General Chemistry 1 Lessons
General Chemistry 1 Lessons
General Chemistry 1 Lessons
Matter
Lesson objectives:
1. Define matter. 6. Classify the pure substances.
2. Identify the three phases/states of matter. 7. Discuss elements.
3. Enumerate the properties of matter. 8. Discuss compounds.
4. Explains the changes of matter. 9. Enumerate three kinds of
mixtures.
5. Contrast two general classes of matter. 10. Describe separation techniques
for mixtures and compounds
Discussion:
Matter is everything that occupy space and has mass. It is composed of tiny particles
called atoms. Atoms are the smallest unit of an element. Combination of two elements
in forming compounds, the smallest particles of two or more elemental unit is called
molecule.
I. Composition of Matter
A. Three Phases/States of Matter
Examples: water, soda, juices, tea, etc., Microscopic view of liquid particles.
Examples: air, steam, carbon dioxide, etc., Microscopic view of gas particles.
(1) Qualitative Observation – matter can be observed based on some of its quality.
Qualitative observation can be described as color, shape, texture, shininess, and
physical state.
(2) Quantitative Observation – matter is observed based on numerical values.
Qualitative observation is either numbers or measurements. It should be carefully made
and reported since some of them involved very large or very small numbers. They are
often expressed using scientific or exponential notations.
(1) Intrinsic properties – this is also called as intensive properties. These are properties
that are inherent characteristics of matter and depend on the kind of material. Intrinsic
properties are permanent properties of matter. Some of these properties are taste,
odor, color, transparency, solubility, miscibility, boiling point, melting point, viscosity,
hardness, malleability, ductility, conductivity.
(2) Extrinsic properties – this is also called as extensive properties. These properties are
the outside appearance of matter. It depends on the amount of matter present. Some
of these properties are size, length, mass, and weight.
Examples: odor, taste, hardness, mass, volume, density, magnetism, conductivity and
temperature.
*Changing liquid water into water vapor. It could represent by the symbol: H2O (l) →
H2O (g)
(1) Pure Substance – is a matter in which all samples have identical and fixed
composition and properties. It cannot be separated by physical means.
(a) Homogeneous Mixture - has uniform composition throughout and often called
solution. It is in single phase only.
Examples: antibiotic drugs, shampoo, dust in air, salad, water and oil
I. Atoms
Democritus – a Greek philosopher believed that there was a limit to how far matter
could be broken down into smaller pieces. He was the one who name this smallest
particle as ‘atomos’ which means unbreakable.
John Dalton – a math teacher who offered the first convincing argument for the existing
of atoms based on well – established experimental results which he published in a book
in 1808. He proposed the first atomic theory.
Laws of Matter
determine the percent by mass of hydrogen, for example, simply divide the mass of
hydrogen by the sample mass and multiply by 100%. For each sample you will obtain
the same result: 11.59% H.
John Dalton, expanding on Proust’s work, found out that 100 g of tin can react
with either 13.5 g or 27 g of oxygen. He noticed that the amount of oxygen used is in a
ratio of 1:2. This result, along with countless others, led Dalton to propose the law of
multiple proportions.
The law of multiple proportions states that if two elements can combine to
form more than one compound, the masses of one of the elements that combine with
the mass of the other element are in fixed ratios of whole numbers.
Atoms are not the small particles Dalton envisioned. They consist of even smaller
particles that together make up all atoms.
Ernest Rutherford – he experimented on the gold foil that led to the nuclear model of
atom. Also, he discovered nucleus through this experiment along with the positively
charged atom called proton in 1907.
+ 0 -
Element Atomic Mass
Number Number p n e
16
8 16 8 8 8
O
24
12 24 12 12 12
Mg
Isotopes - have the same number of protons (atomic number) but different atomic
masses.
Examples:
11 H Hydrogen
14
C
6
2
1 H Deuterium
12 3
C
6 1
H Tritium
Classifications of Ions
(1) Cation – a positively charged ion that contains fewer electrons than the number of
protons in the nucleus.
(2) Anion – a negatively charged ion that contains more electrons than the number of
protons in the nucleus.
Naming of Compounds
A. Chemical Formula - formula use to express the composition of molecules and ionic
compounds in terms of chemical symbols.
Example: Water – H2 O
Molecular Formula - gives the exact number of each element present in smallest unit of
substance
Example:
N2 O4 - Dinitrogen tetroxide
H2 O - Dihydrogen oxide (water)
Empirical Formula - shows which elements are present and in what ratio they exist in
the substance.
Example:
Example:
Naming Ionic Compound - in ionic compounds that are formed from two elements, the
cation is named first, then the anion next. The name of the anion is derived from the
first part of the name of the element together with the suffix “-ide”
Formula Writing for Ionic Compounds - for an ionic compound to be neutral, the sum of
positive charges and negative charges must be equal to zero.
Naming Molecular Compound - first element in the formula is named first, then the
second element with the suffix “-ide” is named next.
Formula Writing for Molecular Compound – writing the formula for two non – metal
compounds, the less electronegative element is written first (name of element farther to
the left in the periodic table is written first). Followed by the next element using the
greek prefixes:
PREFIX MEANING
Mono 1
di- 2
tri- 3
tetra- 4
penta- 5
hexa- 6
hepta- 7
octa- 8
nona- 9
deca- 10
Example:
Carbon oxide - CO
Phosphorus chloride - PCl
Naming Metals more One Cations - the charge of cation is written in Roman Numerals
and enclose in parentheses.
Example: Manganese (Mn) can form three cations: Mn2+, Mn3+, Mn4+. These
corresponding cations are manganese (II), manganese (III), manganese (IV).
Mn O is named Manganese (II) oxide
Mn2O3 is named Manganese (III) oxide
Common metals with varying oxidation numbers
Example:
1. NaHCO3-1 4. ScCn
Sodium bicarbonate Scandium cyanide
2.H2PO4-3 5. NH4S
Dihydrogen phosphate Ammonium sulfide
3. HCO3-2 6. CuO2
Hydrogen carbonate Cupric peroxide
Lesson 3
Stoichiometry
Take note that the representative particle for glucose (C6H12O6) is a molecule.
The number of molecules of glucose is composed by multiplying 5.50 moles of glucose
by the conversion factor, which is the Avogadro’s number.
Molar Mass
The molar mass of a compound is the mass of 1 mole of the compound in
grams.
The molar masses of water, salt (NaCl), and table sugar (C12H22O11) are calculated as
follows:
Mass % of Element = (Total mass of that Element per mole of compound) x 100%
Molar mass of the Compound
For example, the mass percents of hydrogen, nitrogen, and oxygen in
ammonium nitrate, NH4NO3 are calculated as follows:
The empirical formula is a formula that shows the simplest mole ratio of
elements in the compound.
The molecular formula of a compound is the true formula that indicates the
actual number of atoms of each type in a molecule (or the number of moles of
each type of element in 1 mole of the compound). But it does not tell how atoms
are connected in the molecule.
The structural formula shows how atoms are bonded in the actual molecule.
Solution: The masses of carbon and hydrogen in the sample are calculated as follows:
If a 100-g sample of the compound, there will be 34.1 g C, 6.7 g H, 13.3 g N, and 45.9
g O.
Converting the mass of each element into its moles, we will obtain
Mole of C = (34.1 g C)/(12.0 g/mol) = 2.84 mol
Mole of H = (6.7 g)/(1.008 g/mol) = 6.6 mol
Mole of N = (13.3 g N)/(14.0 g/mol) = 0.950 mol N
Mole of O = (45.9 g O)/(16.0 g/mol) = 2.87 mol O
Dividing throughout by the mole of N (the smallest mole in the compound), we obtain
the following mole ratio: 2.99 C-to-6.9 H-to-1 N-to-3.02 O, which simplifies to
3C:7H:1N:3O;
This mole ratio yields the empirical formula C3H7NO3.
