Principles of Physical Chemistry - Lionel M - Raff Raff - 2nd Ed, Upper Saddle River, NJ, 2004 - Globe Fearon Educational Publishing - 9780131463448 - Anna's Archive
Principles of Physical Chemistry - Lionel M - Raff Raff - 2nd Ed, Upper Saddle River, NJ, 2004 - Globe Fearon Educational Publishing - 9780131463448 - Anna's Archive
PRINCIPLES
OF
PHYSICAL
CHEMISTRY
William H. Hamill
University of Notre Dame
Russell R. Williams, Jr.
formerly of Haverford College
Colin MacKay
Haverford College
PRINCIPLES
OF
PHYSICAL
CHEMISTRY
second edition
PRENTICE-HALL, INC.
englewood cliffs, new jersey
© 1959, 1966
by Prentice-Hall, Inc., Englewood Cliffs, N.J.
In this revision we have tried to remain true to the original philosophy of the
text. Our aim is at the average student of physical chemistry. We do not try to
provide him with an encyclopedic overview of all of physical chemistry. Instead
we strive for an examination in some detail of the basic areas of physical chemis-
try, and we try to conduct this examination with as much rigor as his preparation
will permit. We hope that from this he will learn how physical chemists think
about and attack chemical problems.
Thermodynamics is the core of introductory physical chemistry, and it is
presented first to leave no doubts that it is self-contained.. A minimum of non-
thermodynamic explanation has been interpolated because it obscures the
monolithic simplicity of thermodynamics and implies a fatal deficiency. The
presentation of atomic and molecular structure has been treated more extensive-
ly, in both depth and scope. The final chapter on chemical statistics synthesizes
these separate disciplines, but this order of presentation can be modified, and
even inverted.
The chapter on Nuclear Chemistry has been eliminated, with regrets, be-
cause it is becoming more nearly identified with inorganic chemistry. We _re-
commend to those who prefer to retain this topic the recent text by Bernard
Harvey, Introduction to Nuclear Physics and Chemistry (Englewood Cliffs, N.J.:
Prentice-Hall, Inc., 1962).
A mathematical appendix has been added to provide a handy review of some
basic mathematical techniques. These include partial differentiation, series
expansion, and some material required for specific developments in the later
chapters.
We are indebted to many and acknowledge particularly the students of the
University of Notre Dame and of Haverford College who have also taught and
tested us. We solicit their continued cooperation. Professor George Strauss of
Rutgers University and Professor Richard Fink of Amherst College read much
of the text and made invaluable comments. The treatment of X-ray crystallog-
raphy benefits from the good advice of Professor Alexander Tulinsky of-
Michigan State University. Professor Richard Wolfgang of Yale University
contributed the problem sets he uses in his course there.
Digitized by the Internet Archive
in 2024
https://archive.org/details/principlesofohysOO02unse_p9y3
CONTENTS
3 Thermochemistry 56
§ Equilibria in Solutions of
Nonelectrolytes 196
10 Electrochemistry 235
Appendixes 551
index 567
PRINCIPLES
OF
PHYSICAL
CHEMISTRY
INTRODUCTION AND
DESCRIPTION OF THE
PROPERTIES OF MATTER
still employs such terms as spin, angular momentum, and orbit in referring
to electrons. The “model” of this theory is now taken to be the Schrédinger
wave equation, and insofar as it yields accurate descriptions of the behavior
of matter in terms of measurable quantities such as mass, charge, time, and
energy, it may be regarded as a satisfactory model. The thermodynamic ap-
proach to the correlation of chemical and physical phenomena is quite different
from that of wave mechanics.
1.2 Thermodynamics
treated is often ideal, so that the problem can be treated more simply. An ideal
system is always partly but not completely imaginary, since it is intended to
serve as a simplified but still realistic substitute for the actual system. This
simplification invariably introduces some inaccuracies and the equations which
result are often valid only under special “limiting conditions,” as, for exam-
ple, Boyle’s law, which is valid only in the limit of vanishingly small pressures.
(See section 1.4.)
Other systems in the immediate environment which interact by exchange
of matter or energy with the system of primary interest are treated collectively
as the surroundings. A closed system does not exchange matter with the sur-
roundings and an isolated system exchanges neither matter nor energy.
A system is characterized by describing its state, specifying a sufficient
number of its properties, such as mass, volume, temperature, or pressure,
so that another investigator can construct a duplicate. Such properties are
of two types: (1) extensive properties, which vary with the size of the system,
such as mass, volume, heat capacity, or electric charge; and (2) intensive prop-
erties, which do not depend on the size of the system, such as density,
temperature, or refractive index.
It is not necessary to specify every possible property of a system in order
to characterize its state. For a given mass of a known pure substance it suffices
to specify any two independent variables, usually pressure and temperature.
Once these are known, other properties, such as volume, refractive index, and
heat capacity, are automatically fixed. In a system of m components it is neces-
sary to specify, in addition to three properties such as mass, pressure, and
temperature, n — 1 additional properties, that is, quantities describing the
composition of the system. The total number of properties to be specified is
therefore n + 2.
The preceding statement requires some qualification since it is only for
“well behaved” systems of m components that the specification of n + 2 in-
dependent properties is sufficient to characterize the state of the system.
Except when rate phenomena are of primary interest, it is to be understood
that the system is stable and that its properties are reproducible and indepen-
dent of its history. Some systems may attain a given state very rapidly, as, for
example, the system: one gram of oxygen gas at 50° and one atmosphere pres-
sure. However, the properties of a system such as one gram of acetic acid plus
one gram of ethanol at 25° and one atmosphere may change with time until
a final equilibrium state has been reached in which acetic acid, ethanol, ethyl
acetate, and water are all present. This final state can be characterized with
precision and can be reproduced indefinitely. Both of the preceding types of
systems can be said to be “well behaved.” On the other hand, a solution of
egg albumin is more difficult to characterize, and some types of measurements
on it give results depending on its previous history. The stress-strain relation
of rubber depends on the degree and duration of prior elongations, and this
system behaves reproducibly only within a restricted range of conditions.
A change in state consists of any change in one or more properties of
4 |ntroduction and Description of the Properties of Matter Chapel
1.3 Temperature
erties a standard is adopted and assigned a value for the property, e.g., the
standard kilogram. The value of the property for other bodies or events is
measured by direct or indirect comparison and given in multiple or submultiple
units. In the case of temperature this method of comparison with a single stan-
dard is not possible.
We usually say that temperature is measured with a thermometer but in
actual fact the observation which we make is one of change in volume, that
is, the change in the length of a column of fluid in a capillary tube. We
presume that this measurement of volume bears some unique relation to the
temperature of the system, such as
t=kV-+c
where ¢ is temperature, V is volume, & and c are constants. The thermometer
scale is defined by calibration at fixed points, i.e., systems of reproducible
temperature, such as ice-water at | atm. (the ice point), and boiling water at 1
atm. (the steam point). These points are defined as 0° and 100° on the centi-
grade scale. Unless otherwise specified, all temperatures in this text are given
in the centigrade scale, and the symbol C is omitted.
The choice of fixed points in the centigrade scale implies that the temper-
ature interval between them is to be divided into one hundred equal parts
and that temperatures above and below the fixed points may be measured
by suitable extrapolation. However, it must be kept in mind that such a scale
is divided into units of volume rather than temperature. This shortcoming
may be illustrated by comparing the measurements obtained with two different
thermometric fluids: mercury, the fluid used in most common thermometers,
and amyl alcohol. Suppose that the mercury thermometer and the alcohol
thermometer are each marked at the ice point and the steam point. Then on
each thermometer the space between these marks is divided into one hundred
equal intervals of volume called degrees. While the two thermometers neces-
sarily agree in their readings at 0° and 100°, if they are placed in a bath of such
a temperature that the mercury thermometer reads 50° it will be found that
the amyl alcohol thermometer reads 42.4°. Its volume has increased by less
than half the change observed between the fixed points. In general, two
thermometers using different fluids cannot be expected to agree except at
common points of calibration, although in some instances the disagreements
are small. It is apparent that the temperature scale may depend on the choice
of thermometric fluid as well as upon the choice of fixed points. In a later sec-
tion a method of avoiding this limitation of temperature measurement will
be discussed.
as Boyle’s law: For a given mass of gas at a constant temperature, the volume
is inversely proportional to the pressure.
PV, = P.V, = const. at const-temp. Ca)
Boyle’s law may be expressed in graphical form as shown in Fig. 1.1, in
which the data for a sample of gas at a given temperature lie on a rectangular
hyperbola called an isotherm. The curves in the figure refer to several tem-
peratures for the same sample.
EXERCISE 1.1
This regularity was first stated by Avogadro as a hypothesis (now given the
status of a law) that equal volumes of gases under the same conditions of tem-
perature and pressure contain the same number of molecules.
EXERCISE 1.3
A bulb filled with oxygen at ¢° and p atm. weighs 220.400 g. Evacuated it
weighs 219.850 g. Filled with an unknown gas, also at f° and p atm., the bulb
weighs 221.005 g. Find the molecular weight of the gas. Ans. 67.
is proportional to the number of moles of gas in the sample, equation (1.1) may
be written
PV — nie (1.2)
where n is the number of moles of gas in the sample and the proportionality
constant k is very nearly the same for all gases at the same temperature, but
varies with temperature. Its value is 22.4 1. atm. at 0° and 30.6 |. atm. at 100°.
EXERCISE 1.4
A sample of gas at 100° and 0.80 atm. has a density 1.15 g/l. What is the
molecular weight? Ans. 44.
Evaluate the constant b of equation (1.3) from the slope of the curve for hydrogen
at 100° in Fig. 1.2. From this measurement find the % difference between
(PVx,)° and PV, at 1 atm. Ans. b = 0.0164 1./mole; 0.073 %.
8 Introduction and Description of the Properties of Matter Chap. 1
30
20
(Piy)°
(L. atm.)
10
with the horizontal axis. This can be found by evaluating the constants a and
b in equation (1.4) as follows: Since (PV,,)° = 30.6192 1. atm. at 100°C and
22.4136 1. atm. at 0°C, solution of the simultaneous equations
EXERCISE 1.6
Charles’ law defines a method of measuring temperature. A sample of gas
occupies 100 cc. at 0° and 1 atm. What is the value of ¢ when the gas volume is
80 cc. at 1 atm? Ans. —55°.
The value of R, the gas constant, has been obtained from previous data
on the value of (PV,,)° at various temperatures.
R = (PV x,)°/T = 22.4136/273.15 = 0.082056 1. atm. mole~! deg.~!
It is useful to note that the liter-atmosphere has the dimension of energy and
therefore, in general, the dimensions of the gas constant must be energy
mole deg.~!
EXERCISE 1.8
Compute the volume of 1 mole of hydrogen at 200 atm. and 100°, using the
ideal gas law. Compare the measured volume indicated by the value of PV, at
100° and 200 atm. in Fig. 1.2. Ans. Ideal = 0.153 1.; real = 0.168 1.
EXERCISE 1.9
A compound analyzes 24.5% C, 4.1% H, 71.4% Cl. Find the empirical formula.
The gas density is approximately 2.8 g./l. at 700 mm. and 100°. Find the molecular
formula. Ansa@,
tt, Ely
£=($) +P (1.10)
where a is a constant. Since real gases approach ideal behavior at low pressures
(see Fig. 1.2), the intercept of the curve should give an accurate value of the
molecular weight.
p\?_ M _ (#2)
(5) =ar Ma (p) RP St
In Fig. 1.4, a straight line drawn through the data gives an intercept corre-
sponding to M = 64.06 for sulfur dioxide.
Gas mixtures can be treated by applica-
tion of Dalton’s law, which states that in a
mixture of gases the total pressure is equal to
the sum of the partial pressures of the separate
components. The partial pressure of a com-
ponent is defined as the pressure that it would
exert if present alone in the system, and for
the ith component of an ideal gas,
Py Bee:
=m (1.12)
where V is the volume of the system. Dalton’s
law is
IS Cag See ae Bow 05 10
( 13) P (atm)
where P, is the total pressure. Taking each FIG. 1.4. Molecular weight of
component to be ideal, sulfur dioxide from gas density
measurements at 0°C [after E.
a RT Sn 1 RT (1.14) Moles, T.E Toral,| and A. Escri-
seri
Vee V bano, Trans. Far. Soc., 35, 1451
(1939)].
where m, = 7, tn, +n, +--- = DN
i
EXERCISE 1.10
An equimolar mixture of hydrogen and oxygen occupies 80 cc. at 0.24 atm. and
100°. What will be the pressure of the gaseous mixture after ignition at 100°?
Ans, 0.18 atm.
12 Introduction and Description of the Properties of Matter Chap. 1
TABLE 1.1
PROPERTIES OF GASES
The van der Waals equation can describe pressure-volume data more ac-
curately than the ideal gas law since it allows for variation of PV with pressure.
14 Introduction and Description of the Properties of Matter Chap. 1
The information most readily obtained from the van der Waals equation is
the pressure of a gas sample at a given volume and temperature. Solving equa-
tion (1.17) for the pressure, we have
P = nRT/(V
— nb) — an*/[V? (1.18)
or for comparison with the data in Fig. 1.2, we may compute PV,, by taking
n = 1 and multiplying by V,,
PV peel ae (1.19)
i= b/ Vix Vin
With increasing pressure (decreasing V,,) the first term on the right hand side
of equation (1.19) leads to increasing values of PV,,, while the second term tends
to diminish PV,,. Therefore it should be possible to describe, at least qual-
itatively, a curve such as that for oxygen at 0° in Fig. 1.2.
EXERCISE 1.11
From the data in Table 1.1 and equation (1.19), evaluate PV,, for oxygen
at 0° and 800 atm. by successive approximations. Compare this result with the
fact that z = 1.50.
EXERCISE 1.12
At 200 atm. and 0° the ideal gas law predicts the molar volume of oxygen to be
0.1120 1. Use the van der Waals equation (1.18) to predict the pressure of 1 mole
of oxygen in this volume. Ans. 171 atm.
For comparison with the result obtained in the preceding exercise, we find
from Fig. 1.2 that in this region of pressure PV,, = 20.5 |. atm. (z = 0.915)
for oxygen and therefore the observed pressure of 1 mole in 0.1120 1. is 184
atm. The results of similar computations at other volumes are given in Table
1.2. It is evident from these values that the van der Waals equation is of some
use in describing deviations of a few per cent, but that it is quite inadequate
at high pressures.
Several more complex and more nearly accurate equations of state are
available. One of the most useful of these is the virial equation, which may
be regarded as an extended form of equation (1.3). It contains a series of cor-
rection terms in increasing powers of 1/V or P which take account of the
change in PV with increasing gas density. Written in a form for computation
of the pressure of n moles of gas in volume V, the equation is
P =nRTIV+ Bn2[V?2 + yn3/V3 + Snt/V4 (1.20)
The coefficients 8, y, and 6 depend on the nature of the gas and on the tem-
perature. J. A. Beattie [Chem. Rev., 44, 144 (1949)] has developed the following
relations for the temperature dependence of the virial coefficients.
8 = RTB, — A, — Rc/T’
y = —RTB, + Ava — RB,c/T? (1.21)
6 = RB,bc/T?*
Sec. 1.8 Introduction and Description of the Properties of Matter 15
The several constants have the following values for oxygen: A, = 1.4911 1.2 atm.,
a = 0.02562 1., By = 0.04624 1., 6 = 0.004208 1., and c = 4.80 x 10! 1. deg.’.
Inserting these values in equation (1.21) we may compute the virial coefficients
per mole of oxygen at 0°.
8 = —0.4300 1.? atm.
yy = 0.03270 1.3 atm.
O 49/2 105° atm:
EXCERCISE 1.13
Show that equation (1.20) and the values of @, y, and & for oxygen at 0° permit
one to locate a minimum in PV ys. P (cf. Fig. 1.2).
Values of the pressures given by the virial equation for several other vol-
umes are given in Table 1.2 for comparison with the observed pressure and
with the predictions of other equations of state. It is evident that the virial
equation is useful over a much greater range of pressure than the van der Waals
equation. This improvement follows from the increased number of adjustable
parameters.
TABLE 1.2
TEST OF EQUATIONS OF STATE; OXYGEN AT 0°
V (ce gram')
other; a qualitative comparison of the critical constants in Table 1.1 with the
compressibility factors of Fig. 1.2 shows that the degree of deviation from
ideality is greatest when the pressure and temperature of the gas are near the
critical pressure and temperature of that substance. This suggests that for the
purpose of considering deviations from ideality, the relation of the temperature
and pressure of the system to the critical temperature and pressure are more
important than their absolute values. To assist in this correlation we define
the reduced temperature T, and reduced pressure P, by
=r, PPP: (1.22)
When the compressibility factors are plotted as a function of the reduced
pressure at a given reduced temperature, as in Fig. 1.6, the points for various
substances fall on the same curve. This regularity permits the prediction of the
compressibility factor over a wide range of temperature and pressure from
a knowledge of the critical constants of the substance concerned. For example,
to find the compressibility factor for carbon dioxide at 122° and 300 atm., we
obtain the critical properties of this substance from Table 1.1 and find
T, = 395/304.3 = 130; P= 300/73.0 = 4.11
Since the curve for T, = 1.3 is not given in Fig. 1.6, the value of z is estimated
by interpolation, yielding z = 0.70, that is,
PV/nRT = 0.70 for CO, at 122° and 300 atm.
Secale? Introduction and Description of the Properties of Matter 17
EXERCISE 1.14
From the data given in Fig. 1.6 and Table 1.1 compute the molar volume of
ammonia at 172° and 167 atm. Ans. 0.11 1.
PV/pRT
factor
Compressibility
z=-
Nitrogen n- Butane
Methane Isopentane
Ethane n-Heptane
Ethylene Carbon dioxide
Propane
FIG. 1.6. Compressibility factor
as function of reduced state
variables [from Gouq-Jen Su,
Ind, Eng. Chem., 38, 803 (1946)].
Reduced pressure, P
18 Introduction and Description of the Properties of Matter Chap. 1
'
600 ma
Vapor
pressure
(mm)
500
400 :
300
200f
100
Vapor
pressure hoy
pure liquid
(mm) solvent
solution
4.58
FIG. 1.9. Vapor pressure lower-
pure solid
solvent
ing (schematic, not drawn to
scale),
Temperature (°C)
20 Introduction and Description of the Properties of Matter Chap. 1
the solid and the liquid is metastable. If the vapor pressure of the solid ex-
ceeds that of the atmosphere at a temperature which is less than the melting
point, then the solid sublimes rather than melts. This is the case with solid
carbon dioxide (Dry Ice), which has a vapor pressure of 1 atm. (760 mm.) at
—78°, while the melting point is —56°. At the latter temperature the vapor
pressure is 5.25 atm.
1.10 Solutions
NGeake; (1.23)
X, is the mole fraction of dissolved gas, and in the dilute solutions to which
Henry’s law applies it may be considered to be proportional to the molarity.
A simple consequence of Henry’s law is that the volume solubility of a
gas is independent of pressure. For example, in dilute solution the number
of moles of solute dissolved in 1 mole of solvent is nearly equal to the mole
fraction of solute. Therefore the number of moles dissolved is given by
Secs E10 Introduction and Description of the Properties of Matter 21
TABLE 1.3
SOLUBILITIES OF GASES IN WATER
(Values given are solubility coefficients, i.e., volume of gas, measured at 273°K, ab-
sorbed by 1 volume of water when the partial pressure of gas is 1 atm.)
Temperature
Substance 0° USS 50°
No 0.02354 0.01434 0.01088
O, 0.04889 0.02831 0.02090
H, 0.02148 0.01754 0.01608
CO 0.03537 0.02142 0.01615
CO,* eS 0.759 0.436
H.S* 4.670 2.282 1.392
A solute in a liquid has a significant effect upon the properties of the liquid,
particularly its vapor pressure and related properties. The vapor pressure of
the solvent is lowered by the addition of a solute; in dilute solutions the low-
ering is proportional to the concentration of solute.
This effect may be represented qualitatively as shown in Fig. 1.9, where
the lower curve represents the pressure of solvent vapor over a liquid phase
containing a small concentration of nonvolatile solute. This representation
indicates that the boiling point of the solution (7,) is higher than that of the
pure solvent (73) and that the freezing point of the solution (7) is lower than
that of the pure solvent (T?). The generalizations assume that the solute is
present only in the solution and not in the second phase (solid or vapor).
For so-called ideal solutions the vapor pressure of a volatile component
is described by the relation
P= BPX (1.24)
where P, is the vapor pressure of the substance over the solution in which
its mole fraction is X,. P is the vapor pressure of the pure substance at the
temperature of interest. Equation (1.24) is Raoult’s law and defines an ideal
solution.
Equation (1.24) can be rearranged to show that the vapor pressure lowering
22 = Introduction and Description of the Properties of Matter Chap. 1
and therefore
P, = PA(l — X3)
Rearranging, we have
EDS (1.25)
For example, the vapor pressure of toluene (C,H;CH;) at 50° is 92.6 mm.
If sufficient naphthalene is added to make a solution containing 10 mole per
cent naphthalene the vapor pressure of toluene over the solution will be low-
ered to
P=92.6'x 0.9 =:33.3 mm.
EXERCISE 1.16
The ideal solution law, equation (1.24), also includes Henry’s law, equation
(1.23). In an ideal solution the Henry’s law constant is 1/P%.
The vapor pressure of a solvent over an ideal solution as a function of
composition is shown in Fig. 1.10. When the solute is also volatile, its vapor
pressure over an ideal solution may also be represented by a straight line cor-
responding to Raoult’s law, as indicated in Fig. 1.10. The total vapor pressure
over such solutions
100 ;— 100
iE
P?® (benzene)
P (mm) L
ae P (total) 50
P (benzene)
L
P° (toluene)
P (toluene)
O O
(0) O5 10 FIG. 1.10. Vapor pressure of toluene-benzene
Mole fraction benzene solutions at 20°.
Sec. 1.10 Introduction and Description of the Properties of Matter 23
P,= Tuciao a:
Many real solutions show rather marked deviations from the simple be-
havior predicted by Raoult’s law. The measured total vapor pressure may
be either greater than or less than that predicted by equation (1.26), as is il-
lustrated in Fig. 1.11. It is found, however, that Raoult’s law is still a good
approximation to the observed behavior of the major component when the
mole fraction of that component is near unity. In the region labelled A in Fig.
600 600
P° (carbon
disulfide)
500
P (mm)
400 400
P° (acetone)
300 300
200 200
100 100
O
FIG. 1.11. Vapor pressure of acetone-carbon O 05 1.0
disulfide solutions at 35.2°. Mole fraction
24 Introduction and Description of the Properties of Matter Chap. 1
1.11, the observed curve for vapor pressure of acetone approaches asymptotical-
ly the Raoult’s law line; the same is true for carbon disulfide in region B.
It is also found in real solutions that the minor component obeys Henry’s
law as its mole fraction approaches zero. In region A of Fig. 1.11, the curve
representing the observed vapor pressure of carbon disulfide becomes nearly
a straight line, as predicted by equation (1.23), although the value of the con-
stant k in that equation is not 1/P°, as would be the case if the solution
were ideal.
There is a large class of solutions which conduct electric current, the elec-
trolyte solutions, which will be given special consideration in a later chapter.
SUMMARY, CHAPTER 1
1. Temperature: t= kV +c
Fixed points: ice point, 0°; steam point, 100°
2. Isothermal gas expansion
Boyle’s law
PY — const. at const. temp.
Avogadro’s law
PV Snk(k = 22.41. atm: at 0°)
Real gases
PVy = f(P)
(PV,,)° depends only on temperature
3. Thermal expansion of gases
(2) U0: 082056n1"(CK)
Charles’ law
V = kT (K) at const. pressure
4. Ideal gas law: PV = nRT
R = 0.082056 |. atm. deg.-! mole=!
Ideal gas density
p = PM/RT
Real gas density
p/P = (M/RT)° + aP
Dalton’s law
P,= X;P,
6. Liquefaction of gases
Critical properties
ivandee.
Reduced properties
p= tT and P= PFs
Compressibility factor
Zila)
7. Properties of liquids and solids
Vapor pressure
Normal boiling point
Freezing point
8. Solutions
Henry’s law
Raoult’s law
PROBLEMS, CHAPTER 1
1. The tank of a pressure type van der Graaf generator has a volume of 2400 1.
and it is filled with nitrogen to a pressure of 8 atm. at 25°. Use the ideal gas law
to estimate the weight of nitrogen required. Ans. 22.0 kg.
2. By means of a gas density balance it was found that oxygen at 419 mm.
produces the same buoyancy as an unknown gas at 304mm. What is the
molecular weight of the unknown? Ans. 44.
4. A glass bulb was evacuated and weighed, then filled with oxygen and re-
weighed. The difference in weights was 0.250 g. The operation was repeated
under the same conditions of temperature and pressure with an unknown gas
and it weighed 0.375 g. Find the molecular weight of the unknown gas. Ans. 48.
5. Suppose that a hot air balloon is constructed with a volume of 100,000 1.
and that the contained air is maintained at a temperature of 200°. If the exterior
air has a temperature of 20° and a density of 1.2 g./l., what mass could the
balloon lift? Ans. 46 kg.
12. How many 361. cylinders of nitrogen at 147 atm. and —25° would be
required to fill an evacuated tank of 10‘ 1. capacity to 5.5 atm. at —25°? Obtain
z from Fig. 1.6.
13. A sealed one-liter glass bulb weighs 245.60 g. when suspended in an at-
mosphere of pure oxygen and 245.10 g. in an atmosphere of an unknown, both at
20° and P = | atm. What is the gram-molecular weight of the unknown?
14. A glass bulb connected to a mercury manometer of negligible volume con-
tains helium at P mm. and 0°. When it is used as a thermometer it is desirable
that the pressure become 323 mm. at 50°. Specify the value of P at O°.
15. Referring to the balloon in problem 5, what mass could be lifted if the
balloon were filled with hydrogen at 1 atm. and 20°?
16. A sample of unknown liquid weighing 0.180 g. is volatilized to displace an
equal volume of air which is collected over water at 25° and 740 mm. total
pressure. The observed volume is 60 cc. and the vapor pressure of water at 25°
is 24 mm. Find the molecular weight of the unknown.
17. A mixture of water vapor, carbon dioxide, and excess oxygen resulting
from combustion of a hydrocarbon occupies 85.0 cc. at 100° and 760mm. At
—20° (negligible water vapor pressure) and constant volume the pressure becomes
296 mm. At —130° (negligible carbon dioxide vapor pressure) the pressure be-
comes 105mm. What is the empirical formula of the compound?
18. A soft drink bottle with a total volume of 200 cc. is ;% filled with a water
solution saturated with carbon dioxide at 2 atm. and 25°. The solubility of gases
in ice is negligible. Using the ideal gas law, estimate the gas pressure developed
when the contents are frozen at —10°. The density of ice at —10° is 0.917 g. cc.-!.
19. The total vapor pressure over a solution of heptane-toluene at 50° is found
to be 110 mm. What is the composition of the solution?
Problems, Chap. 1 Introduction and Description of the Properties of Matter 27
20. Use Raoult’s law to predict the volume solubility of CH;F in benzene at
25° and 1 atm. The vapor pressure of CH,F at 25° is 40 atm. The density of
liquid benzene at this temperature is 0.875 g./cc.
21. To 24.00 cc. of gas, known to contain only hydrogen, nitrogen, and excess
water, is added 15.50 cc. of oxygen, both measured at 250 mm. and 0°. Following
ignition the residual gases occupy 39.50 cc. at 22.3 mm. and 0° C. Find % H, in the
original gas mixture.
22. The apparent molecular weight of a gas mixture is defined by M,,. =
WRT/PV. Find M,,, for air (20% O,, 80% N,).
23. A mixture of H, and CH, has a density 0.25 as great as the density of O,
at the same temperature and pressure. Find the % CH, in the mixture.
24. An oil diffusion vacuum pump has a speed of 501./sec. at 10-4 mm. (a)
How long would it take at this rate to remove | cc. of gas originally measured
at 1 atm. and at the same temperature?
In order to operate the diffusion pump described above, its exhaust pressure
must be maintained at 10-'! mm. or better by a mechanical fore pump. (b) What
is the required speed of the fore pump ?
25. A sealed glass bulb terminates in a capillary tube of uniform diameter and
length L. It contains gas at an unknown pressure P and a liquid of density d. The
bulb is rotated to a vertical position in a constant temperature bath with the
capillary tube down. The confined column of gas has length /,, the liquid /,.
Solve for P, ignoring the effect of surface tension.
26. A capillary glass tube of uniform diameter and temperature contains gas
samples A and B, separated by a short column of mercury, L mm. in length.
The ends are sealed. In horizontal position the confined gases occupy volumes
amm. and 5 mm. in length at the common unknown pressure P. In vertical
position the lengths become a’ and b’. Solve for P.
27. The dependence of barometric pressure upon altitude is described by the
barometric formula. It can be derived by considering a vertical column of unit
cross-section area containing an ideal gas of molecular weight M at uniform
temperature 7. The weight of gas above compresses the gas below. Representing
gas density at altitude h by p, dP = pgdh. Find P as a function of h.
FIRST LAW
OF THERMODYNAMICS
titatively, into heat. One can then compare the heat liberated when a given
amount of gasoline burns in an open vessel with the combined heat + light +
work when the same net change occurs in an internal combustion engine. Or
one can conduct the process
Zn | Ca** (aqg.) = Zn** (aq) -- Cu
in a beaker, measuring the heat liberated, and compare this with the total
energy liberated when the same process occurs in a galvanic cell. In any such
comparison of two or more processes which bring about exactly the same change
in state, it is always found that the total energy produced or consumed is precisely
the same.
In the study of any process, the energy term is measured by the effects pro-
duced in the immediate surroundings. In a calorimetric process such as
12 g. C (25°) + 32 g. O, (1 atm., 25°) + 10 kg. water (25°)
= 44 ¢. CO, (34.6°, 1 atm.) + 10 kg. water (34.6°)
“12g. C and 32g. O,” is the system and “10 kg. water” the immediate sur-
roundings. Our repeated experience is consistent with the interpretation that
when a system is in effective contact with its immediate surroundings, so that
energy can be interchanged, and when these are isolated from the rest of the
laboratory and universe, then for any process
Heat as such cannot be measured at all, but the effects which it pro-
duces are measurable. When a system absorbs energy as heat, its temperature
increases. On this basis, heat was formerly measured
by the increased temperature of a body of water. The
amount of heat required to raise the temperature of
1 g. of water from 14.5° to 15.5° became a unit of
heat, the calorie (cal.). Since we shall demonstrate
later by rigorous thermodynamic development that
temperature is fundamentally defined and measured
in terms of heat, the 15° cal. is no longer suitable.
There are also experimental difficulties which arise
from using the 15° cal. Another definition, which
avoids circular logic, is that heat is that form of
energy which melts ice. This definition has an experi-
mental advantage, since quantity of heat can be
related to quantity of ice melted in an ice calorimeter
(to be described in section 2. 3). In turn, the amount
of electric energy required to melt a given amount of
ice can be accurately measured. The defined calorie
equals 4.1840 joules.’
Measurements of temperature change vs. energy
input are often carried out in a calorimeter of the
type indicated in Fig. 2.1. Heat is delivered by dis-
sipation of electric energy in a resistance heater at a
measured voltage and current for a measured time.
FIG. 2.1. Adiabatic calori: he temperature of the outer jacket is manually or
meter (schematic). T, sample automatically adjusted so that it is as close as possible
thermometer; 1’, jacket to that of the sample. This minimizes heat transfer
thermometer; S, sample; R, between the sample and its surroundings; a process
sample heating resistance;
Riedie tincestianie: carried out under such conditions is said to be adia-
A, ammeter; V, voltmeter; tic. For reasons of experimental convenience, such
t, timer. measurements are often made with the sample sub-
jected to a constant pressure, usually that of the
atmosphere.
‘Modern physical theory regards mass as a form of energy, quantitatively related through
the equation due to Albert Einstein, energy = mass x (velocity of light)?, and the con-
servation law is broadened to include mass as well as energy.
*The magnitude chosen for the defined calorie is such that this amount of energy will
raise the temperature of 1 g. of water from 14.5° to 15.5° within the precision of ordinary
measurements.
Secu 2.2 First Law of Thermodynamics 31
EXERCISE 2.1
An electric timer ft, an ammeter A, a voltmeter V, a resistance heater R, and an
adiabatic calorimeter are arranged as shown in Fig. 2.1. Express in defined calories
the heat generated within the calorimeter system when a current of 2.250 amp.
at 1.500 v. is passed through the heater for 1000 sec. Ans. 806 cal.
slope = 1.0 ee
liquid
100
|
|
Current x voltage x time |
= Energy (calories) |
.
gl solid
50 o|
= |
a
liquid
10
fe
===
=5=F
-
O 500 1000 1500
32> First Law of Thermodynamics Chap. 2
physical state. It is described by the property called the heat capacity, C, which
is defined by
dq _ (2.1)
where g is the quantity of heat. The heat capacity is given by the slope of the
curve in Fig. 2.2. The observation ofg versus T may be made either at constant
pressure or at constant volume. The heat capacity of one phase varies
smoothly with temperature, as shown in Fig. 2.3
A mean value of the heat capacity, C, for a finite temperature interval may
be computed from the temperature increase resulting from a finite heat input.
cope SVS
C= AT (2.2)
Cr aera
TT, ao IE een5 Me
(2.3)
For instance, we observe that the quantity of energy required to raise the
temperature of the specimen described in Fig. 2.2 from —15° to 0° is 7.5 cal.
By equation (2.3) we obtain
C = 0.5 cal. deg.~!
for the temperature interval —15° to 0°. This may also be taken as the ap-
proximate value of C at the mean temperature —7.5°.
EXERCISE 2.2
In Exercise 2.1 if the calorimeter consists of a block of metallic lead weighing
1000 g. and its temperature rose from 0.0° to 26.3°, what is the atomic heat capa-
city of lead in defined calories? Ans. 6.36 cal. deg.~! (g. atom.)~!.
EXERCISE 2.3
Evaluate the constants a and b in equation (2.4) for the case of H,O (g) from the
data in Fig. 2.3 over the temperature range 373-1500°K.
LVS, = TAY toy = 3h) SO
In order to represent the data for other substances, such as CH, (g) shown
in Fig. 2.3, it may be necessary to add a third term to equation (2.4), yielding
Cs=a+bT + cf?
or Ce=a+br+
cr
or C,=¢--
br + clo} (2.5)
and again the constants a, b, and c are chosen for best fit to the experimental
values of Cp. The curve drawn through the data for methane in Fig. 2.3 cor-
responds to the equation
Cp = 3.381 + 18.044 x 10°-°T— 43.00 x 10-7T?
Table 2.1 contains a representative selection of heat capacity data given in
terms of the constants in equation (2.5).
TABLE 2.1
MOLAR HEAT CAPACITIES AT CONSTANT PRESSURE*
Ce = a+ bT + cT? (300-1500°K)
Substance a bx 102 e x 10° Cp at 25°
Hy (g) 6.947 —0,200 4.808 6.89
O» (g) 6.148 3.102 —9,23 7.02
Nz (g) 6.524 1.250 —0.01 6.90
HO (g) 7.256 2.298 2.83 8.03
CO, (g) 6.214 10.396 —35.45 8.87
CO (g) 6.420 1.665 —1.96 6.97
CH, (g) 3.381 18.044 —43.00 8.54
CoH, (g) 2.830 28.601 —87.26 10.41
CoH, (g) 2.247 38.201 —110.49 12.59
NH; (g)t 6.189 7.887 =7.28 8.52
*From H. M. Spencer and J. L. Justice, J. Am. Chem. Soc., 56, 2311 (1934), H. M. Spencer
and G. N. Flannagan, ibid., 64, 2511 (1942), H. M. Spencer, ibid., 67, 1859 (1945), H. M.
Spencer, Ind. Eng. Chem., 40, 2152 (1948).
+300-1000°K.
From tabulated data on heat capacities one may reverse the operations
described above to compute the heat required to increase the temperature of
a substance over any interval within the range of validity of the data. Equation
(2.1) integrated over a finite interval of temperature becomes
the area under the curve between the temperature limits is measured as indicated
in the case of heating methane from 300°K to 1000°K. For this process gp =
9121 cali mole=7.
When the heat capacity is known as a function of temperature, the integral
in equation (2.6) becomes, e.g.,
q I [@+or+ cT?) aT
T\
(2.7)
aT, —T) + (3 —T) + (TT)
EXERCISE 2.4
Using the empirical equation for the heat capacity of methane, compute the
energy required to raise the temperature of one mole of this substance from 300°K
to 1000°K. Compare with the result of the graphical integration indicated in
Fig. 2.3.
EXERCISE 2.5
Using the heat capacities and latent heats given above, compute the energy
required to carry out the following process with 10g. of H,O: H,O (s, 263°K,
1 atm.) = H,O (g, 383°K, 1 atm.). Ans. 7290 cal.
The latent heat of fusion of ice can be measured precisely, and then used as
a calorimetric standard. The ice calorimeter, illus-
trated in Fig. 2.4, utilizes this method. The process
to be measured occurs in chamber D, in good thermal
contact with an ice-water mixture. Ice melts in the
well-insulated vessel A by absorbing heat from the
isothermal reaction system. Since the densities of ice
and water at 0° are 0.917 g./cc. and 1.000 g./cc., res-
pectively, the volume of the ice-water mixture de-
creases by 0.01 cc. for each 0.12 g. of ice melted by
absorbing 9.5 cal.
EXERCISE 2.6
That is
work = force X distance
pressure = force < areay*
work = pressure X area X distance
= pressure X volume
Then
dw = Pop AV (2.8)
and in a finite process of expansion the work will be given by the definite
integral of equation (2.8).
ee ayy (2.9)
Vi
EXERCISE 2.7
The density of liquid water at 1 atm. and 0° is 0.99984 g. cc.-!; and at 1 atm. and
100°, 0.95835 g. cc.-! Evaluate the work of expansion when 18 g. of water is
heated from 0° to 100° at a constant pressure of 1 atm. Ans. 7.85 xX 10-41. atm.
When only the initial and final temperatures are known, the coefficient of
thermal expansion is needed. This may be obtainable from an equation of
state, as in the case of an ideal gas where, at constant pressure, PdV = nRdT.
When P,,; and Psu, are actually the same, and they often are not, equation
(2.8) becomes dw = Prag dVqa,, or simply PdV. Since V and T are the variables
and are related by PV = nRT, it follows that dw = nRdT, or
EXERCISE 2.8
Calculate the expansion work done when 1 mole of ideal gas is heated at a con-
stant pressure of 1 atm. from 0° to 100°. Ans. 8.2 1. atm.
Vi
The same overall change from an initial volume of 2.241. to a final volume
of 22.41. may, however, be accomplished in an infinite variety of ways, each
yielding a different amount of work by the system. For example, reduce the
pressure of the surroundings to 5 atm.; this will result in an increase in volume
to 4.48 1. Next reduce P,,,. to 1 atm.; the volume now increases to the desired
final value of 22.41. The total work done in this two-stage process is given by
the sum of two terms corresponding to equation (2.12).
Wes (44s = 24) 2 02.4 — 4)
= 9 nel atinsmoles:
The same total change in volume may be carried out in a many-stage process,
in which the pressure of the surroundings is decreased in small steps until the
final value is reached. The course of such a process is indicated by path 3 in
Fig. 2.5, along with the course of the two processes described above. The third
path would obviously result in a larger value of w than either path 1 or 2. This
is most easily seen if we observe that the value of w is given by the area under
38 First Law of Thermodynamics Chap. 2
w= | Pav
Vo
We see that paths 1, 2, and 3 enclose successively larger areas between the
fixed initial and final volumes.
Since the gas will not expand unless the pressure of the surroundings is
less than that of the gas, there is evidently a limit to the work which can be
obtained in expansion between these fixed initial and final states. This limiting or
maximum work is obtained when the pressure of the surroundings is only infini-
tesimally less than the pressure of the gas at all times during the expansion from
V, to V,. Substituting P,,, for Purr, in equation (2.9) we have
ee li ieencl (2.13)
Vo
Vi
(2.14)
The maximum work in an isothermal expan-
sion is also measured by the area under the
smooth curve, whose equation is PV =
const., In) big. .2:5. Since Pin Payton
isothermal expansion of an ideal gas, equa-
tion (2.14) may also be written in the form
Vi Sane (2.15)
Py
EXERCISE 2.9
V (liters) Calculate the maximum work for the expan-
FIG. 2.5. Isothermal gas ex- sion of 1 mole of ideal gas from 2.241. to
pansion, one mole ideal gas 22.41. at 0°; repeat at 100°.
at 273°K. Ans. 51.5 and 70.4 1. atm.;
1247 and 1705 cal.
area under this curve generated from right to left. This area has a negative value,
indicating that work is done on the gas in compression. The work done by
the surroundings would be less than that computed above if a path such as 5,
consisting of a series of small pressure increments, were followed. The work
done on the gas would evidently be minimized if the pressure of the surroundings
were only infinitesimally greater than the pressure of the gas, and the work done
by the gas in compression (which is negative) would be the maximum. This
limiting case is obviously described by equations (2.13), (2.14), and (2.15), but
in compression the value of V, is less than that of V, and the value of Wyyay
is negative.
It has been shown that there is a maximum amount of work which a gas
can do on its surroundings in an isothermal expansion. The value of the maxi-
mum work is determined by the initial and final volumes, by the temperature,
and of course by the amount of gas involved. The maximum work done by the
gas in compression is given by the same relation. The process which yields
maximum work in expansion may also be described as reversible, since each
infinitesimal state of the system can be traversed in reverse order in a subsequent
compression which utilizes only the potential energy stored in the surroundings
during the first half of the cycle. The system and its surroundings can both be
restored to their initial states, and therein lies the test of reversibility. On the
other hand, a process such as that described by path | in Fig 2.5 is an irreversible
expansion, since a succeeding compression cannot retrace that path. Further-
more, although the system can be restored to its initial state by utilizing potential
energy from the surroundings, over and above that stored during expansion,
the surroundings do not regain the initial state.
The maximum work of isothermal expansion can be computed for real
gases by returning to equation (2.13) and substituting the appropriate relation
for P as a function of V. If the equation of state (see Chapter 1) is
PV = nRT + nbP
the integrated equation is
= VO RIE = V, — nb
si ING
Wmnax. I (y — =) Boer! Ip V, —nb ( )
If experimental values of P vs. V for a gas are available these may be graphed
and the value of
“ale
Vy
obtained from the area under the curve between the limits V, and Vy.
EXERCISE 2.10
From data given in Fig. 1.2 plot P vs. Vx, for oxygen and find, by graphical
integration, Wy,ax. for the expansion of 1 mole of oxygen from a pressure of 500
atm. to a pressure of 200 atm at 0°. Compare with the result of application of
equation (2.15). Ans. Real Wmax. = 151. atm.; ideal Wax, = 20.61. atm.
40 First Law of Thermodynamics Chap. 2
and w evaluated.
w = 1 atm. x (—1.49 x 10°-*) 1. mole"!
= —1.49 x 10°-* 1. atm. mole“!
The work of expansion in vaporization is much larger, since a gas is formed.
For example, the molar volume of liquid water at 100° is 18.8 cc., while the
molar volume of water vapor at 100° and | atm. is 30.21. Thus the change in
molar volume on vaporization under these conditions is practically equal to
the molar volume of the vapor, 30.2 1., and the work of vaporization obtained
from equation (2.17) is
w= 1 x 30.21. atm. mole7!
EXERCISE 2.12
Evaluate w approximately at 25° and 1 atm. for the process CaCO,(s) = CaO(s)
+ CO,(g) Ans. 24.41. atm. mole~!.
The positive electrode, at which electrons from the external circuit are con-
sumed, is called the cathode, and at this electrode copper ions are reduced to
copper metal.
2e=—— Cut = Cu
involves either heat or work. When a system absorbs heat, q, this energy is not
lost; it is stored, and can be recovered. Absorption of heat increases the internal
energy, or the energy content, E, of the system. When a system performs
work, w, this is done at the expense of the total energy of the system. These
two ideas can be combined in a single equation, applicable to any change in
state of any system.
E, — E, = AE=q—w (2.19)
E, and E, represent the values of the internal energy of the system in the
initial and final states, whether for reversible or irreversible processes. Equation
(2.19) is our most useful statement of the first law of thermodynamics. Only g
and w can be measured. The internal energy is an idea, or concept, which can-
not be measured, as such. Nor can changes in the internal energy, AE, be
measured, as such. But measurement of g — w establishes AE, by definition.
The concept of internal energy cannot be explained within the framework of ther-
modynamics, and it does not lend itself to representation by models or pic-
tures. It is an abstraction, but it has operational significance, namely EF, —
E, =q — w. When a weight is lifted, the internal energy of the system increases.
To say that this represents an increase in potential energy does not explain it,
since potential energy is only another aspect of the same concept.
EXERCISE 2.14
processes of gas expansion described in Fig. 2.5. It was shown there that the
change in state
gas (2.241., 10 atm., 0°) = gas (22.41., 1 atm., 0°)
can be carried out by a variety of paths with various values of w. The first law
tells us that AF is the same for all these processes, and therefore equation (2.19)
tells us that g must increase or decrease as w decreases or increases.
The statement that AE, the change in internal energy, depends only on the
initial and final state and not on the path applies to all kinds of processes, both
physical and chemical. A characteristic of a system such as the internal energy
E, whose value depends only on the state of the system and not on its history,
is known as a thermodynamic property of state. Other useful properties which
meet this requirement will be introduced later. The change in internal energy,
AE, may be called a function of state. As indicated above q and w are not, in
general, functions of state, since their values depend on the path described by
the process.
Many processes of interest occur cyclically. In a simple cycle, the final
state of the system is identical with the initial state, regardless of the path, while
the initial and final states of the surroundings may or may not be the same.
The restoring process need not retrace the forward process. A cyclic process
which produces maximum work is also said to be reversible. It must not be sup-
posed that this necessarily entails retracing the forward process, but only that
INE corwatd — NBs oe or,
Wrorward == —~ Wreyerse
Dorward — — reverse
and therefore
Aeycle = Weycle (2.20)
This places no restriction on the values of g and w except that they must be
equal. Referring again to Fig. 2.5, consider the expansion-compression cycle
which is described by paths 1 plus 4. The total work done by the gas in this cycle
is 20.2 — 202.0 = —181.81. atm. The sign indicates that the net result is work
done on the gas. Since AE for the cycle is zero, equation (2.19) indicates that the
heat absorbed by the gas would be —181.81. atm. = —4400 cal., and again
the sign indicates that heat flows from the system to the surroundings.
The device of a cycle is useful to evaluate AF and other functions of state
for a variety of noncyclic physical and chemical processes. Imagine a com-
pletely generalized process of interest for which the initial state is A and the final
state is B, as indicated in Fig. 2.7. No restrictions as to the nature of the pro-
cess are implied. It may not be possible to measure g,4, and w,,, and thus com-
pute AF,,, for the particular process of interest. However, it is often possible to
measure or compute q’s and w’s for another process which produces the same
Sec. 2.10 First Law of Thermodynamics 45
IND ee {_ Cy dT
To
(2.22)
If experimental values of C, as a function of 7 are available, graphical integra-
tion may be employed, or an empirical equation for C, may be inserted in equa-
46 First Law of Thermodynamics Chap. 2
tion (2.22) and integrated (see section 2.2). For approximate results or for
small temperature change, C, may be considered constant and
NE ve CaN (2.23)
Ao Hh Chee Wy)
AH = AE + APV (2.26)
and the value of AH depends only on the initial and final states of the system,
not the path.
For a process at constant pressure
AH, = AE, + PAV (2:27)
and substituting by equation (2.24),
AH> =e dp (2.28)
(5),
OE
=0 (2.30)
In other words, the internal energy of an ideal gas is a function of the temper-
ature alone, given by
GE ==Gr df [see equation (2.22)]
For a change in state of an ideal gas,
gas (P,, V;, T,) = gas (P2, V2, T,); AB rota
we can devise two partial processes:
gas (Re Vi, T,) = gas cc Vs Ls AE,, = 0
ideal gas
dH, = dE + Pdv [see equation (2.25)]
dH, = C, dT + PdV
For one mole of gas at constant pressure, with V and TJ as variables, the work
of thermal expansion is PdV = RdT. Substituting into the preceding equation
gives
The work of isothermal gas expansion was considered in section 2.6. Since
E is invariant at constant T
q=— "= [Poet dV
how the final temperature may be computed. This relation may be formulated
from the first law, with q = 0, thus
w= —AE
P,V,/T, = P,V,/T,
lOexe2245 sex Vs
Or
HEIL OG
Solution of these simultaneous equations gives T, = 218°K and V, = 17.91.
Comparison of this result with that of the isothermal expansion from 10 to
1 atm. indicated in Fig. 2.5 shows that the final volume in the adiabatic expan-
sion is smaller, corresponding to the decrease in temperature. Also, since the
volume increase is less, the work of expansion is less.
Y= Panels ae V,)
cf
ra
RYaoe
hs Ve
Cy liees
ie V; (2.36)
Crlnzt ao
pas= Ring (237)
EXERCISE 2.18
Apply equation (2.37)to determine the final temperature in the adiabatic, reversible
expansion of an ideal gas with Cy = 8.0 cal. mole~! from 10 atm. and 0° to 1 atm.
Ans. 173°K.
SUMMARY, CHAPTER 2
4. Electric work
Ye NEF
5. Applications of first law
a ha = 0; Qeycle == Weycle
AE, = Gr
AHp
= qp
52 First Law of Thermodynamics Chap. 2
PROBLEMS, CHAPTER 2
1. Evaluate Cp for one mole of nitrogen at 125° using the parameters of Table
Dale Find the corresponding value of Cy.
Ans. Cp = 7.02 cal. mole=! deg=1, C, = 5.03 cal. mole=! deg;!.
2. The average Cp of carbon dioxide from 0-100° is 8.90 cal. deg.-! mole7!.
For heating 10 moles of this gas from 0° to 100° at constant pressure, find q, w,
AH, KE. Ans. q = AH = 8.90 kcal., w = 1.99 kcal., AE = 6.91 kcal.
3. Use the data of the preceding problem to find g, w, AH, AE for heating 10
moles of carbon dioxide at constant volume from 0° to 100°.
Ans. q = AE = 6.91 kcal., w = 0, AH = 8.90 kcal.
4. How many grams of ice at 0° would have to be added to 100 g. of water at
50° to bring the final temperature of the system to 10°? Ans. 44.5 g.
5. Benzene melts at 5.48° and its heat of fusion at this temperature is 30.3 cal.
g.-!. Cp for the solid is 0.30 cal. g.~! deg.~! and Cp for the liquid is 0.40 cal. g.~!
deg.-1. Find AA for the process
Gc; (S102 atm) = (CH, (50s. lratm>)
Ans. 3.94 kcal.
6. The heat of vaporization of benzene at 80.2° (n.b.p.) and 1 atm. is 94.3 cal.
g.~! For the process
GH, des02eslkatm:)i— CAH. (ess02ealkatms)
compute Wmax. AH, and AE. Ans. w = 695 cal.
7. A 10g. sample of helium initially at 50 atm. and 25° is allowed to expand
isothermally against a constant external pressure of 1 atm. Assume that the gas
is ideal and compute qg, w, Wmax. AH and AE.
Ans. gq = w = 1.45 keal’, Wax. = 5.79 kcal:
AH = AE=0
8. One hundred g. of nitrogen gas (Cp = 6.96 cal. mole! deg.~!) initially at
25° and 10 atm. is allowed to expand adiabatically against a constant pressure of
1 atm. Assume that the gas is ideal and compute the final temperature, AF and AH.
Ans. T = 222°K, AE = —1349 cal., AH = —1889 cal.
9. One mole of helium gas (Cy = 3 cal. deg.~! mole~!) is confined in a cylinder
and piston at 25°. The area of the piston is 100 cm.? and the gas is subjected to
a total external force of 10 kg. acted upon by the acceleration of gravity. (The
acceleration of gravity = 980 cm. sec.~?) To what extent must the gas be heated
Problems, Chapter 2 First Law of Thermodynamics 53
in order to raise the piston 5 cm. ? How much heat will be absorbed in the process ?
Ans. 0.59°, 2.95 cal.
10. For the constant pressure process
C,H,6(g) + 70.(g) = 2CO,(g) + 3H,0()
AA, = dp = —372.800 kcal. mole!
(a) Find w and AE for the process.
(b) The heat of vaporization of water at 25° and 1 atm. is 10.519 kcal. mole=!.
Find AAy,, and AF, for the process
C,H,(g) + $0.(g) = 2CO,(g) + 3H,O(g)
Ans. (a) —1.48 keal., AE = —371.325 kcal.
(b) AH = —341.240 kceal., AE = —341.535 kcal.
11. When the process
H,(g) + 2AgCl(s) = 2HCI (1 molar sol.) + 2Ag(s)
occurs at 25° in a galvanic cell a zero current e.m.f. of 0.2329 v. is observed.
When the same change in state is carried out in such a way that no electric work
is done, g = —17,400 cal. per mole of hydrogen consumed. Evaluate wWynax and the
corresponding grey. for operation of the galvanic cell.
ANS. Wmax. = 10.74 kcal.
Grey. = — 6.67 kcal.
12. From the empirical equation given in Table 2.1 compute the heat capacity
at constant pressure of carbon dioxide at 50°. Compute the heat capacity at
constant volume at this temperature, assuming that the gas is ideal.
13. Use the values of the heat capacities computed above as constants over
the temperature range 0-100° in the following calculations: A 10g. sample of
carbon dioxide initially at 1 atm. and 0° is heated to 100°. Assume that the gas
is ideal.
(a) For the case that pressure is constant, compute the final volume, qg, w, AE,
and AH.
(b) For the case that volume is constant, compute the final pressure, g, w, AE,
and AH.
14. Insert the empirical equation for the heat capacity of carbon dioxide as a
function of temperature in the expression for q, integrate and repeat the com-
putation of g, w, AE, and AH for the process described in problem 13 (a).
15. To determine the heat capacity of methanol vapor, a steady flow of vapor
is maintained over an electric heater at constant voltage and constant current.
In an experiment at 76.5° the heater dissipated 0.04674 cal. sec.-!; 94.45 g. of
CH;OH passed through the calorimeter during 45 min. at an average tempera-
ture rise of 2.464°. Find the specific heat of methanol.
16. (a) One mole of an ideal gas at 1 atm. and 0° is to be compressed isothermally
until its volume is 1 1. What is the minimum work which must be done on the gas?
(b) If the initial state is 1 atm. and 100°, what is the minimum work to compress
the gas sample to 1 1. at 100°?
17. What is the lowest temperature that can be attained by adiabatic expansion
54 First Law of Thermodynamics Chap. 2
temperature rise T, — T,. Work effects are negligible. The same combination of
system and surroundings is raised from 7, to T, by a measured input of electrical
energy,n Fé.
(a) Find a relation expressing Q in measured quantities.
(b) Show explicitly how the first law has been invoked.
30. For a gas which obeys the equation of state PV = RT + bP find a relation
between 7 and Vfor adiabatic reversible expansion.
31. In gas mixtures, each constituent contributes to the total heat capacity as
though it were present alone. Find the empirical equation for the heat capacity
at constant pressure of a mixture containing 0.5 mole CHy,, 1.0 mole C,H,, and
2 moles C,Hg.
32. A 201. vessel contains dry gas at 20° and 785mm. A large quick-acting
valve opens briefly, establishing equalization of pressure with the atmosphere at
723 mm. After adiabatic expansion, the residual confined gas recovers its original
temperature. The final pressure is 745 mm. What is the value of Cp?
33. The equation PV = RT + bP describes the behavior of hydrogen at 100°
over a range of 10° atm. when b = 0.01641. Evaluate w,,,, for the isothermal
expansion of 1 mole of hydrogen from 10? atm. to 1 atm.
34. With reference to the preceding problem, evaluate Cp — C, for one mole
of hydrogen at 100° and 10 atm.
35. Water vapor at 100° and 500 mm. is expanded reversibly and adiabatically.
The vapor pressure of water can be expressed by log Prin. = —2120/T + 8.63 and
the mean molar heat capacity taken as 8.0 cal. mole“! deg-!. At what pressure
will condensation begin?
36. A very large mass of dry air (take Cp as 7 cal. mole“! deg™!) at 20° and
750 mm. is pushed by steady winds over the top of a mountain range where the
pressure is 550 mm. Estimate the air temperature on the mountain top.
37. Real gases at high pressures are non-ideal and both H and E become
functions of P (or V) at constant 7. For the free, adiabatic expansion (AH = 0)
of oxygen from P; = 200 atm. and 7, = 298° to P, = 0.1 atm., the final temper-
ature is T, = 248°. Devise a simple cycle from which AA7, for the isothermal
expansion O, (200 atm., 248°K) = O, (0.1 atm., 248°K) can be evaluated. Find
AAr,.
38. Two gases are confined in a long cylinder and separated by a membrane at
a uniform temperature and unequal pressures. Their mean heat capacities are
C, and C,. The membrane is broken and A expands, compressing B. Find an
expression relating AT, and AT>3.
39. Derive a relation between Cp and 7for adiabatic reversible expansion of an
ideal gas starting with the equation dH = dE + d(PV).
THERMOCHEMISTRY
Therefore, under these two specific conditions the heats of reaction are
thermodynamic functions whose values depend only on the initial and final
states of the systems under study. This important application of the first law
of thermodynamics will enable us to develop several useful relationships a-
mong heats of reaction.
3.2 Calorimetry
to form carbon dioxide and liquid water the following data were obtained:
Weight of benzoic acid, 0.9350 g.; weight of water in contact with the bomb,
1235 g.; heat capacity of bomb, 535 cal. deg.-!; observed temperature increment,
3.340°; average temperature, 25°. Ignore minor corrections and compute the molar
heat of combustion of benzoic acid at constant volume.
Ans. qy = —771.5 kcal. mole-!.
Since the work done by the system in bomb calorimetry is zero, the heats
of reaction observed in such a device are to be equated to AE for the process,
as in equation (3.1). For example, the combustion of a gram ofliquid methanol
in a bomb calorimeter at 25° liberates 5410 cal. Therefore, for the constant
volume, constant temperature process
CH;OH (I) + 30, (g) = CO; (g) + 2 H,0 (1)
Qv = AEgog = — 173.34 kcal. mole“!
Many processes are carried out under constant pressure, often that of the
atmosphere, and the heat effect is given by AH rather than AE, as shown in
equation (3.2). For a constant pressure process, the relation between AH and
AE is given by equation (2.27).
The initial and final volumes of the system will be largely determined by the
volumes of gases present in the initial and final states, and
NVA REP. An = nr; — Ni;
where n, and n, denote the final and initial numbers of moles of gas in the
system. Substitution in equation (3.3) yields
qr = AHp = AE + AnRT (3.4)
While the processes which take place in a bomb calorimeter are not con-
stant pressure processes, the internal energies of the products and reactants
are practically independent of pressure. For the gases this is equivalent to say-
ing that they are approximately ideal. Therefore we may state as a good ap-
proximation that for the constant pressure, constant temperature process
TABLE 3.1
HEATS OF COMBUSTION AT 25°*
(to form liquid water and carbon dioxide)
A Agog (combustion)
Substance (kcal. mole-')
Hy (g) —68.317
C (s) (graphite) —94.05
CO (g) — 67.64
CH, (g) —212.798
C2Hg (g) — 372.820
C2Hy (g) —337.234
C.Hy (g) —310.615
CgHe (1) —780.98
* From Selected Values of the Properties of Hydro-
carbons, Circular C461, National Bureau of Standards,
1947.
EXERCISE 3.2
From the data given in Table 3.1 find AE,., for the constant pressure combus-
tion of ethylene. Ans. —336.054 kcal. mole7!.
AH(A—M)
aA+bB+.-.-- >mM-+nN-+.---
a. Vhs
JJ+kK+-:--
It is evident that
A Figg (iii) = AA (i) — AA (i)
= — 94.05 + 67.64
A Hog (iii) = — 26.41 kcal.
EXERCISE 3.4
Using data given above compute the value of AAyo, for
2¢ 23H C6. H.
and therefore the value of AH... for this process is given by
products. In each case the molar heat of combustion is multiplied by the ap-
propriate stoichiometric coefficient.
EXERCISE 3.5
For the combustion of liquid ethanol to form carbon dioxide and liquid water,
Afio9, = —326.70 kcal. mole. Find AAyo, for the process
2 C (graph.) + 3 H; (g) + 40, (g) = C,H;OH (1)
Ans. AH = —66.36 kcal.
TABLE. 3.2
HEATS OF FORMATION AT 25°*
(from the elements in their standard states)
A598 AH x98
in Table 3.2, we may compute the value of AH... for the oxidation of ethane
to ethanol by application of Hess’s law (equation 3.8).
The formation of ethane and ethanol from the elements corresponds to the
following processes:
64 Thermochemistry Chap. 3
and find AH. for the combustion from the data of Table 3.2. From equation
(3.9)
ANG los oe 2 AH,(CO,) = 3 AH,(H,O, 1) a AH,(C,He)
In order for equation (3.9) to be valid, it is evident that the heats of formation
of the compounds involved in a process must refer to identical states of their
constituent elements. Only in this case will the heat contents of the elements
disappear in the equation. In the case of
HI(g) + NaC\(s) = Nal(s) + HC\(g)
the heats of formation of both HI(g) and Nal(s) must refer to a single state
of the element iodine. In the example given, the formation processes for HI(g)
and Nal(s) both refer to I,(s), although one could just as well have used the
values of AH,., for the formation processes from gaseous iodine. In such a
case, AH,(Nal, s) and AH,(HI, g) would both be altered by exactly the same
amount, the AH,,. for sublimation of 4I,. Therefore no particular state of
an element need be chosen, but the state chosen must be used consistently
within any set of calculations.
In order to make the tabulated values of heats of formation consistent
in this respect, it is convenient to specify standard states for the elements. At
each temperature the standard state of the element is that physical state (g, 1, s)
which is most stable, at a pressure of | atm. By this convention the standard
state of hydrogen at 25° is the gaseous state (H,, g) at 1 atm.;' the standard
state of mercury is the liquid at 1 atm. With crystalline solids which exhibit
1This definition of the standard state of gases assumes that they are ideal. For greater
precision in dealing with real gases, the standard state is defined as the state of unit
fugacity, a property which may be described as the ideal pressure of a real gas.
66 Thermochemistry Chap. 3
more than one crystalline form it is also necessary to further specify the form.
Thus the standard state of carbon at 25° is graphite at 1 atm.; the standard
state of sulfur at 25° is the rhombic form at 1 atm., but at 100° the standard
state is the monoclinic form at 1 atm.
The specification of the standard state is also extended to compounds, ex-
cept that the restriction as to physical state is not used and the standard state
of a compound is simply that of 1atm. Thus water vapor at | atm. is the stand-
ard state of water vapor at any given temperature and liquid water at | atm.
is the standard state of liquid water at any given temperature.
A superscript o attached to the symbol for the change in heat content thus,
AH?, denotes the standard states. For example, the standard change in heat
content for the formation of liquid water from its elements refers to
H,(g, 1 atm.) + 40,(g, 1 atm.) = H,O(, | atm.)
AHS, = — 68.32 Keal.
and the standard change in heat content for the formation of water vapor
refers to
H,(g, 1 atm.) + 40,(g, 1 atm.) = H,O(g, | atm.)
AH, = — 57.80 kcal.
In practice, the pressures of the reactants and products have small influence
on AH except for pressures considerably above | atm. For an ideal gas the
value of AH for isothermal compression or expansion is zero, and for real
gases, liquids, and solids it is negligible for all except extreme changes. The
specification of the phase or state of aggregation of the elements, however, is
an important consideration in reactions of formation.
In the example above, the value of AH; (H) is one-half the heat of dis-
sociation of H, or 51.5 kcal/g. atom. For iodine, however, the AH, (I) corre-
sponds, by definition, to
31,(s) = I(g)
It is therefore necessary to include the heat of sublimation of iodine
I,(s) = I,(g), AAs = 14.88 kcal.
EXERCISE 3.9
Show that AH,(I) is 25.5 kcal./g. atom.
EXERCISE 3.10
From information given in this chapter find AAs for HI(g) = H(g) + I(g).
Ans. AH bo = 70.8 kcal.
The bond dissociation energy should not be confused with a quantity some-
times called the bond strength. A single illustration will suffice. For the com-
plete atomization of CH,,
CH,(g) = C(g) + 4H(g)
the value of AH can be obtained from
C(graphite) = C(g), AH%s = 171 kcal.
C(graphite) + 2H,(g) = CH,(g), AHS 3 = — 18 kcal.
H,(g) = 2H(g), AA%s = 103 kcal.
Combining these data, for atomization of CH,
AH%, = 171 + (2 x 103) — (— 18) = 395 kcal.
and the bond strength is #22 = 98.7 kcal.
EXERCISE 3.11
Show that the thermochemical quantity just found is the sum of the four individual
bond dissociation energies:
Dou,-u + Dow,-a + Dou-n + Do-n = 395 kcal.
Gaseous ions and solvated ions differ greatly in their chemistry. That will
be considered later, but let us be careful to distinguish between them as, eg.,
H*(g) and H*(aq.). In either case, Hess’s law can apply. If it could not, the
first law would be invalidated.
The common ions form spontaneously in aqueous media, and often very
exothermically, as with sodium. Ionization in water (and other polar media)
is a complicated business. The difficulties may be glossed over by writing
Na(s) = Na*(aq) + e(aq)
and we do not at all understand how H, forms. In the gaseous state the forma-
tion of ions is much simpler in principle, but invariably very endothermic.
