction rate is at
m

half-maximum.
From the last two terms in Equation 10.2.11, we can express V max in terms of a turnover number (k
cat ):
Vmax = kcat [E ]o (10.2.12)

where [E] is the enzyme concentration and k is the turnover number, defined as the maximum number of substrate molecules
0 cat

converted to product per enzyme molecule per second. Hence, the turnover number is defined as the maximum number of chemical
conversions of substrate molecules per second that a single catalytic site will execute for a given enzyme concentration [E] . o

10.2.2 https://chem.libretexts.org/@go/page/41362
Example 10.2.1: Turnover number of acetylcholinesterase

Acetylcholinesterase (AChE) may be one of the fastest enzymes. It hydrolyzes acetylcholine to choline and an acetate group.
One of the earliest values of the turnover number was 3 × 10 (molecules of acetylcholine) per minute per molecule of
7

enzyme. A more recent value at 25°C, pH = 7.0, acetylcholine concentration of 2.5 × 10 M , was found to be −3

7.4 × 10 mi n
5
(J Biol Chem. 236 (8): 2292–5.).
−1

There may be some 30 active centers per molecule. AChE is a serine hydrolase that reacts with acetylcholine at close to the
diffusion-controlled rate. A diffusion-controlled reaction occurs so quickly that the reaction rate is the rate of transport of the
reactants through the solution; as quickly as the reactants encounter each other, they react.

The Significance of K M and V


max

The Michaelis-Menten model is used in a variety of biochemical situations other than enzyme-substrate interaction, including
antigen-antibody binding, DNA-DNA hybridization, and protein-protein interaction. It can be used to characterize a generic
biochemical reaction, in the same way that the Langmuir equation can be used to model generic adsorption of biomolecular
species. When an empirical equation of this form is applied to microbial growth. The experimentally determined parameters values
vary wildly between enzymes (Table 10.2.1):
Table 10.2.1 : Enzyme Kinetic parameters
Enzyme Km (M) kcat (1/s) kcat / Km (1/M.s)

Chymotrypsin 1.5 × 10−2 0.14 9.3

Pepsin 3.0 × 10−4 0.50 1.7 × 103

Tyrosyl-tRNA synthetase 9.0 × 10−4 7.6 8.4 × 103

Ribonuclease 7.9 × 10−3 7.9 × 102 1.0 × 105

Carbonic anhydrase 2.6 × 10−2 4.0 × 105 1.5 × 107

Fumarase 5.0 × 10−6 8.0 × 102 1.6 × 108

While K is equal to the substrate concentration at which the enzyme converts substrates into products at half its maximal rate and
m

hence is related to the affinity of the substrate for the enzyme. The catalytic rate k is the rate of product formation when the
cat

enzyme is saturated with substrate and therefore reflects the enzyme's maximum rate. The rate of product formation is dependent
on both how well the enzyme binds substrate and how fast the enzyme converts substrate into product once substrate is bound. For
a kinetically perfect enzyme, every encounter between enzyme and substrate leads to product and hence the reaction velocity is
only limited by the rate the enzyme encounters substrate in solution. From Equation 10.2.8, the catalytic efficiency of a protein can
be evaluated.
kcat k2 k1 k2
= = (10.2.13)
Km Km k−1 + k2

This k /K ratio is called the specificity constant measure of how efficiently an enzyme converts a substrate into product. It has
cat m

a theoretical upper limit of 108 – 1010 /M.s; enzymes working close to this, such as fumarase, are termed superefficient (Table
10.2.1).

Determining V and K from experimental data can be difficult and the most common way is to determine initial rates, v , from
m m 0

experimental values of [P ] or [S] as a function of time. Hyperbolic graphs of v vs. [S] can be fit or transformed as we explored
0

with the different mathematical transformations of the hyperbolic binding equation to determine K . These included:
d

nonlinear hyperbolic fit (e.g., Figure 10.2.1)


double reciprocal plot (e.g., Lineweaver–Burk plot discussed below
Eadie-Hofstee plot

Lineweaver–Burk plot
Another commonly-used plot in examining enzyme kinetics is the Lineweaver-Burk plot, in with the inverse of the reaction rate,
1/r, is plotted against the inverse of the substrate concentration 1/ [S]. Rearranging Equation 10.2.10,

10.2.3 https://chem.libretexts.org/@go/page/41362
1 KM + [S] KM 1 1
= = + (10.2.14)
r k2 [E] [S] k2 [E] [S] k2 [E]
0 0 0

Tthe Lineweaver–Burk plot (or double reciprocal plot) is a graphical representation of the Lineweaver–Burk equation of enzyme
kinetics, described by Hans Lineweaver and Dean Burk in 1934 (Figure 10.2.2). The Lineweaver-Burk plot results in a straight line
with the slope equal to K /k [E] and y -intercept equal to 1/k [E] which is 1/V
M 2 0 2 0
via Equation 10.2.10. max

Figure 10.2.2 : Lineweaver–Burk plot of Michaelis–Menten kineitcs.


The plot provides a useful graphical method for analysis of the Michaelis–Menten equation:
Vmax [S]
V = (10.2.15)
Km + [S]

Taking the reciprocal gives


1 Km + [S] Km 1 1
= = + (10.2.16)
V Vmax [S] Vmax [S] Vmax

where
V is the reaction velocity (the reaction rate),
Km is the Michaelis–Menten constant,
V max is the maximum reaction velocity, and
[S] is the substrate concentration.

The Lineweaver–Burk plot was widely used to determine important terms in enzyme kinetics, such as K and V , before the
m max

wide availability of powerful computers and non-linear regression software. The y-intercept of such a graph is equivalent to the
inverse of V max; the x-intercept of the graph represents −1/K . It also gives a quick, visual impression of the different forms of
m

enzyme inhibition.

Example 10.2.2

The reaction between nicotineamide mononucleotide and ATP to form nicotineamide–adenine dinucleotide and pyrophosphate
is catalyzed by the enzyme nicotinamide mononucleotide adenylyltransferase. The following table provides typical data
obtained at a pH of 4.95. The substrate, S, is nicotinamide mononucleotide and the initial rate, v, is the μmol of nicotinamide–
adenine dinucleotide formed in a 3-min reaction period.

[S] (mM) v (μmol) [S] (mM) v (μmol)

0.138 0.148 0.560 0.324

0.220 0.171 0.766 0.390

0.291 0.234 1.460 0.493

Determine values for Vmax and Km.


Solution

10.2.4 https://chem.libretexts.org/@go/page/41362
Figure 13.12 shows the Lineweaver–Burk plot for this data and the resulting regression equation. Using the y-intercept, we
calculate Vmax as
Vmax= 1 / y−intercept = 1 / 1.708 mol = 0.585 mol

and using the slope we find that Km is


Km = slope × Vmax = 0.7528 molimM × 0.585 mol = 0.440 mM

Figure 13.12: Linweaver–Burk plot and regression equation for the data in Example 13.6.

-diphenyl oxidase
The following data are for the oxidation of catechol (the substrate) to o-quinone by the enzyme o-diphenyl oxidase. The
reaction is followed by monitoring the change in absorbance at 540 nm. The data in this exercise are adapted from jkimball.

[catechol] (mM) 0.3 0.6 1.2 4.8

rate (∆AU/min) 0.020 0.035 0.048 0.081

Determine values for Vmax and Km.


Click here to review your answer to this exercise.

The double reciprocal plot distorts the error structure of the data, and it is therefore unreliable for the determination of enzyme
kinetic parameters. Although it is still used for representation of kinetic data, non-linear regression or alternative linear forms of the
Michaelis–Menten equation such as the Hanes-Woolf plot or Eadie–Hofstee plot are generally used for the calculation of
parameters.

Problems with the Method


The Lineweaver–Burk plot is classically used in older texts, but is prone to error, as the y-axis takes the reciprocal of the rate of
reaction – in turn increasing any small errors in measurement. Also, most points on the plot are found far to the right of the y-
axis (due to limiting solubility not allowing for large values of [S] and hence no small values for 1/[S]), calling for a large
extrapolation back to obtain x- and y-intercepts.

When used for determining the type of enzyme inhibition, the Lineweaver–Burk plot can distinguish competitive, non-competitive
and uncompetitive inhibitors. Competitive inhibitors have the same y-intercept as uninhibited enzyme (since V is unaffected by
max

competitive inhibitors the inverse of V also doesn't change) but there are different slopes and x-intercepts between the two data
max

10.2.5 https://chem.libretexts.org/@go/page/41362
sets. Non-competitive inhibition produces plots with the same x-intercept as uninhibited enzyme (K is unaffected) but different
m

slopes and y-intercepts. Uncompetitive inhibition causes different intercepts on both the y- and x-axes but the same slope.

Eadie–Hofstee Plot
The Eadie–Hofstee plot is a graphical representation of enzyme kinetics in which reaction rate is plotted as a function of the ratio
between rate and substrate concentration and can be derived from the Michaelis–Menten equation (10.2.9) by inverting and
multiplying with V :
max

Vmax Vmax (Km + [S]) Km + [S]


= = (10.2.17)
v Vmax [S] [S]

Rearrange:

vKm v[S] vKm


Vmax = + = +v (10.2.18)
[S] [S] [S]

Isolate v:
v
v = −Km + Vmax (10.2.19)
[S]

Figure 10.2.3 : The Eadie-Hofstee plot is a more accurate linear plotting method with v is plotted against v/[S].
A plot of v against v/[S] will hence yield V max as the y-intercept, V /Kmax as the x-intercept, and K as the negative slope
m m

(Figure 10.2.3). Like other techniques that linearize the Michaelis–Menten equation, the Eadie-Hofstee plot was used historically
for rapid identification of important kinetic terms like K and V
m , but has been superseded by nonlinear regression methods that
max

are significantly more accurate and no longer computationally inaccessible. It is also more robust against error-prone data than the
Lineweaver–Burk plot, particularly because it gives equal weight to data points in any range of substrate concentration or reaction
rate (the Lineweaver–Burk plot unevenly weights such points). Both Eadie-Hofstee and Lineweaver–Burk plots remain useful as a
means to present data graphically.

Problems with the Method


One drawback from the Eadie–Hofstee approach is that neither ordinate nor abscissa represent independent variables: both are
dependent on reaction rate. Thus any experimental error will be present in both axes. Also, experimental error or uncertainty
will propagate unevenly and become larger over the abscissa thereby giving more weight to smaller values of v/[S]. Therefore,
the typical measure of goodness of fit for linear regression, the correlation coefficient R, is not applicable.

Contributors and Attributions


David Harvey (DePauw University)
World Public Library
Wikipedia
Dr. S.K. Khare (IIT Delhi) via NPTEL

10.2: The Equations of Enzyme Kinetics is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

10.2.6 https://chem.libretexts.org/@go/page/41362
10.3: Chymotrypsin- A Case Study
Chymotrypsin is a digestive enzyme belonging to a super family of enzymes called serine proteases. It uses an active serine residue
to perform hydrolysis on the C-terminus of the aromatic amino acids of other proteins. Chymotrypsin is a protease enzyme that
cleaves on the C-terminal phenylalanine (F), tryptophan (W), and tyrosine (Y) on peptide chains. It shows specificity for aromatic
amino acids because of its hydrophobic pocket.

Introduction
Chymotrypsin is one of the most studied enzymes due to its two phase kinetics: pre-steady-state and steady state. The study of
these two kinetic states gives evidence of the "Ping-Pong" mechanism, the formation of covalent complexes leading to covalent
hydrolysis reactions, and the rate of the catalyzed reactions. Synthesis of chymotrypsin occurs primarily in the pancreas. Instead of
the active form, however, it is produced as an inactive zymogen called chymotrypsinogen to prevent its protease activity from
digesting the pancreas. Upon secretion into the lumen of the small intestine, it is converted to its active form by another enzyme
called trypsin. This dependence of a different enzyme for the activation of a protease is a common way for the body to prevent the
digestion of organs and other harmful enzymatic side-effects.
Chymotrypsin operates through a general mechanism known as the ping-pong mechanism (Figure 10.3.1) whereby the enzyme
reacts with a substrate to form an enzyme intermediate. This intermediate has different properties than the initial enzyme, so to
regenerate the initial enzymatic activity, it must react with a secondary substrate. This process is illustrated below:

Figure 10.3.1 : Generic Ping-Pong Mechanism


More specifically, chymotrypsin operates through a particular type of ping-pong mechanism called covalent hydrolysis. This means
that the enzyme first forms a covalent bond with the target substrate, displacing the more stable moiety into solution. This enzyme-
substrate complex is called the enzyme intermediate. The intermediate then reacts with water, which displaces the remaining part of
the initial substrate and reforms the initial enzyme.

Chymotrypsin, like most enzymes, is specific in the types of substrates with which it reacts. As a protease, it cleaves polypeptides,
and its inherent specificity allows it to act only on the carboxy-terminal of aromatic residues. It is a somewhat complicated
mechanism, and is best explained in a series of steps.

Step 1: The target enters the active site of chymotrypsin, and it is held there by hydrophobic interactions between exposed non-
polar groups of enzyme residues and the non-polar aromatic side-chain of the substrate. It is important to note the hydrogen bond
between the Schiff nitrogen on histidine-57 and the oxygen side-chain of serine-195.

10.3.1 https://chem.libretexts.org/@go/page/41363
Step 2: Aided by the histidine-serine hydrogen bonding, the hydroxyl group on serine-195 performs a nucleophilic attack on the
carbonyl carbon of an aromatic amino acid while simultaneously transferring the hydroxyl hydrogen to the histidine Schiff
nitrogen. This attack pushes the pi carbonyl electrons onto the carbonyl oxygen, forming a short-lived intermediate consisting of a
c-terminal carbon with four single bonds: an oxygen anion, the beta-carbon of the aromatic amino acid, the n-terminus of the
subsequent amino acid of the substrate protein, and the serine-195 side-chain oxygen.

Step 3: This intermediate is short-lived, as the oxyanion electrons reform the pi bond with the c-terminus of the aromatic amino
acid. The bond between the carboxy-terminus of the aromatic amino acid and the n-terminus of the subsequent residue is cleaved,
and its electrons are used to extract the hydrogen of the protonated Schiff nitrogen on histidine-57. The bonds between the carbonyl
carbon and the serine-195 oxygen remain in an ester configuration. This is called the acyl-enzyme intermediate. The c-terminal side
of the polypeptide is now free to dissociate from the active site of the enzyme.

Step 4: Water molecules are now able to enter and bind to active site through hydrogen bonding between the hydrogen atoms of
water and the histidine-57 Schiff nitrogen.

Step 5: The water oxygen now makes a nucleophilic attack on the carbonyl carbon of the acyl-enzyme intermediate, pushing the
carbonyl’s pi electrons onto the carbonyl carbon as histidine-57 extracts one proton from water. This forms another quaternary

10.3.2 https://chem.libretexts.org/@go/page/41363
carbon covalently bonded with serine, a hydroxyl, an oxyanion, and the aromatic amino acid. The proton on the recently protonated
histidine-57 is now able to make a hydrogen bond with the serine oxygen.

Step 6: The oxyanion electrons reform the carbonyl pi bond, cleaving the bond between the carbonyl carbon and the serine
hydroxyl. The electrons in this bond are used by the serine oxygen to deprotonate the histidine Schiff nitrogen and reform the
original enzyme. The substrate no longer has affinity for the active site, and it soon dissociates from the complex.

Kinetics
Experiments were conducted in 1953 by B.S. Hartley and B.A. Kilby to investigate the kinetics of chymotrypsin-catalyzed
hydrolysis. Instead of using a poly-peptide chain as a substrate, they used a nitro-phenyl ester, p-nitrophenyl acetate, that resembles
an aromatic amino acid. Hydrolysis of this compound by chymotrypsin at the carbonyl group yields acetate and nitrophenolate, the
latter of which absorbs near 400 nm light and its concentration can thus be measured by spectrophotometry (Figure 10.3.2).

Figure 10.3.2 : Catalytic activity of he hydrolysis of p-nitropenolate under


Spectrophotometric analysis of chymotrypsin acting on nitrophenylacetate showed that nitrophenolate was produced at a rate
independent of substrate concentration, proving that the only factor contributing to the rate of product formation is the
concentration of enzyme; this is typical for enzyme-substrate kinetics. However, when the slope of the 0-order absorbance plot was
traced back to the starting point (time = 0), it was found that the initial concentration of nitrophenolate was not 0. In fact, it showed
a 1:1 stoichiometric ratio with the amount of chymotrypsin used in the assay. This can only be explained by the fact that hydrolysis
by chymotrypsin is biphasic in nature, meaning that it proceeds in two distinct steps.
1. The first step, which describes the initial burst of nitrophenolate seen in Hartley and Kilby’s absorbance plot, is the fastest. The
attack of the nitrophenyl acetate substrate by chymotrypsin immediately cleaves the nitrophenolate moiety and leaves the
acetate group attached to chymotrypsin, rendering the enzyme inactive.
2. The second step has been deduced to involve the hydrolysis of the acetate group from the inactivated chymotrypsin to
regenerate the original enzyme.

10.3.3 https://chem.libretexts.org/@go/page/41363
To analytically determine the rate of catalysis, all substrates, products, and intermediates must be defined. Refer to the figure
below:

Using these abbreviations, kinetic analysis becomes less cumbersome.


1. The initial amount of enzyme can be represented as the sum of the free enzyme, the bound enzyme, and the inactive
intermediate.

[E ]o = [ES] + [ ES] + [E]

2. Assuming the initial step is the faster than the subsequent steps, the rate of nitrophenolate production can be described as:
d[ P1 ]
= k2 [ES]
dt

3. Likewise, the rate of acetate formation can be represented by the equation:


d[ P2 ]

= k3 [ ES]
dt

4. Therefore, the net change in concentration of the inactive intermediate can be deduced:
d[∗ES]

= k2 [ES] − k3 [ ES]
dt

5. The last inference that can be made from analysis of the measured kinetics data (Figure 10.3.2) is that the first step of the
reaction equilibrates rapidly, and thus the change in bound substrate can be described in the following equation. This is a principal
tenet in analyzing the kinetics of chymotrypsin and is a ubiquitous mechanism in biological enzyme catalysis.
d[ES]
= k1 [E][S] − k−1 [ES] = 0
dt

6. Where:
k−1 [E][S]
= Ks =
k1 [ES]

7. Combining all of these quantities, we can deduce the catalytic rate constant as:
k2 k3
kcat =
k2 + k3

8. In ester hydrolysis, k3 >> k2 , so the resultant catalytic rate constant simplifies to:

kcat = k2

which is in agreement with the observed zeroth-order kinetics of Figure 10.3.2.


Table 10.3.1 : Kinetic Constants of the Chymotrypsin-Catalyzed Hydrolysis of p-Nitrophenyl Trimethylacetate at pH 8.2. 0.01 M tris-HCL
buffer, ionic strength 0.06, 25.6 ± 0.1 °C, 1.8% (v/v) acetonitrile-water. From M.L. Bender, F.J. Kezdy and F.C. Wedler, J. Chem. Educ. 44,
84 (1967

10.3.4 https://chem.libretexts.org/@go/page/41363
k2 0.37 ± 0.11 s-1

k3 (1.3 ± 0.02) × 10-4 s-1

Ks (1.6 ± 0.5) × 10-3

kcat 1.3 × 10-4 s-1

KM 5.6 × 10-7

 Exercise 10.3.1

Speculate on how the catalytic rate constant can be determined from the spectrophotogram.

Answer
The catalytic rate constant can be deduced from the graph by simply determining the slope of the line where the reaction
demonstrates 0-order kinetics (the linear part)

 Exercise 10.3.2
How can product be consistently produced if the rate of change of the ES complex is 0?

Answer
This is pre-equilibrium kinetics in action. The ES complex is formed from E and S at a faster rate than any other step in the
reaction. As soon as ES is converted to *ES, another mole of ES is produced from an infinite supply of E + S. This means
that the amount of ES and E + S is constantly at equilibrium, and thus the change of either with respect to time is 0.

 Exercise 10.3.3

How would the rate of product formation change if:


a. the substrate concentration was doubled.
b. the enzyme concentration was doubled.
c. The reaction was carried out in mono-deuterated water instead of H2O (comment qualitatively).

Answer
a. No change.
b. Two fold increase.
c. Since water is involved in the final, slowest step of the mechanism, deuterating the water would decrease the rate of the
overall reaction from 5 to 30-fold.

 Exercise 10.3.4

Explain the role of hydrogen bonding in protein hydrolysis catalyzed by chymotrypsin.

Answer
Initially, hydrogen bonding between the enzymes histidine and serine side chains weakens the bond of serine’s O-H. This
allows for a facilitated nucleophilic attack of the hydroxyl Oxygen on the substrates carbonyl group. Conversely, in the
final step of the reaction, the bound serine oxygen forms a hydrogen bond with a protonated histidine, which allows for
easier cleavage from the substrate.

10.3.5 https://chem.libretexts.org/@go/page/41363
 Exercise 10.3.5

What would the spectrophotogram look like if the reaction proceeded via a steady-state mechanism instead of pre-equilibrium.

Answer
The graph would show similar 0-order kinetics, but the line would intercept the Y-axis at an absorbance of 0 instead of the
1:1 mole ratio of nitrophenolate to enzyme.

References
1. Chang, Raymond. Physical Chemistry for the Biosciences. Sansalito, CA: University Science, 2005.
2. Segel, Irwin H., Enzyme Kinetics: Behavior and Analysis of Rapid Equilibrium and Steady-State Enzyme Systems (Wiley
Classics Library)
3. Garret, Reginald. Grishman, Charles. Biochemistry. University of Virginia, 2007
4. Whitaker, John., Princles of Enzymology for the Food Sciences. University of California, Davis. 1994
5. “Chymotrypsin Mechanism” viewable at: www.sumanasinc.com/webcontent...motrypsin.html
6. Sumanas Inc, viewed on February 25th, 2009.
7. M.L. Bender, F.J. Kezdy and F.C. Wedler, J. Chem. Educ. 44, 84 (1967)

10.3: Chymotrypsin- A Case Study is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
7.2: Chymotrypsin by Elissa Nysetvold is licensed CC BY-NC-SA 4.0.

10.3.6 https://chem.libretexts.org/@go/page/41363
10.4: Multisubstrate Systems
The Michaelis –Menten model of enzyme kinetics was derived for single substrate reactions. Enzymatic reactions requiring
multiple substrates and yielding multiple products are more common and yielding multiple products are more common than single-
substrate reaction. In these types of reactions, the all the substrates involved are bound to the enzyme before catalysis of the
reaction takes place to release the products. Sequential reactions can be either ordered or random. In contrast to the Michealis-
Menton kinetics where a binary Enzyme-Substrate complex is generated in the mechanism ([ES], in bisubstrate enzyme reactions,
a ternary complex of the enzyme and two substrates is generated:
E

A + B−
↽⇀
−P + Q (10.4.1)

Bisubstrate reactions account for ~ 60% of the known enzymatic reactions. Multi-substrate reactions follow complex rate equations
that describe how the substrates bind and in what sequence. The analysis of these reactions is much simpler if the concentration of
substrate A is kept constant and substrate B varied. Under these conditions, the enzyme behaves just like a single-substrate enzyme
and a plot of v by [S] gives apparent K and VM constants for substrate B. If a set of these measurements is performed at
max

different fixed concentrations of A, these data can be used to work out what the mechanism of the reaction is. For an enzyme that
takes two substrates A and B and turns them into two products P and Q, there are two types of mechanism: ternary complex and
ping–pong.
How do you resolve the enzymes kinetics of these more complicated systems? The answer is fairly straightforward. You keep one
of the substrates (B, for example) fixed, and vary the other substrate (A) and obtain a series of hyperbolic plots of v vs A at o

different fixed B concentrations. This would give a series of linear 1/v vs. 1/A double-reciprocal plots (Lineweaver-Burk plots) as
well. The pattern of Lineweaver-Burk plots depends on how the reactants and products interact with the enzyme.

Sequential Mechanism
In this mechanism, both substrates must bind to the enzyme before any products are made and released.The substrates might bind
to the enzyme in a random fashion (A first then B or vice-versa) or in an ordered fashion (A first followed by B). An abbreviated
notation scheme developed by W.W. Cleland is shown in Figure 10.4.1 for the sequential random and sequential ordered
mechanisms (Figure 10.4.1). For both mechanisms, Lineweaver-Burk plots at varying A and different fixed values of B give a
series of intersecting lines. Derivative curves can be solved to obtain appropriate kinetic constants. In ordered sequential reactions,
all the substrates are first bound to the enzyme in a defined order or sequence. The products, too, are released after catalysis in a
defined order or sequence.
An example is the lactate dehydrogenase enzyme, which is a protein that catalyzes glucose metabolism. In this ordered mechanism,
the coenzyme, NADH, always binds first, with pyruvate binding afterward. During the reaction, the pyruvate is reduced to lactate
while NADH is oxidized to NAD+ by the enzyme. Lactate is then released first, followed by the release of NAD+.

Figure 10.4.2 : Reaction of the lactate dehydrogenase: pyruvate (left) is oxidized to lactate (right) by expense of NADH

Figure 10.4.3 : Ordered Sequential Mechanism for the lactate dehydrogenase enzyme
This is a characteristic of a ternary complex, which consists of three molecules that are bound together. Before catalysis, the
substrates and coenzyme are bound to the enzyme. After catalysis, the complex consists of the enzyme and products, NAD+ and
lactate.
In random sequential reactions, the substrates and products are bound and then released in no preferred order, or "random" order
(Figure 10.4.1). An example is the creatine kinase enzyme, which catalyzes creatine and ATP (the two substrates) to form
phosphocreatine and ADP (teh Products) in Figure 10.4.4. In this case, either substrates may bind first and either products may be

10.4.1 https://chem.libretexts.org/@go/page/41364
released first. A ternary complex is still observed for random sequential reactions. Before catalysis, the complex is generated that
includes the enzyme, ATP, and creatine. After catalysis, the complex consists of the enzyme, ADP, and phosphocreatine.

Figure 10.4.4 : The metabolism of creatine follows a random sequential mechanism.

Ping-Pong Mechanism
In this mechanism, one substrate bind first to the enzyme followed by product P release. Typically, product P is a fragment of the
original substrate A.The rest of the substrate is covalently attached to the enzyme E, which is designated as E'. Now the second
reactant, B, binds and reacts with the enzyme to form a covalent adduct with the A as it is covalentattached to the enzyme to form
product Q. This is now released and the enzyme is restored to its initial form, E. This represents a ping-pong mechanism. An
abbreviated notation scheme is shown below for the ping-pong mechanisms. For this mechanism, Lineweaver-Burk plots at varying
A and different fixed values of B give a series of parallel lines. An example of this type of reaction might be low molecular weight
protein tyrosine phosphatase against the small substrate p-initrophenylphosphate (A) which binds to the enzyme covalently with the
expulsion of the product P, the p-nitrophenol leaving group. Water (B) then comes in and covalently attacks the enzyme, forming an
adduct with the covalently bound phosphate releasing it as inorganic phosphate. In this particular example, however, you cannot
vary the water concentration and it would be impossible to generate the parallel Lineweaver-Burk plots characteristic of ping-
pong kinetics.

Figure 10.4.5 ): The Ping-Pong mechanism and associated Lineweaver-Burk.

Contributors and Attributions


Prof. Henry Jakubowski (College of St. Benedict/St. John's University)

10.4: Multisubstrate Systems is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

10.4.2 https://chem.libretexts.org/@go/page/41364
10.5: Enzyme Inhibition
Enzymes can be regulated in ways that either promote or reduce their activity. There are many different kinds of molecules that
inhibit or promote enzyme function, and various mechanisms exist for doing so. In some cases of enzyme inhibition, for example,
an inhibitor molecule is similar enough to a substrate that it can bind to the active site and simply block the substrate from binding.
When this happens, the enzyme is inhibited through competitive inhibition, because an inhibitor molecule competes with the
substrate for active site binding. On the other hand, in noncompetitive inhibition, an inhibitor molecule binds to the enzyme in a
location other than an allosteric site and still manages to block substrate binding to the active site.

Elucidating Mechanisms for the Inhibition of Enzyme Catalysis


When an inhibitor interacts with an enzyme it decreases the enzyme’s catalytic efficiency. An irreversible inhibitor covalently
binds to the enzyme’s active site, producing a permanent loss in catalytic efficiency even if we decrease the inhibitor’s
concentration. A reversible inhibitor forms a noncovalent complex with the enzyme, resulting in a temporary decrease in catalytic
efficiency. If we remove the inhibitor, the enzyme’s catalytic efficiency returns to its normal level.
There are several pathways for the reversible binding of an inhibitor to an enzyme, as shown in Figure 10.5.1. In competitive
inhibition the substrate and the inhibitor compete for the same active site on the enzyme. Because the substrate cannot bind to an
enzyme–inhibitor complex, EI, the enzyme’s catalytic efficiency for the substrate decreases. With noncompetitive inhibition the
substrate and the inhibitor bind to different active sites on the enzyme, forming an enzyme–substrate–inhibitor, or ESI complex.
The formation of an ESI complex decreases catalytic efficiency because only the enzyme–substrate complex reacts to form the
product. Finally, in uncompetitive inhibition the inhibitor binds to the enzyme–substrate complex, forming an inactive ESI
complex.

Figure 10.5.1 : Mechanisms for the reversible inhibition of enzyme catalysis. E: enzyme, S: substrate, P: product, I: inhibitor, ES:
enzyme–substrate complex, EI: enzyme–inhibitor complex, ESI: enzyme–substrate–inhibitor complex.
We can identify the type of reversible inhibition by observing how a change in the inhibitor’s concentration affects the relationship
between the rate of reaction and the substrate’s concentration. As shown in Figure 13.14, when we display kinetic data using as a
Lineweaver-Burk plot it is easy to determine which mechanism is in effect. For example, an increase in slope, a decrease in the x-
intercept, and no change in the y-intercept indicates competitive inhibition. Because the inhibitor’s binding is reversible, we can
still obtain the same maximum velocity—thus the constant value for the y-intercept—by adding enough substrate to completely
displace the inhibitor. Because it takes more substrate, the value of Km increases, which explains the increase in the slope and the
decrease in the x-intercept’s value.

Figure 10.5.2 : Linweaver–Burk plots for competitive inhibition, noncompetitive inhibition, and uncompetitive inhibition. The thick
blue line in each plot shows the kinetic behavior in the absence of inhibitor, and the thin blue lines in each plot show the change in
behavior for increasing concentrations of the inhibitor. In each plot, the inhibitor’s concentration increases in the direction of the
green arrow.

10.5.1 https://chem.libretexts.org/@go/page/41365
Example 10.5.1

Practice Exercise 13.3 provides kinetic data for the oxidation of catechol (the substrate) to o-quinone by the enzyme o-diphenyl
oxidase in the absence of an inhibitor. The following additional data are available when the reaction is run in the presence of p-
hydroxybenzoic acid, PBHA. Is PBHA an inhibitor for this reaction and, if so, what type of inhibitor is it?

[catechol] (mM) 0.3 0.6 1.2 4.8

rate (∆AU/min) 0.011 0.019 0.022 0.060

Solution
Figure 10.5.3 shows the resulting Lineweaver–Burk plot for the data in Practice Exercise 13.3 and Example 13.7. Although the
y-intercepts are not identical in value—the result of uncertainty in measuring the rates—the plot suggests that PBHA is a
competitive inhibitor for the enzyme’s reaction with catechol.

Figure 10.5.3 : Lineweaver–Burk plots for the data in Practice Exercise 13.3 and Example 13.7.

Exercise 10.5.1

Practice Exercise 13.3 provides kinetic data for the oxidation of catechol (the substrate) to o-quinone by the enzyme o-diphenyl
oxidase in the absence of an inhibitor. The following additional data are available when the reaction is run in the presence of
phenylthiourea. Is phenylthiourea an inhibitor for this reaction and, if so, what type of inhibitor is it? The data in this exercise
are adapted from jkimball.

[catechol] (mM) 0.3 0.6 1.2 4.8

rate (∆AU/min) 0.010 0.016 0.024 0.040

Click here to review your answer to this exercise.

Contributors and Attributions


David Harvey (DePauw University)

10.5: Enzyme Inhibition is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

10.5.2 https://chem.libretexts.org/@go/page/41365
10.6: Allosteric Interactions
Learning Objectives
When a substrate binds to one enzymatic subunit, the rest of the subunits are stimulated and become active. Ligands can
either have non-cooperativity, positive cooperativity or negative cooperativity.

A significant portion of enzymes function such that their properties can be studied using the Michaelis-Menten equation. However,
a particular class of enzymes exhibit kinetic properties that cannot be studied using the Michaelis-Menten equation. The rate
equation of these unique enzymes is characterized by an “S-shaped” sigmoidal curve, which is different from the majority of
enzymes whose rate equation exhibits hyberbolic curves. Allosteric regulation is the regulation of an enzyme or other protein by
binding an effector molecule at the protein's allosteric site (that is, a site other than the protein's active site). Effectors that enhance
the protein's activity are referred to as allosteric activators, whereas those that decrease the protein's activity are called allosteric
inhibitors. The term allostery refers to the fact that the regulatory site of an allosteric protein is physically distinct from its active
site. Allosteric regulations are a natural example of control loops, such as feedback from downstream products or feedforward from
upstream substrates. Long-range allostery is especially important in cell signaling.

Allosteric Modulation (Cooperativity)


Cooperativity is a phenomenon displayed by enzymes or receptors that have multiple binding sites where the affinity of the binding
sites for a ligand is increased, positive cooperativity, or decreased, negative cooperativity, upon the binding of a ligand to a binding
site. We also see cooperativity in large chain molecules made of many identical (or nearly identical) subunits (such as DNA,
proteins, and phospholipids), when such molecules undergo phase transitions such as melting, unfolding or unwinding. This is
referred to as subunit cooperativity (discussed below).
An example of positive cooperativity is the binding of oxygen to hemoglobin. One oxygen molecule can bind to the ferrous iron of
a heme molecule in each of the four chains of a hemoglobin molecule. Deoxy-hemoglobin has a relatively low affinity for oxygen,
but when one molecule binds to a single heme, the oxygen affinity increases, allowing the second molecule to bind more easily, and
the third and fourth even more easily. The oxygen affinity of 3-oxy-hemoglobin is ~300 times greater than that of deoxy-
hemoglobin. This behavior leads the affinity curve of hemoglobin to be sigmoidal, rather than hyperbolic as with the monomeric
myoglobin. By the same process, the ability for hemoglobin to lose oxygen increases as fewer oxygen molecules are bound.
Negative allosteric modulation (also known as allosteric inhibition) occurs when the binding of one ligand decreases the affinity
for substrate at other active sites. For example, when 2,3-BPG binds to an allosteric site on hemoglobin, the affinity for oxygen of
all subunits decreases. This is when a regulator is absent from the binding site.
Another instance in which negative allosteric modulation can be seen is between ATP and the enzyme Phosphofructokinase within
the negative feedback loop that regulates glycolysis. Phosphofructokinase (generally referred to as PFK) is an enzyme that
catalyses the third step of glycolysis: the phosphorylation of Fructose-6-phosphate into Fructose 1,6-bisphosphate. PFK can be
allosterically inhibited by high levels of ATP within the cell. When ATP levels are high, ATP will bind to an allosteric site on
phosphofructokinase, causing a change in the enzyme's three-dimensional shape. This change causes its affinity for substrate
(fructose-6-phosphate and ATP) at the active site to decrease, and the enzyme is deemed inactive. This causes glycolysis to cease
when ATP levels are high, thus conserving the body's glucose and maintaining balanced levels of cellular ATP. In this way, ATP
serves as a negative allosteric modulator for PFK, despite the fact that it is also a substrate of the enzyme.
Sigmoidal kinetic profiles are the result of enzymes that demonstrate positive cooperative binding. cooperativity refers to the
observation that binding of the substrate or ligand at one binding site affects the affinity of other sites for their substrates. For
enzymatic reactions with multiple substrate binding sites, this increased affinity for the substrate causes a rapid and coordinated
increase in the velocity of the reaction at higher [S] until V maxis achieved. Plotting the V vs. [S] for a cooperative enzyme, we
0

observe the characteristic sigmoidal shape with low enzyme activity at low substrate concentration and a rapid and immediate
increase in enzyme activity to V max as [S] increases. The phenomenon of cooperativity was initially observed in the oxygen-
hemoglobin interaction that functions in carrying oxygen in blood. Positive cooperativity implies allosteric binding – binding of the
ligand at one site increases the enzyme’s affinity for another ligand at a site different from the other site. Enzymes that demonstrate
cooperativity are defined as allosteric. There are several types of allosteric interactions: homotropic (positive) and heterotropic
(negative).

10.6.1 https://chem.libretexts.org/@go/page/41366
Figure 10.6.1 : Rate of Reaction (velocity) vs. Substrate Concentration.
Positive and negative allosteric interactions (as illustrated through the phenomenon of cooperativity) refer to the enzyme's binding
affinity for other ligands at other sites, as a result of ligand binding at the initial binding site. When the ligands interacting are all
the same compounds, the effect of the allosteric interaction is considered homotropic. When the ligands interacting are different,
the effect of the allosteric interaction is considered heterotropic. It is also very important to remember that allosteric interactions
tend to be driven by ATP hydrolysis.

The Hill Equation


The degree of cooperativity is determined by Hill equation (Equation 10.6.1) for non-Michaelis-Menten kinetics. The Hill equation
accounts for allosteric binding at sites other than the active site. n is the "Hill coefficient."
n n
[L] [L]
θ = = (10.6.1)
n n n
Kd + [L] Ka + [L]

where
θ is the fraction of ligand binding sites filled
[L] is the ligand concentration

K is the apparent dissociation constant derived from the law of mass action (equilibrium constant for dissociation)
d

K is the ligand concentration producing half occupation (ligand concentration occupying half of the binding sides), that is also
a

the microscopic dissociation constant


n is the Hill coefficient that describes the cooperativity

Taking the logarithm of both sides of the equation leads to an alternative formulation of the Hill Equation.
θ
log( ) = n log[L] − log Kd (10.6.2)
1 −θ

when n < 1 , there is negative cooperativity


when n = 1 , there is no cooperativity
when n > 1 , there is positive cooperativity

Allosteric Models
Currently, there are 2 models for illustrating cooperativity: the concerted model and the sequential model. Most allosteric effects
can be explained by the concerted MWC model put forth by Monod, Wyman, and Changeux, or by the sequential model described
by Koshland, Nemethy, and Filmer. Both postulate that enzyme subunits exist in one of two conformations, tensed (T) or relaxed
(R), and that relaxed subunits bind substrate more readily than those in the tense state. The two models differ most in their
assumptions about subunit interaction and the preexistence of both states.

10.6.2 https://chem.libretexts.org/@go/page/41366
The Concerted model
The concerted model of allostery, also referred to as the symmetry model or MWC model, postulates that enzyme subunits are
connected in such a way that a conformational change in one subunit is necessarily conferred to all other subunits. Thus, all
subunits must exist in the same conformation. The model further holds that, in the absence of any ligand (substrate or otherwise),
the equilibrium favours one of the conformational states, T or R. The equilibrium can be shifted to the R or T state through the
binding of one ligand (the allosteric effector or ligand) to a site that is different from the active site (the allosteric site).

The Sequential model


The sequential model of allosteric regulation holds that subunits are not connected in such a way that a conformational change in
one induces a similar change in the others. Thus, all enzyme subunits do not necessitate the same conformation. Moreover, the
sequential model dictates that molecules of substrate bind via an induced fit protocol. In general, when a subunit randomly collides
with a molecule of substrate, the active site, in essence, forms a glove around its substrate. While such an induced fit converts a
subunit from the tensed state to relaxed state, it does not propagate the conformational change to adjacent subunits. Instead,
substrate-binding at one subunit only slightly alters the structure of other subunits so that their binding sites are more receptive to
substrate. To summarize:
subunits need not exist in the same conformation
molecules of substrate bind via induced-fit protocol
conformational changes are not propagated to all subunits

Note: Allosteric database

Allostery is a direct and efficient means for regulation of biological macromolecule function, produced by the binding of a
ligand at an allosteric site topographically distinct from the orthosteric site. Due to the often high receptor selectivity and lower
target-based toxicity, allosteric regulation is also expected to play an increasing role in drug discovery and bioengineering. The
AlloSteric Database (ASD, provides a central resource for the display, search and analysis of the structure, function and related
annotation for allosteric molecules. Currently, ASD contains allosteric proteins from more than 100 species and modulators in
three categories (activators, inhibitors, and regulators). Each protein is annotated with detailed description of allostery,
biological process and related diseases, and each modulator with binding affinity, physicochemical properties and therapeutic
area. Integrating the information of allosteric proteins in ASD should allow the prediction of allostery for unknown proteins, to
be followed with experimental validation. In addition, modulators curated in ASD can be used to investigate potential allosteric
targets for a query compound, and can help chemists to implement structure modifications for novel allosteric drug design.

Summary
Allosteric enzymes are an exception to the Michaelis-Menten model. Because they have more than two subunits and active sites,
they do not obey the Michaelis-Menten kinetics, but instead have sigmoidal kinetics. Since allosteric enzymes are cooperative, a
sigmoidal plot of V versus [S] results: There are distinct properties of Allosteric Enzymes that makes it different compared to
0

other enzymes.
1. One is that allosteric enzymes do not follow the Michaelis-Menten Kinetics. This is because allosteric enzymes have multiple
active sites. These multiple active sites exhibit the property of cooperativity, where the binding of one active site affects the
affinity of other active sites on the enzyme. As mentioned earlier, it is these other affected active sites that result in a sigmoidal
curve for allosteric enzymes.

10.6.3 https://chem.libretexts.org/@go/page/41366
2. Allosteric Enzymes are influenced by substrate concentration. For example, at high concentrations of substrate, more enzymes
are found in the R state. The T state is favorite when there is an insufficient amount of substrate to bind to the enzyme. In other
words, the T and R state equilibrium depends on the concentration of the substrate.
3. Allosteric Enzymes are regulated by other molecules. This is seen when the molecules 2,3-BPG, pH, and CO2 modulates the
binding affinity of hemoglobin to oxygen. 2,3-BPG reduces binding affinity of O2 to hemoglobin by stabilizing the T- state.
Lowering the pH from physiological pH=7.4 to 7.2 (pH in the muscles and tissues) favors the release of O . Hemoglobin is
2

more likely to release oxygen in C O rich areas in the body.


2

There are two primary models for illustrating cooperativity. The concerted model (also called the Monod-Wyman-Changeux
model) illustrates cooperativity by assuming that proteins have two or more subunits, and that each part of the protein molecule is
able to exist in either the relaxed (R) state or the tense (T) state - the tense state of a protein molecule is favored when it doesn't
have any substrates bound. All aspects, including binding and dissociation constants are the same for each ligand at the respective
binding sites. The sequential model aims to demonstrate cooperativity by assuming that the enzyme/protein molecule affinity is
relative and changes as substrates bind. Unlike the concerted model, the sequential model accounts for different species of the
protein molecule.

10.6: Allosteric Interactions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by World Public Library.

10.6.4 https://chem.libretexts.org/@go/page/41366
10.7: The Effect of pH on Enzyme Kinetics
In the same way that every enzyme has an optimum temperature, so each enzyme also has an optimum pH at which it works best.
For example, trypsin and pepsin are both enzymes in the digestive system which break protein chains in the food into smaller bits -
either into smaller peptide chains or into individual amino acids. Pepsin works in the highly acidic conditions of the stomach. It has
an optimum pH of about 1.5. On the other hand, trypsin works in the small intestine, parts of which have a pH of around 7.5.
Trypsin's optimum pH is about 8.
Table 10.7.1 : pH for Optimum Activity
Enzyme Optimal pH Enzyme Optimal pH

Lipase (pancreas) 8.0 Invertase 4.5

Lipase (stomach) 4.0 - 5.0 Maltase 6.1 - 6.8

Lipase (castor oil) 4.7 Amylase (pancreas) 6.7 - 7.0

Pepsin 1.5 - 1.6 Amylase (malt) 4.6 - 5.2

Trypsin 7.8 - 8.7 Catalase 7.0

Urease 7.0

If you think about the structure of an enzyme molecule, and the sorts of bonds that it may form with its substrate, it isn't surprising
that pH should matter. Suppose an enzyme has an optimum pH around 7. Imagine that at a pH of around 7, a substrate attaches
itself to the enzyme via two ionic bonds. In the diagram below, the groups allowing ionic bonding are caused by the transfer of a
hydrogen ion from a -COOH group in the side chain of one amino acid residue to an -NH2 group in the side chain of another.
In this simplified example, that is equally true in both the substrate and the enzyme.

Now think about what happens at a lower pH - in other words under acidic conditions. It won't affect the -NH3+ group, but the -
COO- will pick up a hydrogen ion. What you will have will be this:

You no longer have the ability to form ionic bonds between the substrate and the enzyme. If those bonds were necessary to attach
the substrate and activate it in some way, then at this lower pH, the enzyme won't work. What if you have a pH higher than 7 - in
other words under alkaline conditions. This time, the -COO- group won't be affected, but the -NH3+ group will lose a hydrogen ion.
That leaves . . .

10.7.1 https://chem.libretexts.org/@go/page/41367
Again, there is no possibility of forming ionic bonds, and so the enzyme probably won't work this time either. At extreme pH's,
something more drastic can happen. Remember that the tertiary structure of the protein is in part held together by ionic bonds just
like those we've looked at between the enzyme and its substrate. At very high or very low pH's, these bonds within the enzyme can
be disrupted, and it can lose its shape. If it loses its shape, the active site will probably be lost completely. This is essentially the
same as denaturing the protein by heating it too much.

Kinetics
The rates of enzyme-catalysed reactions vary with pH and often pass through a maximum as the pH is varied. If the enzyme obeys
Michaelis-Menten kinetics the kinetic parameters k0 and kA often behave similarly. The pH at which the rate or a suitable
parameter is a maximum is called the pH optimum and the plot of rate or parameter against pH is called a pH profile. Neither the
pH optimum nor the pH profile of an enzyme has any absolute significance and both may vary according to which parameter is
plotted and according to the conditions of the measurements.
If the pH is changed and then brought back to its original value, the behavior is said to be reversible if the original properties of the
enzyme are restored; otherwise it is irreversible. Reversible pH behavior may occur over a narrow range of pH, but effects of large
changes in pH are in most cases irreversible. The diminution in rate as the pH is taken to the acid side of the optimum can be
regarded as inhibition by hydrogen ions. The diminution in rate on the alkaline side can be regarded as inhibition by hydroxide
ions. The equations describing pH effects are therefore analogous to inhibition equations. For single-substrate reactions the pH
behavior of the parameters k0 and kA can sometimes be represented by an equation of the form
kopt
k = (10.7.1)
+
[H ] K2
1+ +
+
K1 [H ]

in which k represents k0 or kA, and k is the value of the same parameter that would be observed if the enzyme existed entirely in
opt

the optimal state of protonation; it may be called the pH-independent value of the parameter. The constants K1 and K2 can
sometimes be identified as acid dissociation constants for the enzyme. substrates or other species in the reaction mixture. The
identification is, however, never straight forward and has to be justified by independent evidence. The behavior is frequently much
more complicated than represented by Equation 10.7.1.
It is not accidental that this section has referred exclusively to pH dependences of k0 and kA. The pH dependence of the initial rate
or, worse, the extent of reaction after a given time is rarely meaningful; the pH dependence of the Michaelis constant is often too
complex to be readily interpretable.

The pH dependence of the Michaelis constant is often too complex to be readily


interpretable.

Quenching Enzyme Activity


In principle, we can use any of the range of possible analytical techniques to follow a reaction’s kinetics provided that the reaction
does not proceed to any appreciable extent during the time it takes to make a measurement. As you might expect, this requirement
places a serious limitation on kinetic methods of analysis. If the reaction’s kinetics are slow relative to the analysis time, then we
can make our measurements without the analyte undergoing a significant change in concentration. When the reaction’s rate is too
fast—which often is the case—then we introduce a significant error if our analysis time is too long.
One solution to this problem is to stop, or quench the reaction by adjusting experimental conditions. For example, many reactions
show a strong pH dependency, and may be quenched by adding a strong acid or a strong base. Figure 13.7 shows a typical example
for the enzymatic analysis of p-nitrophenylphosphate using the enzyme wheat germ acid phosphatase to hydrolyze the analyte to p-
nitrophenol.

10.7.2 https://chem.libretexts.org/@go/page/41367
The reaction has a maximum rate at a pH of 5. Increasing the pH by adding NaOH quenches the reaction and converts the colorless
p-nitrophenol to the yellow-colored p-nitrophenolate, which absorbs at 405 nm.

Figure 13.7: Initial rate for the enzymatic hydrolysis of p-nitrophenylphosphate using wheat germ acid phosphatase. Increasing the
pH quenches the reaction and coverts colorless p-nitrophenol to the yellow-colored p-nitrophenolate, which absorbs at 405 nm. The
data are adapted from socrates.hunter.cuny.edu.

This page titled 10.7: The Effect of pH on Enzyme Kinetics is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by Jim Clark & David Harvey via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit
history is available upon request.
13.2: Chemical Kinetics by David Harvey is licensed CC BY-NC-SA 4.0.

10.7.3 https://chem.libretexts.org/@go/page/41367
10.8: The Effect of Temperature on Enzyme Kinetics
Enzymes are generally globular proteins, acting alone or in larger complexes. Like all proteins, enzymes are linear chains of amino
acids that fold to produce a three-dimensional structure. The sequence of the amino acids specifies the structure which in turn
determines the catalytic activity of the enzyme. Although structure determines function, a novel enzyme's activity cannot yet be
predicted from its structure alone. Enzyme structures unfold (denature) when heated or exposed to chemical denaturants and this
disruption to the structure typically causes a loss of activity.
Protein folding is key to whether a globular protein or a membrane protein can do its job correctly. It must be folded into the right
shape to function. But hydrogen bonds, which play a big part in folding, are rather weak, and it does not take much heat, acidity, or
other stress to break some and form others, denaturing the protein. This is one reason why tight homeostasis is physiologically
necessary in many life forms.

Denaturation
Denaturation is a process in which proteins or nucleic acids lose the quaternary structure, tertiary structure and secondary structure
which is present in their native state, by application of some external stress or compound such as a strong acid or base, a
concentrated inorganic salt, an organic solvent (e.g., alcohol or chloroform), radiation or heat. If proteins in a living cell are
denatured, this results in disruption of cell activity and possibly cell death. Denatured proteins can exhibit a wide range of
characteristics, from conformational change and loss of solubility to aggregation due to the exposure of hydrophobic groups.

Figure 10.8.1 : Enzyme activity initially increases with temperature until the enzyme's structure unfolds (denaturation), leading to
an optimal rate of reaction at an intermediate temperature. (CC-BY-SA-4.0; Thomas Shafee)
Enzyme denaturation is normally linked to temperatures above a species' normal level; as a result, enzymes from bacteria living in
volcanic environments such as hot springs are prized by industrial users for their ability to function at high temperatures, allowing
enzyme-catalyzed reactions to be operated at a very high rate.

10.8.1 https://chem.libretexts.org/@go/page/52395
Contributors and Attributions
Wikipedia

10.8: The Effect of Temperature on Enzyme Kinetics is shared under a CC BY-SA license and was authored, remixed, and/or curated by
LibreTexts.

10.8.2 https://chem.libretexts.org/@go/page/52395
10.E: Exercises
10.1: General Principles of Catalysis
Q10.1a
One day in class about enzyme kinetics, Jack comes over to you and asks, "I know enzymes are biological catalysts, but I do not
understand how it works. Can you explain how enzymes make reactions go faster? And is it only faster in one direction?"

S10.1a
Because the activation energy is the energy hill between reactants and products, enzymes decreasing the size of the hill also
decreases the amount of energy needed for reactions to go in either direction. A smaller energy hill allows reactants and products to
overcome the barrier quicker, resulting a faster reaction rate.

Q10.1b
If your student colleagues argues that a catalysts affects only the rate of only one direction of a reaction. Explain why he is correct
or not.

S10.1b
False. Catalysts affect rate by providing an alternative mechanism which has a lower transition state energy. It's impossible to lower
transition state energy for only one direction of a reaction. (It'd be like making a hill shorter from the north, but keeping it the same
height from the south.)

Q10.1c
How does the enzyme catalysis affect both forward and reverse reaction?

S10.1c
The enzyme catalyst lowers the Gibb energy of transition state, which reduces the activation energy of both reactions. Therefore, it
makes reactions occur faster.

Q10.2a
Given enzyme-catalyzed reaction k1 = 4x106 M-1 s-1 , k-1 =6x104 s-1 and k2= 2.0x103 s-1. Determine if the enzyme –substrate
binding follow the equilibrium or not ?

S10.2a
The dissociation constant ks = k-1 / k1 =6x104 s-1/ 4x106 M-1 s-1 = 0.15 M
The Michaelis constant kM = k-1 +k2 / k1= 6x104 s-1+ 2.0x103 s-1/ 4x106 M-1 s-1 =0.155M
The two constant are not equal. Therefore, the binding does not follow the equilibrium scheme.

Q10.2b
Is it appropriate to use the rapid equilibrium scheme to model the kinetics of a catalyzed reaction with the following rate constants?
7 −1 −1
k1 = 7 × 10 M s (10.E.1)

5 −1
k−1 = 8 × 10 s (10.E.2)

4 −1
k2 = 5 × 10 s (10.E.3)

S10.2b
Assuming rapid equilibrium is appropriate when k ≫ k , so that KM = KS. In this case, the values of k-1 and k2 are not
−1 2

dramatically different, so we can calculate KS and KM and compare.


1. Calculate K :
M

k−1 + k2
KM = (10.E.4)
k1

10.E.1 https://chem.libretexts.org/@go/page/41336
5 −1 4 −1
8 × 10 s + 5 × 10 s
KM = (10.E.5)
7 −1 −1
7 × 10 M s

KM = 0.01 M (10.E.6)

2. Calculate K :
S

k−1
KS = (10.E.7)
k1

5 −1
8 × 10 s
KM = (10.E.8)
7 −1 −1
7 × 10 M s

KM = 0.01 M (10.E.9)

Since K M = KS , it is appropriate to assume rapid equilibrium.

Q10.2c
Given k1= 7 x 10^6 M-1s-1, k-1=6 x 10^4 s-1, k3= 2 x 10^3 s-1, determine if the enzyme substance binding follow the equilibrium
or steady state scheme?
k−1
Ks = (10.E.10)
k1
4 −1
6 × 10 s
= (10.E.11)
6 −1
7 × 10 M s−1
−3
= 8.8 × 10 M (10.E.12)

KM = k−1 + k2 (10.E.13)
4 −1 3 −1
6 × 10 s + 2 × 10 s
= (10.E.14)
6 −1 −1
7 × 10 M s
−3
= 8.9 × 10 M (10.E.15)

Q10.3a
The substrate N-acetylglycine ethyl ester can be catalyzed by the enzyme carbonic anhydrase. This enzyme has a turnover rate of
30,000 s-1. Determine how long it will take carbonic anhydrase to cleave the substrate.

S10.3a
We already know the turnover number (k cat ). The amount of time necessary to cleave the substrate is the reciprocal of the turnover
rate.
1 1 −5
t = = = 3.33 × 10 (10.E.16)
−1
k 30, 000 s

−6 −5
(5 minutes)(7.1 × 10 ) = 3.5 × 10 (10.E.17)

Q10.3b
RuBisCO is an enzyme in the Calvin cycle that fixes atmospheric carbon and has a turnover rate of 3.3 s-1. How long does it take
RuBisCO to fix one molecule of carbon dioxide?

10.E.2 https://chem.libretexts.org/@go/page/41336
Spacefilling structure of RuBisCO created using Rasmol and the 8RUC file from the Protein Data Bank.

S10.3b
The turnover number is the number of molecules of substrate per unit time (when the enzyme is fully saturated). So simply take the
reciprocal to find the time per molecule of substrate.
1
= 0.30 s (10.E.18)
−1
3.3 s

Note: RuBisCO is a notoriously slow enzyme.

Q10.3c
The catalyze of acetylcholine has a rate 50000 s-1 . Calculate the time for the enzyme to cleave one Ach molecule.

S10.3c
t= 1/k2 =1/ 50000 = 2.0 x10-5s

Q10.3d
Carbonic anhydrase, an enzyme that catalyzes the dehydration of carbonic acid to form carbonic acid, has the turnover rate of kcat
4.0x 105 s-1. Calculate how long does it take does it take for the enzyme to cleave one molecule carbonic acid?

S10.3d
The time required for the enzyme to cleave one molecule carbonic acid:
t=1/kcat = 1/ 4x105s-1 =2.5x10-6s = 2.5µs

Q10.3e
p-nitrophenyl acetate(PNPA) is catalyzed by chymotrypsin to yield p-nitrophenolate ion and acetate ion. The turnover rate of that
enzyme is 40,000 s-1. How long will it take for the enzyme to produce 1 mole of Nitrophenyl acetate?

S10.3e
40, 000
It takes 1 second to convert all 40,000 molecule substrate into the product, so: t = 5
= 3.8 sec.
9.5x10

10.2: The Equations of Enzyme Kinetics


Q10.4
What is plotted on the x and y axes on a Lineweaver-Burk plot? Show how to derive the equation for the plot from the equation
Vmax [S]
v0 = (10.E.19)
KM + [S]

and explain how Vmax and KM can be found from the graph's intercepts. Hint: A Lineweaver-Burk plot is also sometimes called a
double reciprocal plot.

10.E.3 https://chem.libretexts.org/@go/page/41336
S10.4
The x-axis is 1/ν0, and the x-axis is 1/Vmax.
1. Take the reciprocal of both sides of the equation.
1 KM + [S]
= (10.E.20)
v0 Vmax [S]

1 KM 1 1
= + (10.E.21)
v0 Vmax [S] Vmax

2. Set 1/[S] = 0 to find the y-intercept, and show that it relates to Vmax.
1
Y int = (10.E.22)
Vmax

1
Vmax = (10.E.23)
Y int

3. Set 1/ν0 = 0 to find the x-intercept, and show that it relates to KM.
KM 1
0 = Xint + (10.E.24)
Vmax Vmax

KM 1
Xint = − (10.E.25)
Vmax Vmax

1
Xint = − (10.E.26)
KM

1
KM = − (10.E.27)
Xint

Q10.4a
Derive the Michaelis-Menten equation by assuming rapid equilibrium.
K1 K2
E + S ⇌ ES → E + P (10.E.28)

Vmax [S]
v0 = (10.E.29)
KM + [S]

S10.4a
[E][S] [ES] [E][S] [ES] k−1 [E][S]
= K−1 [ES] = K1 [E][S] = K−1 [ES] = K1 [E][S] KM =
dt dt dt dt K1 [ES]

given that [E ] 0 = [E] + [ES]

[E ]0 − [ES][S] [E ]0 [S] − [ES][S]


KM = KM = KM × [ES] = [E ]0 [S] − [ES][S] KM × [ES] + [ES][S] = [E ]0 [S]
[ES] [ES]

[ E0 ][S]
(KM + [S])[ES] = [ E0 ][S] [ES] =
KM + [S]

d[P ]
since = k2 [ES]
dt

K2 [ E0 ][S]
v0 =
KM + [S]

since v max = k2 [E ]0

Vmax [S]
v0 =
KM + [S]

10.E.4 https://chem.libretexts.org/@go/page/41336
Q10.4b
We know that the Michaelis Menten derivation for the following reaction:

E + S ⇌ ES ⟶ E + P (10.E.30)

However, what if the reaction took place in a different scenario whereby:

E + S ⇌ ES ⟶ ES ⟶ E+P (10.E.31)
1 2

What would be the corresponding Michaelis-Menten Ebe quation now?

S10.4b
This is the an outline for determining an expression for the rate of substrate conversion in the given case:
1. Set up the reaction with rate constants, assuming k−2 ≈ k−3 ≈ 0 :
k1 k2 k3

E+S ⇌ ES −
→ ES −
→ E+P (10.E.32)
1 2
k−1

2. Set up the differential equations describing the reaction, i.e. the rate of change for each component with time. The rate of
substrate change, for example, will be
d[S]
= −k1 [E][S] + k−1 [ ES ] (10.E.33)
1
dt

3. Choose initial conditions and set up two equations for conservation of mass. For example, the initial concentration of enzyme
must equal the sum of the concentrations of E, ES1 and ES2.
4. Make the steady-state assumpt: assume that the concentrations of the intermediate complexes do not change on the time-scale
of product formation, i.e.
d[ ES ] d[ ES ]
1 2
≈ ≈0 (10.E.34)
dt dt

5. Solve for −r , the negative rate of substrate conversion, obtaining the Michaelis-Menten expression describing the kinetics of
S

the given situtation.

Q10.4c
Prove that K equals the concentration S when the initial rate is half its maximum value.
s

S10.4c

We have:

Divided both said by 1

When the initial rate is half its maximum value:

10.E.5 https://chem.libretexts.org/@go/page/41336
Q10.5a
An enzyme that has a K value of 4.6 × 10 M is studied at an initial substrate concentration of 0.041 M. After a minute, it is
m
−5

found that 7.3 uM of product has been produced. Calculate the value of Vmax and the amount of product formed after 4.5 minutes.

Q10.5b
An solution initially contains a catalytic amount of an enzyme with KM = 1.5 mM, 0.25 M of substrate, and no product. After 45
seconds, the solution contains 25 µM of product. Find Vmax and the concentration of product after 2.0 minutes.
Hint: [S] >> KM

S10.5b
1. Find the initial velocity.
25 μM
v0 = (10.E.35)
0.75 min

v0 = 33.3 μM /min (10.E.36)

2. Use v0, [S], and KM to solve for Vmax.


Vmax [S]
v0 = (10.E.37)
KM + [S]

KM
Vmax = v0 ( + 1) (10.E.38)
[S]

1.5 mM M
Vmax = 33.3 μM /min ( × + 1) (10.E.39)
0.25 M 1000 mM

Vmax = 33.3 μM /min (10.E.40)

Notice that ν0 = Vmax. Since [S] >>KM, the reaction will continue with a velocity of Vmax for the remainder of the two minutes.
3. Predict [P] after 2.0 minutes at the rate Vmax.
[P ] = Vmax × t (10.E.41)

[P ] = 33.3 μM /min × 2.0 min (10.E.42)

[P ] = 66.6 μM (10.E.43)

Q10.5c
A particular enzyme at a research facility is being studied by a group of graduate students. This enzyme has a Km value of 5.0 X
10-6 M. The students study this enzyme with an initial substrate concentration of 0.055 M. At one minute, 7 µM of product was
made. What is the amount of product produced after 5 minutes. What is the Vmax ?

S10.5c
7.0 X 10-6 M = Vmax (0.055 M) / (5.0 X 10-6 M + .055 M) Vmax = 7.1 X 10-6 M/min At 5
minutes the amount of product formed is:

Q10.5d
Calculate the value if an enzyme has , value and

S10.5d

We have:

10.E.6 https://chem.libretexts.org/@go/page/41336
Q10.6a
Given the values, [S]/10^-4 M 3.0 4.6 10.5 16.5
vo/10^-6 M * min^-1. 2.64 3.5 6.2 7.8
construct a Lineweaver-Burk plot, and assuming Michaelis-Menten kinetics, calculate the values of Vmax, Km, and k2 using the
constructed plot.

Q10.6b
The data below represents the data recorded after the hydrolysis of a substrate by an enzyme.

[S]/10-4 M 2.1 4.2 9.3 14.2

vo / 10-6 M · min-1 1.2 3.1 6.3 9.1

Calculate Vmax, Km and K2 using a Lineweaver-Burk plot. Assume Michaelis. Given [E}0 is 5.0 X 10-6 M

S10.6b
(1/[S])/10-3 M-1 (1/vo )/ 10-5 M-1 · min

4.8 8.3

2.4 3.2

1.1 1.6

0.7 1.1

Figure: Lineweaver-Burk Plot The linear equation for the


graph above is: y=176x+4.23X104 To solve for Vmax use: Intercept= 1/Vmax Vmax = 1/(.4.23X104) = 2.36 X 10-5 To solve for Km
use Slope= Km/Vmax (176)( 2.36 X 10-5) = 4.2 X 10-5 To solve for K2 use K2 = Vmax/ [E}0 = (2.36 X 10-5)(5.0 X 10-6 M)= 4.72
min-1

Q10.6c
Using the table below, calculate the KM, Vmax, and slope.

1 sec 1
−3 2 −1
(10 ) (10.E.44) (10 M ) (10.E.45)
V0 −1 [S]
M

2.9 0.3

3.2 1.2

4.4 1.8

10.E.7 https://chem.libretexts.org/@go/page/41336
1 sec 1
−3 2 −1
(10 ) (10.E.44) (10 M ) (10.E.45)
V0 −1 [S]
M

5.1 3.3

S10.6c
−3 −3
5.1 × 10 − 2.9 × 10 1 1
slope= 2 2
= 6.73 × 10
−6
6.73 × 10
−6
× 0.3 × 10
2
+ = 2.9 × 10
−3
= 0.00270
3.3 × 10 − 0.3 × 10 Vmax Vmax

−1 −1
−1 −6
Vmax = 370.62M s 6.73 × 10 × + 0.00270 = 0 = −0.00943 KM = 106.044
KM KM

Q10.6d
Given the value = 0.00032 and . Find the ratio between and

S10.6d

Divided both side by 1:

Divided both side by 1 again:

Q10.7a
From this graph determine the KM and Vmax?

S10.7b
From his graph we can see that the value KM is 2. Then we look to see where KM is half. At that point, we see that KM/2 is 1 and
the x-value for that coordinate is 1. This means Vmax is 1.

Q10.7b
Neutral sphingomyelinase 2 converts sphingomyelin into ceramide and phosphocholine. Assume its Vmax is 35 µM min-1. When
you provide 3 x 10-5 M of sphingomyelin, you observe an initial velocity of 6.0 µM min-1. Calculate the KM for this reaction,
rounding to 3 significant figures.

10.E.8 https://chem.libretexts.org/@go/page/41336
S10.7b
1 KM 1
= + (10.E.46)
Vo Vmax [S] Vmax

1 KM 1
= + (10.E.47)
6.0 (35)(30) 35

1 1 KM
− = (10.E.48)
6 35 1050

KM = 145 µM

Q10.8a
We are given a second order equation: r=k[A][B]. The concentration of A is 0.05g and the concentration of B is 2.5g. We decide
this difference is great enough to treat this as a pseudo first reaction? What concentration is held constant and why? Write the new
equation.

S10.8a
We hold B constant because the concentration is so much larger, so it should be close to constant for the reaction. The new equation
would be r=k'[a]

Q10.9a
Name and briefly describe two types of reactions that do not follow Michaelis-Menten kinetics.

S10.9a
Irreversible inhibition - the inhibitor binds covalently and irreversibly to the enzyme Allosteric interactions - the binding of
effectors at allosteric sites (away from the active site) influence substrate binding.

Q10.9b
For the reaction mechanism below, how does the concentration of C affect the concentration of B ?
A ⇋ B → C (10.E.49)

S10.9b
Because the second state is irreversible (i.e., since arrow), it does not matter if you have a large or little concentration of C , it
would not affect B and hence the kinetics of the reaction. However, the concentration of
A\0wouldclearlyaf f ecttheconcentrationof \(B.

10.3: Chymotrypsin: A Case Study


Q1
1. Speculate on how the catalytic rate constant can be determined from the spectrophotogram.
2. How can product be consistently produced if the rate of change of the ES complex is 0?
3. How would the rate of product formation change if:
a. the substrate concentration were doubled?
b. the enzyme concentration were doubled?
c. The reaction was carried out in mono-deuterated water instead of H2O (comment qualitatively)?
4. Explain the role of hydrogen bonding in protein hydrolysis catalyzed by chymotrypsin.
5. What would the spectrophotogram look like if the reaction proceeded via a steady-state mechanism instead of pre-equilibrium.

S2
1. The catalytic rate constant can be deduced from the graph by simply determining the slope of the line where the reaction
demonstrates 0-order kinetics (the linear part).
2. This is pre-equilibrium kinetics in action. The ES complex is formed from E and S at a faster rate than any other step in the
reaction. As soon as ES is converted to *ES, another mole of ES is produced from an infinite supply of E + S. This means that
the amount of ES and E + S is constantly at equilibrium, and thus the change of either with respect to time is 0.

10.E.9 https://chem.libretexts.org/@go/page/41336
a. No change.
b. Two-fold increase.
c. Because water is involved in the final, slowest step of the mechanism, deuterating the water would decrease the rate of the
overall reaction from 5- to 30-fold.
3. Initially, hydrogen bonding between the enzymes histidine and serine side chains weakens the bond of serine’s O-H. This allows
a facilitated nucleophilic attack of the hydroxyl oxygen on the substrates carbonyl group. Conversely, in the final step of the
reaction, the bound serine oxygen forms a hydrogen bond with a protonated histidine, which allows for easier cleavage from the
substrate.
4. The graph would show similar 0-order kinetics, but the line would intercept the Y-axis at an absorbance of 0 instead of the 1:1
mole ratio of nitrophenolate to enzyme.

10.4: Multisubstrate Systems

10.5: Enzyme Inhibition

10.6: Allosteric Interactions

10.7: The Effect of pH on Enzyme Kinetics

10.4: Multisubstrate Systems

10.5: Enzyme Inhibition


10.6: Allosteric Interactions

10.7: The Effect of pH on Enzyme Kinetics

10.E: Exercises is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

10.E.10 https://chem.libretexts.org/@go/page/41336
CHAPTER OVERVIEW

11: Quantum Mechanics and Atomic Structure


Quantum mechanics is a fundamental branch of physics concerned with processes involving small particles (e.g, atoms and
photons). Such processes as said to be quantized with properties that are observed in only in integer multiples of the Planck
constant. This is utterly inexplicable in classical physics which is a continuum approach.
11.1: The Wave Theory of Light
11.2: Planck's Quantum Theory
11.3: The Photoelectric Effect
11.4: Bohr's Theory of the Hydrogen Emission Spectrum
11.5: de Broglie's Postulate
11.6: The Heisenberg Uncertainty Principle
11.7: The Schrödinger Wave Equation
11.8: Particle in a One-Dimensional Box
11.9: Quantum-Mechanical Tunneling
11.10: The Schrödinger Wave Equation for the Hydrogen Atom
11.11: Many-Electron Atoms and the Periodic Table
11.E: Quantum Mechanics and Atomic Structure (Exercises)

11: Quantum Mechanics and Atomic Structure is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.

1
11.1: The Wave Theory of Light
Learning Objectives
to learn about the characteristics of electromagnetic waves. Light, X-Rays, infrared and microwaves among the types of
electromagnetic waves.

Scientists discovered much of what we know about the structure of the atom by observing the interaction of atoms with various
forms of radiant, or transmitted, energy, such as the energy associated with the visible light we detect with our eyes, the infrared
radiation we feel as heat, the ultraviolet light that causes sunburn, and the x-rays that produce images of our teeth or bones. All
these forms of radiant energy should be familiar to you. We begin our discussion of the development of our current atomic model
by describing the properties of waves and the various forms of electromagnetic radiation.

Figure 1.1.1 A Wave in Water When a drop of water falls onto a smooth water surface, it generates a set of waves that travel
outward in a circular direction.

Properties of Waves
A wave is a periodic oscillation that transmits energy through space. Anyone who has visited a beach or dropped a stone into a
puddle has observed waves traveling through water (Figure 1.1.1). These waves are produced when wind, a stone, or some other
disturbance, such as a passing boat, transfers energy to the water, causing the surface to oscillate up and down as the energy travels
outward from its point of origin. As a wave passes a particular point on the surface of the water, anything floating there moves up
and down.

Figure 1.1.2: Important Properties of Waves (a) Wavelength (λ in meters), frequency (ν, in Hz), and amplitude are indicated on this
drawing of a wave. (b) The wave with the shortest wavelength has the greatest number of wavelengths per unit time (i.e., the
highest frequency). If two waves have the same frequency and speed, the one with the greater amplitude has the higher energy.
Waves have characteristic properties (Figure 1.1.2). As you may have noticed in Figure 1.1.1, waves are periodic, that is, they
repeat regularly in both space and time. The distance between two corresponding points in a wave—between the midpoints of two
peaks, for example, or two troughs—is the wavelength (λ), distance between two corresponding points in a wave—between the
midpoints of two peaks or two troughs. λ is the lowercase Greek lambda, and ν is the lowercase Greek nu. Wavelengths are
described by a unit of distance, typically meters. The frequency (ν), the number of oscillations (i.e., of a wave) that pass a particular
point in a given period of time. of a wave is the number of oscillations that pass a particular point in a given period of time. The

11.1.1 https://chem.libretexts.org/@go/page/41368
usual units are oscillations persecond (1/s = s−1), which in the SI system is called the hertz (Hz). It is named after German physicist
Heinrich Hertz (1857–1894),a pioneer in the field of electromagnetic radiation. The amplitude (the vertical height of a wave, which
is defined as half the peak-to-trough height), or vertical height, of a wave is defined as half the peak-to-trough height; as the
amplitude of a wave with a given frequency increases, so does its energy. As you can see in Figure 1.1.2 , two waves can have the
same amplitude but different wavelengths and vice versa. The distance traveled by a wave per unit time is its speed (v), the distance
traveled by a wave per unit time, which is typically measured in meters per second (m/s). The speed of a wave is equal to the
product of its wavelength and frequency:
(wavelength)(frequency) = speed

λν = v (1.1.1a)

meters wave meters


( )( ) = (1.1.1b)
wave second second

Be careful not to confuse the symbols for the speed, v , with the frequency, ν . Water waves are slow compared to sound waves,
which can travel through solids, liquids, and gases. Whereas water waves may travel a few meters per second, the speed of sound in
dry air at 20°C is 343.5 m/s. Ultrasonic waves, which travel at an even higher speed (>1500 m/s) and have a greater frequency, are
used in such diverse applications as locating underwater objects and the medical imaging of internal organs.

Electromagnetic Radiation
Water waves transmit energy through space by the periodic oscillation of matter (the water). In contrast, energy that is transmitted,
or radiated, through space in the form of periodic oscillations of electric and magnetic fields is known as electromagnetic radiation,
which is energy that is transmitted, or radiated, through space in the form of periodic oscillations of electric and magnetic fields.
(Figure 1.1.3). Some forms of electromagnetic radiation are shown in Figure 1.1.4. In a vacuum, all forms of electromagnetic
radiation—whether microwaves, visible light, or gamma rays—travel at the speed of light (c), which is the speed with which all
forms of electromagnetic radiation travel in a vacuum, a fundamental physical constant with a value of 2.99792458 × 108 m/s
(which is about 3.00 ×108 m/s or 1.86 × 105 mi/s). This is about a million times faster than the speed of sound.