Deriving Empirical and Molecular Formula from percent Composition and Molar Mass
Example:
A compound with molar mass 180 g/mol consists of 40.0% carbon, 6.7%
hydrogen and 53.3% oxygen. Determine the empirical and molecular formula of the
compound.
Solution: Assuming a 100-g sample, the masses of carbon hydrogen and oxygen in the
sample are: 40.0 g, 6.7 g, and 53.3 g, respectively.
Moles of C = (40.0 g C)/(12.0 g/mol) = 3.33 mol C;
Moles of H = (6.7 g H)/(1.008 g/mol) = 6.6 mol H;
Moles of O = (53.3 g O)/(16.0 g/mol) = 3.33 mol O.
Dividing by the smallest mole yields a mole ratio: 1C:2H:1O, which yields an empirical
formula of CH2O.
Reaction: Ammonia gas reacts with oxygen gas to form nitrogen monoxide and water;
Unbalanced equation: NH3(g) + O2(g) NO(g) + H2O(g)
Balanced equation: 4NH3(g) + 5 O2(g) 4NO(g) + 6H2O(g);
When substances react, old bonds are broken and new ones formed; atoms are
reorganized in different ways. The equation must show that the total number of each
type of atoms stays the same after the reaction.
A chemical equation provides two types of information: the nature of the reactants and
products and the relative numbers of each type of atoms.
An equation must also indicate the physical states of the substances: (s) for solid, (l) for
liquid, (g) for gas, and (aq) for aqueous solution.
For example, the combustion of ethanol (C2H5OH) to form carbon dioxide and water
vapor is represented by the following equation (not balanced):
Since carbon and hydrogen appear only once on each side of the equation, we balance
these elements first by introducing coefficients 2 and 3 in front of CO2 and H2O,
respectively.
To remove this fraction, we multiply the entire equation by 2, which yield the proper
equation:
Exercise-3: Balance the following equations using the simplest integer coefficient
possible:
1. C6H6(l) + O2(g) CO2(g) + H2O(g);
At molecular quantity, this equation implies that 1 N2 molecule will react with 3 H2
molecules to form 2 NH3 molecules. In molar quantity, it implies 1 mole of N2 react with
3 moles of H2 to yield 2 moles of NH3. If the amounts of nitrogen and hydrogen are
given, we can calculate how much ammonia will be formed at the end of the reaction.
Likewise, knowing how much NH3 we want to produce, we can calculate the amount of
each reactant that will be required.
For example, if we want to know how many moles of H2 are needed to react with
5 moles of N2 and how many moles of NH3 will be formed, we can perform the
calculation as follows:
The factor (3 mol H2/1 mol N2) and (2 mol NH3/1 mol N2) are called the
stoichiometric ratios, or stoichiometric factors.
2. To calculate the mol of N2 needed and mol of NH3 expected from the given moles of
H2:
a) ? mol N2 = Given mol of H2 x (1 mol N2/3 mol H2)
b) ? mol NH3 = Given mol of H2 x (2 mol NH3/3 mole H2)
3. To calculate the mol of H2 and N2 needed, respectively, from the given moles of
NH3::
a) ? mol N2 = Given mol NH3 x (1 mol N2/2 mol NH3)
b) ? mol H2 = Given mol NH3 x (3 mol H2/2 mol NH3)
Exercise-4:
1. According to the equation:
2 C8H18(l) + 25 O2(g) 16 CO2(g) + 18 H2O(g),
(a) how many moles of CO2 are produced when 5.5 moles of octane (C8H18) are
completely burned in air? (b) How many moles of water are also formed? (c)
How many moles of oxygen are reacted by 5.5 moles of octane?
(Answer: (a) 44 moles of CO2; (b) 50. moles of H2O)
(a) how many moles of NaN3 are formed when 2.30 moles of Na is reacted with an
excess amount of N2 gas? (b) How many moles of N2 are reacted by 2.30 moles
of Na? (c) How many moles of Na and N2, respectively, are needed to produce 2.5
moles of NaN3 ?
(Answer: (a) 2.30 mol NaN3; (b) 3.45 mol N2; (c) 2.5 mol Na and 3.75 mol N2)
Mass-to-Mole-to-Mole-to-Mass Relationships
When actual reactions are carried out, the quantities of substances are measured
in mass units. We have to convert their masses into moles so that they can be directly
related to each other using the coefficients in the balanced equation.
Again consider the reaction: N2(g) + 3H2(g) 2NH3(g)
Suppose that 454 g N2 is available and we want to know how many grams of H 2 are
needed to react completely with this amount of N2 and how many grams of NH3 will be
produced.
Chemical reactions carried out in laboratories or manufacturing plants are not always
carried out such that reactants are mixed in the exact stoichiometric quantities. One of
the reactants maybe present in a limited amount and get used up first. It is referred to
as the limiting reagent. When a reaction involves a limiting reagent, the amount of
products formed is dependent on the quantity of the limiting reagent that is available.
Consider the reaction:
H2SO4(aq) + 2NH3(g) (NH4)2SO4(s)
The amount of ammonium sulfate, (NH4)2SO4, produced will depend on whether H2SO4,
or NH3 is the limiting reagent. If H2SO4 is the limiting reagent, we will produce one
mole of (NH4)2SO4 for every mole of H2SO4 reacted. If NH3 is the limiting reagent, we
will produce one-half mole of (NH4)2SO4 for every mole of NH3 reacted.
When the amounts of both reactants are given, it is important that we know which of
them is the limiting reagent, because it is the one that determine how much product
will be formed. For example, when 5 moles each of H 2SO4 and NH3 are used, the
limiting reactant would be NH3, since 5 moles of NH3 would react with 2.5 moles of
H2SO4, but 5 moles of H2SO4 would require 10 mole of NH3 (which is not available).
If the masses of reactants are given, we will need to convert them into moles of the
substances. This will allow us to find the mole ratio of the reactants and then determine
the limiting reactant. For example, suppose that 35 g of H 2SO4 and 15 g of NH3 are
allowed to react. To determine the limiting reactant we calculate the mole of each
reactant as follows:
mole ratio: 0.88 mol NH3 = 2.4 mol NH3 > 2/1
0.36 mol H2SO4 1 mol H2SO4
Since the mole ratio of NH3 to H2SO4 is larger than the stoichiometric ratio according to
the balanced equation, it means that there are more NH 3 than is needed, which makes
H2SO4 as the limiting reagent. The amount of ammonium sulfate, (NH 4)2SO4 is
calculated as follows:
On the other hand, if a reaction mixture contains 45 g H2SO4 and 15 g NH3, their
number of moles are:
mole of NH3 = 15 g NH3 x 1 mol NH3 = 0.88 mol NH3
17.03 g NH3
The mole ratio = 0.88 mol NH3 = 1.9 /1 < 2/1, which makes NH3 as the limiting
reagent
0.46 mol H2SO4
Since the calculation based on available H2SO4 yields less ammonium sulfate, it means
that H2SO4 is the limiting reagent, and the maximum amount of (NH 4)2SO4 expected is
94 g.
Electrons are negatively charged, and thus are opposite the charge of the nucleus.
If objects are with opposite charges attract one another, why are electrons not attracted
toward the nucleus? How are electrons arranged, and how can they still move around
the nucleus despite the electrostatic attraction between them and the nucleus? These are
the questions to be answered in this lesson.
