Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

앳킨스 물리화학 10판 기말범위

Download as pdf or txt
Download as pdf or txt
You are on page 1of 56

16C Liquids

At low enough temperatures the molecules of a gas have insuf-


Contents ficient kinetic energy to escape from each other’s attraction and
they stick together. But although molecules attract each other
16C.1 Molecular interactions in liquids 680
when they are a few diameters apart, as soon as they come into
(a) The radial distribution function 680
(b) The calculation of g(r) 681
contact they repel each other. This repulsion is responsible
for the fact that liquids and solids have a definite bulk and do
Brief illustration 16C.1: The radial distribution
function 682 not collapse to an infinitesimal point. The molecules are held
(c) The thermodynamic properties of liquids 682 together by molecular interactions, but their kinetic energies
16C.2 The liquid–vapour interface 683
are comparable to their potential energies. As a result, although
(a) Surface tension 683 the molecules of a liquid are not free to escape completely from
Example 16C.1: Using the surface tension 683
the bulk, the whole structure is very mobile and we can speak
(b) Curved surfaces 684 only of the average relative locations of molecules. In this Topic
Brief illustration 16C.2: The Laplace equation 685
we build on those concepts and add thermodynamic argu-
(c) Capillary action 685 ments to describe the surface of a liquid and the condensation
Brief illustration 16C.3: Capillary action 685
of a gas into a liquid.
16C.3 Surface films 686
(a) Surface pressure 686
The thermodynamics of surface layers 687
(b)

Example 16C.2: Calculating the surface excess 688


16C.1 Molecular interactions in liquids
16C.4 Condensation 689
The starting point for the discussion of solids is the well-
Checklist of concepts 689
ordered structure of a perfect crystal, which is discussed in
Checklist of equations 690
Topic 18A. The starting point for the discussion of gases is the
completely disordered distribution of the molecules of a perfect
gas (Topic 1A). Liquids lie between these two extremes. We see
that the structural and thermodynamic properties of liquids
depend on the nature of intermolecular interactions and that
an equation of state can be built in a similar way to that just
➤ Why do you need to know this material? demonstrated for real gases.
Many chemical reactions, including those in biological
cells and chemical reactors, occur in liquids, so you need
to understand the structure of liquids, how molecular
(a) The radial distribution function
interactions foster condensation of a gas into a liquid, The average relative locations of the particles of a liquid are
the thermodynamic properties of liquids, and how liquid expressed in terms of the radial distribution function, g(r).
surfaces are formed. This function is defined so that Ng(r)r2dr, where N is the
number density of molecules (N = N/V), is the probability that
➤ What is the key idea? a molecule will be found in the range dr at a distance r from
The attractions between molecules are responsible for the another molecule. In a perfect crystal, g(r) is a periodic array
condensation of gases into liquids at low temperatures. of sharp spikes, representing the certainty (in the absence of
defects and thermal motion) that molecules (or ions) lie at
➤ What do you need to know already? definite locations. This regularity continues out to the edges
You need to understand the nature of molecular of the crystal, so we say that crystals have long-range order.
interactions (Topic 16B), the concepts of Helmholtz and When the crystal melts, the long-range order is lost and,
Gibbs energies (Topic 3C), and the Boltzmann distribution wherever we look at long distances from a given molecule,
(Topic 15A). there is equal probability of finding a second molecule. Close
to the first molecule, though, the nearest neighbours might
16C Liquids  681

still adopt approximately their original relative positions


(b) The calculation of g(r)
and, even if they are displaced by newcomers, the new parti-
cles might adopt their vacated positions. It is still possible to Because the radial distribution function can be calculated by
detect a sphere of nearest neighbours at a distance r1, and per- making assumptions about the intermolecular interactions, it
haps beyond them a sphere of next-nearest neighbours at r2. can be used to test theories of liquid structure. However, even
The existence of this short-range order means that the radial a fluid of hard spheres without attractive interactions (a col-
distribution function can be expected to oscillate at short lection of ball-bearings in a container) gives a function that
distances, with a peak at r1, a smaller peak at r2, and perhaps oscillates near the origin (Fig. 16C.2), and one of the factors
some more structure beyond that. influencing, and sometimes dominating, the structure of a
The radial distribution function of the oxygen atoms in liquid liquid is the geometrical problem of stacking together reason-
water is shown in Fig. 16C.1. Closer analysis shows that any ably hard spheres. Indeed, the radial distribution function of a
given H2O molecule is surrounded by other molecules at the liquid of hard spheres shows more pronounced oscillations at
corners of a tetrahedron. The form of g(r) at 100 °C shows that a given temperature than that of any other type of liquid. The
the intermolecular interactions (in this case, principally by attractive part of the potential modifies this basic structure,
hydrogen bonds) are strong enough to affect the local structure but sometimes only quite weakly. One of the reasons behind
right up to the boiling point. Raman spectra indicate that in the difficulty of describing liquids theoretically is the similar
liquid water most molecules participate in either three or four importance of both the attractive and repulsive (hard core)
hydrogen bonds. Infrared spectra show that about 90 per cent components of the potential.
of hydrogen bonds are intact at the melting point of ice, falling There are several ways of building the intermolecular poten-
to about 20 per cent at the boiling point. tial into the calculation of g(r). Numerical methods take a box
The formal expression for the radial distribution function of about 103 particles (the number increases as computers grow
for molecules 1 and 2 in a fluid consisting of N particles is the more powerful), and the rest of the liquid is simulated by sur-
somewhat fearsome equation rounding the box with replications of the original box (Fig.
16C.3). Then, whenever a particle leaves the box through one of

g (r ) =
∫∫ ∫
N (N −1)  e − βV dτ 3dτ 4 …dτ N
N its faces, its image arrives through the opposite face. When cal-
culating the interactions of a molecule in a box, it interacts with
N2 ∫∫∫e− βVN
dτ 1dτ 2 …dτ N all the molecules in the box and all the periodic replications of
those molecules and itself in the other boxes.
Radial distribution function (16C.1)
In the Monte Carlo method, the particles in the box are
where β = 1/kT and VN is the N-particle potential energy. moved through small but otherwise random distances, and the
Although fearsome, this expression is nothing more than the change in total potential energy of the N particles in the box,
Boltzmann distribution for the relative locations of two mol- ΔVN, is calculated. Whether or not this new configuration is
ecules in a field provided by all the other molecules in the accepted is then judged from the following rules:
system.
t If the potential energy is not greater than before the
change, then the configuration is accepted.
2 If the potential energy is greater than before the change, then it
Radial distribution function, g(r)

4 °C is necessary to check if the new configuration is reasonable and


25 °C 2
100 °C
Radial distribution function, g(r)

1
High density

Low density
1

0
0 200 400 600 800 1000
Radius, r/pm

Figure 16C.1 The radial distribution function of the oxygen 0


0 1 2 3 4
atoms in liquid water at three temperatures. Note the Radius, r/d
expansion as the temperature is raised. (Based on A.H. Narten,
M.D. Danford, and H.A. Levy, Discuss. Faraday. Soc. 43, 97 Figure 16C.2 The radial distribution function for a simulation
(1967).) of a liquid using impenetrable hard spheres (ball bearings).
682 16 Molecular interactions

1.5

g(r)
1

0.5
1 2 3 r/r0 4 5 6
Figure 16C.3 In a two-dimensional simulation of a liquid that
uses periodic boundary conditions, when one particle leaves Figure 16C.4 An example of a radial distribution function
the cell its mirror image enters through the opposite face. for two particles with an interaction energy given by the
Lennard-Jones potential energy function.
can exist in equilibrium with configurations of lower potential
energy at a given temperature. To make progress, we use the for r ≥ r0 and g(r) = 0 for r < r0. Here the parameter r0 is the sep-
result that, at equilibrium, the ratio of populations of two states aration at which the Lennard-Jones potential energy function
with energy separation ΔVN is e −ΔV /kT . Because we are testing
N (eqn 16B.13; V = 4ε{(r0/r)1/2 − (r0/r)6} is equal to zero. The func-
the viability of a configuration with a higher potential energy tion g(r) is plotted in Fig. 16C.4, and we see that it resembles
than the previous configuration in the calculation, ΔVN > 0 and the form shown in Fig. 16C.2.
the exponential factor varies between 0 and 1. In the Monte Self-test 16C.1 Use mathematical software to plot the function
Carlo method, the second rule, therefore, is: v2(r) = r(dV/dr). The significance of this function will become
apparent soon.
t The exponential factor is compared with a random Answer: See Fig. 16C.5.
number between 0 and 1; if the factor is larger than the
random number, then the configuration is accepted; if
the factor is not larger, the configuration is rejected.
(c) The thermodynamic properties of liquids
The configurations generated with Monte Carlo calculations
can be used to construct g(r) simply by counting the number Once g(r) is known it can be used to calculate the thermody-
of pairs of particles with a separation r and averaging the result namic properties of liquids. For example, the contribution of
over the whole collection of configurations. the pairwise additive intermolecular potential, V2, to the inter-
In the molecular dynamics approach, the history of an ini- nal energy is given by the integral
tial arrangement is followed by calculating the trajectories of all
the particles under the influence of the intermolecular poten- 2πN 2 ∞

tials and the forces they exert. The calculation gives a series of U interaction (T ) =
V ∫ 0
g (r )V2r 2dr
snapshots of the liquid, and g(r) can be calculated as before. The Contribution of pairwise interactions to the internal energy (16C.3)
temperature of the system is inferred by computing the mean
kinetic energy of the particles and using the equipartition result That is, Uinteraction is essentially the average two-particle potential
(Foundations B) that energy weighted by g(r)r2dr, which is the probability that the pair
of particles have a separation between r and r + dr. Likewise, the
〈 12 mvq2 〉 = 12 kT contribution that pairwise interactions make to the pressure is
(16C.2)
pV 2πN ∞ dV2
for each coordinate q. nRT
= 1−
3kTV ∫
0
g (r )v2r 2dr v2 = r
dr
(16C.4a)

Brief illustration 16C.1 The quantity v2 is called the virial (hence the term ‘virial equa-
The radial distribution function
tion of state’). The dependence of v2 on r is shown for a simple
A simple pair distribution function has the form form of g(r) in Fig. 16C.5. To understand the physical content
of this expression, we rewrite it as
⎛ 4r ⎞
g (r ) = 1 + cos ⎜ − 4⎟ e −(r /r0 −1)
⎝ r0 ⎠ nRT 2π ⎛ N ⎞
2


Pressure in
p= − ⎜ ⎟ g (r )v2r 2dr terms of g(r) (16C.4b)
V 3 ⎝V ⎠ 0
16C Liquids  683

1 Table 16C.1* Surface tensions of liquids at 293 K, γ/(mN m−1)


0
γ/(mN m−1)
–1
Benzene 28.88
–2 Mercury 472
v2(r)

–3 Methanol 22.6
Water 72.75
–4
* More values are given in the Resource section. Note that 1 N m−1 = 1 J m−2.
–5

–6
1 2 3 r/r0 4 5 6 sphere is the shape with the smallest surface-to-volume ratio.
However, there may be other forces present that compete
Figure 16C.5 The virial v2(r) associated with the Lennard-Jones against the tendency to form this ideal shape, and in particular
potential energy function. gravity may flatten spheres into puddles or oceans.
Surface effects may be expressed in the language of
and then note that Helmholtz and Gibbs energies (Topic 3C). The link between
these quantities and the surface area is the work needed to
t The first term on the right is the kinetic pressure, change the area by a given amount, and the fact that dA and
Physical interpretation

the contribution to the pressure from the impact dG are equal (under different conditions) to the work done in
of the molecules in free flight. changing the energy of a system. The work needed to change
t The second term is essentially the internal pressure, the surface area, σ, of a sample by an infinitesimal amount dσ is
πT = (∂U/∂V)T (Topic 2D), representing the proportional to dσ, and we write
contribution to the pressure from the dw =γ dσ Definition Surface tension (16C.5)
intermolecular forces.

To see the connection to the internal pressure, we should rec- The constant of proportionality, γ, is called the surface ten-
ognize –dV2/dr (in v2) as the force required to move two mol- sion; its dimensions are energy/area and its units are typically
ecules apart, and therefore –r(dV2/dr) as the work required to joules per metre squared (J m−2). However, as in Table 16C.1,
separate the molecules through a distance r. The second term values of γ are usually reported in newtons per metre (because
is therefore the average of this work over the range of pair- 1 J = 1 N m, it follows that 1 J m−2 = 1 N m−1). The work of surface
wise separations in the liquid as represented by the probability formation at constant volume and temperature can be identi-
of finding two molecules at separations between r and r + dr, fied with the change in the Helmholtz energy, and we can write
which is g(r)r2dr. In brief, the integral, when multiplied by the
square of the number density, is the change in internal energy dA =γ dσ (16C.6)
of the system as it expands, and therefore is equal to the inter-
nal pressure. Because the Helmholtz energy decreases (dA < 0) if the surface
area decreases (dσ < 0), surfaces have a natural tendency to
contract. This is a more formal way of expressing what we have
already described.
16C.2 The liquid–vapour interface
We turn our attention to the physical boundary between phases, Example 16C.1 Using the surface tension
such as the surface where solid is in contact with liquid or liquid
is in contact with its vapour. In this Topic we concentrate on Calculate the work needed to raise a wire of length l and
the liquid–vapour interface, which is interesting because it is so to stretch the surface of a liquid through a height h in the
mobile. Topic 22A deals with solid surfaces. arrangement shown in Fig. 16C.6. Disregard gravitational
potential energy.
Method According to eqn 16C.5, the work required to create
(a) Surface tension
a surface of area σ given that the surface tension does not vary
Liquids tend to adopt shapes that minimize their surface area, as the surface is formed is w = γσ. Therefore, all we need do is
for then the maximum number of molecules is in the bulk to calculate the surface area of the two-sided rectangle formed
and hence surrounded by and interacting with neighbours. as the frame is withdrawn from the liquid.
Droplets of liquids therefore tend to be spherical, because a
684 16 Molecular interactions

rise of internal pressure which would then result. When the


l Force
pressure inside a cavity is pin and its radius is r, the outward
force is
Total area
= 2hl pressure × area = 4πr 2 pin
h

The force inwards arises from the external pressure and the
surface tension. The former has magnitude 4πr2pout. The latter
is calculated as follows. The change in surface area when the
radius of a sphere changes from r to r + dr is

dσ = 4π(r + dr )2 − 4 πr 2 = 8πrdr
Figure 16C.6 The model used for calculating the work of
forming a liquid film when a wire of length l is raised and (The second-order infinitesimal, (dr)2 , is ignored.) The work
pulls the surface with it through a height h. done when the surface is stretched by this amount is therefore

Answer When the wire of length l is raised through a height dw = 8πγ rdr
h it increases the area of the liquid by twice the area of the
rectangle (because there is a surface on each side). The total As force × distance is work, the force opposing stretching
increase is therefore 2lh and the work done is 2γlh. through a distance dr when the radius is r is
The expression 2γlh can be expressed as force × distance by
writing it as 2γl × h, and identifying γl as the opposing force F = 8πγ r
on the wire of length l. This interpretation is why γ is called a
The total inward force is therefore 4πr2pout + 8πγr. At equilib-
tension and why its units are often chosen to be newtons per
rium, the outward and inward forces are balanced, so we can
metre (so γl is a force in newtons).
write
Self-test 16C.2 Calculate the work of creating a spherical cav-
ity of radius r in a liquid of surface tension γ. 4πr 2 pin = 4πr 2 pout + 8πγ r
Answer: 4πr2γ
which rearranges into eqn 16C.7.

(b) Curved surfaces


The minimization of the surface area of a liquid may result in The Laplace equation shows that the difference in pressure
the formation of a curved surface. A bubble is a region in which decreases to zero as the radius of curvature becomes infinite
vapour (and possibly air too) is trapped by a thin film; a cav- (when the surface is flat, Fig. 16C.7). Small cavities have small
ity is a vapour-filled hole in a liquid. What are widely called radii of curvature, so the pressure difference across their sur-
‘bubbles’ in liquids are therefore strictly cavities. True bubbles face is quite large.
have two surfaces (one on each side of the film); cavities have
only one. The treatments of both are similar, but a factor of 2 is
required for bubbles to take into account the doubled surface
area. A droplet is a small volume of liquid at equilibrium sur-
rounded by its vapour (and possibly also air).
Pressure inside, pin

The pressure on the concave side of an interface, pin, is always pout


Increasing
greater than the pressure on the convex side, pout. This relation surface tension, γ pin
is expressed by the Laplace equation, which is derived in the
following Justification:
r

pin = pout + Laplace equation (16C.7)
r
pout
0 Radius, r
Justification 16C.1 The Laplace equation
Figure 16C.7 The dependence of the pressure inside a curved
The cavities in a liquid are at equilibrium when the ten- surface on the radius of the surface, for two different values of
dency for their surface area to decrease is balanced by the the surface tension.
16C Liquids  685

p
Brief illustration 16C.2 The Laplace equation p – 2γ /r p

The pressure differential of water across the surface of a spher-


ical droplet of radius 200 nm at 20 °C is r
γ at 20 ° C
 
water
 p – 2γ /r + ρgh
2 × (72.75 ×10−3 N m −1) h
pin − pout = p p
2.00 ×10−7 m
 
r
= 7.28 × 105 N m −2 = 728 kPa

Self-test 16C.3 Calculate the pressure differential of ethanol p p p


across the surface of a spherical droplet of radius 220 nm at
20 °C. The surface tension of ethanol at that temperature is Figure 16C.8 When a capillary tube is first stood in a liquid, the
22.39 mN m−1. latter climbs up the walls, so curving the surface. The pressure just
Answer: 204 kPa under the meniscus is less than that arising from the atmosphere
by 2γ/r. The pressure is equal at equal heights throughout the
liquid provided the hydrostatic pressure (which is equal to ρgh)
cancels the pressure difference arising from the curvature.
(c) Capillary action
78
The tendency of liquids to rise up capillary tubes (tubes of
Surface tension, γ /(mN m–1)

narrow bore; the name comes from the Latin word for ‘hair’), 74
which is called capillary action, is a consequence of surface
tension. Consider what happens when a glass capillary tube 70
is first immersed in water or any liquid that has a tendency to
adhere to the walls. The energy is lowest when a thin film covers as 66
much of the glass as possible. As this film creeps up the inside
wall it has the effect of curving the surface of the liquid inside 62
the tube. This curvature implies that the pressure just beneath
the curving meniscus is less than the atmospheric pressure by 58
approximately 2γ/r, where r is the radius of the tube and we 0 20 40 60 80 100
Temperature, θ/°C
assume a hemispherical surface. The pressure immediately
under the flat surface outside the tube is p, the atmospheric Figure 16C.9 The variation of the surface tension of water with
pressure; but inside the tube under the curved surface it is only temperature.
P − 2γ/r. The excess external pressure presses the liquid up the
tube until hydrostatic equilibrium (equal pressures at equal
depths) has been reached (Fig. 16C.8).
Brief illustration 16C.3 Capillary action
To calculate the height to which the liquid rises, we note that
the pressure exerted by a column of liquid of mass density ρ If water at 25 °C rises through 7.36 cm in a capillary of radius
and height h is 0.20 mm, its surface tension at that temperature is
γ = 12 ρghr
p = ρgh (16C.8)
= 12 × (997.1 kg m −3 ) × (9.81m s −2 ) × (7.36 × 10−2 m)
This hydrostatic pressure matches the pressure difference 2γ/r × (2.0 × 10−4 m)
at equilibrium. Therefore, the height of the column at equilib- 1kg m s−2 =1N

rium is obtained by equating 2γ/r and ρgh, which gives = 72 mN m −1

2γ Self-test 16C.4 Calculate the surface tension of water at


h= (16C.9)
ρgr 30 °C given that at that temperature water climbs to a height
of 9.11 cm in a clean glass capillary tube of internal radius
This simple expression provides a reasonably accurate way 0.320 mm. The density of water at 30 °C is 0.9956 g cm−3.
of measuring the surface tension of liquids. Surface tension Answer: 142 mN m−1
decreases with increasing temperature (Fig. 16C.9).
686 16 Molecular interactions

γsg
+1

cos θc
0
γlg θc

γsl

–1
0 1 2
Figure 16C.10 The balance of forces that results in a contact wad/γlg
angle, θc.
Figure 16C.11 The variation of contact angle (shown by the
semaphore-like object) as the ratio wad/γ lg changes.
When the adhesive forces between the liquid and the material
of the capillary wall are weaker than the cohesive forces within For mercury in contact with glass, θc = 140°, which corresponds
the liquid (as for mercury in glass), the liquid in the tube retracts to wad/γlg = 0.23, indicating a relatively low work of adhesion of
from the walls. This retraction curves the surface with the con- the mercury to glass on account of the strong cohesive forces
cave, high pressure side downwards. To equalize the pressure at within mercury.
the same depth throughout the liquid the surface must fall to
compensate for the heightened pressure arising from its curva-
ture. This compensation results in a capillary depression.
In many cases there is a nonzero angle between the edge 16C.3 Surface films
of the meniscus and the wall. If this contact angle is θc, then
eqn 16C.9 should be modified by multiplying the right-hand The compositions of surface layers have been investigated by
side by cos θc. The origin of the contact angle can be traced to the simple but technically elegant procedure of slicing thin lay-
the balance of forces at the line of contact between the liquid ers off the surfaces of solutions and analysing their composi-
and the solid (Fig. 16C.10). If the solid–gas, solid–liquid, and tions. The physical properties of surface films have also been
liquid–gas surface tensions (essentially the energy needed to investigated. Surface films one molecule thick are called mono-
create unit area of each of the interfaces) are denoted γsg, γsl, and layers. When a monolayer has been transferred to a solid
γlg, respectively, then the vertical forces are in balance if support, it is called a Langmuir–Blodgett film, after Irving
Langmuir and Katherine Blodgett, who developed experimen-
γ sg = γ sl + γ lg cosθ c (16C.10) tal techniques for studying them.