Figure 1.1.3: The Nature of Electromagnetic Radiation. All forms of electromagnetic radiation consist of perpendicular oscillating
electric and magnetic fields.
Because the various kinds of electromagnetic radiation all have the same speed (c), they differ in only wavelength and frequency.
As shown in Figure 1.1.4 and Table 1.1.1 , the wavelengths of familiar electromagnetic radiation range from 101 m for radio waves
to 10−12 m for gamma rays, which are emitted by nuclear reactions. By replacing v with c in Equation 6.1.1, we can show that the
frequency of electromagnetic radiation is inversely proportional to its wavelength:
c = λν

c (1.1.2)
ν =
λ

For example, the frequency of radio waves is about 10 Hz, whereas the frequency of gamma rays is about 1020 Hz. Visible light,
8

which is electromagnetic radiation that can be detected by the human eye, has wavelengths between about 7 × 10−7 m (700 nm, or
4.3 × 1014 Hz) and 4 × 10−7 m (400 nm, or 7.5 × 1014 Hz). Note that when frequency increases, wavelength decreases; c being a
constant stays the same. Similarly when frequency decreases, the wavelength increases.

11.1.2 https://chem.libretexts.org/@go/page/41368
Figure 1.1.4: The Electromagnetic Spectrum. (a) This diagram shows the wavelength and frequency ranges of electromagnetic
radiation. The visible portion of the electromagnetic spectrum is the narrow region with wavelengths between about 400 and 700
nm. (b) When white light is passed through a prism, it is split into light of different wavelengths, whose colors correspond to the
visible spectrum.
Within this visible range our eyes perceive radiation of different wavelengths (or frequencies) as light of different colors, ranging
from red to violet in order of decreasing wavelength. The components of white light—a mixture of all the frequencies of visible
light—can be separated by a prism, as shown in part (b) in Figure 1.1.4. A similar phenomenon creates a rainbow, where water
droplets suspended in the air act as tiny prisms.
Table 1.1.1: Common Wavelength Units for Electromagnetic Radiation
Unit Symbol Wavelength (m) Type of Radiation

picometer pm 10−12 gamma ray

angstrom Å 10−10 x-ray

nanometer nm 10−9 x-ray

micrometer μm 10−6 infrared

millimeter mm 10−3 infrared

centimeter cm 10−2 microwave

meter m 100 radio

As you will soon see, the energy of electromagnetic radiation is directly proportional to its frequency and inversely proportional to
its wavelength:
E ∝ ν (1.1.3)

1
E ∝ (1.1.4)
λ

Whereas visible light is essentially harmless to our skin, ultraviolet light, with wavelengths of ≤ 400 nm, has enough energy to
cause severe damage to our skin in the form of sunburn. Because the ozone layer absorbs sunlight with wavelengths less than 350
nm, it protects us from the damaging effects of highly energetic ultraviolet radiation.

The energy of electromagnetic radiation increases with increasing frequency and decreasing wavelength.

Example 1.1.1

Your favorite FM radio station, WXYZ, broadcasts at a frequency of 101.1 MHz. What is the wavelength of this radiation?
Given: frequency
Asked for: wavelength
Strategy:
Substitute the value for the speed of light in meters per second into Equation 1.1.2 to calculate the wavelength in meters.
Solution:

11.1.3 https://chem.libretexts.org/@go/page/41368
From Equation 1.1.2 , we know that the product of the wavelength and the frequency is the speed of the wave, which for
electromagnetic radiation is 2.998 × 108 m/s:
8
λν = c = 2.998 × 10 m/s

Thus the wavelength λ is given by


8
c 2.988 × 10 m/ s 1 M Hz
λ = =( )( ) = 2.965 m
ν 101.1 M Hz
6 −1
10 s

Exercise 1.1.1
As the police officer was writing up your speeding ticket, she mentioned that she was using a state-of-the-art radar gun
operating at 35.5 GHz. What is the wavelength of the radiation emitted by the radar gun?

Click to Check Answer


8.45 mm

External Videos & Examples


Light Speed, Wavelength and Frequency by Doc Schuster - Explains wavelength and frequency
Waves in General by JaHu Productions - a bit faster. Also discusses sound waves
Calculating Frequency Given Wavelength - Johnny Cantrell
Wavelength Frequency Problems - Cayer Chem
Quantum Chemistry - Ohio State
Quantum Chemistry Quizzes - mhe education
AP Chemistry Chapter 7 Review - Science Geek
Quantum Theory of the Atom Practice Quiz - Northrup
Answers for these quizzes are included. There are also questions covering more topics in Chapter 6.

Summary
Understanding the electronic structure of atoms requires an understanding of the properties of waves and electromagnetic
radiation.
A basic knowledge of the electronic structure of atoms requires an understanding of the properties of waves and electromagnetic
radiation. A wave is a periodic oscillation by which energy is transmitted through space. All waves are periodic, repeating regularly
in both space and time. Waves are characterized by several interrelated properties: wavelength (λ), the distance between successive
waves; frequency (ν), the number of waves that pass a fixed point per unit time; speed (v), the rate at which the wave propagates
through space; and amplitude, the magnitude of the oscillation about the mean position. The speed of a wave is equal to the product
of its wavelength and frequency. Electromagnetic radiation consists of two perpendicular waves, one electric and one magnetic,
propagating at the speed of light (c). Electromagnetic radiation is radiant energy that includes radio waves, microwaves, visible
light, x-rays, and gamma rays, which differ in their frequencies and wavelengths.

Contributors
Modified by Joshua Halpern (Howard University)

11.1: The Wave Theory of Light is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
1.1: The Wave Nature of Light is licensed CC BY-NC-SA 4.0.

11.1.4 https://chem.libretexts.org/@go/page/41368
11.2: Planck's Quantum Theory
 Learning Objectives
To understand how energy is quantized in blackbody radiation

By the late 19th century, many physicists thought their discipline was well on the way to explaining most natural phenomena. They
could calculate the motions of material objects using Newton’s laws of classical mechanics, and they could describe the properties
of radiant energy using mathematical relationships known as Maxwell’s equations, developed in 1873 by James Clerk Maxwell, a
Scottish physicist. The universe appeared to be a simple and orderly place, containing matter, which consisted of particles that had
mass and whose location and motion could be accurately described, and electromagnetic radiation, which was viewed as having no
mass and whose exact position in space could not be fixed. Thus matter and energy were considered distinct and unrelated
phenomena. Soon, however, scientists began to look more closely at a few inconvenient phenomena that could not be explained by
the theories available at the time.
One experimental phenomenon that could not be adequately explained by classical physics was blackbody radiation (Figure 1.2.1 ).
Attempts to explain or calculate this spectral distribution from classical theory were complete failures. A theory developed by
Rayleigh and Jeans predicted that the intensity should go to infinity at short wavelengths. Since the intensity actually drops to zero
at short wavelengths, the Rayleigh-Jeans result was called the ultraviolet catastrophe (Figure 1.2.1 dashed line). There was no
agreement between theory and experiment in the ultraviolet region of the blackbody spectrum.

Figure 1.2.1 : Relationship between the temperature of an object and the spectrum of blackbody radiation it emits. At relatively low
temperatures, most radiation is emitted at wavelengths longer than 700 nm, which is in the infrared portion of the spectrum. As the
temperature of the object increases, the maximum intensity shifts to shorter wavelengths, successively resulting in orange, yellow,
and finally white light. At high temperatures, all wavelengths of visible light are emitted with approximately equal intensities. The
white light spectrum shown for an object at 6000 K closely approximates the spectrum of light emitted by the sun. Note the sharp
decrease in the intensity of radiation emitted at wavelengths below 400 nm, which constituted the ultraviolet catastrophe. The
classical prediction fails to fit the experimental curves entirely and does not have a maximum intensity. (CC BY-SA-NC;
anonymous).

Quantizing Electrons in the Radiator


In 1900, the German physicist Max Planck (1858–1947) explained the ultraviolet catastrophe by proposing that the energy of
electromagnetic waves is quantized rather than continuous. This means that for each temperature, there is a maximum intensity of
radiation that is emitted in a blackbody object, corresponding to the peaks in Figure 1.2.1 , so the intensity does not follow a
smooth curve as the temperature increases, as predicted by classical physics. Thus energy could be gained or lost only in integral
multiples of some smallest unit of energy, a quantum (the smallest possible unit of energy). Energy can be gained or lost only in
integral multiples of a quantum.

11.2.1 https://chem.libretexts.org/@go/page/41369
 Quantization
Although quantization may seem to be an unfamiliar concept, we encounter it frequently in quantum mechanics (hence the
name). For example, US money is integral multiples of pennies. Similarly, musical instruments like a piano or a trumpet can
produce only certain musical notes, such as C or F sharp. Because these instruments cannot produce a continuous range of
frequencies, their frequencies are quantized. It is also similar to going up and down a hill using discrete stair steps rather than
being able to move up and down a continuous slope. Your potential energy takes on discrete values as you move from step to
step. Even electrical charge is quantized: an ion may have a charge of −1 or −2, but not −1.33 electron charges.

A continuous vs. a quantized (gravitationaly) potential energy system. In the continuous case (left) a system can have any
potential energy, but in the quantized case (right), a system can only have certain values (other values are not allowed). (CC
BY-NC; Ümit Kaya via LibreTexts)

Planck's quantization of energy is described by the his famous equation:

E = hν (11.2.1)

where the proportionality constant h is called Planck’s constant, one of the most accurately known fundamental constants in
science
−34
h = 6.626070040(81) × 10 J⋅s

However, for our purposes, its value to four significant figures is sufficient:
−34
h = 6.626 × 10 J⋅s

As the frequency of electromagnetic radiation increases, the magnitude of the associated quantum of radiant energy increases. By
assuming that energy can be emitted by an object only in integral multiples of hν, Planck devised an equation that fit the
experimental data shown in Figure 1.2.1 . We can understand Planck’s explanation of the ultraviolet catastrophe qualitatively as
follows: At low temperatures, radiation with only relatively low frequencies is emitted, corresponding to low-energy quanta. As the
temperature of an object increases, there is an increased probability of emitting radiation with higher frequencies, corresponding to
higher-energy quanta. At any temperature, however, it is simply more probable for an object to lose energy by emitting a large
number of lower-energy quanta than a single very high-energy quantum that corresponds to ultraviolet radiation. The result is a
maximum in the plot of intensity of emitted radiation versus wavelength, as shown in Figure 1.2.1 , and a shift in the position of the
maximum to lower wavelength (higher frequency) with increasing temperature.
At the time he proposed his radical hypothesis, Planck could not explain why energies should be quantized. Initially, his hypothesis
explained only one set of experimental data—blackbody radiation. If quantization were observed for a large number of different
phenomena, then quantization would become a law. In time, a theory might be developed to explain that law. As things turned out,
Planck’s hypothesis was the seed from which modern physics grew.
Max Planck explain the spectral distribution of blackbody radiation as result from oscillations of electrons. Similarly, oscillations
of electrons in an antenna produce radio waves. Max Planck concentrated on modeling the oscillating charges that must exist in the
oven walls, radiating heat inwards and—in thermodynamic equilibrium—themselves being driven by the radiation field. He found
he could account for the observed curve if he required these oscillators not to radiate energy continuously, as the classical theory

11.2.2 https://chem.libretexts.org/@go/page/41369
would demand, but they could only lose or gain energy in chunks, called quanta, of size hν , for an oscillator of frequency ν

(Equation 11.2.1).
With that assumption, Planck calculated the following formula for the radiation energy density inside the oven:

dρ(ν , T ) = ρν (T )dν (11.2.2)

3
2hν 1
= ⋅ dν (11.2.3)
2
c hν
exp( ) −1
kB T

with
π = 3.14159

h = 6.626 × 10 J ⋅ s −34

c = 3.00 × 10 m/s
8

ν = 1/s

k = 1.38 × 10 J/K
B
−23

T is absolute temperature (in Kelvin)

Planck's radiation energy density (Equation 11.2.3) can also be expressed in terms of wavelength λ .
2
2hc 1
ρ(λ, T )dλ = dλ (11.2.4)
5
λ hc
exp( ) −1
λ kB T

Planck's equation (Equation 11.2.4 ) gave an excellent agreement with the experimental observations for all temperatures (Figure
1.2.2 ).

Figure 1.2.2 : The Sun is an excellent approximation of a blackbody. Its effective temperature is ~5777 K. (CC-SA-BY 3.0; Sch).

 Max Planck (1858–1947)

Planck made many substantial contributions to theoretical physics, but his fame as a physicist rests primarily on his role as the
originator of quantum theory. In addition to being a physicist, Planck was a gifted pianist, who at one time considered music as
a career. During the 1930s, Planck felt it was his duty to remain in Germany, despite his open opposition to the policies of the
Nazi government.

11.2.3 https://chem.libretexts.org/@go/page/41369
(left) The German physicist Max Planck had a major influence on the early development of quantum mechanics, being the first
to recognize that energy is sometimes quantized. Planck also made important contributions to special relativity and classical
physics. (Public Domain; Library of Congress via Wikimedia) (left) The society's logo features Minerva, the Roman goddess
of wisdom. (Fair use)
One of his sons was executed in 1944 for his part in an unsuccessful attempt to assassinate Hitler and bombing during the last
weeks of World War II destroyed Planck’s home. After WWII, the major German scientific research organization was renamed
the Max Planck Society.

 Exercise 1.2.1

Use Equation 11.2.4 to show that the units of ρ(λ, T ) dλ are J/m as expected for an energy density.
3

The near perfect agreement of this formula with precise experiments (e.g., Figure 1.2.3 ), and the consequent necessity of energy
quantization, was the most important advance in physics in the century. His blackbody curve was completely accepted as the
correct one: more and more accurate experiments confirmed it time and again, yet the radical nature of the quantum assumption did
not sink in. Planck was not too upset—he didn’t believe it either, he saw it as a technical fix that (he hoped) would eventually prove
unnecessary.

Part of the problem was that Planck’s route to the formula was long, difficult and
implausible—he even made contradictory assumptions at different stages, as Einstein
pointed out later. However, the result was correct anyway!
The mathematics implied that the energy given off by a blackbody was not continuous, but given off at certain specific
wavelengths, in regular increments. If Planck assumed that the energy of blackbody radiation was in the form

E = nhν

where n is an integer, then he could explain what the mathematics represented. This was indeed difficult for Planck to accept,
because at the time, there was no reason to presume that the energy should only be radiated at specific frequencies. Nothing in
Maxwell’s laws suggested such a thing. It was as if the vibrations of a mass on the end of a spring could only occur at specific
energies. Imagine the mass slowly coming to rest due to friction, but not in a continuous manner. Instead, the mass jumps from one
fixed quantity of energy to another without passing through the intermediate energies.

To use a different analogy, it is as if what we had always imagined as smooth inclined


planes were, in fact, a series of closely spaced steps that only presented the illusion of
continuity.

Summary
The agreement between Planck’s theory and the experimental observation provided strong evidence that the energy of electron
motion in matter is quantized. In the next two sections, we will see that the energy carried by light also is quantized in units of hν̄ .
These packets of energy are called “photons.”

Contributors and Attributions


Michael Fowler (Beams Professor, Department of Physics, University of Virginia)
David M. Hanson, Erica Harvey, Robert Sweeney, Theresa Julia Zielinski ("Quantum States of Atoms and Molecules")

11.2: Planck's Quantum Theory is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

11.2.4 https://chem.libretexts.org/@go/page/41369
1.2: Quantum Hypothesis Used for Blackbody Radiation Law is licensed CC BY-NC-SA 4.0.

11.2.5 https://chem.libretexts.org/@go/page/41369
11.3: The Photoelectric Effect
 Learning Objectives
To be familiar with the photoelectron effect for bulk materials
Understand how the photoelectron kinetic energy and intensity vary as a function of incident light wavelength
Understand how the photoelectron kinetic energy and intensity vary as a function of incident light intensity
Describe what a workfunction is and relate it to ionization energy
Describe the photoelectric effect with Einstein's quantized photon model of light

Nature, it seemed, was quantized (non-continuous, or discrete). If this was so, how could Maxwell’s equations correctly predict the
result of the blackbody radiator? Planck spent a good deal of time attempting to reconcile the behavior of electromagnetic waves
with the discrete nature of the blackbody radiation, to no avail. It was not until 1905, with yet another paper published by Albert
Einstein, that the wave nature of light was expanded to include the particle interpretation of light which adequately explained
Planck’s equation.
The photoelectric effect was first documented in 1887 by the German physicist Heinrich Hertz and is therefore sometimes referred
to as the Hertz effect. While working with a spark-gap transmitter (a primitive radio-broadcasting device), Hertz discovered that
upon absorption of certain frequencies of light, substances would give off a visible spark. In 1899, this spark was identified as
light-excited electrons (called photoelectrons) leaving the metal's surface by J.J. Thomson (Figure 1.3.1 ).

Figure 1.3.1 : The photoelectric effect involves irradiating a metal surface with photons of sufficiently high energy to cause the
electrons to be ejected from the metal. (CC BY-SA-NC; anonymous)
The classical picture underlying the photoelectron effect was that the atoms in the metal contained electrons, that were shaken and
caused to vibrate by the oscillating electric field of the incident radiation. Eventually some of them would be shaken loose, and
would be ejected from the cathode. It is worthwhile considering carefully how the number and speed of electrons emitted would be
expected to vary with the intensity and color of the incident radiation along with the time needed to observe the photoelectrons.
Increasing the intensity of radiation would shake the electrons more violently, so one would expect more to be emitted, and they
would shoot out at greater speed, on average.
Increasing the frequency of the radiation would shake the electrons faster, so it might cause the electrons to come out faster. For
very dim light, it would take some time for an electron to work up to a sufficient amplitude of vibration to shake loose.

Lenard's Experimental Results (Intensity Dependence)


In 1902, Hertz's student, Philipp Lenard, studied how the energy of the emitted photoelectrons varied with the intensity of the light.
He used a carbon arc light and could increase the intensity a thousand-fold. The ejected electrons hit another metal plate, the
collector, which was connected to the cathode by a wire with a sensitive ammeter, to measure the current produced by the
illumination (Figure 1.3.2 ). To measure the energy of the ejected electrons, Lenard charged the collector plate negatively, to repel
the electrons coming towards it. Thus, only electrons ejected with enough kinetic energy to get up this potential hill would
contribute to the current.

11.3.1 https://chem.libretexts.org/@go/page/41370
incoming blue light incoming blue light

collector plate emitter plate collector plate emitter plate

vacuum vacuum
electrons get to collector plate electrons get to collector plate

+ - mA
+ - mA

Ammeter Ammeter
V V
Figure 11.3.2 : Lenard's photoelectric experiment. (left) High light intensity increase photocurrent (number of collected
photoelectrons). (right) Low light intensity has reduced photocurrent. However, the kinetic energy of the ejected electrons is
independent of incident light intensity. (CC BY-NC; Ümit Kaya via LibreTexts)
Lenard discovered that there was a well defined minimum voltage that stopped any electrons getting through (V ). To Lenard's stop

surprise, he found that V did not depend at all on the intensity of the light! Doubling the light intensity doubled the number of
stop

electrons emitted, but did not affect the kinetic energies of the emitted electrons. The more powerful oscillating field ejected more
electrons, but the maximum individual energy of the ejected electrons was the same as for the weaker field (Figure 1.3.2 ).

Millikan's Experimental Results (Wavelength Dependence)


The American experimental physicist Robert Millikan followed up on Lenard's experiments and using a powerful arc lamp, he was
able to generate sufficient light intensity to separate out the colors and check the photoelectric effect using light of different colors.
He found that the maximum energy of the ejected electrons did depend on the color - the shorter wavelength, higher frequency light
eject photoelectrons with greater kinetic energy (Figures 1.3.3 ).
incoming blue light incoming red light

collector plate emitter plate collector plate emitter plate

vacuum vacuum
electrons get to collector plate electrons don't get to plate

+ - mA
+ - mA
no current
Ammeter Ammeter
V V
Figure 1.3.3 : Millikan's photoelectric experiment. (left) Incident high-energy blue light. The battery represents the potential Lenard
used to charge the collector plate negatively, which would actually be a variable voltage source. Since the electrons ejected by the
blue light are getting to the collector plate, the potential supplied by the battery is less than V , for blue light. (right) Indicent
stop

low-energy red light. Since the electrons ejected by the red light are not getting to the collector plate, the potential supplied by the
battery exceeds V for red light. (CC BY-NC; Ümit Kaya via LibreTexts)
stop

As shown in Figure 1.3.4 , just the opposite behavior from classical is observed from Lenard's and Millikan's experiments. The
intensity affects the number of electrons, and the frequency affects the kinetic energy of the emitted electrons. From these sketches,
we see that
the kinetic energy of the electrons is linearly proportional to the frequency of the incident radiation above a threshold value of
ν (no current is observed below ν ), and the kinetic energy is independent of the intensity of the radiation, and
0 0

the number of electrons (i.e. the electric current) is proportional to the intensity and independent of the frequency of the incident
radiation above the threshold value of ν (i.e., no current is observed below ν ).
0 0

11.3.2 https://chem.libretexts.org/@go/page/41370
Figure 1.3.4 : Schematic drawings showing the characteristics of the photoelectric effect from Lenard's and Millikan's experiments.
(A) The kinetic energy of any single emitted electron increases linearly with frequency above some threshold value (B) The
electron kinetic energy is independent of the light intensity above the threshold frequency and zero below. (C) The number of
electrons emitted per second (i.e. the electric current) is independent of light frequency above the threshhold frequency and zero
below. (D) The number of electrons increases linearly with the light intensity. (CC BY-NC; Ümit Kaya via LibreTexts)

 Classical Theory does not Describe Experiment

Classical theory predicts that energy carried by light is proportional to its amplitude independent of its frequency, and this fails
to correctly explain the observed wavelength dependence in Lenard's and Millikan's observations.

As with most of the experimental results we discuss in this text, the behavior described above is a simplification of the true
experimental results observed in the laboratory. A more complex description involves a greater introduction of more complex
physics and instrumentation, which will be ignored for now.

Einstein's Quantum Picture


In 1905 Einstein gave a very simple interpretation of Lenard's results and borrowed Planck's hypothesis about the quantized energy
from his blackbody research and assumed that the incoming radiation should be thought of as quanta of energy hν , with ν the
frequency. In photoemission, one such quantum is absorbed by one electron. If the electron is some distance into the material of the
cathode, some energy will be lost as it moves towards the surface. There will always be some electrostatic cost as the electron
leaves the surface, which is the workfunction, Φ. The most energetic electrons emitted will be those very close to the surface, and
they will leave the cathode with kinetic energy
KE = hν − Φ (11.3.1)

On cranking up the negative voltage on the collector plate until the current just stops, that is, to Vstop , the highest kinetic energy
electrons (KE ) must have had energy eV
e upon leaving the cathode. Thus,
stop

eVstop = hν − Φ (11.3.2)

Thus, Einstein's theory makes a very definite quantitative prediction: if the frequency of the incident light is varied, and V stop

plotted as a function of frequency, the slope of the line should be (Figure 11.3.4A). It is also clear that there is a minimum light
h

frequency for a given metal ν , that for which the quantum of energy is equal to Φ (Equation 11.3.1). Light below that frequency,
o

no matter how bright, will not eject electrons.

According to both Planck and Einstein, the energy of light is proportional to its frequency
rather than its amplitude, there will be a minimum frequency ν needed to eject an 0

electron with no residual energy.


Since every photon of sufficient energy excites only one electron, increasing the light's intensity (i.e. the number of photons/sec)
only increases the number of released electrons and not their kinetic energy. In addition, no time is necessary for the atom to be
heated to a critical temperature and therefore the release of the electron is nearly instantaneous upon absorption of the light. Finally,
because the photons must be above a certain energy to satisfy the workfunction, a threshold frequency exists below which no

11.3.3 https://chem.libretexts.org/@go/page/41370
photoelectrons are observed. This frequency is measured in units of Hertz (1/second) in honor of the discoverer of the photoelectric
effect.
Einstein's Equation 11.3.1 explains the properties of the photoelectric effect quantitatively. A strange implication of this
experiment is that light can behave as a kind of massless "particle" now known as a photon whose energy E = hν can be
transferred to an actual particle (an electron), imparting kinetic energy to it, just as in an elastic collision between to massive
particles such as billiard balls.

Robert Millikan initially did not accept Einstein's theory, which he saw as an attack on the wave theory of light, and worked
for ten years until 1916, on the photoelectric effect. He even devised techniques for scraping clean the metal surfaces inside
the vacuum tube. For all his efforts he found disappointing results: he confirmed Einstein's theory after ten years. In what he
writes in his paper, Millikan is still desperately struggling to avoid this conclusion. However, by the time of his Nobel Prize
acceptance speech, he has changed his mind rather drastically!

Einstein's simple explanation (Equation 11.3.1) completely accounted for the observed phenomena in Lenard's and Millikan's
experiments (Figure 1.3.4 ) and began an investigation into the field we now call quantum mechanics. This new field seeks to
provide a quantum explanation for classical mechanics and create a more unified theory of physics and thermodynamics. The study
of the photoelectric effect has also lead to the creation of new field of photoelectron spectroscopy. Einstein's theory of the
photoelectron presented a completely different way to measure Planck's constant than from black-body radiation.

The Workfunction (Φ)


The workfunction is an intrinsic property of the metal. While the workfunctions and ionization energies appear as similar
concepts, they are independent. The workfunction of a metal is the minimum amount of energy (E) necessary to remove an
electron from the surface of the bulk (solid) metal (sometimes referred to as binding energy).
+ −
M(s) + Φ → M (s) + e (free with no kinetic energy)

The workfunction is qualitatively similar to ionization energy (IE), which is the amount of energy required to remove an electron
from an atom or molecule in the gaseous state.
+ −
M(g) + IE → M (g) + e (free with no kinetic energy)

However, these two energies differ in magnitude (Table 1.3.1 ). For instance, copper has a workfunction of about 4.7 eV, but has a
higher ionization energy of 7.7 eV. Generally, the ionization energies for metals are greater than the corresponding workfunctions
(i.e., the electrons are less tightly bound in bulk metal).
Table 1.3.1 : Workfunctions and Ionization Energies of Select Elements
Element workfunction Φ (eV) Ionization Energy (eV)

Copper (Cu) 4.7 7.7

Silver (Ag) 4.72 7.57

Aluminum (Al) 4.20 5.98

Gold (Au) 5.17 9.22

Boron (B) 4.45 8.298

Beryllium (Be) 4.98 9.32

Bismuth (Bi) 4.34 7.29

Carbon (C) 5.0 11.26

Cesium (Cs) 1.95 3.89

Iron (Fe) 4.67 7.87

Gallium (Ga) 4.32 5.99

(Hg) liquid 4.47 10.43

11.3.4 https://chem.libretexts.org/@go/page/41370
Element workfunction Φ (eV) Ionization Energy (eV)

Sodium (Na) 2.36 5.13

Lithium (Li) 2.93 5.39

Potassium 2.3 4.34

Selenium (Se) 5.9 9.75

Silicon (Si) 4.85 8.15

Tin (Sn) 4.42 7.34

Germanium (Ge) 5.0 7.89

Arsenic (As) 3.75 9.81

 Example 1.3.1 : Calcium


a. What is the energy in joules and electron volts of a photon of 420-nm violet light?
b. What is the maximum kinetic energy of electrons ejected from calcium by 420-nm violet light, given that the workfunction
for calcium metal is 2.71 eV?
Strategy
To solve part (a), note that the energy of a photon is given by E = hν . For part (b), once the energy of the photon is calculated,
it is a straightforward application of Equation 11.3.1 to find the ejected electron’s maximum kinetic energy, since Φ is given.
Solution for (a)
Photon energy is given by

E = hν

Since we are given the wavelength rather than the frequency, we solve the familiar relationship c = νλ for the frequency,
yielding
c
ν =
λ

Combining these two equations gives the useful relationship


hc
E =
λ

Now substituting known values yields


−34 8
(6.63 × 10 J ⋅ s)(3.00 × 10 m/s)
E =
−9
420 × 10 m

−19
= 4.74 × 10 J

Converting to eV, the energy of the photon is


1 eV
−19
E = (4.74 × 10 J) ( )
−19
1.6 × 10 J

= 2.96 eV .

Solution for (b)


Finding the kinetic energy of the ejected electron is now a simple application of Equation 11.3.1. Substituting the photon
energy and binding energy yields

11.3.5 https://chem.libretexts.org/@go/page/41370
KEe = hν – Φ

= 2.96 eV – 2.71 eV

= 0.246 eV .

Discussion
The energy of this 420-nm photon of violet light is a tiny fraction of a joule, and so it is no wonder that a single photon would
be difficult for us to sense directly—humans are more attuned to energies on the order of joules. But looking at the energy in
electron volts, we can see that this photon has enough energy to affect atoms and molecules. A DNA molecule can be broken
with about 1 eV of energy, for example, and typical atomic and molecular energies are on the order of eV, so that the UV
photon in this example could have biological effects.
The ejected electron (called a photoelectron) has a rather low energy, and it would not travel far, except in a vacuum. The
electron would be stopped by a retarding potential of 0.26 eV. In fact, if the photon wavelength were longer and its energy less
than 2.71 eV, then the formula would give a negative kinetic energy, an impossibility. This simply means that the 420-nm
photons with their 2.96-eV energy are not much above the frequency threshold. You can show for yourself that the threshold
wavelength is 459 nm (blue light). This means that if calcium metal is used in a light meter, the meter will be insensitive to
wavelengths longer than those of blue light. Such a light meter would be insensitive to red light, for example.

 Exercise 1.3.1 : Silver

What is the longest-wavelength electromagnetic radiation that can eject a photoelectron from silver? Is this in the visible
range?

Answer
Given that the workfunction is 4.72 eV from Table 1.3.1 , then only photons with wavelengths lower than 263 nm will
induce photoelectrons (calculated via E = hν ). This is ultraviolet and not in the visible range.

 Exercise 1.3.2

Why is the workfunction of an element generally lower than the ionization energy of that element?

Answer
The workfunction of a metal refers to the minimum energy required to extract an electron from the surface of a (bulk)
metal by the absorption a photon of light. The workfunction will vary from metal to metal. In contrast, ionization energy is
the energy needed to detach electrons from atoms and also varies with each particular atom, with the valence electrons
require less energy to extract than core electrons (i.e., from lower shells) that are more closely bound to the nuclei. The
electrons in the metal lattice there less bound (i.e., free to move within the metal) and removing one of these electrons is
much easier than removing an electron from an atom because the metallic bonds of the bulk metal reduces their binding
energy. As we will show in subsequent chapters, the more delocalized a particle is, the lower its energy.

Summary
Although Hertz discovered the photoelectron in 1887, it was not until 1905 that a theory was proposed that explained the effect
completely. The theory was proposed by Einstein and it made the claim that electromagnetic radiation had to be thought of as a
series of particles, called photons, which collide with the electrons on the surface and emit them. This theory ran contrary to the
belief that electromagnetic radiation was a wave and thus it was not recognized as correct until 1916 when Robert Millikan
experimentally confirmed the theory
The photoelectric effect is the process in which electromagnetic radiation ejects electrons from a material. Einstein proposed
photons to be quanta of electromagnetic radiation having energy E = hν is the frequency of the radiation. All electromagnetic
radiation is composed of photons. As Einstein explained, all characteristics of the photoelectric effect are due to the interaction of
individual photons with individual electrons. The maximum kinetic energy KE of ejected electrons (photoelectrons) is given by
e

11.3.6 https://chem.libretexts.org/@go/page/41370
K Ee = hν – Φ , where hν is the photon energy and Φ is the workfunction (or binding energy) of the electron to the particular
material.

Conceptual Questions
1. Is visible light the only type of electromagnetic radiation that can cause the photoelectric effect?
2. Which aspects of the photoelectric effect cannot be explained without photons? Which can be explained without photons? Are
the latter inconsistent with the existence of photons?
3. Is the photoelectric effect a direct consequence of the wave character of electromagnetic radiation or of the particle character of
electromagnetic radiation? Explain briefly.
4. Insulators (nonmetals) have a higher Φ than metals, and it is more difficult for photons to eject electrons from insulators.
Discuss how this relates to the free charges in metals that make them good conductors.
5. If you pick up and shake a piece of metal that has electrons in it free to move as a current, no electrons fall out. Yet if you heat
the metal, electrons can be boiled off. Explain both of these facts as they relate to the amount and distribution of energy
involved with shaking the object as compared with heating it.

Contributors and Attributions


Michael Fowler (Beams Professor, Department of Physics, University of Virginia)
Mark Tuckerman (New York University)
David M. Hanson, Erica Harvey, Robert Sweeney, Theresa Julia Zielinski ("Quantum States of Atoms and Molecules")
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley (Stephen F. Austin State University) with contributing authors. Textbook
content produced by OpenStax College is licensed under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-2bd...a7ac8df6@9.110).

11.3: The Photoelectric Effect is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
1.3: Photoelectric Effect Explained with Quantum Hypothesis is licensed CC BY-NC-SA 4.0.

11.3.7 https://chem.libretexts.org/@go/page/41370
11.4: Bohr's Theory of the Hydrogen Emission Spectrum
Overview
To introduce the concept of absorption and emission line spectra and describe the Balmer equation to describe the visible
lines of atomic hydrogen.
Describe Rydberg's theory for the hydrogen spectra.
Interpret the hydrogen spectrum in terms of the energy states of electrons.

The first person to realize that white light was made up of the colors of the rainbow was Isaac Newton, who in 1666 passed
sunlight through a narrow slit, then a prism, to project the colored spectrum on to a wall. This effect had been noticed previously, of
course, not least in the sky, but previous attempts to explain it, by Descartes and others, had suggested that the white light became
colored when it was refracted, the color depending on the angle of refraction. Newton clarified the situation by using a second
prism to reconstitute the white light, making much more plausible the idea that the white light was composed of the separate colors.
He then took a monochromatic component from the spectrum generated by one prism and passed it through a second prism,
establishing that no further colors were generated. That is, light of a single color did not change color on refraction. He concluded
that white light was made up of all the colors of the rainbow, and that on passing through a prism, these different colors were
refracted through slightly different angles, thus separating them into the observed spectrum.

Atomic Line Spectrum


The spectrum of hydrogen, which turned out to be crucial in providing the first insight into atomic structure over half a century
later, was first observed by Anders Angstrom in Uppsala, Sweden, in 1853. His communication was translated into English in
1855. Angstrom, the son of a country minister, was a reserved person, not interested in the social life that centered around the court.
Consequently, it was many years before his achievements were recognized, at home or abroad (most of his results were published
in Swedish).
Most of what is known about atomic (and molecular) structure and mechanics has been deduced from spectroscopy. Figure 1.4.1
shows two different types of spectra. A continuous spectrum can be produced by an incandescent solid or gas at high pressure
(blackbody radiation, for example, is a continuum). An emission spectrum can be produced by a gas at low pressure excited by heat
or by collisions with electrons. An absorption spectrum results when light from a continuous source passes through a cooler gas,
consisting of a series of dark lines characteristic of the composition of the gas.

Figure 11.4.1 : Continuous spectrum and two types of line spectra. from Wikipedia.

In 1802, William Wollaston in England had discovered (perhaps by using a thinner slit or a better prism) that in fact the solar
spectrum itself had tiny gaps - there were many thin dark lines in the rainbow of colors. These were investigated much more
systematically by Joseph von Fraunhofer, beginning in 1814. He increased the dispersion by using more than one prism. He found
an "almost countless number" of lines. He labeled the strongest dark lines A, B, C, D, etc. Frauenhofer between 1814 and 1823
discovered nearly 600 dark lines in the solar spectrum viewed at high resolution and designated the principal features with the
letters A through K, and weaker lines with other letters (Table 11.4.1). Modern observations of sunlight can detect many thousands
of lines. It is now understood that these lines are caused by absorption by the outer layers of the Sun.
Table 11.4.1 : Major Fraunhofer lines, and the elements they are associated with
Designation Element Wavelength (nm)

y O2 898.765

Z O2 822.696

11.4.1 https://chem.libretexts.org/@go/page/41371
Designation Element Wavelength (nm)

A O2 759.370

B O2 686.719

C H 656.281

a O2 627.661

D1 Na 589.592

D2 Na 588.995

D3 or d He 587.5618

Fraunhofer Absorption Lines


The Fraunhofer lines are typical spectral absorption lines. These dark lines are produced whenever a cold gas is between a
broad spectrum photon source and the detector. In this case, a decrease in the intensity of light in the frequency of the incident
photon is seen as the photons are absorbed, then re-emitted in random directions, which are mostly in directions different from
the original one. This results in an absorption line, since the narrow frequency band of light initially traveling toward the
detector, has been turned into heat or re-emitted in other directions.

Figure 11.4.2 : Spectrum of blue sky. Dips are present at the Fraunhofer line wavelengths. from Wikipedia.
By contrast, if the detector sees photons emitted directly from a glowing gas, then the detector often sees photons emitted in a
narrow frequency range by quantum emission processes in atoms in the hot gas, resulting in an emission line. In the Sun,
Fraunhofer lines are seen from gas in the outer regions of the Sun, which are too cold to directly produce emission lines of the
elements they represent.

Gases heated to incandescence were found by Bunsen, Kirkhoff and others to emit light with a series of sharp wavelengths. The
emitted light analyzed by a spectrometer (or even a simple prism) appears as a multitude of narrow bands of color. These so called
line spectra are characteristic of the atomic composition of the gas. The line spectra of several elements are shown in Figure
11.4.3.

11.4.2 https://chem.libretexts.org/@go/page/41371
Figure 11.4.3 : The Emission Spectra of Elements Compared with Hydrogen. These images show (a) hydrogen gas, which is
atomized to hydrogen atoms in the discharge tube; (b) helium; and (c) mercury. The strongest lines in the hydrogen spectrum are in
the far UV Lyman series starting at 124 nm and below. The strongest lines in the mercury spectrum are at 181 and 254 nm, also in
the UV. These are not shown.

The Balmer Series of Hydrogen


Obviously, if any pattern could be discerned in the spectral lines for a specifc atom (in contract to the mixture that Fraunhofer lines
represent), that might be a clue as to the internal structure of the atom. One might be able to build a model. A great deal of effort
went into analyzing the spectral data from the 1860's on. The big breakthrough was made by Johann Balmer, a math and Latin
teacher at a girls' school in Basel, Switzerland. Balmer had done no physics before, and made his great discovery when he was
almost sixty.

Figure 11.4.4 : The visible hydrogen emission spectrum lines in the Balmer series. The red line at the right is the Hα line and the
two leftmost lines are considered to be ultraviolet as they have wavelengths less than 400 nm. from Wikipedia.
Balmer decided that the most likely atom to show simple spectral patterns was the lightest atom, hydrogen. Angstrom had
measured the four visible spectral lines to have wavelengths 656.21, 486.07, 434.01 and 410.12 nm (Figure 11.4.4). Balmer
concentrated on just these four numbers, and found they were represented by the formula:
2
n
2
λ =b( ) (11.4.1)
2
n −4
2

where
b= 364.56 nm and
n = 3, 4, 5, 6.
2

The first four wavelengths of Equation 11.4.1 (with n_2=3,4,5,6) were in excellent agreement with the experimental lines from
Angstrom (Table 11.4.2). Balmer predicted that other lines exist in the infrared that correspond to n = 7, 8, etc., and in fact some of
them had already been observed, unbeknown to Balmer.
Table 11.4.2 : The Balmer Series of Hydrogen Emission Lines
n2 3 4 5 6 7 8 9 10

λ 656 486 434 410 397 389 383 380

color red teal blue indigo violet not visible not visible not visible

The n integer in the Balmer series extends theoretically to infinity and the series resents a monotonically increasing energy (and
2

frequency) of the absorption lines with increasing n values. Moreover, the energy difference between successive lines decreased
2

as n increases (Figure 1.4.4). This behavior converges to a highest possible energy as Example 11.4.1 demonstrates. If the lines
2

are plot according to their λ on a linear scale, you will get the appearance of the spectrum in Figure 11.4.4; these lines are called
the Balmer series.

11.4.3 https://chem.libretexts.org/@go/page/41371
Balmer's general formula (Equation 11.4.1) can be rewritten in terms of the inverse wavelength typically called the wavenumber (
ν̃ ).

1 1 1
ν̃ = = RH ( − ) (11.4.2)
2
λ 4 n
2

where
n2 = 3, 4, 5, 6 and
RH is the Rydberg constant (discussed below) and is equal to 109,737 cm-1.
He further conjectured that the 4 could be replaced by 9, 16, 25, … and this also turned out to be true - but these lines, further into
the infrared, were not detected until the early twentieth century, along with the ultraviolet lines.

Wavenumbers
The relation between wavelength and frequency for electromagnetic radiation is

λν = c (11.4.3)

In the SI system of units the wavelength, \lambda\) is measured in meters (m) and since wavelengths are usually very small one
often uses the nanometer (nm) which is 10 m. The frequency (ν ) in the SI system is measured in reciprocal seconds 1/s −
−9

which is called a Hertz (after the discover of the photoelectron effect) and is represented by Hz.
It is common to use the reciprocal of the wavelength in centimeters as a measure of the frequency of radiation. This unit is
called a wavenumber and is represented by (ν̃ ) and is defined by
1 ν
ν̃ = = (11.4.4)
λ c

Wavenumbers is a convenient unit in spectroscopy because it is directly proportional to energy.


hc
E = (11.4.5)
λ

1
E = hc × (11.4.6)
λ

E = hcν̃ (11.4.7)

E ∝ ν̃ (11.4.8)

Balmer Series
Calculate the longest and shortest wavelengths (in nm) emitted in the Balmer series of the hydrogen atom emission spectrum.
Solution
From the behavior of the Balmer equation (Equation 11.4.1 and Table 11.4.2 ), the value of n that gives the longest (i.e.,
2

greatest) wavelength (λ ) is the smallest value possible of n , which is n


2 2 =3 for this series. This results in
9
λlongest = (364.56 nm) ( ) (11.4.9)
9 −4

λlongest = (364.56 nm) (1.8) = 656.2 nm (11.4.10)

This is also known as the H line of atomic hydrogen and is bight red (Figure 1.4.4).
α

For the shortest wavelength, it should be recognized that the shortest wavelength (greatest energy) is obtained at the limit of
greatest (n_2\):
2
n
2
λshortest = lim (364.56 nm) ( ) (11.4.11)
n2 →∞ 2
n −4
2

11.4.4 https://chem.libretexts.org/@go/page/41371
This can be solved via L'Hôpital's Rule, or alternatively the limit can be expressed via the equally useful energy expression
(Equation 1.4.2) and simply solved:

1 1
ν̃ greatest = lim RH ( − ) (11.4.12)
2
n2 →∞ 4 n
2

1 −1
ν̃ greatest = lim RH ( ) = 27, 434 c m (11.4.13)
n2 →∞ 4

1
Since =λ in units of cm, this converts to 364 nm as the shortest wavelength possible for the Balmer series.
ν̃

The Balmer series is particularly useful in astronomy because the Balmer lines appear in numerous stellar objects due to the
abundance of hydrogen in the universe, and therefore are commonly seen and relatively strong compared to lines from other
elements.

The Generalized Rydberg Equation


In an amazing demonstration of mathematical insight, in 1885 Balmer came up with a simple formula for predicting the wavelength
of any of the lines in atomic Hydrogen in what we now know as the Balmer series (Equation 11.4.2). Three years later, Rydberg
generalized this so that it was possible to determine the wavelengths of any of the lines in the hydrogen emission spectrum.
Rydberg suggested that all atomic spectra formed families with this pattern (he was unaware of Balmer's work). It turns out that
there are families of spectra following Rydberg's pattern, notably in the alkali metals, sodium, potassium, etc., but not with the
precision the hydrogen atom lines fit the Balmer formula, and low values of n predicted wavelengths that deviate considerably.
2

Rydberg's general equation is as follows:

1 1 1
ν̃ = = RH ( − ) (11.4.14)
2 2
λ n n
1 2

where
RH is the Rydberg constant and is equal to 109,737 cm-1.
n1 and n are integers (whole numbers) with n > n .
2 2 1

For the Balmer lines, n = 2 and n can be any whole number between 3 and infinity. The various combinations of numbers that
1 2

can be substituted into this formula allow the calculation the wavelength of any of the lines in the hydrogen emission spectrum;
there is close agreement between the wavelengths generated by this formula and those observed in a real spectrum.

Other Series
The results given by Balmer and Rydberg for the spectrum in the visible region of the electromagnetic radiation start with n 2 =3 ,
and n = 2 . Is there a different series with the following formula (e.g., n = 1 ?
2
1 1

1 1 1
= RH ( − ) (11.4.15)
2 2
λ 1 n

The values for n and wavenumber ν̃ for this series would be:
2

Table 11.4.3 : The Lyman Series of Hydrogen Emission Lines (n 1 = 1 )


n2 2 3 4 5 ...

λ (nm) 121 102 97 94 ...

ν̃ (cm-1) 82,2291 97,530 102,864 105,332 ...

These lines are in the UV region, and they are not visible, but they are detected by instruments; these lines form a Lyman series.
The existences of the Lyman series and Balmer's series suggest the existence of more series. For example, the series with n = 3 2
2

and n = 4, 5, 6, 7, ... is called Pashen series.


2
1

11.4.5 https://chem.libretexts.org/@go/page/41371
Note
The spectral lines are grouped into series according to n values. Lines are named sequentially starting from the longest
1

wavelength/lowest frequency of the series, using Greek letters within each series. For example, the (n = 1/n = 2 ) line is 1 2

called "Lyman-alpha" (Ly-α), while the (n = 3/n = 7 ) line is called "Paschen-delta" (Pa-δ). The first six series have
1 2

specific names:
Lyman series with n = 1 1

Balmer series with n = 21

Paschen series (or Bohr series) with n 1 =3

Brackett series with n = 4 1

Pfund series with n = 5


1

Humphreys series with n = 6 1

The spectral series of hydrogen based of the Rydberg Equation (on a logarithmic scale).

Example 11.4.2: The Lyman Series

The so-called Lyman series of lines in the emission spectrum of hydrogen corresponds to transitions from various excited states
to the n = 1 orbit. Calculate the wavelength of the lowest-energy line in the Lyman series to three significant figures. In what
region of the electromagnetic spectrum does it occur?
Given: lowest-energy orbit in the Lyman series
Asked for: wavelength of the lowest-energy Lyman line and corresponding region of the spectrum
Strategy:
A. Substitute the appropriate values into Equation 1.5.1 (the Rydberg equation) and solve for λ .
B. Locate the region of the electromagnetic spectrum corresponding to the calculated wavelength.
Solution:
We can use the Rydberg equation to calculate the wavelength:

1 1 1
= R( − ) (11.4.16)
2 2
λ n n
1 2

A For the Lyman series, n 1 =1 .

1 1 1 1 1
−1 6 −1
= R( − ) = 1.097 × m ( − ) = 8.228 × 10 m (11.4.17)
2 2
λ n n 1 4
1 2

Spectroscopists often talk about energy and frequency as equivalent. The cm-1 unit is particularly convenient. The infrared
range is roughly 200 - 5,000 cm-1, the visible from 11,000 to 25.000 cm-1 and the UV between 25,000 and 100,000 cm-1. The
units of cm-1 are called wavenumbers, although people often verbalize it as inverse centimeters. We can convert the answer in
part A to cm-1.

1 m
6 −1 −1
ν̃ = = 8.228 × 10 m ( ) = 82, 280 c m (11.4.18)
λ 100 cm

and
−7
λ = 1.215 × 10 m = 122 nm (11.4.19)

11.4.6 https://chem.libretexts.org/@go/page/41371
This emission line is called Lyman alpha. It is the strongest atomic emission line from the sun and drives the chemistry of the
upper atmosphere of all the planets producing ions by stripping electrons from atoms and molecules. It is completely absorbed
by oxygen in the upper stratosphere, dissociating O2 molecules to O atoms which react with other O2 molecules to form
stratospheric ozone
B This wavelength is in the ultraviolet region of the spectrum.

Exercise 11.4.2: The Pfund Series

The Pfund series of lines in the emission spectrum of hydrogen corresponds to transitions from higher excited states to the \
(n_1 = 5\) orbit. Calculate the wavelength of the second line in the Pfund series to three significant figures. In which region of
the spectrum does it lie?
Answer: 4.65 × 103 nm; infrared

The above discussion presents only a phenomenological description of hydrogen emission lines and fails to provide a probe of the
nature of the atom itself. Clearly a continuum model based on classical mechanics is not applicable.

Contributors and Attributions


Michael Fowler (Beams Professor, Department of Physics, University of Virginia)
Chung (Peter) Chieh (Professor Emeritus, Chemistry @ University of Waterloo)

11.4: Bohr's Theory of the Hydrogen Emission Spectrum is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by LibreTexts.

11.4.7 https://chem.libretexts.org/@go/page/41371
11.5: de Broglie's Postulate
 Learning Objectives
To introduce the wave-particle duality of light extends to matter
To describe how matter (e.g., electrons and protons) can exhibit wavelike properties, e.g., interference and diffraction
patterns
To use algebra to find the de Broglie wavelength or momentum of a particle when either one of these quantities is given

The next real advance in understanding the atom came from an unlikely quarter - a student prince in Paris. Prince Louis de Broglie
was a member of an illustrious family, prominent in politics and the military since the 1600's. Louis began his university studies
with history, but his elder brother Maurice studied x-rays in his own laboratory, and Louis became interested in physics. After
World War I, de Broglie focused his attention on Einstein's two major achievements, the theory of special relativity and the
quantization of light waves. He wondered if there could be some connection between them. Perhaps the quantum of radiation really
should be thought of as a particle. De Broglie suggested that if waves (photons) could behave as particles, as demonstrated by the
photoelectric effect, then the converse, namely that particles could behave as waves, should be true. He associated a wavelength λ
to a particle with momentum p using Planck's constant as the constant of proportionality:
h
λ = (11.5.1)
p

which is called the de Broglie wavelength. The fact that particles can behave as waves but also as particles, depending on which
experiment you perform on them, is known as the wave-particle duality.

 Deriving the de Broglie Wavelength


From the discussion of the photoelectric effect, we have the first part of the particle-wave duality, namely, that electromagnetic
waves can behave like particles. These particles are known as photons, and they move at the speed of light. Any particle that
moves at or near the speed of light has kinetic energy given by Einstein's special theory of relatively. In general, a particle of
mass m and momentum p has an energy
−−−−−−−−−−
2 2 2 4
E = √p c +m c (11.5.2)

Note that if p = 0 , this reduces to the famous rest-energy expression E = mc . However, photons are massless particles
2

(technically rest-massless) that always have a finite momentum p. In this case, Equation 11.5.2 becomes

E = pc.

From Planck's hypothesis, one quantum of electromagnetic radiation has energy E = hν . Thus, equating these two expressions
for the kinetic energy of a photon, we have
hc
hν = = pc (11.5.3)
λ

Solving for the wavelength λ gives Equation 11.5.1:


h h
λ = =
p mv

where v is the velocity of the particle. Hence, de Broglie argued that if particles can behave as waves, then a relationship like
this, which pertains particularly to waves, should also apply to particles.

Equation 11.5.1 allows us to associate a wavelength λ to a particle with momentum p. As the momentum increases, the
wavelength decreases. In both cases, this means the energy becomes larger. i.e., short wavelengths and high momenta correspond
to high energies.

11.5.1 https://chem.libretexts.org/@go/page/41372
It is a common feature of quantum mechanics that particles and waves with short
wavelengths correspond to high energies and vice versa.
Having decided that the photon might well be a particle with a rest mass, even if very small, it dawned on de Broglie that in other
respects it might not be too different from other particles, especially the very light electron. In particular, maybe the electron also
had an associated wave. The obvious objection was that if the electron was wavelike, why had no diffraction or interference effects
been observed? But there was an answer. If de Broglie's relation between momentum and wavelength also held for electrons, the
wavelength was sufficiently short that these effects would be easy to miss. As de Broglie himself pointed out, the wave nature of
light is not very evident in everyday life. As the next section will demonstrate, the validity of de Broglie’s proposal was confirmed
by electron diffraction experiments of G.P. Thomson in 1926 and of C. Davisson and L. H. Germer in 1927. In these experiments it
was found that electrons were scattered from atoms in a crystal and that these scattered electrons produced an interference pattern.
These diffraction patterns are characteristic of wave-like behavior and are exhibited by both electrons (i.e., matter) and
electromagnetic radiation (i.e., light).

 Example 1.6.1 : Electron Waves

Calculate the de Broglie wavelength for an electron with a kinetic energy of 1000 eV.

Solution
To calculate the de Broglie wavelength (Equation 11.5.1), the momentum of the particle must be established and requires
knowledge of both the mass and velocity of the particle. The mass of an electron is 9.109383 × 10 g and the velocity is
−28

obtained from the given kinetic energy of 1000 eV:


2
mv
KE =
2

2
p
= = 1000 eV
2m

Solve for momentum


−−−−−−
p = √2mKE

convert to SI units
−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−− −

−19
 1.6 × 10 J −31
p =  (1000 eV ) ( ) (2)(9.109383 × 10 kg)
⎷ 1 eV

expanding definition of joule into base SI units and cancel


−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−
−16 2 2 −31
p = √ (3.1 × 10 kg ⋅ m / s )(9.109383 × 10 kg)

−−−−−−−−−−−−−−−−−
−40 2 2 2
= √ 2.9 × 10 kg m /s

−23
= 1.7 × 10 kg ⋅ m/s

Now substitute the momentum into the equation for de Broglie's wavelength (Equation ) with Planck's constant (
11.5.1

J ⋅ s ). After expanding units in Plank's constant


−34
h = 6.626069 × 10

h
λ =
p

−34 2
6.626069 × 10 kg ⋅ m /s
=
−23
1.7 × 10 kg ⋅ m/s

−11
= 3.87 × 10 m

= 38.9 pm

11.5.2 https://chem.libretexts.org/@go/page/41372
 Exercise 1.6.1 : Baseball Waves

Calculate the de Broglie wavelength for a fast ball thrown at 100 miles per hour and weighing 4 ounces. Comment on whether
the wave properties of baseballs could be experimentally observed.

Answer
Following the unit conversions below, a 4 oz baseball has a mass of 0.11 kg. The velocity of a fast ball thrown at 100 miles
per hour in m/s is 44.7 m/s.
0.0283 kg
m = (4 oz ) ( ) = 0.11kg
1 oz

100 mi 1609.34 m 1 hr
v=( )( )( ) = 44.7 m/s
hr mi 3600 s

The de Broglie wavelength of this fast ball is:


−34 2
h 6.626069 × 10 kg ⋅ m /s
−34
λ = = = 1.3 × 10 m
mv (0.11 kg)(44.7 m/s)

 Exercise 1.6.2 : Electrons vs. Protons

If an electron and a proton have the same velocity, which would have the longer de Broglie wavelength?
a. The electron
b. The proton
c. They would have the same wavelength

Answer
Equation 11.5.1shows that the de Broglie wavelength of a particle's matter wave is inversely proportional to its momentum
(mass times velocity). Therefore the smaller mass particle will have a smaller momentum and longer wavelength. The
electron is the lightest and will have the longest wavelength.

This was the prince's Ph.D. thesis, presented in 1924. His thesis advisor was somewhat taken aback, and was not sure if this was
sound work. He asked de Broglie for an extra copy of the thesis, which he sent to Einstein. Einstein wrote shortly afterwards: "I
believe it is a first feeble ray of light on this worst of our physics enigmas" and the prince got his Ph.D.

Contributors and Attributions


Michael Fowler (Beams Professor, Department of Physics, University of Virginia)
Mark Tuckerman (New York University)
David M. Hanson, Erica Harvey, Robert Sweeney, Theresa Julia Zielinski ("Quantum States of Atoms and Molecules")

11.5: de Broglie's Postulate is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
1.6: Matter Has Wavelike Properties is licensed CC BY-NC-SA 4.0.

11.5.3 https://chem.libretexts.org/@go/page/41372
11.6: The Heisenberg Uncertainty Principle
 Learning Objectives
To understand that sometime you cannot know everything about a quantum system as demonstrated by the Heisenberg
uncertainly principle.

In classical physics, studying the behavior of a physical system is often a simple task due to the fact that several physical qualities
can be measured simultaneously. However, this possibility is absent in the quantum world. In 1927 the German physicist Werner
Heisenberg described such limitations as the Heisenberg Uncertainty Principle, or simply the Uncertainty Principle, stating that it is
not possible to measure both the momentum and position of a particle simultaneously.
The Heisenberg Uncertainty Principle is a fundamental theory in quantum mechanics that defines why a scientist cannot measure
multiple quantum variables simultaneously. Until the dawn of quantum mechanics, it was held as a fact that all variables of an
object could be known to exact precision simultaneously for a given moment. Newtonian physics placed no limits on how better
procedures and techniques could reduce measurement uncertainty so that it was conceivable that with proper care and accuracy all
information could be defined. Heisenberg made the bold proposition that there is a lower limit to this precision making our
knowledge of a particle inherently uncertain.

Probability
Matter and photons are waves, implying they are spread out over some distance. What is the position of a particle, such as an
electron? Is it at the center of the wave? The answer lies in how you measure the position of an electron. Experiments show that
you will find the electron at some definite location, unlike a wave. But if you set up exactly the same situation and measure it
again, you will find the electron in a different location, often far outside any experimental uncertainty in your measurement.
Repeated measurements will display a statistical distribution of locations that appears wavelike (Figure 1.9.1 ).

Figure 1.9.1 : The building up of the diffraction pattern of electrons scattered from a crystal surface. Each electron arrives at a
definite location, which cannot be precisely predicted. The overall distribution shown at the bottom can be predicted as the
diffraction of waves having the de Broglie wavelength of the electrons (CC BY 4.0; OpenStax).
After de Broglie proposed the wave nature of matter, many physicists, including Schrödinger and Heisenberg, explored the
consequences. The idea quickly emerged that, because of its wave character, a particle’s trajectory and destination cannot be
precisely predicted for each particle individually. However, each particle goes to a definite place (Figure 1.9.1 ). After compiling
enough data, you get a distribution related to the particle’s wavelength and diffraction pattern. There is a certain probability of
finding the particle at a given location, and the overall pattern is called a probability distribution. Those who developed quantum
mechanics devised equations that predicted the probability distribution in various circumstances.
It is somewhat disquieting to think that you cannot predict exactly where an individual particle will go, or even follow it to its
destination. Let us explore what happens if we try to follow a particle. Consider the double-slit patterns obtained for electrons and

11.6.1 https://chem.libretexts.org/@go/page/41373
photons in Figure 1.9.2 . The interferrence patterns build up statistically as individual particles fall on the detector. This can be
observed for photons or electrons—for now, let us concentrate on electrons. You might imagine that the electrons are interfering
with one another as any waves do. To test this, you can lower the intensity until there is never more than one electron between the
slits and the screen. The same interference pattern builds up!
This implies that a particle’s probability distribution spans both slits, and the particles actually interfere with themselves. Does this
also mean that the electron goes through both slits? An electron is a basic unit of matter that is not divisible. But it is a fair
question, and so we should look to see if the electron traverses one slit or the other, or both. One possibility is to have coils around
the slits that detect charges moving through them. What is observed is that an electron always goes through one slit or the other; it
does not split to go through both.
But there is a catch. If you determine that the electron went through one of the slits, you no longer get a double slit pattern—
instead, you get single slit interference. There is no escape by using another method of determining which slit the electron went
through. Knowing the particle went through one slit forces a single-slit pattern. If you do not observe which slit the electron goes
through, you obtain a double-slit pattern. How does knowing which slit the electron passed through change the pattern? The answer
is fundamentally important—measurement affects the system being observed. Information can be lost, and in some cases it is
impossible to measure two physical quantities simultaneously to exact precision. For example, you can measure the position of a
moving electron by scattering light or other electrons from it. Those probes have momentum themselves, and by scattering from the
electron, they change its momentum in a manner that loses information. There is a limit to absolute knowledge, even in principle.

Heisenberg’s Uncertainty Principle


It is mathematically possible to express the uncertainty that, Heisenberg concluded, always exists if one attempts to measure the
momentum and position of particles. First, we must define the variable “x” as the position of the particle, and define “p” as the
momentum of the particle. The momentum of a photon of light is known to simply be its frequency, expressed by the ratio h/λ,
where h represents Planck’s constant and λ represents the wavelength of the photon. The position of a photon of light is simply its
wavelength (λ ). To represent finite change in quantities, the Greek uppercase letter delta, or Δ, is placed in front of the quantity.
Therefore,
h
Δp = (11.6.1)
λ

Δx = λ (11.6.2)

By substituting Δx for λ into Equation 11.6.1, we derive


h
Δp = (11.6.3)
Δx

or,
ΔpΔx = h (11.6.4)
early form of uncertainty principle

 A Common Trend in Quantum Systems

Equation 11.6.4can be derived by assuming the particle of interest is behaving as a particle, and not as a wave. Simply let
Δp = mv , and Δx = h/(mv) (from De Broglie’s expression for the wavelength of a particle). Substituting in Δp for mv in
the second equation leads to Equation 11.6.4.

Equation 11.6.4 was further refined by Heisenberg and his colleague Niels Bohr, and was eventually rewritten as
h ℏ
Δpx Δx ≥ = (11.6.5)
4π 2

h
with ℏ = = 1.0545718 × 10
−34
m
2
⋅ kg/s .

Equation 11.6.5 reveals that the more accurately a particle’s position is known (the smaller Δx is), the less accurately the
momentum of the particle in the x direction (Δp ) is known. Mathematically, this occurs because the smaller Δx becomes, the
x

11.6.2 https://chem.libretexts.org/@go/page/41373
larger Δp must become in order to satisfy the inequality. However, the more accurately momentum is known the less accurately
x

position is known (Figure 1.9.2 ).

Figure 1.9.2 : The animation shows the relevant spreads in the uncertainty for position and momentum of light/photons (light
wave's corresponding photon particle). From the result of de Broglie, we know that for a particle with known momentum, p will
have a precise value for its de Broglie wavelength can be determined (and hence a specific color of the light).

 What is the Proper Definition of Uncertainty?

Equation 11.6.5 relates the uncertainty of momentum and position. An immediate questions that arise is if Δx represents the
full range of possible x values or if it is half (e.g., ⟨x⟩ ± Δx ). Δx is the standard deviation and is a statistic measure of the
spread of x values. The use of half the possible range is more accurate estimate of Δx. As we will demonstrated later, once we
construct a wavefunction to describe the system, then both x and Δx can be explicitly derived. However for now, Equation
11.6.5 will work.

For example: If a problem argues a particle is trapped in a box of length, L, then the uncertainly of it position is ±L/2. So the
value of Δx used in Equation 11.6.5 should be L/2, not L.

 Example 1.9.1

An electron is confined to the size of a magnesium atom with a 150 pm radius. What is the minimum uncertainty in its
velocity?

Solution
The uncertainty principle (Equation 11.6.5):

ΔpΔx ≥
2

can be written

Δp ≥
2Δx

and substituting Δp = mΔv since the mass is not uncertain.



Δv ≥
2 m Δx

the relevant parameters are


mass of electron m = m = 9.109383 × 10
e
−31
kg

uncertainty in position: Δx = 150 × 10 m −12

−34 2
1.0545718 × 10 kg m /s
Δv ≥
−31 −12
(2) (9.109383 × 10 kg ) (150 × 10 m )

5
= 3.9 × 10 m/s

11.6.3 https://chem.libretexts.org/@go/page/41373
 Exercise 1.9.1
What is the maximum uncertainty of velocity the electron described in Example 1.9.1 ?

Answer
Infinity. There is no limit in the maximum uncertainty, just the minimum uncertainty.

Understanding the Uncertainty Principle through Wave Packets and the Slit Experiment
It is hard for most people to accept the uncertainty principle, because in classical physics the velocity and position of an object can
be calculated with certainty and accuracy. However, in quantum mechanics, the wave-particle duality of electrons does not allow us
to accurately calculate both the momentum and position because the wave is not in one exact location but is spread out over space.
A "wave packet" can be used to demonstrate how either the momentum or position of a particle can be precisely calculated, but not
both of them simultaneously. An accumulation of waves of varying wavelengths can be combined to create an average wavelength
through an interference pattern: this average wavelength is called the "wave packet". The more waves that are combined in the
"wave packet", the more precise the position of the particle becomes and the more uncertain the momentum becomes because more
wavelengths of varying momenta are added. Conversely, if we want a more precise momentum, we would add less wavelengths to
the "wave packet" and then the position would become more uncertain. Therefore, there is no way to find both the position and
momentum of a particle simultaneously.

Quantum Mechanics The Uncertainty Pr…


Pr…

Several scientists have debated the Uncertainty Principle, including Einstein. Einstein created a slit experiment to try and disprove
the Uncertainty Principle. He had light passing through a slit, which causes an uncertainty of momentum because the light behaves
like a particle and a wave as it passes through the slit. Therefore, the momentum is unknown, but the initial position of the particle
is known. Here is a video that demonstrates particles of light passing through a slit and as the slit becomes smaller, the final
possible array of directions of the particles becomes wider. As the position of the particle becomes more precise when the slit is
narrowed, the direction, or therefore the momentum, of the particle becomes less known as seen by a wider horizontal distribution
of the light.

 Example 1.9.2

The speed of a 1.0 g projectile is known to within 10 −6


m/s.
a. Calculate the minimum uncertainty in its position.
b. What is the maximum uncertainty of its position?

Solution
a
From Equation 11.6.5, the Δp x = mΔvx with m = 1.0 g. Solving for Δx to get

11.6.4 https://chem.libretexts.org/@go/page/41373

Δx =
2mΔv

−34 2
1.0545718 × 10 m ⋅ kg/s
=
−6
(2)(0.001 kg)(10 m/s)

−26
= 5.3 × 10 m

This negligible for all intents and purpose as expected for any macroscopic object.
b
Unlimited (or the size of the universe). The Heisenberg uncertainty principles does not quantify the maximum uncertainty.

 Exercise 1.9.2

Estimate the minimum uncertainty in the speed of an electron confined to a hydrogen atom within a diameter of 1 × 10 −10
m ?

Answer
We need to quantify the uncertainty of the electron in position. We can estimate that as ±5 × 10 −10
m . Hence, substituting
the relavant numbers into Equation 11.6.5and solving for Δv we get
6
Δv = 1.15 × 10 km/s

Notice that the uncertainty is significantly greater for the electron in a hydrogen atom than in the magnesium atom
(Example 1.9.1 ) as expected since the magnesium atom is appreciably bigger.

Heisenberg’s Uncertainty Principle not only helped shape the new school of thought known today as quantum mechanics, but it
also helped discredit older theories. Most importantly, the Heisenberg Uncertainty Principle made it obvious that there was a
fundamental error in the Bohr model of the atom. Since the position and momentum of a particle cannot be known simultaneously,
Bohr’s theory that the electron traveled in a circular path of a fixed radius orbiting the nucleus was obsolete. Furthermore,
Heisenberg’s uncertainty principle, when combined with other revolutionary theories in quantum mechanics, helped shape wave
mechanics and the current scientific understanding of the atom.

 Humor: Heisenberg and the Police


Heisenberg get pulled over for speeding by the police. The officer asks him "Do you know how fast you were going?"
Heisenberg replies, "No, but we know exactly where we are!"
The officer looks at him confused and says "you were going 108 miles per hour!"
Heisenberg throws his arms up and cries, "Great! Now we're lost!"

11.6: The Heisenberg Uncertainty Principle is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.
1.9: The Heisenberg Uncertainty Principle by Kris Baumgartner, Sarah Woods is licensed CC BY 4.0.

11.6.5 https://chem.libretexts.org/@go/page/41373
11.7: The Schrödinger Wave Equation
Beginning in the early 20th century, physicists began to acknowledge that matter--much like electromagnetic radiation--possessed
wave-like behaviors. While electromagnetic radiation were well understood to obey Maxwell's Equations, matter obeyed no known
equations.

Introduction
In 1926, the Austrian physicist Erwin Schrödinger formulated what came to be known as the (time-dependent) Schrödinger
Equation:
∂ −ℏ
2
iℏ Ψ(x, t) = ∇ Ψ(x, t) + V (x)Ψ(x, t) (11.7.1)
∂t 2m

Equation 11.7.1 effectively describes matter as a wave that fluctuates with both displacement and time. Since the imaginary portion
of the equation dictates its time dependence, it is sufficed to say that for most purposes it can be treated as time-independent. The
result is seen in Equation 11.7.2:
2 2
−ℏ d ψ(x)
+ V (x)ψ(x) = Eψ(x) (11.7.2)
2
2m dx

Although this time-independent Schrödinger Equation can be useful to describe a matter wave in free space, we are most interested
in waves when confined to a small region, such as an electron confined in a small region around the nucleus of an atom. Several
different models have been developed that provide a means by which to study a matter-wave when confined to a small region: the
particle in a box (infinite well), finite well, and the Hydrogen atom. We will discuss each of these in order to develop a greater
understanding for how a wave behaves when it is in a bound state.

Figure 11.7.1 : Two unacceptable wavefunctions. (left) This is not a single-valued function and (right) this is a non-continuous
function
There are four general aspects that are applicable to an acceptable wavefunction
1. An acceptable wavefunction will be the solution of the Schrödinger equation (either Equations 11.7.1 or 11.7.2).
2. An acceptable wavefunction must be normalizable so will approaches zero as position approaches infinity.
3. An acceptable wavefunction must be a continuous function of position.
4. An acceptable wavefunction will have a continuous slope (Figure 11.7.1).

11.7.1 https://chem.libretexts.org/@go/page/41374
Figure 11.7.2 : Wavefunction continuity in space

Interpretation of the Wavefunction


Since wavefunctions can in general be complex functions, the physical significance cannot be found from the function itself
−−−
because the √−1 is not a property of the physical world. Rather, the physical significance is found in the product of the
wavefunction and its complex conjugate, i.e. the absolute square of the wavefunction, which also is called the square of the
modulus.
∗ 2
ψ (r, t)ψ(r, t) = |ψ(r, t)| (11.7.3)

where r is a vector (x, y, z) specifying a point in three-dimensional space. The square is used, rather than the modulus itself, just
like the intensity of a light wave depends on the square of the electric field. At one time it was thought that for an electron
described by the wavefunction ψ(r), the quantity eψ (r i)ψ(r )dτ was the amount of charge to be found in the volume dτ located

i

at r . However, Max Born found this interpretation to be inconsistent with the results of experiments.
i

The Born interpretation, which generally is accepted today, is that ψ (r )ψ(r ) dτ is the probability that the electron is in the

i i

volume dτ located at r . The Born interpretation therefore calls the wavefunction the probability amplitude, the absolute square of
i

the wavefunction is called the probability density, and the probability density times a volume element in three-dimensional space (
dτ ) is the probability. The idea that we can understand the world of atoms and molecules only in terms of probabilities is disturbing

to some, who are seeking more satisfying descriptions through ongoing research.

Normalization of the Wavefunction


A probability is a real number between 0 and 1. An outcome of a measurement which has a probability 0 is an impossible outcome,
whereas an outcome which has a probability 1 is a certain outcome. The probability of a measurement of x yielding a result
between −∞ and +∞ is

2
Px∈−∞:∞ (t) = ∫ |ψ(x, t)| dx. (11.7.4)
−∞

However, a measurement of x must yield a value between −∞ and +∞ , since the particle has to be located somewhere. It follows
that P (t) = 1 , or
x∈−∞:∞


2
∫ |ψ(x, t)| dx = 1, (11.7.5)
−∞

which is generally known as the normalization condition for the wavefunction.

11.7.2 https://chem.libretexts.org/@go/page/41374
Example 11.7.1: Normalizing a Gaussian wavefunction

Normalize the wavefunction of a Gaussian wave packet, centered on x = x , and of characteristic width σ: i.e., o

2
2
−(x−x0 ) /(4 σ )
ψ(x) = ψ0 e . (11.7.6)

Solution
To determine the normalization constant ψ , we simply substitute Equation 11.7.6 into Equation 11.7.5, to obtain
0


2 2
2 −(x−x0 ) /(2 σ )
| ψ0 | ∫ e dx = 1. (11.7.7)
−∞


Changing the variable of integration to y = (x − x 0 )/(
√2σ) , we get

2 – −y
2

| ψ0 | √2σ ∫ e dy = 1. (11.7.8)
−∞

However,

−y
2

∫ e dy = √π , (11.7.9)
−∞

which implies that


1
2
| ψ0 | = . (11.7.10)
2 1/2
(2πσ )

Hence, a general normalized Gaussian wavefunction takes the form



e 2
2
−(x−x0 ) /(4 σ )
ψ(x) = e (11.7.11)
2 1/4
(2πσ )

where ϕ is an arbitrary real phase-angle.

Contributors and Attributions


David M. Hanson, Erica Harvey, Robert Sweeney, Theresa Julia Zielinski ("Quantum States of Atoms and Molecules")

11.7: The Schrödinger Wave Equation is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

11.7.3 https://chem.libretexts.org/@go/page/41374
11.8: Particle in a One-Dimensional Box
A particle in a 1-dimensional box is a fundamental quantum mechanical approximation describing the translational motion of a
single particle confined inside an infinitely deep well from which it cannot escape.

Introduction
The particle in a box problem is a common application of a quantum mechanical model to a simplified system consisting of a
particle moving horizontally within an infinitely deep well from which it cannot escape. The solutions to the problem give possible
values of E and ψ that the particle can possess. E represents allowed energy values and ψ(x) is a wavefunction, which when
squared gives us the probability of locating the particle at a certain position within the box at a given energy level.

recipe for Quantum Mechanics


To solve the problem for a particle in a 1-dimensional box, we must follow our Big, Big recipe for Quantum Mechanics:
1. Define the Potential Energy, V
2. Solve the Schrödinger Equation
3. Solve for the wavefunctions
4. Solve for the allowed energies

Step 1: Define the Potential Energy V


The potential energy is 0 inside the box (V=0 for 0<x<L) and goes to infinity at the walls of the box (V=∞ for x<0 or x>L). We
assume the walls have infinite potential energy to ensure that the particle has zero probability of being at the walls or outside the
box. Doing so significantly simplifies our later mathematical calculations as we employ these boundary conditions when solving
the Schrödinger Equation.

Step 2: Solve the Schrödinger Equation


The time-independent Schrödinger equation for a particle of mass m moving in one direction with energy E is
2 2
ℏ d ψ(x)
− + V (x)ψ(x) = Eψ(x) (11.8.1)
2
2m dx

with
ℏ is the reduced Planck constant where ℏ = 2π
h

m is the mass of the particle

ψ(x) is the stationary time-independent wavefunction

V (x) is the potential energy as a function of position

E is the energy, a real number

This equation can be modified for a particle of mass m free to move parallel to the x-axis with zero potential energy (V = 0
everywhere) resulting in the quantum mechanical description of free motion in one dimension:
2 2
ℏ d ψ(x)
− = Eψ(x) (11.8.2)
2
2m dx

This equation has been well studied and gives a general solution of:
ψ(x) = A sin(kx) + B cos(kx) (11.8.3)

11.8.1 https://chem.libretexts.org/@go/page/41375
where A, B, and k are constants.