The first, and best understood ionization is
H(g)— (2) 4
To avoid misconceptions arising from the symbol H*, it should be understood
that gaseous ions bear little relation to aqueous ions. Well-known gaseous
ions include H;, H}, OH*, CH}, CH. The relevent thermochemical quantity
in cases of simple electron removal is the ionization potential, I, for products
in their lowest energy states. It can be measured from the least energetic
Sec. 3.6 Thermochemistry 69
quantum of light which just produces ions from an atom, radical, or molecule
[K. Watanabe, J. Chem. Phys., 26, 542 (1957)]. It can also be identified with
the minimum electron accelerating voltage required to produce ions by elec-
tron impact in a mass spectrometer. By this method it is possible to distinguish
between simple electron removal.
CH,(g) + e = CHi(g) + 2e, AH = J
and formation of fragment ions,
CH,(g) + e = CH;(g) + H(g) + 2e, AH = A
The thermochemical quantity A is the appearance potential. For the example
chosen A(CH3) = Dou,-n + Icu,
Although the determination of appearance potentials is less precise than
photo-ionization measurements, a wider variety of processes can be studied.
For example, the appearance potentials? of C,H; ions from C,H, and from
C,H, have been determined [J.J. Mitchell and F.F. Coleman, J. Chem. Phys.,
17, 44 (1949)]:
(i) GH,--e=C,H? --H £e
A(i) = 12.7le.v.
(ii) CH 6 C,H CH. 2.
‘A, Gi) =A 203-€, Vv:
When combined, by use of Hess’s law, with the process
(iii) C,H, + CH, = C;Hs + 2H
AA (iii) = 116.3 kcal. mole7!
whose heat effect is obtained from standard heats of formation, a value of
Dou,-1 is obtained
(iv) CH, = CH, 2H
AH = AH (iii) + AGi) — AQ)
= 116.3 + 277.3 — 293.0 = 100.6 kcal. mole“!
EXERCISE 3.13
The appearance potentials for the processes
CH, + e = CH; + H + 2e; A= 144e.v.
CH, +-e = CH; + 2e; A —Oiliexvs
have been observed as indicated. From these data find a value for Dou,-un.
Ans. 99 kcal. mole-!.
2One electron volt (e. v.) per molecule = 23.061 kcal. mole“.
7O Thermochemistry Chap. 3
-AH
EXERCISE 3.14
From the data in Fig. 3.3 compute the heat effect of adding 45 moles of water
to a solution containing 1 mole of ethanol and 5 moles of water at 25° and
1 atm. Ans. q = —1300 cal.
Integral heats of solution or dilution may also be used to compute the heat
effects accompanying the mixing of two solutions of different concentrations.
For example, when a solution of 1 mole of ethanol in 5 moles of water is mixed
with a solution of 2 moles of ethanol in 20 moles of water, the result will be
a solution of 3 moles of ethanol in 25 moles of water and the process of in-
terest is
C,H;OH (5H,O) + 2[C,H;OH (10H,O)] = 3[C,H;OH (2H,O)]
From Fig. 3.3
C,H;OH + 5H,0 = C,H;OH (5H,0),
AF (sol., SH,O) = — 1120 cal.
C,H;OH + 10H,O = C,H;OH (10H,O),
AHF (sol., 10H,O) = — 1760 cal.
C,H;OH + 8.33H,O = C,H;OH (8.33H,0),
AH (sol., 8.33H,O) = — 1650 cal.
The value of AH for the process of interest is obtained by application of Hess’s
law.
AH = 3 x AH(8.33) — 2 x AH(10) — AH(5)
E1650) -— 2(- 1760) (1120) = — 310 cal.
3.7 Thermochemistry of lonic Solutions
The heat evolved in the neutralization of metaboric acid, HBO,, with strong
base is 10,000 cal. mole~!. Compute the heat of dissociation of metaboric acid.
Ans. AH = 3360 cal. mole“.
can be obtained. For example, from the heat of formation of HCl (g) and the
heat of solution to form H*(aq.) and Cl-(aq.) we obtain
EXERCISE 3.17
Taking the heat content of OH~(aq.) to be zero, find the heat content of Ag*(aq.)
Ans. —29.65 kcal. mole~.
TABLE 3.3
ENTHALPIES OF AQUEOUS IONS AT 25°
Engg Fino
Cations (kcal. mole-!) Anions (kcal. mole-!)
H* (aq.) 0 OH- (aq.) — 54.957
Leia (ade) — 66.554 Clee (aa) —40.023
Na* (aq.) —57.279 Bre (aq) —28.90
K+ (aq.) — 60.04 I= (aq.) —13.37
Agt (aq.) 25.31 CN- (aq.) 36.1
Znt++ (aq.) — 36.43 NO;- (aq.) —49.372
Cu**+ (aq.) 15.39 SO,- (aq.) —216.90
Compute AH for
Br-(aq.) + AgCl (s) = AgBr (s) + Cl-(aq.)
Ans. —4.54 kcal.
All discussion up to this point has been concerned with the value
of AH for an isothermal process, usually at 298°K. We now examine the varia-
tion of the isothermal AH with temperature. For the generalized process
aA + bB+...=mM+
N+...
(3.10)
AH, = mH,(M) + nH,(N) — aH,(A) — bH;,(B)
At constant pressure, H,(M), etc., represent molar heat contents of the prod-
ucts and reactants at some particular temperature, T. To find the variation
Sec. 3.8 Thermochemistry 75
with temperature of the isothermal heat of reaction, AH,, we take its partial
derivative with respect to temperature at constant pressure.’
where AH, is for the isothermal process at 7, and AH, is for the same
process at 7).
The relation specified by equation (3.14) may also be described by the di-
agram:
aA + bB+... mM nN +.
| T2
AH,
aA+bB+... >mM+nmN-+...
T\
T)
where AH (react.) = fi [aC,(A) + bC,(B)]dT
T.
T2
and AH (prod.) = i [mC,(M) + 2Cp(N)|dT
T)
QAHr)\
( oT ie
_ lim,
j (AHa
—AH
T» = Wh lace
4Note that ACp = mCp(M) +... —aCp(A)... and that AH = mH(M) + ...—aH(A)
....Both C and H are properties of a system and for any process, including chemical
reactions, the expression Aproperty will always be used similarly as new thermo-
dynamic properties arise.
76 Thermochemistry Chap. 3
TAIN Fp (i
“(Aa + ABT + AcT?) aT
(3.19)
= Aa(T,
—7,) +927 — 79+ (7; — TH
Applying this relation to the computation of Ajo, for
AH oo = AHoys + [| ACp aT
= —11.04 — 2.27 = —13.31 kcal. mole™!
For some purposes, it is more convenient to use
AH, = AH ie"AC» dT
(3.20)
AH, = AH, + AaT + (Ab/2)T?+ (Ac/3)T!
The constant of integration, AH,, may be computed from a known value of
AH at some temperature, usually 298°K. In the case of ammonia synthesis
ioze lOa? 14.49 x 1077
AH, = —11.04 + 7.493(298) — 2
(298)? + 3
(298)8
AH, = —9.16 kcal. mole!
Since the empirical equations for heat capacity do not describe the ex-
perimental heat capacities down to absolute zero, the value of AH, obtained
from equation (3.20) is not to be regarded as the change in heat content for the
process at the absolute zero, but rather as a constant of integration applicable
only with a particular set of empirical heat capacity equations.
EXERCISE 3.20
Substitute the value of AH, for the process
4N, (g) + 3 H, (g)= NH; (g)
in equation (3.20), use previously given values of Aa, Ab, and Ac, and compute
AF for the process. Ans. —13.35 kcal. mole.
liquid water is the stable form at 25° and 1 atm., but at temperatures above
100° water vapor will be produced in a real combustion. This change in state
of the product water may be taken into account in computing AH at an
elevated temperature either by using the appropriate values of AHy,, for
formation of the substances in the hypothetical process
Ad (H,20, 1)
AH 298
CH, (g) + 20,(g) —— CO,(g) + 2H,0 (1)
AH(O,) =2 |_ Cp(O.) aT
298
AH(CO,) = {” Cp(CO,) aT
298
AH(H,O, l) = 2 i CAMO ar
373
298
373 Tt
From the heat of combustion of methane given in Table 3.1 and the following
values of Cp compute the heat evolved in combustion of methane at 300°. Cp
(cal. deg.-! mole-'!): CH, (g), 8.6; O, (g), 7.0; CO, (g), 8.9; H,O (J), 18.0; H,O
(g), 8.1; AA37; (vap. H,O) = 9.71 kcal. mole-. Ans. —191 kcal. mole-.
or evolve heat, is instead carried out in such a fashion that heat is not ex-
changed with the surroundings, the final temperature will differ from the initial
temperature. If the heat capacities of the products and reactants are known,
the principle of conservation of energy may be applied to obtain a relation
between the observed temperature change and the isothermal heat of reaction.
If the generalized process at constant pressure
aA + bB+...=mM+mN-+...
I |
II
4H ,(T;)
aA + b6B+...—~mM+maN-+...
II
The change in enthalpy for the process is independent of path and is zero, since
de — 0 — ial
When the heat capacities of the products are known, a relation between
the change in temperature and AH, may be obtained.
The final temperature is above 2000°, well beyond the range of applicability
of the empirical heat capacity equations given in Table 2.1, and an accurate
solution is not practicable.
An approximate flame temperature may be obtained by estimating an
average value for the heat capacity of the products. We will take C,(H,O,g) =
12 cal. deg.-! and C,(Nz, g) = 8 cal. deg.-'. Substitution in equation (3.21) yields
= )/,000 === (12 228) AT
AT 20652
or 1] = 2:4-.10°-K
With a given fuel the adiabatic flame temperature is strongly influenced
by the composition of the mixture. When air is used the maximum flame tem-
perature is obtained when the proportions are such as to provide the exact
stoichiometric requirement of oxygen. If the proportion of air is increased
above this value, then additional nitrogen and unburned oxygen must be in-
cluded in the products and, on the other hand, if an excess of hydrogen is pre-
sent, this must be included in assessing the heat capacity of the products. The
highest flame temperatures will be obtained when pure oxygen is mixed with
the fuel gas in stoichiometric proportion, when the heat capacity of the
products has its minimum value.
EXERCISE 3.22
Take Cp(O,) = 8 cal. deg.-! mole™! and estimate the flame temperature when
hydrogen is burned with air in the volume ratio | : 3, i.e., a 20% excess of air.
VAT Sle— "2sas amlOsa
SUMMARY, CHAPTER 3
1. Heats of reaction
qv = AE qe = SH
= AE -—An,,skl
2. Hess’s law
aA+b6B+...=jJ+kK+...=mM+4+mN-+...
AH(A — M) = AA(A — J) + AH(J — M)
AH(A — M) = mH(M) + nH(N) — aH(A) — bH(B)
independent of path.
3. Heats of formation
aA + bB+...=mM+mN-...
AH? = mAH+(M) + nAH5(N) — aAH?(A) — bAH?7(B)
Problems, Chapter 3 Thermochemistry 81
Standard state: pressure of 1 atm. and, for elements, physical form most
stable at | atm. and given temperature. Heat content of an element in its
standard state is taken to be zero
4. Heats of solution
solute + solvent = solution
AH = integral heat of solution
solute + co solvent = oo dilute solution
AH = heat of solution at infinite dilution
solution + co solvent = co dilute solution
AH = heat of dilution
5. Ionic solutions
Heat of neutralization
H*(aq.) + OH-(aq.) = H.O; Ago, = — 13,360 cal. mole-!
By convention HH) = 0
H(OH-) = A(H,O) — AA (neut.), etc.
6. Temperature dependence of AH
ie Noe J” AC» aT
T
+ 3°(TiTY
or: AH, = AH, + daT + S72 4 SE7°
approx. form: AH, = AH, + AC,(7, — T;)
7. Adiabatic processes
1S = J"Cp (prod.) aT
Te
T.
AA —— |) -C.(react.) dt
T,
PROBLEMS, CHAPTER 3
1. When 1.000 g. of a substance is burned completely in oxygen in an adiabatic
calorimeter, the temperature of the system rises from 23.67° to 26.90°. The same
change in temperature of the calorimeter system is then reproduced by passing
an electric current of 1.000 amp. through a heating coil in the calorimeter for
15 min. at a steady potential difference of 19.80 v. Evaluate the heat of combus-
tion of the substance in cal. gram~!. Ans. 4.26 kcal. g.-}
82 Thermochemistry Chap. 3
2. Use the heats of combustion in Table 3.1 to find AAy9, and A£Foo, for the
following constant temperature, constant pressure processes:
(a) 2 CH, (g) = C,H, (g) + 3 H; (g) Ans. (a)89.97, 88.79
(b) 3 C,H, (g) = CeH, (1) (b) —150.88, —149.11
(c) C,H, d) = 6 C (graph.) + 3 H, (g) (c)—11.72, —13.49 kcal.
3. Use the heats of formation in Table 3.2 to find the heats of combustion
per gram for the gaseous normal alkanes CH, through C,H,,. Which is the best
fuel in terms of weight?
Ans. —13.3, —12.4, —12.2, —11.9, —11.7, —11.7 kcal. g.}
4. If the heat content of each elementary substance in its standard state at
298°K were arbitrarily chosen to be 10,000 cal. (g. atom)~!, what would be the
value of the molar heat content of CH,OH(l), CO,(g) and H,O(]) at 298°K?
Show that the use of these values yields the correct heat of combustion for
CH,OH (I). Ans. —22.02, —74.05, —53.32 kcal. mole@!.
5. Ina series of experiments by C. A. Kraus and J. A. Ridderhof [J. Am. Chem.
Soc., 56, 79 (1934)] the heat effects of reactions in liquid ammonia at —33° were
measured by observing the quantity of liquid ammonia vaporized by the process
of interest. The heat of vaporization of ammonia at —33° is 327 cal. g.-' When
0.835 g. of NH,Br was dissolved in 20 g. of liquid ammonia 0.221 g. of ammonia
was vaporized.
(a) Find the molar heat of solution of NH,Br in NH, (1) at this concentration.
When 0.948 g. of NH,Br was dissolved in 20 g. of liquid ammonia containing
an equimolar amount of KNH,, then 0.845 g. of ammonia was vaporized.
(b) Find AH,,, for the reaction
NHy (NH;) + NHz (NH;) = 2 NH; (1)
Ans. (a) —8.6 kcal. mole™!,
(b) —20.0 kcal. mole™!.
6. Find the heat of formation of Mn,O, from the following data of C. H.
Shomate [/. Am. Chem. Soc., 65, 785 (1943)]:
(a) Manganese metal was dissolved in dilute aqueous sulfuric acid, potassium
iodide solution:
Mn -- 2H* = Mn" -— Ha, AH = —53,900 cal. (g. atom)~!.
(b) Solid iodine was dissolved in the solution resulting from expt. (a):
hy a VE = Ie. AH = 1100'cal. mole
(c) Mn,Q, is dissolved in a solution such that the final composition was identical
with that resulting from expts. (a) and (b):
Mn,O, + 8Ht + 3I- = 3Mn*+ + 1; +4H,O
AH = —78,200 cal. mole
(d) The heat of solution of gaseous HI in the solution of interest was determined:
IBN) == 18h? SE JI=, AH = —18,700 cal. mole-!
Ans. AH; (Mn;0,) = —330,600 cal. mole!
7. From the appropriate heats of formation evaluate AH,,, for the following
processes:
(a) Agt(aq.) + Br-(aq.) = AgBr (s) Ans. (a) —20.19 kcal. mole!
(b) SO, (g) + H,O () = 2H*(aq.) + SO; “(aq.) (b) —54.13 kcal. mole-
(c) 2NH; (g) + 30, (g) = 2 NO, (g) + 3 H,O (g) (c)—135.14 kcal. mole-'.
Problems, Chapter 3 Thermochemistry 83
9. Given the bond dissociation energy of CH,—H as 103 kcal. mole-', find the
standard heat of formation of the methyl radical. Ans. 33.6 kcal. mole-.
10. From the data in Fig. 3.3 compute the heat effect when 67 ml. of water
(p = 1.00 g. ml.-') is added to SO ml. of 10 mole per cent ethanol in water
(p = 0.97 g. ml.-") at 25°. Ans. —150 cal.
11. From the heats of formation of liquid and gaseous water at 25° and other
data in the text find the heat of vaporization of water at 75°. Assume heat ca-
pacities independent of temperature. Ans. 10.02 kcal. mole~!.
15. From the emf of the Daniell galvanic cell at 25° (1.1 volt) and the heats of
formation for Zn**+ (aq.) and Cut*(aq.)
(a) find AH and AE for Zn (s) + Cutt(ag.) = Zn**(aq.) + Cu (s).
(b) Find qrey for the cell reaction.
16. Within the context of Chapters 1 to 3, how does the internal combustion
engine convert the energy of a chemical reaction into work?
Given the measured heat effect g and given the temperature-independent mean
molar heat capacities C,, Cp, etc., all at constant volume, show how to evaluate
the isothermal heat of reaction at T and constant volume. Use diagram and
equation to answer.
18. Given S(rhombic) = S(monoclinic), AH; = 96 cal. and Cp(rh.) = 3.52
+ 6.3 x 10-°7, Cp(mon.) = 3.64 + 6.8 x 10-°T
(a) Find AA, for the transition.
(b) Let S(mon.) at 20° convert adiabatically to S(rh.). What will be the final
temperature? Explain clearly the principle involved.
19. From the appropriate heats of formation, find the heat of combustion of
normal hexane (g) at 25°.
20. Find AA, for the thermite reaction:
2Al + Fe,O, > 2Fe + AI,O,
21. The value of the bond dissociation energy of water is Dy_oy = 119 kcal.
mole~. Find the heat of formation of the OH radical.
22. Da, is 28.5 kcal.; Dy-on is 119 kcal. Assuming that the heats of solution
of Cl and OH in water are the same, find AH5o, for
OH aq.) + Cliag) > Oa) + Cli
23. Bond dissociation can be induced by electron impact.
(a) What is the minimum electron energy in e.v. which just suffices to
decompose hydrogen into atoms?
(b) Find the heats of formation of H and of Ht. The ionization potential of H
is 13.595 e.v.
24. From information in the text find AH, 3, for the ion-molecule reaction
CH3(g) + CH,(g) > C.Hs(g) + H,(g).
25. For the reaction NH,(g) + Ht(g) > NHi(g), AAo9, = —220 kcal. Find
AG, for C,H (g) + NH;(g) ~ NHi(g) + C.H,(g). The ionization potential of
H is 13.595 e.v.
26. A water heater is required to heat 2001. of water per hour from 20° to 60°.
(a) Using methane as fuel and 14 times the stoichiometric requirement of air,
how many liters per hour of methane at 20° and 1 atm. will be required? Assume
perfect heat exchange and ignore heat losses. Use values of Cp from Table 2.1.
(b) How many kilowatts of electric energy would be required?
27. From the bond dissociation energy Dcoy,-1 = 103 kcal. mole-! and heats
of formation find the C—C bond dissociation energy in ethane.
28. Given the bond dissociation energies
Ini = ABI. AH = 103 kcal. mole!
H,O = H + OH, AH = 119 kcal. mole“
(0), == 0), Ad = 118 kcal. mole!
find (a) the standard heat of formation of the OH radical and (b) the bond dis-
sociation energy of the OH radical.
29. The following standard heats of formation are available for urea, CH,ON,:
State AF? 598
crystal —79.634 kcal. mole-!
Problems, Chapter 3 Thermochemistry 85
CH, (g) + 20, (g) + 8N» (g) = CO, (g) + 2 H,0 (g) + 8 N; (g)
Since the reaction is fast it may be considered adiabatic. Take Cp(H,O, g) = 12
calideg.= Gp (Ns, 2) —Sical. deg, GC, (CO, g) = 13 calidegs+
34. Estimate the adiabatic flame temperature when methane is burned at con-
stant pressure with twice the stoichiometric requirement of air. Use values of
Cp given in problem 33 and Cp (Oy, g) = 8 cal. deg. mole=!.
37. The standard heat of formation of tin telluride, SnTe, is —14.65 kcal.
mole-!, and its heat of sublimation is 53.1 kcal. mole~!. The heats of sublimation
of Sn(s) and Te(s) are 72.0 and 46.5 kcal. mole~! respectively, all at 298°K. Find
the bond dissociation energy, Dsnre [C. Hirayama, Y. Ichikawa, and A. M.
DeRoo, J. Phys. Chem., 67, 1039 (1963)].
SECOND LAW
OF THERMODYNAMICS
observed to take place in finite time. Among these is the reaction of hydrogen
and oxygen at room temperature, the classic example of a chemical reaction
which is “spontaneous” but in the absence of a catalyst proceeds at an infinitesi-
mally small rate. In order to convey the proper breadth of meaning we shall
use the term permitted change to characterize spontaneous gas expansion and
chemical reactions as well as the “spontaneous” phenomena which may have
vanishingly small rates. In a thermodynamic context permitted must be under-
stood as not forbidden, although the process may not occur to a measurable
extent.
There are other processes which cannot occur and are, therefore, called
forbidden. For every change in state which can occur spontaneously, the corres-
ponding reverse change cannot occur. If a system changes from state A to state
B without external aid, then it never changes from state B to state A without
external aid. This statement holds both for the isolated system (¢ = w = 0)
and for the isothermal system at constant volume (w = 0, q # 0).
In reality, processes either do or do not occur. In ideality, the intermediate
reversible process can be imagined. An example of reversibility, involving
isothermal gas expansion which produces maximum work, was discussed in
Chapter 2.
Other possible paths of irreversible expansion can be readily imagined, but in all
such cases it is apparent that the work done by the gas would be less than the
maximum work which could be obtained by expansion between the fixed initial
and final volumes, since in an irreversible or permitted expansion the external
pressure must be appreciably less than the gas pressure. In a forbidden expan-
sion against pressure greater than gas pressure, work done by the gas would
have to be greater than the maximum work. Such an expansion is never
observed.
We recognize a quantitative distinction between the three classes of events:
Permitted process w < Wynax.
Reversible process w = Wynax. (4.1)
Forbidden process w > Wax.
This distinction applies also to isothermal gas compression, if proper regard to
sign is observed. In compression, work done by the gas is negative. In an
irreversible compression the pressure on the gas is at all times greater than the
gas pressure and, therefore, the work done on the gas is greater than the mini-
mum required, which is to say that work done by the gas is less (more negative)
than W,nax.-
It should be recalled from the discussion in Chapter 2 that the variable
amount of work done by the system in an isothermal gas expansion does not
contradict the principle of energy conservation. For given initial and final
states AE has a unique value (zero for isothermal expansion of an ideal gas) and
q, the heat absorbed by the system, varies according to the manner of expansion
in such a fashion that AE = g — wis independent of path.
Let the system change from state A to state B via process (1) and return to the
initial state via process (2). In the first half of the cycle work is done by the
system, and in the second half work is done on the system. The difference between
these two quantities is the net work output for the cycle. This will reach a
limiting maximum value when the processes (1) and (2) are carried out reversibly.
Sec. 4.3 Second Law of Thermodynamics 91
That is, the work done by the system will be maximized at w/,,, and the work
done on the system will be minimized at w/’,, Since we have postulated that
Winax. > Wimax.» the cycle has resulted in net work being done by the system on
the surroundings.' This, of course, must be compensated by net heat absorption
(Gtev. > Yrev.), Since AE = 0 for the cycle. There is nothing to prevent indefinite
repetition of the cycle, since the system always returns to its initial state. Each
repetition converts heat from the surroundings to work on the surroundings.
We conclude that if Wax. > Wiax., a cyclic process can convert heat into work
isothermally. This result is not contrary to the first law, which requires only
that g = w in a cyclic process, but a little consideration will show that it is
contrary to experience. If permitted, it should be possible to supply the work
to drive an ocean liner from the thermal energy of
the ocean; it should be possible to supply work to
operate a refrigerator from the thermal energy of the
room and thus to create a temperature gradient Generalized
without depending on an external source of energy; sd
it should be possible to use work withdrawn as heat
from a thermal reservoir at one temperature to be
dissipated as heat (perhaps by friction) in a body Generalized displacement
at a higher temperature and thus to bring about the Fig. 4.3. Indicator diagram
flow of heat from a lower to a higher temperature for an isothermal change
without supplying work. Experience denies all of in state.
these possibilities, and these denials are forms of the
second law of thermodynamics as applied to isothermal processes, another
form of which is that an isothermal process whose only result is the conversion
of heat into work is impossible. Also equivalent to this is the answer to the
question proposed above: The maximum work obtainable from an isothermal
process is a function only of the initial and final states and not of the path. The
unlimited isothermal conversion of heat to work is termed perpetual motion of
the second kind. The second law forbids such a conversion.
The equivalent conclusions reached above represent the most important
step in our discussion of permitted processes, for it was previously shown that
for a specific path (such as simple expansion) from initial to final state, the
change is permitted if w << Wyyax.. Now it has been shown that if we can compute
Wax. for any one isothermal path from initial to final state, we know its value for
all possible paths. That is, Wax, is a function of state for the isothermal process.
For example, isothermal expansion of an ideal gas from V, to V, can yield
maximum work nRT In V,/V,, and therefore any process which achieves the
same result (and no other result) has the same value of w,,.x.. In other words,
for any equivalent process, that is, any other process which brings about the
same change in state, the work performed by that process may be less than the
maximum work for that change in state but cannot exceed it (Ww < Wyax.). We
‘For the purpose of this discussion we could as well have postulated that W’max. < W”max.
and reversed the direction of the cycle.
92 Second Law of Thermodynamics Chap. 4
will illustrate later how Wynax, may be calculated indirectly and then used to
predict the permissibility of chemical reaction.
4.4 Recapitulation
The ability to produce work, measured by w,,,x, for the same change in state
under reversible conditions, is not necessarily zero. The classification of iso-
thermal processes at constant volume becomes:
Permitted process 0 < Wyax.
Reversible process 0 = Wynax. (4.2)
Forbidden process 0 > Wynax.
Since W,,ax, is determined by the initial and final states, within our experience,
there must be a correlative thermodynamic property of the system, the Helmholtz
free energy, or the work content. Designating this property by A, we postulate that
A, — A, = AA = —Waax. (isothermal) (4.3)
oF dA, = SL
for the specified change. { PdV is usually negligible for solids and liquids. For
the ideal gas
SARE
Je = i
and
94 Second Law of Thermodynamics Chap. 4
Ne pe —{" — dV
That is, a permitted process is inefficient, and the work done (always PAV) is
less than the maximum work for the specified change in state. When a permitted
process occurs under these conditions, the ability to do work decreases.
Rearranging the above relations and using the previously defined property
AA, we may state the classification thus:
Permitted process AAp» + PAV <0
Reversible process AAp p + PAV = 0 (4.7)
Forbidden process AA; p + PAV > 0
ee A A leg a a
AF = AA + A(PV) (4.9)
As in the case of E, H, and A, the absolute value of F cannot be known, but
changes in F, that is, values of AF for specified changes in state, can in principle
be calculated or measured.
hee | VdP
P
EXERCISE 4.2
Evaluate AF yo, for
H,O (1, 1 atm.) = H,O (1, 100 atm.)
Ans. 43 cal.
96 Second Law of Thermodynamics Chap. 4
AF, = | as
P»
=
= Loge
ee a Ue
V, 4
(4.11)
For any change in which the initial and final pressures are identical,
equation (4.9) simplifies to
AF,» = AA -— PAV (4.12)
and the classification given in equation (4.7) simplifies to
Permitted process AF; p <0
Reversible process AF,» = 0 (4.13)
Forbidden process AF,» > 0
Since, for a process at constant temperature and pressure
stitute a contradiction, since w,,,.. is a property of the system whereas the net
work includes work contributed by the surroundings.
Let us evaluate AF,» = AA + APV for
HO (1), atm: 373°K) = H.O\(g, atm. 373°K)
The value of AA,;;; = —W,,,, is obtained from the fact that the saturation
vapor pressure of water at 373°K is | atm., and, therefore,
Neglecting V, in comparison with V,, and assuming the gas to be ideal, we have
Winex, = PAV = 1 x RT/molé vaporized
For the second term, APV, we see that this also is PAV, since P is constant
and has the same value as w,,,,. Therefore,
AF373 = DA373 et Ay Wea ate JPIN| —— (0)
Since AF = 0 for the process described, it follows that for any path leading from
the specified initial state to the specified final state AF,,, = 0. That is to say,
when water vapor at 100° expands against the atmosphere at 760 mm., it
cannot perform any additional work.
To compute AF, for any process it is necessary to know AA> or Wax. In
the example just given, the process was in fact a reversible one and Wyyax. Was
easily and directly evaluated. To compute AF, for an irreversible process it is
necessary and sufficient to accomplish the same change in state by any reversible
process for which w,,,x, is calculable.
An example of an irreversible process for which we can calculate AF is
(i) H,O (1, 373°K, 700 mm.) = H,O (g, 373°K, 700 mm.)
To accomplish this change in reversible fashion, the following sequence of
processes may be used, each of which is to be conducted reversibly:
(ii) Isothermal compression of liquid water
H,O(1, 373°K, 700 mm.) = H,O (1, 373°K, 760 mm.)
(iii) Isothermal vaporization of water
H,0O (1, 373°K, 760 mm.) = H,O (g, 373°K, 760 mm.)
(iv) Isothermal expansion of water vapor
H,O (g, 373°K, 760 mm.) = H,O (g, 373°K, 700 mm.)
Since AF is a function of state, it follows that
AF (i) = AF (ii) + AF (iii) + AF (iv)
AF may be evaluated for each of the reversible steps as follows:
where P is the pressure exerted by the surroundings on the system and P.,. is
the equilibrium vapor pressure of water at the specified temperature. In the
example described above, AF (vaporization H,O at 700 mm., 373°K) = —2.52
l. atm. mole™! and the sign indicates that the process is permitted.
EXERCISE 4.4
Evaluate AF for the hypothetical process
H,0 (1, 100°, 800 mm.) = H,O (g, 100°, 800 mm.)
and show that the result is consistent with experience.
Ans. AF37, = -+-1.57 1. atm. mole=!.
This process is forbidden.
AF =n X 0.082 x 298 In lw
= 20 = 2611. atm. mole
— aa
A chemical change with specified initial and final pressures may be forbidden
at one temperature but permitted at another. Thus, the decomposition of
Mg(OH), is forbidden at 25° and 1 atm. The equilibrium vapor pressure of
water in this system increases with increasing temperature and at temperatures
above 260° exceeds | atm. Therefore, at 1 atm. and a temperature above 260°,
AF is negative and the decomposition is permitted. The value of the isothermal
free energy change is in general a function of temperature, and changing temper-
ature is an important way to change forbidden processes to permitted processes.
Another means of accomplishing a given change in state when it is normally
forbidden is to couple the process for which AF(a) > 0 with another for which
AF(b) < 0 such that AF(a) + AF(b) < 0. For example, consider the reaction
CaO (s, 20 mm., 25°) + H,O (g, 20 mm., 25°) = Ca(OH),(s, 20 mm., 25°)
In this system, the equilibrium vapor pressure of water is 4.7 x 10°-° mm. and
20
A Fos, 20mm. — — i xX 0.082 x 298 In ( )= — 5421. atm. mole“
4.7 x 107°
Now, if this process is coupled with the dehydration of Mg(OH), in a single
system at 25° and 20 mm., the net result is
The pressure on the system, P, disappears from the expression, and the value
of AF depends only on the relative equilibrium water vapor pressures of the
two hydrates.
This method of computing AF depends on the principle that F is an extensive
property of state. For any reaction
aA + b6B+...=mM+aN-+...
it follows that
AF = mF(M) + nF(N) +... — aF(A) — bF(B) —...
When two or more processes are combined, as
aA + bB+...=jJ+kK4+...=mM+mN-+...
the net free energy change is the algebraic sum of the free energy changes for
the individual steps.
AF(A — M) = AF(A — J) + AF(J — M)
This operation is completely analogous to Hess’s law of thermochemistry,
equation (3.8).
EXERCISE 4.6
The equilibrium vapor pressure of water in a system containing CdBr,-4H,O (s)
and CdBr, (s) at 25° is 30 mm. Evaluate AF), for
CdBr,-4 H,O (s) + 4 CaO = CdBr, (s) + 4 Ca(OH), (s)
Ans. AF 98 — —53.6 kcal.
Since
Wimax. = W (elect.) + w (exp.)
AF, ja = VV. (elect.)
the reversible emf is 0.04 v. (iron anode), The corresponding free energy
change per unit reaction is given by
AF = —2 x 96,487 x 0.04 = —7.7 x 10° joules
This is to say that if we mix the substances at the specified concentrations with
no provision for electric work, the reaction written above is permitted. If the
rate is not too small for observation, the concentration of Cd** will decrease and
the concentration of Fe** will increase until eventually a state is reached where
no further net change occurs. This condition is reached when the concentration
of Fe** is about 20 times the concentration of Cd**. This is the state of chemical
equilibrium in which there is no further tendency to react.
When a galvanic cell is constructed with the reactants and products at this
equilibrium) concentration: ratio of 20/1) (eie., He =) MW) Cd" == 01005017):
the reversible emf will be zero. This is an experimental fact, but it is also
clearly predictable from the second law of thermodynamics and its consequence
which we have developed, namely, that if AF is zero for a constant pressure
isothermal process, then the process is neither permitted nor forbidden. This
means that there is no tendency for either the forward or reverse reaction to
occur spontaneously, as is in fact the case.
If a galvanic cell with a concentration ratio Fe**/Cd** > 20 is constructed,
the reversible emf of the cell will again have a finite value, but of a polarity
opposite to that of the original cell, i.e., the Cd electrode will be the anode in
spontaneous discharge, and cadmium will be oxidized. Since cadmium is
being oxidized, the reaction
Fe (s) + Cd** (aq.) = Fe** (aq.) + Cd (s)
with (Fe**t)/(Cd**) > 20 is proceeding from right to left. The free energy
change for the forward reaction must be positive, since it is the reverse of the
permitted process. To be consistent with the sign convention adopted, we must
assign a negative emf to the cell for the reaction as written. A resulting cor-
ollary is that a negative emf denotes a forbidden process. The reverse reaction
Cd (s) + Fet* (aq.) = Cd** (aq.) + Fe (s)
(Fe**) = 20
(Ca**)
will now have a negative free energy change, a positive emf, and is a permitted
process, with AF = —n¥Fé <0. The relation between AF and electric work
is a very useful and important one which will be used repeatedly.
We can also use galvanic cells to illustrate further the principle that a process,
which by itself has a positive value of AF and is, therefore, forbidden, may be
made to proceed through the expedient of coupling it with another process
having a negative AF whose magnitude is greater than the first. This principle
of the additivity of free energy changes applies even more evidently to the
additivity of emf’s. That is, since for processes (a), (b),...
AF (net) = AF (a) + AF(b) +...
Sec. 4.1] Second Law of Thermodynamics 103
6 (nett) =Gé(a)+€é(b)+...
This result is consistent with the common observation that when cells are
connected in series their emf’s combine algebraically. This may be illustrated
by combining the two cells previously discussed,
as indicated in Fig. 4.4. There are two possible
ways of connecting the cells. When we connect
terminals 2 and 3 and then measure the reversible
emf across | and 4, experience leads us to expect a
value of 1.10 + 0.04 = 1.14 v., and this is in fact
the observation. On the other hand, when we
reverse the polarity of one cell, connecting ter-
minals 2 and 4 and then measure the reversible emf
across 1 and 3, we find 1.10 — 0.04 = 1.06 Vv. In
the latter case the net process which occurs on
spontaneous discharge (no electric work) is
Zn (s) + Cut* (1 M) + Cd (s) + Fe** (1 M)
FIG. 4.4. Coupled galvanic cells.
= Zn**(1 M) + Cu(s) + Cd*t(1 M) + Fe (s)
and the AF for this process is given by
AF = —2(1.10 — 0.04)96,487
= —2.05 x 10° joules
Note that in this instance the permitted reaction involves oxidation of cadmium
and reduction of iron, which was forbidden when the iron-cadmium cell was
considered alone.
EXERCISE 4.7
Mee aR in(7)
V,
Let the initial and final volumes V, and V, be fixed and take the temperature
derivative of AA,.
dMAr _ __ Ve
= —nRIln (y)
dT
Since
AA = —Wrax.
and since it has been shown that in the isothermal expansion of an ideal gas
E20)
then
= (4.15)
The temperature derivative of the isothermal quantity AA, is to be understood
in the sense
dT ee? earners. 5
dA, ee lim. (94: ia 7)
=
AF =nRTin (5)
P
Remembering that T is constant, we express this relation by
Since
[ aAFIT) = — {AHS
AF 2/T2 T2 ak
AF,/T, Ye
AE Al (+ on F)
T, Co Au aT
which rearranges to
AF = AH NT (Ba ee (4.19)
1
GAT
(S$), Veen,
Ra ae — (— 30,220) 49.2 cal. mole~!Z deg.~!d
"F (55),
a€\
= n¥&€
ae+ AH 4.20
= kei aé
a7 gai a, ah G2)
provided AH is constant over the temperature interval involved. We see that
here also the value of the reversible emf at various temperatures may be
calculated if the value at some one temperature is known, together with the
AH for the process. Examples of this application will be given in a later chapter
on electromotive force.
SUMMARY, CHAPTER 4
(Aap) = 9
PROBLEMS, CHAPTER 4
1. Evaluate w and AA for the expansion of one mole of ideal gas at 300°K from
10 atm. to | atm.
(a) By free expansion into an evacuated vessel.
(b) Against a constant Psu, = 1 atm.
(c) Reversibly. Ans. (a) AA = —1371 cal. (b) w = 537 cal.
2. For the processes in Problem 1, find AE and AH.
3. A lead storage battery generates an emf of 6v.; its internal resistance is
0.05 ohm.
(a) What fraction of the energy it expends is wasted by ohmic internal resistance
at a current of 1 ma.?
(b) What fraction at a current of 1 amp.? Ans. (a)8.35 x 107°, (b)8.35 = 107%.
4. A galvanic cell is constructed so that the cell reaction is
18l, (U2) = Il asin) = 16 C2 = Oull anton)
(a) What is the maximum electric work obtainable from the cell at 25°?
Ans. (a) 5680 joules.
(b) What is the emf of the cell? (b) 0.0296 v.
(c) What is the temperature coefficient of the emf? (CSIP S10m vadegas:
5. In the case of a cell such as that described in the preceding problem, what
ratio of pressures P,/P, would be required to produce an emf of 0.1 v.?
Ans. 2.41 x 10°.
6. Using the data in the text (p. 37), evaluate AA for
(a) H,O (s, 0°, 1 atm.) = H,O (1, 0°, 1 atm.). Ans. (a) 1.42 x 10-7? 1. atm.
(b) H,O (I, 100°, 1 atm.) = H,O (g, 100°, 1 atm.). (b) —30.6 1. atm.
(c) H,O (g, 100°, 1 atm.) = H,O (g, 100°, 0.1 atm.). (c) —70.4 1. atm.
7. For gases which obey the equation of state PV = RT + DP, find expressions
for AA and AF for isothermal expansion
(a) In terms of V, and Vj.
(b) In terms of P, and P,. Ans. AA = RT In (V, — 5)/(V, — 6).
8. At 30° the equilibrium vapor pressure of water is 31.82 mm. Hg. Compute
AF for the process
H,0 (1, 1 atm., 30°) = H,O (g, 1 atm., 30°) Ans. 1.90 kcal.
Problems, Chapter 4 Second Law of Thermodynamics 109
in state occurs isothermally with no provision for electric work, 18 kcal. of heat
is lost to the surroundings. Determine @yey..
19. For the cell described in the preceding problem,
Ca— OT Vaated on.
(a) Find A Fos.
(b) Find & at 50°.
20. (a) Devise and describe an isothermal reversible procedure for transferring
one mole of volatile component A from a binary solution obeying Raoult’s law
from mole fraction X’, to X”’,, when the second component is nonvolatile. The
volumes of solution are taken so large that composition is sensibly constant.
(b) Express AFy as a function of X,.
21. According to the principle of Le Chatelier, increasing the pressure on a
system in a state of equilibrium will shift the position of equilibrium in that
direction which will tend to offset the increased pressure. Explain this principle
in terms of AF.
22. For the reaction 2C (graphite) + 3H, (g) + S (rhombic) = C,H;SH(), for
all substances in their standard states, AH 3 = —17.61 kcal., AF 93 = —1.36
kcal.
(a) For the preceding reaction, find Py, at which AF), = 0.
(b) Taking AH® constant, find the temperature at which AF°® = 0. Integrate
equation (4.18) for the general case that
AH7 = AH, + AaT + (Ab/2)T? + (Ac/3)TF?
23. At —5S° the vapor pressure of ice is 3.01 mm. Find AF for
H,O (g, 1 mm., —5°) = H,O (s, 100 mm., —5°)
24. Two galvanic cells are constructed as follows:
(a) H, (1 atm.) + 2 AgCl (s) = 2 HCI (0.0050 m) + 2 Ag (s)
(b) H, (1 atm.) + 2 AgCl (s) = 2 HCI (0.050 m) + 2 Ag (s)
and the emf’s at 25° are 0.49844 and 0.38589 v., respectively. Couple the two
cells in opposition so that cell a runs forward and cell b in reverse.
(a) What is the net emf of the combination?
(b) Combine the opposed cell reactions and write the net process.
(c) Evaluate AF for the net process.
25. At 25° the partial pressure of H,S over its aqueous solutions is 1.00 atm.
at 0.10 molal and 2.00 atm. at 0.20 molal. Describe a procedure to effect the
process
HES O:1 m) = HS 02m)
in a reversible fashion and evaluate AFyo, for the change in state. The amounts
of solution are so large that transfer of one mole of H,S does not affect the
concentration.
26. The heat effects which attend a given change in state when carried out
reversibly and irreversibly are not, in general, the same. This can be demonstrated
by using equation (4.16) to obtain grey, and (4.17) to obtain AH. The emf of the
cell whose reaction is
H, + 2 AgCl = 2 HCI (0.100 m) + 2 Ag
is 0.35240 v. at 25° and d&/dT = —0.0019 v. deg.-! Find AH and @yrey..
Problems, Chapter 4 Second Law of Thermodynamics 111
In the preceding chapter it was shown that the value of the free
energy change (AF) in a constant temperature, constant pressure process in
an isolated system can be used to classify the process as permitted, reversible,
or forbidden. If the specified temperature and pressures of substances present
are such that the free energy change in the process of interest is zero, there is
no tendency for the process to occur in either direction, and the system is in an
equilibrium state.
Consider a system consisting of gaseous hydrogen, nitrogen, and ammonia.
In order to characterize the state of a mixture, it is necessary to specify not only
the temperature and the total pressure of the system, but also the composition.
If the total pressure and temperature are fixed, and if the system exchanges no
work with the surroundings other than expansion work, then the sign of the
free energy change for the process is a criterion of permissibility.
If one chooses an initial composition of the system quite at random, the
composition almost certainly changes with time. Regardless of the composition
of the starting mixture, if reaction is possible, the composition of the system
eventually reaches a value which no longer changes with time. That is, the
system comes to equilibrium.
There is no unique composition of the system at equilibrium. In fact, an
infinite number of different equilibrium compositions are possible at constant
total pressure and constant temperature. To illustrate, let the initial composition
of the system be a moles of nitrogen, b moles of hydrogen, and c moles of
112
Sec. 5.2 Chemical Equilibrium in Gaseous Systems 113
ammonia. Let the number of moles of ammonia formed in the course of equi-
libration be x.
Initial moles a b c
N:. = 73Hy = .2NH.
Equil. moles a—4+x b—8x c+x
The value of x is found from experiment, but the values of a, b, and c can be
selected arbitrarily by the experimenter. They have no relation to the stoichio-
metric coefficients. Thus, with an infinite choice of initial compositions, it is
possible to have an infinite variety of equilibrium states. Application of the
second law of thermodynamics reveals a useful relationship describing these
equilibrium states which will now be examined.
EXERCISE 5.1
1.165 moles of nitrogen and 0.465 mole of hydrogen are equilibrated in a 1 1.
vessel at 450°. At equilibrium 0.0845 mole of ammonia is present. Find the
amounts of nitrogen and hydrogen and the partial pressures of all components
at equilibrium. Ans. hy, = 1.123; ny, = 0.338; Px, = 66.5 atm;
Py, = 20 atm.; Pru, = 5 atm.
iPiB
Previously, this expression was integrated between the limits P, and P, to obtain
114 Chemical Equilibrium in Gaseous Systems Chap. 5
the change in free energy as a function of pressure. For the present purpose the
indefinite integral will be more suitable. For one mole of ideal gas the constant
of integration is the molar free energy of the gas at unit pressure (when P = 1,
InP = 0), a value denoted by F7.
Fy= RE InP ee (5.2)
Equation (5.2), therefore, does not yield values of F;, since Fp is not known,
but rather gives the relationship between the free energy at an arbitrary pressure
P and the free energy at unit pressure. The same relation could have been
obtained by equation (4.11), thus, for one mole
AF, = F,(state 2) F,, (state 1) = Ryan (PP)
Let P, = 1; that is, let state 1 be the standard state
F, (State 2). = REIn P; -- Fe
Equation (5.2) applies to each of the components of a system and may be
used to expand the expression for AF;.
AF, = mF3(M) + nF3.(N) + ... — aF?(A) — bF}(B) — ... + mRT In Py
+ nRT In Py +... — aRT ln P, — bRT InP; —... (3:3)
Representing the coefficient of the ith component by v; and summing over all
terms gives
AF, == DS) Dyer + » URE In lees (5.4)
The first term on the right-hand side of equation (5.4) is a collection of constants
of integration from equation (5.2) for each substance.
Dy mb (M) nk (N) 4-2. ais (A) oF, (By... 29)
Given AF%,, = 7.16 kcal. for the synthesis of 1 gram mole of ammonia from
the elements, find the corresponding standard free energy of formation of 1
pound mole of ammonia. Ans. 3245 kcal.
AF?,, = 7160 cal mole', and AF,,; for formation of 2 moles of ammonia when
Pyy, = 10 atm., Py, = 1 atm. and Py, = 2 atm. can be computed as follows.
10?
Op = T- 3 (atm. ~*)
1The partial pressures in the reaction function are dimensionless and, therefore, Q>p is
dimensionless. Nevertheless, its numerical value depends on the choice of standard
states, and it is helpful to specify this choice by assigning appropriate pressure units to
Qp, which will be shown in parentheses.
Sec. 5.3 Chemical Equilibrium in Gaseous Systems 117
RES — 1 160\cal:
QO> for this process is formulated
Pyu,
Op = PEP? (atm. ~')
and the free energy change when Pyy, = 10 atm., Py, = 1 atm., and Py, = 2
atm. is given by
P%,Py,
PN Bre aS — 14,320 +- RT \n
P?
NH;
118 Chemical Equilibrium in Gaseous Systems Chap. 5
EXERCISE 5.4
If Pyx, = 1 atm. and Py, = 2 atm., what value must Py, have in order that AF 73
shall be zero? Ans. 2.66 * 10% atm.
ae Bye (5.10)
This expression is often called the mass action law.
It was shown that the standard free energy change AF’ has a fixed value
for a given process at a given temperature. It follows from equation (5.9)
that Kg also has a fixed value for a given process at a given temperature. For
this reason, Kg is called the equilibrium constant. The relation between Kg and
AF? is given by
AF; = —RT|n Kg (Sa)
To illustrate, let us consider the process
A+B=C
taking Kg = 10(atm.-'), and let the initial partial pressures be P,, Pz with
P, = 0. The partial pressures at equilibrium become
Dad ON lS 8 F,—=P,—F-
Po Z
= |) etna.”
(P, — Pj Ps — Po) a
Since an infinite variety of initial pressures P,, Ps, and P, are permitted, no
two of which lead to the same equilibrium state, there is an infinite number of
sets of values for P,, Pz, and Y,. For instance, if P, = 0.1 atm., Ps = 0.5 atm.,
>The previous remarks concerning units in Qp apply equally to Kg and appropriate units
will be given in parentheses.
Sec. 5.5 Chemical Equilibrium in Gaseous Systems 119
Kg _= DFP
Pru,
TABLE 5.1
VALUES OF AF73; FOR THE PROCESS
No aF 3H, — 2NH;
Comparison of case 2 with case 3 shows that decreasing the nitrogen and
hydrogen pressures below the equilibrium values also makes the reverse reaction
permitted.
Comparison of case 5 with case 3 shows that increase of the nitrogen pres-
sure above the equilibrium value makes the forward reaction permitted.
Comparison of the fifth and sixth columns of the table shows that the
critical factor in the determination of permissibility is the relation of Q, to the
fixed value of Kz.
EXERCISE 5.7
The volume of a system in the equilibrium state described in case 3 of Table 5.1
is increased by a factor of two, thus decreasing temporarily the pressures of all
components to one-half their equilibrium values. What process will be permitted
and what will be the value of AF,,,?
Ans. 2 NH, = N, + 3 He; AF 723 = —1990 cal.
Av=m+n--t+.... a—b
It is to be observed that Ky is not an “equilibrium constant” with regard to
variations in FY, except for Av = 0. Kyis simply the value of the equilibrium
function of the mole fractions.
Equation (5.13) may be modified by remembering that
1 i Seeni
Again, by analogy, K, is the value of I(n‘), and this quantity is not an “equili-
brium constant” with regard to variations in P, except for Av = 0. K, is simply
the value of the equilibrium function of the mole numbers.
EXERCISE 5.8
The dissociation process
A, =2A
occurs to the extent that X, = 0.01 at 1 atm. total pressure. Compute the
approximate value of X, at 10 atm. total pressure. (Let X4, = 1.)
Jai SMG S< WO
The equilibrium mole fractions may be inserted in equation (5.13) and the
equilibrium mole numbers in equation (5.14).
ne Py (5.15)
8 (6 8
RK — Ke —Av Av __ (iw =F mx)"(v Si nx)”
This expression relates the equilibrium constant and the various experimental
characteristics of a system. In general, the measurement of an equilibrium
constant is accomplished by mixing known amounts of reagents (a, f, |, v, etc.,
are known), equilibrating, and measuring the extent of reaction (x measured) or
the equilibrium amounts of reagents (ju + mx, etc.) at a known total pressure
F,. On the other hand, the value of Kg may be known and used to find the
equilibrium composition of a system with a chosen initial composition.
The equilibrium between N,O, and NO, is conveniently measurable in the
124 Chemical Equilibrium in Gaseous Systems Chap. 5
interval 25—45° [F. H. Verhoeck and F. Daniels, J. Am. Chem. Soc., 53, 1250
(1931)]. Some of the data appear in Table 5.2.
TABLZ 5.2
THE DISSOCIATION EQUILIBRIUM N,O4 = 2NO,
In these experiments the initial amount of product was always zero (u = 0),
and equation (5.15) simplifies to
RT
Ff,
6.42 x 10-° 2:3) mole 1. =}!
FP; - |) ) =
(7. O38 X< IO 3 irnoile II,i
Initial Equilibrium
NoO, pail | a—-x=1-x
NO, (i— 0 pp + 2x = 2x
nm=1+x
0:649.(atm = 2)"
— OMe sj atm, —F = 0374
eens
EXERCISE 5.10
At what total pressure is N,O, 50% dissociated at 45°? Ans. 0.49 atm.
TABLE 5.3
THE EQUILIBRIUM 2HI = Hy + Ip
lls I, H, HI KP Ke
Temp. (init. m/cc. (equil. m/cc. (equil.m/cc. (equil.m/cc. obs.) (corr.)
(°K) x 10°) x 10°) x 10°) <aO2) x 10? x 10?
763.8 Cc 1.173 0.1185 0.4262 1.494 2.244 2.196
763.8 d 0 0.2424 0.2424 1.641 2.182 DAD
730.8 Cc 1.228 0.1524 — 1.687 — 2.018
730.8 d 0 0.1696 0.1696 1.181 2.063 2.007 *
698.6 c 1.134 0.0738 0.4565 1.354 1.835 1.812
698.6 d 0 0.0479 0.0479 0.353 1.840 1.812
666.8 Cc 1.119 0.1295 0.3258 1.587 1.676 1.642
666.8 d 0 0.1395 0.1395 1.079 1.672 1.644
126 Chemical Equilibrium in Gaseous Systems Chap. 5
sample was analyzed for hydrogen iodide and iodine. These data, in Table
5.3, are designated by c (for combination). In these experiments a small cor-
rection was necessary for loss of hydrogen by diffusion through the walls of
the vessel. This loss will be largest at the highest temperatures, where the rate
of diffusion is greatest, and in mixtures containing the most hydrogen. The
values of the equilibrium constants have been corrected for this loss.
In this reaction Av = 0, and equation (5.15) reduces to
(ee Oa)
Ore (5.17)
For the decomposition experiments we have
Initial Equilibrium
HI a CO 2x = nat
I, [Ui 0 x = ny,
H, v=0 xX = ny,
K — ae — Ny,
He Te
4° @= 2x)
The equilibrium values ny; and n,, = my, are determined experimentally, and
for the second line in Table 5.3 we obtain
(0.2424)? Dalen On:
eS (1.641)?
For the combination experiments we have
Initial Equilibrium
HI a=0 0 — 2x = nur
I, LL b+ xX = mM,
H, v y+ % = ny,
The equilibrium values of ng, and m,, are again measured, and the value of
v, the initial amount of hydrogen, is known. Since ny; = —2x (this is a case in
which “products” are consumed), the value of ny, is
a 1
Sst ae oe al fA
That is, each mole of hydrogen iodide formed consumes 4 mole of hydrogen.
Ky = Wilt, = om)
Nxt
The first line of data in Table 5.3 gives
— 0.1185(1.173 — 1.494/2) _
Kg 14942 = 2.244 x 10°
EXERCISE 5.11
Initial Equilibrium
HI C2 a— 2x =2—2x
I, (a— 0 ptx=x
H, — v+x=2+4+x
Kg Za
= 1.812 x 10 =o x(2— +x)
@ 2x)?
101033
EXERCISE 5.12
TABLE 5.4
THE EQUILIBRIUM EN» + 3H. = NH;
Initial cormp.: 76.2% Ho, 23.5% Noe, 0.3% Ar
Temp. P, % NH3 KZ
(°C) (atm.) at equil. (atm.~—!)
350 10 7.35 0.0266
350 50 Dell —
400 10 3.85 0.0129
400 50 Syatlit 0.0130
450 10 2.04 0.00659
450 50 9.17 0.00690
500 10 1.20 0.00381
500 50 5.58 0.00388
128 Chemical Equilibrium in Gaseous Systems Chap. 5
Chem. Soc., 45, 2918 (1923)]. The small trend in the values of Kg with pressure
is an effect of gas imperfection which will be discussed in a later section.
For the process
Ke [ae (5.18)
(@ = By"8 — 3)” Ui
To evaluate Ky» for one of the mixtures described in Table 5.4 it is first
necessary to establish the relation between Xx, and the various mole numbers
at equilibrium. That is,
Xu TNH __ (ju aX)
ny Gl = B35 =,
For the cases being considered pp = 0 and
BGNHg ee ceeBX) A
a+B—x
X xu, = 0.0120 —= f a
1 0:0019
Initial Equilibrium
No 10235 a — 4x = 0.229
Hy B= OS B — 3x = 0.744
NH; p= p+ x = 0.0119
n, = 0.985
Kp ~ (0.229)'2(0.744)3?
Une x 0.985 x 10-1
EXERCISE 5.13
Compute the value of Kg in line 2 of Table 5.4. Ans. 0.0278.
=te =e aim =
— = a =,
ob —_ — <—
0 + 0) —
we jAR? 7.493
2.3 log 0.0270 = 1987 633 ~ 1987 2.3 log 623
which applies to the temperature range over which the empirical heat capacity
equations are valid.
EXERCISE 5.16
From the information given above, find Kg and AF® at 1000°K for
KFS) = RTA Bp
AF%(ii) = —RT In 2%
NO:
AF°%,.(iii)
0 (49
= —RT In DP,
Px 2
x a
Ze ¥204
A Foil)== RT In
—RT "GP,
P
EXERCISE 5.17
From study of the further dissociation of NO, formed from N.O, it has been
found that for the reaction
NO (g) + 4 O, (g)= NO, (g)
AFoo = —8.329 kcal.
Compute the value of AF%,, for the reaction
+N, (g) + 4 O, (g) = NO (g)
Ans. 20.719 kcal.
C2He He
SOC 5.9 Chemical Equilibrium in Gaseous Systems 133
TABLE 5.5
STANDARD MOLAR FREE ENERGIES OF FORMATION AT 25°*
DNS AF 598
Substance (kcal. mole-!) Substance (kcal. mole!)
H,O (g) —54.6352 methane, CH, (g) —12.140
H,0 (1) — 56.6902 ethane, C2He (g) —7.860
HCl (g) — 22.769 propane, C;Hs (g) —5.614
HBr (g) —12.72 normal butane, C,H) (g) — 3.754
HI (g) 0.31 isobutane, CH (g) —4.296
SO, (g) —71.79 normal pentane, C;H)p» (g) —1.96
SO; (g) —88.52 normal hexane, C,H, (g) 0.05
HS (g) —7.892 benzene, CoHg (g) 30.989
NO (g) 20.719 benzene, CeH,g (1) 29.756
NO, (g) 12.390 ethylene, CsH, (g) 16.282
NHs (g) —3.976 acetylene, CoH» (g) 50.000
CO (g) — 32.8079 formaldehyde, HCHO (g) —26.3
CO, (g) —94.2598 acetaldehyde, CH;CHO (g) —31.96
AgcCl (s) — 26.224 methanol, CH;0OH (1) —39.73
AgBr (s) —22.930 ethanol, C,H;OH (1) —41.77
Fe,O; (s) —177.1 formic acid, HCOOH (1) —82.7
Al,Oz (s) —376.77 acetic acid, CH;COOH (1) —93.8
NaCl (s) —91.785
Obviously the equilibrium state in this system at 25° lies very far in the
direction of complete hydrogenation, although the rate of hydrogenation at
room temperature is negligible in the absence of a catalyst. Conversely, ethane
has very little tendency to lose hydrogen at room temperature, and no catalyst
can change this situation so long as the products of the decomposition are not
removed.
Let us examine the possibility that ethylene may disproportionate to ethane
and acetylene by computing the equilibrium constant for the process
Ky = FeatFoae
PF om,
— 9,6 x 10-8 (atm)
Very large or very small values of equilibrium constants such as those
found above often permit considerable simplification of the calculation. For
example, to compute the extent of disproportionation of ethylene at one
atmosphere pressure we have
Initial Equilibrium
C,H, a=2 a — 2x =2 — 2x
C,H, [= 0 LL + x=x
C,H, y= 0 Ve ox
=8 x
EXERCISE 5.18
Compute the value of AF for the process
, CH, = Calsl. aa lels
discussed in Chapter 6, we can evaluate the free energy function from heat
capacity data. For any reaction, including formation of compounds from the
elements,
fie WN eee
A T T r (5.24)
It follows that
ee NES Fi. — H°
T= oF +A 7 (5.25)
TABLE 5.6
THE THERMODYNAMIC FUNCTIONS*
pletely realistic, since there are no ideal gases, but it has the advantage of
providing a simple working model. We have seen that many gaseous systems do,
in fact, approach rather closely to the behavior ofideal gaseous systems, and we
may, therefore, consider the preceding equations as good first approximations
for real equilibrium reaction mixtures at moderate pressures. Whenever such
approximations are not permissible, and this will tend to be the case at high
pressure and low temperature, a more nearly correct treatment must be used.
There are two evident approaches to the problem of improving the de-
scription of a gaseous system at equilibrium. The first adopts the previous
approach to equilibrium through the free energy function, and we have the
usual relation between the free energy change at arbitrary pressures AF, and
the free energy change at unit pressures AF?, the standard free energy change.
A A
we have
SR Tig) ah ee)
at
Continuing by analogy with the prior development, we find that
< ae JOURS
EME Ae ~ wa
AF, — AFR RE in Pepe si LUNN == den) ge daly
al des) (5.27)
AF pitear = RT In P,
P,
for an ideal gas. Let us once more examine the problem of the equilibrium
process, beginning with the familiar diagram but replacing all P;’s by the
analogous f;’s.
=aRTint-+...+mRTIin +...
Sa
= RT in Lh:
Sife---
= RT In (f%)
At equilibrium AF= 0
and Ari —=KT In TLC equi.
Since AF? is a fixed quantity it follows that the value of the equilibrium
function of fugacities,
K,;= I(F? equ.
138 Chemical Equilibrium in Gaseous Systems Chap. 5
is a constant in real gas systems. The data in Table 5.7 show that Kg is not
a constant in the ammonia equilibrium when the pressure is high, but that at
low pressures the value of Kg approaches a fixed limit. That is, in any real gas
system at sufficiently low pressures
IT(f equa. = U(F%) X const. (5.30)
It is common practice to complement the definition of fugacity given in
equation (5.28) by the additional stipulation that in the limit of very low pressure
the fugacity approaches the pressure (f— P as P — 0); f and P are now com-
pletely interchangeable at sufficiently low pressure, and the constant in equa-
tion (5.30) becomes unity.
The ratio of the fugacity to the pressure is called the activity coefficient y.
4 ar (5.31)
At low pressures the activity coefficient approaches unity.
It remains to be shown that f; can be evaluated for pure gases or mixtures of
gases. Let us first consider the simple case of a pure gas. From equation (5.28)
RT In relat:
4 =| (4S
> =— @\dP
)
1
fp = RT In
RT In} Py, — |: aap (5.33)
or
RT nL = RT InP, — { wdP
ti 1
RT
inf= RT InP — fa dP 0
or
oa ee=
Iny=In5 ar |, «4P (5.34)
Summary, Chapter 5 Chemical Equilibrium in Gaseous Systems 139
TABLE 5.7
EQUILIBRIUM IN THE AMMONIA SYNTHESIS
ie
(atm.)
%equil.
NH; Kg x 103 Ky x 103
SUMMARY, CHAPTER 5
Kg = Il(FP?)
SCRE
PP...
5. The reaction isotherm
AF, = RT \n (Q>/Kg)
QO,» < Kg forward process permitted
6. The equilibrium function of mole fractions
K, = Kg X P;*”, where Av =m+n+...—a—b
Ko = Kn “Pe whem Av 0) Ko = Ke ke
7. Relation of initial and equilibrium states
Initial Equilibrium
A a a — ax
B B B — bx
M lL pL + mx
N v vp + nx
PROBLEMS, CHAPTER 5
1. For each of the following reactions
(a) Evaluate AF ‘og.
(b) Evaluate A$...
(c) State whether Kg increases or decreases with increasing temperature.
(d) State the effect of increasing the volume of the system on Ky.
Problems, Chapter 5 Chemical Equilibrium in Gaseous Systems 141
3. Using data from Table 5.6, find the value of Kg at 25° for
HI (g) = ¢ H. (g) + £1 @&)
ANS Sila alOmee
4. Given that AF%,, = 16.77 kcal. for
2 1, (g) = 1g)
find the extent of dissociation of iodine vapor to atoms
(a) At 1 mm.
(b) At 10-° mm. VATISS (GQ) ie —s lee aolOnn ((b) Ca lanl Ome
5. Using data from Table 5.5, find the fractional extent of reaction for SO; (g)
= SO, (g) + 4 O, (g) at 25° when the partial pressure
of oxygen is maintained at 10-° mm.
PAT Seales nl Ones
6. Find the equilibrium composition for n-C,H,) =
i-C,H,, atwZow Ans. XiogHio = (NN7/i1S),
the value of Kg at 100° is 6.7 x 10-° (atm.) Find the fraction of phosgene dis-
sociated at 100° under the following circumstances:
(a) 1 mole of phosgene in a 100 1. vessel.
(b) 1 mole of phosgene in a 1 1. vessel.
(c) 1 mole of phosgene in a 1001. vessel containing chlorine at a partial
pressure of 1 atm.
(d) 1 mole of phosgene in a 100 1. vessel containing nitrogen at a partial
pressure of | atm.
Ans: @) 148) 1045 (b) F483 <5 1On2(C)iGw/s 10a (GES 10m:
12. A vessel having a volume of 503 ml. was filled at 50° with methanol vapor
at a partial pressure of 37.6 mm. and with 0.0686 g. of nitrosyl chloride. After
equilibration the partial pressure of nitrosyl chloride was found to be 20.7 mm.
Find Kg for the reaction
CH,OH + NOC! = CH,ONO + HCl
Ans. 1.33.
13. Given the standard free energies of formation of the three isomeric pentanes,
find their mole fractions in the equilibrium mixture at 600°K.
AF (n-pentane) = 33,790 cal. mole!
AF (isopentane) = 32,660 cal. mole!
AFgo (neopentane) = 35,080 cal. mole-!
AES 2G, = ULSD, Gey
2 = OS
14. One g. of iodine and 100 mm. partial pressure of hydrogen is introduced
into a 11. vessel at 25°. The vessel is then heated to 666.8°K. From data in
Table 5.3 find the amount of hydrogen iodide at equilibrium.
Ans. 3.4 x 10-° mole.
15. From the data in Tables 5.5 and 3.2 find Kg at 200° for
CO (g) + H,O (g) = CO, (g) + Hy (g)
Ans. 213.
18. A reaction mixture contains 0.2 mole A, 1.0 mole AB initially. At equili-
brium, 0.1 mole B is present at 2 atm. total pressure, unmeasured amounts of A
and AB are also present. How much B will be present at equilibrium when the
same mixture is expanded to | atm. total pressure?
19. For the reaction system 3A (g) + 2B (g) = 4C (g) + D(g) a mixture at
500°K contains initially 0.65 moles A, 0.40 moles B, 0.80 moles C and 0 moles
D in a volume of 101. Letting np = x at equilibrium,
Fi
(a) Express my, mp; No in terms of x.