This expression solves to


(a) Surface pressure
γ sg − γ sl
cos θ c = (16C.11) The principal apparatus used for the study of surface mono-
γ lg
layers is a surface film balance (Fig. 16C.12). This device
If we note that the superficial work of adhesion of the liquid to consists of a shallow trough and a barrier that can be moved
the solid (the work of adhesion divided by the area of contact) is along the surface of the liquid in the trough, and hence com-
press any monolayer on the surface. The surface pressure, π,
wad = γ sg + γ lg − γ sl (16C.12)
the difference between the surface tension of the pure sol-
eqn 16C.11 can be written vent and the solution (π = γ * − γ ) is measured by using a tor-
wad sion wire attached to a strip of mica that rests on the surface
cosθ c = −1 Contact angle (16C.13) and pressing against one edge of the monolayer. The parts
γ lg
of the apparatus that are in touch with liquids are coated
We now see that: in polytetrafluoroethene to eliminate effects arising from
the liquid–solid interface. In an actual experiment, a small
t The liquid ‘wets’ (spreads over) the surface,
amount (about 0.01 mg) of the surfactant under investigation
corresponding to 0 < θc < 90°, when 1 < wad/γ lg < 2
interpretation

is dissolved in a volatile solvent and then poured on to the


(Fig. 16C.11).
surface of the water; the compression barrier is then moved
Physical

t The liquid does not wet the surface, corresponding across the surface and the surface pressure exerted on the
to 90° < θc < 180°, when 0 < wad/γ lg < 1. mica bar is monitored.
16C Liquids  687

Liquid + surfactant 50

Surface pressure, π /(mN m–1)


40
Stearic
30 acid
Liquid Liquid

20 Isostearic
acid Tri-p-cresyl
10 phosphate
Compression
barrier Mica float
0
Figure 16C.12 A schematic diagram of the apparatus used 0 0.2 0.4 0.6 0.8 1
to measure the surface pressure and other characteristics of Area per molecule/nm2
a surface film. The surfactant is spread on the surface of the
Figure 16C.13 The variation of surface pressure with the area
liquid in the trough, and then compressed horizontally by
occupied by each surfactant molecule. The collapse pressures
moving the compression barrier towards the mica float. The
are indicated by the horizontal lines.
latter is connected to a torsion wire, so the difference in force
on either side of the float can be monitored.
phosphate film is more compressible than the stearic acid films,
which is consistent with their different molecular structures.
Some typical results are shown in Fig. 16C.13. One param- A third feature of the isotherms is the collapse pressure, the
eter obtained from the isotherms is the area occupied by the highest surface pressure. When the monolayer is compressed
molecules when the monolayer is closely packed. This quan- beyond the point represented by the collapse pressure, the
tity is obtained from the extrapolation of the steepest part of monolayer buckles and collapses into a film several molecules
the isotherm to the horizontal axis. As can be seen from the thick. As can be seen from the isotherms in Fig. 16C.13, stearic
illustration, even though stearic acid (1) and isostearic acid (2) acid has a high collapse pressure, but that of tri-p-cresyl phos-
are chemically very similar (they differ only in the location of phate is significantly smaller, indicating a much weaker film.
a methyl group at the end of a long hydrocarbon chain), they
occupy significantly different areas in the monolayer. Neither,
though, occupies as much area as the tri-p-cresyl phosphate
(b) The thermodynamics of surface layers
molecule (3), which is like a wide bush rather than a lanky tree. A surfactant is a species that is active at the interface between
two phases, such as at the interface between hydrophilic and
O hydrophobic phases. A surfactant accumulates at the inter-
HO face, and modifies its surface tension and hence the surface
pressure. To establish the relation between the concentration
1 Stearic acid, C17H35COOH
of surfactant at a surface and the change in surface tension it
brings about, we consider two phases α and β in contact and
O suppose that the system consists of several components J, each
one present in an overall amount nJ. If the components were
HO
distributed uniformly through the two phases right up to the
2 Isostearic acid, C17H35COOH interface, which is taken to be a plane of surface area σ, the
total Gibbs energy, G, would be the sum of the Gibbs energies
of both phases, G = G(α) + G(β). However, the components are
not uniformly distributed because one may accumulate at the
interface. As a result, the sum of the two Gibbs energies differs
from G by an amount called the surface Gibbs energy, G(σ):
O
O P O
Surface
O G(σ ) = G − {G(α) + G(β)} Definition Gibbs (16C.14)
energy
3 Tri-p-cresylphosphate
Similarly, if it is supposed that the concentration of a species J
The second feature to note from Fig. 16C.13 is that the tri- is uniform right up to the interface, then from its volume we
p-cresyl phosphate isotherm is much less steep than the stearic would conclude that it contains an amount nJ(α) of J in phase α
acid isotherms. This difference indicates that the tri-p-cresyl and an amount nJ(β) in phase β. However, because a species may
688 16 Molecular interactions

accumulate at the interface, the total amount of J differs from The comparison implies that, at constant temperature,
the sum of these two amounts by nJ(σ) = nJ − {nJ(α) + nJ(β)}.
This difference is expressed in terms of the surface excess, ΓJ:
σ dγ + ∑n (σ )dμ = 0
J
J J

nJ (σ )
ΓJ = Definition Surface excess (16C.15) Division by σ then gives eqn 16C.16.
σ
The surface excess may be either positive (an accumulation of J
at the interface) or negative (a deficiency there). Now consider a simplified model of the interface in which
The relation between the change in surface tension and the the ‘oil’ and ‘water’ phases are separated by a geometrically flat
composition of a surface (as expressed by the surface excess) surface. This approximation implies that only the surfactant,
was derived by Gibbs. In the following Justification we derive S, accumulates at the surface, and hence that Γoil and Γ water are
the Gibbs isotherm, between the changes in the chemical both zero. Then the Gibbs isotherm equation becomes
potentials of the substances present in the interface and the
dγ = − Γ S dμ S (16C.17)
change in surface tension:
For dilute solutions,
dγ = − ∑Γ dμ
J
J J Gibbs isotherm (16C.16) dμS = RT ln c (16C.18)

where c is the molar concentration of the surfactant. It follows


that
Justification 16C.2 The Gibbs isotherm dc
dγ = −RT Γ S
A general change in G is brought about by changes in T, p, and c
the nJ: at constant temperature, or

dG = − SdT + Vdp + γ dσ + ∑μ dn
J
J J ⎛ ∂γ ⎞ RT Γ S
⎜⎝ ∂c ⎟⎠ = − c
Dependence of the
surface tension on (16C.19)
T surfactant concentration
When this relation is applied to G, G(α), and G(β) we find
If the surfactant accumulates at the interface, its surface excess
dG(σ ) = −S(σ )dT + γ dσ + ∑ J
μJdnJ (σ ) is positive and eqn 16C.19 implies that (∂γ /∂c)T < 0. That is,
the surface tension decreases when a solute accumulates at a
surface. Conversely, if the concentration dependence of γ is
because at equilibrium the chemical potential of each compo- known, then the surface excess may be predicted and used
nent is the same in every phase, μ J(α) =μ J(β) = μ J(σ). Just as in to infer the area occupied by each surfactant molecule on the
the discussion of partial molar quantities (Chapter 5), the last surface.
equation integrates at constant temperature to

G(σ ) = γσ + ∑μ n (σ )
J
J J Example 16C.2 Calculating the surface excess
Calculate the surface excess of 1-aminobutanoic acid in a
We are seeking a connection between the change of surface 0.10 mol dm−3 aqueous solution at 20 °C given that dγ/d(ln c)
tension dγ and the change of composition at the interface. = −40 μN m−1. Convert the answer to the number of molecules
Therefore, we use the argument that in Topic 5A led to the per square metre.
Gibbs–Duhem equation (eqn 5A.12b), but this time we com-
Method Use the relation d(ln x) = (1/x)dx to convert eqn
pare the expression
16C.19 into an expression for ∂γ/∂(ln c), then rearrange it to
obtain an expression for the surface excess Γ S . Multiplying
dG(σ ) = γ dσ + ∑μ dn (σ )
J
J J
the surface excess by Avogadro’s constant gives the number of
molecules per square metre.
(which is valid at constant temperature) with the expres- Answer Because d(ln c) = (1/c)dc and dc = cd(ln c), eqn 16C.19
sion for the same quantity but derived from the preceding may be written as
equation:
⎛ ∂γ ⎞
⎜⎝ ∂(ln c ) ⎟⎠ = − RT Γ S
dG(σ ) = γ dσ + σdγ + ∑μ dn (σ )+ ∑n (σ )dμ
J
J J
J
J J T
16C Liquids  689

about 10 molecules in diameter and the basis of the calculation


It follows that
is suspect. The first figure shows that the effect is usually small;
1 ⎛ ∂γ ⎞ nevertheless it may have important consequences.
ΓS =
RT ⎜⎝ ∂(ln c ) ⎟⎠ T Consider the formation of a cloud. Warm, moist air rises into
1 the cooler regions higher in the atmosphere. At some altitude
= × (−4.0 ×10−5 N m −1)
(8.314 JK −1 mol −1) × (293 K) the temperature is so low that the vapour becomes thermody-
= 1.6 ×10−8 mol m −2 namically unstable with respect to the liquid and we expect it
to condense into a cloud of liquid droplets. The initial step can
The number of molecules per square metre, N, is be imagined as a swarm of water molecules congregating into
a microscopic droplet. Because the initial droplet is so small, it
N = N A Γ S = (6.02 ×1023 mol −1) × (1.6 ×10− 8 mol m − 2) has an enhanced vapour pressure. Therefore, instead of growing
= 9.6 ×1015 m − 2 it evaporates. This effect stabilizes the vapour because an initial
tendency to condense is overcome by a heightened tendency to
Self-test 16C.5 Use the result from Example 16C.2 to calculate evaporate. The vapour phase is then said to be supersaturated.
the area occupied by a molecule. It is thermodynamically unstable with respect to the liquid but
Answer: 1.0 × 102 nm2 not unstable with respect to the small droplets that need to
form before the bulk liquid phase can appear, so the formation
of the latter by a simple, direct mechanism is hindered.
Clouds do form, so there must be a mechanism. Two pro-
16C.4 Condensation cesses are responsible. The first is that a sufficiently large num-
ber of molecules might congregate into a droplet so big that the
We now bring together concepts from this Topic and Topic 4B enhanced evaporative effect is unimportant. The chance of one
to explain the condensation of a gas to a liquid. We saw in Topic of these spontaneous nucleation centres forming is low, and
4B that the vapour pressure of a liquid depends on the pressure in rain formation it is not a dominant mechanism. The more
applied to the liquid. Because curving a surface gives rise to a important process depends on the presence of minute dust
pressure differential of 2γ/r, we can expect the vapour pressure particles or other kinds of foreign matter. These nucleate the
above a curved surface to be different from that above a flat sur- condensation (that is, provide centres at which it can occur) by
face. By substituting this value of the pressure difference into providing surfaces to which the water molecules can attach.
eqn 4B.3 ( p = p * eV ΔP /RT , where p* is the vapour pressure when
m
Liquids may be superheated above their boiling tem-
the pressure difference is zero) we obtain the Kelvin equation peratures and supercooled below their freezing tempera-
for the vapour pressure of a liquid when it is dispersed as drop- tures. In each case the thermodynamically stable phase is not
lets of radius r: achieved on account of the kinetic stabilization that occurs in
the absence of nucleation centres. For example, superheating
p = p *e2γ V m /rRT
Kelvin equation (16C.20) occurs because the vapour pressure inside a cavity is artifi-
cially low, so any cavity that does form tends to collapse. This
The analogous expression for the vapour pressure inside a cav- instability is encountered when an unstirred beaker of water
ity can be written at once. The pressure of the liquid outside the is heated, for its temperature may be raised above its boiling
cavity is less than the pressure inside, so the only change is in point. Violent bumping often ensues as spontaneous nuclea-
the sign of the exponent in the last expression. For droplets of tion leads to bubbles big enough to survive. To ensure smooth
water of radius 1 μm and 1 nm the ratios p/p* at 25 °C are about boiling at the true boiling temperature, nucleation centres, such
1.001 and 3, respectively. The second figure, although quite as small pieces of sharp-edged glass or bubbles (cavities) of air,
large, is unreliable because at that radius the droplet is less than should be introduced.

Checklist of concepts
☐ 1. The radial distribution function, g(r), is a probability ☐ 3. The internal energy and pressure of a f luid may be
density in the sense that g(r)dr is the probability that a expressed in terms of the radial distribution function.
molecule will be found in the range dr at a distance r ☐ 4. Liquids tend to adopt shapes that minimize their sur-
from another molecule. face area.
☐ 2. The radial distribution function may be calculated with ☐ 5. The minimization of surface area results in the forma-
Monte Carlo and molecular dynamics techniques. tion of bubbles, cavities, and droplets.
690 16 Molecular interactions

☐ 6. Capillary action is the tendency of liquids to rise up ☐ 9. A surfactant modifies the surface tension and surface
narrow tubes. pressure.
☐ 7. The surface pressure is the difference between the sur- ☐ 10. Nucleation provides surfaces to which molecules can
face tension of the pure solvent and the solution. attach and thereby induce condensation.
☐ 8. The collapse pressure is the highest surface pressure
that a surface film can sustain.

Checklist of equations
Property Equation Comment Equation number

Radial distribution function g (r12 ) = A /B, 16C.1

A = N (N −1)
∫∫∫e dτ dτ …dτ− βVN
3 4 N,

B=N
∫∫ ∫e dτ dτ …dτ
2 − βVN
1 2 N


Contribution of interactions to the internal
energy
U (T ) = (2πN /V )
interaction
2
∫ g(r)V r dr
0
2
2 V2 is the intermolecular potential
energy
16C.3


Pressure in terms of g(r) p = nRT /V − (2π /3)(N /V )
∫ g(r)v r dr
2
0
2
2 16C.4b

Laplace equation pin = pout + 2γ/r γ is the surface tension 16C.7

Contact angle cos θc = (wad/γ1g) − 1 16C.13

Surface Gibbs energy G(σ) = G − {G(α) + G(β)} Definition 16C.14

Excess energy ΓJ = nJ(σ)/σ Definition 16C.15

Gibbs isotherm dγ = − ∑ Γ J dμ J 16C.16


J

Dependence of the surface tension on (∂γ/∂c)T = −RTΓS/c 16C.19


surfactant concentration

Kelvin equation p = p*e2γ Vm /rRT 16C.20


Exercises and problems  691

CHAPTER 16 Molecular interactions

TOPIC 16A Electric properties of molecules


Discussion questions
16A.1 Explain how the permanent dipole moment and the polarizability of a 16A.3 Describe the experimental procedures available for determining the
molecule arise. electric dipole moment of a molecule.
16A.2 Explain why the polarizability of a molecule decreases at high
frequencies.

Exercises
16A.1(a) Which of the following molecules may be polar: CIF3, O3, H2O2? 16A.5(b) At 0 °C, the molar polarization of a liquid is 32.l6 cm3 mol−1 and its
16A.1(b) Which of the following molecules may be polar: SO3, XeF4, SF4? density is 1.92 g cm−3. Calculate the relative permittivity of the liquid. Take
M = 85.0 g mol−1.
16A.2(a) Calculate the resultant of two dipole moments of magnitude 1.5 D
and 0.80 D that make an angle of 109.5° to each other. 16A.6(a) The refractive index of CH2I2 is 1.732 for 656 nm light. Its density
16A.2(b) Calculate the resultant of two dipole moments of magnitude 2.5 D at 20 °C is 3.32 g cm−3. Calculate the polarizability of the molecule at this
and 0.50 D that make an angle of 120° to each other. wavelength.
16A.6(b) The refractive index of a compound is 1.622 for 643 nm light. Its
16A.3(a) Calculate the magnitude and direction of the dipole moment of the
density at 20 °C is 2.99 g cm−3. Calculate the polarizability of the molecule at
following arrangement of charges in the xy-plane: 3e at (0,0), −e at (0.32 nm,
this wavelength. Take M = 65.5 g mol−1.
0), and −2e at an angle of 20° from the x-axis and a distance of 0.23 nm from
the origin. 16A.7(a) The polarizability volume of H2O at optical frequencies is
16A.3(b) Calculate the magnitude and direction of the dipole moment of the 1.5 × 10−24 cm3: estimate the refractive index of water. The experimental value
following arrangement of charges in the xy-plane: 4e at (0, 0), −2e at (162 pm, is 1.33; what may be the origin of the discrepancy?
0), and −2e at an angle of 30° from the x-axis and a distance of 143 pm from 16A.7(b) The polarizability volume of a liquid of molar mass 72.3 g mol−1
the origin. and density 865 kg m−3 at optical frequencies is 2.2 × 10−30 m3: estimate the
refractive index of the liquid.
16A.4(a) The molar polarization of fluorobenzene vapour varies linearly with
T−1, and is 70.62 cm3 mol−1 at 351.0 K and 62.47 cm3 mol−1 at 423.2 K. Calculate 16A.8(a) The dipole moment of chlorobenzene is 1.57 D and its polarizability
the polarizability and dipole moment of the molecule. volume is 1.23 × 10−23 cm3. Estimate its relative permittivity at 25 °C, when its
16A.4(b) The molar polarization of the vapour of a compound was found to density is 1.173 g cm−3.
vary linearly with T−1, and is 75.74 cm3 mol−1 at 320.0 K and 71.43 cm3 mol−1 at 16A.8(b) The dipole moment of bromobenzene is 5.17 × 10−30 C m and its
421.7 K. Calculate the polarizability and dipole moment of the molecule. polarizability volume is approximately 1.5 × 10−19 m3. Estimate its relative
permittivity at 25 °C, when its density is 1491 kg m−3.
16A.5(a) At 0 °C, the molar polarization of liquid chlorine trifluoride
is 27.18 cm3 mol−1 and its density is 1.89 g cm−3. Calculate the relative
permittivity of the liquid.

Problems
16A.1 The electric dipole moment of toluene (methylbenzene) is 0.4 D. acid has a magnitude that increases with increasing temperature. Suggest an
Estimate the dipole moments of the three xylenes (dimethylbenzene). About interpretation of this observation.
which answer can you be sure?
16A.2 Plot the magnitude of the electric dipole moment of hydrogen peroxide O H
as the HeOeOeH (azimuthal) angle φ changes from 0 to 2π. Use the
dimensions shown in 1.

H
φ 2
97 pm
16A.4‡ D.D. Nelson et al. (Science 238, 1670 (1987)) examined several weakly
O 149 pm bound gas-phase complexes of ammonia in search of examples in which the
H atoms in NH3 formed hydrogen bonds, but found none. For example, they
1 found that the complex of NH3 and CO2 has the carbon atom nearest the
nitrogen (299 pm away): the CO2 molecule is at right angles to the CeN ‘bond’,
16A.3 Acetic acid vapour contains a proportion of planar, hydrogen bonded
dimers (2). The apparent dipole moment of molecules in pure gaseous acetic ‡ These problems were supplied by Charles Trapp and Carmen Giunta
692 16 Molecular interactions

and the H atoms of NH3 are pointing away from the CO2. The magnitude of
T/K 292.2 309.0 333.0 387.0 413.0 446.0
the permanent dipole moment of this complex is reported as 1.77 D. If the N
and C atoms are the centres of the negative and positive charge distributions, Pm/(cm3 mol−1) 57.57 55.01 51.22 44.99 42.51 39.59
respectively, what is the magnitude of those partial charges (as multiples of e)? The refractive index of ammonia at 273 K and 100 kPa is 1.000 379 (for
16A.5 The polarizability volume of NH3 is 2.22 × 10−30 m3; calculate the dipole yellow sodium light). Calculate the molar polarizability of the gas at this
moment of the molecule (in addition to the permanent dipole moment) temperature and at 292.2 K. Combine the value calculated with the static
induced by an applied electric field of strength 15.0 kV m−1. molar polarizability at 292.2 K and deduce from this information alone the
molecular dipole moment.
16A.6 The magnitude of the electric field at a distance r from a point charge
Q is equal to Q/4πε0r2. How close to a water molecule (of polarizability 16A.10 Values of the molar polarization of gaseous water at 100 kPa as
volume 1.48 × 10−30 m3) must a proton approach before the dipole moment it determined from capacitance measurements are given below as a function of
induces has a magnitude equal to that of the permanent dipole moment of the temperature.
molecule (1.85 D)?
T/K 384.3 420.1 444.7 484.1 521.0
16A.7 The relative permittivity of chloroform was measured over a range of
temperatures with the following results: Pm/(cm3 mol−1) 57.4 53.5 50.1 46.8 43.1

θ/°C −80 −70 −60 −40 −20 0 20 Calculate the dipole moment of H2O and its polarizability volume.
ε 3.1 3.1 7.0 6.5 6.0 5.5 5.0 16A.11 From data in Table 16A.1 calculate the molar polarization, relative
ρ/(g cm−3) 1.65 1.64 1.64 1.61 1.57 1.53 1.50 permittivity, and refractive index of methanol at 20 °C. Its density at that
temperature is 0.7914 g cm−3.
The freezing point of chloroform is −64 °C. Account for these results and 16A.12 Show that, in a gas (for which the refractive index is close to 1), the
calculate the dipole moment and polarizability volume of the molecule. refractive index depends on the pressure as nr = 1 + const × p, and find the
16A.8 The relative permittivities of methanol (with a melting point of −95 °C) constant of proportionality. Go on to show how to deduce the polarizability
corrected for density variation are given below. What molecular information volume of a molecule from measurements of the refractive index of a gaseous
can be deduced from these values? Take ρ = 0.791 g cm−3 at 200 °C. sample.
16A.13 Acetic acid vapour contains a proportion of planar, hydrogen-bonded
θ/°C −185 −170 −150 −140 −110 −80 −50 −20 0 20
dimers. The relative permittivity of pure liquid acetic acid is 7.14 at 290 K
εr 3.2 3.6 4.0 5.1 67 57 49 43 38 34 and increases with increasing temperature. Suggest an interpretation of the
latter observation. What effect should isothermal dilution have on the relative
16A.9 In his classic book Polar molecules, Debye reports some early permittivity of solutions of acetic acid in benzene?
measurements of the polarizability of ammonia. From the selection below,
determine the dipole moment and the polarizability volume of the molecule.

TOPIC 16B Interactions between molecules


Discussion questions
16B.1 Identify the terms in and limit the generality of the following 16B.5 Account for the hydrophobic interaction and discuss its manifestations.
expressions: (a) V = −Q2μ1/4πε 0r2, (b) V = −Q2μ1cos θ/4πε 0r2, and
16B.6 Some polymers have unusual properties. For example, Kevlar (3) is
(c) V = μ 2μ1(1 − 3 cos2 θ)/4πε 0r3.
strong enough to be the material of choice for bulletproof vests and is stable
16B.2 Draw examples of the arrangements of electrical charges that at temperatures up to 600 K. What molecular interactions contribute to the
correspond to monopoles, dipoles, quadrupoles, and octupoles. Suggest a formation and thermal stability of this polymer?
reason for the different distance dependencies of their electric fields.
16B.3 Account for the theoretical conclusion that many attractive interactions O
H N NH
between molecules vary with their separation as 1/r6. H
16B.4 Describe the formation of a hydrogen bond in terms of (a) electrostatic O O H
interactions and (b) molecular orbitals. How would you identify the better n
model? 3 Kevlar

Exercises
16B.1(a) Calculate the molar energy required to reverse the direction of an 16B.1(b) Calculate the molar energy required to reverse the direction of an
H2O molecule located 100 pm from a Li+ ion. Take the magnitude of the HCl molecule located 300 pm from a Mg2+ ion. Take the magnitude of the
dipole moment of water as 1.85 D. dipole moment of HCl as 1.08 D.
Exercises and problems  693

16B.2(a) Calculate the potential energy of the interaction between two linear 16B.3(b) Estimate the energy of the dispersion interaction (use the London
quadrupoles when they are collinear and their centres are separated by a formula) for two Ar atoms separated by 1.0 nm. Relevant data can be found in
distance r. the Resource section.
16B.2(b) Calculate the potential energy of the interaction between two linear
16B.4(a) How much energy (in kJ mol−1) is required to break the hydrogen
quadrupoles when they are parallel and separated by a distance r.
bond in a vacuum (εr = 1)? Use the electrostatic model of the hydrogen bond.
16B.3(a) Estimate the energy of the dispersion interaction (use the London 16B.4(b) How much energy (in kJ mol−1) is required to break the hydrogen
formula) for two He atoms separated by 1.0 nm. Relevant data can be found in bond in water (εr ≈ 80.0)? Use the electrostatic model of the hydrogen bond.
the Resource section.

Problems
16B.1 An H2O molecule is aligned by an external electric field of strength and evaluating {V(r + δr) − V(r)}/δr. At the end of the calculation, let δr
1.0 kV m−1 and an Ar atom (α ′ = 1.66 × 10 −30 m3) is brought up slowly from become vanishingly small.)
one side. At what separation is it energetically favourable for the H 2O
16B.5 Given that F = −dV/dr, calculate the distance dependence of the force
molecule to flip over and point towards the approaching Ar atom?
acting between two non-bonded groups of atoms in a polymer chain that have
16B.2 Suppose an H2O molecule (μ = 1.85 D) approaches an anion. What is a London dispersion interaction with each other.
the favourable orientation of the molecule? Calculate the electric field (in
16B.6 Consider the arrangement shown in 5 for a system consisting of an
volts per metre) experienced by the anion when the water dipole is (a) 1.0 nm,
OeH group and an O atom, and then use the electrostatic model of the
(b) 0.3 nm, (c) 30 nm from the ion.
hydrogen bond to calculate the dependence of the molar potential energy of
16B.3 Phenylalanine (Phe, 4) is a naturally occurring amino acid. What is interaction on the angle θ.
the energy of interaction between its phenyl group and the electric dipole
moment of a neighbouring peptide group? Take the distance between the O –0.83e
groups as 4.0 nm and treat the phenyl group as a benzene molecule. The 200 pm
dipole moment of the peptide group is μ = 2.7 D and the polarizability volume θ H
of benzene is α ′ = 1.04 × 10−29 m3. –0.83e
+0.45e
95.7 pm
O
5
H2N OH 16B.7 Suppose you distrusted the Lennard-Jones (12,6) potential for assessing
a particular polypeptide conformation, and replaced the repulsive term by an
4 Phenylalanine exponential function of the form e −r /r0 . (a) Sketch the form of the potential
energy and locate the distance at which it is a minimum. (b) Identify the
16B.4 Now consider the London interaction between the phenyl groups of
distance at which the exponential-6 potential is a minimum.
two Phe residues (see Problem 16B.3). (a) Estimate the potential energy of
interaction between two such rings (treated as benzene molecules) separated 16B.8 The cohesive energy density, U, is defined as U/V, where U is the mean
by 4.0 nm. For the ionization energy, use I = 5.0 eV. (b) Given that force is the potential energy of attraction within the sample and V its volume. Show that
negative slope of the potential, calculate the distance-dependence of the force U = 12 N ∫ V (R)dτ where N is the number density of the molecules and V(R)
acting between two non-bonded groups of atoms, such as the phenyl groups of is their attractive potential energy and where the integration ranges from d to
Phe, in a polypeptide chain that can have a London dispersion interaction with infinity and over all angles. Go on to show that the cohesive energy density of
each other. What is the separation at which the force between the phenyl groups a uniform distribution of molecules that interact by a van der Waals attraction
(treated as benzene molecules) of two Phe residues is zero? (Hint: Calculate the of the form −C6/R6 is equal to (2π/3)(N A2 /d3M2)ρ2C6, where ρ is the mass
slope by considering the potential energy at r and r + δr, with δr ≪ r, density of the solid sample and M is the molar mass of the molecules.