Step 3: Define the Wavefunction


The solution to the Schrödinger equation we found above is the general solution for a 1-dimensional system. We now need to apply
our boundary conditions to find the solution to our particular system. According to our boundary conditions, the probability of
finding the particle at x = 0 or x = L is zero. When x = 0 , then sin(0) = 0 and cos(0) = 1 ; therefore, B must equal 0 to fulfill
this boundary condition giving:

ψ(x) = A sin(kx) (11.8.4)

We can now solve for our constants (A and k ) systematically to define the wavefunction.
Solving for k
Differentiate the wavefunction with respect to x:

= kA cos(kx) (11.8.5)
dx

Differentiate the wavefunction algain with respect to x:


2
d ψ
2
= −k A sin(kx) (11.8.6)
2
dx

Since ψ(x) = A sin(kx) , then


2
d ψ
2
= −k ψ (11.8.7)
2
dx

If we then solve for k by comparing with the Schrödinger equation above, we find:
1/2
2
8 π mE
k =( ) (11.8.8)
h2

Now we plug k into our wavefunction (Equation 11.8.4):


1/2
2
8 π mE
ψ = A sin ( ) x (11.8.9)
2
h

Solving for A
To determine A, we have to apply the boundary conditions again. Recall that the probability of finding a particle at x =0 or
x = L is zero.

When x = L:
1/2
2
8 π mE
0 = A sin ( ) L (11.8.10)
2
h

This is only true when


1/2
2
8 π mE
( ) L = nπ (11.8.11)
2
h

where n = 1, 2, 3, …
Plugging this back in gives us:

ψ = A sin x (11.8.12)
L

To determine A , recall that the total probability of finding the particle inside the box is 1, meaning there is no probability of it
being outside the box. When we find the probability and set it equal to 1, we are normalizing the wavefunction.

11.8.2 https://chem.libretexts.org/@go/page/41375
L
2
∫ ψ dx = 1 (11.8.13)
0

For our system, the normalization looks like:


L

2 2
A ∫ sin ( )x dx = 1 (11.8.14)
0
L

Using the solution for this integral from an integral table, we find our normalization constant, A :
−−
2
A =√ (11.8.15)
L

Which results in the normalized wavefunctions for a particle in a 1-dimensional box:


−−
2 nπ
ψn = √ sin x (11.8.16)
L L

where n = 1, 2, 3, …

Step 4: Determine the Allowed Energies


Solving for the energy of each ψ requires substituting Equation 11.8.16 into Equation 11.8.2 to get the allowed energies for a
particle in a box:
2 2
n h
En = (11.8.17)
2
8mL

Equation 11.8.17 is a very important result and tells us that:


1. The energy of a particle is quantized.
2. The lowest possible energy of a particle is NOT zero. This is called the zero-point energy and means the particle can never be
at rest because it always has some kinetic energy.
This is also consistent with the Heisenberg Uncertainty Principle: if the particle had zero energy, we would know where it was in
both space and time.

What does all this mean?


The wavefunction for a particle in a box at the n = 1 and n = 2 energy levels look like this:

The probability of finding a particle a certain spot in the box is determined by squaring ψ . The probability distribution for a particle
in a box at the n = 1 and n = 2 energy levels looks like this:

11.8.3 https://chem.libretexts.org/@go/page/41375
Notice that the number of nodes (places where the particle has zero probability of being located) increases with increasing energy
n. Also note that as the energy of the particle becomes greater, the quantum mechanical model breaks down as the energy levels get
closer together and overlap, forming a continuum. This continuum means the particle is free and can have any energy value. At
such high energies, the classical mechanical model is applied as the particle behaves more like a continuous wave. Therefore, the
particle in a box problem is an example of Wave-Particle Duality.

Example 11.8.1

What is the ΔE between the n = 4 and n = 5 states for an F molecule trapped within in a one-dimension well of length 3.0
2

cm? At what value of n does the energy of the molecule reach ¼k T at 450 K, and what is the separation between this energy
B

level and the one immediately above it?


Solution
Since this is a one-dimensional particle in a box problem, the particle has only kinetic energy (V = 0), so the permitted energies
are:
2 2
n h
En =
2
8mL

with n = 1, 2, . . .
The energy difference between n = 4 and n = 5 is then
2 2 2 2
5 h 4 h
ΔE = E5 − E4 = −
8mL2 8mL2

Using Equation 11.8.17 with the mass of F2 (37.93 amu = 6.3 × 10


−26
kg ) and the length of the box (
L = 3 × 3.0 × 10
−2
m
2
:
2 −34 2 −1 2
9h 9(6.626 × 10 kg ⋅ m ⋅s )
ΔE = =
2 −26 −2 2 2
8mL 8(6.30938414 × 10 kg)(3.0 × 10 m )

−39
ΔE = 8.70 × 10 J

The n value for which the energy reaches 1

4
kB T :
2 2
n h 1
= kB T
2
8mL 4

9
n = 1.79 × 10

The separation between n + 1 and n :


2 2 2 2
(n + 1 ) h − (n) h
ΔE = En+1 − En =
2
8mL

2
(2n + 1)h
ΔE =
8mL2

−30
ΔE = 3.47 × 10 J

Important Facts to Learn from the Particle in the Box


The energy of a particle is quantized. This means it can only take on discreet energy values.
The lowest possible energy for a particle is NOT zero (even at 0 K). This means the particle always has some kinetic energy.
The square of the wavefunction is related to the probability of finding the particle in a specific position for a given energy level.
The probability changes with increasing energy of the particle and depends on the position in the box you are attempting to
define the energy for
In classical physics, the probability of finding the particle is independent of the energy and the same at all points in the box

11.8.4 https://chem.libretexts.org/@go/page/41375
Helpful Links
Provides a live quantum mechanical simulation of the particle in a box model and allows you to visualize the solutions to the
Schrödinger Equation: www.falstad.com/qm1d/

References
1. Chang, Raymond. Physical Chemistry for the Biosciences. Sansalito, CA: University Science, 2005.

11.8: Particle in a One-Dimensional Box is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

11.8.5 https://chem.libretexts.org/@go/page/41375
11.9: Quantum-Mechanical Tunneling
Tunneling is a quantum mechanical phenomenon when a particle is able to penetrate through a potential energy barrier that is
higher in energy than the particle’s kinetic energy. This amazing property of microscopic particles play important roles in
explaining several physical phenomena including radioactive decay. Additionally, the principle of tunneling leads to the
development of Scanning Tunneling Microscope (STM) which had a profound impact on chemical, biological and material science
research.

Violating Classical Mechanics


Consider a ball rolling from one valley to another over a hill (Figure 11.9.1). If the ball has enough energy (E ) to overcome the
potential energy (V ) at the top of the barrier between each valley, then it can roll from one valley to the other. This is the classical
picture and is controlled by the simple Law of Conservation of Energy approach taught in beginning physics courses. However, If
the ball does not have enough kinetic energy (E < V ), to overcome the barrier it will never roll from one valley to the other. In
contrast, when quantum effects are taken into effect, the ball can "tunnel" through the barrier to the other valley, even if its kinetic
energy is less than the potential energy of the barrier to the top of one of the hills.

Figure 11.9.1 : Contrasting classical (over the barrier) motion vs. quantum (through the barrier) motion. (CC BY-NC 4.0; Ümit
Kaya via LibreTexts)

 The Heisenberg Uncertainty Principle in Action


The reason for the difference between classical and quantum motion comes from wave-particle nature of matter. One
interpretation of this duality involves the Heisenberg uncertainty principle, which defines a limit on how precisely the position
and the momentum of a particle can be known at the same time. This implies that there are no solutions with a probability of
exactly zero (or one), though a solution may approach infinity if, for example, the calculation for its position was taken as a
probability of 1, the other, i.e. its speed, would have to be infinity. Hence, the probability of a given particle's existence on the
opposite side of an intervening barrier is non-zero, and such particles will appear on the 'other' (a semantically difficult word in
this instance) side with a relative frequency proportional to this probability.

Microscopic particles such as protons, or electrons would behave differently as a consequence of wave-particle duality. Consider a
particle with energy E that is confined in a box which has a barrier of height V . Classically, the box will prevent these particles
from escaping due to the insufficiency in kinetic energy of these particles to get over the barrier. However, if the thickness of the
barrier is thin, the particles have some probability of penetrating through the barrier without sufficient energy and appear on the
other side of the box (Figure 11.9.2).

11.9.1 https://chem.libretexts.org/@go/page/41376
Figure 11.9.2 : Quantum tunneling through a barrier. The energy of the tunneled particle is the same but the amplitude is decreased.
(CC BY-NC 4.0; Ümit Kaya via LibreTexts)
When it reaches a barrier it cannot overcome, a particle's wave function changes from sinusoidal to exponentially diminishing in
form. The solution for the Schrödinger equation in such a medium (Figure 11.9.2; blue region) is:
−βx
ψ = Ne (11.9.1)

where
N is a normalization constant and
−−−−−−−−−−
2m(V − E)
β =√
2

For a quantum particle to appreciably tunnel through a barrier three conditions must be met (Figure 11.9.2):
1. The height of the barrier must be finite and the thickness of the barrier should be thin.
2. The potential energy of the barrier exceeds the kinetic energy of the particle (E < V ).
3. The particle has wave properties because the wavefunction is able to penetrate through the barrier. This suggests that quantum
tunneling only apply to microscopic objects such protons or electrons and does not apply to macroscopic objects.
If these conditions are met, there would be some probability of finding the particles on the other side of the barrier. Beginning as a
sinusoidal wave, a particle begins tunneling through the barrier and goes into exponential decay until it exits the barrier and gets
transmitted out the other side as a final sinusoidal wave with a smaller amplitude. The act of tunneling decreases the wave
amplitude due the reflection of the incident wave when it comes into the contact with the barrier but does not affect the wave
equation.

11.9.2 https://chem.libretexts.org/@go/page/41376
Figure 11.9.3 : Quantum tunnelling of alion through a barrier is a quantum effect with no classical analog. (CC BY-NC 4.0; Ümit
Kaya via LibreTexts)
The probability, P , of a particle tunneling through the potential energy barrier is derived from the Schrödinger Equation and is
described as,
−4aπ −−−−−−−−−
P = exp( √ 2m(V − E) ) (11.9.2)
h

with E < V
where
V is the potential barrier,
E is the kinetic energy possessed by the particle, and
a is the thickness of the barrier.

m is mass of the particle

h is Planks Constant (6.6260 × 10 m kg/s )


−34 2

Therefore, the probability of an object tunneling through a barrier decreases with the object's increasing mass and with the
increasing gap between the energy of the object and the energy of the barrier. And although the wave function never quite reaches 0
(as can be determined from the functionality), this explains how tunneling is frequent on nanoscale, but negligible at the
macroscopic level. Equation 11.9.2 argues that the probability of tunneling decreases exponentially with the square root of the
particles mass (particles with a small mass can effectively tunnel through barriers more easily than those with larger mass) and the
thickness of the barrier (a )

 Example 11.9.1

An electron having total kinetic energy E of 4.50 eV approaches a rectangular energy barrier with V = 5.00 eV and
L = 950 pm. Classically, the electron cannot pass through the barrier because E < V . Calculate probability of tunneling of

this electron through the barrier.

Solution
This is a straightforward application of Equation 11.9.2.

11.9.3 https://chem.libretexts.org/@go/page/41376
−4aπ −−−−−−−−−
P = exp( √ 2m(V − E) )
h

The electronvolt (eV) is a unit of energy that is equal to approximately 1.6 × 10 −19
J , which is the conversion used below.
The mass of an electron is 9.10 × 10 −31
kg

−12
(4)(950 × 10 m)(π) −−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−
−31 −19
P = exp[(− ) √ (2)(9.10 × 10 kg)(5.00 − 4.50 eV )(1.60 × 10 J/ eV )]
−34
6.6260 × 10 m2 kg/s

−6.88 −3
= exp = 1.03 × 10

There is a ~0.1% probability of the electrons tunneling though the barrier.

Alpha Decay Radioactivity


Protons and neutrons in a nucleus have kinetic energy, but it is about 8 MeV less than that needed to get out from attractive nuclear
potential (Figure 11.9.4). Hence, they are bound by an average of 8 MeV per nucleon. The slope of the hill outside the bowl is
analogous to the repulsive Coulomb potential for a nucleus, such as for an α particle outside a positive nucleus. In α decay, two
protons and two neutrons spontaneously break away as a 4He unit. Yet the protons and neutrons do not have enough kinetic energy
to classically get over the rim.

Figure 11.9.4 : Nucleons within an atomic nucleus are bound or trapped by the attractive nuclear force, as shown in this simplified
potential energy curve. An particle outside the range of the nuclear force feels the repulsive Coulomb force. The α particle inside
the nucleus does not have enough kinetic energy to get over the rim, yet it does manage to get out by quantum mechanical
tunneling. (CC BY; OpenStax)
The α article tunnels through a region of space it is forbidden to be in, and it comes out of the side of the nucleus. Like an electron
making a transition between orbits around an atom, it travels from one point to another without ever having been in between
(Figure 11.9.5). The wave function of a quantum mechanical particle varies smoothly, going from within an atomic nucleus (on
one side of a potential energy barrier) to outside the nucleus (on the other side of the potential energy barrier). Inside the barrier, the
wave function does not become zero but decreases exponentially, and we do not observe the particle inside the barrier. The
probability of finding a particle is related to the square of its wave function, and so there is a small probability of finding the
particle outside the barrier, which implies that the particle can tunnel through the barrier.

Figure 11.9.5 : The wave function representing a quantum mechanical particle must vary smoothly, going from within the nucleus
(to the left of the barrier) to outside the nucleus (to the right of the barrier). Inside the barrier, the wave function does not abruptly
become zero; rather, it decreases exponentially. Outside the barrier, the wave function is small but finite, and there it smoothly
becomes sinusoidal. Owing to the fact that there is a small probability of finding the particle outside the barrier, the particle can
tunnel through the barrier. (CC BY; OpenStax)

11.9.4 https://chem.libretexts.org/@go/page/41376
Scanning Tunneling Microscopy (STM)
A metal tip usually made out of tungsten is placed between a very small distance above a conducting or semiconducting surface.
This distance acts as a potential barrier for tunneling. The space between the tip and the surface normally is vacuum. When
electrons tunnel from the metal tip to the surface, a current is created and monitored by a computer (Figure 11.9.6). The current
depends on the distance between the tip and the surface, which is controlled by a piezoelectric cylinder. If there is a strong current,
the tip will move away from the surface. The increase of the potential barrier will decrease the probability of tunneling and
decrease the current. If the current becomes too weak, the tip moves closer to the surface. The potential barrier will be reduced and
the current will increase. The variations in the current as the tip moves over the sample are reconstructed by the computer to
produce topological image of the scanned surface.

Figure 11.9.6 : Schematic diagram of a scanning tunneling microscope. (CC BY-SA 2.0 Austria; Michael Schmid via Wikipedia).

References
1. Chang, Raymond. Physical Chemistry for the Biosciences. Sansalito, CA: University Science, 2005. Print.
2. Engel, Thomas. Quantum Chemistry and Spectroscopy.Upper Saddle River, NJ. Pearson, 2006. Print.

Contributors and Attributions


Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
2bd...a7ac8df6@9.110).

11.9: Quantum-Mechanical Tunneling is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
Tunneling is licensed CC BY 4.0.

11.9.5 https://chem.libretexts.org/@go/page/41376
11.10: The Schrödinger Wave Equation for the Hydrogen Atom
The hydrogen atom, consisting of an electron and a proton, is a two-particle system, and the internal motion of two particles around their
center of mass is equivalent to the motion of a single particle with a reduced mass. This reduced particle is located at r, where r is the vector
specifying the position of the electron relative to the position of the proton. The length of r is the distance between the proton and the electron,
and the direction of r and the direction of r is given by the orientation of the vector pointing from the proton to the electron. Since the proton
is much more massive than the electron, we will assume throughout this chapter that the reduced mass equals the electron mass and the proton
is located at the center of mass.

Figure 11.10.1: a) The proton (p ) and electron (e ) of the hydrogen atom. b) Equivalent reduced particle with reduced mass μ at distance r
+ −

from center of mass.


Since the internal motion of any two-particle system can be represented by the motion of a single particle with a reduced mass, the description
of the hydrogen atom has much in common with the description of a diatomic molecule discussed previously. The Schrödinger Equation for
the hydrogen atom
^
H (r, θ, φ)ψ(r, θ, φ) = Eψ(r, θ, φ) (11.10.1)

employs the same kinetic energy operator, T^, written in spherical coordinates. For the hydrogen atom, however, the distance, r, between the
two particles can vary, unlike the diatomic molecule where the bond length was fixed, and the rigid rotor model was used. The hydrogen atom
Hamiltonian also contains a potential energy term, V^, to describe the attraction between the proton and the electron. This term is the Coulomb
potential energy,
2
e
^
V (r) = − (11.10.2)
4π ϵ0 r

where r is the distance between the electron and the proton. The Coulomb potential energy depends inversely on the distance between the
electron and the nucleus and does not depend on any angles. Such a potential is called a central potential.
It is convenient to switch from Cartesian coordinates x, y, z to spherical coordinates in terms of a radius r, as well as angles ϕ , which is
measured from the positive x axis in the xy plane and may be between 0 and 2π, and θ , which is measured from the positive z axis towards the
xy plane and may be between 0 and π.
z
(r, θ, φ)

r
θ

y
φ

x
Figure 11.10.2: Spherical Coordinates
The time-indepdent Schrödinger equation (in spherical coordinates) for a electron around a positively charged nucleus is then
2 2 2
ℏ ∂ ∂ 1 ∂ ∂ 1 ∂ e
2
{− [ (r )+ (sin θ )+ ]− } ψ(r, θ, φ) = Eψ(r, θ, φ) (11.10.3)
2 2 2
2μr ∂r ∂r sin θ ∂θ ∂θ sin θ ∂φ 4π ϵ0 r

Since the angular momentum operator does not involve the radial variable, r, we can separate variables in Equation 11.10.3 by using a
product wavefunction. We know that the eigenfunctions of the angular momentum operator are the Spherical Harmonic functions (Table
M4), Y (θ, φ) , so a good choice for a product function is

ψ(r, θ, φ) = R(r)Y (θ, φ) (11.10.4)

The Spherical Harmonic Y (θ, φ) functions provide information about where the electron is around the proton, and the radial function R(r)
describes how far the electron is away from the proton. A solution for both R(r) and Y (θ, φ) with E that depends on only one quantum n

11.10.1 https://chem.libretexts.org/@go/page/41377
number n , although others are required for the proper description of the wavefunction:
4
me e
En = − (11.10.5)
2
8ϵ h n2 2
0

with n = 1, 2, 3. . . ∞
The hydrogen atom wavefunctions, ψ(r, θ, ϕ), are called atomic orbitals. An atomic orbital is a function that describes one electron in an
atom. The wavefunction with n = 1, l l = 0 is called the 1s orbital, and an electron that is described by this function is said to be “in” the ls
orbital, i.e. have a 1s orbital state. The constraints on n, l l), and m that are imposed during the solution of the hydrogen atom Schrödinger
l

equation explain why there is a single 1s orbital, why there are three 2p orbitals, five 3d orbitals, etc. We will see when we consider multi-
electron atoms, these constraints explain the features of the Periodic Table. In other words, the Periodic Table is a manifestation of the
Schrödinger model and the physical constraints imposed to obtain the solutions to the Schrödinger equation for the hydrogen atom.

The Three Quantum Numbers


Schrödinger’s approach requires three quantum numbers (n , l, and m ) to specify a wavefunction for the electron. The quantum numbers
l

provide information about the spatial distribution of an electron. Although n can be any positive integer (NOT zero), only certain values of l
and m are allowed for a given value of \(n).
l

The principal quantum number (n): One of three quantum numbers that tells the average relative distance of an electron from the nucleus.
indicates the energy of the electron and the average distance of an electron from the nucleus
n = 1, 2, 3, 4, . . . (11.10.6)

Asn increases for a given atom, so does the average distance of an electron from the nucleus. A negatively charged electron that is, on
average, closer to the positively charged nucleus is attracted to the nucleus more strongly than an electron that is farther out in space. This
means that electrons with higher values of n are easier to remove from an atom. All wave functions that have the same value of n are said to
constitute a principal shell. All the wave functions that have the same value of n because those electrons have similar average distances from
the nucleus. because those electrons have similar average distances from the nucleus. As you will see, the principal quantum number n
corresponds to the n used by Bohr to describe electron orbits and by Rydberg to describe atomic energy levels.
The Azimuthal Quantum Number: The second quantum number is often called the azimuthal quantum number (l). One of three quantum
numbers that describes the shape of the region of space occupied by an electron.. The value of l describes the shape of the region of space
occupied by the electron. The allowed values of l depend on the value of n and can range from 0 to n − 1:

l = 0, 1, . , 2, 3, . . (n − 1) (11.10.7)

For example, if n = 1, l can be only 0; if n = 2, l can be 0 or 1; and so forth. For a given atom, all wave functions that have the same values of
both n and l form a subshell. A group of wave functions that have the same values of n and l. The regions of space occupied by electrons in
the same subshell usually have the same shape, but they are oriented differently in space.
The Magnetic Quantum Number: The third quantum number is the magnetic quantum number (m ). One of three quantum numbers that
l

describes the orientation of the region of space occupied by an electron with respect to an applied magnetic field.. The value of m describes
l

the orientation of the region in space occupied by an electron with respect to an applied magnetic field. The allowed values of m depend on
l

the value of l: m can range from −l to l in integral steps:


l

m = −l, −l + 1, . . .0, . . . l − 1, l (11.10.8)

For example, if l = 0 , ml can be only 0; if l = 1, m can be −1, 0, or +1; and if l = 2, m can be −2, −1, 0, +1, or +2.
l l

Each wave function with an allowed combination of n, l, and ml values describes an atomic orbital A wave function with an allowed
combination of n, l and ml quantum numbers., a particular spatial distribution for an electron. For a given set of quantum numbers, each
principal shell has a fixed number of subshells, and each subshell has a fixed number of orbitals.

Example 11.10.1: Number of subshells

How many subshells and orbitals are contained within the principal shell with n = 4?
Given: value of n
Asked for: number of subshells and orbitals in the principal shell
Strategy:
A Given n = 4, calculate the allowed values of l. From these allowed values, count the number of subshells.

11.10.2 https://chem.libretexts.org/@go/page/41377
B For each allowed value of l, calculate the allowed values of ml. The sum of the number of orbitals in each subshell is the number of
orbitals in the principal shell.
Solution
A We know that l can have all integral values from 0 to n − 1. If n = 4, then l can equal 0, 1, 2, or 3. Because the shell has four values of l,
it has four subshells, each of which will contain a different number of orbitals, depending on the allowed values of ml.
B For l = 0, ml can be only 0, and thus the l = 0 subshell has only one orbital. For l = 1, ml can be 0 or ±1; thus the l = 1 subshell has three
orbitals. For l = 2, ml can be 0, ±1, or ±2, so there are five orbitals in the l = 2 subshell. The last allowed value of l is l = 3, for which ml
can be 0, ±1, ±2, or ±3, resulting in seven orbitals in the l = 3 subshell. The total number of orbitals in the n = 4 principal shell is the sum
of the number of orbitals in each subshell and is equal to n2:
2
1 + 3 + 5 + 7 = 16 orbitals = (4 principal shells)
(l=0) (l=1) (l=2) (l=3)

Exercise 11.10.1

How many subshells and orbitals are in the principal shell with n = 3?
Answer: three subshells; nine orbitals

Rather than specifying all the values of n and l every time we refer to a subshell or an orbital, chemists use an abbreviated system with
lowercase letters to denote the value of l for a particular subshell or orbital:

l= 0 1 2 3

Designation s p d f

The principal quantum number is named first, followed by the letter s, p, d, or f as appropriate. These orbital designations are derived from
corresponding spectroscopic characteristics of lines involving them: sharp, principle, diffuse, and fundamental. A 1s orbital has n = 1 and l =
0; a 2p subshell has n = 2 and l = 1 (and has three 2p orbitals, corresponding to ml = −1, 0, and +1); a 3d subshell has n = 3 and l = 2 (and has
five 3d orbitals, corresponding to ml = −2, −1, 0, +1, and +2); and so forth.
We can summarize the relationships between the quantum numbers and the number of subshells and orbitals as follows (Table 11.10.1):
Each principal shell has n subshells. For n = 1, only a single subshell is possible (1s); for n = 2, there are two subshells (2s and 2p); for n =
3, there are three subshells (3s, 3p, and 3d); and so forth. Every shell has an ns subshell, any shell with n ≥ 2 also has an np subshell, and
any shell with n ≥ 3 also has an nd subshell. Because a 2d subshell would require both n = 2 and l = 2, which is not an allowed value of l
for n = 2, a 2d subshell does not exist.
Each subshell has 2l + 1 orbitals. This means that all ns subshells contain a single s orbital, all np subshells contain three p orbitals, all nd
subshells contain five d orbitals, and all nf subshells contain seven f orbitals.

Note

Each principal shell has n subshells, and each subshell has 2l + 1 orbitals.

Table 11.10.1: Allowed values of n , l, and m through n = 4


l

11.10.3 https://chem.libretexts.org/@go/page/41377
N
u
N
m
u
b
m
e
b
r
e
o
r
f
o
O
f
r
O
b
r
i
b
tm
n l Subshell Designation i
al
t
l
a
s
l
i
s
n
i
S
n
u
S
b
h
s
e
h
l
e
l
l
l

1 0 1s 10

0 2s 01


1
2 4,
1 2p 3
0
,
1

0 3s 01


1
,
1 3p 3
0
,
1


3 92
,

1
2 3d 5,
0
,
1
,
2

4 0 4s 10
6

1
,
1 4p 3
0
,
1

11.10.4 https://chem.libretexts.org/@go/page/41377

2
,

1
2 4d 5,
0
,
1
,
2


3
,

2
,

1
3 4f 7
,
0
,
1
,
2
,
3

The Radial Component


The first six radial functions are provided in Table 11.10.2. Note that the functions in the table exhibit a dependence on Z , the atomic number
of the nucleus. Other one electron systems have electronic states analogous to those for the hydrogen atom, and inclusion of the charge on the
nucleus allows the same wavefunctions to be used for all one-electron systems. For hydrogen, Z = 1 and for helium, Z = 2 .
Zr
Table 11.10.2: Hydrogen-like atomic wavefunctions for n values 1, 2, 3 : Z is the atomic number of the nucleus, and ρ = , where a is the Bohr
0
a0

radius and r is the radial variable.


n ℓ m Radial Component
3

1 Z 2

n = 1 ℓ = 0 m = 0 −ρ
ψ100 = ( ) e

√π a0

3 −ρ
1 Z 2
n = 2 ℓ = 0 m = 0 2
ψ200 = ( ) (2 − ρ)e
−−−
√32π a0

1 Z 2

ℓ = 1 m = 0 −ρ/2
ψ210 = ( ) ρe cos(θ)
−−−
√32π a0

ℓ = 1 m = ±1 1 Z 2 −ρ/2 ± iϕ
ψ21± 1 = ( ) ρe sin(θ)e
−−−
√64π a0

1 Z 2

n = 3 ℓ = 0 m = 0 2
ψ300 = ( ) (27 − 18ρ + 2 ρ )e
−−
81 √3π a0

3
−−
ℓ = 1 m = 0 1 2 Z 2 2 −ρ/3
ψ310 = √ ( ) (6r − ρ )e co
81 π a0

1 Z 2

ℓ = 1 m = ±1 2 −r/3
ψ31± = ( ) (6ρ − ρ )e sin
1 −
81 √π a0

1 Z 2

ℓ = 2 m = 0 2 −ρ/3 2
ψ320 = ( ) ρ e (3co s (θ)
−−
81 √6π a0

1 Z 2

ℓ = 2 m = ±1 2 −ρ/3
ψ32± = ( ) ρ e sin(θ) cos
1 −
81 √π a0

11.10.5 https://chem.libretexts.org/@go/page/41377
n ℓ m Radial Component
3

1 Z 2

ℓ = 2 m = ±2 2 −ρ/3 2 ±
ψ32± = ( ) ρ e sin (θ)e
2 −
162 √π a0

Visualizing the variation of an electronic wavefunction with r,θ , and ϕ is important because the absolute square of the wavefunction depicts
the charge distribution (electron probability density) in an atom or molecule. The charge distribution is central to chemistry because it is
related to chemical reactivity. For example, an electron deficient part of one molecule is attracted to an electron rich region of another
molecule, and such interactions play a major role in chemical interactions ranging from substitution and addition reactions to protein folding
and the interaction of substrates with enzymes.
Methods for separately examining the radial portions of atomic orbitals provide useful information about the distribution of charge density
within the orbitals. Graphs of the radial functions, R(r), for the 1s, 2s, and 2p orbitals plotted in Figure 11.10.2 left). The quantity R(r) R(r) ∗

gives the radial probability density; i.e., the probability density for the electron to be at a point located the distance r from the proton. Radial
probability densities for three types of atomic orbitals are plotted in Figure 11.10.2 (right).

11.10.6 https://chem.libretexts.org/@go/page/41377
Figure 11.10.2: (left) Radial function, R(r), for the 1s, 2s, and 2p orbitals. (right) Radial probability densities for the 1s, 2s, and 2p orbitals.
For the hydrogen atom, the peak in the radial probability plot occurs at r = 0.529 Å (52.9 pm), which is exactly the radius calculated by Bohr
for the n = 1 orbit. Thus the most probable radius obtained from quantum mechanics is identical to the radius calculated by classical
mechanics. In Bohr’s model, however, the electron was assumed to be at this distance 100% of the time, whereas in the Schrödinger model, it

11.10.7 https://chem.libretexts.org/@go/page/41377
is at this distance only some of the time. The difference between the two models is attributable to the wavelike behavior of the electron and the
Heisenberg uncertainty principle.
Figure 11.10.3 compares the electron probability densities for the hydrogen 1s, 2s, and 3s orbitals. Note that all three are spherically
symmetrical. For the 2s and 3s orbitals, however (and for all other s orbitals as well), the electron probability density does not fall off
smoothly with increasing r. Instead, a series of minima and maxima are observed in the radial probability plots (part (c) in Figure 11.10.3).
The minima correspond to spherical nodes (regions of zero electron probability), which alternate with spherical regions of nonzero electron
probability.

Figure 11.10.3: Probability Densities for the 1 s, 2 s, and 3s Orbitals of the Hydrogen Atom. (a) The electron probability density in any plane
that contains the nucleus is shown. Note the presence of circular regions, or nodes, where the probability density is zero. (b) Contour surfaces
enclose 90% of the electron probability, which illustrates the different sizes of the 1s, 2s, and 3s orbitals. The cutaway drawings give partial
views of the internal spherical nodes. The orange color corresponds to regions of space where the phase of the wave function is positive, and
the blue color corresponds to regions of space where the phase of the wave function is negative. (c) In these plots of electron probability as a
function of distance from the nucleus (r) in all directions (radial probability), the most probable radius increases asn increases, but the 2s and
3s orbitals have regions of significant electron probability at small values of r.

The Angular Component


The angular component of the wavefunction Y (θ, ϕ) in Equation 11.10.4 does much to give an orbital its distinctive shape. Y (θ, ϕ) is
typically normalized so the the integral of Y (θ, ϕ) over the unit sphere is equal to one. In this case, Y (θ, ϕ) serves as a probability
2 2

function. The probability function can be interpreted as the probability that the electron will be found on the ray emitting from the origin that
is at angles (θ, ϕ) from the axes. The probability function can also be interpreted as the probability distribution of the electron being at
position (θ, ϕ) on a sphere of radius r, given that it is r distance from the nucleus. Y (θ, ϕ) are also the wavefunction solutions to
l,ml

Schrödinger’s equation for a rigid rotor consisting of rotating bodies, for example a diatomic molecule. These are called Spherical Harmonic
functions (Table M4).

s Orbitals (l=0)
Three things happen to s orbitals as n increases (Figure 6.6.2):
1. They become larger, extending farther from the nucleus.
2. They contain more nodes. This is similar to a standing wave that has regions of significant amplitude separated by nodes, points with zero
amplitude.
3. For a given atom, the s orbitals also become higher in energy as n increases because of their increased distance from the nucleus.
Orbitals are generally drawn as three-dimensional surfaces that enclose 90% of the electron density. Although such drawings show the relative
sizes of the orbitals, they do not normally show the spherical nodes in the 2s and 3s orbitals because the spherical nodes lie inside the 90%
surface. Fortunately, the positions of the spherical nodes are not important for chemical bonding.

p Orbitals (l=1)
Only s orbitals are spherically symmetrical. As the value of l increases, the number of orbitals in a given subshell increases, and the shapes of
the orbitals become more complex. Because the 2p subshell has l = 1, with three values of ml (−1, 0, and +1), there are three 2p orbitals).

11.10.8 https://chem.libretexts.org/@go/page/41377
Figure 6.6.2, the colors correspond to regions of space where the phase of the wave function is positive (orange) and negative (blue).
The electron probability distribution for one of the hydrogen 2p orbitals is shown in Figure 11.10.4. Because this orbital has two lobes of
electron density arranged along the z axis, with an electron density of zero in the xy plane (i.e., the xy plane is a nodal plane), it is a 2pz orbital.
As shown in Figure 11.10.5, the other two 2p orbitals have identical shapes, but they lie along the x axis (2px) and y axis (2py), respectively.
Note that each p orbital has just one nodal plane. In each case, the phase of the wave function for each of the 2p orbitals is positive for the lobe
that points along the positive axis and negative for the lobe that points along the negative axis. It is important to emphasize that these signs
correspond to the phase of the wave that describes the electron motion, not to positive or negative charges.

Figure 11.10.5: The Three Equivalent 2 p Orbitals of the Hydrogen Atom


The surfaces shown enclose 90% of the total electron probability for the 2px, 2py, and 2pz orbitals. Each orbital is oriented along the axis
indicated by the subscript and a nodal plane that is perpendicular to that axis bisects each 2p orbital. The phase of the wave function is
positive (orange) in the region of space where x, y, or z is positive and negative (blue) where x, y, or z is negative. Just as with the s orbitals,
the size and complexity of the p orbitals for any atom increase as the principal quantum number n increases. The shapes of the 90%
probability surfaces of the 3p, 4p, and higher-energy p orbitals are, however, essentially the same as those shown in Figure Figure 11.10.5.

d Orbitals (l=2)
Subshells with l = 2 have five d orbitals; the first principal shell to have a d subshell corresponds to n = 3. The five d orbitals have ml values of
−2, −1, 0, +1, and +2 (Figure 11.10.6).

Figure 11.10.6: The Five Equivalent 3d Orbitals of the Hydrogen Atom. The surfaces shown enclose 90% of the total electron probability for
the five hydrogen 3d orbitals. Four of the five 3d orbitals consist of four lobes arranged in a plane that is intersected by two perpendicular
nodal planes. These four orbitals have the same shape but different orientations. The fifth 3d orbital, 3d , has a distinct shape even though it
z
2

is mathematically equivalent to the others. The phase of the wave function for the different lobes is indicated by color: orange for positive and
blue for negative.
The hydrogen 3d orbitals have more complex shapes than the 2p orbitals. All five 3d orbitals contain two nodal surfaces, as compared to one
for each p orbital and zero for each s orbital. In three of the d orbitals, the lobes of electron density are oriented between the x and y, x and z,
and y and z planes; these orbitals are referred to as the 3d , \)3d_{xz}\_, and \)3d_{yz}\) orbitals, respectively. A fourth d orbital has lobes
xy

11.10.9 https://chem.libretexts.org/@go/page/41377
lying along the x and y axes; this is the 3dx2−y2 orbital. The fifth 3d orbital, called the 3d orbital, has a unique shape: it looks like a 2p
z
2
z

orbital combined with an additional doughnut of electron probability lying in the xy plane. Despite its peculiar shape, the 3d orbital is
z
2

mathematically equivalent to the other four and has the same energy. In contrast to p orbitals, the phase of the wave function for d orbitals is
the same for opposite pairs of lobes. As shown in Figure 11.10.6, the phase of the wave function is positive for the two lobes of the dz 2

orbital that lie along the z axis, whereas the phase of the wave function is negative for the doughnut of electron density in the xy plane. Like
the s and p orbitals, as n increases, the size of the d orbitals increases, but the overall shapes remain similar to those depicted in Figure 6.6.5.

f Orbitals (l=3)
Principal shells with n = 4 can have subshells with l = 3 and ml values of −3, −2, −1, 0, +1, +2, and +3. These subshells consist of seven f
orbitals. Each f orbital has three nodal surfaces, so their shapes are complex (not shown).

Energies
The constraint that n be greater than or equal to l + 1 also turns out to quantize the energy, producing the same quantized expression for
hydrogen atom energy levels that was obtained from the Bohr model of the hydrogen atom.
2
Z
E =− Rhc (11.10.9)
2
n

or
2 4
Z μe
En = − (11.10.10)
2 2 2
8ϵ h n
0

The relative energies of the atomic orbitals with n ≤ 4 for a hydrogen atom are plotted in Figure 11.10.7 ; note that the orbital energies depend
on only the principal quantum number n. Consequently, the energies of the 2s and 2p orbitals of hydrogen are the same; the energies of the 3s,
3p, and 3d orbitals are the same; and so forth. The orbital energies obtained for hydrogen using quantum mechanics are exactly the same as
the allowed energies calculated by Bohr. In contrast to Bohr’s model, however, which allowed only one orbit for each energy level, quantum
mechanics predicts that there are 4 orbitals with different electron density distributions in the n = 2 principal shell (one 2s and three 2p
orbitals), 9 in the n = 3 principal shell, and 16 in the n = 4 principal shell.

Note
For a single electron system, the energy of that electron is a function of only the principal quantum number (Equation 11.10.9).

The different values of l and ml for the individual orbitals within a given principal shell are not important for understanding the emission or
absorption spectra of the hydrogen atom under most conditions, but they do explain the splittings of the main lines that are observed when
hydrogen atoms are placed in a magnetic field. As we have just seen, however, quantum mechanics also predicts that in the hydrogen atom, all
orbitals with the same value of n (e.g., the three 2p orbitals) are degenerate, meaning that they have the same energy. Figure 11.10.7 shows
that the energy levels become closer and closer together as the value of n increases, as expected because of the 1/n2 dependence of orbital
energies.

Figure 11.10.7: Orbital Energy Level Diagram for the Hydrogen Atom. Each box corresponds to one orbital. Note that the difference in
energy between orbitals decreases rapidly with increasing values of n.
In general, both energy and radius decrease as the nuclear charge increases. Thus the most stable orbitals (those with the lowest energy) are
those closest to the nucleus. For example, in the ground state of the hydrogen atom, the single electron is in the 1s orbital, whereas in the first

11.10.10 https://chem.libretexts.org/@go/page/41377
excited state, the atom has absorbed energy and the electron has been promoted to one of the n = 2 orbitals. In ions with only a single electron,
the energy of a given orbital depends on only n, and all subshells within a principal shell, such as the px, py, and pz orbitals, are degenerate.

Note: Bohr vs. Schrödinger


It is interesting to compare the results obtained by solving the Schrödinger equation with Bohr’s model of the hydrogen atom. There are
several ways in which the Schrödinger model and Bohr model differ.
1. First, and perhaps most strikingly, the Schrödinger model does not produce well-defined orbits for the electron. The wavefunctions
only give us the probability for the electron to be at various directions and distances from the proton.
2. Second, the quantization of angular momentum is different from that proposed by Bohr. Bohr proposed that the angular momentum is
1

quantized in integer units of ℏ , while the Schrödinger model leads to an angular momentum of (l(l + 1)ℏ ) 2 .
2

3. Third, the quantum numbers appear naturally during solution of the Schrödinger equation while Bohr had to postulate the existence of
quantized energy states. Although more complex, the Schrödinger model leads to a better correspondence between theory and
experiment over a range of applications that was not possible for the Bohr model.

Electron Spin: The Fourth Quantum Number


The quantum numbers n, l, m are not sufficient to fully characterize the physical state of the electrons in an atom. In 1926, Otto Stern and
Walther Gerlach carried out an experiment that could not be explained in terms of the three quantum numbers n, l, m and showed that there
is, in fact, another quantum-mechanical degree of freedom that needs to be included in the theory. The experiment is illustrated in the Figure
11.10.8. A beam of atoms (e.g. hydrogen or silver atoms) is sent through a spatially inhomogeneous magnetic field with a definite field

gradient toward one of the poles. It is observed that the beam splits into two beams as it passes through the field region.

Figure 11.10.8: The Stern-Gerlach apparatus. from Wikipedia


The fact that the beam splits into 2 beams suggests that the electrons in the atoms have a degree of freedom capable of coupling to the
magnetic field. That is, an electron has an intrinsic magnetic moment M arising from a degree of freedom that has no classical analog. The
magnetic moment must take on only 2 values according to the Stern-Gerlach experiment. The intrinsic property that gives rise to the magnetic
moment must have some analog to a spin, S ; unlike position and momentum, which have clear classical analogs, spin does not. The
implication of the Stern-Gerlach experiment is that we need to include a fourth quantum number, m in our description of the physical state of
s

the electron. That is, in addition to give its principle, angular, and magnetic quantum numbers, we also need to say if it is a spin-up electron or
a spin-down electron.

Note: Spin is a Quantum Phenomenon


Unlike position and momentum, which have clear classical analogs, spin does not.

George Uhlenbeck (1900–1988) and Samuel Goudsmit (1902–1978), proposed that the splittings were caused by an electron spinning about
its axis, much as Earth spins about its axis. When an electrically charged object spins, it produces a magnetic moment parallel to the axis of
rotation, making it behave like a magnet. Although the electron cannot be viewed solely as a particle, spinning or otherwise, it is indisputable
that it does have a magnetic moment. This magnetic moment is called electron spin.

11.10.11 https://chem.libretexts.org/@go/page/41377
Figure 11.10.9: Electron Spin. In a magnetic field, an electron has two possible orientations with different energies, one with spin up, aligned
with the magnetic field, and one with spin down, aligned against it. All other orientations are forbidden.
In an external magnetic field, the electron has two possible orientations (Figure 11.10.9). These are described by a fourth quantum number
(ms), which for any electron can have only two possible values, designated +½ (up) and −½ (down) to indicate that the two orientations are
opposites; the subscript s is for spin. An electron behaves like a magnet that has one of two possible orientations, aligned either with the
magnetic field or against it. The implications of electron spin for chemistry were recognized almost immediately by an Austrian physicist,
Wolfgang Pauli (1900–1958; Nobel Prize in Physics, 1945), who determined that each orbital can contain no more than two electrons whoe
developed the Pauli exclusion principle.

Note: Pauli Exclusion Principle


No two electrons in an atom can have the same values of all four quantum numbers (n , l, m , m ). l s

By giving the values of n, l, and ml, we also specify a particular orbital (e.g., 1s with n = 1, l = 0, ml = 0). Because ms has only two possible
values (+½ or −½), two electrons, and only two electrons, can occupy any given orbital, one with spin up and one with spin down. With this
information, we can proceed to construct the entire periodic table, which was originally based on the physical and chemical properties of the
known elements.

Example 11.10.2

List all the allowed combinations of the four quantum numbers (n, l, ml, ms) for electrons in a 2p orbital and predict the maximum number
of electrons the 2p subshell can accommodate.
Given: orbital
Asked for: allowed quantum numbers and maximum number of electrons in orbital
Strategy:
A. List the quantum numbers (n, l, ml) that correspond to an n = 2p orbital. List all allowed combinations of (n, l, ml).
B. Build on these combinations to list all the allowed combinations of (n, l, ml, ms).
C. Add together the number of combinations to predict the maximum number of electrons the 2p subshell can accommodate.
Solution:
A For a 2p orbital, we know that n = 2, l = n − 1 = 1, and ml = −l, (−l +1),…, (l − 1), l. There are only three possible combinations of (n, l,
ml): (2, 1, 1), (2, 1, 0), and (2, 1, −1).
B Because ms is independent of the other quantum numbers and can have values of only +½ and −½, there are six possible combinations
of (n, l, ml, ms): (2, 1, 1, +½), (2, 1, 1, −½), (2, 1, 0, +½), (2, 1, 0, −½), (2, 1, −1, +½), and (2, 1, −1, −½).
C Hence the 2p subshell, which consists of three 2p orbitals (2px, 2py, and 2pz), can contain a total of six electrons, two in each orbital.

Exercise 11.10.2

List all the allowed combinations of the four quantum numbers (n, l, ml, ms) for a 6s orbital, and predict the total number of electrons it
can contain.
Answer: (6, 0, 0, +½), (6, 0, 0, −½); two electrons

Contributors and Attributions


David M. Hanson, Erica Harvey, Robert Sweeney, Theresa Julia Zielinski ("Quantum States of Atoms and Molecules")

11.10.12 https://chem.libretexts.org/@go/page/41377
11.10: The Schrödinger Wave Equation for the Hydrogen Atom is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.

11.10.13 https://chem.libretexts.org/@go/page/41377
11.11: Many-Electron Atoms and the Periodic Table
Quantum mechanics can account for the periodic structure of the elements, by any measure a major conceptual accomplishment for
any theory. Although accurate computations become increasingly more challenging as the number of electrons increases, the
general patterns of atomic behavior can be predicted with remarkable accuracy.
Figure 11.11.1 shows a schematic representation of a helium atom with two electrons whose coordinates are given by the vectors
r and r . The electrons are separated by a distance r
1 2 = | r − r | . The origin of the coordinate system is fixed at the nucleus. As
12 1 2

with the hydrogen atom, the nuclei for multi-electron atoms are so much heavier than an electron that the nucleus is assumed to be
the center of mass. Fixing the origin of the coordinate system at the nucleus allows us to exclude translational motion of the center
of mass from our quantum mechanical treatment.

Figure 11.11.1 : a) The nucleus (++) and electrons (e-) of the helium atom. b) Equivalent reduced particles with the center of mass
(approximately located at the nucleus) at the origin of the coordinate system. Note that μ and μ ≈ m .
1 2 e

The Schrödinger equation operator for the hydrogen atom serves as a reference point for writing the Schrödinger equation for
atoms with more than one electron. Start with the same general form we used for the hydrogen atom Hamiltonian
^
^+V
(T )ψ = Eψ (11.11.1)

Include a kinetic energy term for each electron and a potential energy term for the attraction of each negatively charged electron for
the positively charged nucleus and a potential energy term for the mutual repulsion of each pair of negatively charged electrons.
The He atom Schrödinger equation is
2

2 2
(− (∇ + ∇ ) + V (r1 ) + V (r2 ) + V (r12 )) ψ = Eψ (11.11.2)
1 2
2me

where
2
2e
V (r1 ) = − (11.11.3)
4πϵ0 r1

2
2e
V (r2 ) = − (11.11.4)
4πϵ0 r2

2
e
V (r12 ) = − (11.11.5)
4πϵ0 r12

Equation 11.11.2 can be extended to any atom or ion by including terms for the additional electrons and replacing the He nuclear
charge +2 with a general charge Z; e.g.
2
Ze
V (r1 ) = − (11.11.6)
4πϵ0 r1

Equation 11.11.2 then becomes


2

2
(− ∑∇ + ∑ V (ri ) + ∑ V (rij )) ψ = Eψ (11.11.7)
i
2me
i i i≠j

Given what we have learned from the previous quantum mechanical systems we’ve studied, we predict that exact solutions to the
multi-electron Schrödinger equation would consist of a family of multi-electron wavefunctions, each with an associated energy
eigenvalue. These wavefunctions and energies would describe the ground and excited states of the multi-electron atom, just as the

11.11.1 https://chem.libretexts.org/@go/page/41378
hydrogen wavefunctions and their associated energies describe the ground and excited states of the hydrogen atom. We would
predict quantum numbers to be involved, as well.
The fact that electrons interact through their electron-electron repulsion means that an exact wavefunction for a multi-electron
system would be a single function that depends simultaneously upon the coordinates of all the electrons; i.e., a multi-electron
wavefunction:

Ψ(r1 , r2 , ⋯ ri ) (11.11.8)

Unfortunately, the electron-electron repulsion terms make it impossible to find an exact solution to the Schrödinger equation for
many-electron atoms. The most basic ansatz to the exact solutions involve writing a multi-electron wavefunction as a simple
product of single-electron wavefunctions

ψ(r1 , r2 , ⋯ , ri ) = φ1 (r1 )φ2 (r2 ) ⋯ φi (ri ) (11.11.9)

Obtaining the energy of the atom in the state described by that wavefunction as the sum of the energies of the one-electron
components.
By writing the multi-electron wavefunction as a product of single-electron functions in Equation 11.11.9, we conceptually
transform a multi-electron atom into a collection of individual electrons located in individual orbitals whose spatial characteristics
and energies can be separately identified. For atoms these single-electron wavefunctions are called atomic orbitals. For molecules,
as we will see in the next chapter, they are called molecular orbitals. While a great deal can be learned from such an analysis, it is
important to keep in mind that such a discrete, compartmentalized picture of the electrons is an approximation, albeit a powerful
one.

Electron Configurations
The specific arrangement of electrons in orbitals of an atom determines many of the chemical properties of that atom and is
formulated via the Aufbau principle, which means "building-up" in German. Aufbau principles determine the order in which atomic
orbitals are filled as the atomic number is increased. For the hydrogen atom, the order of increasing orbital energy is given by 1s <
2s = 2p < 3s = 3p = 3d, etc. The dependence of energy on n alone leads to extensive degeneracy, which is however removed for
orbitals in many-electron atoms. Thus 2s lies below 2p, as already observed in helium. Similarly, 3s, 3p and 3d increase energy in
that order, and so on. The 4s is lowered sufficiently that it becomes comparable to 3d. The general ordering of atomic orbitals is
summarized in the following scheme:
1s < 2s < 2p < 3s < 3p < 4s ∼ 3d < 4p < 5s ∼ 4d (11.11.10)

< 5p < 6s ∼ 5d ∼ 4f < 6p < 7s ∼ 6d ∼ 5f

and illustrated in Figure 11.11.2. This provides enough orbitals to fill the ground states of all the atoms in the periodic table. For
orbitals designated as comparable in energy, e.g., 4s ∼ 3d, the actual order depends which other orbitals are occupied. The energy
of atomic orbitals increases as the principal quantum number, n , increases. In any atom with two or more electrons, the repulsion
between the electrons makes energies of subshells with different values of l differ so that the energy of the orbitals increases within
a shell in the order s < p < d < f. Figure 11.11.2 depicts how these two trends in increasing energy relate. The 1s orbital at the
bottom of the diagram is the orbital with electrons of lowest energy. The energy increases as we move up to the 2s and then 2p, 3s,
and 3p orbitals, showing that the increasing n value has more influence on energy than the increasing l value for small atoms.
However, this pattern does not hold for larger atoms. The 3d orbital is higher in energy than the 4s orbital. Such overlaps continue
to occur frequently as we move up the chart.

11.11.2 https://chem.libretexts.org/@go/page/41378
Figure 11.11.2: Generalized energy-level diagram for atomic orbitals in an atom with two or more electrons (not to scale).
Electrons in successive atoms on the periodic table tend to fill low-energy orbitals first. The arrangement of electrons in the orbitals
of an atom is called the electron configuration of the atom. We describe an electron configuration with a symbol that contains three
pieces of information ( Figure 11.11.3):
1. The number of the principal quantum shell, n,
2. The letter that designates the orbital type (the subshell, l), and
3. A superscript number that designates the number of electrons in that particular subshell.
For example, the notation 2p4 (read "two–p–four") indicates four electrons in a p subshell (l = 1) with a principal quantum number
(n) of 2. The notation 3d8 (read "three–d–eight") indicates eight electrons in the d subshell (i.e., l = 2) of the principal shell for
which n = 3.

Figure 11.11.3: The diagram of an electron configuration specifies the subshell (n and l value, with letter symbol) and superscript
number of electrons.
To determine the electron configuration for any particular atom, we can “build” the structures in the order of atomic numbers.
Beginning with hydrogen, and continuing across the periods of the periodic table, we add one proton at a time to the nucleus and
one electron to the proper subshell until we have described the electron configurations of all the elements. This procedure is called
the Aufbau principle, from the German word Aufbau (“to build up”). Each added electron occupies the subshell of lowest energy
available (in the order shown in Figure 11.11.4), subject to the limitations imposed by the allowed quantum numbers according to
the Pauli exclusion principle. Electrons enter higher-energy subshells only after lower-energy subshells have been filled to capacity.
Figure 11.11.3 illustrates the traditional way to remember the filling order for atomic orbitals.

Figure 11.11.4: The arrow leads through each subshell in the appropriate filling order for electron configurations. This chart is
straightforward to construct. Simply make a column for all the s orbitals with each n shell on a separate row. Repeat for p, d, and f.
Be sure to only include orbitals allowed by the quantum numbers (no 1p or 2d, and so forth). Finally, draw diagonal lines from top
to bottom as shown.

11.11.3 https://chem.libretexts.org/@go/page/41378
We will now construct the ground-state electron configuration and orbital diagram for a selection of atoms in the first and second
periods of the periodic table. Orbital diagrams are pictorial representations of the electron configuration, showing the individual
orbitals and the pairing arrangement of electrons. We start with a single hydrogen atom (atomic number 1), which consists of one
proton and one electron. Referring to either Figure 11.11.4, we would expect to find the electron in the 1s orbital. By convention,
1
the m s =+ value is usually filled first. The electron configuration and the orbital diagram are:
2

Following hydrogen is the noble gas helium, which has an atomic number of 2. The helium atom contains two protons and two
1
electrons. The first electron has the same four quantum numbers as the hydrogen atom electron (n = 1, l = 0, ml = 0, ms = + ).
2
The second electron also goes into the 1s orbital and fills that orbital. The second electron has the same n, l, and ml quantum
1
numbers, but must have the opposite spin quantum number, ms = − . This is in accord with the Pauli exclusion principle: No
2
two electrons in the same atom can have the same set of four quantum numbers. For orbital diagrams, this means two arrows go in
each box (representing two electrons in each orbital) and the arrows must point in opposite directions (representing paired spins).
The electron configuration and orbital diagram of helium are:
In this figure, the element symbol H e is followed by the electron configuration, “1 s superscript 2.” An orbital diagram is provided that consists of a single square. The square is labeled below as “1 s.” It contains a
pair of half arrows: one pointing up and the other down.

The n = 1 shell is completely filled in a helium atom.


The next atom is the alkali metal lithium with an atomic number of 3. The first two electrons in lithium fill the 1s orbital and have
the same sets of four quantum numbers as the two electrons in helium. The remaining electron must occupy the orbital of next
lowest energy, the 2s orbital (Figure 11.11.4 ). Thus, the electron configuration and orbital diagram of lithium are:

An atom of the alkaline earth metal beryllium, with an atomic number of 4, contains four protons in the nucleus and four electrons
surrounding the nucleus. The fourth electron fills the remaining space in the 2s orbital.
In this figure, the element symbol B e is followed by the electron configuration, “1 s superscript 2 2 s superscript 2.” An orbital diagram is provided that consists of two individual squares. The first square is labeled
below as, “1 s.” The second square is similarly labeled, “2 s.” Both squares contain a pair of half arrows: one pointing up and the other down.

An atom of boron (atomic number 5) contains five electrons. The n = 1 shell is filled with two electrons and three electrons will
occupy the n = 2 shell. Because any s subshell can contain only two electrons, the fifth electron must occupy the next energy level,
which will be a 2p orbital. There are three degenerate 2p orbitals (ml = −1, 0, +1) and the electron can occupy any one of these p
orbitals. When drawing orbital diagrams, we include empty boxes to depict any empty orbitals in the same subshell that we are
filling.

Carbon (atomic number 6) has six electrons. Four of them fill the 1s and 2s orbitals. The remaining two electrons occupy the 2p
subshell. We now have a choice of filling one of the 2p orbitals and pairing the electrons or of leaving the electrons unpaired in two
different, but degenerate, p orbitals. The orbitals are filled as described by Hund’s rule: the lowest-energy configuration for an atom
with electrons within a set of degenerate orbitals is that having the maximum number of unpaired electrons. Thus, the two electrons
in the carbon 2p orbitals have identical n, l, and ms quantum numbers and differ in their ml quantum number (in accord with the
Pauli exclusion principle). The electron configuration and orbital diagram for carbon are:
In this figure, the element symbol C is followed by the electron configuration, “1 s superscript 2 2 s superscript 2 2 p superscript 2.” The orbital diagram consists of two individual squares followed by 3 connected
squares in a single row. The first blue square is labeled below as, “1 s.” The second is similarly labeled, “2 s.” The connected squares are labeled below as, “2 p.” All squares not connected to each other contain a pair of
half arrows: one pointing up and the other down. The first two squares in the group of 3 each contain a single upward pointing arrow.

Nitrogen (atomic number 7) fills the 1s and 2s subshells and has one electron in each of the three 2p orbitals, in accordance with
Hund’s rule. These three electrons have unpaired spins. Oxygen (atomic number 8) has a pair of electrons in any one of the 2p

11.11.4 https://chem.libretexts.org/@go/page/41378
orbitals (the electrons have opposite spins) and a single electron in each of the other two. Fluorine (atomic number 9) has only one
2p orbital containing an unpaired electron. All of the electrons in the noble gas neon (atomic number 10) are paired, and all of the
orbitals in the n = 1 and the n = 2 shells are filled. The electron configurations and orbital diagrams of these four elements are:

Figure 11.11.5). Since the core electron shells correspond to noble gas electron configurations, we can abbreviate electron
configurations by writing the noble gas that matches the core electron configuration, along with the valence electrons in a
condensed format. For our sodium example, the symbol [Ne] represents core electrons, (1 s22s22p6) and our abbreviated or
condensed configuration is [Ne]3s1.

Figure 11.11.5: A core-abbreviated electron configuration (right) replaces the core electrons with the noble gas symbol whose
configuration matches the core electron configuration of the other element.
Similarly, the abbreviated configuration of lithium can be represented as [He]2s1, where [He] represents the configuration of the
helium atom, which is identical to that of the filled inner shell of lithium. Writing the configurations in this way emphasizes the
similarity of the configurations of lithium and sodium. Both atoms, which are in the alkali metal family, have only one electron in a
valence s subshell outside a filled set of inner shells.
1
Li : [He] 2s (11.11.11)

1
Na : [Ne] 3s

The alkaline earth metal magnesium (atomic number 12), with its 12 electrons in a [Ne]3s2 configuration, is analogous to its family
member beryllium, [He]2s2. Both atoms have a filled s subshell outside their filled inner shells. Aluminum (atomic number 13),
with 13 electrons and the electron configuration [Ne]3s23p1, is analogous to its family member boron, [He]2s22p1.
The electron configurations of silicon (14 electrons), phosphorus (15 electrons), sulfur (16 electrons), chlorine (17 electrons), and
argon (18 electrons) are analogous in the electron configurations of their outer shells to their corresponding family members
carbon, nitrogen, oxygen, fluorine, and neon, respectively, except that the principal quantum number of the outer shell of the
heavier elements has increased by one to n = 3. Figure
Beginning with the transition metal scandium (atomic number 21), additional electrons are added successively to the 3d subshell.
This subshell is filled to its capacity with 10 electrons (remember that for l = 2 [d orbitals], there are 2l + 1 = 5 values of ml,
meaning that there are five d orbitals that have a combined capacity of 10 electrons). The 4p subshell fills next. Note that for three
series of elements, scandium (Sc) through copper (Cu), yttrium (Y) through silver (Ag), and lutetium (Lu) through gold (Au), a
total of 10 d electrons are successively added to the (n – 1) shell next to the n shell to bring that (n – 1) shell from 8 to 18 electrons.
For two series, lanthanum (La) through lutetium (Lu) and actinium (Ac) through lawrencium (Lr), 14 f electrons (l = 3, 2l + 1 = 7
ml values; thus, seven orbitals with a combined capacity of 14 electrons) are successively added to the (n – 2) shell to bring that
shell from 18 electrons to a total of 32 electrons.

11.11.5 https://chem.libretexts.org/@go/page/41378
Example 11.11.1

Quantum Numbers and Electron Configurations What is the electron configuration and orbital diagram for a phosphorus
atom? What are the four quantum numbers for the last electron added?
Solution
The atomic number of phosphorus is 15. Thus, a phosphorus atom contains 15 electrons. The order of filling of the energy
levels is 1s, 2s, 2p, 3s, 3p, 4s, . . . The 15 electrons of the phosphorus atom will fill up to the 3p orbital, which will contain
three electrons:

The last electron added is a 3p electron. Therefore, n = 3 and, for a p-type orbital, l = 1. The ml value could be –1, 0, or +1.
The three p orbitals are degenerate, so any of these ml values is correct. For unpaired electrons, convention assigns the
1 1
value of + for the spin quantum number; thus, m s =+ .
2 2

Exercise 11.11.1

Identify the atoms from the electron configurations given:


a. [Ar]4s23d5
b. [Kr]5s24d105p6
Answer
(a) Mn (b) Xe

Effective Charge, Shielding and Penetration


For an atom or an ion with only a single electron, we can calculate the potential energy by considering only the electrostatic
attraction between the positively charged nucleus and the negatively charged electron. When more than one electron is present,
however, the total energy of the atom or the ion depends not only on attractive electron-nucleus interactions but also on repulsive
electron-electron interactions. When there are two electrons, the repulsive interactions depend on the positions of both electrons at
a given instant, but because we cannot specify the exact positions of the electrons, it is impossible to exactly calculate the repulsive
interactions. Consequently, we must use approximate methods to deal with the effect of electron-electron repulsions on orbital
energies.
If an electron is far from the nucleus (i.e., if the distance r between the nucleus and the electron is large), then at any given moment,
most of the other electrons will be between that electron and the nucleus. Hence the electrons will cancel a portion of the positive
charge of the nucleus and thereby decrease the attractive interaction between it and the electron farther away. As a result, the
electron farther away experiences an effective nuclear charge (Zeff) that is less than the actual nuclear charge Z (Figure 11.11.6).
This effect is called electron shielding.
As the distance between an electron and the nucleus approaches infinity, Zeff approaches a value of 1 because all the other (Z − 1)
electrons in the neutral atom are, on the average, between it and the nucleus. If, on the other hand, an electron is very close to the
nucleus, then at any given moment most of the other electrons are farther from the nucleus and do not shield the nuclear charge. At
r ≈ 0, the positive charge experienced by an electron is approximately the full nuclear charge, or Zeff ≈ Z. At intermediate values of
r, the effective nuclear charge is somewhere between 1 and Z: 1 ≤ Zeff ≤ Z. Thus the actual Zeff experienced by an electron in a
given orbital depends not only on the spatial distribution of the electron in that orbital but also on the distribution of all the other
electrons present. This leads to large differences in Zeff for different elements, as shown in Figure 2.5.1 for the elements of the first
three rows of the periodic table. Notice that only for hydrogen does Zeff = Z, and only for helium are Zeff and Z comparable in
magnitude.

11.11.6 https://chem.libretexts.org/@go/page/41378
Figure 11.11.6: Relationship between the Effective Nuclear Charge Zeff and the Atomic Number Z for the Outer Electrons of the
Elements of the First Three Rows of the Periodic Table. Except for hydrogen, Zeff is always less than Z, and Zeff increases from left
to right as you go across a row.
Because of the effects of shielding and the different radial distributions of orbitals with the same value of n but different values of l,
the different subshells are not degenerate in a multielectron atom. For a given value of n, the ns orbital is always lower in energy
than the np orbitals, which are lower in energy than the nd orbitals, and so forth. As a result, some subshells with higher principal
quantum numbers are actually lower in energy than subshells with a lower value of n; for example, the 4s orbital is lower in energy
than the 3d orbitals for most atoms.

Note
Except for the single electron containing hydrogen atom, in every other element Z ef f is always less than Z .

Figure 11.11.7: Orbital Penetration. A comparison of the radial probability distribution of the 2 s and 2p orbitals for various states
of the hydrogen atom shows that the 2s orbital penetrates inside the 1s orbital more than the 2p orbital does. Consequently, when an
electron is in the small inner lobe of the 2s orbital, it experiences a relatively large value of Zeff, which causes the energy of the 2s
orbital to be lower than the energy of the 2p orbital.

Ionization Energy
Ionization energy is the energy required to remove an electron from a neutral atom in its gaseous phase. Conceptually, ionization
energy is the opposite of electronegativity. The lower this energy is, the more readily the atom becomes a cation. Therefore, the
higher this energy is, the more unlikely it is the atom becomes a cation. Generally, elements on the right side of the periodic table
have a higher ionization energy because their valence shell is nearly filled. Elements on the left side of the periodic table have low
ionization energies because of their willingness to lose electrons and become cations. Thus, ionization energy increases from left to
right on the periodic table.
Another factor that affects ionization energy is electron shielding. Electron shielding describes the ability of an atom's inner
electrons to shield its positively-charged nucleus from its valence electrons. When moving to the right of a period, the number of
electrons increases and the strength of shielding increases. As a result, it is easier for valence shell electrons to ionize, and thus the
ionization energy decreases down a group. Electron shielding is also known as screening.

11.11.7 https://chem.libretexts.org/@go/page/41378
Note
The ionization energy of the elements within a period generally increases from left to right. This is due to valence shell
stability.
The ionization energy of the elements within a group generally decreases from top to bottom. This is due to electron
shielding.
The noble gases possess very high ionization energies because of their full valence shells as indicated in the graph. Note
that helium has the highest ionization energy of all the elements.

Some elements have several ionization energies; these varying energies are referred to as the first ionization energy, the second
ionization energy, third ionization energy, etc. The first ionization energy is the energy requiredto remove the outermost, or highest,
energy electron, the second ionization energy is the energy required to remove any subsequent high-energy electron from a gaseous
cation, etc. Below are the chemical equations describing the first and second ionization energies:
First Ionization Energy:
+ −
X(g) → X +e (11.11.12)
(g)

Second Ionization Energy:


+ 2+ −
X → X +e (11.11.13)
(g) (g)

Generally, any subsequent ionization energies (2nd, 3rd, etc.) follow the same periodic trend as the first ionization energy.
Ionization energies decrease as atomic radii increase. This observation is affected by n (the principal quantum number) and Z ef f

(based on the atomic number and shows how many protons are seen in the atom) on the ionization energy (I). The relationship is
given by the following equation:
2
RH Z
ef f
I = (11.11.14)
2
n

Across a period, Z increases and n (principal quantum number) remains the same, so the ionization energy increases.
ef f

Down a group, n increases and Z increases slightly; the ionization energy decreases.
ef f

Figure 11.11.8: Periodic trends in ionization energy.


The periodic structure of the elements is evident for many physical and chemical properties, including chemical valence, atomic
radius, electronegativity, melting point, density, and hardness. The classic prototype for periodic behavior is the variation of the
first ionization energy with atomic number, which is plotted in Figure 11.11.8.

Electron Affinity
The electron affinity (EA) of an element E is defined as minus the internal energy change associated with the gain of an electron by
a gaseous atom, at 0 K :
− −
E(g) + e → E (11.11.15)
(g)

11.11.8 https://chem.libretexts.org/@go/page/41378
Unlike ionization energies, which are always positive for a neutral atom because energy is required to remove an electron, electron
affinities can be positive (energy is released when an electron is added), negative (energy must be added to the system to produce
an anion), or zero (the process is energetically neutral).

Figure 11.11.9 : Periodic trends in electron affinities.


The periodic trends of electron affinity (Figure 11.11.9) shows that chlorine has the most positive electron affinity of any element,
which means that more energy is released when an electron is added to a gaseous chlorine atom than to an atom of any other
element, EA= 348.6 kJmol-1 and the Group 17 elements have the largest values overall. The addition of a second electron to an
element is expected to be much less favored since there will be repulsion between the negatively charged electron and the overall
negatively charged anion. For example, for O the values are:
→ − −1
O(g) + e O EA = +141 kJmol (11.11.16)
(g)

− − 2− −1
O +e → O )(g) EA = −798 kJmol (11.11.17)
(g)

David M. Hanson, Erica Harvey, Robert Sweeney, Theresa Julia Zielinski ("Quantum States of Atoms and Molecules")

11.11: Many-Electron Atoms and the Periodic Table is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.

11.11.9 https://chem.libretexts.org/@go/page/41378
11.E: Quantum Mechanics and Atomic Structure (Exercises)
11.1: The Wave Theory of Light
Q11.1
What is the wavelength associated with a photon of a light with the energy is 3.6 × 10 −19
J ?

S11.1
hc
E = hv =
λ

−34 8 −1
(6.626 × 10 J s)(3 × 10 ms )
−19
3.6 × 10 J =
λ

−7
λ = 0.55 × 10 m = 550 nm

Q11.2
Calculate the energy of a photon of a light with the frequency is 6.5 × 10 −14
s
−1
?

S11.2
E = hv

−34 −14 −1
E = (6.626 × 10 J s)(6.5 × 10 s )

−19
E = 4.3 × 10 J

Q11.3a
What are the wavelength of the light emitted by atomic hydrogen when the electron move from n =5 to n =3 orbit in the Bohr
model?

S11.3a
Bohr model:

1 1 1
= RH ( − )
2 2
λ na n
b

1 −2 −1
1 1
= (1.09678 × 10 nm )( − )
λ 9 25

λ = 1, 282 nm

Q11.3b
Calculate the different energy between n = 3 and n = 4 of atomic hydrogen based on Bohr orbits

S11.3b
1 1
ΔE = −hc RH ( − )
2 2
3 4

−34 8 −1
1 1
ΔE = −6.626 × 10 J. s)(2.9979 × 10 m/s)(109, 737 c m )( − ))
2 2
3 4

−21
ΔE = 1.8 × 10 J

Q11.4
Calculate the wavenumber of the wavelength of the light emitted from the n = 8 to n = 6 transition.

11.E.1 https://chem.libretexts.org/@go/page/41337
S11.4
This is a simple application of the Rydberg equation

1 1
~
ν RH ( − )
2 2
n n
f i

with ni =6 and n
f =8 .
1 1
~ −1 −1
ν = 109, 737c m ( − ) = 1333.61 c m
2 2
6 8

Q11.5
Calculate the wavelength associated with a 42 g baseball with speed of 80 m/s.

S11.5
h h
λ = =
p mv

34
6.626 × 10 Js
−34
λ = = 1.97 × 10 m
(0.042 kg)(80m/s)

This is a very small wavelength as expected since a ball behaves rather classically (i.e., non-quantum).

Q11.6
Calculate the energy of a 530-nm photon.

S11.6
E
p =
c

h
λ =
p

hc
λ =
E

−34 8
hc (6.626 × 1 − )(3 × 10 )
−9
E = = = 3.75 × 10 J
−9
λ 530 × 10

11.3: The Photoelectric Effect


Q11.7
Describe Planck's three experimental observations that explain the photoelectric effect.

S11.7
"When light of a certain frequency shines on a clean metal surface, electrons are ejectedd from the metal. Experimentally, it is found
that:
1. The number of electrons ejected is proportional to the intensity of light
2. The kinetic energy of the ejected electrons is proportional to the frequency of the ejected light
3. No electrons can be ejected if the frequency of the light is lower than a certain value, called the threshold frequency (nu ). o

Q11.8a
What is the difference between classical and quantum mechanics? Provide the equation relating the energy of emitted radiation to
frequency.

S11.8a
Classical mechanics predicts that the radiant energy produced by oscillating objects is continuous. Quantum mechanics predicts that
their energy can be viewed as existing in discrete levels. The equation is

11.E.2 https://chem.libretexts.org/@go/page/41337
E = hν

. Energy is emitted in discrete multiples of hν , where h = Planck's constant.

Q11.8b
Do you expect the first ionization of an atom to be an endothermic process or an exothermic process?

S11.8b
It will be endothermic because energy is needed to ionize the atom.

11.2: Planck's Quantum Theory


Q11.9a
Calculate the wavelength of light with an energy of 5.22 × 10 −19
J .

S11.9a
Rearrange the equation
hc
E =
λ

so that the unknown value is the wavelength of light


hc
λ =
E

−34 8
(6.626 × 10 Js)(3.0 × 10 m/s)
−7
λ = = 3.81 × 10 m = 381nm
−19
5.22 × 10 J

Q11.9b
You realize after class that you forgot to write down the final wavelength of your proton. You do, however, remember that the
potential difference was 114 V. From this information, figure out the final wavelength:

S11.9b
1 2
mv = eV
2

e = 1.6x 10
−19
C This is the charge of the electron
m=9.11x10-31kg
−−−
2eV
v=√
m
−−−−−−−−−−−
−19
2(1.6⋅ 10 )(114)

−31
1.67⋅10

=2.0x109m/s
h
λ =
mv
−34
6.626⋅10
λ =
−31 9
(9.11⋅ 10 )(2⋅ 10 )

−13
λ = 3.64 ⋅ 10 m

Q11.10a
given the uncertainty for position of an electron circling an atom is 0.5A0. Find uncertainty in its velocity?

S11.10a
Δp ≥ h/ 4πΔx = 6.626 x 10-34 Js/ 4π(0.5 x 10-10m) = 1.311.05 x 10-24kg m s-1
Δ v= Δp/ m ≥ 1.3 x 10-24 kg m s-1/ 9.109 x 10-31 kg = 1 x 106 m s-1

Q11.10
If the uncertainty of measuring the position of an electron is 2.0 Å, what is the uncertainty of simultaneously measuring its velocity?
Hint: What formula deals with uncertainty of measurements?

11.E.3 https://chem.libretexts.org/@go/page/41337
11.5: de Broglie's Postulate

11.6: The Heisenberg Uncertainty Principle


11.7: The Schrödinger Wave Equation
Q11.11
A typical mass for a horse is 510 kg, and a typical galloping speed is 22 kilometers per hour. Use these values to answer the following
questions.
a. What is the momentum of a galloping horse? What is its wavelength?
b. If a galloping horse's velocity and position are simultaneously measured, and the velocity is measured to within ± 1.0%, what is the
uncertainty of its position?
c. Suppose Planck's constant was actually 0.01 J s. How would that change your answers to (a) and (b)? Which values would be
unchanged?
Hints:
de Broglie's postulate deals with the wave-like properties of particles.
Heisenberg's uncertainty principle deals with uncertainty of simultaneous measurements.

S11.11
(a)
1. Use the relationship between velocity and momentum to find the momentum.

p = mv

km 1000 m hr
p = 510 kg × 22 × ×
hr km 3600 s

kg m
3
p = 3.1 × 10
s

2. Find the de Broglie wavelength.


h
λ =
p

2
kg m
−34
6.626 × 10 J s 2
s
λ = ×
kg m J
3
3.1 × 10
s

−37
λ = 2.1 × 10 m

(b)
1. Find the uncertainty of momentum from the uncertainty of velocity.

Δp = mΔv

Δv = 0.01 × v

km 1000 m hr
Δp = 510 kg × 0.01 × 22 × ×
hr km 3600 s

kg m
Δp = 31
s

2. Use Heisenberg's uncertainty principle to find the uncertainty of position.


h
ΔxΔp ≥

h
Δx ≥
4πΔp

11.E.4 https://chem.libretexts.org/@go/page/41337
2
kg m
−34
6.626 × 10 J s s
2

Δx ≥ ×
kg m J
3
4π × 3.1 × 10
s

−36
Δx ≥ 1.7 × 10 m

(c)
The calculations that involve Planck's constant (h) will change. This much larger value will make the wave-like properties of the horse
more important; its wavelength and uncertainty of position will increase dramatically:


h 0.010 J s
λ = =
2
p kg m

kg m s2
3
3.1 × 10 ×
s J

′ −6
λ = 3.2 × 10 m = 3200 nm

2
kg m

h 0.010 J s 2
′ s
Δx ≥ = ×
4πΔp 3
kg m J
4π × 3.1 × 10
s

′ −5
Δx ≥ 2.6 × 10 m = 26 μm

The momentum would not change.