Scientists have tried to explain the behaviour and occurrence of atoms and
molecules. Initially, they have assumed that atoms were like balls (i.e., macroscopic
objects) following the conditions specified by classical mechanics, the study of motion of
objects in relation to Newton’s Law of Motion. Though they were able to establish the
connection of some macroscopic observations on properties of matter such as the
pressure of gases, they failed to explain how atoms attain their stability and how force
holds atoms together to form compounds. They concluded that classical mechanics, which
govern the behaviour of larger object, was not applicable to, minute entities such as
atoms and molecules. Thus, quantum mechanics was introduced.
Quantum mechanics refers to the study of motion of minute entities, such as
electrons, in relation to waves and the specific amount of energy they can hold. It
describes the electrons as particles with wavelike properties such as immense speed.
Speed is directly proportional to kinetic energy. Thus, electrons have high kinetic energy
as they revolve around the nucleus, and this energy prevents them from collapsing into
the nucleus of the atom.
Electromagnetic Radiation (EMR)
Electromagnetic radiation is a form of energy that is made up of electric and
magnetic fields travelling in a wavelike pattern. Both fields oscillate in perpendicular
planes. As EMR travels in a wavelike pattern, it can be characterized by its amplitude,
wavelength, and frequency. Amplitude refers to the height of the crest. It determines the
intensity of the radiation, which is usually manifested in terms of brightness. Thus it can
be concluded that an EMR with high amplitude gives bright light, whereas an EMR with
low amplitude produces dim light.
On the other hand, wavelength, which is represented by the lowercase version of
the Greek letter lambda(𝜆), is defined as the distance between two successive maxima
(crest) or two minima (trough). Usually, wavelength is expressed in terms of nanometers
(nm), which is given as 1m = 1 × 109nm.
It can also be expressed in terms of centimeters (sm)meters (m)or micrometers (μm)
The wavelength of an EMR gives its characteristics color. It can be described by
the following the ROYGBIV rainbow color sequence wherein red light has relatively
longest wavelength , wherein violet light exhibits the shortest wavelength (in visible
spectrum of EMR).
Frequency (v) is the number of oscillations per unit of time, which is usually
expressed as per second hertz (Hz). Wavelength is related to frequency as shown in the
given equation
𝑐
𝑣=
𝜆
where c is the speed of light given by the constant value of 3. 00 × 108m/s. The
equation shows that wavelength is inversely proportional to frequency. Hence, longer
frequencies of light has shorter wavelengths.
Electromagnetic Spectrum
The electromagnetic spectrum is the range of all types of EMR. The other types of
EMR that make up the electromagnetic frequency are radio frequency (RF) microwaves,
infrared (IR), ultraviolet (UV), X-rays, and gamma ray radiation. These are shown in the
figure below.
Light as a Particle
The photoelectric effect is a phenomenon wherein metals emit electrons when
bombarded with light or EMR. However, this phenomenon takes only when the EMR
exceeds that of the work function, ∅ or the binding energy that holds the electrons of the
metal atom. No electrons will be further ejected from the metal’s surface regardless of
the intensity. In 1905, Albert Einstein proposed an explanation to the photoelectric effect.
Einstein proposed that light must be made up of packets of energy called quanta
(singular: quantum) or photons. Each packet is quantized and has energy equal to
𝐸 = ℎ𝑣
where h is the Planck’s constant equal to 6.626 × 10^ − 34 J (s). The above equation can
also be expressed as
ℎ𝑐
𝐸=
𝜆
The excess energy after the ejection of electron is further transferred to the
electron in the form of kinetic energy (KE), which can be mathematically expressed as
𝐾𝐸 = ℎ𝑣 − ∅
This explains the dual wavelike and particle-like property of light.
The restrictions of Bohr’s planetary model led scientists to propose another quantum
mechanical model of an atom. This model is based on Max Planck’s quanta and Albert
Einstein’s relativity mechanics.
From these notions, Louis de Broglie proposed dual nature of matter, primarily based
on the dual nature of light, wherein electrons manifest particle and wave properties
similar to that of light. This idea of de Broglie was proven when electrons were allowed
to pass through a narrow opening and produced a diffraction pattern.
The wavelike property of electron can be described using the de Broglie equation
which is given by
ℎ
𝜆=
𝑝
where 𝜆 is the de Broglie wavelength, ℎ is the Planck’s constant and 𝑝 is the linear
momentum of the particle traveling in a wavelike pattern. Meanwhile, 𝑝 can be
determined using the equation
𝑝 = 𝑚𝑣
where 𝑚 is the mass of the electron (9.1 × 10−31kg) and 𝑣 is the velocity.
Mathematically, such uncertainty in momentum or uncertainty in position as
mentioned by Heisenberg can be expressed as
ℎ ℎ
(∆𝑥)(∆𝑝) ≥ or (∆𝑥)(∆𝑝) ≥
𝜋 4𝜋
The dual nature of matter, specifically of electrons, presented a dilemma. The idea
that electrons can be regarded as both an ideal particle and an ideal wave function seems
contradicting. The electron as an ideal particle and an ideal wave function implies that its
location and speed can be accurately predicted with respect to time. However, this
implication was not agreeing with the experimental results.
Werner Heisenberg tried to resolve the dilemma posed by dual nature of matter.
He proposed that it is impossible to measure the momentum or determine the velocity of
a very small entity such as electron and at the same time determine the precise location
with utmost certainty. This is called the Heisenberg ‘s uncertainty principle.
where,
ℎ is the reduced Plank’s constant equal ℎ,
2𝜋
Similar to Bohr’s model, an energy level called the principal or main energy level
(n), written as integer from 1,2,3, onward, is the region in space occupied by a single
electron, as in the case of the H atom. By mere approximation, this can be applied to
atoms having many electrons.
Energy varies with respect to the probable location of the electron. The amount of
kinetic energy of a moving electron is directly proportional to its distance from the
nucleus. The opposite charges of the electron and the nucleus result in an electrostatic
force that affects the movement of the electron by slightly attracting it toward the
nucleus. Thus, the nearer it is to the nucleus, the lower the energy level of an electron.
The farther it is from the nucleus, the higher the energy level of an electron.
Sublevels
Wave mechanics has established that every atom has a principal energy level. Each
energy level has one or more divisions or subshells. The subshells were accounted from
Bohr’s experiment on spectral emission. He observed that the spectrum was separated
by lines very close to each other. Variations of the lines were identified based on their
appearance in the spectrum such as sharp (s), principal (p), diffused (d), and fundamental
(i). Other possible sublevels are g, h, i and so on. Sublevels depict the shape of the region
where an electron can be found (e.g., s has a spherical shape, p has two lobes, d has
four lobes and f has approximately seven lobes).
The subshells can also be arranged in term of the amount of energy they have.
Electrons are normally distributed starting from the subshell with the lowest amount of
energy to the subshell with the highest amount of energy.
Fig. 4.5 Plotting elements in the periodic table based on the sublevels of the valence
electrons
Source:
https://www.angelo.edu/faculty/kboudrea/general/quantum_numbers/Qua
Orbitals
Orbitals take into consideration the shape of the sublevels and how they are
orientated in space. The s sublevel has one orbital, the p sublevel has three orbitals and
d sublevel has five orbitals, and the f sublevel has seven orbitals. Only two electrons can
occupy an orbital. Thus, the s sublevel has two electrons, the p sublevel has six electrons
the d sublevel has 10 electrons, and the f sublevel has 14 electrons.
