You are on page 1of 352

Chemistry 351 and 352

Physical Chemistry I and II


Darin J. Ulness
Fall 2006 2007

Contents
I

Basic Quantum Mechanics

15

1 Quantum Theory
1.1 The Fall of Classical Physics . . . . . . . . . . . . . . . . . . . .
1.2 Bohrs Atomic Theory . . . . . . . . . . . . . . . . . . . . . . . .
1.2.1 First Attempts at the Structure of the Atom . . . . . . . .

16
16
17
17

2 The Postulates of Quantum Mechanics


2.1 Postulate I . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

22
22

2.2 How to normalize a wavefunction . . . . . . . . . . . . . . . . . .


2.3 Postulates II and II . . . . . . . . . . . . . . . . . . . . . . . . . .

23
24

3 The
3.1
3.2
3.3
3.4

Setup of a Quantum Mechanical Problem


The Hamiltonian . . . . . . . . . . . . . . . . .
The Quantum Mechanical Problem . . . . . . .
The Average Value Theorem . . . . . . . . . . .
The Heisenberg Uncertainty Principle . . . . . .

.
.
.
.

27
27
27
29
30

4 Particle in a Box
4.1 The 1D Particle in a Box Problem . . . . . . . . . . . . . . . . . .
4.2 Implications of the Particle in a Box problem . . . . . . . . . . .

31
31
34

5 The Harmonic Oscillator


5.1 Interesting Aspects of the Quantum Harmonic Oscillator . . . . .

38
40

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

5.2 Spectroscopy (An Introduction) . . . . . . . . . . . . . . . . . . .

II

Quantum Mechanics of Atoms and Molecules

6 Hydrogenic Systems
6.1 Hydrogenic systems . . . . . . . . . . . . . . . . . .
6.2 Discussion of the Wavefunctions . . . . . . . . . . .
6.3 Spin of the electron . . . . . . . . . . . . . . . . . .
6.4 Summary: the Complete Hydrogenic Wavefunction
7 Multi-electron atoms
7.1 Two Electron Atoms: Helium
7.2 The Pauli Exclusion Principle
7.3 Many Electron Atoms . . . .
7.3.1 The Total Hamiltonian

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

42

45

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

46
46
49
51
52

.
.
.
.

55
55
56
58
59

8 Diatomic Molecules and the Born Oppenheimer Approximation


8.1 Molecular Energy . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.1.1 The Hamiltonian . . . . . . . . . . . . . . . . . . . . . . .
8.1.2 The BornOppenheimer Approximation . . . . . . . . . .
8.2 Molecular Vibrations . . . . . . . . . . . . . . . . . . . . . . . . .
8.2.1 The Morse Oscillator . . . . . . . . . . . . . . . . . . . . .
8.2.2 Vibrational Spectroscopy . . . . . . . . . . . . . . . . . . .

60
60
61
62
63
64
66

9 Molecular Orbital Theory and Symmetry


9.1 Molecular Orbital Theory . . . . . . . . . . . . . . . . . . . . . .
9.2 Symmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

67
67
68

10 Molecular Orbital Diagrams


10.1 LCAOLinear Combinations of Atomic Orbitals . . . . . . . . .
10.1.1 Classification of Molecular Orbitals . . . . . . . . . . . . .

72
72
73

10.2 The Hydrogen Molecule . . . . . . . . . . . . . . . . . . . . . . .


10.3 Molecular Orbital Diagrams . . . . . . . . . . . . . . . . . . . . .
10.4 The Complete Molecular Hamiltonian and Wavefunction . . . . .
11 An Aside: Light ScatteringWhy the Sky is Blue
11.1 The Classical Electrodynamics Treatment of Light Scattering
11.2 The Blue Sky . . . . . . . . . . . . . . . . . . . . . . . . . .
11.2.1 Sunsets . . . . . . . . . . . . . . . . . . . . . . . . .
11.2.2 White Clouds . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

74
76
78
79
79
81
82
83

III Statistical Mechanics and The Laws of Thermodynamics


88
12 Rudiments of Statistical Mechanics
12.1 Statistics and Entropy . . . . . . . . . . . . . . . . . . . . . . . .
12.1.1 Combinations and Permutations . . . . . . . . . . . . . . .
12.2 Fluctuations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

89
89
90
92

13 The Boltzmann Distribution


13.1 Partition Functions . . . . . . . . . . . . . . . . . . . . . . . . . .
13.1.1 Relation between the Q and W . . . . . . . . . . . . . . .
13.2 The Molecular Partition Function . . . . . . . . . . . . . . . . . .

94
96
97
99

14 Statistical Thermodynamics

103

15 Work
15.1 Properties of Partial Derivatives
15.1.1 Summary of Relations .
15.2 Definitions . . . . . . . . . . . .
15.2.1 Types of Systems . . . .
15.2.2 System Parameters . . .

107
107
107
108
108
109

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

15.3 Work and Heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109


15.3.1 Generalized Forces and Displacements . . . . . . . . . . . 110
15.3.2 P V work . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
16 Maximum Work and Reversible changes
16.1 Maximal Work: Reversible versus Irreversible changes . .
16.2 Heat Capacity . . . . . . . . . . . . . . . . . . . . . . . .
16.3 Equations of State . . . . . . . . . . . . . . . . . . . . .
16.3.1 Example 1: The Ideal Gas Law . . . . . . . . . .
16.3.2 Example 2: The van der Waals Equation of State
16.3.3 Other Equations of State . . . . . . . . . . . . . .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

113
113
115
116
116
117
118

17 The Zeroth and First Laws of Thermodynamics


119
17.1 Temperature and the Zeroth Law of Thermodynamics . . . . . . . 119
17.2 The First Law of Thermodynamics . . . . . . . . . . . . . . . . . 121
17.2.1 The internal energy state function . . . . . . . . . . . . . . 121
18 The Second and Third Laws of Thermodynamics
18.1 Entropy and the Second Law of Thermodynamics .
18.1.1 Statements of the Second Law . . . . . . . .
18.2 The Third Law of Thermodynamics . . . . . . . . .
18.2.1 The Third Law . . . . . . . . . . . . . . . .
18.2.2 Debyes Law . . . . . . . . . . . . . . . . . .
18.3 Times Arrow . . . . . . . . . . . . . . . . . . . . .

IV

Basics of Thermodynamics

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

124
124
127
127
128
129
130

134

19 Auxillary Functions and Maxwell Relations


135
19.1 The Other Important State Functions of Thermodynamics . . . . 135
19.2 Enthalpy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
19.2.1 Heuristic definition: . . . . . . . . . . . . . . . . . . . . . . 137

19.3 Helmholtz Free Energy . . . . . . . . . . . .


19.3.1 Heuristic definition: . . . . . . . . . .
19.4 Gibbs Free Energy . . . . . . . . . . . . . .
19.4.1 Heuristic definition: . . . . . . . . . .
19.5 Heat Capacity of Gases . . . . . . . . . . . .
19.5.1 The Relationship Between CP and CV
19.6 The Maxwell Relations . . . . . . . . . . . .
20 Chemical Potential
20.1 Spontaneity of processes . . . . . . . . . . .
20.2 Chemical potential . . . . . . . . . . . . . .
20.3 Activity and the Activity coecient . . . . .
20.3.1 Reference States . . . . . . . . . . .
20.3.2 Activity and the Chemical Potential

.
.
.
.
.

.
.
.
.
.
.
. .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

137
138
138
139
139
139
140

.
.
.
.
.

142
142
144
146
147
148

21 Equilibrium
151
21.0.3 Equilibrium constants in terms of KC . . . . . . . . . . . . 153
21.0.4 The Partition Coecient . . . . . . . . . . . . . . . . . . . 153
22 Chemical Reactions
22.1 Heats of Reactions . . . . . . .
22.1.1 Heats of Formation . . .
22.1.2 Temperature dependence
22.2 Reversible reactions . . . . . . .
22.3 Temperature Dependence of Ka
22.4 Extent of Reaction . . . . . . .

. . . . . . . .
. . . . . . . .
of the heat of
. . . . . . . .
. . . . . . . .
. . . . . . . .