Q11.12a
how fast must a person weighing 42 kg move through a door 0.5 m wide in order to be diffracted?

S11.12a
p= h/ λ = 6.626 x 10 -34 J s/ 0.5
= 3.313 x 10-34 kg m s-1
V= p/m = 3.313 x 10-34 / 42 kg = 7.88 x 10-36 ms-1

Q11.12b
When a particle passes through a slit, diffraction occurs if the particle's wavelength is on the same order of magnitude as the width of
the slit. At approximately what velocity would a 1670 kg car have to move through the Lincoln Tunnel (6.6 m width) for diffraction to
occur? Hint: How can you find the wavelength of the car?

S11.12b
1. Use de Broglie's equation to relate velocity to wavelength; solve for velocity.
h
λ =
mv

h
v=

2. Set the wavelength equal to the tunnel width and solve.


2
kg m
−34
6.626 × 10 J s 2
s
v= ×
1670 kg × 6.6 m J

Answer: the car would have to be driving very slowly.


−38
v = 6.0 × 10 m/s

11.E.5 https://chem.libretexts.org/@go/page/41337
11.4: Bohr's Theory of the Hydrogen Emission Spectrum
Q11.13a
The Paschen emission spectrum is the collection of spectral lines emitted by H atoms, where the final state is n = 3. What are the
shortest and longest wavelengths (in nm) of Paschen spectral lines? Hints: The minimum value for ni is 4, the maximum value for ni is
infinity and the Rydberg formula describes the emission spectra of hydrogen

Q11.13b
Spectral lines of Lyman and Balmer series do not overlap. Verify this statement.

S11.13b
A photon has a specific energy according to the energy between the excited and ground state.
E = (1/nf2 - 1/ni2), A = hv
The lyman series is when nf is the ground state or nf =1 and Balmer series is when nf =2
nf = 1 the value , ni from 2 to infinity with the values between 1 and ½. For the balmer series ni is from 3 to infinity the range between
1/4 and 5/36. Thus, these intervals don not overlap --> the energies do not overlap --> the lines in the series cant overlap

Q11.14
Assume the Rydberg constant for He+ is 9.72x10-18nm, calculate the wavelength of He+ ions from n=3 to n=2.

S11.14
∣ ∣
1 1 1
∣ ∣
= RH −
∣ 2 2 ∣
λ n n
∣ i f ∣

1 ∣ 1 1 ∣
−18
= 9.72 × 10 ∣ − ∣
λ ∣ 32 2
2 ∣

17
λ = 7.41 × 10 nm

Q11.15
Derive from the following equation to solve for wavelength.
1 1
ΔE = hc RH ( − )
2 2
n n
i f

S11.15
∣ ∣
1 1
∣ ∣
ΔE = RH −
∣ 2 2 ∣
∣ n n ∣
i f

∣ ∣
v 1 1
∣ ∣
= ΔE = RH −
∣ 2 2 ∣
c
∣ ni n ∣
f

∣ ∣
1 v 1 1
∣ ∣
= = ΔE = RH −
∣ 2 2 ∣
λ c n n
∣ i f ∣

1
λ =
∣ ∣
1 1
∣ ∣
RH −
∣ 2 2 ∣
∣ ni n ∣
f

Q11.16
Photosynthesis occurs when light excites the electrons in the chloroplasts of leaves. If you flash a light of 500 nm at the plant and
assuming your flashlight emits 4.5x1020 numbers of photons, how much energy in joules was that one flash?

11.E.6 https://chem.libretexts.org/@go/page/41337
Q11.50
What is an antiatom? How is it different from a regular hydrogen? What will happen if these two collided with one another?

S11.50
An antiatom is a hydrogen atom with reversed electrical charges compared to that of a hydrogen atom. The proton is called an
antiproton and it bears a negative charge rather than a positive charge. The electron, or antielectron, has a positive charge rather than a
negative charge. As a result, an antiatom and regular hydrogen would obliterate one another should they collide. Subsequent energy
would then be released.

Q11.51
A scientist determined the kinetic energy released electrons of cesium metal through a photoelectric experiment. Determine h and the
\work function for cesium graphically from the following results:

λ/nm 410 420 475 500 550 630

V/volt 1.6 1.4 1.1 0.91 0.72 0.37

S11.51
First, convert wavelength to frequency and volt to energy in J.

f (1/sec) Energy (J)

7.31707 × 10
15
1.6

7.14286 × 10
15
1.4

6.3157 × 10
15
1.1

6 × 10
15
0.91

5.45455 × 10
15
0.72

4.7619 × 10
15
0.32

Then graph the values and determine the linear equation.

The linear equation is


−16
y = 4.7 × 10 f − 1.89

Use the equation below:


hf- =Ek (max kinetic energy)
h=slope
= y-intercept

11.E.7 https://chem.libretexts.org/@go/page/41337
h= 4.7 X 10-16
= 1.89

Q11.52
Calculate the de Broglie wavelength of a Cl2 molecule at 300 K.

S11.52
−−−−−
3RT
vrms =√
M

=((3)(8.314)(300)/(70.8X10-3 kg) = 325 m/s


h
λ =
mvrms

=(6.626X10-34 J*s)/[(70.8 amu)(1.66X10-27 kg/amu)(325 m/s)] = 1.73 X 10-11 m

Q11.53
Consider a balloon with a diameter of 2.5 × 10
−5
m . What is the uncertainty of the velocity of an oxygen molecule that is trapped
inside.

S11.53
Use the direct application of the uncertainty principle:
h
ΔxΔp ≥

Let's consider the molecule has an uncertainly that is ± radius of the balloon.
−5
Δx = 1.3 × 10 m

The uncertainty of the momentum of the molecule can be estimated via the uncertainty principle.
−34
h (6.626 × 10 J ∗ s)
−30 −1
Δp = = = 4.05 × 10 kg m s
−5
4πΔx (4π)(1.3 × 10 m)

This can be converted to uncertainty in velocity via

p = mv

or

Δp = me Δv

with the electron mass m equal to 9.109 × 10


e
−31
kg . since the mass of the molecule is not uncertain.
−30
Δp 4.05 × 10
Δv = = = 4.5 m/s
−31
me 9.109 × 10 kg

11.8: Particle in a One-Dimensional Box

Q11.55
What conditions must be satisfied for the Schrodinger's equation for ψ to be an acceptable wavefunction. What are some examples of
unacceptable wave functions?
Q11.56
The equation for calculating the energies of the electron in a hydrogen atom or a hydrogenlike ion is given by En = -(2.18 x10^-18
J)Z^2(1/n^2), where Z is the atomic number of the element. One way to modify this equation for many-electron atoms is to replace Z
with (Z- [sigma]), where [sigma] is a positive dimensionless quantity called the shielding constant. Consider the lithium atom as an
example. The physical significance of sigma that it represents the extent of shielding that the two 1s electrons exert on each other.

11.E.8 https://chem.libretexts.org/@go/page/41337
Thus the quantity (Z - [sigma]) is appropriated called the "effective nuclear change." Use the first ionization energy of lithium to
calculate the value of [sigma].

Q11.57
Calculate the wavelength and frequency of an emitted gamma particle that has energy of 3.11 x 10^12 J mol^-1.

11.9: Quantum-Mechanical Tunneling


11.10: The Schrödinger Wave Equation for the Hydrogen Atom

11.11: Many-Electron Atoms and the Periodic Table


11.E: Quantum Mechanics and Atomic Structure (Exercises) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by LibreTexts.

11.E.9 https://chem.libretexts.org/@go/page/41337
CHAPTER OVERVIEW

12: The Chemical Bond


Chemical bond refers to the forces holding atoms together to form molecules and solids. This force is of an electric nature, and the
attraction between electrons of one atom to the nucleus of another atom contributes to what is known as chemical bonds. A
chemical bond is a lasting attraction between atoms that enables the formation of chemical compounds and may result from the
electrostatic force of attraction between atoms with opposite charges, or through the sharing of electrons as in the covalent bonds.
The strength of chemical bonds varies considerably.
12.1: Lewis Structures
12.2: Valence Bond Theory
12.3: Hybridization of Atomic Orbitals
12.4: Electronegativity and Dipole Moment
12.5: Molecular Orbital Theory
12.6: Diatomic Molecules
12.7: Resonance and Electron Delocalization
12.8: Coordination Compounds
12.9: Coordination Compounds in Biological Systems
12.E: The Chemical Bond (Exercises)

Thumbnail: Boundary surface of a delta chemical bond. (Public Domain; Ben Mills via Wikipedia)

12: The Chemical Bond is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1
12.1: Lewis Structures
Chemical bond refers to the forces holding atoms together to form molecules and solids. This force is of an electric nature, and the
attraction between electrons of one atom to the nucleus of another atom contributes to what is known as chemical bonds. Various
theories regarding chemical bonds have been proposed over the past 300 years, during which our interpretation of the world has
also changed. Some old concepts such as Lewis dot structure and valency are still rather useful in our understanding of the
chemical properties of atoms and molecules, and new concepts involving quantum mechanics of chemical bonding interpret
modern observations very well. You learn new concepts such as bond length, bond energy, bond order, covalent bond, ionic bond,
polar and non-polar bond etc. These concepts help you understand the material world at the molecular level.
Of the three primary theories of bonding discussed in this class (Lewis Theory, Valance Bond Theory and Molecular orbital
theory), Lewis theory does not take full advantage of the quantum mechanical concepts discussed previously. Key to this theory is
the Lewis Structure, which is a very simplified representation of the electrons in a molecule and is use to show how the electrons
are arranged around individual atoms in a molecule. Between 1916 and 1919, Gilbert Newton Lewis, Walther Kossel, and Irving
Langmuir formulated a theory to explain chemical bonding. This theory is now called Lewis Theory and it is based on the
following five principles:
1. Valence electrons, or the electrons in the outermost electron shell, have an essential role in chemical bonding.
2. Ionic bonds are formed between atoms when electrons are transferred from one atom to another. Ionic bond is a bond between
nonmetals and metals .
3. Covalent bonds are formed between atoms when pairs of electrons are shared between atoms. A covalent bond is between two
nonmetals.
4. Electrons are transferred/shared so that each atom may reach a more stable electron configuration, i.e., the noble gas
configuration which contains 8 valence electrons; this is called octet rule.
5. Although many different Lewis structures can be constructed, those that are most stable have the lowest formal charge on the
atoms.
Writing out Lewis structures can be at times, tricky and somewhat difficult. Moreover, a compound can have multiple Lewis
Structures that contribute to the shape of the overall compound, so one Lewis structure of a compound may not necessarily be
exactly what the compound looks like. If you need a refresher on how to constructing Lewis Structures consult the following
Modules: Writing Lewis Structures, Lewis Structures I and Lewis Structures II.
In Lewis Structures, valence electrons are represented as dots and only outer-shell electrons are considered (e.g., the valence
electrons and not the core electrons). Single (covalent) bonds are made up of two shared electrons and are represented by two dots
between the bonded atoms or a single line. Double bonds and triple bonds are represented as two and three lines/(pairs of
electrons), respectively. Lone pairs on the outer rims of an atom are represented as two dots. This is because they are the ones
involved in chemical reactions. For the 2nd and 3rd periods elements, the number of valence electrons range from 1 to 8. Lewis dot
structure for them are as indicated:
. . . . . . .. ..
Li Be ⋅B⋅ ⋅C ⋅ ⋅N : : O : : F : : Ne : (12.1.1)
. . . . . ..

Lewis made the valence electron visible with dots. Other examples include the following molecules:
H
.. .. H
.. .. .. .. .. .. ..
..
H : H H : O : H H : F : : N ::: N : : O :: O : : F : F : : O :: C :: O :
H : C : H
.. .. H : N : H ..
..
H

Note
Lewis structure does NOT attempt to explain the geometry of molecules, how the bonds form, or how the electrons are shared
between the atoms. It is the simplest and most limited theory on electronic structure.

Principle 1: Valence Electrons


The electrons of an atom can divided into two categories: valence and core electrons. Valence electrons are those occupying the
outermost shell or highest energy level of an atom while core electrons are those occupying the innermost shell or lowest energy
level. This difference greatly influences the role of the two types of electrons in a chemical reaction. Generally, valence electrons

12.1.1 https://chem.libretexts.org/@go/page/41379
can participate in the formation of chemical bonding, but core electrons cannot. While core electrons are not involved in bonding,
they do have an influence on the chemical reactivity of an atom.
The electron configuration of the oxygen atoms is
2 2 4
O : 1s 2s 2p (12.1.2)

As discussed before, this configuration may be shorted


2 4
O : [H e]2 s 2 p (12.1.3)

where the [H e] stands for the configuration of helium (1s 2s ). The 1s electrons in oxygen (Equation 12.1.2) do not participate in
2 2

bonding, i.e., chemistry, and are called core electrons. The other electrons (i.e., the 2s 2p part) are call valence electrons and are
2 4

responsible for the making and breaking of chemical bonds.


The configuration of calcium with 20 electrons can be written
2
C a : [Ar]4s (12.1.4)

where the [Ar] stands for the configuration of argon (1s 2s 2p 3s 3p ). Electronic configurations that are the same as noble gases
2 2 6 2 6

are very stable since they have a full octet (except helium with a full 1s orbital). The electrons in the argon-like closed shell of
Equation 12.1.4 are the core electrons and the two electrons in the 4s orbital are valence electrons.

Example 12.1.1: Cobalt

What are the core and valence electrons in cobalt?


Solution
Start by writing the electron configuration of cobalt (i.e., 27 electrons):
2 2 6 2 6 2 7
1s 2s 2p 3s 3p 4s 3d

However, argon has the electronic structure 1s 2 2 6


2s 2p 3s 3p
2 6
, so we can write the configuration of cobalt as
2 7
[Ar]4 s 3 d

The two electrons in the 4s orbital and the seven electrons in the 3d are the valence electrons: all others are core electrons.

Principle 2: Ionic Bonding


In ionic bonding, electrons are transferred from one atom to another resulting in the formation of positive and negative
ions. The electrostatic attractions between the positive and negative ions hold the compound together. The predicted
overall energy of the ionic bonding process, which includes the ionization energy of the metal and electron affinity of the nonmetal,
is usually positive, indicating that the reaction is endothermic and unfavorable. However, this reaction is highly favorable because
of their electrostatic attraction.
The energy of the electrostatic attraction (E ), a measure of the force’s strength, is inversely proportional to the internuclear distance
between the charged particles (r):
Q1 Q2
E =k (12.1.5)
r

where each ion’s charge is represented by the symbol Q. The proportionality constant k is equal to 2.31 × 10−28 J·m.

Example 12.1.2: Ionic Bonding in NaCl

For example, in the reaction of Na (sodium) and Cl (chlorine), each Cl atom takes one electron from a Na atom. Therefore each
Na becomes a Na+ cation and each Cl atom becomes a Cl- anion. Due to their opposite charges, they attract each other to form
an ionic lattice. The formula (ratio of positive to negative ions) in the lattice is NaCl.

2N a(s) + C l2(g) → 2N aC l(s)

12.1.2 https://chem.libretexts.org/@go/page/41379
Sodium has one valence electron and chlorine has seven valence electrons; the two elements react such that the chlorine atom
takes the valence electron from the sodium atom leaving the chlorine atom with one extra electron and thus negatively charged
and the Sodium atom without an electron and thus positively charged. The two atoms then become ions and because of their
opposite charges the ions are held together in an ionic bond (Equation 12.1.5).
The chlorine has a high affinity for electrons, and the sodium has a low ionization potential. Thus the chlorine gains an electron
from the sodium atom. The Lewis Structure of this reaction is (here we will consider one chlorine atom, rather than Cl2) is:

The arrow indicates the transfer of the electron from sodium to chlorine to form the Na+ metal ion and the Cl- chloride ion

Principle 3: Covalent Bonding


Covalent bonding occurs when pairs of electrons are shared by atoms. Atoms will covalently bond with other atoms in order to gain
more stability, which is gained by forming a full electron shell. By sharing their outer most (valence) electrons, atoms can fill up
their outer electron shell and gain stability. Nonmetals will readily form covalent bonds with other nonmetals in order to obtain
stability, and can form anywhere between one to three covalent bonds with other nonmetals depending on how many valence
electrons they posses. Although it is said that atoms share electrons when they form covalent bonds, they do not usually share the
electrons equally.
The diatomic hydrogen molecule (H2) is the simplest model of a covalent bond, and is represented in Lewis structures as:

The shared pair of electrons provides each hydrogen atom with two electrons in its valence shell (the 1s) orbital. In a sense, it has
the electron configuration of the noble gas helium. When two chlorine atoms covalently bond to form C l C l , the following
2 2

sharing of electrons occurs:

Each chlorine atom shared the bonding pair of electrons and achieves the electron configuration of the noble gas argon. In Lewis
structures the bonding pair of electrons is usually displayed as a line, and the unshared electrons as dots:

The shared electrons are not located in a fixed position between the nuclei. In the case of the H
2 H2 compound, the electron density
is concentrated between the two nuclei:

The two atoms are bound into the H H molecule mainly due to the attraction of the positively charged nuclei for the negatively
2 2

charged electron cloud located between them.

Example 12.1.3: The Chlorate Ion

Draw the Lewis Structure for the chlorate ion (ClO3-).


Solution
First, lets find the how many valence electrons chlorate has:

ClO3- : 7 e-(from Cl) + 3(6) e-(from 3 O atoms) + 1 (from the total charge of -1) = 26
There are 26 valence electrons. Next lets draw the basic framework of the molecule:

12.1.3 https://chem.libretexts.org/@go/page/41379
The molecule uses covalent bonds to hold together the atoms to the central Chlorine. The remaining electrons become non-
bonding electrons. Since 6 electrons were used for the bonds, the 20 others become those un-bonding electrons to complete the
octet:

The oxygen atom's shells fill up with 18 electrons, and the other 2 complete Chlorine's octet

Example 12.1.4: Formaldehyde

Constructing the Lewis Structure of the formaldehyde (H2CO) molecule.


Solution
First find number of valence electrons:
H2CO: 2(1) e- (from the H atoms) + 4 e- (from the C atom) + 6 e- (from the O atom)
There are 12 valence electrons. Next draw out the framework of the molecule:

To satisfy the octet of Carbon, one of the pairs of electrons on Oxygen must be moved to create a double bond with Carbon.
Therefore our Lewis Structure would look as it does below:

The Hydrogen atoms are each filled up with their two electrons and both the Carbon and the Oxygen atoms' octets are filled.

Principle 4: The Octet Rule


The octet rule is a chemical rule of thumb that reflects observation that atoms of main-group elements tend to combine in such a
way that each atom has eight electrons in its valence shell, giving it the same electronic configuration as a noble gas. The Octet
Rule requires all atoms in a molecule to have 8 valence electrons--either by sharing, losing or gaining electrons--to become stable.
For covalent bonds, atoms to share their electrons with each other to satisfy the octet rule. It requires eight electrons because that is
the amount of electrons needed to fill a s- and p- orbital (electron configuration); also known as a noble gas configuration.
Although it is important to remember the "magic number" is 8, but there are many octet rule exceptions.

Octet Rule
A stable arrangement is attended when the atom is surrounded by eight electrons. This octet can be made up by own electrons
and some electrons which are donated (e.g, ionic bonding) or shared (covalent bonding). An atom continues to form bonds
until an octet of electrons is reached.

For the first rows in the periodic table, the magic number 8 can easily be explained from quantum mechanics of the multiple
electron atoms discussed previously. In particular, electrons possess four quantum numbers (usually n , l, m , and m ), which are
l s

restricted as such:
n = 1, 2, 3, . . . (12.1.6)

l = 0, 1, 2, . . . , n − 1 (12.1.7)

ml = −l, −l + 1, . . . , 0, . . . , l − 1, l (12.1.8)

1 1
ms = − , (12.1.9)
2 2

12.1.4 https://chem.libretexts.org/@go/page/41379
where
n is the principal quantum number (dealing with energy states),
lis the orbital quantum number (dealing with subshells),
m is the orbital magnetic quantum number (direction of l) and
l

m is the spin quantum number (direction of spin).


s

By the Pauli's Exclusion Principle all of these numbers cannot be the same for any two electrons in an atom. So in the n = 2 shell
you can have two possible values for l, one possible value for m when l = 0 , three possible values for m when l = 1 , and two
l l

possible values for m . This sums to a total of eight possible value vectors:
s

(1 + 3) ⋅ 2 = 8 (12.1.10)

In the n = 3 shell, the Octet rule also holds for l < 2 , which gives you sort of the same combinatorics all over again.
Since the quantum numbers cannot be the same for any electrons in an atom, adding an extra electron to an atom with eight
electrons in its outermost shell forces the electron to occupy an orbital with a higher principal quantum number (n ). Since the
binding energy of the electron increases with n , this is energetically unfavorable to binding with an atom that has more space in its
outermost shell. For example, for the nonmetals (and the 's' block metals) to achieve an octet, the number of valence electrons is
equal to the group number (Table 12.1.1).
Table 12.1.1 : Select electronic properties of p-block elements of the 2nd row.
Number of Covalent Bonds needed to form
Element Group Valence electrons
valence octet

F 17 7 1

O 16 6 2

N 15 5 3

C 14 4 4

Note: The 18 Electron Rule


The octet rule is not a law and it only works for a handful of elements, mainly the second-row p-block ones; there are many
exceptions. Transition metals compounds often have 18 electrons on the metal atom because eighteen is the number of
electrons in the nearest noble gas configuration in transition metals, which includes s + p + d electrons. The second and third
row of transition metals have f electrons, too, but we usually simplify and don't worry about those electrons for electron
counting purposes; we treat them like core electrons, not valence electrons.

Lewis theory successfully describe the covalent interactions between various nonmetal elements. Examples of hydride compounds
of the elements in Table 12.2.1 (covalent compounds with hydrogen):

The sharing of a pair of electrons represents a single covalent bond, usually just referred to as a single bond. However, in many
molecules atoms attain complete octets by sharing more than one pair of electrons between them.
Two electron pairs shared a double bond
Three electron pairs shared a triple bond

Because each nitrogen contains five valence electrons, they need to share three pairs to each achieve a valence octet. The strong
nitrogen triple bond makes the molecule fairly inert. Moreover, the N - N bond distance in N2 is 1.10 Å, which is appreciably
shorter than the average N-N single bonds (14.7Å).

12.1.5 https://chem.libretexts.org/@go/page/41379
Principle 5: Formal Charges
It is sometimes possible to write more than one Lewis structure for a substance that does not violate the octet rule, as we saw for
CH2O, but not every Lewis structure may be equally reasonable. In these situations, we can choose the most stable Lewis structure
by considering the formal charge on the atoms, which is the difference between the number of valence electrons in the free atom
and the number assigned to it in the Lewis electron structure. The formal charge is a way of computing the charge distribution
within a Lewis structure; the sum of the formal charges on the atoms within a molecule or an ion must equal the overall charge on
the molecule or ion. A formal charge does not represent a true charge on an atom in a covalent bond but is simply used to predict
the most likely structure when a compound has more than one valid Lewis structure. To calculate formal charges, we assign
electrons in the molecule to individual atoms according to these rules:
Nonbonding electrons are assigned to the atom on which they are located.
Bonding electrons are divided equally between the bonded atoms.
For each atom, we then compute a formal charge:

bonding e
− −
formal charge= valence e − (nonbonding e + )
2 (12.1.11)

(f ree atom) (atom in Lewis structure)

The most favorable Lewis Structure has the smallest formal charge for the atoms, and negative formal charges tend to come from
more electronegative atoms. An example of determining formal charge can be seen below with the nitrate ion, NO3-:

The double bonded O atom has 6 electrons: 4 non-bonding and 2 bonding (one electron for each bond). Since O should have 6
electrons, the formal charge is 0.
The two singly bonded O atoms each have 7 electrons: 6 non-bonding and 1 bonding electron. Since O should have 6 electrons,
and there is one extra electron, those O atoms each have formal charges of -1.
The N atom has 4 electrons: 4 bonding and 0 non-bonding electrons. Since N should have 5 electrons and there are only 4
electrons for this N, the N atom has a formal charge of +1.
The charges add up to the overall charge of the ion. 0 + (-1) + (-1) + 1 = -1. Thus, these charges are correct, as the overall
charge of nitrate is -1.
In general you want the fewest number of formal charges possible, i.e. formal charges of 0 for as many of the atoms in a structure
as possible. Also the formal charges should match the electronegativity of the atom, that is negative charges should be on the more
electronegative atoms and positive charges on the least electronegative atoms if possible. Charges of -1 and +1 on adjacent atoms
can usually be removed by using a lone pair of electrons from the -1 atom to form a double (or triple) bond to the atom with the +1
charge.

Summary
Lewis Structures are visual representations of the bonds between atoms and illustrate the lone pairs of electrons in molecules. They
can also be called Lewis dot diagrams and are used as a simple way to show the configuration of atoms within a molecule.
Electrons are shown as "dots" or for bonding electrons as a line between the two atoms. The goal is to obtain the "best" electron
configuration, i.e. the octet rule and formal charges need to be satisfied. Lewis structures can also be useful in predicting molecular
geometry in conjunction with hybrid orbitals. A compound may have multiple resonance forms that are also all correct Lewis
structures. This section will discuss the rules for writing out Lewis structures correctly.

Contributors and Attributions


Mike Blaber (Florida State University)
Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

12.1.6 https://chem.libretexts.org/@go/page/41379
Andreas Hagen (via StackExchange)

12.1: Lewis Structures is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

12.1.7 https://chem.libretexts.org/@go/page/41379
12.2: Valence Bond Theory
Valence bond (VB) theory is one of two basic theories, along with molecular orbital (MO) theory, that were developed to use the
methods of quantum mechanics to explain chemical bonding. It focuses on how the atomic orbitals of the dissociated atoms
combine to give individual chemical bonds when a molecule is formed. In contrast, molecular orbital theory, which will be
discussed elsewhere, predict wavefunctions that cover the entire molecule.
The simplest case to consider is the hydrogen molecule, H . When we say that the two hydrogen nuclei share their electrons to
2

form a covalent bond, what we mean in VB theory terms is that the two spherical 1s orbitals (the grey spheres in Figure 12.2.1)
overlap and contain two electrons.

Figure 12.2.1 : These two electrons are now attracted to the positive charge of both of the hydrogen nuclei, with the result that they
serve as a sort of ‘chemical glue’ holding the two nuclei together. from Wikipedia (credit: Jacek FH)
How far apart are the two nuclei? If they are too far apart, their respective 1s orbitals cannot overlap, and thus no covalent bond can
form - they are still just two separate hydrogen atoms. As they move closer and closer together, orbital overlap begins to occur, and
a bond begins to form. This lowers the potential energy of the system, as new, attractive positive-negative electrostatic
interactions become possible between the nucleus of one atom and the electron of the second (Figure 12.2.2). However, something
else is happening at the same time: as the atoms get closer, the repulsive positive-positive interaction between the two nuclei also
begins to increase.

Figure 12.2.2 : A Plot of Potential Energy versus Internuclear Distance for the Interaction between Two Gaseous Hydrogen Atoms.
At long distances, both attractive and repulsive interactions are small. As the distance between the atoms decreases, the attractive
electron–proton interactions dominate, and the energy of the system decreases. At the observed bond distance, the repulsive
electron–electron and proton–proton interactions just balance the attractive interactions, preventing a further decrease in the
internuclear distance. At very short internuclear distances, the repulsive interactions dominate, making the system less stable than
the isolated atoms.
At first this repulsion is more than offset by the attraction between nuclei and electrons, but at a certain point, as the nuclei get even
closer, the repulsive forces begin to overcome the attractive forces, and the potential energy of the system rises quickly. When the
two nuclei are ‘too close’, we have an unstable, high-energy situation. There is a defined optimal distance between the nuclei in
which the potential energy is at a minimum, meaning that the combined attractive and repulsive forces add up to the greatest
overall attractive force. This optimal internuclear distance is the bond length; the distance is 74 pm for H . Likewise, the 2

difference in potential energy between the lowest energy state (at the optimal internuclear distance) and the state where the two
atoms are completely separated is called the bond dissociation energy, or, more simply, bond strength; the H bond strength is 2

435 kJ/mol.
A comparison of some bond lengths and energies is shown in Table 12.2.1. We can find many of these bonds in a variety of
molecules, and this table provides average values. For example, breaking the first C–H bond in CH4 requires 439.3 kJ/mol, while

12.2.1 https://chem.libretexts.org/@go/page/41380
breaking the first C–H bond in H– CH 2
C H
6 5
(a common paint thinner) requires 375.5 kJ/mol.
Table 12.2.1 : Representative Bond Energies and Lengths
Bond Length (pm) Energy (kJ/mol) Bond Length (pm) Energy (kJ/mol)

H–H 74 436 C–O 140.1 358

H–C 106.8 413 C = O 119.7 745

H–N 101.5 391 C ≡ O 113.7 1072

H–O 97.5 467 H–Cl 127.5 431

C–C 150.6 347 H–Br 141.4 366

C = C 133.5 614 H–I 160.9 298

C ≡ C 120.8 839 O–O 148 146

C–N 142.1 305 O = O 120.8 498

C = N 130.0 615 F–F 141.2 159

C ≡ N 116.1 891 Cl–Cl 198.8 243

Example 12.2.1: A Morse Potential Approximation

The Morse potential, named after physicist Philip M. Morse, is a convenient interatomic interaction approximation for the
potential energy of a diatomic molecule.
2
−β(r−re )
V (r) = De [1 − e ]

Here r is the distance between the atoms, r is the equilibrium bond distance, D is the well depth (defined relative to the
e e

dissociated atoms), and β controls the 'width' of the potential (the smaller β is, the larger the well). For H , the equilibrium
2

bond length r = 74.1 pm, the bond energy D = 7.61 × 10 J , and the Morse potential parameter β = 0.0193 pm .
e e
−19 −1

The force on the system is given by


dV (r)
F (r) = −
dr

Calculate the force for H and explain the status of the system (e.g., compressing, expanding, or stationary) at:
2

a. 45 pm,
b. 74 pm,
c. 150 pm, and
d. 500 pm.

Overlap
Valence bond theory describes a covalent bond as the overlap of half-filled atomic orbitals (each containing a single electron) that
yield a pair of electrons shared between the two bonded atoms. We say that orbitals on two different atoms overlap when a portion
of one orbital and a portion of a second orbital occupy the same region of space. According to valence bond theory, a covalent bond
results when two conditions are met:
1. an orbital on one atom overlaps an orbital on a second atom and
2. the single electrons in each orbital combine to form an electron pair.
The mutual attraction between this negatively charged electron pair and the two atoms’ positively charged nuclei serves to
physically link the two atoms through a force we define as a covalent bond. The strength of a covalent bond depends on the extent
of overlap of the orbitals involved. Orbitals that overlap extensively form bonds that are stronger than those that have less overlap.
In addition to the distance between two orbitals, the orientation of orbitals also affects their overlap (other than for two s orbitals,
which are spherically symmetric). Greater overlap is possible when orbitals are oriented such that they overlap on a direct line

12.2.2 https://chem.libretexts.org/@go/page/41380
between the two nuclei. Figure 12.2.3 illustrates this for two p orbitals from different atoms; the overlap is greater when the
orbitals overlap end to end rather than at an angle.

Figure 12.2.3 : (a) The overlap of two p orbitals is greatest when the orbitals are directed end to end. (b) Any other arrangement
results in less overlap. The dots indicate the locations of the nuclei.
The overlap of two s orbitals (as in H2), the overlap of an s orbital and a p orbital (as in HCl), and the end-to-end overlap of two p
orbitals (as in Cl2) all produce sigma bonds (σ bonds), as illustrated in Figure 12.2.4. A σ bond is a covalent bond in which the
electron density is concentrated in the region along the internuclear axis; that is, a line between the nuclei would pass through the
center of the overlap region. Single bonds in Lewis structures are described as σ bonds in valence bond theory.
Three diagrams are shown and labeled “a,” “b,” and “c.” Diagram a shows two spherical orbitals lying side by side and overlapping. Diagram b shows one spherical and one peanut-shaped orbital lying near one
another so that the spherical orbital overlaps with one end of the peanut-shaped orbital. Diagram c shows two peanut-shaped orbitals lying end to end so that one end of each orbital overlaps the other.

Figure 12.2.4 : Sigma (σ) bonds form from the overlap of the following: (a) two s orbitals, (b) an s orbital and a p orbital, and (c)
two p orbitals. The dots indicate the locations of the nuclei.
A pi bond (π bond) is a type of covalent bond that results from the side-by-side overlap of two p orbitals, as illustrated in Figure
12.2.5. In a π bond, the regions of orbital overlap lie on opposite sides of the internuclear axis. Along the axis itself, there is a node,

that is, a plane with no probability of finding an electron.

Figure 12.2.5 : Pi (π) bonds form from the side-by-side overlap of two p orbitals. The dots indicate the location of the nuclei.
While all single bonds are σ bonds, multiple bonds consist of both σ and π bonds. As the Lewis structures in suggest, O2 contains a
double bond, and N2 contains a triple bond. The double bond consists of one σ bond and one π bond, and the triple bond consists of
one σ bond and two π bonds. Between any two atoms, the first bond formed will always be a σ bond, but there can only be one σ
bond in any one location. In any multiple bond, there will be one σ bond, and the remaining one or two bonds will be π bonds.
These bonds are described in more detail later in this chapter.

As seen in Table 12.2.2, an average carbon-carbon single bond is 347 kJ/mol, while in a carbon-carbon double bond, the π bond
increases the bond strength by 267 kJ/mol. Adding an additional π bond causes a further increase of 225 kJ/mol. We can see a
similar pattern when we compare other σ and π bonds. Thus, each individual π bond is generally weaker than a corresponding σ
bond between the same two atoms. In a σ bond, there is a greater degree of orbital overlap than in a π bond.

Counting σ and π Bonds


Butadiene, C6H6, is used to make synthetic rubber. Identify the number of σ and π bonds contained in this molecule.

Butadiene
Solution

12.2.3 https://chem.libretexts.org/@go/page/41380
There are six σ C–H bonds and one σ C–C bond, for a total of seven from the single bonds. There are two double bonds that
each have a π bond in addition to the σ bond. This gives a total nine σ and two π bonds overall.

Exercise 12.2.2

Identify each illustration as depicting a σ or π bond:


a. side-by-side overlap of a 4p and a 2p orbital
b. end-to-end overlap of a 4p and 4p orbital
c. end-to-end overlap of a 4p and a 2p orbital

Answer
(a) is a π bond with a node along the axis connecting the nuclei while (b) and (c) are σ bonds that overlap along the axis.

The Valence Bond Wavefunction


In the hydrogen molecule, VB theory uses two electrons (electron 1 and electron 2) in two 1s atomic orbitals (ψ and ψ ) on the a b

hydrogen atoms(atom a and atom b ). The simplest description of a H molecule with two electrons and two nuclei is
2

ψsimple = ψa (1)ψb (2) (12.2.1)

Equation 12.2.1 means electron 1 is in ψ of atom a and electron 2 is in ψ on the atom b . Where ψ and ψ may be the 1s
a a a b

orbitals as the simplest wavefunction (n = 1 , l = 0 , m ). The ψ


l wavefunction predicts an energy that is about 6% of the
simple

experimental value at a bond length of 0.90 pm, which is considerably longer than the experimental value of 0.74 pm (Table
12.2.1). However, Equation 12.2.1 does not address the indistinguishability of the problem, i.e, we cannot distinguished each

electron from the other.

Distinguishable vs. Indistinguishable Particles


Assume you have two particle A and B in states 1 and 2. If the two particle are distinguishable, then by exchanging the
particles A and B, you will obtain a new state that will have the same properties as the old state. However, for indistinguishable
particles, exchanging A and B is a transformation that does nothing and you have the same physical state. This means that for
indistinguishable particles, particle labels are unphysical and they represent a redundancy in describing the physical state.

To account for the indistinguishable nature of the two electrons in the system, a covalent wavefunction that is better than Equation
12.2.1 can be constructed (ignoring spin effect):

ψcovalent = ψa (1)ψb (2) + ψa (2)ψb (1) (12.2.2)

Equation 12.2.2 means electron 1 is sometimes in ψ on atom a and electron 2 is on the atom b (left term of right hand side of
a

Equation) and sometime electron 1 is sometimes in ψ on atom b and electron 2 is in ψ on the atom a (right term of right hand
b a

side of Equation). Notice that in contrast to Equation 12.2.1, if the two electron labels are switched, the wavefunction in Equation
12.2.2 does not change. The potential energy curve for the ψ wavefunction predicts a bond energy that is about 72% at a
covalent

bond length of 0.87 pm (Table 12.2.2), so the ψ wavefunction is more accurate than ψ
covalent . simple

To be more accurate within VB theory, the possibility that both electrons may be on atom a (ψ a (1)ψa (2)) or on atom b (
ψ (1)ψ (2)) should also be taken into account. This introduces an ionic contribution:
b b

ψionic = ψa (1)ψa (2) + ψb (2)ψb (1) (12.2.3)

Notice that Equation 12.2.3 addresses indistinguishability since the wavefunction is unchanged if the two electron labels are
switched. The more complete VB wavefunction is a linear combination of both ψ and ψ :
covalent ionic

ψAB = Ψcovalent + λ ψionic (12.2.4)

12.2.4 https://chem.libretexts.org/@go/page/41380
where λ quantifies the amount of ionic character to the bond, e.g., if λ = 0 , then the bond is purely covalent. The potential energy
curve for the ψAB wavefunction predicts a bond energy that is ~80% of the experimental value and a reasonably accurate bond
length of 0.77 pm (Table 12.2.1). Additional tinkering with the wavefunction leads to even better results, but is beyond the scope of
this discussion. It should be noted that the wavefunctions in Equations 12.2.2, 12.2.3, and 12.2.4 are unnormalized.
Table 12.2.2 : Valence Bond Theory Results for H 2

Wavefunction Bond Energy (kJ mol-1) Bond Length (pm)

Experiment 435 0.74

ψsimple 25 0.90

ψcovalent 301 0.87

ψAB 335 0.77

Two-Atom Bonds
The central gist of VB theory is that a covalent bond is formed between the two atoms by the overlap of half filled valence
atomic orbitals of each atom containing one unpaired electron.

Example 12.2.3: Percent Ionic Character

Calculate the relative fraction of the contribution of the ionic character to the VB wavefunction. Assuming that the ionic
character in H Br bond is 20%.
Solution
From Equation 12.2.4, the percent ionic character is given by:
2
λ
× 100%
2
1 +λ

Therefore,
2
λ
20% = 100%
2
1 +λ

2 2
20% + 20%λ = 100%λ

2 2
20% = 100%λ – 20 λ

2
20% = 80%λ

Therefore,

2
20
λ =
80

−−−−−
Hence, λ = √20/80 = 0.5 and the fraction of contribution of ionic character to ψ AB
is 0.5 or 50%.

Linus Pauling used the pair bonding ideas of Lewis together with Heitler–London theory to develop two key concepts in VB
theory: resonance and orbital hybridization. These will be addressed in later Modules.

Summary
Valence bond theory describes bonding as a consequence of the overlap of two separate atomic orbitals on different atoms that
creates a region with one pair of electrons shared between the two atoms. When the orbitals overlap along an axis containing the
nuclei, they form a σ bond. When they overlap in a fashion that creates a node along this axis, they form a π bond.

Contributors and Attributions


Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)

12.2.5 https://chem.libretexts.org/@go/page/41380
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
2bd...a7ac8df6@9.110).
http://physics.stackexchange.com

12.2: Valence Bond Theory is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

12.2.6 https://chem.libretexts.org/@go/page/41380
12.3: Hybridization of Atomic Orbitals
Hybridization was introduced to explain molecular structure when the valence bond theory failed to correctly predict them. It is
experimentally observed that bond angles in organic compounds are close to 109°, 120°, or 180°. According to Valence Shell
Electron Pair Repulsion (VSEPR) theory, electron pairs repel each other and the bonds and lone pairs around a central atom are
generally separated by the largest possible angles.
In this section, we will use a model called valence bond theory to describe bonding in molecules. In this model, bonds are
considered to form from the overlapping of two atomic orbitals on different atoms, each orbital containing a single electron. In
looking at simple inorganic molecules such as H2 or HF, our present understanding of s and p atomic orbitals will suffice. In order
to explain the bonding in organic molecules, however, we will need to introduce the concept of hybrid orbitals.

Hybrid orbitals: sp3 hybridization and tetrahedral bonding


Now let’s turn to methane, the simplest organic molecule. Recall the valence electron configuration of the central carbon:

This picture, however, is problematic. How does the carbon form four bonds if it has only two half-filled p orbitals available for
bonding? A hint comes from the experimental observation that the four C-H bonds in methane are arranged with tetrahedral
geometry about the central carbon, and that each bond has the same length and strength. In order to explain this observation,
valence bond theory relies on a concept called orbital hybridization. In this picture, the four valence orbitals of the carbon (one 2s
and three 2p orbitals) combine mathematically (remember: orbitals are described by equations) to form four equivalent hybrid
orbitals, which are named sp3 orbitals because they are formed from mixing one s and three p orbitals. In the new electron
configuration, each of the four valence electrons on the carbon occupies a single sp3 orbital.

The sp3 hybrid orbitals, like the p orbitals of which they are partially composed, are oblong in shape, and have two lobes of
opposite sign. Unlike the p orbitals, however, the two lobes are of very different size. The larger lobes of the sp3 hybrids are
directed towards the four corners of a tetrahedron, meaning that the angle between any two orbitals is 109.5o.

This geometric arrangement makes perfect sense if you consider that it is precisely this angle that allows the four orbitals (and the
electrons in them) to be as far apart from each other as possible. This is simply a restatement of the Valence Shell Electron Pair
Repulsion (VSEPR) theory that you learned in General Chemistry: electron pairs (in orbitals) will arrange themselves in such a
way as to remain as far apart as possible, due to negative-negative electrostatic repulsion.
Each C-H bond in methane, then, can be described as an overlap between a half-filled 1s orbital in a hydrogen atom and the larger
lobe of one of the four half-filled sp3 hybrid orbitals in the central carbon. The length of the carbon-hydrogen bonds in methane is
1.09 Å (1.09 x 10-10 m).

12.3.1 https://chem.libretexts.org/@go/page/41381
While previously we drew a Lewis structure of methane in two dimensions using lines to denote each covalent bond, we can now
draw a more accurate structure in three dimensions, showing the tetrahedral bonding geometry. To do this on a two-dimensional
page, though, we need to introduce a new drawing convention: the solid / dashed wedge system. In this convention, a solid wedge
simply represents a bond that is meant to be pictured emerging from the plane of the page. A dashed wedge represents a bond that
is meant to be pictured pointing into, or behind, the plane of the page. Normal lines imply bonds that lie in the plane of the page.

This system takes a little bit of getting used to, but with practice your eye will learn to immediately ‘see’ the third dimension being
depicted.
interactive 3D model of the bonding in methane

Exercise
Imagine that you could distinguish between the four hydrogens in a methane molecule, and labeled them Ha through Hd. In the
images below, the exact same methane molecule is rotated and flipped in various positions. Draw the missing hydrogen atom
labels. (It will be much easier to do this if you make a model.)

In the ethane molecule, the bonding picture according to valence orbital theory is very similar to that of methane. Both carbons are
sp3-hybridized, meaning that both have four bonds arranged with tetrahedral geometry. The carbon-carbon bond, with a bond
length of 1.54 Å, is formed by overlap of one sp3 orbital from each of the carbons, while the six carbon-hydrogen bonds are formed
from overlaps between the remaining sp3 orbitals on the two carbons and the 1s orbitals of hydrogen atoms. All of these are sigma
bonds.

Because they are formed from the end-on-end overlap of two orbitals, sigma bonds are free to rotate. This means, in the case of
ethane molecule, that the two methyl (CH3) groups can be pictured as two wheels on a hub, each one able to rotate freely with
respect to the other.

12.3.2 https://chem.libretexts.org/@go/page/41381
In another module we will learn more about the implications of rotational freedom in sigma bonds, when we discuss the
‘conformation’ of organic molecules.
The sp3 bonding picture is also used to described the bonding in amines, including ammonia, the simplest amine. Just like the
carbon atom in methane, the central nitrogen in ammonia is sp3-hybridized. With nitrogen, however, there are five rather than four
valence electrons to account for, meaning that three of the four hybrid orbitals are half-filled and available for bonding, while the
fourth is fully occupied by a (non-bonding) pair of electrons.

The bonding arrangement here is also tetrahedral: the three N-H bonds of ammonia can be pictured as forming the base of a
trigonal pyramid, with the fourth orbital, containing the lone pair, forming the top of the pyramid.

Recall from your study of VSEPR theory in General Chemistry that the lone pair, with its slightly greater repulsive effect, ‘pushes’
the three N-H sigma bonds away from the top of the pyramid, meaning that the H-N-H bond angles are slightly less than
tetrahedral, at 107.3˚ rather than 109.5˚.
VSEPR theory also predicts, accurately, that a water molecule is ‘bent’ at an angle of approximately 104.5˚. It would seem logical,
then, to describe the bonding in water as occurring through the overlap of sp3-hybrid orbitals on oxygen with 1sorbitals on the two
hydrogen atoms. In this model, the two nonbonding lone pairs on oxygen would be located in sp3 orbitals.

Some experimental evidence, however, suggests that the bonding orbitals on the oxygen are actually unhybridized 2p orbitals rather
than sp3 hybrids. Although this would seem to imply that the H-O-H bond angle should be 90˚ (remember that p orbitals are
oriented perpendicular to one another), it appears that electrostatic repulsion has the effect of distorting this p-orbital angle to
104.5˚. Both the hybrid orbital and the nonhybrid orbital models present reasonable explanations for the observed bonding
arrangement in water, so we will not concern ourselves any further with the distinction.

Exercise
Draw, in the same style as the figures above, an orbital picture for the bonding in methylamine.
Solution

12.3.3 https://chem.libretexts.org/@go/page/41381
Formation of pi bonds - sp2 and sp hybridization
The valence bond theory, along with the hybrid orbital concept, does a very good job of describing double-bonded compounds such
as ethene. Three experimentally observable characteristics of the ethene molecule need to be accounted for by a bonding model:
1. Ethene is a planar (flat) molecule.
2. Bond angles in ethene are approximately 120o, and the carbon-carbon bond length is 1.34 Å, significantly shorter than the 1.54
Å single carbon-carbon bond in ethane.
3. There is a significant barrier to rotation about the carbon-carbon double bond.

Clearly, these characteristics are not consistent with an sp3 hybrid bonding picture for the two carbon atoms. Instead, the bonding in
ethene is described by a model involving the participation of a different kind of hybrid orbital. Three atomic orbitals on each
carbon – the 2s, 2px and 2py orbitals – combine to form three sp2 hybrids, leaving the 2pz orbital unhybridized.

The three sp2 hybrids are arranged with trigonal planar geometry, pointing to the three corners of an equilateral triangle, with
angles of 120°between them. The unhybridized 2pz orbital is perpendicular to this plane (in the next several figures, sp2 orbitals
and the sigma bonds to which they contribute are represented by lines and wedges; only the 2pz orbitals are shown in the 'space-
filling' mode).

another view
2
The carbon-carbon double bond in ethene consists of one sigma bond, formed by the overlap of two sp orbitals, and a second
bond, called a π (pi) bond, which is formed by the side-by-side overlap of the two unhybridized 2pz orbitals from each carbon.

another view
interactive model
The pi bond does not have symmetrical symmetry. Because they are the result of side-by-side overlap (rather then end-to-end
overlap like a sigma bond), pi bonds are not free to rotate. If rotation about this bond were to occur, it would involve disrupting the
side-by-side overlap between the two 2pz orbitals that make up the pi bond. The presence of the pi bond thus ‘locks’ the six atoms
of ethene into the same plane. This argument extends to larger alkene groups: in each case, the six atoms of the group form a single
plane.

12.3.4 https://chem.libretexts.org/@go/page/41381
Conversely, sigma bonds such as the carbon-carbon single bond in ethane (CH3CH3) exhibit free rotation, and can assume many
different conformations, or shapes.

Example
Circle the six atoms in the molecule below that are ‘locked’ into the same plane.

Exercise
What kinds of orbitals are overlapping in bonds a-d indicated below?

A similar picture can be drawn for the bonding in carbonyl groups, such as formaldehyde. In this molecule, the carbon is sp2-
hybridized, and we will assume that the oxygen atom is also sp2 hybridized. The carbon has three sigma bonds: two are formed by
overlap between two of its sp2 orbitals with the 1sorbital from each of the hydrogens, and the third sigma bond is formed by
overlap between the remaining carbon sp2 orbital and an sp2 orbital on the oxygen. The two lone pairs on oxygen occupy its other
two sp2 orbitals.

interactive 3D model
The pi bond is formed by side-by-side overlap of the unhybridized 2pz orbitals on the carbon and the oxygen. Just like in alkenes,
the 2pz orbitals that form the pi bond are perpendicular to the plane formed by the sigma bonds.
Finally, the hybrid orbital concept applies well to triple-bonded groups, such as alkynes and nitriles. Consider, for example, the
structure of ethyne (common name acetylene), the simplest alkyne.

This molecule is linear: all four atoms lie in a straight line. The carbon-carbon triple bond is only 1.20Å long. In the hybrid orbital
picture of acetylene, both carbons are sp-hybridized. In an sp-hybridized carbon, the 2s orbital combines with the 2px orbital to
form two sp hybrid orbitals that are oriented at an angle of 180°with respect to each other (eg. along the x axis). The 2py and 2pz
orbitals remain unhybridized, and are oriented perpendicularly along the y and z axes, respectively.

12.3.5 https://chem.libretexts.org/@go/page/41381
another view
The C-C sigma bond, then, is formed by the overlap of one sp orbital from each of the carbons, while the two C-H sigma bonds are
formed by the overlap of the second sp orbital on each carbon with a 1s orbital on a hydrogen. Each carbon atom still has two half-
filled 2py and 2pz orbitals, which are perpendicular both to each other and to the line formed by the sigma bonds. These two
perpendicular pairs of p orbitals form two pi bonds between the carbons, resulting in a triple bond overall (one sigma bond plus two
pi bonds).

another view
The hybrid orbital concept nicely explains another experimental observation: single bonds adjacent to double and triple bonds are
progressively shorter and stronger than ‘normal’ single bonds, such as the one in a simple alkane. The carbon-carbon bond in
ethane (structure A below) results from the overlap of two sp3 orbitals.

In alkene B, however, the carbon-carbon single bond is the result of overlap between an sp2 orbital and an sp3 orbital, while in
alkyne C the carbon-carbon single bond is the result of overlap between an sp orbital and an sp3 orbital. These are all single bonds,
but the bond in molecule C is shorter and stronger than the one in B, which is in turn shorter and stronger than the one in A.
The explanation here is relatively straightforward. An sp orbital is composed of one s orbital and one p orbital, and thus it has 50%
s character and 50% p character. sp2 orbitals, by comparison, have 33% s character and 67% p character, while sp3 orbitals have
25% s character and 75% p character. Because of their spherical shape, 2s orbitals are smaller, and hold electrons closer and
‘tighter’ to the nucleus, compared to 2p orbitals. Consequently, bonds involving sp + sp3 overlap (as in alkyne C) are shorter and
stronger than bonds involving sp2 + sp3 overlap (as in alkene B). Bonds involving sp3-sp3overlap (as in alkane A) are the longest
and weakest of the group, because of the 75% ‘p’ character of the hybrids.

Example
For each of the bonds indicated by arrows a-e in the figures below, describe the bonding picture by completing this sentence:
"The sigma bond indicated by this arrow is formed by the overlap of an ________ orbital of a _________atom and an
________orbital of a _______atom."

12.3.6 https://chem.libretexts.org/@go/page/41381
Solution

Contributors and Attributions


Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)

12.3: Hybridization of Atomic Orbitals is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

12.3.7 https://chem.libretexts.org/@go/page/41381
12.4: Electronegativity and Dipole Moment
Within a group of the periodic table, bond lengths tend to increase with increasing atomic number Z . Consider the Group 17
elements:
F2 d = 141.7 pm

C l2 d = 199.1 pm

Br2 d = 228.6 pm

I2 d = 266.9 pm

which corresponds to an increased valence shell size, hence increased electron-electron repulsion. An important result from
experiment, which has been corroborated by theory, is that bond lengths tend not to vary much from molecule to molecule. Thus, a
C H bond will have roughly the same value in methane, C H as it will in aspirin, C H O .
4 9 8 4

Bond dissociation energies. ΔE measured in kJ/mol, measure the energy required to break a mole of a particular kind of bond.
d

A similar periodic trend exists for bond dissociation energies. Consider the hydrogen halides:
HF ΔEd = 565 kJ/mol d = 0.926 pm

HC l ΔEd = 429 kJ/mol d = 128.4 pm

H Br ΔEd = 363 kJ/mol d = 142.4 pm

HI ΔEd = 295 kJ/mol d = 162.0 pm

Thus, as bond lengths increase with increasing Z , there is a corresponding decrease in the bond dissociation energy.
CC bonds are an exception to the the rule of constancy of bond lengths across different molecules. Because C C bonds can be
single, double, or triple bonds, some differences can occur. For example, consider the C C bond in the molecules ethane (C H ) , 2 6

ethylene (C H ) and acetylene (C H ) :


2 4 2 2

C2 H6 (single) d = 1.536 A ΔEd = 345 kJ/mol

C2 H4 (double) d = 133.7 pm ΔEd = 612 kJ/mol

C2 H2 (triple) d = 126.4 pm ΔEd = 809 kJ/mol

The greater the bond order, i.e., number of shared electron pairs, the greater the dissociation energy. The same will be true for any
kind of bond that can come in such different ``flavors'', e.g., N N bonds, OO bonds, N O bonds, C O bonds, etc.

Figure 12.4.1 : The Electron Distribution in a Nonpolar Covalent Bond, a Polar Covalent Bond, and an Ionic Bond Using Lewis
Electron Structures. In a purely covalent bond (a), the bonding electrons are shared equally between the atoms. In a purely ionic
bond (c), an electron has been transferred completely from one atom to the other. A polar covalent bond (b) is intermediate between
the two extremes: the bonding electrons are shared unequally between the two atoms, and the electron distribution is asymmetrical
with the electron density being greater around the more electronegative atom. Electron-rich (negatively charged) regions are shown
in blue; electron-poor (positively charged) regions are shown in red.
The two idealized extremes of chemical bonding:
1. Ionic bonding—in which one or more electrons are transferred completely from one atom to another, and the resulting ions are
held together by purely electrostatic forces—and
2. Covalent bonding, in which electrons are shared equally between two atoms.
Most compounds, however, have polar covalent bonds, which means that electrons are shared unequally between the bonded
atoms. Figure 12.4.1 compares the electron distribution in a polar covalent bond with those in an ideally covalent and an ideally
ionic bond. Recall that a lowercase Greek delta (δ ) is used to indicate that a bonded atom possesses a partial positive charge,

12.4.1 https://chem.libretexts.org/@go/page/41382
indicated by δ , or a partial negative charge, indicated by δ , and a bond between two atoms that possess partial charges is a polar
+ −

bond.
Most real chemical bonds in nature are neither truly covalent nor truly ionic. Only homonuclear bonds are truly covalent, and
nearly perfect ionic bonds can form between group I and group VII elements, for example, KF. Generally, however, bonds are
partially covalent and partially ionic, meaning that there is partial transfer of electrons between atoms and partial sharing of
electrons. To quantify how much ionic character (and how much covalent character) a bond possesses, electronegativity differences
between the atoms in the bond can be used. In 1936, Linus Pauling came up a method for estimating atomic electronegativities
forms the basis of our understanding of electronegativity today.

Only homonuclear bonds are truly covalent, or as covalent as a bond can get. As discussed in Section 12.2, a quantum-
mechanical treatment has shown that the two ionic structures (e.g., H H and H H for H ) also contribute via a
+ − − +
2

resonance with the covalent structure H − H . However, this is often to a small extent, for example in H , each ionic structure
2

contributes only ~2% to the bonding of the molecule.

Pauling Electronegativity
Linus Pauling described electronegativity as “the power of an atom in a molecule to attract electrons to itself.” Basically, the
electronegativity of an atom is a relative value of that atom's ability to attract election density toward itself when it bonds to another
atom. The higher the electronegative of an element, the more that atom will attempt to pull electrons towards itself and away from
any atom it bonds to. The main properties of an atom dictate it's electronegativity are it's atomic number as well as its atomic
radius. The trend for electronegativity is to increase as you move from left to right and bottom to top across the periodic table. This
means that the most electronegative atom is Fluorine and the least electronegative is Francium.
Recall the Mulliken's method was based on the arithmetic average of the first ionization energy I E and the electron affinity EA.
1

Both of these energies are properties of individual atoms, hence this method is appealing in its simplicity. However, there is no
information about bonding in the Mulliken method. Pauling's method includes such information, and hence is a more effective
approach.
To see how the Pauling method works, consider a diatomic AB, which is polar covalent. Let ΔE and ΔE be the dissociation
AA BB

energies of the diatomics A and B , respectively. Since A and B are purely covalent bonds, these two dissociation energies can
2 2 2 2

be used to estimate the pure covalent contribution to the bond AB. Pauling proposed the geometric mean of ΔE and ΔE , AA BB

this being more sensitive to large differences between these energies than the arithmetic average:
−−−−−−−−−−
pure covalent contribution = √ΔEAA ΔEBB (12.4.1)

If ΔE AB is the true bond dissociation energy, then the difference


−−−−−−−−−−
ΔEAB − √ΔEAA ΔEBB (12.4.2)

is a measure of the ionic contribution. Let us define this difference to be Δ:


−−−−−−−−−−
Δ = ΔEAB − √ΔEAA ΔEBB (12.4.3)

Then Pauling defined the electronegativity difference χ A − χB between atoms A and B to be




χA − χB = 0.102 √Δ (12.4.4)

where Δ is measured in kJ/mol, and the constant 0.102 has units mol /kJ , so that the electronegativity difference is
1/2 1/2

dimensionless. Thus, with some extra input information, he was able to generate a table of atomic electronegativities that are still
used today and is Tablated in Table A2 (Figure 12.4.2).

12.4.2 https://chem.libretexts.org/@go/page/41382
Figure 12.4.2 : Pauling Electronegativity Values of the s-, p-, d-, and f-Block Elements. Values for most of the actinides are
approximate. Elements for which no data are available are shown in gray. Source: Data from L. Pauling, The Nature of the
Chemical Bond, 3rd ed. (1960).
To use the electronegativities to estimate degree of ionic character, simply compute the absolute value of the difference for the two
atoms in the bond. As an example, consider again the hydrogen halides:

HF | χF − χH | = 1.78

HC l | χC l − χH | = 0.96

H Br | χBr − χH | = 0.76

HI | χI − χH | = 0.46

As the electronegativity difference decreases, so does the ionic character of the bond. Hence its covalent character increases. It is
possible to predict whether a given bond will be non-polar, polar covalent, or ionic based on the electronegativity difference, since
the greater the difference, the more polar the bond (Figure 12.4.3)

Electronegativity difference, ΔχP Bond

Δχ < 0.4 covalent

0.4 < Δχ < 1.7 polar covalent

Δχ > 1.7 ionic

Figure 12.4.3 : Range of Bonds Polarity based on difference of Electronegativites

As the electronegativity difference decreases, so does the ionic character of the bond.
Hence its covalent character increases.
Dipole moment
In a nearly perfect ionic bond, such as KF , where electron transfer is almost complete, representing the molecule as
+ −
K F (12.4.5)

is a very good approximation, since the charge on the potassium will be approximately 1e and the charge on the fluorine will be
approximately −1e . For a polar covalent bond, such as H F , in which only partial charge transfer occurs, a more accurate
representation would be

12.4.3 https://chem.libretexts.org/@go/page/41382
+δ −δ
H F (12.4.6)

where δ , expressed in units of \(e\), is known as a partial charge. It suggests that a fraction of an electron is transferred, although
the reality is that there is simply a little more electron density on the more electronegative atom and a little less on the
electropositive atom.
How much charge is actually transferred can be quantified by studying the electric dipole moment of the bond, which is a quantity
that can be measured experimentally. The electric dipole moment for a diatomic with charges Q = Q = δe and Q = −Q = −δe
1 2

on atoms 1 and 2, respectively, is


μ = Q1 r1 + Q2 r2

= Q r1 − Q r2

= Q(r1 − r2 )

Hence, the magnitude of the dipole moment is


μ = |μ| = Q| r1 − r2 | = QR (12.4.7)

where R is the bond length. As an example, consider H F , which has a partial charge on H of 0.41 e, which means δ = 0.41, and

a bond length of 0.926 A . Thus, the magnitude of the dipole moment is


−19 −10 −30
|μ| = 0.41(1.602 × 10 C )(0.926 × 10 m) = 6.08 × 10 C ⋅m (12.4.8)

Thus, the units of the dipole moment are Coulomb-meters. However, as this example makes clear, this is a very large unit and
awkward to work with for molecules. A more convenient unit is the Debye (D), defined to be
−30
1 D = 3.336 × 10 Coulomb ⋅ meters (12.4.9)

Historically, the Debye was defined in terms of the dipole moment resulting from two equal charges of opposite sign and
separated by 1 Ångstrom (10 m ) as 4.801 D from Equation 12.4.7. This value arises from
−10

−19 −10
(1.602 × 10 )(1 × 10 )

−30
3.336 × 10

where
−30
D = 3.336 × 10 C m

or
29
1 C m = 2.9979 × 10 D

Thus, for a diatomic with partial charges +δ and −δ, the dipole moment in D is given by

δ ∗ R(A)
μ(D) = (12.4.10)

−1
0.2082 A D

and the percent ionic character is defined in terms of the partial charge δ by
percent ionic character = 100% ∗ δ (12.4.11)

Typical dipole moments for simple diatomic molecules are in the range of 0 to 11 D (Table 12.4.1 ). Equation 12.4.11 can be
expressed differently in terms of the expected dipole assuming a full charge separation (μ ionic ) compared to the experimental
dipole moment (μ )exp

μexp
percent ionic character = 100% ∗ (12.4.12)
μionic

12.4.4 https://chem.libretexts.org/@go/page/41382
Example 12.4.1: KBr

Calculate a theoretical dipole moment for the KBr molecule, assuming opposite charges of one fundamental unit located at
each nucleus, and hence the percentage ionic character of KBr.
Solution
From Table 12.4.1, the observed dipole moment of KBr is given as 10.41 D, (3.473 x 10-29 Coulomb-meters), which being
close to the upper level of 11 indicates that it is a highly polar molecule. The interatomic distance between K+ and Br- is 282
pm. From this it is possible to calculate a theoretical dipole moment for the KBr molecule, assuming opposite charges of one
fundamental unit located at each nucleus, and hence the percentage ionic character of KBr.
Dipole moment
μ = q × e × d (in Coulomb-meters) (12.4.13)

q = 1 for complete separation of unit charge


e = 1.602 x 10-19 C
d = 2.82 x 10-10 m for KBr (282 pm)
Hence calculated ionic dipole moment is
−19 −10 −29
μ = (1)(1.602 × 10 )(2.82 × 10 ) = 4.518 × 10 C m = 13.54 D
KBr

The observed dipole moment is


−29
μKBr = 3.473 × 10 C m = 10.41 D

the % ionic character from Equation 12.4.12 is


−29
3.473 × 10 10.41 D
KBr = × 100% = × 100% = 76.87%
−29
4.518 × 10 13.54 D

and the % covalent character is therefore about 23% (100% - 77%).


Given the observed dipole moment is 10.41 D (3.473 x 10-29) it is possible to estimate the charge distribution from the same
equation by now solving for q.
Dipole moment μ = q * e * d Coulomb metre, but since q is no longer 1 we can substitute in values for μ and d to obtain an
estimate for it.
q = μ /(e * d) = 3.473 x 10-29 / (1.602 x 10-19 * 2.82 x 10-10)
thus q = 3.473 x 10-29 / (4.518 x 10-29) = 0.77 and the δ- and δ+ are -0.8 and +0.8 respectively.

Example 12.4.2: NaCl

In the gas phase, NaCl has a dipole moment of 9.001 D and an Na–Cl distance of 236.1 pm. Calculate the percent ionic
character in NaCl.
Given: chemical species, dipole moment, and internuclear distance
Asked for: percent ionic character
Strategy:
A Compute the charge on each atom using the information given and Equation 8.4.2.
B Find the percent ionic character from the ratio of the actual charge to the charge of a single electron.
Solution:
A The charge on each atom is given by
−30
μ 3.3356 × 10 C ⋅ m 1 1 pm
−19
Q = = 9.001 D ( )( )( ) = 1.272 × 10 C (12.4.14)
−12
r 236.1 pm 10 m
1 D

12.4.5 https://chem.libretexts.org/@go/page/41382
Thus NaCl behaves as if it had charges of 1.272 × 10−19 C on each atom separated by 236.1 pm.
B The percent ionic character is given by the ratio of the actual charge to the charge of a single electron (the charge expected
for the complete transfer of one electron):
−19
1.272 × 10 C
% ionic character = ( ) (100) = 79.39% ≃ 79% (12.4.15)
−19
1.6022 × 10 C

Exercise 12.4.2: AgCl

In the gas phase, silver chloride (AgCl) has a dipole moment of 6.08 D and an Ag–Cl distance of 228.1 pm. What is the percent
ionic character in silver chloride?
Answer
55.5%

Table 12.4.1 : Bond characteristics of select diatomics


diatomic Δχ %ionic bond dist (pm) μ
exp
(D) μ
ionic
(D)

Cl2 0.0 0.0 200 0.00 9.60

IBr 0.3 5.9 247 0.70 11.86

HI 0.4 5.7 161 0.44 7.73

ICl 0.5 5.4 232 0.60 11.14

HBr 0.7 12.1 141 0.82 6.77

HCl 0.9 17.7 127 1.08 6.10

ClF 1.0 11.2 163 0.88 7.83

BrF 1.2 15.1 178 1.29 8.55

LiI 1.5 65.0 238 7.43 11.43

HF 1.9 41.2 92 1.82 4.42

LiBr 1.8 69.8 217 7.27 10.42

KI 1.7 73.7 305 10.80 14.65

LiCl 2.0 73.5 202 7.13 9.70

KBr 2.0 76.9 282 10.41 13.54

NaCl 2.1 79.4 236 9.00 11.33

KCl 2.2 80.1 267 10.27 12.82

CsCl 2.3 74.6 291 10.42 13.97

LiF 3.0 86.7 152 6.33 7.30

KF 3.2 82.5 217 8.60 10.42

CsF 3.3 64.4 255 7.88 12.25

Pauling proposed an empirical relationship (instead of the defintion in Equation 12.4.12) which relates the percent ionic character
in a bond to the electronegativity difference.
2
(Δχ/2)
percent ionic character = 100% (1 − e ) (12.4.16)

This is shown as the curve in Figure 12.4.4 and is compared to the values for some diatomic molecules calculated from observed
and calculated dipole moments.

12.4.6 https://chem.libretexts.org/@go/page/41382
Figure 12.4.4 : A Plot of the Percent Ionic Character of a Bond as Determined from Measured Dipole Moments versus the
Difference in Electronegativity of the Bonded Atoms.In the gas phase, even CsF, which has the largest possible difference in
electronegativity between atoms, is not 100% ionic. Solid CsF, however, is best viewed as 100% ionic because of the additional
electrostatic interactions in the lattice.

As an example, consider H F again, for which δ = 0.41. The bond length is R = 0.926 A . Thus, its dipole moment will be

0.41 ∗ 0.926 A
μ(D) = = 1.82D (12.4.17)

−1
0.2082 A D

and its percent ionic character is 41%.

Example 12.4.3: Bond Polarity

Without consulting the table of electronegativities (use the periodic table), arrange the following bonds in order of decreasing
polarity:
B—Cl, Ba—Cl, Be—Cl, Br—Cl, Cl—Cl.
Solution
We first need to arrange the elements in order of increasing electronegativity. Since the electronegativity increases in going up
a column of the periodic table, we have the following relationships:

Ba < Be and Br < Cl

Also since the electronegativity increases across the periodic table, we have

Be < B

Since B is a group III element on the borderline between metals and non-metals, we easily guess that

B < Br

which gives us the complete order

Ba < Be < B < Br < Cl

Among the bonds listed, therefore, the Ba—Cl bond corresponds to the largest difference in electronegativity, i.e., to the most
nearly ionic bond. The order of bond polarity is thus

Ba—Cl > Br—Cl > B—Cl > Br—Cl > Cl—Cl

12.4.7 https://chem.libretexts.org/@go/page/41382
where the final bond, Cl—Cl,is, of course, purely covalent.

Contributors
Mark Tuckerman (New York University)
Ed Vitz (Kutztown University), John W. Moore (UW-Madison), Justin Shorb (Hope College), Xavier Prat-Resina (University of
Minnesota Rochester), Tim Wendorff, and Adam Hahn.
Prof. Robert J. Lancashire (The Department of Chemistry, University of the West Indies)

12.4: Electronegativity and Dipole Moment is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.

12.4.8 https://chem.libretexts.org/@go/page/41382
12.5: Molecular Orbital Theory
None of the approaches we have described so far can adequately explain why some compounds are colored and others are not, why
some substances with unpaired electrons are stable, and why others are effective semiconductors. These approaches also cannot
describe the nature of resonance. Such limitations led to the development of a new approach to bonding in which electrons are not
viewed as being localized between the nuclei of bonded atoms but are instead delocalized throughout the entire molecule. Just as
with the valence bond theory, the approach we are about to discuss is based on a quantum mechanical model.
Previously, we described the electrons in isolated atoms as having certain spatial distributions, called orbitals, each with a
particular orbital energy. Just as the positions and energies of electrons in atoms can be described in terms of atomic orbitals
(AOs), the positions and energies of electrons in molecules can be described in terms of molecular orbitals (MOs) A particular
spatial distribution of electrons in a molecule that is associated with a particular orbital energy.—a spatial distribution of electrons
in a molecule that is associated with a particular orbital energy. As the name suggests, molecular orbitals are not localized on a
single atom but extend over the entire molecule. Consequently, the molecular orbital approach, called molecular orbital theory is a
delocalized approach to bonding.
Although the molecular orbital theory is computationally demanding, the principles on which it is based are similar to those we
used to determine electron configurations for atoms. The key difference is that in molecular orbitals, the electrons are allowed to
interact with more than one atomic nucleus at a time. Just as with atomic orbitals, we create an energy-level diagram by listing the
molecular orbitals in order of increasing energy. We then fill the orbitals with the required number of valence electrons according to
the Pauli principle. This means that each molecular orbital can accommodate a maximum of two electrons with opposite spins.

The Hydrogen Molecule-Ion


Molecular orbital theory is a conceptual extension of the orbital model, which was so successfully applied to atomic structure. As
was once playfully remarked, "a molecule is nothing more than an atom with more nuclei." This may be overly simplistic, but we
do attempt, as far as possible, to exploit analogies with atomic structure. Our understanding of atomic orbitals began with the exact
solutions of a prototype problem – the hydrogen atom. We will begin our study of homonuclear diatomic molecules beginning with
another exactly solvable prototype, the hydrogen molecule-ion H . This species actually has a transient existence in electrical
+

discharges through hydrogen gas and has been detected by mass spectrometry. It also has been detected in outer space. The
Schrödinger equation for H can be solved exactly within the Born-Oppenheimer approximation. For fixed internuclear distance R,
+

this reduces to a problem of one electron in the field of two protons, designated A and B. We can write
1 2
1 1 1
{− ∇ − + } ψ(r) = Eψ(r) (1)
2 rA rB R

where rA and rB are the distances from the electron to protons A and B, respectively. This equation was solved by Burrau (1927),
after separating the variables in prolate spheroidal coordinates.

Molecular Orbitals Involving Only ns Atomic Orbitals


We begin our discussion of molecular orbitals with the simplest molecule, H2, formed from two isolated hydrogen atoms, each with
a 1s electron configuration. As discussed before, electrons can behave like waves. In the molecular orbital approach, the
1

overlapping atomic orbitals are described by mathematical equations called wave functions. The 1s atomic orbitals on the two
hydrogen atoms interact to form two new molecular orbitals, one produced by taking the sum of the two H 1s wave functions, and
the other produced by taking their difference:
M O(1) = AO(atom A) + AO(atomB)
(12.5.1)
M O(1) = AO(atom A) − AO(atomB)

The molecular orbitals created from the sum and the difference of two wavefunctions (atomic orbitals) from Equation 12.5.1 are
called Linear Combinations of Atomic Orbitals (LCAOs). A molecule must have as many molecular orbitals as there are atomic
orbitals.
Adding two atomic orbitals corresponds to constructive interference between two waves, thus reinforcing their intensity; the
internuclear electron probability density is increased. The molecular orbital corresponding to the sum of the two H 1s orbitals is
called a σ1s combination (pronounced “sigma one ess”) (part (a) and part (b) in Figure 12.5.1). In a sigma (σ) orbital, A bonding
molecular orbital in which the electron density along the internuclear axis and between the nuclei has cylindrical symmetry, the

12.5.1 https://chem.libretexts.org/@go/page/41383
electron density along the internuclear axis and between the nuclei has cylindrical symmetry; that is, all cross-sections
perpendicular to the internuclear axis are circles. The subscript 1s denotes the atomic orbitals from which the molecular orbital was
derived: The ≈ sign is used rather than an = sign because we are ignoring certain constants that are not important to our argument.

Figure 12.5.1 : Molecular Orbitals for the H2 Molecule. (a) This diagram shows the formation of a bonding σ1s molecular orbital for
H2 as the sum of the wave functions (Ψ) of two H 1s atomic orbitals. (b) This plot of the square of the wave function (Ψ2) for the
bonding σ1s molecular orbital illustrates the increased electron probability density between the two hydrogen nuclei. (Recall that
the probability density is proportional to the square of the wave function.) (c) This diagram shows the formation of an antibonding
σ

1s
molecular orbital for H2 as the difference of the wave functions (Ψ) of two H 1s atomic orbitals. (d) This plot of the square of
the wave function (Ψ2) for the σ antibonding molecular orbital illustrates the node corresponding to zero electron probability

1s

density between the two hydrogen nuclei.


σ1s ≈ 1s (A) + 1s (B) (12.5.2)

Conversely, subtracting one atomic orbital from another corresponds to destructive interference between two waves, which reduces
their intensity and causes a decrease in the internuclear electron probability density (part (c) and part (d) in Figure 12.5.1 ). The
resulting pattern contains a node where the electron density is zero. The molecular orbital corresponding to the difference is called
σ

1s
(“sigma one ess star”). In a sigma star (σ*) orbital An antibonding molecular orbital in which there is a region of zero electron
probability (a nodal plane) perpendicular to the internuclear axis., there is a region of zero electron probability, a nodal plane,
perpendicular to the internuclear axis:

σ ≈ 1s (A) − 1s (B) (12.5.3)
1s

A molecule must have as many molecular orbitals as there are atomic orbitals.