TOPIC 16C Liquids


Discussion question
16C.1 Describe the process of condensation.

Exercises
16C.1(a) Calculate the vapour pressure of a spherical droplet of water of 16C.2(a) The contact angle for water on clean glass is close to zero. Calculate
radius 10 nm at 20 °C. The vapour pressure of bulk water at that temperature the surface tension of water at 20 °C given that at that temperature water
is 2.3 kPa and its density is 0.9982 g cm−3. climbs to a height of 4.96 cm in a clean glass capillary tube of internal radius
16C.1(b) Calculate the vapour pressure of a spherical droplet of water of radius 0.300 mm. The density of water at 20 °C is 998.2 kg m−3.
20.0 nm at 35.0 °C. The vapour pressure of bulk water at that temperature is 16C.2(b) The contact angle for water on clean glass is close to zero. Calculate
5.623 kPa and its density is 994.0 kg m−3. the surface tension of water at 30 °C given that at that temperature water
694 16 Molecular interactions

climbs to a height of 9.11 cm in a clean glass capillary tube of internal 16C.3(b) Calculate the pressure differential of ethanol across the surface of a
diameter 0.320 mm. The density of water at 30 °C is 0.9956 g cm−3. spherical droplet of radius 220 nm at 20 °C. The surface tension of ethanol at
that temperature is 22.39 mN m−1.
16C.3(a) Calculate the pressure differential of water across the surface of a
spherical droplet of radius 200 nm at 20 °C.

Problem
16C.1 The surface tensions of a series of aqueous solutions of a surfactant A Calculate the surface excess concentration.
were measured at 20 °C, with the following results:

[A]/(mol dm−3) 0 0.10 0.20 0.30 0.40 0.50


γ/(mN m−1) 72.8 70.2 67.7 65.1 62.8 59.8

Integrated activities
16.1 Show that the mean interaction energy of N atoms of diameter part (a) to 20 kJ mol−1, a typical value for the energy of a hydrogen bonding
d interacting with a potential energy of the form C 6/R6 is given by interaction in biological systems.
U = −2N2C6/3Vd3, where V is the volume in which the molecules are
confined and all effects of clustering are ignored. Hence, find a connection O
between the van der Waals parameter a and C6, from n2a/V2 = (∂U/∂V)T. CH 3
N
16.2‡ F. Luo et al. (J. Chem. Phys. 98, 3564 (1993)) reported experimental H
observation of the He2 complex, a species which had escaped detection
for a long time. The fact that the observation required temperatures in the 8 trans-N-methylacetamide
neighbourhood of 1 mK is consistent with computational studies which
suggest that hcD e for He2 is about 1.51 × 10−23 J, hcD 0 about 2 × 10−26 J, and 16.5 This problem gives a simple example of a quantitative structure–
R about 297 pm. (a) Determine the Lennard-Jones parameters r0, and a and activity relation (QSAR). The binding of nonpolar groups of amino acid
plot the Lennard–Jones potential for HeeHe interactions. (b) Plot the Morse to hydrophobic sites in the interior of proteins is governed largely by
potential given that a = 5.79 × 1010 m−1. hydrophobic interactions. (a) Consider a family of hydrocarbons ReH. The
hydrophobicity constants, π, for RaCH3, CH2CH3, (CH2)2CH3, (CH2)3CH3,
16.3 Molecular orbital calculations may be used to predict structures of and (CH2)4CH3 are, respectively, 0.5, 1.0, 1.5, 2.0, and 2.5. Use these data to
intermolecular complexes. Hydrogen bonds between purine and pyrimidine predict the π value for (CH2)6CH3. (b) The equilibrium constants KI for the
bases are responsible for the double helix structure of DNA (see Topic 17A). dissociation of inhibitors (9) from the enzyme chymotrypsin were measured
Consider methyladenine (6, with RaCH3) and methylthymine (7, with for different substituents R:
RaCH3) as models of two bases that can form hydrogen bonds in DNA.
(a) Using molecular modelling software and the computational method of
your choice, calculate the atomic charges of all atoms in methyladenine and R CH3CO CN NO2 CH3 Cl
methylthymine. (b) Based on your tabulation of atomic charges, identify the π −0.20 −0.025 0.33 0.5 0.9
atoms in methyladenine and methylthymine that are likely to participate
log KI −1.73 −1.90 −2.43 −2.55 −3.40
in hydrogen bonds. (c) Draw all possible adenine–thymine pairs that can
be linked by hydrogen bonds, keeping in mind that linear arrangements of
the AdH…B fragments are preferred in DNA. For this step, you may want Plot log KI against π. Does the plot suggest a linear relationship? If so, what
to use your molecular modelling software to align the molecules properly. are the slope and intercept to the log KI axis of the line that best fits the data?
(d) Consult Topic 17A and determine which of the pairs that you drew in (c) Predict the value of KI for the case RaH.
part (c) occur naturally in DNA molecules. (e) Repeat parts (a)–(d) for
cytosine and guanine, which also form base pairs in DNA (see Topic 17A R NH
for the structures of these bases). CHO

O 9
NH2 CH 3
HN 16.6 Derivatives of the compound TIBO (10) inhibit the enzyme reverse
N N
O transcriptase, which catalyses the conversion of retroviral RNA to DNA. A
N N QSAR analysis of the activity A of a number of TIBO derivatives suggests the
R
R following equation:
6 7
log A = b0 + b1S + b2W
16.4 Molecular orbital calculations may be used to predict the dipole moments
of molecules. (a) Using molecular modelling software and the computational where S is a parameter related to the drug’s solubility in water and W is a
method of your choice, calculate the dipole moment of the peptide link, parameter related to the width of the first atom in a substituent X shown in 10.
modelled as a trans-N-methylacetamide (8). Plot the energy of interaction (a) Use the following data to determine the values of b0, b1, and b2. Hint: The
between these dipoles against the angle θ for r = 3.0 nm (see eqn 16B.4). (b) QSAR equation relates one dependent variable, log A, to two independent
Compare the maximum value of the dipole–dipole interaction energy from variables, S and W. To fit the data, you must use the mathematical procedure of
Exercises and problems  695

multiple regression, which can be performed with mathematical software or an


electronic spreadsheet.

X H Cl SCH3 OCH3 CN CHO Br CH3 CCH

log A 7.36 8.37 8.3 7.47 7.25 6.73 8.52 7.87 7.53
S 3.53 4.24 4.09 3.45 2.96 2.89 4.39 4.03 3.80
W 1.00 1.80 1.70 1.35 1.60 1.60 1.95 1.60 1.60

(b) What should be the value of W for a drug with S = 4.84 and log A = 7.60?

O
HN
N

N
X C

10
CHAPTER 17

Macromolecules and
self-assembly

Macromolecules are built from covalently linked components. 17C Self-assembly


They are everywhere, inside us and outside us. Some are natu-
ral: they include polysaccharides (such as cellulose), polypep- Atoms, small molecules, and macromolecules can form large
tides (such as protein enzymes), and polynucleotides (such as aggregates, sometimes by processes involving self-assembly,
deoxyribonucleic acid, DNA). Others are synthetic (such as that are held together by one or more of the molecular interac-
nylon and polystyrene). Molecules both large and small may tions described in Topic 16B. In this Topic we explore ‘colloids’,
also gather together in a process called ‘self-assembly’ and give ‘micelles’, and biological membranes, which are assemblies
rise to aggregates that to some extent behave like macromol- with some of the typical properties of molecules but also with
ecules. One example is the assembly of the protein actin into their own characteristic features. We also consider examples in
filaments in muscle tissue. In this chapter we examine the which the controlled design of new materials with enhanced
structures and properties of macromolecules and aggregates. properties is informed by understanding of the principles
underlying self-assembly.

17A The structures of macromolecules


17D Determination of size and shape
Macromolecules adopt shapes that are governed by the molecu-
lar interactions described in Topic 16B. The overall shape of a Macromolecules, whether natural or synthetic, and aggregates
protein, for instance, is maintained by van der Waals interac- need to be characterized in terms of their molar mass, their
tions, hydrogen bonding, and the hydrophobic effect. In this size, and their shape. This Topic considers how these features
Topic we consider a range of structures, beginning with a struc- are determined experimentally.
tureless ‘random coil’, partially structured coils, and then the
structurally precise proteins and nucleic acids.

What is the impact of this material?


17B Properties of macromolecules The impact of this material is immense as it underlies the dis-
cussion of biological phenomena and the properties of many
Natural macromolecules differ in certain respects from syn- modern materials. However, the applications are embedded
thetic macromolecules, particularly in their composition and in the development of the concepts and are not found on the
the resulting structure, but the two share a number of common web site.
properties. In this Topic we concentrate on mechanical, ther-
mal, and electrical properties.
17A The structures of macromolecules

Macromolecules are very large molecules assembled from


Contents smaller molecules biosynthetically in organisms, by chemists
17A.1 The different levels of structure 697
in the laboratory, or in an industrial reactor. Naturally occur-
ring macromolecules include polysaccharides such as cellulose,
17A.2 Random coils 698
polypeptides such as protein enzymes, and polynucleotides
(a) Measures of size 699
such as deoxyribonucleic acid (DNA). Synthetic macromol-
Brief illustration 17A.1: One-dimensional
random coils 699
ecules include polymers such as nylon and polystyrene that are
manufactured by stringing together and in some cases cross-
Brief illustration 17A.2: Three-dimensional
random coils 700 linking smaller units known as monomers (Fig. 17A.1).
Brief illustration 17A.3: Measures of size of
a random coil 701
(b) Constrained chains 702
(c) Partly rigid coils 702 17A.1 The different levels of structure
Example 17A.1: Calculating the root-mean-
square separation of a partly rigid coil 703 The concept of the ‘structure’ of a macromolecule takes on dif-
17A.3 Biological macromolecules 703 ferent meanings at the different levels at which we think about
(a) Proteins 704 the arrangement of the chain or network of monomers. The
(b) Nucleic acids 705 primary structure of a macromolecule is the sequence of small
Checklist of concepts 706 molecular residues making up the polymer. The residues may
Checklist of equations 706 form either a chain, as in polyethene, or a more complex net-
work in which cross-links connect different chains, as in cross-
linked polyacrylamide. In a synthetic polymer, virtually all the
residues are identical and it is sufficient to name the monomer
used in the synthesis. Thus, the repeating unit of polyethene
and its derivatives is eCHXCH2e, and the primary structure
➤ Why do you need to know this material? of the chain is specified by denoting it as e(CHXCH2)ne.
Macromolecules give rise to special problems that include The concept of primary structure ceases to be trivial in the
the investigation and description of their shapes and the case of synthetic copolymers and biological macromolecules,
determination of their sizes. You need to know how to for in general these substances are chains formed from different
describe the structural features of macromolecules in order
to understand their physical and chemical properties.

➤ What is the key idea? (a)


The structure of a macromolecule takes on different Monomer

meanings at the different levels at which the arrangement


of the chain or network of its building blocks is considered.

➤ What do you need to know already? (b)


You need to be familiar with statistical arguments
(Topic 15A). The discussion of the shapes of biological
(c)
macromolecules depends on an understanding of the
nonbonding interactions that act between molecules Figure 17A.1 Three varieties of polymer: (a) a simple linear
(Topic 16B). polymer, (b) a cross-linked polymer, and (c) one variety of
copolymer.
698 17 Macromolecules and self-assembly

molecules. For example, proteins are polypeptides formed


from different amino acids (about twenty occur naturally)
strung together by the peptide link, eCONHe. The determi-
nation of the primary structure is then a highly complex prob-
lem of chemical analysis called sequencing. The degradation
of a polymer is a disruption of its primary structure, when the
chain breaks into shorter components.
The term conformation refers to the spatial arrangement
of the different parts of a chain, and one conformation can be
changed into another by rotating one part of a chain around
a bond. The conformation of a macromolecule is relevant at
three levels of structure. The secondary structure of a mac-
Figure 17A.3 Several subunits with specific tertiary structures
romolecule is the (often local) spatial arrangement of a chain.
pack together, providing an example of quaternary structure.
The secondary structure of a molecule of polyethene in a good
solvent is typically a random coil in the absence of a solvent
polyethene forms crystals consisting of stacked sheets with a
hairpin-like bend about every 100 monomer units, presumably 17A.2 Random coils
because for that number of monomers the intermolecular (in
this case intramolecular) potential energy is sufficient to over- The most likely conformation of a chain of identical units not
come thermal disordering. The secondary structure of a pro- capable of forming hydrogen bonds or any other type of spe-
tein is a highly organized arrangement determined largely by cific bond is a random coil. Polyethene is a simple example. The
hydrogen bonds, and taking the form of random coils, helices random coil model is a helpful starting point for estimating the
(Fig. 17A.2a), or sheets in various segments of the molecule. orders of magnitude of the hydrodynamic properties of poly-
The tertiary structure is the overall three-dimensional struc- mers and denatured proteins in solution.
ture of a macromolecule. For instance, the hypothetical protein The simplest model of a random coil is a freely-jointed
shown in Fig. 17A.2b has helical regions connected by short chain, in which any bond is free to make any angle with
random-coil sections. The helices interact to form a compact respect to the preceding one (Fig. 17A.4). We assume that the
tertiary structure. Denaturation may also occur at this level, residues occupy zero volume, so different parts of the chain
with tertiary structure lost but secondary structure largely can occupy the same region of space. The model is obviously
retained. an oversimplification because a bond is actually constrained
The quaternary structure of a macromolecule is the man- to a cone of angles around a direction defined by its neighbour
ner in which large molecules are formed by the aggregation of (Fig. 17A.5) and real chains are self-avoiding in the sense that
others. Figure 17A.3 shows how four molecular subunits, each distant parts of the same chain cannot fold back and occupy
with a specific tertiary structure, aggregate. Quaternary struc- the same space.
ture can be very important in biology. For example, the oxygen- In a hypothetical one-dimensional freely jointed chain
transport protein haemoglobin consists of four myoglobin-like all the residues lie in a straight line, and the angle between
subunits that work cooperatively to take up and release O2. neighbours is either 0° or 180°. The residues in a three-
dimensional freely jointed chain are not restricted to lie in a
line or a plane.

Arbitrary
= angle

Arbitrary
angle
(a) (b)

Figure 17A.2 (a) A polymer adopts a highly organized helical


conformation, an example of a secondary structure. The
helix is represented as a cylinder. (b) Several helical segments Figure 17A.4 A freely-jointed chain is like a three-dimensional
connected by short random coils pack together, providing an random walk, each step being in an arbitrary direction but of
example of tertiary structure. the same length.
17A The structures of macromolecules  699

Arbitrary
angle Self-test 17A.1 What is the probability that the ends of a poly-
θ θ ethene chain of molar mass 85 kg mol−1 are 10 nm apart when
θ the polymer is treated as a one-dimensional freely jointed
θ
chain?
Answer: 9.8 × 10 −3
θ

Figure 17A.5 A better description is obtained by fixing the


bond angle (for example, at the tetrahedral angle) and allowing
Justification 17A.1 One-dimensional random coils
free rotation about a bond direction.
Consider a one-dimensional freely jointed polymer. We can
specify its conformation by stating the number of bonds
(a) Measures of size pointing to the right (NR) and the number pointing to the left
As shown in the following Justification, the probability, P, that (NL). The distance between the ends of the chain is (NR – NL)l,
the ends of a long one-dimensional freely jointed chain com- where l is the length of an individual bond. We write n = NR –
NL and the total number of bonds as N = NR + NL . Later in the
posed of N units of length l (and therefore of total length Nl) are
calculation we use N R = 12 (N + n) and N L = 12 (N − n).
a distance nl apart is
The probability, P, that the end-to-end separation of a ran-
1/2 domly selected polymer is nl is
⎛ 2 ⎞ One-dimensional Probability
P =⎜ e − n /2 N
2

⎝ πN ⎟⎠ random coil distribution (17A.1)


number of polymers with end-to-end distance nl
P=
total num
mber of possible conformations
This function is plotted in Fig. 17A.6.
Each of the N bonds of the polymer may in principle lie to the
1 left or the right, so the total number of possible conforma-
tions is 2 N. The total number of ways, W, of forming a chain
0.8 of N bonds with the end-to-end distance nl is the number of
Probability, P/(2/πN)1/2

ways of having N R right-pointing bonds, the rest being left-


0.6 pointing bonds. Therefore, to calculate W we need to deter-
mine the number of ways of achieving N R right-pointing
0.4 bonds given a total of N bonds. This is the same problem as
selecting N R objects from a collection of N objects (see Topic
0.2 15A), and is

0 N! N! N!
W= = =
{
N R !(N − N R )! N R ! N L ! { 1 (N + n)}! 1 ( N − n) ! }
–4 –2 0 2 4
n/N1/2 2 2

Figure 17A.6 The probability distribution for the separation of It now follows that
the ends of a one-dimensional random coil. The separation of
the ends is Nl, where l is the bond length. W N!
P= =
2N { 12 (N + n)}!{ 12 (N − n)}! 2N
Brief illustration 17A.1 One-dimensional random coils
When the chain is compact in the sense that n≪N, it is more
Suppose that N = 1000 and l = 150 pm, then the probability that convenient to evaluate ln P: the factorials are then large and
the ends of a one-dimensional random coil are nl = 3.00 nm we can use Stirling’s approximation (Topic 15A) in the form
apart is given by eqn 17A.1 by setting n = (3.00 × 103 pm)/
(150 pm) = 20.0: ln x ! ≈ ln(2π)1/2 + ( x + 12 ) ln x − x
1/2
⎛ 2 ⎞ The result, after quite a lot of algebra (see Problem 17A.7), is
P =⎜ e −20.0 /(2×1000) = 0.0207
2

⎝ π ×1000 ⎟⎠
1/2
⎛ 2 ⎞
ln P ≈ ln ⎜ − 12 (N + n + 1)ln(1 + ␯) − 12 (N − n + 1)ln(1 − ␯)
⎝ πN ⎟⎠
meaning that there is a 1 in 48 chance of being found there.
700 17 Macromolecules and self-assembly

where ν = n/N. For a compact coil (ν ≪ 1) we use the approxi- If x is regarded as continuously variable, we need to replace this
mation ln(1 ± ␯) ≈ ± ␯ − 12 ␯2 and so obtain probability by a probability density f(x) such that f(x)dx is the
probability that the ends of the chain will be found between
⎛ 2 ⎞
1/2
x and x + dx. Because dx = 2(l/31/2)dn (for the factor 2, see the
ln P ≈ ln ⎜ − 12 N ␯2
⎝ πN ⎟⎠ remark at the end of Justification 17A.1), dn = (31/2/2l)dx, so
1/2
which rearranges into eqn 17A.1. 1⎛ 6 ⎞
f (x ) = e −3 x
2
/2 Nl 2
To confirm that the total probability of the chain ends 2l ⎜⎝ πN ⎟⎠
being at any separation is 1, we integrate P over all values of
n. However, because n can change only in steps of 2, the inte- Because the probabilities of making steps along all three coor-
gration step size is 12 dn, not dn itself. Then (with N allowed to dinates are independent, the probability of finding the ends of
become infinite), the chain in a region of volume dV = dxdydz at a distance r is
the product of these densities:

N ∞ 1/ 2 ∞ Integral G..1
⎛ 2
∑P →∫
n=− N
−∞
P(n)( 12 dn) = 12 ⎝
πN ∫
−∞
e−n
2
/2 N
dn = 1
f (x , y , z )dV = f (x) f ( y ) f (z )dxdydz =
1 ⎛ 6 ⎞
8l 3 ⎜⎝ πN ⎟⎠
3/2
e −3r
2
/2 Nl 2
dV

The volume of a spherical shell at a distance r is 4πr2 , so the


total probability of finding the ends at a separation between r
We show in the following Justification that eqn 17A.1 can be and r + dr, regardless of orientation, is
used to calculate the probability that the ends of a long three- 3/2
4π ⎛ 6 ⎞
f (r )dr = r 2e −3r
2
/2 Nl 2
dimensional freely jointed chain lie in the range r to r + dr. We 8l 3 ⎝⎜ πN ⎠⎟
dr
write this probability as f(r)dr, where
from which f(r) can be identified (in blue), as in eqn 17A.2.
3
⎛ a ⎞
f (r ) = 4 π ⎜ 1/2 ⎟ r 2 e − a r
2 2

⎝π ⎠ Three-
Probability
dimensional In some coils, the ends may be far apart whereas in others
1/2 distribution (17A.2)
random coil
⎛ 3 ⎞ their separation is small. Here and elsewhere we are ignoring
a=⎜
⎝ 2Nl 2 ⎟⎠ the fact that the chain cannot be longer than Nl. Although eqn
17A.2 gives a nonzero probability for r > Nl, the values are so
For a narrow range of distances δr, the probability density can be small that the errors in pretending that r can range up to infin-
treated as a constant and the probability calculated from f(r)δr. ity are negligible.
An alternative interpretation of this expression is to regard each
coil in a sample as ceaselessly writhing from one conformation Brief illustration 17A.2 Three-dimensional random coils
to another; then f(r)dr is the probability that at any instant the
chain will be found with the separation of its ends between r Consider the chain described in Brief illustration 17A.1, with
and r + dr. N = 1000 and l = 150 pm. If the coil is three dimensional, we set
1/2
⎛ 3 ⎞
a=⎜ = 2.58…× 10−4 pm −1
Justification 17A.2 Three-dimensional random coils ⎝ 2 ×1000 × (150 pm)2 ⎟⎠

The formation of a three-dimensional random coil can be Then the probability density at r = 3.00 nm is given by eqn
regarded as the outcome of a three-dimensional random walk, 17A.2 as
in which each bond of length l represents a step of length 3
l taken in a random direction. The length of the step can be ⎛ 2.58…× 10−4 pm −1 ⎞
f (3.00 nm) = 4 π × ⎜ ⎟⎠
expressed in terms of its projections on each of three orthog- ⎝ π1/2
onal axes as l 2 = l x2 + l 2y + lz2 . The average values of l x2 , l 2y , and lz2 × (3.00 × 103 pm)2 × e −(2.58…×10
−4
pm −1 )2 (3.00×103 pm )2

are all the same in a spherically symmetric environment, so = 1.92 × 10−4 pm −1


the average length of a step in the x-direction (or any of the
other two directions) can be obtained by writing l 2 = 3〈l x2 〉, and The probability that the ends will be found in a narrow range
is x = 〈l x2 〉1/2 = l /31/2 . The probability that the random walk will of width δr = 10.0 pm at 3.00 nm (regardless of direction) is
end up at a distance x from the origin is given by eqn 17A.1 therefore
with n = x/(l/31/2) = 31/2x/l:
f (3.00 nm)δr = (1.92 ×10−4 pm −1) × (10.0 pm) = 1.92 ×10−3 v
1/2
⎛ 2 ⎞
P(x ) = ⎜ e −3 x
2
/2 Nl 2 or about 1 in 5200.
⎝ πN ⎟⎠
17A The structures of macromolecules  701

Root-mean-square separation, Rrms/l


Self-test 17A.2 What is the probability that the ends of a poly-
ethene chain of molar mass 85 kg mol−1 are between 15.0 nm 60
and 15.1 nm apart when the polymer is treated as a three-
dimensional freely jointed chain?
40
Answer: 5.9 × 10 −3

20
There are several measures of the geometrical size of a random
coil. The contour length, Rc , is the length of the macromolecule
0
measured along its backbone from atom to atom. For a polymer 0 1000 2000 3000 4000
of N monomer units each of length l, the contour length is Number of monomers, N