. . . . .
. . . . .
reaction
. . . . .
. . . . .
. . . . .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

156
156
157
157
158
159
160

23 Ionics
161
23.1 Ionic Activities . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
23.1.1 Ionic activity coecients . . . . . . . . . . . . . . . . . . . 162
23.2 Theory of Electrolytic Solutions . . . . . . . . . . . . . . . . . . . 163

23.3 Ion Mobility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164


23.3.1 Ion mobility . . . . . . . . . . . . . . . . . . . . . . . . . . 165
24 Thermodynamics of Solvation
24.1 The Born Model . . . . . . . . . . . . . . .
24.1.1 Free Energy of Solvation for the Born
24.1.2 Ion Transfer Between Phases . . . . .
24.1.3 Enthalpy and Entropy of Solvation .
24.2 Corrections to the Born Model . . . . . . . .

. . . .
Model
. . . .
. . . .
. . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

25 Key Equations for Exam 4

169
170
173
174
174
175
177

Quantum Mechanics and Dynamics

180

26 Particle in a 3D Box
181
26.1 Particle in a Box . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
26.2 The 3D Particle in a Box Problem . . . . . . . . . . . . . . . . . . 183
27 Operators
187
27.1 Operator Algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
27.2 Orthogonality, Completeness, and the Superposition Principle . . 191
28 Angular Momentum
28.1 Classical Theory of Angular Momentum . .
28.2 Quantum theory of Angular Momentum . .
28.3 Particle on a Ring . . . . . . . . . . . . . . .
28.4 General Theory of Angular Momentum . . .
28.5 Quantum Properties of Angular Momentum
28.5.1 The rigid rotor . . . . . . . . . . . .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

192
192
193
194
195
199
200

29 Addition of Angular Momentum


29.1 Spin Angular Momentum . . . . . . . . . . .
29.2 Addition of Angular Momentum . . . . . . .
29.2.1 The Addition of Angular Momentum:
29.2.2 An Example: Two Electrons . . . . .
29.2.3 Term Symbols . . . . . . . . . . . . .
29.2.4 Spin Orbit Coupling . . . . . . . . .

. . . . .
. . . . .
General
. . . . .
. . . . .
. . . . .

. . . . .
. . . . .
Theory
. . . . .
. . . . .
. . . . .

.
.
.
.
.
.

.
.
.
.
.
.

201
201
202
202
203
204
205

30 Approximation Techniques
207
30.1 Perturbation Theory . . . . . . . . . . . . . . . . . . . . . . . . . 207
30.2 Variational method . . . . . . . . . . . . . . . . . . . . . . . . . . 209
31 The Two Level System and Quantum Dynamics
211
31.1 The Two Level System . . . . . . . . . . . . . . . . . . . . . . . . 211
31.2 Quantum Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . 214

VI

Symmetry and Spectroscopy

32 Symmetry and Group Theory


32.1 Symmetry Operators . . . . . . . . . . . . . . .
32.2 Mathematical Groups . . . . . . . . . . . . . . .
32.2.1 Example: The C2v Group . . . . . . . .
32.3 Symmetry of Functions . . . . . . . . . . . . . .
32.3.1 Direct Products . . . . . . . . . . . . . .
32.4 Symmetry Breaking and Crystal Field Splitting

220

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

221
222
222
223
223
225
225

33 Molecules and Symmetry


228
33.1 Molecular Vibrations . . . . . . . . . . . . . . . . . . . . . . . . . 228
33.1.1 Normal Modes . . . . . . . . . . . . . . . . . . . . . . . . 229
33.1.2 Normal Modes and Group Theory . . . . . . . . . . . . . . 229

34 Vibrational Spectroscopy and Group Theory


231
34.1 IR Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
34.2 Raman Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . 233
35 Molecular Rotations
35.1 Relaxing the rigid rotor . . . . . . . . . . . . . . . . . . . . . . . .
35.2 Rotational Spectroscopy . . . . . . . . . . . . . . . . . . . . . . .
35.3 Rotation of Polyatomic Molecules . . . . . . . . . . . . . . . . . .

235
236
236
237

36 Electronic Spectroscopy of Molecules


36.1 The Structure of the Electronic State
36.1.1 Absorption Spectra . . . . . .
36.1.2 Emission Spectra . . . . . . .
36.1.3 Fluorescence Spectra . . . . .
36.2 FranckCondon activity . . . . . . .
36.2.1 The FranckCondon principle

240
240
241
241
242
243
243

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

37 Fourier Transforms
245
37.1 The Fourier transformation . . . . . . . . . . . . . . . . . . . . . 245

VII

Kinetics and Gases

249

38 Physical Kinetics
250
38.1 kinetic theory of gases . . . . . . . . . . . . . . . . . . . . . . . . 250
38.2 Molecular Collisions . . . . . . . . . . . . . . . . . . . . . . . . . . 252
39 The Rate Laws of Chemical Kinetics
39.1 Rate Laws . . . . . . . . . . . . . . . . . . . . . .
39.2 Determination of Rate Laws . . . . . . . . . . . .
39.2.1 Dierential methods based on the rate law
39.2.2 Integrated rate laws . . . . . . . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

254
254
258
259
259

40 Temperature and Chemical Kinetics


40.1 Temperature Eects on Rate Constants . . . . . .
40.1.1 Temperature corrections to the Arrhenious
40.2 Theory of Reaction Rates . . . . . . . . . . . . .
40.3 Multistep Reactions . . . . . . . . . . . . . . . . .
40.4 Chain Reactions . . . . . . . . . . . . . . . . . . .

. . . . . . .
parameters
. . . . . . .
. . . . . . .
. . . . . . .

41 Gases and the Virial Series


41.1 Equations of State . . . . . . . . . . . . . . . . . . . .
41.2 The Virial Series . . . . . . . . . . . . . . . . . . . . .
41.2.1 Relation to the van der Waals Equation of State
41.2.2 The Boyle Temperature . . . . . . . . . . . . .
41.2.3 The Virial Series in Pressure . . . . . . . . . . .
41.2.4 Estimation of Virial Coecients . . . . . . . . .
42 Behavior of Gases
42.1 P, V and T behavior . . . . . . . . . . . . . . . .
42.1.1 and T for an ideal gas . . . . . . . . . .
42.1.2 and T for liquids and solids . . . . . . .
42.2 Heat Capacity of Gases Revisited . . . . . . . . .
42.2.1 The Relationship Between CP and CV . .
42.3 Expansion of Gases . . . . . . . . . . . . . . . . .
42.3.1 Isothermal and Adiabatic expansions . . .
42.3.2 Heat capacity CV for adiabatic expansions
42.3.3 When P is the more convenient variable .
42.3.4 Joule expansion . . . . . . . . . . . . . . .
42.3.5 Joule-Thomson expansion . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

261
261
262
262
265
267

.
.
.
.
.
.

269
269
270
271
272
272
273

.
.
.
.
.
.
.
.
.
.
.

274
274
275
275
276
276
279
279
280
281
282
283

43 Entropy of Gases
286
43.1 Calculation of Entropy . . . . . . . . . . . . . . . . . . . . . . . . 286
43.1.1 Entropy of Real Gases . . . . . . . . . . . . . . . . . . . . 288

VIII

More Thermodyanmics

292

44 Critical Phenomena
44.1 Critical Behavior of fluids . . . . . . . . . . . . . . . . . . . . .
44.1.1 Gas Laws in the Critical Region . . . . . . . . . . . . . .
44.1.2 Gas Constants from Critical Data . . . . . . . . . . . . .
44.2 The Law of Corresponding States . . . . . . . . . . . . . . . . .
44.3 Phase Equilibrium . . . . . . . . . . . . . . . . . . . . . . . . .
44.3.1 The chemical potential and T and P . . . . . . . . . . .
44.3.2 The Clapeyron Equation . . . . . . . . . . . . . . . . . .
44.3.3 Vapor Equilibrium and the Clausius-Clapeyron Equation
44.4 Equilibria of condensed phases . . . . . . . . . . . . . . . . . . .
44.5 Triple Point and Phase Diagrams . . . . . . . . . . . . . . . . .
45 Transport Properties of Fluids
45.1 Diusion . . . . . . . . . . . . .
45.2 Viscosity . . . . . . . . . . . . .
45.3 Thermal conductivity . . . . . .
45.3.1 Thermal Conductivity of
45.3.2 Thermal Conductivity of
46 Solutions
46.1 Measures of Composition . . . .
46.2 Partial Molar Quantities . . . .
46.2.1 Notation . . . . . . . . .
46.2.2 Partial Molar Volumes .
46.3 Reference states for liquids . . .
46.3.1 Activity (a brief review)
46.3.2 Raoults Law . . . . . .
46.3.3 Ideal Solutions (RL) . .
46.3.4 Henrys Law . . . . . . .