Problems, Chapter 5 Chemical Equilibrium in Gaseous Systems 143
(a) Find the mole per cent of A, B, and x (where x = nq) at equilibrium.
(b) Find Kg.
(c) How is the position of equilibrium affected by adding D at constant total
pressure?
21. The compounds A, B, and C are interconvertible in the gaseous state. That
is, (i) A (g) = B(g), and (ii) A (g) = C(g). Given AF% (A) = 10 kcal., AF (B)
= 11 kcal., and AF% (C) = 14 kcal., all at 298°K,
(a) Evaluate Kg for reactions (i) and (ii) at 298°K.
(b) Find the equilibrium composition.
22. In the vapor of HCN both of the following reactions occur: 2 HCN =
(HCN),, Kg = 0.095; 3 HCN = (HCN);, Kg = 0.055.
(a) Find the composition, in mole percentage, of hydrogen cyanide vapor at
Pyota1 = 100 mm.
(b) Find Kg for (HCN), (g) = HCN (g) + (HCN), (g) [W. F. Giauque and
R. A. Ruehrwein, J. Am. Chem. Soc., 61, 2626 (1939)].
23. From the data in Table 5.5, find the vapor pressure of benzene at 25°.
25. Vapor density measurements on acetic acid vapor yield apparent molecular
weights which are in excess of the formula weight, 60, indicating that the acid
dimerizes. The following values of r = (obs. mol. wt./formula wt.) have been
obtained at the indicated temperatures and total pressures.
tC) 110 132 156 184
ip eau! 1.33 1t9 1.10
Pmm. 453 403 473 553
Find Kg at each temperature and obtain AH for
2 CH,;COOH (g) = (CH;COOH), (g)
26. For the reaction
Np (g) + O (g) = 2 NO (g)
Kg = 1.21 x 10-4 at 1800 °K and Kg = 4.08 x 10-* at 2000°K. Find
(a) AF%00 per mole of NO.
(b) AAi99) per mole of NO.
(c) Kg at 2500°K, taking AH independent of temperature.
144 Chemical Equilibrium in Gaseous Systems Chap. 5
30. For a reaction of the type A (g) + B(g) = M (g) + N (g), and for the case
that P(V — b) = RT is applicable, under what condition is PyPyx/Ar»Pz =
constant ?
31. Consider that P = (n/V) RT = cRT, where c is concentration of gas (e.g.,
AMOLeM Items) e
(a) Derive an expression for K, in terms of Kg.
(b) Derive an expression for dln K,/dT.
32. Ina reacting system described by A (g) + 2B (g) = 3C (g) the mole numbers
at equilibrium aren, = 4.5, ng = 6.5, nc = 2.0 at 400°K. Given AHfo = 46 kcal.,
(a) Evaluate the equilibrium constant at 400°K.
(b) Find AF ip.
(c) Find, within 1 per cent, K4o1/K4oo.
33. The fractional extent of dissociation, a, of F, has been measured with the
following results:
ae 513 565 604 658
a 0.007 0.035 0.097 0.247
Jerse, OS OS 6.89 UE 7.47 7.80
Evaluate Kg at each temperature and find AH for
zF, (g) = F (g).
[H. Wise, J. Phys. Chem., 58, 389 (1954)].
34. The vapor pressure of water at 0° and 1 atm. is 4.58 mm. Find AF for
(a) H,O (1, 0°, 1 atm.) = H,O (1, 0°, 4.58 mm.).
(b) H,O (1, 0°, 4.58 mm.) = H,O (g, 0°, 4.58 mm.).
(©) 1ELO CW’, tl etian,)) == 18 OG, Os Il aim),
(d) H,O (s, 0°, 4.58 mm.) = H,O (g, 0°, 4.58 mm.).
(e) H,O (1, 0°, 1 atm.) = H,O (g, 0°, 4.58 mm.).
Problems, Chapter 5 Chemical Eqiulibrium in Gaseous Systems 145
(f) For the process H,O(s, ¢°, 4.58 mm.) = H,O (1, ¢°, 4.58 mm.), AF =0.
Find ¢°.
35. Show by thermodynamic considerations that the equilibrium composition
of a chemical reaction mixture cannot depend upon the presence or absence of
a catalyst (e.g., platinum).
36. Show mathematically that two equilibrium compositions for the reaction
system aA + bB = cC + dD can be the same for different initial mole number
ratiosa:B:¥:6.
37. Consider the reaction
aA + bB = mM
and its equilibrium function in the form
nit (P, We
Kg=
ni n\ ny,
Using previously defined symbols and for the case that Kg < 1, so that at con-
stant total pressure
K’=nma*B
show that the yield of M is a maximum for @ + 8 = constant when the ratio
p = a/B is the same as the stoichiometric ratio a/b.
38. At somewhat elevated temperatures elemental iodine reacts with paraffinic
hydrocarbons to produce olefins, diolefins, and acetylenes of the same carbon
skeleton as the reactant.
CnHon+2(g) + I: (8) = CnHen (g) + 2HI (8); Ka
CnHon (g) + I; (8) = CaHen-2(g) + 2HI(g); = Kp
(a) Using the data of Table 5.6, find K, and Kg at 500°K and at 1000°K
for C,H,, C.Hy, and C,Hy.
(b) By interpolating with In K = —AA/RT + const., find K, and Kz at 958°K.
(c) In an actual run at 685° and 1 atm. the initial ratio I,/C,H, was 4.6 and
24 per cent of I, subsequently reacted. On the basis of 100 moles C,H, initially
present, 1.9 C,H,, 72.0 C,H,, and 10.0 C,H, were present at equilibrium.
Evaluate K, and Kp from these measurements [J. H. Raley, R. D. Mullineux,
and C. W. Bittner, J. Am. Chem. Soc., 85, 3174 (1963)].
ENTROPY
AND THE THIRD LAW
OF THERMODYNAMICS
a ee)
os [Winar | —_ loner rs Gamer"
iim.
Apparently, nature has placed some limitations on our ability to convert heat
into work.
Another type of device which belongs in the general class of cyclic heat
engines is more commonly known as a refrigerator or heat pump. In this de-
vice, heat is absorbed by a working fluid from a low temperature region (cold
box), work is done on the fluid, and heat is released to a high temperature
region (surroundings). That is, each of the operations indicated in Fig. 6.1 is
reversed. Again, the first law of thermodynamics requires that the heat de-
livered to the surroundings must be equal to the sum of work done on the
fluid plus heat absorbed from the cold box.
pe
tall. (ala See
1The symbols |q.| and |q;| are used to denote the absolute amount of heat absorbed
or evolved by the system without regard to sign. The same applies to |w].
148 Entropy and the Third Law of Thermodynamics Chap. 6
decreases as the device is operated in more and more nearly reversible fashion.
That is, |w| decreases, and the limiting value of |w|/|q.| is
Weel. eee
Nim.
= Idi) eee eel oe)
Practical experience indicates that the limiting value of the coefficient of
performance of a heat pump is not zero. Indeed, this would correspond to flow
of heat from 7, to T, with no expenditure of work. All experience denies this
possibility. Even with the most efficient machine operating in reversible fash-
ion, a finite amount of work must be done to remove heat |q,| from the cold
box.
We now inquire into the nature of this limit on |w|/|q.| and ask what ex-
perimental circumstances govern its value.
Let the heat engine B operate as a heat pump so as to exactly consume the
work output of engine A, as indicated in Fig. 6.2. That is,
| Wrnax. |(A) a | Wrnax.|(B) (6.8)
Vo
q1 rev == Wmax. = nRT, In Vp (6.12)
and, therefore,
Vg Va
An
or
== u (6.16)
Equation (6.14) simplifies to
|Wmax.|(cycle) = (J, — T))nR In (Vz/V 4) (6.17)
and the limiting efficiency is given by equations (6.10) and (6.17),
i, Wet
awe lesT. 6.18 )
(6.
150 Entropy and the Third Law of Thermodynamics Chap. 6
For a refrigerator which keeps the cold box temperature at 0° and delivers
heat to the room at 25°, the coefficient of performance is
298° == 273°
Miim. me 298° = 0.084
Since this is
|Wax. | = |Wax. |
the reciprocal of mim, will give the relation of|Wyax.| and |q; rey.|-
1 — 1 a lq: at
Mim. |Wmax. |
6 19
1 1 = 10 8
Ch =e —
; ( Hi )
[Wiranl iim.
Equation (6.19) shows that in the present example the minimum work input
to the refrigerator must be 1/10.8 cal. per calorie of heat removed from the
cold box.
EXERCISE 6.1
|G» coy | T,
(6.20)
ldo rey. | = 19 rey. |
iis ae
Since gq = 0 for the adiabatic steps (BC and DA) in the Carnot cycle, it follows
that for the complete cycle
Sec. 6.3 Entropy and the Third Law of Thermodynamics 151
» $
Greet
T,
exer.
jishaa 0
i@iteay
(6.21)
Although this relation has been demonstrated only for the Carnot cycle,
which consists of two isothermal and two adiabatic steps, it can be shown
that it also applies to any other cyclic process. Furthermore, the relation is
independent of the nature of the working fluid, and consequently we shall take
equation (6.21) to be applicable to all cyclic processes. That is, for any cyclic
process
0h
lee il ie Y
as —= ie
Wrev. (6.22)
As stated, the change in the value of this property in a cyclic process is zero.
AS ycie
=at Advev. _
ioe Tey ae ad 0 (6.23)
It will be remembered that AH,,.., and AF,,.1. are also zero. By analogy to
those functions of state, the entropy change for a given change in state can
be measured or computed, but the absolute value of S cannot be determined.
The value of AS for a given change in state is given by integrating equation
(6.22).
S,—S,= | Mae
-{
= a
dV — (nkaVy
a.
=—nhiln vs
Lone nR In P,
P, (6.28)
AS iz 22.4
= OomR Ines
AS One in Pie=
Sec. 6.3 Entropy and the Third Law of Thermodynamics 153
The entropy change is a function only of the mole numbers and volumes. For
mixing any amounts of gases A and B at a common temperature and pressure,
it follows that
RS nk in es eR ins (6.29)
Van Vz
The problem of finding AS for
aA (V4, Ty) + 5B (Vz, Tz) = (a + 6) M (Vu, Tu)
where M refers to mixture, is resolvable into the component entropy effects
for changing temperature (by equations 6.25 or 6.26), changing pressure (by
equation 6.28), and finally mixing (by equation 6.29). For one mole of mixture
this is more conveniently expressed in terms of mole fractions as
Soe === DEX R In DK a XG R In XG (6.30)
Similarly, for mixing any number of gases at a common temperature and pres-
sure, the entropy of mixing, per mole of mixture, can easily be shown to be
ASais = == /® »y X; In X; (6.31)
It can also be shown that equations (6.29)-(6.31) apply as well to mixing two
liquids which form a solution obeying Raoult’s law.
As a final instance of the computation of entropy changes in simple physical
processes, consider the phase changes, such as fusion and vaporization. For
each change, such as
H,O (1, P, T) = H,O (g, P, ff)
there is for any given pressure only one temperature at which the process is
reversible. At a pressure of | atm. the vaporization of water is reversible only
at 100°. At the equilibrium pressure and temperature
AH = Frey.
and, therefore,
N= oe (6.32)
Equation (6.32) applies to reversible, isothermal processes such as
H,O(I, 1 atm., 100°) = H,O(g, 1 atm., 100°)
for which
JENS)
AS37. = = 26,0'cal. deg.- moles
373°
At 25°, the equilibrium vapor pressure of water is 23.6 mm. and, therefore,
equation (6.32) applies to
H,O(1, 23.6 mm.) = H,O(g, 23.6 mm)
for which
ANHses—= 103520 cal. mole
154 Entropy and the Third Law of Thermodynamics Chap. 6
Therefore,
INSii = 10,520
293°
== 3),.3:Cal, devas moles:
EXERCISE 6.3
Compute AS at —10° for
H,O (s, 1 atm.) = H,O (J, 1 atm.)
Take Cp (H,O,s) = 9 cal. deg.-! mole and Ad,,,; (fusion) = 1440 cal. mole-!.
Ans. ASoe; = 4.93 cal. deg.=! mole-!.
system surroundings.
Let us now restore the system reversibly to its initial state, transferring heat
Gre. and work w,., from the surroundings to the system. The net entropy
change in the combined, isolated, complex is zero. Although the system in its
final state is unchanged, the surroundings have been degraded. For this revers-
ible process AS% + AS, = 0. Also, the system has been returned to its
initial state, and AS{,.. + AS... = 0. Consequently, AS... for the entire opera-
tion is
DSS soi aad (AS evan se SS eee ae ae AS =F NS es )ee.
or
AShe = See ae AS = ASe = DS core (6.35)
The amount of work w,.,. required to restore the system equals in magni-
tude (but differs in sign from) the maximum work w’,,, which could have been
produced by the system in the forward process:
(Wier + Wrox Jaya: = 0 (6.36)
Correspondingly,
(Gre. Grew ayes 0 (6.37)
Combining equations (6.34) and (6.37) gives
(Yrov. + Jirvev.)ayst. < 0 (6.38)
and, with changed signs, for the surroundings:
(Grov. + Jirrev.)surr, > 9 (6.39)
For changes at constant temperature, we divide both sides of the inequality
(6.39) by T and obtain, by equations (6.35) and (6.39),
ASret= (ASirev. + ASrev.)surre, > 0 (6.40)
For nonisothermal processes it would be necessary to resolve the reversible
steps into differential isothermal steps, with dq,.,./T = dS, and connect them
by differential adiabatic reversible steps with dg... = 0, dS =0. Again, it
can be shown, both system and surroundings being taken into consideration,
that for any irreversible change there is a net increase in entropy.
EXERCISE 6.4
One mole of steam at 100°, and 1 atm. and 100 g. of ice at 0°, 1 atm., comprise the
initial state of an isolated system. What is the final state, and what is the net
entropy change? Ans. Final temp.30°, AS = 9.9 cal. deg.
It has been shown in Chapter 4 that the Gibbs free energy F and
the Helmholtz free energy A are important thermodynamic functions of state.
These properties can now be fully defined in terms of the entropy, the earlier
restriction to isothermal processes being removed:
A=££—TS (6.41)
F=H—TS (6.42)
Since E, H, and S are properties of state, it follows that A and F are also. For
156 Entropy and the Third Law of Thermodynamics Chap. 6
any given process, the values of AA and AF depend only upon initial and final
states, and not on the path.
A, a A, — E, a E, a (T.S'5 am T,S;)
NINE VES (6.43)
Be Fi i dg eS)
INT SIR ILS (6.44)
Although AF can now be computed, in principle, for any change in state
including one in which the initial and final temperatures are not the same,
this thermodynamic function is principally used for constant pressure, constant
temperature changes. For an isothermal change in state
AE, — AH, =a AS, (6.45)
Since
AH = AE + APV
then
AF, = AE, + A(PV), — T ASp (6.46)
For an isothermal change
T AS = Grey,
and from the first law AE = qrey. — Wmax.. Therefore,
— AFy = Wmax, — APV (6.47)
For a constant pressure change in state
— AFy p = Wmax. — P AV (6.48)
which conforms with the limited definition of AF given in Chapter 4. Wax.
includes all possible kinds of work, that is, electric work as well as expansion
work. If we separate the work term into expansion work w and all other kinds
Of WOIK, Wras ne) SO that
SINCE Wexp, = P AV at constant pressure. This is to say that the decrease in free
energy is a measure of the maximum net work obtainable from the specified
change.
EXERCISE 6.5
Starting with equation (6.43), show that, for an isothermal change in state
—AA = Wmax.:
there still occurs the unavailable work of expansion against the constant pres-
sure of the surroundings, P AV. Therefore, in such a system a process for
which Wyyax. (a function of state) exceeds P AV is a permitted change. That is,
AF < 0. If Winax. = P AV, AF = 0 and the process is reversible. When AF > 0,
the process is forbidden.
Returning to the relation given for AF in a constant temperature change
(equation 6.45), we note that the term TAS is sometimes called the unavailable
energy. Since
AF = — Wax. + APV = AH — TAS
and since
AE = AH — APY
then
— Wnex = AE— T AS (6.50)
This equation shows that the maximum work available from a change in state
is less than the internal energy change AE by the amount 7 AS. That is, an
amount of energy T AS is not available for conversion into work, even when
the change is carried out reversibly. This situation is similar to that which is
found with heat engines, in which only part of the heat absorbed from the
source can be converted into work.
For any isothermal change in state, the value of AF may be computed
by equation (6.45). For example, it has been shown that for
H,Od, 1 atm., 100°) H,O(g, 1 atm., 100°)
Sie. — 9713 cal mole-*
and
AS-i — 20.0 cal. des-* mole>*
Therefore,
Nig — 9713 — 373.16: X 26:0 = 0 cal: mele
This process is, of course, reversible.
On the other hand, for
EL Oe! atime 255) =H,O (2,1 atm. 252)
Wil. = 010,520 cal. mole>
and
AS5os— 28.0 cal. deg. *mole: !
Therefore,
AF9g = 10,520 — 298° x 28.0 = 2170 cal. mole~*
If AF and AH are known for some isothermal change, equation (6.45) may
be used to compute AS. It was shown in Chapter 5 that evaluation of the
equilibrium function yields a value of AF? = — RT In K. For example, for
N,(g, 1 atm.) + 3H,(g, 1 atm.) = 2NH;(g, | atm.)
NES. = — 1 OS kcal:
From Chapter 3
AHS, = — 22.08 kcal,
Therefore,
[NS(ksiy ==
Pes AF bo Je AF
798.99 og
= — 47.45 cal. deg.~!
EXERCISE 6.7
From data given in Chapters 3 and 5 compute AS% for
H, (g, 1 atm.) + I, (s, 1 atm.) = 2 HI (g, 1 atm.)
Ans. AS% = 39.5 cal. deg.—.
(=), By (6.56)
Sec. 6.6 Entropy and the Third Law of Thermodynamics 159
At constant temperature
(OA St)
Fp,r) ps
= —Sr,» SANT by 7 RG ok (6.59)
This is the Gibbs-Helmholtz equation previously given as equation (4.17), since
AS = Grey./[T. However, the chain of reasoning which has been used in the
present development contains none of the limitations which were present in
the arguments which lead to equation (4.17). Equation (6.54) applies to all
types of substances, and to chemical change as well as physical change.
The remainder of the development of useful forms of the Gibbs-Helmholtz
equation follows much the same pattern as used in Chapter 4, except that
we now have AS instead of g,.y,/T. The principal utility of the Gibbs-Helmholtz
equation lies in its application to the temperature dependence of the equil-
ibrium constant (see section 5.8), which is related to the standard free energy
change by
AF? = — RT1\n Kg
or
“is she
The temperature derivative for this property is obtained as follows: Since
@ (Az B lca(AFs) «ARE
WMC an Am fi T?
160 Entropy and the Third Law of Thermodynamics Chap. 6
éInKe _ AH%
a(/T) R
For the case that AH’ may be considered to be independent of temperature,
integration gives
VieehES ee (6.61)
For large m we may use Stirling’s approximation (see Appendix 1)
Inn! =nInn—n (6.62)
from which it follows that
InW=nlnn—n— Ya Inn, — n) =nIlnn — Dn, Inn, (6.63)
Sec. 6.8 Entropy and the Third Law of Thermodynamics 161
Clearly,
ninn=n Inn+n,Inn+...=Dn,
Inn
and equation (6.63) can be written
4Sinw=- 5% X, (6.65)
The similarity of this equation to equation (6.31) suggests that entropy is a
function of probability. Multiplying both sides of equation (6.65) by the gas
constant R and taking n = N = 6 x 10” gives
*inW=kinW=—REXInX,
These relations suggest, but do not prove,
Sk in WW (6.66)
This is Boltzmann’s equation relating entropy and probability; k is known
as Boltzmann’s constant.
When one mole of a gas, confined in volume V,, is given access to an ad-
ditional volume V,, it expands to fill the available volume. The entropy of the
initial state of the system is related to the chance W, of finding all N mole-
cules in V,, [V,/(V, + V,)]”. The chance of finding a given molecule in V, is
V/V, + V.). The chance of finding all N molecules in V, + V,, correspond-
ing to the final state of the system, is W, = 1. Then, by equation (6.66),
See
Oat Ta
ae Sain one Spi eae
V, (6.67)
The preceding considerations apply equally well to a superficially dissimilar
problem, that of arranging molecules in a crystal. Suppose that small, nearly
symmetric molecules can be arranged head-to-tail, as (I): AB — AB —AB—.
or, randomly, (II): AB — BA — AB — AB— .... For N molecules, W; = 1” = 1
and W,, = (4)”. The AS between these two states is
= {Cran
ree + So (6.68)
For a pure substance, invoking the third law, we have
Sie= Ia
fi Grdr
(6.69)
The great importance of equation (6.69) in its application to chemistry lies
in the possibility of evaluating the entropy change for a chemical reaction:
AS; = mS>(My -2nS; (NE. = aSA) SB) — ee EGO)
Sec. 6.9 Entropy and the Third Law of Thermodynamics 163
cal. mole
This is indicated in Fig. 6.5, where Cp is plotted versus In T. The area under
the curve between limits is 16.57 cal. deg.-? mole-'.
The measured AH,,,,.of fusion is 1531 cal. mole~! and, therefore,
1531
ASi71.19 (fusion) a 171 (ae = 9.90 cal. deg. “mole:
From 171.12°K to 239.05°K, the entropy change for heating the liquid is
evaluated by measuring the area
239.05
under the curve between the given limits. The value of AS for this change is
5-25'cal, deg. =smole==
EXERCISE 6.8
From Fig. 6.4 estimate the mean value of Cp (Cl,, 1) between 171°K and 239°K.
Find AS for this temperature change by application of equation (6.27) and com-
pare with the graphical result.
20 x ~
ale
= (on)
Buh
ale) y Z 7
If
op A-8 is
cal, mole Ye)4
"
10 Pe
eae RS)oi
5 «Xl
m
area = |Cod ln7 =16.57
ba
: Siz
°
Temp. °K Method AS
0-15 Debye theory 0.33
15-171.12 graphical 16.57
7c? AA/T 8.90
171.12-239.05 graphical 3323
239.05 AA/T 20.41
51.44 cal. deg.-! mole
This is the measured third law entropy of gaseous chlorine at 1 atm. and
239.05°K. The standard entropy at this temperature differs by a small correc-
tion of + 0.12 cal. deg.-’ mole! for the gas imperfection of chlorine at this
temperature. With this correction,
Sears) —=01.50. cal. des. mole
To find S% 1, Cp for chlorine gas from 239.05°K to 298.10°K is needed.
This is practically constant at 8.20 cal. deg.-' mole~' over this temperature
range, and therefore
TABLE 6.1
THIRD LAW ENTROPIES*
S598 Soo
Substance (cal. deg.-! mole-!) Substance (cal. deg.-! mole)
Oz (g) 49.003 H,S (g) 49.15
Hy (g) Salil NO (g) 50.339
N> (g) 45.767 NO, (g) 57.47
C (graphite) 1.3609 NH; (g) 46.01
Cl, (g) 53.286 CO (g) 47.301
Bry (g) 58.639 CO; (g) 51.061
I, (s) 27.9 methane, CH, (g) 44.50
S (rhombic) 7.62 ethane, CoH, (g) 54.85
H,0 (g) 45.106 propane, C3Hs (g) 64.51
H,0 (1) 16.716 normal butane, C,Hyjp (g) 74.10
HCl (g) 44.617 isobutane, Cy Hp (g) 70.42
HBr (g) 47.437 normal pentane, C,H, (g) 83.27
HI (g) 49.314 normal hexane, CgHy, (g) 92.45
SO, (g) 59.40 benzene, CogHe (g) 64.34
SO; (g) 61.24 ethylene, CoH, (g) 52.45
acetylene, CoH, (g) 47.997
The value of AS° for a chemical change at some temperature other than
25° is obtained by application of equation (6.27) to each of the reactants and
products. When the heat capacities may be considered constant, the general
expression is
ASA S UNG a
EXERCISE 6.9
Use data from Table 6.1 and the heat capacities from Table 2.1 to find AS;,;
for the combustion of methane to form carbon dioxide and water vapor.
Ans. —.71 cal- deg. mole.
SUMMARY, CHAPTER 6
ie Limiting efficiency of cyclic heat engines
W max. Lae Le = T;,
Ulin ==
qo T,
Seok
For mixing
AS = — RD XxX, In X;
3: The Helmholtz free energy, defined by
A=E—TS
The Gibbs free energy, defined by
r= HTS
For an isothermal change
Rae TAS = = Wins, tot
AF = AH — TAS = — Wyascmee
For pressure and temperature changes of a pure substance
dF = (55),
oP
dp (55)
oT] p
Gis gPeas ah
At constant temperature
dF= V dP
168 Entropy and the Third Law of Thermodynamics Chap. 6
At constant pressure
dF = — SdT
For a chemical reaction or phase transformation
ser | BESS
oT IP.
C(AE/ DA
oT If
é ln Kg ue AH
OL/T® “2 ER
4. The third law of thermodynamics
Probability
S=kinW
Third law
So = 0
Third law entropy
gee i CdinT
ya
PROBLEMS, CHAPTER 6
1. If the heat sink of a cyclic heat engine is at 0°, what must be the temperature
of the heat source for the engine to have a limiting efficiency of 0.5?
Ans. 273°.
2. What is the minimum work per hr. which must be expended to keep a
refrigerator at a temperature of —5° when the room temperature is 30° and heat
flows into the refrigerator at a rate of 1000 kcal. hr.-'? How much heat hr.~-!
is liberated to the room? Ans. 130 kcal. hr.-!, 1130 kcal.hr.-!.
3. A sample of ideal gas is carried through a reversible Carnot cycle. It absorbs
1000 cal. at 400°K (step I), expands adiabatically to 300°K (step II), releases
x cal. at 300°K (step III), and is restored adiabatically to the initial state (step
IV).
(a) Evaluate AS for each step.
(b) Evaluate x (step III).
(c) Evaluate Wmax,. for the cycle.
Ans. (a) AS(I) = 2.50 cal. deg.-!, AS(D = 0, ASCID = —2.50 cal. deg.-',
AS(IV) = 0 cal. deg.-!, (b) —750 cal., (c) 250 cal.
Problems, Chapter 6 Entropy and the Third Law of Thermodynamics 169
4. Find AS», for H,O (1, —5°) = H,O (s, —5°), using Cp (H,O, 1) = 18 cal.
deg.-! mole-!, Cp (H,O,s) = 9 cal. deg.-! mole!, and AdA,,; (H,O, fusion) =
1440 cal. mole. Ans. —5.11 cal. deg.-!.
5. Using data in Table 6.1, find AS%s for combustion of ethane to form carbon
dioxide and liquid water. Compare T AS for this reaction with the heat of com-
bustion from Table 3.1. Ans. AS = —74.1 cal. deg.-! mole-.
6. Evaluate AS for
H,O d, 1 atm., 125°) = H,O (g, 1 atm., 125°)
Some of the following data may be useful: Cp (H,O, 1) = 18 cal. deg.-! mole;
Cp (HO, g) = 9 cal. deg.-! mole!; AH3;3 (H,O, vap.) = 9713 cal. mole; A Agog
(H,O, vap.) = 9400 cal. mole; vapor pressure at 125°, 1740.5 mm.
Ans. 25.4 cal. deg.—.
7. In the temperature range 300-1000°K, the molar heat capacity at constant
pressure of water vapor is given by 7.0 + 3.0 x 10-* T cal. deg.-! (see Exercise
2.3). Find Soo for water vapor. Ans. 55.6 cal. deg.—! mole.
8. Devise and describe a reversible path for the process
H,O (1, 3 atm. 125°) = H,O (g, 1 atm., 100°)
Use data from Problem 6 to evaluate AS for the system.
Ans. 24.9 cal. deg.-! mole.
9. From the data of Problem 5, find AF: for the combustion of ethane.
Ans. —351 kcal. mole.
10. One mole of helium at 1 atm. and 25° expands isothermally into a con-
necting 101. evacuated container. Evaluate g, AH, AF, and AS.
Ans. q = AH = 0; AF = —203 cal., AS = 0.681 cal. deg.+.
15. From data in Table 6.1, find the change in AF° per degree at 298°K for
2 N2 (g) + 3 He (g) = NH; (g)
16. Find AS for
N, (g, 1 atm., 25°) = N, (g, 10 atm., 100°)
17. Find the molar entropy of N,O at its normal boiling point from the follow-
ing data [R. W. Blue and W. F. Giauque, J. Am. Chem. Soc., 57, 991 (1935)]:
S,, = 0.214 cal. deg.-! mole-!; melting point = 182.26°K; heat of fusion = 1563
cal. mole-!; normal boiling point = 184.59°K; heat of vaporization = 3958 cal.
mole-!. The heat capacity at various temperatures is
T CK) 20 30 40 50 60 70
Cp (cal. deg.-! mole-!) 1.51 354 OS OrS2 an 5 OMS,
80 90 100 110 120 130 140 150
8.95 9.44 9.90 10.32. 10.77 IWS) a2 IAL
160 170 180 182.26 — 184.59°K
WT 13.30 13.98 Cp (I) 18.57
18. Evaluate AS for releasing 1 mole O, at 25°, 1 atm., into the air at the same
temperature and pressure.
19. It is notoriously difficult to reduce systems to low temperatures and keep
them there. Liquids with low boiling points are commonly used as refrigerants.
Heat leaking into the system is offset by the latent heat of vaporization. According
to Trouton’s rule, AS\a,). = 21 cal. deg.-! mole“! at the boiling point. How
many moles of refrigerant must evaporate per kcal. leaking into the system when
the refrigerant is
(a) Liquid nitrogen, b. p. = 77°K?
(b) Liquid hydrogen, b. p. = 20.4°K ?
(c) Liquid helium, b. p. = 4.2°K?
20. At 473°K, Ag,O is in equilibrium with oxygen when Po, = 1.35 atm. Evalu-
ate AS,,, for the process
2 Ag,O (s) = 4 Ag (s) + O, (g)
at a constant pressure of | atm., given AH,;, = —14.4 kcal.
21. For two liquids A and B which obey Raoult’s law, AAnix. = 0 and con-
sequently AH.» (A) and AH,,,. (B) are the same for the pure liquids and for
the mixture. Devise and describe a reversible process for mixing two liquids,
and show that @rey,/T = —R (X4 In X4 + Xz 1n Xz).
22. An ideal, monatomic gas expands reversibly and adiabatically from P,, V;,
T, to P,, Vz, T, and AS = 0. Replace this process by a sequence of reversible
processes, none of which is adiabatic, and show that the sum of the component
entropy changes is zero.
23. Ina refrigerator, or heat pump, the work wyey, is required to deliver heat
Grey, from the system at 7, to the surroundings at 7,, where T, > 7,. Show that
|Wrev.|/ |rev.| = (72 — T,)/T}.
24. Find wyey. required to remove grey, = 1 kcal. to a room at 298°K from a
system to be maintained at the temperature of
(a) Boiling nitrogen, 77°K.
(b) Boiling hydrogen, 20.4°K.
(c) Boiling helium, 4.2°K.
(d) Absolute zero.
Problems, Chapter 6 Entropy and the Third Law of Thermodynamics 171
This is a specific example of equation (7.2) and in this case K = #. For exam-
ple, the vapor pressure of water at 25° is 23.6 mm., or 0.0311 atm.
AF2,(vap. H,O) = — 1.987 x 298° In 0.0311
= 2057 cal. mole™!
This is the free energy change for
H,O(, 1 atm.) = H,O(g, | atm.)
which is evidently a forbidden change at 25°.
EXERCISE 7.1
The vapor pressure of ice at —10° is 2.0 mm. Compute AF, for
H,0 (s) = H,0 (g)
when the standard state of the vapor is taken to be 1 atm. Ans. 3105 cal. mole™!.
Since
__ —AF? (vap.)
Le eR Ts
then
dinP _ AH’ (vap.) (7.9)
dT RT?
where AH°(vap.) is the change in heat content for the standard change in state
Ae pl =A (Sree 2)
When AH*(vap.) may be considered independent of temperature, equation
(7.9) integrates to
_ —AH? (vap.)
InP + const.
RT
or
inZz — SH? (vap.) (= = *)
F, R AROie
(7.10)
Equation (7.10) may be demonstrated graphically by plotting log A versus
1/T (°K), as shown in Fig. 7.1. The experimental points lie on a practically
straight line of slope — AH°(vap.)/2.3 R. AH°(vap.) of both liquids and solids
is invariably positive, and the curves have negative slope.
log @
(mm)
EXERCISE 7.2
From the slope of the appropriate curve in Fig. 7.1 estimate AH (vap.) for carbon
tetrachloride. Ans. 7.5 kcal. mole™!.
EXERCISE 7.3
Use equation (7.10) and the result of the preceding example to compute the vapor
pressure of water at 50°. Ans. 90 mm.
176 Phase Equilibria and Colligative Properties Chap. 7
According to Trouton’s rule (cf. problem 6.19), the molar heat of vapori-
zation in calories is approximately 21 times the absolute boiling point. Equi-
valently, AS(vap.) = 21 cal. deg.-! mole-'. Values of AH(vap.) and AS(vap.)
in Table 7.1 show that this is often, but not always, obeyed. Water and several
other substances do not fit this generalization very well and are said to be
abnormal liquids. Hydrocarbons usually have normal constants. The rule
permits estimation of the vapor pressures of a liquid over a range of tem-
perature from a knowledge of its normal boiling point.
TABLE 7.1
VAPORIZATION OF LIQUIDS
Normal AS°(vap.)
boiling AH (vap.) (cal.
point (cal. decnns
Substance (°C) mole-!) mole-!)
Water, H,O 100.0 9717 26.04
Carbon tetrachloride, CCl, 76.7 7170 20.5
Acetic acid, CH; COOH 118.2 5830 14.9
Ethanol, C,H;0H 78.5 9220 26.22
Normal butane, C,H; —0.5 5320 19.63
Normal pentane, C;Hj,. 36.1 6160 19.92
Normal hexane, CgH;,4 68.7 6896 20.17
Benzene, CgHg, 80.1 7353 20.81
Toluene, C;Hs 110.6 8000 20.85
Ethyl benzene, CgH;, 136.2 8600 21.01
Cyclohexane, CgH;, 80.7 7190 20.3
EXERCISE 7.4
The normal boiling point of 2-methyl hexane is 90°. Estimate its vapor pressure
at 0°. Ans. 23.4 mm.
at constant temperature and where X, is the mole fraction of the volatile solvent
in the solution. Employing the same type of cycle used previously, namely,
A
A (soln., X4, P) = AER)
AF| =0 AF = RT'In (7)
A 0
It can be shown that solutions which obey Raoult’s law (P = P°X, equa-
tion 1.24) are produced isothermally without change in enthalpy. That is,
ABH, = O0tor
nxA(l) + ngB(l) = soln. A, BCX,, Xp)
If Raoult’s law applies, then
InFA,=nPApt+in X (7.14)
and the temperature derivative at fixed composition is
7) ae) =! (2oe)
( OL T)xx OL Jxa Ce)
That is, the temperature derivative of the logarithm of the vapor pressure of
A(or B) over the solution is the same as that for the pure liquid, and it follows
that
AH°(vap. soln.) = AH°(vap. solvent)
Combining the thermochemical equations, we have
A (pure liq.) = A(g, | atm.); AH? (vap. solvent)
A (g, 1 atm.) = A (soln.); —AH” (vap. soln.)
Op = PX/P3
is the reaction function of pressures. It resembles the function used for all-gas
systems, but the reason that P, and Py do not appear in the functions Q and
K is not that these pressures are negligibly small, since solids may have vapor
pressures of many atm. (e.g., solid CO,). Rather, P, and Py do not enter into
Q> because the terms in { VdP for solids (and liquids) contribute little to the
summation of free energy terms.
For a set of pressures corresponding to an equilibrium state, AF, = 0
for the process, and
0 = AF}+ RT In Kg
AF}? = — RT 1|n Kg (7.17)
The decomposition of calcium carbonate illustrates the application of
equations (7.16) and (7.17).
CaCO,(s) = CaO(s) + CO,(g)
At 800°, Ago, = 167 mm., provided both solid phases are also present, and
Kg == P wor
Kg == Pryu, Pus
oF OF
(a). T, No... ae (on). T, Mm... ld a
= VdP — SdT+ j1,dn, + ptodnn +... (7.19)
The coefficient jz is called the chemical potential or the partial molar free
energy.
At constant temperature and pressure,
dO a Oe the
PI XP?
Equation (7.25), for n = 1, becomes
KE = i Ri (7.26)
When one state refers to the pure solvent, then equation (7.26), for any solvent
composition X,, becomes
i — Ve, = RE In, (7.27)
and
Ate RT ain oG (7.28)
EXERCISE 7.10
Show, by analogy with the preceding development, that when the volatile com-
ponent obeys Henry’s law (Chapter 1), equation (7.26) is valid for the solute.
Show also that equation (7.27) is invalid.
When one component of a binary solution obeys Raoult’s law and equation
(7.28) applies, equation (7.24) can be used to discover how P, depends upon
X, for the second component, or solute. By substitution,
Ny RT din X + N» d [Ly = 0
Divide each term by n,, then replace n,/n, by X,/X,, and d In X, by dX,/X,,
to give
dil, _= _ —RT
pp AX;
ze
Ain_ RT.
pp dX2
x,
Its— wg = RT In (Ft) = RT In X,
P}
or
d
eee) GY ea
That is, when one component of a binary solution obeys Raoult’s law, the
second component must also.
EXERCISE 7.11
Evaluate AFyo3 for mixing 2 moles of A and 8 moles of B to form an ideal solution.
Consider both
DAC — li) = ZACK = 0:2)
and
8B(Xz = 1) = 8B(X3 = 0.8)
184 Phase Equilibria and Colligative Properties Chap. 7
A=) ard)
Step (i) is the reverse of dilution of the solvent from the pure state to mole
fraction X,. The free energy change is given by equation (7.26). Step (ii) is the
vaporization of the pure solvent (X, = 1) at | atm. AF(ii) is the standard free
energy of vaporization as defined in section 7.1. Therefore, for the change
of interest
and since this is an equilibrium phenomenon, that is, the temperature is ad-
justed until the vapor pressure of the solution is | atm., AF, = 0.
eels
xX;
and its temperature dependence is obtained by applying the Gibbs-Helmholtz
equation.
(44) __ AH® (vap.) (4 — r)
In a= R Tz (7.31)
We are concerned here with the relation between the boiling points of solvent
Sec. 7.6 Phase Equilibria and Colligative Properties 185
(Xj = 1 at 7’) and solution (Xj! <1 at 7”). The boiling point raising is
AT, = T” — T' and equation (7.31) becomes
= Rie pe
Ripe AH (vap.)
In XY’ (7.32)
A simpler, approximate equation is usually adequate. The boiling points of
solvent and solution differ but little and we replace T’’ T’ by T? where T, now
represents, for convenience, the boiling point of pure solvent. Also, for small
solute concentrations X,< 1; —In X¥, = —In(1 — X,) may be expanded in
series
= In(dl — X,) = X, + SXGE EXE ++
and the higher terms neglected.
With these approximations, we have
AT SRT (7.33)
- AH? (vap.)
That is, the boiling point elevation is proportional to the mole fraction of solute.
In dilute solutions n, <n, and n, may be neglected in the denominator of the
mole fraction yielding
ss RT; ve RT} W,M,
AT, Avan) poe AA \(Vvap.) “Wi
(7.34)
Equation (7.34) is often used to estimate the molecular weight of a solute,
assuming that this property is fixed and unchanged by the process of dissolu-
tion. For example, suppose that 5 g. of an unknown solute is dissolved in 100
g. of water and that this solution is found to have a normal boiling point of
100.40°. AT, = 0.40° and all quantities in equation (7.34) are known except Mg.
RT} W, x M, 1
M
‘> AHO wap.) ~ «Wy ~~ AT,
LOST Oe@SIS oo X18 1
—oiea| rior 040
It should be noted that if any kind of reaction occurs between solvent
molecules or between solvent and solute, the “molecular weight” obtained by
use of equation (7.34) will not necessarily have a simple meaning. However,
equation (7.33) correctly gives the mole fraction of solute, which may consist
of more than one species, or a species different from that used in making up the
solution.
Since a single solvent A may be employed in boiling point measurements
with several different solutes, the properties of the solvent appearing on the
right-hand side of equation (7.34) are collected for convenience into a single
factor. It is also customary to express the concentration of solute in molal
units. (A 1 m solution contains 1 mole of solute in 1000 g. of solvent.) The
molality m = (W,/M,)/(W,/1000) and therefore by substitution in equation
(7.34)
186 Phase Equilibria and Colligative Properties Chap. 7
Der ee
BP a Hogan O00 Gey)
INTE = K, m
The factor RT?M,/[AH°(vap.) x 1000] is called the molal boiling point elevation
constant K,, that is, the boiling elevation for a 1 m ideal solution. In the case
of water, its value is
K, =
_ 1.987 x 373°? x 18 = 0.51° molal!
u 9717 x 1000
TABLE 7.2
MOLAL BOILING POINT AND FREEZING POINT CONSTANTS
EXERCISE 7.12
Compute the molecular weight of a solute which in a 6 weight per cent solution
in carbon tetrachloride elevates the boiling point by 1.815°. VATSewVie—el lel
T’ is the freezing temperature of the pure solvent (X, = 1), T” is the freezing
temperature of the solution (X,; < 1), and T’ — 7” is the freezing point de-
pression AT,, given by
al aA
RUT
(fusion) me
(7.39)
Equation (7.39) is completely analogous to equation (7.32) and for dilute
solutions (X, near unity) the same approximations may be made. Replacing
T' and T” by T,, the freezing point of the pure solvent, we obtain
ae RT?
le 7.40
ee NIE (fusion) x: Ca)
and
RT’ WM, (7.41)
EXERCISE 7.13
The freezing point of pure benzene, C,Hg, is 5.45° and its heat of fusion is
2351 cal. mole~!. Compute the freezing point of a solution containing 1 weight
per cent of naphthalene, C,,Hs, in benzene. Ans. 5.04°.
As in the case of boiling point elevation, the properties of the solvent which
appear in equation (7.41) may be collected into a molal freezing point depression
constant, K,.
AT, = K,;m
where
Ke a~ RT; M,
AH? (fusion)- 1000
(7.42)
Values of K, for various liquids are given in Table 7.2.
EXERCISE 7.14
Compute the molal freezing point constant for water. AH° (fusion) = 1440 cal.
mole™!. Ans. K;(H;O) = 1.85° mol.7}.
iz
Ne = LI (7.46)
Measurement of osmotic pressure is far more sensitive for measuring
molecular weight at very small solute concentrations than either boiling point
raising or freezing point lowering.
EXERCISE 7.15
An aqueous solution containing 1 per cent of protein by weight has an osmotic
pressure of 1.0 mm. Hg at 25°. Estimate the molecular weight of the protein.
Estimate the freezing point lowering of the solution. Ans; M=2 < 10°:
since the phase transition involves no change in free energy for the reversible
process at constant pressure and temperature.
Let the pressure be changed from P to P + dP, and the system restored
to equilibrium by an appropriate change in temperature from T to T + dT.
Then
Fp7(@) = dF(a) —- Fp (8) a dF(8)
This is the Clapeyron equation, which describes the relation between change
in pressure and change in temperature for a univariant phase equilibrium of
any kind. For example, the AH of fusion of ice is 1440 cal. mole! or 59.6 1.
atm. mole~'. AV is obtained by comparing the density of ice at 0° (0.917 g.
cc.-') and that of water at 0° (1.000 g. cc.~').
aie
ap 0.0075 deg. atm.
Ss
The freezing point of water will decrease by 0.0075° for each atmosphere of
pressure increase. (Over a small range of temperature AH and AV are prac-
tically independent of pressure and temperature and dT = AT, dP = AP.)
The freezing point of water is ordinarily observed with the system at 1 atm.
and this defines 0°C. However, if the freezing point is observed for water in
a closed vessel from which all gases (e.g., air) except water vapor are excluded,
the pressure on the system is that of the vapor, or 4.58 mm. at 0°. According
to the result obtained above, the decrease in pressure will increase the melting
point by 0.0075°. An additional increase of 0.0025° is to be expected due to
the removal of dissolved air. The equilibrium temperature at which solid,
liquid, and vapor coexist as pure phases is called the triple point.
EXERCISE 7.16
At —38.87°, the melting point of mercury at 1 atm., the density of the liquid is
13.69 g. cc.~! and the density of the solid is 14.19 g. cc.-! The heat of fusion is
2.33 cal. g.-! Estimate the melting point of mercury at 1000 atm. Ans. —32.6°.
SUMMARY, CHAPTER 7
1. Vaporization of pure solids, liquids, and solutions
AF3(vap.) = —RTIng
inZ? _ AH? (vap.) (4 — *)
Fy R oda
2. Heterogeneous chemical equilibrium
aA(s) + bB(g) = mM(s) + nN(g)
AF, = AF’ -- RT ln O>
AP? =" Rink
Kg CaS5
B
Apa ee
Si Km
5. Osmotic pressure (ideal dilute solution)
RX te
6. Clapeyron equation (any two phases)
dP_
ae.
AH
Clapeyron-Clausius equation (ideal vapor)
din? — AH
She ar
PROBLEMS, CHAPTER 7
1. The vapor pressure of water is 23.76 mm. at 25°. Given AF? of H,O (1) equal
to —56.6902 kcal. mole™! at 25°, find AF? of H,O (g) and compare with the result
of Table 5.5.
192 Phase Equilibria and Colligative Properties Chap. 7
2. The heat of vaporization of water is 9.7 kcal. mole~! at 100°. Find the vapor
pressure of water at 99°. Ans. 733.3 mm.
3. The vapor pressure of benzene is 115 mm. at 30° and 181 mm. at 40°. Find
the vapor pressure of benzene at 20°. Ans. 70.8 mm.
4. From the data of Table 7.1, calculate the vapor pressure of carbon tetra-
chloride at 70°. Ans. 622 mm.
5. The vapor pressure of solid carbon dioxide is 1.0 atm. at —78° and 2.0 atm.
at —69°. The vapor pressure of liquid carbon dioxide is 6.7 atm. at —50° and
14.1 atm. at —30°. Find the temperature and pressure at the triple point of
carbon dioxide. Ans. —53.5°, 5.9 atm.
6. The normal boiling point of cyclohexane is 81.4°. Use Trouton’s rule to
estimate the vapor pressure at 25°. Ans. 103 mm.
7. Show that a simple expression for correction of observed boiling points to
normal boiling points of the form
AT = 1.25 X 10! Tops. [P(mm.) — 760]
can be obtained from equation (7.9) and Trouton’s rule by use of approximations
appropriate for small changes in 7 and P.
8. The vapor pressure of water is 355 mm. at 80°, and for 50 weight per cent
LiBr solution it is 114 mm. at the same temperature. Calculate AF;;, for adding
one mole of H,O (1) to a very large volume of 50 per cent LiBr solution.
Ans. —798 cal.
9. The vapor pressure of water for 50 weight per cent LiBr aqueous solution is
256 mm. at 100°. Referring to problem 8, find AH (vap.) of water from this solu-
tion at 90°. Ans. 10.60 kcal.
10. From the result of Problem 9, find the enthalpy change for adding one mole
of H,O (1) to a very large amount of 50 per cent LiBr solution. Ans. .88 kcal.
11. One g. of gaseous H,S is admitted to a 11. bulb initially containing NH, at
1 atm. At 25°, how much solid NH,HS will be present at equilibrium?
Ans. 0.0210 moles.
12. In the reduction of iron oxide according to
Fe;0,(s) + H,(g) = 3FeO(s) + H;O(g)
at a total pressure of 1 atm., the equilibrium composition of the gas phase is
44 mole per cent H, at 678° and 27 mole per cent H, at 772°. Find AA for the
process. Ans. 15.8 kcal.
13. The equilibrium 2 FeCl,(s) = Fe,Cl(g) obeys the equation log P(mm.)
= —6887 T-! + 14.52.
(a) Calculate the equilibrium pressure at 120°.
(b) Find AA (vap.).
[R. R. Hammer and N. W. Gregory, J. Phys. Chem., 66, 1705 (1962).]
Ans. (a) P = 1.0 x 10-* mm (b) 31.5 kcal.
14. A solution containing 122 g. of benzoic acid, C,H;COOH, in 1000 g. of
benzene, C,H,, boils at 81.5°. Find the apparent molecular weight of benzoic
acid (which dimerizes) in the solution and the degree of dimerization.
Ans. 81 per cent dimer.
Problems, Chapter 7 Phase Equilibria and Colligative Properties 193
18. Find AF%,; for vaporization of water when the standard state of the vapor
is chosen to be 1 mm. pressure.
19. From data in Table 7.1 find the vapor pressure of n-butane at 25°.
20. The vapor pressure of supercooled liquid water is 2.149 mm. at —10° and
4.579 mm. at 0°. The vapor pressure of ice at —10° is 1.950 mm. From these data
find the heat of fusion of ice.
21. Find AFos for a system in which 100 g. of cyclohexane, C,H», is mixed with
100 g. of benzene, C,H,, assuming that an ideal solution is formed.
22. Iodine reacts slowly with cyclohexene, C,H, , to form an unstable diiodide.
If 0.01 mole of iodine is added to 0.25 mole of cyclohexene, what is the boiling
point elevation
(a) If no reaction occurs ?
(b) If substantially complete reaction occurs ?
For cyclohexene, ¢, = 83°. Use Trouton’s rule to find AH(vap.) and neglect
the volatility of the solutes.
23. When one mole of A(solvent) is added to a very large amount of its solution
with B(nonvolatile solute) at Xz and temperature 7, the attendant heat effect is
AHs. When pure A is vaporized at 7, the heat effect is AH;. The P, — T depen-
dence of the solution obeys the Clapeyron-Clausius equation, from which one
obtains a heat effect AAcc.
(a) Complete the accompanying diagram which demonstrates the relationships
among the various heat effects AH,, AAs, etc.
(b) State the fundamental principle employed to establish the relationships.
(c) Identify AH, ... As.
1
A (lig., X, = 1, 7) ——————
A (soln. X,, T)
A( A )
24. Use equation (7.24) and the result of exercise 7.9 to establish a relation be-
tween dP, and dP, for solvent and solute at constant temperature with com-
position variable. Use Fig. 1.11 to test approximately the result obtained, for
XG eX, — 0:5.
194 Phase Equilibria and Colligative Properties haps
25. A substance B(s) melts at T;, absorbing heat AH, (B), and B (1) forms an
ideal solution with A (1). When B (s) dissolves in A (1) at a temperature just below
T,, what is the integral heat of solution?
26. When the solid phase which separates from a solution on cooling is a solid
solution, equation (7.40) does not apply. The binary system A-B forms a solid
solution as well as a liquid solution. Both solutions obey Raoult’s law. Show that
the equilibrium function corresponding to
A(solid soln., X{) = A(liquid soln., XW
is
K = X4/Xx
Consider A the solvent and use the method of section 7.7 to find an expression
for AT; in terms of these compositions and of physical constants of the solvent.
27. A brief, correct derivation of the equation for b.p. raising begins: when
P, = PiX, and P, = Pum. =const., then In P, =In X, +1n P° and din
P,/dT =0 =d\n X,/dT + d\n P°/dT. Complete the derivation, remembering
that X, is the independent variable.
28. In an experimental measurement of molecular weight by boiling point
raising in Exercise 7.12, the boiling point of the solvent was measured at a baro-
metric pressure 756.8 mm. and that of the solution, somewhat later, at 755.6 mm.
What correction should be applied to AT, ?
29. The m.p. of Hg at | atm. is —38.87°; the heat of fusion is 2.33 cal./g.; the
density of the liquid is 13.69 g./cc.-!; the density of the solid is 14.19 g.cc.~!.
Calculate the change of melting point for a pressure increase of 2 atm.
30. It has been established that the decomposition pressure of CO, in equi-
librium with CaCO, and CaO is independent of the relative amounts of the two
solid phases, as long as some of each is present. From this fact, show that CaCO,(s)
decomposes at phase boundary surfaces, and does not decompose homogeneously.
31. The vapor pressure P of a solution of nonvolatile solute is a function of
temperature 7 and of solvent mole fraction X,. That is,
(5) at + (ee
ap =(& gy) ah
Complete each of the following equations in the most nearly appropriate manner
as applied to the measurement of boiling point raising for determination of
molecular weight.
dP=
(57),, >
(4) ey
OX\)/7
Use the preceding relations to obtain a differential equation suitable to describe
b.p. raising.
32. Monoclinic sulfur is stable above 368.5°K at 1 atm. and AF 5. = 18 cal.
mole! for S (rhombic) = S (monoclinic). The density of rhombic sulfur is 2.07 g.
cc.~'. At what applied pressure will rhombic sulfur have the same vapor pressure as
monoclinic sulfur has at an applied pressure of 1 atm. Consider both at 25°?
Problems, Chapter 7 Phase Equilibria and Colligative Properties 195
InP, =Ink+I1n X,
196
Sec. 8.1 Equilibria in Solutions of Nonelectrolytes 197
and
dit, RT din xX, (8.3)
In sufficiently dilute solution all measures of composition of solute are
proportional to each other. Accordingly, Henry’s law may be expressed in
terms of molality (P = k,m) or molarity (P = kM) as well as in terms of
mole fraction. It will be convenient to represent any measure of composition of
solute in dilute solution by the generalized symbol c. Then, provided P is
proportional to c, equation (8.1) gives
du + RT dlinc (8.4)
and
iL =e Rin (8.5)
where pl? pertains to solute at c = 1. Each choice of concentration unit es-
tablishes a corresponding standard state. For solutions which obey Henry’s
law up to 1m, py}, in the molal scale refers to the experimentally realizable
1 m solution, at each temperature. When Henry’s law is not obeyed, 1°, (or 2
in general) is best considered simply as a constant of integration which need
not be visualized (cf. section 8.6).
Henry’s law has limited validity for real solutions and often applies only
at rather low concentrations. Also, the state at unit concentration may not be
experimentally accessible because of a limited solubility. In either case equation
(8.4) is limited to the concentration range to which Henry’s law applies, and this
restriction applies equally to equation (8.5). The standard state is no longer the
real solution at unit concentration, say m = 1, but a hypothetical state which
can be partly described as having the solute vapor pressure which a | m solution
would have if Henry’s law were valid at 1 m. This hypothetical state is indicated
by the point A in Fig. 8.1, which corresponds to the extrapolation of the Henry’s
law line to 1 m solute concentration. Since by definition this hypothetical state
satisfies Henry’s law, for the process
A(hyp. std. state) = A
(m, Henry’s law region) FIG. 8.1. Standard states for
solutes.
m
}L = Hm + RT In =
When composition is expressed in terms :
of mole fraction, the standard state of the |
solute will be experimentally accessible only
for solutions which obey Raoult’s law over ? | ey
the entire range ¥,=0 to X, =1. For all |
other solutions it refers to the hypothetical
state described by extrapolation from the very |
dilute solution to pure liquid solute at X, = I.
One such state is indicated by point B in
Fig. 8.1. For the change in state
198 Equilibria in Solutions of Nonelectrolytes Chap. 8
= be + RT In 42
For any real change in state within the Henry’s law region, as
A(m') = A(m")
or
A(X) = A(Xs/)
the isothermal change in free energy, the free energy of dilution, is readily
obtained from equations (8.6) and (8.7).
Find AF% 3;for the solute when a Henry’s law solution containing 0.1 mole of
solute in 100 g. of solvent is diluted by addition of 900 g. of solvent.
Ans. A Foe — —136 cal.
8.2 Solubility
Many solutions exhibit a natural upper limit of solute concentration
called the solubility. It is characterized by the fact that at constant temperature
and pressure the pure solute is in equilibrium with the dissolved solute. The free
energy change for dissolving solute in solution approaches zero as the con-
centration approaches saturation. Representing the molar free energy of pure
substance at | atm. by 3, and that of solute in the saturated solution by
[bo cara thenewe have (i; —="[5 cae.
The significant free energy changes for solutions can be related by the cycle
AF
A (pure subs.) —— A (soln., c)
AF (i) = AF? =p — 4;
In step (ii) the solute is transferred from its standard state to any arbitrarily
chosen concentration. For solutions which obey Henry’s law,
Sec. 8.2 Equilibria in Solutions of Nonelectrolytes 199
Equation (8.8) is a familiar relation between the standard free energy change
and the corresponding equilibrium function. In this instance, K = c,.. Ap-
plication of the Gibbs-Helmholtz equation leads to an expression of the solu-
bility (equilibrium constant) as a function of temperature:
Insc ee
aT RT ee)
The enthalpy term in various applications of van’t Hoff’s equation is always the
heat effect for the given process involving standard states. In this case it is the
standard heat of solution.
For solid solutes which obey Raoult’s law, AH®° has special significance.
The vapor pressure P%,)4 of a pure solid substance at any temperature below
its freezing point is less than that of the supercooled liquid, Piouia (cf. section
1.9) at the same temperature. Raoult’s law applies only to liquids and in such
instances must be understood in the sense P = XPiiquia. The partial pressure of
the dissolved substance over any saturated solution is necessarily equal to that
of the pure solid solute and Pouia = Xeat, Piiquiae AN immediate conclusion is
that the solubility of a given solute is the same in all solutions obeying Raoult’s
law at a given temperature, since P$\4 and Piqua are functions of the solute
alone. Further, the heat of solution at any composition is the heat of fusion
of the pure solute, since the net process can be resolved into two steps:
A (pure solid) = A (pure liquid), AH (i) = AM nsion
A (pure liquid) = A (solution), Ad (ii)= 0
1000 grams solvent. a where Q and R represent the two immiscible liquids
ethyl acetate in water, b : 9
toe ee and cg and cg are not necessarily the saturation values?
Smaee Ata . .
acid in HoO, d—CyHy in Lhe change of interest is related to the standard change
H,O, e—ly in H,O, f— _ by the isothermal cycle
CoHg in HO. ;
: AF (dist.) :
A (soln. in Q, Cg) _ A (soln. in R, cg)
AF (ii) = AF °®(dist.)
A (soln. in Q, std. state) ee
A (soln. in R, std. state)
Steps (i) and (iii) are dilution processes. Step (ii) is the standard process, namely,
the transfer of solute from its standard state in solvent Q to its standard state
in solvent R.
0 = AF% (dist.) + RT In 2
Q
That is, the ratio of concentrations in the two phases is constant for a given
system and temperature.
Equation (8.11) may be valid in dilute solution even when it fails (because of
deviations from Henry’s law) at higher concentrations. When the solute is
sparingly soluble, then Henry’s law and equation (8.11) may be valid for the
saturated solutions as well. If this is the case, then in the saturated solutions
ee__ Cpee
(Sat.
(8.13)
and the distribution constant may be evaluated from the ratio of solubilities
in the two solvents.
Some solutes have different molecular forms in different solvents. For
example, picric acid, C,H,(NO,),;OH, exists largely as the undissociated mo-
lecular species in benzene solution, but dissociates as an acid in water forming
a positive and a negative ion. In this case the distribution equilibrium is
H* (aq.) -+ P- (aq.) = HP (benzene)
and the equilibrium function is
K = Cup (benzene)
Z Cie (aq.)
Table 8.1 gives some results for this system. The constancy of the values
obtained in the fourth column supports the proposed dimerization. Note that
the ratio of concentrations in the two phases (third column) is not constant.
TABLE 8.1
DISTRIBUTION OF BENZOIC ACID BETWEEN WATER AND BENZENE AT 6°
CHB (benzene)
CHB (aq.) CHB (benzene) CHB (benzene) c*HB (aq.)
m/l. m/l. CHB (aq.) (m/l1.)-1
EXERCISE 8.3
AFr =| bRT
In (1/cp) AFr= | nRT In (cyn/1)
AFr
aA (std. st.) + bB (std. st.) + ... — > mM (std. st.) + nN (std. st.) + ...
The change in free energy for the over-all reaction at arbitrary concentrations is
AF; = My - N[Ly | aco als bus = oare (8.14)
Since AF} is the free energy change for a uniquely defined change in state,
namely, that involving the standard states of all reactants and products, it
follows that K, has a unique numerical value for a given system at a given
temperature. This value is of course the equilibrium constant for the reaction.
Equation (8.19) is valid for any of the usual concentration units, including
molality, molarity, and mole fraction, so long as the solutes obey Henry’s law.
Of course, the numerical value of the equilibrium constant will depend on the
Sec. 8.4 Equilibria in Solutions of Nonelectrolytes 203
TABLE 8.2
DISSOCIATION OF N2O,4 IN CCly AT 25°
X?xo,
Xno, X 10° K = 3 x 10°
1.1 5.65
9.7 5.55
38.6 6.34
96.5 6.49
145.0 5.68
Since only rather dilute solutions are involved, where nyo, Nom, and
Ny,o, XK Nou, the equilibrium function of mole fractions may be written
:
k= X Xo, — Mxo. , it
; Xy.0. Mx.o. Noon
The number of moles of carbon tetrachloride is given by
Acco, =
Voc,
CCl4
XoePon414 — Vou, « 10.4
where Vo, is the volume of solvent (or « ution) in liters and poq, is the
density. Letting the degree of dissociation 0: «,O, be @ and n, the initial number
of moles of N,O,, we have
_ 4a? ie
oe l1—a Vou, X 10.4
4a’ Ky x 10.4
1 erat a ant Co
where ¢, is the initial concentration of N,O, in molar units. For example, when
c, = 0:01 M,
4a? 60x 10-5 x 10.4
ae OL0L :
CA Oi,
From the form of this equation, it is evident that the degree of dissociation
204 Equilibria in Solutions of Nonelectrolytes Chap. 8
will increase with decreasing N,O, concentration, in keeping with the principle
of Le Chatelier.
EXERCISE 8.4
Estimate the initial concentration of NO, at which 50% dissociation will occur
in carbon tetrachloride solution at 25°. AUS Sl? SOF?
Starting with ¢o,o,u,, = 0.01 M, what initial concentration of iodine will be re-
quired to produce 50% conversion of cyclohexene into the iodide at 25° in carbon
tetrachloride? Ans. 0.055 M.
where AH? is the change in heat content for the standard change in state, taken
to be independent of temperature.
EXERCISE 8.6
The equilibrium constant for the iodination of cyclohexene in carbon tetrachloride
solution is 13 (M)~! at 35°. Compute AH? for the process.
Ans. —7.9 kcal. mole!.
x tN +.%)
“(a= x\(8 — x)
When | mole of acetic acid is mixed with 2 moles of ethanol at 100°, the amount
of acetic acid at equilibrium is 0.150 mole. That is, x = 0.850, and
(0.850)(0.850)
k= (1 — 0.850)(2 — 0.850)
=A?
If large amounts of ester and water, with small amounts of acetic acid and
alcohol are present initially, the approach to equilibrium may increase the
amount of acetic acid and alcohol in the system. In this case the value of x, as
defined above, would be negative.
EXERCISE 8.7
Find the equilibrium composition of the system at 100° when 1 mole of ethyl
acetate and 1 mole of water are mixed.
Ans. Xester = XH,0 = 0.33; Xacia = Xalcohol = 0.17.
standard states of solutes has taken into account deviations from these ideal
relations, the application of equilibrium functions has been restricted to situ-
ations in which approximately ideal behavior has been found.
The limitation of the simple descriptions of equilibria in solution which have
been examined so far arises from the necessity to express the free energy of
dilution in terms of some measurable property and from the mathematical
convenience of inaccurate laws. Let us reconsider the problem of evaluating the
free energy change for isothermally transferring a component of a solution
from one composition to another. In the isothermal cycle
AF
A (soln., X’) A (soln.,. X"’)
AF) =0 AF |=0
AF =nRTIn(P"/P’) i
P’' and P” are the equilibrium vapor pressures over the solutions containing
A at mole fractions X’ and X”’. To be entirely correct it would be necessary to
use fugacity, not pressure, but for the majority of processes of interest the
inaccuracy is not great. The net change in free energy per mole of substance
transported from one solution to the other—whether by osmosis, electrolysis,
or vaporization—is given by
AE een
If either Henry’s law or Raoult’s law is invoked at this point, replacing P’’/P’
by c’’/c’, in whatever units, the consequence is the same as replacing equa-
tion (8.1) by (8.2). It would be possible to correlate pressure and composition
by using empirical descriptions, such as P= aX + bX? +..., but the equi-
librium functions would become unmanageable. The device which has been
generally adopted closely resembles the one employed for gaseous systems
(section 5.11). We define the activity a for a component of a solution so that the
equilibrium function of the activities will be correct. That is,
Katee HOES
Gane
(8.22)
and the value of K, will be constant without respect tothe applicability of
Raoult’s law or Henry’s law.
We f (8.25)
For a gas, f? = 1 atm. and a=f. The activity and fugacity are, therefore,
interchangeable for gases, except that fis dimensioned, whereas a is dimension-
less.
For the solvent in solution it is necessary to make a further choice in order
to relate activity to concentration. Recalling that f= P for substances in the
gaseous state at moderate pressure, we find that equation (8.25) becomes
ax = (8.26)
The resemblance of equation (8.26) to Raoult’s law is striking and suggests that,
for the solvent, the standard state be chosen for convenience as a, = 1 when
X, = 1 at each temperature. On this basis a, = X, for solutions which obey
Raoult’s law, apart from deviations of the vapor from ideal behavior. Even for
highly nonideal solutions (cf. Fig. 1.11), a = P/P° for moderate vapor pressures.
Equation (8.25) is evidently a complete analog of Raoult’s law, and the similarity
is altogether intentional.
EXERCISE 8.8
The vapor pressure of water over 5 m sucrose (Xy,0 = 0.918) is 20.7 mm. at
25° and Pt,o = 23.6 mm. Find ay,o for the solution and compare with Xy,0.