Figure 17A.8 The variation of the root-mean-square


Rc = Nl Random coil Contour length (17A.3)
separation of the ends of a three-dimensional random coil, Rrms,
The root-mean-square separation, Rrms, is a measure of the with the number of monomers.
average separation of the ends of a random coil: it is the square
root of the mean value of R2. To determine its value we note same moment of inertia (and therefore rotational characteris-
that the vector joining the two ends of the chain is the vector tics) as the actual molecule of the same mass. We show in the
N
sum of the vectors joining neighbouring monomers: R = ∑ ri following Justification that
i =1
(Fig. 17A.7). The mean square separation of the ends of the One-dimensional Radius of
Rg = N 1/2 l random coil gyration (17A.5)
chain is therefore
A similar calculation for a three-dimensional random coil gives
N N
2
l N
〈 R2 〉 = 〈 R⋅ R〉 = ∑ 〈ri ⋅rj 〉 = ∑ 〈ri2 〉 + ∑ 〈ri ⋅rj 〉
1/2
i. j i i≠ j ⎛N⎞ Three-dimensional Radius of
Rg = ⎜ ⎟ l random coil (17A.6)
⎝ 6⎠ gyration
When N is large (which we assume throughout) the second
sum (in blue) vanishes because the individual vectors all lie The radius of gyration is smaller in this case because the extra
in random directions. The remaining sum is equal to Nl2 as dimensions enable the coil to be more compact. The radius
all bond lengths are the same (and equal to l); so, after taking of gyration may also be calculated for other geometries. For
square roots to obtain Rrms = 〈R2〉1/2, we conclude that example, a solid uniform sphere of radius R has Rg =( 35 )1/2R, and
a long thin uniform rod of length l has Rg = l/(12)1/2 for rota-
Rrms = N 1/2 l Random coil Root-mean-square separation (17A.4) tion about an axis perpendicular to the long axis. A solid sphere
with the same radius and mass as a random coil has a greater
We see that, as the number of monomer units increases, the radius of gyration as it is entirely dense throughout.
root-mean-square separation of its end increases as N1/2 (Fig.
17A.8), and consequently its volume increases as N3/2. The Brief illustration 17A.3 Measures of size of a random coil
result must be multiplied by a factor when the chain is not
freely jointed (see next section). With a powerful microscope it is possible to see that a long
Another convenient measure of size is the radius of gyra- piece of DNA is flexible and writhes as if it were a random coil.
tion, Rg, which is the radius of a hollow sphere that has the However, small segments of the macromolecule resist bend-
ing, so it is more appropriate to visualize DNA as a freely
N jointed chain with N and l as the number and length, respec-
R rN–1
tively, of these rigid units. The length l of a rigid unit is approx-
imately 45 nm. It follows that for a piece of DNA with N = 200,
r3 we estimate (by using 103 nm = 1 μm)
3
r2 From eqn 17A.3: Rc = 200 × 45 nm = 9.0 μm
2 r1
From eqn 17A.4: R rms = (200)1/2 × 45 nm = 0.64 μm
1 1/2
From eqn 17A.6: Rg = ⎛⎜ 200 ⎞⎟ × 45 nm = 0.26 μm
Figure 17A.7 A schematic illustration of the calculation of the ⎝ 6 ⎠
root-mean-square separation of the ends of a random coil.
702 17 Macromolecules and self-assembly

1/2
Self-test 17A.3 Calculate the contour and root-mean-square ⎛ 1 − cosθ ⎞
F =⎜
⎝ 1 + cosθ ⎟⎠
(17A.7)
lengths of a polymer chain modelled as a random coil with
N = 1000 and l = 150 pm.
For tetrahedral bonds, for which cosθ = 13 (that is, θ = 109.5°),
Answer: Rc = 150 nm, R rms = 4.74 nm
F = 21/2. Therefore:
1/2
⎛N⎞
Rrms = (2N )1/2 l Rg = ⎜ ⎟ l
Justification 17A.3 ⎝ 3⎠
The radius of gyration
Dimensions of a tetrahedrally constrained chain (17A.8)
For a one-dimensional random coil with N + 1 identical mon-
omers (and therefore N bonds) each of mass m, the moment of
The model of a randomly coiled molecule is still an approxima-
inertia around the centre of the chain (which is also at the first
tion, even after the bond angles have been restricted, because
monomer, because steps occur in equal numbers to left and
it does not take into account the impossibility of two or more
right) is
atoms occupying the same place. Such self-avoidance tends to
N N swell the coil, so (in the absence of solvent effects) it is better to
I= ∑
i =0
mi ri2 = m ∑r
i =0
i
2
regard Rrms and Rg as lower bounds to the actual values.

This moment of inertia is set equal to mtot Rg2 , where mtot is (c) Partly rigid coils
the total mass of the polymer, mtot = (N + 1)m. Therefore, after
averaging over all conformations, An important measure of the flexibility of a chain is the persis-
tence length, lp, a measure of the length over which the direc-
N

∑ 〈r 〉
1 tion of the first monomer–monomer direction is sustained. If
Rg2 = 2
N +1 i
the chain is a rigid rod, then the persistence length is the same
i =0
as the contour length. For a freely-jointed random coil, the per-
For a linear random chain, 〈ri2 〉 = Nl 2 , and as there are N + 1 sistence length is just the length of the monomer–monomer
such terms in the sum, we find bond. Therefore, the persistence length can be regarded as a
measure of the stiffness of the chain. In general, the persistence
Rg2 = Nl 2
length of a chain of identical monomers of length l is defined as
Equation 17A.5 then follows after taking the square root of the average value of the projection of the end-to-end vector on
each side. the first bond of the chain (Fig. 17A.9):
N −1
r1
∑ 〈r ⋅ r 〉
1
lp = ⋅R = 1 i Definition Persistence length (17A.9)
The random coil model ignores the role of the solvent: a poor l l
i =1
solvent will tend to cause the coil to tighten so that solute–
solvent contacts are minimized; a good solvent does the oppo- (The sum ends at N − 1 because the last atom is atom N and the
site. Therefore, calculations based on this model are better last bond is from atom N − 1 to atom N.) Experimental values of
regarded as lower bounds to the dimensions for a polymer in persistence lengths are as follows:
a good solvent and as an upper bound for a polymer in a poor
poly(glycine) poly(l-alanine) poly(l-proline)
solvent. The model is most reliable for a polymer in a bulk solid
sample, where the coil is likely to have its natural dimensions. 0.6 nm 2 nm 22 nm
These values suggest that the stiffness of the chain increases
from left to right along the series.
(b) Constrained chains The mean square distance between the ends of a chain that
The freely jointed chain model is improved by removing the has a persistence length greater than the monomer length can
freedom of bond angles to take any value. For long chains, we be expected to be greater than for a random coil because the
can simply take groups of neighbouring bonds and consider partial rigidity of the coil does not let it roll up so tightly. We
the direction of their resultant. Although each successive indi- show in the following Justification that
vidual bond is constrained to a single cone of angle θ relative
1/2
to its neighbour, the resultant of several bonds lies in a random ⎛ 2lp ⎞
Rrms = N 1/2 lF where F = ⎜ −1⎟ (17A.10)
direction. By concentrating on such groups rather than indi- ⎝ l ⎠
viduals, it turns out that for long chains the expressions for the
root-mean-square separation and the radius of gyration given For a random coil, lp = l, so Rrms = N1/2l, as we have already
above should be multiplied by found. For lp > l, F > 1, so the coil has swollen, as we anticipated.
17A The structures of macromolecules  703

N
Method When lp = l, the chain is a random coil. From eqn
R 17A.4, write the root-mean-square separation of the ends of
the chain in the random coil limit as R rms,random coil = N1/2l. Then
it follows from eqn 17A.10 that the root-mean-square separa-
lP tion of the ends of the chain, R rms, of the partly rigid coil with
l
persistence length lp is
r1 1/2
⎛ 2lp ⎞
Rrms = Rrms ,random coil ⎜ −1⎟
1 ⎝ l ⎠
Figure 17A.9 The persistence length is defined as the average
From eqn 17A.3, we write R c = Nl and we are given that
value of the projection of the end-to-end vector on the first
lp = 0.025Rc, which can therefore be interpreted as 0.025Nl. From
bond of the chain.
these expressions, calculate the fractional change in the root-
mean-square separation, (Rrms − Rrms ,random coil )/ Rrms ,random coil ,
Justification 17A.4 Partly rigid coils and express the result as a percentage.

In each of the following steps we use N → ∞ when necessary. Answer We write the fractional change in the root-mean-
We start from square separation as


l2 Rrms − Rrms ,random coil Rrms
N
= −1
〈 R2 〉 = ∑ 〈ri2 〉 + ∑
i≠ j
〈 ri ⋅ rj 〉 Rrms ,random coil Rrms ,random coil
1/2
i
⎛ 2lp ⎞
=⎜ −1⎟ −1
The first term on the right is Nl 2 regardless of the rigidity of ⎝ l ⎠
1/2
the coil. The second term can be written as follows: ⎛ 2 × 0.025Nl ⎞
=⎜ −1⎟ −1
N N N
⎝ l ⎠

∑〈r ⋅r 〉 = 2∑〈r ⋅r 〉 + 2∑〈r ⋅r 〉 +…


i≠ j
i j
i =2
1 i
i =3
2 i = (0.050N − 1) −1
1/2

With N = 1000, the fractional change is 6.00, so the root-mean-


There are N − 1 such terms, and provided we allow N to become
square separation increases by 600 per cent.
infinite, all the sums on the right have the same value, so
Self-test 17A.4 Calculate the fractional change in the volume
N N N

∑〈r ⋅r 〉 = 2(N −1)∑〈r ⋅r 〉 ≈ 2N ∑〈r ⋅r 〉


i≠ j
i j
i =2
1 i
i =2
1 i
of the same coil.
Answer: 340

The final (blue) sum on the right is close to being the square of
the persistence length. Specifically, from eqn 17A.9,
N N 17A.3 Biological macromolecules
∑j =2
〈 r1 ⋅ rj 〉 = ∑
j =1
〈 r1 ⋅ rj 〉 − 〈r12 〉 = llp − l 2
A protein is a polypeptide composed of linked α-amino acids,
NH2CHRCOOH, where R is one of about 20 groups. For a
Now we bring the three pieces of the calculation together:
protein to function correctly, it needs to have the correct con-
〈 R 2 〉 = Nl 2 + 2N (llp − l 2 ) = 2Nllp − Nl 2 formation. For example, an enzyme has its greatest catalytic
efficiency only when it is in a specific conformation. The amino
which, on taking the square root of both sides, is eqn 17A.10. acid sequence of a protein contains the necessary information
to create the active conformation of the protein as it is formed.
However, the prediction of the observed conformation from
the primary structure, the so-called protein folding problem, is
Example 17A.1 Calculating the root-mean-square extraordinarily difficult and is still the focus of much research.
separation of a partly rigid coil The other class of biological macromolecules we consider are
By what percentage does the root-mean-square separation of the nucleic acids, which are key components of the mechanism
the ends of a polymer chain with N = 1000 increase or decrease of storage and transfer of genetic information in biological
when the persistence length changes from l (the bond length) cells. Deoxyribonucleic acid (DNA) contains the instructions
to 2.5 per cent of the contour length? for protein synthesis, which is carried out by different forms of
ribonucleic acid (RNA).
704 17 Macromolecules and self-assembly

(a) Proteins
The origin of the secondary structures of proteins is found in ψ
the rules formulated by Linus Pauling and Robert Corey in
1951 that seek to identify the principal contributions to the φ
lowering of energy of the molecule by focusing on the role of
hydrogen bonds and the peptide link, eCONHe. The latter
can act both as a donor of the H atom (the NH part of the link)
and as an acceptor (the CO part). The Corey–Pauling rules are
as follows (Fig. 17A.10):
Figure 17A.11 The definition of the torsional angles ψ and ϕ
1. The four atoms of the peptide link lie in a relatively rigid between two peptide units. In this case (an α-L-polypeptide)
plane. the chain has been drawn in its all-trans form, with ψ = ϕ = 180°.
The planarity of the link is due to delocalization of π electrons
over the O, C, and N atoms and the maintenance of maximum ψ are equal. For a right-handed helix, an α helix (Fig. 17A.12),
overlap of their p orbitals. all ϕ = 57° and all ψ = −47°. For a left-handed helix, both angles
2. The N, H, and O atoms of a hydrogen bond lie in a are positive. The torsional contribution to the total potential
straight line (with displacements of H tolerated up to not energy is
more than 30° from the NeO direction).
Vtorsion = A(1+ cos 3φ ) + B(1+ cos 3ψ ) (17A.11)
3. All NH and CO groups of the peptide links are engaged
in hydrogen bonding.
in which A and B are constants of the order of 1 kJ mol−1.
The rules are satisfied by two structures. One, in which hydro- Because only two angles are needed to specify the conformation
gen bonding between peptide links leads to a helical structure, is of a helix, and they range from −180° to +180°, the torsional
a helix, which can be arranged as either a right-or a left-handed potential energy of the entire molecule can be represented on a
screw. The other, in which hydrogen bonding between peptide Ramachandran plot, a contour diagram in which one axis rep-
links leads to a planar structure, is a sheet; this form is the sec- resents ϕ and the other represents ψ.
ondary structure of the protein fibroin, the constituent of silk. Figure 17A.13 shows the Ramachandran plots for the heli-
Because the planar peptide link is relatively rigid, the geom- cal form of polypeptide chains formed from the non-chiral
etry of a polypeptide chain can be specified by the two angles amino acid glycine (R = H) and the chiral amino acid l-ala-
that two neighbouring planar peptide links make to each other. nine (R]CH3). The glycine map is symmetrical, with minima
Figure 17A.11 shows the two angles ϕ and ψ commonly used to of equal depth at ϕ = −80°, ψ = +90° and at ϕ = +80°, ψ = −90°.
specify this relative orientation. The sign convention is that a In contrast, the map for l-alanine is unsymmetrical, and there
positive angle means that the front atom must be rotated clock- are three distinct low-energy conformations (marked I, II, III).
wise to bring it into an eclipsed position relative to the rear The minima of regions I and II lie close to the angles typical of
atom. For an all-trans form of the chain, all ϕ and ψ are 180°.
A helix is obtained when all the ϕ are equal and when all the

R
Rotational
freedom
H H
116°
C 151 132.5 102
N
C N
124 145.5
O 122°
123.5°
O
Rotational
R freedom

Figure 17A.10 The dimensions that characterize the peptide Figure 17A.12 The polypeptide α helix, with poly-L-glycine as
link. The CeNHeCOeC atoms define a plane (the CeN bond an example. There are 3.6 residues per turn, and a translation
has partial double-bond character), but there is rotational along the helix of 150 pm per residue, giving a pitch of 540 pm.
freedom around the CeCO and NeC bonds. The diameter (ignoring side chains) is about 600 pm.
17A The structures of macromolecules  705

180° CeO bond of one chain is aligned with a CeO bond of another
chain. These structures are not common in proteins.
Covalent and non-covalent interactions may cause poly-
peptide chains with well-defined secondary structures to fold
ψ into tertiary structures. Although the rules that govern protein
folding are still being elucidated, a few general conclusions
may be drawn from X-ray diffraction studies of water-soluble
natural proteins and synthetic polypeptides. In an aqueous
–180° environment, the chains fold in such a way as to place nonpolar
–180° φ 180° –180° φ 180°
(a) (b)
R groups in the interior (which is often not very accessible to
solvent) and charged R groups on the surface (in direct con-
Figure 17A.13 Contour plots of potential energy against the tact with the polar solvent). Other factors that promote the
angles ψ and ϕ, also known as a Ramachandran diagram, for folding of proteins include covalent disulfide (eSeSe) links,
(a) a glycyl residue of a polypeptide chain and (b) an alanyl Coulombic interactions between ions (which depend on the
residue. The glycyl diagram is symmetrical, but that for alanyl degree of protonation of groups and therefore on the pH), van
is unsymmetrical and the global minimum corresponds to an der Waals interactions, and hydrophobic interactions (Topic
α-helix. (T. Hovmoller et al., Acta Cryst. D58, 768 (2002).) 16B). The clustering of nonpolar, hydrophobic, amino acids
into the interior of a protein is driven primarily by hydrophobic
right- and left-handed helices, but the former has a lower mini- interactions.
mum. This result is consistent with the observation that poly-
peptides of the naturally occurring l-amino acids tend to form
right-handed helices.
(b) Nucleic acids
A β-sheet (also called the β-pleated sheet) is formed by Both DNA and RNA are polynucleotides (1), in which base–
hydrogen bonding between two extended polypeptide chains sugar–phosphate units are linked by phosphodiester bonds.
(large absolute values of the torsion angles ϕ and ψ). In an anti- In RNA the sugar is β-d-ribose and in DNA it is β-d-2-
parallel β-sheet (Fig. 17A.14a), ϕ = 139°, ψ = 113°, and the deoxyribose (as shown in 1). The most common bases are
NeH…O atoms of the hydrogen bonds form a straight line. This adenine (A, 2), cytosine (C, 3), guanine (G, 4), thymine (T,
arrangement is a consequence of the antiparallel arrangement of found in DNA only, 5), and uracil (U, found in RNA only, 6).
the chains: every NeH bond on one chain is aligned with a CeO At physiological pH, each phosphate group of the chain carries
bond from another chain. Antiparallel β-sheets are very com- a negative charge and the bases are deprotonated and neutral.
mon in proteins. In a parallel β-sheet (Fig. 17A.14b), ϕ = 119°, This charge distribution leads to two important properties. One
ψ = 113°, and the NeH…O atoms of the hydrogen bonds are is that the polynucleotide chain is a polyelectrolyte, a macro-
not perfectly aligned. This arrangement is a result of the paral- molecule with many different charged sites, with a large and
lel arrangement of the chains: each NeH bond on one chain is negative overall surface charge. The second is that the bases can
aligned with a NeH bond of another chain and, as a result, each interact by hydrogen bonding, as shown for AeT (7) and CeG
base pairs (8). The secondary and tertiary structures of DNA
and RNA arise primarily from the pattern of this hydrogen
bonding between bases of one or more chains.

O Base
O NH2
H H NH2
(a) H H
O R N N
N

O P O NH O
NH N
O Base
O 2 Adenine, A 3 Cytosine, C
H H
(b) H H
O R O O

Figure 17A.14 The two types of β-sheets: (a) antiparallel O P O N
NH NH
(ϕ = −139°, ψ = 113°), in which the NeHeO atoms of the O
hydrogen bonds form a straight-line; (b) parallel (ϕ = −119°, NH N NH2 NH O
ψ = 113°) in which the NeHeO atoms of the hydrogen bonds 1 D-ribose (R = OH)
are not perfectly aligned. and 2′-deoxy-D-ribose (R = H) 4 Guanine, G 5 Thymine, T
706 17 Macromolecules and self-assembly

H
N N H O C
O A
G
NH N N H N
G
N N
A T
NH O O
C
6 Uracil 7 A-T base pair T

H
N H O N
Figure 17A.15 The DNA double helix, in which two
N polynucleotide chains are linked together by hydrogen bonds
N H N
N
between adenine (A) and thymine (T) and between cytosine (C)
N
and guanine (G).
O H N
H
8 C-G base pair parallel to each other and perpendicular to the axis of the helix.
The structure is stabilized further by interactions between the
In DNA, two polynucleotide chains wind around each planar π systems of the bases. In B-DNA, the most common
other to form a double helix (Fig. 17A.15). The chains are held form of DNA found in biological cells, the helix is right-handed
together by links involving AeT and CeG base pairs that lie with a diameter of 2.0 nm and a pitch of 3.4 nm.

Checklist of concepts
☐ 1. The primary structure of a macromolecule is the ☐ 8. The secondary structure of a protein is the spatial
sequence of small molecular residues making up the arrangement of the polypeptide chain and includes
polymer. helices and the β-sheet.
☐ 2. The secondary structure is the spatial arrangement of a ☐ 9. Helical and sheet-like polypeptide chains are folded
chain of residues. into a tertiary structure by bonding influences between
☐ 3. The tertiary structure is the overall three-dimensional the residues of the chain.
structure of a macromolecule. ☐ 10. Some proteins have a quaternary structure as aggre-
☐ 4. The quaternary structure is the manner in which large gates of two or more polypeptide chains.
molecules are formed by the aggregation of others. ☐ 11. In DNA, two polynucleotide chains held together by
☐ 5. In a freely jointed chain any bond in a polymer is free hydrogen bonded base pairs wind around each other to
to make any angle with respect to the preceding one. form a double helix.
☐ 6. The freely jointed chain model is improved by remov- ☐ 12. In RNA, single chains fold into complex structures by
ing the freedom of bond angles to take any value. formation of specific base pairs.
☐ 7. The least structured conformation of a macromolecule
is a random coil, which can be modelled as a freely
jointed chain.

Checklist of equations
Property Equation Comment Equation number

P = (2/πN )1/2 e −n /2 N
2
Probability distribution One-dimensional random coil 17A.1

f (r ) = 4 π(a /π1/2 )3 r 2 e − a r
2 2
Three-dimensional random coil 17A.2
a = ( 23 Nl 2 )1/2

Contour length of a random coil Rc = Nl 17A.3


17A The structures of macromolecules  707

Property Equation Comment Equation number

Root-mean-square separation of a random coil Rrms = N1/2l Unconstrained chain 17A.4

Radius of gyration of a random coil Rg = N1/2l Unconstrained one-dimensional chain 17A.5

Rg = (N/6)1/2l Unconstrained three-dimensional chain 17A.6

Root-mean-square separation of a random coil Rrms = (2N)1/2l Constrained tetrahedral chain 17A.8

Radius of gyration of a random coil Rg = (N/3)1/2l Constrained tetrahedral chain 17A.8


17B Properties of macromolecules

Contents (a) Conformational entropy


17B.1 Mechanical properties 708 The random coil is the least structured conformation of a poly-
(a) Conformational entropy 708 mer chain and corresponds to the state of greatest entropy. Any
Brief illustration 17B.1: Conformational entropy 708 stretching of the coil introduces order and reduces the entropy.
(b) Elastomers 709 Conversely, the formation of a random coil from a more
Brief illustration 17B.2: The restoring force 710 extended form is a spontaneous process (provided enthalpy
17B.2 Thermal properties 710 contributions do not interfere). As shown in the following
Example 17B.1: Predicting the melting Justification, we can use the same model to deduce that the
temperature of DNA 711 change in conformational entropy, the statistical entropy aris-
17B.3 Electrical properties 712 ing from the arrangement of bonds, when a one-dimensional
Checklist of concepts 712 chain containing N bonds of length l is stretched or compressed
Checklist of equations 713 by nl is

ΔS = − 12 kN ln{(1+ ␯)1+␯ (1− ␯)1−␯ } Random Conformational


coil entropy (17B.1)
␯ =n/N
➤ Why do you need to know this material?
Macromolecules are important in modern technology. They This function is plotted in Fig. 17B.1, and we see that minimum
are also building blocks of biological cells. To understand extension corresponds to maximum entropy.
why this is so, you need to explore the characteristic
physical properties of macromolecules.
Brief illustration 17B.1 Conformational entropy
➤ What is the key idea?
Suppose that N = 1000 and l = 150 pm. The change in entropy
The unique properties of macromolecules are related to
when the (one-dimensional) random coil is stretched through
their unique structural features.
1.5 nm (corresponding to n = 10 and ν = 1/100) is
➤ What do you need to know already?
⎪⎧⎛ ⎪⎫
1+(1/100) 1−(1/100)
1 ⎞ ⎛ 1 ⎞
You need to be familiar with the structural features of ΔS = − 12 k × (1000) × (ln ⎨⎜ 1 + ⎟⎠ ⎜⎝ 1 − 100 ⎟⎠ ⎬
⎝ 100
macromolecules (Topic 17A), particularly the properties ⎩⎪ ⎭⎪
of a random coil. You also need to be familiar with the = −0.050k
statistical interpretation of entropy (Topic 15E) and the
Because R = NAk, the change in molar entropy is ΔSm = −0.050R,
concept of internal energy (Topic 2A). or −0.42 J K−1 mol−1.
Self-test 17B.1 What is the change in conformational entropy
Macromolecules have special physical properties that arise when the same random coil is stretched from fully coiled by
from details of their structures. In this Topic we explore the 10 per cent?
physical mechanical, thermal, and electrical properties of syn- Answer: −0.042 J K−1 mol−1

thetic and biological macromolecules.