. . . . . .
. . . . . .
. . . . . .
Gases and
Solids . .

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

. . . . .
. . . . .
. . . . .
Liquids
. . . . .

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

293
293
294
295
296
296
297
298
298
299
300

.
.
.
.
.

301
301
303
305
306
307

.
.
.
.
.
.
.
.
.

308
308
308
309
310
311
311
312
314
316

46.4 Colligative Properties . . . . . . . . . . . . . . . . . . . . . . . . . 318


46.4.1 Freezing Point Depression . . . . . . . . . . . . . . . . . . 318
46.4.2 Osmotic Pressure . . . . . . . . . . . . . . . . . . . . . . . 319
47 Entropy Production and Irreverisble Thermodynamics
47.1 Fundamentals . . . . . . . . . . . . . . . . . . . . . . . .
47.2 The Second Law . . . . . . . . . . . . . . . . . . . . . .
47.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . .
47.3.1 Entropy Production due to Heat Flow . . . . . .
47.3.2 Entropy Production due to Chemical Reactions .
47.4 Thermodynamic Coupling . . . . . . . . . . . . . . . . .
47.5 Echo Phenonmena . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

322
322
324
325
326
328
330
331

Chemistry 351: Physical


Chemistry I

1
1

Solved Problems
I make-up most of the problems on the problems sets, so it might be helpful to
you to see some of these problems worked out.
Even though there arent many book problems assigned during the year, you can
still learn a lot be working these and looking that their solutions in the solution
manual.
Keep in mind this chapter provides some examples of how to solve problems for
both physical chemistry I and physical chemistry II. Consequently early in the
course some of the examples might seem very itimidating. Simply skip those
examples as you scan through this chapter.

Tips for solving problems


Working problem sets is the heart and sole of learning physical chemistry. The
only way that you can be sure that you understand a concept at to be able to
solve the problems associated with it.
This takes time and hard work.
But there are some things that you can do to help yourself with these problems.
Tips

2
2

1. Remember nobody cares if you solve any particular problem on the problem
set. They have all been solved before, so if you solve them you will not
become famous nor will you save the world. The only reason you work them
is to learn.
2. Budget your time so that you dont have to work on an overwhelming number
of problems at a time. Try to whip-o a few on the same day that you get
the problem set. Then work on them consistently during the week. This
will make the problem sets much more ecient at helping you learn.
3. You can do the problem. I dont assign problems that you cannot do. If you
think you cant do the problem then maybe you need try a dierent way of
thinking about it.
4. Part of the trouble is simply understanding what the problem is asking you
to do. There is a tendency to try to start solving the problem before fully
understanding the question.
Read the question carefully
Try to think about what topic(s) in lecture and in the notes the problem
is dealing with.
Do not worry about not knowing how to solve it yet.
Just identify the general ideas that you think you might need.
Determine wether you need to approach the problem mathematically
or conceptually or both.
If the question is long, try to identify subsections of it.
5. For problems that require a mathematical approach...
Do not be afraid. Try to figure out what mathematical techniques you
need to express the solution to the problem.

Do the math; either you will be able to do this or you wont. It might
take some review on your part.
Always check to see if the math makes sense when you are done.
6. For problems that require a conceptual approach...
Make sure that the physical idea that you are using in your argument is
correct. If you are not sure, start with a related concept that is better
known by you.
Look for self-consistency. Does you final answer jive with what you
know.

Problems Dealing With Quantum Mechanics


Problem: What is the periodicity of the following functions
f (x) = sin2 x
f (x) = cos x
f(x) = e2ix
Solution: For the first function it is easiest to see the periodicity by writing the
function as f (x) = (sin x)(sin x). We know that this function will repeat zeros
when ever sin x = 0. This occurs at x = n, n = 0, 1, 2 . . ., so the periodicity
is . The second function we should remember from trig as having a period of 2.
Finally for the last function it is best to used Eulers identity and write
e2ix = cos 2x + i sin 2x

(1)

The real part of this function, cos 2x, has a period of as does the imaginary
part, sin 2x. Therefore the entire function has a period of .

Problem: Which of the following functions are eigenfunction of the momentum


d
operator, px = i~ dx
.
(x) = eikx
2

(x) = ex

(x) = cos kx
Solution: We need to determine if px (x) = (x) where is a constant. If
this equation is true then the function is an eigenfunction with eigenvalue . For
the case of momentum all we need to do is take the derivative of each function,
multiply by i~ and check to see if the eigenvalue equation holds.
For the first function
px (x) = i~

deikx
d(x)
= i~
= ~keikx = ~k(x),
dx
dx

(2)

so, yes, this function is an eigenfunction of the momentum operator.


For the second function
2

px (x) = i~

dex
d(x)

2
= i~
= 2i~xex = 2i~x(x),
dx
dx

(3)

so, no, this function is not and eigenfunction of the momentum operator.
For the last function
6=cos kx

z }| {
d cos kx
d(x)
= i~
= i~k sin kx,
px (x) = i~
dx
dx

(4)

so, no, this function is not an eigenfunction of the momentum operator.


2

Problem: A quantum object is described by the wavefunction (x) = ex .


What is the probability of finding the object further than away from the origin
( x = 0)?

Solution: First of all we do not know if this wavefunction is normalized, so we


should assume that it isnt. We could normalize this wavefunction, but we wont.
We are interested in finding the probability that the object is outside of the region
< x < . To do this using an unnormalized wavefunction we must evaluate
R
R
2
|(x)|
dx
+
|(x)|2 dx

P (|x| > ) =
R
.
(5)
|(x)|2 dx

The first integral in the numerator gives the probability that the object is at a
position x < and the second integral in the numerator gives the probability
for x > . The denominator accounts for the fact that the wavefunction is unnormalized. The limits of the integral in the denominator represent all space for
the object. If you were working with a normalized wavefunction the denominator
would be equal to 1 and hence not needed. Plugging in the wavefunctions we have
R 2x2
R
2
e
dx + e2x dx

R
P (|x| > ) =
.
(6)

e2x2 dx

Mathematica can assist with these integrals to give the final answer of
3
P (|x| > ) = erfc[ 2 2 ].

(7)

Problem: A quantum object is described by the wavefunction (x) = ex over


the range 0 x < . Normalize this wavefunction.
Solution: Following our general procedure from the notes if we have some unnormalized wavefunction, unnorm we know that this function must simply be some
constant N multiplied by the normalized version of this function:
(8)

unnorm = Nnorm
We have shown generally that N is given by
sZ
N=
|unnorm (x)|2 dx.

(9)

space

Which for this case is


N=

sZ

|ex |2

dx =

sZ

e2x dx

1
2

(10)

So finally we get the normalized wavefunction by rearanging unnorm = Nnorm :


p
(11)
norm (x) = 2ex .
Problem: A quantum object is described by the wavefunction (x) = ex over
the range 0 x < . What is the average position of the object?
Solution: We need to work with the normalized wavefunction that we found in

the previous problem, (x) = 2ex . Generally and average is calculated as


Z
(x)
o(x),
(12)
h
oi =
space

which in this case is


Z p
Z
p x
x
h
xi =
2e x 2e dx = 2
0

xe2x dx =

So on average you will find the object at x =

1
.
2

(13)

1
.
2

Problem: What is the probability of finding an electron in the 1s state of hydrogen


further than one Bohr radius away from the nucleus?
Solution: We need to evaluate
Z
P (r > a0 ) =

a0

|1s |2 r2 sin drdd.