The electron density in the σ1s molecular orbital is greatest between the two positively charged nuclei, and the resulting electron–
nucleus electrostatic attractions reduce repulsions between the nuclei. Thus the σ1s orbital represents a bonding molecular orbital. A
molecular orbital that forms when atomic orbitals or orbital lobes with the same sign interact to give increased electron probability
between the nuclei due to constructive reinforcement of the wave functions. In contrast, electrons in the σ orbital are generally

1s

found in the space outside the internuclear region. Because this allows the positively charged nuclei to repel one another, the σ ⋆
1s

orbital is an antibonding molecular orbital (a molecular orbital that forms when atomic orbitals or orbital lobes of opposite sign
interact to give decreased electron probability between the nuclei due to destructive reinforcement of the wave functions).

Antibonding orbitals contain a node perpendicular to the internuclear axis; bonding


orbitals do not.

Energy-Level Diagrams
Because electrons in the σ1s orbital interact simultaneously with both nuclei, they have a lower energy than electrons that interact
with only one nucleus. This means that the σ1s molecular orbital has a lower energy than either of the hydrogen 1s atomic orbitals.

12.5.2 https://chem.libretexts.org/@go/page/41383
Conversely, electrons in the σ orbital interact with only one hydrogen nucleus at a time. In addition, they are farther away from

1s

the nucleus than they were in the parent hydrogen 1s atomic orbitals. Consequently, the σ molecular orbital has a higher energy

1s

than either of the hydrogen 1s atomic orbitals. The σ1s (bonding) molecular orbital is stabilized relative to the 1s atomic orbitals,
and the σ (antibonding) molecular orbital is destabilized. The relative energy levels of these orbitals are shown in the energy-

1s

level diagram (a schematic drawing that compares the energies of the molecular orbitals (bonding, antibonding, and nonbonding)
with the energies of the parent atomic orbitals) in Figure 12.5.2

Figure 12.5.2 : Molecular Orbital Energy-Level Diagram for H2. The two available electrons (one from each H atom) in this
diagram fill the bonding σ1 s molecular orbital. Because the energy of the σ1s molecular orbital is lower than that of the two H 1s
atomic orbitals, the H2 molecule is more stable (at a lower energy) than the two isolated H atoms.

A bonding molecular orbital is always lower in energy (more stable) than the component atomic orbitals, whereas an
antibonding molecular orbital is always higher in energy (less stable).

To describe the bonding in a homonuclear diatomic molecule (a molecule that consists of two atoms of the same element) such as
H2, we use molecular orbitals; that is, for a molecule in which two identical atoms interact, we insert the total number of valence
electrons into the energy-level diagram (Figure 12.5.2). We fill the orbitals according to the Pauli principle and Hund’s rule: each
orbital can accommodate a maximum of two electrons with opposite spins, and the orbitals are filled in order of increasing energy.
Because each H atom contributes one valence electron, the resulting two electrons are exactly enough to fill the σ1s bonding
molecular orbital. The two electrons enter an orbital whose energy is lower than that of the parent atomic orbitals, so the H2
molecule is more stable than the two isolated hydrogen atoms. Thus molecular orbital theory correctly predicts that H2 is a stable
molecule. Because bonds form when electrons are concentrated in the space between nuclei, this approach is also consistent with
our earlier discussion of electron-pair bonds.

Bond Order in Molecular Orbital Theory


In the Lewis electron structures, the number of electron pairs holding two atoms together was called the bond order. In the
molecular orbital approach, bond order One-half the net number of bonding electrons in a molecule. is defined as one-half the net
number of bonding electrons:
number of bonding electrons − number of antibonding electrons
bond order = (12.5.4)
2

To calculate the bond order of H2, we see from Figure 12.5.2 that the σ1s (bonding) molecular orbital contains two electrons, while
the σ (antibonding) molecular orbital is empty. The bond order of H2 is therefore

1s

2 −0
=1 (12.5.5)
2

This result corresponds to the single covalent bond predicted by Lewis dot symbols. Thus molecular orbital theory and the Lewis
electron-pair approach agree that a single bond containing two electrons has a bond order of 1. Double and triple bonds contain
four or six electrons, respectively, and correspond to bond orders of 2 and 3.
We can use energy-level diagrams such as the one in Figure 12.5.2 to describe the bonding in other pairs of atoms and ions where n
= 1, such as the H2+ ion, the He2+ ion, and the He2 molecule. Again, we fill the lowest-energy molecular orbitals first while being
sure not to violate the Pauli principle or Hund’s rule.

12.5.3 https://chem.libretexts.org/@go/page/41383
Figure 12.5.3 : Molecular Orbital Energy-Level Diagrams for Diatomic Molecules with Only 1 s Atomic Orbitals. (a) The H2+ ion,
(b) the He2+ ion, and (c) the He2 molecule are shown here.
Part (a) in Figure 12.5.3 shows the energy-level diagram for the H2+ ion, which contains two protons and only one electron. The
single electron occupies the σ1s bonding molecular orbital, giving a (σ1s)1 electron configuration. The number of electrons in an
orbital is indicated by a superscript. In this case, the bond order is (1-0)/2=1/2 Because the bond order is greater than zero, the H2+
ion should be more stable than an isolated H atom and a proton. We can therefore use a molecular orbital energy-level diagram and
the calculated bond order to predict the relative stability of species such as H2+. With a bond order of only 1/2 the bond in H2+
should be weaker than in the H2 molecule, and the H–H bond should be longer. As shown in Table 12.5.1, these predictions agree
with the experimental data.
Part (b) in Figure 12.5.3 is the molecular orbital energy-level diagram for He2+. This ion has a total of three valence electrons.
Because the first two electrons completely fill the σ1s molecular orbital, the Pauli principle states that the third electron must be in
2 1
the σ antibonding orbital, giving a (σ ) (σ ) electron configuration. This electron configuration gives a bond order of (2-

1s 1s

1s

1)/2=1/2. As with H2+, the He2+ ion should be stable, but the He–He bond should be weaker and longer than in H2. In fact, the He2+
ion can be prepared, and its properties are consistent with our predictions (Table 12.5.1).
Table 12.5.1 : Molecular Orbital Electron Configurations, Bond Orders, Bond Lengths, and Bond Energies for some Simple Homonuclear
Diatomic Molecules and Ions
Molecule or Ion Electron Configuration Bond Order Bond Length (pm) Bond Energy (kJ/mol)

H2+ (σ1s)1 1/2 106 269

H2 (σ1s)2 1 74 436

He2+
1
2
(σ1s ) (σ
1s

) 1/2 108 251
2
He2 2
(σ1s ) (σ
1s

) 0 not observed not observed

Finally, we examine the He2 molecule, formed from two He atoms with 1s2 electron configurations. Part (c) in Figure 12.5.3 is the
molecular orbital energy-level diagram for He2. With a total of four valence electrons, both the σ1s bonding and σ antibonding ⋆
1s
2 1
orbitals must contain two electrons. This gives a (σ ) (σ ) electron configuration, with a predicted bond order of (2 − 2) ÷ 2 =
1s

1s

0, which indicates that the He2 molecule has no net bond and is not a stable species. Experiments show that the He2 molecule is
actually less stable than two isolated He atoms due to unfavorable electron–electron and nucleus–nucleus interactions.
In molecular orbital theory, electrons in antibonding orbitals effectively cancel the stabilization resulting from electrons in bonding
orbitals. Consequently, any system that has equal numbers of bonding and antibonding electrons will have a bond order of 0, and it
is predicted to be unstable and therefore not to exist in nature. In contrast to Lewis electron structures and the valence bond
approach, molecular orbital theory is able to accommodate systems with an odd number of electrons, such as the H2+ ion.

Note
In contrast to Lewis electron structures and the valence bond approach, molecular orbital theory can accommodate systems
with an odd number of electrons.

Example 12.5.1

Use a molecular orbital energy-level diagram, such as those in Figure 12.5.3, to predict the bond order in the He22+ ion. Is this
a stable species?
Given: chemical species
Asked for: molecular orbital energy-level diagram, bond order, and stability

12.5.4 https://chem.libretexts.org/@go/page/41383
Strategy:
A. Combine the two He valence atomic orbitals to produce bonding and antibonding molecular orbitals. Draw the molecular
orbital energy-level diagram for the system.
B. Determine the total number of valence electrons in the He22+ ion. Fill the molecular orbitals in the energy-level diagram
beginning with the orbital with the lowest energy. Be sure to obey the Pauli principle and Hund’s rule while doing so.
C. Calculate the bond order and predict whether the species is stable.
Solution:
A Two He 1s atomic orbitals combine to give two molecular orbitals: a σ1s bonding orbital at lower energy than the atomic
orbitals and a σ antibonding orbital at higher energy. The bonding in any diatomic molecule with two He atoms can be

1s

described using the following molecular orbital diagram:

B The He22+ ion has only two valence electrons (two from each He atom minus two for the +2 charge). We can also view
He22+ as being formed from two He+ ions, each of which has a single valence electron in the 1s atomic orbital. We can now fill
the molecular orbital diagram:

The two electrons occupy the lowest-energy molecular orbital, which is the bonding (σ1s) orbital, giving a (σ1s)2 electron
configuration. To avoid violating the Pauli principle, the electron spins must be paired. C So the bond order is
2 −0
=1
2

He22+ is therefore predicted to contain a single He–He bond. Thus it should be a stable species.

Exercise 12.5.1

Use a molecular orbital energy-level diagram to predict the valence-electron configuration and bond order of the H22− ion. Is
this a stable species?
Answer:

H22− has a valence electron configuration of (σ


2
1s )
2

1s

) with a bond order of 0. It is therefore predicted to be unstable.

So far, our discussion of molecular orbitals has been confined to the interaction of valence orbitals, which tend to lie farthest from
the nucleus. When two atoms are close enough for their valence orbitals to overlap significantly, the filled inner electron shells are
largely unperturbed; hence they do not need to be considered in a molecular orbital scheme. Also, when the inner orbitals are
completely filled, they contain exactly enough electrons to completely fill both the bonding and antibonding molecular orbitals that
arise from their interaction. Thus the interaction of filled shells always gives a bond order of 0, so filled shells are not a factor when
predicting the stability of a species. This means that we can focus our attention on the molecular orbitals derived from valence
atomic orbitals.
A molecular orbital diagram that can be applied to any homonuclear diatomic molecule with two identical alkali metal atoms (Li2
and Cs2, for example) is shown in part (a) in Figure 12.5.4, where M represents the metal atom. Only two energy levels are
important for describing the valence electron molecular orbitals of these species: a σns bonding molecular orbital and a σ*ns
antibonding molecular orbital. Because each alkali metal (M) has an ns1 valence electron configuration, the M2 molecule has two

12.5.5 https://chem.libretexts.org/@go/page/41383
valence electrons that fill the σns bonding orbital. As a result, a bond order of 1 is predicted for all homonuclear diatomic species
formed from the alkali metals (Li2, Na2, K2, Rb2, and Cs2). The general features of these M2 diagrams are identical to the diagram
for the H2 molecule in Figure 12.5.2. Experimentally, all are found to be stable in the gas phase, and some are even stable in
solution.

Figure 12.5.4 : Molecular Orbital Energy-Level Diagrams for Alkali Metal and Alkaline Earth Metal Diatomic (M2) Molecules. (a)
For alkali metal diatomic molecules, the two valence electrons are enough to fill the σ ns (bonding) level, giving a bond order of 1.
(b) For alkaline earth metal diatomic molecules, the four valence electrons fill both the σns (bonding) and the σns* (nonbonding)
levels, leading to a predicted bond order of 0.
Similarly, the molecular orbital diagrams for homonuclear diatomic compounds of the alkaline earth metals (such as Be2), in which
each metal atom has an ns2 valence electron configuration, resemble the diagram for the He2 molecule in part (c) in Figure 12.5.3
As shown in part (b) in Figure 12.5.4, this is indeed the case. All the homonuclear alkaline earth diatomic molecules have four
valence electrons, which fill both the σns bonding orbital and the σns* antibonding orbital and give a bond order of 0. Thus Be2,
Mg2, Ca2, Sr2, and Ba2 are all expected to be unstable, in agreement with experimental data.In the solid state, however, all the
alkali metals and the alkaline earth metals exist as extended lattices held together by metallic bonding. At low temperatures, Be is 2

stable.

Example 12.5.2

Use a qualitative molecular orbital energy-level diagram to predict the valence electron configuration, bond order, and likely
existence of the Na2− ion.
Given: chemical species
Asked for: molecular orbital energy-level diagram, valence electron configuration, bond order, and stability
Strategy:
A. Combine the two sodium valence atomic orbitals to produce bonding and antibonding molecular orbitals. Draw the
molecular orbital energy-level diagram for this system.
B. Determine the total number of valence electrons in the Na2− ion. Fill the molecular orbitals in the energy-level diagram
beginning with the orbital with the lowest energy. Be sure to obey the Pauli principle and Hund’s rule while doing so.
C. Calculate the bond order and predict whether the species is stable.
Solution:
A Because sodium has a [Ne]3s1 electron configuration, the molecular orbital energy-level diagram is qualitatively identical to
the diagram for the interaction of two 1s atomic orbitals.
B The Na2− ion has a total of three valence electrons (one from each Na atom and one for the negative charge), resulting in a
1
filled σ3s molecular orbital, a half-filled σ3s* and a (σ ) (σ ) electron configuration.
3s
2 ⋆
3s

C The bond order is (2-1)÷2=1/2 With a fractional bond order, we predict that the Na2− ion exists but is highly reactive.

12.5.6 https://chem.libretexts.org/@go/page/41383
Exercise 12.5.2

Use a qualitative molecular orbital energy-level diagram to predict the valence electron configuration, bond order, and likely
existence of the Ca2+ ion.

Answer: Ca2+ has a (σ


1
4s )
2

4s

) electron configurations and a bond order of 1/2 and should exist.

Contributors
Seymour Blinder (Professor Emeritus of Chemistry and Physics at the University of Michigan, Ann Arbor)

12.5: Molecular Orbital Theory is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

12.5.7 https://chem.libretexts.org/@go/page/41383
12.6: Diatomic Molecules
For almost every covalent molecule that exists, we can now draw the Lewis structure, predict the electron-pair geometry, predict the molecular
geometry, and come close to predicting bond angles. However, one of the most important molecules we know, the oxygen molecule O2, presents a
problem with respect to its Lewis structure. We would write the following Lewis structure for O2:

This electronic structure adheres to all the rules governing Lewis theory. There is an O=O double bond, and each oxygen atom has eight electrons
around it. However, this picture is at odds with the magnetic behavior of oxygen. By itself, O2 is not magnetic, but it is attracted to magnetic fields.
Thus, when we pour liquid oxygen past a strong magnet, it collects between the poles of the magnet and defies gravity. Such attraction to a magnetic
field is called paramagnetism, and it arises in molecules that have unpaired electrons. And yet, the Lewis structure of O2 indicates that all electrons
are paired. How do we account for this discrepancy?
Magnetic susceptibility measures the force experienced by a substance in a magnetic field. When we compare the weight of a sample to the weight
measured in a magnetic field (Figure 12.6.1), paramagnetic samples that are attracted to the magnet will appear heavier because of the force exerted
by the magnetic field. We can calculate the number of unpaired electrons based on the increase in weight.

Figure 12.6.1 : A Gouy balance compares the mass of a sample in the presence of a magnetic field with the mass with the electromagnet turned off to
determine the number of unpaired electrons in a sample.
Experiments show that each O2 molecule has two unpaired electrons. The Lewis-structure model does not predict the presence of these two unpaired
electrons. Unlike oxygen, the apparent weight of most molecules decreases slightly in the presence of an inhomogeneous magnetic field. Materials
in which all of the electrons are paired are diamagnetic and weakly repel a magnetic field. Paramagnetic and diamagnetic materials do not act as
permanent magnets. Only in the presence of an applied magnetic field do they demonstrate attraction or repulsion.

levitating frog

Video 12.6.1 : Water, like most molecules, contains all paired electrons. Living things contain a large percentage of water, so they demonstrate
diamagnetic behavior. If you place a frog near a sufficiently large magnet, it will levitate. You can see videos of diamagnetic floating frogs,
strawberries, and more (https://www.youtube.com/watch?v=A1vyB-O5i6E)

12.6.1 https://chem.libretexts.org/@go/page/41384
Molecular orbital theory (MO theory) provides an explanation of chemical bonding that accounts for the paramagnetism of the oxygen molecule. It
also explains the bonding in a number of other molecules, such as violations of the octet rule and more molecules with more complicated bonding
(beyond the scope of this text) that are difficult to describe with Lewis structures. Additionally, it provides a model for describing the energies of
electrons in a molecule and the probable location of these electrons. Unlike valence bond theory, which uses hybrid orbitals that are assigned to one
specific atom, MO theory uses the combination of atomic orbitals to yield molecular orbitals that are delocalized over the entire molecule rather than
being localized on its constituent atoms. MO theory also helps us understand why some substances are electrical conductors, others are
semiconductors, and still others are insulators. Table 12.6.1 summarizes the main points of the two complementary bonding theories. Both theories
provide different, useful ways of describing molecular structure.
Table 12.6.1 : Comparison of Bonding Theories
Valence Bond Theory Molecular Orbital Theory

considers bonds as localized between one pair of atoms considers electrons delocalized throughout the entire molecule

creates bonds from overlap of atomic orbitals (s, p, d…) and hybrid orbitals (sp,
combines atomic orbitals to form molecular orbitals (σ, σ*, π, π*)
sp2, sp3…)

forms σ or π bonds creates bonding and antibonding interactions based on which orbitals are filled

predicts molecular shape based on the number of regions of electron density predicts the arrangement of electrons in molecules

needs multiple structures to describe resonance

Molecular Orbitals Formed from ns and np Atomic Orbitals


Atomic orbitals other than ns orbitals can also interact to form molecular orbitals. Because individual p, d, and f orbitals are not spherically
symmetrical, however, we need to define a coordinate system so we know which lobes are interacting in three-dimensional space. Recall that for
each np subshell, for example, there are npx, npy, and npz orbitals. All have the same energy and are therefore degenerate, but they have different
spatial orientations.

σnp = npz (A) − npz (B) (12.6.1)


z

Just as with ns orbitals, we can form molecular orbitals from np orbitals by taking their mathematical sum and difference. When two positive lobes
with the appropriate spatial orientation overlap, as illustrated for two npz atomic orbitals in part (a) in Figure 12.6.2, it is the mathematical difference
of their wave functions that results in constructive interference, which in turn increases the electron probability density between the two atoms. The
difference therefore corresponds to a molecular orbital called a σ bonding molecular orbital because, just as with the σ orbitals discussed
npz

previously, it is symmetrical about the internuclear axis (in this case, the z-axis):
σnp = npz (A) − npz (B) (12.6.2)
z

The other possible combination of the two npz orbitals is the mathematical sum:
σnp = npz (A) + npz (B) (12.6.3)
z

In this combination, shown in part (b) in Figure 12.6.2, the positive lobe of one npz atomic orbital overlaps the negative lobe of the other, leading to
destructive interference of the two waves and creating a node between the two atoms. Hence this is an antibonding molecular orbital. Because it, too,
is symmetrical about the internuclear axis, this molecular orbital is called a σ = np (A) − np (B) antibonding molecular orbital. Whenever
npz z z

orbitals combine, the bonding combination is always lower in energy (more stable) than the atomic orbitals from which it was derived, and the
antibonding combination is higher in energy (less stable).

Figure 12.6.2 : Formation of Molecular Orbitals from npz Atomic Orbitals on Adjacent Atoms.(a) By convention, in a linear molecule or ion, the z-
axis always corresponds to the internuclear axis, with +z to the right. As a result, the signs of the lobes of the npz atomic orbitals on the two atoms
alternate − + − +, from left to right. In this case, the σ (bonding) molecular orbital corresponds to the mathematical difference, in which the overlap
of lobes with the same sign results in increased probability density between the nuclei. (b) In contrast, the σ* (antibonding) molecular orbital
corresponds to the mathematical sum, in which the overlap of lobes with opposite signs results in a nodal plane of zero probability density
perpendicular to the internuclear axis.

Overlap of atomic orbital lobes with the same sign produces a bonding molecular orbital, regardless of whether it corresponds to the sum or the
difference of the atomic orbitals.

The remaining p orbitals on each of the two atoms, npx and npy, do not point directly toward each other. Instead, they are perpendicular to the
internuclear axis. If we arbitrarily label the axes as shown in Figure 12.6.3, we see that we have two pairs of np orbitals: the two npx orbitals lying in
the plane of the page, and two npy orbitals perpendicular to the plane. Although these two pairs are equivalent in energy, the npx orbital on one atom

12.6.2 https://chem.libretexts.org/@go/page/41384
can interact with only the npx orbital on the other, and the npy orbital on one atom can interact with only the npy on the other. These interactions are
side-to-side rather than the head-to-head interactions characteristic of σ orbitals. Each pair of overlapping atomic orbitals again forms two molecular
orbitals: one corresponds to the arithmetic sum of the two atomic orbitals and one to the difference. The sum of these side-to-side interactions
increases the electron probability in the region above and below a line connecting the nuclei, so it is a bonding molecular orbital that is called a pi (π)
orbital (a bonding molecular orbital formed from the side-to-side interactions of two or more parallel np atomic orbitals). The difference results in
the overlap of orbital lobes with opposite signs, which produces a nodal plane perpendicular to the internuclear axis; hence it is an antibonding
molecular orbital, called a pi star (π*) orbital An antibonding molecular orbital formed from the difference of the side-to-side interactions of two or
more parallel np atomic orbitals, creating a nodal plane perpendicular to the internuclear axis..
πnp = npx (A) + npx (B) (12.6.4)
x


πnp = npx (A) − npx (B) (12.6.5)
x

The two npy orbitals can also combine using side-to-side interactions to produce a bonding π molecular orbital and an antibonding π
npy

npy

molecular orbital. Because the npx and npy atomic orbitals interact in the same way (side-to-side) and have the same energy, the π and npx

πnpymolecular orbitals are a degenerate pair, as are the π and π



npx molecular orbitals.

npy

Figure 12.6.3 : Formation of π Molecular Orbitals from npx and npy Atomic Orbitals on Adjacent Atoms.(a) Because the signs of the lobes of both
the npx and the npy atomic orbitals on adjacent atoms are the same, in both cases the mathematical sum corresponds to a π (bonding) molecular
orbital. (b) In contrast, in both cases, the mathematical difference corresponds to a π* (antibonding) molecular orbital, with a nodal plane of zero
probability density perpendicular to the internuclear axis.
Figure 12.6.4 is an energy-level diagram that can be applied to two identical interacting atoms that have three np atomic orbitals each. There are six
degenerate p atomic orbitals (three from each atom) that combine to form six molecular orbitals, three bonding and three antibonding. The bonding
molecular orbitals are lower in energy than the atomic orbitals because of the increased stability associated with the formation of a bond. Conversely,
the antibonding molecular orbitals are higher in energy, as shown. The energy difference between the σ and σ* molecular orbitals is significantly
greater than the difference between the two π and π* sets. The reason for this is that the atomic orbital overlap and thus the strength of the interaction
are greater for a σ bond than a π bond, which means that the σ molecular orbital is more stable (lower in energy) than the π molecular orbitals.

Figure 12.6.4 : The Relative Energies of the σ and π Molecular Orbitals Derived from npx, npy, and npz Orbitals on Identical Adjacent Atoms.
Because the two npz orbitals point directly at each other, their orbital overlap is greater, so the difference in energy between the σ and σ* molecular
orbitals is greater than the energy difference between the π and π* orbitals.
Although many combinations of atomic orbitals form molecular orbitals, we will discuss only one other interaction: an ns atomic orbital on one atom
with an npz atomic orbital on another. As shown in Figure 12.6.5, the sum of the two atomic wave functions (ns + npz) produces a σ bonding
molecular orbital. Their difference (ns − npz) produces a σ* antibonding molecular orbital, which has a nodal plane of zero probability density
perpendicular to the internuclear axis.

Figure 12.6.5 : Formation of Molecular Orbitals from an ns Atomic Orbital on One Atom and an npz Atomic Orbital on an Adjacent Atom.(a) The
mathematical sum results in a σ (bonding) molecular orbital, with increased probability density between the nuclei. (b) The mathematical difference
results in a σ* (antibonding) molecular orbital, with a nodal plane of zero probability density perpendicular to the internuclear axis.

12.6.3 https://chem.libretexts.org/@go/page/41384
Second Row Diatomic Molecules
If we combine the splitting schemes for the 2s and 2p orbitals, we can predict bond order in all of the diatomic molecules and ions composed of
elements in the first complete row of the periodic table. Remember that only the valence orbitals of the atoms need be considered; as we saw in the
cases of lithium hydride and dilithium, the inner orbitals remain tightly bound and retain their localized atomic character.
We now describe examples of systems involving period 2 homonuclear diatomic molecules, such as N2, O2, and F2. When we draw a molecular
orbital diagram for a molecule, there are four key points to remember:
1. The number of molecular orbitals produced is the same as the number of atomic orbitals used to create them (the law of conservation of orbitals).
2. As the overlap between two atomic orbitals increases, the difference in energy between the resulting bonding and antibonding molecular orbitals
increases.
3. When two atomic orbitals combine to form a pair of molecular orbitals, the bonding molecular orbital is stabilized about as much as the
antibonding molecular orbital is destabilized.
4. The interaction between atomic orbitals is greatest when they have the same energy.

The number of molecular orbitals is always equal to the total number of atomic orbitals we started with.

We illustrate how to use these points by constructing a molecular orbital energy-level diagram for F2. We use the diagram in part (a) in Figure
12.6.6; the n = 1 orbitals (σ1s and σ1s*) are located well below those of the n = 2 level and are not shown. As illustrated in the diagram, the σ2s and

σ2s* molecular orbitals are much lower in energy than the molecular orbitals derived from the 2p atomic orbitals because of the large difference in
energy between the 2s and 2p atomic orbitals of fluorine. The lowest-energy molecular orbital derived from the three 2p orbitals on each F is σ 2pz

and the next most stable are the two degenerate orbitals, π and π . For each bonding orbital in the diagram, there is an antibonding orbital, and
2px 2py

the antibonding orbital is destabilized by about as much as the corresponding bonding orbital is stabilized. As a result, the σ orbital is higher in

2p
z

energy than either of the degenerate π and π orbitals. We can now fill the orbitals, beginning with the one that is lowest in energy.

2p

2p
x y

Each fluorine has 7 valence electrons, so there are a total of 14 valence electrons in the F2 molecule. Starting at the lowest energy level, the electrons
are placed in the orbitals according to the Pauli principle and Hund’s rule. Two electrons each fill the σ2s and σ2s* orbitals, 2 fill the σ orbital, 4 2pz

fill the two degenerate π orbitals, and 4 fill the two degenerate π* orbitals, for a total of 14 electrons. To determine what type of bonding the
molecular orbital approach predicts F2 to have, we must calculate the bond order. According to our diagram, there are 8 bonding electrons and 6
antibonding electrons, giving a bond order of (8 − 6) ÷ 2 = 1. Thus F2 is predicted to have a stable F–F single bond, in agreement with experimental
data.

Figure 12.6.6 : Molecular Orbital Energy-Level Diagrams for Homonuclear Diatomic Molecules.(a) For F2, with 14 valence electrons (7 from each F
atom), all of the energy levels except the highest, σ are filled. This diagram shows 8 electrons in bonding orbitals and 6 in antibonding orbitals,

2pz

resulting in a bond order of 1. (b) For O2, with 12 valence electrons (6 from each O atom), there are only 2 electrons to place in the (π , π )

npx

npy

pair of orbitals. Hund’s rule dictates that one electron occupies each orbital, and their spins are parallel, giving the O2 molecule two unpaired
electrons. This diagram shows 8 electrons in bonding orbitals and 4 in antibonding orbitals, resulting in a predicted bond order of 2.
We now turn to a molecular orbital description of the bonding in O2. It so happens that the molecular orbital description of this molecule provided an
explanation for a long-standing puzzle that could not be explained using other bonding models. To obtain the molecular orbital energy-level diagram
for O2, we need to place 12 valence electrons (6 from each O atom) in the energy-level diagram shown in part (b) in Figure 12.6.6. We again fill the
orbitals according to Hund’s rule and the Pauli principle, beginning with the orbital that is lowest in energy. Two electrons each are needed to fill the
σ2s and σ2s* orbitals, 2 more to fill the σ orbital, and 4 to fill the degenerate π and π orbitals. According to Hund’s rule, the last 2 electrons
2pz

2px

2py

must be placed in separate π* orbitals with their spins parallel, giving two unpaired electrons. This leads to a predicted bond order of (8 − 4) ÷ 2 = 2,
which corresponds to a double bond, in agreement with experimental data (Table 4.5): the O–O bond length is 120.7 pm, and the bond energy is
498.4 kJ/mol at 298 K.
None of the other bonding models can predict the presence of two unpaired electrons in O2. Chemists had long wondered why, unlike most other
substances, liquid O2 is attracted into a magnetic field. As shown in Video 12.6.2, it actually remains suspended between the poles of a magnet until
the liquid boils away. The only way to explain this behavior was for O2 to have unpaired electrons, making it paramagnetic, exactly as predicted by

12.6.4 https://chem.libretexts.org/@go/page/41384
molecular orbital theory. This result was one of the earliest triumphs of molecular orbital theory over the other bonding approaches we have
discussed.

Paramagnetism of Oxygen

Video 12.6.2 : Liquid O2 Suspended between the Poles of a Magnet.Because the O2 molecule has two unpaired electrons, it is paramagnetic.
Consequently, it is attracted into a magnetic field, which allows it to remain suspended between the poles of a powerful magnet until it evaporates.
Full video can be found at https://www.youtube.com/watch?featur...&v=Lt4P6ctf06Q.
The magnetic properties of O2 are not just a laboratory curiosity; they are absolutely crucial to the existence of life. Because Earth’s atmosphere
contains 20% oxygen, all organic compounds, including those that compose our body tissues, should react rapidly with air to form H2O, CO2, and
N2 in an exothermic reaction. Fortunately for us, however, this reaction is very, very slow. The reason for the unexpected stability of organic
compounds in an oxygen atmosphere is that virtually all organic compounds, as well as H2O, CO2, and N2, have only paired electrons, whereas
oxygen has two unpaired electrons. Thus the reaction of O2 with organic compounds to give H2O, CO2, and N2 would require that at least one of the
electrons on O2 change its spin during the reaction. This would require a large input of energy, an obstacle that chemists call a spin barrier.
Consequently, reactions of this type are usually exceedingly slow. If they were not so slow, all organic substances, including this book and you,
would disappear in a puff of smoke!
For period 2 diatomic molecules to the left of N2 in the periodic table, a slightly different molecular orbital energy-level diagram is needed because
the σ molecular orbital is slightly higher in energy than the degenerate π
2pz

npx and π

npy orbitals. The difference in energy between the 2s and 2p
atomic orbitals increases from Li2 to F2 due to increasing nuclear charge and poor screening of the 2s electrons by electrons in the 2p subshell. The
bonding interaction between the 2s orbital on one atom and the 2pz orbital on the other is most important when the two orbitals have similar
energies. This interaction decreases the energy of the σ2s orbital and increases the energy of the σ orbital. Thus for Li2, Be2, B2, C2, and N2, the
2pz

σ2pz orbital is higher in energy than the σ orbitals, as shown in Figure 12.6.7 Experimentally, it is found that the energy gap between the ns and
3pz

np atomic orbitals increases as the nuclear charge increases (Figure 12.6.7 ). Thus for example, the σ molecular orbital is at a lower energy than
2pz

the π pair.
2px,y

12.6.5 https://chem.libretexts.org/@go/page/41384
Figure 12.6.7 : Molecular Orbital Energy-Level Diagrams for the Diatomic Molecules of the Period 2 Elements. Unlike earlier diagrams, only the
molecular orbital energy levels for the molecules are shown here. For simplicity, the atomic orbital energy levels for the component atoms have been
omitted. For Li2 through N2, the σ orbital is higher in energy than the π
2pz 2px, y orbitals. In contrast, the σ orbital is lower in energy than the
2pz

π orbitals for O2 and F2 due to the increase in the energy difference between the 2s and 2p atomic orbitals as the nuclear charge increases
2px, y

across the row.

Completing the diagram for N2 in the same manner as demonstrated previously, we find that the 10 valence electrons result in 8 bonding electrons
and 2 antibonding electrons, for a predicted bond order of 3, a triple bond. Experimental data show that the N–N bond is significantly shorter than
the F–F bond (109.8 pm in N2 versus 141.2 pm in F2), and the bond energy is much greater for N2 than for F2 (945.3 kJ/mol versus 158.8 kJ/mol,
respectively). Thus the N2 bond is much shorter and stronger than the F2 bond, consistent with what we would expect when comparing a triple bond
with a single bond.

Example 12.6.1

Use a qualitative molecular orbital energy-level diagram to predict the electron configuration, the bond order, and the number of unpaired
electrons in S2, a bright blue gas at high temperatures.
Given: chemical species
Asked for: molecular orbital energy-level diagram, bond order, and number of unpaired electrons
Strategy:
A. Write the valence electron configuration of sulfur and determine the type of molecular orbitals formed in S2. Predict the relative energies of
the molecular orbitals based on how close in energy the valence atomic orbitals are to one another.
B. Draw the molecular orbital energy-level diagram for this system and determine the total number of valence electrons in S2.
C. Fill the molecular orbitals in order of increasing energy, being sure to obey the Pauli principle and Hund’s rule.
D. Calculate the bond order and describe the bonding.
Solution:
A Sulfur has a [Ne]3s23p4 valence electron configuration. To create a molecular orbital energy-level diagram similar to those in Figure 12.6.6
and Figure 12.6.7, we need to know how close in energy the 3s and 3p atomic orbitals are because their energy separation will determine
whether the π or the σ > molecular orbital is higher in energy. Because the ns–np energy gap increases as the nuclear charge increases
3px,y 3pz

(Figure 12.6.7), the σ molecular orbital will be lower in energy than the π
3pz pair. 3px,y

B The molecular orbital energy-level diagram is as follows:

Each sulfur atom contributes 6 valence electrons, for a total of 12 valence electrons.

12.6.6 https://chem.libretexts.org/@go/page/41384
C Ten valence electrons are used to fill the orbitals through π 3px and π 3py , leaving 2 electrons to occupy the degenerate π ⋆
3px
and π
3py

pair. From
Hund’s rule, the remaining 2 electrons must occupy these orbitals separately with their spins aligned. With the numbers of electrons written as
2 2 4 2
superscripts, the electron configuration of S2 is (σ 3s )
2

3s

) (σ3p ) (π3p
z x,y
) (π3p⋆x,y ) with 2 unpaired electrons. The bond order is (8 − 4) ÷ 2 =
2, so we predict an S=S double bond.

Exercise 12.6.1

Use a qualitative molecular orbital energy-level diagram to predict the electron configuration, the bond order, and the number of unpaired
electrons in the peroxide ion (O22−).
Answer
2 2 4 4
2
(σ2s ) (σ
2s

) (σ2p ) (π2p
z x,y
) (π2p⋆x,y ) bond order of 1; no unpaired electrons

Molecular Orbitals for Heteronuclear Diatomic Molecules


Diatomic molecules with two different atoms are called heteronuclear diatomic molecules. When two nonidentical atoms interact to form a chemical
bond, the interacting atomic orbitals do not have the same energy. If, for example, element B is more electronegative than element A (χB > χA), the
net result is a “skewed” molecular orbital energy-level diagram, such as the one shown for a hypothetical A–B molecule in Figure 12.6.8. The
atomic orbitals of element B are uniformly lower in energy than the corresponding atomic orbitals of element A because of the enhanced stability of
the electrons in element B. The molecular orbitals are no longer symmetrical, and the energies of the bonding molecular orbitals are more similar to
those of the atomic orbitals of B. Hence the electron density of bonding electrons is likely to be closer to the more electronegative atom. In this way,
molecular orbital theory can describe a polar covalent bond.

Figure 12.6.8 : Molecular Orbital Energy-Level Diagram for a Heteronuclear Diatomic Molecule AB, Where χB > χA. The bonding molecular orbitals
are closer in energy to the atomic orbitals of the more electronegative B atom. Consequently, the electrons in the bonding orbitals are not shared
equally between the two atoms. On average, they are closer to the B atom, resulting in a polar covalent bond.

A molecular orbital energy-level diagram is always skewed toward the more electronegative atom.

An Odd Number of Valence Electrons: NO


Nitric oxide (NO) is an example of a heteronuclear diatomic molecule. The reaction of O2 with N2 at high temperatures in internal combustion
engines forms nitric oxide, which undergoes a complex reaction with O2 to produce NO2, which in turn is responsible for the brown color we
associate with air pollution. Recently, however, nitric oxide has also been recognized to be a vital biological messenger involved in regulating blood
pressure and long-term memory in mammals.
Because NO has an odd number of valence electrons (5 from nitrogen and 6 from oxygen, for a total of 11), its bonding and properties cannot be
successfully explained by either the Lewis electron-pair approach or valence bond theory. The molecular orbital energy-level diagram for NO
(Figure 12.6.9) shows that the general pattern is similar to that for the O2 molecule (Figure 12.6.7). Because 10 electrons are sufficient to fill all the
bonding molecular orbitals derived from 2p atomic orbitals, the 11th electron must occupy one of the degenerate π* orbitals. The predicted bond
order for NO is therefore (8-3) ÷ 2 = 2 1/2 . Experimental data, showing an N–O bond length of 115 pm and N–O bond energy of 631 kJ/mol, are
consistent with this description. These values lie between those of the N2 and O2 molecules, which have triple and double bonds, respectively. As we
stated earlier, molecular orbital theory can therefore explain the bonding in molecules with an odd number of electrons, such as NO, whereas Lewis
electron structures cannot.

12.6.7 https://chem.libretexts.org/@go/page/41384
Figure 12.6.9 : Molecular Orbital Energy-Level Diagram for NO. Because NO has 11 valence electrons, it is paramagnetic, with a single electron
occupying the (π , π ) pair of orbitals.

2px

2py

Molecular orbital theory can also tell us something about the chemistry of N O . As indicated in the energy-level diagram in Figure 12.6.9, NO has a
single electron in a relatively high-energy molecular orbital. We might therefore expect it to have similar reactivity as alkali metals such as Li and
Na with their single valence electrons. In fact, N O is easily oxidized to the N O cation, which is isoelectronic with N and has a bond order of 3,
+
2

corresponding to an N≡O triple bond.

Nonbonding Molecular Orbitals


Molecular orbital theory is also able to explain the presence of lone pairs of electrons. Consider, for example, the HCl molecule, whose Lewis
electron structure has three lone pairs of electrons on the chlorine atom. Using the molecular orbital approach to describe the bonding in HCl, we can
see from Figure 12.6.10 that the 1s orbital of atomic hydrogen is closest in energy to the 3p orbitals of chlorine. Consequently, the filled Cl 3s
atomic orbital is not involved in bonding to any appreciable extent, and the only important interactions are those between the H 1s and Cl 3p
orbitals. Of the three p orbitals, only one, designated as 3pz, can interact with the H 1s orbital. The 3px and 3py atomic orbitals have no net overlap
with the 1s orbital on hydrogen, so they are not involved in bonding. Because the energies of the Cl 3s, 3px, and 3py orbitals do not change when
HCl forms, they are called nonbonding molecular orbitals. A nonbonding molecular orbital occupied by a pair of electrons is the molecular orbital
equivalent of a lone pair of electrons. By definition, electrons in nonbonding orbitals have no effect on bond order, so they are not counted in the
calculation of bond order. Thus the predicted bond order of HCl is (2 − 0) ÷ 2 = 1. Because the σ bonding molecular orbital is closer in energy to the
Cl 3pz than to the H 1s atomic orbital, the electrons in the σ orbital are concentrated closer to the chlorine atom than to hydrogen. A molecular
orbital approach to bonding can therefore be used to describe the polarization of the H–Cl bond to give H − C l . δ+ δ−

Figure 12.6.10: Molecular Orbital Energy-Level Diagram for HCl. The hydrogen 1 s atomic orbital interacts most strongly with the 3pz orbital on
chlorine, producing a bonding/antibonding pair of molecular orbitals. The other electrons on Cl are best viewed as nonbonding. As a result, only the
bonding σ orbital is occupied by electrons, giving a bond order of 1.

Electrons in nonbonding molecular orbitals have no effect on bond order.

Example 12.6.2

Use a “skewed” molecular orbital energy-level diagram like the one in Figure 12.6.8 to describe the bonding in the cyanide ion (CN−). What is
the bond order?
Given: chemical species
Asked for: “skewed” molecular orbital energy-level diagram, bonding description, and bond order
Strategy:

12.6.8 https://chem.libretexts.org/@go/page/41384
A. Calculate the total number of valence electrons in CN−. Then place these electrons in a molecular orbital energy-level diagram like Figure
12.6.8 in order of increasing energy. Be sure to obey the Pauli principle and Hund’s rule while doing so.

B. Calculate the bond order and describe the bonding in CN−.


Solution:
A The CN− ion has a total of 10 valence electrons: 4 from C, 5 from N, and 1 for the −1 charge. Placing these electrons in an energy-level
diagram like Figure 12.6.8 fills the five lowest-energy orbitals, as shown here:

Because χN > χC, the atomic orbitals of N (on the right) are lower in energy than those of C. B The resulting valence electron configuration gives
a predicted bond order of (8 − 2) ÷ 2 = 3, indicating that the CN− ion has a triple bond, analogous to that in N2.

Exercise 12.6.2

Use a qualitative molecular orbital energy-level diagram to describe the bonding in the hypochlorite ion (OCl−). What is the bond order?
Answer
All molecular orbitals except the highest-energy σ* are filled, giving a bond order of 1.

Although the molecular orbital approach reveals a great deal about the bonding in a given molecule, the procedure quickly becomes computationally
intensive for molecules of even moderate complexity. Furthermore, because the computed molecular orbitals extend over the entire molecule, they
are often difficult to represent in a way that is easy to visualize. Therefore we do not use a pure molecular orbital approach to describe the bonding in
molecules or ions with more than two atoms. Instead, we use a valence bond approach and a molecular orbital approach to explain, among other
things, the concept of resonance, which cannot adequately be explained using other methods.

12.6: Diatomic Molecules is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

12.6.9 https://chem.libretexts.org/@go/page/41384
12.7: Resonance and Electron Delocalization
Resonance is a mental exercise and method within the Valence Bond Theory of bonding that describes the delocalization of
electrons within molecules. It compares and contrasts two or more possible Lewis structures that can represent a particular
molecule. Resonance structures are used when one Lewis structure for a single molecule cannot fully describe the bonding that
takes place between neighboring atoms relative to the empirical data for the actual bond lengths between those atoms. The net sum
of valid resonance structures is defined as a resonance hybrid, which represents the overall delocalization of electrons within the
molecule. A molecule that has several resonance structures is more stable than one with fewer. Some resonance structures are more
favorable than others.

Introduction
Electrons have no fixed position in atoms, compounds and molecules (see image below) but have probabilities of being found in
certain spaces (orbitals). Resonance forms illustrate areas of higher probabilities (electron densities). This is like holding your hat
in either your right hand or your left. The term Resonance is applied when there are two or more possibilities available. Chemists
use Lewis diagrams to depict structure and bonding of covalent entities, such as molecules and polyatomic ions, henceforth,
molecules. The Lewis diagram of many a molecule, however, is not consistent with the observed properties of the molecule.

Example 12.7.1: Nitrate Ion

The nitrate ion, according to its Lewis diagram, has two types of nitrogen-oxygen bonds, one double bond and two single
bonds, suggesting that one nitrogen-oxygen bond in the nitrate ion is shorter and stronger than each of the other two. Also, the
Lewis structure implies, with respect to formal charge, that there are two types of oxygen atoms in the nitrate ion, one formally
neutral and each of the other two bearing a formal charge of –1. Experimentally, however, the three nitrogen-oxygen bonds in
the nitrate ion have the same bond length and the same bond energy, and the three oxygen atoms are indistinguishable. The
Lewis diagram fails to explain the structure and bonding of the nitrate ion satisfactorily.
Two additional Lewis diagrams can be drawn for the nitrate ion.

However, none of them are consistent with the observed properties of the nitrate ion and, therefore, does not correctly depict
the nitrate ion.

Example 12.7.2: Benzene

Benzene, according to its Lewis diagram, has two types of carbon-carbon bonds, three double bonds and three single bonds,
suggesting that each of three carbon-carbon bonds in benzene is shorter and stronger than each of the other three.
Experimentally, however, the six carbon-carbon bonds in benzene have the same bond length and the same bond energy. The
Lewis diagram fails to explain the structure and bonding of benzene satisfactorily.
An attribute of molecules of which the classical Lewis diagram is not consistent with the observed properties is that other valid
Lewis diagrams can be generated for them. One additional Lewis diagram can be drawn for benzene.

However, none of them are consistent with the observed properties of benzene and, therefore, does not correctly depict
benzene.

Resonance theory is an attempt to explain the structure of a species, like the nitrate ion or benzene, no Lewis diagram of which is
consistent with the observed properties of the species. The major advantage of resonance theory is that, although based on rigorous

12.7.1 https://chem.libretexts.org/@go/page/41385
mathematical analysis, resonance theory can be applied successfully invoking little or no math. Resonance theory is explained
below using the nitrate ion as the example.
According to resonance theory, the structure of the nitrate ion is not 1 nor 2 nor 3 but the average of all three of them, weighted by
stability. Lewis diagrams 1, 2, and 3 are called resonance forms, resonance structures, or resonance contributors of the nitrate ion.
The structure of the nitrate ion is said to be a resonance hybrid or, simply, hybrid of resonance forms 1, 2, and 3. Whenever it is
necessary to show the structure of the nitrate ion, resonance forms 1, 2, and 3 are drawn, connected by a double-headed arrows.

The three resonance forms of the nitrate ion, 1, 2, and 3, are identical, so they have the same stability and, therefore, contribute
equally to the hybrid. Since the exact extent to which each resonance form of the nitrate ion contributes to the hybrid is known, the
bond order of each nitrogen-oxygen bond as well as the formal charge on each oxygen atom in the hybrid can be easily determined:

According to resonance theory, each bond in the nitrate ion is one and one-third of a bond, which is consistent with the observation
that the three bonds in the nitrate ion have the same bond length and the same bond energy.

According to resonance theory, each oxygen atom in the nitrate ion has a formal charge of –2/3, which, in conjunction with the fact
that the three nitrogen-oxygen bonds are identical, is consistent with the observation that the three oxygen atoms in the nitrate ion
are indistinguishable. In each resonance form of the nitrate ion, there are two π electrons, and they are shared only by two atoms.
An electron shared only by two atoms is said to be localized. Thus, the two π electrons in each resonance form of the nitrate ion are
localized. The nitrate ion, as represented by the hybrid, has two π electrons:
# electrons in one π bond = 2
# electrons in one-third of a π bond = 2/3
# electrons in three of them = 3 x (2/3) = 2

12.7.2 https://chem.libretexts.org/@go/page/41385
The two π electrons in the nitrate ion are shared by a total of four atoms, one nitrogen atom and three oxygen atoms. An electron
shared by more than two atoms is said to be delocalized. Thus, the two π electrons in the nitrate ion are delocalized. Delocalization
of π electrons in the nitrate ion requires that the four atoms be on the same plane, allowing lateral overlap of the p orbitals on them.

If the energy of the nitrate ion were the weighted average of the energies of its three resonance forms, just as the structure of the
nitrate ion is the weighted average of the structures of its three resonance forms, it should be equal to the energy of one of the three
identical resonance forms:

If the energy of the hybrid were equal to that of a resonance form, given that all chemical entities (elementary particles, atoms,
molecules, etc.) naturally tend to be in the lowest possible energy state, there would be no advantage for the nitrate ion to exist as
the hybrid; it could simply exist as a resonance form. Since the nitrate ion exists as the hybrid, not as a resonance form, it can be
inferred that the energy of the hybrid is lower than that of any of the resonance forms.

According to resonance theory then, the energy of a molecule is lower than that of the lowest-energy resonance form. Since the
nitrate ion has lower energy and, therefore, is more stable than any of its resonance forms, the nitrate ion is said to be resonance
stabilized.

Common Misconceptions

There are two misconceptions about resonance theory among beginning students, likely due to literal interpretation of the word
resonance. They are described below, using the nitrate ion as the example.
Misconception 1: The nitrate ion exists as resonance form 1 for a moment and then changes either to resonance form 2 or to
resonance form 3, which interconvert, or revert to 1.

The structure of the nitrate ion is not 1 nor 2 nor 3 but the hybrid and does not change with time unless undergoing a reaction.
Misconception 2: In a sample of nitrate ions, at a given moment, one-third of the ions exist as resonance form 1, another one-
third as resonance form 2, and the remaining one-third as resonance form 3.

12.7.3 https://chem.libretexts.org/@go/page/41385
In a sample of nitrate ions, at a given moment, all ions have the same structure, which is the hybrid.

Mules and Rhinoceros


The classic analogy used to clarify these two misconceptions is the mule (Morrison, R. T.; Boyd, R.N. Organic Chemistry, fifth
edition; Allyn and Bacon: Boston, 1987, pg. 373). Biologically, a mule is a hybrid of a horse and a donkey. This does not mean
that a mule resembles a horse for a moment and then changes to resemble a donkey. The appearance of a mule is a combination
of that of a horse and that of a donkey and does not change with time. Nor does it mean that, in a herd, some mules resemble a
horse and the others a donkey. In a herd, all mules have the same appearance, which is a combination of a horse and a donkey.
The weakness of this analogy is that horses and donkeys do exist, whereas resonance forms are strictly hypothetical. A better
analogy, cited in Morrison and Boyd, is the rhinoceros. Upon seeing a rhinoceros, one could describe it as the hybrid of a
dragon and a unicorn, two creatures that do not exist.

Peptide Bonds
A consideration of resonance contributors is crucial to any discussion of the amide functional group. One of the most important
examples of amide groups in nature is the ‘peptide bond’ that links amino acids to form polypeptides and proteins.

Critical to the structure of proteins is the fact that, although it is conventionally drawn as a single bond, the C-N bond in a peptide
linkage has a significant barrier to rotation, almost as if it were a double bond.

This, along with the observation that the bonding around the peptide nitrogen has trigonal planar geometry, strongly suggests that
the nitrogen is sp2-hybridized. An important resonance contributor has a C=N double bond and a C-O single bond, with a
separation of charge between the oxygen and the nitrogen.

Although B is a minor contributor due to the separation of charges, it is still very relevant in terms of peptide and protein structure
– our proteins would simply not fold up properly if there was free rotation about the peptide C-N bond.

Contributors and Attributions


Gamini Gunawardena from the OChemPal site (Utah Valley University)
Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)

12.7: Resonance and Electron Delocalization is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.

12.7.4 https://chem.libretexts.org/@go/page/41385
12.8: Coordination Compounds
 Learning Objectives
To know the most common structures observed for metal complexes.
To predict the relative stabilities of metal complexes with different ligands

One of the most important properties of metallic elements is their ability to act as Lewis acids that form complexes with a variety
of Lewis bases. A metal complex consists of a central metal atom or ion that is bonded to one or more ligands (from the Latin
ligare, meaning “to bind”), which are ions or molecules that contain one or more pairs of electrons that can be shared with the
metal. Metal complexes can be neutral, such as Co(NH3)3Cl3; positively charged, such as [Nd(H2O)9]3+; or negatively charged,
such as [UF8]4−. Electrically charged metal complexes are sometimes called complex ions. A coordination compound contains one
or more metal complexes.
Coordination compounds are important for at least three reasons. First, most of the elements in the periodic table are metals, and
almost all metals form complexes, so metal complexes are a feature of the chemistry of more than half the elements. Second, many
industrial catalysts are metal complexes, and such catalysts are steadily becoming more important as a way to control reactivity.
For example, a mixture of a titanium complex and an organometallic compound of aluminum is the catalyst used to produce most
of the polyethylene and polypropylene “plastic” items we use every day. Finally, transition-metal complexes are essential in
biochemistry. Examples include hemoglobin, an iron complex that transports oxygen in our blood; cytochromes, iron complexes
that transfer electrons in our cells; and complexes of Fe, Zn, Cu, and Mo that are crucial components of certain enzymes, the
catalysts for all biological reactions.

History of the Coordination Compounds


Coordination compounds have been known and used since antiquity; probably the oldest is the deep blue pigment called Prussian
blue: KFe (CN) . The chemical nature of these substances, however, was unclear for a number of reasons. For example, many
2 6

compounds called “double salts” were known, such as AlF ⋅ 3 KF , Fe(CN) ⋅ 4 KCN , and ZnCl ⋅ 2 CsCl , which were
3 2 2

combinations of simple salts in fixed and apparently arbitrary ratios. Why should AlF ⋅ 3 KF exist but not AlF ⋅ 4 KF or
3 3

AlF ⋅ 2 KF ? And why should a 3:1 KF:AlF3 mixture have different chemical and physical properties than either of its
3

components? Similarly, adducts of metal salts with neutral molecules such as ammonia were also known—for example,
CoCl ⋅ 6 NH , which was first prepared sometime before 1798. Like the double salts, the compositions of these adducts exhibited
3 3

fixed and apparently arbitrary ratios of the components. For example, CoCl ⋅ 6 NH , CoCl ⋅ 5 NH , CoCl ⋅ 4 NH , and
3 3 3 3 3 3

CoCl ⋅ 3 NH
3
were all known and had very different properties, but despite all attempts, chemists could not prepare
3

CoCl ⋅ 2 NH
3
or CoCl ⋅ NH .
3 3 3

Although the chemical composition of such compounds was readily established by existing analytical methods, their chemical
nature was puzzling and highly controversial. The major problem was that what we now call valence (i.e., the oxidation state) and
coordination number were thought to be identical. As a result, highly implausible (to modern eyes at least) structures were
proposed for such compounds, including the “Chattanooga choo-choo” model for CoCl3·4NH3 shown here.

The modern theory of coordination chemistry is based largely on the work of Alfred Werner (1866–1919; Nobel Prize in Chemistry
in 1913). In a series of careful experiments carried out in the late 1880s and early 1890s, he examined the properties of several
series of metal halide complexes with ammonia. For example, five different “adducts” of ammonia with PtCl4 were known at the
time: PtCl4·nNH3 (n = 2–6). Some of Werner’s original data on these compounds are shown in Table 12.8.1. The electrical
conductivity of aqueous solutions of these compounds was roughly proportional to the number of ions formed per mole, while the
number of chloride ions that could be precipitated as AgCl after adding Ag+(aq) was a measure of the number of “free” chloride
ions present. For example, Werner’s data on PtCl4·6NH3 in Table 12.8.1 showed that all the chloride ions were present as free
chloride. In contrast, PtCl4·2NH3 was a neutral molecule that contained no free chloride ions.

12.8.1 https://chem.libretexts.org/@go/page/41386
 Alfred Werner (1866–1919)

Werner, the son of a factory worker, was born in Alsace. He developed an interest in chemistry at an early age, and he did his
first independent research experiments at age 18. While doing his military service in southern Germany, he attended a series of
chemistry lectures, and he subsequently received his PhD at the University of Zurich in Switzerland, where he was appointed
professor of chemistry at age 29. He won the Nobel Prize in Chemistry in 1913 for his work on coordination compounds,
which he performed as a graduate student and first presented at age 26. Apparently, Werner was so obsessed with solving the
riddle of the structure of coordination compounds that his brain continued to work on the problem even while he was asleep. In
1891, when he was only 25, he woke up in the middle of the night and, in only a few hours, had laid the foundation for modern
coordination chemistry.

Table 12.8.1 : Werner’s Data on Complexes of Ammonia with P tC l 4

Number of Cl− Ions Precipitated


Complex Conductivity (ohm−1) Number of Ions per Formula Unit
by Ag+

PtCl4·6NH3 523 5 4

PtCl4·5NH3 404 4 3

PtCl4·4NH3 299 3 2

PtCl4·3NH3 97 2 1

PtCl4·2NH3 0 0 0

These data led Werner to postulate that metal ions have two different kinds of valence: (1) a primary valence (oxidation state) that
corresponds to the positive charge on the metal ion and (2) a secondary valence (coordination number) that is the total number of
ligand-metal bonds bound to the metal ion. If Pt had a primary valence of 4 and a secondary valence of 6, Werner could explain
the properties of the PtCl ⋅ NH adducts by the following reactions, where the metal complex is enclosed in square brackets:
4 3

4+ −
[Pt(NH3 )6 ]C l4 → [Pt(NH3 )6 ] (aq) + 4C l (aq)

3+ −
[Pt(NH3 )5 Cl]C l3 → [Pt(NH3 )5 Cl ] (aq) + 3C l (aq)

2+ −
[Pt(NH3 )4 C l2 ]C l2 → [Pt(NH3 )4 C l2 ] (aq) + 2C l (aq)

+ −
[Pt(NH3 )3 C l3 ]Cl → [Pt(NH3 )3 C l3 ] (aq) + C l (aq)

0
[Pt(NH3 )2 C l4 ] → [Pt(NH3 )2 C l4 ] (aq)

Further work showed that the two missing members of the series—[Pt(NH3)Cl5]− and [PtCl6]2−—could be prepared as their mono-
and dipotassium salts, respectively. Similar studies established coordination numbers of 6 for Co3+ and Cr3+ and 4 for Pt2+ and
Pd2+.
Werner’s studies on the analogous Co3+ complexes also allowed him to propose a structural model for metal complexes with a
coordination number of 6. Thus he found that [Co(NH3)6]Cl3 (yellow) and [Co(NH3)5Cl]Cl2 (purple) were 1:3 and 1:2 electrolytes.
Unexpectedly, however, two different [Co(NH3)4Cl2]Cl compounds were known: one was red, and the other was green (Figure
12.8.1a). Because both compounds had the same chemical composition and the same number of groups of the same kind attached

to the same metal, there had to be something different about the arrangement of the ligands around the metal ion. Werner’s key
insight was that the six ligands in [Co(NH3)4Cl2]Cl had to be arranged at the vertices of an octahedron because that was the only
structure consistent with the existence of two, and only two, arrangements of ligands (Figure 12.8.1b. His conclusion was
corroborated by the existence of only two different forms of the next compound in the series: Co(NH3)3Cl3.

12.8.2 https://chem.libretexts.org/@go/page/41386
Figure 12.8.1 : Complexes with Different Arrangements of the Same Ligands Have Different Properties. The [Co(NH3)4Cl2]+ ion
can have two different arrangements of the ligands, which results in different colors: if the two Cl− ligands are next to each other,
the complex is red (a), but if they are opposite each other, the complex is green (b).

 Example 12.8.1

In Werner’s time, many complexes of the general formula MA4B2 were known, but no more than two different compounds
with the same composition had been prepared for any metal. To confirm Werner’s reasoning, calculate the maximum number
of different structures that are possible for six-coordinate MA4B2 complexes with each of the three most symmetrical possible
structures: a hexagon, a trigonal prism, and an octahedron. What does the fact that no more than two forms of any MA4B2
complex were known tell you about the three-dimensional structures of these complexes?
Given: three possible structures and the number of different forms known for MA4B2 complexes
Asked for: number of different arrangements of ligands for MA4B2 complex for each structure
Strategy:
Sketch each structure, place a B ligand at one vertex, and see how many different positions are available for the second B
ligand.
Solution
The three regular six-coordinate structures are shown here, with each coordination position numbered so that we can keep track
of the different arrangements of ligands. For each structure, all vertices are equivalent. We begin with a symmetrical MA6
complex and simply replace two of the A ligands in each structure to give an MA4B2 complex:

For the hexagon, we place the first B ligand at position 1. There are now three possible places for the second B ligand: at
position 2 (or 6), position 3 (or 5), or position 4. These are the only possible arrangements. The (1, 2) and (1, 6) arrangements
are chemically identical because the two B ligands are adjacent to each other. The (1, 3) and (1, 5) arrangements are also
identical because in both cases the two B ligands are separated by an A ligand.
Turning to the trigonal prism, we place the first B ligand at position 1. Again, there are three possible choices for the second B
ligand: at position 2 or 3 on the same triangular face, position 4 (on the other triangular face but adjacent to 1), or position 5 or
6 (on the other triangular face but not adjacent to 1). The (1, 2) and (1, 3) arrangements are chemically identical, as are the (1,
5) and (1, 6) arrangements.
In the octahedron, however, if we place the first B ligand at position 1, then we have only two choices for the second B ligand:
at position 2 (or 3 or 4 or 5) or position 6. In the latter, the two B ligands are at opposite vertices of the octahedron, with the
metal lying directly between them. Although there are four possible arrangements for the former, they are chemically identical
because in all cases the two B ligands are adjacent to each other.

12.8.3 https://chem.libretexts.org/@go/page/41386
The number of possible MA4B2 arrangements for the three geometries is thus: hexagon, 3; trigonal prism, 3; and octahedron, 2.
The fact that only two different forms were known for all MA4B2 complexes that had been prepared suggested that the correct
structure was the octahedron but did not prove it. For some reason one of the three arrangements possible for the other two
structures could have been less stable or harder to prepare and had simply not yet been synthesized. When combined with
analogous results for other types of complexes (e.g., MA3B3), however, the data were best explained by an octahedral structure
for six-coordinate metal complexes.

 Exercise 12.8.1
Determine the maximum number of structures that are possible for a four-coordinate MA2B2 complex with either a square
planar or a tetrahedral symmetrical structure.

Answer
square planar, 2; tetrahedral, 1

Structures of Metal Complexes


The coordination numbers of metal ions in metal complexes can range from 2 to at least 9. In general, the differences in energy
between different arrangements of ligands are greatest for complexes with low coordination numbers and decrease as the
coordination number increases. Usually only one or two structures are possible for complexes with low coordination numbers,
whereas several different energetically equivalent structures are possible for complexes with high coordination numbers (n > 6).
The following presents the most commonly encountered structures for coordination numbers 2–9. Many of these structures should
be familiar to you from our discussion of the valence-shell electron-pair repulsion (VSEPR) model because they correspond to the
lowest-energy arrangements of n electron pairs around a central atom.

Compounds with low coordination numbers exhibit the greatest differences in energy
between different arrangements of ligands.
Coordination Number 2
Although it is rare for most metals, this coordination number is surprisingly common for d10 metal ions, especially Cu+, Ag+, Au+,
and Hg2+. An example is the [Au(CN)2]− ion, which is used to extract gold from its ores. As expected based on VSEPR
considerations, these complexes have the linear L–M–L structure shown here.

Coordination Number 3
Although it is also rare, this coordination number is encountered with d10 metal ions such as Cu+ and Hg2+. Among the few known
examples is the HgI3− ion. Three-coordinate complexes almost always have the trigonal planar structure expected from the VSEPR
model.

Coordination Number 4
Two common structures are observed for four-coordinate metal complexes: tetrahedral and square planar. The tetrahedral structure
is observed for all four-coordinate complexes of nontransition metals, such as [BeF4]2−, and d10 ions, such as [ZnCl4]2−. It is also
found for four-coordinate complexes of the first-row transition metals, especially those with halide ligands (e.g., [FeCl4]− and
[FeCl4]2−). In contrast, square planar structures are routinely observed for four-coordinate complexes of second- and third-row

12.8.4 https://chem.libretexts.org/@go/page/41386
transition metals with d8 electron configurations, such as Rh+ and Pd2+, and they are also encountered in some complexes of Ni2+
and Cu2+.

Coordination Number 5
This coordination number is less common than 4 and 6, but it is still found frequently in two different structures: trigonal
bipyramidal and square pyramidal. Because the energies of these structures are usually rather similar for most ligands, many five-
coordinate complexes have distorted structures that lie somewhere between the two extremes.

Coordination Number 6
This coordination number is by far the most common. The six ligands are almost always at the vertices of an octahedron or a
distorted octahedron. The only other six-coordinate structure is the trigonal prism, which is very uncommon in simple metal
complexes.

Coordination Number 7
This relatively uncommon coordination number is generally encountered for only large metals (such as the second- and third-row
transition metals, lanthanides, and actinides). At least three different structures are known, two of which are derived from an
octahedron or a trigonal prism by adding a ligand to one face of the polyhedron to give a “capped” octahedron or trigonal prism. By
far the most common, however, is the pentagonal bipyramid.

12.8.5 https://chem.libretexts.org/@go/page/41386
Coordination Number 8
This coordination number is relatively common for larger metal ions. The simplest structure is the cube, which is rare because it
does not minimize interligand repulsive interactions. Common structures are the square antiprism and the dodecahedron, both of
which can be generated from the cube.

Coordination Number 9
This coordination number is found in larger metal ions, and the most common structure is the tricapped trigonal prism, as in
[Nd(H2O)9]3+.

Key Takeaways
Coordination compounds are a major feature of the chemistry of over half the elements.
Coordination compounds have important roles as industrial catalysts in controlling reactivity, and they are essential in
biochemical processes.

Summary
Transition metals form metal complexes, polyatomic species in which a metal ion is bound to one or more ligands, which are
groups bound to a metal ion. Complex ions are electrically charged metal complexes, and a coordination compound contains one or
more metal complexes. Metal complexes with low coordination numbers generally have only one or two possible structures,
whereas those with coordination numbers greater than six can have several different structures. Coordination numbers of two and
three are common for d10 metal ions. Tetrahedral and square planar complexes have a coordination number of four; trigonal
bipyramidal and square pyramidal complexes have a coordination number of five; and octahedral complexes have a coordination
number of six. At least three structures are known for a coordination number of seven, which is generally found for only large
metal ions. Coordination numbers of eight and nine are also found for larger metal ions.

12.8.6 https://chem.libretexts.org/@go/page/41386
12.8: Coordination Compounds is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
24.1: Werner’s Theory of Coordination Compounds by Anonymous is licensed CC BY-NC-SA 4.0.

12.8.7 https://chem.libretexts.org/@go/page/41386
12.9: Coordination Compounds in Biological Systems
12.9: Coordination Compounds in Biological Systems is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.

12.9.1 https://chem.libretexts.org/@go/page/41387
12.E: The Chemical Bond (Exercises)
12.1: Lewis Structures
Q12.1
Which of the following molecules has the shortest bond length?
O2

C l2

Br2

I2

Q12.2
Draw a plausible Lewis structure for 1,3-Bisphosphoglyerate (1,3-BPG), (C3H8O10P2).

Q12.3
Draw resonance structures for NO3- and ClO3-, please show formal charges in drawings.

Q12.4
Calculate the dipole moment (μ ) of a two ions of +2e and -2e that are separated by 100 pm. Express dipole moment in both units of
C m and D.

S12.4
−19 −10 −28
μ = Q × r = 2(1.602 × 10 )(8 × 10 ) = 2.56 × 10 C m (12.E.1)

−28
1D
(2.56 × 10 C m )( ) = 76.8 D (12.E.2)
−30
3.336 × 10 C m

12.2: Valence Bond Theory


Q12.5
What are the resonance of N O ?
0
3

S12.5
Resonance structure in basic means that there are multiple way to draw a Lewis structure. For example:

*Source from Michael Blaber

Q12.6
Draw resonance structures for the carbonate ion (C O −2
3
).

S12.6

12.E.1 https://chem.libretexts.org/@go/page/41338
Q12.7a
How is it that this structure, PF5, have P have 5 bonds to it?

S12.7a
The d orbitals are available to make extra bonds allowing the P to bond to 5 Flourines instead of just Four.

Q12.7b
Draw the Lewis structure for nitric acid (H N O ). This may or may not include resonance structures.
3

S12.7b

12.3: Hybridization of Atomic Orbitals


Q12.8a
What is the hybridization of carbon in CH4?

S12.8a
Carbon has the electron configuration 1s22s2sp2. In the excited state, because the 2s and 2p orbitals are so close one of the s
electrons is able to fill the empty 2p orbital. All the carbon bonds are identical so the four unpaired electrons will be in a sp3
hybridization. The sp3 hybridization is the only hybridization that accounts for four identical orbitals.

Q12.8b
Describe the pattern of valence electrons across the periodic table (left to right and top to bottom), excluding the transition metals.

S12.8b
The number of valence electrons increases going left to right along the same row and stays constant going from top to bottom along
the same column.

Q12.9a
What is the difference between covalent bond and ionic bond wave functions?

S12.9a
An ionic bond wave function takes into account the probability that the electrons of a multiatomic molecule might exist on the
same atom; the covalent bond wave function assumes they exist on separate atoms.

Q12.9b
What is the hybridization of HCl?

S12.9b
There is no orbital hybridization in HCl. The 3p of Cl can overlap with the 1s of H just fine; they are only 0.06 eV apart in energy
z

However, the 3s of Cl and the 1s of H are too far away in energy to interact, so there is no reason why the 3p of Cl can interact with
the 3s of Cl anyway.

12.E.2 https://chem.libretexts.org/@go/page/41338
Q13.7
Calculate the potential energy of interaction between an electron situated 5.5 Å away from an HF molecule with a dipole moment
of 2.5 D.

S13.7
30
3.3356 × 10 Cm −30
μ = 2.5D × = 8.39 × 10 Cm (12.E.3)
D

−10
r = 5.5 angstroms = 5.5 × 10 m (12.E.4)

−19 −30
qμ (+1.602 × 10 C )(8.39 × 10 C m)
−11
V = = = 2.196 × 10 J (12.E.5)
2 −12 −10
4πεo r 4π(8.851 × 10 C −2 N −1
m−2 )(5.5 × 10 )

Q12.10
Give two types of bonding that present in the molecule CH2. how do you do this?

S12.10
1 is C atom to be not hydrid, no paired electrons and the molecule will be diamagnetic.
2 is C atom is sp3 hybbridized with 2 of the hybrid orbitals used for bonding to H's , unpaired electron. this is paramagnetic.

Q12.11
Describe bonding CO2 and C3H4 as hybridization. Draw diagram of forming of sigma and pi bond of C3H4 there is a double bond
CO2 and its hydridization is sp and C3H4 is sp2 belong to CH2 and 1 sp of C=C=C.

Q12.12
Give the bonding scheme in term of molecular orbital H2+, H2, He2+, He. Besides, give the species in the following decreasing
stability.

12.E.3 https://chem.libretexts.org/@go/page/41338
Q12.13
Between F2 and F2+ which one has longer bond?

S12.13
F
2
has bond order 1.5, F has bond order 1.0. The bond lengths are inverse to the bond order, so the order is F2+ < F2 .There is 1
+
2

unpaired electron in F2+, 0 unpaired electrons in F2

Q12.14
Which of the following molecules have the shortest bond: F2, N2, O2? Why?

S12.14
N2 has the shortest bond because the bond order is 3 while O2 is 2 and F2 is 1.

12.5: Molecular Orbital Theory


Q12.15
Draw the molecular orbital diagram for F2 with the atomic orbitals labeled and find the bond order.

Q12.16
Comparing benezene and cyclopentadiene, which one has a lesser degree of π -electron delocalization?

Q12.19
Draw the molecular orbital energy diagram for the F2 Then compare the relative stability of the molecule to F2+.
Answer

12.E.4 https://chem.libretexts.org/@go/page/41338
BOND ORDER OF F2 = (8-6)/2 =1
BOND ORDER OF F2+ (8-5)/2 =1.5
Based on bond order F2+ is stable than F2

Q12.21
Diatomic carbon (C2) exists at very high temperatures. What type or types of bonds do you expect a molecule of diatomic carbon to
have?

S12.21
1. Draw a molecular orbital diagram.
Molecular orbital diagram for diatomic carbon

Answer: The molecular orbital diagram predicts that there will be a double bond consisting of two π bonds.

Q12.21
Determine the second period for molecule O2 by using molecular orbital?

S12.21

12.E.5 https://chem.libretexts.org/@go/page/41338
Based on general diagram for molecular orbital of second row of O2 molecule, the 2p MO is below the 2p molecular orbitals.

Q12.30
Though there are some exceptions, the double bond rule states that atoms of period 3 and greater do not form double or triple
bonds. Explain why this is generally true. Hint: What properties of elements change with increasing period?

S12.30
Pi bonds are usually weaker than sigma bonds because there is less orbital overlap. The larger atomic radius of period 3 and greater
elements means that there is even less overlap between the orbitals that would form pi bonds. There is not enough overlap to form
the pi bonds that would make for double or triple bonds.

Q12.30
Explain why the Si=Si double bond is not stable than the C=C. Given the covalent radius of carbon is 77 pm and Silicon is 111pm?

S12.30
Silicon is significantly bigger than carbon. Thus the Si-Si bond is longer than C-C bond. However, the longer bond result in
lowering the effective overlap of 3p orbital to form a π bonds. In contrast, C atoms can form the π bond from 2p orbital. Thus the
Si=Si is less stable despite its larger size.

Q12.33
Myohemerythrin is an iron-containing protein that binds oxygen in marine invertebrates. The protein is monomeric (has one
subunit) and contains 0.8107% iron by mass. Estimate the molar mass of myohemerythrin. Is your answer the same as the known
molar mass of myohemerythrin, 13,780 grams per mole? If it is not, account for the difference between your estimate and the
known value.

Myohemerythrin ribbon schematic, a 4-helix bundle protein structure with a bound iron. Hand drawn by Jane Richardson in 1980,
from PDB file 2MHR.

S12.33
1. Find the mass of the protein if there is one atom of iron per protein.
100.0 g protein
55.85 g F e/mol × (12.E.6)
0.8107 g F e

12.E.6 https://chem.libretexts.org/@go/page/41338
6889g/mol (12.E.7)

2. This is much less than the known molar mass. In fact, it is half.
13780 g/mol
= 2.000 (12.E.8)
6889 g/mol

We know the protein is monomeric, so there must be two atoms of iron in each protein.

Q12.19
Neon is a noble gas. Draw the molecular orbital diagram of N e2 and N e . From that, describe the bonding scheme of those two
2

molecules based on molecular orbital theory.

S12.19
Based on the molecular orbital diagram, there are more anti-bonding than bonding state. The bonding order of Ne2 is:
BO = 1/2 (8-8) = 0
Therefore, N e does not exist in nature.
2

Q12.21
What is the difference between the last three group, including O2, F2 and the other diatomic molecules in the second row of
Periodic table in term of Molecular orbital theory?

S12.21
For O2 and F2, the pi bonds(2p) have higher energy than the sigma bonds (2p). Whereas, a group of B2, C2, and N2, the sigma
bonds(2p) have higher energy than the pi bonds(2p)

Q12.30
Be is in the second period, and very rare Be2 exist in nature. Explain the instability of Be2 , and calculate the bond order.

S12.30
Be2 molecule has 2 valence electrons. The bonding order, BO= 1/2 (2-2)= 0 Therefore, diatomic Be2 is not stable! Not exist in
nature

12.4: Electronegativity and Dipole Moment


Q12.36
Will the dipole moment of a cis-dichloroethylene molecule increase or decrease upon heating? Why?

12.E.7 https://chem.libretexts.org/@go/page/41338
S12.36
The dipole moment will decrease because cis-trans isomerization will occur when heated. This is because the trans isomer of the
molecule does not have a dipole moment present.

Q12.37
Draw the σ and σ molecular orbital of

CO . Draw the MO energy level diagram and write the electron coefficient. Calculate the
Bond order.