Ans. a = 0.878.
To correlate activity and concentration units for solutes, one may choose
mole fraction, molality, molarity, or other units. To establish a suitable cor-
relation it is well to remember that solutes exhibit their simplest behavior in
very dilute solutions. Accordingly, we choose the infinitely dilute solution as the
correlation reference state for given solute and solvent at each temperature. In
practical terms we say that a, > X, as X,— 0, or a, > mas m-— 0 and the
like. In terms of f and a, Henry’s law is exactly
disk a; (8.27)
at all concentrations, where k is some constant. As a frequently admissible
approximation
P, = kya, (8.28)
or
Po ka (8.29)
where the values of the proportionality constants depend upon whether a,
and X, or a, and mare correlated. To the same degree of approximation
208 Equilibria in Solutions of Nonelectrolytes Chap. 8
Pa eX (8.30)
or
P, = Kpm (8.31)
within the range of validity of Henry’s law. That is, a = X or a = m in the limit
of very dilute solutions for which ky or k,, can be evaluated. These experimental
constants, in turn, lead to values of activities by equations (8.28) and (8.29) and
measured pressures.
The system ethanol-chloroform, for which data appear in Table 8.3, is
rather nonideal and may be used to illustrate some of the methods of calculating
activities. It should be observed that as Xo,n,o, approaches unity, do,n,on
= Pon,on/Pe.uon approaches the mole fraction. That is, Raoult’s law
applies to the solvent and do,u,on = Xc.n,on- In the same range of composi-
tion, but for the minor component, Poya,/Xcuca, 1S approximately constant.
The solute nearly obeys Henry’s law. The limiting value of P/ X can be estimated
graphically, by extrapolating to the infinitely dilute solution (Fig. 8.3), to be
428. At any composition, therefore, douq, = Pouc,/428. As might be expected,
values of doyq, and Xoucq, correspond fairly well in dilute solutions but differ
considerably at higher concentrations. On this same basis dgym, at Xcoua, = |
is 0.690, whereas the alternative choice (a — X,
X — 1) would give dcouq, = 1. The choice of
FIG. 8.3. Extrapolation of : ; , ;
ee eat is correlation reference state is arbitrary. In either
infinitelyadilitte ssolotion, case activity is proportional to the partial vapor
pressure.
It is often convenient to resolve a_ther-
modynamic function of activities into the cor-
500
responding function of the concentrations and
another function which corrects for nonideality.
The simplest correlation of activity and com-
position is one of the form
475
Ge (8.32)
where ¥ is the defined activity coefficient. Analo-
gous relations apply to other composition units.
This same device has been employed previously
P/X
450 for converting pressure to fugacity.
EXERCISE 8.9
Evaluate dcouc, at 35° from the data of Table
8.3 at Xouci, = 0.0400 for the correlation
reference state Xoum, = 1.
425
Ans. acucl, = 0.968.
EXERCISE 8.10
Show that the difference in chemical potential
of a solute between solutions at X$, ai and X%’,
400
0.01 0.02 0.03 0.04 0.05 0.06 ay’ is given by AF = ly! — Jy = RT In ay//a.
XCHCI;
Sec. 8.7 Equilibria in Solutions of Nonelectrolytes 209
TABLE 8.3
VAPOR PRESSURES OF CHLOROFORM AND ETHANOL AT 35°
0 0 295.11 0.690
0.0384 17.81 286.10 0.1733 297.5 0.668
0.0400 18.13 285.56 0.1764 297.5 0.667
0.0414 18.69 285.48 0.1818 297.8 0.667
0.0440 19.42 285.45 0.1889 298.6 0.667
The use of the property activity, rather than one of the more direct
measures of composition such as mole fraction, molality, or molarity, permits
the redevelopment of the equilibrium function in a form precisely applicable
to real solutions. Since the equation for the free energy of dilution in terms of
activities has exactly the same mathematical form as that applying to ideal
solutions [Exercise 8.10 and equations (8.6), (8.7)], the development is precisely
the same as that for equation (8.19).
The generalized chemical change is related to the standard change by the
isothermal cycle
aA(a,) + ObB(as)+... Beate mM (ay) + m”mN(ay)+...
AF
rp = | aRT In (1/aa) sr erin armen ste ert
and AF®, is the standard free energy change, that is, for unit activities of re-
actants and products. Q, is the reaction function of activities, each of which
can be chosen arbitrarily. When the activities of reactants and products are
those corresponding to an equilibrium state, AF, = 0.
0=AF?Z+
RT In K,
210 = Equilibria in Solutions of Nonelectrolytes Chap. 8
or
AF? = — RT ink, (8.34)
where K, is the value of the equilibrium function of activities.
K, = (hat
Ay Ag...
-)/equil. (8.35)
The activity of each substance factors into a concentration and an activity
coefficient, for example a = Xv, and the activity coefficients are collected sepa-
rately.
Be Se. en Hi ono
oe (8.36)
For convenience let
K, = Ky K,
Since it is K,, not K;, which is related to the standard free energy change
by equation (8.34), it is K, which is the equilibrium constant, having a unique
value for a given system and temperature. The value of K, is expected to depend
on the composition of the system, since the activity coefficients vary with com-
position. For example, in the esterification equilibrium, treated as an ideal
system in section 8.4, if various mixtures of acetic acid, alcohol, and water are
equilibrated at 25°, and the value of K, = K, computed for the equilibrium
composition, the results shown in Table 8.4 are obtained. The variation in K,
is to be attributed to variations in the activity coefficients with the composition
(equilibrium) of the system. The value of K, and hence K,, is also affected by the
presence of salts such as sodium bromide and sodium chloride which do not
appear in the chemical equation.
TABLE 8.4
THE ESTERIFICATION EQUILIBRIUM
1 1 0 3.79
3 1 0 4.73
1 3 0 2.45
1 1 28 3.56
SUMMARY, CHAPTER 8
1. Free energy of dilution
A x AEE Ps
AF, = nRTIn S
Standard states X¥= 1, c = 1 may be hypothetical.
2. Solubility
AF? (soln.) = —RT In ¢ (sat.)
aes
Cyne OR
ieee
This
JE,
4. Chemical equilibrium
AF, = AF} + RT In Q
AF?, = —RT1|n K
K= GG F* (dilute solution)
ae ee
In & —_ AH (2 =e *)
TERE ORS eT;
5. Activity
dj-= RI dina
w= p+ RTIna
PROBLEMS, CHAPTER 8
1. Succinic acid, C,H,O,, is practically undissociated in its saturated aqueous
solution. Its solubility given in grams per 100 g. of water is
Temp. (C) 0° 12:5: DS Bihon 50°
Solubility DS 4.92 8.35 14.00 21.40
Plot log solubility versus 1/T and find AA (soln.). Ans. 7.3 kcal.
2. The solubility of succinimide hydrate, C,H;O;N, in alcohol is 3.4 g./100 g.
at 10° and 6.6 g./100 g. at 25°. Find AH, AF, and AS at 25° for the solution of
1 mole of this substance in a large amount of solution containing 1.0 g./100 g.
Ans. AH = 7.4 kcal., AF = —1.15 kcal., AS = 28.7 cal. deg.~!.
and the solubility of I, in CCl,, which is 20 g./1000 g. at 25°, find AF%Qs for
layer. If the same amount of benzoic acid is distributed between 5 1. of water and
5 1. of benzene, what per cent will remain in the water layer?
9. For the reaction
C.Hio ar I, = CeHiols
16. From the result of the preceding problem, show that the entropy of mixing
for a binary solution which obeys Raoult’s law is given by
ASmixing = —X, Rin X, — Xz Rin Xz
The Gibbs phase rule can now be formulated as follows: Consider a system
of p phases (A, B, C, etc.) and c components (1, 2, 3, etc.) in an equilibrium state,
as indicated in Fig. 9.1. To assess the number of independent variables, we begin
by specifying the concentration of each component in each phase, p X c
quantities. In addition, we specify the temperature and pressure of the system
as a whole, these properties being uniform throughout in an equilibrium state.
The total number of variables v specified is, therefore,
Ope (9.1)
However, these are not all independent variables. In
each phase containing c components it is necessary
Components to specify only the concentration of c — 1 com-
ponents, since only one possible concentration of
the remaining component is consistent with this
specification. For example, if the composition of the
phases is given in mole fraction units, the sum
Mia As ewe ce. == be. Therefore. we reduce tne
number of variables by one for each phase, obtaining
v=pe+2—p (9.2)
Next, when the concentrations of the components in
the various phases, Aq... \Gie. qos oe Fepresemt an
equilibrium distribution, then a specification of X,,
automatically determines X,,, X,,;, etc. (see section
FIG. 9.1. Phase equilibria. § 3), Therefore, for each one of the c components,
p —1 composition variables are not independent,
and we obtain
rhombic
P (atm.)
3x1075
(Ome
97 114 154
if (EC)
9.3. Again, the scale of values on the pressure and temperature axes is nonlinear.
Two crystal forms (rhombic and monoclinic) of solid sulfur are observed, the
former being the stable form at room temperature.
The curves again describe two-phase equilibria in this one-component
system and, therefore, describe systems with one degree of freedom, by ap-
plication of equation (9.4). Curve CD describes the vapor pressure of liquid
sulfur as a function of temperature, curve BC the vapor pressure of monoclinic
sulfur, and curve AB the vapor pressure of rhombic sulfur. These three curves
can be quantitatively described by the Clausius-Clapeyron equation (7.9).
Curve CE describes the equilibrium between liquid and monoclinic crystals,
that is, the variation of the melting point of monoclinic sulfur with pressure,
according to equation (7.50). At pressures above 1288 atm. the monoclinic
form is no longer stable, and the rhombic form melts directly to the liquid, an
equilibrium described by curve EF. At pressures below 1288 atm. the rhombic
form equilibrates with the monoclinic form at temperatures and pressures
described by curve BE. In this case the curve describes the equilibrium between
two solid phases. In such a change the equilibrium state is attained rather
slowly, and it is possible to superheat the rhombic form and observe its vapor
pressure along curve BG and its melting point along curve GE. The slow at-
tainment of the equilibrium state in the rhombic-monoclinic transformation
is responsible for the observation of an indefinite melting point of sulfur.
When two solid phases occur in a one-component system, four triple points
can arise: s’-s’’-1; s’-s’’-v; s'-1-v; s’’-l-v. The phase rule does not predict whether
Sec. 9.3 Phase Diagrams 219
or not all triple points are experimentally accessible. In the present case experi-
ment shows that S,-l-v is metastable. The stable triple points are B, which
describes the equilibrium S,-S,,-v; the point C, which describes the equilibrium
Sm-l-v; and the point E, which describes the equilibrium S,,-S,-l. Each of these
points represents a one-component system containing three phases, and there-
fore each has no degrees of freedom. That is, there is only one temperature and
pressure at which each of these equilibria can be established.
EXERCISE 9.1
From the slopes of the curves in Fig. 9.3 establish the relative densities of mono-
clinic, rhombic, and liquid sulfur by application of equation (7.50).
Ans. py < Pm < pr.
liquid
liquid
phase at equilibrium, while the lower curve relates the pressure and the composi-
tion of the vapor at equilibrium. That is, any pair of points on the two curves
connected by a horizontal line represent the compositions of the liquid and
vapor at equilibrium at the chosen pressure. For an ideal system, such pairs of
points obey Raoult’s law. For a real system, this relationship must be determined
by experiment, but in either case it is unique. That is, for a liquid of given com-
position there is only one possible vapor composition at equilibrium. The gross
composition of the system will vary with the relative amounts of the two phases
and is, therefore, not an intensive property within our present usage. In Fig.
9.4(b) the same system is described at a fixed pressure, showing the relation
between vapor-liquid equilibrium temperature and composition, and again, any
pair of points on the two curves connected by a horizontal line represent the
equilibrium compositions of the liquid and vapor at the chosen temperature.
For two-component systems, the phase rule shows that when two phases
are present there are two degrees of freedom (f= 2 — 2 + 2 = 2). In Fig. 9.4
one variable has already been fixed, the temperature in (a) and the pressure in
(b). There remains, in either instance, one variable subject to arbitrary choice
if the system is to contain two phases at equilibrium. In (a), for a given choice
of liquid and corresponding vapor composition, there is only one equilibrium
pressure. In (b) when the composition is chosen, the equilibrium temperature is
established. Conversely, if in case (a) a pressure is selected (within the range
P*,to P%), the compositions of the two phases are determined, and if in case (b)
a temperature is selected (within the range 7%, to 7%), the compositions of the
two phases are again fixed.
Many two-component vapor-liquid systems deviate considerably from the
ideal behavior described in Fig. 9.4. Two extreme cases are shown in Figs. 9.5
and 9.6. For each system (a) is a pressure-composition diagram at constant
400 liquid =
i . 80
300 4
vapor | |7
P (mm) Lp.
500 ‘] vapor (°C)
- 470
1a 1 L liquid
q 65
400 70
100 55
(0)
(6) 05 10 (0) 05 1.0
Mole fraction chloroform Mole fraction chloroform
If a still pot is charged with a mixture containing 25 mole per cent of benzene in
toluene, estimate the composition of the vapor at the head of a column with three
theoretical plates. (Assume that the composition of the material in the still pot
does not change.) AN SeeXGensenet 04.
1 phase 200
50 al 150
T (°C) |
2 phases
100
ie) 1 eee O 50
O 20 40 60 80 100 O 20 40 60 80 100 O 20 40 60 80 100°
Wt. % phenol Wt. % triethylamine Wt. % nicotine
in water in water in water
mation can be expressed in the form of phase diagrams such as those in Fig. 9.8.
The information shown refers to a constant pressure of | atm.
The curves of Fig. 9.8 describe the limits of solubility, and the region en-
closed is called a two-phase region, since systems of gross composition and
temperature such as those represented by the points A will form two phases of
compositions given by points A, and A,, the intersections of a horizontal line
with the solubility curve. For example, at 30° a mixture of 60% phenol with
water will yield two liquid layers, one containing 9% phenol and the other
70% phenol.
EXERCISE 9.4
Sixty grams of phenol is mixed with 40 g. of water at 30°. How many grams of
each liquid phase will be formed? Ans. 16.4 g.9% phenol; 83.6 g. 70% phenol.
Figure 9.8 shows that the mutual solubilities of phenol and water increase
with increasing temperature and that the two liquids become miscible in all
proportions at 66°. This is called the upper consolute temperature. The system
triethylamine-water shows the opposite (and less common) behavior, an increase
in mutual solubilities with descreasing temperature. This system has a lower
consolute temperature of 18.5°. The system nicotine-water is quite unusual in
that it has both an upper and a lower consolute temperature.
We shall begin discussion of the vaporization equilibria involving two-
phase liquid systems by considering the extreme case of an immiscible liquid
pair, that is, a liquid pair whose mutual solubilities are so small that the vapor
pressures of the two mutually saturated liquid phases are practically those of
the pure substances. In such a case, the total vapor pressure over the two-phase
liquid system is simply equal to the sum of the vapor pressures of the two pure
substances. The presence of a second liquid phase does not affect the vaporiza-
tion equilibrium of the first.
It is evident that the sum of the two vapor pressures will reach that of the
atmosphere at a temperature below the boiling point of either pure liquid, and,
Sec. 9.5 Phase Diagrams 225
therefore, the boiling point of the two-phase system is less than that of either
component. This behavior is the basis for the technique of steam distillation in
which an organic substance immiscible with water may be distilled at a temper-
ature less than 100°, even though its own boiling point may be considerably
higher. This is often advantageous in purification without the thermal decom-
position which might occur in simple distillation of the pure substance.
Partially miscible liquids illustrate an extreme case of deviation from ideali-
ty, and for those ranges of concentrations in which a homogeneous liquid phase
is formed the phase diagram has the usual appearance, as seen in the areas out-
side the vertical lines in Fig. 9.9. In Fig. 9.9 (a), showing equilibrium pressure as
a function of composition, the curve AB describes the composition of the liquid
in equilibrium with vapor of composition given by the point at the same pressure
on curve AC. In Fig. 9.9(b), showing the equilibrium temperature as a function
of composition at | atm., the curve AB describes the equilibrium vaporization
temperature of liquid solutions of isobutanol in water, and the composition of
the vapor formed at a given temperature is given by the curve AC. The solu-
bility of isobutanol in water varies slightly with temperature, as indicated by the
curve DB. The point B represents the composition of a saturated solution of
isobutanol in water at the equilibrium vaporization temperature, and the point C
represents the composition of the vapor in equilibrium with this solution.
Figure 9.9 describes the course of events when a liquid mixture of isobutanol
and water is warmed. For any mixture containing less than 8.5% or more than
78% isobutanol, the liquid phase is homogeneous with boiling points given by
curves AB and EF, respectively. The composition of the vapor in equilibrium
with these liquids at the boiling points is given by curves AC and CF, respective-
ly. However, a system of gross composition between 8.5% and 77% isobutanol
will have two liquid layers, and regardless of their relative amounts, so long as
two liquids are present, boiling will occur at 89° and 1 atm. The composition
of the vapor in equilibrium with the two liquid phases is given by point C, 67%
isobutanol. According to the phase rule, at this point f= 2 — 3 + 2 = 1. Since
the pressure has been arbitrarily selected, no degrees of freedom remain, and
a constant boiling point will be observed until one of the liquid phases dis-
appears. For example, if the gross composition of the system is 50% isobutanol,
400
P (mm)
as indicated by the vertical dashed line, the 77% butanol liquid phase will dis-
appear first, leaving the 8.5% butanol liquid phase. At this time only two phases
remain, one liquid and one vapor, and the equilibrium temperature is no longer
fixed. Further boiling results in an increase in temperature with change in the
liquid composition along line BA until, at approximately 94°, the last of the
liquid disappears.
EXERCISE 9.5
From the data given in Fig. 9.9 (a) describe the number and nature of the phases
present at various temperatures as a mixture of 50% isobutanol and water vapor
is compressed at 75°.
Ans. First liquid, 7°% isobutanol forms at 375 mm.; 2 liquid phases, 8.5% isobu-
tanol, 77% isobutanol form at 400 mm. Above 400 mm., no vapor phase.
1As shown previously, solid-liquid equilibria are insensitive to pressure changes, and only
large variations have an appreciable effect on the temperature-composition relation.
228 Phase Diagrams Chap. 9
solid solution contains 5% Bi. Further cooling produces more of this solid
phase containing increasing proportions of Bi, until at ca. 220° complete
solidification has occurred. At no time are more than two phases present,
and therefore no halt is observed. The eutectic halt will be observed only
for systems containing between 37% and 97.3% Bi, and the two phases of the
eutectic mixture are the two solid solutions rather than the pure substances.
The phase diagram for the system Pb-Bi shows that the solubility of Bi in
Pb changes with temperature and that solid solutions containing 18-37% Bi
become supersaturated at low temperatures. A slow transformation to a more
dilute solid solution of Bi in Pb with the formation of a second phase should
occur and should be microscopically recognizable.
A
two liquid phases are formed from that
in which a single homogeneous liquid
phase is formed. Although phenol and
water at this temperature and pressure are
ANSI
only partly miscible, the addition of ace-
tone increases their mutual solubility until,
AO
are miscible. (Very small additions of ace-
tone actually diminish the mutual solu-
AAV AAV IN
So
bilities of phenol and water to a slight
Phenol ALlater
SCCwI 7) Phase Diagrams 231
constant temperature is obtained only when four phases are present, in this case
three solids and one liquid, for according to the phase rule f= 3 —4 4 2= 1,
and the pressure has been arbitrarily fixed.
SUMMARY, CHAPTER 9
I. The Gibbs phase rule: f=c—p+2
2: One-component systems: f =3— p
One phase. =
Two phases, f= |
Three phases, triple point, f= 0
3: Two-component vapor-liquid systems: f= 4— p
For vapor-one liquid equilibria, constant T or P, f = 1
For vapor-two liquid equilibria, constant T or P, f = 0
4. Solid-liquid equilibria
Two components, pressure fixed: f= 3 — p
Liquid, one solid, f= 1
Liquid, two solids; eutectic, peritectic, f = 0
Three components, pressure fixed, f= 4 — p
Liquid, two solids, f= 1
Liquid, three solids; ternary eutectic, f = 0
PROBLEMS, CHAPTER 9
1. In the equilibrium
CaCO, (s) == CaO (s) + CO, (g)
we have seen that at each temperature there is a characteristic dissociation pressure
P =K. Does the system consist of pure crystals of CaCO, mixed with pure
crystals of CaO, or does it contain crystals in which CaCO, and CaO are molecu-
larly dispersed (a solid solution)? Only one of these alternatives is thermody-
namically allowed. Explain.
2. From the data given below construct a graph of liquid and vapor composition
versus boiling point and describe the result of (a) free boiling and (b) fractional
distillation of a mixture containing 50 mole per cent of carbon tetrachloride in
ethyl alcohol.
B.p. (°C) TS 12 C30 GO Go Ch Do
Mole per cent eeu 0 6.4 17.6 33.6 63 72.8 100
CCl, in C,H;,0H Jvapor 0 DS 45 55 63 67 100
3. Use the data given below for the liquid-vapor system chloroform in benzene
to estimate the composition of the first distillate when a large quantity of 50%
chloroform in benzene is distilled in a fractionating column of 4 theoretical
plates.
(ee) SOLO 90S E35 5:3 4 Ol 7196S Olmo ted
Mole per cent eee 0 15 29 44 54 66 79 100
CHCl, in C,Hg)/vapor 0 20 40 60 70 80 90 100
Ses / Phase Diagrams 233
4. At 20° the mutual solubility of methyl ethyl ketone and water is such that
one layer contains 22.6 wt. % and the second contains 90.1 wt. °%% methyl ethyl
ketone. How much of each layer will be formed when 50 g. of water and 50 g. of
methyl ethyl ketone are mixed?
5. From the data given below for the vapor-liquid system furfural-water con-
struct a phase diagram showing vapor and liquid composition at the boiling point
and also the mutual solubility of the two liquids below the boiling point.
B.p. (C) LOOPS Ss 979 TO esLOOL6 122255 S55 162
Mole per “es liquid 0 2 O20 80 92 96 100
furfural vapor 0 8 i OD 11 S2 81 100
Temp. (C) 90 =680 60
Mutual rsa |
Mole per cent L; 345 3.0) 2.2
furfural Ly 55 60 67
Describe the number and nature of phases present when a mixture containing
60 mole per cent furfural is warmed.
6. Obtain an expression for the composition of the vapor X{ as a function of
the composition of the liquid X, at constant temperature in a two-component
system forming an ideal solution.
7. From the following set of cooling curves for the system PbSO,-K,SO, con-
struct the phase diagram for the system. The melting points of the pure sub-
stances are 1070° and 1080°C, respectively. What is the formula of the compound
formed?
Mole % PbSO,
60 66.7 70 90
1100
1000
iP (6)
900
800
8. Using Figs. 9.12, 9.13, and 9.14, sketch the cooling curves for melts contain-
ing (a) 10 wt. % Mgin Zn. (b) 20 wt. % Mgin Zn. (c) 20 wt. % K in Na.
(d) 70 wt. % Kin Na. (e) 20 wt. % Biin Pb. (f)40 wt. % Biin Pb
234 Phase Diagrams Chap. 9
9. Magnesium melts at 651° and nickel at 1450°. These elements form two
compounds in the solid state, Mg,Ni, which shows an incongruent melting point
at 770° yielding liquid containing 38 wt. % Ni and the compound MgNi,. The
latter compound melts at 1180°. Two eutectics are found; one at 510° and 28%
Ni, the other at 1080° and 88% Ni. Use this information to sketch the phase
diagram.
10. The following phase diagram describes the system silver-tin. What is the
formula of the compound formed? Construct cooling curves for melts containing
30, 60, and 90% silver.
solid
solution |
11. The data given below are the weight per cent compositions of the two
layers formed in various mixtures of water, ether, and methanol. Use the data to
construct the ternary solubility diagram, indicating the two-phase region.
Wt. % methanol 0) 10 20 30
Ly 93 82 70 45
Wt. % water te 1 6 155 40
IN
ELECTROCHEMISTRY
In 1887 Svante Arrhenius proposed that acids, bases, and salts disso-
ciate in aqueous solutions to form positive and negative ions. Upon appli-
cation of an electric field, conduction of electric current in such electrolyte
solutions takes place through the migration of these ions, rather than by migra-
tion of electrons as in the metals.
It is the nature of electrolytic conduction that electric charge is transferred
between ions in solution and the electrodes which are immersed in the solution
and connect with the external circuit. When negative electric charge enters the
solution, it effects chemical reactions of reduction. The electrode at which reduc-
tion occurs is the cathode, and the positively charged ions which migrate toward
it are cations. At the other electrode, the anode, reactions of oxidation provide
a mechanism for transferring electrons from the solution to the metallic circuit.
The negative ions which migrate toward this electrode are anions. For example,
if two metallic silver electrodes are dipped into a solution of silver nitrate, as
shown in Fig. 10.1, and an electric current is passed through the system, the
following processes occur: Current is carried through the electrolyte by the
migration of anions (NO;) toward the anode and cations (Ag*) toward the
cathode. At the cathode silver ions are discharged by acquisition of an electron
from the cathode, and deposit on the cathode as metallic silver. At the anode
silver enters the solution as silver ions, liberating electrons to the external circuit.
The processes which occur at electrodes depend on the sign of the electrode,
its composition, the nature and concentrations of the ions in solution, and the
235
236 Electrochemistry Chap. 10
a-c source
an electrolytic cell necessarily involves Avogadro’s number
of electrons in reduction at the cathode and a like number
in oxidation at the anode. An ion with a positive valence
number of 2, such as Cut*, requires two units of electrons
for reduction to the metal, and therefore one unit of electrons
will reduce | gram-equivalent weight, or 4 gram-atomic
weight. The experimental determination of the faraday, which
is equal to 96,487 coulombs, may be combined with the experi-
mental value of the charge on the electron (see Chapter 13),
1.60210 x 107!® coulomb, to obtain a value of Avogadro’s
number.
96,487
= 6.0225 <x 107
16021051077?
This is the number of electronic charges in 1 faraday and, FIG. 10.2. Wheat-
therefore, the number of atoms in | gram-atomic weight. li ie
10.2 Conductance
ps 74 (10.2)
TABLE 10.1
EQUIVALENT CONDUCTANCES OF STRONG ELECTROLYTES
IN AQUEOUS SOLUTION AT INFINITE DILUTION AND 25°*
* From G.F.A. Kortum and J.O’M. Bockris, Textbook of Electrochemistry, Elsevier Press,
Inc., New York, 1951.
we expect that
A = aX (10.7)
This relation is valid at vanishingly small ion concentrations and a useful first
approximation in fairly dilute solutions. The value of A, for a weak electrolyte
such as acetic acid may be obtained by application of Kohlrausch’s law to
measurements of A, for strong electrolytes involving the same ions. For example,
since for acetic acid
A,(HOAc) = 2,(H*) + A,.(OAc-)
its value may be obtained from
A,.(HCl) + A,(NaOAc) — A,(NaCl)
Using values from Table 10.1, we obtain
A.(HOAc) = 426.16 + 91.0 — 126.45
= 390.) mim eg--*em,"
Comparison of this calculated value of A,(HOAc) with the observed value of
A(HOAc) at some finite concentration yields the degree of dissociation a of
the acid. For example, the observed value of A(HOAc) at 0.01 M is 16.3 ohm™!
eqd.s cms heretore:
21032
= ase a 0.0417
EXERCISE 10.4
Combine the values of A, for a strong acid, a strong base, and the corresponding
salt from Table 10.1 to compute the sum
A(H*) + A(OH-)
Compare this value with the observed conductance of pure water (L = 6.2 x 107
ohm-'cm.~!) to obtain a value for the degree of dissociation of water.
Ansoe— 2 0alOm:
os 2 10.8
ry
=>
nN, i + pan n_ ie = A ( )
indicated. We may imagine that these two electrolytes are a strong acid and
its salt. Then the position of the boundary can be observed with an acid-base
indicator or by the change of refractive index of the solution at the boundary.
An electric field is applied to the cell in such a sense that the faster ion (H*)
precedes the slower ion (M*) toward the cathode. Under these circumstances the
boundary will remain sharply defined, because any tendency of the faster ion
(H*) to outrun the slower ion (M*) will reduce the concentration of electrolyte
at the boundary, thus increasing the resistance and the potential gradient along
the tube in that region. This will tend to speed up the slower ion (M*). On the
other hand, if the slower ions (M*) diffuse ahead of the boundary they enter a
region of lower resistance and lower potential gradient and, therefore, tend to
slow down. Thus, the boundary formed with the ion of higher mobility preceding
the one of lower mobility remains sharply defined.
A steady current / is passed through the solution for ¢t seconds, causing the
boundary to migrate toward the cathode, sweeping out a volume V liters. This
is accompanied by the migration of ¥c,V coulombs of positive charge past an
arbitrary reference plane in the acid solution. The transference number of the
cation (H*) in this solution is obtained from
n, ACif (10.9)
and
n_=1—n,
Table 10.2 shows that the transference number is not a strong function of
concentration.
TABLE 10.2
CATION TRANSFERENCE NUMBERS IN AQUEOUS SOLUTION AT 25%
* From G.F.A. Kortum and J.O’M. Bockris, Textbook of Electrochemistry, Elsevier Press,
Inc., New York, 1951.
EXERCISE 10.5
Use data from Table 10.2 to find the linear velocity of H+ in 0.1 N HCl when
subjected to a potential gradient of 1 v. cm.~!. The specific conductance of 0.1 N
HCl is 0.0391 ohm=!cm.~!. Ans. 3.37 < 107? cm.sec.*!.
Sec. 10.5 Electrochemistry 243
Since X,/A =n,, the fraction of current carried by the cation, the trans-
ference numbers may be used to obtain values of equivalent ionic conductances.
NN
Vee (10.10)
TABLE 10.3
EQUIVALENT IONIC CONDUCTANCES IN AQUEOUS SOLUTION AT INFINITE
DILUTION AND 25°
* From G.F.A. Kortum and J.O’M. Bockris, Textbook of Electrochemistry, Elsevier Press,
Inc., New York, 1951.
(+) The factor + is used as a reminder that this is the conductance of | gram-eqguivalent.
Zu = Zn X(aq;)i-= 2e
and at the cathode it is
2e + Cutt (aq.) = Cu
The algebraic sum of the half cell reactions is the net cell reaction:
Zn + Cu*t (aq.) = Cu + Zn** (aq.)
It is convenient to describe galvanic cells or single electrodes by an abbrevi-
ated notation together with appropriate conventions concerning the relationship
between these descriptions and the corresponding chemical reactions. The most
important single consideration throughout is this: the galvanic cell is a device
for effecting electrochemical processes in a reversible manner and for measuring
the free energy change from the emf. Any conventions used for cells must be
compatible with other thermodynamic conventions. It will be necessary to inter-
convert the representation of a galvanic cell and a corresponding chemical
equation which describes the cell process. There is an important difference
between them, since a chemical equation connotes extensive properties, such
aeat of reaction, whereas the essential characteristic of a galvanic cell is its
emf, an intensive property. Accordingly, the representation of a galvanic cell
must clearly indicate substances, their states and concentrations, as well as any
other factor which may affect the emf. A common form of notation is now
described.
In this cell the junction of the two solutions is a source of emf as exemplified
by the fact that the observed potential depends to a small extent on the nature
of the anions present in the two solutions. The emf which develops at a
liquid junction corresponds to the free energy change of transferring the migrat-
ing ions from one environment to the other. This junction potential is often
diminished by use of a salt bridge, which consists of a tube filled with a solution
of potassium chloride connecting the two electrode compartments. This device
permits flow of current by migration of ions and reduces the net junction poten-
tial to a very small value. The salt bridge is indicated by two vertical lines, and
when it is used, the nature of the anions has practically no effect on the emf
and need not be specified, as in
Zn|Zn*? (aq., m,)||Cu*? (aq., m,)|Cu
Although cells with liquid junctions or salt bridges are commonly used for
practical measurements, it is preferable to use cells without liquid junctions for
thermodynamic measurements. A single electrolyte solution may furnish the
ions for both electrodes, as in the cell
Pe Ee ce) et (age 7m,), Cle (agai) WEG (P;). Pt
for which the cell reaction is found by adding the electrode reactions thus:
A galvanic cell consisting of two electrodes of the same type which differ
only in concentrations of reagents is called a concentration cell. For example,
in the cell
Pt, H, (P,)| H* (m,) ||H* (m,) |H2 (P2), Pt
with cell reaction
$ H, (P,) + H* (m,) = 3 Hy (P2) + Ht (mm)
an emf is developed only when P, 4 P, or m, #™m,. The liquid junctions in
this cell make it unsatisfactory for precise thermodynamic measurements,
and for such purposes the double concentration cell
Pia; CP) Hel (7,) ||Cl @.),-Pt-Pt. Cl. (Pe) VHC (n,)i(HaCe), Pt
where all gas pressures are equal and the emf depends only on the relative
values of m, and m,, may be used.
EXERCISE 10.8
Write the reaction occurring in the double concentration cell formulated above.
Ans. HCl (m,) = HCI (m,).
it follows that
AFy = — Winax, (€lect.) = —nFér (10.16)
248 Electrochemistry Chap. 10
EXERCISE 10.9
The emf at 25° for the cell
Pt, H, (1 atm.) |HCl (aq., 1 m)| Cl, (1 atm.) Pt
is 1.36 v. Use this information to compute the free energy of formation of hydro-
chloric acid in 1 M solution at 25°. Ans. —31.4 kcal.
(dAF;\
at) = —ASr.» (10.17)
Therefore, the derivative of é with respect to temperature (at constant pressure)
oer) ANS
(OL Je nF oe
yields a value of AS, for the cell reaction. Since
AH, Ss AF, ++ T AS,
EXERCISE 10.10
For the cell
Zn| ZnCl, (0.555 m) |AgCl, Ag
6573 = 1.015 v. and d&/8T = —0.000492 v. deg.-}
Calculate AF>73, AS273, and AH,;; for the cell reaction
Zn + 2 AgCl = ZnCl, (0.555 m) + 2 Ag
Ans. AFy73 = —46.8 kcal.; AS>7;, = —22.7 cal. deg.-!;
AH,73 = —53.0 kcal.
Although the cycle and equation (10.20) are not completely general, the exten-
sion to other types of reactions should be obvious. It is assumed that the gas
X, is ideal and that the free energy of the pure solid M is fixed. (To be precise,
the fugacity should be used instead of the pressure.)
As shown in Chapter 8, the activity a is defined in such a way as to make
equation (10.20) completely rigorous. The reference state used in the consider-
ation of electrolytic solutions is the infinitely dilute solution, and the concen-
tration unit is usually the molality m (approximately equal to the molarity M
in dilute aqueous solutions) so that
a—m as m—-0O0
3 In ee
whereas if m = 1, we have
Cyn ax
In
Pe
and the value of the second term of equation (10.23) is the same in either case.
This is a consequence of the fact that the emf is an intensive property, inde-
pendent of the amounts of materials involved. On the other hand, the free
energy change, AF, = —n¥é& >, represents a change in an extensive property,
and as such its value depends on n.
Methods for evaluating a and &° will be described later, but an approximate
application of equation (10.23) is possible when a = m, that is, for dilute solu-
tions. Consider again the cell
Pt, H,| HCl (aq.)|Cl,, Pt
When ayo) = 1 and Py, = Po, = 1, 1.e., for all reacting substances in their
standard states, the emf is &%3 = 1.36 v. For any other composition, such as
Pt, (?atm,) |H2 (10-272) Cl (0s my
iiChd0s2atm. are
we find by equation (10.23)
Secli0zs Electrochemistry 251
= 00592 me.mée-
Orxee = 1236 ee log Pepe
a= my:
For other types of electrolytes analogous definitions of the activity of the
electrolyte and the mean activity coefficient are used. For example, in the case
of a 2-1 electrolyte, M**?X;,
C= G,02— Migam
and
ye = (ys¥2)"”
252 Electrochemistry Chap. 10
so
a = 4m'y’.
The value of &° in equation (10.23) is determined by a procedure which takes
advantage of the fact that, by definition, a— m as m— 0. Emf measure-
ments are made at a series of lower concentrations, and the value of &° obtained
by extrapolation to infinitely dilute solution. For example, the emf of the cell
Pt, H, (1 atm.)| HCl (m)| AgCl (s), Ag
has been accurately determined [H. S. Harned and R. W. Ehlers, J. Am. Chem.
Soc., 54, 1350 (1932)] for various concentrations of HCl. The cell reaction is
H, (1 atm.) + 2 AgCl (s) = 2 H* (aq., m) + 2 Cl (aq., m) + 2 Ag (s)
and, therefore, the expression for the activity dependence of the emf is
RT
Oa = Ge ney We In Geader
0.2250
0.2200
V7
&o9g
0.1183
00602
log
+m—
Sey Oy Electrochemistry 253
TABLE 10.4
POTENTIALS OF STANDARD HALF CELLS*
6° = oxidation potential (emf), “~° = electrode potential
Electrode Reaction Cases IP So
Li (s) = Lit (aq.) +e 3.045 —3.045
K (s) = K+ (aq.) +e 2.925 —2.925
Na (s) = Nat (aq.) +e 2.714 —2.714
Mg (s) = Mgt? (aq.) + 2e 2.37 —2.37
Mn (s) = Mn?*? (aq.) + 2e 1.18 —1.18
Zn (s) = Znt? (aq.) + 2e 0.763 —0.763
Cr (s) = Crt? (aq.) + 2e 0.74 —0.74
Fe (s) = Fe*? (aq.) + 2e 0.440 —0.440
Crt? (aq.) = Crt? (aq.) te 0.41 —0.41
Cd (s) = Cd*? (aq.) + 2e 0.403 —0.403
Pb (s) + 2 1- (aq.) = PbI, (s) + 2e 0.365 —0.365
Ni (s) = Ni*? (aq.) + 2e 0.250 —0.250
Ag (s) + I- (aq.) = Agl (s) +e 0.151 —0.151
Sn (s) = Sn*? (aq.) + 2e 0.136 —0.136
Pb (s) = Pb*? (aq.) + 2e 0.126 —0.126
4H, (g) = H+ (aq.) +e 0.000 0.000
Tit3 (aq.) = Ti*4 (aq.) +e —0.04 0.04
Ag (s) + Br- (aq.) = AgBr (s) +e —0.071 0.071
Sn*? (aq.) = Sn*4 (aq.) + 2e —0.15 0.15
Hg (1) + Cl-(aq., 1 N) = 4 Hg,Cl, (s) + e (N calomel) —0.2802 0.2802
Cut (aq.) = Cut? (aq.) + e —0.153 0.153
Ag (s) + Cl-(aq.) = AgCl (s) +e —0.2224 0.2224
I-(aq.) =i]hL()+e —0.5355 0.5355
Fe*? (aq.) = Fet* (aq.) +e —0.771 0.771
4 CgHeO, (aq.) = + CgH.O> (aq.) + H* (aq.) + e (quinhydrone) —0.6996 0.6996
Hg (1) = + Hef? (aq.) + —0.789 0.789
Ag (s) = Ag* (aq.) +e —0.7991 0.7991
Br- (aq.) = + Br. (1) +e —1.0652 1.0652
2 Crt8 (aq.) + 7 H,O = Cr,O,;= (aq.) + 14 H+ (aq.) + 6e —1.33 133
Cl- (aq.) = + Cl, (2g) +e —1.3595 1.3595
Mn*? (aq.) + 4H,O = MnO; (aq.) + 8 Ht (aq.) + Se —1.51 151
Cet’ (aq.) = Cet4 (aq.) + e —1.61 1.61
Cot? (aq.) = Cot? (aq.) + e —1.82 1.82
In an actual cell the sign of &..., and of AF (reaction) depend both upon the
values of &°.,, and of the activities of all reacting substances. For example,
let us compute the emf of the cell
Pike 2(¢ — 10" are (a= il Hs. (ge is
for which the cell reaction is
2 He** =} Hg? = 2 He - 2 Fe**
Aire? Ayg+2
It has already been shown that the measured emf of a galvanic cell
at several temperatures gives important information about the thermodynamics
of the corresponding cell reaction. The AF for the cell reaction is directly pro-
portional to the emf and AS is directly proportional to its temperature deriva-
tive. Likewise, AF° for the cell reaction is directly proportional to €&°. Now, since
AF. = —RTInK (10.29)
it follows that
and
IK == 5) S< Oe
EXERCISE 10.14
From data in Table 10.4 estimate the equilibrium constant in the oxidation of
Cet? to Ce*4 by Cl. Ans. 5.85 X 107* (moles kg.~! atm.~!”).
258 Electrochemistry Chap. 10
Galvanic cells are often used to determine the activities of ions in aqueous
solutions, particularly of hydrogen ion activity. This is usually given in terms of
the pH scale,
pH = — log ay. (10.31)
Any one of several cells in which the emf depends on the hydrogen ion
activity may be used. The most obvious is a cell using a hydrogen electrode.
The other electrode may be any convenient one, and for this purpose the
normal calomel electrode is often used. The electrode potential of the latter is
included in Table 10.4. The value given there is not the standard electrode
potential, but rather the potential of the electrode in which the potassium chloride
concentration is | normal. The potassium chloride solution can also serve as
a salt bridge. The complete cell is
Pt, H, (1 atm.)| H* (ay) ||KCI (1 V)| Hg,Cl,, Hg
normal calomel electrode
If the hydrogen pressure in the cell described above is 0.9 atm. instead of 1 atm.
what error in pH will result? Ans. 0.023 pH units.
Another electrode often used for hydrogen ion measurements is the quin-
hydrone electrode. This consists of an inert metal such as platinum in contact
with a saturated solution of quinhydrone. The latter is a compound consisting
of equimolar portions of quinone, C,H,O,, and hydroquinone, C,H,(OH),.
It may also be regarded as the dimer of semiquinone, C,H,O,H, which in solu-
tion establishes the equilibrium
(C,H,0,H), = C,H,0, an C,H,(OH),
Sec. 10.10 Electrochemistry 259
The function a,-@,- cannot be resolved, and ay. is undefined, because it cannot
be measured. To the extent that the difference between yy-y.,- and unity can
be ignored, my: serves to measure acidity for dilute aqueous solutions of fully
ionized acids.
Galvanic cells may also be used to follow the course of many kinds of
oxidation-reduction titrations. For example, in the oxidation of ferrous ion by
ceric ion the ferrous-ferric electrode, in conjunction with a reference electrode,
Pt|Fe*?, Fet$||KCI(1 N)|Hg.Cl,, Hg
260 Electrochemistry Chap. 10
and so forth. The variation of the emf of the cell with the extent of oxidation
of the ion is shown in Fig. 10.8. For points corresponding to oxidation of Fe*?,
the cell is to be regarded as though the anode is Ce**, Ce*‘, and its emf will
be given by
6 = 6°(Ce*, Ce**) — 0.0592 log ae — & (Ncalomel)
Cet3
For example, when a 10% excess of Ce** has been added, doen = fodces and
& = —1.61 — 0.0592 log (1/10) — (—0.280) = —1.27 v.
Note in Fig. 10.8 that the most rapid change in emf versus reagent added
occurs at the equivalence point. This may be emphasized by plotting A@/A%
Sec. 10.11 Electrochemistry 261
as shown by the dashed line. This function exhibits a sharp maximum at the
equivalence point.
EXERCISE 10.16
Show that the cell
Pt| Ce*?, Cet ||KCI (1 N)|Hg.Cl, Hg
has the same emf as the cell
Pt| Fet?, Fet* ||KCI (1 N)|Hg,Cl,, Hg,
when the activities of Ce**, Ce**, Fe*? and Fe** have their equilibrium values
related by
Ke ed Aer: CeAcers
per? Acers
where m; is the concentration of each ion present in the solution and Z; is the
charge on that ion. The correlation between activity coefficient and ionic strength
becomes increasingly accurate with increasing dilution.
It was pointed out in connection with Fig. 10.7 that the curves of activity
coefficient versus concentration for electrolytes of a given charge type form a
family which converges with decreasing concentration. For example, among
the 1-1 electrolytes at 0.1 m some mean activity coefficients are 0.78 for NaCl,
0.77 for KCl. The curves are practically indistinguishable at lower concentra-
tions, but above 0.1 m the differences become increasingly significant.
The ionic strength is a property of the solution as a whole and at constant
low ionic strength the activity coefficient of a given electrolyte is found to have
a nearly constant value regardless of the actual composition of the solution.
For example, the mean activity coefficient of sodium chloride 0.001 m in NaCl
is 0.966, but in a solution 0.001 m in NaCl and 0.099 m in KBr, jp = 0.1 and
we may expect the mean activity coefficient of sodium chloride to be approxi-
mately 0.78. (See Fig. 10.7.)
The activity coefficient of an electrolyte also depends on its charge type.
For example, in the case of a 2-1 electrolyte, for which
262 Electrochemistry Chap. 10
m=m,=4m_
b= 2? 2 non
an ionic strength of 0.1 is achieved at a concentration of 0.033 m. Figure 10.7
shows that the activity coefficient of calcium chloride at this concentration is
approximately 0.60. By the same token, we may predict that the activity coef-
ficient of magnesium bromide in 0.033 m solution will be approximately the
same. Furthermore, in a solution containing 0.01 m barium bromide and 0.07 m
sodium chloride the ionic strength is
x [2 + Tie x 12 =| Me- xX 1?)
= S (tga x 22 +- Mp,-
Ye = (yoy) (10.40)
Substitution of equation (10.39) in (10.40) yields
—log y. = 0.509 Z,.Z_/ (10.41)
The validity of this equation may be tested by plotting —logy. versus \/j1
which should give a straight line of slope 0.51 for a 1-1 electrolyte, a slope
0.51 x 2 for a 2-1 electrolyte and so on. Data which can be subjected to this
test will be given in the next section.
EXERCISE 10.17
Estimate the mean activity coefficient of barium bromide and of sodium chloride
in a solution of ionic strength 0.1. Compare with data given above.
Ans. y+ (BaBr,) = 0.48; y. (NaCl) = 0.69.
(s), which is the standard state, its activity is unity and does not appear in the
equilibrium function.
Factoring the activities
Ky = mim yee, (10.45)
and substituting the mean activity coefficient, we have
K, = mim? (ys)°”’ (10.46)
For example, in the case of silver chloride
Estimate the solubility of silver chloride in 0.1 m KCI at 20°, (a) assuming yi =
1, and (b) using ys = 0.78. Ans. (a) 10 x 105? 77, (b) 1-81 x 10s? 7.
HA
+ A” = HA’
+ A-
for which the equilibrium constant is Ky,/Ky,, and does not involve the basicity
of B. In aqueous media water is the common proton acceptor and for strong
acids the equilibrium
HA -- H,O= H,0* ——A™
lies immeasurably far to the right. In such cases Ky, cannot be measured, and
strong acids cannot be compared in terms of Ky,/Ky,,, since neither constant,
nor their ratio, can be measured. The only strong acid species present is H,O*,
the hydronium ion. The species H,O* is known in the attenuated gaseous state,
and the reaction Ht + H,O = H,O* is exothermic by approximately 170 kcal.
Other gaseous species H* (H,O), with n = 2 to 5 also exist, but the energy
of binding each additional H,O decreases very rapidly and the value of n is
unknown for aqueous H*(H,O),. The symbol H* is used for the solvated proton
unless the proton-accepting role of the solvent B is to be explicitly shown, when
the proton is represented by HB*, and so forth.
The extent to which an acid HA undergoes proton donation to the solvent
B depends upon the energetics of several processes:
HA(g) = H(g) + A(g) bond dissociation energy
H(g) = H*(g) +e ionization of H
A(g) +e = A-(g) electron affinity of A
H*(g) + B(g) = HB*(g) proton affinity of B
Sec. 10.14 Electrochemistry: 267
i les
er?
(10.47)
‘
IK Ss =Q1,0+ Gon-
Q1,0
Ge (10.48)
O14 4,0
In dilute aqueous solutions ay,o = 1 and
Nag = Es (10.50)
aya
Acetic acid is a typical example, for which Ky, = 1.8 x 10~-> at 25° in water.
Provided no other electrolytes are present, the ionic strength is small, and it is
often permissible to replace activities by concentrations. Another useful approxi-
mation, valid for all but very weak acids, is to neglect the self-dissociation of
water and to set
My = Moae-
Thus, in 0.1 m acetic acid my. = moy.- = 1.4 x 107°. The contribution to
my from water itself can be established by estimating the equivalent moy-. At
25° the value of Ky is 1.00 x 10~'* and
10-14
Mon- = Va Se 10
Ky, = y+ CHiGose
Ao rc- — My+Moxc-
Murttosc Yur Yur Yosc (10.51)
AyOAc Myoac YVHOAc
The emf of the cell is given by
= 8° — F In MorMo in Yor
toa — yin Kas (1032)
Since HOAc is electrically neutral, yyo,, = 1 in sufficiently dilute solutions.
Also, the Debye-Hiickel limiting law gives yo- = Youc-, and the term involving
Sec. 10.14 Electrochemistry 269
Solve equation (10.52) using values of mi, m) and m, given in the text to evaluate
Kya for acetic acid. VANS Kgyae— led 2 Oma
Ky=> 0. (10.53)
2
TABLE 10.5
DISSQCIATION CONSTANTS OF WEAK ACIDS AND
HYDROLYSIS CONSTANTS OF WEAK BASES (25°)
Acids Ka Bases Kp
Acetic acid, CH;COOH RIB <0 Ammonia, NH; Sia On?
Arsenic acid, H3;AsO, (1) IES) Sales sec. Butylamine, C,H,,N 4:4 >< 1034
(2) 5.6 x 10-8 Diethylamine, C,H,,N 126105"
(3) 3.0 x 10-8 Dimethylamine, C.H;N 54 S< Oe
Boric acid, H;BO; 6.0 x 10-!° Ethylamine, C,H;N S16 al Ome
Carbonic acid, H.CO; (1) Ae al Ome Hydrazine, N.H,.H,O 3 SO"
(2) 4.8 x 10-!! Methylamine, CH;N Sn On
Formic acid, HCOOH 1.77 x 10-* Phenylhydrazine, CsHsNe 1k6 1052
Hydrogen sulfide, H,S (1) IO se Oe" Pyridine, C;H;N gy Se NOR
(2) 1.3 x 10-% Silver hydroxide, AZOH <a 10s"
Hypochlorous acid, HCIO 3) Se Ore Urea, CH,ON> ss} S< TORE
Oxalic acid, HyC,O, (1) BiSm al Ome
(2) S02 D105
Phosphoric acid, H;PO, (1) 7.5 x 10-3
(2) G2 a 05°
(yi 1052
and K, is the hydrolysis constant. The hydrolysis of bases does not depend
270 Electrochemistry Chap. 10
upon the electric charge type, and it is worth restating the preceding relations
for any base B’, where Z is electric charge:
B2 => HO = HBAs One
Ky = “itsApeGou- (10.55)
Equation (10.55) applies as well to ammonia as to acetate ion.
EXERCISE 10.20
Show that equations (10.50) and (10.55) are related by KpK,y = Kw.
SUMMARY, CHAPTER 10
1. Faraday’s law
1 gram-equivalent of electrolysis requires 96,487 coulombs.
2. Conductance
ee eee
eeirey,
Specific conductance
Bad
b=) 7
More
= ng (FF
(22):
2) =
AH, = = nF +1nF (=)
0€ p
oT /p
& ip= 64 — RT In 0)
[= AS
Summary, Chapter 10 Electrochemistry 271
One
7 = 0, (_)
oo) =0
=
= ae InK
pH measurement
& (cell)
=a xX pH+ 6b
where a and b depend on nature of cell used
Redox titration
ie SORTS ass
& (cell) = const. nF In a
5. Electrolyte activities
a=ym
Ionic strength
b=3DaZi
Debye-Hiickle limiting law at 25°
—log y= 01509 ZZ _A/ jr
6. Solubility product
A,B; (s) = aA” (aq.) + 5B” (aq.)
K, = m§ m (y.)**”
7. Acid dissociation
HA + H,0 = H;0* + A>
Re a Sen
+Q4-
aya
8. Base hydrolysis
B + H,O = BH* + OH™
K, = a
@ou- +
Gon
a, —
ag
9. Ion product of water
2 HO = H,0* -- OH-
Ken => ay,0+ Qor-
PROBLEMS, CHAPTER 10
1. A certain divalent metal salt solution is electrolyzed in series with a silver
coulometer. After 1 hr. the weight of silver deposited is 0.5094 g. and the weight
of the unknown metal deposited is 0.2653 g.
(a) What is the atomic weight of the unknown?
(b) What was the average current during the electrolysis ?
Ans. (a) 112.3, (b) 0.1266 amp.
3. From data in Table 10.1 find A, for the weak electrolyte NH,OH. In 0.1 M@
solution this substance is 1.34% dissociated. What is the specific conductance
of such a solution? AES, Ng = Zilla 1b = BES. 3% NO*,
4. In a moving boundary apparatus a 0.1 N solution of potassium acetate is
electrolyzed at a steady current of 0.1 amp. After 1000 sec. it is found that the
cation boundary has swept out a volume of 6.84 cc. What is the transference
number of the anion in this solution? Ans. n_ = 0.340.
5. From the data in Tables 10.1 and 10.2 find the equivalent ionic conductances
of Lit and Cl at infinite dilution. ANS Na Soe10:
6. Use the data of Table 10.3 to estimate the specific conductance of 0.01 M
Ca(NO,),. Assume that at this concentration A = A,. Find the transference
number of the cation. SIG == P20, SK NOES, Ps, SOLS
7. A-sample of water having a specific conductance of 1.12 « 10-® ohm! cm.7!
was saturated with silver chloride, whereupon the specific conductance rose to
2.85 <x 10-§ ohm7! cm.~!. Find the solubility of silver chloride, a strong electro-
lyte. AVR: © = IPS SX NO? IML
8. The equivalent conductance of the sodium salt of crotonic acid at infinite
dilution is 88.30 ohm~'eq.~!cm.?. The equivalent conductance of crotonic acid at
1.7 x 10-3 N is 39.47 ohm~'!eq.~'!cm.?. Find the degree of dissociation of crotonic
acid in this solution. Ans. 10%.
9. Formulate the galvanic cells corresponding to the following reactions and
calculate AF'%3 for each reaction:
(a) Ag (s) + H* (aq.) = Ag* (aq.) + 3 H; (g)
(b) Cd (s) + 2 H* (aq.) = Cd*? (aq.) + Hz (g)
(c) Pb (s) + 2 Agl (Ss) = PbI, (s) + 2 Ag (s)
(d) Br, (1) + 2 Cet? (aq.) = 2 Br- (aq.) + 2 Ce*4 (aq.)
Ans. (a) 18.43 keal., (c) —9.87 kcal.
10. Find &,,, for the following cells:
(a) Zn|Zn*? (a = 0.001) ||Fe+? (a = 0.001)
Pt
|Res1(Ge—105)))
(b) Ag|Agt (a = 0.1) ||Cd*? (a = 0.1)| Cd
(c) Ag, AgBr (s) |HBr (a = 0.01)| H, (1 atm.), Pt Ans. (a) 1.74 v.
Problems, Chapter 10 Electrochemistry 273
11. Estimate the mean activity coefficient of the first salt mentioned in each of
the following mixed salt solutions:
(a) 0.01 m CaCl,, 0.05 m NaCl.
(b) 0.01 m FeSO,, 0.02 m Na,SO,.
(c) 0.02 m NaCl, 0.01 m LaCl, Ans. (a) y+ = 0.52.
12. The solubility of silver sulfate, Ag,SO,, in water at 25° is 0.79 g./100 g.
Find K, (a) neglecting activity considerations, and (b) using Debye-Hiickel limit-
ing law to estimate y.. Ans. (a) 6.5 < 1075, (b) 9:4 x 107°.
14. Find the pH at the first and second end points in the titration of 0.01 M
H,CO, with 0.01 M strong base. Ans. First end point, pH =9.0.
16. From the data in Table 10.4 find the equilibrium constants for the following
reactions at 25°: {le
(a) AgBr (s) = Ag* (aq.) + Bro (aq.) 5017 410 cau
(b) Snt? (aq.) + I, (Ss) = Sn*t4 (aq.) + 21° (aq.)
(c) 5 Ce+? (aq.) + MnO, (aq.) + 8 H+ (aq.)
= 5 Ce*4 (aq.) + Mn*? (aq.) + 4H,O
17. Find @49, for the following cell
Pt| Q, QH;, H,O ||KCI (1 N)| Hg,Cl,, Hg
from the fact that in water with no added acid or base ay: = 107°.
18. Show that in the cell
Pt |Fet?, Fet?|| KCl (1 N)| Hg,Cl,, Hg
when the ferrous ion is titrated to equivalence with ceric ion (so that total Ce
= total Fe) that the emf at equivalence is
&€ = 1[€(Cet?, Ce+4) + &Fet?, Fet*)] — &(N.C.E.)
Find &49, for the cell at the equivalence point.
19. The cell
Cd |CdCl, (1 m)| AgCl, Ag
has @59, = 0.675 v. and d&/dT = —0.00065 v. deg.-! Find AHoog and ASyog for
the reaction
Cd + 2 AgCl = Cd*? + 2Cl +2 Ag
20. The emf of the cell
Pt, H, (1 atm.) |HBr (™) |AgBr (s), Ag
274 Electrochemistry Chap. 10
28. Show in general that the minimum slope of the curve of pH versus per cent
neutralization is found at 50% neutralization.
RATES OF
CHEMICAL REACTION
in a chemical system does not depend upon reaction paths. It is easily shown
that the rate of approach to equilibrium does depend strongly upon the reaction
path. For example, the immeasurably slow reaction in the dark between pure
hydrogen and chlorine becomes explosive when a trace of sodium vapor is
introduced. In either case the standard free energy of formation of hydrogen
chloride gas is —23 kcal. per mole. We may well ask why a reaction which is
quite slow under one set of conditions becomes quite rapid under different
conditions.
It is one of the chief functions of chemical kinetics to propose and test a
mechanism, or detailed description of the reaction path, as an explanation of
what is observed. To illustrate, in the case described above it appears that
reaction by simple collision between hydrogen molecules and chlorine molecules
does not produce hydrogen chloride at an appreciable rate. On the other hand,
sodium vapor reacts readily with chlorine to produce chlorine atoms
Na + Cl, — NaCl + Cl
which react efficiently with hydrogen,
Cl + H, — HCl + H
producing the product, hydrogen chloride. From the facts (a) that the reaction
is extremely rapid and (b) that the amount of hydrogen chloride produced
is much greater than the amount of sodium initially present, it is possible to
conclude that the hydrogen atoms produced in the second step react efficiently
with molecular chlorine
H+ Cl, — HCl + Cl
to produce another molecule of product and another chlorine atom. The latter
can react with hydrogen as before, thus leading to a chain reaction in which
many molecules of product are produced for each chlorine atom initially gen-
erated by the sodium.
An indispensable prerequisite to the study of the chemical kinetics of a
reaction system is a careful test of the nature of the reaction which is occurring.
Changes in concentrations should be measured quantitatively for as many
of the reactants and products as are required to establish the process or proces-
ses involved. When the stoichiometry of a reaction under study has been estab-
lished as
aA + bB+...=mM+xN-+...
the next step is to observe the rate of the reaction by repeated measurements
of the concentration of a reactant or product at known times and at constant
temperature. The choice of substance to be measured is arbitrary and may be
dictated by experimental convenience, since the stoichiometric coefficients
give relations between the rates of consumption and production of reactants
and products,
= ldGy eden
a. ad be dt.
=e (11.1)
Sec. 11.2 Rates of Chemical Reaction 277
where Cy, Cp, Cy and cy are the instantaneous concentrations and f is time.
In some instances the rate law takes the form of a simple proportionality
between the rate and the concentrations of reactants, but it will not be of such
a simple form if there are two or more simultaneous or competing processes.
In such cases the experimenter may attempt to isolate one process by a proper
choice of temperature, concentration, solvent, or other condition. If the reac-
tion is known to be reversible, the back reaction can usually be avoided by
measuring rates during the early stages of the reaction when c,, cy are near
their maximum values and cy, cy are still small. The reverse reaction can be
similarly isolated by beginning with M and N. The experimenter eventually
succeeds, we shall suppose, in securing an adequate rate law for the widest
practicable range of experimental conditions.
aA + b6B+...=mM+aN-+...
TABLE 11.1
DECOMPOSITION OF NsO; IN CCl, AT 45°
eae
—dcx,o, oa
slope
is seen to vary with the value of cy,o,. The relationship between rate and con-
centration may be examined by plotting Ac/At against the average concentra-
tion for each interval, as in Fig. 11.2. With the exception of one point, evidently
due to an experimental error, the relation is linear and
—dcy,o, ea key -
2U5
Letting x and x.. denote moles of oxygen produced at times ¢ and co, show that
for the decomposition of nitrogen pentoxide
dax|dt = k(x. — x)
RS
20
4.0F
ENZO,
(moles lit.)
3.0 (moles fri
1 lit.)
1.0
20
0.5
FIG. 11.1. Decomposition of NoOs. 10
@) concentration of NoO; (left hand scale)
(-] oxygen yield (right hand scale) of
(0) 1000 2000 3000 4000
Time (sec.)
40
3.0
RIE
The rate is first order in concentration of ester and first order in concentration
of base. It is a second-order reaction.
The reaction between nitric oxide and oxygen
2NO + O, = 2NO,
280 Rates of Chemical Reaction Cinefos 1h
when the reverse reaction may be ignored, the rate of formation of ethylene
diiodide is given by
The rate is proportional to the concentration of acid, but since the acid is not
consumed it does not appear in the stoichiometric equation. Water is consumed
in the chemical reaction, but in dilute aqueous solution the fractional change
in its concentration during the reaction is very small. For purposes of rate
measurements the concentration of water is constant, and the factor cy, 9may
be replaced by the number it represents; more frequently it is combined with
the specific velocity constant, thus: kcy,o = k’. The factor cy. is also constant
within any single experiment, and we may write
/
kCy,0 Cy —— k Cyt
Mopserce’ =
oan CGR
t = K ses. Csucrose
Then k,,;. will change from one experiment to another as cy. is varied while
k' will remain constant.
EXERCISE 11.2
From the following data for the inversion of sucrose, show that the rate is first-
order in cy:.
Inversion velocity 18.3 13.0 9.2 3.8
(min.~! x 104)
Hydrogen ion concentration 9.84 6.99 4.98 2.06
(moles l.-! x 10%)
Sec. 11.3 Rates of Chemical Reaction 281
—-G= (11.4)
It follows that the fractional rate of change in concentration does not vary with
time or concentration. The concentration can be replaced by the amount, as
moles or molecules, irrespective of the volume. The specific velocity constant
has the meaning of fractional change per unit time, and its dimension is recipro-
cal time, €.2., Sec..*.
EXERCISE 11.3
For N,O, at 35°, k = 1.4 x 10-4 sec.~!. At any moment, what is the rate of reaction
expressed as per cent per minute?
Ans. 0.84% min.7!.
—(f=kfar (11.5)
=Ine—kit I (11.6)
where / is the constant of integration. Equation (11.6) shows that a graph of
In c is linear in ¢ with a slope equal to —k. If log c is employed, the slope is
—k/2.303. Figure 11.3 demonstrates this relationship for the decomposition
of nitrogen pentoxide.
EXERCISE 11.4
From the slope of the line in Fig. 11.3 obtain a value of the first-order rate con-
stant for the reaction and compare with the values given in Table 11.1.
282 Rates of Chemical Reaction Ghapealil
In ¢, Inc = In = kt (11.7)
EXERCISE 11.5
iy tee Wy (11.8)
which shows that the half life is independent of initial concentration or amount.
For radioactive decay it is customary to characterize the decay rate by giving
the half life rather than the rate constant.
EXERCISE 11.6
Show that for radioactive decay n = n,e~** where n is the number of atoms.
io Gap Te
For gaseous substances c = P/RT and c,/c = P,/P, where P is the pressure
of reactant, but not necessarily the total pressure. For example, in the first-
order gas phase reaction
50;Cl, = SO, + Cl,
Sec. 11.3 Rates of Chemical Reaction 283
the total pressure (at constant V and T) approaches twice the initial pressure
of SO,Cl, as the pressure of SO,Cl, approaches zero. Since two moles of
products are produced for each mole of reactant consumed, the total pressure
at any instant is given by
Po Po
1 rs 24 a _— Protai
t (min.) 0 6 14 20
P,(mm.) 1795 198.6 PEMD 237.3
t 26 34 40 46
P, 252.5 payllles 284.9 29 Tel
ape me (11.10)
TABLE 11.3
REACTION OF n-PROPYL BROMIDE WITH THIOSULFATE
tion of thiosulfate ion is given by the iodine titer multiplied by the volumetric
factor Nj, Voom, = 0.00257. The various factors appearing in equation (11.11)
are obtained from the experimental data as follows:
a— x = 000257 V,= 33.63 < 0.00257 = 86.5 x 10-* at 2010\sec:
0 0025 1a) 3:1.03> <0:/00257 ==96.8 3 1074
DE=(V5 = V,,) 0:00257'= 15.39: x 0.00257 = 39.5-x 107°
G2. 000257 == 22.24<-0.00257 =3572-% 10-*
b— x = (a— x) — (a — 5) = (/, — V..) 0.00257
—L39 x 0:00257 = 29.2, x 10° at 2010 sec:
The values of the rate constant so obtained appear in the third column of
Table 11.3.
Such data can also be represented graphically, as shown in Fig. 11.5. In-
spection of equation (11.11) shows that a plot ofthe left-hand side of the equation
versus time should be a straight line with zero intercept. For simplicity, the
ordinate can be In (a — x)/(6 — x), for which an intercept of In (a/b) is obtained
as shown in Fig. 11.5. The slope of the line is A(a — b) or, if logarithms to the
base 10 are used, k(a — b)/2.303.