(14)

Remember the extra r2 sin is needed when integrating in spherical polar coordinates. The normalized 1s wavefunction is
1
1s = p 3 er/a0 .
a0

(15)

We can do this integral by hand or have Mathematica help us to give


P (r > a0 ) =

5
= 0.677.
e2

(16)

So, about 68% of the time the electron would be found at some distance greater
then one Bohr radius from the proton.
Problem: A free particle in three dimensions is described by the Hamiltonian,
= ~2 2 . Express the wavefunction (in Cartesian coordinates) as a product
H
2m
state.
Solution: This problem appears hard at first since we are not studying three
dimensional systems, but all it is asking is to express the wavefunction, which is
a function of the three spatial dimensions, (x, y, z) as a product state. We know
that if the wavefunction is to be a product state then the Hamiltonian must be
made up of a sum of independent terms. To see this we write out the Laplacian
to get
2

2
2
2
~

=
H
.
(17)
+
+
2m x2 y 2 z 2
We see that indeed the Hamiltonian is a sum of term that depends only on x,
a term depending only on y and a term that depends only on z. Therefore the
appropriate product state is
(18)

(x, y, z) = (x)(y)(z).

Problem: Expand the Morse potential in a Taylors series about Req . Verify that
the coecient for the linear term is zero. What is the force constant associated
with the Morse potential?
Solution: The Morse potential is

V (x) = De 1 e(RReq ) .

(19)

The Taylor series about Req for this function is

dV (x)
1 d2 V (x)
(R Req ) +
(R Req )2 + . (20)
V (x) = V (x)|Req +
dx Req
2! dx2 Req
| {z }
| {z }
|
{z
}
=0
=0

= 2 De

So, yes the coecient of the linear term (the term involving (R Req ) to the
first power) is zero. This will always be true when you perform a Taylor series
expansion about a minimum (or maximum). The force constant is given by the
coecient of the quadratic term so in this case k = 2 De .
Problem: Without performing any calculations, compare hRi as a function of
the vibrational quantum number for a diatomic modelled as a harmonic oscillator
versus a Morse oscillator.
Solution: This problem requires the we think qualitatively about the wavefunctions and the potentials for the harmonic oscillator and the Morse oscillator. The
potential for the harmonic oscillator is described by a parabola centered about the
equilibrium bond length. Hence no mater what the vibrational quantum number is
there is just as much of the wavefunction on either side equilibrium thus hRi = Req
for any quantum number. The Morse potential does not have this symmetry. It
is steeper on the short side of equilibrium and softer on the long side of equilibrium and this softness increases with increasing quantum number. Therefore
without performing any calculations we can at least say that hRi increases as the
quantum number increases.

Problems Dealing With Statistical Mechanics and Thermodynamics


Problem: A vial containing 10 20 benzene molecules is at 300K. How many molecules are in the first excited state of the ring breathing mode (992 cm 1 )? How

many are in the first excited state of the symmetric CH vibrational mode (3063
cm 1 )?
Solution: This is a problem that deals with the Boltzmann distribution. So,

3992
992
rb
e 2208 1020 = 8.41 1017
Nv=1 = 2 sinh
(21)
2 208
and

CH
Nv=1

3063
33063
2208
1020 = 4.02 1013 .
e
= 2 sinh
2 208

(22)

17

We see that about 8.4110


100% = 0.841% of the benzene molecules are in the
1020
13
first vibrational excited state for the ring breathing mode and 4.0210
100% =
1020
0.0000402% of the benzene molecules are in the first excited state for the CH
stretching mode.
Problem: Consider a linear chain of N atoms. Each of the atoms can be in one
of three states A, B or C, except that an atom in state A can not be adjacent to
an atom in state C. Find the entropy per atom for this system as N . To
solve this problem it is useful to define the set of three dimensional column vectors
V (j) such that the three elements are the total number of allowed configurations of
a j-atom chain having the j th atom in state A, B or C. For example,



1
2
5



(1)
(2)
(3)
(23)
V = 1 , V = 3 , V = 7 , .
1

The V (j+1) can be found from the V (j) vector using the matrix equation,

where for this example

V (j+1) = MV (j) ,

(24)

1 1 0

M = 1 1 1 .
0 1 1

(25)

10

The matrix M is the so-called transfer matrix for this system. It can be shown
that the number of configurations W = Tr[M N ]. Now for large N, Tr[M N ] N
max ,
where max is the largest eigenvalue of M. So
W = lim N
max .

(26)

1.

1. Use M to find V (4)


2. Verify V (3) explicitly by drawing all the allowed 3-atom configurations.
3. Verify W = Tr[M N ] for N = 1 and N = 2.
4. Use Boltzmanns equation to find the entropy per atom for this chain
as N goes to infinity.

Solution: For part (a) we simply


problem (we are given V (3) ):

1 1

(4)
V = 1 1
0 1

use the transfer matrix as directed in the


12
5
0


1 7 = 17 .
12
5
1

For part (b) we need to list all states for the case of N = 3 and verify the we get
the same result as calculated using the transfer matrix. Remembering that V (3)
gives us the number of sequences that end in a given state we should organize our
list in the same manner
States ending in A States ending in B
AAA
ABA
BAA
BBA
CBA

5 states

AAB
ABB
BAB
BBB
BCB
CBB
CCB

7 states

States ending in C
ABC
BBC
BCC
CBC

5 states

11

States like AAC are not allowed because A and C are neighbors.
For part (c) we evaluate W = Tr[M N ] for N = 1 and 2. For N = 1, W =
Tr[M] = 3 This corresponds to the three distinguishable microstates A, B, and
C. For N = 2,

2 2 1
1 1 0
1 1 0

W = Tr[M 2 ] = Tr 1 1 1 1 1 1 = Tr 2 3 2 = 7 (27)
0 1 1

0 1 1

1 2 2

This corresponds to the seven distinguishable microstates AA, AB, BA, BB, BC,
CB and CC (Remember C and A cannot be neighbors).
For part (d) we use
S
k
k
k
=
ln W = lim
ln N
N ln max = k ln max .
max = lim
N N
N N
N
N

(28)

So, we simply need to find the maximum eigenvalue of the Transfer matrix. Using

Mathematica we find max = 1 + 2. Therefore the limiting entropy per atom


is


S
= k ln 1 + 2 .
(29)
N
Problem: Using the classical theory of light scattering, calculate the positions of
the Rayleigh, Stokes and anti-Stokes spectral lines for benzene. Assume benzene
has only two active modes (992cm 1 and 3063cm 1 ) and assume the Laser light
used to do the scattering is at 20000cm 1 (this is 500nmgreen light).
Solution: Since there are two vibrational modes we expect two Stokes lines to
the red of 20000cm1 , one at 20000cm1 992cm1 = 19008cm1 and one at
20000cm1 3063cm1 = 16937cm1 . Likewise we expect two anti-Stokes lines,
one at 20000cm1 + 992cm1 = 20992cm1 and one at 20000cm1 + 3063cm1 =
23063cm1 . There is only one Rayleigh line and it is at the same frequency at the
input laser beam which, in this case, is 20000cm1 .

12

Problem: A simple model for a crystal is a gas of harmonic oscillators. Determine A, S, and U from the partition function for this model.
Solution: For this model the crystal is modelled as a collection of harmonic
oscillators so we need the partition function for the harmonic oscillator.