Justification 17B.1 The conformational entropy of a
17B.1 Mechanical properties freely jointed chain
The conformational entropy of the chain is given by the
Significant insight into the consequences of stretching and con- Boltzmann formula, S = k ln W (eqn 15E.7). In the present
tracting a polymer can be obtained on the basis of the freely case, we identify W with the number of ways of achieving a
jointed chain as a model (Topic 17A). coil with a given extension, the W calculated in eqn 17A.2:
17B Properties of macromolecules  709

N!
W=
{ 12 (N + n)}!{ 12 (N − n)}!

Stress
Plastic
deformation
Therefore, Yield
point
S /k = ln N !− ln { 12 (N + n)}!− ln { 12 (N − n)}!
Elastic
Because the factorials are large (except for large exten- deformation
sions), we can use Stirling’s approximation (Topic 15A,
ln x ! ≈ ( x + 12 ) ln x − x + 12 ln 2π ) to obtain Strain

Figure 17B.2 A typical stress–strain curve.


S /k = ln(2π)1/2 + ( N + 1) ln 2 − ( N + 12 ) ln N −
1
2 ln{(N + n)N +n+1(N + n)N −n+1}
The stress–strain curve in Fig. 17B.2 shows how a material
We have seen that the most probable conformation of a one-
responds to stress. The region of elastic deformation is where
dimensional chain is the one with the ends close together
the strain is proportional to the stress and is reversible: when
(n = 0). This conformation also corresponds to maximum
entropy, as may be confirmed by differentiation. Therefore, the
the stress is removed, the sample returns to its initial shape. As
maximum entropy is we see in more detail in Topic 18C, the slope of the stress–strain
curve in this region is ‘Young’s modulus’, E, for the material. At
S /k = ln(2π)1/2 + (N + 1)ln 2 + 12 ln N the yield point, the reversible, linear deformation gives way to
plastic deformation, where the strain is no longer linearly pro-
The change in entropy when the chain is stretched or com- portional to the stress and the initial shape of the sample is not
pressed by nl is therefore the difference of these two quantities, recovered when the stress is removed. Thermosetting plastics
and the resulting expression, after some algebraic manipula- have only a very short elastic range; thermoplastics typically
tion, is eqn 17B.1. (but not universally) have a long plastic range. An elastomer is
specifically a polymer with a long elastic range. They typically
have numerous cross-links (such as the sulfur links in vulcan-
0
ized rubber) that pull them back into their original shape when
the stress is removed.
Change of entropy, ΔS/Nk

–0.1 Although practical elastomers are typically extensively cross-


linked, even a freely jointed chain behaves as an elastomer for
small extensions. It is a model of a perfect elastomer, a polymer
–0.2
in which the internal energy is independent of the extension. In
the following Justification we also see that the restoring force, F,
–0.3 of a one-dimensional random coil when the chain is stretched
or compressed by nl is
–0.4 One-
kT 1+ ␯ Restoring
–0.8 –0.4 0 0.4 0.8 F= ln ␯ = n /N dimensional (17B.2a)
Extension, ν = n/N 2l 1− ␯ random coil
force

Figure 17B.1 The change in molar entropy of a perfect where N is the total number of bonds each of length l. This
elastomer as its extension changes; ν = 1 corresponds to function is plotted in Fig. 17B.3. At low extensions, when ␯ << 1
complete extension; ν = 0, the conformation of highest entropy, we can use ln (1 + x) = x – 12 x 2 + … and find (retaining only linear
corresponds to a random coil. terms) that

␯kT nkT One-dimensional


(b) Elastomers F≈ = random coil, small
Restoring
force (17B.2b)
l Nl extensions
The fundamental concepts for the discussion of the mechani-
cal properties of solids are stress and strain. The stress on an That is, for small displacements the sample obeys Hooke’s law:
object is the applied force divided by the area to which it is the restoring force is proportional to the displacement (which
applied. The strain is the resulting distortion of the sample. The is proportional to n). For small displacements, therefore, the
general field of the relations between stress and strain is called whole coil shakes with simple harmonic motion. When this
rheology. equation is rearranged to
710 17 Macromolecules and self-assembly

6 In a perfect elastomer, as in a perfect gas, the internal energy


is independent of the dimensions (at constant temperature), so
4 Hooke’s law
Restoring force, F/(kT/2l)

(∂U/∂x)T = 0. The restoring force is therefore


2
⎛ ∂S ⎞
F = −T ⎜ ⎟
0
⎝ ∂x ⎠ T

–2
If now we substitute eqn 17B.1 into this expression, we obtain

T ⎛ ∂S ⎞ T ⎛ ∂S ⎞ kT ⎛ 1 + ␯ ⎞
F =− = = ln
l ⎜⎝ ∂n ⎟⎠ T Nl ⎜⎝ ∂␯ ⎟⎠ T 2l ⎜⎝ 1 − ␯ ⎟⎠
–4

–6
–1 –0.5 0 0.5 1 as in eqn 17B.2a.
Extension, ν = n/N

Figure 17B.3 The restoring force, F, of a one-dimensional


perfect elastomer. For small strains, F is linearly proportional to
the extension, corresponding to Hooke’s law.
17B.2 Thermal properties
The crystallinity of synthetic polymers can be destroyed by ther-
mal motion at sufficiently high temperatures. This change in
⎛ Nl 2 ⎞ crystallinity may be thought of as a kind of intramolecular melt-
nl = ⎜ F (17B.2c)
⎝ kT ⎟⎠ ing from a crystalline solid to a more fluid random coil. Polymer
melting also occurs at a specific melting temperature, Tm, which
we see that for small displacements, the strain, as measured increases with the strength and number of intermolecular inter-
by the extension nl, is proportional to the applied force, actions in the material. Thus, polyethene, which has chains that
as is characteristic of the elastic deformation region of an interact only weakly in the solid, has Tm = 414 K and nylon-66
elastomer. fibre, in which there are strong hydrogen bonds between chains,
has Tm = 530 K. High melting temperatures are desirable in most
Brief illustration 17B.2 The restoring force practical applications involving fibres and plastics.
All synthetic polymers undergo a transition from a state of
Consider a polymer chain with N = 5000 and l = 0.15 nm. If high to low chain mobility at the glass transition temperature,
the ends of the chain are moved apart by 1.5 nm, then
Tg. To visualize the glass transition, we consider what happens
n = (1.5 nm)/(0.15 nm) = 10 and ν = n/N = 10/5000 = 2.0 × 10 −3.
to an elastomer as we lower its temperature. There is sufficient
Because ν ≪ 1, the restoring force at 293 K is given by eqn
energy available at normal temperatures for limited bond rota-
17B.2b as
tion to occur and the flexible chains writhe. At lower tempera-
N
m
tures, the amplitudes of the writhing motion decrease until a
10 × (1.381 ×10−23 J K −1) × (293 K)
F= = 5.4 × 10−14 N specific temperature, Tg, is reached at which motion is frozen
5000 × (1.5 × 10−10 m) completely and the sample forms a glass. Glass transition tem-
peratures well below 300 K are desirable in elastomers that are
or 54 fN.
to be used at normal temperatures. Both the glass transition
Self-test 17B.2 Repeat the calculation for N = 6.0 × 103 and a temperature and the melting temperature of a polymer may
displacement of 2.0 nm at 298 K. be measured by calorimetric methods. Because the motion of
Answer: 61 fN the segments of a polymer chain increase at the glass transition
temperature, Tg may also be determined from a plot of the spe-
cific volume of a polymer (the reciprocal of its mass density)
Justification 17B.2 Hooke’s law against temperature (Fig. 17B.4).
Proteins and nucleic acids are relatively unstable towards
The work done on an elastomer when it is extended through a chemical and thermal denaturation, the loss of structure.
distance dx is Fdx, where F is the restoring force. The change Thermal denaturation is similar to the melting of synthetic
in internal energy is therefore polymers. Denaturation is a cooperative process in the sense
dU = TdS + Fdx that the biopolymer becomes increasingly more susceptible
to denaturation once the process begins. This cooperativity is
It follows that
observed as a sharp step in a plot of fraction of unfolded poly-
⎛ ∂U ⎞ ⎛ ∂S ⎞ mer against temperature. The melting temperature, Tm, is the
⎜⎝ ∂x ⎟⎠ = T ⎜⎝ ∂x ⎟⎠ + F temperature at which the fraction of denatured polymer is 0.5
T T
17B Properties of macromolecules  711

Example 17B.1 Predicting the melting temperature


of DNA
Specific volume, Vs

The melting temperature of a DNA molecule (in the sense of


the temperature at which it undergoes denaturation) can be
determined by calorimetric methods. The following data were
obtained in 0.010 m Na3PO4(aq) for a series of DNA molecules
with varying base pair composition, with f the fraction of
Tg GeC base pairs:

f 0.375 0.509 0.589 0.688 0.750


Temperature, T Tm/K 339 344 348 351 354
Figure 17B.4 The variation of specific volume with Estimate the melting temperature of a DNA molecule contain-
temperature of a synthetic polymer. The glass transition ing 40.0 per cent GeC base pairs.
temperature, Tg, is at the point of intersection of extrapolations
Method Look for a quantitative relationship between the
of the two linear parts of the curve.
melting temperature and the composition of DNA. Begin by
plotting Tm against fraction of GeC base pairs and examining
(Fig. 17B.5). For example, Tm = 320 K for ribonuclease T1 (an the shape of the curve. If visual inspection of the plot suggests
enzyme that cleaves RNA in the cell), which is not far above a linear relationship, then the melting point at any composi-
the temperature at which the enzyme must operate (close to tion can be predicted from the equation of the straight line
body temperature, 310 K). More surprisingly, the Gibbs energy that fits the data.
for the denaturation of ribonuclease T1 at pH = 7.0 and 298 K is
only 19.5 kJ mol−1, which is comparable to the energy required Answer Figure 17B.6 shows that Tm varies linearly with the
to break a single hydrogen bond (about 20 kJ mol−1). The sta- fraction of GeC base pairs, at least in this range of composi-
bility of a protein does not increase in a simple way with the tion. The equation of the line that fits the data is
number of hydrogen bonding interactions. While the reasons Tm /K = 325 + 39.7f
for the low stability of proteins are not known, the answer prob-
ably lies in a delicate balance of the intra- and intermolecular It follows that Tm = 341 K for 40.0 per cent GeC base pairs (at
interactions that allow a protein to fold into its active confor- f = 0.400). The thermal stability of DNA increases with the
mation, and the role of the aqueous environment. By contrast, number of CeG base pairs in the sequence because each GeC
the melting temperature of DNA can be predicted with reason- base pair has three hydrogen bonds, whereas each TeA base
able accuracy by examining its structure, as we see in the fol- pair has only two (Topic 17A). More energy, and therefore a
lowing Example. higher temperature, is required to unravel a double helix that

1 360
Melting temperature, Tm/K
Fraction of unfolded protein

350

0.5

340

Tm 330
0 0.4 0.6 0.8
0
Temperature, T Fraction of GC base pairs

Figure 17B.5 A protein unfolds as the temperature of the Figure 17B.6 Data for Example 17B.1 showing the variation
sample increases. The sharp step in the plot of fraction of of the melting temperature of DNA molecules with the
unfolded protein against temperature indicated that the fraction of GdC base pairs. All the samples also contain
transition is cooperative. The melting temperature, Tm, is the 1.0 × 10−2 mol dm−3 Na3PO4.
temperature at which the fraction of unfolded polymer is 0.5.
712 17 Macromolecules and self-assembly

has a higher proportion of hydrogen bonding interactions per Oxidation


.
base pair.
+ .
A note on good practice In this example we do not have
a good theory to guide us in the choice of a mathemati- + .
cal model to describe the behaviour of the system over
a wide range of conditions. We are limited to finding a +
purely empirical relation—in this case a simple first-order
Figure 17B.7 The mechanism of migration of a partially
polynomial equation—that fits the available data. It follows
localized cation radical, or polaron, in polyacetylene.
that we should not attempt to predict the property of a
system that falls outside the narrow range of the data used
to generate the fit because the mathematical model may
conduction along the polymer chain. The 2000 Nobel Prize in
have to be enhanced (for example, by using higher-order
chemistry was awarded to A.J. Heeger, A.G. MacDiarmid, and
polynomial equations) to describe the system over a wider
range of conditions. In the present case, we should not
H. Shirakawa for their pioneering work in the synthesis and
attempt to predict the Tm of DNA molecules outside the characterization of conducting polymers.
range 0.375 < f < 0.750. One example of a conducting polymer is polyacetylene
(polyethyne, Fig. 17B.7). Whereas the delocalized π bonds do
Self-test 17B.3 The following calorimetric data were obtained suggest that electrons can move up and down the chain, the
in solutions containing 0.15 m NaCl(aq) for the same series of electrical conductivity of polyacetylene increases significantly
DNA molecules studied in Example 17B.1. Estimate the melt- when it is partially oxidized by I2 and other strong oxidants.
ing temperature of a DNA molecule containing 40.0 per cent The product is a polaron, a partially localized cation radical
GeC base pairs under these conditions.
that travels virtually (by exchanging its identity with a neigh-
f 0.375 0.509 0.589 0.688 0.750 bour) through the chain, as shown in Fig. 17B.7. Further oxida-
Tm/K 359 364 368 371 374 tion of the polymer forms either bipolarons, a di-cation that
moves virtually (in a similar way) as a unit through the chain,
Answer: 360 K or solitons, two separate cation radicals that move indepen-
dently. Polarons and solitons contribute to the mechanism of
charge conduction in polyacetylene.
Conducting polymers are slightly better electrical conduc-
17B.3 Electrical properties tors than silicon semiconductors but are far worse than metal-
lic conductors. They are currently used in a number of devices,
Most of the macromolecules and self-assembled structures such as electrodes in batteries, electrolytic capacitors, and sen-
considered in this chapter are insulators, or very poor electrical sors. Recent studies of photon emission by conducting poly-
conductors. However, a variety of newly developed macromo- mers may lead to new technologies for light-emitting diodes
lecular materials have electrical conductivities that rival those and flat-panel displays. Conducting polymers also show prom-
of silicon-based semiconductors and even metallic conductors. ise as molecular wires that can be incorporated into nanome-
We examine one example in detail: conducting polymers, in tre-sized electronic devices.
which extensively conjugated double bonds permit electron

Checklist of concepts
☐ 1. The elastic properties of a material are summarized by ☐ 5. The melting temperature, Tm, of a protein or nucleic
a stress–strain curve. acid is the temperature at which the fraction of dena-
☐ 2. A perfect elastomer is a polymer for which the internal tured polymer is 0.5.
energy is independent of the extension. ☐ 6. In conducting polymers conjugated double bonds per-
☐ 3. The disruption of long-range order in a polymer occurs mit electron conduction along the chain.
at a melting temperature.
☐ 4. Synthetic polymers undergo a transition from a state
of high to low chain mobility at the glass transition
temperature.
17B Properties of macromolecules  713

Checklist of equations
Property Equation Comment Equation number

Conformational entropy of a random coil ΔS = − 12 kN ln{(1+ ␯)1+␯ (1− ␯)1−␯ } ν = n/N 17B.1

Restoring force of a one-dimensional random coil F = (kT/2l)ln{(1 + ν)/(1 − ν)} 17B.2a

F ≈ nkT/Nl Small extensions 17B.2b


17C Self-assembly

of self-assembly include the formation of liquid crystals, of


protein quaternary structures from two or more polypeptide
Contents
chains (Topic 17A), and (by implication) of a DNA double helix
17C.1 Colloids 714 from two polynucleotide chains (Topic 17A). Here we concen-
(a) Classification and preparation 714 trate on the specific properties of additional self-assembled
(b) Structure and stability 715 systems, including some that are becoming important in the
(c) The electrical double layer 715 development of nanotechnology.
Example 17C.1: Determining the isoelectric
point of a protein 717
17C.2 Micelles and biological membranes 717
(a) Micelle formation 717 17C.1 Colloids
Brief illustration 17C.1: The fraction of surfactant
molecules in micelles 718
A colloid, or disperse phase, is a dispersion of small particles of
(b) Bilayers, vesicles, and membranes 719
one material in another that does not settle out under gravity. In
Brief illustration 17C.2: The melting temperatures
this context, ‘small’ means that one dimension at least is smaller
of membranes 720
than about 500 nm in diameter (about the wavelength of visible
(c) Self-assembled monolayers 720
Checklist of concepts 720
light). Many colloids are suspensions of nanoparticles (particles
Checklist of equations 721
of diameter up to about 100 nm). In general, colloidal particles
are aggregates of numerous atoms or molecules, but are com-
monly but not universally too small to be seen with an ordinary
optical microscope. They pass through most filter papers, but can
be detected by light-scattering and sedimentation (Topic 17D).

➤ Why do you need to know this material?


Aggregates of small and large molecules form the
(a) Classification and preparation
basis of many technologies (such as detergents and The name given to the colloid depends on the two phases
nanotechnology) and are abundant in biological cells. involved:
To see why this is the case, you need to understand their
structures and properties. t A sol is a dispersion of a solid in a liquid (such as clusters
of gold atoms in water) or of a solid in a solid (such as
➤ What is the key idea? ruby glass, which is a gold-in-glass sol, and achieves its
Colloids, micelles, and biological membranes form colour by light scattering).
spontaneously by self-assembly of molecules or t An aerosol is a dispersion of a liquid in a gas (like fog and
macromolecules and are held together by molecular many sprays) or a solid in a gas (such as smoke): the
interactions. particles are often large enough to be seen with a
microscope.
➤ What do you need to know already?
t An emulsion is a dispersion of a liquid in a liquid (such
You need to be familiar with molecular interactions (Topic
as milk). A foam is a dispersion of a gas in a liquid.
16B), the formation of liquids (Topic 16C), and interactions
between ions (Topic 5F). A further classification of colloids is as lyophilic, or solvent
attracting, and lyophobic, solvent repelling. If the solvent is
water, the terms hydrophilic and hydrophobic, respectively,
Self-assembly is the spontaneous formation of complex struc- are used instead. Lyophobic colloids include the metal sols.
tures of molecules or macromolecules that are held together Lyophilic colloids generally have some chemical similarity to
by molecular interactions, such as Coulombic, dispersion, the solvent, such as eOH groups able to form hydrogen bonds.
hydrogen bonding, or hydrophobic interactions. Examples A gel is a semi-rigid mass of a lyophilic sol.
17C Self-assembly  715

The preparation of aerosols can be as simple as sneezing


(which produces an imperfect aerosol). Laboratory and com-
mercial methods make use of several techniques. Material (for
example, quartz) may be ground in the presence of the disper-
sion medium. Passing a heavy electric current through a cell
may lead to the sputtering (crumbling) of an electrode into
colloidal particles. Arcing between electrodes immersed in the
support medium also produces a colloid. Chemical precipita-
tion sometimes results in a colloid. A precipitate (for example,
R
silver iodide) already formed may be dispersed by the addition
of a peptizing agent (for example, potassium iodide). Clays may
Figure 17C.1 Although the attraction between individual
be peptized by alkalis, the OH− ion being the active agent.
molecules is proportional to 1/R6, more molecules are
Emulsions are normally prepared by shaking the two com-
within range at large separations (pale region) than at small
ponents together vigorously, although some kind of emulsify- separation (dark region), so the total interaction energy
ing agent usually has to be added to stabilize the product. This declines more slowly and is proportional to a lower power of R.
emulsifying agent may be a soap (the salt of a long-chain car-
boxylic acid) or other surfactant (surface active) species, or a
lyophilic sol that forms a protective film around the dispersed kinetics of collapse: colloids are thermodynamically unstable
phase. In milk, which is an emulsion of fats in water, the emul- but kinetically non-labile.
sifying agent is casein, a protein containing phosphate groups. At first sight, even the kinetic argument seems to fail: colloi-
It is clear from the formation of cream on the surface of milk dal particles attract each other over large distances, so there is a
that casein is not completely successful in stabilizing milk: the long-range force that tends to condense them into a single blob.
dispersed fats coalesce into oily droplets which float to the sur- The reasoning behind this remark is as follows. The energy of
face. This coagulation may be prevented by ensuring that the attraction between two individual atoms i and j separated by
emulsion is dispersed very finely initially: intense agitation a distance Rij, one in each colloidal particle, varies with their
with ultrasonics brings this dispersion about, the product being separation as 1/Rij6 (Topic 16B). The sum of all these pairwise
‘homogenized’ milk. interactions, however, decreases only as approximately 1/R2
One way to form an aerosol is to tear apart a spray of liquid (the precise variation depending on the shape of the particles
with a jet of gas. The dispersal is aided if a charge is applied and their closeness), where R is the separation of the centres of
to the liquid, for then electrostatic repulsions help to blast it the particles. The change in the power from 6 to 2 stems from
apart into droplets. This procedure may also be used to produce the fact that at short distances only a few molecules interact but
emulsions, for the charged liquid phase may be directed into at large distances many individual molecules are at about the
another liquid. same distance from one another, and contribute equally to the
Colloids are often purified by dialysis, the process of squeez- sum (Fig. 17C.1), so the total interaction does not fall off as fast
ing the solution though a membrane. The aim is to remove as the single molecule–molecule interaction.
much (but not all, for reasons explained later) of the ionic Several factors oppose the long-range dispersion attraction.
material that may have accompanied their formation. A mem- For example, there may be a protective film at the surface of
brane (for example, cellulose) is selected that is permeable to the colloid particles that stabilizes the interface and cannot be
solvent and ions, but not to the colloid particles. Dialysis is very penetrated when two particles touch. Thus the surface atoms
slow, and is normally accelerated by applying an electric field of a platinum sol in water react chemically and are turned into
and making use of the charges carried by many colloidal parti- ePt(OH)3H3, and this layer encases the particle like a shell.
cles; the technique is then called electrodialysis. A fat can be emulsified by a soap because the long hydrocarbon
tails penetrate the oil droplet but the carboxylate head groups
(or other hydrophilic groups in synthetic detergents) surround
(b) Structure and stability the surface, form hydrogen bonds with water, and give rise to
Colloids are thermodynamically unstable with respect to the a shell of negative charge that repels a possible approach from
bulk. This instability can be expressed thermodynamically by another similarly charged particle.
noting that because the change in Helmholtz energy, dA, when
the surface area of the sample changes by dσ at constant tem-
perature and pressure is dA = γ dσ, where γ is the interfacial sur-
(c) The electrical double layer
face tension (Topic 16C), it follows that dA < 0 if dσ < 0. That is, A major source of kinetic non-lability of colloids is the exist-
the contraction of the surface (dσ < 0) is spontaneous (dA < 0). ence of an electric charge on the surfaces of the particles. On
The survival of colloids must therefore be a consequence of the account of this charge, ions of opposite charge tend to cluster
716 17 Macromolecules and self-assembly

nearby, and an ionic atmosphere is formed, just as for ions


(Topic 5F). rD >> a
We need to distinguish two regions of charge. First, there is rD << a
a fairly immobile layer of ions that adhere tightly to the sur-

Potential energy, V
face of the colloidal particle, and which may include water mol-
ecules (if that is the support medium). The radius of the sphere
that captures this rigid layer is called the radius of shear and is
the major factor determining the mobility of the particles. The 0

Coagulation
electric potential at the radius of shear relative to its value in
the distant, bulk medium is called the zeta potential, ζ, or the
Flocculation
electrokinetic potential. Second, the charged unit attracts an Separation, s
oppositely charged atmosphere of mobile ions. The inner shell
of charge and the outer ionic atmosphere is called the electrical Figure 17C.2 The potential energy of interaction as a function
double layer. of the separation of the centres of the two particles and its
The theory of the stability of lyophobic dispersions was variation with the ratio of the particle size to the thickness a of
developed by B. Derjaguin and L. Landau and independently the electric double layer, rD. The regions labelled coagulation
by E. Verwey and J.T.G. Overbeek, and is known as the DLVO and flocculation show the dips in the potential energy curves
theory.1 It assumes that there is a balance between the repulsive where these processes occur.
interaction between the charges of the electric double layers
on neighbouring particles and the attractive interactions aris- permittivity, ε = εrε0). The potential energy arising from the
ing from van der Waals interactions between the molecules in attractive interaction has the form
the particles. The potential energy arising from the repulsion of
double layers on particles of radius a has the form B
Vattraction = − (17C.4)
s
Aa ζ − s/r
2 2
Vrepulsion = + e D (17C.1) where B is another constant. The variation of the total potential
R
energy with separation is shown in Fig. 17C.2.
where A is a constant, ζ is the zeta potential, R is the separation At high ionic strengths, the ionic atmosphere is dense and
of centres, s is the separation of the surfaces of the two parti- the potential shows a secondary minimum at large separations.
cles (s = R − 2a for spherical particles of radius a), and rD is the Aggregation of the particles arising from the stabilizing effect
thickness of the double layer. This expression is valid for small of this secondary minimum is called flocculation. The floccu-
particles with a thick double layer (a ≪ rD). When the double lated material can often be redispersed by agitation because the
layer is thin (rD ≪ a), the expression is replaced by well is so shallow. Coagulation, the irreversible aggregation of
distinct particles into large particles, occurs when the separa-
Vrepulsion = + 12 Aa2ζ 2 ln(1+ e − s/r ). D (17C.2) tion of the particles is so small that they enter the primary min-
imum of the potential energy curve and van der Waals forces
In each case, the thickness of the double layer can be estimated are dominant.
from an expression like that derived for the thickness of the The ionic strength is increased by the addition of ions, par-
ionic atmosphere in the Debye–Hückel theory (Topic 5F) in ticularly those of high charge type, so such ions act as floc-
which there is a competition between the assembling influ- culating agents. This increase is the basis of the empirical
ences of the attraction between opposite charges and the dis- Schulze–Hardy rule, that hydrophobic colloids are flocculated
ruptive effect of thermal motion: most efficiently by ions of opposite charge type and high charge
number. The Al3+ ions in alum are very effective, and are used
1/2
⎛ εRT ⎞ to induce the congealing of blood. When river water containing
rD = ⎜ Thickness of the electrical double layer (17C.3)
⎝ 2 ρF 2 Ib < ⎟⎠ colloidal clay flows into the sea, the salt water induces floccula-
tion and coagulation, and is a major cause of silting in estuaries.
where I is the ionic strength of the solution, ρ its mass den- Metal oxide sols tend to be positively charged whereas sulfur
sity, and b< = 1 mol kg−1 (F is Faraday’s constant and ε is the and the noble metals tend to be negatively charged. Naturally
occurring macromolecules also acquire a charge when dis-
persed in water, and an important feature of proteins and other
1 The derivation of the expressions quoted is too complicated to include natural macromolecules is that their overall charge depends
here. For a full description, see Volume 1 of R. J. Hunter, Foundations of col- on the pH of the medium. For instance, in acidic environ-
loid science, Oxford University Press (1987). ments protons attach to basic groups, and the net charge of
17C Self-assembly  717

the macromolecule is positive; in basic media the net charge is of charge. This disruption may occur at high temperatures,
negative as a result of proton loss. At the isoelectric point the which is one reason why sols precipitate when they are heated.
pH is such that there is no net charge on the macromolecule.