EXERCISE 11.8
Measure the slope of the line in Fig. 11.5; compute the value of k and compare
with the average value given in Table 11.3.
ee (11.12)
and the integral is
Lae
(G
(11.13)
These equations also apply to the general second-
order rate equation (11.10), when c, = cg. Equa-
tion (11.13) shows that a graph of 1/c versus ¢
will be linear with slope & and intercept 1/c), as
indicated in Fig. 11.6.
A very well-known example of a second-order
reaction is the decomposition of hydrogen iodide.
For a complete description of the rate process
Jc
over a wide range of experimental conditions, it
is necessary to consider both forward and reverse
reactions. At early stages of the decomposition,
however, the reverse reaction may be ignored
without serious error. The rate of this reaction was measured [G.B. Kistia-
kowsky, J. Am. Chem. Soc., 50, 2315 (1928)] by heating silica bulbs containing
hydrogen iodide for various times in a molten lead bath. The samples were
rapidly cooled and analyzed for hydrogen iodide and for iodine. Some typical
results appear in Table 11.4.
TABLE 11.4
DECOMPOSITION OF HYDROGEN IODIDE AT 321.4°
EXERCISE 11.10
Lees (11.14)
0
In contrast to a first-order reaction, the half life depends on the initial con-
centration. .
To illustrate the contrast between first-order and second-order kinetics,
consider the following hypothetical example: A substance A, initially present
at a concentration of | mole per |., decomposes at such a rate that after | hr.
its concentration has fallen to + mole per |. Evid-
ently the half life of this reaction is 1 hr. under
these conditions. If the reaction is first order,
we can predict that, since the half life is in-
dependent of concentration, the concentration
aver 2 ire will be mole per l., after 3 hr... +
mole per |. and so forth, as indicated in Table
11.5. However, if the reaction is second order,
equation (11.14) indicates that the half life for
an initial concentration of 4 mole per |. will be
twice as great as the original half life. There-
fore, it will require 2 additional hours for the
Time (hr.)
concentration to fall to mole per 1., and so forth, as shown in Table 11.5.
The two cases are illustrated graphically in Fig. 11.7, which also shows that
interpretation would be unreliable for a limited period of observation.
TABLE 11.5
CONTRAST OF FIRST- AND SECOND-ORDER KINETICS
0 1.000 1.000
1 0.500 (observed) 0.500
2 0.250 (predicted) —
3 0.125 0.250
EXERCISE 11.11
For the system described in Table 11.5, calculate the concentration of reactant at
2 hrs. if the reaction is second order. Ans. 0.333 moles/I.
288 Rates of Chemical Reaction Chap. 11
b ases (a — Xe) — Xe
eerie (11.19)
7 Ws
aG == 8G
ky
If the value of the equilibrium constant for the reaction is known, the ob-
served rate constant, k, + k_,, can be separated to yield the values of the in-
dividual rate constants, since at equilibrium
i b+ x,
= eos
Therefore,
k, ea son
Robs oa k, mis and k, —
1 Oe 1 Se Kee
where Kova, = ky = ky
In order to find the individual values of ky, and kx, it is necessary only to meas-
ure the ratio c),/cy at any time during the course of the reaction.
EXERCISE 11.13
Derive the integrated rate law for production of M in the case described above.
k in
Ans. (M) = hoe ad e— Gem tent)
Note that the second term in equation (11.24) describes the first-order decom-
position of M for the special case that o,, = 0.
Sec. 11.6 Rates of Chemical Reaction 291
In many instances, the only substance initially present will be A, that is,
Com = Coz = 9. In this event, the first term of equation (11.24) describes the
time dependence of cy. Since at any time cy + cy + cz = Co4, Solving for c,
by substitution with equations (11.22) and (11.24) yields
RM
ket
MZ Time (min)
where k,<k,< k_,. Since for ‘the first step
FIG. 11.9. Time depend-
Ky, = ki/k_,, the second inequality corresponds
ence of concentrations
to Keo, <1. That is, very little of the reactant A is in a complex reaction.
at any time present in the form M. Therefore, the
loss of A is measured by the gain in Z,
Ai as KGa (11.26)
—dc, = %, dcy,
Kee
ee
ae cy
Therefore,
SO
dt
Go
= at
Rc, (11.27)
292 Rates of Chemical Reaction Chap. 11
and the reaction follows a first-order rate law in spite of the complexity. The
observed rate constant ko; = k.Keq,
An example of this type of complexity is the reaction between nitric oxide
and oxygen
2NO + O, = 2NO,
which follows a third-order law [M. Bodenstein, Z. physik. Chem., 100, 68
(1922)|F 1:65
—dcxo 2
= KCxo0Co,
dt
The reaction was interpreted as being termolecular, but the possibility
exists that the reaction is complex, involving the equilibrium
ZNO== NO; Ka,
followed by the rate-determining step
NOP) On ONO:
This mechanism yields
—dcxo a dcxo,
ie “ad = ketv,0,Co,
Since
Key CN.0»
iQ.
CX
where
Kops. ae Ka Kean:
principle which predicts a, 8, {1, v or even the form of the rate equation. It can
only be said of the forward and reverse rate equations that they may be equated
at equilibrium and that the resulting function of the concentrations, e.g.,
CMON ees
Cree ake
must resolve to the appropriate equilibrium function
CMON ox
c% ee. = eq.
AB
K,=een
ai
and
tate = key = KKaty,
If cu,< cy, then the concentration of passive molecules may be replaced
by the concentration of all M molecules cy and the rate is directly propor-
tional to the total concentration,
fate = Koen
The observed rate constant includes the equilibrium constant for activation,
294 Rates of Chemical Reaction Chap. 11
K,, asa factor. It is postulated that the rate constant for decomposition of the
activated complex, k, has a negligible temperature dependence. Since
Lee = kK,
and
In kone — neko lnk
it follows, after applying van’t Hoff’s equation, that
dinky: «4d ky) E,
dT AT we RT
(11.30)
where E, is the energy of activation per mole for converting passive molecules
to active molecules. If both rate and equilibrium are measured at constant
pressure, it would be correct to employ H,, the heat or enthalpy of activation.
Arrhenius considered that the active and passive forms were related
somewhat as isomeric species. It is now thought that activation of a molecule
consists in raising it to a high energy level whether by thermal collision, ab-
sorption of light, or otherwise.
The Arrhenius equation, (11.28), integrates to
a3 4 In-A (11.31)
or
ee ee Deeg (11.32)
lose = 3303R TT,
and a plot of log k versus 1/T should be a straight line with slope — E,/2.303 R.
EXERCISE 11.14
If the rate constant for a reaction is twice as great at 35° as at 25°, estimate the
energy of activation by equation (11.32). Ans. 12.6 kcal. mole}.
The first stage of a kinetic study is commonly to find a suitable rate equa-
tion to describe the time dependence of the course of reaction. The measurement
of the rate is only the tool to be used. It is seldom the primary objective. The
next step is to determine the temperature dependence of the rate constant and to
measure the energy of activation. (The final and most important step, to explain
the results, will concern us for the rest of this chapter.) To illustrate,
consider again the decomposition of di-t-butyl peroxide for which measurements
TABLE 11.6
THE DECOMPOSITION OF DI-t-BUTYL PEROXIDE AT VARIOUS TEMPERATURES
k X 105 sec.-!
Temp, 125° 11352 145° Be
Solvent kcal. mole~! I< 10718 sec.~
Cumene 1.6 Si2 15.6 37.5 0.63
t-Butyl benzene iL) 5.0 15,1 38.0 iil
tri-n-Butyl amine ie, 4.2 16.0 37.0 0,35
Vapor 1,1 3,6 11.5 39.1 3.2
at different temperatures in several
solvents are given in Table 11.6 and
in Fig. 11.10 [J.H. Raley, F.F. Rust,
and W.E. Vaughan, J. Am. Chem. log&
Soc., 70, 1336 (1948)]. In each
solvent the rate constants conform
to the linear dependence of log k
versus 1/T as required by the Arrhe-
nius equation (11.28). It is worth
noting that the rate constant and
the activation energy are scarcely
affected by the solvent. 2.55
EXERCISE 11.15 ia, 10°
between A—B and C—D have stretched and weakened in the complex, while
those between A—C and B—D have begun to form.
LAL ob
The reverse reaction proceeds through the same activated complex as that
for the forward reaction.
It is postulated that the activated complex, once formed from reactants
or products, has approximately equal probabilities of decomposing into re-
actants or products. That is, the rate-determining step for the reaction is the
formation of the activated complex, a bimolecular process in this instance.
The decomposition of the activated complex however, is necessarily a unim-
olecular process, regardless of the mechanism of formation. It was shown
previously that this type of complexity will lead to a second-order rate law.
At equilibrium the rates of the forward and reverse reaction are equal.
/
K'CapCop = k’'Cactgp
sf Pa Sag
a
(11.33)
k CapCcp
et U6
—+#\
AE
Cy« =
kc =
M k,
Therefore, at sufficiently low pressure the measured rate is that of the bimole-
cular activation process which varies as the square of the pressure.
fate —=h.c% (11.39)
CH,NNCH,; = N, + C,H,
(0)
10) 100 200 300 400 500
Azomethane pressure (mm)
Sec. 11.9 Rates of Chemical Reaction 299
The first-order rate constant was measured by H.C. Ramsperger [J. Am. Chem.
Soc., 49, 1495 (1927)] over a range of pressures with the results indicated in
Fig. 11.12. It is seen that the rate constant does indeed fall off with decreasing
pressure. It has also been shown that addition of an inert gas to the system
increases the rate constant toward the high pressure value.
shown that
Cyne) SEG aCe (11.40)
In fact, we are only concerned here with the magnitude of the velocity and,
therefore, also use c to represent molecular speed.
The pressure which the gas exerts on the walls of the container is due to
molecular impacts. In a perfectly elastic collision of a molecule with a wail,
the component of velocity perpendicular to that wall simply reverses its sign.
That is, a molecule having a positive component of velocity c, which collides
with the wall parallel to the yz plane will, after collision, have an equal but
negative component of velocity —c,. Its components of velocity c, and c,
will be unchanged in the collision.
The collision described produces a change in momentum of 2mc,, where m
is the mass of the molecule. The rate of collision of a molecule with both walls
parallel to the yz plane is c,//, since the distance / must be traversed between
collisions. The rate of collisions at one wall is c,/2/ and the rate of momentum
change is
Bia a (11.41)
Individual molecular velocities will be shown to have a wide distribution.
Therefore, in computing the total rate of momentum change at the wall for n
molecules in the container we must use the average value of c2 denoted by c?.
(This is distinct from the square of the average velocity.) The average rate of
momentum change at one wall is, therefore, nmc?/I, and this is the force exerted
by the molecules on the wall. Since pressure is force per unit area
OF =p)
pees Cee ee (11.42)
and from equation (11.40) 3c2 = c?. Therefore, equation (11.42) may be written:
This is not greatly different from the average molecular speed é given by
= 8RT
Use equations (11.48) and (11.49) to calculate the root-mean-square velocity and
the average speed of nitrogen molecules at 0° in meters sec.7!.
Ans. u = 493 meters sec.~}; ¢€ = 454 meters sec.7}.
In equation (11.49) we see the basis for Graham’s law of molecular effusion
which states that if two gases effuse through a small hole, it is found that their
relative rates of effusion are inversely proportional to the square root of their
molecular weights. Evidently the rate of effusion is proportional to the average
molecular speed, and, therefore,
Cy= V8RT/xM, _ | /M, (11.50)
Cy. V8RI/xM, VM,
The use of an average value of the molecular speed in the preceding rela-
tions implies that this property of the system is independent of time as must
be the case for a sample having a constant temperature. On the other hand,
the velocities of the individual molecules are continually changing, both in
magnitude and direction, as a result of molecular collisions. Although the total
kinetic energy and momentum of the two particles involved in a collision does
not change, in general energy will be transferred between them, because most
collisions are not simple “head on” collisions, but glancing blows of molecules
moving at odd angles. It is easy to imagine that a particular molecule in a suc-
302 Rates of Chemical Reaction Chap. 11
cession of collisions may be reduced to a very small kinetic energy or, con-
versely, may acquire a very high kinetic energy. Thus, there must exist at any
instant a distribution of molecular velocities and energies which can be de-
scribed in a statistical manner. This distribution function D can be given in
the form
U0 D dc
n
where dn/n is the fraction of molecules having speeds between the values c
and c + dc.
J. C. Maxwell and L. Boltzmann treated this situation by considering
the effect of random collisions upon the distribution of molecular velocities.
They demonstrated that there is a unique distribution which does not change
with time and which must therefore describe the system at thermal equilibrium.
The distribution function of molecular speeds is derived in Chapter 17 and
is found to be
dn 4x ( m Sy
=e sy] GRE 2 Ole (11.51)
25
Nitrogen at 273°K
4 oh x 403 ——— Nitrogen at 373°K
n aw
responding to the maximum of the curve is the most probable speed. Since
the curve is not symmetrical, the average speed ¢ is slightly greater than the
most probable speed. The fraction of molecules having speeds greater than
a particular value, such as c’, is given by the ratio of the area under the curve
from c’ to infinity to the total area under the curve.
Secs Rates of Chemical Reaction 303
where Z is the collision rate and where P(E) is the fraction of collisions in which
the relative energy E, is equalled or exceeded.
To evaluate Z, the number of collisions per
cubic centimeter per second in a system of iden-
tical molecules, we consider that a collision occurs
when the centers of two molecules approach
within distance d, the molecular diameter (in
centimeters). In one second a molecule sweeps &
out a volume zd’¢, where € is the average mole- fe ‘oe Te
cular speed (in cm. sec.~'). That is, it will collide af \ ae /
. . . ° et Ee ee a a
with any other molecules whose centers lie within 71 } CS \
this volume, as indicated in Fig. 11.15. The number wed a ee ae a ee —
The collision diameter of the hydrogen iodide molecule is 3(10)-*§ cm. Find the
number of molecular collisions per cubic centimeter of gas at 0° and 1 atm. in
1 sec. FR IP SROs
304 Rates of Chemical Reaction Chap. 11
The average time between gas collisions for a single molecule is a useful
number. Dividing the answer to Exercise 11.17 by the number of molecules per
cubic centimeter, we obtain
Del Ors
= 8.08 « 10° collisions per second
6.02 x 10??/22,400
or approximately 10-° second average time between collisions. This is a very
short time, but it is long compared to the rate of molecular vibrations, which
occur on a time scale of 10~!* second. That is, approximately 10‘ internal mole-
cular vibrations occur for each collision in the gas phase. On the other hand,
an electronically excited molecule usually requires a time greater than 10~® sec.
to emit light (see Problem 26) and can be completely “quenched” by collision
at ordinary pressure.
It has been postulated that reaction occurs between two molecules when
they collide with a relative translational energy equal to or greater than the
Arrhenius activation energy E£,. For distribution of relative kinetic energy
in such a collision, Maxwell’s distribution law (see Chapter 17) takes the form
dn
7 — a
RTS —E/RT dE (11.53)
The collision theory does not permit the complete prediction of reaction
rates from molecular properties, since the theory does not provide a method
of predicting the activation energy. However, the collision number Z in equa-
tion (11.55) can be compared with the observed frequency factor A in equation
(11.29). For example, the rate constant for decomposition of hydrogen iodide
is observed to be 7.26 x 10! e-¥*/”" at 700°K where the value of E, is 43,700
cal. mole~'. The units of this rate constant are mole~' |. sec.~!. That is, the
rate is measured in moles reacting per liter per second. The collision number
at 700°K when n corresponds to | mole 1.~! is given by
8 X18.50< 10752700 € x =
Zz 5 (3 x 107-8)? 3.14 x 128 10°
= 1.75 « 10* collistonstees=*sec.=!
Converting to moles |.~! sec.~’ of HI consumed, we have
iS XO:
rate = 2 X = 5.8 x 10!°
6x 10%
which is rather good agreement with the experimental value, in view of the
approximations. Only a few simple second-order gas phase reactions are
adequately described by this theory. In some cases the frequency factor is
substantially less than the collision number, and it was formerly proposed that
some geometrical requirement has to be fulfilled in the collision, in a kind of
lock-and-key effect. This may be described by a steric factor, a number less
than unity which gives the probability that a given collision will satisfy this
requirement. This factor is taken to be independent of the energy requirement.
When the observed frequency factor is substantially greater than the pre-
dicted value, one suspects that the reaction is not a simple bimolecular process
and that a chain mechanism such as that-described in section 11.1 applies.
The next major development of the theory of reaction rates is found in the
work of Eyring and others (see S. Glasstone, K.J. Laidler, and H. Eyring, The
Theory of Rate Processes, McGraw-Hill, New York, 1941) who considered the
activation process in detail. This theory of the transition state postulates that
the products are formed by decomposition of an activated complex formed by
the reacting species. If the activated complex is in equilibrium with the reactants,
then for the bimolecular process
A + B = AB* — products
Capt — Ke CiCpz
The rate constant for the unimolecular decomposition of the activated com-
plex is defined by the equation
rate of product formation = k*c,p+
If the activated molecule AB* requires a mean time ¢* = 1/k* to attain a
critical configuration which always decomposes to product, the rate constant
may be expressed roughly as k* = 1/t* = v*. That is, v~ is a fictitious vibration
frequency of AB* which always leads to rupture of the bond. The energy
306 Rates of Chemical Reaction Chap. 11
intermediates. These species usually have such great reactivity that they survive
relatively few collisions and rapidly react either with molecules to generate
still other radicals, or with other atoms or free radicals. Until recently, the
evidence for their existence and importance in reaction mechanisms depended
largely on interpretation of kinetic data. However techniques such as flash
photolysis and intense radiolysis (Chapter 16) now produce such high concen-
trations that direct measurements are possible by absorption spectroscopy
and electron spin resonance.
The best understood example of a reaction mechanism demonstrating the
importance of atoms is the classic hydrogen-bromine reaction. In contrast to
the hydrogen-iodine reaction, which follows a simple second-order rate law
[rate = k(H,)(1,)], it was found by M. Bodenstein and S.C. Lind [Z. Physik.
Chem., 57, 168 (1907)] that the rate of reaction is described by the empirical
equation
d(HBr) _—sk((H,.)(Br,.)'”
dt 1+ k’(HBr)/(Br,)
(11.62)
While k’ is found to be temperature-independent and to have a value of - »-
proximately 0.1, k has a temperature dependence corresponding to an activation
energy of 40,200 cal. mole™'.
The complexity of the rate law was understood only when, thirteen years
later, the following mechanism was independently proposed by J.A Christiansen,
K.F. Herzfeld, and M. Polanyi. The essence of the mechanism is the proposal
that reaction is initiated by the dissociation of molecular bromine.
Br, — 2Br k,
Br+H, —HBr+H k,
H + Br, — HBr+Br k,
H + HBr—H, + Br k,
Br+ Br — Br, kK;
The bromine atoms so formed attack hydrogen, forming the product hydrogen
bromide and a hydrogen atom. The latter, in turn, attacks bromine, forming
another molecule of product and another bromine atom. Thus, the process
can be a chain reaction in which many molecules of product are formed for
each bromine atom originally produced. However, when the concentration
of product becomes significant the inhibition reaction, step 4, may compete
for hydrogen atoms and thus decrease the rate. Bromine atoms which do not
react with hydrogen recombine by step 5.
That the proposed mechanism is consistent with the observed rate law
can be shown as follows: Hydrogen bromide, the product, is formed in steps
2 and 3, and is consumed in step 4 to give the net rate.
d(H)
dt
=r) d(Br)
di
_ 0 (11.66)
and this permits solution of equations (11.64) and (11.65) as follows: From equa-
tion (11.65)
k,(Br)(H,) = k,(H)(Br,) + &,(H)(ABr)
Substituting in equation (11.64) yields
k,(Br,) = k;(Br)’
A 1/2
= Ay (4) e7 (Bal?) + (1/2) [Ba(l) — £a(5))}/RT
A;
That is,
E,(obs.) = E,(2) + 3[E.(1) — £.(5)] (11,71)
Combining equations (11.70) and (11.71) yields a value for the activation energy
for reaction of a bromine atom with molecular hydrogen.
E,(2) = 40.2 — 4(45.2) = 17.6 kcalmole"!
It has been found that organic free radicals are important intermediates
in the thermal decomposition of many organic compounds. A technique for
detection of such free radicals developed by F. Paneth and W. Hofeditz [Ber-
ichte, 62, 1335 (1929)] was applied by F.O. Rice and his collaborators [J. Am.
Chem. Soc., 53, 1959 (1932)] to this problem. The method depends on the fact
that many simple free radicals, such as CH;, react readily with a film or mirror
of metallic lead on the wall of a glass or silica tube to form the corresponding
gaseous lead alkyl. This compound is collected for analysis in a cold trap as
indicated in Fig. 11.16. The rate of removal of the mirror is a measure of the
carrier
—_
gas
lead
mirror
furnace
for decomposition
of hydrocarbon
lead alkyl
hydrocarbon collects here
FIG. 11.16. Paneth mirror
technique for detection
of free radicals. cold trap
concentration of free radicals in the gas stream and their nature can be inferred
from the composition of the lead alkyl collected.
An important advance in understanding the mechanism of pyrolysis of
many organic compounds was achieved by F.O. Rice and K.F. Herzfeld [J/.
Am. Chem. Soc., 56, 284 (1934)]. The decomposition of butane will serve as
an illustration.
Reaction is initiated by the slow dissociation of a molecule of reactant into
free radicals, for example,
310 Rates of Chemical Reaction Chap. 11
to yield another free radical and a stable molecule. The chain is eventually
terminated by combination of free radicals by such reactions as
CH, -+ lale aeacg @He k,
Ke ee OE 2 Cap oY sgt
(11.75)
Asap CCR VAY8
and according to the activated complex theory the reaction rate is
rate k = KT Ks we (11.76)
h YAR
now contains a factor in the activity coefficients not previously included.
Taking the logarithm of the rate constant and substituting by the Debye-
Hiickel limiting law (equation 10.39) for log y, we have
A+X<2AX AX +B—product
+X
Kk
ee) (11.80)
k_, + k,(B)
A simple rate law is obtained if k_, and k,(B) have very different values. If
k_,<k,(B),
rate = k,(A)(X) (11.81)
and if k_,>k,(B)
ratei== Tn? (A)(B)(X) (11.82)
In either event, the rate depends on the concentration of catalyst, although
it is neither consumed nor produced in the net reaction.
Many organic reactions are catalyzed by acids and bases. The mutarotation
of glucose, described in section 11.5, has been extensively studied from this
viewpoint. In any given experiment the rate is first order in glucose, as previ-
ously shown, and the value of the observed first-order rate constant varies with
pH, as indicated in Fig. 11.18. In the pH range 3-6 the rate is practically inde-
pendent of pH and can be described by a rate constant k, for the uncatalyzed
reaction. The increase in rate at pH < 3 can be described by a term which is
linear in concentration of H,O* and the increase in rate at pH > 6 by a term
which is linear in concentration of OH~. The net rate constant is
Kops, ae k,(H;0*) =F ky te ky(OH)
SUMMARY, CHAPTER 11
1. Reaction order
Order
— G4 Bares
Des First-order reactions
Ts | Teg Co
apa’ In > = kt,
_ des = KG. Gs
dt
Wher(a — x) =a 4 on = © — x)
] Og —)
Fah ae
When cy ¢s =
Be le
© Co
Half life: ts = 1/kés
314 Rates of Chemical Reaction Chap. 11
= (ky + kx = ¥)
(k, +k_)t= ng
2 Complex reactions
Simultaneous first order
ky ho
Successive first order A— M — Z
og S= KGa
— 1P =Kyt1 Cm kot2 Cc C= kot2
M ky as k, ( ) ae 0,M
dink
d(1/T)
_ EsR
E,(forward) — E,(reverse) = AE(thermodynamic)
1p Kinetic theory of gases
nme
io 3V
Root-mean-square molecular velocity
Vesey eet
Maxwell-Boltzmann distribution law for relative kinetic energy
Ci ee er
aaeRT Ae gas
8. Theory of reaction rate
Collision theory
Fate —9Z P(E)
Zz = 5 da’én?
(ECE) — e7 2a/RT
rate k = a Ke
= +
Kea — eo /R eA /RL
Problems, Chapter 11 Rates of Chemical Reaction 315
d(radical) _ 0
dint
10. Reactions in solution
A + B — AB* — products
rate k = KT ke Yas
h PYAR
Acid-base catalysis
PROBLEMS, CHAPTER 11
1. The rate of disintegration of a radioactive sample gave the following results:
Days 1 3 5 7 9
Rates 6200 4700 53) 2800 2100
Show that the data obey a first-order rate law. Find the half life and evaluate the
rate att = 5. Ans. ti. = 5.4 days.
2. A solution of N,O, in CCl, at 45° produces 5.02 cc. of O, in 1198 sec., 7.33
cc. of O, in 2315 sec., and a maximum of 9.58 cc. of O, after a long time. Show
that these results conform to a first-order rate law for decomposition of N,O;
and find the first-order rate constant. VAN Saka —1 Olea OmaiSCCrm
3. The rate constant for the decomposition of hydrogen iodide is 3.96 « 10-°1.
mole-! sec™!. at 321.4°, and c, = 1 m.l.7}
(a) Find the extent of decomposition after 20 hrs.
(b) Find the half life. Anse(a) 227. (b) 2-5) « 10sec:
4. What are the dimensions of the rate constants in liters, moles and seconds
for first-, second-, and third-order reactions?
5. The following data were obtained for the decomposition of di-t-butyl peroxide
at 154.6°:
Time (min.) 0 3 6 9 12 15 18
P,.¢ (mm.) 17355 193.4 2113 228.6” 244.4 2592 273.9
Plot the appropriate function of pressure versus time to demonstrate first-order
kinetics, and obtain the rate constant. FOS, SMOG IO See
54, 3863 (1932)]. Test the following data to determine whether the reaction is of
the first or second order, and evaluate k graphically.
Time (min.) 0 35) 8.02 14.30—24:55 42:50) 90:05 oe)
P (mm.) 632.0 618.5 599.4 576.1 546.8 509.3 453.3 316.0
Ansa 24 10s nine mile
19, 85 (1951)] indicates that the reaction 2 CH, — C,H, occurs at every collision
in the gas phase. Methyl radicals are produced by decomposition of azomethane
(CH;NNCH;) at 600°K at the rate of 10!° molecules cc.~! sec.~!. Since in the
steady state the rate of removal of CH; radicals must equal the rate of formation,
find the concentration of CH, in the reaction system if the collision diameter of
CH, with CH, is2.3 x 107° cm.
16. The half life of radioactive iodine,!”°I, is 25 min.
(a) Find the rate constant for the decay reaction.
(b) A sample of radioactive iodine is found to be disintegrating at a rate of
1500 atoms sec.~!. How many '”5 I atoms are present in the sample?
(c) What would be the number of 128 I atoms present in the sample 50 min.
after the observation given in (b)?
17. In acid solutions the following data were obtained for the rate of the reaction
NHi + NO, = N, + 2H,0 [J. H. Dusenbury and R. E. Powell, J. Am. Chem.
Soc., 73, 3266 (1951)]:
HNO,, mole |.~! 0.0092 0.0092 0.0488 0.0249
NH, mole 1.-! 0.098 0.049 0.196 0.196
rate.moleiles secs << 10s? 34.9 16.6 335 156
Express the rate law.
18. The value of A in the Arrhenius equation is often taken as 10!* sec.~! for
rough calculations of unimolecular reactions. Estimate the temperature at which
ti. = 1 sec. for the reaction
CHA CH.
if the C-I bond dissociation energy is 54.0 kcal. mole™'.
19. A mixture of H, and C,H,, initially equimolar is admitted to a mass spec-
trometer from a storage bulb at low pressure through a small orifice.
(a) What is the composition of the gas which emerges from the orifice?
(b) The rate of effusion of either gas is proportional to its partial pressure. After
a time it is found that the partial pressure of butane has decreased by 1%. By
how much will the hydrogen pressure have decreased ?
Integrate this rate equation, find a general expression for t,,., and show that
when log t,/2 is plotted versus log a, a straight line with slope (1 — 7) will be
obtained.
318 Rates of Chemical Reaction Chap. 11
22. Apply the steady-state treatment to the following mechanism for the decom-
position of ozone:
and obtain a general differential equation for the rate of disappearance of ozone.
Except at the very beginning of reaction it is found that the experimental rate
law is —dco,/dt = kco,/co,. What does this imply concerning the relative values
of k_, and k,?
23. The decomposition of gaseous NO,
N,O; = 2 NO, + 40,
is found to be first order and was long presumed to be a simple unimolecular
reaction. R.A. Ogg [J. Chem. Phys., 18, 573 (1950)] has shown that the mechanism
ky
N,O; = NO, in NO, fast
kt
ke
NO, + NO, —NO, + O, + NO slow
25. The rate constants for the second-order reaction of hydrogen with iodine are
T(°K) 556 629 700 781
k(mole™! |. sec.~!) LID 3 IO 61/62 l0me loge < 10" 3.58
(a) Plot log & versus 1/7 and find E,.
(b) The equilibrium constant for the reaction is found to be 3.73 at 629°K and
2.43 at 700°K. Find the rate constant and the activation energy for the reverse
reaction at 700°.
26. Electronically excited NO fluoresces with a half life approximating 2 x 10-7
sec. at very low pressure of NO. Every collision with added CO, “quenches”
fluorescence. What pressure of CO, is required at 25° to change the apparent
half life to 1 x 10-’sec.? Let dyo = dco, =4 x 10-§cm. [A.B. Callear and
I.W.M. Smith, Trans. Faraday Soc., 59, 1720 (1964).]
SURFACE
CHEMISTRY
12.1 Colloids
TABLE 12.1
PARTICLE SIZE IN COLLOIDAL SYSTEMS
ae :¢ re typical colloidal
10-7 6 x 107 suspensions
Oss 6 x 108
* Assuming spherical particles
surface separating the two phases. For this reason surface phenomena such as
adsorption are very important in colloidal systems.
EXERCISE 12.1
The radius of a gold atom is 1.4 x 107° cm. Estimate the number of gold atoms
in a spherical particle of 20 « 10-® cm. diameter (neglect voids). Estimate the
fraction of atoms lying in the surface of the particle. | Ans. 360 atoms; ca. 57%.
The most frequently encountered colloidal systems are those in which the
dispersed phase is a liquid or solid and the dispersion medium is a liquid, that
is, emulsions and sols. Two general categories of such colloids are recognizable.
The /yophobic (solvent repelling) colloids such as those formed by sulfur and
gold are inherently unstable systems which tend eventually to coagulate into
the gross solid phase. The maximum concentration of disperse phase for reason-
able stability is quite small, and the stability depends strongly on the presence of
minor components in the dispersion medium, e.g., electrolytes in water suspen-
sions. Coagulation is not usually reversible. That is, the colloidal system cannot
be reformed by simple mixing of the two phases. As the term lyophobic implies,
there is little or no attraction between the molecules of the two phases.
There is a second large group of colloidal systems called /yophilic (solvent
attracting) in which attractive forces between the two phases render the sus-
pension intrinsically stable. Coagulation does not usually occur spontaneously.
In fact, the reverse process of dispersion may occur sponténeously when the
two phases are mixed. The colloidal systems formed by proteins in aqueous
solution and high polymers in organic solvents are typical members of this class.
It has been shown in Chapter 7 that the osmotic pressure for an ideal
See, 122 Surface Chemistry 321
WSAc-
Be? = Ce =...
Simultaneous solution of equations (12.3) and (12.4) gives the equilibrium dis-
tribution of electrolytes.
(12.5)
Cnacl (ID) ; ous CNnax (1)
(a) Be ieacce(D
This equation shows that the diffusible electrolyte Na*Cl~ will not be equally
322 Surface Chemistry Chapa
distributed between the two compartments in the presence of NaX. For example,
if Cyax (I) = Cyaci () then cyac; (HH) = 1.4 Cyc (D-
It might be expected that the osmotic pressure difference between the two
solutions would be simply that corresponding to the concentration of Na*X~
in compartment I, but since, when concentrations are expressed in moles
per unit volume.
Ht() — dp _
RT 2Cnax (1) + 2Cyacr (D) — 2exacr (ID) (12.6)
and Cyac, ID) > Cyac (1) this is not the case. A simpler relation can be obtained
if a large excess of the diffusible electrolyte is used. When Cyac (ID) > Cyax (D,
by (12.3) and 12.4 cyacy (I) = Cxaci (ID) and
ell)
=I]
(12.7)
The latter technique can be used in osmotic pressure measurements on protein
systems.
EXERCISE 12.2
Find the osmotic pressure difference to be expected when cyax (J) = Cram Cd)
wi OnmViEatZon Ans. 0.0294 atm.
F=,% (12.8)
the viscosity coefficient, which has the dimensions gram cm.~' sec.~! or poise.
The viscosity of water at 25° is 0.00895 poise, and that of glycerine is 9.54
poise.
The viscosity equation (12.8), in differential form, dv/dx = F/An, is appli-
cable to a variety of experimental situations. Examples are rotating concentric
cylinders, rotating coplanar circular plates, the terminal velocity of a sphere
falling through a fluid, or the rate of flow of a fluid through a capillary tube
(viscometer). In the last instance the result is
Perch
aes (12.9)
Setes. AS Surface Chemistry 323
where V is the volume of liquid passed through the tube in time ¢, r is the radius
of the tube of length L, and P is the pressure drop through the tube.
Viscosity measurements can be made by comparing the time of flow of an
unknown fluid with a liquid of known viscosity in a viscometer. In this case
the ratio of pressures is the ratio of densities of the two fluids. Equation (12.9)
yields
M1
as Pill
Oe (12.10)
where p is the density and ¢ the time for flow of the same volume of each fluid.
EXERCISE 12.3
A certain volume of water passes through a viscometer in 30 sec. at 25°. How
long will it take the same volume of glycerine (p = 1.26 g.cc.~!) to pass? The
capillary is to be replaced with another suitable for measuring viscosities of
liquids similar to glycerine. By what factor should the radius be increased to give
approximately the same time interval? VANSE2 Sy MLOMSEC COME Ie
is usually a linear function of concentration for 7), < 2, and it may be obtained
accurately as the intercept of a plot of 7,,,/¢ versus c.
It has been found that for high polymers of a given type the intrinsic viscosity
bears a simple relation to the molecular weight of the dispersed particles, as
shown in Fig. 12.3. This relationship is summarized in the equation:
[7] = kM2 (12.13)
where kK is a proportionality constant characteristic of the dispersed phase and
dispersion medium and a (the slope of the line in Fig. 12.3) depends on the
shape of the particles. For rigid rodlike particles a = 1.0 whereas for flexible
chainlike molecules a = 0.7.
Equation (12.13) has been extensively used to determine the molecular
weight of high polymers. For polydisperse suspensions the value obtained is
evidently an average, and when a = 1.0 (rodlike particles) it can be shown that
this is the weight average molecular weight, defined by
1M; MM;
M, (12.14)
i 1 M; : DG
i
from left to right is 4n,A and from right to left is n,A. The net transfer
rate is ;
dn A(n, — ny) 12.15
dimes 2t a)
The difference in concentrations n, and n, is, in terms of the concentration
gradient dn/dx,
er (2) A (12.16)
Substitution in equation (12.15) gives
dn
a A?)\ dn
-(5)¢ OREN)
The negative sign shows, as must be the case, that the net transfer is in the
direction of decreasing concentration.
Equation (12.17) has the form of Fick’s law of diffusion, 1.e., the rate of
transfer of material across a plane is proportional to the concentration gradient
perpendicular to the plane. The proportionality constant is called the diffusion
coefficient D, and from equation (12.17)
L2
: a
j=
This equation was verified in 1909 by J. Perrin who observed the Brownian
movement of relatively large particles, ca. 5(10)-* cm. diameter, and obtained
an early value of the Avogadro number.
EXERCISE 12.5
Find A? for a time interval of 1000 sec. for the gold particles described in Exercise
12.1 at 25° in water. Ans. 4.87 x 107% cm.
Sec. 12.4 Surface Chemistry 327
(a)
o =xepyneo (12.22)
where c, is the concentration of the solution before diffusion.
For a given time of diffusion the shape of the dc/dx curve is determined by
the exponential factor. In fact, the diffusion coefficient can be determined from
a measured value of x corresponding to dc/dx equal to half its maximum value
at a particular time of diffusion, as in Fig. 12.7 (b). This quantity, x,,., is
related to D by
Le
its —x2/,/ADt
X42 23
ain? ae
Comparison of equation (12.23) with equation (12.18) shows that x, is
approximately equal to the root-mean-square distance of Brownian move-
ment, / A®.
Some values of diffusion coefficients at 20° in water are given in Table 12.3.
EXERCISE 12.6
From the diffusion coefficient of insulin given in Table 12.2 estimate the time
required for x,/2 to become 1 mm. at 20° in water. Ans. 4.4(10°) sec.
328 Surface Chemistry Giitejas |
TABLE 12.2
DIFFUSION AND SEDIMENTATION PROPERTIES OF PROTEINS
12.5 Sedimentation
(i pe (12.26)
where
/ is the frictional coefficient, given by Stokes’s law, equation (12.20).
The velocity of sedimentation increases until the force of resistance to
motion is equal to the centrifugal force, at which point the particle has achieved
its terminal velocity. This condition is established in a very short time. Equating
the right-hand sides of equations (12.25) and (12.26), we have
Vipeie) ae pe (12.27)
For spherical particles Vv= 4zr°, f = 6anr, and
5 —at axldt
EE _= =
2r°(p — Pd
po) (12.28)
wx 9n
Sec. 12.5 Surface Chemistry 329
The left-hand side of equation (12.28), the rate of sedimentation in a unit cen-
trifugal field, is called the sedimentation coefficient, symbol S. It has the dimen-
sion of time and is usually given in units of 10~!° sec., called a svedberg.
To illustrate equation (12.28), let us ask what magnitude of centrifugal field
will be required to give a velocity of sedimentation of 1 mm. per minute for
a typical colloidal suspension with particles of 100A radius, p = 1.5, suspended
in water. Solving for w?x yields
Ou
2, (dx/dt)9n
iG = AS
— (0.1/60) x 9 x 0.00895
2x 10-"(1.5 — 1.0) ——
1.35(10)8 8
cm. sec. -2
Sot { dt = ™ dx
0 xy Xx
S=Pat iy
et! In X»
a (12.29)
EXERCISE 12.7
The boundary in a monodisperse colloidal suspension in water at 20° moves from
a radial distance of 4.4 cm. to 5.0 cm. in 1000 sec. at an angular velocity of 10+
radians per sec. What is the sedimentation coefficient? Ans. 12.8 svedbergs
common basis, usually water at 20°. Some values of S and D so corrected are
given in Table 12.2.
EXERCISE 12.8
t= Or (4 ~ 5) (1 + cos* 8) (12.32)
7 4 6 2 2
where n is the number of particles per cc., r is the radius of the particles, d is
the distance between the observer and the scattering system, X’ is the wave length
of light in the dispersion medium, and m is the ratio of the refractive index of
the particles to that of the dispersion medium.
The factor 1 + cos? 6 (@ is the angle of observation with respect to the
beam axis) shows that the scattered intensity is a function of angle 6. A maxi-
mum intensity is observed at 6 = 0° and 180° and falls to one-half the maximum
at 6 = 90° and 270°. In fact, the scattered light can be resolved into two plane-
polarized components: i, polarized perpendicular to the plane containing the
source of light, the scattering center, and the point of observation; and i,
polarized in this plane. Figure 12.8 represents the angular dependence of
ig = i, + i, in graphical form. It is
seen that i, is independent of angle,
FIG. 12.8. Intensity of scattering
by small spherical particles.
corresponding to the term 1 in
(1 + cos? 6), while i, goes to zero
at 90° and 270°, as does cos? 6.
Se 90° a5°
The light that we see coming
from the sky at an angle from the
sun on a Clear day is scattered from
submicroscopic particles including
the molecules of the air. Equation
(12.32) shows why this light is pre-
dominantly blue, for the factor (\’)!
in the denominator requires that the
scattered intensity increase with
decreasing wave length. Conversely,
when we observe the sun directly at
sunset through a long atmospheric
path, the predominant loss of blue
See. 1257; Surface Chemistry 331
light by scattering gives the transmitted light its characteristic red color. When
the size of the particle becomes greater than 1/20 of the wavelength, the
angular dependence of the scattered light intensities is much more complex
than that described by equation (12.32), and is a function of particle size
as shown in Fig. 12.9.
270°
is, energy must be supplied to increase the surface area of a drop of liquid.
The origin of this characteristic of liquids lies in the attractive forces existing
between the molecules. For a molecule in the body of the liquid, these forces
tend to cancel as to direction so that there is no net force. On the other hand,
a molecule in the surface is acted upon by the attractive forces of the molecules
below it, but the opposing forces are absent. This results in a net force directed
perpendicular to the surface and into the body of the liquid. In order to increase
332 Surface Chemistry Chap. 12
the surface area of a body of liquid, molecules must move from the body of the
liquid to the surface against the forces of molecular attraction. Therefore, work
must be done to increase the surface area.
A simple method of defining the work re-
quired to increase the surface area of a body of
liquid is indicated in Fig. 12.10, which shows a
liquid film on a rectangular frame of fine wire.
Some force F on the movable side is required to
maintain the film at a given area. This force is
parallel to the surface of the film and perpendi-
cular to the edge at which the surface meets the
movable wire. If the force is increased by an
infinitesimal amount, the surface area of the film
will increase (if frictional resistance to move-
ment of the wire is neglected). Suppose that the
wire moves out a distance d, increasing the
FIG 1210-@ Surfacel eneroy, surface area of the film by 2d/. (There are two
liquid surfaces, one on each side of the film.)
The work done is Fd, and this is proportional
to the increase in surface area. The propor-
FIG. 12.1]. Capillary rise.
tionality constant is the surface tension of the
liquid, y.
Fd = 2dly
y = Fd/2dl = energy/area
y = F/2l = force/length (12.33)
Thus, we see that the surface tension is the work
required to produce a unit increase in surface
area. In the c.g.s. system the units are ergs cm.~?
or dynes cm“.
The rise or depression of liquids in capillary
tuves 1s a common manifestation of surface
tension. A liquid rises if it forms a surface which
is concave upward, for in this case the pressure
above the surface is less than that below. Since
the pressure of the gas above the surface is
essentially constant and the same as that on the
liquid surface in the main reservoir, the liquid
rises until the hydrostatic head is equal to AP.
The radius of curvature of the liquid, r, may
not be the same as that of the capillary tube, ro,
as indicated in Fig. 12.11. That is, the contact
angle of the liquid surface with the wall of the
capillary, 6, is not necessarily zero. The rela-
tionship between r and r,, by a simple geometric
Sec. 12:8 Surface Chemistry 333
construction, is found to be
zB
COSIG ==
r
and the pressure difference across the surface is
p= aes (12.34)
0
218-8 = hg
Ds
0
In many cases @ is practically zero, and the use of the approximate form
y = thpgr, is justified.
EXERCISE 12.9
A liquid of density 1.5 gm. cm™' rises to a height of 1 cm. in a capillary of 0.5 mm.
radius. Given that the contact angle is zero, find y. Ans. 36.7 dynes cm.~?.
The presence of a solute often has little effect upon the surface ten-
sion of a liquid. However, it is found that a few classes of substances such as
organic acids, alcohols, and esters exhibit a strong depression of the surface
tension even at small solute concentrations. Soaps, the alkali metal salts of fatty
acids, are especially active in this respect. For example, a 0.002 N aqueous
solution of sodium oleate has a surface tension of 25 dynes cm.~' as compared
to 72.8 dynes cm.~! for pure water.
The strong effect of soaps on the surface tension suggests that these sub-
stances must be present at much higher concentration in the surface than in the
bulk of the liquid phase. They may be said to be adsorbed at the interface. J.
Willard Gibbs obtained, by rigorous thermodynamic treatment (see W. J.
Moore, Physical Chemistry, 3rd ed., Prentice-Hall, Inc., Englewood Cliffs,
N. J., 1963, p. 737), a relation between the change of surface tension with
concentration of solute, dy/dc, and the excess concentration of solute in the
surface, I’, in moles cm.’
nite c dy
iD myer (12.36)
which is present at lower concentration in the surface than in the main body of
liquid (I‘ negative) can produce an increase in surface tension (dy/dc positive).
In the light of modern concepts of bond polarity it is not difficult to under-
stand why the long chain fatty acids and their salts should concentrate in the
surface of an aqueous solution. The carboxyl group is highly polar and as such
is compatible with water. In the low molecular weight acids the influence of this
group dominates the character of the molecule. Such acids are soluble in water
and do not have a strong effect on the surface tension. As the molecular weight
and, therefore, the length of the —CH,— chain increase, the solubility decreases,
and the effect on the surface tension increases. This is to be attributed to the
increasing influence of the nonpolar hydrocarbon portion of the molecule.
W. D. Harkins suggested that the fatty acid molecules in the surface of an
aqueous solution tend to be oriented in such a fashion that the water soluble
carboxyl group or “head” of the molecule is in the body ofthe solution, while the
insoluble hydrocarbon “tail” projects above the surface. This picture received
considerable support from the study of the surface films formed by insoluble
substances such as the high molecular
FIG. 12.12. Film pressure weight fatty acids. The experimental tech-
vemarec: nique, due to J. Langmuir, consists of
allowing an oily film to spread on a clear
LOir water surface enclosed by one movable
and one fixed barrier. The latter is at-
tached to a delicate torsion balance for
40
measurement of the horizontal force ex-
erted by the film as it is compressed by
30 the movable barrier. The force per unit
Film pressure
(dynes cm’) length of barrier as measured by the
20 | torsion balance is called the film pressure,
since it is the two-dimensional analog of
gas pressure.
When the film pressure is measured
as a function of film area, curves such as
ARES Ap 40 50 that in Fig. 12.12 are obtained. As the
Sine area area is decreased, the film pressure shows
(cm? mole” x 10°°) an abrupt increase at an area of ca.
18 x 10° cm’ mole~!. Langmuir concluded
that at this point the surface consists of a tightly packed monomolecular
layer of molecules and that further compression must result in “crumpling”
the film and formation of multimolecular layers.
If the critical area represents a monomolecular layer, the area per molecule
becomes (,)
om foam = 30A?/molecule
The area per molecule is nearly constant for fatty acids over a considerable
20
20
12.9 Adsorption
area per unit mass is exposed to a gas or Cg (g. HAc per 100 mi. soln.)
Such curves often look like Fig. 12.13, where the concentration of interest
is the equilibrium concentration of adsorbate.
The shape of the adsorption iso-
therm suggests that the capacity of the
Fig. 12.14. Test of the Langmuir surface for adsorbed material is limited
adsorption isotherm. and that at increasing concentrations
of adsorbate this capacity is more and
more closely approached. Langmuir
developed a simple theory of adsorp-
tion which satisfactorily describes such
systems by assuming that the surface
is capable of adsorbing a layer one
molecule thick and no more. When
adsorption equilibrium is reached, the
rate of desorption is equal to the rate
of adsorption.
This theory is most easily consid-
ered in its application to adsorption
of a gas. The rate of adsorption is
proportional to the rate of collisions
with the surface. This rate, in turn, is
proportional to gas pressure and to the
fraction of the surface which is unoc-
cupied by adsorbed molecules.
rate of adsorption = k,P(1 — 0) (12.37)
where 6 is the fraction of the surface occupied.
The rate of desorption will be proportional to the fraction of surface oc-
cupied :
rate of desorption = k,@ (12.38)
Equating the right-hand sides of equations (12.37) and (12.38) as the condition
for adsorption equilibrium and solving for @ yields
(12.39)
(kK,P + kp)
The weight of substance adsorbed per unit weight of adsorbent, x/m, is
proportional to 6, and
x aP
7a Ene) (12.40)
or
IP l bP
xm a sir ae (12.41)
This shows that a plot of P/(x/m) versus P (or c in the case of adsorption from
solution) should be linear. The data of Fig. 12.13 are expressed in this way
in Fig. 12.14.
See, 250 Surface Chemistry 337
(12.42)
where V is the volume adsorbed at pressure P, P, is the saturation vapor pres-
sure of the adsorbate, V,, is the volume of adsorbed gas corresponding to a
monomolecular layer, and C is a constant. For purposes of testing, the equation
may be put in the form
fis 1 (6. Sele
VPeeaP)ae Va.Ge 4,10. PB, (12.43)
A plot of the left-hand side of this equation versus P/P, should give a straight
line with slope (C — 1)/V,C and intercept 1/V,,C. Evaluation of these two
quantities and simultaneous solution gives V,,, the volume of gas required to
form a monomolecular layer. Using a value of the cross-sectional area per
molecule obtained from liquid or solid densities, this figure can be used to
estimate the surface area of finely divided solids which exhibit this type of
adsorption.
(12.45)
Sec. 12:10 Surface Chemistry 339
The properties of the diffuse double layer described by g and d are often
given in terms of the zeta potential €, the potential difference of a condenser of
charge q per unit area and distance d between plates
ped. 28 (12.46)
where e is the dielectric constant in the double layer. Substituting for gd in
equation (12.44) yields
y= ose (12.47)
The existence of zeta potential at the surface of colloidal particles gives rise
to two complementary phenomena. One is the Dorn effect: the development of
a potential difference upon sedimentation of a suspension. This is analogous to
the streaming potential, except that in this case the solid phase moves rather
than the liquid. The second phenomenon is more closely related to electro-
osmosis and is called electrophoresis. In this case a potential gradient applied
to a suspension causes migration of the suspended particles in a direction corre-
sponding to their charge. The charge on the particles may be due to absorbed
ions, or, as in the case of the proteins, due to the inherent ionizability of certain
groups in the molecule.
Since proteins contain both —NH, and —-COOH groups, it is expected
that at low pH acidic dissociation will be repressed and base hydrolysis to form
—NH); will be favored. Under these conditions the protein molecules should
exhibit electrophoretic migration toward the cathode. Conversely, at high pH
acidic ionization to form COO~ is favored and base hydrolysis repressed.
Migration, therefore, occurs toward the anode. For each protein there will be
a particular pH at which equal numbers of positive and negative charges
are present, resulting in zero electrophoretic mobility. This is the isoelectric
point.
The existence of adsorbed ions on the surface of lyophobic colloids has
a great deal to do with their stability. The tendency of such particles to coagulate
is counteracted by coulombic repulsion of the charged surfaces when such
particles approach each other. The stability of such suspensions is strongly
dependent on the type and amounts of electrolytes present in the solution.
Lyophobic colloids are coagulated by addition of relatively small amounts
of electrolyte to the suspension medium. The critical concentration for coagu-
lation of a particular colloid depends on the nature of the electrolyte. Those
most effective are those having multiply charged ions of sign opposite to that
of the colloidal particles. For example, the negatively charged colloid arsenic
trisulfide is coagulated by 0.040 M KCI and 0.033 M K,SO,, but only 0.00023
M PbCl, and only 0.000093 M AICI, are required. A positively charged sus-
pension is equally sensitive to multiply charged negative ions.
340 Surface Chemistry Chap. 12
SUMMARY, CHAPTER 12
1. Kinetic properties of colloids
Osmotic pressure
im py RE
c0 (6 = M
: ; vA
Viscosity, defined by F = » rae
G70)
sp. No
dt ax
kT
ONS
At a plane boundary
dc £0 —22/4D¢
dx 2(x Diy?“
Sedimentation
ies dx/dt _— 2r?(p — ps)
moe 9n
2. Optical properties
Rayleigh scattering
; 8 4 6 2 ==] 2
7 = Toy fe is3) Ca
Dissymmetry of scattering, i,;-/i,3;, depends on particle size and shape.
3. Surface properties
Surface tension
Capillary rise
Problems, Chapter 12 Surface Chemistry 341
Ee ne
m (1+ 5P)
Brunauer-Emmett-Teller adsorption isotherm (multimolecular layer)
eee VAGR
eee
(= Pile DPI
4. Electrical properties—electrocapillarity, electroosmosis, electrophoresis,
streaming potential
Zeta potential,
c= 4nqd
€
14. Find the radius of spherical gold particles which fall 1 cm. in water at 20°.
(a) In 0.5 sec.
(b) In 7 hrs.
(c) In 29 days.
The density of gold is 19.3 g. cc.7!.
15. A nondiffusible electrolyte NaX is present on one side of a semipermeable
membrane at a concentration 0.1 M. Find the ratios of NaCl concentrations on
the two sides of the membrane for the case that the concentration of chloride ion
in the solution of NaX is (a) 10-3 M, (b) 10-? M, (c) 107! M, and (d) 1.0 M.
16. The following data were obtained for the surface tension at 18° of aqueous
solutions of isobutyric acid.
Conc. (M) 0 0.0187 0.0250 0.0500 0.100 0.250
ry (dynes cm.~!) 73.0 68.6 67.3 63.3 Sst 48.3
0.500 1.00
40.7 32.6
(a) Use these data to test the empirical equation of Szyszkowski [Z. physik.
Chem., 64, 385 (1908).
%—)
Yo
— pin(£A +1)
where c is concentration, A (= 0.051) is constant for all concentrations of iso-
butyric acid, and Bis found to be the same for all fatty acids. Evaluate B.
(b) Use the value of B in the preceding equation to evaluate dy/dInc at high
concentrations where c > A.
(c) Use the preceding result to evaluate I in the Gibbs equation for high con-
centrations.
(d) Find the effective area per molecule in the surface layer of isobutyric acid.
17. Fora tube of the construction shown in the figure a sample of
liquid assumes an equilibrium position. Explain the effect and show
that the equation y = hpgRr/2(R — r) applies where h is the overall
height of the liquid column, p is the density of the liquid, g is the
acceleration of gravity, R and r are the radii of the large and small
bore, respectively. Assume that @ = 0. [See. S Natelson and A. H.
Pearl, J. Am. Chem. Soc., 57, 1520 (1935).]
@rAt 26 fork — 0109 7iem.. 7 01015 cma—/r— 83cm. for
water. Find vy.
(b) Show that when the tube is held in a horizontal position and
a pressure P applied at the narrow end, y = PRr/2(R — r). Find P
for water.
18. At sedimentation equilibrium the rate of sedimentation across
any plane in the suspension is equal and opposite to the rate of
diffusion across that plane. The rate of sedimentation is given by
or
dt /s if
where v is the number of particles per unit volume, m, is the effec- Prob. Fig. 12.17.
tive mass per particle and is equal to $zr*(p — po), g is the accel-
eration of gravity, and f is the frictional coefficient. The rate of
diffusion is
Surface Chemistry Chap. 12
344
ee ae
Equate these expressions and integrate between limits n, and n,, x, and x, to obtain
detail in Fig. 13.3. In the region between the horizontal lines a uniform
electric field exerts a force at right angles to the direction of the ray. The force
acting on a particle of mass m and of charge e in an electric field of strength E is
Ee. This force produces an acceleration Ee/m directed toward the positive
electrode and at right angles to the direction of the ray. If the angle of deflection
is small, the time of transit through the field, during which transverse acceleration
occurs, is //v, where v is the velocity of the particles in the ray and / is the distance
FIG. 13.3. Curvature of cath- kk Z >
ode rays in an electric field. eee =
traveled in the field. Applying the simple ee of motion for distance traveled
under a constant acceleration (distance = 4 acc. X time’), we have, for the
vertical displacement, cee
x= 52 (+) (13.1)
This may be solved for e/m, the charge-to-mass ratio of the particles composing
the cathode ray.
eo 2x
Fe le (1322)
eee
m Hel?
(13.4)
A consistent set of units must be used for the quantities in equation (13.4). Such
a set would be the mks (meter-kilogram-second) system in which £ is in volts
per meter and H in webers per square meter (1 weber meter~” = 10* gauss).
In this case e/m is obtained in coulombs per kilogram. The most recent value of
e/m for cathode rays is 1.7588 x 10'! coulomb kg.~’
In the type of experiment which has just been described it is significant that
all particles describe the same trajectory. It follows from equation (13.4) that
they must all have a common value of e/m. It would appear highly probable that
these particles have common values of e and of m and consist of but one species.
348 Atomic Structure Chap. 13
strength that it neither rises nor falls, then the electric force Ee may be equated
to the gravitational force and we have
m
1100797 = 1.67366 x 10>" ¢.
1 ~ 6.02252 x10%
This is the lightest type of atom known.
The mass of the electron is much less than that of any atom. It is only about
zsiyz as heavy as a hydrogen atom. Therefore, the electron, the negatively
charged particle common to all atoms, constitutes only a very minor fraction of
the mass of an atom. The positive ion formed by electron loss possesses practi-
cally all of the mass of the atom.
EXERCISE 13.2
Find e/m for the helium ion, He++, and compare with that for the electron.
Ans. 4.80 x 10° coulomb kg.~!.
Having described the charge and mass of the positive and negative
electric particles obtainable from atoms, we will now examine information
350 Atomic Structure Ghap eas
bearing on the spatial distribution of these particles in the atom. Early concepts
of this distribution were, of course, rather indefinite, but tended to view the
positive portion as extending throughout the volume occupied by the neutral
atom, with the light, negative electrons embedded in this matrix.
In 1911 Ernest Rutherford proposed the nuclear theory of atomic structure,
based on an analysis of experiments on the scattering of positive ions by metal
foils. The positive ions used in the original experiments were alpha particles,
a “radiation” emitted by radioactive substances. The alpha particles had been
shown to be very energetic helium ions He**, with an atomic mass four times as
great as a hydrogen atom. A collimated beam of these particles was directed
normal to the surface of a thin sheet of a metal such as gold, and by means of
a fluorescent screen placed in various positions, the angular distribution of
the scattered particles was observed. The arrangement is indicated in Fig.
13%:
The most striking qualitative
observation made in these experi-
ments was that while most of the
Gold |foil alpha particles passed through the
foil with little or no deflection, a
few were deflected through angles
greater than 90°. Although the foil
++ tt+¢t¢¢]¢ + 4+ 4+ 44+ +444 Detector was only 4(10)"4 cm. thick, it nev-
1 ertheless constituted a layer of
+ about 10‘ atoms. To understand
the significance of the large angle
+ scattering, imagine that rifle bullets
are being fired through a thick slab
of plastic foam. What would you
FIG. 13.5. Rutherford scattering experiment.
conclude if something like one in
ten thousand of the bullets was
deflected back in the general direc-
tion of the marksman? Superficially, the target looks like a uniform layer of
matter of low density, but the occasional deflection of the projectiles through
a large angle indicates small bodies of high density. Since the scattering is a
very infrequent occurrence, it is not likely that it can be the result of two or
more encounters by the same projectile.
From momentum considerations scattering cannot be due to encounters
between the a particle and the electron. The mass of the a@ particle is about
7300 times that of the electron, and, therefore, the a particle cannot be deflected
appreciably by an encounter with an electron. The observed deflection must
result from an interaction with the heavy positive portion of the atom. Quali-
tatively, the fact that deflection is a relatively rare event indicates that this heavy
positive portion of the atom occupies very little space. Furthermore, the obser-
vation of a small number of very large deflections implies that the positive
charge is very highly concentrated in space, thus producing an electrical field of
great strength in its immediate vicinity.
Sec. 13.5 Atomic Structure 351
On the assumption that the interaction between the a particle and the
nucleus can be described by Coulomb’s law, Rutherford computed the distri-
bution of scattered a@ particles as a function of the angle through which they
were deflected. He found excellent agreement between calculation and experi-
ment down to interaction distances of ~10°-!2cm. The agreement between
model and experiment is taken as evidence that the positive heavy portion of the
atom is indeed concentrated in a very small fraction of the total volume oc-
cupied by the atom. This portion of the atom is called the nucleus. It must be
less than 10°!2 cm. in size. Apparently the heavy positive nucleus has an ex-
tension in space only 107° times that of the whole atom, which extends approxi-
mately 10°‘ cm. In terms of volume, which is proportional to the cube of a
linear dimension, the nucleus occupies only about 10~'° of the total volume of
the atom. The positive charge and most of the mass of the atom is concentrated
in this tiny volume, giving the nucleus a density of about 10'!° gm. cm.~*, or 10°
metric tons per cubic millimeter.
13.5. Isotopes
Hev = a (13.11)
For any given ion the velocity v may be eliminated between (13.10) and (13.11),
Secin liste Atomic Structure 353
yielding
wal sas
ag oi err (1342)
This equation shows that for a given accelerating potential and magnetic field
strength, the radius of curvature is determined by the charge-to-mass ratio of
the ions. The conditions of ionization in the source region are such that most
of the ions are formed by loss of a single electron, and, therefore, most
of them have the same charge,
1.602 « 10°'® coulomb. However,
the masses of the ions formed,
even from a single substance, may
have several values, due to the
presence of isotopes. These mass
varieties give rise to a mass spec-
trum such as that shown in Fig.
13.7. =
current
In the mass spectrometer a sin-
gle exit slit and collecting electrode
are provided so that only ions
having a predetermined radius of
curvature pass through and are
registered as a current on the col-
lecting electrode. Variation of V
20 21
or H brings ions of various values
of e/m to focus on the exit slit and
thus the mass spectrum is scanned. FIG. 13.7. Mass spectrum of neon.
By rearrangement of equation
(13.12),
”m — 72
x (13.13)
i) ~
EXERCISE 13.4
Typical operating values for a modern mass spectrometer are r = 15.0 cm. and
H = 0.200 webers m.~2. What accelerating voltage will be required to detect”? Ne*
ions ? (Use consistent units, as suggested in section 13.2.) Ariss 2x 10s ve
TABLE 13.1
ISOTOPES OF THE ELEMENTS
The three isotopes of neon differ in mass by almost exactly the mass of the
hydrogen atom, and this indicates a general regularity found among isotopes of
a given element: their masses differ by almost exact integral multiples of a unit
mass (the hydrogen mass). When the isotopes of one element are compared
with those of another element, the same regularity is observed. The isotopic
masses are approximately integral multiples of the hydrogen mass. The ap-
proximation becomes poorer as the comparison is extended to the full length
of the system of elements, and this mass defect has an important meaning which
is discussed in texts on nuclear chemistry.
Sec. 13.7 Atomic Structure 355
The wave lengths or frequencies of these lines shift in a regular manner with
the position of the element in the periodic system. In 1916 H. G. J. Moseley
showed that for a particular line such as the K, line, the square root of the
frequency v is a linear function of atomic number.
It is evident that isotopic nuclides differ only in the number of neutrons in their
nuclei.
Although it has been found that the relative amounts of the various
isotopes of an element are practically independent of source or treatment, some
very small differences in chemical and physical properties are to be expected.
The difference in properties of the isotopes of an element will in general depend
on the ratio of the isotopic masses. Therefore, the greatest differences in isotopic
properties are to be found in the light elements where the isotopes exhibit the
greatest fractional difference in mass. The isotopes of hydrogen, 'H and °H,
differ in mass by a factor of two, and this element will furnish the most out-
standing illustration of isotopic differences.
Many physical properties of atoms and molecules depend on atomic mass.
For example, gas density depends directly on the masses of the atoms consti-
tuting the molecules. This property has been used as a basis for analysis of
hydrogen isotope mixtures. Many other gas properties, such as rate of effusion,
viscosity and thermal conductivity also depend on molecular mass. For ex-
ample, the relative rates of effusion of molecular hydrogen containing only 'H
(called protium) as compared to molecular hydrogen containing only *H (called
deuterium, symbol D) are given by Graham’s law.
Ru,
Rp, = nes
5 aS 1.414
The uranium isotope ?*°U is concentrated for atomic energy use in large
amounts by this kind of process, employing the gas UF, diffusing repeatedly
through very small pores in a metallic barrier. The ideal separation factor a@ for
a single stage in such a process is given by
(Xo3s/Xoss)nnar — [352.04
a= al = 1.0043
(Xo35/Xo38)initial 349.03
If n stages of effusion are employed, the overall change in concentration is
given by
Z) 23 (3) a” 13.15
(338/ final AG 38/ initial ( ‘ )
where X,;, and X,3, are the mole fractions of **°UF, and ***UF,, respectively.
EXERCISE 13.5
What is the minimum number of stages of effusion required to give an equimolar
mixture of 2°5UF, and 7°°UF,, starting with the natural mixture of isotopes?
Ans. n = 1150.
358 Atomic Structure Chap. 13
TABLE 13.2
VAPOR PRESSURES OF ISOTOPIC LIQUIDS
Compounds P,/P»
MNNp, Nz 1.0081
WNH,, NH; 1.00246
NH,, ND; 1.110
H, 0, H,80 1.0046
H,0, D,O 1.051
The isotope effect on chemical equilibrium constants is usually quite small
but it is significant for the lighter elements. It is best observed by examination
of the equilibrium constant for isotopic exchange. For example, if two com-
pounds of hydrogen HX and HY containing significant amounts of both
isotopes (the isotopic molecules being designated by DX and DY) are capable of
exchanging hydrogen atoms, the process
HX -- DY = AY += DX
occurs until a state of equilibrium is reached. This state is described by the
equilibrium function
K = Gar Ox (13.16)
Cyx* Coy
(taking concentration as the measure of activity). This equation may be rear-
ranged to show the relation between the isotopic constitution of the two mo-
lecular species
Cay _ Gax (13.17)
Coy Cpx
Now if the chemical properties of the isotopic molecules are not identical,
that is, if F°(HX) 4 F°(DX) and F°(HY) + F°(DY), the equilibrium constant
in equations (13.16) and (13.17) will differ from unity. This means that at
equilibrium the isotopic composition in the Y compound will differ from that
in the X compound. Several cases in which the equilibrium constant for isotopic
exchange has been measured are given in Table 13.3. In the case of the nitrogen
isotopic equilibrium, the value indicates that when gaseous ammonia is equi-
librated with an ammonium salt solution, the ratio '‘N/!°N in the gas phase
will be 1.023 times that ratio in the solution.