!
1
N
Qcrystal = qHO
=
(30)
2 sinh ~
2
From our formulas for statistical thermodynamics

~
,
A = kT ln Qcrystal = +NKT ln 2 sinh
2

(31)

where we used properties of logs to pull the N out front and move the sinh term
from to the numerator,
Qcrystal
+ k ln Qcrystal

~
~
Nk~
coth
k ln 2 sinh
=
2
2
2

S = k

(32)

and

N~
~
Qcrystal
=
coth
.
(33)

2
2
Problem: Express the equation of state for internal energy for a Berthelot gas.
U =

Solution: The equation representing a Berthelot gas is


P =

nRT
n2 a

.
V nb T V 2

(34)

We are interesting in an equation of state for U(T, V ). Writing out the total
derivative of U(T, V ) we get

U
U
dT +
dV.
(35)
dU =
T V
V T

13


U
Now U
is just heat capacity, CV , but V
is nothing convenient so we must
T V
T
proceed. We employ the useful relation

U
P
=T
P
(36)
V T
T V
to eliminate U in favor of P so that we can use the equation of state for a Berthelot
gas. One obtains

n2 a
n2 a
P
nRT
nR
2n2 a
+ 2 2
+
T
P =T
=
.
(37)
T V
V nb T V
V nb T V 2
TV 2
Hence the equation of state for internal energy of a Berthelot gas is
dU = CV dT +

2n2 a
dV
TV 2

(38)

Problem: Use the identities for partial derivatives to eliminate the


in

P
V
Cp = Cv + T
T P T V

T V

factor
(39)

so that all derivatives are at constant pressure or temperature.


Solution: Here we either remember an identity or turn to our handout of partial

derivative identities to employ the cyclic rule to P
:
T V

P
T

P
=
V

V
T

(40)

This eliminates the constant V term and so,


2

P
V
.
Cp = Cv T
T P V T

(41)

14

Part I
Basic Quantum Mechanics

15
15

1. Quantum Theory
The goal of science is unification.
Many phenomena described by minimal and general concepts.

1.1. The Fall of Classical Physics


A good theory:
explain known experimental results
self consistent
predictive
minimal number of postulates
Around the turn of the century, experiments were being performed in which the results defied explanation by means of the current understanding of physics. Among
these experiments were
1. The photoelectric eect
2. Low temperature heat capacity
3. Atomic spectral lines
4. Black body radiation and the ultraviolet catastrophe
16
16

5. The two slit experiment


6. The Stern-Gerlach experiment

See Handouts

1.2. Bohrs Atomic Theory


1.2.1. First Attempts at the Structure of the Atom
The solar system model.
The electron orbits the nucleus with the attractive coulomb force balanced
by the repulsive centrifugal force.

Flaws of the solar system model


Newton: OK

Maxwell: problem

17

As the electron orbits the nucleus, the atom acts as an oscillating dipole

The classical theory of electromagnetism states that oscillating dipoles


emit radiation and thereby lose energy.
The system is not stable and the electron spirals into the nucleus. The
atom collapses!

Bohrs model: Niels Bohr (18851962)

18

Atoms dont collapse = what are the consequences


Experimental clues
Atomic gases have discrete spectral lines.
If the orbital radius was continuous the gas would have a continuous spectrum.
Therefore atomic orbitals must be quantized.

4 0 N 2 ~2
(1.1)
Zme e2
where Z is the atomic number, me and e are the mass and charge of the
electron respectively and 0 is the permittivity of free space. N is a positive
real integer called the quantum number. ~ = h/2 is Plancks constant
divided by 2.
r=

0~
The constant quantity 4
appears often and is given the special symbol a0
me e2
2
4 0 ~
= 0.52918 and is called the Bohr radius.
me e2

The total energy of the Bohr atom is related to its quantum number
2
e
1
2
.
EN = Z
2a0 N 2

(1.2)

Tests of the Bohr atom

Ionization energy of Hydrogen atoms


The Ionization energy for Hydrogen atoms (Z = 1) is the minium
energy required to completely remove an electron form it ground state,
i.e., N = 1 N =

1
e2
Z 2 e2
1
=

(1.3)
Eionize = E E1 =
2a0
2 12
2a0

19

e
Eionize = 2a
= 13.606 eV = 109,667 cm1 = R. R is called the Rydberg
0
constant.

Eionize experimentally observed from spectroscopy is 13.605 eV (very


good agreement)
Spectroscopic lines from Hydrogen represent the dierence in energy between
the quantum states
Bohr theory: Dierence energies

1
1
e2
1
1
Ej Ek =
=R

2a0 Nj2 Nk2


Nj2 Nk2

Initial state Nk

Final States Nj

Series Name

1
2
3
4
5

2,3,4,
3,4,5,
4,5,6,
5,6,7,
6,7,8,

Lyman
Balmer
Pachen
Brackett
Pfund

(1.4)

Since the orbitals are quantized, the atom may only change its orbital
radius by discrete amounts.
Doing this results in the emission or absorption of a photon with energy
v =

4E
hc

(1.5)

Failure of the Bohr model


No fine structure predicted (electron-electron coupling)
No hyperfine structure predicted (electron-nucleus coupling)
No Zeeman eect predicted (response of spectrum to magnetic field)

20

Spin is not included in theory


The Bohr quantization idea points to a wavelike behavior for the electron.
The wave must satisfy periodic boundary conditions much like a vibrating ring
See Fig. 11.9 Laidler&Meiser
The must be continuous and single valued
Particles have wave-like characteristics
The Bohr atom was an important step towards the formulation of quantum theory
Erwin Schrdinger (18871961): Wave mechanics
Werner Heisenberg (19021976): Matrix mechanics
Paul Dirac (19021984): Abstract vector space approach

21

2. The Postulates of Quantum


Mechanics
2.1. Postulate I
Postulate I: The state of a system is defined by a wavefunction, , which contains all the information that can be known about the system.
We will normally take to be a complex valued function of time and coordinates: (t, x, y, z) and, in fact, we will most often deal with time independent
stationary states (x, y, z)
Note: In general the wavefunction need not be expressed as a function of coordinate. It may, for example, be a function of momentum.
The wavefunction represents a probability amplitude and is not directly observable.
However the mod-square of the wavefunction, = ||2 , represents a probability
distribution which is directly observable.
That is, the probability of finding a particle which is described by (x, y, z) at the
position between x and x+dx, y and y +dy and z and z +dz is |(x, y, z)|2 dxdydz
(or |(r, , )|2 r2 sin drdd in spherical coordinates).
22
22

Properties of the wavefunction


Single valueness
continuous and finite
continuous and finite first derivative
R
space |(x, y, z)|2 dxdydz <
Normalization of the wavefunction
In order for |(x, y, z)|2 to be exactly interpreted as a probability distribution, (x, y, z) must be normalizable.
qR
That is, unnorm = Nnorm , where N =
|unnorm (x, y, z)|2 dxdydz
space
R
This assures that space |norm |2 dxdydz = 1 as expected for a probability distribution
From now on we will always normalize our wavefunctions.

2.2. How to normalize a wavefunction


If we have some unnormalized wavefunction, unnorm we know that this function
must simply be some constant N multiplied by the normalized version of this
function:
unnorm = Nnorm .
(2.1)
Now, we take the mod-square of both sides and then integrate both sides of this
equation over all space
Z
Z
2
|unnorm | dxdydz =
|Nnorm |2 dxdydz.
(2.2)
space

space

23

The N is just a constant so it can be pulled out of both the mod-square and the
integral
Z
Z
2
2
|unnorm | dxdydz = N
|norm |2 dxdydz,
(2.3)
space

space

but

space

|norm |2 dxdydz = 1

(2.4)

because that is the very definition of a normalized wavefunction. Thus wherever


R
we see space |norm |2 dxdydz we can replace it with 1. So,
Z
|unnorm |2 dxdydz = N 2 1 = N 2 .
(2.5)
space

This gives us an expression for N. Taking the square root of both sides gives.
sZ
N=

space

|unnorm (x, y, z)|2 dxdydz.

(2.6)

So finally we get the normalized wavefunction by reagranging unnorm = Nnorm :


norm =

.
N unnorm

(2.7)

Notice that no where did we ever specify what unnorm or norm actually were,
therefore this is a general procedure that will work for any wavefunction.
To find the probability for the particle to be in a finite region of space we simple
evaluate (here a 1D case)
R x2
Z x2
|(x)|2 dx if (x)
x1
P (x1 < x < x2 ) = R
=
|(x)|2 dx
(2.8)
2
normalized
|(x)|
dx
x
1

2.3. Postulates II and II

Postulate II: Every physical observable is represented by a linear (Hermitian)


operator.