Example 17C.1 Determining the isoelectric point of a Micelles and biological


17C.2
protein membranes
The speed with which bovine serum albumin (BSA) moves
through water under the inf luence of an electric field was In aqueous solutions surfactant molecules or ions can cluster
monitored at several values of pH, and the data are listed together as micelles, which are colloid-sized clusters of mol-
below. What is the isoelectric point of the protein? ecules, for their hydrophobic tails tend to congregate (through
hydrophobic interactions: see Topic 16B), and their hydrophilic
pH 4.20 4.56 5.20 5.65 6.30 7.00
head groups provide protection (Fig. 17C.4).
Velocity/(μm s −1) 0.50 0.18 −0.25 −0.65 −0.90 −1.25

Method If we plot speed against pH, we can use interpolation (a) Micelle formation
to find the pH at which the speed is zero, which is the pH at
which the molecule has zero net charge. Micelles form only above the critical micelle concentration
(CMC) and above the Krafft temperature. The CMC is detected
Answer The data are plotted in Fig. 17C.3. The velocity passes by noting a pronounced change in physical properties of the
through zero at pH = 4.8; hence pH = 4.8 is the isoelectric solution, particularly the molar conductivity (Fig. 17C.5). There
point. is no abrupt change in properties at the CMC; rather, there is a
1.0

0.5
4.8
Speed, s/(μm s–1)

–0.5

–1

–1.5
4 5 pH 6 7
Figure 17C.4 A schematic version of a spherical micelle.
Figure 17C.3 The plot of the speed of a moving The hydrophilic groups are represented by spheres and the
macromolecule against pH allows the isoelectric point to be hydrophobic hydrocarbon chains are represented by the stalks;
detected as the pH at which the speed is zero. The data are these stalks are mobile.
from Example 17C.1.

Self-test 17C.1 The following data were obtained for another Molar
protein: conductivity
Surface
Physical property

pH 3.5 4.5 5.0 5.5 6.0 tension


Velocity/(μm s−1) 0.10 −0.10 −0.20 −0.30 −0.40

Estimate the pH of the isoelectric point.


Answer: 4.0
Osmotic
pressure
CMC
The primary role of the electric double layer is to confer
Concentration of surfactant
kinetic non-lability. Colliding colloidal particles break through
the double layer and coalesce only if the collision is sufficiently Figure 17C.5 The typical variation of some physical properties
energetic to disrupt the layers of ions and solvating molecules, of an aqueous solution of sodium dodecylsulfate close to the
or if thermal motion has stirred away the surface accumulation critical micelle concentration (CMC).
718 17 Macromolecules and self-assembly

transition region corresponding to a range of concentrations


that for large N, there is a reasonably sharp transition in the
around the CMC where physical properties vary smoothly but relative concentrations of surfactant molecules that are pre-
nonlinearly with the concentration. The hydrocarbon interior sent in micelles, which corresponds to the existence of a CMC.
of a micelle is like a droplet of oil. Nuclear magnetic resonance
shows that the hydrocarbon tails are mobile, but slightly more Self-test 17C.2 Equation 17C.5b is surprisingly tricky to solve,
restricted than in the bulk. Micelles are important in industry but it is possible to make good progress with simple cases.
and biology on account of their solubilizing function: matter With N = 2 and K = 1, find an expression for [M2].
can be transported by water after it has been dissolved in their Answer: [M2 ] =[M total ]− 14 {(1+ 8[Mtotal ])1/2 −1}
hydrocarbon interiors. For this reason, micellar systems are
used as detergents, for organic synthesis, froth flotation, and
petroleum recovery. Non-ionic surfactant molecules may cluster together in
The self-assembly of a micelle has the characteristics of a clumps of 1000 or more, but ionic species tend to be disrupted
cooperative process in which the addition of a surfactant mol- by the electrostatic repulsions between head groups and are
ecule to a cluster that is forming becomes more probable the normally limited to groups of less than about 100. However, the
larger the size of the aggregate, so after a slow start there is a disruptive effect depends more on the effective size of the head
cascade of formation of micelles. If we suppose that the domi- group than the charge. For example, ionic surfactants such
nant micelle MN consists of N monomers M, then the dominant as sodium dodecyl sulfate (SDS) and cetyltrimethylammo-
equilibrium we have to consider is nium bromide (CTAB) form rods at moderate concentrations
whereas sugar surfactants form small, approximately spherical
[M N ]
NM M N K= (17C.5a) micelles. The micelle population is often polydisperse, and the
[M]N
shapes of the individual micelles vary with shape of the con-
We have assumed, probably dangerously on account of the large stituent surfactant molecules, surfactant concentration, and
sizes of monomers, that the solution is ideal and that activities temperature. A useful predictor of the shape of the micelle the
can be replaced by molar concentrations. The total concentra- surfactant parameter, Ns, defined as
tion of surfactant is [M]total = [M] + N[MN] because each micelle
consists of N monomer molecules. Therefore, V
Ns = Definition Surfactant parameter (17C.6)
Al
[M N ]
K= (17C.5b)
([M]total − N[MN ])N where V is the volume of the hydrophobic surfactant tail, A is
the area of the hydrophilic surfactant head group, and l is the
maximum length of the surfactant tail. Table 17C.1 summa-
Brief illustration 17C.1 The fraction of surfactant rizes the dependence of aggregate structure on the surfactant
molecules in micelles parameter.
In aqueous solutions spherical micelles form, as shown in Fig.
Equation 17C.5b can be solved numerically for the micelle
17C.4, with the polar head groups of the surfactant molecules on
concentration as a function of the total surfactant concentra-
tion and some results for K = 1 are shown in Fig. 17C.6. We see
the micellar surface and interacting favourably with solvent and
ions in solution. Hydrophobic interactions stabilize the aggre-
0.4 gation of the hydrophobic surfactant tails in the micellar core.
Under certain experimental conditions, a liposome may form,
Fraction present as micelles

with an inward pointing inner surface of molecules surrounded


0.3
by an outward pointing outer layer (Fig. 17C.7). Liposomes may
be used to carry nonpolar drug molecules in blood.
N=3
0.2

Table 17C.1 Variation of micelle shape with the surfactant


0.1 N = 30 (magnified × 10) parameter

Ns Micelle shape
0 <0.33 Spherical
0 2 4 6 8 10
Total number of surfactant molecules 0.33 to 0.50 Cylindrical rods
0.50 to 1.00 Vesicles
Figure 17C.6 The micelle concentration as a function of the
1.00 Planar bilayers
total surfactant concentration for K = 1.
>1.00 Reverse micelles and other shapes
17C Self-assembly  719

Bilayers show a close resemblance to biological membranes,


and are often a useful model on which to base investigations
of biological structures. However, actual membranes are
highly sophisticated structures. The basic structural element
of a membrane is a phospholipid, such as phosphatidyl cho-
line (1), which contains long hydrocarbon chains (typically
in the range C14–C24) and a variety of polar groups, such as
–CH2 CH2 N(CH3 )3+ . The hydrophobic chains stack together to
Figure 17C.7 The cross-sectional structure of a spherical form an extensive layer about 5 nm across. The lipid molecules
liposome. form layers instead of micelles because the hydrocarbon chains
are too bulky to allow packing into nearly spherical clusters.
Increasing the ionic strength of the aqueous solution reduces
O
repulsions between surface head groups, and cylindrical (CH) 7 (CH 2)7 CH 3
micelles can form. These cylinders may stack together in rea- cis
O
sonably close-packed (hexagonal) arrays, forming lyotropic O– O
mesomorphs and, more colloquially, ‘liquid crystalline phases’. +
(H3C) 3N O P O O
Reverse micelles form in nonpolar solvents, with small polar O (CH) 14 CH 3
surfactant head groups in a micellar core and more volumi-
nous hydrophobic surfactant tails extending into the organic 1 Phosphatidyl choline
bulk phase. These spherical aggregates can solubilize water in
organic solvents by creating a pool of trapped water molecules The bilayer is a highly mobile structure. Not only are the
in the micellar core. As aggregates arrange at high surfactant hydrocarbon chains ceaselessly twisting and turning in the
concentrations to yield long-range positional order, many region between the polar groups, but the phospholipid and
other types of structures are possible including cubic and hexa- cholesterol molecules migrate over the surface. It is better to
gonal shapes. think of the membrane as a viscous fluid rather than a perma-
The enthalpy of micelle formation reflects the contributions nent structure, with a viscosity about 100 times that of water.
of interactions between micelle chains within the micelles and Typically, a phospholipid molecule in a membrane migrates
between the polar head groups and the surrounding medium. through about 1 μm in about 1 min.
Consequently, enthalpies of micelle formation display no read- All lipid bilayers undergo a transition from a state of
ily discernible pattern and may be positive (endothermic) or high to low chain mobility at a temperature that depends
negative (exothermic). Many non-ionic micelles form endo- on the structure of the lipid. To visualize the transition, we
thermically, with ΔH of the order of 10 kJ per mole of surfactant consider what happens to a membrane as we lower its tem-
molecules. That such micelles do form above the CMC indicates perature (Fig. 17C.8). There is sufficient energy available at
that the entropy change accompanying their formation must normal temperatures for limited bond rotation to occur and
then be positive, and measurements suggest a value of about the flexible chains writhe. However, the membrane is still
+140 J K−1 mol−1 at room temperature. The fact that the entropy highly organized in the sense that the bilayer structure does
change is positive, even though the molecules are clustering
together, shows that hydrophobic interactions are important in
the formation of micelles (in the sense that water molecules are
released to become more disordered in the process and hence
make an overwhelming contribution to the entropy change).

(b) Bilayers, vesicles, and membranes


Some micelles at concentrations well above the CMC form
extended parallel sheets two molecules thick, called planar
(a) (b)
bilayers. The individual molecules lie perpendicular to the
sheets, with hydrophilic groups on the outside in aqueous solu- Figure 17C.8 A depiction of the variation with temperature of
tion and on the inside in nonpolar media. When segments of the flexibility of hydrocarbon chains in a lipid bilayer.
planar bilayers fold back on themselves, unilamellar vesicles (a) At physiological temperature, the bilayer exists as a liquid
may form where the spherical hydrophobic bilayer shell sepa- crystal, in which some order exists but the chains writhe.
rates an inner aqueous compartment from the external aque- (b) At a specific temperature, the chains are largely frozen and
ous environment. the bilayer is said to exist as a gel.
720 17 Macromolecules and self-assembly

not come apart and the system is best described as a liquid


(c) Self-assembled monolayers
crystal. At lower temperatures, the amplitudes of the writh-
ing motion decrease until a specific temperature is reached at Molecular self-assembly can be used as the basis for manipula-
which motion is largely frozen. The membrane is then said to tion of surfaces on the nanometre scale. Of current interest are
exist as a gel. Biological membranes exist as liquid crystals at self-assembled monolayers (SAMs), ordered molecular aggre-
physiological temperatures. gates that form a single layer of material on a surface. To under-
Phase transitions in membranes are often observed as ‘melt- stand the formation of SAMs, consider exposing molecules
ing’ from gel to liquid crystal by calorimetric methods. The data such as alkyl thiols, RSH, where R represents an alkyl chain,
show relations between the structure of the lipid and the melting to an Au(0) surface. The thiols react with the surface, forming
temperature. Interspersed among the phospholipids of biological RS−Au(I) adducts:
membranes are sterols, such as cholesterol (2), which is largely
hydrophobic but does contain a hydrophilic −OH group. Sterols, RSH + Au(0)n → RS− Au(I)⋅ Au(0)n −1 + 12 H2
which are present in different proportions in different types of
cells, prevent the hydrophobic chains of lipids from ‘freezing’ If R is a sufficiently long chain, van der Waals interactions
into a gel and, by disrupting the packing of the chains, spread between the adsorbed RS units lead to the formation of a
the melting point of the membrane over a range of temperatures. highly ordered monolayer on the surface (Fig. 17C.9). The
Gibbs energy of formation of SAMs increases with the length
of the alkyl chain, with each methylene group contributing
0.4–4 kJ mol−1.
A self-assembled monolayer alters the properties of the
surface. For example, a hydrophilic surface may be rendered
hydrophobic once covered with a SAM. Furthermore, attaching
HO functional groups to the exposed ends of the alkyl groups may
2 Cholesterol impart specific chemical reactivity or ligand-binding proper-
ties to the surface, leading to applications in chemical (or bio-
chemical) sensors and reactors.
Brief illustration 17C.2 The melting temperatures of
membranes
To predict trends in melting temperatures we need to assess
the strengths of the interactions between molecules. Longer
chains can be expected to be held together more strongly by
hydrophobic interactions than shorter chains, so we should
expect the melting temperature to increase with the length
of the hydrophobic chain of the lipid. On the other hand, any
structural elements that prevent alignment of the hydropho- S S S S
bic chains in the gel phase lead to low melting temperatures. Au surface
Indeed, lipids containing unsaturated chains, those contain-
ing some Ca C bonds, form membranes with lower melting
Figure 17C.9 Self-assembled monolayers of alkylthiols formed
temperatures than those formed from lipids with fully satu-
onto a gold surface by reaction of the thiol groups with the
rated chains, those consisting of CeC bonds only.
surface and aggregation of the alkyl chains.
Self-Test 17C.3 Why do bacterial and plant cells grown at low
temperatures synthesize more phospholipids with unsatu-
rated chains than do cells grown at higher temperatures?
Answer: Insertion of lipids with unsaturated chains
lowers the plasma membrane’s melting temperature to a value
that is close to the lower ambient temperature.

Checklist of concepts
☐ 1. A disperse system is a dispersion of small particles of ☐ 2. Colloids are classified as lyophilic and lyophobic.
one material in another.
17C Self-assembly  721

☐ 3. A surfactant is a species that accumulates at the inter- ☐ 11. A micelle is a colloid-sized cluster of molecules that
face of two phases or substances. forms at and above the critical micelle concentration
☐ 4. Many colloid particles are thermodynamically unstable and the Krafft temperature.
but kinetically non-labile. ☐ 12. Micelles can assume a number of shapes, depending on
☐ 5. The radius of shear is the radius of the sphere that temperature, shape and concentration of constituent
captures the rigid layer of charge attached to a colloid molecules.
particle. ☐ 13. Planar bilayers are micelles that exist as extended par-
☐ 6. The zeta potential is the electric potential at the radius allel sheets two molecules thick that are extended.
of shear relative to its value in the distant, bulk medium. ☐ 14. Unilamellar vesicles are micelles that exist as extended
☐ 7. The inner shell of charge and the outer atmosphere parallel sheets two molecules thick that fold back on
jointly constitute the electric double layer. themselves.
☐ 8. Flocculation is the reversible aggregation of colloidal ☐ 15. Self-assembled monolayers are ordered molecular
particles. aggregates that form a single layer of material on a sur-
☐ 9. Coagulation is the irreversible aggregation of colloidal face spontaneously.
particles.
☐ 10. The Schultze–Hardy rule states that hydrophobic col-
loids are flocculated most efficiently by ions of opposite
charge type and high charge number.

Checklist of equations
Property Equation Comment Equation number

Thickness of the electrical double layer rD = (εRT/2ρF2Ib<)1/2 Debye–Hückel theory 17C.3

Surfactant parameter Ns = V/Al Definition 17C.6


17D Determination of size and shape

lengths and extents of cross-linking, in which case sharp


Contents X-ray images are not obtained. Even if all the molecules in
the sample are identical, it might prove impossible to obtain
17D.1 Mean molar masses 722
a single crystal, which is essential for diffraction studies
Example 17D.1: Calculating number and mass
because only then does the electron density (which is respon-
averages 723
sible for the scattering) have a large-scale periodic varia-
17D.2 The techniques 724
tion. Furthermore, although work on proteins and DNA has
(a) Mass spectrometry 724
shown how immensely interesting and motivating the data
Example 17D.2: Interpreting the mass spectrum
can be, the information is incomplete. For instance, what can
of a polymer 725
(b) Laser light scattering 725
be said about the shape of the molecule in its natural environ-
ment, a biological cell? What can be said about the response
Example 17D.3: Determining the size
of a polymer by light scattering 726 of its shape to changes in its environment?
(c) Sedimentation 726
Example 17D.4: Determining a sedimentation
constant 727
(d) Viscosity 728 17D.1 Mean molar masses
Example 17D.5: Using intrinsic viscosity
to measure molar mass 729 A pure protein is monodisperse, meaning that it has a single,
Checklist of concepts 730 definite molar mass. There may be small variations, such as one
Checklist of equations 730 amino acid replacing another, depending on the source of the
sample. A synthetic polymer, however, is polydisperse, in the
sense that a sample is a mixture of molecules with various chain
lengths and molar masses. The various techniques that are used
➤ Why do you need to know this material? to measure molar mass result in different types of mean values
of polydisperse systems.
To appreciate modern work on macromolecules in
The mean obtained from the determination of molar mass
technology and biochemistry, you need to understand
by osmometry (Topic 5B) is the number-average molar mass,
several experimental techniques that are used to
M n , which is the value obtained by weighting each molar mass
determine the molar masses and shapes of synthetic and
by the number of molecules of that mass present in the sample:
biological polymers.

∑N M = 〈 M 〉
1 Number-average
➤ What is the key idea? Mn = i i Definition
molar mass (17D.1)
N
i
Mass spectrometry, laser light scattering, ultracentrifugation,
and viscosity measurements are useful techniques for the
where Ni is the number of molecules with molar mass Mi and
determination of size and shape of macromolecules.
there are N molecules in all. The notation 〈X〉 denotes the usual
➤ What do you need to know already? (number average) of a property X, and we shall use it again
below. For reasons related to the ways in which macromole-
You need to be familiar with structures of macromolecules
cules contribute to physical properties, viscosity measurements
(Topic 17A) and aggregates (Topic 17C).
give the viscosity-average molar mass, M v , light-scattering
experiments give the weight-average molar mass, M w , and
sedimentation experiments give the Z-average molar mass,
X-ray diffraction (Topic 18A) can reveal the position of M Z . (The name is derived from the z-coordinate used to depict
almost every atom other than hydrogen even in very large data in a procedure for determining the average.) Although
molecules. However, there are several reasons why other such averages are often best left as empirical quantities, some
techniques must also be used. In the first place, the sam- may be interpreted in terms of the composition of the sample.
ple might be a mixture of molecules with different chain Thus, the weight-average molar mass is the average calculated
17D Determination of size and shape  723

by weighting the molar masses of the molecules by the mass of


Answer The amounts in each interval are as follows:
each one present in the sample:
Interval 5–9 10–14 15–19 20–24 25–29 30–35
∑m M
1
Mw = i i Definition Weight-average molar mass (17D.2a)
m Molar mass/ 7.5 12.5 17.5 22.5 27.5 32.5
i
(kg mol−1)
In this expression, mi is the total mass of molecules of molar
Amount/mmol 1.3 0.70 0.51 0.25 0.11 0.052
mass Mi and m is the total mass of the sample. Because
mi = NiMi/NA, we can also express this average as Total: 2.92

∑N M i
2
i
〈M 2〉 Weight-average
The number-average molar mass is therefore
Mw = i
= Interpretation (17D.2b) 1
∑N M i i
〈M 〉 molar mass M n /(kg mol −1) =
2.92
(1.3 × 7.5 + 0.70 × 12.5 + 0.51 × 17.5
i
+ 0.25 × 22.5 + 0..11 × 27.5 + 0.052 × 32.5) = 13
This expression shows that the weight-average molar mass is
proportional to the mean square molar mass. Similarly, the The weight-average molar mass is calculated directly from the
Z-average molar mass turns out to be proportional to the mean data after noting that the total mass of the sample is 37.6 g:
cubic molar mass: 1
M w /(kg mol −1) = (9.6 × 7.5 + 8.7 × 12.5 + 8.9 × 17.5
37.6
∑N M i
3
i
〈M 3〉 Z-average
+ 5.6 × 22.5 + 3.1 × 27.5 + 1.7 × 32.5) = 16
MZ = i
= Interpretation
molar mass (17D.2c)
∑N M
i
i
2
i
〈M 2〉 Note the different values of the two averages. In this instance,
M w /M n =1.2.
Self-test 17D.1 Evaluate the Z-average molar mass of the
Example 17D.1 sample.
Calculating number and mass averages
Answer: 19 kg mol−1
Determine the number-average and the weight-average molar
masses of a sample of poly(vinyl chloride) from the following
data:
The ratio M w /M n is called the (molar-mass) dispersity (pre-
Molar mass Average molar mass within Mass of sample
viously the ‘polydispersity index’, PDI) and denoted Ð (read
interval/(kg mol−1) interval/(kg mol−1) within interval/g ‘D-stroke’). It follows from eqns 17D.1 and 17D.2 that
5–9 7.5 9.6
Mw
10–14 12.5 8.7 Ð= Definition Dispersity (17D.3a)
Mn
15–19 17.5 8.9
20–24 22.5 5.6 It then follows from the interpretation of the weight and num-
25–29 27.5 3.1 ber averages that
30–35 32.5 1.7
〈M 2 〉
Ð= Interpretation Dispersity (17D.3b)
Method The relevant equations are eqns 17D.2a and 17D.2b. 〈 M 〉2
Calculate the two averages by weighting the molar mass
within each interval by the number and mass, respectively, of That is, the dispersity is proportional to the ratio of the mean
the molecules in each interval. Obtain the numbers in each square molar mass to the square of the mean molar mass. In the
interval by dividing the mass of the sample in each interval determination of protein molar masses we expect the various
by the average molar mass for that interval. Because the num- averages to be the same because the sample is monodisperse
ber of molecules is proportional to the amount of substance (unless there has been degradation). A synthetic polymer nor-
(the number of moles), the number-weighted average can be mally spans a range of molar masses and the different averages
obtained directly from the amounts in each interval. That is, yield different values. Typical synthetic materials have Ð≈ 4
on dividing the numerator and denominator by Avogadro’s but much recent research has been devoted to developing
constant NA, and writing n = N/NA, eqn 17D.1 becomes methods that give much lower dispersities. The term ‘monodis-
perse’ is conventionally applied to synthetic polymers in which
∑(N /N ∑n M
1 1
Mn = i A )M i = i i the dispersity is less than 1.1; commercial polyethene samples
N /N A n
i i might be much more heterogeneous, with a dispersity close to
30. One consequence of a narrow molar mass distribution for
724 17 Macromolecules and self-assembly

synthetic polymers is often a higher degree of three-dimen- High


sional long-range order in the solid and therefore higher den- potential Laser
sity and melting point. The spread of values is controlled by difference
the choice of catalyst and reaction conditions. In practice, it is
Detector
found that long-range order is determined more by structural
factors (branching, for instance) than by molar mass.

d l
17D.2 The techniques
Average molar masses may be determined by osmotic pressure Figure 17D.1 A matrix-assisted laser desorption/ionization
of polymer solutions. The upper limit for the reliability of mem- time-of-flight (MALDI-TOF) mass spectrometer. A laser beam
brane osmometry is about 1000 kg mol−1. A major problem for ejects macromolecules and ions from the solid matrix. The
macromolecules of relatively low molar mass (less than about ionized macromolecules are accelerated by an electrical
10 kg mol−1) is their ability to percolate through the membrane. potential difference over a distance d and then travel through
One consequence of this partial permeability is that membrane a drift region of length l. Ions with the smallest mass to charge
osmometry tends to overestimate the average molar mass of a ratio (m/z) reach the detector first.
polydisperse mixture. Several techniques for the determination
of molar mass and dispersity that are not so limited include macromolecule is ionized by collisions and complexation with
mass spectrometry, laser light scattering, ultracentrifugation, small cations, such as H+, Na+, and Ag+.
and viscosity measurements. Figure 17D.2 shows the MALDI-TOF mass spectrum of a
polydisperse sample of poly(butylene adipate) (PBA, 1). The
A note on good practice The masses of macromolecules MALDI technique produces mostly singly charged molecular
are often reported in daltons (Da), where 1 Da = mu (with ions that are not fragmented. Therefore, the multiple peaks in
mu = 1.661 × 10−27 kg). Note that 1 Da is a measure of molecular the spectrum arise from polymers of different lengths, with the
mass not of molar mass. We might say that the mass (not the intensity of each peak being proportional to the abundance of
molar mass) of a certain macromolecule is 100 kDa (that is, its each polymer in the sample. Values of M n , M w and the disper-
mass is 100 × 103 × mu); we could also say that its molar mass sity can be calculated from the data. It is also possible to use the
is 100 kg mol−1; we should not say (even though it is common mass spectrum to verify the structure of a polymer, as shown in
practice) that its molar mass is 100 kDa. the following example.