S12.37

Electron Configuration:
( 1s)2( *1s)2( 2s)2( *2s)2( x)2( y)2( 2sp)2
Bond Order: 1/2(Bonding Electrons-Antibonding Electrons)
=1/2(8-2) = 3

12.6: Diatomic Molecules

12.7: Resonance and Electron Delocalization

12.8: Coordination Compounds

12.9: Coordination Compounds in Biological Systems

12.E: The Chemical Bond (Exercises) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

12.E.8 https://chem.libretexts.org/@go/page/41338
CHAPTER OVERVIEW

13: Intermolecular Forces


Intermolecular forces are forces of attraction or repulsion which act between neighboring particles (atoms, molecules, or ions).
They are weak compared to the intramolecular forces, the forces which keep a molecule together.
13.1: Intermolecular Interactions
13.2: The Ionic Bond
13.3: Types of Intermolecular Forces
13.4: Hydrogen Bonding
13.5: The Structure and Properties of Water
13.6: Hydrophobic Interaction
13.E: Intermolecular Forces (Exercises)

13: Intermolecular Forces is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1
13.1: Intermolecular Interactions
Learning Objectives
Classify intermolecular forces as ionic, covalent, London dispersion, dipole-dipole, or hydrogen bonding.
Explain properties of material in terms of type of intermolecular forces.
Predict the properties of a substance based on the dominant intermolecular force.

Forces binding atoms in a molecule are due to chemical bonding. The energy required to break a bond is called the bond-energy.
For example, the average bond-energy for O−H bonds in water is 463 kJ/mol. On average, 463 kJ is required to break 6.023x1023
O−H bonds, or 926 kJ to convert 1.0 mole of water into 1.0 mol of O and 2.0 mol of H atoms. The forces holding molecules

together are generally called intermolecular forces. The energy required to break molecules apart is much smaller than a typical
bond-energy, but intermolecular forces play important roles in determining the properties of a substance. Intermolecular forces are
particularly important in terms of how molecules interact and form biological organisms or even life. This link gives an excellent
introduction to the interactions between molecules.

Classifying Intermolecular Forces


In general, intermolecular forces can be divided into several categories. The four prominent types are:
Ion-Ion Interactions: Recall lattice energy and its relation to properties of solids. The more ionic, the higher the lattice energy.
Examine the following list and see if you can explain the observed values by way of ionic attraction: LiF = 1036; LiI = 737 ;
KF = 821 ; Mg F = 2957 kJ/mol
2

Dipole-dipole Interactions: Substances whose molecules have dipole moment have a higher melting point or boiling point than
those of similar molecular mass, but whose molecules have no dipole moment.
London dispersion forces or van der Waal's force: These forces always operate in any substance. The force arisen from
induced dipole and the interaction is weaker than the dipole-dipole interaction. In general, the heavier the molecule, the stronger
the van der Waal's force of interaction. For example, the boiling points of inert gases increase as their atomic masses increase
due to stronger London dispersion interactions.
Hydrogen bonds: Certain substances such as H O , HF, and NH form hydrogen bonds, which affects properties (mp, bp,
2 3

solubility) of the substance. Other compounds containing OH and NH groups also form hydrogen bonds. Molecules of many
2

organic compounds such as alcohols, acids, amines, and amino acids contain these groups, and thus hydrogen bonding plays a
important role in biological science.
Covalent bonding: Covalent bonding is really intramolecular force rather than intermolecular force. It is mentioned here,
because some solids are formed due to covalent bonding. For example, in diamond, silicon, quartz etc., all the atoms in the
entire crystal are linked together by covalent bonding. These solids are hard, brittle, and have high melting points. Covalent
bonding holds atoms tighter than ionic attraction.
Metallic bonding: Forces between atoms in metallic solids belong to another category. Valence electrons in metals are rampant.
They are not restricted to certain atoms or bonds. Rather they run freely in the entire solid, providing good conductivity for heat
and electric energy. This behavior of electrons gives special properties such as ductility and mechanical strength to metals.
The division into types is for convenience in their discussion. Of course all types can be present simultaneously for many
substances. Usually, intermolecular forces are discussed together with The States of Matter. Intermolecular forces also play
important roles in solutions, a discussion of which is given in Hydration, solvation in water. A summary of the interactions is
illustrated in the following diagram:

13.1.1 https://chem.libretexts.org/@go/page/41388
See if you can answer the following questions.
What are dipoles?
What are dipole moments?
How do dipoles interact?
Why do molecules attract one another?
How do London dispersion forces come about?
What parameters cause an increase of the London dispersion forces?
What is a hydrogen bond?
What type of hydrogen bonds are strong?
What chemical groups are hydrogen acceptors for hydrogen bonds?
If you are looking for specific information, your study will be efficient. Some answers can be found in the Confidence Building
Questions. Consider carefully the purpose of each question, and figure out what there is to be learned in it.

Confidence Building Questions


1. Which of the following molecules have a permanent dipole moment?
a. H O 2

b. CO 2

c. CH 4

d. N 2

e. CO
f. NH 3

Hint: a e f
Discussion -
CO , CH , and N are symmetric, and hence they have no permanent dipole moments. A molecule with polar bonds
2 4 2

unsymmetrically arranged will possess a permanent dipole.


2. Which has the higher boiling point, Br or ICl?
2

Hint: ICl
Discussion -
They have similar molecular weights: Br 2 = 160 ; ICl = 162. Their boiling points are 332 K and 370 K respectively.
3. An atom or molecule can be temporarily polarized by a nearby species. Polarization separates centers of charge giving
a. permanent dipole

13.1.2 https://chem.libretexts.org/@go/page/41388
b. temporary charges
c. hydrogen bonding
d. induced dipole
e. induced ions
f. radicals
Hint: d.
Discussion -
Induced dipoles are responsible for the London dispersion forces. The heavier the molecule, the larger the induced dipole will
be. Thus, London dispersion forces are strong for heavy molecules.
4. Which has a higher boiling point, I or Br ?
2 2

Hint: iodine, I . 2

Discussion -
Atomic weights for Br and I are 80 and 127 respectively. The higher the molecular weight, the stronger the London dispersion
forces.
5. If only London dispersion forces are present, which should have a lower boiling point, H 2
O or H
2
S ?
Hint: water, H 2
O

Discussion -
The b.p. for H O is 100 deg C, and that of H
2 2
S is -70 deg C. Very strong hydrogen bonding is present in liquid H O
2
, but no
hydrogen bonding is present in liquid H S . 2

6. Contrary to most other substances, the density of water decreases as temperature decreases between 4 and 0 deg C. This is due
to
a. dipole-dipole interaction
b. London dispersion
c. decreasing number of hydrogen bonds
d. increasing number of hydrogen bonds formed
Hint: d.
Discussion -
As more hydrogen bonds form when the temperature decreases, the volume expands, causing a decrease in density. Above 4
deg C, the thermal expansion is more prominent than the effect of hydrogen bonds.
7. Ethanol (C 2
H OH
5
, molar mass 46) boils at 351 K, but water (H 2
O , molar mass 18) boils at higher temperature, 373 K. This is
because:
a. water is denser
b. water has stronger London dispersion forces
c. water has stronger hydrogen bonds
d. water molecules contain no carbon
Hint: c.
Discussion -
A hydrogen atom between two small, electronegative atoms (such as F , O, N ) causes a strong intermolecular interaction known
as the hydrogen bond. The strength of a hydrogen bond depends upon the electronegativities and sizes of the two atoms.
8. Ethanol (C 2
H OH
5
) and methyl ether (CH 3
OCH
3
) have the same molar mass. Which has a higher boiling point?
Hint: Ethanol has a higher boiling point.
Discussion -
R−OH group is both proton donor and acceptor for hydrogen bonding. Methyl groups have very weak hydrogen bonding, if

any.

13.1.3 https://chem.libretexts.org/@go/page/41388
Contributors and Attributions
Chung (Peter) Chieh (Professor Emeritus, Chemistry @ University of Waterloo)

13.1: Intermolecular Interactions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

13.1.4 https://chem.libretexts.org/@go/page/41388
13.2: The Ionic Bond
The interactions between ions (ion - ion interactions) are the easiest to understand: like charges repel each other and opposite
charges attract. These Coulombic forces operate over relatively long distances in the gas phase. The force depends on the product
of the charges (Z , Z ) divided by the square of the distance of separation (d ):
1 2
2

−Z1 Z2
F ∝ (1)
2
d

Two oppositely-charged particles flying about in a vacuum will be attracted toward each other, and the force becomes stronger and
stronger as they approach until eventually they will stick together and a considerable amount of energy will be required to separate
them.

Figure 1: Ion - Ion Interactions in the Gas Phase


They form an ion-pair, a new particle which has a positively-charged area and a negatively-charged area. There are fairly strong
interactions between these ion pairs and free ions, so that these the clusters tend to grow, and they will eventually fall out of the gas
phase as a liquid or solid (depending on the temperature).

Contributors and Attributions


Gary L Bertrand, Professor of Chemistry, University of Missouri-Rolla

13.2: The Ionic Bond is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
Ion - Ion Interactions is licensed CC BY-NC-SA 4.0.

13.2.1 https://chem.libretexts.org/@go/page/41389
13.3: Types of Intermolecular Forces
13.3: Types of Intermolecular Forces is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

13.3.1 https://chem.libretexts.org/@go/page/41390
13.4: Hydrogen Bonding
This page explains the origin of hydrogen bonding - a relatively strong form of intermolecular attraction.

The evidence for hydrogen bonding


Many elements form compounds with hydrogen. If you plot the boiling points of the compounds of the Group 4 elements with
hydrogen, you find that the boiling points increase as you go down the group.

The increase in boiling point happens because the molecules are getting larger with more electrons, and so van der Waals
dispersion forces become greater. If you repeat this exercise with the compounds of the elements in Groups 5, 6 and 7 with
hydrogen, something odd happens.

Although for the most part the trend is exactly the same as in group 4 (for exactly the same reasons), the boiling point of the
compound of hydrogen with the first element in each group is abnormally high. In the cases of NH3, H2O and HF there must be
some additional intermolecular forces of attraction, requiring significantly more heat energy to break. These relatively powerful
intermolecular forces are described as hydrogen bonds.

The origin of hydrogen bonding


The molecules which have this extra bonding are:

The solid line represents a bond in the plane of the screen or paper. Dotted bonds are going back into the screen or paper away from
you, and wedge-shaped ones are coming out towards you. Notice that in each of these molecules:
The hydrogen is attached directly to one of the most electronegative elements, causing the hydrogen to acquire a significant
amount of positive charge.
Each of the elements to which the hydrogen is attached is not only significantly negative, but also has at least one "active" lone
pair.
Lone pairs at the 2-level have the electrons contained in a relatively small volume of space which therefore has a high density
of negative charge. Lone pairs at higher levels are more diffuse and not so attractive to positive things.
Consider two water molecules coming close together.

13.4.1 https://chem.libretexts.org/@go/page/41391
The + hydrogen is so strongly attracted to the lone pair that it is almost as if you were beginning to form a co-ordinate (dative
covalent) bond. It doesn't go that far, but the attraction is significantly stronger than an ordinary dipole-dipole interaction.
Hydrogen bonds have about a tenth of the strength of an average covalent bond, and are being constantly broken and reformed in
liquid water. If you liken the covalent bond between the oxygen and hydrogen to a stable marriage, the hydrogen bond has "just
good friends" status.

Water as a "perfect" example of hydrogen bonding


Notice that each water molecule can potentially form four hydrogen bonds with surrounding water molecules. There are exactly the
right numbers of + hydrogens and lone pairs so that every one of them can be involved in hydrogen bonding. This is why the
boiling point of water is higher than that of ammonia or hydrogen fluoride. In the case of ammonia, the amount of hydrogen
bonding is limited by the fact that each nitrogen only has one lone pair. In a group of ammonia molecules, there aren't enough lone
pairs to go around to satisfy all the hydrogens. In hydrogen fluoride, the problem is a shortage of hydrogens. In water, there are
exactly the right number of each. Water could be considered as the "perfect" hydrogen bonded system.
The diagram shows the potential hydrogen bonds formed to a chloride ion, Cl-. Although the lone pairs in the chloride ion are at the
3-level and wouldn't normally be active enough to form hydrogen bonds, in this case they are made more attractive by the full
negative charge on the chlorine.

However complicated the negative ion, there will always be lone pairs that the hydrogen atoms from the water molecules can
hydrogen bond to.

Hydrogen bonding in alcohols


An alcohol is an organic molecule containing an -O-H group. Any molecule which has a hydrogen atom attached directly to an
oxygen or a nitrogen is capable of hydrogen bonding. Such molecules will always have higher boiling points than similarly sized
molecules which don't have an -O-H or an -N-H group. The hydrogen bonding makes the molecules "stickier", and more heat is
necessary to separate them.
Ethanol, CH3CH2-O-H, and methoxymethane, CH3-O-CH3, both have the same molecular formula, C2H6O.

They have the same number of electrons, and a similar length to the molecule. The van der Waals attractions (both dispersion
forces and dipole-dipole attractions) in each will be much the same. However, ethanol has a hydrogen atom attached directly to an
oxygen - and that oxygen still has exactly the same two lone pairs as in a water molecule. Hydrogen bonding can occur between

13.4.2 https://chem.libretexts.org/@go/page/41391
ethanol molecules, although not as effectively as in water. The hydrogen bonding is limited by the fact that there is only one
hydrogen in each ethanol molecule with sufficient + charge.
In methoxymethane, the lone pairs on the oxygen are still there, but the hydrogens aren't sufficiently + for hydrogen bonds to
form. Except in some rather unusual cases, the hydrogen atom has to be attached directly to the very electronegative element for
hydrogen bonding to occur. The boiling points of ethanol and methoxymethane show the dramatic effect that the hydrogen bonding
has on the stickiness of the ethanol molecules:

ethanol (with hydrogen bonding) 78.5°C

methoxymethane (without hydrogen bonding) -24.8°C

The hydrogen bonding in the ethanol has lifted its boiling point about 100°C.
It is important to realize that hydrogen bonding exists in addition to van der Waals attractions. For example, all the following
molecules contain the same number of electrons, and the first two are much the same length. The higher boiling point of the butan-
1-ol is due to the additional hydrogen bonding.

Comparing the two alcohols (containing -OH groups), both boiling points are high because of the additional hydrogen bonding due
to the hydrogen attached directly to the oxygen - but they are not the same. The boiling point of the 2-methylpropan-1-ol isn't as
high as the butan-1-ol because the branching in the molecule makes the van der Waals attractions less effective than in the longer
butan-1-ol.

Hydrogen bonding in nitrogen containing organic molecules


Hydrogen bonding also occurs in organic molecules containing N-H groups - in the same sort of way that it occurs in ammonia.
Examples range from simple molecules like CH3NH2 (methylamine) to large molecules like proteins and DNA. The two strands of
the famous double helix in DNA are held together by hydrogen bonds between hydrogen atoms attached to nitrogen on one strand,
and lone pairs on another nitrogen or an oxygen on the other one.

Contributors and Attributions


Jim Clark (Chemguide.co.uk)

13.4: Hydrogen Bonding is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
Hydrogen Bonding by Jim Clark is licensed CC BY-NC 4.0.

13.4.3 https://chem.libretexts.org/@go/page/41391
13.5: The Structure and Properties of Water
With 70% of our earth being ocean water and 65% of our bodies being water, it is hard to not be aware of how important it is in our
lives. There are 3 different forms of water, or H2O: solid (ice), liquid (water), and gas (steam). Because water seems so ubiquitous,
many people are unaware of the unusual and unique properties of water, including:
Boiling Point and Freezing Point
Surface Tension, Heat of Vaporization, and Vapor Pressure
Viscosity and Cohesion
Solid State
Liquid State
Gas State

Boiling Point and Freezing Point


If you look at the periodic table and locate tellurium (atomic number: 52), you find that the boiling points of hydrides decrease as
molecule size decreases. So the hydride for tellurium: H2Te (hydrogen telluride) has a boiling point of -4°C. Moving up, the next
hydride would be H2Se (hydrogen selenide) with a boiling point of -42°C. One more up and you find that H2S (hydrogen sulfide)
has a boiling point at -62°C. The next hydride would be H2O (WATER!). And we all know that the boiling point of water is
100°C. So despite its small molecular weight, water has an incredibly big boiling point. This is because water requires more energy
to break its hydrogen bonds before it can then begin to boil. The same concept is applied to freezing point as well, as seen in the
table below. The boiling and freezing points of water enable the molecules to be very slow to boil or freeze, this is important to the
ecosystems living in water. If water was very easy to freeze or boil, drastic changes in the environment and so in oceans or lakes
would cause all the organisms living in water to die. This is also why sweat is able to cool our bodies.

COMPOUND BOILING POINT FREEZING POINT

Hydrogen Telluride -4°C -49°C

Hydrogen Selenide -42°C -64°C

Hydrogen Sulfide -62°C -84°C

Water 100°C 0 °C

Surface Tension, Heat of Vaporization, and Vapor Pressure


Besides mercury, water has the highest surface tension for all liquids. Water's high surface tension is due to the hydrogen bonding
in water molecules. Water also has an exceptionally high heat of vaporization. Vaporization occurs when a liquid changes to a gas,
which makes it an endothermic reaction. Water's heat of vaporization is 41 kJ/mol. Vapor pressure is inversely related to
intermolecular forces, so those with stronger intermolecular forces have a lower vapor pressure. Water has very strong
intermolecular forces, hence the low vapor pressure, but it's even lower compared to larger molecules with low vapor pressures.
Viscosity is the property of fluid having high resistance to flow. We normally think of liquids like honey or motor oil being
viscous, but when compared to other substances with like structures, water is viscous. Liquids with stronger intermolecular
interactions are usually more viscous than liquids with weak intermolecular interactions.
Cohesion is intermolecular forces between like molecules; this is why water molecules are able to hold themselves together in a
drop. Water molecules are very cohesive because of the molecule's polarity. This is why you can fill a glass of water just barely
above the rim without it spilling.

Solid State (Ice)


All substances, including water, become less dense when they are heated and more dense when they are cooled. So if water is
cooled, it becomes more dense and forms ice. Water is one of the few substances whose solid state can float on its liquid state!
Why? Water continues to become more dense until it reaches 4°C. After it reaches 4°C, it becomes LESS dense. When freezing,
molecules within water begin to move around more slowly, making it easier for them to form hydrogen bonds and eventually
arrange themselves into an open crystalline, hexagonal structure. Because of this open structure as the water molecules are being
held further apart, the volume of water increases about 9%. So molecules are more tightly packed in water's liquid state than its
solid state. This is why a can of soda can explode in the freezer.

13.5.1 https://chem.libretexts.org/@go/page/41392
Liquid State (Liquid Water)
It is very rare to find a compound that lacks carbon to be a liquid at standard temperatures and pressures. So it is unusual for water
to be a liquid at room temperature! Water is liquid at room temperature so it's able to move around quicker than it is as solid,
enabling the molecules to form fewer hydrogen bonds resulting in the molecules being packed more closely together. Each water
molecule links to four others creating a tetrahedral arrangement, however they are able to move freely and slide past each other,
while ice forms a solid, larger hexagonal structure.

Gas State (Steam)


As water boils, its hydrogen bonds are broken. Steam particles move very far apart and fast, so barely any hydrogen bonds have the
time to form. So, less and less hydrogen bonds are present as the particles reach the critical point above steam. The lack of
hydrogen bonds explains why steam causes much worse burns that water. Steam contains all the energy used to break the hydrogen
bonds in water, so when steam hits your face you first absorb the energy the steam has taken up from breaking the hydrogen bonds
it its liquid state. Then, in an exothermic reaction, steam is converted into liquid water and heat is released. This heat adds to the
heat of boiling water as the steam condenses on your skin.

Water as the "Universal Solvent"


Because of water's polarity, it is able to dissolve or dissociate many particles. Oxygen has a slightly negative charge, while the two
hydrogens have a slightly positive charge. The slightly negative particles of a compound will be attracted to water's hydrogen
atoms, while the slightly positive particles will be attracted to water's oxygen molecule; this causes the compound to dissociate.
Besides the explanations above, we can look to some attributes of a water molecule to provide some more reasons of water's
uniqueness:
Forgetting fluorine, oxygen is the most electronegative non-noble gas element, so while forming a bond, the electrons are pulled
towards the oxygen atom rather than the hydrogen. This creates 2 polar bonds, which make the water molecule more polar than
the bonds in the other hydrides in the group.
A 104.5° bond angle creates a very strong dipole.
Water has hydrogen bonding which probably is a vital aspect in waters strong intermolecular interaction

Why is this important for the real world?


The properties of water make it suitable for organisms to survive in during differing weather conditions. Ice freezes as it expands,
which explains why ice is able to float on liquid water. During the winter when lakes begin to freeze, the surface of the water
freezes and then moves down toward deeper water; this explains why people can ice skate on or fall through a frozen lake. If ice
was not able to float, the lake would freeze from the bottom up killing all ecosystems living in the lake. However ice floats, so the
fish are able to survive under the surface of the ice during the winter. The surface of ice above a lake also shields lakes from the
cold temperature outside and insulates the water beneath it, allowing the lake under the frozen ice to stay liquid and maintain a
temperature adequate for the ecosystems living in the lake to survive.

Resources
1. Cracolice, Mark S. and Edward Peters I. Basics of Introductory Chemistry. Thompson, Brooks/Cole Publishing Company. 2006
2. Petrucci, et al. General Chemistry: Principles & Modern Applications: AIE (Hardcover). Upper Saddle River: Pearson/Prentice
Hall, 2007.

Contributors and Attributions


Corinne Yee (UCD), Desiree Rozzi (UCD)

13.5: The Structure and Properties of Water is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.
Unusual Properties of Water is licensed CC BY-NC-SA 4.0.

13.5.2 https://chem.libretexts.org/@go/page/41392
13.6: Hydrophobic Interaction
Hydrophobic interactions describe the relations between water and hydrophobes (low water-soluble molecules). Hydrophobes are
nonpolar molecules and usually have a long chain of carbons that do not interact with water molecules. The mixing of fat and water
is a good example of this particular interaction. The common misconception is that water and fat doesn’t mix because the Van der
Waals forces that are acting upon both water and fat molecules are too weak. However, this is not the case. The behavior of a fat
droplet in water has more to do with the enthalpy and entropy of the reaction than its intermolecular forces.

Causes of Hydrophobic Interactions


American chemist Walter Kauzmann discovered that nonpolar substances like fat molecules tend to clump up together rather than
distributing itself in a water medium, because this allow the fat molecules to have minimal contact with water.

The image above indicates that when the hydrophobes come together, they will have less contact with water. They interact with a
total of 16 water molecules before they come together and only 10 atoms after they interact.

Thermodynamics of Hydrophobic Interactions


When a hydrophobe is dropped in an aqueous medium, hydrogen bonds between water molecules will be broken to make room for
the hydrophobe; however, water molecules do not react with hydrophobe. This is considered an endothermic reaction, because
when bonds are broken heat is put into the system. Water molecules that are distorted by the presence of the hydrophobe will make
new hydrogen bonds and form an ice-like cage structure called a clathrate cage around the hydrophobe. This orientation makes the
system (hydrophobe) more structured with an decrease of the total entropy of the system; therefore ΔS < 0 .

The change in enthalpy (ΔH ) of the system can be negative, zero, or positive because the new hydrogen bonds can partially,
completely, or over compensate for the hydrogen bonds broken by the entrance of the hydrophobe. The change in enthalpy,
however, is insignificant in determining the spontaneity of the reaction (mixing of hydrophobic molecules and water) because the
change in entropy (ΔS ) is large.
According to the Gibbs Energy formula

ΔG = ΔH − T ΔS (13.6.1)

with a small unknown value of ΔH and a large negative value of ΔS , the value of ΔG will turn out to be positive. A positive ΔG
indicates that the mixing of the hydrophobe and water molecules is not spontaneous.

13.6.1 https://chem.libretexts.org/@go/page/41393
Formation of Hydrophobic Interactions
The mixing hydrophobes and water molecules is not spontaneous; however, hydrophobic interactions between hydrophobes are
spontaneous. When hydropobes come together and interact with each other, enthalpy increases ( ΔH is positive) because some of
hydrogen bonds that form the clathrate cage will be broken. Tearing down a portion of the clathrate cage will cause the entropy to
increase ( ΔS is positive), since forming it decreases the entropy.
According to the Equation 13.6.1
ΔH = small positive value
ΔS = large positive value
Result: ΔG is negative and hence hydrophobic interactions are spontaneous.

Strength of Hydrophobic Interactions


Hydrophobic interactions are relatively stronger than other weak intermolecular forces (i.e., Van der Waals interactions or
Hydrogen bonds). The strength of Hydrophobic Interactions depend on several factors including (in order of strength of influence):
1. Temperature: As temperature increases, the strength of hydrophobic interactions increases also. However, at an extreme
temperature, hydrophobic interactions will denature.
2. Number of carbons on the hydrophobes: Molecules with the greatest number of carbons will have the strongest hydrophobic
interactions.
3. The shape of the hydrophobes: Aliphatic organic molecules have stronger interactions than aromatic compounds. Branches on
a carbon chain will reduce the hydrophobic effect of that molecule and linear carbon chain can produce the largest hydrophobic
interaction. This is so because carbon branches produce steric hindrance, so it is harder for two hydrophobes to have very close
interactions with each other to minimize their contact to water.

Biological Importance of Hydrophobic Interactions


Hydrophobic Interactions are important for the folding of proteins. This is important in keeping a protein stable and biologically
active, because it allow to the protein to decrease in surface are and reduce the undesirable interactions with water. Besides from
proteins, there are many other biological substances that rely on hydrophobic interactions for its survival and functions, like the
phospholipid bilayer membranes in every cell of your body!

13.6.2 https://chem.libretexts.org/@go/page/41393
Illustration of how protein changes shape to allow polar regions (blue) to interact with water while non-polar hydrophobic regions
(red) do not interact with the water. (CC BY-SA 3.0; Treshphrd via Wikipedia).

References
1. Atkins, Peter and Julio de Paula. Physical Chemistry for the Life Sciences. Oxford, UK: Oxford University Press. 2006. 95.
2. Chang, Raymond. Physical Chemistry for the Biosciences. Sausalito, CA: Edwards Brothers, Inc. 2005. 508-510.
3. Garrett, Reginald H. and Charles M. Grisham. Biochemistry. Belmont, CA: Thomas Brooks/Cole. 2005. 15.

13.6: Hydrophobic Interaction is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
Hydrophobic Interactions by Justin Than is licensed CC BY-SA 4.0.

13.6.3 https://chem.libretexts.org/@go/page/41393
13.E: Intermolecular Forces (Exercises)
13.1: Intermolecular Interactions
Q13.1a
List all the intermolecular interactions that take place in each of the follow kings of molecules: C C l3 F , C C l2 F2 , , and
C C lF3

CF .
4

Q13.1b
Determine what type of intermolecular forces exist in the following molecules: LiF, MgF2, H2O, and HF.

S13.1b
H2O: London Force, Dipole-Dipole interaction, Hydrogen bonds.
HF: Dipole-Dipole intermolecular forces, Hydrogen bonds.
MgF2 and LiF: strong ionic attraction.

Q13.2a
Arrange the follow species in order of decreasing melting points: CsBr, KI, KCL, MgF2.

Q13.2b
Which has the highest boiling point I2, Br2, and Cl2. Explain why?

S13.2b
The atomic weigh of Iodine = 127, Bromine = 80, and Chlorine = 35.5. The weigh is proportion to the London dispersion force,
and the higher molecular weigh, the larger the force. Thus, I2 has a highest boiling point.

Q13.3
1-Propanol C3H7OH and methoxyethane CH3O C2H5 have the same molecular weigh. Which has the higher boiling point?

S13.3
The 1-Propanol can form London Force, Dipole- Dipole, and H- bonding due to the H bonded to O atom of OH group, whereas the
methoxyethane can not form the H-bonding. Therefore, the 1-Propanol has higher intermolecular attractive force and thus a higher
boiling point.

Q13.3
Why do the lightest compounds such as NH3, H2O, and HF have the highest boiling points?

Q13.4a
What kind of attractive interaction exists between atoms and between nonpolar molecules?

Q13.4b
Why nature gas CH4 is a good choice to storage tank in winter?

S13.4b
It's about boiling point. The methane has the boiling point at -161 °C, making it to be a good choice for winter season.

Q13.4
Explain why methane (CH_4\) is used as the primary heating gas in Alaska during wintertime instead of the more commonly used
butant or propane gases use in the lower 48 states.

S13.4
Methane (C H ) remains gas because its boiling point is about -160°C. Other gases, such as propane or butane, would liquefy
4

under freezing condition. Therefore, methane is more likely to be used during wintertime at Alaska

13.E.1 https://chem.libretexts.org/@go/page/41339
Q13.5
Define types of intermolecular forces and give example for each.

S13.5
There are 3 types of intermolecular force: London Dispersion, Dipole-Dipole (Example: Two N aC l) and Ion-Dipole (Example:
Mg
+
and H C l)
Dipole- Dipole occurs between polar molecules
Ion- Dipole occurs between an ion and polar molecules
London Dispersion occurs between the nonpolar molecules.

Q13.6
How does the intermolecular determine the boiling point?

S13.6
The weakest intermolecular, the lowest boiling point.

Q13.7a
What is their dipole-dipole interaction of wo HCl molecules are co-linear head-to tail.
Given: The dipole moment of HF is 1.86 D. The dipole moment of HCl is 1.05 D. The distance between the two is 1.78
H+-F- - - - - H+-Cl-

S13.7a
2μA μB
V =− (13.E.1)
3
4πε0 r

2(1.05)(1.86)
V =− (13.E.2)
−12 3
4π(8.854187817 ⋅ 10 )(1.78 )

13.2: The Ionic Bond


Q13.7b
Calculate an ion-ion interaction energy between K +
and C l at a distance of 600 pm.

S13.7b
The ion-ion interaction energy is given by Coulomb's law.
q1 q2
V = (13.E.3)
4π ϵo r

−19 −19
−(1.602 × 10 C )(1.602 × 10 C )
−19
V = = −3.84 × 10 J (13.E.4)
−12 2 −1 −2 −10
4π(8.853 × 10 C ⋅N ⋅m )(6 × 10 m )

Q13.7
Calculate an ion-dipole interaction energy between K +
and H C l at a distance of 600 pm. H C l has a dipole moment of 1.08 D.

S13.7
30
3.3356 × 10 C ⋅m −30
μ = 1.08 D × = 3.6 × 10 C ⋅m (13.E.5)
1 D

−19 −30
−q μ −(1.602 × 10 C )(3.6 × 10 C ⋅ m )
−20
V = = = −1.44 × 10 J (13.E.6)
4πϵo r2 −12 2 −1 −2 −10 2
4π(8.853 × 10 C ⋅N ⋅m )(6 × 10 m )

As expected this is appreciably smaller in energy than covalent bonds (e..g, H C l has a bond enthalpy of 7.0 × 10 −19
J ).

13.E.2 https://chem.libretexts.org/@go/page/41339
13.3: Types of Intermolecular Forces
Q13.8a
Rank the interactions from weakest to strongest:
a. H2O - - OH2
b. Li+ - - F-
c. Li+ - - OH2

S13.8a
a. ion-ion interaction: Li+ - - F-
b. ion-dipole interaction: Li+ - - OH2
c. dipole-dipole interaction: H2O - - OH2

Q13.8b
A low concentration electrolytic solution behaves non-ideally while a high concentration of the same solution behaves ideally.
Explain this phenomenon in terms of forces, noting that Coulomb forces depend on 1/r2 while van der Waals forces depend on 1/r7.
Which of these forces are low concentration electrolytic solutions likely to follow? High concentration?

S13.8b
The interatomic distances in a low concentration electrolytic solution are greater than those in a high concentration solution. They
follow van der Waals forces and thus behave less ideally. High concentration electrolytic solutions follow Coulomb forces.

Q13.9a
Calculate the ion-dipole interaction between H2O and Li+. You are given the dipole moment of H2O is 1.82 D. The distance
between these two is 2 Å.

S13.9a

V =− (13.E.7)
2
4πε0 r

−30
3.3356⋅ 10 C m)
1.82D ⋅ (
1D
= = 1.36 kJ/mol (13.E.8)
−12 −10 2

4π(8.85 ⋅ 10 )(2 ⋅ 10 m)

Q13.9b
Calculate the potential energy of interaction between a Cl- ion situated 120 pm away from an H 2O molecule with a dipole moment
of 1.85 D.

S13.9b
30
3.3356 × 10 C ⋅m
−30
μ = 1.85 D × = 6.18 × 10 C ⋅m (13.E.9)
1 D

−10
r = 1.2 × 10 m (13.E.10)

−19 −30
qμ (−1.602 × 10 C )(6.18 × 10 C ⋅ m)
V = = (13.E.11)
2 −12 −2 −1 −2 −10 2
4πεo r 4π(8.851 × 10 C N m )(1.2 × 10 m)

Q13.10
Do you expect a greater dipole-dipole interaction between two molecules that are antiparallel or between two molecules that are co-
linear head-to-tail?

S13.10
You expect a stronger interaction when the two are co-linear head-to tail. This can be seen by looking at the formula or in the
images of the two.

13.E.3 https://chem.libretexts.org/@go/page/41339
Q13.10
Express the equilibrium distance re in term ð and show V = - €

Q13.12
a. Determine Vander Waals radius of Argon.
b. Use this radius to find fraction of volume by 2 mole of argon at room temperature at 1 atm.

S13.12
a. r= σ/2 = 3.40 A0 /2 = 1.70 A0
b. Volume of 2 mole of Ar
4/3 πr^3 ((6.022 x 10^23)/(2 mol))= 4/3 π (1.70 x 10^(-10) m)^3 ((6.022 x 10^23)/(2 mol))
= 6.19 x 10-6 L mol-1
V/n=RT/P= ((0.08206 L atm K^(-1) mol^(-1) (298.2 K))/1atm
= 24.5 L mol-1
The fraction of this volume occupied by 2 mole of Ar
(1.239 x 10^-2 L mol-1)/ 24.47 L mol-1 = 2.5 x10 -7

Q13.15
a) What is the original of polarity in a molecule?
b) Is CO2 polar? Explain.

Q13.16
Of the following compound, which one(s) is/are soluble?
a. CH4
b. NCl3
c. C6H6
d. CO(NH2)2

Q13.17
What makes a compound soluble in water? Explain using examples.

13.E.4 https://chem.libretexts.org/@go/page/41339
Q13.18
Explain why does water have a high specific heat.

Q13.20
The energy of a hydrogen bond for each base pair in DNA is 15 kJ/mol. Two complimentary strands has 50 base pairs each. What
is the ratio of the 2 different strands to hydrogen double helix in a solution given a temperature of 300 K.

S13.20
First calculate the ratio of the two different strands for just one pair.
ΔE/RT 3 −3
e = exp[(15 × 10 J/mol)/(8.314 J/K ∗ mol)(300 K) = 2.4 × 10 (13.E.12)

Since there are 50 base pairs, we need to multiply by 50 to account for all the base pairs.
exp[100X(15X103 J/mol)/(8.314 J/K*mol)(300K) = 0

Q13.24
Consider two pure liquids. One has strong intermolecular interactions, and the other has relatively weak intermolecular
interactions. For the following properties, indicate which of the liquids you would expect to have a higher value (answer with
"strong" or "weak").
a. viscosity
b. vapor pressure
c. freezing point
d. surface tension

S13.24
a. Strong. Higher viscosity results from stronger interactions between the liquid molecules.
b. Weak. The liquid with weaker bonds takes less energy to turn into vapor, so it will exert a higher vapor pressure.
c. Strong. The freezing point is the same as the melting point; it takes more energy to melt a solid with stronger intermolecular
interactions.
d. Strong. Surface tension is a result of intermolecular interactions. The stronger these interactions, the greater the surface tension.

Q13.25
Fun fact: if the DNA in a single human cell were stretched out (but still in its familiar double helix conformation), it would be
approximately 2 meters long. The distance, along the helix, between nucleotides is 3.4 Å.
a. Estimate the number of basepairs in the haploid human genome, from the 2 meter fun fact.
b. The human body contains about 100 trillion cells. About a quarter of these are erythrocytes (red blood cells) and contain no
genomic DNA. Use the average molar mass for a basepair, 650 grams per mole, to estimate how much of a human's mass is
human genomic DNA.
c. At its closest, Pluto is 4.28 billion km from Earth. Do you have enough DNA to reach Pluto?
Hint: Humans are diploid.

S13.25
(a)
$$ 2\ m/cell \times \dfrac{bp}{3.4\ Å} \times \dfrac{10^{10}\ Å}{m} \times \dfrac{cell}{2\ haploid\ genomes} = 3 \times 10^{9}
\dfrac{bp}{haploid\ genome} \]
Three billion basepairs.
(b) 75 trillion of the human cells in your body have genomic DNA.
$$ 75 \times 10^{20}\ cells \times \dfrac{haploid\ genomes}{cell} \times \dfrac{3 \times 10^9\ bp}{haploid\ genome} \times
\dfrac{mol}{6.022 \times 10^{23}} \times 650 \dfrac{g}{mol\ bp} = 200\ g \]
That's about half a pound.

13.E.5 https://chem.libretexts.org/@go/page/41339
(c)
$$ \dfrac{2\ m}{cell} \times 75 \times 10^{12}\ cells \times \dfrac{km}{1000\ m} = 2 \times 10^{11}\ km \]
Yes, you have way more DNA than you need to stretch it from Earth to Pluto.

13.4: Hydrogen Bonding

13.5: The Structure and Properties of Water

13.6: Hydrophobic Interaction

13.E: Intermolecular Forces (Exercises) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

13.E.6 https://chem.libretexts.org/@go/page/41339
CHAPTER OVERVIEW

14: Spectroscopy
The focus of this chapter is on the interaction of ultraviolet, visible, and infrared radiation with matter. Because these techniques
use optical materials to disperse and focus the radiation, they often are identified as optical spectroscopies. For convenience we will
use the simpler term spectroscopy in place of optical spectroscopy; however, you should understand that we are considering only a
limited part of a much broader area of analytical techniques.
14.1: Vocabulary
14.2: Microwave Spectroscopy
14.3: Infrared Spectroscopy
14.4: Electronic Spectroscopy
14.5: Nuclear Magnetic Resonance
14.6: Electron Spin Resonance
14.7: Fluorescence and Phosphorescence
14.8: Lasers
14.9: Optical Rotatory Dispersion and Circular Dichroism
14.E: Spectroscopy (Exercises)

14: Spectroscopy is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1
14.1: Vocabulary
Electromagnetic radiation—light—is a form of energy whose behavior is described by the properties of both waves and particles.
Some properties of electromagnetic radiation, such as its refraction when it passes from one medium to another are explained best
by describing light as a wave. Other properties, such as absorption and emission, are better described by treating light as a particle.
The exact nature of electromagnetic radiation remains unclear, as it has since the development of quantum mechanics in the first
quarter of the 20th century.1 Nevertheless, the dual models of wave and particle behavior provide a useful description for
electromagnetic radiation.
Electromagnetic radiation consists of oscillating electric and magnetic fields that propagate through space along a linear path and
with a constant velocity. In a vacuum electromagnetic radiation travels at the speed of light, c, which is 2.997 × 108 m/s. When
electromagnetic radiation moves through a medium other than a vacuum its velocity, v, is less than the speed of light in a vacuum.
The difference between v and c is sufficiently small (<0.1%) that the speed of light to three significant figures, 3.00 × 108 m/s, is
accurate enough for most purposes.
The oscillations in the electric and magnetic fields are perpendicular to each other, and to the direction of the wave’s propagation.
Figure 14.1.1 shows an example of plane-polarized electromagnetic radiation, consisting of a single oscillating electric field and a
single oscillating magnetic field.

Figure 14.1.1 : Plane-polarized electromagnetic radiation showing the oscillating electric field in red and the oscillating magnetic
field in blue. The radiation’s amplitude, A, and its wavelength, λ, are shown. Normally, electromagnetic radiation is unpolarized,
with oscillating electric and magnetic fields present in all possible planes perpendicular to the direction of propagation.
An electromagnetic wave is characterized by several fundamental properties, including its velocity, amplitude, frequency, phase
angle, polarization, and direction of propagation.2 For example, the amplitude of the oscillating electric field at any point along the
propagating wave is

At = Ae sin(2πνt + ϕ) (14.1.1)

where At is the magnitude of the electric field at time t, Ae is the electric field’s maximum amplitude, ν is the wave’s frequency—
the number of oscillations in the electric field per unit time—and ϕ is a phase angle, which accounts for the fact that At need not
have a value of zero at t = 0. The identical equation for the magnetic field is

At = Am sin(2πνt + ϕ) (14.1.2)

where Am is the magnetic field’s maximum amplitude.

Units
Other properties also are useful for characterizing the wave behavior of electromagnetic radiation. The wavelength, λ, is defined as
the distance between successive maxima (Figure 14.1.1). For ultraviolet and visible electromagnetic radiation the wavelength is
usually expressed in nanometers (1 nm = 10–9 m), and for infrared radiation it is given in microns (1 μm = 10–6 m). The
relationship between wavelength and frequency is
c
λ = (14.1.3)
ν

Another unit useful unit is the wavenumber, ν~ , which is the reciprocal of wavelength
1
~
ν = (14.1.4)
λ

Wavenumbers are frequently used to characterize infrared radiation, with the units given in cm–1.

14.1.1 https://chem.libretexts.org/@go/page/41394
Example 14.1.1

In 1817, Josef Fraunhofer studied the spectrum of solar radiation, observing a continuous spectrum with numerous dark lines.
Fraunhofer labeled the most prominent of the dark lines with letters. In 1859, Gustav Kirchhoff showed that the D line in the
sun’s spectrum was due to the absorption of solar radiation by sodium atoms. The wavelength of the sodium D line is 589 nm.
What are the frequency and the wavenumber for this line?
Solution
The frequency and wavenumber of the sodium D line are
8
c 3.00 × 10 m/s
14 −1
ν = = = 5.09 × 10 s (14.1.5)
−9
λ 589 × 10 m

1 1 1 m
~ 4 −1
ν = = × = 1.70 × 10 cm (14.1.6)
−9
λ 589 × 10 m 100 cm

Exercise 14.1.1

Another historically important series of spectral lines is the Balmer series of emission lines form hydrogen. One of the lines
has a wavelength of 656.3 nm. What are the frequency and the wavenumber for this line?
Click here to review your answer to this exercise.

Regions of the Spectrum


The frequency and wavelength of electromagnetic radiation vary over many orders of magnitude. For convenience, we divide
electromagnetic radiation into different regions—the electromagnetic spectrum—based on the type of atomic or molecular
transition that gives rise to the absorption or emission of photons (Figure 14.1.2). The boundaries between the regions of the
electromagnetic spectrum are not rigid, and overlap between spectral regions is possible.

Figure 14.1.2 : The electromagnetic spectrum showing the boundaries between different regions and the type of atomic or
molecular transition responsible for the change in energy. The colored inset shows the visible spectrum. Source: modified from
Zedh (www.commons.Wikipedia.org).
Above, we defined several characteristic properties of electromagnetic radiation, including its energy, velocity, amplitude,
frequency, phase angle, polarization, and direction of propagation. A spectroscopic measurement is possible only if the photon’s
interaction with the sample leads to a change in one or more of these characteristic properties. We can divide spectroscopy into two
broad classes of techniques. In one class of techniques there is a transfer of energy between the photon and the sample. Table
14.1.1 provides a list of several representative examples.

Table 14.1.1 : Examples of Spectroscopic Techniques Involving an Exchange of Energy Between a Photon and the Sample
Region of
Type of Energy Transfer Electromagnetic Spectrum Spectroscopic Techniquea

14.1.2 https://chem.libretexts.org/@go/page/41394
Region of
Type of Energy Transfer Electromagnetic Spectrum Spectroscopic Techniquea

absorption γ-ray Mossbauer spectroscopy

X-ray X-ray absorption spectroscopy

UV/Vis spectroscopy atomic absorption


UV/Vis
spectroscopy

IR infrared spectroscopy raman spectroscopy

Microwave microwave spectroscopy

electron spin resonance spectroscopy nuclear


Radio wave
magnetic resonance spectroscopy

emission (thermal excitation) UV/Vis atomic emission spectroscopy

photoluminescence X-ray X-ray fluorescence

fluorescence spectroscopy phosphorescence


UV/Vis
spectroscopy atomic fluorescence spectroscopy

chemiluminescence UV/Vis chemiluminescence spectroscopy

Line Widths
A spectral line extends over a range of frequencies, not a single frequency (i.e., it has a nonzero linewidth). There are multiple
reasons for this broadening and shifts and only two are discussed below.
Lifetime Broadening. This mechanism ordinates directly from the Heisenberg Principle, which the lifetime of an excited state
(due to the spontaneous radiative decay) with the uncertainty of its energy Δt. Since the system is changing in time, it is
impossible to estimate the energies of wavefunctions exactly (ΔE.

ΔEΔt ≥ (14.1.7)
2

Hence, systems with excited-states that have short lifetimes will have a large energy uncertainty and a broad emission. This
broadening effect results in an broadened profile. As expected, the lifetime broadening can be experimentally if the decay rates
can be artificially suppressed or enhanced.
Doppler Broadening. This mechanism is intrinsic to atoms in a gas which are emitting radiation will have a distribution of
velocities. Each photon emitted will be "red"- or "blue"-shifted by the Doppler effect depending on the velocity of the atom
relative to the observer. The Doppler effect is the change in frequency of a wave (or other periodic event) for an observer
moving relative to its source. It is commonly heard when a vehicle sounding a siren or horn approaches, passes, and recedes
from an observer. Compared to the emitted frequency, the received frequency is higher during the approach, identical at the
instant of passing by, and lower during the recession.

Doppler Broadening example with sound, but applies to light too. (left): Stationary sound source produces sound waves at a
constant frequency u, and the wave-fronts propagate symmetrically away from the source at a constant speed c (assuming speed of
sound, c = 330 m/s), which is the speed of sound in the medium. The distance between wave-fronts is the wavelength. All observers
will hear the same frequency, which will be equal to the actual frequency of the source where u = u − o (right): The same sound

14.1.3 https://chem.libretexts.org/@go/page/41394
source is radiating sound waves at a constant frequency in the same medium. However, now the sound source is moving with a
speed so the center of each new wavefront is now slightly displaced to the right. As a result, the wave-fronts begin to bunch up on
the right side (in front of) and spread further apart on the left side (behind) of the source. An observer in front of the source will
hear a higher frequency. Images used with permission from Wikipedia (credit Lookang).
If the speeds are small compared to the speed of the wave, the relationship between observed frequency u and emitted frequency
o

u is approximately

Δv
u = (1 + ) v0 (14.1.8)
c

The Doppler effect also applies for spectroscopy. For example, the higher the temperature of the gas, the wider the distribution of
velocities in the gas (via the Maxwell-Bolztmann Distribution). Since the spectral line is a combination of all of the emitted
radiation, the higher the temperature of the gas, the broader the spectral line emitted from that gas. If the average velocity of a gas
is non-zero, then a Doppler shift will be observed that is correlated with the amplitude of this average velocity.

Redshift of spectral lines in the optical spectrum of a supercluster of distant galaxies (right), as compared to that of the Sun (left)
since the sun is not moving with respect to Earth (or weakly) and the galaxy is moving away. from Wikipedia (Credit Georg Wiora).
Doppler broadening is one of the explanations for the broadening of spectral lines, and as such gives an indication for the
temperature of observed material. Other causes of velocity distributions may exist, though, for example due to turbulent motion.
Doppler broadening can also be used to determine the velocity distribution of a gas given its absorption spectrum. In particular, this
has been used to determine the velocity distribution of interstellar gas clouds.

Example 14.1.2

Find the uncertainty of simultaneously measuring the frequency and wavelength of an emission from an excited molecule, if
the wavelength is 430 nm and the excited state lifetime is 0.50 nanoseconds.
Soluton
1. Use Heisenberg's uncertainty principle and the relationship between energy and frequency to find the uncertainty of
frequency.
h
ΔEΔt ≥

ΔE = hΔu

h
hΔuΔt ≥

1
Δu ≥
4πΔt

The maximum value for Δt is the lifetime of the excited state.

14.1.4 https://chem.libretexts.org/@go/page/41394
1
Δu ≥
s
4π0.50ns ×
9
10 ns

8 −1
Δu ≥ 1.6 × 10 s

2. Use the uncertainty of frequency and the relationship between frequency and wavelength to find the uncertainty of the
wavelength.
c
λ =
u

c
|Δλ| = |Δu|
2
u

2
λ |Δu|
|Δλ| =
c

2 8 −1
(430 nm ) × 1.6 × 10 s m
|Δλ| = ×
8 9
2.998 × 10 m/s 10 nm

−7
|Δλ| = 2.3 × 10 nm

Absorption and Emission


In absorption spectroscopy a photon is absorbed by an atom or molecule, which undergoes a transition from a lower-energy state to
a higher-energy, or excited state (Figure 14.1.3). The type of transition depends on the photon’s energy. The electromagnetic
spectrum in Figure 14.1.2, for example, shows that absorbing a photon of visible light promotes one of the atom’s or molecule’s
valence electrons to a higher-energy level. When an molecule absorbs infrared radiation, on the other hand, one of its chemical
bonds experiences a change in vibrational energy.

Figure 14.1.3 : Simplified energy diagram showing the absorption and emission of a photon by an atom or a molecule. When a
photon of energy hν strikes the atom or molecule, absorption may occur if the difference in energy, ∆E, between the ground state
and the excited state is equal to the photon’s energy. An atom or molecule in an excited state may emit a photon and return to the
ground state. The photon’s energy, hν, equals the difference in energy, ∆E, between the two states.
When it absorbs electromagnetic radiation the number of photons passing through a sample decreases. The measurement of this
decrease in photons, which we call absorbance, is a useful analytical signal. Note that the each of the energy levels in Figure
14.1.3 has a well-defined value because they are quantized. Absorption occurs only when the photon’s energy, hν, matches the

difference in energy, ∆E, between two energy levels. A plot of absorbance as a function of the photon’s energy is called an
absorbance spectrum. Figure 14.1.4, for example, shows the absorbance spectrum of cranberry juice.

14.1.5 https://chem.libretexts.org/@go/page/41394
Figure 14.1.3 ). As a result, the juice appears red.
When an atom or molecule in an excited state returns to a lower energy state, the excess energy often is released as a photon, a
process we call emission (Figure 14.1.4). There are several ways in which an atom or molecule may end up in an excited state,
including thermal energy, absorption of a photon, or by a chemical reaction. Emission following the absorption of a photon is also
called photoluminescence, and that following a chemical reaction is called chemiluminescence. A typical emission spectrum is
shown in Figure 14.1.6.

Figure 14.1.5 : Photoluminescence spectrum of the dye coumarin 343, which is incorporated in a reverse micelle suspended in
cyclohexanol. The dye’s absorbance spectrum (not shown) has a broad peak around 400 nm. The sharp peak at 409 nm is from the
laser source used to excite coumarin 343. The broad band centered at approximately 500 nm is the dye’s emission band. Because
the dye absorbs blue light, a solution of coumarin 343 appears yellow in the absence of photoluminescence. Its photoluminescent
emission is blue-green. Source: data from Bridget Gourley, Department of Chemistry & Biochemistry, DePauw University).

Selection Rules
The transition probability is defined as the probability of particular spectroscopic transition to take place. When an atom or
molecule absorbs a photon, the probability of an atom or molecule to transit from one energy level to another depends on two
things: the nature of initial and final state wavefunctions and how strongly photons interact with an wavefunction. Transition
strengths are used to describe transition probability. Selection rules are utilized to determine whether a transition is allowed or not.
Electronic dipole transitions are by far the most important for the topics covered in this module.
In an atom or molecule, an electromagnetic wave (for example, visible light) can induce an oscillating electric or magnetic
moment. If the frequency of the induced electric or magnetic moment is the same as the energy difference between one
wavefunction ψ and another wavefunction ψ , the interaction between an atom or molecule and the electromagnetic field is
1 2

resonant (which means these two have the same frequency). Typically, the amplitude of this (electric or magnetic) moment is called
the transition moment. In quantum mechanics, the transition probability of one molecule from one wavefunction ψ to another 1

wavefunction ψ is given by |M⃗ | , and M⃗ is called the transition dipole moment, or transition moment, from ψ to ψ . M⃗
2 21
2
21 1 2 21

can be written as


M 21 = ∫ ψ2 μ⃗ ψ1 dτ (14.1.9)

where ψ and ψ are two different wavefunctions in one molecule, and M⃗ is the electric dipole moment operator. If we have a
1 2 21

system with n atoms and each has charge q , and the dipole moment operator is can be written as
n

14.1.6 https://chem.libretexts.org/@go/page/41394
μ⃗ = ∑ qn r n
⃗ (14.1.10)

the r ⃗ is the position vector operator for the ith charge. The nature of ψ and ψ change (e..g, the quantum numbers associated
n 1 1

with each wavefunction) M⃗ . Large values of M⃗ signify transitions with strong probabilities and small M⃗ values represent
21 21 21

weak probabilities. A zero probability for a transition is a forbidden transition.


For electronic wavefucntion (either atoms or molecules), the two primary selection rules governing transitions between electronic
energy wavefunctions are:
1. ΔS = 0 (The Spin Rule)
2. Δl = ±1 (The Orbital Rule (or Laporte rule))
The spin multiplicity can be calculated from the quantum number S of the total spin or from the number of unpaired electrons (like
when determining paramagnetic properties of molecules). The spin-multiplicity is (2S + 1) , where

S = ∑s (14.1.11)

spin quantum #

The Spin Rule says that allowed transitions must involve the promotion of electrons without a change in their spin. The Orbital
Rule says that transitions within a given set of p or d orbitals (i.e. those which only involve a redistribution of electrons within a
given subshell) are forbidden. The orbital rule can be used to construct Grotrian diagrams, which show the allowed electronic
transitions between the energy levels of atoms (Figure 14.1.6) by taking into account the specific selection rules related to the
system (e..g, the Orbital or Spin Rules).

Figure 14.1.6 : A Grotrian diagram of the hydrogen atom. Only transitions between adjacent columns are allowed, as per the
selection rule Δl = ±1 . from Tufts OCW (CC SA-BY; Gary Goldstein).

Transitions not permitted by selection rules are said forbidden, which means that theoretically they must not occur. However,
in practice they may occur, but very low probabilities (see 14.1.3 below).

The Beers-Lambert Law


The Beer-Lambert law relates the attenuation of light to the properties of the material through which the light is traveling. This
page takes a brief look at the Beer-Lambert Law and explains the use of the terms absorbance and molar absorptivity relating to
UV-visible absorption spectrometry. For each wavelength of light passing through the spectrometer, the intensity of the light
passing through the reference cell is measured. This is usually referred to as I - that's I for Intensity.
o

Figure 14.1.6 : Light absorbed by sample in a cuvetter


The intensity of the light passing through the sample cell is also measured for that wavelength - given the symbol, I . If I is less
than I , then the sample has absorbed some of the light (neglecting reflection of light off the cuvetter surface). A simple bit of math
o

14.1.7 https://chem.libretexts.org/@go/page/41394
is then done in the computer to convert this into something called the absorbance of the sample - given the symbol, A .

Note

The absorbance of a transition depends on two external assumptions.


1. The absorbance is directly proportional to the concentration (c ) of the solution of the the sample used in the experiment.
2. The absorbance is directly proportional to the length of the light path (l), which is equal to the width of the cuvette.

Assumption one relates the absorbance to concentration and can be expressed as


A ∝c (14.1.12)

The absorbance (A ) is defined via the incident intensity I and transmitted intensity I by
o

Io
A = log10 ( ) (14.1.13)
I

Assumption two can be expressed as

A ∝l (14.1.14)

Combining Equations 14.1.12 and 14.1.14:

A ∝ cl (14.1.15)

This proportionality can be converted into an equality by including a proportionality constant.

A = ϵcl (14.1.16)

This formula is the common form of the Beer-Lambert Law, although it can be also written in terms of intensities:
Io
A = log10 ( ) = ϵlc (14.1.17)
I

The constant ϵ is called molar absorptivity or molar extinction coefficient and is a measure of the probability of the electronic
transition. On most of the diagrams you will come across, the absorbance ranges from 0 to 1, but it can go higher than that. An
absorbance of 0 at some wavelength means that no light of that particular wavelength has been absorbed. The intensities of the
sample and reference beam are both the same, so the ratio Io/I is 1. Log10 of 1 is zero.

Example 14.1.3

In a sample with an absorbance of 1 at a specific wavelength, what is the relative amount of light that was absorbed by the
sample?
Solution
This question does not need Beer-Lambert Law (Equation 14.1.16 ) to solve, but only the definition of absorbance (Equation
14.1.13)

Io
A = log ( )
10
I

The relative loss of intensity is


I − Io I
=1−
Io Io

Equation 14.1.13 can be rearranged using the properties of logarithms to solved for the relative loss of intensity:
Io
A
10 =
I

−A
I
10 =
Io

14.1.8 https://chem.libretexts.org/@go/page/41394
−A
I
1 − 10 =1−
Io

Substituting in A = 1
I −1
1
1− = 1 − 10 =1− = 0.9
Io 10

Hence 90% of the light at that wavelength has been absorbed and that the transmitted intensity is 10% of the incident intensity
To confirm, substituting these values into Equation 14.1.13 to get the absorbance back:
Io 100
= = 10
I 10

and

log10 10 = 1

You will find that various different symbols are given for some of the terms in the equation - particularly for the concentration and
the solution length.

The Greek letter epsilon in these equations is called the molar absorptivity - or sometimes the molar absorption coefficient. The
larger the molar absorptivity, the more probable the electronic transition. In uv spectroscopy, the concentration of the sample
solution is measured in molL-1 and the length of the light path in cm. Thus, given that absorbance is unitless, the units of molar
absorptivity are L mol-1 cm-1. However, since the units of molar absorptivity is always the above, it is customarily reported without
units.

Example 14.1.4

Guanosine has a maximum absorbance of 275 nm. ϵ = 8400M cm 275 and the path length is 1 cm. Using a
−1 −1

spectrophotometer, you find the that A = 0.70 . What is the concentration of guanosine?
275

Solution
To solve this problem, you must use Beer's Law.

A = ϵlc

0.70 = (8400 M-1 cm-1)(1 cm)(c )


Next, divide both side by [(8400 M-1 cm-1)(1 cm)]
c = 8.33x10-5 mol/L

Example 14.1.5

There is a substance in a solution (4 g/liter). The length of cuvette is 2 cm and only 50% of the certain light beam is
transmitted. What is the extinction coefficient?
Solution
Using Beer-Lambert Law, we can compute the absorption coefficient. Thus,
It 0.5
− log( ) = − log( ) = A = 8ϵ
Io 1.0

Then we obtain that

14.1.9 https://chem.libretexts.org/@go/page/41394
ϵ = 0.0376

Example 14.1.6

In Example 3 above, what is the molar absorption coefficient if the molecular weight is 100?
Solution
It can simply obtained by multiplying the absorption coefficient by the molecular weight. Thus,
ϵ = 0.0376 x 100 = 3.76 L·mol-1·cm-1

The proportion of the light absorbed will depend on how many molecules it interacts with. Suppose you have got a strongly colored
organic dye. If it is in a reasonably concentrated solution, it will have a very high absorbance because there are lots of molecules to
interact with the light. However, in an incredibly dilute solution, it may be very difficult to see that it is colored at all. The
absorbance is going to be very low. Suppose then that you wanted to compare this dye with a different compound. Unless you took
care to make allowance for the concentration, you couldn't make any sensible comparisons about which one absorbed the most
light.

Example 14.1.7

In Example 14.1.3 above, how much is the beam of light is transmitted when 8 g/liter ?
Solution
Since we know ϵ, we can calculate the transmission using Beer-Lambert Law. Thus,
log(1) − log(It ) = 0 − log(It ) = 0.0376 x 8 x 2 = 0.6016
log(It ) = -0.6016
Therefore, I = 0.2503 = 25%
t

Example 14.1.7

The absorption coefficient of a glycogen-iodine complex is 0.20 at light of 450 nm. What is the concentration when the
transmission is 40 % in a cuvette of 2 cm?
Solution
It can also be solved using Beer-Lambert Law. Therefore,

− log(It ) = − log10 (0.4) = 0.20 × c × 2

Then c = 0.9948

The Beer-Lambert law Equation 14.1.16 can be rearranged to obtain an expression for ϵ (the molar absorptivity):
A
ϵ= (14.1.18)
lc

Remember that the absorbance of a solution will vary as the concentration or the size of the container varies. Molar absorptivity
compensates for this by dividing by both the concentration and the length of the solution that the light passes through. Essentially,
it works out a value for what the absorbance would be under a standard set of conditions - the light traveling 1 cm through a
solution of 1 mol dm-3. That means that you can then make comparisons between one compound and another without having to
worry about the concentration or solution length.
Values for molar absorptivity can vary hugely. For example, ethanal has two absorption peaks in its UV-visible spectrum - both in
the ultra-violet. One of these corresponds to an electron being promoted from a lone pair on the oxygen into a pi anti-bonding
orbital; the other from a π bonding orbital into a π anti-bonding orbital. Table 14.1.2 gives values for the molar absorptivity of a
solution of ethanal in hexane. Notice that there are no units given for absorptivity. That's quite common since it assumes the length
is in cm and the concentration is mol dm-3, the units are mol-1 dm3 cm-1.

14.1.10 https://chem.libretexts.org/@go/page/41394
Table 14.1.2 :
electron jump wavelength of maximum absorption (nm) molar absorptivity

lone pair to π anti-bonding orbital 290 15

π bonding to π anti-bonding orbital 180 10,000

The ethanal obviously absorbs much more strongly at 180 nm than it does at 290 nm. (Although, in fact, the 180 nm absorption
peak is outside the range of most spectrometers.) You may come across diagrams of absorption spectra plotting absorptivity on the
vertical axis rather than absorbance. However, if you look at the figures above and the scales that are going to be involved, you are
not really going to be able to spot the absorption at 290 nm. It will be a tiny little peak compared to the one at 180 nm. To get
around this, you may also come across diagrams in which the vertical axis is plotted as log10(molar absorptivity). If you take the
logs of the two numbers in the table, 15 becomes 1.18, while 10,000 becomes 4. That makes it possible to plot both values easily,
but produces strangely squashed-looking spectra!
Table 14.1.3 : Expected intensities of electronic transitions
Transition type Typical values of ε /m2mol-1

Spin forbidden and Laporte forbidden 0.1

Spin allowed and Laporte forbidden 1 - 10

Spin allowed and Laporte allowed


1,000 - 106
e.g. charge transfer bands

Contributors and Attributions


Jim Clark (Chemguide.co.uk)
Gamini Gunawardena from the OChemPal site (Utah Valley University)
David Harvey (DePauw University)

14.1: Vocabulary is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

14.1.11 https://chem.libretexts.org/@go/page/41394
14.2: Microwave Spectroscopy
Microwave rotational spectroscopy uses microwave radiation to measure the energies of rotational transitions for molecules in the
gas phase. It accomplishes this through the interaction of the electric dipole moment of the molecules with the electromagnetic field
of the exciting microwave photon.

Introduction
To probe the pure rotational transitions for molecules, scientists use microwave rotational spectroscopy. This spectroscopy utilizes
photons in the microwave range to cause transitions between the quantum rotational energy levels of a gas molecule. The reason
why the sample must be in the gas phase is due to intermolecular interactions hindering rotations in the liquid and solid phases of
the molecule. For microwave spectroscopy, molecules can be broken down into 5 categories based on their shape and the inertia
around their 3 orthogonal rotational axes. These 5 categories include diatomic molecules, linear molecules, spherical tops,
symmetric tops and asymmetric tops.

Classical Mechanics
The Hamiltonian solution to the rigid rotor is

H =T (14.2.1)

since,

H = T +V (14.2.2)

Where T is kinetic energy and V is potential energy. Potential energy, V , is 0 because there is no resistance to the rotation (similar
to a particle in a box model).
Since H = T , we can also say that:
1
2
T = ∑ mi v (14.2.3)
i
2

However, we have to determine v in terms of rotation since we are dealing with rotation. Since,
i

v
ω = (14.2.4)
r

where ω = angular velocity, we can say that:

vi = ωX ri (14.2.5)

Thus we can rewrite the T equation as:


1
T = ∑ mi vi (ωX ri ) (14.2.6)
2

Since ω is a scalar constant, we can rewrite the T equation as:


ω ω L
T = ∑ mi (vi X ri ) = ∑ li = ω (14.2.7)
2 2 2

where l is the angular momentum of the ith particle, and L is the angular momentum of the entire system. Also, we know from
i

physics that,
L = Iω (14.2.8)

where I is the moment of inertia of the rigid body relative to the axis of rotation. We can rewrite the T equation as,
Iω 1 2
T =ω = Iω (14.2.9)
2 2

Quantum Mechanics
The internal Hamiltonian, H, is:

14.2.1 https://chem.libretexts.org/@go/page/41395
2 2
i ℏ
H = (14.2.10)
2I

and the Schrödinger Equation for rigid rotor is:


2 2
i ℏ
ψ = Eψ (14.2.11)
2I

Thus, we get:
2
J(J + 1)h
En = (14.2.12)
8π 2 I

where J is a rotational quantum number and ℏ is the reduced Planck's constant. However, if we let:
h
B = (14.2.13)
2
8π I

where B is a rotational constant, then we can substitute it into the E equation and get:
n

En = J(J + 1)Bh (14.2.14)

Considering the transition energy between two energy levels, the difference is a multiple of 2. That is, from J =0 to J =1 , the
ΔE 0→1 is 2Bh and from J = 1 to J = 2, the ΔE is 4Bh.
1→2

Figure 14.2.1 : Energy levels and line positions calculated in the rigid rotor approximation. This diagram illustrates how transitions
between the rotational energy levels of molecules map onto the energies at which these transitions are observed during laboratory
experiments. (CC CS-BY 3.0; Nnrw).

Theory
When a gas molecule is irradiated with microwave radiation, a photon can be absorbed through the interaction of the photon’s
electronic field with the electrons in the molecules. For the microwave region this energy absorption is in the range needed to cause
transitions between rotational states of the molecule. However, only molecules with a permanent dipole that changes upon rotation
can be investigated using microwave spectroscopy. This is due to the fact that their must be a charge difference across the molecule
for the oscillating electric field of the photon to impart a torque upon the molecule around an axis that is perpendicular to this
dipole and that passes through the molecules center of mass.
This interaction can be expressed by the transition dipole moment for the transition between two rotational states

Probability of Transition = ∫ ^ ψrot (I )dτ


ψrot (F )μ (14.2.15)

14.2.2 https://chem.libretexts.org/@go/page/41395
Where Ψrot(F) is the complex conjugate of the wave function for the final rotational state, Ψrot(I) is the wave function of the initial
rotational state , and μ is the dipole moment operator with Cartesian coordinates of μx, μy, μz. For this integral to be nonzero the
integrand must be an even function. This is due to the fact that any odd function integrated from negative infinity to positive
infinity, or any other symmetric limits, is always zero.
In addition to the constraints imposed by the transition moment integral, transitions between rotational states are also limited by the
nature of the photon itself. A photon contains one unit of angular momentum, so when it interacts with a molecule it can only
impart one unit of angular momentum to the molecule. This leads to the selection rule that a transition can only occur between
rotational energy levels that are only one quantum rotation level (J) away from another1.

ΔJ = ±1 (14.2.16)

The transition moment integral and the selection rule for rotational transitions tell if a transition from one rotational state to another
is allowed. However, what these do not take into account is whether or not the state being transitioned from is actually populated,
meaning that the molecule is in that energy state. This leads to the concept of the Boltzmann distribution of states. The Boltzmann
distribution is a statistical distribution of energy states for an ensemble of molecules based on the temperature of the sample2.
(−Erot (J)/RT )
nJ e
= (14.2.17)
J=n
n0 ∑ e
(−Erot (J)/RT )
J=1

where Erot(J) is the molar energy of the J rotational energy state of the molecule,
R is the gas constant,
T is the temperature of the sample.
n(J) is the number of molecules in the J rotational level, and
n0 is the total number of molecules in the sample.
This distribution of energy states is the main contributing factor for the observed absorption intensity distributions seen in the
microwave spectrum. This distribution makes it so that the absorption peaks that correspond to the transition from the energy state
with the largest population based on the Boltzmann equation will have the largest absorption peak, with the peaks on either side
steadily decreasing.

Degrees of Freedom
A molecule can have three types of degrees of freedom and a total of 3N degrees of freedom, where N equals the number of atoms
in the molecule. These degrees of freedom can be broken down into 3 categories3.
Translational: These are the simplest of the degrees of freedom. These entail the movement of the entire molecule’s center of
mass. This movement can be completely described by three orthogonal vectors and thus contains 3 degrees of freedom.
Rotational: These are rotations around the center of mass of the molecule and like the translational movement they can be
completely described by three orthogonal vectors. This again means that this category contains only 3 degrees of freedom.
However, in the case of a linear molecule only two degrees of freedom are present due to the rotation along the bonds in the
molecule having a negligible inertia.
Vibrational: These are any other types of movement not assigned to rotational or translational movement and thus there are 3N
– 6 degrees of vibrational freedom for a nonlinear molecule and 3N – 5 for a linear molecule. These vibrations include bending,
stretching, wagging and many other aptly named internal movements of a molecule. These various vibrations arise due to the
numerous combinations of different stretches, contractions, and bends that can occur between the bonds of atoms in the
molecule.
Each of these degrees of freedom is able to store energy. However, In the case of rotational and vibrational degrees of freedom,
energy can only be stored in discrete amounts. This is due to the quantized break down of energy levels in a molecule described by
quantum mechanics. In the case of rotations the energy stored is dependent on the rotational inertia of the gas along with the
corresponding quantum number describing the energy level.

Rotational Symmetries
To analyze molecules for rotational spectroscopy, we can break molecules down into 5 categories based on their shapes and their
moments of inertia around their 3 orthogonal rotational axes:4
1. Diatomic Molecules

14.2.3 https://chem.libretexts.org/@go/page/41395
2. Linear Molecules
3. Spherical Tops
4. Symmetrical Tops
5. Asymmetrical Tops

Diatomic Molecules
The rotations of a diatomic molecule can be modeled as a rigid rotor. This rigid rotor model has two masses attached to each other
with a fixed distance between the two masses.

It has an inertia (I) that is equal to the square of the fixed distance between the two masses multiplied by the reduced mass of the
rigid rotor.
2
Ie = μ re (14.2.18)

m1 m2
μ = (14.2.19)
m1 + m2

Using quantum mechanical calculations it can be shown that the energy levels of the rigid rotator depend on the inertia of the rigid
rotator and the quantum rotational number J2.
E(J) = Be J(J + 1) (14.2.20)

h
Be = (14.2.21)
2
8 π c Ie

However, this rigid rotor model fails to take into account that bonds do not act like a rod with a fixed distance, but like a spring.
This means that as the angular velocity of the molecule increases so does the distance between the atoms. This leads us to the
nonrigid rotor model in which a centrifugal distortion term (D ) is added to the energy equation to account for this stretching
e

during rotation.
−1 2 2
E(J)(c m ) = Be J(J + 1)– De J (J + 1 ) (14.2.22)

This means that for a diatomic molecule the transitional energy between two rotational states equals
′ ′ ′′ ′′ ′2 ′ 2 ′′2 ′ 2
E = Be [ J (J + 1) − J (J + 1)] − De [ J (J + 1) −J (J + 1) ] (14.2.23)

Where J’ is the quantum number of the final rotational energy state and J’’ is the quantum number of the initial rotational energy
state. Using the selection rule of ΔJ = ±1 the spacing between peaks in the microwave absorption spectrum of a diatomic
molecule will equal
′′ ′′3
ER = (2 Be − 4 De ) + (2 Be − 12 De )J − 4 De J (14.2.24)

Linear Molecules
Linear molecules behave in the same way as diatomic molecules when it comes to rotations. For this reason they can be modeled as
a non-rigid rotor just like diatomic molecules. This means that linear molecule have the same equation for their rotational energy
levels. The only difference is there are now more masses along the rotor. This means that the inertia is now the sum of the distance
between each mass and the center of mass of the rotor multiplied by the square of the distance between them2.

14.2.4 https://chem.libretexts.org/@go/page/41395
n

2
Ie = ∑ mj r (14.2.25)
ej

j=1

Where mj is the mass of the jth mass on the rotor and rej is the equilibrium distance between the jth mass and the center of mass of
the rotor.

Spherical Tops
Spherical tops are molecules in which all three orthogonal rotations have equal inertia and they are highly symmetrical. This means
that the molecule has no dipole and for this reason spherical tops do not give a microwave rotational spectrum.

Figure 14.2.2 : Geometrical example of a spherical top


Examples:

Symmetrical Tops
Symmetrical tops are molecules with two rotational axes that have the same inertia and one unique rotational axis with a different
inertia. Symmetrical tops can be divided into two categories based on the relationship between the inertia of the unique axis and the
inertia of the two axes with equivalent inertia. If the unique rotational axis has a greater inertia than the degenerate axes the
molecule is called an oblate symmetrical top. If the unique rotational axis has a lower inertia than the degenerate axes the molecule
is called a prolate symmetrical top. For simplification think of these two categories as either frisbees for oblate tops or footballs for
prolate tops.

Figure 14.2.3 : Symmetric Tops: (Left) Geometrical example of an oblate top and (right) a prolate top. Images used with
permission from Wikipedia.com.

14.2.5 https://chem.libretexts.org/@go/page/41395
Figure 14.2.4 : Examples of symmetric tops. Benzene (oblate) XeF4 (oblate) ClCH3 (prolate) NH3 (prolate)
In the case of linear molecules there is one degenerate rotational axis which in turn has a single rotational constant. With
symmetrical tops now there is one unique axis and two degenerate axes. This means an additional rotational constant is needed to
describe the energy levels of a symmetrical top. In addition to the rotational constant an additional quantum number must be
introduced to describe the rotational energy levels of the symmetric top. These two additions give us the following rotational
energy levels of a prolate and oblate symmetric top
−1 2
E(J,K) (c m ) = Be ∗ J(J + 1) + (Ae − Be )K (14.2.26)

Where Be is the rotational constant of the unique axis, Ae is the rotational constant of the degenerate axes, J is the total rotational
angular momentum quantum number and K is the quantum number that represents the portion of the total angular momentum that
lies along the unique rotational axis. This leads to the property that K is always equal to or less than J . Thus we get the two
selection rules for symmetric tops
ΔJ = 0, ±1 (14.2.27)

ΔK = 0 (14.2.28)

when K ≠ 0
ΔJ = ±1 (14.2.29)

ΔK = 0 (14.2.30)

when K = 0
However, like the rigid rotor approximation for linear molecules, we must also take into account the elasticity of the bonds in
symmetric tops. Therefore, in a similar manner to the rigid rotor we add a centrifugal coupling term, but this time we have one for
each quantum number and one for the coupling between the two.
−1 2 2 2
E(J,K) (c m ) = Be J(J + 1) − DeJ J (J + 1 ) + (Ae − Be ) ∗ K (14.2.31)

4 2
−Dek K − Dejk J(J + 1)K (14.2.32)

Asymmetrical Tops
Asymmetrical tops have three orthogonal rotational axes that all have different moments of inertia and most molecules fall into this
category. Unlike linear molecules and symmetric tops these types of molecules do not have a simplified energy equation to
determine the energy levels of the rotations. These types of molecules do not follow a specific pattern and usually have very
complex microwave spectra.