24

An operator takes a function and turns it into another function


(x) = g(x)
Of

(2.9)

This is just like how a function takes a number and turns it into another number.
So in quantum mechanics operators act on the wavefunction to produce a new
wavefunction
The two most important operators as far as we are concerned are
x = x

px = i~ x

and of course the analogous operators for the other coordinates (y, z) and coordinate systems (spherical, cylindrical, etc.).
Nearly all operators we will need are algebraic combinations of the above.
Postulate III: The measurement of a physical observable will give a result that
is one of the eigenvalues of the corresponding operator.
There is a special operator equation called the eigenvalue equation which is
(x) = f (x)
Of

(2.10)

where is just a number.


For a given operator only a special set of function satisfy this equation. These
functions are called eigenfunctions.

25

The number that goes with each function is called the eigenvalue.
So solution of the eigenvalue equation gives a set of eigenfunctions and a set of
eigenvalues.
Example
in the eignevalue equation be the operator that takes the derivative: O
=
Let O
d
.
d = dx
So we want a solution to
(x) = f (x)
df
df (x)
= f (x)
dx

(2.11)

So, we ask ourselves what function is proportional to its own derivative?


f (x) = ex .
So the eigenfunctions are the set of functions f (x) = ex and the eigenvalues are
the numbers

26

3. The Setup of a Quantum


Mechanical Problem
3.1. The Hamiltonian
The most important physical observable is that of the total energy E.
The operator associated with the total energy is called the Hamiltonian operator

(or simply the Hamiltonian) and is given the symbol H.


The eigenvalue equation for the Hamiltonian is
= E.
H

(3.1)

This equation is the (time independent) Schrdinger equation.


This equation is the most important equation of the course and we will use it many
times throughout our discussion of quantum mechanics and statistical mechanics.

3.2. The Quantum Mechanical Problem


Nearly every problem one is faced with in elementary quantum mechanics is handled by the same procedure as given in the following steps.
1. Define the classical Hamiltonian for the system.
27
27

The total energy for a classical system is


Ecl = T + V,

(3.2)

where T is the kinetic energy and V is the potential energy.


The kinetic energy is always of the form
T =

1 2
px + p2y + p2z
2m

(3.3)

The potential energy is almost always a function of coordinates only


V = V (x, y, z)

(3.4)

Note: Some quantum systems dont have classical analogs so the Hamiltonian operator must be hypothesized.
2. Use Postulate II to replace the classical variables, x, px etc., with their
appropriate operators. Thus,
~2 2 ~2 2
T =
,
=
2m
2m
where 2

2
x2

2
y2

2
,
z 2

(3.5)

and

V = V (
x, y, z) = V (x, y, z).

(3.6)

2
= T + V = ~ 2 + V (x, y, z)
H
2m

(3.7)

So,

= E, which is now a second order


3. Solve the Schrdinger equation, H
dierential equation of the form

2
~ 2
+ V (x, y, z) = E
2m
~2 2
+ (V (x, y, z) E) = 0 (3.8)

2m

28

Note: It is solely the form of V (x, y, z) which determines whether this


is easy or hard to do.
For one-dimensional problems
~2 d2
+ (V (x) E) = 0
2m dx2

(3.9)

3.3. The Average Value Theorem


Postulate III implies that if is an eigenfunction of a particular operator representing a physical observable, then all measurements of that physical property
will yield the associated eigenvalue.
However, If is not an eigenfunction of a particular operator, then all measurements of that physical property will still yield an eigenvalue, but we cannot predict
for certain which one.
We can, however, give an expectation, or average, value for the measurement.
This is given by
Z
h
i =

dxdydz
(3.10)
space

For example,
h
xi =
and
h
px i =

xdxdydz =

space

space

space

px dxdydz = i~

x ||2 dxdydz

space

dxdydz
x

(3.11)

(3.12)

29

3.4. The Heisenberg Uncertainty Principle


In quantum mechanics certain pairs of variables can not, even in principle, be
simultaneously known to arbitrary precision. Such variables are called complimentary.
This idea is the Heisenberg uncertainty principle and is of profound importance.
The general statement of the Heisenberg uncertainty principle is
1 Dh iE

, ,
(3.13)
2
h
i

The commutator is
where the notation
, means the commutator of
and .
defined as
h
i

,

.
(3.14)
The most important example of complimentary variables is position and momentum. We see
1
1
|h[
px , x]i| = |h
px x xpx i|
2 Z
2

1
~

x
dx
=

2
i x
x

~ ~
= = .
2i
2

px x

(3.15)

So, at the very best we can only hope to simultaneously know position and momentum such that the product of the uncertainty in each is ~2 . (n.b., px y = 0, we can
know, for example, the y position and the x momentum to arbitrary precision.)
Suppose we know the position of a particle perfectly, what can we say about its
momentum?

30

4. Particle in a Box
We now will apply the general program for solving a quantum mechanical problem
to our first system: the particle in a box.
This system is very simple which is one reason for beginning with it. It also can
be used as a zeroth order model for certain physical systems.
We shall soon see that the particle in a box is a physically unrealistic system and,
as a consequence, we must violate one of our criteria for a good wavefunction.
Nevertheless it is of great pedagogical and practical value.

4.1. The 1D Particle in a Box Problem


Consider the potential, V (x), shown in the figure and given by

x0


V (x) =
0
0<x<a .

xa

(4.1)

Because of the infinities at x = 0 and x = a, we need to partition the x-axis into


the three regions shown in the figure.

31
31

Now, in region I and III, where the potential is infinite, the particle can never
exist so, must equal zero in these regions.
The particle must be found only in region II.
The Schrdinger equation in region II is (V (x) = 0)
~2 d2 (x)

= E,
H = E =
2m dx2

(4.2)

which can be rearranged into the form


d2 (x) 2mE
+ 2 (x) = 0.
dx2
~

(4.3)

The general solution of this dierential equation is


(x) = A sin kx + B cos kx,
where k =

(4.4)

2mE
.
~2

Now must be continuous for all x. Therefore it must satisfy the boundary
conditions (b.c.): (0) = 0 and (a) = 0.

32

From the (0) = 0 b.c. we see that the constant B must be zero because
cos kx|x=0 = 1.
So we are left with (x) = A sin kx for our wavefunction.
As can be inferred from the following figure, the second b.c., (a) = 0, places
certain restrictions on k.

In particular,

n
,
n = 1, 2, 3, .
a
The values of k are quantized. So, now we have
nx
.
n (x) = A sin
a
kn =

The constant A is the normalization constant. We obtain A from


Z
Z a
nx
nx

sin
dx.
n (x) n (x) = 1 =
A2 sin
a
a

0
Letting u =

x
,
a

du = a dx, this becomes


Z

A2 a
a/
2a
=
.
1=A
sin2 nudu = A2
0
2
/ 2

(4.5)

(4.6)

(4.7)

(4.8)

33

Solving for A gives


A=

2
.
a

(4.9)

Thus our normalized wavefunctions for a particle

0q
2
n (x) =
sin nx
a
a

in a box are
I
II

(4.10)

III

Is this wavefunction OK?