These effects are not large enough to cause concern in ordinary chemical
manipulation, and the isotopes may be said to have practically identical prop-
erties. Nevertheless, it is possible by countercurrent exchange techniques to
Seca sall0 Atomic Structure 359
TABLE 13.3
ISOTOPIC EXCHANGE EQUILIBRIA
Equilibrium
Equilibrium constant
continuum
14 Vy.
1S wave wave
length number
12 (A) (cm')
11 g
2
Energy 10 3
(electron volts cS
per atom) 9 \ 30,000
: a
Pesos
7 ae
o
5 2 20,000
ie]
4 (oa)
1 15,000
0 iM 7000
When an atom emits light, its energy is decreased. When it absorbs light, its
energy is increased, The change in the energy of the atom, AE,, can be equated
to the energy, E, of the light quantum either absorbed or emitted.
AE hy 321)
Since only certain frequencies are observed, the atom undergoes only certain
definite changes in energy. This implies the existence of discrete, sharply
defined energy levels in the atom. The emission and absorption of radiation may
then be understood as resulting from transitions between fixed energy states
characteristic of the atom. This was first recognized by Niels Bohr in 1913.!
EXERCISE 13.6
One of the important lines in the spectrum of atomic hydrogen lies at 7X = 4861
A, v = 6.17 x 10! sec.~!. What is the corresponding change in the energy content ey }, yy
of the hydrogen atom? Ans. 4.09 x 107'!? erg atom™!; 59 kcal. (g.-atom)~!. ‘ AS ‘s
wa
¥
1For a fascinating description of the history of this discovery, see W. J. Moore, Physical
Chemistry, 3rd ed. (Prentice-Hall, Inc., Englewood Cliffs, N. J., 1962), p. 471. The whole
topic of the rise of quantum theory is discussed in detail in A. D’Abro, The Rise of the
New Physics (Dover Publications, Inc., New York, N.Y. 1951).
362 Atomic Structure Chap: 13
The Bohr model of the hydrogen atom may be based on the following
postulates.
I. The Rutherford model is assumed to be correct. The atom consists
of a heavy positively charged nucleus around which orbits a light negatively
charged electron much as the planets orbit the sun.
II. These orbits are assumed to be stable, and while the electron oc-
cupies one of them it cannot radiate energy. This postulate is required because,
according to classical physics, electrical charges undergoing accelerated motion
radiate energy. Any particle executing circular motion is accelerating, since it
is constantly changing its direction and therefore its velocity. (Recall that
velocity is a vector quantity and is specified by both a magnitude and a di-
rection.)
Ill. Emission of radiation occurs whenever the electron moves from an
orbit of higher energy to one of lower energy. The frequency of this radiation
is given by the relation
AE
ie
where AE is the difference in energy between the orbits.
IV. Only orbits corresponding to certain values of angular momentum,
mor, are allowed. These allowed angular momenta meet the condition
nh
Mery = Be (13.22)
Secoeals: Atomic Structure 363
where m is the mass of the electron, v is its velocity, r, is the radius of the orbit
corresponding to a given n, n is any integer except 0, and / is Planck’s constant.
The required concept of quantization is thus introduced as the a priori assump-
tion that only certain values of angular momentum are allowed, or that angular
momentum isquantized. Aa
On the basis of these postulates the Bohr model can be developed as fol-
lows. In stable orbits the centrifugal force arising from the electron’s motion
must be equal in magnitude
to the coulombic force of attraction between the
nucleus of charge Ze electron
and the of charge e,
(Ze)(e) _ mv’
Pr ln
Therefore,
pee Lez
earner (13.23)
ee meZ ae?
For n = | this equation yields r = a, = 0.529 x 10-8 cm. a, has been named
the Bohr radius.
The total energy F of the electron in a given orbit is equal to the sum of
its kinetic energy T and its potential energy V.
Bat a try? =
Using (13.23) to substitute for v’, we have
es meee Le, Ze.
F=f V7 = ee (13225)
(The kinetic and potential energies are opposite in sign in agreement with the
general relation that force equals the rate of change of energy with distance,
Since the forces act in opposite directions, they must be opposite in sign. There-
fore, the values of dE/dx are also opposite in sign.)
We can restate the information contained in equation (13.25) as follows:
E=-T=5 (13.26)
We shall take r = co where both JT and V are zero as our reference point. At
r = co dissociation is complete. As r decreases, V_will decrease and 7 will in-
creas the total energy, also decreases (equation 13.26), since the absolute
364 Atomic Structure Chap. 13
eee (13.27)
This relationship cannot be checked directly. However, atomic spectral emis-
sion lines are postulated to arise from transitions of an electron from a higher
energy orbit (designated n’’) to one of lower energy (designated n’). The
absolute value of the energy difference between these orbits, |AZ|, can be used
to calculate a value for the Rydberg constant which can then be compared to
experiment. From (13.27)
we Nee Ge ne fic
and substituting, we obtain
pf = EEch (aa
\n? —gn) (13.29)
2 4 °F 2
Se me Z (] | )
hi 1D”
(13.30)
However, equation (13.29) was derived from an expression for the permitted
orbits of a single electron in the field of a nucleus. In the many-electron atoms
the nuclear field is somewhat influenced by the electrons; that is, a screening
effect occurs which influences the energy content attributed to the various orbits.
This can be allowed for by including a parameter b which reduces the effective
nuclear charge. This effect amounts to only about 0.5 units of charge for the K
radiation. Thus, we obtain from equation (13.30) an expression of the same form
as the Moseley equation (13.14).
ss — (Z— b)*(0.75)
Jo =./0.15 2% We Z—b) (13.31)
The computed value of the parameter a agrees satisfactorily with that obtained
from the Moseley equation.
EXERCISE 13.8
With the value of R given previously, use equation (13.31) to predict the frequency
and wave length of the K, line for the element calcium. Assume b = 0.5.
AS — OS Se ulOMSecia Sey 33220 nl Onercmis
to the plate P, and only electrons with sufficient energy to overcome this
potential can reach P. The current reaching P is measured by some device
such as an electrometer, C, attached to P.
Electrons emitted from F can be involved in two kinds of collisions, elastic
and inelastic. In an elastic collision total kinetic energy is conserved. Since the
mass of the electron is much less than that of the struck atom, it will lose little
energy in the collision. In an inelastic collision kinetic energy of the electron
can be converted into electronic excitation energy of the atom; that is, an
electron in the atom can go from a lower to a higher energy level. In such a
collision the impinging electron loses such a large amount of kinetic energy
that it cannot overcome the negative retarding potential on the plate P and is
returned to the grid.
Since the energy content of the atom and its orbital electrons is sharply
quantized, inelastic collisions occur only at energies in excess of well-defined
values corresponding to transitions between permitted energy states. That is,
with increasing accelerating potential from F to G, it is observed that the plate
current drops sharply at certain definite potentials. These critical potentials
measure the onset of permitted changes in the energy content of the atoms
bombarded. For example, helium shows critical potentials at 19.75, 20.55,
AD 2 2 Oe ane 245.2
The highest critical potential is the ionization potential; that is, it measures
the minimum energy required to remove an electron completely from the atom.
The energy corresponding to this or other critical potentials may also be supplied
by absorption of electromagnetic radiation in the form of light, X rays, etc.
By suitable modifications of the apparatus, the positive ions thus produced can
also be detected.
25
20
= 1)
Ionization
(v.)
potential
O 10 20 30 40 50 60 70 80 90 100
Sec. lonlo Atomic Structure 367
pf ,_2n'me'Z?
= EE (- (1 )I (13.32)
== [== NOSIS ean
or
RE hc = 217850" ere atom
= 13.595 electron volts atom7!
This transition is indicated as the upper limit of converging energy states in
Fig. 13.9 and is confirmed by direct observation.
The measured ionization potentials of the atoms according to atomic
number are indicated in Fig. 13.11 and show a definite periodicity which may be
correlated with chemical properties. The elements of highest ionization potential
are helium and neon, which seem to be of low chemical reactivity. Each of
these elements is immediately followed by an element of very low ionization
potential, an alkali metal. These elements are very reactive, easily forming
electrovalent compounds in which the alkali metal atom loses an electron to
form a positive ion. [The similar ionization potentials of Xe and O, provided
a clue which led to the first synthesis of a noble gas compound by N. Bartlett
(Proc. Chem. Soc., 1962, 218).]
EXERCISE 13.9
From the ionization potential of the hydrogen atom given above find the ioni-
zation potential of He*, that is, the energy term in the process
Het = He** + e
Ans. 54.380 electron volt atom7!.
2xr = nr (13-35)
ERY
Now, since p = mv, where v is the particle velocity, by equation (13.34)
Therefore, 2xmur = nh, and since p,, the angular momentum, is equal to mur,
we have pz = nh/2z, which is Bohr’s postulate of quantization of angular mo-
mentum. Thus, de Broglie’s work indicated the possibility of a description of
the hydrogen atom based specifically on a consideration of the wave properties
Sec: 13.17 Atomic Structure 369
(n= 2)
es
n
(13.38)
where 7 is any integer except 0. Only those funda-
mental vibrations which satisfy equation (13.38) are
allowed. What is the source of this limitation on the
modes of vibration? It arises from a physical condi-
TG > je) \) EY tion imposed by the fact that the two ends of the
string are fastened. Only those vibrations are allowed
ae) x=a, Which are compatible with the condition that the ends
of the string cannot move.
It is clear that the waves which we have dia-
grammed can all be represented mathematically as
sine functions. We shall represent such wave functions
by the symbol yr. For the waves graphed in Fig. 13.13 a suitable form of the
function Jr is
py = A sin (360°) = A sin ae (13.39)
where A is a constant governing the amplitude or maximum height of the wave,
x is the distance from the left origin, and 2» is the wave length of the wave.
If we now employ equation (13.38), (13.39) becomes
Other functions can also be used to describe waves. For example, the wave
shown in Fig. 13.13(a) can be described by a cosine function with the origin
selected to coincide with the maximum in the wave. Combinations of sine and
cosine functions can also be employed. We are led therefore to seek some
property common to these various functions.
A common property is found in their second derivatives. Consider the
second derivative of equation (13.39).
Secs Sl Atomic Structure 371
oor
Ox?
+ Se + Ge
oz
= (13.44)
This equation can be Zeros to give
— gem
(At De
(Go at oy? a
Oe ae
geet
Eee
An often-used shorthand notation for this equation is’? Hy = Er where
point quite different in character from that of the Bohr planetary model. In the
Bohr theory a mechanical model consisting of an electron circling a nucleus
was a basic postulate. However, wave mechanics begins not with the postulate
of a mechanical model, but rather with the postulate of a mathematical relation-
ship, the wave equation. Since we feel (usually incorrectly) that we have a better
intuitive grasp of mechanical models, we are somewhat annoyed by the choice
of the more abstract starting point which a mathematical equation provides.
However, the ultimate test of a model is the agreement of its predictions with
experimental observations, and judged by this criterion the wave mechanical
model is an outstanding success.
We shall begin by making a formal statement of the postulates upon which
we shall base our discussion of a simple one-dimensional problem.‘
Postulate I. To each system there belongs a function, yy, which is a solution
of the Schrodinger equation for that system.
Postulate If. The function xy and its slope dyr/dx are continuous, finite,
and single-valued at every point in space. These are requirements which are
met by all functions describing physical waves, for example water waves, sound
waves, and electromagnetic waves.
Postulate II. wy*dx represents the probability of finding the particle
associated with the wave function yy between x and x + dx. As a consequence
of this assumption
[tax = 1
(See Section 13.18).
Our next task is to apply these postulates to some system of interest, and to
derive the consequences of this application.
4The set of postulates which we make here is not the only possible set, nor is it a com-
plete set; however, it is adequate to our purposes.
374 Atomic Structure Chap. 13
the question of the value of V later.) The potential V is assumed to affect the
particle only at the walls of the box. Inside the box V = 0, and therefore the
wave equation describing this system inside the box will be
(13.46)
2 2
a = re 4)
For convenience we shall define our origin as the left wall of the box. The
coordinate of the right wall is then x = a,. Our problem now is to find the wr
functions which reduce equation (13.46) to an identity. The term 82?mE/h’ is
constant for any particular value of the energy £. Since the potential energy
inside the box is 0, E represents kinetic energy only.
We can rewrite equation (13.46) for a particular value of E as
ON = — kip (13.47)
where
5 oeThe
Ke R (13.48)
ra (13.52)
ap = Asin — (13.53)
There is thus a family of acceptable functions corresponding to n = 1, n = 2,
n = 3, etc. These functions are called eigenfunctions.
In order to derive the energy formula, equations (13.48) and (13.52) must be
combined. The result is
n2 h?
| pe (13.54)
8m a.
Equation (13.54) gives the allowed values of the energy in terms of a set of
integers (nm = 1, 2, 3, etc.), the quantum numbers for the system. The allowed
values of the energy are called eigenvalues.
The value of the constant A may be determined by the use of postulate III.
Our problem requires that the particle be inside the box, that is, somewhere
between 0 and a,. Therefore,
[asin es
0
1 (13.55)
x
A=,/2
ay
The wave function now becomes
eee = sm (13.56)
x ay
The process which we have just carried out is called normalization, and the wave
function yr, is now said to be normalized.
The family of functions which we are now considering has another im-
self [ose
d(sn2] =
Evaluating, we have
2 2
Seee a
376 Atomic Structure Chap. 13
portant property. Consider any two of them, ,(m = /), and r,,(m = m) where
l=ams Then®
ifWirlrm dx = 0
Functions which show this type of behavior are said to be orthogonal to one
another. Orthogonal functions are independent of each other. Exact solutions
of the Schrédinger equation are always orthogonal.
There are several other points of general interest to be mentioned. First of
all, graphs of the wave functions for n = 1, 2,3 are presented in Fig. 13.13.
The function for n = 1 represents the lowest allowed state of the system. The
wave length cannot be longer than that shown if the wave function is to be zero
at both x = 0 and x = a,. The energy cannot then be lower than that for this
state, and therefore cannot be 0. The energy of this lowest state is called the
zero point energy. It is characteristic of systems executing to-and-fro motion
(vibrations) that the energy of their lowest allowed state is greater than 0,
that is, that a zero point energy exists.
Next, we should note that there is a relation between the kinetic energy of
the particle and the curvature of the wave function. For a box of fixed size, as
the curvature increases, the number of nodes increases, the wave length there-
fore decreases, and the kinetic energy increases. These phenomena are a direct
consequence of the form of the Schrédinger equation, and so these types of
behavior are general.
Our final consideration will be the magnitude of the potential energy V
required to keep the particle in the box. At the wall of the box the differential
equation which the wave function must fit is
d?yp 14
ar = h(E Vy (13.57)
The solution of equation (13.57) is
ap = Cre aa ae Cre (eae
Wp, Dy pees
a sin(ae ) and ¢, =A
= . (7)
a sin (max
Assume / > m.
| bibm d ee 2 [si n (22) sinn (22) le
Ie Di. abe = =e
= [cos (1 m) as
= — cos (i+ m)==| dx
re
le sin (« — mB) - (Ea
eeae
ae sin (« + m) ax)"
az
Since we now wish to consider the possibility that the particle can escape
from the box we shall require only that it be somewhere in space, that is be-
tween x = 0 and x = oo, rather than assuming, as previously, that it will be
confined between x = 0 and x = a,. If the particle is to be kept in the box, the
restraining potential V must be greater than E, the energy of the particle, and
therefore as x goes to infinity the term C,e-*”~"” goes to infinity. Clearly if the
condition of postulate III is to be met, C, must be zero and the acceptable
function becomes
ap =i Cee ee (13.58)
As we have seen earlier, if the particle is to be always in the box, vw as given
by equation (13.58) must equal 0 at the wall. This can only occur if the expo-
nent equals — co. Therefore, V must equal co. This is a result completely without
parallel in classical mechanics, where any potential greater than the kinetic
energy of the particle would confine it completely to a classically defined box.
However, it is a consequence of our formulation of
quantum mechanics that even when the energy of
the particle is less than the energy of the confining
potential, there is a finite probability that the par- wanes
ticle will be found outside any quantum mechanical
box with a wall of finite thickness. There is much
~*~ ‘box
—
experimental evidence in agreement with this re-
quirement of quantum mechanics. For example, in
nuclear decay q@ particles of energy lower than the
energy of the potential barrier of the nucleus have
been observed. These particles could not have sur-
mounted the barrier, because to do this their energy
must at least equal the potential energy of the FIG. 13.14. One wave
function for particle
barrier. Therefore, they are said to have “tunneled” NR ah ed |
through it. box with walls of
When the wall potential is finite, the solution of finite potential.
the wave equation consists of two parts. One part
describes the wave function within the field free
region of the box and consists of a sine or cosine function. The second is an
exponential function and describes the behavior of the particle under the
influence of the potential. These functions must join smoothly at the wall. One
such function is shown in Fig. 13.14.
potential at the walls is infinite. With these assumptions we can write the
appropriate wave equation.
oeie ~
he ee
OF 87° mE (13.59)
Next we shall assume that the function wy can be written as the product of
three functions, one dependent only on x, a second dependent only on y, and
a third dependent only on z. w(x, y, z) = X(x) Y(y) Z(z). Upon substitution for
ar equation (13.59) takes the form
yZ a?EXx + ALES
Ca he
+ XY OL,
SE A+ OT TE XZ
0
Dividing by XYZ, we find
2
ee a ey ee UP eRe
S2?mX Ox? | 8x2*mY dy? | 8n*mZ az? oe rey
The first term of equation (13.60) is a function of x only, the second term a
function of y only, the third term a function of z only, and the fourth term is
a constant. This equation must be true for all possible combinations of the values
of x, y, and z, since it represents an identity. (Note the contrast to the type of
algebraic equation which is solved only by specific values of variables.) The only
way that this can hold true is for the values of each of the terms to be constant,
that is, to be independent of the choice of particular values of x, y, and z. To
put it another way, we can choose a particular value of y and a particular
value of z. The second (y) and third (z) terms then become constant. Since the
fourth term is already constant, the first term becomes equal to a constant for
any value of x that we choose. Let us designate this constant as —F,. Then
We At
8x?m X Ox? = 2, (13.61)
As the result of similar arguments, we can equate the second term of equation
(13.60) to a constant —£, and the third term to a constant —E,. It is clear
from equation (13.60) that
E=E, +5, +E, (13.62)
Equation (13.61) can be rearranged to have the same form as equation (13.46),
the equation describing the particle in a one-dimensional box. Therefore, the
function X has the form
Exenn, ne
= x (5 co)
ny, n
(13.65)
Because this problem involves three variables, there are three quantum numbers.
These quantum numbers specify the components into which the kinetic energy
can be divided, each component being associated with one of the three axes.
The lowest energy level will be that for which n, = n, =n, = 1. (Recall that
n=O is not allowed.) For this level the energy is given by the relation
E, , ; = 3h?/8ma’. For the next higher level a possible set of quantum numbers
iS. 2, 1, and yn, == 1.and E,,, = 6h7/8ma*.. However, there are two
other sets of quantum numbers which will give the same energy: n, = l,
pie le n,n, == 2. Lheseasets of «quantum, numbers
correspond to different wave functions, X(x) Y(y) Z(z), and, therefore, to differ-
ent states of the particle in a box. Since there are three such states having the same
energy, this is said to be a friply or three-fold degenerate level. With this we
conclude our only detailed discussion of a wave mechanical problem. For the
rest of this chapter we shall content ourselves with a presentation of the results
of the application of wave mechanics to problems of chemical interest.
EXERCISE 13.12
What is the degeneracy of the level for which E = 14 h?/8ma?? Ans. 6 fold.
Hydrogen is the simplest atomic system that we know and the only
one for which the Schrédinger equation can provide a wave function without
resort to approximations. Consequently, it plays a key role in our discussion of
atomic structure. The hydrogen atom consists of a proton and an electron.
The potential energy V for the interaction of these two particles is given by
Coulomb’s law
ee = (13.66)
where Z represents the number of unit charges on the nucleus (one for hydrogen),
e represents the charge of the electron in electrostatic units, and r represents the
distance of separation of the two charges. Since the problem is three-dimensional,
three coordinates must be used to describe the position of the electron relative
to the proton. The form of Coulomb’s law indicates that polar coordinates
(r, and two angles, 6 and ¢) are a more convenient choice than Cartesian
380 Atomic Structure Chap. 13
coordinates (x, y, and z). The solution to the Schrédinger equation can then be
expressed as the product of three functions, one depending on r alone, R; one
on the angle 6 alone, O; and one on the angle ¢ alone, ®.
ir = ROD (13.67)
A few of these functions are given in Table 13.4. In this table a, is the Bohr
radius, and p = (2Z/na,)r where these symbols are defined in the text.
TABLE 13.4
SOME COMPONENT WAVE FUNCTIONS FOR HYDROGEN
(ZG.
n = 2 ; 1 == 1 R Wes
0.
pe p/2
— 0 arn, Oy = ve
m=
= (0) Dy
a= ol
2a
i= 1! OO = piled cos ¢
J1
m= -+1
10: im =0 L=1 <m=_.0
m= -—1
For hydrogen these levels have the same energy; that is, they are degenerate.
For n = 3 there are nine eigenfunctions:
m=+t2
m= -1 m=-+1
L=-0,-m=0 [sree (W) 2
m= —1 m= —!1
m= —2
We shall refer to single-particle wave functions which are exact or approximate
solutions of the Schrédinger equation as orbitals. Each of the eigenfunctions
mentioned above can thus be called an atomic hydrogen orbital.
The distribution of the electron in space is of particular interest to chemists.
As we have discussed earlier, the value of Wyr* dr for any small element of
space dt, gives the probability of the electrons occupying this element of space.
Since the wave functions for hydrogen are complex, we shall begin with a dis-
cussion of one component of this wave function, the radial function R. The
values of R® for some values of n and / are graphed as a function of r, the distance
from the nucleus in Fig. 13.15, where it can be seen that R® depends on two
quantum numbers, » and /. Contrast this to the particle in a box where the
corresponding functions X(x), Y(y), and Z(z) each depended on only one
quantum number.
8See C.W. Sherwin, Introduction to Quantum Mechanics, Holt, Rinehart, & Winston, Inc.,
New York (1959), p. 77.
°The limits on the integration over ¢ are 0 and z because the ring swept out by integrating
over 6 from 0 to 2z need only be rotated through z radians to generate a spherical shell.
Since the original functions are normalized for these limits, each integration evaluates to 1.
If 7%? is spherically symmetrical it will not contain ¢ and @ explicitly, and the result of this
integration can then be expressed in the alternate form, #4zr?dr. In this form 4zr2dr is
clearly recognizable as the volume of the spherical shell lying between r and r + dr, that is
4/3 xr? — 4/3x(r + dr)? = 4xrdr, since terms containing products of differentials are negligi-
bly small and may be dropped.
Seca gis Atomic Structure 383
FIG. 13.16. a) Graphs of O®@ for the hydrogen atom |!=0 and |=1 orbitals. b) Graphs of
[O®]}?, which together with R2 determines the electron density in any small volume of space [from
W.J. Moore, Physical Chemistry, 3rd ed., Prentice-Hall, Inc., Englewood Cliffs, N.J., 1962].
Zz
this general scheme. For example, in the spectra of the alkali metals, where the
scheme predicts that only one line, a singlet, should be expected, high resolu-
tion shows two closely spaced lines, doublets. Something other than the three
quantum numbers that we have so far discussed is required 1n order adequately
to categorize alkali metal electrons. An explanation of this occurrence of dou-
blets was proposed by G. E. Uhlenbeck and S. Goudsmit in 1925. They proposed
that the electron has an intrinsic spin, and therefore an intrinsic angular mo-
mentum which can either add to or subtract from the angular momentum
generated by the movement of the electron through space. About a selected
axis this spin angular momentum can have only two values, +4(h/27), and
—3(h/27). There is thus a fourth quantum number s, the spin quantum number,
which can have two possible values, +4 and —4. In order to specify the state
of an electron completely we must give not only the three quantum numbers
characterizing its orbital, but also must state whether the spin quantum number
is +4 or —3.
Since an electron possesses an angular momentum, it has associated with it
a magnetic dipole moment. Note the analogy here to a circulating electrical
current. Thus, any system with only one electron must be paramagnetic, that is,
Sec. 13.24 Atomic Structure 385
for more complex atoms the quantum number / is also important. As a first
approximation such atoms can be thought of as consisting of two parts, (1)
a core made up of the nucleus and all electron shells or subshells which are
completely full, and (2) the electron of interest. Assuming now that the orbitals
of all atoms are qualitatively similar to those of hydrogen having the same
quantum numbers n and / we see an important difference between hydrogen and
the other atoms. Consider the lithium atom, which has three protons in its
nucleus, and three electrons distributed about that nucleus. The first two of
these electrons will occupy the m = 1 level, the lowest energy level available.
The core of this atom thus consists of a three proton two electron system with
a net charge of +1. At the 7 = 2 level we have two types of orbitals available, s
orbitals and p orbitals. As Fig. 13.15 indicates, an electron in a 2s orbital
(J = 0) has a much higher probability of penetrating into this core than does an
electron in a 2p orbital (/ = 1); therefore, the average positive charge to which
a 2s electron is subjected is higher than that to which a 2p electron is subjected.
The potential energy term for the 2s electron is thus lower (more negative),
than that for the 2p electron. The third electron will, therefore, occupy the 2s
orbital, and the configuration of Li is 1s”, 2s'!. Similar considerations lead to the
conclusion that in general, for more complex atoms, orbitals of the same main
quantum number increase in energy as / increases.
The electron configuration of Be is written as Is’, 2s, for B as 1s’, 257, 2p',
and for C as Is’, 2s’, 2p’. Carbon poses two new problems. There are three
equivalent p orbitals. (1) Will the second electron enter the occupied p orbital
or a different p orbital? (2) If it enters a different p orbital, will its spin be the
same as, or different from, that of the first electron? The first question seems
easy to answer. Simple electrostatics tells us that the most stable arrangement
of two electrons is one which puts them as far apart as possible, thereby mini-
mizing the repulsive potential between them. Since the p orbitals occupy different
regions in space (Fig. 13.16), the most stable arrangement of the two electrons
will be that which places them in different orbitals. Now the electron spin must
be considered. When electrons are placed in different orbitals, the arrangement
in which they have parallel spins is energetically favored over that in which the
spins are opposed.'® Therefore in the lowest energy state the two p electrons of
carbon should occupy different orbitals and have the same spin. The atom is
thus paramagnetic with a paramagnetism corresponding to that for two electron
spins. States in which two electrons have the same spin are known as friplet
states, because they are triply degenerate. It should be clear to the student that
two of the three degenerate states have spin (+4, +4), and (—+4, —4). The
resultant spins are +1 and —1. The intermediate state of resultant spin 0 is
also possible (see Chapter 14).
We have seen above that when two electrons are to be distributed in two
equivalent atomic orbitals, the triplet state will be lower in energy. This is an
10Rlectrons of the same spin tend to keep apart, an effect described as being due to
spin correlation.
Secailice24. Atomic Structure 387
example of Hund’s first rule, which may be stated as follows: Electrons with
the same n and / values will occupy orbitals of different m values, and their spins
will be unpaired so far as possible.
In the elements N through Ne the rest of the 2p orbitals are occupied.
With Na the 3s orbital is occupied, and the elements after Na follow a pattern
similar to that of the first row elements ending with Ar, Is”, 25°, 2p, 3s”, 3p°.
The parallel between the periodicity of electronic configuration as given by
wave mechanics and the chemical properties of the elements is immediately
evident. The origin of the first three periods is clear, and certain types of elec-
tronic structure can be associated with distinctive chemical properties. For
example, after the first short “period” associated with n = 1, one s electron,
whether 2s, 3s, 45s, 5s, or 6s, evidently gives rise to alkali metal properties. On
the other hand the configuration s?, p® is that ofa rare gas.
In potassium and calcium the 4s orbitals are filled. However, beginning
with the next element, scandium, the 3d rather than the 4p shell begins to fill.
We thus enter the first group of transition elements. These elements are
characterized by multiple valency and by colored compounds. Both of these
properties are associated with the fact that the 3d and 4s levels are close in energy.
As a result, the number of electrons involved in bond formation can vary, and
electronic transitions can take place at an energy corresponding to that available
in a quantum of visible light.
The electronic structures of the transition elements cannot always be
predicted on the basis of the simple model that we have presented. The use of
hydrogen-like orbitals and the assumption of a definite order of energies which
persists from atom to atom for these orbitals leads to a greatly oversimplified
model. For example, the configuration of vanadium is 1s’, 25°, 2p°, 3s”, 3p°, 3d’,
4s", whereas that of chromium, the next element, is 1s’, 25°, 2p®, 35°, 3p°, 3d°, 4s".
This distribution is favored in chromium because it minimizes unfavorable
interactions due both to coulomb repulsion and to spin correlation effects
relative to the alternative configuration, Is’, 25°, 2p°, 3s”, 3p°, 3d‘, 45°, which
puts two electrons in the 4s orbital. Evidently the reduction in the magnitude
of the two unfavorable energy terms is sufficient to outweigh the “normal”
difference in energy between the two orbitals. This could not have been pre-
dicted in advance from our model. (Note that we have been talking here in
terms of a model which makes a useful but completely arbitrary division of the
energy of the electron into separate parts.)
After filling the 3d orbitals the 4p orbitals are filled in the usual way.
Then, after strontium, which has a completely filled 5s orbital, the 4d orbital is
filled, beginning a second transition series.
Here we must call attention to a fourth type of orbital with / = 3, designated
an f orbital. Since, for / = 3, m can have any integral value from +3 to —3,
there are seven f orbitals. These seven orbitals can accommodate fourteen
electrons. These 4f orbitals are evidently higher in energy than both the 5s and
6s orbitals and so do not begin to fill until after element 57, lanthanum, which
has the electron structure Is”, 2s”, 2p°, 3s”, 3p°, 3d°, 45°, 4p®, 4d?°, 5s°, 5p®, Sd’,
388 Atomic Structure Chap. 13
6s?. The next element, cerium, has the structure 1s’, 25°, 2p°, 35°, 3p°, 3d'°, 45°,
4p®, 4d’, 4f?, 5s”, 5p°, 6s”. The following elements in which the 4f shell is being
filled are called the lanthanides, or rare earths. When the 4f shell is completed,
there is a transition series in which the 5d shell is filled. Finally the 6p shell is
filled, and a rare gas configuration is reached.
In the last period the first two elements, francium and radium, have the
expected electron structure. The next two elements, actinium and thorium, have
electrons in the 6d shell. Then with protoactinium the Sf shell begins to fill.
Thus, we have an actinide series analagous to the lanthanide series, although
some of the elements do have 6d electrons. The 5f shell should be completed in
nobelium, 1s?,}2s?,.Qp°.93s7) 3p°, 347 4s"5 Ap ed Aj28) 552 Sp*. Od 2.0) 1 OSes
Gps s*.
A considerable amount of information regarding the relative energy states
of the various orbitals is summarized in Fig. 13.18.
SUMMARY, CHAPTER 13
Ov
O¢
09
Og
08
OL
OS
06
Ol
000‘O!
2: Nuclear composition
Atomic number, Z = no. of protons
Mass number, A = no. of neutrons + no. of protons
Isotopes—same Z, different A
Separation factor
_ [M,
QS ive for gaseous effusion
ieee
mv
=
sma.2
n
PROBLEMS, CHAPTER 13
\k A beam of protons is accelerated from rest to a velocity of 3 x 10° meters
Seca:
(a) What potential difference would be required to achieve this velocity ?
(b) If the proton beam is then passed through an electric field such as that
described in Exercise 13.1, what deflection (x) would result?
Ans. (a) 4.69 < 104 volts, (b) 5.3 « 107" meters.
3. A drop of oil of density 0.9 g.cm.~* and 10-*° cm. radius has one electronic
unit of charge.
(a) What will be the velocity of free fall in air? The viscosity of air is 1.87 x 10-4
poise.
(b) What electric field strength will be required to hold the drop stationary?
Ans. (a) 1.05 x 10;*'cm. sec:—, (b) 2.32 volts em.c*
5. A gaseous sample containing 99 mole per cent D, and | mole per cent H, is
admitted to a mass spectrometer through a pinhole leak. What will be the relative
intensities of the ions D} and H; assuming equal ionization efficiencies?
.6. The atomic weight of lead obtained from thorium ores is found to be 207.71.
Thorium decays to 2°8Pb. Assuming that the abundance ratio Pb: ?°°Pb: 7°'Pb
is the same as that in ordinary lead, find the isotopic composition of the lead in
thorium ore. Ans. ?°®Pb = 83 atom per cent.
7. The chlorine isotopes are to be fractionated by effusion of HCI gas through
a system of 20 stages. Starting with the natural mixture, what is the maximum
per cent of **°Cl which can be expected? Ans. 84 per cent.
8. From data in Table 13.3 estimate the minimum number of separation stages
required to enrich natural nitrogen to °N = 10% by exchange of NH; (g) with
NH}? (aq.). Ans. 148 stages.
392 Atomic Structure Chap. 13
19. For the ground state of the H atom the radial distribution function is
almost symmetric about a), the Bohr radius, and the uncertainty in electron
position can be approximated as a). (See Fig. 13.15.) An alternate statement
of the uncertainty principle is Ap Aq = A/2x. Use this to show that T = 27?
me'/h?, where T is the average kinetic energy of the electron. Compare to the
Bohr model.
20. Assume an electron trapped in a one-dimensional box of the dimensions of
a typical small nucleus (-~107!% cm.). Calculate the energy of the lowest allowed
level in electron volts. Discuss the possibility of electrons being bound in nuclei.
(The energies of 8 particles emitted from nuclei are seldom more than 10 “e.v.)
Problems, Chapter 13 Atomic Structure 393
(22. Although positive and negative electrons eventually annihilate each other
with resultant production of energy equivalent to their masses, they can form
a short-lived complex called positronium. Assume a model for this complex
similar to the Bohr model for the hydrogen atom.
(a) Compute the energy difference between the ground state and the first
excited state. (Use of the electron mass, m,, in our development of the Bohr model,
was an approximation. Properly the reduced mass, 4, must be used. For this
problem » = m.m,/(m, + m,), where m, is the mass of the positive electron
(positron).)
(b) Compute the energy required to completely separate the positron and the
electron.
(c) Repeat (b) for mesonium, a proton and a negative {4 meson of mass 207
times that of the electron.
23. The 1s wave function for the hydrogen atom is
1
Vivo a mane
me ?.q,7/
(24. For / = 0 the radial part of the Schrodinger equation for hydrogen reduces to
AR eGR women
ai eae ae cee
The n = 1 eigenfunction has the form
JR = IN| Qe
\25. Show that for the ground state of the hydrogen atom the quantum mechanical
treatment requires that the most probable value of the radius, r, is a,, the Bohr
radius.
ah Sem
Aa OObg
where ¢@ is the angular coordinate. Assume that the potential energy is zero and
show the following:
; s Potential
energy 1s assumed to provide the energy
best description of the system, (electron volts)
=e
There are various systematic pro-
cedures for choosing initial forms
for these approximate wave func- -
tions. Later in this chapter sim-
plified versions of two will be =6
discussed, the molecular orbital
method, and the valence bond e ? me Pe. See ze
Internuclear separation (A)
method.
As we discuss these proce-
FIG. 14.1. Potential energy of NatCl-.
dures the student should keep
clearly in mind that one is not
“right” and the other “wrong.” They are merely different approximations
which emphasize different aspects of the problem. In some problems one of
these methods is often found to provide a simpler and more accurate descrip-
tion than the other. On these grounds it is therefore preferred for that
particular problem.
In the last section it has been pointed out that the problem of bonding
in compounds such as NaCl could be treated through wave mechanics. However,
there is a simpler approach to an adequate description of this type of bonding.
In fact, historically the treatment of this problem antedated wave mechanics.
The clue to a satisfactory model was provided by observations such as that of
electrical conduction by molten NaCl. Such electrical conduction implied the
existence of the charged particles Na* and Cl- in molten NaCl and, by extension,
in crystalline NaCl as well.
In the most simple model Nat and Cl~ ions may be treated as point
particles of charge +e and —e. The potential energy decrease when these ions
are brought to the internuclear separatic’ , which is 2.5 x 10°§cm. for
gaseous NaCl, may then be calculated by using the coulombic relation:
a= re [rn (4.80 x 105s)? /O-54-X 10°-*cm.) = =917 x 10°” ergs
O17 10 eres) (6.24 =< 10™e.v. ere-+) = —5.73 en.
This simple model thus indicates strong bonding between Na* and Cl.
According to the coulombic law, the most energetically favorable arrange-
ment of positive and negative point charges would be that in which they are
superimposed. However, real ions are not point charges. They consist of a
positively charged core surrounded by mobile electrons. As the ions, drawn by
the coulombic attraction of their net opposite charges, come close together,
repulsion between the groups of electrons belonging to each ion begins to be-
come important. This repulsion shows an r~” dependence where n is in the
range from 6 to 12. The interaction between coulombic attraction and this
short-range repulsive force results in a minimum in the potential energy curve
as shown in Fig. 14.1. This topic is discussed further in Chapter 15.
395
396 Molecular Structure Chap. 14
As with any model this simple electrostatic model for bonding in NaCl
must be compared to experimental observation. The prediction of a 5.73 e.v.
binding energy between Na* and Cl in the gas phase cannot be directly verified.
However, the basic assumptions of coulombic (r~') energy terms corresponding
to attractions and repulsions, and of strong short-range (r~”) energy terms corre-
sponding to repulsion can be used to calculate the energy evolved in the for-
mation of the NaCl crystal lattice from gaseous Na* and Cl- ions. The cal-
culated value, 7.80 e.v., is to be compared to an experimental value of 8.02 e.v.
derived by the use of the Born-Haber cycle (Chapter 15). The model is evidently
quite satisfactory.
On the basis of this simple electrostatic model the bond energy of gaseous
NatCl- may be estimated. The reaction of interest is Na(g) + Cl(g) — Na*Cl-
(g). For convenience in calculation, the energy of this reaction may be con-
sidered to be the sum of the energies of three individual processes: (1) the energy
required to remove an electron from a sodium atom and convert it to a sodium
ion, measured by the ionization potential I of sodium; (2) the energy evolved
when an electron is attached to the chlorine atom to form the chloride ion,
measured by the electron affinity A of chlorine; (3) the energy evolved when
the two ions interact, calculated, as we have seen, by the coulombic law.
TABLE 14.1
ELECTRON AFFINITIES
Element A (e.v.)
18 3.448 + 0.005!
Cl 3.613 + 0.003!
Br 3.363 + 0.003!
I 3.063 + 0.003!
O 1.465 + 0.005?
S 2.07 + 0.072
Consider coulombic attraction and repulsions and show that the energy of the
array described in the preceding paragraph is —41 e.v.
Despite this favorable potential energy, this array is, of course, unrealistic as
a model for Ht where the negative particle is an electron. However, we shall
explore it further by asking how it might be made more realistic. The answer to
this question comes from wave mechanics. The basic difficulty with the proposed
electrostatic array as a model for H} is that the electron must be represented
by a highly localized wave function. This results in a very high calculated
398 Molecular Structure Chap. 14
kinetic energy (Exercise 14.2), the total energy of the system is very high, and
the model system is actually unstable.
EXERCISE 14.2
mate wave functions must be used. These approximate wave functions are
often synthesized from atomic wave functions, and in the next two sections we
shall examine two such methods of synthesis. The results obtained with any
wave function are in part a function of the method of synthesis employed, and,
therefore, the student should be wary of attaching too much physical signi-
ficance to these results.
state, and it is called an antibonding orbital. Since the molecular orbital functions
were constructed from a Linear Combination of Atomic Orbitals, this is known
as the L.C.A.O.-M.O. approach.
If we turn now to the problem of the hydrogen molecule, the molecular
orbitals should be the same as for the hydrogen molecule ion. H, has two
electrons, and the Pauli principle allows the placing of both in the bonding
orbital so long as their spins are opposed. Therefore, the wave function for this
system would be
[yu) + Wn) fru) + W(2))
The L.C.A.O. method can also be applied to the hydrogen orbitals of main
quantum number 2. The 2s atomic orbitals combine to give both a bonding and
an anti-bonding orbital. These are designated as o2s and o*2s orbitals. The
Pz atomic orbitals, which we shall designate as those extending along the
internuclear axis, also combine to give molecular orbitals of o symmetry,
o2p,,and o*2p,. In contrast, the combination of two 2p, or 2p, orbitals gives rise
to molecular orbitals which have a node in a plane passing through the two
nuclei. These are called z orbitals. Thus we have the bonding orbitals, 2p,
and z2p., and the anti-bonding orbitals, 7*2p, and 2*2p., which have spatial
characteristics quite different from those of the o orbitals. Schematic diagrams
of these orbitals are given in Fig. 14.2. The theoretical treatment also indicates
the relative energy of the various orbitals to be in the order ols < o*ls <
CLS AUOSt ODD. = ROD pt p< 2) = LPO lp y.
With this information it is possible to predict the electronic configuration
of some simple diatomic molecules, two electrons of opposite spins being placed
in each orbital, according to the Pauli principle, as we have seen. H, in its
lowest electronic energy state ({| in Fig. 14.5) will be expected to have two
electrons in the orbital of lowest energy. This configuration is abbreviated
H,(ols)?. The first excited state (ff in Fig. 14.5) must be H,(cls)!(o*ls)',
having one bonding and one antibonding electron. The effects of these two
electrons cancel, and no bond exists. The hypothetical molecule He, would
have the configuration (cls)?(o*ls)’, that is, two bonding electrons and two
antibonding electrons, and again no significant net attraction is observed. For
the succeeding diatomic molecules it is presumed that the K shell of the atoms
is not involved in the bond, and only the molecular orbitals arising from the
second shell are specified. Thus for three of the well-known gaseous diatomic
elements we have,
N,: KK(o2s)? (¢*2s)? (o2p)? (2p)!
O,: KK(o2s)? (c*2s)? (o2p)? (22p)! (x*2p)?
F,: KK(o2s)? (c*2s)? (c2p)? (x2p)! (x *2p)*
There are six net bonding electrons in N, which can be said to constitute a triple
bond. However, one pair occupies a o orbital, and the others occupy z orbitals.
In O, there are only four net bonding electrons, two o and two zw and O, is
doubly bonded. In both of these cases of multiple bond formation, the
molecular orbital treatment indicates that the several bonds are not equivalent.
~
Sec. 14.4 Molecular Structure 401
2 Py 2 Py LA Le a) Ta
Te 2p,
A Vag = 0*/tgp 8
Sec. 14.5 Molecular Structure 403
Ca 9 ES ean anaa
yi Oz? 0x3 Oy5 0z3
8x*m (z tee es ree eee:
— + — 4+ — 4 = — = _ e=(() 14.1
- = += } Le:y) et rp lio ane, ( )
Equation (14.1) cannot be solved exactly for all possible internuclear distances.
However, at large separations the system described by this equation must be
two normal hydrogen atoms. We shall now verify by direct substitution into
equation (14.1) that at large separations the function [yr,(2) y,(1)] assigning
electron | to nucleus B and electron 2 to nucleus A is a solution if the potential
energy terms involving r4,, go, 12, and r,, are all assumed to be small. These
are valid assumptions with this electron assignment at large separations of A
and B. Substituting [y,(2) W,(1)] into equation (14.1), we have
Seale eat echt) 3 etsy hee
xy
a Ee geeitieay
s saa ie+ & + Z) yes) = 0
(14.2)
Recalling that »r,(2) contains only the coordinates of electron 2 and is there-
fore constant with respect to variations in x,, y,, and z,; that the same is true
with respect to yr,(1) and the coordinates of electron 2; and that £, the total
energy of the two-atom system must equal E, + E,, the sum of the energies
of atoms A and B, we rewrite equation (14.2) as
spi) |2h
AD) Spl) 4.Papel) cL ee2 (1)
+ roll) Eze
Bg) OI) OPI) 4 BA (Ey + H)ap4(2)|=0
(14.3)
The terms in brackets are simply wave equations for hydrogen atoms and are,
therefore, equal to 0. Thus, we have an identity and [yr,(2) v,(1)] is indeed a
solution of equation (14.1) at large separations.
However, this is not the only solution. An argument similar to that which
we have just given shows that [y-,(1) w,(2)] is also an acceptable solution to
equation (14.1) at large separations, and so are the sum function [yr,(2) y,(1)
+ r4(1) YWr,(2)] and the difference function [yr4(2) wx(1) — yu) Wz (2)].
EXERCISE 14.4
Verify that [Yru(1) Wa(2)], [YuQ) Wa) + Yu) We], and Yr) Wa) — Yu)
af,(2)] are all solutions of equation (14.1) at large separations.
We now have four exact solutions for the system of two hydrogen atoms
at infinite separation [Wu(1) W2(2)], ru) a), Para) + ra rn),
and [vr,(2) W,(1) — Yru(1) Wrz(2)]. These four functions can be tested as solutions
404 Molecular Structure Chap. 14
of the wave equation at finite separations. It can be shown [W. Heitler and
F. London, Z. Physik., 44, 455 (1927)] that each ofthe first three leads to a build-
up of electron density in the internuclear region, and to a lowering of potential
energy relative to the separated atoms. The first two give bonding energies of
only 0.4 e.v., whereas the function [Wr4(2) (1) + Yr) r,(2)] gives a maximum
bonding energy of 3.14 e.v.! (Fig. 14.5). The experimental binding energy is
4.75 e.v., and that calculated from the simple molecular orbital approach is
2.68 e€.v.
That the function [yr,(2),(1) + Yr)
alr,(2)] is superior to the functions [yr,(1)
ar,(2)] and [vr,(2)r,(1)] is a consequence
of the fact that it is the only function of
the three which takes into account the
basic fact that the two electrons must be
indistinguishable, and each must therefore
be distributed about both nuclei in the
Energy
electron volts) same way. Since in the model which we
are considering this was done by the device
of exchanging electrons (1) and (2) between
nuclei A and B the energy difference, 2.74
ev., between [yr(2)yr,(1) + aD v2(2)]
and [yr4(1)W,(2)], is often referred to as an
exchange energy. It should be clear from
the discussion that exchange energy is not
0 1 2 3 4 inherent in nature, but arises in the Heitler-
Internucleor separation (A) London treatment as a result of the par-
ticular method chosen for the synthesis of
FIG. 14.5. Binding energy of the approximate wave function. The differ-
ve ie eoc rete! ence function [y,(2hb,(1) — Wru(1r,(2)]
experimental
has not yet been considered. This function
shows a steady increase in energy as inter-
nuclear separation decreases (Fig. 14.5). It
therefore represents an unbound state. Finally the question of the electron spins
in the two states [r4(2yra(1) + Wu(lyrx(2)] and fyra(2yra(1) — ra(Drra(2)]
must be considered. As we have mentioned earlier, only two spin states are
possible for an electron. We shall designate these as a and 8. For a two-
electron system two possible combined spin functions are [a(2)a(1)], and
[8(2)8(1)], in which each electron has the same spin function. By analogy to
the Heitler-London treatment of wave functions, two other spin functions are
TABLE 14.2
COMPLETE HEITLER-LONDON WAVE FUNCTION FOR Ho.
=
1. | Fy oso) + Tz ba wa ||Fy e@AW) — Fy (DE
1 ail
ee es Q)|
2. [Fy babu) — Fy kaya) |[Deca]
3. |ay baal) — Ty 800) || @<a |
(os
TABLE 14.3
HYBRID ATOMIC ORBITALS
pounds are about 120 deg., one possible description begins with the assignment of
sp? hybridization to the carbon atoms. This leaves a p orbital projecting at
right angles to the plane of the hybridized orbitals. An orbital bonding the
two C atoms can then be constructed by combining one sp? orbital from each
atom. This orbital has o symmetry, and is therefore referred to as a o bond.
Combination of the two atomic p orbitals leads to a second bond of z-type
symmetry (Fig. 14.7).
Although the « — z description of the double bond is the more familiar,
109°28'
109°28'
benzene. The delocalized orbitals are lower in energy (more stable) than the
localized ethylenic-like orbitals of the Kekule structure, and the energy differ-
ence between them is called the delocalization energy. The model meets the
requirement that all the C — C bonds be equivalent, as evidenced by the fact
that they all have the same length, 1.39 A, intermediate between the C — C bond
length in ethane, 1.54 A, and the C = C bond length in ethylene, 1.30 A. The
general treatment of delocalized electrons represents one of the great successes
of the molecular orbital theory. :
The valence bond model also provides a treatment of delocalized electron
systems through Pauling’s very fruitful concept of resonance. According to this
proposal, if a molecule can be represented by several classical valence bond
structures (1) of roughly the same energy, (2) differing only in the position of
their electrons, and (3) containing the same number of paired and unpaired
electrons, then a wave function x describing this molecule can be made up of
a linear combination of the wave functions Wy, Wo, W3..., describing the
possible valence bond structures.
The coefficients, c,, Cs, C3, etc., are chosen to minimize the total energy cor-
responding to the wave function yr.
Consider benzene from this viewpoint. Two obvious valence bond struc-
tures to be considered are the Kekule structures, I and II. (See diagram below.)
Other forms, such as the Dewar structures, III, IV, and V, have a higher energy
content and, therefore, would have less weight in the final hybrid wave function.
Still other structures can be imagined, but they are of even higher energy
content and therefore of lesser importance in the final hybrid.
The final hybrid wave function con-
structed from the wave functions for
structures I, II, III, 1V, and V describes
the electrons as being extensively delo-
calized, since the various structures con-
I I ping W Vv
sidered place these electrons between
different nuclei. In a qualitative sense,
what the molecular orbital treatment
seeks to achieve by a combination of atomic wave functions, the resonance
treatment seeks to achieve by a combination of “molecular” (valence bond)
wave functions. It should be clear from this discussion that the resonance
model does not imply that some molecules of benzene have one structure and
some another, nor does it imply that at one time the properties of the mole-
cule are those of structure I, at another time those of structure II, etc. Rather,
the attempt is to synthesize a single wave function which provides some basis
for an understanding of the properties of the benzene molecule as these are
revealed by experiment.
Sec. 14.8 Molecular Structure 411
Ty + Ay > Iy + Ax (14.5)
it is seen that 7+ A is a measure of electronegativity. Since both J and A
increase in going from the beginning to the end of a period, it is evident that
the electronegativity will similarly increase.
The normalized relation
i+53 A eV: 14.6
(14.6)
Be
@
Us}
Al Si
@ @
1S 18
|
TiiGe As
e| @ @
iusy| hs} 20)
|
|
Zr|Sn Sb | Te
@| @@ @
141718 |21 25
| ae
“ transition elements
EXERCISE 14.5
Use the ionization potential (Fig. 13.11) and the electron affinity (Table 14.1) of
chlorine to estimate its electronegativity. Compare with Fig. 14.10.
The observed bond dissociation energy of HCl is 4.43 e.v., and the difference
Dao. — Dac. = 4.4 — 3.5 = 0.9 ev. (14.7)
Sec. 14.8 Molecular Structure 413
is taken as a measure of the ionic character of the bond, and the electronega-
tivity difference between hydrogen and chlorine.
Using an empirical method of computing electronegativity difference, and
arbitrarily adjusting the scale so that the
elements in the first period range from 2.5
to 4.0, the values indicated in Fig. 14.10
have been obtained by Pauling. Note that
electronegativity tends to increase toward
the upper right-hand corner of the periodic
table.
The electronegativity difference be-
tween the atoms of a diatomic molecule
is a measure of the degree of ionic char-
acter in the bond. Pauling has given an
ionic
%
character
approximate relation, which is shown in
Fig. 14.11. From this figure it may be
deduced that HI, with an electronegativity
difference of 0.4e.v., has approximately
5 per cent ionic character in its bond,
whereas HF, with an _ electronegativity Electronegativity difference
difference of 1.9 e.v., has 55 per cent ionic
character, that is, in HI the shared electron FIG. 14.11. lonic character vs.
electronegativity difference (Paul-
pair is almost equally distributed around ing).
the two nuclei, whereas in HF the shared
pair strongly favors the fluorine.
EXERCISE 14.6 >
Use the table of electronegativity values and Fig. 14.11 to estimate the per cent
ionic character in the O—H bond of water. Ans. 35 per cent.
414
Sec. 14.9 Molecular Structure 415
I- < Br- < Cl-< OH- < OAc~ < F- < C,H,OH < H,O < NH,
< ethylenediamine < NO,” < CN™
(b) dy2_y2
(a) day
Dy2—y2 Ge
a A
: _————ee
a)
Apy1 ez» Ye
hey, Obs Of
iS
~ Aa
5 = ise eae
fou ——_—
-
ic
—
SE
ieee
D2 y21%G2
Energy levels
dy2 —y?
orbitals. The final electron arrangement represents a balance between these two
factors. Consider a weak ligand field. The splitting A between the two sets
of d orbitals is small (Fig. 14.13). As a result, factor (2) above dominates, and
the available electrons are placed in both high- and low-energy d orbitals in
accord with the rule of maximum multiplicity. The result is a “spin-free” com-
plex with a maximum number of unpaired electrons. Such a complex is highly
paramagnetic. An example would be FeF;* in which the central Fe** ion has
five unpaired electrons. If the ligand field is strong, then the splitting A between
the two sets of orbitals is great. As a result, factor (1) dominates, and the lower-
energy d orbitals must be filled in accord with the rule of maximum multiplicity
before any electrons are placed in the high-energy orbitals. The result is a
maximum amount of spin pairing for a given number of electrons. Such com-
plexes are called “spin-paired.” An example is Fe(CN)5°, in which the central
Fe** ion has only one unpaired electron.
EXERCISE 14.7
Predict the number of unpaired electrons for
Mn(CN)s5*, Mn(H,0)??, Mn(CN)5*, Mn(H,0)*3.
Assume that H,O provides a weak ligand field. Ans. 1,5,2,4.
G2 _y2
5 ———
a ny
cc
uw
Nez Vez
a2
They are now the lower-energy orbitals. Thus, we now have one low-energy
group of two orbitals and one high-energy group of three orbitals (Fig. 14.14a).
If four alternating ligands are removed from the cube we have discussed above,
the result is tetrahedral arrangement. However, the separation of the d orbitals
into a low-energy group of two (Fig. 14.14a) and a high-energy group of three
still holds.
The other important arrangement of ligands often encountered in complex
ions is the square planar arrangement. Despite the fact that this can be de-
rived from the octahedral arrangement by simply removing two ligands,
there is no simple way of arriving at the order of the energies of the d
orbitals. However, it is clear that one of them (say, the d,._,,.) will be
directed at the ligands and will be a very high-
EERE Ay i OSI a energy orbital. A possible energy ordering is
tion of d-p multiple bonding shown in Fig. 14.14b.
in a cyano complex. The original assumption that the bonding in
these complex ions is electrostatic cannot be
correct. There must be a substantial amount of
covalent character. This can be introduced by
adopting the idea that the positive charge on the
central ion distorts or “polarizes” electron clouds
of the ligands towards itself. This introduces to
some degree a sharing of electrons. In addition,
it is possible for multiple bonding to take place
by way of an added interaction between p-type
orbitals in the ligand and d orbitals in the central
atom (Fig. 14.15). Another way of treating bond-
Sec. 14.10 Molecular Structure 419
ing in complex ions is by use of the molecular orbital method. This is dis-
cussed in advanced texts.
H,. The dielectric constant of the medium is due to both permanent and
induced dipoles in the molecule. In order to measure the permanent dipole
moment, the effect of the induced moment must be evaluated.
The dipole moment induced by an electric field acting on a molecule is
proportional to the field strength and the proportionality constant is called the
distortion polarizability ay. In the absence of permanent dipoles, it can be
shown that the dielectric constant is related to the distortion polarizability by
the Clausius-Mosotti equation
e—1
ao 4nd,
ees (14.10)
where n is the number of molecules per cc. Multiplying both sides of equation
(14.10) by the ratio of molecular weight to density M/p yields
e—1 M _
4xnMa,_ 4
nN, == 12 y, (14.11)
eae ats
where P,, is the molar polarization.
When the medium contains permanent dipoles, these will contribute to the
molar polarizability and the expression for P,, must then contain a term which
accounts for this contribution. In contrast to the induced polarization term,
this term is temperature dependent, decreasing as temperature increases. This
dependence arises from the fact
that the molecules of the medium
FIG. 14.17. Test of the Debye equation [from are in continuous random motion;
W.J. Moore, Physical Chemistry, 2d ed., Prentice- their collisions with other molecules
Hall, Inc., Englewood Cliffs, N.J., 1956]. tend to destroy the orientation
Note that the temperature appears in the second term on the right-hand side
of equation (14.12). Fig. 14.17 shows a test of this equation for HI, HBr, and
HCl, which have permanent dipole moments increasing in that order. The
dipole moment is obtained from the slopes of the curves shown. Some values
of dipole moments at room temperature are given in Table 14.4.
EXERCISE 14.8
Determine the slope of the line for HCl in Fig. 14.17 and obtain the dipole
moment of HCl.
The dipole moment often gives considerable insight into the structure of
a molecule. It is a vector property, and in polyatomic molecules the net dipole
moment will be the vector sum of the dipoles associated with each bond.
For example, the CO, molecule seems to have no dipole moment despite the
difference in electronegativity between carbon and
oxygen. This is consistent with a linear arrange-
ment of the atoms in which the two bond moments
cancel. On the other hand, the dipole moment of
H,O is 1.82 debye, indicating a non-linear con-
figuration. From spectroscopic observations (see
section 14.16) it is known that the H— O—H
bond angle is 105 deg. A simple vector diagram
(at right) allows calculation of the electric moment
associated with the O — H bond, 1.49 debye.
TABLE 14.4
DIPOLE MOMENTS OF SOME MOLECULES IN THE GAS PHASE*
*Taken from A. L. McClellan, Tables of Experimental Dipole Moments (W. H. Freeman and
Co., San Francisco, 1963).
+Value of C. T. Zahn, Phys. Rev., 27, 455-9 (1926), and F. G. Keyes and J. G. Kirkwood,
Phys. Rev., 36, 754 (1930).
magnetic moments, like dipole moments, are resultant properties. Thus, for
many molecules the intrinsic magnetic moment is zero. However, a magnetic
field always induces in all molecules a magnetic moment which opposes it and
tends to reduce it, arising from polarization of the molecule by the imposed
magnetic field. Substances which have only induced magnetic moments are
repelled by an inhomogeneous field and are said to be diamagnetic.
Molecules? which have a permanent magnetic moment tend to orient parallel
to an applied magnetic field and therefore to reinforce it. Such substances are
attracted into an inhomogeneous magnetic field and are said to be paramag-
netic. Paramagnetism is often large in magnitude compared to diamagnetism,
and is temperature-dependent, whereas diamagnetic polarization is temperature-
independent.
The measure of response to a magnetic field is the magnetic susceptibility
x, which is negative for diamagnetic substances and positive for paramagnetic
substances. The magnetic susceptibility is measured by suspending a sample
partly in the field of a strong electromagnet. A diamagnetic substance tends to
be repelled and a paramagnetic substance attracted, and a measurement of the
force of attraction or repulsion can be translated into a value of the magnetic
susceptibility. An equation which is similar in form to equation (14.12), and
which relates the magnetic susceptibility, the magnetic polarizability, a), and
the permanent magnetic moment /1),, has been derived on the assumption that
each magnetic dipole is completely independent of all others.
since [1;,/3k is a constant for a given substance. This is called the Curie equation
after Pierre Curie. Many solid materials do not obey this law, but instead.
follow a law of the form
ar GC:
Xpara — iie A’
»the Curie-Weiss law. A is known as the Weiss constant after Pierre Weiss.
An equation of this form has been derived on the assumption that the various
magnetic dipoles of a substance do actually influence each other, in contrast
*Although we use the term “‘molecule’’ in this discussion, the discussion applies to
radicals and atoms as well.
Sec. 14.11 Molecular Structure 423
to the assumption on which (14.13) is based. The Weiss constant can thus be
thought of as taking into account interactions between species. In some cases,
however, this simple interpretation is known to be incorrect and the Curie-
Weiss law is inapplicable.
As we mentioned above, classically magnetic moments can be associated
with circulating electrical charges. Therefore, it should not be surprising that
there is a relationship between the resultant angular momentum of the elec-
trons of an atom or molecule and its magnetic moment. Electrons have both
an orbital angular momentum (related to the quantum number /), and a spin
angular momentum (related to the quantum number s). Both of these can
contribute to the observed magnetic moment, but we shall consider only cases
where the spin contribution alone effectively decides the magnetic moment.
The magnetic moment associated with one electron spin is 1.73 Bohr magnetons.
(The Bohr magneton = he/4amc or 9.273 < 10°?! erg gauss~!.) The resultant
magnetic moment for n unpaired electrons which occupy several different
orbitals is
[ly = /n(n + 2) Bohr magnetons (14.14)
Thus, the magnetic moment of the FeF;* complex to which we referred earlier
would be 5.92 Bohr magnetons, corresponding to five unpaired electrons, where-
as that of the Fe(CN);* complex should be 1.73 Bohr magnetons corresponding
to one unpaired electron. It is this ability to relate paramagnetic properties to
electronic structures that makes the measurement of bulk paramagnetism one of
the most valuable tools available for the study of transition metal compounds.
(Some typical experimental results are given in Table 14.5.)
TABLE 14.5
PARAMAGNETIC MOMENTS OF SOME TRANSITION METAL IONS
No. of Calculated
Configu- unpaired moment Observed
Jon ration electrons (spin only) moment
Sct d° 0) 0
Tit? d 2 2.84 2.76
Vit? a 3 3.87 3.86
(Cee d‘ 4 4.90 4.80
Mn?*? d° 3) 5:92 5.96
Fet? d6 4 4.90 5.0 —5.5
Cot? da 3 3.87 4.4 —5.2
Nit? d® 22 2.84 2.9 —3.4
Cut a 1 LB 1.8 —2.2
Znt? GES 0 0 0
bulk magnetic properties. We shall defer our brief discussion of this technique
until after the section on nuclear magnetic resonance, since in many respects
the two techniques are analogous.
Ey = —8UnM,H, (14.15)
Therefore, the M = +4 sublevel is lowered in energy by 4 g4,H, while
the M = —4 sublevel is raised in energy by the same amount (Fig. 14.18).
[ln = (e/2m,c)(h/27), where e is the charge on the proton, m, is the proton
mass, c is the velocity of light, and A is Planck’s constant. Since g for the
proton is a constant, equation (14.15) can be rewritten in the form
h
Foc ay (5) MH, (14.16)
where y represents a proportionality constant.
Transitions from the M = +4 to the M = —4 sublevel shown in Fig. 14.18
involve a reorientation of the nucleus with respect to the lines of force of the
applied magnetic field. The frequency associated with the energy of such transi-
tions can be computed from equation (14.16).
=— hy = yhH,
— Eas ae Eft
AE
2n
and
Segoe
_ YH, (14.17)
In a field of 10‘ gauss, for the hydrogen nucleus vy = 42.57 Mc. sec.~'. Thus,
the transition corresponds to a frequency in the short-wave radio band.
If now both radio-frequency energy and a magnetic field are simultaneously
applied to the nucleus, transitions from the lower to the higher level should
occur when the condition set by equation (14.17) is met. The system is then said
to be in resonance, hence the name nuclear magnetic resonance (n.m.r.). Since
the probability of absorption of radiation by nuclei in the lower sublevel is
equal to the probability of emission of energy by nuclei in the upper sublevel,
the net absorption of energy depends on there being an excess population of
nuclei in the lower level. At a field of 10,000 gauss the separation of the two
sublevels is only 1.76 x 10-* e.v. (equivalent to 4.06 x 10~° kcal. mole’).
426 Molecular Structure Chap. 14
magnetic field at the nucleus, H,, will be the sum of the applied magnetic field
H and the local magnetic field, H,,...
A, — ef = 15bes
FIG. 14.20. 60 mc. n.m.r. spectra of ethyl alcohol (courtesy of Dr. D.J. Pasto, University of
Notre Dame). Peak positions are given relative to the tetramethylsilane standard and
are in cps. No value is given for the OH proton since the position of this peak is variable.
In (a) (above) a trace of acid was present to catalyze exchange of the hydroxyl hydrogen.
Since this exchange is rapid on the n.m.r. time scale, the environment is averaged out for the
OH hydrogen, and the OH peak does not show splitting by the CH»y hydrogens. In (b) no acid is
present, the exchange rate is much slower, and the OH peak is a triplet.
428 Molecular Structure Chap. 14
nucleus can either add to or subtract from the applied field, depending upon
which of the two spin states the perturbing nucleus occupies. Since the two
spin states are quite close in energy, there will be an equal number of nuclei
in each of the two possible states. As a result, the resonance due to the nucleus
of interest will show two peaks of equal intensity whose field separation, AH,
is given by the relation AH = 2 H,,,.. Proton spectra in solids show this kind
of behavior. Characteristically, peaks are broad, ~1 gauss wide, since a given
nucleus interacts with many other nuclei.
In liquids the frequency of reorientation of molecules due to their natural
motions (~10!° sec.~!) is more than 10? times higher than the frequency of
nuclear resonance. As a result, a molecule containing a given nucleus can
reorient itself many times during a resonance absorption. The local fields due
to other magnetic nuclei are thus averaged out, and absorption lines become
much sharper, of the order of fractions of a milligauss. The sharpness of these
lines allows for precise experimentation, and as a result two very important
effects can be observed.