Figure 14.2.5 : Molecular examples of asymmetrical tops. (left) water and (right) acetone

14.2.6 https://chem.libretexts.org/@go/page/41395
Additional Rotationally Sensitive Spectroscopies
In addition to microwave spectroscopy, IR spectroscopy can also be used to probe rotational transitions in a molecule. However, in
the case of IR spectroscopy the rotational transitions are coupled to the vibrational transitions of the molecule. One other
spectroscopy that can probe the rotational transitions in a molecule is Raman spectroscopy, which uses UV-visible light scattering
to determine energy levels in a molecule. However, a very high sensitivity detector must be used to analyze rotational energy levels
of a molecule.

References
1. Harris, D. C.; Bertolucci, M. D., Symmetry and Spectroscopy. University Press: Oxford, 1978.
2. McQuarrie, D. A.; Simon, J. D., Physical Chemistry: A Molecular Approach. University Science Books: 1997.
3. Shoemaker, D. P.; W., G. C.; W., N. J., Experiments in Physical Chemistry. 8th ed.; McGraw Hill: New York, 2009.
4. Hollas, M. J., Basic Atomic and Molecular Spectroscopy. Royal Society of Chemistry: Cambridge, 2002.

14.2: Microwave Spectroscopy is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
Microwave Rotational Spectroscopy by Nicholas Houghton is licensed CC BY 4.0.

14.2.7 https://chem.libretexts.org/@go/page/41395
SECTION OVERVIEW

14.3: Infrared Spectroscopy


Infrared Spectroscopy is the analysis of infrared light interacting with a molecule. This can be analyzed in three ways by measuring
absorption, emission and reflection. The main use of this technique is in organic and inorganic chemistry. It is used by chemists to
determine functional groups in molecules. IR Spectroscopy measures the vibrations of atoms, and based on this it is possible to
determine the functional groups.5 Generally, stronger bonds and light atoms will vibrate at a high stretching frequency
(wavenumber).

14.3: Infrared Spectroscopy is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

14.3.1 https://chem.libretexts.org/@go/page/41396
14.4: Electronic Spectroscopy
This page explains what happens when organic compounds absorb UV or visible light, and why the wavelength of light absorbed
varies from compound to compound.

Molecular Absorption of Light


When we were talking about the various sorts of molecular orbitals present in organic compounds earlier, you will have come
across this diagram showing their relative energies:

Figure 14.4.1 : The diagram is not intended to be to scale - it just shows the relative placing of the different orbitals.
When light passes through the compound, energy from the light is used to promote an electron from a bonding or non-bonding
orbital into one of the empty anti-bonding orbitals. The possible electron jumps that light might cause are:

Figure 14.4.2 : Promotion of an electron via different electron transitions


In each possible case, an electron is excited from a full orbital into an empty anti-bonding orbital. Each jump takes energy from the
light, and a big jump obviously needs more energy than a small one. Each wavelength of light has a particular energy associated
with it. If that particular amount of energy is just right for making one of these energy jumps, then that wavelength will be absorbed
- its energy will have been used in promoting an electron.
An absorption spectrometer works in a range from about 200 nm (in the near ultra-violet) to about 800 nm (in the very near infra-
red). Only a limited number of the possible electron jumps absorb light in that region. Look again at the possible jumps. This time,
the important jumps are shown in black, and a less important one in grey. The grey dotted arrows show jumps which absorb light
outside the region of the spectrum we are working in.

Figure 14.4.3 : Promotion of an electron via different electron transitions scaled by energies.
Remember that bigger jumps need more energy and so absorb light with a shorter wavelength. The jumps shown with grey dotted
arrows absorb UV light of wavelength less that 200 nm. The important jumps are:
from π bonding orbitals to π anti-bonding orbitals;
from non-bonding orbitals to π anti-bonding orbitals;
from non-bonding orbitals to sigma anti-bonding orbitals.

14.4.1 https://chem.libretexts.org/@go/page/41397
That means that in order to absorb light in the region from 200 - 800 nm (which is where the spectra are measured), the molecule
must contain either π bonds or atoms with non-bonding orbitals. Remember that a non-bonding orbital is a lone pair on, say,
oxygen, nitrogen or a halogen. Groups in a molecule which absorb light are known as chromophores.

What does an absorption spectrum look like


The diagram below shows a simple UV-visible absorption spectrum for buta-1,3-diene - a molecule we will talk more about later.
Absorbance (on the vertical axis) is just a measure of the amount of light absorbed. The higher the value, the more of a particular
wavelength is being absorbed.

Figure 14.4.4 : UV-visible absorption spectrum for buta-1,3-diene


You will see that absorption peaks at a value of 217 nm. This is in the ultra-violet and so there would be no visible sign of any light
being absorbed - buta-1,3-diene is colorless. In buta-1,3-diene, CH2=CH-CH=CH2, there are no non-bonding electrons. That means
that the only electron jumps taking place (within the range that the spectrometer can measure) are from π bonding to π anti-
bonding orbitals.
A chromophore such as the carbon-oxygen double bond in ethanal, for example, obviously has π electrons as a part of the double
bond, but also has lone pairs on the oxygen atom. That means that both of the important absorptions from the last energy diagram
are possible. You can get an electron excited from a π bonding to a π anti-bonding orbital, or you can get one excited from an
oxygen lone pair (a non-bonding orbital) into a π anti-bonding orbital.

Figure 14.4.4 : A simplified diagram showing n → π and π → π transitions


∗ ∗

The non-bonding orbital has a higher energy than a π bonding orbital. That means that the jump from an oxygen lone pair into a π
anti-bonding orbital needs less energy. That means it absorbs light of a lower frequency and therefore a higher wavelength. For
example, ethanal can therefore absorb light of two different wavelengths:
the π bonding to π anti-bonding absorption peaks at 180 nm. These n → π transitions involve moving an electron from a

nonbonding electron pair to a antibonding \*pi^*\) orbital. They tend to have molar absorbtivities less than 2000
the non-bonding to π anti-bonding absorption peaks at 290 nm. These π → π transitions involve moving an electron from a

bonding π∗ orbital to an antibonding π orbital. They tend to have molar absorptivities on the order of 10,000.

Both of these absorptions are in the ultra-violet, but most spectrometers will not pick up the one at 180 nm because they work in
the range from 200 - 800 nm.

Conjugation and Delocalization


Consider these three molecules:

Ethene contains a simple isolated carbon-carbon double bond, but the other two have conjugated double bonds. In these cases, there
is delocalization of the π bonding orbitals over the whole molecule. Now look at the wavelengths of the light which each of these
molecules absorbs.

14.4.2 https://chem.libretexts.org/@go/page/41397
molecule wavelength of maximum absorption (nm)

ethene 171

buta-1,3-diene 217

hexa-1,3,5-triene 258

All of the molecules give similar UV-visible absorption spectra - the only difference being that the absorptions move to longer and
longer wavelengths as the amount of delocalization in the molecule increases. Why is this? You can actually work out what must be
happening.
The maximum absorption is moving to longer wavelengths as the amount of delocalization increases.
Therefore maximum absorption is moving to shorter frequencies as the amount of delocalization increases.
Therefore absorption needs less energy as the amount of delocalization increases.
Therefore there must be less energy gap between the bonding and anti-bonding orbitals as the amount of delocalization
increases.
. . . and that's what is happening.
Compare ethene with buta-1,3-diene. In ethene, there is one π bonding orbital and one π anti-bonding orbital. In buta-1,3-diene,
there are two π bonding orbitals and two π anti-bonding orbitals. This is all discussed in detail on the introductory page that you
should have read.

The highest occupied molecular orbital is often referred to as the HOMO - in these cases, it is a π bonding orbital. The lowest
unoccupied molecular orbital (the LUMO) is a π anti-bonding orbital. Notice that the gap between these has fallen. It takes less
energy to excite an electron in the buta-1,3-diene case than with ethene. In the hexa-1,3,5-triene case, it is less still.

If you extend this to compounds with really massive delocalization, the wavelength absorbed will eventually be high enough to be
in the visible region of the spectrum, and the compound will then be seen as colored. A good example of this is the orange plant
pigment, beta-carotene - present in carrots, for example.

Why is beta-carotene orange?


Beta-carotene has the sort of delocalization that we've just been looking at, but on a much greater scale with 11 carbon-carbon
double bonds conjugated together. The diagram shows the structure of beta-carotene with the alternating double and single bonds
shown in red.

The more delocalization there is, the smaller the gap between the highest energy π bonding orbital and the lowest energy π anti-
bonding orbital. To promote an electron therefore takes less energy in beta-carotene than in the cases we've looked at so far -
because the gap between the levels is less. Remember that less energy means a lower frequency of light gets absorbed - and that's
equivalent to a longer wavelength. Beta-carotene absorbs throughout the ultra-violet region into the violet - but particularly

14.4.3 https://chem.libretexts.org/@go/page/41397
strongly in the visible region between about 400 and 500 nm with a peak about 470 nm. The wavelengths associated with the
various colors are approximately:

color region wavelength (nm)

violet 380 - 435

blue 435 - 500

cyan 500 - 520

green 520 - 565

yellow 565 - 590

orange 590 - 625

red 625 - 740

So if the absorption is strongest in the violet to cyan region, what color will you actually see? It is tempting to think that you can
work it out from the colors that are left - and in this particular case, you wouldn't be far wrong. Unfortunately, it is not as simple as
that! Sometimes what you actually see is quite unexpected. Mixing different wavelengths of light does not give you the same result
as mixing paints or other pigments. You can, however, sometimes get some estimate of the color you would see using the idea of
complementary colors. If you arrange some colors in a circle, you get a "color wheel". The diagram shows one possible version of
this. An internet search will throw up many different versions!

colors directly opposite each other on the color wheel are said to be complementary colors. Blue and yellow are complementary
colors; red and cyan are complementary; and so are green and magenta. Mixing together two complementary colors of light will
give you white light. What this all means is that if a particular color is absorbed from white light, what your eye detects by mixing
up all the other wavelengths of light is its complementary color. In the beta-carotene case, the situation is more confused because
you are absorbing such a range of wavelengths. However, if you think of the peak absorption running from the blue into the cyan, it
would be reasonable to think of the color you would see as being opposite that where yellow runs into red - in other words, orange.

Franck-Condon: Electronic and Vibrational Coupling


So far, we have come across one big rule of photon absorbance. To be absorbed, a photon's energy has to match an energy
difference within the compound that is absorbing it.

14.4.4 https://chem.libretexts.org/@go/page/41397
In the case of visible or ultraviolet light, the energy of a photon is roughly in the region that would be appropriate to promote an
electron to a higher energy level. Different wavelengths would be able to promote different electrons, depending on the energy
difference between an occupied electronic energy level and an unoccupied one. Other types of electromagnetic radiation would not
be able to promote an electron, but they would be coupled to other events. For example, absorption of infrared light is tied to
vibrational energy levels. Microwave radiation is tied to rotational energy levels in molecules. Thus, one reason a photon may or
may not be absorbed has to do with whether its energy corresponds to the available energy differences within the molecule or ion
that it encounters.
Photons face other limitations. One of these is a moderate variation on our main rule. It is called the Frank Condon Principle.
According to this idea, when an electron is excited from its normal position, the ground state, to a higher energy level, the optimal
positions of atoms in the molecule may need to shift. Because electronic motion is much faster than nuclear motion, however, any
shifting of atoms needed to optimize positions as they should be in the excited state will have to wait until after the electron gets
excited. In that case, when the electron lands and the atoms aren't yet in their lowest energy positions for the excited state, the
molecule will find itself in an excited vibrational state as well as an excited electronic state.

That means the required energy for excitation does not just correspond to the difference in electronic energy levels; it is fine-tuned
to reach a vibrational energy level, which is quantized as well.
The Franck Condon Principle states that electronic transitions are vertical.
A vertical transition is one in which non of the nuclei move while the electron journeys from one state to another.

14.4.5 https://chem.libretexts.org/@go/page/41397
A vertical transition may begin in a vibrational ground state of an electronic ground state and end in a vibrational excited state
of an electronic excited state.

Contributors and Attributions


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)
Jim Clark (Chemguide.co.uk)

14.4: Electronic Spectroscopy is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

14.4.6 https://chem.libretexts.org/@go/page/41397
14.5: Nuclear Magnetic Resonance
Nuclear Magnetic Resonance (NMR) is a nuceli (Nuclear) specific spectroscopy that has far reaching applications throughout the
physical sciences and industry. NMR uses a large magnet (Magnetic) to probe the intrinsic spin properties of atomic nuclei. Like all
spectroscopies, NMR uses a component of electromagnetic radiation (radio frequency waves) to promote transitions between
nuclear energy levels (Resonance). Most chemists use NMR for structure determination of small molecules.

Introduction
In 1946, NMR was co-discovered by Purcell, Pound and Torrey of Harvard University and Bloch, Hansen and Packard of Stanford
University. The discovery first came about when it was noticed that magnetic nuclei, such as 1H and 31P (read: proton and
Phosphorus 31) were able to absorb radio frequency energy when placed in a magnetic field of a strength that was specific to the
nucleus. Upon absorption, the nuclei begin to resonate and different atoms within a molecule resonated at different frequencies.
This observation allowed a detailed analysis of the structure of a molecule. Since then, NMR has been applied to solids, liquids and
gasses, kinetic and structural studies, resulting in 6 Nobel prizes being awarded in the field of NMR. More information about the
history of NMR can be found in the NMR History page. Here, the fundamental concepts of NMR are presented.

Spin and Magnetic Properties


The nucleus consists of elementary particles called neutrons and protons, which contain an intrinsic property called spin. Like
electrons, the spin of a nucleus can be described using quantum numbers of I for the spin and m for the spin in a magnetic field.
Atomic nuclei with even numbers of protons and neutrons have zero spin and all the other atoms with odd numbers have a non-zero
spin. Furthermore, all molecules with a non-zero spin have a magnetic moment, μ , given by

μ = γI (14.5.1)

where γ is the gyromagnetic ratio, a proportionality constant between the magnetic dipole moment and the angular momentum,
specific to each nucleus (Table 14.5.1).
Table 14.5.1 : The gyromagnetic ratios for several common nuclei
Nuclei Spin Gyromagetic Ratio (MHz/T) Natural Abundance (%)
1H 1/2 42.576 99.9985
13C 1/2 10.705 1.07
31P 1/2 17.235 100
27Al 5/2 11.103 100
23Na 3/2 11.262 100
7Li 3/2 16.546 92.41
29Si 1/2 -8.465 4.68
17O 5/2 5.772 0.038
15N 1/2 -4.361 0.368

The magnetic moment of the nucleus forces the nucleus to behave as a tiny bar magnet. In the absence of an external magnetic
field, each magnet is randomly oriented. During the NMR experiment the sample is placed in an external magnetic field, B , which0

forces the bar magnets to align with (low energy) or against (high energy) the B . During the NMR experiment, a spin flip of the
0

magnets occurs, requiring an exact quanta of energy. To understand this rather abstract concept it is useful to consider the NMR
experiment using the nuclear energy levels.

14.5.1 https://chem.libretexts.org/@go/page/41398
Figure 14.5.1 : Application of a magnetic field to a randomly oriented bar magnet. The red arrow denotes magnetic moment of
the nucleus. The application of the external magnetic field aligns the nuclear magnetic moments with or against the field.

Nuclear Energy Levels


As mentioned above, an exact quanta of energy must be used to induce the spin flip or transition. For any m, there are 2m+1 energy
levels. For a spin 1/2 nucleus, there are only two energy levels, the low energy level occupied by the spins which aligned with B 0

and the high energy level occupied by spins aligned against B . Each energy level is given by
0

E = −mℏγB0 (14.5.2)

where m is the magnetic quantum number, in this case +/- 1/2. The energy levels for m > 1/2 , known as quadrupolar nuclei, are
more complex and information regarding them can be found here.
The energy difference between the energy levels is then
ΔE = ℏγB0 (14.5.3)

where ℏ is Planks constant.


A schematic showing how the energy levels are arranged for a spin=1/2 nucleus is shown below. Note how the strength of the
magnetic field plays a large role in the energy level difference. In the absence of an applied field the nuclear energy levels are
degenerate. The splitting of the degenerate energy level due to the presence of a magnetic field in known as Zeeman Splitting.

Figure 14.5.1: The splitting of the degenerate nuclear energy levels under an applied magnetic field. The green spheres represent
atomic nuclei which are either aligned with (low energy) or against (high energy) the magnetic field.

Energy Transitions (Spin Flip)


In order for the NMR experiment to work, a spin flip between the energy levels must occur. The energy difference between the two
states corresponds to the energy of the electromagnetic radiation that causes the nuclei to change their energy levels. For most
NMR spectrometers, B is on the order of Tesla (T) while γ is on the order of 10 . Consequently, the electromagnetic radiation
0
7

required is on the order of Hz. The energy of a photon is represented by


E = hν (14.5.4)

and thus the frequency necessary for absorption to occur is represented as:
γB0
ν = (14.5.5)

Hence, NMR experiment measures the resonant frequency that causes a spin flip. For the more advanced NMR users, the sections
on NMR detection and Larmor frequency should be consulted.

Figure 14.5.3 : Absorption of radio frequency radiation to promote a transition between nuclear energy levels, called a spin flip.

14.5.2 https://chem.libretexts.org/@go/page/41398
Nuclear Shielding
The power of NMR is based on the concept of nuclear shielding, which allows for structural assignments. Every atom is
surrounded by electrons, which orbit the nucleus. Charged particles moving in a loop will create a magnetic field which is felt by
the nucleus. Therefore the local electronic environment surrounding the nucleus will slightly change the magnetic field experienced
by the nucleus, which in turn will cause slight changes in the energy levels! This is known as shielding. Nuclei that experinece
differnet magnetic fields due to the local electronic interactions are known as inequivalent nuclei. The change in the energy levels
requires a different frequency to excite the spin flip, which as will be seen below, creates a new peak in the NMR spectrum. The
shielding allows for structural determination of molecules.

Figure 14.5.4 : The effect that shielding from electrons has on the splitting of the nuclear energy levels. Electrons impart their
own magnetic field which shields the nucleus from the externally applied magnetic field. This effect is greatly exaggerated in this
illustration.
The shielding of the nucleus allows for chemically inequivalent environments to be determined by Fourier Transforming the NMR
signal. The result is a spectrum, shown below, that consists of a set of peaks in which each peak corresponds to a distinct chemical
environment. The area underneath the peak is directly proportional to the number of nuclei in that chemical environment.
Additional details about the structure manifest themselves in the form of different NMR interactions, each altering the NMR
spectrum in a distinct manner. The x-axis of an NMR spectrum is given in parts per million (ppm) and the relation to shielding is
explained here.

Figure 14.5.5 : 31P spectrum of phosphinic acid. Each peak corresponds to a distinct chemical environment while the area under
the peak is proportional to the number of nuclei in a given environment.

Relaxation
Relaxation refers to the phenomenon of nuclei returning to their thermodynamically stable states after being excited to higher
energy levels (Figure 14.5.4). The energy absorbed when a transition from a lower energy level to a high energy level occurs is
released when the opposite happens. This can be a fairly complex process based on different timescales of the relaxation. The two
most common types of relaxation are spin lattice relaxation (T1) and spin spin relaxation (T2). A more complex treatment of
relaxation is given elsewhere.

Figure 14.5.6 : The process of relaxation


To understand relaxation, the entire sample must be considered. By placing rhe nuclei in an external magnetic field, the nuclei
create a bulk magnetization along the z-axis. The spins of the nuclei are also coherent. The NMR signal may be detected as long as
the spins are coherent with one another. The NMR experiment moves the bulk magnetization from the z-axis to the x-y plane,
where it is detected.

14.5.3 https://chem.libretexts.org/@go/page/41398
Spin-Lattice Relaxation (T ): T1 is the time it takes for the 37% of bulk magnetization to recovery along Z-axis from the x-y
1

plane. The more efficient the relaxation process, the smaller relaxation time (T1) value you will get. In solids, since motions
between molecules are limited, the relaxation time (T1) values are large. Spin-lattice relaxation measurements are usually
carried out by pulse methods.
Spin-Spin Relaxation (T ): T2 is the time it takes for the spins to lose coherence with one another. T2 can either be shorter or
2

equal to T1.

Applications
The two major areas where NMR has proven to be of critical importance is in the fields of medicine and chemistry, with new
applications being developed daily
Nuclear magnetic resonance imaging, better known as magnetic resonance imaging (MRI) is an important medical diagnostic tool
used to study the function and structure of the human body. It provides detailed images of any part of the body, especially soft
tissue, in all possible planes and has been used in the areas of cardiovascular, neurological, musculoskeletal and oncological
imaging. Unlike other alternatives, such as computed tomography (CT), it does not used ionized radiation and hence is very safe to
administer.

Figure 14.5.7 : 1H MRI of a human head showing the soft tissue such as the brain and sinuses. The MRI also clearly shows the
spinal column and skull. from Wikipedia
In many laboratories today, chemists use nuclear magnetic resonance to determine structures of important chemical and biological
compounds. In NMR spectra, different peaks give information about different atoms in a molecule according specific chemical
environments and bonding between atoms. The most common isotopes used to detect NMR signals are 1H and 13C but there are
many others, such as 2H, 3He, 15N, 19F, etc., that are also in use.
NMR has also proven to be very useful in other area such as environmental testing, petroleum industry, process control, earth’s
field NMR and magnetometers. Non-destructive testing saves a lot of money for expensive biological samples and can be used
again if more trials need to be run. The petroleum industry uses NMR equipment to measure porosity of different rocks and
permeability of different underground fluids. Magnetometers are used to measure the various magnetic fields that are relevant to
one’s study.

References
Atta-ur-Rahman. Nuclear Magnetic Resonance. New York: Springer-Verlag, 1986.
Freeman, Ray. Magnetic Resonance in Chemistry and Medicine. New York: Oxford University Press, 2003.
Lambert, Joseph B and Eugene P Mazzola. Nuclear Magnetic Resonance Spectroscopy: An Introduction to Princliples,
Applications, and Experimental Methods. Upper Saddle River: Pearson Education, 2004.
Chang, Raymond. Physical Chemistry for the Biosciences. University Science Books, 2005

Contributors and Attributions


Derrick Kaseman (UC Davis) and Revathi Srinivasan Ganesh Iyer (UCD)

14.5: Nuclear Magnetic Resonance is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

14.5.4 https://chem.libretexts.org/@go/page/41398
14.6: Electron Spin Resonance
Though less used than Nuclear Magnetic Resonance (NMR), Electron Paramagnetic Resonance (EPR) is a remarkably useful form
of spectroscopy used to study molecules or atoms with an unpaired electron. It is less widely used than NMR because stable
molecules often do not have unpaired electrons. However, EPR can be used analytically to observe labeled species in situ either
biologically or in a chemical reaction.

Introduction
Electron Paramagnetic Resonance (EPR), also known as Electron Spin Resonance (ESR). The sample is held in a very strong
magnetic field, while electromagnetic (EM) radiation is applied monochromatically (Figure 1).

Figure 1(3)-monochromatic electromagnetic beam


This portion of EPR is analogous to simple spectroscopy, where absorbance by the sample of a single or range of wavelengths of
EM radiation is monitored by the end user ie absorbance. The unpaired electrons can either occupy +1/2 or -1/2 ms value (Figure
2). From here either the magnetic field "B0" is varied or the incident light is varied. Today most researchers adjust the EM radiation
in the microwave region, the theory is the find the exact point where the electrons can jump from the less energetic ms=-1/2 to
ms=+1/2. More electrons occupy the lower ms value (see Boltzmann Distribution).

Figure 2: Resonance of a free electron.


Overall, there is an absorption of energy. This absorbance value, when paired with the associated wavelength can be used in the
equation to generate a graph of showing how absorption relates to frequency or magnetic field.

ΔE = hν = ge βB B0 (14.6.1)

-24 -1
where ge equals to 2.0023193 for a free electron; β is the Bohr magneton and is equal to 9.2740 * 10
B J T ; and B0 indicates
the external magnetic field.

Theory
Like NMR, EPR can be used to observe the geometry of a molecule through its magnetic moment and the difference in electron
and nucleus mass. EPR has mainly been used for the detection and study of free radical species, either in testing or anylytical
experimentation. "Spin labeling" species of chemicals can be a powerfull technique for both quantification and investigation of
otherwise invisible factors.
The EPR spectrum of a free electron, there will be only one line (one peak) observed. But for the EPR spetrum of hydrogen, there
will be two lines (2 peaks) observed due to the fact that there is interaction between the nucleus and the unpaired electron. This is

14.6.1 https://chem.libretexts.org/@go/page/41399
also called hyperfine splitting. The distance between two lines (two peaks) are called hyperfine splitting constant (A).
By using (2NI+1), we can calculate the components or number of hyperfine lines of a multiplet of a EPR transtion, where N
indicates number of spin, I indicates number of equivalent nuclei. For example, for nitroxide radicals, the nuclear spin of 14N is 1,
N=1, I=1, we have 2 x 1 + 1 = 3, which means that for a spin 1 nucleus splits the EPR transition into a triplet.
To absorb microwave, there must be unpaired electrons in the system. no EPR signal will be observed if the system contains only
paired electrons since there will be no resonant absorption of microwave energy. Molecules such as NO, NO2, O2 do have unpaired
electrons in groud states. EPR can be also performed on proteins with paramagnetic ions such as Mn2+, Fe3+ and Cu2+.
Additionally, molecules containing stable nitroxide radicals such as 2,2,6,6-tetramethyl-1-piperidinyloxyl (TEMPO, Figure 3) and
di-tert-butyl nitroxide radical.

Figure 3-The nitroxide radical TEMPO


Examples of EPR spectra:

Figure 4 -Stimulated EPR spectrum of CH3 radical

Figure 5 - Stimulated EPR spectrum of methoxymethl ( H2C(OCH3) )radical

References
1. G. Maksina, Yu. S. Arkhangel'skaya and B. A. Dainyak. Russian State Medical University, Moscow. Translated from
Meditsinskaya Tekhnika, No. 5, pp. 32–34, September–October, 1995. external link:
www.springerlink.com/content/n73312406t6941l4/fulltext.pdf
2. E.F. Block. The Role in Coherent Resonance in Human Fair: Part One-Electromagnetic and Gravity. December 2010.
http://journalinformationalmedicine.org/cr1.htm
3. Thomas Engel, Gary Drobny and Philip Reid. Physical Chemistry for the Life Science.s 2008 Pearson Education, Inc. Upper
Saddle River, NJ 07458. pp. 514-516
4. Cortel, Adolf. "Demonstrations on Paramagnetism with an Electronic Balance." J. Chem. Educ. 1998 75 61.
5. Geselbracht, Margaret J.; Cappellari, Ann M.; Ellis, Arthur B.; Rzeznik, Maria A.; Johnson, Brian J. "Rare Earth Iron Garnets:
Their Synthesis and Magnetic Properties." J. Chem. Educ. 1994, 71, 696.
6. Shimada, Hiroshi; Yasuoka, Takashi; Mitsuzawa, Shunmei. "Observation of paramagnetic property of oxygen by simple
method: A simple experiment for college chemistry and physics courses (TD)." J. Chem. Educ. 1990, 67, 63.

14.6.2 https://chem.libretexts.org/@go/page/41399
14.6: Electron Spin Resonance is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
EPR: Introduction is licensed CC BY 4.0.

14.6.3 https://chem.libretexts.org/@go/page/41399
14.7: Fluorescence and Phosphorescence
Singlet and Triplet Excited State
Understanding the difference between fluorescence and phosphorescence requires the knowledge of electron spin and the
differences between singlet and triplet states. The Pauli Exclusion principle states that two electrons in an atom cannot have the
same four quantum numbers (n , l, m , m ) and only two electrons can occupy each orbital where they must have opposite spin
l s

states. These opposite spin states are called spin pairing. Because of this spin pairing, most molecules do not exhibit a magnetic
field and are diamagnetic. In diamagnetic molecules, electrons are not attracted or repelled by the static electric field. Free radicals
are paramagnetic because they contain unpaired electrons have magnetic moments that are attracted to the magnetic field.
Singlet state is defined when all the electron spins are paired in the molecular electronic state and the electronic energy levels do
not split when the molecule is exposed into a magnetic field. A doublet state occurs when there is an unpaired electron that gives
two possible orientations when exposed in a magnetic field and imparts different energy to the system. A singlet or a triplet can
form when one electron is excited to a higher energy level. In an excited singlet state, the electron is promoted in the same spin
orientation as it was in the ground state (paired). In a triplet excited stated, the electron that is promoted has the same spin
orientation (parallel) to the other unpaired electron. The difference between the spins of ground singlet, excited singlet, and excited
triplet is shown in Figure \(\PageIndex{1}\). Singlet, doublet and triplet is derived using the equation for multiplicity, 2S+1, where
S is the total spin angular momentum (sum of all the electron spins). Individual spins are denoted as spin up (s = +1/2) or spin
down (s = -1/2). If we were to calculated the S for the excited singlet state, the equation would be 2(+1/2 + -1/2)+1 = 2(0)+1 = 1,
therefore making the center orbital in the figure a singlet state. If the spin multiplicity for the excited triplet state was calculated, we
obtain 2(+1/2 + +1/2)+1 = 2(1)+1 =3, which gives a triplet state as expected.

Figure 14.7.1 : Spin in the ground and excited states


The difference between a molecule in the ground and excited state is that the electrons is diamagnetic in the ground state and
paramagnetic in the triplet state.This difference in spin state makes the transition from singlet to triplet (or triplet to singlet) more
improbable than the singlet-to-singlet transitions. This singlet to triplet (or reverse) transition involves a change in electronic state.
For this reason, the lifetime of the triplet state is longer the singlet state by approximately 104 seconds fold difference.The radiation
that induced the transition from ground to excited triplet state has a low probability of occurring, thus their absorption bands are
less intense than singlet-singlet state absorption. The excited triplet state can be populated from the excited singlet state of certain
molecules which results in phosphorescence. These spin multiplicities in ground and excited states can be used to explain transition
in photoluminescence molecules by the Jablonski diagram.

Jablonski Diagrams
The Jablonski diagram that drawn below is a partial energy diagram that represents the energy of photoluminescent molecule in its
different energy states. The lowest and darkest horizontal line represents the ground-state electronic energy of the molecule which
is the singlet state labeled as S . At room temperature, majority of the molecules in a solution are in this state.
o

14.7.1 https://chem.libretexts.org/@go/page/41400
Figure 14.7.2 : Partial Jablonski Diagram for Absorption, Fluorescence, and Phosphorescence. from Bill Reusch.
The upper lines represent the energy state of the three excited electronic states: S1and S2 represent the electronic singlet state (left)
and T1 represents the first electronic triplet state (right). The upper darkest line represents the ground vibrational state of the three
excited electronic state.The energy of the triplet state is lower than the energy of the corresponding singlet state.
There are numerous vibrational levels that can be associated with each electronic state as denoted by the thinner lines. Absorption
transitions (blues lines in Figure 14.7.2) can occur from the ground singlet electronic state (So) to various vibrational levels in the
singlet excited vibrational states. It is unlikely that a transition from the ground singlet electronic state to the triplet electronic state
because the electron spin is parallel to the spin in its ground state (Figure 14.7.1). This transition leads to a change in multiplicity
and thus has a low probability of occurring which is a forbidden transition. Molecules also go through vibration relaxation to lose
any excess vibrational energy that remains when excited to the electronic states (S and S ) as demonstrated in wavy lines in
1 2

Figure 14.7.2. The knowledge of forbidden transition is used to explain and compare the peaks of absorption and emission.

Relaxation and Fluorescence


Sometimes, when an excited state species relaxes, giving off a photon, the wavelength of the photon is different from the one that
initially led to excitation. When this happens, the photon is invariably red-shifted; its wavelength is longer than the initial one. This
situation is called "fluorescence".

How can that be? Isn't energy quantized? How is the molecule suddenly taking a commission out of the energy the original photon
brought with it? This discrepancy is related to the Franck-Condon principle from the previous page. When an electron is promoted
to an electronic excited state, it often ends up in an excited vibrational state as well. Thus, some of the energy put into electronic
excitation is immediately passed into vibrational energy. Vibrational energy, however, doesn't just travel in photons. It can be
gained or lost through molecular collisions and heat transfer.

14.7.2 https://chem.libretexts.org/@go/page/41400
The electron might simply drop down again immediately; a photon would be emitted of exactly the same wavelength as the one
that was previously absorbed. On the other hand, if the molecule relaxes into a lower vibrational state, some of that initial energy
will have been lost as heat. When the electron relaxes, the distance back to the ground state is a little shorter. The photon that is
emitted will have lower energy and longer wavelength than the initial one.

Just how does a molecule undergo vibrational relaxation? Vibrational energy is the energy used to lengthen or shorten bonds, or to
widen or squeeze bond angles. Given a big enough molecule, some of this vibrational energy could be transferred into bond lengths
and angles further away from the electronic transition. Otherwise, if the molecule is small, it may transfer some of its energy in
collisions with other molecules.

Note

There are many examples of energy being transferred this way in everyday life. In a game of pool, one billiard ball can transfer
its energy to another, sending it toward the pocket. Barry Bonds can transfer a considerable amount of energy through his bat
into a baseball, sending it out of the park, just as Serena Williams can send a whole lot of energy whizzing back at her sister.

Exercise 1

How does the energy of an electronic absorption compare to other processes? To find out, you might consider the excitation of
an entire mole of molecules, rather than a sinle molecule absorbing a single photon. Calculate the energy in kJ/mol for the
following transitions.
a. absorbance at 180 nm (ultraviolet)
b. absorbance at 476 nm (blue)

14.7.3 https://chem.libretexts.org/@go/page/41400
c. absorbance at 645 nm (red)

Exercise 2
How does the energy of an excitation between vibrational states compare to that of an electronic excitation? Typically, infrared
absorptions are reported in cm-1, which is simply what it looks like: the reciprocal of the wavelength in cm. Because
wavelength and frequency are inversely related, wavenumbers are considered a frequency unit. Calculate the energy in kJ/mol
for the following transitions.
a. absorbance at 3105 cm-1
b. absorbance at 1695 cm-1
c. absorbance at 963 cm-1

In molecules, as one molecule drops to a lower vibrational state, the other will hop up to a higher vibrational state with the energy it
gains. In the drawing below, the red molecule is in an electronic excited and vibrational state. In a collision, it transfers some of its
vibrational energy to the blue molecule.

Radiationless Transitions: Internal Conversion


If electrons can get to a lower energy state, and give off a little energy at a time, by hopping down to lower and lower vibrational
levels, do they need to give off a giant photon at all? Maybe they can relax all the way down to the ground state via vibrational
relaxation. That is certainly the case. Given lots of vibrational energy levels, and an excited state that is low enough in energy so
that some of its lower vibrational levels overlap with some of the higher vibrational levels of the ground state, the electron can hop
over from one state to the other, without releasing a photon.

This event is called a "radiationless transition", because it occurs without release of a photon. The electron simply slides over from
a low vibrational state of the excited electronic state to a high vibrational state of the electronic ground state. If the electron simply
keeps dropping a vibrational level at a time back to the ground state, the process is called "internal conversion".
Internal conversion has an important consequence. Because the absorption of UV and visible light can result in energy transfer into
vibrational states, much of the energy that is absorbed from these sources is converted into heat. That can be a good thing if you
happen to be a marine iguana trying to warm up in the sun after a plunge in the icy Pacific. It can also be a tricky thing if you are a

14.7.4 https://chem.libretexts.org/@go/page/41400
process chemist trying to scale up a photochemical reaction for commercial production of a pharmaceutical, because you have to
make sure the system has adequate cooling available.

Radiationless Transitions: Intersystem Crossing


There is a very similar event, called "intersystem crossing", that leads to the electron getting caught between the excited state and
the ground state. Just as, little by little, vibrational relaxation can lead the electron back onto the ground state energy surface, it can
also lead the electron into states that are intermediate in energy.
For example, suppose an organic molecule undergoes electronic excitation. Generally, organic molecules have no unpaired
electrons. Their ground states are singlet states. According to one of our selection rules for electronic excitation, the excited state
must also have no unpaired electrons. In other words, the spin on the electron that gets excited is the same after excitation as it was
before excitation.
However, that's not the lowest possible energy state for that electron. When we think about atomic orbital filling, there is a rule that
governs the spin on the electrons in degenerate orbitals: in the lowest energy state, spin is maximized (Hund's rule). In other
words, when we draw a picture of the valence electron configuration of nitrogen, we show nitrogen's three p electrons each in its
own orbital, with their spins parallel.

The picture with three unpaired electrons, all with parallel spins, shows a nitrogen in the quartet spin state. Having one of those
spins point the other way would result in a different spin state. One pair of electrons in the p level would be spin-paired, one up and
one down, even though they are in different p orbitals. That would leave one electron without an opposite partner. The nitrogen
would be in a doublet spin state. That is not what happens. The spin state on the left is lower in energy than the state on the right.
That's just one of the rules of quantum mechanics (Hund's rule): maximize spin when orbitals are singly occupied.
It's the same in a molecule with the triplet state lower in energy than the singlet state. Why didn't the electron get excited to the
triplet state in the first place? That's against the rules. But sliding down vibrationally onto the triplet state from the singlet excited
state is not, because it doesn't involve absorption of a photon.

Intersystem crossing can have important consequences in reaction chemistry because it allows access to triplet states that are not
normally avaiable in many molecules. Because triplet states feature unpaired electrons, their reactivity is often typified by radical
processes. That means an added suite of reactions can be accessed via this process.

Phosphorescence: A Radiationless Transition Followed by Emission


Intersystem crossing is one way a system can end up in a triplet excited state. Even though this state is lower in energy than a
singlet excited state, it cannot be accessed directly via electronic excitation because that would violate the spin selection rule
(\Delta S=0\). That's where the electron gets stuck, though. The quick way back down to the bottom is by emitting a photon, but

14.7.5 https://chem.libretexts.org/@go/page/41400
because that would involve a change in spin state, it is not allowed. Realistically speaking, that means it takes a long time. By "a
long time", we might mean a few seconds, several minutes, or possibly even hours. Eventually, the electron can drop back down,
accompanied by the emission of a photon. This situation is called "phosphorescence".

Molecules that display phosphorescence are often incorporated into toys and shirts so that they will glow in the dark.

Contributors and Attributions


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

14.7: Fluorescence and Phosphorescence is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

14.7.6 https://chem.libretexts.org/@go/page/41400
14.8: Lasers
LASER is an acronym for Light Amplification by Stimulated Emission of Radiation. Laser is a type of light source which has the
unique characteristics of directionality, brightness, and monochromaticity. The goal of this module is to explain how a laser
operates (stimulated or spontaneous emission), describe important components, and give some examples of types of lasers and their
applications.

Introduction
The word LASER is an acronym for Light Amplification by Stimulated Emission of Radiation. In 1916, Albert Einstein discovered
the physical principle responsible for this amplification, and the foundation principle is called stimulated emission. It was widely
accepted at that time that laser would represent a big leap in science and technology, even before Theodore H. Maiman built the
first one in 1960. The 1951 Nobel Prize in physics was shared by Charles H. Townes, Nokolay Basov, and Aleksandr Prokhorov, in
citation, “For Fundamental work in the field of quantum electronics, which has led to the construction of oscillator and amplifiers
based on the maser-laser principle”.
The early lasers developed in the 1950s by Charles H. Townes and Arthur Shawlow were gas and solid-state lasers for use in
spectroscopy. The principles of lasers were adapted from masers. MASER is an acronym that stands for Microwave Amplification
by Stimulated Emission of Radiation. It uses the idea of stimulated emission and population inversion to produce a coherent
amplified radiation of light in the microwave region. Stimulated emission is when an electron in an excited state falls back to
ground state after absorbing energy from an incident photon. Amplified radiation or light is produced with the same direction and
energy as the incident light. Population inversion is when you have a greater population of electrons in the excited state than in
ground state. Population inversion is achieved through various pumping mechanisms. The laser uses these same ideas except that
the electromagnetic wave created is in the visible light region. When emission begins the light oscillates within the resonant cavity
and gains magnitude. Once enough light has been acquired, the laser beam is produced. This allows lasers to be used as a powerful
light source. Three unique characteristics of a laser are its properties of monochromaticity, directionality, and brightness.
The monochromaticity of lasers is due to the fact that lasers are highly selective in the wavelength of light produced, which in itself
is due to the resonant frequency inside the active material. Resonant frequency means that the light is oscillating in a single mode
creating a monochromatic beam of light. The property of directionality depends on the angle of which the light propagates out of
the source. Since lasers have large spatial and temporal coherence directionality is maximized. Temporal coherence is when there
are small fluctuations in the phase. Spatial coherence has small changes in the amplitude of the emitted light. Like
monochromaticity, directionality is dependent on the resonant cavity of the active material. The property of brightness is a result of
the directionality and the coherence of the light. Due to these properties, lasers today are used in simple laser pointers, cutting
devices, the development of military technologies, spectroscopy, and medical treatments. Their direct application to spectroscopy
has allowed scientists to measure lifetimes of excited state molecules, structural analysis, probing far regions of the atmosphere,
photochemistry and their use as ionization sources.

History
One of the most important characteristics of light is that it has wave-like properties and that it is an electromagnetic wave.
Experiments on the blackbody radiation demonstrated a comprehensive idea of emission and absorption of electromagnetic waves.
In 1900, Max Plank developed the theory that electromagnetic waves can only exist in distinct quantities of energy, which are
directly proportional to a given frequency (ν ). In 1905, Albert Einstein proposed the dual nature of light, having both wave-like
and particle-like properties. He used the photoelectric effect to show that light acts as a particle, with energy inversely proportional
to the wavelength of light. This is important, because the number of particles is directly related to how intense a light beam will be.
In 1915, Einstein introduced the idea of stimulated emission- a key concept to lasers.
In 1957, Townes and Shawlow proposed the concept of lasers in the infrared and optical region by adapting the concept of masers
to produce monochromatic and coherent radiation. In 1953, Townes was the first to build a maser with an ammonia gas source.
Masers use stimulated emission to generate microwaves. Townes and other scientists wanted to develop the optical maser to
generate light. Optical masers would soon adopt the name LASER: Light Amplification by Stimulated Emission of Radiation. An
optical maser would need more energy than what can be provided by microwave frequencies and a resonant cavity of the order of
1μm or less. Townes and Shawlow proposed the use of a Fabry-Pérot interferometer equipped with parallel mirrors, where
Interference of radiation which is traveling back and forth between parallel mirrors in the cavity allowed for selection of certain
wavelengths. Townes built an optical maser with potassium gas. That failed because the mirrors degraded over time. In 1957,

14.8.1 https://chem.libretexts.org/@go/page/41401
Gordon Gould improved upon Townes' and Shawlow's laser concept. It was Gould who renamed the optical maser to laser. In April
1959, Gould filed a patent for the laser and later in March 1960, Townes and Shawlow had also made a request for a patent. Since
Gould’s notebook was officially dated the idea was his first, but he did not receive the patent until 1977.

Components
A laser consists of three main components: a lasing medium, a resonant or optical cavity, and an output coupler. The lasing medium
consists of a group of atoms, molecules , or ions in solid, liquid or gaseous form, which acts as an amplifier for light waves. For
amplification, the medium has to achieve population inversion, which means in a state in which the number of atoms in the upper
energy level is greater than the number of atoms in the lower energy level. The output coupler serves as energy source which
provides for obtaining such a state of population inversion between a pair of energy levels of the atomic system. When the active
medium is placed inside an optical resonator, the system acts as an oscillator.

Figure 1. Diagram of a typical laser, showing the three major parts.

Lasing Medium
The lasing medium is the component used to achieve lasing, such as chromium in the aluminum oxide crystal-found in a ruby laser.
Helium and neon gas are two materials most commonly used in gas lasers. These are only a few examples of lasing mediums or
materials that have been used in the past and present states of the laser. For further information about different types of lasing
mediums please refer to the section where Types of Lasers is discussed.

Optical Cavity and Output Coupler


Rays of light moving along an optical path tend to diverge over time, because the energy of radiation has very high frequency.
Therefore an optical cavity is needed to refocus the light. Figure 1, represents the basics of an optical cavity where the light inside
moves back and forth between two mirrors. These redirect and focus the light each time it hits the surface of the mirrors. There are
two types of cavities: stable cavities and unstable cavities. A stable cavity is when the ray of light does not diverge far from the
optical axis. An unstable cavity is when the ray of light bounces off and away from a mirrors surface. The importance of the
cavities is that it allows for the laser to have properties of directionality, monochromaticity and brightness.

Figure 2. an optical cavity where the light inside moves back and forth between two mirrors
Light oscillating between the first mirror (Mo) and the second mirror (M1) separated by distance, d, will have a round-trip phase
shift (RTPS) of 2θ=2kd=q2π- ϕ. In Figure 1, a round-trip can be described as the beam traveling from Mo to M1back to Mo.
Resonance occurs in the cavity because the light propagating between the two mirrors is uniform. The ABCD law describes that an
optical cavity has a field distribution, because it reproduces itself as it is making these round-trips between the two parallel mirrors.
The ABCD law was first applied to a Gaussian beam with a beam parameter, q, which is described as
(Aq1 + B)
q2 = (14.8.1)
(C q1 + D)

This law states that the beam, which is oscillating through an optical system, will experience some changing as it moves in the
cavity. Fields that are created in an optical cavity have analogous shape and phase as they make each trip back and forth. However,
the one thing that changes is the size of the field because the electromagnetic wave is unrestricted, unlike a wave in a short-
circuited coaxial cable used to build a resonator or a microwave cavity mode. Since there is a field distribution of Eoat the surface
of Mo, it can be said that there is a field distribution at the surface of M1. Since there will be a change in size of the field this means
that the electromagnetic wave will have change in amplitude by ρ0*ρ1 and a phase factor of e jk2d , creating additional fields.

14.8.2 https://chem.libretexts.org/@go/page/41401
This is an example of a phasorial addition of all fields between Mo and M1, creating a total field ET (Figure 2). Phasorial addition is
described by RTPS, where each additional En will have a delay of angle ϕ which is related to kd.

Figure 3. Totle field ET


ET will always be greater than Eoonly if ρ0 and ρ1are not greater than 1 and ϕ=0. In this case, when ϕ=0 resonance is enhanced
because all factors such as ET travelling between Mo and M1, the intensity of the electromagnetic waves, number of photons
traveling between Moand M1, and the amount of energy that is stored are maximized. The resonant wavelength can also be
determined by using the relationship between RTPS, and because
ωn 2π
k = =
c λ

Using 2θ = 2kd = q2π


2π2d
= q2π
λ


d =
2

Where the wavelength of interest is given by λ = λ 0 /n , where n is the index of refraction and λ is the free-space wavelength .
0

Since we are dealing with light as a wave, the light in the resonant cavity can be described in terms of frequency, ν. Where
2nd 2nd
k2d = ω = 2πν = q(2π)
c c

c
ν =q
2nd

A Fabry-Pérot interferometer is a prime example of an optical cavity used in a laser. The Fabry-Pérot is equipped with two parallel
mirrors, one that is completely reflective and the other that is partially reflective. As light is accumulating in the cavity after taking
several round trips between the two mirrors, some light is transmitted through the partially reflective mirror and a laser beam is
produced. The beam can be in pulsed mode or continuous-wave (CW) mode. To increase the performance of the resonant cavity,
the length of the cavity (d) must be considered as a way to avoid a decrease in the laser beam intensity due to any diffraction losses.
The size of the aperture of the cavity is also important because it determines the strength or the intensity of the laser beam. In fact,
determining the best length of a resonant cavity will enhance the coupling conditions of the output coupler by producing a
frequency that is stable, which ultimately generates a laser beam that is coherent and has high power.

14.8.3 https://chem.libretexts.org/@go/page/41401
Figure 4. The first stage shows that all molecules present are in the ground state since there is no excitation energy present in the
cavity. In the second stage, molecules become excited causing spontaneous emission of photons. In the third stage, photons collide
with excited molecules causing an amplification of the light. This results in stimulated emission where the photon that is released
has the same energy and direction of the photon that created it. Last is the fourth stage, where stimulated emission causes
amplification of the light, and those photons that are traveling parallel to the cavities axis will hit one of the parallel mirrors and
will be reflected back along the same axis. As the light is building up from the back and forth movement from one mirror to
another, some of the light will pass through the partial mirror or the output coupler as a beam of rays. The remaining light in the
cavity will continue producing more photons as long as there is a population inversion.
There are essentially six stages in the lasing process. First is the ground state where there is no excitation of the lasing medium.
Second is pumping, which is applied to the medium where spontaneous emission occurs. Then the third stage is when emitted
photons collide with an excited molecule where stimulated emission occurs. In the fourth stage the photons are produced in
multiples, however those moving parallel in the cavity will hit a mirror and then hit the second mirror. During the fifth stage this
process continues until there is an accumulation of light that is coherent and of a specific frequency. Finally, the sixth stage is when
the light or laser beam exits the partially reflective mirror which is also known as the output coupler. An output coupler is the last
important component of a laser because it must be efficient to produce an output of light with maximum intensity. If the output
coupler is too transparent than there is much more loss of electromagnetic waves and this will decrease lasing significantly because
population inversion will no longer be maintained. If the output coupler or partially reflective mirror is too reflective, then all the
accumulated light that is built up in the resonant cavity will be trapped in the cavity. The beam will not pass through the output
coupler, producing little to no light making the laser ineffective.

Emission
Lasers create a high energy beam of light by stimulated emission or spontaneous emission. Within in a molecule there are discrete
energy levels. A simple molecular description has a low energy ground state (E1) and a high energy excited state (E2). When an
electromagnetic wave, referred to as the incident light, irradiates a molecule there are two processes that can occur: absorption and
stimulated emission.
Absorption occurs when the energy of the incident light matches the energy difference between the ground and excited state,
causing the population in the ground state to be promoted to the excited state. The rate of absorption is given by the equation:
dN1
= −W12 N1
dt

Where N1 is the population in E1, and W12 is the probability of this transition. The probability of the transition can also be related
to the photon flux (intensity of incident light):
W12 = σ12 F

Where F is the photon flux and σ12 is the cross section of the transition with units of area. When absorption occurs photons are
removed from the incident light and the intensity of the light is decreased.
Stimulated emission is the reverse of absorption. Stimulated emission has two main requirements: there must be population in the
excited state and the energy of the incident light must match the difference between the excited and ground state. When these two
requirements are met, population from the excited state will move to the ground energy level. During this process a photon is
emitted with the same energy and direction as the incident light. Unlike absorption, stimulated emission adds to the intensity of the
incident light. The rate for stimulated emission is similar to the rate of absorption, except that it uses the population of the higher
energy level:
W21 = σ21 F

14.8.4 https://chem.libretexts.org/@go/page/41401
Like absorption the probability of the transition is related to the photon flux of the incident light through the equation:
dN2
= −W21 N2
dt

When absorption and stimulated emission occur simultaneously in a system the photon flux of the incident light can increase or
decrease. The change in the photon flux is a combination of the rate equations for absorption and stimulated emission. This is given
by the equation:
dF = σF (N2 − N1 )dτ (14.8.2)

Spontaneous emission has the same characteristics as stimulated emission except that no incident light is required to cause the
transition from the excited to ground state. Population in the excited state is unstable and will decay to the ground state through
several processes. Most decays involve non-radiative vibrational relaxation, but some molecules will decay while emitting a photon
matching the energy of the energy difference between the two states. The rate of spontaneous emission is given by:
dN2
= −AN2 (14.8.3)
dt

Where A is the spontaneous emission probability which depends on the transition involved. The coefficient A is an Einstein
coefficient obtained from the spontaneous emission lifetime. Since spontaneous emission is not competing with absorption, the
photon flux is based solely on the rate of spontaneous emission.

Figure 5. Diagram of spontaneous emission, stimulated emission and absorption in a two energy level system
The population ratio of a molecule or atom is found using the Boltzmann distribution and the energy of the ground state (E1) and
the excited state (E2):
−(E2 − E1 )
N2
=e kT (14.8.4)
N1

Under normal conditions, the majority, if not all, of the population is in the lower energy level (E1). This is because the energy of
the excited is greater than the ground state. Normal thermal energy available (kT) is not enough to overcome the difference, and the
ratio of population favors the ground state. For example, if the difference in energy between two states absorbes light at 500nm, the
ratio of N1 to N2 is 5.1x1041:1. The photon flux of the incident light is directly proportional to the difference in populations. Since
the ground state has more populations, the photon flux decreases: there is more absorption occurring than stimulated emission. In
order to increase the photon flux there must be more population in the excited state than in the ground state, generally known as a
population inversion.
In a two level energy system it is impossible to create the population inversion needed for a laser. Instead three or four level energy
systems are generally used (Figure 5).

14.8.5 https://chem.libretexts.org/@go/page/41401
Figure 6. Three and four level energy system
Three level processes involve pumping of population from the lowest energy level to the highest, third energy state. The population
can then decay down to the second energy level or back down to the first energy level. The population that makes it to the second
energy level is available for stimulated emission. Light matching the energy difference between the second and first energy level
will cause a stimulated emission. Four level systems follow roughly the same process except that population is moved from the
lowest state to the highest fourth level. Then it decays to the third level and lasing happens when the incident light matches the
energy between the third and second level. After lasing there is decay to the first level.

Pumping Process
Pumping is the movement of population from the ground state to a higher excited state. The general rate at which this is done is
given by:
dNg
( )p = Wp Ng
dt

Where Ng is the population in the ground level and Wp is the pump rate. Pumping can be done optically, electronically, chemically
(see chemical laser), using gases at high flow rates, and nuclear fission. Only optical and electrical pumping will be discussed in
detail.

Optical Pumping
Optical pumping uses light to create the necessary population inversion for a laser. Usually high pressure xenon or krypton lamps
are used to excite solid or liquid laser systems. The active material in the laser absorbs the light from the pump lamp, promoting the
population from the ground state to the higher energy state. The material used in the laser can be continuously exposed to the
pumping light which creates a continuous wave laser (CW). A pulsed laser can be created by using flashes of pumping light.
In optical pumping there are three types of efficiency: transfer, lamp radiative, and pump quantum efficiency. Transfer efficiency is
the ratio of the energy created by the lamp and the power of the light emitted by the laser. The lamp radiative efficiency is the
measure of how much electrical power is converted into light in the optical lamp. Pump quantum efficiency accounts for the ratio
of population that decays to the correct energy level and population that decays either back to the ground state or another incorrect
energy level. For example, the overall pumping rate of the first ruby laser was around 1.1%.
The average pump rate for optical pumping depends on the total efficiency of the pump (η ), volume of the laser material (V),
p

ground state population (Ng), power input (P), and frequency of the lasing transition (ν0):
⟨Wp ⟩ = ηp (P /(V Ng ℏυ0 ))

Electrical Pumping
Electrical pumping is a much more complicated process than optical pumping. Usually used for gas and semiconducting lasers,
electrical pumping uses electrical current to excite and promote the ground state population. In a simple gas laser that contains only

14.8.6 https://chem.libretexts.org/@go/page/41401
one species (A), current passes through the gas medium and creates electrons that collide with the gas molecules to produce excited
state molecules (A*):

A+e ⟶ A +e

During electron impact either an ion or an excited state can be created. The ability to make the excited state depends mostly on the
material used in the laser and not the electrical pumping source making it difficult to describe the efficiency of the pumping. Total
efficiencies have been calculated and tabulated for most active materials used in electrical pumping. Where eficiencies range from
a < 0.1% N2 gas laser to 70% for some CO2 gas lasers.
Like the pumping rate of optical pumps, the rate of electrical pumping is found using the overall efficiency of the pump, power
applied, and population of the ground state. However instead of using the frequency of the ground to upper state transition,
electrical pumping uses the energy of the upper state (ħωp) and the volume of the electron discharge (V):
⟨Wp ⟩ = ηp (P /(V Ng ℏωp ))

Pulsed operation
Q-Switching
The technique of Q switching allows the generation of laser pulses of short duration from a few nanoseconds to a few tens of
nanoseconds and high peak power from a few megawatts to a few tens of megawatts.
Suggest we put a shutter into the laser cavity. If the shutter is closed, laser action cannot occur and the population inversion can be
very high. If the shutter is opened suddenly, the stored energy will be released in a short and intense light pulse. This technique is
known as Q-switching. Q here denotes the ratio of the energy stored to the energy dissipated in the cavity. This technique is used in
many types of solid-stat lasers and CO2 lasers to get a high-power pulsed output.

Figure 7. The bandwidth, ∆f , or f1 to f2, of a damped oscillator is shown on a graph of energy versus frequency. The Q factor of
the damped oscillator, or filter, is fc/∆f. The higher the Q, the narrower and 'sharper' the peak is.
To produce high inversion required for Q-switching, four requirements must be satisfied.
1. The lifetime of the upper level must be longer than cavity buildup time.
2. The pumping flux duration must be longer than the cavity build up time.
3. The initial cavity losses must be high enough during the pumping duration to prevent oscillation occurring.
4. The cavity losses must be reduced instantaneously.

Mode-Locking
The technique of mode locking allows the generation of laser pulses of ultrashort duration from less than a picosecond to
femtoseconds and very high peak, a few gigawatts.
Mode-locking is achieved by inducing the different longitudinal modes of a laser to a locked mode. When combining the
electromagnetic waves modes with different frequencies and random phases, they produce a random and average output. When the
modes are added in phase, they combine to produce a total amplitude and intensity output with a repeated pulse.

14.8.7 https://chem.libretexts.org/@go/page/41401
FIgure 8:Laser mode structure

Types of Lasers
There are many different types of lasers with a wide range applications, and below is a brief description of some of the main types.

Solid State Lasers


A solid-state laser is one that uses a solid active medium generally in a rod shape. Inside the active material is a dopant that acts as
the light emitting source. Optical pumping is used to create population inversion of the active material. Solid-state lasers generally
use stimulated emission as the mechanism for creating the high energy beam.
Ruby Laser

The ruby laser was the first operating laser and was built in 1960. It has a three-level (Figure 6) energy system that uses aluminum
oxide with some of the aluminum atom replaced with chromium as its active material. The chromium in the aluminum oxide
crystal is the active part of the laser. Electrons in the ground state of chromium absorb the incident light and become promoted to
higher energy states. The short lived excited state relaxes down to a metastable state with a longer lifetime. Laser emission happens
when there is relaxation from the metastable state back to the ground state.

Figure 9. Schematic of Ruby laser


A xenon flash lamp emitting light at wavelengths of 6600Å and 4000Å (matching the energy needed to excite the chromium
atoms). In order to create resonance of the incident light in the active material silver platting was put at both ends of the ruby rod.
One end was completely covered while the other end was partially covered so lasing light could exit the system.
Nd: YAG Laser
Nd: YAG laser are the most popular type of solid state laser. The laser medium is a crystal of Y3Al5O12 which are commonly called
YAG, an acronym for yttrium aluminum garnet. A simplified energy-level scheme for Nd:YAG is shown in Fig. 9. The λ=1.06 μm
laser transition is the strongest of the 4F3/2→4I11/2 transitions.

14.8.8 https://chem.libretexts.org/@go/page/41401
Figure 10. Simplified energy level diagram of the4F3/2---4I11/2 laser transition in Nd:YAG showing the four- level nature of the
system
The major application of the Nd laser is in various form of material processing: drilling, spot welding, and laser marking. Because
they can be focused to a very small spot, the laser are also used in resistor trimming and in circuit mask , memory repair and also in
cutting out specialized circuits. Medical applications include many types of surgeries. Many medical applications take advantage of
low-loss optical fiber delivery systems that can be inserted into the body to wherever is needed. Nd lasers are also used in military
applications such as range finding and target designation. High power pilsed versions are also used for X-ray spectral regions. In
addition, Nd lasers are used in scientific lab as good sources for pumping dye laser and other types of lasers.
Semiconductor Laser
The semiconductor laser is another type of solid state laser that uses a semiconducting material like germanium or silicon. When
the temperature of the semiconducting material is increased, electrons move from the valence band to the conducting band creating
holes in the valence band (Figure 7). In between the conducting band and valence band is a region where there are no energy
levels, called the band gap. Applying a voltage to the semiconductor causes electrons to move to the conduction band creating a
population inversion. Irradiating a semiconductor with incident light matching the energy of the forbidden area causes a large
transition from the conduction band to the valence band, increasing and amplifying the incident light.

Figure 11. Schematic of semicondecor laser

Gas Lasers
A gas laser contains active material composed of a mixture of gases with similar energy states inside a small gas chamber.
Electrical pumping is used to create the population inversion where one gas is excited through collisions with electrons and in turn
excites the other gas through collisions.
Helium-Neon Laser
The helium-neon laser was the first gas laser. It consists of a long narrow tube that contains He and Ne gas. Mirrors are placed at
both ends of the gas tube to form the resonant cavity with one of the mirrors partially reflecting the incident light. Stimulated
emission of the gas mixture is carried out by first exciting the He gas to a higher energy state through electron collision with
electrons from the electronic pumping source (electrical pumping). Then the excited He atoms collide with the Ne atoms
transferring their energy and exciting them to a higher energy level. The Ne atoms in the higher energy level will then relax to a
lower metastable energy state. Lasing occurs when there is relaxation from the metastable state to a lower energy state causing
spontaneous emission. The Ne gas then returns to the ground state when it collides with the outer walls of the gas tube (Figure 8).

14.8.9 https://chem.libretexts.org/@go/page/41401
Figure 12. Schematic of gas laser
Carbon Dioxide Laser

The carbon dioxide laser is a gas laser that uses the energy difference between rotational-vibrational energy levels. Within the
vibrational levels of CO2 there are rotational sub-energy levels. A mixture of N2 and CO2 gas are placed inside a chamber. The N2
atoms are excited through an electrical pumping mechanism. The excited atoms then collide with the CO2 atoms transfer energy.
This transfer of energy causes the CO2 to go into a higher vibrational level. The excited CO2 molecules then go through
spontaneous emission when they are relaxed to lower rotational-vibrational levels increasing the signal of the incident light (Figure
9). Carbon dioxide lasers are extremely efficient, around 70%, and powerful compared to other gas lasers making them useful for
welding and cutting.

Figure 13. Schematic of carbondioxide laser

Liquid Lasers
Liquid lasers consist of a liquid active material usually composed of an organic dye compound. The most common type of liquid
laser uses rhodamine 6G (Figure 10) dye mixed with alcohol and is excited by different types of lasers, such as an argon-ion laser
or a nitrogen laser. Organic dyes are large compounds that have absorption bands in the UV or visible region with a strong intense
fluorescence spectrum. The free π electrons of the dye are excited using an optical pumping source and the transition from the S1 to
the S0 state creates the lasing light (see Jablonski diagrams). Liquids are generally used because they can easily be tuned to emit a
certain wavelength by changing the resonant frequency within the cavity. Wavelengths from the visible to the infrared can be
covered. There are many benefits of liquid lasers, some include that they can be cooled in a relative amount of time, they cannot be
damaged unlike a solid-laser, and their production is cost-effective. The efficiency of liquid lasers is low because the lifetime of the
excited state is relatively short; there are many non-radiative decay processes, and the material degrades over time. Liquid lasers
tend to be used only as a pulse laser when tunability is required. Liquid lasers can be used for high-resolution spectroscopy since
they are easily tuned over a wide range of wavelengths. They can also be used because they have concentrations which are
manageable when dissolved in solids or other liquids.

14.8.10 https://chem.libretexts.org/@go/page/41401
Figure 14. Rhodamine 6G molecule

Chemical Lasers
Chemical lasers are different from other lasers because the population inversion is the direct product of a chemical reaction when
energy is released as a result of an exothermic reaction. Usually reactions involve gases where the energy created is used to make
vibrationally excited molecules. Light used for lasing is then created from vibrational-rotational relaxation like in the CO2 gas laser.
An example of a chemical laser is the HF gas laser. Inside the gas chamber fluorine and hydrogen react to form an excited HF
molecule:
F + H2 → HF + H
The excess energy from the reaction allows HF to stay in its excited state. As it relaxes, light is emitted through spontaneous
emission. Deuterium can also be used in place of hydrogen. Deterium fluoride is useds for applications that require high-power. For
example, MIRACL was built for military research and was known to produce 2.2 megawatts of power. The uniqueness of a
chemical laser is that the power required for lasing is produced in the reaction itself.

Laser Applications
The applications of lasers are numerous and cover scientific and technological fields. In general, these applications are a direct
consequence of the special characteristics of lasers. Below are a few examples of the laser applications, for a complete list please
go to en.Wikipedia.org/wiki/List_of...ons_for_lasers

Lidar
Lidar is short for light detection and ranging which is an optical remote sensing technology can be used for monitoring the
environment. A typical lidar system involves a transmitter of laser radiation and a receiver for the detection and analysis of
backscattered light. A beam expander is usually used at transmitter to reduce divergence of the laser beam before it propagates into
the atmosphere. The receiver includes a wavelength filter, a photo detector, and computers and electronics for data acquisition and
analysis.
Lidar system dates back to the 1930, because of the laser, it has become one of the primary tools in atmospheric and environmental
research. Other than that, Lidar has been put into various uses. In agriculture, lidar can be used to create topographic map to help
farmer to decide appropriate amount of fertilizing to achieve a better crop yield. In Archaeology, lidar can be used to create a
geographic information system to help archaeologists to find sites. In transportation, lidar has been used in autonomous cruise
control system to prevent road accident and policemen are also using lidar speed gun to enforce the speed limit regulation.

Laser in Material Processing


The beam of a laser is usually a few millimeters in diagram. For most material processing applications, lenses are used to increase
the intensity of the beam. The beam from a laser is either plane or spherical. After passing through a lens, the beam should get
focused to a point. But in actual practice, diffraction effects have to be taken into consideration, the incoming will focus into a
region of radius. If λ is the wavelength of the laser light, a is the radius of the beam, and f is the focal length of the lens, then the
radius of the region is
λf
b =
a

If P represents the power of the laser beam, the intensity I, obtained at the focused region would be given by,

14.8.11 https://chem.libretexts.org/@go/page/41401
2
P Pa
I = =
2 2
πb ππλ f 2

The high-power (P>100w) laser are widely used in material processing such as welding, drilling, cutting, surface treatment, and
alloying. The main advantage of the laser beam can be summarized as follow: (1) The heating produced by the laser is less than
that in conventional process. Material distortion is considerably reduced. (2) Possibility of working in inaccessible region. Any
region which can be seen can be processed by a laser. (3) The process can be better controlled and easily automatized. However,
against all these advantages, the disadvantages are: (1) high cost of the laser system. (2) Reliability and reproducibility problems of
the laser system. (3) Safety problems.

Laser in Medicine
In field of medicine, the major use of lasers is for surgery such as laser eye surgery commonly known as LASIK. Besides that,
there are also a few diagnosetic applications such as clinical use of flow microfluormeters, Doppler velocimetry to measure the
blood velocity, laser fluorescence bronchoscope to detect tumors in their early phase.
For surgery, the laser beams are used instead of a conventional scalpel. The infrared beam from the CO2 laser is strongly absorbed
by water molecules in the tissue. It produces a rapid evaporation of these molecules, consequently cutting the tissue. The main
advantage of laser beam surgery can be summarized as follows: (1) High precision. The incision can be made with a high precision
particularly when the beam is directed by means of a microscope. (2) Possibility of operating in inaccessible region. Laser surgery
can be operated in any region of the body which can be observed by means of an optical system. (4) Limited damage to blood
vessel and adjacent tissue. However, the disadvantages are: (1) considerable cost. (2) Smaller velocity of the laser scalpel. (3)
Reliability and safety problems associated with the laser procedure.

References
1. Arias, P., Jaeckey, J., Redondo, J., & Ringwald, A. (2010). Optimizing light-shining-through-a-wall experiments for axion and
other weakly interacting slim particle searches. The American Physical Society, 82(11), DOI: 10.1103/PhysRevD.82.11501
2. Bertolotti, M. (1999). The History of the Laser. Philadelphia, PA: Institute of Physics Publishing.
3. Ferrai, G., et al. (2010). Coherent Addition of Laser Beams in Resonant Passive Optical Cavities. OpticsInfoBase, 35(18), DOI:
10.1364/OL.35.003105
4. Hecht, J. (1992). Laser pioneers. San Diego, CA: Academic Press Limited.
5. Horn, D. T. (1988). Laser experimenter's handbook. Blue Ridge Summit, PA: Tab Books Inc.
6. Kenichi, I. (1988). Fundamentals of Laser Optics. Yokohama, Japan: Plenum Press.
7. Khanin, Y. (1995). Principles of Laser Dynamics. (P. Press, Ed.) New York, NY: Elsevier.
8. Verdeyen, J. (1989). Laser Electronics. Englewood Cliffs, NJ: Prentice-Hall, Inc.
9. Milonni,P.,Eberly, J., (2010) Laser Physics.Hoboken NJ:John Wiley&Sons, Inc.
10. Silfvast,W.,(1996) Laser Fundamentals. New York,NY: the Press Syndicate of the university of Cambridge.
11. Thyagarajan, K. (1981) Lasers: Theory and Application. New York, NY: Plenum Press.
12. Hooker,S., Webb,C., (2010) Laser Physics. Oxford, NY:Oxford University Press.
13. Svelto, O. (1982). Principles of Lasers. (2nd, Ed.) New York, NY: Plenum Press.
14. Coherent, I. (Ed.). (1980). Lasers Operation, Equipment, Application and Design. Palo Alto, CA: McGraw-Hill.

Outside Links
http://en.Wikipedia.org/wiki/Laser
en.Wikipedia.org/wiki/List_of_laser_types
en.Wikipedia.org/wiki/Q_factor
en.Wikipedia.org/wiki/Mode-locking

Problems
1. Determine the free-space wavelength (λ ) in Å, and frequency of the resonant cavity for a beam parameter q1 that is 632,110
0

and q2 that is 632,111 in a helium-neon gas laser at 1 atm. The index of refraction n is 1.00, the length of the resonant cavity is
20 cm and the wavelength region of interest is 6328Å.
2. What wavelength of light will be released by the spontaneous emission of Ne gas, where the the energy difference between the
excited and ground state is 9.9 x 10-19 J.
3. What is the population ratio of the above question at 300K.

14.8.12 https://chem.libretexts.org/@go/page/41401
Answers
1. For q1 λ = 6328.0125\angstromÅ, q2 λ
0 0 = 6328.0025\angstrom Å, and v=474 THz for both q values
2. 200 nm
3. N2/N1=1.9 x 10-105

Contributors and Attributions


Greg Allen (UCD), Arpana Vaniya (UCD), Zheng Zhang (UC Davis)

14.8: Lasers is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

14.8.13 https://chem.libretexts.org/@go/page/41401
14.9: Optical Rotatory Dispersion and Circular Dichroism
Circular Dichroism, an absorption spectroscopy, uses circularly polarized light to investigate structural aspects of optically active
chiral media. It is mostly used to study biological molecules, their structure, and interactions with metals and other molecules.

Introduction
Circular Dichroism (CD) is an absorption spectroscopy method based on the differential absorption of left and right circularly
polarized light. Optically active chiral molecules will preferentially absorb one direction of the circularly polarized light. The
difference in absorption of the left and right circularly polarized light can be measured and quantified. UV CD is used to determine
aspects of protein secondary structure. Vibrational CD, IR CD, is used to study the structure of small organic molecules, proteins
and DNA. UV/Vis CD investigates charge transfer transitions in metal-protein complexes.

Circular Polarization of Light


Electromagnetic radiation consists of oscillating electric and magnetic fields perpendicular to each other and the direction of
propagation. Most light sources emit waves where these fields oscillate in all directions perpendicular to the propagation vector.
Linear polarized light occurs when the electric field vector oscillates in only one plane. In circularly polarized light, the electric
field vector rotates around the propagation axis maintaining a constant magnitude. When looked at down the axis of propagation
the vector appears to trace a circle over the period of one wave frequency (one full rotation occurs in the distance equal to the
wavelength). In linear polarized light the direction of the vector stays constant and the magnitude oscillates. In circularly polarized
light the magnitude stays constant while the direction oscillates.

Figure 1: Diagram of linearly polarized and circularly polarized light


As the radiation propagates the electric field vector traces out a helix. The magnetic field vector is out of phase with the electric
field vector by a quarter turn. When traced together the vectors form a double helix.

Light can be circularly polarized in two directions: left and right. If the vector rotates counterclockwise when the observer looks
down the axis of propagation, the light is left circularly polarized (LCP). If it rotates clockwise, it is right circularly polarized
(RCP). If LCP and RCP of the same amplitude, they are superimposed on one another and the resulting wave will be linearly
polarized.

Figure 2: The superposition of LCP and RCP light of the same amplitude produces linearly polarized light

Interaction with Matter


As with linear polarized light, circularly polarized light can be absorbed by a medium. An optically active chiral compound will
absorb the two directions of circularly polarized light by different amounts

ΔA = Al − Ar (14.9.1)

14.9.1 https://chem.libretexts.org/@go/page/41402
This can be extended to the Beer-Lambert Law. The molar absorpitivty of a medium will be different for LCP and RCP. The Beer-
Lambert Law can be rewritten as
A = (εl − εr )cl (14.9.2)

The difference in molar absorptivity is also known as the molar circular dichroism
Δε = εl − εr (14.9.3)

The molar circular dichroism is not only wavelength dependent but also depends on the absorbing molecules conformation, which
can make it a function of concentration, temperature, and chemical environment.
Any absorption of light results in a change in amplitude of the incident wave; absorption changes the intensity of the light and
intensity of the square of the amplitude. In a chiral medium the molar absorptivities of LCP and RCP light are different so they will
be absorbed by the medium in different amounts. This differential absorption results in the LCP and RCP having different
amplitudes which means the superimposed light is no longer linearly polarized. The resulting wave is elliptically polarized.

Figure 3: The diagrams of the superposition of LCP and RCP light when viewed down the axis of propagation. On the left the two
circular waves (red and green) have the same amplitude which produces linearly polarized light (blue). On the right the LCP (red)
has a larger amplitude than the RCP (green), the superposition of the two waves (blue) forms an ellipse.

Molar Ellipticity
The CD spectrum is often reported in degrees of ellipticity, θ , which is a measure of the ellipticity of the polarization given by:
El − Er
tanθ = (14.9.4)
El + Er

where E is the magnitude of the electric field vector.

Figure 4: Elliptically polarized light (purple) is the superposition of LCP (red) and RCP (blue) light. θ is the angle between the
magnitude of the electric field vector at its maximum and its minimum
The change in polarization is usually small and the signal is often measured in radians where θ = 2.303

4
(Al − Ar ) and is a function
of wavelength. θ can be converted to degrees by multiplying by which gives θ = 32.98ΔA
180

The historical reported unit of CD experiments is molar ellipticity, [θ] , which removes the dependence on concentration and path
length

[θ] = 3298Δε (14.9.5)

where the 3298 converts from the units of molar absorptivity to the historical units of degrees⋅ cm2⋅dmol-1.

Applications

14.9.2 https://chem.libretexts.org/@go/page/41402
Instrumentation
Most commercial CD instruments are based on the modulation techniques introduced by Grosjean and Legrand. Light is linearly
polarized and passed through a monochromator. The single wavelength light is then passed through a modulating device, usually a
photoelastic modulator (PEM), which transforms the linear light to circular polarized light. The incident light on the sample
switches between LCP and RCP light. As the incident light swtches direction of polarization the absorption changes and the
differention molar absorptivity can be calculated.