We can get the energy levels from kn =


En =

2mEn
~2

and kn =

n
:
a

h
n2 2 ~2 ~= 2
n2 h2
=
.
2ma2
8ma2

(4.11)

4.2. Implications of the Particle in a Box problem


Zero Point Energy

34

The smallest value for n is 1 which corresponds to an energy of


E1 =

h2
6= 0.
8ma2

(4.12)

That is, the lowest energy state, or ground state, has nonzero energy. This residual
energy is called the zero point energy and is a consequence of the uncertainty
principle.
If the energy was zero then we would conclude that momentum was exactly zero,
p = 0. But we also know that the particle is located within a finite region of
space, so
x 6= .
Hence,
xp = 0 which violates the uncertainty principle.
Features of the Particle in a Box Energy Levels
The energy level spacing is
4E = En+1 En =
4E = (2n + 1)

h2
8ma2

(n + 1)2 h2
n2 h2
h2
2
2

=
(n
/
+
2n
+
1

n
/
)
8ma2
8ma2
8ma2
(4.13)

This spacing increases linearly with quantum level n


This spacing decreases with increasing mass
This spacing decreases with increasing a
It is this level spacing that is what is measured experimentally
The Curvature of the Wavefunction

35

The operator for kinetic energy is T =

~ 2 d2
.
2m dx2

The important part of this is

d2
.
dx2

From freshman calculus we know that the second derivative of a function describes
its curvature so, a wavefunction with more curvature will have a larger second
derivative and hence it will posses more kinetic energy.
This is an important concept for the qualitative understanding of wavefunctions
for any quantum system.
Applying this idea to the particle in a box we an anticipate both zero point energy
and the behavior of the energy levels with increasing a.
We know the wavefunction is zero in regions I and III. We also know that
the wave function is not zero everywhere. Therefore it must do something
between x = 0 and x = a. It must have some curvature and hence some zero
point energy.
As a is increased, the wavefunction is less confined and so the curvature does
not need to be as great to satisfy the boundary conditions. Therefore the
energy levels decrease in energy as does their dierence.
The particle in a box problem illustrates some of the many strange features of
quantum mechanics.
We have already seen such nonclassical behavior as quantized energy and zero
point energy.
As another example consider the expectation value of position for a particle in
the second quantum level:
Z
Z
a
2 a
2

hxi =
(4.14)
2 (x)x 2 (x)dx =
x sin2 [ x]dx =
a 0
a
2

36

yet the probability of finding the particle at x = a2 is zero: 2 ( a2 ) = 0. There is


a node at x = a2 . So even though the particle may be found anywhere else in the
box and it may get from the left side of the node to the right side, it can never
be found at the node.

37

5. The Harmonic Oscillator


The harmonic oscillator model which is simply a mass undergoing simple harmonic
motion. The classical example is a ball on a spring

The harmonic oscillator is arguably the single most important model in all of
physics.
We shall begin by reviewing the classical harmonic oscillator and than we will
turn our attention to the quantum oscillator.
The force exerted by the spring in the above figure is F = k(R Req ), where k
is the spring constant and Req is the equilibrium position of the ball.
Setting x = R Req we can measure the displacement about the equilibrium
position.
38
38

From Newtons law of motion F = ma = m ddt2x , we get


d2 x
k
d2 x
m 2 = kx 2 + x = 0
dt
dt
m

(5.1)

This is second order dierential equation which we already know the solutions to:
x = A sin t + B cos t,
where =
conditions.

k
m

(5.2)

and A and B are constants which are determined by the initial

For quantum mechanics it is much more convenient to talk about energy rather
than forces, so in going to the quantum oscillator, we need to express the force of
the spring in terms of potential energy V . We know
Z
1
V = F dx = kx2 + C.
(5.3)
2
Since energy is on an arbitrary scale we can set C = 0. Thus V = 12 kx2 .
By postulate III the Schrdinger equation becomes

~2 d2
1

= E
H
+ kx2 = E.

2
|2m{zdx } |2 {z }
K.E.

(5.4)

P.E.

This can be rearrange into the form


~2 d2
+
2m dx2

1 2
kx E = 0
2

(5.5)

This dierential equation is not easy to solve (you can wait to solve it in graduate
school).

39

The equation is very close to the form of a know dierential equation called Hermites dierential equation the solutions of which are called the Hermite polynominals.
As it turns out, the solutions (the eigenfunctions) to the Schrdinger equation for
the harmonic oscillator are
1

2
km 4
1
y2
n (y) = An Hn (y)e , y =
(5.6)
x, An = p
,
2
~
2n n!

where An is the normalization constant for the nth eigenfunction and Hn (y) are
the Hermite polynomials.
The eigenvalues (the energy levels) are
1
En = (n + )~,
2
where again =

(5.7)

k
.
m

Note the energy levels are often written as


1
En = (n + )h 0 ,
2
where 0 =

1
2

k
m

(5.8)

and is called the vibrational constant.


See Fig. 11.12 Laidler&Meiser

5.1. Interesting Aspects of the Quantum Harmonic Oscillator


It is interesting to investigate some of the unintuitive properties of the oscillator
as we have gone quantum mechanical

40

1. Consider the ground state (the lowest energy level)


There is residual energy in the ground state because
1
E0 = (0 + )~.
2
Just like for the particle in a box, this energy is called the zero point
energy.
It is a consequence of uncertainty principle
If the ground state energy was really zero, then we would conclude
that the momentum of the oscillator was zero.
On the other hand, we would conclude the particle was located at
the bottom of the potential well (at x = 0)
Thus we would have p = 0, x = 0, so px = 0 Not allowed!
The uncertainty principle forces there to be some residual zero
point energy.
2. Consider the wavefunctions.
The wavefunctions penetrate into the region where the classical particle
is forbidden to go
The wavefunction is nonzero past the classical turning point.
The probability distribution ||2 becomes more and more like what is
expected for the classical oscillator when v .
This is a manifestation of the correspondence principle which
states that for large quantum numbers, the quantum system must
behave like a classical system. In other words the quantum mechanics must contain classical mechanics as a limit.
3. Interpretation of the wavefunctions and energy levels

41

Remember the wavefunctions are time independent and the energy levels are stationary
If a molecule is in a particular vibrational state it is NOT vibrating.

5.2. Spectroscopy (An Introduction)


The primary method of measuring the energy levels of a material is through the
use of electromagnetic radiation.
Experiments involving electromagnetic radiationmatter interaction are called
spectroscopies.
Atoms and molecules absorb or emit light only at specific (quantized) energies.
These specific values correspond to the energy level dierence between the initial
and final states.

42

Key Equations for Exam 1


Listed here are some of the key equations for Exam 1. This section should not
substitute for your studying of the rest of this material.
The equations listed here are out of context and it would help you very little to
memorize this section without understanding the context of these equations.
The equations are collected here simply for handy reference for you while working
the problem sets.

Equations
The short cut for getting the normalization constant (1D, see above for 3D).
sZ
N=

space

|unnorm (x)|2 dx.

(5.9)

The normalized wavefunction:


norm =

.
N unnorm

(5.10)

The Schrdinger equation (which should be posted on your refrigerator),


= E.
H

(5.11)

43
43

The Schrdinger equation for 1D problems as a dierential equation,


~2 d2
+ (V (x) E) = 0.
2m dx2
How to get the average value for some property (1D version),
Z

dx.
h
i =

(5.12)

(5.13)

space

The momentum operator

px = i~

.
x

(5.14)

Normalized wavefunctions for the 1D particle in a box,


r
nx
2
n (x) =
sin
.
a
a

(5.15)

The energy levels for the 1D particle in a box,


En =

h
n2 2 ~2 ~= 2
n2 h2
=
.
2ma2
8ma2

(5.16)

The energy level spacing for the 1D particle in a box,


4E = (2n + 1)

h2
8ma2

(5.17)

The wavefunctions for the harmonic oscillator are


1

2
km 4
1
y2
x, An = p
n (y) = An Hn (y)e , y =
,
2
n
~
2 n!

(5.18)

where An is the normalization constant for the nth eigenfunction and Hn (y)
are the Hermite polynomials.

The energy levels are


1
En = (n + )~, =
2

k
m

(5.19)

44

Part II
Quantum Mechanics of Atoms
and Molecules

45
45

6. Hydrogenic Systems
Now that we have developed the formalism of quantum theory and have discussed
several important systems, we move onto the quantum mechanical treatment of
atoms.
Hydrogen is the only atom for which we can exactly solve the Schrdinger equation
for. So this will be the first atomic system we discuss.
The Schrdinger equation for all the other atoms on the periodic table must be
solved by approximate methods.

6.1. Hydrogenic systems


Hydrogenic systems are those atomic systems which consist of a nucleus and one
electron. The Hydrogen atom (one proton and one electron) is the obvious example
Ions such as He+ and Li2+ are also hydrogenic systems.
These system are centrosymmetric. That is they are completely symmetric about
the nucleus.
The obvious choice for the coordinate system is to use spherical polar coordinates

46
46

with the origin located on the nucleus.