Paired electrons are diamagnetic; that is to say, a magnetic field induces a
moment in a paired electron system in a direction opposed to its own moment.
This induced moment is directly proportional to the strength of the applied
field. Thus, both bonding and nonbonding electrons in the vicinity of a nucleus
reduce the magnetic field at that nucleus by an amount which depends on their
distribution about that nucleus. Since these electrons in turn are influenced by
their environment, n.m.r. is a very sensitive probe for distinguishing two or more
protons in different locations in a molecule. Thus, ethanol, which contains
different kinds of hydrogen atoms, shows three groups of peaks at high resolu-
tion (Fig. 14.20). The areas of these groups of peaks are in the ratio 3 to 2 to 1,
corresponding to the number of hydrogen atoms at each location. This pheno-
menon is called the chemical shift.
Chemical shifts are reported in terms of a chemical shift parameter 6
which is expressed in terms of a reference substance.
__ H(sample) — H (reference) ;
: H (reference) 7e10
(14.18)
In equation (14.18) H refers to the applied field strength at which a given peak is
observed. A common proton reference standard is tetramethylsilane, which is
usually mixed with the sample whose spectrum is to be determined.
As Fig. 14.20 shows, groups of peaks rather than single peaks can be
observed for each type of H atom. This so-called splitting is due to the magnetic
moment of the adjacent nuclei in the same molecule. Consider the protons of
the CH, group, which we shall designate as A and B. There are four possible
ways that these protons can align themselves relative to a given proton in the
CH, group.
(1) Spins of A and B parallel to that of the proton.
(2) Spins of A and B antiparallel to that of the proton.
(3) Spin of A parallel and of B antiparallel to that of the proton.
(4) Spin of A antiparallel and of B parallel to that of the proton.
Seen 4as Molecular Structure 429
(3) and (4) are energetically equivalent. Therefore, we would expect the CH,
proton peak to be split into three components, with the middle component,
representing (3) plus (4), being twice the height of each of the other two. Figure
14.20 bears this out. In contrast to the chemical shift this spin-spin splitting
is not dependent on magnetic field strength, at least to a first approximation.
EXERCISE 14.9
Show that the methyl hydrogen atoms will split the CH, peak into four com-
ponents of statistical weights 1: 3: 3: 1. Why are there eight peaks in Fig. 14.20?
working at lower temperatures, where the relative population of the lower state
is increased. (The contribution of these unpaired electrons to bulk paramagnetic
behavior also depends on the excess population of the lower state, and this is
reflected in the temperature dependence of paramagnetism. Recall equation
(14.13).)
The advantages of electron spin resonance (e.s.r.) are several. It is extremely
sensitive, being capable under optimum conditions of detecting less than 10'”
unpaired electrons per gram. It provides a precise method for the determination
of g, and significant variations of g from 2, the spin only value, can be interpreted
in terms of contributions from the orbital motion of electrons. In turn, infor-
mation about orbital motions can be interpreted in terms of orbital populations
and hybridizations. As with n.m.r., the magnetic field of nuclei can split e.s.r.
lines. The study of such splittings leads to valuable information which can be
interpreted in terms of the electron density of the unpaired electron at the
magnetic nucleus. As an example of splitting of e.s.r. lines by a nucleus, consider
an electron in the field of a proton. The proton can exist in two states, either
aligned with the field (M, = +4) or aligned against it (MW, = —4). Therefore,
each electron level is split into two components by the protons. Those protons
with M, = +4 reinforce the effect of the magnetic field, whereas those with
M, = —+ are opposed to the magnetic field. Thus, four energy levels arise.
Since nuclei do not alter their orientation during the short time required for
electron transitions, only transitions for which AM, = 0 are allowed, and thus
only two lines arise, as illustrated in Fig. 14.21.
s=+1/2 _ M, = +\/2
V2 ~
¢ M, ms =\72
XN
SGESW2 Mp Ne
M, = +\/2
Slightly more complex spectra, those of the methyl and ethyl free radicals,
are shown in Fig. 14.22. For the methyl free radical the four observed peaks
in the intensity ratios |: 3: 3: 1 can be explained in terms of the proton splitting
diagram of Fig. 14.22c and the previously mentioned selection rule. (For the
"C isotope of carbon J = 0. See problem 29 for further discussion of the ethyl
radical spectrum.) e.s.r. is an invaluable tool for investigating the reaction
properties of such radicals.
Sec. 14.13 Molecular Structure 431
FIG. 14.22. e. s.r. spectra. These spectra are presented as the second derivative of the absorp-
tion peak. Field increases from left to right [from R. W. Fessenden and R. H. Schuler, J. Chem.
Phys., 39, 2147 (1963)].
K- 23.06>|
K— 26.96
—>
k- 22.46>
O GD C60 C68 of ©
(b) 12 line @.s.7. spectrum of the ethyl radical at -180°C.
|
: |
“ NS
“ < Ph, |
—_—< —— |
aes =
base oe SS
. es |
|
|
|
|
Pee |
Z Sy
———_« —— |
7 .= & A |
.. ’ =—=— |
S = |
<a |
LE SsNoerl
| CH
Number of equivalent O | 2 3 Wetrcnciions
protons
(c) Splitting diagram for one, two, and three equivalent protons
432 Molecular Structure Chap. 14
SSE (14.20)
where E, is the energy of the emitted gamma ray, M is the mass of the recoiling
nucleus, and c is the velocity of light. For °’Fe, E, is 14.4 k.e.v.
If there were no recoil effect, the energy of the gamma ray would be £, the
energy corresponding to the full energy difference between the excited and
ground states of the nucleus. However, the energy of the gamma ray is reduced
to (E — R) by the energy absorbed by the nucleus in the recoil. Now, consider
the absorption of a gamma ray by a free °’Fe atom. The energy required for
this is (E + R), since the absorbing
atom must recoil in the direction of
FIG. 14.23. Illustration of the energy distribution :
for resonant absorption. (a) represents the energy the ae ray with an rey R.
distribution in the excited state. In (b) are shown The question of the probability
the energy distribution of the emitted protons, Of a gamma ray emitted by the
(E —R), and the energy (E +R) required to nucleus of one °*’Fe atom being
both excite a second nucleus and provide it absorbed by the nucleus of a second
with a recoil energy, R. Resonant absorption
can occur only when these overlap.
*"Fe atom can now be discussed.
The energy of the gamma ray is
distributed about a central energy
with a certain spread I’, as shown
| in Fig. 14.23. The gamma rays
r (a) emitted by the first nucleus are dis-
tributed about the energy (EF — R),
ae a I es ie ee ee while the gamma rays which can be
absorbed by the second nucleus are
distributed about (£ + R). Only if
the two distributions overlap (Fig.
14.23) can the gamma rays emitted
by the first nucleus be captured by
the second in a process called
resonant absorption. Although the
recoil loss R for free atoms is very
small, it is large enough to prevent
resonant absorption. R. L. Moss-
bauer (1958) discovered that if the
Sec. 14.14 Molecular Structure 433
emitting nucleus were bound into a crystal, for some events M in equation
(14.20) is effectively infinite, and the recoil loss is therefore so drastically
reduced that resonant absorption can occur.
The significance of this discovery lies in the following fact. The natural
line width IT‘ of the *’Fe gamma ray is 4.8 x 10-°e.v. Thus, any change in the
environment of a *'Fe atom which causes a change in the energy of the emitted
gamma ray by ~ 107" to 10-* e.v. effectively prevents resonant absorption. The
Mossbauer effect, therefore, provides an extremely sensitive probe for measure-
ment of very small energy changes. This measurement is accomplished in the
following way. The energy of a photon can be slightly altered by moving the
source of the photon with the change in photon energy, AE,, given by the
relation
where v is the source velocity and c is the velocity of light. A change in gamma
ray energy of 10~* e.v. corresponds to a velocity of ~ 2mm. sec.~!. Experimentally,
a movable *’Fe source whose velocity can be continuously varied is provided.
The gamma rays from this source are passed through an absorber containing
57Fe nuclei, and then allowed to fall on a detector as shown in Fig. 14.24. The
source velocity is varied, and a curve of counting rate versus source velocity
is determined. It is found that at certain source velocities the condition for
resonant absorption is met, and the counting rate decreases as illustrated in
Fig. 14.25.
Movable
Source Absorber Detector
ip Counter-
Analyzer FIG. 14.24. Schematic
diagram of a Méss-
baver apparatus.
2
oO
oO
”
~
FIG. 14.25. A Méss- Fa
bauer spectrum for ©
absorption by Fe,O3 2
of gamma rays from a
ce.
a Fe source em- 2
bedded in stainless So
steel. Sy
Aloe =8 6-6 =4. =2 OF 942. ee are +8 +10
EXERCISE 14.10
The half life of the 14.4 k.e.v. excited state of °’Fe is 10-7 sec. Use the uncertainty
principle to show that the corresponding line width should be between 10-7 and
NOr? av
The energy content of all particles such as atoms and molecules may
be divided into two parts: (1) energy of motion of the center of mass in space,
called translational energy, and (2) internal energy. For atoms the potential
energy associated with electronic structure falls into the second class. (We dis-
regard nuclear energy, since our subsequent discussions will be confined to
processes in which there is no change in intrinsic nuclear properties.) Changes
in the internal energy of an atom arising from changes in its electronic structure
have been discussed in Chapter 13. Similar changes in the electronic structure of
molecules may occur, leading to the absorption or emission of radiation.
A molecule consisting of two or more atoms has two varieties of internal
energy not found in atoms. These arise from the fact that each atom in a
molecule preserves its identity, at least so far as its inner electronic structure
and nucleus are concerned. Therefore, a molecule consists of two or more
massive nuclei situated at distances which are large compared to the diameter
of the nucleus. These nuclei may execute vibrational motions with respect to each
other or may rotate about the center of mass of the molecule. Both of these
types of motion are independent of the motion of the center of mass of the
molecule in space.
Sec. 14.15 Molecular Structure 435
Ain,
/ \
f= i(An) (14.24)
The proportionality constant is called the force constant of the oscillator. The
restoring force at any point is the negative of the derivative of the potential
energy U, with respect to Ar at that point.
This is called the zero point energy, and represents the minimum energy which
the oscillator must have even when the temperature of the system is so low
that all translational motions have ceased.
EXERCISE 14.12
Typical values of the fundamental vibration frequencies of diatomic molecules
lie in the region of 5 x 10!*sec.-!. Find the corresponding value of the “quantum”
of vibrational energy ho. Ans. 3.3 X 107} erg.
The very simple model which we have just discussed is, of course, not com-
pletely adequate. One difficulty is that the potential function of equation (14.26)
and Fig. 14.27 is a poor approximation to the actual molecular potential func-
tion at all but the lowest energies, as Fig. 14.5 shows. The actual molecular
potential is not a simple parabolic function.
This difference requires that for accurate description of molecular vibrations
correction terms must be added to equation (14.29). More important, it causes
the permitted vibrational energy states to become more and more closely spaced
with increasing v, eventually approaching an energy limit corresponding to
dissociation of the molecule, as indicated in Fig. 14.29. In that figure the bond
dissociation energy D, corresponds to the vertical distance between the v = 0
vibrational energy state and the energy level corresponding to the separated
438 Molecular Structure Chap. 14
atoms. This is less than D,., the dissociation energy from the bottom of the
potential energy curve, by E£%.
So
absorption
50
ill yur
O
2500 2600 2700 2800 2900 3000 3100
v' (cm)
called the fundamental (0, 1) band. The origin of the band is that frequency
corresponding to AJ = 0, a forbidden transition, and lies at the position of the
missing line in the center of the band. To the right of the origin (higher fre-
quencies) lie the successive lines for AJ = +1, that is, for J = 0 in the lower
vibrational state to J = | in the higher vibrational state; J = 1 to J=2, J=2
to J = 3, etc. To the left of the origin lie the successive lines for AJ = —1;
IN WSS
OI = AiO d/ = lly Ge:
Assuming harmonic vibration, equation (14.28) may be used to obtain the
force constant of the bond. Since @ = cw’ and for the (0, 1) transition w’ = 2886
cm.~!, from equation (14.28) we obtain
k = 47’? ’o"
At what wave length and frequency will the 0,2 band of H*°Cl be found?
Ans. 1.673 x 1074 cm.; 5978 cm.7!.
The foregoing analysis of the rotational structure assumes that the molecule
is a perfectly rigid dumbbell rotator. In actual fact, the rotational moment of
inertia increases slightly with increasing J, due to stretching of the bond. Further-
more, the internuclear separation and hence the moment
of inertia also increases with increasing vibrational
energy, since the vibrations are not perfectly harmonic. I
A more precise calculation taking these effects into (1595 cm')
ee 2
S $ 3
= =
ee g
% fe) 2
ae ae " ats
eo © O Oo
80
60
n-hexadecane
methanol
Transmission
%o
———— 1 1 te
80
60
40 acetophenone
20
fees l ] i eS. L
benzaldehyde
molecular electronic states such as those indicated in Fig. 14.35. It can be seen
from that figure that the energy difference of such states is of the order of
magnitude of several electron volts. Since 2 e.v. = 16.2 x 103 cm.~! = 6200 A,
these spectra are usually found in the visible and ultraviolet ranges (1000-
8000 A).
Each attractive (stable) electronic state of a molecule has its set of permitted
vibrational energy states and each of these a set of permitted rotational energy
states. However, the vibrational force constant and rotational moment of inertia
will usually not be the same in the various electronic energy states of a molecule.
In the course of an electronic transition, changes in the rotational and
vibrational quantum numbers may also occur, giving rise to a band spectrum
such as that shown in Fig. 14.34. All of the bands shown arise from a single
electronic transition with various changes in v, the vibrational quantum
number, as indicated. The fine structure of each band (not resolved in the pho-
tograph) is due to changes in J, the rotational quantum number. The selection
rule for AJ in electronic spectra allows AJ = 0, +1; there is no selection
rule for Av. The bands in Fig. 14.34 arise from transitions from v = 0 in the
lower electronic energy state, since at ordinary temperatures this is the most
highly populated vibrational energy state.
Electronic emission spectra of molecules originate in transitions from excited
electronic states. These excited states can be produced by electron impact, by
heating to high temperature, or by light absorption. Emission spectra are
more complex than absorption spectra, in the sense that more bands, includ-
ing those corresponding to transitions to v = 1, 2, 3, etc. in the ground elec-
tronic state, can be seen. A series of bands such as those in Fig. 14.34, which
arise from transitions to or from a single vibrational state, is called a progres-
sion. A larger variety of band progressions is usually seen in emission spectra.
It is found that a regular change in intensity of absorption occurs in a given
progression. That is, the probability for transition from v = 0 in the lower
electronic state varies with the value of v in the higher state. This phenomenon
can be understood with the aid of the Franck-Condon principle, which states
that the most probable transitions are those in which separation and kinetic energy
of the nuclei do not change. This arises from the fact that electronic transitions
occur in a time much less than the frequency of molecular vibrations, and,
therefore, the molecule, upon being excited, must momentarily retain the same
nuclear separation and kinetic energy. The Franck-Condon principle may be
illustrated by reference to case a in Fig. 14.35, which represents, qualitatively,
Cc
v 25 30 35 40 45 |
FIG. 14.34. Absorption spec- | - | |
trum of ly [from G. Herzberg,
Molecular Spectra and Mole-
cular Structure, D. Van Nos-
trand Company, Inc., Princeton,
N. J., 1950].
Case a LO ee the electronic states involved in absorp-
Ip _
tion by I,.
Absorption of energy by a molecule
144 ev.
in the lower electronic state with v = 0
produces an electronically excited mole-
cule with the same internuclear separa-
tion. The instantaneous final state, there-
fore, lies vertically above the initial state
in the region enclosed by the dashed
lines in the figure. The requirement that
Case 6 2.2 ev.
the vibrational kinetic energy shall not
HI — change discontinuously is met by having
the instantaneous final state be one of
3.0ev. small vibrational kinetic energy, but large
vibrational potential energy. That is, the
instantaneous final state may be regarded
as one in which the nuclei are at an
extreme displacement from the new equi-
librium internuclear separation rj. This
is the most probable transition and
evidently will give rise to a large value
Case ¢ of v in the upper electronic state. Transi-
redissociation
ihe tions to other vibrational states become
less probable as the required change in
internuclear separation or _ vibrational
kinetic energy increases.
Absorption of radiation by a molecule
—>
in its ground state can lead to dissocia-
FIG. 14.35. Electronic energy states.
tion if the potential energy curve for the
upper state lies in such a position that
vertical displacement gives an energy
above the dissociation limit for the upper state as is the case with I, at wave
lengths below 4995 A. The absorption spectrum of I, is banded at long wave
lengths, indicating transitions to quantized vibrational states of the upper
electronic energy state. However, at wave lengths less than 4995 A (2.44 e.v.)
the absorption is continuous rather than banded. This behavior is evidence of
dissociation into I + I*, where I* represents an excited iodine atom. The
large potential energy of the instantaneous final state becomes relative kinetic
energy of the separated atoms. The wave length at which continuous absorption
begins corresponds to the energy required for the process
Ness head
It is known that the excitation energy of the iodine atoms is 0.94 e.v.; and
therefore, we obtain for the dissociation energy of I,
D,, = 2.44 — 0.94 = 1.5ev.
= 35 kcal. mole“
444
Sec. 14.17 Molecular Structure 445
Some bond dissociation energies obtained from analysis of spectra are included
in Table 14.6.
For an optical transition to a repulsive electronic state of the molecule,
there will be no band structure, but only a continuous absorption leading to
dissociation. This is so for HI, illustrated in case b, Fig. 14.35. In such instances
the Franck-Condon principle indicates that absorption will occur only for
energies considerably greater than the bond dissociation energy. The large
potential energy of the instantaneous final state becomes relative kinetic energy
of the separated atoms. In the case illustrated, the absorption maximum occurs
at about 2500 A (5.2 e.v. per quantum) whereas the bond dissociation energy
is only 3.0 e.v. Therefore, at this wavelength, the relative kinetic energy of the
H and IJ atoms will be 2.2 e.v.
In some instances it is observed that a molecule exhibits banded absorption,
indicating an attractive upper state, but that in a certain wavelength region
the bands become diffuse. That is, the rotational lines are not sharp. This is
an indication of the phenomenon of predissociation and one of several possible
combinations of potential energy curves which can lead to this effect is shown
in case c, Fig. 14.35. Absorption gives rise to an attractive state (1) at such a
vibrational level that dissociation to A + B cannot occur. If the repulsive state
(2) “crosses” the state (1) at an energy level corresponding to that formed in
the absorption act, an internal conversion from state (1) to state (2) may occur
and result in dissociation. Such a conversion is governed by the Franck-Condon
principle, and the two states must have the same instantaneous nuclear separa-
tion and kinetic energy.
TABLE 14.6
PROPERTIES OF MOLECULES OBTAINED FROM SPECTRA*
SUMMARY, CHAPTER 14
1. Types of bonds
Ionic: Between atoms of low ionization potential and atoms of high
electron affinity.
Covalent: Electron pair shared equally in homonuclear molecules.
Polar covalent: Degree of ionic character increases with increasing
electronegativity difference. Electronegativity increases toward upper right
of periodic table.
2. Covalent bond
Necessity for detailed description of electron distribution by use of wave
mechanics.
Molecular orbital theory: Construction of one type of molecular orbital
by linear combination of atomic orbitals. Requirement that combination
of two atomic orbitals give two molecular orbitals, one bonding and one
antibonding. Occurrence of bonding when number of bonding electrons
exceeds number of antibonding electrons. Classification of orbitals by
symmetry, o orbitals being cylindrically symmetrical about internuclear
axis, in contrast to z orbitals, with node in one plane.
Valence bond theory: Description of a covalent bond by overlapping
of atomic orbitals.
Construction of hybridized atomic orbitals to describe directional prop-
erties of bonds (sp? in carbon, for example).
Construction of a resonance hybrid of valence bond structures to repre-
sent the structure of molecules which cannot be described by single valence
bond structures.
Summary, Chapter 14 Molecular Structure 447
N Lx
x= Nan + 3c)
x, is positive for paramagnetic substances. When paramagnetism arises
only from unpaired electron spins
E, = ho (v + 4)
Permitted rotational energy states
ea B I (I - )
Molecular electronic spectra consists of bands corresponding to transi-
tions between attractive electronic states. Continuous absorption denotes
transition to a repulsive electronic state. The Franck-Condon principle
states that the most probable transitions are those in which the relative
position and vibrational kinetic energy do not change. The spin conservation
rule states that transitions between states of different multiplicity are for-
bidden.
448 Molecular Structure Chap. 14
PROBLEMS, CHAPTER 14
1. The equilibrium internuclear separation in gaseous diatomic Nal is 2.90 A.
Find the energy term in the process
Na(g) -1@)= > Na ip
neglecting the energy of repulsion. Ans. 4.97 e.v. molecule=!
2. From the bond dissociation energy of I, (page 444) and the electron affinity
of I (Table 14.1), find AE for the process
e+I1,—-I1-4T-. Ans. —1.56 e.v.
3. From data in Fig. 13.11 and Table 14.1 estimate the electronegativity of the
halogens (cf. Exercise 14.5). Compare with the empirical values given in Fig. 14.10.
4. From data in Fig. 13.11 and Table 14.1 show that the process
M(g) + X(g) — M*(g) + X(g)
is endothermic for all possible combinations of alkali metal and halogen.
5. Use the molecular orbital approach to predict the stability of the diatomic
molecules Li,, Be,, B,, and C,.
6. Use the concept of hybrid bond orbitals to predict the geometry of BH3.
What will the polarity of the B — H bonds be? Will the molecule have a dipole
moment ?
7. The bond angle in H,S is 97°. From the dipole moment of the molecule given
in Table 14.4 find the S — H bond moment. Ans. 0.72 debye.
8. The dipole moments of HCl, HBr, and HI are 1.07, 0.78, and 0.38 debye,
respectively. Show that these values are approximately proportional to the elec-
tronegativity differences between hydrogen and the respective halogens.
9. The total molar polarization of HCl gas at 200°K is 41 cc. mole~!. Use the
dipole moment in Table 14.4 to find the value of the distortion polarizability.
Compare with j1?/3kT. Ans 0 — 02555 1052cm: molecules ™
[?/3kT = 1.38 x 10-23 cm.* molecule=!.
AO. From the dipole moment of HCl given in Table 14.4 and the internuclear
“separation, 1.275 A, find the effective charge on each atom in electronic charge
units. Compute the % ionic character of the bond.
_11. R. P. Bell and I. E. Coop [Trans. Faraday Soc., 34, 1209 (1938)] have found
that the dipole moment of deuterium chloride is greater than that for hydrogen
chloride. Explain.
12. The magnetic moments ££ of some transition metal complexes are listed.
Where possible, use these data as a criterion and show how the d orbitals (Fig.
14.12) are occupied. If it is not possible to reach a decision on the basis of mag-
netic moments alone, state this and state why this is so.
Complex Lt (Bohr magnetons)
Co(NO,)g * 1.9
Co(NHs)3” 5.0
CoF;? 5.3
Fe(CN)s° 2.3
Fe(H,0)¢” Bee)
Problems, Chapter 14 Molecular Structure 449
| 13. From the following data for the dielectric constant of gaseous HCl as a
function of temperature, estimate its dipole moment. The pressure for each
measurement was one atmosphere. Assume that the ideal gas law holds.
15. R.A. Ogg [Discussions of the Faraday Society, 17, 215 (1954)] has used
n.m.r. to investigate the exchange reaction
NH; + NHs — NH; + NH;
in liquid NH. Sketches of his spectra in the order of increasing concentration
of NH» follow.
(a) Explain what is happening.
PROB. FIG. 14.15.
(b) The separation of these peaks
is 46 cycles per second. Compute r,
the mean life for the exchange reac-
tion. Ans.¢r = 2.2 x 10-2sec. 4°: BS,
16. Draw sketches showing how
the n.m.r. spectrum of ethyl alcohol
would be expected to change as_ c. ee D. a EP
temperature is increased.
18. Pure rotational spectra are observed in the far infrared. Find the frequency
in cm.~! corresponding to the changes J = OtoJ = 1,J = 1toJ =2,andJ =2
to J = 3 for H*Cl. (Take J from. Table 14.6.) <Ans.J =0to J =1, 2h.2cm.*.
~19. The spacing of the rotational lines in the infrared spectrum of CO is approxi-
mately 3.86 cm.~!. Find the internuclear separation in the molecule.
20. The molecule HI has a force constant of 3.0 x 10° dynes cm.~! and an
internuclear separation of 1.60 A. Calculate
(a) The fundamental vibration frequency of the molecule.
(b) Its zero point energy in calories mole™!.
(c) The position of its first vibrational band in wave numbers.
(21. In the near infrared spectrum of CO there is an intense band at 2168 cm.~'.
Calculate
(a) The fundamental vibrational frequency of CO.
(b) The period of vibration.
(c) The force constant.
(d) The zero point energy of CO in calories per mole.
Ans-(a)70.) X< 10! secs1,
(d) 3.10 kcal. mole™!.
450 Molecular Structure Chap. 14
_ 26. The heat of dissociation of Cl, in the gas phase is 56,800 calories mole.
Calculate the work in calories required to separate the two ?5Cl atoms a distance
of 0.1 A from their equilibrium position in the molecule. The equilibrium inter-
atomic distance in the Cl, molecule is 1.98 A, and the fundamental vibrational
frequency is 1.603 = 101% sec.7}.
27. The particle in a box treatment provides a very simple model for the elec-
trons in a system of conjugated double bonds, for example CH;—(CH=CH),—
CH,. Assume that the conjugated double bond system may be approximated as an
infinite walled one-dimensional box of effective length, L = 9.8 A, and calculate
(a) The minimum kinetic energy of an electron in this model.
(b) The wavelength of the radiation required to excite an electron to form
the first excited state.
(c) The addition of a single CH=CH group increases L by 1.4 A. What is the
corresponding increase in the wavelength of the radiation required to reach the
first excited state? Ans. (a) 6.27 < 107!’ ergs,
(b) ~ 3520 A.
28. Sketch a splitting diagram for an unpaired electron interacting with a
magnetic nucleus of J = 1. How many e.s.r. lines would be expected?
29. The intensity ratios of the peaks in the ethyl radical spectrum shown in
Fig. 14.22a are 1: 2: 3:1: 6:3: 3:6: 1: 3:2: 1. Show how such intensity ratios
can arise in a radical which has one group of three equivalent protons and a second
group of two equivalent protons. In the ethyl radical spectrum, which group of
protons causes the greater splitting?
80. The dipole moment of nitrobenzene is 3.93 debye, and the dipole moment
of p-nitrotoluene is 4.39 debye (A. L. McClellan, op. cit.). Predict the dipole
moments of m-nitrotoluene and o-nitrotoluene. Ans. 4.2 and 3.8 debye.
SOME PROPERTIES
OF SOLIDS
AND LIQUIDS
ny
1. TRICLINIC
6
{\
ARPA
2. SIMPLE
MONOCLINIC
Ae /
3. SIDE-CENTERED
MONOCLINIC
4. SIMPLE
ORTHORHOMBIC
FIG.
lattices
sical
Hall,
1962].
15.2.
[from
The
Chemistry,
fourteen
W. J. Moore,
3rd
Inc., Englewood
ed.,
Bravais
Phy-
Prentice-
Cliffs, N. J.,
CIOS
oa
ne
10. SIMPLE
TETRAGONAL
11. BODY-CENTERED
TETRAGONAL
12. SIMPLE
CUBIC
13. BODY-
CENTERED
CUBIC
14. FACE-
CENTERED
CUBIC
Sec. 15.2 Some Properties of Solids and Liquids 453
lattices are shown in Fig. 15.2. To generate a complete space lattice, each of the
representations in 15.2 must be repeated in all directions without limit. With
three-dimensional lattices, three vectors, usually designated a,b, and ec, are
used to establish unit cells. As in the two-dimensional case a primitive cell must
contain only one lattice point. Thus, in Fig. 15.2, cells 1, 2, 4,9, 10, and 12 are
primitive cells.
An infinite number of planes can be passed through any space lattice; some
of them are shown in the two-dimensional cut of Fig. 15.3. The various lines in
this figure may be considered to represent edges of planes. These planes can be
defined in terms of the x, y, and z axes using three numbers which represent
the simplest whole-number ratio of
its intercepts on these axes. This
leads to the sets of indices shown in FIG. 15.3. Some possible planes in a cubic lattice.
Planes are viewed edge on.
Fig. 15.3. Since all planes shown are
parallel to the z axis, the index for
this axis is oo.
Although the indices of Fig.
15.3 can be used to define planes,
a more convenient set is usually
chosen, the Miller indices, which E(o, 4, w)
can be derived as follows. Take the
reciprocals ofthe coefficients of a, b,
ande in Fig. 15.3. Clear fractions
by multiplying by the lowest com-
mon multiple, excluding infinity,
and the resulting three numbers are
the Miller indices of the plane in
question.
The operations are illustrated in
Table 15.1, where the Miller in-
dices are represented by the sym- A(2a, 26, o)
Dia, ©, ©)
bols A, k, and / referring to the x, &(2a, 56, 0) *C(4a, 66, o)
y, and z axes respectively. Each set
of three numbers represents an infi-
nite set of parallel planes with a
fixed interplanar spacing.
TABLE 15.1
aiid
pig
py The perpendicular distance
oo=9 thio=3V20 of separation d between adja-
(a) cent members of the set of
parallel planes represented by
bee a Se
FIG. 15.4. Spacings in cubic lattices: (a) simple cubic; a a/ h? + k? + [? ( )
(b) body-centered cubic ; (c) face-centered cubic ; [from ; f :
W. J. Moore. Physical Chemistry, 3rd ed., Prentice-Hall, Figure 15.4 illustrates spacings
Inc., Englewood Cliffs, N. J., 1956]. between some sets of planes in
cubic lattices.
EXERCISE 15.1
For the two-dimensional array of Fig. 15.3 use a geometrical argument to show
that the distance of separation of the parallel set of AA planes is given by equation
(15.1) with / = 0.
EXERCISE 15.2
Show that the distance between 210 planes in the simple cubic lattice is a/,/ 5.
| I
| I
I |
I I
| |
|
(c) Orthorhombic (d) Hexagonal
| |
TABLE 15.2
THE CRYSTAL SYSTEMS
sources of scattered X rays. The X rays scattered from different point sources
may reinforce or cancel each other, giving rise to a typical diffraction pattern.
The condition for reinforcement may be derived most easily by considering a
reflection which has an equivalent effect. As we have already seen, parallel
planes defined by sets of Miller indices, h, k, and /, may be passed through these
points. Consider Fig. 15.6, which represents a cut through a set of such parallel
planes. A ray incident at an angle 6 to one of the planes defined by a given row
of points can be considered to be reflected at the same angle. A ray reflected
from the second plane must travel a distance ABC greater than a ray reflected
from the first plane. The increased path length is given by
where 7 is an integer, and X is the wave length of the scattered radiation. This
is the famous Bragg equation, named for W. L. Bragg, who first developed
this approach.
If the angle 6 is varied by rotating the crystal, an intense scattering will be
observed at values of sin 6 corresponding to n = 1, 2, 3, etc. The reflection with
n = | is referred to as first order, that with m = 2 as second order, etc. For
second-order reflections the Bragg equation becomes 2 = 2d sin 6, where 6,
refers to the angle at which second-order scattering occurs. This equation may
be rewritten as \ = 2(d/2) sin 6,. Thus, second-order reflection from a set of
planes of spacing d is formally equivalent to first-order reflection from a set
of planes with spacing d/2. As a specific example, if the spacing of the 111 set
of planes is d, that of the 222 set of planes is d/2. Therefore second-order reflec-
tion from the 111 planes and first-order reflection from the 222 planes are
equivalent. In general, then, an alternate statement of Bragg’s law (equation
15.4) is
N= 2dair
SID nx (1535)
Lead
X-ray stop |
beam -
Film
3% X-ray X-ray
film beam
iia Cae aN l
111 200 220 311 222 400 331 420 422
SODIUM CHLORIDE
| | | | | |
200 220 222 400 420 422 620
600
POTASSIUM CHLORIDE
¢
FIG. 15.7. X ray diffraction by NaCl and by KCI powders [from W. J. Moore, Physical Chemis-
try, 3rd ed., Prentice-Hall, Inc., Englewood Cliffs, N.J., 1962; photograph courtesy of Dr.
Arthur Lessor, IBM Laboratories].
where 7 is an integer equal to (h? + k* + /°); that is, the observed values of
sin’ 6,,, should progress in even steps. However, there are some important ex-
ceptions. For example, (4? + k*? + /*) cannot equal seven. Thus, the value of
sin? 6,,, for n= 7 should be absent (Table 15.3). Similarly, the fifteenth,
twenty-third, twenty-eighth, thirty-first, and thirty-ninth lines will be missing.
These patterns of absences enable us both to identify a simple cubic crystal and
to relate the observed lines to appropriate Miller indices.
TABLE 15.3
Body centered
| | | | | | | | | FIG. 15.8. Lines observed with the three
cubic
nike) Ao =~ Zi) 220) BO 321 400 330 420 cubic lattices as a function of angle.
411
The values of A?/4a? are arbitrarily
Face centered || | || | || | | | selected so that the first line will fall
cubic
111200. —-220-311 222 400331 420 422 333 440 «©Mt_:''0 for each lattice.
511
10 14 18 22 26 30 34
Scattering angle (degrees)
from the 200 planes just as for the body-centered cubic crystal. In addition the
110 planes will be interleaved with a set of 220 planes, with the result that the
110 scattering is also cancelled. In general, it can be shown that reflections are
observable only from planes for which the Miller indices are either all even or
all odd.
The characteristics of the three basic cubic systems are compared in Fig. 15.8.
For all three it is assumed that the first line is at 10 deg., and on this basis
the positions of the allowed lines for each system have been calculated by using
equation (15.6). The differences in the patterns serve to identify the type of unit
cell. (This is an extremely simple example. Patterns are usually more complex,
as we shall see below.)
As a concrete example, let us consider the NaCl crystal. It is an easy matter
to show that the square of the sines of the scattering angles corresponding to
the lines shown in Fig. 15.7 can all be expressed in terms of equation (15.7)
and that the crystal is therefore cubic. The value of a, the unit cell parameter,
is 5.63 A. The pattern of lines observed shows only all odd or all even values
of h, k, and /, indicating that the crystal is face-centered. We have one inde-
pendent piece of information. The density of crystalline NaCl is 2.163 g.
cm.~*. The volume occupied by one formula weight is, therefore, 58.45/2.163
= 27.02 cm.*. The volume per ion pair must then be 27.02/(6.02 x 10?)
— 44.9 x 10-4 cm.* or 44.9 A’. a® is(5.63 A)? or 178.5
FIG. 15.9. The unit cell A’®. There/>re, there are 178.5/44.9 = 4 Na*Cl- ion pairs
for NaCl and KCI. per unit ce!
A unit cell which meets these requirements is shown
in Fig. 15.9. This represents two interpenetrating face-
centered arrays, one composed only of Na* ions, the
second only of Cl~ ions. If we recall that units at the
corners of the cells are shared among eight cells, those
on an edge among four cells, and those at the center
of a face between two cells, we see that the require-
ment of four ion pairs per unit cell is met. The scatter-
ing centers in this unit cell are not equivalent, since Cl-
has eight more electrons than does Nat. This has an
effect on the intensity of the line from the 111 planes. If
we assume that the white circles in Fig. 15.9 are Cl- ions,
then we see that there is a set of 111 planes consisting
only of Cl” ions. However, this is interleaved with a set of
Sec. 15.7 Some Properties of Solids and Liquids 461
parallel planes containing only Na* ions. If the scattering condition is met for
the 111 plane of Cl- ions, then the Na+ planes must be scattering radiation out
of phase with that from the Cl- planes, reducing the 111 intensity. Such intensity
variations provide powerful clues to the arrangement of the elements of a unit
cell.
Figure 15.9 also provides a basis for understanding the powder photograph
of KCl. If we assume the same unit cell as for NaCl, then the various planes
have the same Miller indices. Again, we expect the 100 and 110 lines to be
absent because of interference from the interleaved 200 and 220 planes. How-
ever, there will also be no line observed due to scattering by the 111 planes of
Cl- ions. These are now interleaved with planes of K* ions. Since K* ions and
Cl- ions contain equal numbers of electrons, they are almost equivalent in
scattering power. Scattering from the 111 Cl planes is thus effectively cancelled
by out-of-phase scattering from the interleaved K* planes. As Fig. 15.7 shows,
this model agrees with the observed powder pattern.
One of the most important uses of the X ray rotating crystal technique
is for the determination of the struc-
ture of molecules. We shall indicate Fg, 15.10. A two-dimensional lattice consisting
how this might be done by consider- of AB, type units in which scattering from
ing a simple model. Figure 15.10 8 is out of phase with that from A.
alized to other planes. For these the phase difference for an atom 7 along the x
axis will be (hx,/a)27, along the y axis (ky;/b)27, and along the z axis (/z;/¢)27.
For simplicity our discussion now will be limited to the x dimension only.
In Fig. 15.10 we have represented the atoms of the AB, molecule as dots. This
is, of course, unrealistic since the electrons associated with atoms are dis-
tributed in space in some way. In the x dimension this distribution may be de-
scribed by a suitable mathematical function of x, which we shall symbolize as
p(x). It is actually p(x) which we are determining, since X rays are scattered by
electrons. p(x) will vary in some periodic fashion along the x axis, because the
structure is periodic along this axis. It will peak where atoms are centered. The
electron density in any small element between x and (x + dx) of the one-
dimensional unit cell will be p(x) dx, and the intensity of the radiation scattered
from this element will be proportional to p(x) dx.
We shall now define an origin for the coordinate system. Since our AB,
system has a center of symmetry, it is convenient to choose as an origin some
point in the unit cell which reflects this symmetry. We shall select the black
atom inside the square in Fig. 15.10. A ray scattered in the element between
x, and (x, + dx) will be out of phase with one scattered at the origin by an
amount (hx,/a)2z. It will superimpose a displacement at the origin propor-
tional to
integrating p(x) cos 2x(hx/a) dx over half the cell, and doubling this. The
resultant quantity is called the structure factor, F(h), where h signifies an appro-
priate Miller index.
cos 27(hx;/a) + isin 27(hx;,/a), to express both the amplitude and the phase.
Since e’ = cos @ + isin 6 (see Appendix 1), a general form of the structure
factor is
F(hkl) = ¥ f, exp [2 ( Xi KY Ll
i=1 a b (©
The right-hand side of this equation equals 0 for all terms except those for
which n = h. For these terms it equals aC(n). Therefore, C(n) = F(h)/a.! This
elegant result enables us to replace the coefficients, C(n), in the Fourier series
with the corresponding structure factor terms, F(A)/a. The equation for electron
density in one dimension then becomes
TABLE 15.4
IONIC RADII*
At the other extreme, when the cation becomes comparable in size to the
anion, as in CsCl, the halide ion lattice may become simple cubic, with a ca-
tion at the center of the cube, forming a body-centered cubic lattice. Although
the simple cubic lattice of halide ions is not as compact as the face-centered
cubic arrangement usually observed, the number of equidistant ions of opposite
sign is eight instead of six. Apparently, the increased number of ions compen-
sates for the increase in interionic distance.
In ionic crystals no simple unit consisting of two oppositely charged ions
can be recognized. Instead, the whole crystal must be regarded as a single as-
sembly of ions or one giant molecule. Each ion exists in a force field primarily
due to its nearest neighbors, but also partially due to more distant ions, both
positive and negative.
As a specific example, consider the sodium chloride crystal (Fig. 15.9). Con-
tributions to the coulombic energy of a given Na* ion arise from the six neigh-
boring Cl ions at an internuclear separation r, twelve nearest Na* ions at a
distance r,/ 2, the eight next nearest Cl- ions at a distance ra/3, the six next
nearest Na* ions at a distance 2r, the twenty-four next nearest Cl” ions at a
distance r,/ 5, etc. When a Na’ ion is transported from infinity to this lattice
site, energy will be released. This energy is simply the sum of the various
coulombic interaction energies, and is given by
VE
LayZor) +oe
Law) + re Law Zer) + $FCe
This can be transformed as follows:
—eé 12 8 6 24 24
y=—£ (6-7 nom =...) 15.14
r ae ae A / 6 ( )
The term in parenthesis is an infinite series which converges at 1.748. It is called
the Madelung constant for NaCl.
Next we must consider the repulsive forces arising from the proximity of
the electron clouds of Na* and Cl-. It was proposed by Born that the repulsive
energy arising from this source is in-
versely proportional to r”, where r is
the internuclear distance. n, the Born
exponent, can be determined by ex-
periments on the. compressibility of
salts containing various ions. When
both cation and anion are of the
neon configuration it is about 7; for
two ions of the argon configuration,
about 9; for two krypton-like ions, 10; Energy
Potential
attractive forces, and the second term gies arising from attractive and repulsive
forces respectively. The resultant energy is
the potential energy, V,, arising from
V, and the minimum in the V curve corre-
repulsive forces. These relations are sponds to the separation of the ions at
illustrated in Fig. 15.15. equilibrium.
EXERCISE 15.4
Show that r, the equilibrium interionic distance in NaCl, is given by the relation
r = (485/A)"". (At the distance r, V is a minimum.) For NaCl r is 2.82 A.
Evaluate the constant b. Ans. b = 51.7 A.
7 = a
= (15.16)
EXERCISE 15.5
Verify equation (15.16), starting with equation (15.15) and using the result of
exercise 15.4. Compute the crystal lattice energy for NaCl.
Ans. Exaci = 180 kcal.
469
470 Some Properties of Solids and Liquids Chap. 15
As Exercise 15.5 shows, the crystal lattice energy computed from equation
(15.16) for a formula weight of crystalline NaCl is 180 kcal.
The crystal lattice energy is related to other thermochemical properties
through the Born-Haber cycle, an application of the first law of thermody-
namics. In the case of sodium chloride the cycle is
and p? for oxygen. However, the bond is quite polar, with a considerable
contribution from the ionic form Sit‘, O- and the opening of the Si—O—Si
bond to 150 deg. is attributed to ionic repulsion.
Because of the strong three-dimensional forces in these covalent crystals,
they are generally very hard and have very high melting points. In these respects
they are similar to ionic crystals.
When a covalent substance, because of limitations on the bonding power
of its atoms, is not capable of forming a three-dimensional lattice by electron
pair bonds, the crystalline form is relatively soft and has a low melting point.
These qualities characterize the molecular crystals in which the bonding be-
tween units is to be attributed to van der Waals or dipole forces. For example,
the diatomic elements such as O,, Cl,, etc. and many organic compounds such
as CH,, C,Hg, etc. form crystals in which the chief determinant of structure
is the tendency to efficient packing of the individual molecules. On the other
hand, several elements, such as sulfur, are capable of forming two electron-pair
bonds which permit the development of chains or rings. Rhombic sulfur con-
sists of S, rings bound by electron-pair bonds within the ring, while the attrac-
’ tion between rings in the crystal is due to van der Waals forces.
Crystalline water (ice) is a representative of a special type of molecular
crystal. By comparison with the heavier hydrides of its family, H,S, H,Te, etc.,
it has an abnormally high melting point and heat of
sublimation. The lattice energy is much higher than
FIG. 15.18. Synthesis of Could be expected from van der Waals forces and
energy bands by combina:
tion of atomic orbitals.
dipole-dipole forces. This is to be attributed to the
hydrogen bonds formed between the water molecules
(see sections 14.8 and 15.14). The crystal structure is
such that each oxygen atom has disposed about it
four other oxygen atoms at the corners of a tetra-
hedron, with hydrogen atoms lying on the line joining
each pair of oxygen atoms. Each hydrogen atom is
attached to one oxygen atom by an electron pair bond
and to another by hydrogen bonding.
2 3 4
Number of atoms
HTINININT
TIMI
= erties suggests that there are large numbers of con-
Seem Sanz Some Properties of Solids and Liquids 473
duction electrons which are not firmly bound to any particular atom, and are,
therefore, relatively free to move throughout the crystal. The second suggests
a description in terms of delocalized electrons, since the formation of twelve
or more localized bonds is unlikely.
A molecular orbital model for the metallic bond can be constructed in
much the same way as the molecular orbital description of benzene. As an
example, let us consider lithium. We shall assume that the inner Is electrons
are so tightly bonded to the nuclei that they are effectively localized. The re-
quired delocalized orbitals must then be constructed from the outer orbitals.
Two molecular orbitals describing the electron distribution in a system con-
sisting of two lithium atoms designated as A and B, can be constructed from
the 2s atomic orbitals. They are yW4(2s) + W,(2s) and vr,(2s) — y,(2s), and
they differ in energy. If a third Li atom is to be added to the system, three
molecular orbitals may be constructed, and associated with these there will be
three energies, grouped about the original energy associated with yr,. As each
atom is added, a new energy
level is added as illustrated in
FIG. 15.19. Band model for a one-dimensional metal.
Fig. 15.18. When the number
of atoms is very large, there
results a band of closely spaced
energy levels (Fig. 15.18). The ‘
same sort of construction can
go on with the 2p levels of
lithium. If the 2s and 2p atomic
systems are well separated in
energy, or if the atoms are well
separated in space, then the RY 2p
two bands derived from them
will not overlap. Thus we will
have two bands separated by
DOO OX 2s
an energy gap. This gap may
be thought of as a forbidden
energy zone. The available
electrons will be allocated to
the lowest energy levels. Thus A A
if the s and p bands are indeed
Positions of nuclei
separated in Li all electrons
would be in the band derived
from s atomic orbitals.
It is often true that the bands derived from two neighboring atomic states
will overlap. Under these circumstances our description will be more com-
plicated. However, the general features which we have pointed out will still
be found. The energy levels will be banded, and the electrons will be assigned
to the lowest set of energy levels.
Figure 15.19 shows the relationship of the crystal bands and the corre-
474 Some Properties of Solids and Liquids Chap. 15
sponding atomic orbitals for sodium. The solid lines correspond to the peri-
odic potential field which constrains the electrons in a crystalline solid, and
the dashed lines correspond to the field associated with an isolated atom. The
ls, 2s, and 2p bands are completely filled. The 3s and 3p bands overlap con-
siderably. These uppermost energy levels are only partly filled and are effec-
tively delocalized. It is these partially filled delocalized levels which account
for the phenomenon of metallic conduction. When an external electric field
is applied it will tend to make more electrons flow in one direction than in the
other.
By contrast consider a nonconductor such as potassium chloride (Fig.
15.20). Since all low-lying atomic orbitals are completely filled, the correspond-
ing crystal bands will be completely filled. There is now no mechanism for a
net electron flow in the direction of an applied field. However, if a sufficient
amount of energy is supplied to promote electrons into the next empty band,
then conduction can take place. Thus, some substances can show photoconduc-
tivity. Light quanta of appropriate wave length are absorbed, and electrons are
thereby promoted to energy levels in the hitherto unoccupied band. Here
electron migration in the direction of the field can take place as with any partly
occupied band. The contrast between a conductor and a nonconductor is
illustrated in Fig. 15.20.
The existence of a conduction band of closely spaced electronic energy
levels extending through the solid depends strongly on precise regularity of the
arrangement of the atoms, as stated previously. Anything which decreases this
regularity tends to increase the resistance to electron migration. Thus, impurities
in solids, as in alloys, usually increase resistance. Increasing the temperature,
which increases the vibratory motion of the nuclei, also increases the resistance
of metallic conductors.
15.13 Semiconductors
FIG. 15.22. X ray diffraction by iiquid water [from J. Morgan and B.E. Warren, J. Chem. Phys.,
6, 666 (1938)].
Relative
intensity
83°C
SORE
negative to the more electronegative element, and the bond can be adequately
described in terms of the interaction of two ions. When two like atoms are
bound, then the electrons are effectively shared, and a detailed consideration of
the electron distribution using wave mechanics is necessary to evaluate the
strength of bonding. Most covalent bonds are of an intermediate type in which
electrons are unequally shared. There is a net charge separation, so these are
called polar-bonds.
The hydrogen bond constitutes a special case of interatomic force which
arises when a hydrogen atom is covalently bound to an atom of high elec-
tronegativity such as oxygen or fluorine. In this case the high degree of ionic
character of the bond leaves a well-exposed proton whose positive charge
may exert a strong attractive force on other nearby electronegative atoms,
as in liquid water.
In both ionic and macromolecular crystals we find large three-dimensional
structures bound together by coulombic and electron pair bonds, respectively.
On the other hand, in molecular crystals and in the corresponding liquids we
must look to other sources for the forces which bind the molecules together.
Molecules which have a permanent dipole moment attract each other by
electrostatic interaction. In contrast to the energy of interaction between point
charges, which varies inversely with the distance between them, the potential
energy due to dipole-dipole force is given by
PY
V=— a aal (15.18)
where jz is the dipole moment, r is the distance of separation of the dipoles,
and k is Boltzmann’s constant. It is evidently a force effective at somewhat
shorter distances than simple coulombic attraction. It is this type of force which
is largely responsible for the crystal lattice energy of substances like HI, SO,, etc.,
and which must be overcome in the vaporization process.
Nonpolar molecules also exert an attractive force on each other, although
clearly a weak one. The boiling points of nonpolar substances such as N, and
O, are much lower than those of polar substances of comparable molecular
weight. The nature of this force was first explained by F. London. It arises from
the fact that a nonpolar atom or molecule is, in effect, an oscillating system of
electrical charges and has at any instant a small electric moment which varies
with time in such a fashion that the average moment is zero. This moment
induces an opposite moment in nearby molecules and results in a small attrac-
tive force called a dispersion force. The potential energy of attraction due to
this force is given by
3hvy a?
—— 4 fr
(15.19)
where vy, is the frequency of oscillation of the charge cloud and a is the polar-
izability of the molecule or atom. This force is one of the effects responsible
for “van der Waals attraction” in gases. For polar substances the dipole-dipole
force is usually of larger magnitude.
Seca 15a Some Properties of Solids and Liquids 479
V=—2t r2 (15.22)
but this arrangement is not ordinarily achieved since it is opposed by thermal
motions. Ion-dipole forces are often important in solution, but for this appli-
cation equation (15.22) must be modified, for example by considering the ef-
fective dielectric constant of the medium. The appreciable solubility of many
ionic solids in water is due in large part to the fact that the energy resulting
from the interaction between the dissolved ions and the dipolar water mole-
cules is comparable to that of the crystal lattice. (A complete analysis must also
take the entropy of solvation into account.)
On the other hand the energy of interaction of an ion with a nonpolar mol-
ecule, which may be characterized as due to ion-induced dipole force, is given by
— _ 1 ae; (15.23)
and is a weaker and shorter range attraction. This energy of interaction cannot
ordinarily supply the energy requirement for dissolving an ionic substance
and consequently such substances are not readily soluble in nonpolar solvents.
An interaction also occurs between dipolar and nonpolar molecules due
to the induction of a dipole moment in the latter. This dipole-induced dipole
force leads to a potential energy of interaction given by
2a”
V=— “- (15.24)
Although this interaction has the same dependence on distance as does the
dipole-dipole interaction, it is usually of much smaller magnitude, since a is
less than j1?/3KT for molecules with an appreciable moment. (See section 14.10.)
480 Some Properties of Solids and Liquids Chap. 15
It is in part for this reason that polar and nonpolar liquids are often immiscible,
for the formation of a mixture requires substitution of the weaker dipole-in-
duced dipole interaction for the stronger dipole-dipole interaction.
SUMMARY, CHAPTER 15
1. Crystal properties
Crystal systems; cubic, tetragonal, orthorhombic, hexagonal (rhombo-
hedral), monoclinic, triclinic.
Miller indices (hk/): 100, 010, 001 for the faces of a regular cube.
Cubic lattices: simple, body-centered, and face-centered.
Close-packed lattices: hexagonal and face-centered cubic.
2. Xray diffraction
Bragg reflection
nt = 2d sino
Cubic crystal
on 6) = Ss (ee ee
Structure factor for centrosymmetric crystal (0 phase)
PROBLEMS
1. A plane intercepts the x, y, and z axes at 3a, 6b, and 2c. What are its Miller
indices ? Ans. (213).
2. Calculate the longest X ray wave length that may be used to determine a
lattice spacing of 1 A by the Bragg “reflection” method. Ans. 2 A.
3. Find the wave length of the X radiation which shows a second-order
Bragg reflection angle of 14°10’ from the 100 plane of potassium chloride. The
density of KCl is 1.984 gm. cm.~%, and there are four ion pairs in the unit cell.
4. Nickel crystallizes in a face-centered cubic lattice of cell parameter 3.52 A.
(a) Calculate the separation of the 110 planes and that of the 111 planes.
(b) Calculate the distance from a nickel atom to its nearest neighbor in the 100,
110, and 111 planes. Ans. (b) 2.47 A for each.
5. The AH of formation of crystalline lithium bromide from the elements is
—87.6 kcal. mole-!, and the AH of sublimation of lithium metal is 39.0 kcal.
mole-!. The AH for dissociation of Br, is 46.2 kcal. mole~!. Use values of the
ionization potential from Fig. 13.11 and the electron affinity from Table 14.1
to find the crystal lattice energy of LiBr.
6. The AH, of KCl is —104.4 kcal. per gram formula weight. The heat of
sublimation of potassium is 19.8 kcal/g. atom. Use this data, the ionization
potential of potassium from Chapter 13, the KCl internuclear distance (3.14 A),
AH® (disscn) for Cl, (p. 470), and compute a value for the electron affinity of
the chlorine atom.
7. What are the mutual angles formed by the following pairs of intersecting
planes in the cubic system? (a) 100 and 010; (b) 100 and 110; (c) 110 and 101;
(d) 100 and 210; (e) 110 and 111.
8. KBr has a lattice which is equivalent to face-centered cubic with a cube
edge of 6.54 A. The unit cell contains four ions of each kind. Calculate the density
of the crystal.
9. (a) The edge of the cubic unit cell of lead is 4.92A. The density of lead is
11.55 gm./cc. What type of cubic lattice does lead exhibit ?
(b) Calculate the smallest angle with respect to the incident beam at which
482 Some Properties of Solids and Liquids Chap. 15
12. From the ionic radii given in Table 15.4, find the density of RbBr in the
case (a) of a simple cubic lattice and (b) a body-centered cubic lattice, as-
suming that the ions “touch” along the diagonal of the cube.
14. Predict the angles of first-order Bragg reflection of 0.586 A X radiation from
the 100, 110, and 111 planes of CsCl. This substance has a density of 3.97 g.
cc.~!, and a body-centered cubic lattice. Note that the intensity of scattering from
the Cs ions will be much greater than from the Cl ions. To what intensity varia-
tions will this lead?
15. Calcium oxide crystalizes in a cubic latice. The density of calcium oxide is
3.37 gm./cc. By using Mo X rays (x = 0.712 A) lines are observed at the follow-
ing angles: 7-237, 830%, 12°67, 14°12” 14°51’; 17-307 and 18°20”
(a) What type of lattice is this?
(b) How many CaO units are there in one unit cell?
(c) What are the values of dj), di;,, and d,;,?
16. Copper metal shows a cubic crystal habit. X ray diffraction with radiation
of wave length 0.586 A shows maxima at the following angles: 8°5’, 9°19’, 13°15’,
15°35%,. 16°18’, 18°55’, and 20°41’.
(a) What type of lattice is this?
(b) Find the dimensions of the unit cell.
(c) Find the density of copper metal.
Ans. (a) Face-centered cubic, (b) 3.62 TN (c) 8.90 gm./cc.
17. The Copper K. X ray line (WX = 1.541 A) is used to investigate the structure
of an iron crystal. Lines are observed at the following scattering angles: 22°23’,
32°35’, 41°16’, 49°37’, and 58°23’. No other lines are observed.
(a) What type of cubic lattice is this ?
(b) What is the unit cell dimension?
(c) The density of this phase of iron is 7.86 gm./cc. Estimate Avogadro’s
number.
(d) Is the copper K, line the best choice for this analysis?
Ans. (a) Body-centered cubic, (b) 2.86 A.
18. Referring to Fig. 15.16, find the valence bond angle of carbon in diamond.
19. An X ray powder photograph is taken with the Co Ky X ray (A = 1.790 A).
Problems Some Properties of Solids and Liquids 483
The first seven lines are found at the following values of sin? @: 0.0343, 0.0917,
0.1258, 0.1370, 0.1839, 0.2752, 0.3097.
(a) Show that the data correspond to a cubic crystal.
(b) Compute the dimension of the unit cell.
(c) Compute the values of (A? + k? + 1°) for each observed sin? 6.
(d) Compute the allowed values of (h? + k? + 1°) for a face-centered cubic
lattice for all possible combinations of the integers 0, 1, 2, 3, 4, and compare to (c).
What lines are missing?
(e) Could the crystal be face-centered cubic?
(f) Explain the missing lines.
Ans. (b) 8.38 A; (c) 3, 8, 11, 12, 16, 24, 27; (d) 4, 19, 20; (e) yes.
20. The coordination number of a cation in a salt can be related to the relative
sizes of cation and anion. Thus, the sodium ion in NaCl shows a coordination
number of six, whereas the larger Cs ion in CsCl shows a coordination number of
eight.
(a) Draw a cube with eight anions at the corners, and a cation in the center.
Assume that the anions just touch and show that for anion-anion contact the
maximum value of the ratio of the radius of the cation to that of the anion is 0.732.
(b) Show that for an NaCl type structure anion-anion contact occurs at a
radius ratio of 0.414.
21. Verify the following. The Madelung constant for NaCl is the sum of a series
of terms of the form
nH
where n, a, b, and c are all integers. When the sum (a + b+ c) is odd, the term
is positive, and when (a + b + c) is even, it is minus.
22. Neutrons may be used in diffraction experiments since they possess wave
properties. A beam of neutrons is allowed to impinge on a diamond crystal. The
density of diamond is 3.56 gm./cc., and the unit cell is shown in Fig. 15.16.
Reflections are obtained at various angles, the smallest being 10 deg.
(a) Show that the reflection at 10 deg. originates from the 111 rather than
from the 100 (or 200) or the 110 (or 220) planes.
(b) What is the wave length of the neutrons? Ans. 0.712 A.
(c) What is the energy of the neutrons? Ans. 2.8 X 1078 ergs. neutron.
23. The common mineral rutile (TiO,) has a tetragonal unit cell with a = 4.49 A
and c = 2.89 A. Titanium atoms occupy positions at (0, 0, 0), and (4, 4, 4). Oxy-
gen atoms are at (0.31, 0.31, 0); (0.81, 0.19, 0.50); (0.69, 0.69, 0); and (0.19, 0.81,
0.50). [R. W. G. Wyckoff, Crystal Structure, Interscience Publishers, Inc., New
York, (1948- ) Vol. I, Chapt. IV.]
(a) Draw a diagram of the unit cell.
(b) Calculate the density of rutile.
(c) Find the average separation of Ti and the six nearest O atoms.
24. For a tetragonal system two axes are equivalent, and
. Ona = (#5) 8 +) i + P
sin? 45
484 Some Properties of Solids and Liquids Chap. 15
sin? Ou — (AG) (+ 2) = 25
2 272
The first nine lines on an X ray powder photograph of CuAl, have the following
values of sin? @;,;,: 0.0440, 0.0882, 0.1449, 0.1767, 0.1811, 0.2204, 0.2245, 0.3117,
0.3554. [See N. F. M. Henry, H. Lipson, and W. A. Wooster, The Interpretation
of X-ray Diffraction Photographs, Macmillan and Co., Ltd., London (1960).]
(a) Assume that the first line is the 110 line, and compute 2/4a?.
(b) Compute all possible values of sin? @,,%, — (A?/4a?)(h? + k?) for all values
of (A? + k?) less than 16.
(c) What is the value of X2/?/4c?? [This can be recognized as follows. For
some lines, one of the allowable values of sin? @,;, — (A?/4a?)(h? + k?) will be
0. For other lines it will not. For this last class of lines there will be a value
of sin? @n%, — (A2/4a?)(h? + k?) which is C, (2)?C, (3)?C, or (4)2C, where C is
W1?/4e?.
(d) Assign indices to the observed lines.
Ans. (c) 0.0340; (d) 110, 200, 121, 220, 112,
310, 202, 222, 312.
Ih
PHOTOCHEMISTRY
AND RADIATION
CHEMISTRY
16.1 Photochemistry
a — kel (16.1)
where / is the intensity of light and k is a constant of proportionality which
depends on the nature of the absorbing substance and the wave length of the
light. Integrating over a finite thickness of absorber, we obtain the Lambert-
Beer law,
ines (16.2)
I,
where /, and J are the incident and emergent light intensities. When c is molar
concentration, & is the molar extinction coefficient for path length / in centi-
meters.
EXERCISE 16.1
It is found that a substance present at 10-* M absorbs 10% of an incident light
beam in a path length of 1 cm. What concentration will be required to absorb
9075 Ans. 0.022 M.
It has been shown (Chapter 13) that when an atom absorbs light it
exhibits a discrete line spectrum whose well-defined frequencies are associated
with the formation of correspondingly well-defined, excited electronic states of
the atom. An isolated excited atom will eventually re-emit a quantum of light
by fluorescence and return to its ground state. In the presence of other gas
molecules the excited atom may transfer its energy in a collision. Since the de-
excited atom can no longer fluoresce it is said to have been quenched.
488 Photochemistry and Radiation Chemistry Chap. 16
A* — A+ hv ky
= Tear) (16.4)
and when the fluorescence is not forbidden by selection rules, t,,. = 0.693/k,
is typically ~10-* sec.
At moderate gas pressure there is competition between fluorescence and
energy transfer to a quencher, Q
MeO AsO ik
The total rate of loss of excited atoms is given by
I, en
Ni 1 16.7
I, 5 TF EEN ane
The ratio of velocity constants k,/k,; can be evaluated from measurements of
the fluorescence yield y, for known concentrations of quencher. When k;, is
known from the measured half life of fluorescence in the absence of quencher,
then k, can also be evaluated. Quenching occurs in bimolecular collisions be-
tween A* and Q and the rate of such collisions is given by the kinetic theory
of gases (see section 11.10)’ as
Identifying the rate constant for quenching with the factor in brackets, we
find that the effective quenching cross section og =ee is given by
EXERCISE 16.2
As Table 16.1 shows, the efficiency with which hydrocarbons quench excited
mercury atoms is quite high. The mechanisms by which this quenching occurs
are not yet well understood, but a common consequence is the rupturing of a
bond in the quenched molecule. This should not be surprising, since excited
mercury atoms in the *P, state have available 112 kcal. per mole of excitation
energy.
A chemical reaction induced in one substance by energy transferred from
another light-absorbing substance is said to be a photosensitized reaction. The
uranyl oxalate actinometer, mentioned previously, depends on such a process.
Most work has been done with metal vapors, particularly mercury.
Many substances whose free radical reactions would be of interest are not
readily photolyzed directly, because they are quite transparent over the easily
accessible range of wave lengths. This is the case for paraffin hydrocarbons,
which are transparent to wavelengths far below the transmission limit of fused
silica, although the bond dissociation energies for C—H and C—C correspond to
quantum energies at 2700 A or longer wave lengths. Therefore photosensitizers
are used.
The photosensitizer should be relatively inert chemically and able to absorb
light in an effective spectral region. In gaseous systems particularly, mercury is
very satisfactory. The vapor strongly absorbs the 2537 A and 1849 A (reson-
ance) lines of the mercury spectrum.
Hg('S,) + hv(2537 A) — He(?P,)
He('S,) + hv(1849 A) — He(“P,)
Photosensitized decompositions of hydrocarbons have been studied extensively
(see E. W. R. Steacie, Atomic and Free Radical Reactions, 2nd ed., Reinhold,
490 Photochemistry and Radiation Chemistry Chap. 16
New York, 1954, pp. 411 ff.) and it appears that the net result of the primary
process can be represented as
Hg(*P;) + RH — Hg('S,) + R + H
The mercury sensitized decomposition of ethane yields methane, propane,
and butane. The following mechanism has been proposed to account for these
facts:
Hg(?P,) + C,H. — Hg(S,) + C,H; + H
H + C,H, — C,H; + H,
H + C,H; — CH; + CH;
2 CH, — CoH,
CH, + C,H; — C,H,
2 C,H; — C,H
2H — H,
or free radicals may combine without activation energy, the diatomic particle
formed is quite unstable, since it contains sufficient internal energy for disso-
ciation. Unless relieved of at least part of this energy, as by collision with a
third body, it will dissociate within one vibration period, ~ 10~'° sec.
A mechanism consisting of steps (1), (2), and (4) accounts for the observed
quantum yield of 2.0. It may be asked why steps such as
R, _ _k.(HD)G4)
R, — k,(H)(M) C4)
Bimolecular collision frequencies are much greater than termolecular at a
given pressure and, in addition, (H) is quite small under normal experimental
conditions. Similarly, step (8) cannot compete with step (2).
EXERCISE 16.3
Suppose that hydrogen atoms react with hydrogen iodide molecules on every
collision (process 2). According to the kinetic theory of gases, the value of the
bimolecular rate constant for such a process will be ca. 10!° mole™! 1. sec.7}.
Find the steady-state concentration of hydrogen atoms in hydrogen iodide at
25° and | atm. pressure which is absorbing light at the rate of 10!° quanta cc.7!
SeCames Ans. 4.1 < 105° mole ls:
Hydrogen atoms will react appreciably with iodine by step (6), unless (I,) is
kept very small. At room temperature, the small equilibrium vapor pressure of
iodine satisfies this condition, but at elevated temperatures step (6) must be
included in the mechanism.