Figure 5: The instrumentation for a common CD spectrometer showing the polarization of light and the differential absorption of
LCP and RCP light.

Biological molecules
The most widely used application of CD spectroscopy is identifying structural aspects of proteins and DNA. The peptide bonds in
proteins are optically active and the ellipticity they exhibit changes based on the local conformation of the molecule. Secondary
structures of proteins can be analyzed using the far-UV (190-250 nm) region of light. The ordered α -helices, β-sheets, β-turn, and
random coil conformations all have characteristic spectra. These unique spectra form the basis for protein secondary structure
analysis. It should be noted that in CD only the relative fractions of residues in each conformation can be determined but not
specifically where each structural feature lies in the molecule. In reporting CD data for large biomolecules it is necessary to convert
the data into a normalized value that is independent of molecular length. To do this the molar ellipticity is divided by the number of
residues or monomer units in the molecule.
The real value in CD comes from the ability to show conformational changes in molecules. It can be used to determine how similar
a wild type protein is to mutant or show the extent of denaturation with a change in temperature or chemical environment. It can
also provide information about structural changes upon ligand binding. In order to interpret any of this information the spectrum of
the native conformation must be determined.
Some information about the tertiary structure of proteins can be determined using near-UV spectroscopy. Absorptions between
250-300 nm are due to the dipole orientation and surrounding environment of the aromatic amino acids, phenylalanine, tyrosine,
and tryptophan, and cysteine residues which can form disulfide bonds. Near-UV techniques can also be used to provide structural
information about the binding of prosthetic groups in proteins.
Metal containing proteins can be studied by visible CD spectroscopy. Visible CD light excites the d-d transitions of metals in chiral
environments. Free ions in solution will not absorb CD light so the pH dependence of the metal binding and the stoichiometry can
be determined.
Vibrational CD (VCD) spectroscopy uses IR light to determine 3D structures of short peptides, nucleic acids, and carbohydrates.
VCD has been used to show the shape and number of helices in A-, B-, and Z-DNA. VCD is still a relatively new technique and
has the potential to be a very powerful tool. Resolving the spectra requires extensive ab initio calculations, as well as, high
concentrations and must be performed in water, which may force the molecule into a nonnative conformation.

References
1. Woody, R. W. Circular-Dichroim. Methods in Enzymology 246, 34-71 (1995).
2. Johnson, W. C. Protein secondary structure and circular dichroism: A practical guide. Proteins: Structure, Function, and
Genetics 7, 205–214 (1990).

14.9.3 https://chem.libretexts.org/@go/page/41402
3. Drake, A. F. Polarisation modulation-the measurement of linear and circular dichroism. Journal of Physics E: Scientific
Instruments 19, 170–181 (1986).
4. Neidig, M. L., Wecksler, A. T., Schenk, G., Holman, T. R. & Solomon, E. I. Kinetic and spectroscopic studies of N694C
lipoxygenase: a probe of the substrate activation mechanism of a nonheme ferric enzyme. Journal of the American Chemical
Society 129, 7531–7537 (2007).
5. Polavarapu, P. L. & Zhao, C. X. Vibrational circular dichroism: a new spectroscopic tool for biomolecular structural
determination. Fresenius Journal Anaytical Chemistry 366, 727–734 (2000).

14.9: Optical Rotatory Dispersion and Circular Dichroism is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by LibreTexts.
Circular Dichroism by Nick Hurlburt is licensed CC BY 4.0.

14.9.4 https://chem.libretexts.org/@go/page/41402
14.E: Spectroscopy (Exercises)
14.1: Vocabulary
Q14.2
Find the wave number and frequency of light with a wavelength of 700 nm (to three significant figures).

S14.2
1. Wave number is equal to the reciprocal of wavelength. Its units are inverse centimeters.
$$ \tilde \nu = \dfrac{1}{\lambda} \]
$$ \tilde \nu = \dfrac{1}{700\ nm} \times \dfrac{10^9\ nm}{100\ cm} \]
$$ \tilde \nu = 1.43\ \times 10^4\ cm^{-1} \]
2. Use the relationship between frequency, speed, and wavelength to solve for frequency.
$$ \nu = \dfrac{c}{\lambda} \]
$$ \nu = \dfrac{2.998\ \times 10^8\ m/s}{700\ nm \times \dfrac{m}{10^9\ nm}} \]
$$ \nu = 4.28 \times 10^{14}\ s^{-1} \]

Q14.2a
Convert 3 × 10 4
cm
−1
to wavelength. Identify what kind of spectroscopy?

Q14.2a
What is the frequency and wavenumber of a 740 nm photon?

Q14.2b
The wavelength of the red line in the Hydrogen spectrum is 656 nm (656 × 10 −9
m ). What is the wavenumber and frequency of
it?

S14.2b
The wave number =1/λ = 1/656 x10-9 m = 1.5x106 m-1

Q14.2c
Convert 533 nm to wavenumber and frequency.

Q14.4a
Convert the following absorbance to percent transmittance: (a) 0.56, (b) 1.5, (c) 6.8.

Q14.4b
Calculate percent transmittance from the following values of absorbance:
a. 4.0
b. 0.23
c. 1.6

S14.4b
Solve, using the relationship between transmittance and absorbance.
$$ -\log){10}T = A \]
$$ T = 10^{-A} \]
For percent transmittance, multiply T by 100%
(a)
$$ T = 10^{-4.0} \times 100\% \]
$$ T = 0.010\% \]

14.E.1 https://chem.libretexts.org/@go/page/41340
(b)
$$ T = 10^{-0.23} \times 100\% \]
$$ T = 59\%\]
(c)
$$ T = 10^{-1.6} \times 100\% \]
$$ T = 2.5\% \]
Answers: 0.010%, 59%, 2.5%

Q14.4c
Convert the following from percent transmittance to absorbance.
a. 0.10%
b. 23%
c. 84%

Q14.4c
What is the percent transmitter of the following absorbance
a. 0.4
b. 1.2

S14.4c
We have A = 2- log %T
a. 0.4 = 2-log%T -->%T= 10(2-0.4)
%T= 39.8
b. 1.2 = 2-log%T ---> %T = 6.3%

Q14.6a
When molecules are exposed to radiation with frequency, v, such that ΔE = hv , do they travel through a transition from a higher
to lower state, or lower to higher state?

Q14.6b
Find the uncertainty of simultaneously measuring the frequency and wavelength of an emission, if the wavelength is 430 nm and
the excited state lifetime is 0.50 nanoseconds.

S14.6b
1. Use Heisenberg's uncertainty principle and the relationship between energy and frequency to find the uncertainty of frequency.
$$ \Delta E \Delta t \geq \dfrac{h}{4 \pi} \]
$$ \Delta E = h \Delta \nu \]
$$ h \Delta \nu \Delta t \geq \dfrac{h}{4 \pi} \]
$$ \Delta \nu \geq \dfrac{1}{4 \pi \Delta t} \]
The maximum value for Δt is the lifetime of the excited state.
$$ \Delta \nu \geq \dfrac{1}{4 \pi 0.50 ns \times \dfrac{s}{10^9\ ns}} \]
$$ \Delta \nu \geq 1.6 \times 10^8\ s^{-1} \]
2. Use the uncertainty of frequency and the relationship between frequency and wavelength to find the uncertainty of the
wavelength.
$$ \lambda = \dfrac{c}{\nu} \]
$$ | \Delta \lambda | = \dfrac{c}{\nu^2} | \Delta \nu | \]

14.E.2 https://chem.libretexts.org/@go/page/41340
$$ | \Delta \lambda | = \dfrac{\lambda ^2 | \Delta \nu | }{c} \]
$$ | \Delta \lambda | = \dfrac{(430\ nm)^2 \times 1.6 \times 10^8\ s^{-1}}{2.998 \times 10^8\ m/s} \times \dfrac{m}{10^9\ nm} \]
$$ | \Delta \lambda | = 2.3 \times 10^{-7}\ nm \]

Q14.6c
Calculate the wavelength emission of an electronically excited molecule with uncertainties in frequency (∆v) is 5.6x106 s-1and the
wavelength ( ∆λ) 4x10-6 nm? c= 3x108ms-1=3x1017nms-1

S14.6c
We have:
λ| Δ v|
| Δ λ| = (14.E.1)
v

Since
c/
ν = (14.E.2)
λ

then
2
λ | Δ v|
| Δ λ| = (14.E.3)
c

=> 4x10-6nm = (λ2 x 5.6x106 s-1)/ 3x1017nms-1


=> λ = 463 nm

Q14.8a
In the gas phase, using electronic spectroscopy, what is observed on the electronic spectra of diatomic molecules at high resolution
in terms of structure and bands?

Q14.8b
Explain why the decreasing of temperature will increase the resolution of visible and UV spectra?

S14.8b
Decreasing temperature will lower the kinetic energy of molecules. Thus the effect of Doppler and collisional broadening decrease,
making the resolution enhances.

Q14.8c
Measuring the spectrum of the UV light at low temperature is a good way to enhance the resolution of the UV light. Explain. Give
one more example.

Q14.10
In a 5.0 mM solution, a solute absorbs 90% of a visible light as the beam passes through a 80 mm cell. Calculate the molar
absorptivity of this solute.

S14.10
We can calculate the transmittance:

T = 1.00 − 0.9 = 0.10 (14.E.4)

Also, the absorbance is

A = − log T = − log(0.10) = 1.0 (14.E.5)

Next step, use Beer-Lambert law to determine the molar absorptivity,


A
ϵ= (14.E.6)
bc

14.E.3 https://chem.libretexts.org/@go/page/41340
1 −2 −1 −1
ϵ= = 2.5 × 10 L mol cm (14.E.7)
(8.0 cm)(5.0 mM )

Q14.10
Calculate the absorbance (A ) with the molar absorptivity= 6.17 , c= 0.52 M when a certain wavelength passes through
2.3-cm cell.

S14.10
A = ϵbc = (6.17)(2.3)(0.52) = 7.38 (14.E.8)

Q14.12
Calculate E vib for a harmonic oscillator if v = 2 and ν = 3.24 × 10
13
Hz

S14.12
1 1
−34 13 −20
Evib = (v + )hν = (2 + )6.626 × 10 × 3.24 × 10 = 5.37 ∗ 10 Hz (14.E.9)
2 2

Q14.24
Given the following molecules: CO2, H2O. Show the fundamental vibration modes for each of the molecule and explain which one
are IR active or both.

S14.24

Q14.26
Given the following molecules: He2, F2, H2, O2 and Li2. Rank these molecules from lowest to highest fundamental frequency of
vibration? Show works.

S14.26
This can be determined by looking at the mass of the molecules. The one with the lowest mass will have the highest fundamental

frequency of vibration and vice versa. and the similar calculation can be applied for other
molecules, and the rank is as the following: Lowest --- F2 < O2 < Li2 < He2 < H2 ---- highest

14.5: Nuclear Magnetic Resonance


Q14.12
A chemist used the NMR machine to scan his sample and he obtained a signal-to-noise (S/N) ratio of 2.0. How long would it takes
the chemist to generate a spectrum with a S/N ratio of 40? Assumed he spent 5 minutes per scan.

S14.12

14.E.4 https://chem.libretexts.org/@go/page/41340
We are looking at the . This mean the ratio of signal to noise would change according to the number of n scans Thus,

Since 400 scans at the interval of 5 mins, it would take 400 ×


5min = 2000min (1h/60min) = 33.3 h to generate this spectrum

Q14.28
Explain why pentacene crystals are blue in color, but tetracene crystals are orange.

pentacene tetracene

Q14.29
Use the particle in a one-dimensional-box model to calculate the longest wavelength peak in the absorption spectrum of ß-carotene
(structure shown below).

Q14.32a
What does chemical shift measure? What can influence chemical shift?

S14.32a
Chemical shift measures the difference in resonance frequencies between a nucleus of interest and a reference nucleus. It is
influenced by electron shielding; the more electrons are pulled away from a proton of interest (the smaller the electron density
around the nucleus of interest), the greater the chemical shift.

Q14.32b
Assume you are doing an NMR spectroscopy. You operate the spectrometer at 100 Hz, but you find a signal of a compound at an
unknown frequency downfield from TMS peak. However, you know its chemical shift is 6.5 ppm. Calculate the unknown
frequency for your lab report.

Q14.34b
Assuming the precession frequency is 100 MHz and γ = 10.0x105 T-1s-1, calculate the Larmor frequency for 17O .

Q14.32c
Identify the most shielded and deshielded hydrogen for this compound:

S14.32c
The hydrogen groups on the far right end are the most shielded and the hydrogen group on the carbon with the chlorine group is the
most deshielded.

Q14.32d
The NMR signal of coumpound is 280 Hz downfield from TMS is 70 MHZ. find its chemical shifl in ppm.

14.E.5 https://chem.libretexts.org/@go/page/41340
S14.32d
ν − νref
6
δ = × 10 (14.E.10)
νspec

280 H z 6
δ = × 10 = 4.0 ppm (14.E.11)
6
70 × 10

Q14.34
a. Calculate the magnetic field, B0 that corresponds to a precession frequency of 600 MHz for 1H.
b. What is the field strength (in tesla) needed to generate a 1H frequency of 500 MHz?
c. How do spin-spin relaxation and spin-lattice relaxation differ from each other?
d. The 1H NMR spectrum of toluene shows that it has two peaks because of methyl and aromatic protons recorded at 60 MHz and
1.41 T. Given this information, what would be the magnetic field at 400 MHz?
e. What is the difference between 13C and 1H NMR?

S14.34
a. B0= 14.1 T.
b. Using the equation used in problem 1 and solving it for B0we get a field strength of 11.74 T.
c. Look under relaxation.
d. Since we know that the NMR frequency is directly proportional to the magnetic strength, we calculate the magnetic field at 400
MHz: B0 = (400 MHz/60MHz) x 1.41 T = 9.40 T
e. Look under applications.

Q14.34
calculate the field strength in tesla to generate 1H frequency of 300 MHz?

S14.34
B0 = 2πv/ɣ = 2π(300 x 106 s-1)/ 26.75 x 107 T-1s-1
B0 = 7.04 T

Q14.36
Given 4.7 T, calculate the diference in frequency for 2 protons whose § value differ 1.25 and 400 MHz

S14.36
B0 = (Δδ x vspec) / 106 = (1.25 (200 x 106 Hz)) / 106
B0 = 2.5 x 102 Hz

Q14.38
Draw the NMR spectrum of isobutyl alcohol with chemical shift -CH 0.90 ppm, -A-H 1.68ppm, -CH2 3.26 ppm, O-H 4.49 ppm

S14.38

14.E.6 https://chem.libretexts.org/@go/page/41340
14.6: Electron Spin Resonance
Q14.42
You performed an electron spin resonance (ESR) experiment with di-tert-butyl nitroxide radical and get 3 lines of equal intensity.
Then, you combine di-tert-butyl nitroxide radical with ascorbic acid and was about to run another ESR experiment but Jill stopped
you. Why did Jill stop you?

14.7: Fluorescence and Phosphorescence


Q14.43
Someone has handed you data of the luminescence of a material as a function of time. How can you decide whether the
luminescence process was fluorescence or phosphorescence?

14.8: Lasers
14.9: Optical Rotatory Dispersion and Circular Dichroism
14.E: Spectroscopy (Exercises) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

14.E.7 https://chem.libretexts.org/@go/page/41340
CHAPTER OVERVIEW

15: Photochemistry and Photobiology


Photochemistry is the branch of chemistry concerned with the chemical effects of light. Generally, this term is used to describe a
chemical reaction caused by absorption of high energy light. Photobiology is the scientific study of the interactions of the
photochemistry of living organisms and includes the study of photosynthesis, visual processing, circadian rhythms,
bioluminescence, and ultraviolet radiation effects.
15.1: Introduction to Photochemistry
15.2: Photosynthesis
15.3: Vision
15.4: Biological Effects of Radiation

Thumbnail: Photochemical immersion well reactor (50 mL) with a mercury-vapor lamp. (CC BY-SA 4.0, Masohe).

15: Photochemistry and Photobiology is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1
15.1: Introduction to Photochemistry
So far, we have come across one big rule of photon absorbance. In order to be absorbed, a photon's energy has to match an energy
difference within the compound that is absorbing it.

In the case of visible or ultraviolet light, the energy of a photon is roughly in the region that would be appropriate to promote an
electron to a higher energy level. Different wavelengths would be able to promote different electrons, depending on the energy
difference between an occupied electronic energy level and an unoccupied one. Other types of electromagnetic radiation would not
be able to promote an electron, but they would be coupled to other events. For example, absorption of infrared light is tied to
vibrational energy levels. Microwave radiation is tied to rotational energy levels in molecules. Thus, one reason a photon may or
may not be absorbed has to do with whether its energy corresponds to the available energy differences within the molecule or ion
that it encounters.

Franck-Condon: Electronic and Vibrational Coupling


Photons face other limitations. One of these is a moderate variation on our main rule. It is called the Frank Condon Principle.
According to this idea, when an electron is excited from its normal position, the ground state, to a higher energy level, the optimal
positions of atoms in the molecule may need to shift. Because electronic motion is much faster than nuclear motion, however, any
shifting of atoms needed to optimize positions as they should be in the excited state will have to wait until after the electron gets
excited. In that case, when the electron lands and the atoms aren't yet in their lowest energy positions for the excited state, the
molecule will find itself in an excited vibrational state as well as an excited electronic state.

15.1.1 https://chem.libretexts.org/@go/page/41463
That means the required energy for excitation doesn't just correspond to the difference in electronic energy levels; it is fine-tuned to
reach a vibrational energy level, which is quantized as well.
The Franck Condon Principle states that electronic transitions are vertical.
A vertical transition is one in which non of the nuclei move while the electron journeys from one state to another.
A vertical transition may begin in a vibrational ground state of an electronic ground state and end in a vibrational excited state
of an electronic excited state.

LaPorte: Orbital Symmetry


There are other restrictions on electronic excitation. Symmetry selection rules, for instance, state that the donor orbital (from which
the electron comes) and the acceptor orbital (to which the electron is promoted) must have different symmetry. The reasons for this
rule are based in the mathematics of quantum mechanics. What constitutes the same symmetry vs. different symmetry is a little
more complicated than we will get into here. Briefly, let's just look at one "symmetry element" and compare how two orbitals
might differ with respect to that element.
If an orbital is centrosymmetric, one can imagine each point on the orbital reflecting through the very centre of the orbital to a point
on the other side. At the end of the operation, the orbital appears unchanged. That means the orbital is symmetric with respect to a
centre of inversion..

If we do the same thing with a sigma antibonding orbital, things turn out differently.

In the drawing, the locations of the atoms are labelled A and B, but the symmetry of the orbital itself doesn't depend on that. If we
imagine sending each point on this orbital through the very centre to the other side, we arrive at a picture that looks exactly the
opposite of what we started with. These two orbitals have different symmetry. A transition from one to the other is allowed by
symmetry.

Problem RO2.1.
Decide whether each of the following orbitals is centrosymmetric.
a) an s orbital b) a p orbital c) a d orbital d) a π orbital e) a π* orbital

15.1.2 https://chem.libretexts.org/@go/page/41463
Problem RO2.2.
Decide whether each of the following transitions would be allowed by symmetry.
a) π → π* b) p → π* c) p → σ* d) d → d
Symmetry selection rules are in reality more like "strong suggestions." They depend on the symmetry of the molecule remaining
strictly static, but all kinds of distortions occur through molecular vibrations. Nevertheless, these rules influence the likelihood of a
given transition. The likelihood of a transition, similarly, has an influence upon the extinction coefficient, ε.

transition ε, extinction coefficient

π → π* 3,000 - 25,000 M-1 cm-1

p → π* 20 - 150 M-1 cm-1

p → σ* 100 - 7,000

d→d 5 - 400 M-1 cm-1

Spin State
Let's take a quick look at one last rule about electronic emissions. This rule concerns the spin of the excited electron, or more
correctly, the "spin state" of the excited species. The spin state describes the number of unpaired electrons in the molecule or ion.

number of unpaired electrons spin state

0 singlet

1 doublet

2 triplet

3 quartet

The rule says that in an electronic transition, the spin state of the molecule must be preserved. That means if there are no unpaired
electrons before the transition, then the excited species must also have no unpaired electrons. If there are two unpaired electrons
before the transition, the excited state must also have two unpaired electrons.

Contributors
Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

15.1: Introduction to Photochemistry is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
PC2. Rules of Electronic Excitation by Chris Schaller is licensed CC BY-NC 3.0.

15.1.3 https://chem.libretexts.org/@go/page/41463
15.2: Photosynthesis
Learning Objectives
1. Define the following:
photoautotrophenic a. oxyg
photoautotroph anoxygenicb.
c. photon
2. Name the two stages of photosynthesis.
3. State how all radiations in the electromagnetic spectrum travel.
4. State what constitutes visible light.
5. Define photon and describe what happens when photons of visible light energy strike certain atoms of pigments
during photosynthesis and how this can lead to the generation of ATP.
6. Describe the structure of a chloroplast and list the pigments it may contain.
7. Give the overall reaction for photosynthesis.
8. State the reactants and the products for photosynthesis and indicate which are oxidized and which are reduced.
1. Briefly describe the overall function of the light-dependent reactions in photosynthesis and state where in the
chloroplast they occur.
2. State the reactants and the products for the light-dependent reactions.
3. Describe an antenna complex and state the function of the reaction center.
4. Briefly describe the overall function of Photosystem II in the light-dependent reactions of photosynthesis.
5. Briefly describe how ATP is generated by chemiosmosis during the light-dependent reactions of
photosynthesis.
6. Briefly describe the overall function of Photosystem I in the light-dependent reactions of photosynthesis.
7. Compare noncyclic photophosphorylation and cyclic photophosphorylation in terms of Photosystems involved
and products produced.
1. Briefly describe the overall function of the light-independent reactions in photosynthesis and state where in the
chloroplast they occur.
1. State how the light-dependent and light-independent reactions are linked during photosynthesis.
2. State the reactants and the products for the light-independent reactions.
3. Briefly describe the following stages of the Calvin cycle:
CO2 fixation
production of G3P
regeneration of RuBP
4. State the significance of glyceraldehyde-3-phosphate (G3P) in the Calvin cycle.

Autotrophs are organisms that are able to synthesize organic molecules from inorganic materials. Photoautotrophs
absorb and convert light energy into the stored energy of chemical bonds in organic molecules through a process
called photosynthesis. Plants, algae, and bacteria known as cyanobacteria are known as oxygenic photoautotrophs
because they synthesize organic molecules from inorganic materials, convert light energy into chemical energy, use
water as an electron source, and generate oxygen as an end product of photosynthesis. Some bacteria, such as the
green and purple bacteria, are known as anoxygenic phototrophs. Unlike the oxygenic plants, algae, and
cyanobacteria, anoxygenic phototrophs do not use water as an electron source and, therefore, do not evolve oxygen
during photosynthesis. The electrons come from compounds such as hydrogen gas, hydrogen sulfide, and reduced
organic molecules. In this section on photosynthesis, we be concerned with the oxygenic phototrophs.
Photosynthesis is composed of two stages: the light-dependent reactions and the light-independent reactions. The
light-dependent reactions convert light energy into chemical energy, producing ATP and NADPH. The light-independent
reactions use the ATP and NADPH from the light-dependent reactions to reduce carbon dioxide and convert the energy

15.2.1 https://chem.libretexts.org/@go/page/41464
to the chemical bond energy in carbohydrates such as glucose. Before we get to these photosynthetic reactions
however, we need to understand a little about the electromagnetic spectrum and chloroplasts.

The Electromagnetic Spectrum


Visible light constitutes a very small portion of a spectrum of radiation known as the electromagnetic spectrum. All
radiations in the electromagnetic spectrum travel in waves and different portions of the spectrum are catagorized by
their wavelength. A wavelength is the distance from the peak of one wave to that of the next. At one end of the
spectrum are television and radio waves with longer wavelengths and low energy. At the other end of the spectrum are
gamma rays with a very short wavelength and a great deal of energy. Visible light is the range of wavelengths of the
electromagnetic spectrum that humans can see, a mixture of wavelengths ranging from 380 nanometers to 760
nanometers. It is this light that is used in photosynthesis.
Light and other types of radiation are composed of individual packets of energy called photons. The shorter the
wavelength of the radiation, the greater the energy per photon. As will be seen shortly, when photons of visible light
energy strike certain atoms of pigments during photosynthesis, that energy may push an electron from that atom to a
higher energy level where it can be picked up by an electron acceptor in an electron transport chain (see Fig. 15.2.1).
ATP can then be generated by chemiosmosis.

Fig. 15.2.1: Interaction Between a Photon and an Atom. When photons of visible light energy strike certain atoms of
pigments during photosynthesis, that energy may push an electron from that atom to a higher energy level where it can
be picked up by an electron acceptor in an electron transport chain.

Chloroplasts
In eukaryotic cells, photosynthesis takes place in organelles called chloroplasts.

Figure: Chloroplasts visible in the cells of Plagiomnium affine, the many-fruited thyme moss. ( GFDL , credit: Kristian Peters)
Like mitochondria, chloroplasts are surrounded by an inner and an outer membrane. The inner membrane encloses a
fluid-filled region called the stroma that contains enzymes for the light-independent reactions of photosynthesis.
Infolding of this inner membrane forms interconnected stacks of disk-like sacs called thylakoids, often arranged in
stacks called grana. The thylakoid membrane, which encloses a fluid-filled thylakoid interior space, contains
chlorophyll and other photosynthetic pigments as well as electron transport chains. The light-dependent reactions of
photosynthesis occur in the thylakoids. The outer membrane of the chloroplast encloses the intermembrane space
between the inner and outer chloroplast membranes (see Fig. 2).

15.2.2 https://chem.libretexts.org/@go/page/41464
The thylakoid membranes contain several pigments capable of absorbing visible light. Chlorophyll is the primary
pigment of photosynthesis. Chlorophyll absorbs light in the blue and red region of the visible light spectrum and reflects
green light. There are two major types of chlorophyll, chlorophyll a that initiates the light-dependent reactions of
photosynthesis, and chlorophyll b, an accessory pigment that also participates in photosynthesis. The thylakoid
membranes also contain other accessory pigments. Carotenoids are pigments that absorb blue and green light and
reflect yellow, orange, or red. Phycocyanins absorb green and yellow light and reflect blue or purple. These accessory
pigments absorb light energy and transfer it to chlorophyll.
Photosynthetic prokaryotic cells do not possess chloroplasts. Instead, thylakoid membranes are usually arranged
around the periphery of the bacterium as infoldings of the cytoplasmic membrane.

Photosynthesis
As mentioned above, photoautotrophs use sunlight as a source of energy and through the process of photosynthesis,
reduce carbon dioxide to form carbohydrates such as glucose. The radient energy is converted to the chemical bond
energy within glucose and other organic molecules. The overall reaction for photosynthesis is as follows:
6C O2 + 12 H2 Oin the presence of light and chlorophyll yieldsC6 H12 O6 + 6 O2 + 6H 2O (15.2.1)

Note that carbon dioxide (CO2) is reduced to produce glucose (C6H12O6 ), while water (H2O) is oxidized to produce
oxygen (O2). Photosynthesis is composed of two stages: the light-dependent reactions and the light independent
reactions. We will now look at the role of each in the next two sections.

Light-Dependent Reactions
The exergonic light-dependent reactions of photosynthesis convert light energy into chemical energy, producing ATP and NADPH.
These reactions occur in the thylakoids of the chloroplasts. The products of the light-dependent reactions, ATP and NADPH, are
both required for the endergonic light-independent reactions.
The light-dependent reactions can be summarized as follows:
12 H2O + 12 NADP+ + 18 ADP + 18 Pi + light and chlorophyll yields 6 O2 + 12 NADPH + 18 ATP
The light-dependent reactions involve two photosystems called Photosystem I and Photosystem II. These
photosystems include units called antenna complexes composed of chlorophyll molecules and accessory pigments
located in the thylakoid membrane. Photosystem I contain chlorophyll a molecules called P700 because they have
an absorption peak of 700 nanometers. Photosystem II contains chlorophyll a molecules referred to as P680
because they have an absorption peak of 680 nanometers.
Each antenna complex is able to trap light and transfer energy to a complex of chlorophyll molecules and proteins
called the reaction center (see Fig. 3). As photons are absorbed by chlorophyll and accessory pigments, that energy
is eventually transfered to the reaction center where, when absorbed by an excitable electron, moves it to a higher
energy level. Here the electron may be accepted by an electron acceptor molecule of an electron transport chain (see
Fig. 3) where the light energy is converted to chemical energy by chemiosmosis.
The most common light-dependent reaction in photosynthesis is called noncyclic photophosphorylation. Noncyclic
photophosphorylation involves both Photosystem I and Photosystem II and produces ATP and N ADP H . During noncyclic
photophosphorylation, the generation of ATP is coupled to a one-way flow of electrons from H O to N ADP . We will now look
2
+

at Photosystems I and II and their roles in noncyclic photophosphorylation.


1. As photons are absorbed by pigment molecules in the antenna complexes of Photosystem II, excited
electrons from the reaction center are picked up by the primary electron acceptor of the Photosystem II
electron transport chain. During this process, Photosystem II splits molecules of H2O into 1/2 O2, 2H+, and 2
electrons. These electrons continuously replace the electrons being lost by the P680 chlorophyll a molecules in the
reaction centers of the Photosystem II antenna complexes (see Fig. 4).
During this process, ATP is generated by the Photosystem II electron transport chain and chemiosmosis. According to
the chemiosmosis theory, as the electrons are transported down the electron transport chain, some of the energy
released is used to pump protons across the thylakoid membrane from the stroma of the chloroplast to the thylakoid

15.2.3 https://chem.libretexts.org/@go/page/41464
interior space producing a proton gradient or proton motive force. As the accumulating protons in the thylakoid
interior space pass back across the thylakoid membrane to the stroma through ATP synthetase complexes, this proton
motive force is used to generate ATP from ADP and Pi (see Fig. 4 and Fig. 5).

Flash animation illustrating the development of proton motive force as a result of chemiosmosis and ATP production by
ATPsynthase.

2. Meanwhile, photons are also being absorbed by pigment molecules in the antenna complex of Photosystem I
and excited electrons from the reaction center are picked up by the primary electron acceptor of the
Photosystem I electron transport chain. The electrons being lost by the P700 chlorophyll a molecules in the reaction
centers of Photosystem I are replaced by the electrons traveling down the Photosystem II electron transport chain. The
electrons transported down the Photosystem I electron transport chain combine with 2H+ from the
surrounding medium and NADP+ to produce NADPH + H+ (see Fig. 4).

McGraw-Hill Flash animation illustrating photosynthetic electran transport and ATP production by ATPsynthase.

Cyclic photophosphorylation occurs less commonly in plants than noncyclic photophosphorylation, most likely occurring
when there is too little NADP+ available. It is also seen in certain photosynthetic bacteria. Cyclic photophosphorylation
involves only Photosystem I and generates ATP but not NADPH. As the electrons from the reaction center of
Photosystem I are picked up by the electron transport chain, they are transported back to the reaction center
chlorophyll. As the electrons are transported down the electron transport chain, some of the energy released is used to
pump protons across the thylakoid membrane from the stroma of the chloroplast to the thylakoid interior space
producing a proton gradient or proton motive force. As the accumulating protons in the thylakoid interior space pass
back across the thylakoid membrane to the stroma through ATP synthetase complexes, this energy is used to generate
ATP from ADP and Pi (see Fig. 6).

McGraw-Hill Flash animation illustrating cyclic and non-cyclic photophosphorylation.

Light Independent Reactions


The endergonic light-independent reactions of photosynthesis use the ATP and NADPH synthesized during the
exergonic light-dependent reactions to provide the energy for the synthesis of glucose and other organic molecules
from inorganic carbon dioxide and water. This is done by "fixing" carbon atoms from CO2 to the carbon skeletons of
existing organic molecules. These reactions occur in the stroma of the chloroplasts.
The light-independent reactions can be summarized as follows:
12 NADPH + 18 ATP + 6 CO2 yields C6H12O6 (glucose) + 12 NADP+ + 18 ADP + 18 Pi + 6 H2O
Most plants use the Calvin (C3) cycle to fix carbon dioxide. C3 refers to the importance of 3-carbon molecules in the
cycle. Some plants, known as C4 plants and CAM plants, differ in their initial carbon fixation step.

1. The Calvin (C3) Cycle


There are three stages to the Calvin cycle: 1) CO2 fixation; 2) production of G3P;
and 3) regeneration of RuBP. We will now look at each stage.

stage 1: CO2 fixation


To begin the Calvin cycle, a molecule of CO2
reacts with a five-carbon compound called
ribulose bisphosphate (RuBP) producing an
unstable six-carbon intermediate which
immediately breaks down into two molecules of

15.2.4 https://chem.libretexts.org/@go/page/41464
the three-carbon compound phosphoglycerate
(PGA) (see Fig. 7). The carbon that was a part of
inorganic CO2 is now part of the carbon skeleton of
an organic molecule. The enzyme for this reaction
is ribulose bisphosphate carboxylase or Rubisco. A
total of six molecules of CO2 must be fixed this way
in order to produce one molecule of the six-carbon
sugar glucose.
stage 2: Production of G3P from PGA
The energy from ATP and the reducing power of
NADPH (both produced during the light-
dependent reactions) is now used to convert the
molecules of PGA to glyceraldehyde-3-
phosphate (G3P), another three-carbon
compound (see Fig. 7). For every six molecules of
CO2 that enter the Calvin cycle, two molecules of
G3P are produced. Most of the G3P produced
during the Calvin cycle - 10 of every 12 G3P
produced - are used to regenerate the RuBP in
order for the cycle to continue (see Fig. 7). Some of
the molecules of G3P, however, are used to
synthesize glucose and other organic
molecules. As can be seen in Fig. 7, two
molecules of the three-carbon G3P can be used to
synthesize one molecule of the six-carbon sugar
glucose. The G3P is also used to synthesize the
other organic molecules required by
photoautotrophs (see Fig. 8).
stage 3: Regeneration of RuBP from G3P
As mentioned in the previous step, most of the G3P
produced during the Calvin cycle - 10 of every 12
G3P produced - are used to regenerate the RuBP
so that the cycle may continue (see Fig. 7). Ten

15.2.5 https://chem.libretexts.org/@go/page/41464
molecules of the three-carbon compound G3P
eventually form six molecules of the four-
carbon compound ribulose phosphate (RP) (see
Fig. 7). Each molecule of RP then becomes
phosphorylated by the hydrolysis of ATP to
produce ribulose bisphosphate (RuBP), the
starting compound for the Calvin cycle (see Fig. 7).
Contributors and Attributions
Dr. Gary Kaiser (COMMUNITY COLLEGE OF BALTIMORE COUNTY, CATONSVILLE CAMPUS)

15.2: Photosynthesis is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

15.2.6 https://chem.libretexts.org/@go/page/41464
15.3: Vision
Vision is such an everyday occurrence that we seldom stop to think and wonder how we are able to see the objects that surround us.
Yet the vision process is a fascinating example of how light can produce molecular changes. The retina contain the molecules that
undergo a chemical change upon absorbing light, but it is the brain that actually makes sense of the visual information to create an
image.

Introduction
Light is one of the most important resources for civilization, it provides energy as it pass along by the sun. Light influence our
everyday live. Living organisms sense light from the environment by photoreceptors. Light, as waves carry energy, contains energy
by different wavelength. In vision, light is the stimulus input. Light energy goes into eyes stimulate photoreceptor in eyes.
However, as an energy wave, energy is passed on through light at different wavelength.

Light, as waves carry energy, contains energy by different wavelength. From long wavelength to short wavelength, energy increase.
400 nm to 700 nm is visible spectrum.
Light energy can convert chemical to other forms. Vitamin A, also known as retinol, anti-dry eye vitamins, is a required nutrition
for human health. The predecessor of vitamin A is present in the variety of plant carotene. Vitamin A is critical for vision because it
is needed by the retina of eye. Retinol can be convert to retinal, and retinal is a chemical necessary for rhodopsin. As light enters
the eye, the 11-cis-retinal is isomerized to the all-"trans" form.

15.3.1 https://chem.libretexts.org/@go/page/41465
Mechanism of Vision
The molecule cis-retinal can absorb light at a specific wavelength. When visible light hits the cis-retinal, the cis-retinal undergoes
an isomerization, or change in molecular arrangement, to all-trans-retinal. The new form of trans-retinal does not fit as well into the
protein, and so a series of geometry changes in the protein begins. The resulting complex is referred to a bathrhodopsin (there are
other intermediates in this process, but we'll ignore them for now).

The reaction above shows Lysine side-chain from the opsin react with 11-cis-retinal when stimulated. By removing the oxygen
atom form the retinal and two hydrogen atom form the free amino group of the lysine, the linkage show on the picture above is
formed, and it is called Schiff base.

Signal Transduction Pathway


As the protein changes its geometry, it initiates a cascade of biochemical reactions that results in changes in charge so that a large
potential difference builds up across the plasma membrane. This potential difference is passed along to an adjoining nerve cell as
an electrical impulse. The nerve cell carries this impulse to the brain, where the visual information is interpreted.
The light image is mapped on the surface of the retina by activating a series of light-sensitive cells known as rods and cones or
photoreceptors. The rods and cones convert the light into electrical impulses which are transmitted to the brain via nerve fibers. The
brain then determines, which nerve fibers carried the electrical impulse activate by light at certain photoreceptors, and then creates
an image.
The retina is lined with many millions of photoreceptor cells that consist of two types: 7 million cones provide color information
and sharpness of images, and 120 million rods are extremely sensitive detectors of white light to provide night vision. The tops of
the rods and cones contain a region filled with membrane-bound discs, which contain the molecule cis-retinal bound to a protein
called opsin. The resulting complex is called rhodopsin or "visual purple".

In human eyes, rod and cones react to light stimulation, and a series of chemical reactions happen in cells. These cells receive light,
and pass on signals to other receiver cells. This chain of process is class signal transduction pathway. Signal transduction pathway
is a mechanism that describe the ways cells react and response to stimulation.

References
1. Biochemistry, L. Stryer (W.H. Freeman and Co, San Francisco, 1975).
2. The Cambridge Guide to the Material World, Rodney Cotterill (Cambridge University Press, Cambridge, 1985)

Contributors
{{template.ContribOphardt()}}

15.3: Vision is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
Vision and Light is licensed CC BY-NC-SA 4.0.

15.3.2 https://chem.libretexts.org/@go/page/41465
15.4: Biological Effects of Radiation
 Learning Objectives
Describe the biological impact of ionizing radiation.
Define units for measuring radiation exposure.
Explain the operation of common tools for detecting radioactivity.
List common sources of radiation exposure in the US.

The increased use of radioisotopes has led to increased concerns over the effects of these materials on biological systems (such as
humans). All radioactive nuclides emit high-energy particles or electromagnetic waves. When this radiation encounters living cells,
it can cause heating, break chemical bonds, or ionize molecules. The most serious biological damage results when these radioactive
emissions fragment or ionize molecules. For example, alpha and beta particles emitted from nuclear decay reactions possess much
higher energies than ordinary chemical bond energies. When these particles strike and penetrate matter, they produce ions and
molecular fragments that are extremely reactive. The damage this does to biomolecules in living organisms can cause serious
malfunctions in normal cell processes, taxing the organism’s repair mechanisms and possibly causing illness or even death (Figure
15.4.1).

Figure 15.4.1 : Radiation can harm biological systems by damaging the DNA of cells. If this damage is not properly repaired, the
cells may divide in an uncontrolled manner and cause cancer.
A diagram is shown which has a white sphere followed by a right-facing arrow and a large sphere composed of many smaller white
and green spheres. The single sphere has impacted the larger sphere. A right-facing arrow leads from the larger sphere to a pair of
smaller spheres which are collections of the same white and green spheres. A starburst pattern lies between these two spheres and
has three right-facing arrows leading from it to two white spheres and a circle full of ten smaller, peach-colored circles with purple
dots in their centers. An arrow leads downward from this circle to a box that contains a helical shape with a starburst near its top
left side and is labeled “D N A damage.” A right-facing arrow leads from this circle to a second circle, with nine smaller, peach-
colored circles with purple dots in their centers and one fully purple small circle labeled “Cancer cell.” A right-facing arrow leads
to a final circle, this time full of the purple cells, that is labeled “Tumor.”

Ionizing vs. Nonionizing Radiation


There is a large difference in the magnitude of the biological effects of nonionizing radiation (for example, light and microwaves)
and ionizing radiation, emissions energetic enough to knock electrons out of molecules (for example, α and β particles, γ rays, X-
rays, and high-energy ultraviolet radiation) (Figure 15.4.2).
<div data-mt-source="1"

Figure 15.4.2 : Lower frequency, lower-energy electromagnetic radiation is nonionizing, and higher frequency, higher-energy
electromagnetic radiation is ionizing.
Energy absorbed from nonionizing radiation speeds up the movement of atoms and molecules, which is equivalent to heating the
sample. Although biological systems are sensitive to heat (as we might know from touching a hot stove or spending a day at the
beach in the sun), a large amount of nonionizing radiation is necessary before dangerous levels are reached. Ionizing radiation,
however, may cause much more severe damage by breaking bonds or removing electrons in biological molecules, disrupting their
structure and function. The damage can also be done indirectly, by first ionizing H2O (the most abundant molecule in living
organisms), which forms a H2O+ ion that reacts with water, forming a hydronium ion and a hydroxyl radical:

Figure 15.4.3 .

15.4.1 https://chem.libretexts.org/@go/page/41466
Figure 15.4.3 : Ionizing radiation can (a) directly damage a biomolecule by ionizing it or breaking its bonds, or (b) create an H2O+
ion, which reacts with H2O to form a hydroxyl radical, which in turn reacts with the biomolecule, causing damage indirectly.

Biological Effects of Exposure to Radiation


Radiation can harm either the whole body (somatic damage) or eggs and sperm (genetic damage). Its effects are more pronounced
in cells that reproduce rapidly, such as the stomach lining, hair follicles, bone marrow, and embryos. This is why patients
undergoing radiation therapy often feel nauseous or sick to their stomach, lose hair, have bone aches, and so on, and why particular
care must be taken when undergoing radiation therapy during pregnancy.
Different types of radiation have differing abilities to pass through material (Figure 15.4.4). A very thin barrier, such as a sheet or
two of paper, or the top layer of skin cells, usually stops alpha particles. Because of this, alpha particle sources are usually not
dangerous if outside the body, but are quite hazardous if ingested or inhaled (see the Chemistry in Everyday Life feature on Radon
Exposure). Beta particles will pass through a hand, or a thin layer of material like paper or wood, but are stopped by a thin layer of
metal. Gamma radiation is very penetrating and can pass through a thick layer of most materials. Some high-energy gamma
radiation is able to pass through a few feet of concrete. Certain dense, high atomic number elements (such as lead) can effectively
attenuate gamma radiation with thinner material and are used for shielding. The ability of various kinds of emissions to cause
ionization varies greatly, and some particles have almost no tendency to produce ionization. Alpha particles have about twice the
ionizing power of fast-moving neutrons, about 10 times that of β particles, and about 20 times that of γ rays and X-rays.

Figure 15.4.4 : The ability of different types of radiation to pass through material is shown. From least to most penetrating, they are
alpha < beta < neutron < gamma.
A diagram shows four particles in a vertical column on the left, followed by an upright sheet of paper, a person’s hand, an upright
sheet of metal, a glass of water, a thick block of concrete and an upright, thick piece of lead. The top particle listed is made up of
two white spheres and two green spheres that are labeled with positive signs and is labeled “Alpha.” A right-facing arrow leads
from this to the paper. The second particle is a red sphere labeled “Beta” and is followed by a right-facing arrow that passes
through the paper and stops at the hand. The third particle is a white sphere labeled “Neutron” and is followed by a right-facing
arrow that passes through the paper, hand and metal but is stopped at the glass of water. The fourth particle is shown by a squiggly
arrow and it passes through all of the substances but stops at the lead. Terms at the bottom read, from left to right, “Paper,”
“Metal,” “Water,” “Concrete” and “Lead.”
For many people, one of the largest sources of exposure to radiation is from radon gas (Rn-222). Radon-222 is an α emitter with a
half–life of 3.82 days. It is one of the products of the radioactive decay series of U-238, which is found in trace amounts in soil and
rocks. The radon gas that is produced slowly escapes from the ground and gradually seeps into homes and other structures above.

15.4.2 https://chem.libretexts.org/@go/page/41466
Since it is about eight times more dense than air, radon gas accumulates in basements and lower floors, and slowly diffuses
throughout buildings (Figure 15.4.5).

Figure 15.4.5 : Radon-222 seeps into houses and other buildings from rocks that contain uranium-238, a radon emitter. The radon
enters through cracks in concrete foundations and basement floors, stone or porous cinderblock foundations, and openings for
water and gas pipes.
A cut-away image of the side of a house and four layers of the ground it rests on is shown, as well as a second cut-away image of a
person’s head and chest cavity. The house is shown with a restroom on the second floor and a basement with a water heater as the
first floor. Green arrows lead from the lowest ground layer, labeled “radon in ground water,” from the third ground layer, labeled
“Bedrock” and “Fractured bedrock,” from the second layer, labeled “radon in well water,” and from the top layer, labeled “radon in
soil to the inside of the basement area. In the smaller image of the torso, a green arrow is shown to enter the person’s nasal passage
and travel to the lungs. This is labeled “Inhalation of radon decay products.” A small coiled, helical structure next to the torso is
labeled “alpha particle” on one section where it has a starburst pattern and “Radiation damage to D N A” on another segment.
Radon is found in buildings across the country, with amounts dependent on location. The average concentration of radon inside
houses in the US (1.25 pCi/L) is about three times the level found in outside air, and about one in six houses have radon levels high
enough that remediation efforts to reduce the radon concentration are recommended. Exposure to radon increases one’s risk of
getting cancer (especially lung cancer), and high radon levels can be as bad for health as smoking a carton of cigarettes a day.
Radon is the number one cause of lung cancer in nonsmokers and the second leading cause of lung cancer overall. Radon exposure
is believed to cause over 20,000 deaths in the US per year.

Measuring Radiation Exposure


Several different devices are used to detect and measure radiation, including Geiger counters, scintillation counters (scintillators),
and radiation dosimeters (Figure 15.4.6). Probably the best-known radiation instrument, the Geiger counter (also called the Geiger-
Müller counter) detects and measures radiation. Radiation causes the ionization of the gas in a Geiger-Müller tube. The rate of
ionization is proportional to the amount of radiation. A scintillation counter contains a scintillator—a material that emits light
(luminesces) when excited by ionizing radiation—and a sensor that converts the light into an electric signal. Radiation dosimeters
also measure ionizing radiation and are often used to determine personal radiation exposure. Commonly used types are electronic,
film badge, thermoluminescent, and quartz fiber dosimeters.

15.4.3 https://chem.libretexts.org/@go/page/41466
Figure 15.4.6 : Devices such as (a) Geiger counters, (b) scintillators, and (c) dosimeters can be used to measure radiation. (Credit c:
modification of work by “osaMu”/Wikimedia commons.)
Three photographs are shown and labeled “a,” “b” and “c.” Photo a shows a Geiger counter sitting on a table. It is made up of a
metal box with a read-out screen and a wire leading away from the box connected to a sensor wand. Photograph b shows a
collection of tall and short vertical tubes arranged in a grouping while photograph c shows a person’s hand holding a small machine
with a digital readout while standing on the edge of a roadway.
A variety of units are used to measure various aspects of radiation (Table 15.4.1). The SI unit for rate of radioactive decay is the
becquerel (Bq), with 1 Bq = 1 disintegration per second. The curie (Ci) and millicurie (mCi) are much larger units and are
frequently used in medicine (1 curie = 1 Ci = 3.7 × 10 disintegrations per second). The SI unit for measuring radiation dose is
10

the gray (Gy), with 1 Gy = 1 J of energy absorbed per kilogram of tissue. In medical applications, the radiation absorbed dose (rad)
is more often used (1 rad = 0.01 Gy; 1 rad results in the absorption of 0.01 J/kg of tissue). The SI unit measuring tissue damage
caused by radiation is the sievert (Sv). This takes into account both the energy and the biological effects of the type of radiation
involved in the radiation dose.
Table 15.4.1 : Units Used for Measuring Radiation
Measurement Purpose Unit Quantity Measured Description

amount of sample that undergoes 1


becquerel (Bq)
decay/second
activity of source radioactive decays or emissions
amount of sample that undergoes
curie (Ci) 10
3.7 × 10 decays/second

gray (Gy) 1 Gy = 1 J/kg tissue


absorbed dose energy absorbed per kg of tissue
radiation absorbed dose (rad) 1 rad = 0.01 J/kg tissue

sievert (Sv) Sv = RBE × Gy


biologically effective dose tissue damage
roentgen equivalent for man (rem) Rem = RBE × rad

The roentgen equivalent for man (rem) is the unit for radiation damage that is used most frequently in medicine (1 rem = 1 Sv).
Note that the tissue damage units (rem or Sv) includes the energy of the radiation dose (rad or Gy), along with a biological factor
referred to as the RBE (for relative biological effectiveness), that is an approximate measure of the relative damage done by the
radiation. These are related by:

number of rems = RBE × number of rads (15.4.1)

with RBE approximately 10 for α radiation, 2(+) for protons and neutrons, and 1 for β and γ radiation.

15.4.4 https://chem.libretexts.org/@go/page/41466
Figure 15.4.7 : Different units are used to measure the rate of emission from a radioactive source, the energy that is absorbed from
the source, and the amount of damage the absorbed radiation does. (CC by 4.0; OpenStax)
Two images are shown. The first, labeled “Rate of radioactive decay measured in becquerels or curies,” shows a red sphere with ten
red squiggly arrows facing away from it in a 360 degree circle. The second image shows the head and torso of a woman wearing
medical scrubs with a badge on her chest. The caption to the badge reads “Film badge or dosimeter measures tissue damage
exposure in rems or sieverts” while a phrase under this image states “Absorbed dose measured in grays or rads.”

 Example 15.4.1: Amount of Radiation

Cobalt-60 (t1/2 = 5.26 y) is used in cancer therapy since the γ rays it emits can be focused in small areas where the cancer is
located. A 5.00-g sample of Co-60 is available for cancer treatment.
a. What is its activity in Bq?
b. What is its activity in Ci?

Solution
The activity is given by:
ln 2 ln 2 g 60
Activity = λN = ( )N =( ) × 5.00 g = 0.659 of Co that decay
t1/2 5.26 y y

And to convert this to decays per second:


23
g y 1 day 1 h 1 mol 6.02 × 10 atoms 1 decay
0.659 × × × × × ×
y 365 day 24 hours 3, 600 s 59.9 g 1 mol 1 atom

decay
14
= 2.10 × 10
s

decay
(a) Since 1 Bq = 1
s
, the activity in Becquerel (Bq) is:

decay 1 Bq
14 14
2.10 × 10 ×( ) = 2.10 × 10 Bq
decay
s
1
s

decay
(b) Since 1 Ci = 3.7 × 10 11

s
, the activity in curie (Ci) is:

decay 1 Ci
14 2
2.10 × 10 ×( ) = 5.7 × 10 Ci
s 11 decay
3.7 × 10
s

15.4.5 https://chem.libretexts.org/@go/page/41466
 Exercise 15.4.1

Tritium is a radioactive isotope of hydrogen (t = 12.32 years) that has several uses, including self-powered lighting, in
1/2

which electrons emitted in tritium radioactive decay cause phosphorus to glow. Its nucleus contains one proton and two
neutrons, and the atomic mass of tritium is 3.016 amu. What is the activity of a sample containing 1.00mg of tritium (a) in Bq
and (b) in Ci?

Answer a
11
3.56 × 10 Bq

Answer b
0.962 Ci

Effects of Long-term Radiation Exposure on the Human Body


The effects of radiation depend on the type, energy, and location of the radiation source, and the length of exposure. As shown in
Figure 15.4.8, the average person is exposed to background radiation, including cosmic rays from the sun and radon from uranium
in the ground (see the Chemistry in Everyday Life feature on Radon Exposure); radiation from medical exposure, including CAT
scans, radioisotope tests, X-rays, and so on; and small amounts of radiation from other human activities, such as airplane flights
(which are bombarded by increased numbers of cosmic rays in the upper atmosphere), radioactivity from consumer products, and a
variety of radionuclides that enter our bodies when we breathe (for example, carbon-14) or through the food chain (for example,
potassium-40, strontium-90, and iodine-131).

15.4.6 https://chem.libretexts.org/@go/page/41466
Figure 15.4.8 : The total annual radiation exposure for a person in the US is about 620 mrem. The various sources and their relative
amounts are shown in this bar graph. (source: U.S. Nuclear Regulatory Commission).
A bar graph titled “Radiation Doses and Regulatory Limits, open parenthesis, in Millirems, close parenthesis” is shown. The y-axis
is labeled “Doses in Millirems” and has values from 0 to 5000 with a break between 1000 and 5000 to indicate a different scale to
the top of the graph. The y-axis is labeled corresponding to each bar. The first bar, measured to 5000 on the y-axis, is drawn in red
and is labeled “Annual Nuclear Worker Doses Limit, open parenthesis, N R C, close parenthesis.” The second bar, measured to
1000 on the y-axis, is drawn in blue and is labeled “Whole Body C T” while the third bar, measured to 620 on the y-axis, is drawn
in blue and is labeled “Average U period S period Annual Dose.” The fourth bar, measured to 310 on the y-axis, is drawn in blue
and is labeled “U period S period Natural Background Dose” while the fifth bar, measured to 100 on the y-axis and drawn in red
reads “Annual Public Dose Limit, open parenthesis, N R C, close parenthesis.” The sixth bar, measured to 40 on the y-axis, is
drawn in blue and is labeled “From Your Body” while the seventh bar, measured to 30 on the y-axis and drawn in blue reads
“Cosmic rays.” The eighth bar, measured to 4 on the y-axis, is drawn in blue and is labeled “Safe Drinking Water Limit, open
parenthesis, E P A, close parenthesis” while the ninth bar, measured to 2.5 on the y-axis and drawn in red reads “Trans Atlantic
Flight.” A legend on the graph shows that red means “Dose Limit From N R C dash licensed activity” while blue means “Radiation
Doses.”
A short-term, sudden dose of a large amount of radiation can cause a wide range of health effects, from changes in blood chemistry
to death. Short-term exposure to tens of rems of radiation will likely cause very noticeable symptoms or illness; a dose of about 500
rems is estimated to have a 50% probability of causing the death of the victim within 30 days of exposure. Exposure to radioactive
emissions has a cumulative effect on the body during a person’s lifetime, which is another reason why it is important to avoid any
unnecessary exposure to radiation. Health effects of short-term exposure to radiation are shown in Table 15.4.2.
Table 15.4.2 : Health Effects of Radiation
Exposure (rem) Health Effect Time to Onset (Without Treatment)

5–10 changes in blood chemistry —

50 nausea hours

55 fatigue —

70 vomiting —

75 hair loss 2–3 weeks

90 diarrhea —

100 hemorrhage —

400 possible death within 2 months

1000 destruction of intestinal lining —

internal bleeding —

15.4.7 https://chem.libretexts.org/@go/page/41466
Exposure (rem) Health Effect Time to Onset (Without Treatment)

death 1–2 weeks

2000 damage to central nervous system —

loss of consciousness minutes

death hours to days

It is impossible to avoid some exposure to ionizing radiation. We are constantly exposed to background radiation from a variety of
natural sources, including cosmic radiation, rocks, medical procedures, consumer products, and even our own atoms. We can
minimize our exposure by blocking or shielding the radiation, moving farther from the source, and limiting the time of exposure.

Summary
We are constantly exposed to radiation from a variety of naturally occurring and human-produced sources. This radiation can affect
living organisms. Ionizing radiation is the most harmful because it can ionize molecules or break chemical bonds, which damages
the molecule and causes malfunctions in cell processes. It can also create reactive hydroxyl radicals that damage biological
molecules and disrupt physiological processes. Radiation can cause somatic or genetic damage, and is most harmful to rapidly
reproducing cells. Types of radiation differ in their ability to penetrate material and damage tissue, with alpha particles the least
penetrating, but potentially most damaging, and gamma rays the most penetrating.
Various devices, including Geiger counters, scintillators, and dosimeters, are used to detect and measure radiation, and monitor
radiation exposure. We use several units to measure radiation: becquerels or curies for rates of radioactive decay; gray or rads for
energy absorbed; and rems or sieverts for biological effects of radiation. Exposure to radiation can cause a wide range of health
effects, from minor to severe, including death. We can minimize the effects of radiation by shielding with dense materials such as
lead, moving away from the source of radiation, and limiting time of exposure.

Footnotes
1. 1 Source: US Environmental Protection Agency

Glossary
becquerel (Bq)
SI unit for rate of radioactive decay; 1 Bq = 1 disintegration/s.

curie (Ci)
Larger unit for rate of radioactive decay frequently used in medicine; 1 Ci = 3.7 × 1010 disintegrations/s.

Geiger counter
Instrument that detects and measures radiation via the ionization produced in a Geiger-Müller tube.

gray (Gy)
SI unit for measuring radiation dose; 1 Gy = 1 J absorbed/kg tissue.

ionizing radiation
Radiation that can cause a molecule to lose an electron and form an ion.

millicurie (mCi)
Larger unit for rate of radioactive decay frequently used in medicine; 1 Ci = 3.7 × 1010 disintegrations/s.

nonionizing radiation
Radiation that speeds up the movement of atoms and molecules; it is equivalent to heating a sample, but is not energetic enough
to cause the ionization of molecules.

radiation absorbed dose (rad)


SI unit for measuring radiation dose, frequently used in medical applications; 1 rad = 0.01 Gy.

15.4.8 https://chem.libretexts.org/@go/page/41466
radiation dosimeter
Device that measures ionizing radiation and is used to determine personal radiation exposure.

relative biological effectiveness (RBE)


Measure of the relative damage done by radiation.

roentgen equivalent man (rem)


Unit for radiation damage, frequently used in medicine; 1 rem = 1 Sv.

scintillation counter
Instrument that uses a scintillator—a material that emits light when excited by ionizing radiation—to detect and measure
radiation.

sievert (Sv)
SI unit measuring tissue damage caused by radiation; takes energy and biological effects of radiation into account.

15.4: Biological Effects of Radiation is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
21.6: Biological Effects of Radiation by OpenStax is licensed CC BY 4.0. Original source: https://openstax.org/details/books/chemistry-2e.

15.4.9 https://chem.libretexts.org/@go/page/41466
CHAPTER OVERVIEW

16: Macromolecules
Macromolecules are a very large molecules, such as protein, commonly created by polymerization of smaller subunits (monomers).
They are typically composed of thousands or more atoms.
16.1: Size, Shape, and Molar Mass of Macromolecules
16.2: Structure of Synthetic Polymers
16.3: Structure of Proteins and DNA
16.4: Protein Stability

16: Macromolecules is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1
16.1: Size, Shape, and Molar Mass of Macromolecules
16.1: Size, Shape, and Molar Mass of Macromolecules is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.

16.1.1 https://chem.libretexts.org/@go/page/41467
16.2: Structure of Synthetic Polymers
16.2: Structure of Synthetic Polymers is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

16.2.1 https://chem.libretexts.org/@go/page/41468
16.3: Structure of Proteins and DNA
16.3: Structure of Proteins and DNA is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

16.3.1 https://chem.libretexts.org/@go/page/41469
16.4: Protein Stability
16.4: Protein Stability is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

16.4.1 https://chem.libretexts.org/@go/page/41470
Index
A F Phase Equilibria
absolute entropy Flash Photolysis 4.9: Phase Equilibria
4.4: The Third Law of Thermodynamics 9.10: Fast Reactions in Solution phosphorescence
fluorescence 14.7: Fluorescence and Phosphorescence
B 14.7: Fluorescence and Phosphorescence photoelectric effect
11.3: The Photoelectric Effect
blackbody radiation
11.2: Planck's Quantum Theory G
bond enthalpies Gibbs Free Energy Q
3.7: Bond Energies and Enthalpies 4.6: Gibbs Energy Quenched Flow
Graham's law 9.10: Fast Reactions in Solution
C 2.9: Graham's Laws of Diffusion and Effusion
catalytic efficiency R
10.2: The Equations of Enzyme Kinetics H real gases
colligative property hill equation 2.4: Real Gases
5.6: Colligative Properties 10.6: Allosteric Interactions Reversible Reactions (Kinetics)
collision theory 9.4: More Complex Reactions
9.7: Theories of Reaction Rates I
competitive inhibition inhibitor S
10.5: Enzyme Inhibition 10.5: Enzyme Inhibition specificity constant
Consecutive Reactions (Kinetics) Ionic Activity 10.2: The Equations of Enzyme Kinetics
9.4: More Complex Reactions 5.8: Ionic Activity stopped flow
Continuous Flow ionic strength 9.10: Fast Reactions in Solution
9.10: Fast Reactions in Solution 5.8: Ionic Activity
cooperativity T
10.6: Allosteric Interactions K Temperature Dependence of Gibbs
KIE Energy
D 9.8: Isotope Effects in Chemical Reactions 4.8: Dependence of Gibbs Energy on Temperature
de Broglie Wavelength Kirchoff's Law and Pressure
11.5: de Broglie's Postulate 3.6: Thermochemistry Temperature Jump
denaturation 9.10: Fast Reactions in Solution
10.8: The Effect of Temperature on Enzyme Kinetics M The Hill Equation
diffusion mean ionic activity coefficient 10.6: Allosteric Interactions
2.9: Graham's Laws of Diffusion and Effusion 5.8: Ionic Activity
Third Law of Thermodynamics
double reciprocal plot Michaelis constant 4.4: The Third Law of Thermodynamics
10.2: The Equations of Enzyme Kinetics 10.2: The Equations of Enzyme Kinetics
transition state theory
9.7: Theories of Reaction Rates
molecularity
E 9.3: Molecularity of a Reaction
Turnover number
effusion 10.2: The Equations of Enzyme Kinetics
2.9: Graham's Laws of Diffusion and Effusion N
Electron Paramagnetic Resonance U
Noncompetitive Inhibition
14.6: Electron Spin Resonance 10.5: Enzyme Inhibition
Uncompetitive Inhibition
Electron Spin Resonance 10.5: Enzyme Inhibition
14.6: Electron Spin Resonance P
EPR V
Parallel Reactions (Kinetics)
14.6: Electron Spin Resonance
9.4: More Complex Reactions
vision
ESR particle in a box 15.3: Vision
14.6: Electron Spin Resonance
11.8: Particle in a One-Dimensional Box

1 https://chem.libretexts.org/@go/page/212315
Glossary
Sample Word 1 | Sample Definition 1

1 https://chem.libretexts.org/@go/page/279592
Detailed Licensing
Overview
Title: Map: Physical Chemistry for the Biosciences (Chang)
Webpages: 160
Applicable Restrictions: Noncommercial
All licenses found:
CC BY-NC-SA 4.0: 94.4% (151 pages)
Undeclared: 4.4% (7 pages)
CC BY 4.0: 0.6% (1 page)
CC BY-SA 4.0: 0.6% (1 page)

By Page
Map: Physical Chemistry for the Biosciences (Chang) - CC 3.2: The First Law of Thermodynamics - CC BY-NC-
BY-NC-SA 4.0 SA 4.0
Front Matter - CC BY-NC-SA 4.0 3.3: Heat Capacities - CC BY-NC-SA 4.0
TitlePage - Undeclared 3.4: Gas Expansion - CC BY-NC-SA 4.0
InfoPage - Undeclared 3.5: Calorimetry - CC BY-NC-SA 4.0
Table of Contents - Undeclared 3.6: Thermochemistry - CC BY-NC-SA 4.0
Licensing - Undeclared 3.7: Bond Energies and Enthalpies - CC BY-NC-SA
4.0
1: Introduction to Physical Chemistry - CC BY-NC-SA
3.E: Exercises - CC BY-NC-SA 4.0
4.0
4: The Second Law of Thermodynamics - CC BY-NC-SA
1.1: Nature of Physical Chemistry - CC BY-NC-SA
4.0
4.0
4.1: Spontaneous Processes - CC BY-NC-SA 4.0
1.2: Units - CC BY-NC-SA 4.0
4.2: Entropy - CC BY-NC-SA 4.0
1.3: Atomic Mass, Molecular Mass, and the Chemical
4.3: The Second Law of Thermodynamics - CC BY-
Mole - CC BY-NC-SA 4.0
NC-SA 4.0
2: Properties of Gases - CC BY-NC-SA 4.0
4.4: The Third Law of Thermodynamics - CC BY 4.0
2.1: Some Definitions - CC BY-NC-SA 4.0 4.5: Evaluating Entropy and Entropy Changes - CC
2.2: An Operational Definition of Temperature - CC BY-NC-SA 4.0
BY-NC-SA 4.0 4.6: Gibbs Energy - CC BY-NC-SA 4.0
2.3: Ideal Gases - CC BY-NC-SA 4.0 4.7: Standard Molar Gibbs Energy of Formation - CC
2.4: Real Gases - CC BY-NC-SA 4.0 BY-NC-SA 4.0
2.5: Condensation of Gases and the Critical State - 4.8: Dependence of Gibbs Energy on Temperature
CC BY-NC-SA 4.0 and Pressure - CC BY-NC-SA 4.0
2.6: Kinetic Theory of Gases - CC BY-NC-SA 4.0 4.9: Phase Equilibria - CC BY-NC-SA 4.0
2.7: The Maxwell Distribution Laws - CC BY-NC-SA 4.10: Thermodynamics of Rubber Elasticity - CC BY-
4.0 NC-SA 4.0
2.8: Molecular Collisions and the Mean Free Path - 4.E: Exercises - CC BY-NC-SA 4.0
CC BY-NC-SA 4.0
5: Solutions - CC BY-NC-SA 4.0
2.9: Graham's Laws of Diffusion and Effusion - CC
BY-NC-SA 4.0 5.1: Concentration Units - CC BY-NC-SA 4.0
2.E: Properties of Gases (Exercises) - CC BY-NC-SA 5.2: Partial Molar Quantities - CC BY-NC-SA 4.0
4.0 5.3: The Thermodynamics of Mixing - CC BY-NC-SA
4.0
3: The First Law of Thermodynamics - CC BY-NC-SA
5.4: Binary Mixtures of Volatile Liquids - CC BY-NC-
4.0
SA 4.0
3.1: Work and Heat - CC BY-NC-SA 4.0 5.5: Real Solutions - CC BY-NC-SA 4.0
5.6: Colligative Properties - CC BY-NC-SA 4.0

1 https://chem.libretexts.org/@go/page/417258
5.7: Electrolyte Solutions - CC BY-NC-SA 4.0 9.6: Potential Energy Surfaces - CC BY-NC-SA 4.0
5.8: Ionic Activity - CC BY-NC-SA 4.0 9.7: Theories of Reaction Rates - CC BY-NC-SA 4.0
5.9: Colligative Properties of Electrolyte Solutions - 9.8: Isotope Effects in Chemical Reactions - CC BY-
CC BY-NC-SA 4.0 NC-SA 4.0
5.10: Biological Membranes - CC BY-NC-SA 4.0 9.9: Reactions in Solution - CC BY-NC-SA 4.0
5.E: Solutions (Exercises) - CC BY-NC-SA 4.0 9.10: Fast Reactions in Solution - CC BY-NC-SA 4.0
6: Chemical Equilibrium - CC BY-NC-SA 4.0 9.11: Oscillating Reactions - CC BY-NC-SA 4.0
6.1: Chemical Equilibrium in Gaseous Systems - CC 9.E: Chemical Kinetics (Exercises) - CC BY-NC-SA
BY-NC-SA 4.0 4.0
6.2: Reactions in Solutions - CC BY-NC-SA 4.0 10: Enzyme Kinetics - CC BY-NC-SA 4.0
6.3: Heterogeneous Equilibria - CC BY-NC-SA 4.0 10.1: General Principles of Catalysis - CC BY-NC-SA
6.4: The Influence of Temperature, Pressure, and 4.0
Catalysts on the Equilibrium Constant - CC BY-NC- 10.2: The Equations of Enzyme Kinetics - CC BY-
SA 4.0 NC-SA 4.0
6.5: Binding of Ligands and Metal Ions to 10.3: Chymotrypsin- A Case Study - CC BY-NC-SA
Macromolecules - CC BY-NC-SA 4.0 4.0
6.6: Bioenergetics - CC BY-NC-SA 4.0 10.4: Multisubstrate Systems - CC BY-NC-SA 4.0
6.E: Chemical Equilibrium (Exercises) - CC BY-NC- 10.5: Enzyme Inhibition - CC BY-NC-SA 4.0
SA 4.0 10.6: Allosteric Interactions - CC BY-NC-SA 4.0
7: Electrochemistry - CC BY-NC-SA 4.0 10.7: The Effect of pH on Enzyme Kinetics - CC BY-
NC-SA 4.0
7.1: Electrochemical Cells - CC BY-NC-SA 4.0
10.8: The Effect of Temperature on Enzyme Kinetics
7.2: Single Electrode Potentials - CC BY-NC-SA 4.0
- CC BY-SA 4.0
7.3: Thermodynamics of Electrochemical Cells - CC
10.E: Exercises - CC BY-NC-SA 4.0
BY-NC-SA 4.0
7.4: Types of Electrochemical Cells - CC BY-NC-SA 11: Quantum Mechanics and Atomic Structure - CC BY-
4.0 NC-SA 4.0
7.5: Applications of EMF Measurements - CC BY- 11.1: The Wave Theory of Light - CC BY-NC-SA 4.0
NC-SA 4.0 11.2: Planck's Quantum Theory - CC BY-NC-SA 4.0
7.6: Biological Oxidation - CC BY-NC-SA 4.0 11.3: The Photoelectric Effect - CC BY-NC-SA 4.0
7.7: Membrane Potential - CC BY-NC-SA 4.0 11.4: Bohr's Theory of the Hydrogen Emission
7.E: Electrochemistry (Exercises) - CC BY-NC-SA 4.0 Spectrum - CC BY-NC-SA 4.0
8: Acids and Bases - CC BY-NC-SA 4.0 11.5: de Broglie's Postulate - CC BY-NC-SA 4.0
8.1: Definitions of Acids and Bases - CC BY-NC-SA 11.6: The Heisenberg Uncertainty Principle - CC BY-
4.0 NC-SA 4.0
8.2: Acid-Base Properties of Water - CC BY-NC-SA 11.7: The Schrödinger Wave Equation - CC BY-NC-
4.0 SA 4.0
8.3: Dissociation of Acids and Bases - CC BY-NC-SA 11.8: Particle in a One-Dimensional Box - CC BY-
4.0 NC-SA 4.0
8.4: Diprotic and Polyprotic Acids - CC BY-NC-SA 11.9: Quantum-Mechanical Tunneling - CC BY-NC-
4.0 SA 4.0
8.5: Buffer Solutions - CC BY-NC-SA 4.0 11.10: The Schrödinger Wave Equation for the
8.6: Acid-Base Titrations - CC BY-NC-SA 4.0 Hydrogen Atom - CC BY-NC-SA 4.0
8.7: Amino Acids - CC BY-NC-SA 4.0 11.11: Many-Electron Atoms and the Periodic Table -
8.8: Maintaining the pH of Blood - CC BY-NC-SA 4.0 CC BY-NC-SA 4.0
11.E: Quantum Mechanics and Atomic Structure
9: Chemical Kinetics - CC BY-NC-SA 4.0
(Exercises) - CC BY-NC-SA 4.0
9.1: Reaction Rates - CC BY-NC-SA 4.0
12: The Chemical Bond - CC BY-NC-SA 4.0
9.2: Reaction Order - CC BY-NC-SA 4.0
12.1: Lewis Structures - CC BY-NC-SA 4.0
9.3: Molecularity of a Reaction - CC BY-NC-SA 4.0
12.2: Valence Bond Theory - CC BY-NC-SA 4.0
9.4: More Complex Reactions - CC BY-NC-SA 4.0
12.3: Hybridization of Atomic Orbitals - CC BY-NC-
9.5: The Effect of Temperature on Reaction Rates -
SA 4.0
CC BY-NC-SA 4.0

2 https://chem.libretexts.org/@go/page/417258
12.4: Electronegativity and Dipole Moment - CC BY- 14.6: Electron Spin Resonance - CC BY-NC-SA 4.0
NC-SA 4.0 14.7: Fluorescence and Phosphorescence - CC BY-
12.5: Molecular Orbital Theory - CC BY-NC-SA 4.0 NC-SA 4.0
12.6: Diatomic Molecules - CC BY-NC-SA 4.0 14.8: Lasers - CC BY-NC-SA 4.0
12.7: Resonance and Electron Delocalization - CC 14.9: Optical Rotatory Dispersion and Circular
BY-NC-SA 4.0 Dichroism - CC BY-NC-SA 4.0
12.8: Coordination Compounds - CC BY-NC-SA 4.0 14.E: Spectroscopy (Exercises) - CC BY-NC-SA 4.0
12.9: Coordination Compounds in Biological Systems 15: Photochemistry and Photobiology - CC BY-NC-SA
- CC BY-NC-SA 4.0 4.0
12.E: The Chemical Bond (Exercises) - CC BY-NC- 15.1: Introduction to Photochemistry - CC BY-NC-SA
SA 4.0 4.0
13: Intermolecular Forces - CC BY-NC-SA 4.0 15.2: Photosynthesis - CC BY-NC-SA 4.0
13.1: Intermolecular Interactions - CC BY-NC-SA 4.0 15.3: Vision - CC BY-NC-SA 4.0
13.2: The Ionic Bond - CC BY-NC-SA 4.0 15.4: Biological Effects of Radiation - CC BY-NC-SA
13.3: Types of Intermolecular Forces - CC BY-NC-SA 4.0
4.0 16: Macromolecules - CC BY-NC-SA 4.0
13.4: Hydrogen Bonding - CC BY-NC-SA 4.0
16.1: Size, Shape, and Molar Mass of
13.5: The Structure and Properties of Water - CC BY-
Macromolecules - CC BY-NC-SA 4.0
NC-SA 4.0
16.2: Structure of Synthetic Polymers - CC BY-NC-
13.6: Hydrophobic Interaction - CC BY-NC-SA 4.0
SA 4.0
13.E: Intermolecular Forces (Exercises) - CC BY-NC-
16.3: Structure of Proteins and DNA - CC BY-NC-SA
SA 4.0
4.0
14: Spectroscopy - CC BY-NC-SA 4.0 16.4: Protein Stability - CC BY-NC-SA 4.0
14.1: Vocabulary - CC BY-NC-SA 4.0 Back Matter - CC BY-NC-SA 4.0
14.2: Microwave Spectroscopy - CC BY-NC-SA 4.0 Index - Undeclared
14.3: Infrared Spectroscopy - CC BY-NC-SA 4.0 Glossary - Undeclared
14.4: Electronic Spectroscopy - CC BY-NC-SA 4.0 Detailed Licensing - Undeclared
14.5: Nuclear Magnetic Resonance - CC BY-NC-SA
4.0

3 https://chem.libretexts.org/@go/page/417258

You might also like