(a) Mass spectrometry O


O
Mass spectrometry is among the most accurate techniques for HO O OH
the determination of molar masses. The procedure consists of O
n
ionizing the sample in the gas phase and then measuring the 1
mass-to-charge number ratio (m/z; more precisely, the dimen-
sionless ratio m/zmu) of all ions. Macromolecules present a
challenge because it is difficult to produce gaseous ions of large 4

species without fragmentation. However, two techniques have


emerged that circumvent this problem: matrix-assisted laser 3
Relative intensity

desorption/ionization (MALDI) and electrospray ionization. n = 20


We shall discuss MALDI-TOF mass spectrometry, so called
2
because the MALDI technique is coupled to a time-of-flight
(TOF) ion detector.
Figure 17D.1 shows a schematic view of a MALDI-TOF 1
mass spectrometer. The macromolecule is first embedded in
a solid matrix which typically consists of an organic material
0
such as trans-3-indoleacrylic acid and inorganic salts such as 0 4000 8000 12000
m/z
sodium chloride or silver trifluoroacetate. This sample is then
irradiated with a pulsed laser. The laser energy ejects electroni- Figure 17D.2 MALDI-TOF spectrum of a sample of
cally excited matrix ions, cations, and neutral macromolecules, poly(butylene adipate) (1) with Mn = 4525g mol−1 (Adapted
thus creating a dense gas plume above the sample surface. The from D.C. Mudiman et al., J. Chem. Educ. 74, 1288 (1997).)
17D Determination of size and shape  725

Detector
Example 17D.2 Interpreting the mass spectrum
of a polymer
Incident
The mass spectrum in Fig. 17D.2 consists of peaks spaced by ray
200 g mol−1. The peak at 4113 g mol−1 corresponds to the poly- θ
mer for which n = 20. From these data, verify that the sample
consists of polymers with the general structure given by 1. Monochromatic
source
Method Because each peak corresponds to a different value Scattering intensity, I Sample
of n, the molar mass difference, ΔM, between peaks corre-
sponds to the molar mass, M, of the repeating unit (the group Figure 17D.3 Rayleigh scattering from a sample of point-like
inside the brackets in 1). Furthermore, the molar mass of the particles. The intensity of scattered light depends on the angle
terminal groups (the groups outside the brackets in 1) may be θ between the incident and scattered beams.
obtained from the molar mass of any peak by using
I (θ ) 2
M (terminal groups) = M (polymer with n repeating units) R(θ ) = ×r Definition Rayleigh ratio (17D.4)
I0
− nΔM − M (cation)
where I0 is the intensity of the incident laser radiation. The
where the last term corresponds to the molar mass of the cat-
factor r2 occurs in the definition of the Rayleigh ratio because
ion that attaches to the macromolecule during ionization.
the light wave spreads out over a sphere of radius r and surface
Answer The value of ΔM is consistent with the molar mass of area 4πr2, so any sample of the radiation has its intensity I(θ)
the repeating unit shown in 1, which is 200 g mol−1. The molar decreased by a factor proportional to r2.
mass of the terminal group is calculated by recalling that Na+ A detailed examination of the scattering shows that the
is the cation in the matrix: Rayleigh ratio depends on the mass concentration, cP (units: kg
M(terminal group) = 4113g mol −1 − 20(200 g mol −1) m−3), of the macromolecule and its weight-average molar mass
M w as:
− 23g mol −1 = 90 g mol −1

The result is consistent with the molar mass of the R(θ ) = KP(θ )c P M w Relation of Rayleigh ratio to molar mass (17D.5)
dO(CH 2) 4 OH terminal group (89 g mol −1) plus the molar
mass of the dH terminal group (1 g mol−1). where the parameter K depends on the refractive index of the
solution, the incident wavelength, and the distance between
Self-test 17D.2 What would be the molar mass of the n = 20 the detector and the sample, which is held constant during the
polymer if silver trifluoroacetate were used instead of NaCl in experiment. The parameter P(θ) is the structure factor, which
the preparation of the matrix? is related to the size of the molecule. When the radius of gyra-
Answer: 4198 g mol−1 tion, Rg, of the molecule (Topic 17A and Table 17D.1) is much
smaller than the wavelength of the light,

16π 2 Rg2sin 2 12 θ
(b) Laser light scattering P(θ ) ≈ 1− p(θ ) with p(θ ) =
3λ 2
The scattering of light by particles with diameters much smaller Small macromolecules Structure factor (17D.6)
than the wavelength of the incident radiation is called Rayleigh
scattering. The intensity of the scattered light is proportional to Equation 17D.5 applies only to ideal solutions. In practice,
the molar mass of the particle and to λ−4, so shorter-wavelength even relatively dilute solutions of macromolecules can devi-
radiation is scattered more intensely than longer wavelengths. ate considerably from ideality. Being so large, macromolecules
For example, the blue of the sky arises from the more intense displace a large quantity of solvent instead of replacing indi-
scattering of the blue component of white sunlight by the mol- vidual solvent molecules with negligible disturbance. To take
ecules of the atmosphere.
Consider the experimental arrangement shown in Fig. 17D.3 Table 17D.1* Radius of gyration
for the measurement of light scattering from solutions of mac- Rg/nm
M/(kg mol−1)
romolecules. Typically, the sample is irradiated with mono-
chromatic light from a laser. The intensity of scattered light is Serum albumin 66 2.9
then measured as a function of the angle θ that the direction Polystyrene 3.2 × 103 50†
of the laser beam makes with the direction of the detector from DNA 4 × 103 117
the sample at a distance r. Under these conditions, the intensity, * More values are given in the Resource section.
I(θ), of light scattered is written as the Rayleigh ratio: † In a poor solvent.
726 17 Macromolecules and self-assembly

deviations from ideality into account, it is common to rewrite 7.5


eqn 17D.5 as Kc P /R(θ ) = 1/P(θ )M w and to extend it to

Kc P 1
= + Bc P (17D.7)
R(θ ) P(θ )M w 6.5

{102/(R(θ)}/m–2
where B is an empirical constant analogous to the osmotic virial
coefficient (Topic 5B) and indicative of the effect of excluded 5.5
volume.
The preceding discussion shows that structural properties,
such as size and the molar mass of a macromolecule, can be 4.5
obtained from measurements of light scattering by a sample at 0 10 20 30 40 50
{103 sin2 ½θ/(R(θ)}/m–2
several angles θ relative to the direction of propagation on an
incident beam. In modern instruments, lasers are used as the Figure 17D.4 Plot of the data for Example 17D.3.
radiation sources.

θ/° 15.0 45.0 70.0 85.0 90.0


Example 17D.3 Determining the size of a polymer
R(θ)/m2 23.8 22.9 21.6 20.7 20.4
by light scattering
The following data for a sample of polystyrene in butanone In a separate experiment, it was determined that K = 2.40 × 102
were obtained at 20 °C with plane-polarized light at mol m 5 kg−2 . From this information, calculate the radius of
λ = 546 nm. gyration and the molar mass of the protein. Assume the pro-
tein is small enough that eqn 17D.6 holds.
θ/° 26.0 36.9 66.4 90.0 113.6
Answer: Rg = 39.8 nm; M w = 498 kg mol −1
R(θ)/m2 19.7 18.8 17.1 16.0 14.4

I n s e pa r ate ex per i ment s , it wa s deter m i ned t h at


K = 6.42 × 10−5 mol m5 kg−2. From this information, calculate Rg (c) Sedimentation
and M w for the sample. Assume that B is negligibly small, and
that the polymer is small enough that eqn 17D.6 holds. In a gravitational field, heavy particles settle towards the foot
of a column of solution by the process called sedimentation.
Method Substituting the result of eqn 17D.6 into eqn 17D.5
The rate of sedimentation depends on the strength of the field
gives, after some rearrangement
and on the masses and shapes of the particles. Spherical mol-
1 1 ⎛ 16π2 Rg2 ⎞ 1 ecules (and compact molecules in general) sediment faster
= +⎜ sin 2 12 θ
R(θ ) Kc P M w ⎝ 3λ ⎟⎠ R(θ ) than rod-like and extended molecules. When the sample is at
2

equilibrium, the particles are dispersed over a range of heights


Hence, a plot of 1/R(θ) against {1/R(θ )}sin 2 12 θ should be a in accord with the Boltzmann distribution (because the gravi-
straight line with slope 16π2 Rg2 /3λ 2 and y-intercept 1/Kc p M w . tational field competes with the stirring effect of thermal
motion). The spread of heights depends on the masses of the
Answer We construct a table of values of 1/R(θ) and
molecules, so the equilibrium distribution is another way to
(sin 2 12 θ )/R(θ ) and plot the data (Fig. 17D.4).
determine molar mass.
θ/° 26.0 36.9 66.4 90.0 113.6 Sedimentation is normally very slow, but it can be acceler-
ated by ultracentrifugation, a technique that replaces the gravi-
{102/R(θ)}/m−2 5.06 5.32 5.83 6.25 6.96
tational field with a centrifugal field. The effect can be achieved
{10 3
× (sin 2 21 θ )/R(θ )}/ m−2 2.56 5.33 17.5 31.3 48.7 in an ultracentrifuge, which is essentially a cylinder that can be
rotated at high speed about its axis with a sample in a cell near
The best straight line through the data has a slope of 0.391 its outer edge. Modern ultracentrifuges can produce accelera-
and a y-intercept of 5.06 × 10 −2 . From these values and the tions equivalent to about 105 that of gravity (‘105 g’). Initially
value of K, we calculate R g = 4.71 × 10 −8 m = 47.1 nm and the sample is uniform, but the ‘top’ (innermost) boundary of
M w = 987 kg mol −1. the solute moves outwards as sedimentation proceeds.
Self-test 17D.3 The following data for an aqueous solution of A solute particle of mass m has an effective mass meff = bm on
a protein with cP = 2.0 kg m−3 were obtained at 20 °C with laser account of the buoyancy of the medium, with
light at λ = 532 nm:
b = 1− ρvs (17D.8)
17D Determination of size and shape  727

where ρ is the solution density, vs is the partial specific vol- macromolecules, the measurements must be carried out at a
ume of the solute (vs = (∂V/∂mB)T, with mB the total mass of series of concentrations and then extrapolated to zero concen-
solute), and the dimensionless quantity ρvs takes into account tration to avoid the complications that arise from the interfer-
the mass of solvent displaced by the solute. The solute particles ence between bulky molecules.
at a distance r from the axis of a rotor spinning at an angular
velocity ω experience a centrifugal force of magnitude meffrω2. Example 17D.4 Determining a sedimentation constant
The acceleration outwards is countered by a frictional force
proportional to the speed, s = dr/dt, of the particles through The sedimentation of the protein bovine serum albumin
the medium. This force is written fs, where f is the frictional (BSA) was monitored at 25 °C. The initial location of the solute
surface was at 5.50 cm from the axis of rotation, and during
coefficient. The particles therefore adopt a drift speed, a con-
centrifugation at 56 850 r.p.m. it receded as follows:
stant speed through the medium, which is found by equating
the two forces meffrω2 and fs. The forces are equal when t/s 0 500 1000 2000 3000 4000 5000
r/cm 5.50 5.55 5.60 5.70 5.80 5.91 6.01
meff rω 2 bmrω 2
s= = (17D.9)
f f Calculate the sedimentation coefficient.
Method Equation 17D.10 can be interpreted as a differential
The drift speed depends on the angular velocity and the radius, equation for s = dr/dt in terms of r; so integrate it to obtain
and it is convenient to define the sedimentation constant, S, as a formula for r in terms of t. The integrated expression, an
expression for r as a function of t, will suggest how to plot the
s
S= Definition Sedimentation constant (17D.10) data and obtain from it the sedimentation constant.
rω 2
Answer Equation 17D.10 may be written
Then, because the average molecular mass is related to the aver- dr dr
= rω 2S and hence = ω 2 S dt
age molar mass M n through m = M w /N A dt r

bM n If at t = 0 the surface is at r 0 and at a later time t it is at r, this


S= (17D.11) equation integrates to
fN A
r
For a spherical particle of radius a in a solvent of viscosity η, ln =ω 2St
r0
the frictional coefficient f is given by Stokes’ relation:
It follows that a plot of ln(r/r 0) against t should be a straight
f = 6πaη Stokes‘ relation (17D.12) line of slope ω2S. Use ω = 2πν, where ν is in cycles per second,
and draw up the following table:
On substituting this expression into eqn 17D.12, we obtain t/s 0 500 1000 2000 3000 4000 5000
102 ln(r/r0) 0 0.905 1.80 3.57 5.31 7.19 8.87
bM n Spherical Relation between S
S= (17D.13)
6πaηN A polymer and the molar mass The straight-line graph (Fig. 17D.5) has a slope of 1.78 × 10−5;
so ω 2S = 1.78 × 10 −5 s −1. Because ω = 2π × (56 850/60) s −1 =
and S may be used to determine either M n or a. Again, if
the molecules are not spherical, we use the appropriate 10
value of f given in Table 17D.2. As always when dealing with
8

Table 17D.2* Frictional coefficients and molecular geometry†


100 ln(r/r0)

6
a/b Prolate Oblate
4
2 1.04 1.04
3 1.18 1.17
2
6 1.31 1.28
8 1.43 1.37 0
0 1 2 3 4 5
10 1.54 1.46 t/(103 s)
* More values and analytical expressions are given in the Resource section.
† Entries are the ratio f/f , where f = 6πηc, where c = (ab2)1/3 for prolate ellipsoids and Figure 17D.5 A plot of the data for Example 17D.4.
0 0
c = (a 2b)1/3 for oblate ellipsoids; 2a is the major axis and 2b is the minor axis.
728 17 Macromolecules and self-assembly

5.95 × 103 s−1, it follows that S = 5.02 × 10−13 s. The unit 10−13 s (d) Viscosity
is sometimes called a ‘svedberg’ and denoted Sv; in this case
S = 5.02 Sv. The formal definition of viscosity is given in Topic 19A; for
now, we need to know that highly viscous liquids flow slowly
Self-test 17D.4 Calculate the sedimentation constant given the and retard the motion of objects through them. The presence
following data (the other conditions being the same as above): of a macromolecular solute increases the viscosity of a solution.
t/s 0 500 1000 2000 3000 4000 5000 The effect is large even at low concentration, because big mol-
ecules affect the fluid flow over an extensive region surround-
r/cm 5.65 5.68 5.71 5.77 5.84 5.9 5.97
ing them. At low concentrations the viscosity, η, of the solution
Answer: 3.11 Sv is related to the viscosity of the pure solvent, η0, by

η = η0 (1 +[η]c +[η]′c 2 + ...) (17D.15)


The difficulty with using sedimentation rates to measure
molar masses lies in the inaccuracies inherent in the deter- The intrinsic viscosity, [η], is the analogue of a virial coeffi-
mination of diffusion coefficients of disperse systems. This cient like that encountered in the description of real gases (and
problem can be avoided by allowing the system to reach equi- has dimensions of 1/concentration). It follows from eqn 17D.15
librium, for the transport property D is then no longer relevant. that
As we show in the following Justification, the weight-average
molar mass can be obtained from the ratio of concentrations of ⎛ η − η0 ⎞ ⎛ η /η0 −1 ⎞ Intrinsic
[η] = lim ⎜ ⎟ = lim ⎜ Definition
c ⎟⎠
(17D.16)
the macromolecules at two different radii in a centrifuge oper- c →0 ⎝ cη0 ⎠ c →0 ⎝ viscosity
ating at angular frequency ω:
Viscosities are measured in several ways. In the Ostwald vis-
2RT c cometer shown in Fig. 17D.6, the time taken for a solution to
Mw = 2 2 ln 2 (17D.14)
(r2 − r1 )bω 2 c1 flow through the capillary is noted, and compared with a stand-
ard sample. The method is well suited to the determination of
An alternative treatment of the data leads to the Z-average [η] because the ratio of the viscosities of the solution and the
molar mass. The centrifuge is run more slowly in this technique pure solvent is proportional to the drainage time t and t0 after
than in the sedimentation rate method to avoid having all the correcting for different densities ρ and ρ0:
solute pressed in a thin film against the bottom of the cell. At
these slower speeds, several days may be needed for equilib- η t ρ
= × (17D.17)
rium to be reached. η0 t 0 ρ0

Justification 17D.1 This ratio can be used directly in eqn 17D.16. Viscometers in
The weight-average molar mass
the form of rotating concentric cylinders are also used (Fig.
from sedimentation experiments
17D.7), and the torque on the inner cylinder is monitored
The centrifugal force acting on a molecule at a radius r when it while the outer one is rotated. Such rotating rheometers (as
is rotating around the axis of the centrifuge at a frequency ω is in this case, some instruments for the measurement of vis-
mω2r. This force corresponds to a difference in potential energy cosity are also called rheometers, from the Greek word for
(using F = –dV/dr) of – 12 mω 2r 2 . The difference in potential
energy between r1 and r2 (with r2 > r1) is therefore 12 mω 2 (r12 − r22 ).
According to the Boltzmann distribution, the ratio of concen-
trations of molecules at these two radii should therefore be
c 2 − 12 meff ω 2 (r12 −r22 )/kT Measuring
=e
c1 lines

The effective mass, allowing for buoyancy effects, is m(1 – vs ρ),


and m/k can be replaced by M/R, where R = NAk is the gas con-
stant. Then, by taking logarithms of both sides, the last equa- Capillary
tion becomes
c 2 M (1 − vs ρ)ω 2 (r22 − r12 )
ln =
c1 2RT
Figure 17D.6 An Ostwald viscometer. The viscosity is
which rearranges into eqn 17D.14. measured by noting the time required for the liquid to drain
between the two marks.
17D Determination of size and shape  729

Torsion wire Calculate the intrinsic viscosity and estimate the molar mass
Sample of the polymer by using eqn 17D.19 with K = 3.80 × 105 dm3 g−1
and a = 0.63.
Method The intrinsic viscosity is defined in eqn 17D.16; there-
fore, form this ratio at the series of data points and extrapolate
to c = 0. Interpret M v , as M v /(g mol –1) in eqn 17D.18.
Answer We draw up the following table:
c/(g dm−3) 0 2 4 6 8 10
Motor
η/η0 1 1.102 1.208 1.317 1.43 1.549
100[(η/η0) − 1]/(c/ 5.11 5.2 5.28 5.38 5.49
Figure 17D.7 A rotating rheometer. The torque on the inner g dm−3)
drum is observed when the outer container is rotated.
The points are plotted in Fig. 17D.8. The extrapolated intercept
at c = 0 is 0.0504, so [η] = 0.0504 dm3 g−1. Therefore,
‘flow’) have the advantage over the Ostwald viscometer that
1/a
the shear gradient between the cylinders is simpler than in ⎛ [η]⎞
Mv = ⎜ ⎟ = 9.0 ×104 g mol −1
the capillary and effects of the kind discussed shortly can be ⎝K⎠
studied more easily.
There are many complications in the interpretation of vis- 0.055
cosity measurements. Much of the work is based on empirical
observations, and the determination of molar mass is usually 0.054
{(η/η0) – 1}/{c/(g dm–3)}

based on comparisons with standard, nearly monodisperse


0.052 0.053

samples. Some regularities are observed that help in the deter-


mination. For example, it is found that solutions of macro-
molecules that behave nearly ideally often fit the empirical
Mark–Kuhn–Houwink–Sakurada equation:
0.051
[η] = KM va Mark–Kuhn–Houwink–Sakurada equation (17D.18)
0.050
0 2 4 6 8 10
where K and a are constants that depend on the solvent and c/(g dm–3)

type of macromolecule (Table 17D.3); the viscosity-average Figure 17D.8 The plot used for the determination of
molar mass, M v appears in this expression. intrinsic viscosity, which is taken from the intercept at c = 0;
see Example 17D.5.
Table 17D.3* Intrinsic viscosity
Self-test 17D.5 Show that the intrinsic viscosity may also be
Solvent θ/°C K/(cm3 g−1) a obtained as [η] = limc→0 ln(η/η 0) and evaluate the viscosity-
Polystyrene Benzene 25 9.5 × 10−3 0.74 average molar mass by using this relation.
Polyisobutylene Benzene 23 8.3 × 10−2 0.50 Answer: 90 kg mol−1
Various Guanidine hydrochloride + 7.2 × 10−3 0.66
proteins HSCH2CH2OH
* More values are given in the Resource section.
In some cases, the flow is non-Newtonian in the sense
that the viscosity of the solution changes as the rate of flow
Example 17D.5 increases. A decrease in viscosity with increasing rate of flow
Using intrinsic viscosity to measure
indicates the presence of long rod-like molecules that are ori-
molar mass
entated by the flow and hence slide past each other more freely.
The viscosities of a series of solutions of polystyrene in toluene In some somewhat rare cases the stresses set up by the flow are
were measured at 25 °C with the following results: so great that long molecules are broken up, with further conse-
quences for the viscosity.
c/(g dm−3) 0 2 4 6 8 10
η/(10−4 kg m−1 s−1) 5.58 6.15 6.74 7.35 7.98 8.64
730 17 Macromolecules and self-assembly

Checklist of concepts
☐ 1. Macromolecules can be monodisperse, with a sin- ☐ 5. Dynamic light scattering is a technique for the deter-
gle molar mass, or polydisperse, with various molar mination of the diffusion properties and molar masses
masses. of macromolecules and aggregates.
☐ 2. In the MALDI-TOF technique, matrix-assisted laser ☐ 6. The rate of sedimentation in an ultracentrifuge depends
desorption/ionization is coupled with a time-of-flight on molar masses and shapes of the macromolecules in
mass spectrometer to measure the molar masses of the sample.
macromolecules. ☐ 7. The weight-average and Z-average molar mass of a
☐ 3. The intensity of Rayleigh light scattering by a sample sample of macromolecules can be determined from
increases with decreasing wavelength of the incident equilibrium measurements of sedimentation in an
radiation and increasing size of the particles in the ultracentrifuge.
sample. ☐ 8. The viscosity-average molar mass can be determined
☐ 4. Analysis of Rayleigh scattering leads to the determina- from measurements of the viscosity of solutions of
tion of the molar mass of a macromolecule or aggregate. macromolecules.