Fig. 4.6 The orbitals of the s, p, d, and f sublevels
Source: https://chemistry.stackexchange.com/questions/31189/what-is-spdf-
3. Pauli’s exclusion principle. Only two electrons should be placed in a given orbital
and that they should be of opposite spin in space. The placement of electrons
occupying the same orbital with the same spin violates the rule.
For example, to write the electron configuration of potassium (K), first note its
atomic number, which is 19. This means that there are 19 electrons that must be
distributed as follows:
In writing the electron configuration of an atom, you will be able to determine the
atom’s valence electrons. These are the electron that are found in the outermost shell
and participate in chemical bonding. On the other hand, the core electrons are those
electrons in complete principal energy levels. For instance, the valence electron of
potassium is 4𝑠1, whereas its core electrons are1𝑠2 2𝑠2 2𝑝6 3𝑠2 3𝑝6. Core electrons
penetrate into the region near the nucleus (i.e., less shielded) and experience effective
nuclear charge. Moreover, valence electrons do not penetrate, are more shielded by the
core electrons, and therefore experience lower energy.
Orbital diagram:
Orbital diagram:
Based on the orbital diagrams of examples 1 and 2, you can see that some orbitals
are filled with paired electrons, whereas others are not. The presence of an unpaired
electron has some magnetic implications when these atoms are subjected in a magnetic
field. This property is called paramagnetism. On the other hand, atoms having paired
electrons will less likely draw toward any existing magnetic field and later on exhibit weak
repulsion. This property is called diamagnetism. In the examples, 14𝑁
7 is paramagnetic,
and 24𝑀𝑔
12 is diamagnetic.
V. Quantum Numbers
In view of quantum mechanics, orbitals are viewed as three-dimensional spaces
or regions. The dimensions of each orbital are typically adopted from the Cartesian
coordinates, 𝑥, 𝑦, and 𝑧. However, in cases wherein the motion of a single particle is
restricted in a field of force which is a function only of distance from the origin, called
central field problem, the Schrӧdinger equation can be mathematically treated in
separable spherical polar coordinates as shown in Equation 4.1. When this equation is
factored, it gives the same angular parts 𝑃 (θ), F (ϕ) and varies with radial part 𝑅 (𝑟).
The solution of the above equation resulted in the derivation of the distance, r, of each
orbital from the nucleus. A plot of the radial part of the wave function R ( r), against
distance from the nucleus describes the electron density varying distances from the
nucleus. Moreover, as previously mentioned Ψ2 gives the probability density, such that
when these values are considered, it would allow you to define a specific region or
space in which electron will spend at least 90% of its time. On the other hand, the
radial probability function describes the probability density as a function of its distance
from the nucleus, which can be used in rationalizing as to which orbital is most likely
involved in a reaction. The angular functions, θ and ϕ, provide information as to the
shape of an orbital and its orieE
ntqautaiotinonin4s.1pace, determined by the angular momentum
quantum number (which is 1), and magnetic quantum number 𝑚. Accordingly, the
radial function is determined by the quantum number 𝑛, which is the principal quantum
number.
Ψ (𝑟, θ, ϕ) = R (𝑟) 𝑃(θ) F( ϕ)
An atom with a closed shell of valence electrons (corresponding to an electron configuration s2p6)
tends to be chemically inert. Atoms with one or two more valence electrons than are needed for a
"closed" shell are highly reactive because the extra valence electrons are easily removed to form a
positive ion. Atoms with one or two valence electrons fewer than are needed to form a closed shell
are also highly reactive because of a tendency either to gain the missing valence electrons (thereby
forming a negative ion), or to share valence electrons (thereby forming a covalent bond).
Like an electron in an inner shell, a valence electron has the ability to absorb or release energy in the
form of a photon. An energy gain can trigger an electron to move (jump) to an outer shell; this is
known as atomic excitation. Or the electron can even break free from its associated atom's valence
shell; this is ionization to form a positive ion. When an electron loses energy (thereby causing a
photon to be emitted), then it can move to an inner shell which is not fully occupied.
Valence energy levels correspond to the principal quantum numbers (n = 1, 2, 3, 4 ...) or are
labeled alphabetically with letters used in the X-ray notation (K, L, M, …).
44
Valence Electron Configuration and the Periodic Table
REPRESENTATIVE ELEMENTS –
• Representative elements don’t have a full outer energy level.
You can see this in their electron configuration. Their s and p
level won’t be filled.
• Representative elements are in groups 1A-7A.
• Ex. Lithium 1s22s1
• Ex. Carbon 1s22s22p2
• You can see that carbon and lithium don’t have a full outer
45
energy level. This makes them representative elements.
• The number of valence electrons is the same as the rightmost digit of the group number of
the element.
ELEMENT BLOCKS
• Elements are divided into sections that correspond with the
elements highest occupied energy level.
• The s block is 1A and 2A. The p block is 3A-8A. The transition
and inner transition metals are in the d and f block.
• You can figure out an electron configuration by looking at the
periodic table.
• Ex. If in the 4th
period, an element would have levels 1-3 filled.
Then you write the configuration from left to right based on its
position on the table.
Periodic trends are specific patterns that are present in the periodic table that illustrate different
aspects of a certain element, including its size and its electronic properties. Major periodic trends
include: electronegativity, ionization energy, electron affinity, atomic radius, melting
point, and metallic character. Periodic trends, arising from the arrangement of the periodic table,
provide chemists with an invaluable tool to quickly predict an element's properties. These trends exist
46
because of the similar atomic structure of the elements within their respective group families or
periods, and because of the periodic nature of the elements.
Electronegativity can be understood as a chemical property describing an atom's ability to attract and
bind with electrons. Because electronegativity is a qualitative property, there is no standardized
method for calculating electronegativity. However, the most common scale for quantifying
electronegativity is the Pauling scale (Table A2), named after the chemist Linus Pauling. The numbers
assigned by the Pauling scale are dimensionless due to the qualitative nature of electronegativity.
Electronegativity values for each element can be found on certain periodic tables. An example is
provided below.
Electronegativity measures an atom's tendency to attract and form bonds with electrons. This
property exists due to the electronic configuration of atoms. Most atoms follow the octet rule (having
the valence, or outer, shell comprise of 8 electrons). Because elements on the left side of the periodic
table have less than a half-full valence shell, the energy required to gain electrons is significantly
higher compared with the energy required to lose electrons. As a result, the elements on the left side
of the periodic table generally lose electrons when forming bonds. Conversely, elements on the right
side of the periodic table are more energy-efficient in gaining electrons to create a complete valence
shell of 8 electrons. The nature of electronegativity is effectively described thus: the more inclined an
atom is to gain electrons, the more likely that atom will pull electrons toward itself.
If the valence shell of an atom is less than half full, it requires less energy to lose an electron
than to gain one. Conversely, if the valence shell is more than half full, it is easier to pull an electron
into the valence shell than to donate one.
47
This is because atomic number increases down a group, and thus there is an increased
distance between the valence electrons and nucleus, or a greater atomic radius.
Important exceptions of the above rules include the noble gases, lanthanides,
and actinides.
The noble gases possess a complete valence shell and do not usually attract electrons.
The lanthanides and actinides possess more complicated chemistry that does not
generally follow any trends. Therefore, noble gases, lanthanides, and actinides do not
have electronegativity values.
As for the transition metals, although they have electronegativity values, there
is little variance among them across the period and up and down a group.
This is because their metallic properties affect their ability to attract electrons as easily as the other
elements.
According to these two general trends, the most electronegative element is fluorine, with 3.98
Pauling units.
48
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.
· 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.