TABLE 16.2
BOND DISSOCIATION ENERGIES* (kcal. mole)
H CH; Cl Br I
H 103
CH; 101 83
Cl 102 80 57
Br 87 67 52 45
I 71 53 50 42 36
(By = [7
eG) Oi) hs 2
(H) = (Br,)
© + &,(HB5)
d(HBr)
dt
= k,(H,) ,/2k5)'”
(16.9)
‘1 (k,/k,)(HBr)/(Br,)
+
The ratio of rate constants k,/k, also appears in the thermal rate law, has
the same value, ~0.1, and shows no temperature dependence. The temperature
coefficient of the numerator is found to correspond to an activation energy of
17 kcal./mole, and since k; for the recombination of bromine atoms is expected
to have zero activation energy, this is identified with the activation energy for
step (2) in the mechanism. This agrees satisfactorily with the prediction from
the bond dissociation energies of Table 16.2.
EQ) > Daa — Dap, = 103 — 87 = Tokeal-<mole“*
Bond dissociation energies are very useful in understanding many reactions
involving atoms and free radicals. For example, we may compare the various
halogen atom-hydrogen molecule reactions,
xX + H,—
HX +H
where xX == Cl Daag Dee. = wkCabsmoleas
Sis == 16
=| = 32
In view of these figures it is not surprising to find that the photochemical reac- _
tion of chlorine with hydrogen is explosive, while that of bromine occurs at
a moderate rate around 200°, and that of iodine has no appreciable rate.
When the energy of the quantum absorbed exceeds the bond disso-
Sec. 16.4 Photochemistry and Radiation Chemistry 493
ciation energy of the absorbing molecule, we may expect the excess energy
E' = hv — D (16.10)
to appear as relative kinetic energy of the two fragments. When hydrogen iodide
absorbs 2537 A radiation the excess energy amounts to 112 — 71 = 41 kcal.
mole! if the iodine atom is formed in its ground electronic state. Since
momentum must be conserved between the two fragments
MyVy = MyVz (16.11)
En
E! a LO
Teg aLae 27, (16.12)
Evidently the hydrogen atom acquires most of the energy. It is a hot atom and
will be shown as H*.
EXERCISE 16.4
Find the kinetic energy of the hydrogen atom produced by photodissociation
of hydrogen bromide with light of 2537 A wave length. Assume the bromine atom
is in its ground electronic state. Ans. 25 kcal. mole7}.
or lose their excess energy and become thermal atoms. Of the two possible
thermal reactions of deuterium atoms
(3) D+6H,— ADA CH,
(4) D2UDi==D, a
the second has much the smaller activation energy. In this system, then,
deuterium iodide will act as a scavenger for deuterium atoms. Since deuterium
iodide is a minor component, it cannot often react with hot deuterium atoms.
Thus step (4), the thermal reaction, yields D,, and step 1, the hot atom reaction,
yields HD. The observed ratio HD/D, is ca. 0.25 when C,H, is in large excess.
Hot atoms can be produced by nuclear techniques as well. The *He nucleus
reacts with thermalized neutrons to form the radioactive isotope of hydrogen,
tritium (T or *H) and a proton, *He(n, p) *H. This nuclear reaction is exothermic
by 0.78 M.e.v., an enormous energy by chemical standards. This energy is
distributed between the proton and the T atom, with the T atom receiving
0.19 M.e.v., as required by momentum conservation. A 0.19 M.e.v. T atom is
well above the range (below 100 e.v.) where chemical reactions can occur, and
enters this range at an energy inaccessible by other means of activation. As a
result, it can efficiently enter into new types of reactions. [R. L. Wolfgang,
“Progress in Reaction Kinetics, Vol. 3.; F. Schmidt-Bleek and F. S. Rowland,
Angew. Chem. Intern. Ed. Engl., 3, 769 (1964).] For example, although the
characteristic reaction of thermal hydrogen atoms with hydrocarbons is ab-
straction (reaction 3 above), these very energetic tritium atoms can displace a
hydrogen atom. :
MgO reflector
es 16.5 Flash Photolysis
spectroscopic flash reaction vessel With — ordinary
trigger MgO reflector g pals photochemical techniques,
electrode the concentrations of atoms
and free radicals produced
| are too small for detection
vacuum even by the most sensitive
system
Sec. 16.6 Photochemistry and Radiation Chemistry 495
ciently small initial separations. In this case the primary process resembles
a vibrational excitation and the incipient bond rupture is reversed within ~
10-13 sec. When the initial separation is greater and the radicals escape from
the “cage,” viz., by separating by one or more mean free paths, they begin a
random jumping or displacement by diffusion. The jump frequency is ~
10'' sec.~!; the mean free path for diffusion may be estimated at a few tenths of
a molecular diameter.
Diffusion displacements on the average tend to separate the initially con-
tiguous pair of radicals, but in their random motion some radical pairs will
re-encounter each other while they still lie within a fairly small volume. If we
think of the molecules as occupying the positions of a cubic close-packed
lattice, and the radicals as occupying two adjacent sites, a single displacement
of one relative to the other can move it to any of 12 adjacent sites of which
only one is occupied by its former partner. Thus, the probability of recombina-
tion after a second displacement is +4. In each succeeding displacement the
cumulative probability of recombination is somewhat increased, but this rapidly
approaches a finite limit as the most probable displacement from the original
partner rapidly increases with time. R. M. Noyes [J. Chem. Phys., 18, 999
(1950)] has estimated that for dissociation of iodine in hexane solution, the
probability of recombination of original partners is 0.40.5. This has been called
geminate recombination to emphasize that the two recombining atoms came
from the same original molecule, and is to be distinguished from random
recombinations of atoms from different molecules in the system.
In the absence of a scavenger for free radicals, they diffuse in a random way
through the solution and eventually combine completely with each other. A
scavenger added to the system in small concentration prevents steady-state
recombination, but at concentrations less than 107‘ mole fraction it cannot be
expected to interfere measurably with the geminate recombination. At mole
fractions greater than 10~? an efficient scavenger can begin to react with radicals
which would otherwise have recombined with the original partner.
A typical photolysis in solution may be represented by
(1) AB + hv — (A + B) dissociation
(2) (A + B) — AB geminate recombination
(3) (A+ B)—A+B steady state begins
(4) A (or B) + X — AX (or BX) scavenger reaction
(5) A+ B— AB steady-state combination
Approximately one-third of all initial processes will be followed by primary
recombination, which is not included in the above mechanism since it is of
no experimental consequence. Another one-third will undergo geminate recom-
bination unless reaction with a scavenger intervenes. The remaining one-third
escape into the steady state, and their behavior is treated in the usual fashion
as a competition between steps (4) and (5).
Sec. 16.7 Photochemistry and Radiation Chemistry 497
(2) M+ R — M,—
For example, with ethylene
R + CH,—CH, — RCH,—CH,—
The chain reaction propagates by successive addition of monomer.
(3) M-+M,~ —M,.—
(4) M+M,— —M;—
Gi ae
For the case of ethylene, step (5) would be
F M,— + CH,—CH, — M,—CH,—CH,—
Chains may terminate by combination of any two radicals.
(Gi Mia) Me
When a steady state has been established, the rates of initiation R,; and of
termination R, must be equal. The rate of initiation by photochemical dissoci-
ation is given by
p= 2170 (16.13)
where ¢ is the quantum yield* for production of free radicals. The rate of ter-
mination, step (6), is
R, = 2k(Mm—) (M,—) (16.14)
498 Photochemistry and Radiation Chemistry Chap. 16
Ree Ge ym (16.17)
When the reaction is initiated by a unimolecular thermal decomposition of
initiator at concentration (C), R; is given by
Rae) (16.18)
and
: 1/2
Mise co (16.19)
k,
TABLE 16.3
POLYMERIZATION OF METHYL METHACRYLATE*
The collision of an electron with kinetic energy ~ 100 e.v. with the elec-
tronic structure of an atom or molecule produces excited electronic states and,
when an electron is completely ejected from the atom, a positive ion. It has
2 12 13 14 15 16 i?
m/e
been estimated (but not demonstrated) that ionization and excitation events
occur in approximately equal numbers, but it is ionization rather than excitation
which receives the greater attention in radiation chemistry, because it is char-
acteristic and also more susceptible to measurement. Radiation-induced reac-
tions in the gas phase can be described conveniently in terms of the ion pair
yield, M/N, where M is the measured number of molecules produced or de-
composed during an irradiation in which N ion pairs (positive ion and electron)
were formed. In the liquid phase it has become the practice to report yields in
terms of the number of molecules produced or consumed per 100 e.v. of energy
absorbed by the system, and this is called the G value.
The mass spectrometer (see section 13.6) is not only a very useful
analytical instrument but it is also a powerful tool for the examination of
electron impact processes. It measures both the energy of the ionizing electron
and also, within limits, the identity of the ionic reaction product. Actually, the
positive ions are characterized by their mass-to-charge ratio, but since the
Sec. 116.9 Photochemistry and Radiation Chemistry 501
charge is almost always unity, the result is referred to as a mass spectrum. For
example, methane under 70 v. electron bombardment yields ions of masses
16, 15, 14, 13, 12, and 2 as indicated in Fig. 16.6. The small peak at mass 17 is
due to the natural abundance of !°C, the heavy stable isotope of carbon, which
yields '*CH;, mass 17.
Peak
intensity
The intensities of the various mass peaks in the methane spectrum depend
on the kinetic energy of the bombarding electrons as indicated in Fig. 16.7.
For each ionic species there is a characteristic minimum electron energy, called
the appearance potential AP, which is required to produce the given species.
For CH. AP (16) = 13.12 e.v. When the parent molecule ion so formed is in
its lowest energy state, the appearance potential is the same as the ionization
potential. For the general case of producing an ionic fragment in its lowest state,
R,R, + e = Ri + R, + 2e
the appearance potential AP(R;), bond dissociation energy D(R, — R,) and
ionization potential J(R,) of the radical R, are related by
AP (R= I(Ry) + D(Ri = R,) (16.21)
Appearance potentials of ions have been used to evaluate bond strengths.
For methane, AP(CH;) = 14.4 e.v. Similarly, with the use of special techniques,
the appearance potential of CH; from CH; radicals (equivalent to the ionization
potential of CH,) has been measured. J(CH;) = 9.96 e.v. Substitution in equa-
tion (16.21) yields a value for the bond dissociation energy.
D (CH;—H) = AP — I= 4.4e.v. = 101 kcal./mole
Energy measurements from appearance potentials are not yet so precise as
spectroscopic or calorimetric measurements, but they are providing much infor-
mation not otherwise available. Table 16.4 contains some data obtained by
these methods.
502 Photochemistry and Radiation Chemistry Chap. 16
TABLE 16.4
IONIZATION POTENTIALS AND HEATS OF FORMATION OF THE CORRESPONDING
GASEOUS IONS FROM ELECTRON IMPACT MEASUREMENTS{
H 13.62 365
Hy 15.44 356
CH, ES 333
CH; 9.96 262
CH, 13.12 285
CyH, 11.42 317
C,H; 8.69 280
CoH, 10.56 255
C,H; 8.72 224
CyHe 11.65 249
C3H¢ 9.80 231
s-C3H, 7.90 190
C3Hs LI 234
C,Hs 9.24 209
t-C,Hy 6.90 166
n-CyH jo 10.80 219
EXERCISE 16.5
The appearance potential for CH; from methane is 14.4e.v. and from ethane is
13.95 e.v. Use these data together with the CH;—H bond dissociation energy from
Table 16.2 to find the CH;—CH, bond dissociation energy in ethane.
Ans. 3.9 e.v.; 90 kcal. mole™!.
auoyjaus
G9
Ol
Sl
O
O02
SS
Sz (e\9 G¢ Ov Sv iS
09
OL
aupuye
auodoid
aupjng-u
aupjnq-1
@uDjuad-u
auDjued-!
audjUadoau
503
O02 Sz o¢ s¢ Ob Sb Os
OL
Sl
gs
OL
504 Photochemistry and Radiation Chemistry Chap. 16
The centrifugal force is wwv?/r, being the reduced mass and » the relative velo-
city of the particles. Equating these and solving for r,, the radius of the largest
closed orbit
re = (2e’ a/v?)
This result is surprisingly close to that of a more rigorous calculation of s,,
the maximum distance of separation at which collision occurs between an ion
and molecule.
So = (4e? a/v?) (16.22)
The collision cross section, a, is defined as the area swept out by a circle of
radius s,. (See Fig. 11.15.)
o = 182 = n(4e2a/pv?)'” (16.23)
[G. Gioumousis and D. P. Stevenson, J. Chem. Phys., 29, 294 (1958)]
For measurements in a mass spectrometer the molecule may be considered at
rest relative to the accelerated ion, and »v is essentially the velocity acquired by
the primary ion in the electric field of the source. Since molecular polarizabili-
ties are of the order of 10~*4 to 10°?’ cm’, o can be greater than 100 A’, and is
often more than 10 times kinetic theory cross sections.
In order to establish empirical cross sections, analysis of mass spectrometric
data is necessary. Let i, be the current due to primary ions passing through a
source of path length / at an average speed, v. Let i, be the current arising
from secondary ions formed by ion-molecule reactions with molecules of con-
centration c. Clearly, if only a small fraction of the primary ions reacts, then
: i, will be directly proportional to i,, to c, and to the path length /.
i, = G2iplc (16.24)
The proportionality constant, o,, with dimensions of cm.* is called the phe-
nomenological cross section. Thus, measurement of the ratio i,/i, allows evalua-
tion of o,. It is amazing that the calculated collision cross sections and the
experimental reaction cross sections are often almost equal. It must, therefore,
be concluded that in these instances reaction occurs at every collision, and that
the rate of the chemical process is dominated by the physical situation.
Although the expression of these data in terms of cross sections is instructive,
expression in terms of rate constants which are independent of instrumental
parameters is more meaningful. For a beam of monoenergetic ions of mass m,
and charge e accelerated by a potential field E over a distance / this can be done
as follows. v, the velocity of the incident ions, is given by v = (2Eel/m,)'”.
Assuming that the mean residence time of an ion in the reaction zone is given
by //v, the steady-state charge density in the source is i, /(m;/2Eel)'””, and this is
also a measure of the primary ion concentration, c,. The rate of production of
secondary ions, (dS/dt), which is equal to the current i, (rate of charge flow) is
aSidi—=ke,c = ki,l(m,j2Hel)
Ve = i, (16.25)
where c is the concentration of molecules. Comparing (16.25) to (16.24)
k= o(2Eel/m;)”.
5306 Photochemistry and Radiation Chemistry Chap. 16
The situation in a mass spectrometer is more complex, but analysis of the prob-
lem for Maxwellian distributions of ions and reactants gives a similar result.
k = 0.(Eel/2m,)'” (16.26)
EXERCISE 16.6
HBr
B= k,,(H) (Br) = I
k,,(H)(Br,) + &,.(H)(HBr) = 1 + 0.12 Puyp,/Po,,
An expression for the dependence of the ion pair yield on composition was
formulated by Eyring, Hirschfelder, and Taylor as follows: The total number
of H atoms per ion pair is
aNy, + bNupr + AN (16.27)
The number of HBr molecules decomposed directly is
bNup, + AN (16.28)
Since each H atom has the chance B to form HBr (process 11) and the chance
1 — B to decompose HBr (process 12), we obtain for the yield of HBr per
ion pair
4 =([B-(1
M —— 3 as =
B)]|4eS Nu, Nupr
ei pe)
Nusr
102)
B, Ny,, Nap, and N are known for any experiment. A, a, and b are not known
and are chosen for best fit with the experimental data. The best values are
found to be
a= 6.0; 0b == 2.0, Ar"
The calculated values of M/N are compared with experimental values in Table
16.5. Considering the range of conditions covered, the agreement is fairly
good. The value of a indicates that dissociation of hydrogen by excitation must
be a significant process, whereas it is apparently not important in hydrogen
bromide.
Sec. 16.12 Photochemistry and Radiation Chemistry 509
# TABLE 16.5
RADIOLYSIS OF THE HYDROGEN-BROMINE SYSTEM
Pur iP re 1x 2 obs.
(mm. Hg) Sint tie) quate) Cape e ca
12 OTs 84.9 0.24 0.24
2.4 95.5 250.4 0.88 0.84
5.4 56.6 411.3 1.90 1.79
3.8 26.4 424.6 2.76 2.83
162.3 4.5 4.5 —3.98 —4.70
148.3 10.6 10.6 357 —3.44
84.5 1.3 571.1 = 0g Eig fe
(4) OH + OH — H,O,
In addition, the process
(5) H + OH — H,O
must occur, but this has no directly measurable experimental consequences.
Reactions (3) and (4) in spurs are similar to the geminate recombination
processes discussed in section 16.6, and like these cannot be affected by low
concentrations of radical scavengers. The unscavengable yield of H, and H,O,
per 100 e.v. of absorbed energy is called the molecular yield, and is symbolized
by G,,. However, at ordinary radiation intensities, only a fraction of the H atoms
and OH radicals undergo such geminate diffusion-recombination and the re-
mainder diffuse away from the spurs and become more or less homogeneously
distributed through the system at a net concentration of 10°° M. In this case,
the chance of radical-atom encounters is negligible in comparison with the
chance of reaction with a solute present in concentrations even as low as 10°° M.
When the system contains solutes which react efficiently with atoms and
radicals, the extent of such a reaction can be used to measure the radical yield
Gp, that is, the number of radical pairs which escape from the spur per 100 e.v.
of energy absorbed. In aerated water containing ferrous sulfate and 0.8 NV
sulfuric acid, the following sequence of reactions occurs:
H + O, — HO,
OH + Fet? — Fet*? + OH™
Fet? + HO, — Fe** + HO;
HO; + H* — H,O,
Fet? + H,O, — Fet* + OH- + OH
This amounts to oxidation of four ferrous ions by each radical pair. In addition,
each H,O, molecule produced in the spur will oxidize two Fe*’, and therefore
Gyers = 4Gp + 2Gy, (16.30)
The measured value of G,,.; is 15.5, and this system can be used as a dosimeter
for measurement of unknown radiation intensities in a manner analogous to
the use of an actinometer in photochemistry.
The value of G,, is measured by the hydrogen yield in the ferrous sulfate
system.
Gy, = Gy = 0.45 (16.31)
Combining equations (16.30) and (16.31) yields
_ 05.5 q— 0.90) _ , di
Gp
In the absence of solute, the radicals escaping from the spur react with
any products present via the chain
(6) H+H,0,—H.0-+0n
Sec. 16.12 Photochemistry and Radiation Chemistry 511
Oe 0be-
He H,0e- A
and thus prevent the accumulation of appreciable amounts of H, and H,O,.
This accounts for the low yields in electron irradiation of very pure water.
Alpha particle bombardment, in contrast, produces spurs which are very close
together so that they overlap to form a cylinder in which the local concentra-
tion of atoms and radicals is very high. This enhances G,, and diminishes Gp,
making the net yield of products in pure water much higher.
Although the values of G,, and G, depend only on the character of the radi-
ation in dilute solutions, it has been shown by H. A. Schwarz [/. Am. Chem.
Soc., 77, 4960 (1955)] that, with increasing concentrations of various solutes, the
value of G,, diminishes. This is attributed to intrusion of the solute into the
region of geminate diffusion-recombination. Although the efficiencies of the
various solutes differ, it is remarkable that the shapes of the G,, vs. concen-
tration curves are the same, as shown in Fig. 16.10. Here the concentration of
whe
he
a>
WR
ie
G(Hs)
GlHe),
TABLE 16.6
RATE CONSTANTS FOR SOME REACTIONS OF THE HYDRATED ELECTRON*
16.10) have lead to its assignment as the spectrum of the hydrated electron.
[E. J. Hart and J. W. Boag, J. Am. Chem. Soc., 84, 4090 (1962). See also E. J.
. Hart, Science, 146, 19 (1964).] It has proven possible to measure the rate con-
stants for various reactions of the hydrated electron. Some values are summa-
rized in Table 16.6.
These results are consistent with the interpretation that the reactions occuring
are simple C — H and C — C bond ruptures [R. A. Holroyd and G. W. Klein,
J. Am. Chem. Soc., 84, 4000 (1962)]. In the absence of ethylene scavenger, the
primary radicals react by random combination, or by disproportionation, a
reaction in which a hydrogen atom is transferred from one radical to another
to give two stable molecules; for example,
TABLE 16.7
CONSECUTIVE ION-MOLECULE REACTIONS IN METHANE
SUMMARY, CHAPTER 16
1. Fundamental laws of photochemistry
Grotthus-Draper law: Only the light absorbed by a system is effective
in producing chemical change.
Summary, Chapter 16 Photochemistry and Radiation Chemistry 515
A* + RH—A+R-+H
Dissociation
AB + hv — A+B; E(hv)
> D(A — B)
Photochemical hot atoms
K.E. = E(hv) — D(A — B)
Geminate recombination in liquids
AB + hy — (A + B)
(A + B) — AB
Initiation of polymerization
C+thv—2R
R + CH,—CH, — RCH,—CH,—
ay Elementary reactions of radiation chemistry
Electron impact
Excitation
M+e—M*+e
Ionization
M+e— Mt + 2e
Dissociation
AB + e— At + B+ 2e
Ion-molecule reactions
AB*t + CD — ABCt + D
Electron capture
AB+e—A+B-
Charge neutralization
CD* +e—C+D
516 Photochemistry and Radiation Chemistry Chap. 16
or
CDt
+ B-—C+D-+B
Reactions in spurs (water)
H+H— H,
OH + OH — H,O,
PROBLEMS, CHAPTER 16
1. A 2mm. thickness of Vycor glass is found to have a transmission of 70%
for light of 2537 A. What will be the transmission of a 0.55 mm. thickness of this
glass? Ans. 90.7%.
_2. Iodine is to be determined in solution by optical transmission measurements.
In carbon tetrachloride the absorption coefficient defined by k = [log,) (,/Z)]/c/
is 9001. mole-! cm.~! at 5100 A. Find the concentration which will correspond
to 50% absorption in a 1 cm. cell. Ans. 3.35 < 10~* moles liter™!.
_3. Gaseous hydrogen iodide serves as a simple convenient actinometer for
mercury resonance radiation (2537 A). In a given experiment, the radiation from
a monochromatic light source of constant intensity, in the form of a short capil-
lary, passes through a diaphragm 0.5 cm. x 0.5 cm. at a distance of 10 cm. and is
entirely absorbed in a cell filled with hydrogen iodide. After 10 hr. illumination
the cell is cooled in liquid air, and the noncondensable gas (hydrogen) is collected
and found to amount to 2 cc. at 27° and 20 mm. Hg pressure. What is the total
output of the lamp in quanta per second? Ans. 1.8 X 1017 quanta sec.71.
4. Apply the steady-state treatment to the following mechanism for the photol-
ysis of hydrogen iodide:
HI +hy — H+I Rates —sla
H + HI— H, +I ky
H+IJ, —HI+I!1 k
I Si (Fk I =, fs
.__8. From equation 16.29 and the values of the parameters given below that
equation, predict the initial ion pair yield for radiation decomposition of pure
hydrogen bromide. Ans. 6.0.
_ Jil. Use the electron affinity of Cl from Table 14.1 together with data from
Tables 16.2 and 16.4 to find the energy term for the process
48. Using ionization potentials from Table 16.4, evaluate the difference in the
energy change for reactions of the type
(1) R—CH; — Rt + CH,
as opposed to
(2) R—CH; — R + CH;
for the following:
(a) C,Hé.
(b) C,Hio.
Ans. (a) —1.24 e.v.
44. Using values of AH, of gaseous ions in Table 16.4, evaluate AH for the
following reactions:
(a) n-C,Hi, — C,H + H.
(b) n-C,Hij — s-C,H7 + CHs.
(c) n-C,Hi, — C.Hs + C.Hs.
(d) n-C,Hy, — C,H + Hy.
(e) n-C,Hi, — C,H; + CHy.
(f) n-C,Hi, — C.Hi + C.He.
AH,(H) = 51.5, AH;(CH;) = 31, AH,;(C,H;) = 24 kcal. mole’.
Ans. (a) —1.5 kcal. mole“.
tH. Use data from Table 16.4 to show that for ions listed there the general
type of ion-molecule reaction
518 Photochemistry and Radiation Chemistry Chap. 16
17.1. Introduction
(7
OE
) =Car
V
(17.4)
Therefore, according to the simple kinetic theory of gases C,- for an ideal gas
should be 2.98 cal. deg.-! mole~!. Table 17.1 shows that this expectation is
realized for monatomic gases, but that polyatomic gases have consistently
higher values of C,-. Evidently, these more complex molecules possess energy
in forms other than translational. These other forms of energy are, of course,
vibrational and rotational, as discussed in Chapter 14.
Vibrational and rotational motions can be treated by the methods of classical
mechanics. Both types of motion are found to depend on square terms, with the
actual number of such terms depending on the structure of the molecule. Further-
more, it can be shown that there is a contribution of 4 RT to the total energy of
a system of 6.02 « 10* molecules for each square term, whether translational,
vibrational, or rotational, as will be discussed later in this chapter. This classical
result is summed up in the principal o ipartiti which states that
each square term contributes an equal amount, 4RT per mole, to the in er
of a system of molecules. The three square terms for translational energy give,
of course, 3RT per mole for the internal energy of monatomic gases, as shown by
equation (17.3).
Applying the principle of equipartition first to rotational contributions to
the internal energy, we observe that only two coordinates are required to
describe the rotational motions of a diatomic molecule (noting that rotation
about the internuclear axis is of no consequence). That is, there are two square
terms for the rotational energy, and the contribution to the internal energy
should be E, = RT per mole. For linear polyatomic molecules two coordinates
again suffice, and in this case, also, the contribution to the internal energy
should be E, = RT per mole. On the other hand, nonlinear polyatomic mole-
cules have no such symmetry, and the rotational energy expression will accord-
ingly contain three square terms. This will lead to a rotational contribution of
E, = 3RT per mole to the internal energy.
The vibrational energy of a simple diatomic harmonic oscillator was dis-
cussed in section 14.15, where it was shown that two kinds of energy, potential
and kinetic, are involved. Each of these gives
rise to one square term, and, therefore, a dia- 3R
tomic molecule should have a contribution of E,
= RT per mole to the total internal energy. In és 24
a polyatomic molecule there are many possible ar}
vibrational motions, and it can be shown that d
for a molecule of n atoms the number of vibra- aK
tional modes is 31 — 5 for linear molecules al
and 3n — 6 for nonlinear molecules. k
The foregoing analysis can now be com- 27
pared with experiment, taking L |
TABLE 17.1
HEAT CAPACITIES OF GASES
but this is clearly not the case, at least near room temperature, where values of
~ 5cal. deg.~! mole! are most common for diatomic molecules. The observed
values of C,- for polyatomic molecules are also lower than expected. Evidently
the classical equipartition principle does not fully explain these facts.
If C, is examined over a wide range of temperature, as shown for nitrogen
in Fig. 17.1, the nature of the discrepancy becomes clearer. It appears that at
low temperatures C,- = 3R; that is, there is evidently no vibrational or rotational
contribution to the internal energy. At intermediate temperatures values of C,
around $R are observed for several diatomic molecules and at still higher
temperatures C, = 3R. It will be shown that the first increase in C, is due to
the appearance of the rotational contribution to the internal energy, whereas the
second at higher temperatures is due to the vibrational contribution.
Similar variations in heat capacity with temperature are observed for all
522 Chemical Statistics Chap. 17
n!} 100!
W,
In general, when a total of n molecules are arranged with , in the lowest energy
group, n, in the next, etc., the number of arrangements at constant total energy
is given by
n!
a ee ed)
This expression is now used to determine the relative probability W for all
distributions of 100 molecules among five energy levels at a constant total
energy of five equal quanta. The results of these calculations in Table 17.2
can be understood in the following sense. When the system reaches thermal
equilibrium, it changes at random from one distribution to another. The two
distributions a and b have relative probabilities W, = 100 and W, = 9900, so
the latter will occur 99 times as frequently as the former. In general, the
frequency of occurrence of any distribution k will be given by the ratio of W,,
to the sum of all possible values of W, >W;,. It is notable that one of these
distributions occurs with a much greater frequency than any other.
EXERCISE 17.1
Use equation (17.9) to evaluate W, and W, in Table 17.2.
TABLE 17.2
THE DISTRIBUTION OF EQUAL QUANTA AMONG MOLECULES
Number
Entry quanta n; nN Ng Ng ny No logW log >wiw W/s>wcy%)
a 5 1 0 0 0 0 99 2.000 5.964 0.00
b 5 0) 1 0 0) 1 98 3.996 3.968 0.01
Cc 5 0 0 1 1 0 98 3.996 3.968 0.01
d 3} 0 0 1 0 2 97 5.686 2.278 0.53
e 5 0 0 0 2 1 97 5.686 2.278 0.53
if 5 0 0) 0) 1 3 96 7.196 0.768 Nett
g 5 0 0 0) 0 5 95 7.877 0.084 81.9
h 6 0 0 2 0 0 98 3.695 Sola 0.00
i 6 0 0 1 1 1 H/ 5.987 3)7205) 0.06
J 6 0 0 1 0 3 96 7.196 2.016 1.0
Sec. 17.5 Chemical Statistics 525
It can be shown’ that the probability for unequal distribution, W,,.. -— AW,
is related to that for equal distribution, W,,,., by
molecules of energy ¢;, although the population of the ith state is made up of
different molecules at each instant of time as energy is transferred from molecule
to molecule.
The total number of molecules, 1, must remain constant, and is the sum of
the numbers of molecules in each of the available s quantum states. That is,
NW tntnt...tm+...tn=n (17.10)
i=s
The summation equation 17.10 is conveniently represented by }) nm; or, more
i=0
There are a great many ways to distribute the energy E — E, among n mole-
cules. Another distribution is
neo t me, tmet...tnie+...=inieg
= E— E,
Terms in e, equal zero but have been included for the sake of completeness.
The total number of ways to distribute the fixed, available energy is W,
the number of possible arrangements of n distinguishable objects among s
groups, or
n!
Talo. Joo (1712)
In other words, this is the number of ways to arrange n molecules to give n,
with energy €), 7, with energy e,, n, with energy €,, etc.
Using our previous numerical example as a guide, we identify the equilibrium
state of the system with the most probable state for which W = W,,,,. Following
the usual procedure for maximizing the value of a function, let dW = 0. An
equivalent and more convenient choice is to let dIn W = 0.
Taking the logarithm of W, we have
In W = Inn! — Inn,! — Inn,! —Inn,! —...—Inn,!—...
In W = Inn! — > Inn,! jets)
The second term can be simplified by using Stirling’s approximation for
large numbers, In n! = n Inn — a, yielding?
In W = Inn! — (Inn, — n) (17.14)
Differentiating In W, remembering that 7 is a constant, yields
d In YY = —Sd(n; In i == ni) =0 (17.15)
= In n,dn; = 0 TS)
There are two additional conditions to be satisfied, namely
Sih = on = 0 GET}
Zen, = dk =0 (17.18)
That is, the number of particles and the total energy of the system are both
fixed. In order to satisfy these conditions simultaneously, it is necessary to
multiply two of the equations by arbitrary constants? a@ and 8 and combine
the resulting equations to give
S(in n, + a + Be,)dn, = 0 (17.19)
The variables in equation (17.19) are independent of each other for each term
of the summation. That is, for this equation to remain valid for every possible
variation in dn, it must be that each coefficient of dn; individually equals zero.
That is,
Inn, +a+ Be =0
or
De enters (17.20)
The factor e~* can be evaluated from the summation of n; over all states.
21) ne De Cet)
These two equations are then combined to give the fraction of molecules in
a given state.
ny er
= aria (17.22)
The numerical example in the preceding section showed that increasing the
energy content of a system increased the number n; of molecules having energy
€; in excess of ¢€,. It is well known that in molecular systems the total energy
content of a given number of molecules is greater, the greater the temperature.
For equation (17.22) correctly to describe this behavior, 8 must decrease with
increasing temperature. Accordingly, we define a temperature scale by equation
(17.22) and the relation 8 = 1/kT, where k is a proportionality constant.
Without proof, the constant k is identified with Boltzmann’s constant‘, which
is related to the molar gas constant R and Avogadro’s number N by R = Nk.
The equation of energy distribution (17.22) is now written in its final form
—e€/kT
nr @
n Dear
(17523)
The relative population of any two quantum states p and q is given by the
ratio n,/n, obtained by application of equation (17.23).
Ny nes een
Ne nee (17.25)
= e7 (en—€a)/KT
If these two states are degenerate, that is, if e, = e,, it follows that they are
equally populated. If «, > e,, then n, < n,. Note that as long as e, — eg < kT
the populations will be of the same order of magnitude. Furthermore, if there
are many low-lying states of energy substantially lower than kT, they all will
be well populated relative to the ground state. A further consequence of such
uniform distribution is that no one state is heavily populated. Such distribu-
tions correspond to large values of Q. In contrast, widely spaced energy levels
result in very uneven distributions of molecules among states and values of
Q approaching unity. These situations are illustrated in Table 17.3 and Fig.
17.2. Equal energy intervals have been taken for convenience but are not essen-
tial to the argument. The Boltzmann distributions in the figure and table
illustrate that molecules are largely confined to their lowest energy levels when
€ >> kT. Conversely, small energy gradations from level to level result in more
nearly uniform distributions of molecules among the available states.
Secamli/ao Chemical Statistics 529
TABLE 17.3
EFFECT OF QUANTUM SIZE UPON POPULATIONS OF STATES
State e[/kT e7-ekr n/n e;{kT eees/kr n;|n e,/kT e~&x/kL n/n
90
80
70
& 60
xe]
a 50
a
> 40
xs
30
FIG. 17.2. Effect of 20
quantum
size on pop-
ulations of states. 10
O
QO Assoc ry OMIR2SSF 4 t5NGr 7. 889 OMNCRSE4sOnoameseS
State State State
The energy term ¢ in the preceding equations represents the total energy of
the molecule. It includes contributions e; for translation, e, for rotation, e, for
vibration, and e, for electronic excitation. It may be expressed
e=etetate (17.26)
By limiting our considerations to moderate temperatures, it can be assumed
that there is no interaction between the various degrees of freedom. It is
readily seen that the amount of internal energy should not affect the transla-
tional energy. On the other hand, electronic excitation may appreciably affect
force constants and, therefore, the vibrational energy. Also, in higher vibrational
states the internuclear separation differs from that in the ground state. This
530 Chemical Statistics Chap. 17
affects the moment of inertia and therefore the rotational energy. Except at
relatively high temperatures the assumed freedom of interaction does not lead
to serious error. It is then possible to write
e kr == en (ert er tent Ce) /KT
(17.29)
&% =e “*/kL
oF —gie
O & + ole d
This equation can be written in alternative form, since it follows from equation
(17.24) that
a4
C=
Oe
en ee
Si€i
(17.30)
The differentiation is carried out at constant volume, because the energies of
the various states, €9, €,, €2, etc., may be volume dependent. Multiplying each
term in equation (17.30) by kT?/Q yields
5 (52). = Bete
2 —€9/kT
|a
—e,/kT
os. oo
Ap metlicl
Bee TOicih:
The internal energy of the system can now be expressed in terms of the partition
function of the system by combining equations (17.29) and (17.31).
Sec. 17.7 Chemical Statistics 531
It follows at once that the molar heat capacity of a system whose energy
distribution is described by the Boltzmann law is given by
_ (@E\ _f[ @ (RT?é@Ing
ce = (F), =| a oT wk Ces
EXERCISE 17.5
Use the data of Table 17.3 and equations (17.11) and (17.23) to evaluate E — E,
for 1 mole of gas when e/kT = 5, 10, 15, etc., and compare with e/kT= 1, 2, 3,
etc.
The total energy of a system can be conveniently resolved into its separate
contributions from translation, rotation, vibration, and electronic excitation
because the partition function can be correspondingly factored.
Expressing equation (17.28) as
In Q = Inq, + Ing, + Ing, + Ing.
and substituting into equation (17.33) gives
or
Wie OF er Oe (17.38)
EXERCISE 17.6
Derive (17.38) from the expression starting with (17.14) and using the method
outlined in the text.
Substitution of this result into equation (17.37) and using (17.33) gives
E—E,
S=RiInQ+
ip
~ ning ar),
(17:39)
F alnQ
where Stirling’s approximation has been used to substitute for V!. Substituting
equation (17.40) into equation (17.37), the Boltzmann relation, we find that for
the ideal gas
(= E,)
S = Nk ln O — Nk in N+ Nk
T
or
crystal, the internal energy, E, of both is given by equation (17.33). (See exercise
ade)
EXERCISE 17.7
Use equation (17.40) and show that the Boltzmann distribution law, equation
(17.23), which was developed for a crystal, is also valid for an ideal gas, and that
therefore equation (17.33), which is based on (17.23), gives E for an ideal gas.
Show that the pre-exponential factor in equation (17.44) depends upon the
entropy change of the reaction.
Sec. 17.8 Chemical Statistics 535
The partition function for translation is the same for all perfect
gases. Since it is a factor of the total partition function, it is an advantage to
determine it once for all. The increment of energy between adjacent energy
levels of translation is so small [see equation (17.6)] that the partition function
can be evaluated by integration over all quantum states. In other words, the
function is effectively continuous.
For one degree of translational freedom, say along the x-axis, let h?/8ma?
of equation (17.6) be represented by e,. Then, by the definition of the partition
function and considering that n, > 1, we have
—— SS e- WK — Ss e7 (M2? = Nex/kKL
0 Nz=1
Rotational energy levels are sufficiently closely spaced that it is again possible,
where c? = h?/8mkTa,”
see Appendix 1.
536 Chemical Statistics Chap. 17
Ce en Aelk? — e140
No
Still higher states will have still smaller populations, and we conclude that at
298°K practically all molecules of HCI are in the ground vibrational state and
that q, is 1.000001.
Weak valence bonds and heavy atoms make for low fundamental frequencies
and for vibrational excitation at lower temperatures. The fundamental frequency
for I, is w’ = 214.57 cm.~', where ow’ = o/c. If we replace ha/kT by hco'/kT.
= 1.035, the distribution between the ground and first excited vibrational levels
becomes, at 298°K,
My
n __ g-hew'/k?
s eq __ 9-108
~ . __ ()
355
No
ity
Ay _g-*hew
= aanKP g-2.
= — 0)196
No
K — Quo
X Qr 4-azyer
Om X Qa
Referring to equations (17.46) and (17.50) we can write
q, = A,M*”
and
_ Al
Cp a
where the A’s represent collected constants. For both HCl and HI, as with
many diatomic molecules, the vibrational quanta are sufficiently large so that
at ordinary temperatures g, = 1. The electronic partition functions for I and
Cl are the same.
Because of the syminetric nature of the reaction all factors in A, and A,
and the electronic factors will cancel to give
EXERCISE 17.11
Use the preceding data to verify the expression for the equilibrium constant and
evaluate K at 600°K. ANS aes a0:
=
Co =|Fp(RT
ee ee
oT Cae ioe RT Yi es (17.54)
The translational contribution to heat capacity is the same for all gases,
without regard to molecular complexity, and is independent of temperature.
The rotational contribution to heat capacity for linear molecules follows
from equations (17.34) and (17.49) (except at very low temperature).
Inq, = In T + const.
C=Bair
|op(RT oa
are) |ae
= ao(RT?| x +1)Ss
a ie
Similarly, for nonlinear molecules, by equations (17.34) and (17.51)
Sec. 17.8 Chemical Statistics 539
Ing, = In T + const.
ei Mi
Pata 2a
a a 2 (RT 2x sp)| =4R
=|% (17.56)
That is, the rotational contribution to heat capacity may be either 2 R or 2 R,
according to whether one has diatomic and linear polyatomic molecules on
the one hand or nonlinear polyatomic molecules on the other hand.
The vibrational heat capacity can be obtained from equations (17.34) and
(17.53). Again, let U = ha/kT = hco'/kT= 1.439@'/T, where hc/k has the value
1.439 deg. cm. and a’ is in cm.~'. From equation (17.53)
é Inq, Ue="
Or. Tae)
it follows that
U2e"
es tor|RT Td =e") ts= Rap
=e igs 2 Wes? |
(17.57)
To determine the numerical contribution to heat capacity from vibrational
excitation, it is necessary only to have values of the fundamental frequency
from spectroscopic measurements. Some useful values are given in Table 14.6.
Calculations of C, are simplified by using tables of values of the function
U’e"/ (e” — 1)? (M. Dole, op. cit.).
EXERCISE 17.12
Find the vibrational heat capacity for H**Cl at 1000°K. Ans. 0.26R.
Ga Cr GelG, —2R+C,
The molar heat capacities at constant volume for FIG. 17.4. Calculated
N,, Cl,, and I, have been calculated by the pre- heat capacities.
ceding equations and represented in Fig. 17.4
over a range of temperature. All values refer to
the ideal gaseous state, even at low temperature.
The approximate temperature at which the vib-
rational energy begins to contribute to Cy, is
clearly higher, the higher the characteristic fre-
quency. Unless the absolute temperature is at least
rotation
one-fifth as great as hcw'/k, the value of q, is
nearly unity. Thus, over a range of temperature, rare gases
=
its temperature derivative is almost zero, and the
vibrational heat capacity is effectively zero. At
» T
such low temperatures the vibrational degree of
translation
freedom is said to be “frozen in.” |
pacity unless very heavy atoms are present or the temperature is high. On the
other hand, large molecules also have bending or deformation frequencies
which are smaller and may contribute significantly at ordinary temperature.
Internal rotations often involve potential barriers so that motion is “restrict-
ed” and tends to be oscillatory, resembling loose vibrations.
When there are several vibrational modes in the molecule each with its
own characteristic frequency |, @3, @;,..., the vibrational partition function
becomes GS,.. =(G,), < (as ~ (Gy): ~ -. Each” ofthese terms makesna
corresponding contribution to the vibrational heat capacity which may be
expressed
C= (Cy); ate (Cy)» iE (Ci)3 STs
The partition function q, is to be evaluated over the ranges 0 < x < /, and
—oeo < p, < + oo. From equation (17.24) and the condition g = dx dp,/h we
obtain
feet Sheer c|) an|tememe
w=lz PDr=0
me eX a (17.61)
t= jp es
Since we wish to know the number of molecules dn within the interval dp,, at
all x, the numerator of equation (17.60) is to be integrated over the range
Oriya J OF
=e
numerator = { Chat pnax
x=0
nN (
< en Pat /2mMKT Jy
17.62
( )
1 GxkTimy? * (17.64)
Equation (17.64) is a statement of the Maxwell-Boltzmann distribution law
for molecular velocities in one dimen-
sion. The fraction of molecules within
FIG. 17.5. Maxwell-Boltzmann dis-
the velocity interval from v, to v; + dvz tribution in one dimension.
is very small. Accordingly, the fraction
per unit velocity interval dv,, viz., (dn/n) 10
(1/dv.), is chosen in Fig. 17.5 to rep-
resent the distribution as a function of :
the energy. The choice of (mv?,/2kT)'”
as the independent variable is conven-
ient since otherwise a separate numeri-
cal relation would apply for each mo- 05
lecular weight and each temperature. l=
The distribution of velocities in three :
dimensions follows in the preceding
pattern. Letting p* =p: + p, + pi. we
find that the fraction of molecules with he
momenta in the interval dp, dp, dp, is 6 as WS ie 50
(mv2/2KT)2
542 Chemical Statistics Chap. 17
nN oo 3 ; © : ; oo : :
| e era D | e BE || e LEG 14),
Integrating over the energy range from « to co gives the fraction of molecules
having a kinetic energy in excess of a chosen value e. For large values of ¢
n, co Jel/? e-/kT Vn
ay ale wP(kKT de = 2° (5) Cae
EXERCISE 17.13
By analogy with the preceding equations, show that the fraction of molecules
with energy in two dimensions equal to or greater than e is given by n./n = exp.
(—e/kT). This is the relation used in simple collision theory of reaction rates
(Chapter 11).
Average values of the molecular speed and energy are frequently required
in applications of the Maxwell-Boltzmann distribution law. In general, the
average value f of a property r is obtained by
where n; is the number of particles having each particular value of the property,
r;. When n is a continuously varying function of r, the summations are replaced
by integrals, so in the case of the molecular speed we have
Sec. 17.10 Chemical Statistics 543
el Ba es
ee f. cdn
pl NS
c= 4x (step) J, cemenrat
x eevene
(17.69)
This integral is of the form
oS 3 p-ax?
J, xe dx =o= 75
and, therefore,
c ik (eery"
am
(17.70)
This result was given previously (equation 11.49) without proof.
EXERCISE 17.14
Use the procedure described above to verify equation (11.48).
sa
3kT 1/2
(2)? a (
The average molecular velocity in one coordinate, |#|, will be required for
application in the next section. In all of the foregoing discussion it has been
implicitly assumed that the system as a whole is at rest and therefore 5 = 0
along any coordinate. We shall, therefore, take the average without regard to
sign. Applying the procedure described above, we have
: Lye
oe - I.vdn
i ie Ok 3
and, therefore,
: ieee (17.71)
[Ohi sa)
The average kinetic energy may be obtained in analogous fashion.
é= ie edn,
nN Jo
and, therefore,
€ = 3kT (7.72)
It is notable that for three forms of energy, translation, rotation, and
vibration, there is a common limiting behavior. When the energy levels are
separated by approximately kT or less, the average kinetic energy amounts to
4kT for each degree of freedom. This common result is a consequence of the
type of integral involved. Any form of energy which depends upon a square
term of its coordinate or momentum will lead to an average energy of 4kT for
each such term, i.e., for each degree of freedom. This is the principle of the
equipartition of energy which was stated, without proof, at the beginning of
this chapter.
OO, IE) ea
+ Ox. —Eq/RT __ (X*)
where the activation energy E, has been identified with the energy difference
AE} between reactants A, B and complex X*, all in their respective ground
states.
The problem of describing the rate of reaction now becomes the problem
of describing the rate of unimolecular decomposition of the critical complex.
A plausible assumption would be that the rate of this decomposition is the rate
of breaking of some one valence bond in the complex and that in all other
respects X* is an ordinary molecule. From the preceding considerations we
have found that the greater the energy required in any degree of freedom, the
less probable the event. We shall, therefore, assume that the rate-controlling
bond breaking does not produce product molecules with large relative trans-
lational energy. Some energy must be specified in order to proceed, and it
See S. Glasstone, K. J. Laidler, and H. Eyring, The Theory of Rate Processes, McGraw-
Hill Book Company, New York, 1941.
seems plausible to suppose that the aeoi
product molecules Y, Z are formed by
an act which converts their relative
vibrational motion into relative average
translational motion. The coordinate Potential
describing this motion is called the °"%
product
reaction coordinate. The average veloc-
ity of displacement is further assumed a5
Cc
to be described by equation (17.71), =
je)
(S)
5
(>| = (AT /2am)!”. ®o
S activated
complex
The potential energy function which
describes the reaction, shown in Fig.
17.6, may be compared to a barrier Reaction coordinate
whose width at the top, along the reac-
FIG. 17.6. Potential energy along
tion coordinate, is 6. The complex sur-
the reaction coordinate.
mounts the top of the barrier in a time
given by
ay, (ay
aie eke (17.74)
which we take to be the mean lifetime of the complex. Dividing the number of
complexes by the mean lifetime of a complex gives the rate of reaction. That is,
CS) e( kT ieie: K= (kT y
rate — Sig] oe Qa) (A)(B) os 775)
The partition function for the activated complex can be resolved into partial
partition functions. The partial function for the translation leading to decom-
position corresponds to a displacement of magnitude 5 in one dimension. By
equation (17.46), setting a, = 5 gives
1/2
qo = = (17.76)
The total partition function for the complex is factored to remove gp,
O = Q%.gs, and equation (17.73) becomes
rate ——
Ok (2mkT)"8 | (kT/2m)'?(A)(B) .-2aser (17.78)
O.Ozh 6
The factor of the concentrations is the specific rate constant, which is simplified
to give
HS ap es (17.79)
ho QsQs
The factor kT/h is common to all reactions at a given temperature.
The partition functions Q, and Q, refer to ordinary molecules and methods
previously used apply. The activated complex, by definition, is not an ordinary
545
546 Chemical Statistics Chap. 17
molecule, and the necessary data can only be estimated. The simple reaction
of atom-atom combination will be chosen as an example, since little guessing
is involved.
A + B— AB
The partition functions of atoms A and B are simply those for translation.
The activated complex AB is a diatomic molecule for which the vibrational
partition function has been replaced by a translational partition function, and
this has already been factored out in equation (17.79). The partial partition
function Q{, therefore, includes only factors for three degrees of translation
and for two of rotation. Consequently, from equation (17.79)
[27 (m, + my) kT PPh? 827 Lyk Tho? kT ,- eer
rate k =
QrmkT yh? (Qam,kr yeh h
(17.80)
The numerator of equation (17.80) is g, x qg, for the complex. Now replace
Ixp by Uriy, Where pp = mym,/(m, + mpg) and ray is the internuclear separa-
tion. This gives, after simplifying
In the case of the combination of two atoms, the result obtained from
the activated complex theory must, of course, agree with the simpler collision
theory. The latter theory gives, for the rate of collision of two unlike species
having collision radii r, and rz,
SUMMARY, CHAPTER 17
1. Equipartition principle
Each square term can contribute $RT to E and 1R to Cy.
Translational Energy: 3RT
Rotational Energy: linear, RT; nonlinear, 3RT
Vibrational Energy: linear, (37 — 5)RT; nonlinear, (3n — 6)RT
2. The Boltzmann law
Nn; dle Cae
n oe Lge
Q = 4197rFVe
_ QamkT)”V
he
=
Lawn (linear molecules).
1
do = eo)
EE,
=RT (278
7).
Gre alarS in7) |
$= Rin(£) | ae LR
n (2akT/m)” ds
Speed in three dimensions
ag Are me? 2k
cdc
= OxkT/my® “
Average speed
2
é =
Gz)"
|—
mm
548 Chemical Statistics Chap.
— ee j
~ — =
Ss — ay Ne
=
=
=
= ——= 7 =~ ; =.a
a a t
~ == -
= 3 =
SN
a
6 > 2-
— =
=. .
hand =.
i es, ~~
4
i ~
x
Lhe
‘ee ‘
ee .- -
:
tea - —
APPENDIXES
APPENDIX 1
PARTIAL DIFFERENTIATION
Zi (G D
y
x A
dy = f'(dx = (2) dx
With the function z = f(x, y), a total differential can be defined in a similar way,
d=— (32)
(2% ax + (3)
ay dly
EXPANSION IN SERIES
an interval (a, b). The nth derivative, f”(x), can then be integrated n successive
times between the limits a and x, where x is.a point in the interval (a, 5).
Stig LO
Os ecenie
where
Roe le ay £™x)(dxy"
This is Taylor’s formula with remainder. When a = 0, Maclaurin’s formula
results.
e#=1—-@—-Ft Spee,
sea St ee,
Coste ee
Appendix 1 555
SOLUTION OF (1/r)(d*r/dx?) = —k
This equation is of the same form as (1/y)(dy/dx) = —k, the solution of which
is In y= —kx, or y = e-**, Assume a solution of the same form, yp = e*.
Upon substitution into the original differential equation
1
ec
cfe--* = —k, and c= +ik'”
ap ==C 2 Coe
An equivalent form is
ap = A cos kx + Bsin k'?x
(See above.)
STIRLING’S APPROXIMATION
In(x!) =In2+1n3+...+iInx
This is exactly equal to the area under the step curve shown by the dotted line
in App. Fig. 1.2. This area can be approximated by the integral
+00
EVALUATION OF | e~*' dx
Integrals such as this and (ieee: can be evaluated easily if they are
transformed into polar coordinates. Since the function is symmetric, about 0,
+ co co
I Cady — 2 | Cady
= 0
Ps 4 | || 0 0
e-V'dy = 4 | { en (40) dedy
0 0
In polar coordinates
% — sing
v= 1 COS.G, sand
P=xety
eA iei e-"rdé dr
Since
APPENDIX 2
*Constants are based on the atomic weight !2C = 12.0000. From the report of the com-
mittee on fundamental constants of the National Academy of Sciences-National Research
Council.
558
XIGNAddV
€
318VL
SO ADYANA-SSVW NOISHYSANOD
SYOLOV4
$319 WOI}99T9
S}[OA reaie) SIUIO}LSSBU
1-9[Nosjour 1-9[noejour noejour
1-9] *T “Wye {-9[OU aynof ,-9]0ur *[e9 OW 1-9] wes stun
I Cr79
X u0l SEO'S
X siOI pre's
X nOl €70°9
X 9101 6€P'T
X 901 IN|]
28 TaD! IOL9
x sOl
CO9'T
X si-Ol if 990°8
X cOl €7S°6
X 2Ol 679°6
X vOl 90E°C
X vOl €8L'T
X ee-OT PLOT
X 6-Ol
986'T
X gt-Ol OTT
X v-Ol T [8TT
X 1-OT 96T'T
X 101 6$8°C OTT?
X se-Ol TETX et-Ol
C89'T
X ct-Ol OSO'T
X e-OT OLV'8 T €10'T
xX 20 Icyv'~
xX 101 CLET
X ge-Ol T Ler
X at-Ol
099'T
X u-O1 9CO'T
X s-OI 6S€°8
X z-Ol 698°6
X e-OI I! 06€°7
X 1-OI LP8'T
X se-OT €ITT
X nt-Ol
Lv6e'9
X u-Ol 9EEV
X s-OI 867
X 1-OI P 621
X z-Ol y sgl T OEL'L
X se-Ol 989"
X vi-Ol
8868
X o20T O19'S
X zeOI SSP
X 901 CHES
X ceOI €lps
X 20 T PET
X 1201 I €70°9
X es0I
C6r'l
X e-Ol 6 STE
X sOI €IS'L
X siOl 698°8
X u0T L868
X e10I 8P1'7
X e10l 099'T
X s-Ol i
559
APPENDIX 4
15 1761 | 1790 | 1818 | 1847 | 1875 | 1903 | 1931 | 1959 | 1987 | 2014} 3 6 8 11 14
16 D041 20685) 2095n 21225 21485 2175) 22015 2227) 22538) 2279 |3 eS Salle ls
17 2304 | 2330 | 2355 | 2380 | 2405 | 2430 | 2455 | 2480 | 2504 | 2529|2 5 7 1012
18 DDI) ||PST ||PXSOIL ||PEWS ||PASEES |)PASTA ||GXSSEE NNT/IES |)PAED2||O27) ||22S). 9) GY 12
19 2788 | 2810 | 2833 | 2856 | 2878 | 2900 | 2923 | 2945 | 2967 | 2989 | 2 4 7 9 il
20 3010 | 3032 | 3054 | 3075 | 3096 | 3118 | 3139 | 3160 | 3181 | 320112 4 6 811
21 3222 | 3243 | 3263 | 3284 | 3304 | 3324 | 3345 | 3365 | 3385 | 340412 4 6 810
22 3424 | 3444 | 3464 | 3483 | 3502 | 3522 | 3541 | 3560 | 3579 | 3598 |2 4 6 810
23 3617 | 3636 | 3655 | 3674 | 3692 | 3711 | 3729 | 3747 | 3766 | 3784|2 4 5 7 9
24 3802 | 3820 | 3838 | 3856 | 3874 | 3892 | 3909 | 3927 | 3945 | 3962|2 4 5 7 9
25 3979 | 3997 | 4014 | 4031 | 4048 |.4065 | 4082 | 4099 | 4116 | 4133 |2 3 5 7 9
26 4150 | 4166 | 4183 | 4200 | 4216 | 4232 | 4249 | 4265 | 4281 | 4298/2 3 5 7 8
27 4314 | 4330 | 4346 | 4362 | 4378 | 4393 | 4409 | 4425 | 4440 | 4456/2 3 5 6 8
28 4472 | 4487 | 4502 | 4518 | 4533 | 4548 | 4564 | 4579 | 4594 | 4609/2 3 5 6 8
29 4624 | 4639 | 4654 | 4669 | 4683 | 4698 | 4713 | 4728 | 4742 | 4757/1 3 4 6 7
30 4771 | 4786 | 4800 | 4814 | 4829 | 4843 | 4857 | 4871 | 4886 | 4900 |1 3 4 6 7
31 4914 | 4928 | 4942 | 4955 | 4969 | 4983 | 4997 | 5011 | 5024 | 5038}1 3 4 6 7
32 SOSA OOS MES O79 50925 STOS MESO SISDa ola on SOs eSile2 leo aT,
33 SUS 7 S198 S21 1522471) 5237 5250) 5263) 52761) 5289" 5302") 13) 4) 56
34 53157) S328aih 554053534} 5366) |)5378; > S91nil 54035 541605428 Ilo 4. 6
35 5441 | 5453 | 5465 | 5478 | 5490 | 5502 | 5514 | 5527 | 5539 | 555111 2 4 5 6
36 5563) ||557/55) 9587 5599) SELL 15623) | 5635) |||5647) |5658 41-5670 || 12 4 5.6
37 Sih ||SRY |)SWKOS) || SWAG | SPAD | SWE) Sey ||Sikes! | SwASN SSESALTL 92 By Sy. (6
38 5798 | 5809 | 5821 | 5832 | 5843 | 5855 | 5866 | 5877 | 5888 | 5899 |1 2 3 5 6
39 5911 | 5922 | 5933 | 5944 | 5955 | 5966 | 5977 | 5988 | 5999 | 6010/1 2 3 4 6
40 6021 | 6031 | 6042 | 6053 | 6064 | 6075 | 6085 | 6096 | 6107 | 6117/1 2 3 4 5
41 6128 | 6138 | 6149 | 6160 | 6170 | 6180 | 6191 | 6201 | 6212 | 6222}1 2 3 4 5
42 6232 | 6243 | 6253 | 6263 | 6274 | 6284 | 6294 | 6304 | 6314 | 6325/1 2 3 4 5
43 6335 | 6345 | 6355 | 6365 | 6375 | 6385_|_6395 | 6405 | 6415 | 6425/1 2 3 4 5
44 6435 | 6444 | 6454 | 6464 | 6474 | 6484 | 6493 | 6503 | 6513 | 6522/1 2 3 4 5
45 6532 | 6542 | 6551 | 6561 | 6571 | 6580 | 6590 | 6599 | 6609 | 6618 | 1 2 3 4 5
46 6628 | 6637 | 6646 | 6656 | 6665 | 6675 | 6684 | 6693 | 6702 | 6712/1 2 3 4 5
47 6721 | 6730 | 6739 | 6749 | 6758 | 6767 | 6776 | 6785 | 6794 | 6803 }1 2 3 4 5
48 6812 | 6821 | 6830 | 6839 | 6848 | 6857 | 6866 | 6875 | 6884 | 6893/1 2 3 4 4
49 6902 | 6911 | 6920 | 6928 | 6937 | 6946 | 6955 | 6964 | 6972 | 6981|1 2 3 4 4
50 6990 | 6998 | 7007 | 7016 | 7024 | 7033 | 7042 | 7050 | 7059 | 7067/1 2 3 3 4
51 TOTO) O84) 7093) |TAOL | ALON) TAS STLZON 7 1SSe LAS al 525 eee sue
52 TAGON| TLOSMe TT WT L85 7939 720200 72108 |72185572269) 7235) lee ee:
53 T2439 72510 J2599 1267) P27 \ 2845 7292) (07300) 73087316 lesa:
54 7324 | 7332 | 7340 | 7348 | 7356 | 7364 | 7372 | 7380 | 7388 | 7396\|1 2 2 3 4
ee © |e |e ee] alse
Appendix 4 561
N 1 i.2 |3 4 5 6 3 ll 8 9 LP Le
ae |
=
55 7419 | 7427 | 7435 | 7443 | 7451 | 7459 |7466 | 7474] 1 2 2 3
56 TATE ISOS I51391 7/520) 6752807536) |99543)|| 7551 el 2-288
57 7574 | 7582 | 7589 | 7597 | 7604 | 7612 |7619 |7627} 1 2 2 3
58 7649 | 7657 | 7664 | 7672 | 7679 | 7686 |7694 | 7701} 1 1 2 3
59 7723) |1731 | 7738) | 7745 | 7752-\ 7760 |776717774) 1 1 2 3
60 7796 | 7803 |7810 | 7818 | 7825 | 7832 |7839 |7846} 1 1 2 3
61 7868 | 7875 | 7882 | 7889 | 7896 | 7903 | 7910| 7917} 1 1 2 3
62 7938 | 7945 | 7952 | 7959 | 7966 | 7973 | 7980 |7987} 1 1 2 3
63 8007 | 8014 | 8021 | 8028 | 8035 | 8041 | 8048 |8055] 1 1 2 3
64 8075 | 8082 | 8089 | 8096 | 8102 | 8109 | 8116] 8122} 1 1 2 3
65 8142 | 8149 | 8156 | 8162 | 8169 | 8176 | 8182 | 8189} 1 1 2 3
66 8209 | 8215 | 8222 | 8228 | 8235 | 8241 | 8248 | 8254] 1 1 2 3
67 8274 | 8280 | 8287 | 8293 | 8299 | 8306 | 8312] 8319} 1 1 2 3
68 8338 | 8344 | 8351 | 8357 | 8363 | 8370 | 8376 | 8382] 1 1 2 3
69 8401 | 8407 | 8414 | 8420 | 8426 | 8432 | 8439 | 8445] 1 1 2 3
70 8463 | 8470 | 8476 | 8482 | 8488 | 8494 | 8500] 8506} 1 1 2 2
71 8525 | 8531 | 8537 | 8543 | 8549 | 8555 |8561 |8567} 1 1 2 2
IZ 8585 | 8591 | 8597 | 8603 | 8609 | 8615 |8621 |8627} 1 1 2 2
73 8645 | 8651 | 8657 | 8663 | 8669 | 8675 | 8681 |8686} 1 1 2 2
74 8704 | 8710 | 8716 | 8722 | 8727 | 8733 | 8739 |8745} 1 1 2 2
75 8762 | 8768 | 8774 | 8779 | 8785 | 8791 | 8797 |8802} 1 1 2 2
76 8820 | 8825 | 8831 | 8837 | 8842 | 8848 | 8854 | 8859] 1 1 2 2
77 8876 | 8882 | 8887 | 8893 | 8899 | 8904 | 8910} 8915] 1 1 2 2
78 8932 | 8938 | 8943 | 8949 | 8954 | 8960 | 8965 8971] 1 1 2 2
79 8987 | 8993 | 8998 | 9004 | 9009 | 9015 | 9020 |9025} 1 1 2 2
80 9042 | 9047 | 9053 | 9058 | 9063 | 9069 |9074 |9079} 1 1 2 2
81 9096 | 9101 | 9106 | 9112 | 9117 | 9122 | 9128 |9133} 1 1 2 2
82 9149 | 9154 | 9159 | 9165 | 9170 | 9175 | 9180 |9186] 1 1-2 2
83 O20 TN OZ0GN OZ 12921 92225192275) 9232) 92380 leo
84 9253 | 9258 | 9263 | 9269 | 9274 | 9279 | 9284] 9289/1 1 2 2
85 9304 | 9309 | 9315 | 9320 | 9325 | 9330 | 9335 |9340} 1 1 2 2
86 9355 | 9360 | 9365 | 9370 | 9375 | 9380 | 9385 |9390} 1 1 2 2
87 9405 | 9410 | 9415 | 9420 | 9425 | 9430 | 9435 |9440] 0 1 1 2
88 9455 |9460 | 9465 | 9469 | 9474 | 9479 | 9484 |9489/0 1 1 2
89 9504 | 9509 | 9513 | 9518 | 9523 | 9528 | 9533 |9538} 0 1 1 2
90 9552 | 9557 | 9562 | 9566 | 9571 | 9576 | 9581 |9586} 0 1 1 2
91 9600 | 9605 | 9609 | 9614 | 9619 | 9624 | 9628 |9633} 0 1 1 2
92 9647 | 9652 | 9657 | 9661 | 9666 | 9671 | 9675 |9680] 0 1 1 2
93 9694 | 9699 | 9703 | 9708 | 9713 | 9717 | 9722 |9727] 0 1 1 2
94 9741 | 9745 | 9750 | 9754 | 9759 |9763 | 9768 |9773; 0 1 1 2
95 9786 | 9791 | 9795 | 9800 | 9805 | 9809 | 9814} 9818; 0 1 1 2
96 9832 | 9836 | 9841 | 9845 | 9850 | 9854 | 9859 |9863 | 0 1 1 2
97 9877 | 9881 | 9886 | 9890 | 9894 | 9899 | 9903 |9908 | 0 1 1 2
98 9921 | 9926 | 9930 | 9934 | 9939 | 9943 | 9948 |9952] 0 1 1 2
99 9965 | 9969 | 9974 | 9978 | 9983 | 9987 |9991 |9996; 0 1 1 2 NNNNNM
NYNNNWW
NNNNYN
LH
bhRA
WWWWW
WWWWWHW
WHWWWwW
WHWWALA
APPENDIX 5
3. Thermodynamics
4. Phase Equilibria
Adam, N. K., The Physics and Chemistry of Surfaces, Oxford Press, London,
1941.
Adamson, A. W., The Physical Chemistry of Surfaces, Wiley, New York, 1960.
Alexander, A. E. and P. Johnson, Colloid Science, Oxford Press, London, 1949.
7. Chemical Kinetics
8. Atomic Structure
Herzberg, G., Atomic Spectra and Atomic Structure, Dover, New York, 1941.
Richtmeyer, F. K. and E. A. Kennard, Jntroduction to Modern Physics, McGraw-
Hill, New York, 1947.
Semat, H., Introduction to Atomic and Nuclear Physics, Holt, Rhinehart, &
Winston, New York, 1962.
Sherwin, C. W., Introduction to Quantum Mechanics, Holt, Rhinehart, &
Winston, New York, 1959.
Stranathan, J. D., The Particles of Modern Physics, Blakiston, Philadelphia,
1954.
9. Molecular Structure
566
Index 567
a7?