Checklist of equations
Property Equation Comment Equation number

Number-average molar mass M n = (1/ N ) ∑N M


i
i i Definition 17D.1

Weight-average molar mass M w = (1/m) ∑m M


i
i i Definition 17D.2a

Z-average molar mass M Z = 〈 M 3 〉 /〈 M 2 〉 Interpretation 17D.2c

Dispersity Ð= M w /M n Definition 17D.3a

Interpretation 17D.3b
Ð= 〈 M 2 〉 /〈 M 〉2

Rayleigh ratio R(θ) = (I(θ)/I0)r2 Definition 17D.4

R(θ ) = KP(θ )c P M w Experimental implementation; 17D.5


ideal solutions

Structure factor P(θ ) ≈ 1− p(θ ) Small macromolecules 17D.6


p(θ ) = (16π2 Rg2 /3λ 2 )sin2 12 θ

Sedimentation constant S = s / rω 2 Definition 17D.10

Stokes relation f = 6πaη f is the frictional coefficient 17D.12

Relation between the sedimentation constant S = bM n /6πaηN A Spherical polymer 17D.13


and the molar mass of a polymer

Intrinsic viscosity [η] = lim{(η /η0 −1)/ c } Definition 17D.16


c →0

Mark–Kuhn–Houwink–Sakurada equation [η] = KM va Nearly ideal solutions 17D.18


Exercises and problems  731

CHAPTER 17 Macromolecules and self-assembly

TOPIC 17A The structures of macromolecules


Discussion questions
17A.1 Distinguish between the four levels of structure of a macromolecule: 17A.3 Define the terms in, and identify the limits of the generality of, the
primary, secondary, tertiary, and quaternary. following expressions: (a) Rc = Nl, (b) Rrms = N1/2l, (c) Rrms = (2N)1/2l, (d)
Rrms = N1/2lF, (e) Rg = N1/2l, (f) Rg = (N/6)1/2l, (g) Rg = (N/3)1/2l.
17A.2 What are the consequences of there being partial rigidity in an
otherwise random coil? 17A.4 Summarize the Core–Pauling rules and explain how they explain the
helical and sheet-like structures of polypeptides.

Exercises
17A.1(a) A one-dimensional polymer chain consists of 700 segments, each 17A.5(a) What is the probability that the ends of a polyethene chain of molar
0.90 nm long. If the chain were ideally flexible, what would be the r.m.s. mass 65 kg mol−1 are between 10.0 nm and 10.1 nm apart when the polymer is
separation of the ends of the chain? treated as a three-dimensional freely jointed chain?
17A.1(b) A one-dimensional polymer chain consists of 1200 segments, each 17A.5(b) What is the probability that the ends of a polyethene chain of molar
1.125 nm long. If the chain were ideally flexible, what would be the r.m.s. mass 75 kg mol−1 are between 14.0 nm and 14.1 nm apart when the polymer is
separation of the ends of the chain? treated as a three-dimensional freely jointed chain?
17A.2(a) Calculate the contour length (the length of the extended chain) 17A.6(a) By what percentage does the radius of gyration of a one-dimensional
and the root mean square separation (the end-to-end distance) for polymer chain increase (+) or decrease (−) when the bond angle between
polyethylene with a molar mass of 280 kg mol−1, modelled as a one- units is limited to 109°? What is the percentage change in volume of the coil?
dimensional chain. 17A.6(b) By what percentage does the root mean square separation of the ends
17A.2(b) Calculate the contour length (the length of the extended of a one-dimensional polymer chain increase (+) or decrease (−) when the
chain) and the root mean square separation (the end-to-end distance) for bond angle between units is limited to 120°? What is the percentage change in
polypropylene of molar mass 174 kg mol−1, modelled as a one-dimensional volume of the coil?
chain.
17A.7(a) By what percentage does the radius of gyration of a one-dimensional
17A.3(a) The radius of gyration of a long one-dimensional chain molecule is polymer chain increase (+) or decrease (−) when the persistence length
found to be 7.3 nm. The chain consists of CeC links. Assume that the chain is changes from l (the bond length) to 5.0 per cent of the contour length? What
randomly coiled and estimate the number of links in the chain. is the percentage change in volume of the coil?
17A.3(b) The radius of gyration of a long one-dimensional chain molecule 17A.7(b) By what percentage does the root mean square separation of the
is found to be 18.9 nm. The chain consists of links of length 450 pm. Assume ends of a one-dimensional polymer chain increase (+) or decrease (−) when
that the chain is randomly coiled and estimate the number of links in the the persistence length changes from l (the bond length) to 2.5 per cent of the
chain. contour length? What is the percentage change in volume of the coil?
17A.4(a) What is the probability that the ends of a polyethene chain of molar 17A.8(a) The radius of gyration of a three-dimensional partially rigid polymer
mass 65 kg mol−1 are 10 nm apart when the polymer is treated as a one- of 1000 units each of length 150 pm was measured as 2.1 nm. What is the
dimensional freely jointed chain? persistence length of the polymer?
17A.4(b) What is the probability that the ends of a polyethene chain of molar 17A.8(b) The radius of gyration of a three-dimensional partially rigid polymer
mass 85 kg mol−1 are 15 nm apart when the polymer is treated as a one- of 1500 units each of length 164 pm was measured as 3.0 nm. What is the
dimensional freely jointed chain? persistence length of the polymer?

Problems
17A.1 Evaluate the radius of gyration, Rg, of (a) a solid sphere of radius a, 17A.4 Deduce an expression for the radius of gyration of a three-dimensional
(b) a long straight rod of radius a and length l. Show that in the case of a solid freely-jointed chain (eqn 17A.6).
sphere of specific volume vs, Rg/nm ≈ 0.056902 × {(vs/cm3 g−1)(M/g mol−1)}1/3.
17A.5 Derive expressions for the moments of inertia and hence the radii of
Evaluate Rg for a species with M = 100 kg mol−1, vs = 0.750 cm3 g−1, and, in the
gyration of (a) a uniform thin disk, (b) a long uniform rod, (c) a uniform sphere.
case of the rod, of radius 0.50 nm.
17A.6 Construct a two-dimensional random walk by using a random number
17A.2 Use eqn 17A.2 to deduce expressions for (a) the root mean square
generating routine with mathematical software or electronic spreadsheet.
separation of the ends of the chain, (b) the mean separation of the ends, and
Construct a walk of 50 and 100 steps. If there are many people working on the
(c) their most probable separation. Evaluate these three quantities for a fully
problem, investigate the mean and most probable separations in the plots by
flexible chain with N = 4000 and l = 154 pm.
direct measurement. Do they vary as N1/2?
17A.3 Deduce the relation 〈ri 2 〉 = Nl 2 for the mean square distance of a
17A.7 Confirm that for one-dimensional random coils, ln P ≈ ln(2/πN )1/2
monomer from the origin in a freely jointed chain of N units each of length l.
− 12 (N + n +1)ln(1+ ␯) − 12 (N − n +1)ln(1− ␯). Hint: See Justification 17A.1.
Hint: Use the distribution in eqn 17A.2.
732 17 Macromolecules and self-assembly

17A.8 The radius of gyration is defined in Justification 17A.3. Show that an


M/(g mol−1) vs/(cm3 g−1) Rg/nm
equivalent definition is that Rg is the average root mean square distance
of the atoms or groups (all assumed to be of the same mass), that is, that Serum albumin 66 × 103 0.752 2.98
Rg2 = (1/N ) ∑ R2j , where Rj is the distance of atom j from the centre of mass. Bushy stunt virus 10.6 × 106 0.741 12.0
j
17A.9 Use the following information and the expression for Rg of a solid DNA 4 × 106 0.556 117.0
sphere quoted in the text (following eqn 17A.6), to classify the given species
as globular or rod-like.

TOPIC 17B Properties of macromolecules


Discussion questions
17B.1 Distinguish between stress and strain. 17B.3 Distinguish between the melting temperature and the glass transition
temperature of a polymer.
17B.2 Distinguish between elastic and plastic deformation.
17B.4 Describe the mechanism of electrical conductivity in conducting
polymers.

Exercises
17B.1(a) Calculate the change in molar entropy when the ends of a one- 17B.2(a) Calculate the restoring force when the ends of a one-dimensional
dimensional polyethene chain of molar mass 65 kg mol−1 are moved apart by polyethene chain of molar mass 65 kg mol−1 are moved apart by 1.0 nm at
1.0 nm. 20 °C.
17B.1(b) Calculate the change in molar entropy when the ends of a one- 17B.2(b) Calculate the restoring force when the ends of a one-dimensional
dimensional polyethene chain of molar mass 85 kg mol−1 are moved apart polyethene chain of molar mass 85 kg mol−1 are moved apart by 2.0 nm
by 2.0 nm. at 25 °C.

Problems
17B.1 Develop an expression for the fundamental vibrational frequency of 17B.3 The following table lists the glass transition temperatures, Tg, of several
a one-dimensional random coil that has been slightly stretched and then polymers. Discuss the reasons why the structure of the monomer unit has an
released. Evaluate this frequency for a sample of polyethene of molar mass effect on the value of Tg.
65 kg mol−1 at 20 °C. Account physically for the dependence of frequency on
temperature and molar mass. Polymer Poly Polyethene Poly(vinyl Polystyrene
(oxymethylene) chloride)
17B.2 On the assumption that the tension, t, required to keep a sample at a
constant length is proportional to the temperature (t = aT, the analogue of Structure e(OCH2)ne e(CH2CH2)ne e(CH2 e(CH2
p ∝ T), show that the tension can be ascribed to the dependence of the entropy eCHCl)ne eCH(C6H5))ne
on the length of the sample. Account for this result in terms of the molecular Tg/K 198 253 354 381
nature of the sample.

TOPIC 17C Self-assembly


Discussion questions
17C.1 Distinguish between a sol, an emulsion, and a foam. Provide examples 17C.3 What effect is the inclusion of cholesterol likely to have on the transition
of each. temperatures of a lipid bilayer?
17C.2 It is observed that the critical micelle concentration of sodium dodecyl
sulfate in aqueous solution decreases as the concentration of added sodium
chloride increases. Explain this effect.
Exercises and problems  733

Exercises
17C.1(a) The velocity v with which a protein moves through 17C.1(b) The velocity v with which a protein moves through water under
water under the influence of an electric field varied with values the influence of an electric field varied with values of pH in the range
of pH in the range 3.0 < pH < 7.0 according to the expression 3.0 < pH < 5.0 according to the expression v/(μm s−1) = a + b(pH) + c(pH)2 with
v/(μm s−1) = a + b(pH) + c(pH)2 + d(pH)3 with a = 0.50, b = −0.10, c = −3.0 × 10−3, a = 0.80, b = −4.0 × 10−3, and c = −5.0 × 10−2. Identify the isoelectric point of the
and d = 5.0 × 10−4. Identify the isoelectric point of the protein. protein.

Problem
17C.1 Use mathematical software to reproduce the features in Fig. 17C.6.

TOPIC 17D Determination of size and shape

Discussion questions
17D.1 Distinguish between number-average, weight-average, and Z-average 17D.2 Suggest reasons why different techniques produce different molar mass
molar masses. Identify experimental techniques that can measure each of averages.
these properties.

Exercises
17D.1(a) Calculate the number-average molar mass and the weight-average 17D.5(a) Find the drift speed of a particle of radius 20 μm and density
molar mass of a mixture of equal amounts of two polymers, one having 1750 kg m3 which is settling from suspension in water (density
M = 62 kg mol−1 and the other M = 78 kg mol−1. 1000 kg m−3) under the influence of gravity alone. The viscosity of water is
17D.1(b) Calculate the number-average molar mass and the weight-average 8.9 × 10−4 kg m−1 s−1.
molar mass of a mixture of two polymers, one having M = 62 kg mol−1 and 17D.5(b) Find the drift speed of a particle of radius 15.5 μm and density
the other M = 78 kg mol−1, with their amounts (numbers of moles) in the 1250 kg m−3 which is settling from suspension in water (density
ratio 3:2. 1000 kg m−3) under the influence of gravity alone. The viscosity of water is
8.9 × 10−4 kg m−1 s−1.
17D.2(a) A solution consists of solvent, 30 per cent by mass, of a dimer
with M = 30 kg mol−1 and its monomer. What average molar mass would be 17D.6(a) At 20 °C the diffusion coefficient of a macromolecule is found to be
obtained from measurement of (i) osmotic pressure, (ii) light scattering? 8.3 × 10−11 m2 s−1. Its sedimentation constant is 3.2 Sv in a solution of density
17D.2(b) A solution consists of 25 per cent by mass of a trimer with M= 1.06 g cm−3. The specific volume of the macromolecule is 0.656 cm3 g−1.
22 kg mol−1 and its monomer. What average molar mass would be obtained Determine the molar mass of the macromolecule.
from measurement of: (i) osmotic pressure, (ii) light scattering? 17D.6(b) At 20 °C the diffusion coefficient of a macromolecule is found to be
7.9 × 10−11 m2 s−1. Its sedimentation constant is 5.1 Sv in a solution of density
17D.3(a) What is the relative rate of sedimentation for two spherical particles
997 kg m−1. The specific volume of the macromolecule is 0.721 cm3 g−1.
of the same density, but which differ in radius by a factor of 10?
Determine the molar mass of the macromolecule.
17D.3(b) What is the relative rate of sedimentation for two spherical particles
with densities 1.10 g cm−3 and 1.18 g cm−3 and which differ in radius by a 17D.7(a) The data from a sedimentation equilibrium experiment performed
factor of 8.4, the former being the larger? Use ρ = 0.794 g cm−3 for the density at 300 K on a macromolecular solute in aqueous solution show that a graph
of the solution. of ln c against (r/cm)2 is a straight line with a slope of 729. The rotation
rate of the centrifuge was 50 000 r.p.m. The specific volume of the solute is
17D.4(a) Human haemoglobin has a specific volume of 0.749 × 103 m3 kg−1,
0.61 cm3 g−1. Calculate the molar mass of the solute.
a sedimentation constant of 4.48 Sv, and a diffusion coefficient of
17D.7(b) The data from a sedimentation equilibrium experiment performed at
6.9 × 10−11 m2 s−1. Determine its molar mass from this information.
293 K on a macromolecular solute in aqueous solution show that a graph of
17D.4(b) A synthetic polymer has a specific volume of 8.01 × 10−4 m3 kg−1,
ln c against (r/cm)2 is a straight line with a slope of 821. The rotation rate of
a sedimentation constant of 7.46 Sv, and a diffusion coefficient of
the centrifuge was 1080 cycles per second. The specific volume of the solute is
7.72 × 10−11 m2 s−1. Determine its molar mass from this information.
7.2 × 10−4 m3 kg−1. Calculate the molar mass of the solute.

Problems
17D.1 A polymerization process produced a Gaussian distribution of polymers 17D.2 Polystyrene is a synthetic polymer with the structure
in the sense that the proportion of molecules having a molar mass in the e(CH2eCH(C6H5))ne. A batch of polydisperse polystyrene was prepared
range M to M + dM was proportional to e −(M − M ) /2γ . What is the number
2
by initiating the polymerization with t-butyl radicals. As a result, the t-butyl
average molar mass when the distribution is narrow? group is expected to be covalently attached to the end of the final products.
734 17 Macromolecules and self-assembly

A sample from this batch was embedded in an organic matrix containing of the solution. Estimate the effective radius of the haemoglobin molecule
silver trifluoroacetate and the resulting MALDI-TOF spectrum consisted of given that the viscosity of the solution is 1.00 × 103 kg m−1 s−1.
a large number of peaks separated by 104 g mol−1, with the most intense peak
17D.8 The rate of sedimentation of a recently isolated protein was monitored
at 25 578 g mol−1. Comment on the purity of this sample and determine the
at 20 °C and with a rotor speed of 50 000 r.p.m. The boundary receded as
number of (CH2eCH(C6H5)) units in the species that gives rise to the most
follows:
intense peak in the spectrum.
17D.3 Suppose that a rod-like DNA molecule of length 250 nm undergoes a t/s 0 300 600 900 1200 1500 1800
conformational change to a closed-circular (cc) form. (a) Use the information
r/cm 6.127 6.153 6.179 6.206 6.232 6.258 6.284
in Problem 17A.8 and an incident wavelength λ = 488 nm to calculate the ratio
of scattering intensities by each of these conformations, Irod/Icc, when θ = 20°,
Calculate the sedimentation constant and the molar mass of the protein on
45°, and 90°. (b) Suppose that you wish to use light scattering as a technique
the basis that its partial specific volume is 0.728 cm3 g−1 and its diffusion
for the study of conformational changes in DNA molecules. Based on your
coefficient is 7.62 × 10−11 m2 s−1 at 20 °C, the density of the solution then being
answer to part (a), at which angle would you conduct the experiments? Justify
0.9981 g cm−3. Suggest a shape for the protein given that the viscosity of the
your choice.
solution is 1.00 × 103 kg m−1 s−1 at 20 °C.
17D.4 In a sedimentation experiment the position of the boundary as a
17D.9 The concentration dependence of the viscosity of a polymer solution is
function of time was found to be as follows:
found to be as follows:
t/min 15.5 29.1 36.4 58.2
c/(g dm−3) 1.32 2.89 5.73 9.17
r/cm 5.05 5.09 5.12 5.19
η/(g m−1 s−1) 1.08 1.20 1.42 1.73

The rotation rate of the centrifuge was 45 000 r.p.m. Calculate the
The viscosity of the solvent is 0.985 g m−1 s−1. What is the intrinsic viscosity of
sedimentation constant of the solute.
the polymer?
17D.5 Calculate the speed of operation (in r.p.m.) of an ultracentrifuge needed
17D.10 The times of flow of dilute solutions of polystyrene in benzene through
to obtain a readily measurable concentration gradient in a sedimentation
a viscometer at 25 °C are given in the table below. From these data, calculate
equilibrium experiment. Take that gradient to be a concentration at the
the molar mass of the polystyrene samples. Because the solutions are dilute,
bottom of the cell about five times greater than at the top. Use rtop = 5.0 cm,
assume that the densities of the solutions are the same as those of pure
rbot = 7.0 cm, M ≈ 105 g mol−1, ρυs ≈ 0.75, T = 298 K.
benzene. η(benzene) = 0.601 × 103 kg m−1 s−1 (0.601 cP) at 25 °C.
17D.6 In an ultracentrifugation experiment at 20 °C on bovine serum
albumin the following data were obtained: ρ = 1.001 g cm−3, vs = 1.112 cm3 g−1, c/(g dm−3) 0 2.22 5.00 8.00 10.00
ω/2π = 322 Hz,
t/s 208.2 248.1 303.4 371.8 421.3
r/cm 5.0 5.1 5.2 5.3 5.4
17D.11 The viscosities of solutions of polyisobutylene in benzene were
c/(mg cm−3) 0.536 0.284 0.148 0.077 0.039 measured at 23 °C with the following results:

Evaluate the molar mass of the sample. c/(g/102 cm3) 0 0.2 0.4 0.6 0.8 1.0
17D.7 Sedimentation studies on haemoglobin in water gave a sedimentation η/(103 kg m−1 s−1) 0.647 0.690 0.733 0.777 0.821 0.865
constant S = 4.5 Sv at 20 °C. The diffusion coefficient is 6.3 × 10−11 m2 s−1 at
the same temperature. Calculate the molar mass of haemoglobin using vs= Use the information in Table 17D.3 to deduce the molar mass of the polymer.
0.75 cm3 g−1 for its partial specific volume and ρ = 0.998 g cm−3 for the density

Integrated activities
17.1 In formamide as solvent, poly(γ-benzyl-l-glutamate) is found by light ⎛ ∂U ⎞ ⎛ ∂t ⎞
scattering experiments to have a radius of gyration proportional to M; ⎜⎝ ∂l ⎟⎠ = t − ⎜⎝ ∂T ⎟⎠
T l
in contrast, polystyrene in butanone has Rg proportional to M1/2. Present
arguments to show that the first polymer is a rigid rod whereas the second is
17.3 Commercial software (more specifically ‘molecular mechanics’ or
a random coil.
‘conformational search’ software) automate the calculations that lead to
17.2 Consider the thermodynamic description of stretching rubber. The Ramachandran plots, such as those in Fig. 17A.13. In this problem our model
observables are the tension, t, and length, l (the analogues of p and V for for the protein is the dipeptide (1) in which the terminal methyl groups
gases). Because dw = tdl, the basic equation is dU = TdS + tdl. If G = U – TS – tl, replace the rest of the polypeptide chain. (a) Draw three initial conformers
find expressions for dG and dA, and deduce the Maxwell relations of the dipeptide with RaH: one with ϕ = +75°, ψ = −65°, a second with
ϕ = ψ = +180°, and a third with ϕ = +65°, ψ = +35°. Use software of your
instructor’s choice to optimize the geometry of each conformer and find the
⎛ ∂S ⎞ ⎛ ∂t ⎞ ⎛ ∂S ⎞ ⎛ ∂l ⎞
⎜⎝ ∂l ⎟⎠ = − ⎜⎝ ∂T ⎟⎠ ⎜⎝ ∂t ⎟⎠ = − ⎜⎝ ∂T ⎟⎠ final ϕ and ψ angles in each case. Did all the initial conformers converge to
T l T t the same final conformation? If not, what do these final conformers represent?
(b) Use the approach in part (a) to investigate the case R = CH3, with the same
Go on to deduce the equation of state for rubber,
Exercises and problems  735

three initial conformers as starting points for the calculations. Rationalize any where k′ is called the Huggins constant and is typically in the range 0.35–0.40.
similarities and differences between the final conformers of the dipeptides From the fit, determine the intrinsic viscosity and the Huggins constant.
with R = H and R = CH3. (b) Use the empirical Mark–Kuhn–Houwink–Sakurada equation (eqn 17D.18)
to determine the molar mass of polystyrene in the two solvents. For theta
O solvents, a = 0.5 and K = 8.2 × 10−5 dm−3 g−1 for cyclohexane; for the good
NH solvent toluene a = 0.72 and K = 1.15 × 10−5 dm−3 g−1. (c) According to a general
NH theory proposed by Kirkwood and Riseman, the root mean square end-to-end
O H H distance of a polymer chain in solution is related to [η] by Φ 〈r2〉3/2/M, where
Φ is a universal constant with the value 2.84 × 1026 when [η] is expressed
1 in cubic decimetres per gram and the distance is in metres. Calculate this
quantity for each solvent. (d) From the molar masses calculate the average
17.4 The effective radius, a, of a random coil is related to its radius of gyration,
number of styrene (C6H5CHaCH2) monomer units, 〈N〉, (e) Calculate the
Rg, by a = γ Rg, with γ = 0.85. Deduce an expression for the osmotic virial length of a fully stretched, planar zigzag configuration, taking the CeC
coefficient, B (Topic 5B), in terms of the number of chain units for (a) a distance as 154 pm and the CCC bond angle to be 109°. (f) Use eqn 17A.6 to
freely jointed chain, (b) a chain with tetrahedral bond angles. Evaluate B calculate the radius of gyration, Rg. Also calculate 〈r2〉1/2 = N1/2l. Compare this
for l = 154 pm and N= 4000. Estimate B for a randomly coiled polyethylene result with that predicted by the Kirkwood–Riseman theory: which gives the
chain of arbitrary molar mass, M, and evaluate it for M = 56 kg mol−1. Use better fit? (g) Compare your values for M to the results of Problem 17D.2. Is
B = 12 N A vP , where vP is the excluded volume due to a single molecule. there any reason why they should or should not agree? Is the manufacturer’s
17.5 A manufacturer of polystyrene beads claims that they have an average claim valid?
molar mass of 250 kg mol−1. Solutions of these beads are studied by a physical
chemistry student by dilute solution viscometry with an Ostwald viscometer c/(g dm−3 toluene) 0 1.0 3.0 5.0
in both toluene and cyclohexane. The drainage times, tD, as a function of
concentration for the two solvents are given in the table below. (a) Fit the data tD/s 8.37 9.11 10.72 12.52
to the virial equation for viscosity, c/(g dm−3 cyclohexane) 0 1.0 1.5 2.0
tD/s 8.32 8.67 8.85 9.03
η = η0 (1+[η]c + k ′[η]2 c 2 + ...)

You might also like