Figure 3. Graph showing the Ionization Energy of the Elements from Hydrogen to Argon
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 required to 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
49
energies:
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 principle
quantum number) and Zeff (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:
I=RHZ2effn2
Across a period, Zeff increases and n (principal quantum number) remains the same, so the
ionization energy increases.
Down a group, n increases and Zeff increases slightly; the ionization energy decreases.
As the name suggests, electron affinity is the ability of an atom to accept an electron. Unlike
electronegativity, electron affinity is a quantitative measurement of the energy change that occurs
when an electron is added to a neutral gas atom. The more negative the electron affinity value, the
higher an atom's affinity for electrons.
50
Electron affinity generally decreases down a group of elements because each atom is larger
than the atom above it (this is the atomic radius trend, discussed below). This means that an added
electron is further away from the atom's nucleus compared with its position in the smaller atom. With
a larger distance between the negatively-charged electron and the positively-charged nucleus, the
force of attraction is relatively weaker. Therefore, electron affinity decreases. Moving from left to
right across a period, atoms become smaller as the forces of attraction become stronger. This causes
the electron to move closer to the nucleus, thus increasing the electron affinity from left to right
across a period.
· Electron affinity increases from left to right within a period. This is caused by the decrease in
atomic radius.
· Electron affinity decreases from top to bottom within a group. This is caused by the increase
in atomic radius.
The atomic radius is one-half the distance between the nuclei of two atoms (just like a radius
is half the diameter of a circle). However, this idea is complicated by the fact that not all atoms are
normally bound together in the same way. Some are bound by covalent bonds in molecules, some
are attracted to each other in ionic crystals, and others are held in metallic crystals. Nevertheless, it is
possible for a vast majority of elements to form covalent molecules in which two like atoms are held
together by a single covalent bond. The covalent radii of these molecules are often referred to as
atomic radii. This distance is measured in picometers. Atomic radius patterns are observed
throughout the periodic table.
Atomic size gradually decreases from left to right across a period of elements. This is because, within
a period or family of elements, all electrons are added to the same shell. However, at the same time,
protons are being added to the nucleus, making it more positively charged. The effect of increasing
51
proton number is greater than that of the increasing electron number; therefore, there is a greater
nuclear attraction. This means that the nucleus attracts the electrons more strongly, pulling the
atom's shell closer to the nucleus. The valence electrons are held closer towards the nucleus of the
atom. As a result, the atomic radius decreases.
Down a group, atomic radius increases. The valence electrons occupy higher levels due to the
increasing quantum number (n). As a result, the valence electrons are further away from the nucleus
as ‘n’ increases. Electron shielding prevents these outer electrons from being attracted to the
nucleus; thus, they are loosely held, and the resulting atomic radius is large.
· Atomic radius decreases from left to right within a period. This is caused by the increase in
the number of protons and electrons across a period. One proton has a greater effect than one
electron; thus, electrons are pulled towards the nucleus, resulting in a smaller radius.
· Atomic radius increases from top to bottom within a group. This is caused by electron
shielding.
The melting points is the amount of energy required to break a bond(s) to change the solid
phase of a substance to a liquid. Generally, the stronger the bond between the atoms of an element,
the more energy required to break that bond. Because temperature is directly proportional to energy,
a high bond dissociation energy correlates to a high temperature. Melting points are varied and do
not generally form a distinguishable trend across the periodic table. However, certain conclusions can
be drawn from the graph below.
52
· The non-metal carbon possesses the highest boiling point of all the elements. The semi-
metal boron also possesses a high melting point.
The metallic character of an element can be defined as how readily an atom can lose an
electron. From right to left across a period, metallic character increases because the attraction
between valence electron and the nucleus is weaker, enabling an easier loss of electrons. Metallic
character increases as you move down a group because the atomic size is increasing. When the
atomic size increases, the outer shells are farther away. The principle quantum number increases and
average electron density moves farther from nucleus. The electrons of the valence shell have
less attraction to the nucleus and, as a result, can lose electrons more readily. This causes an
increase in metallic character.
· Metallic characteristics decrease from left to right across a period. This is caused by the
decrease in radius (caused by Zeff, as stated above) of the atom that allows the outer electrons to
ionize more readily.
· Metallic characteristics increase down a group. Electron shielding causes the atomic radius to
increase thus the outer electrons ionizes more readily than electrons in smaller atoms.
· Metallic character relates to the ability to lose electrons, and nonmetallic character relates to
the ability to gain electrons.
· Another easier way to remember the trend of metallic character is that moving left and down
toward the bottom-left corner of the periodic table, metallic character increases toward Groups 1
and 2, or the alkali and alkaline earth metal groups. Likewise, moving up and to the right to the
upper-right corner of the periodic table, metallic character decreases because you are passing by to
the right side of the staircase, which indicate the nonmetals. These include the Group 8, the noble
gases, and other common gases such as oxygen and nitrogen.
53
In other words:
Move left across period and down the group: increase metallic character (heading towards
alkali and alkaline metals)
Move right across period and up the group: decrease metallic character (heading towards
nonmetals like noble gases)
54
Lesson 6
Chemical Bonding
There are many types of chemical bonds and forces that bind molecules together. The two most
basic types of bonds are characterized as either ionic or covalent. In ionic bonding, atoms transfer
electrons to each other. Ionic bonds require at least one electron donor and one electron acceptor. In
contrast, atoms with the same electronegativity share electrons in covalent bonds, because neither
atom preferentially attracts or repels the shared electrons.
Ionic Bonding
Ionic bonding is the complete transfer of valence electron(s) between atoms. It is a type of
chemical bond that generates two oppositely charged ions. In ionic bonds, the metal loses electrons to
become a positively charged cation, whereas the nonmetal accepts those electrons to become a
negatively charged anion. Ionic bonds require an electron donor, often a metal, and an electron
acceptor, a nonmetal.
Ionic bonding is observed because metals have few electrons in their outer-most orbitals. By losing
those electrons, these metals can achieve noble gas configuration and satisfy the octet rule. Similarly,
nonmetals that have close to 8 electrons in their valence shells tend to readily accept electrons to
achieve noble gas configuration. In ionic bonding, more than 1 electron can be donated or received to
satisfy the octet rule. The charges on the anion and cation correspond to the number of electrons
donated or received. In ionic bonds, the net charge of the compound must be zero.
This sodium molecule donates the lone electron in its valence orbital in order to achieve octet
configuration. This creates a positively charged cation due to the loss of electron.
55
This chlorine atom receives one electron to achieve its octet configuration, which creates a
negatively charged anion.
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 the electrostatic
attraction between the particles. At the ideal interatomic distance, attraction between these particles
releases enough energy to facilitate the reaction. Most ionic compounds tend to dissociate in polar
solvents because they are often polar. This phenomenon is due to the opposite charges on each ion.
In this example, the sodium atom is donating its 1 valence electron to the chlorine atom. This creates
a sodium cation and a chlorine anion. Notice that the net charge of the resulting compound is 0.
In this example, the magnesium atom is donating both of its valence electrons to chlorine atoms. Each
chlorine atom can only accept 1 electron before it can achieve its noble gas configuration; therefore, 2
atoms of chlorine are required to accept the 2 electrons donated by the magnesium. Notice that the
net charge of the compound is 0.
Covalent Bonding
Atoms make use of their valence electrons in chemical bonding. In ionic bonding the total
number of valence electrons which indicate whether the atom has the tendency to lose or gain
electrons to complete an octet of valence electrons. In covalent bonding, more focus is given to the
number of unpaired valence electron or electrons.
56
From the electron configuration, you can tell whether a valence electron is unpaired or not and
you can count them (example from the table).
No. of Unpaired
Atom Electron Configuration
Electrons
1H 1s1 1
2He 1s2 0
3Li 1s2 2s1 1
4Be 1s2 2s2 0
5B 1s2 2s2 2p1 1
6C 1s2 2s2 2p2 2
7N 1s2 2s2 2p3 3
8O 1s2 2s2 2p4 2
9F 1s2 2s2 2p5 1
10Ne 1s2 2s2 2p6 0
The unpaired electron seeks to be paired and when the pair is formed by sharing with another
atom, a covalent bond result. The attraction of the nuclei of each atom for the shared pair of
electrons keeps them close together.
The number of unpaired electrons of an atom thus indicates the number of covalent bonds it
can form. This is known as its covalency number.
The number of valence electrons of an element depend what group it belongs as illustrated in
the periodic table below.
The numbers 1-8 above indicate the number of valence electrons an element has except helium for it
has only two valence electrons. Take for example Lithium, it has 1-valence electron based on the
group it belongs as shown in the periodic table above. The valence electrons determine how many -
electrons needed or can be shared by an element to achieve the octet rule.
An American chemist Gilbert Lewis (1875-1946) used an observation the noble gases are
unreactive because of their electron configurations. The noble gas neon has the electron
configuration of 1s2 2s2 2p6. It has a full outer shell and cannot incorporate any more electrons into
57
the valence shell. The other noble gases have the same outer shell electron configuration even
though they have different numbers of inner-shell electrons. He called his explanation the octet
rule. The octet rule states that atoms tend to form compounds in ways that give them eight valence
electrons and thus electron configuration of a noble gas. An exception to an octet rule is in the case
of Helium which forms stable compounds by achieving two valence electrons. Sharing of electrons is
one of the two ways on how to achieve the octet rule which form a covalent compound.
Lewis (Electron Dot) Structures, Covalencies, Molecular Shapes
Sharing of electrons can be illustrated using the electron-dot symbols as shown in the
following:
The hydrogen atom has only one valence electron. It needs one more electron to have the
stable electron configuration similar to helium, the noble gas closest to it in the periodic table. Thus,
two hydrogen atoms share a pair of electrons to become stable as the hydrogen molecule H2.
Another example of an atom that reacts by sharing electrons is chlorine. The chlorine atom
has seven valence electrons. It needs just one more electron to complete an octet and attain a stable
configuration similar to that of the noble gas argon. It can share its unpaired electron with another
chlorine atom so that both will have eight electrons in their outermost energy level.
The single particle formed when electrons are shared between two or more atoms is neutral
because it still contains an equal number of protons and electrons. This neutral particle that results
from the union of two or more atoms is called molecule.
As seen in the formulas above, sharing of electrons occurs between atoms of the same kind (H
with H; Cl with Cl) forming diatomic molecules (H2 and Cl2).
Seven elements exist in nature as diatomic molecules. They are H 2, N2, O2, F2, CL2, Br2, and I2.
With the exception of H2, which is stable with two valence electrons, they conform with the octet rule
for stability. Among these seven, oxygen and nitrogen attain the stable octet of electrons by sharing
more than one pair of electrons.
In O2, two pairs of electrons are shared forming a double bond between the two oxygen
atoms.
58
In the case of nitrogen, three pairs of electrons are shared forming a triple bond that holds the
two atoms in the stable N2 molecule.
Sharing of electrons also happens with two or more different kinds of atoms attaining the
stable two or eight electron configuration.
Examples:
or
Water Molecule
Another kind of sharing occurs when both of the shared pair of electrons come originally from
one of the two bonded atoms called dative or coordinate covalent bonding. Example is the formation
of ammonium ion, NH4+, from the ammonia molecule and the hydrogen ion, H+.
The nitrogen in ammonia has a lone pair of electrons (a pair of electrons not used in any
bond) in an orbital that overlaps with the empty orbital of the hydrogen ion. The lone pair of
electrons in the nitrogen atom is thus shared with the hydrogen ions and its positive charge is carried
by the whole unit called the ammonium ion, NH4+.
59
An important thing to remember is that once a coordinate covalent bond is formed, it cannot
be distinguished anymore from a regularly formed covalent bond.
Having the skill to draw the Lewis structures for the polyatomic ions and molecules is helpful in
determining molecular shapes. The following basic rules can serve as a guide to writing Lewis
structures:
1. Get the sum of the valence electrons of the atoms in the molecule.
Example: The number of electrons is 8 for NH3, 17 for NO2, 8 for CH4.
2. Write the sequential arrangement of atoms, (the so-called skeleton) after choosing the central
atom. The central atom is often the first atom in the molecular formula or the atom with the higher
covalency or is capable of higher oxidation number).
Example:
3. Distribute the valence electrons by drawing dots between and around the atoms in the skeleton
formula. (A dash, -, can be used to replace two dots and represent a covalent bond between two
atoms. The dots are arranged so that each atom has eight valence electrons and hydrogen has two
only.
Examples:
60
a. it is formed between to dissimilar atoms (atoms with different electronegativities).
b. the bonding MO is not symmetrical. There is greater electron density near one of the atomic
nuclei involved in the molecular orbital.
c. there is uneven or unequal sharing of electrons (the shared electron pair is more closely
associated with one nucleus than with the other.
d. it produces a partial negative charge on the atom to which the shared electron pair closer, and
the leaves a partial positive charge on the atom to which the shared pair of electrons is farther.
Example: Hδ+-Clδ- Hδ+-Fδ-
Electronegativities can be used to predict the polar character of the bond that holds the atoms,
The bond may be classified as: nonpolar if the electronegativity difference between the bonded
atoms is 0-0.3 units: polar covalent if the difference is between 0.4 – 1.6 units, and the ionic if
the difference is equal to or greater than 1.7.
Covalent Compounds
Covalent compounds are composed of two or more different nonmetal atoms. It is also
possible for the same two elements to form more than one kind of covalent compound. A successful
system for naming covalent compounds must provide a means of identifying different compounds
made of the same elements.
Carbon containing compounds classified as organic compounds are given names following
other rules of nomenclature.
A complete description of a molecule must include a description of its shape, or the three-
dimensional arrangement of its atoms. Prediction of molecular geometries involves the application of
the few basic principles, which, taken together, have come to be called VSEPR theory. These
principles are:
1. Electrons in molecules occupy space and repel each other.
2. The most favorable geometric arrangement for a molecule (where it is most stable) is one in which
electron repulsion is at the minimum.
3. The degree to which electron pairs repel one another depends on whether they are bonding pairs
(BPs) or lone pairs (LPs). The repulsion between two lone pairs is greater than that between a
bonding pair and a lone pair, which is greater than that between two bonding pairs. The strength of
repulsion in decreasing order is given by: [LP-LP > LP-BP > BP-BP].
To predict the geometry of a molecule or ion, we can use the following stepwise procedure:
1. Write the Lewis structure for the molecule.
2. Determine how many groups surround the central atom (the number of surrounding atoms and
the number of lone electron pairs associated with the central atom).
3. Assign a geometry that is consistent with the minimum
electron-pair repulsion. The moat stable arrangement Note: is that
which keeps the groups as far away from each other as 1. The geometry of the molecule is possible.
Name the geometry of the molecule is based on the arrangement of the
based on the spatial relationship between the central bonded atoms only.
atom and
2. Ignore multiple bonds in
the atoms bonded to it. predicting geometry. Count only
atoms and lone pairs.
61
Total Number of Minimum Repulsion
Electron Pairs Geometry
2 linear
3 trigonal
4 tetrahedral
5 trigonal bipyramidal
6 octahedral
Polarity of Molecules
Molecules are classified as polar or nonpolar depending on the polarity of the individual bonds
and the overall shape or geometry of the molecule. Nonpolar molecules generally have no polar
bonds or their individual bond dipoles cancel.
On the other hand, polar molecules generally have only one polar bond or if there are more,
their individual dipoles do not cancel.
The significance of the geometry of the molecule is seen in the case of the carbon dioxide
molecule. Each C-O bond is polar, but because the molecule is linear, the two bond dipoles are
oriented in opposite directions: O=C=O.
As a result, the dipoles cancel each other (their vector sum is zero) and the CO2 molecule is
nonpolar. Examples of polar molecules are water, ammonia and hydrogen chloride.
H2 O NH3
Boron trichloride and carbon tetrachloride are nonpolar molecules because although the bonds
are polar the dipoles are canceled by the symmetry of the molecule.
62
The relationship between polarity and molecular symmetry can be summarized as:
Molecules containing only nonpolar covalent bonds are nonpolar.
Molecules containing polar covalent bonds that have the charges distributed symmetrically are
nonpolar.
Molecules containing polar covalent bonds which have the charges distributed unsymmetrically
are polar.
63
Chapter 7
Organic Compounds
Organic compounds are subdivided into the following groups. Acyclic compounds. They
have unbranched or branched carbon chain, but no rings. In the examples below, the first two
represent compounds with unbranched carbon chain, whereas the third one is a compound with a
branched chain:
64
Carbocyclic compounds. They contain a ring (or rings) of carbon atoms only. The ring may contain
multiple bonds and may have side carbon chains.
Heterocyclic compounds. They contain a cyclic skeleton having at least one heteroatom, an atom
that is not carbon. The most common heteroatoms are nitrogen, oxygen, or sulfur. More than one
heteroatom may be present and these atoms may be identical or different. The structures of some
natural heterocyclic compounds are presented below:
Hydrocarbons are parent compounds in organic chemistry, which, according to their name,
consist of only carbon and hydrogen atoms. Most organic molecules involve functional groups, i. e. an
atom or a group of atoms of non-hydrocarbon origin that determine chemical properties of a compound.
Indeed, chemical changes occur in most reactions at the functional group whereas the molecular
framework remains unchanged. Thus, the knowledge of properties of the functional groups will greatly
help in the study of organic chemistry.
Organic compounds are divided into classes depending on the functional groups present. Some
of the main functional groups and classes are listed in Table 1.1.
65
Table 1.1. Some of the functional groups and the corresponding classes of organic compounds
* The symbol R is usually used for any hydrocarbon radical, the symbol Ar - for an aromatic radical only.
** Multiple bonds in unsaturated compounds are sometimes related to the functional groups. *** Only primary
ones are shown.
NOMENCLATURE
At the earliest stage of organic chemistry, each new compound was usually named on the basis
of its source (caffeine - from coffee-beans, urea - from urine) or its evident properties (glycerol and
glucose - from the Greek glykys, sweet). Such names are known as trivial or common names. Trade
names are widely used in pharmacy and medicine indicating some pharmaceutical effect (anesthesin,
sarcolysin). Trivial and trade names are very convenient because of their brevity, but they give no
information about the structure of a compound and cannot be systematized. Some trivial names went
out of use with time; others have shown their viability and are used now in the systematic
nomenclature.
The first systematic nomenclature appeared as far back as 1892 (Geneva Rules). It was then
perfected by a commission of the International Union of Pure and Applied Chemistry (IUPAC) and is
known now as the IUPAC rules or the IUPAC nomenclature.
66
To minimize confusion the following terms are used in the present rules.
Parent name: a part of the name used for the formation of a particular name according to the
appointed rules. For example, the name ethanol is derived from ethane. The parent name may be
both systematic (hexane from which hexanal is derived) and trivial (benzene and nitrobenzene from
it).
Characteristic group: this term is practically equal to the term functional group, for example, the amino
group -NH2, the carbonyl group >C=O, the oxo group =O, the carboxyl group -COOH.
Principal (senior) group: the characteristic group chosen for expression as a suffix in a particular name.
This group has no other advantages over remainder groups.
Radical: a part of a molecule that remains after removal of one or more hydrogen atoms from it. For
example, the radicals, such as methyl, CH3-, and methylene, -CH2-, are derived from methane,
CH4.
Locant: a numeral or a letter showing a position of a substituent or a multiple bond in a parent structure.
Multiplying affix: syllables di-, tri-, tetra-, etc., which are used to indicate a set of identical
substituents or multiple bonds.
Nomenclature Systems. There are eight basic nomenclature systems from which the most versatile
and common therefore is the substitutive nomenclature. The next in prevalence is the radicofunctional
nomenclature. These two nomenclatures, especially the former, will be considered in greater detail.
Figure 1.1. The scheme for constructing the IUPAC substitutive name. The symbol X represents
multiplying affix(es).
There are two types of characteristic groups. One type is designated in a name only as prefixes. Nitro
group, halogens, and some other groups belong to this type; they are listed in the lower part of Table
2.2.
67
Most of characteristic groups (the upper part of Table 1.2, beyond the coloured line) may be cited
either as suffixes or as prefixes. But only one kind of group (principal group) is to be cited as a suffix.
Within these groups, a conventional order of priority has been established (Table 1.2). It means the
principal group is that which characterizes the class occurring as high as possible in Table 1.2. All other
characteristic groups are then cited as prefixes. Multiplying affixes and locants are added as necessary.
Provided that the characteristic group is univalent (for example, an OH group of alcohols or a halogen
atom of halogen derivatives) the remainder of the molecule attached to that group is expressed in its
radical form as another word, which precedes the class name. When the class name refers to a
characteristic group that is bivalent (for example, the fragment –O of ethers), the two radicals attached
to it are each named as separate words in alphabetic order.
Table 1.2. Suffixes and prefixes used for some important groups in the substitutive nomenclature
IUPAC (in order of decreasing priority)
* Coloured carbon atoms are included in the name of parent structure and not in the suffix or prefix. **
Should be added in front of the name.
*** Phenols have usually common names.
**** Used only with the name of radical R, e. g. ROalkoxy- or RSalkylthio-.
68
This type of nomenclature is the most convenient one for such classes as ethers, sulfides, amines,
and halogen compounds, especially for the compounds with simple radicals.
The formation of a name for a chemical compound usually involves the following steps in the order
given below.
Step 1. From the nature of the compound determine the most pertinent type of nomenclature
(substitutive, radicofunctional, or else).
Step 2. Determine the kind of characteristic group for use as the principal group, if any. It is this group
that stipulates then the choice of a parent structure and its numbering.
Step 31. Determine the parent structure (principal chain or parent ring system2). When in an acyclic
compound there is a choice for principal chain, the following criteria are applied successively, in the
order listed, until a decision is reached:
a) the maximum number of substituents of the highest priority from Table 1.2;
d) the maximum number of substituents cited as prefixes. Step 4. Name the parent structure and the
principal group(s). Step 5. Determine and name prefixes.
Step 6. Complete the numbering. The starting point and direction of numbering must be chosen so as
to give the lowest number to the principal group. If this rule does not effect the choice, a principle of
the lowest locants is used. It means that the locants of substituents must then be as low as possible.
Step 7. Assemble the partial names (according to steps 4 and 5) into a complete name, using the
alphabetic, but not numerical, order. Multiplying affixes are not included in this order. Position of locants
is not strictly determined by the IUPAC rules. In this respect, there are alternative versions in different
languages and, what is more, there are some differences between American and British chemical
English! But usually, locants are placed in front of prefixes and a suffix. Finally, locants are separated
from each other by commas and separated from the letters by hyphens. All parts of a name are written
without space, except for the word acid in the name of carboxylic acids.
Parent structures are presented either by an open carbon chain, or carbocyclic framework, or
heterocyclic one.
Acyclic hydrocarbons.
The generic name of saturated acyclic hydrocarbons (unbranched or branched) is alkanes. The
first four saturated unbranched acyclic hydrocarbons are called methane, ethane, propane, and
butane. Names of the higher members of this series consist of a numerical term (Greek mainly),
69
followed by -ane with elision of the terminal -a from the numerical term. Examples are given in Table
1.3.
1
This step and the following one are related to the substitutive nomenclature.
2
Seniority of ring systems is too much complicated and is not considered as well. It should be
mentioned only that all heterocycles are senior to all carbocycles.
Univalent radicals derived from these hydrocarbons by removal of a hydrogen atom from a
terminal carbon are named by replacing the suffix -ane in the name of the hydrocarbon by -yl (Table
2.3). The following names are used for the unsubstituted radicals only:
The prefixes sec (secondary) and tert (tertiary) refer to the degree of alkyl substitution at a
carbon atom. There are four types of carbons that differ in their alkyl environment. If the carbon atom
is bonded to only one carbon, the former is referred to as a primary carbon. A secondary carbon has
two other carbons bonded to it, and so on, as shown below:
70
Unsaturated unbranched hydrocarbons having one double bond are named by replacing the
suffix -ane in the name of the corresponding alkane by the suffix -ene. If there are two or more double
bonds, the suffix will be -adiene, -atriene, and so on. The generic names of these hydrocarbons are
alkenes, alkadienes, alkatrienes, etc. The chain is so numbered as to give the lowest possible numbers
to the double bonds; only the lower locant is cited in the name.
The generic name of unsaturated hydrocarbons with one triple bond is alkynes. Unbranched
compounds of this series are named similarly to alkenes but using the suffix -yne. The position of the
triple bond is indicated in the same way as for alkenes.
The following non-systematic names are retained: ethylene for CH2=CH2, acetylene for
CH=CH, and isoprene for CH2=C(CH3)-CH=CH2, as well as vinyl for the radical CH2=CH- and allyl
for the radical CH2=CHCH2-.
It should be noted that the numbering system described above is applicable only to
hydrocarbons (including a branched chain) and their derivatives with non-principal characteristic
groups, such as halogens or a nitro group. The numbering of a chain may be changed in the presence
of principal characteristic groups (see step 6 in Sec. 1.2.2 and the following section).
Cyclic hydrocarbons. The names of saturated monocyclic hydrocarbons (with no side chains) are
formed by adding the prefix cyclo- to the name of unbranched alkane with the same number of carbon
atoms. The generic name of these hydrocarbons (with or without side chains) is cycloalkanes.
Unsaturated monocyclic hydrocarbons are named and numbered similarly to acyclic analogues.
This also concerns univalent radicals derived from cyclic hydrocarbons. The carbon with the free
valence is numbered as 1 (but this locant may be omitted) regardless of a characteristic group present
in the radical.
The generic name of monoand polycyclic aromatic hydrocarbons is arenes. The simplest
representatives are called benzene and naphthalene.
71
The numbering in the benzene ring is applied only to derivatives having more than one
substituent. The symbols o- (ortho), m- (meta), and p- (para) may be used in place of 1,2-, 1,3-, and
1,4-, respectively, when only two substituents are present.
Several aromatic hydrocarbons with side chains may be used as parent structures for the
names of compounds having non-principal characteristic groups. These are toluene, cumene,
xylenes (three isomers), and styrene:
The most widespread aromatic radicals are phenyl, C6H5-; benzyl, C6H5CH2-; tolyl (ortho,
meta, and para isomers) derived from toluene; and naphthyl (1- and 2-isomers) - from
naphthalene, for example:
The name of the alcohol (1) is 2-butanol. The only carbon chain consists of four atoms
(butane); the principal characteristic group (OH) is expressed by the suffix -ol. We number the chain
from the right, starting closest to the OH group.
The radicofunctional name of the compound is sec-butyl alcohol, which is derived from the
name of the radical (sec-butyl) and the name of a class. It should be noted that the names such as
72
sec-butanol, as well as isopropanol and tert-butanol are incorrect because there are no hydrocarbons
sec-butane, isopropane, and tert-butane to which the suffix -ol can be added.
The main chain in the ether (2) is a three-carbon chain of propane. The group CH3O- attached
to C-2 represents a combined substituent methyl + oxy = methyloxy, or shortly methoxy that may be
used only as a prefix. Together this gives the name 2-methoxypropane.
The radicofunctional name of the compound (2) is isopropyl methyl ether (two radicals are
set in alphabetic order), which seems to be more convenient than the substitutive one.
The name of the secondary amine (3) is N-methyl-1-propanamine. The locant N means that
substitution with methyl group was performed at nitrogen atom but not at carbon (as is usually in the
substitutive nomenclature).
The systematic name of isovaleric acid, the compound (4), employed in pharmacy is 3-
methylbutanoic acid. The suffix -oic acid is used because carbon of the carboxyl group is a member
of the parent name (butane).
The name of the compound (5) is 3-cyclohexenecarboxylic acid (not 3-cyclohexenoic acid). Carbon
of the carboxyl group cannot be a member of the cyclic parent structure (cyclohexene) in this example.
The cycle is always numbered, starting from a carbon to which the principal characteristic group is
attached.
The systematic name of citral, the compound (6), a component of lemon oil used in treatment of eye-
diseases, is 3,7-dimethyl-2,6-octadienal. It should be mentioned here that the locant 1 is never
used for suffixes -al and -oic acid.
73
The compound (7) is named simply butanone. Pay attention to the absence of a locant for the oxo
group. The atoms C-2 and C-3 are identical whereas the oxo group attached to C-1 gives rise not to a
ketone, but an aldehyde.
The simplest stable glycol is 1,2-ethanediol, the compound (8), known as an antifreeze component.
The multiplying affix di- used in addition to the suffix -ol denotes that there are two hydroxyl groups.
Locants for them must be used here to reject an isomeric structure of 1,1-diol CH3CH(OH)2. The latter
does not exist in a free state but nomenclature rules ignore the problems of stability. This means that
any structure written correctly, i. e. with appropriate valences of all atoms, can be named.
Glutamic acid, the compound (9), is one of natural (protein) amino acids. This trivial name is adopted
by the IUPAC rules as the parent name. Systematically it may be named 2-aminopentanedioic acid.
The multiplying affix di- is also used together with the suffix -oic acid as in Example 2.8.
The heterofunctional compound (10) might be named as either a substituted phenol, C 6H5OH, or a
substituted aniline, C6H5NH2. The former should be taken as the parent name due to priority of phenols
(Table 2.2). Thus, the correct name is 4-aminophenol (or p-aminophenol).
The principal characteristic group in caffeic acid, the compound (11), is the carboxyl group, therefore
the parent name (propene) is assigned to a chain of three carbons. A substituent at C-3 represents
phenyl radical with two own substituents (hydroxyl groups). Together with locants, this leads to the
full name: 3-(3,4-dihydroxyphenyl)propenoic acid.
74
The principal characteristic group in biogenous amine histamine, the compound (12), is situated in a
side chain. The heterocyclic substituent (imidazolyl radical) retains its numbering. Thus, the
systematic name of histamine is 2-(4-imidazolyl)ethanamine.
75