The classical potential energy for these hydrogenic systems is
V (r) =
So the Hamiltonian is

Ze2
.
(4 0 )r

(6.1)

2
2
2 + Ze .
= ~
H
2me
(4 0 )
r

(6.2)

Schrdingers equation (in spherical polar coordinates) becomes

E = H
(6.3)

2
2
Ze
~ 2

+
E =
2me
(4 0 )
r
2

~ 1 2
1

1 2
1
Ze2
E =
r
+
sin +

+
2me r2 r r r2 sin
sin2 2
(4 0 )r
The Hamiltonian is (almost) the sum of a radial part (only a function of r) and
an angular part (only a function of and ):
=H
rad + 1 H
ang ,
H
r2
rad
H
and

(6.4)

Ze2
~2 1 2
r

=
2me r2 r r (4 0 )r

2
ang = ~
H
2me

1 2
1
sin +
sin
sin2 2

(6.5)

(6.6)

Since the Hamiltonian is the sum of two terms, must be a product state.
(r, , ) = rad (r)ang (, )

(6.7)

It turns out that solving the Schrdinger equation,


ang ang (, ) = Eang (, ),
H

(6.8)

47

yields
ang (, ) = Ylm (, ),

(6.9)

where the Ylm (, )s are the spherical harmonic functions characterized by quantum numbers l and m. The spherical harmonics are known functions. (Mathematica
knows them and you can use them just like any other built-in function like sine
or cosine.)
We shall use the spherical harmonics more next semester when we develop the
quantum theory of angular momentum.
ang is found to be
It also turns out that the energy associated with H
E = El =

l(l + 1)~2
.
2me

(6.10)

So,
2
ang ang (, ) = l(l + 1)~ ang (, )
H
2me

(6.11)

Now lets denote the radial part of the wavefunction as rad (r) = R(r).
The full Schrdinger equation becomes

H(r,
, ) = E(r, , )

(6.12)

HR(r)Y
lm (, ) = ER(r)Ylm (, )

rad + 1 H
ang R(r)Ylm (, ) = ER(r)Ylm (, ),
H
r2
ang we get
Operating with H

l(l + 1)~2

R(r)Ylm (, ) = ER(r)Ylm (, )
Hrad +
2me r2

(6.13)

48

The Ylm (, ) can now be cancelled to leave a one dimensional dierential equation:

Ze2
l(l + 1)
~2 1 2
r

R(r) = ER(r).
(6.14)
2me r2 r r 4 0 r
r2
This dierential equation is very similar to a known equation called Laguerres
dierential equation which has as solutions the Laguerre polynomials Lln (x).
In fact, the solutions to our dierential equation are closely related to the Laguerre
polynomials.
l

2
2
/n 2l+1
Rnl () = Anl
,
(6.15)
e
Ln+1
n
n

where the normalization constant, Anl , depends on the n and l quantum numbers
as
s
3
2Z
(n l 1)!
Anl =
(6.16)
na0
2n[(n + l)!]3
The energy eigenvalues, i.e., the energy levels are given by
En =

Z 2R
n2

(6.17)

Note: The energy levels are determined by n alonel drops out.


Also Note: the energy levels are the same as for the Bohr model.
So, the total wavefunction that describes a hydrogenic system (ignoring the spin
of the electron, which will be briefly discussed later) is
nlm (r, , ) = Rnl (r)Ylm (, )

(6.18)

6.2. Discussion of the Wavefunctions


We are now very close to having the atomic orbitals familiar from freshman chemistry.

49

We have explicitly derived the physicists picture of the atomic orbitals


orbital n l

wavefunctions ( = r/a0 )

1s
2s
2p

0
0
0
1
0
1
2

1s = 100 = e

2s = 200 = 1 2 e/2
2p0 = 210 = e/2 cos
2p1 = 211 = e/2 sin ei
3d0 = 320 = 2 e/3 (3 cos2 1)
3d1 = 321 = R32 (r) cos sin ei
3d2 = 322 = R32 (r) sin2 ei2

3d

1
2
2
2
3
3
3

0
0
1
1
2
2
2

The wavefunctions in the physicists picture are complex (they have real and
imaginary components). The wavefunctions that chemists like are pure real. So
one needs to form linear combinations of these orbitals such that these combinations are pure real.
The atomic orbital you are used to from freshman chemistry are the chemists
picture of atomic orbitals
In the above table 1s , 2s , 2p0 , 3d0 are pure real and so these are the same in
the chemists picture as in the physicists picture.
The table below lists the atomic orbitals in the chemists picture as linear combinations of the physicists picture wave functions.

50

orbital n l

1s
2s
2p

0
0
0
1

1
2
2
2
2

3d

3
3
3
3

0
0
1
1

wavefunctions ( = r/a0 )

1s = 1s
2s = 2s
pz = 2p0

2px = 12 2p1 + 2p1

1 1 2py = i1 2 2p1 2p1

2 0

3dz2 = 3d0

2 1 3dxz = 12 3d1 + 3d1

2 1 3dyz = i1 2 3d1 3d1

2 2 3dxy = 12 3d2 + 3d2

3dx2 y2 = i1 2 3d2 3d2

6.3. Spin of the electron

As we know from freshman chemistry, electrons also posses an intrinsic quantity


called spin.
Spin is actually rather peculiar so we will put o a more detailed discussion until
next semester.
For now we must be satisfied with the following:
There are two quantum numbers associated with spin: s and ms
s is the spin quantum number and for an electron s = 1/2 (always).
ms is the spin orientation quantum number and ms = 1/2 for electrons.
The spin wavefunction is a function in spin space not the usual coordinate space,
so we can not write down an explicit function of the coordinate space variables.

51

We simply denote the spin wavefunction generally as s,ms and tack it on as


another factor of the complete wavefunction.
When a particular spin state is needed a further notation is commonly used:
1 , 1 (the spin-up state) and 1 , 1 (the spin-down state)
2 2

6.4. Summary: the Complete Hydrogenic Wavefunction


We are now in position to fully describe all properties of hydrogenic systems
(except for relativistic eects)
The full wave function is
n,l,m,s,ms = n,l,m s,ms

(6.19)

= Rnl (r)Yl,m (, )
The energy is given by
Z 2R
,
(6.20)
n2
where recall. Again note that for a free hydrogenic system the total energy depends
only on the principle quantum number n.
En =

The quantum numbers of the hydrogenic system


The principle quantum number, n: determines the total energy of the systems and the atomic shells.
The principle quantum number, n, can take on values of 1,2,3. . .
The angular momentum quantum numbers, l: determines the total angular
momentum of the system. It also determines the atomic sub-shells

52

The angular momentum quantum number, l, can take on values of 0,


1, . . . (n 1)
For historical reasons l = 0 is called s, l = 1 is called p, l = 2 is called
d, l = 3 is called f etc.
The orientation quantum number, m: determine the projection of the angular momentum onto the z-axis. It also determines the orientation of the
atomic sub-shells
The magnetic quantum number, m, can take on values of 0, 1, . . . l.
The spin quantum number, s: determines the total spin angular momentum.
For electrons s = 1/2.
The spin orientation quantum number, ms : determines the projection of the
spin angular momentum onto the z-axis (i.e., spin-up or spin-down).
For electrons ms = 1/2
We have accomplished quite a bit. We have determined all that we can about the
hydrogen atom within Schrdingers theory of quantum mechanics.
This is not the full story however. The Schrdinger theory is a non-relativistic
one; that is, it can not account for relativistic eects which show up in spectral
data. We also had to add spin in an ad hoc manner to account for what we know
experimentallyspin did not fall out of the theory naturally.
Dirac, in the late 1920s, developed a relativistic quantum theory in which the
well established phenomenon of spin arose naturally. His theory also made the

53

bold prediction of the existence anti-matter that has now been verified time and
again.
The Dirac theory was still not fully complete, because there still existed experimental phenomena that was not properly described. In 1948 Richard Feynman
developed the beginnings of quantum electrodynamics (QED). QED is the best
theory ever developed in terms of matching with experimental data.
Both the relativistic Dirac theory and QED are beyond our reach, so we limit
ourselves to the non-relativistic Schrdinger theory.

54

7. Multi-electron atoms
7.1. Two Electron Atoms: Helium
We now consider a system consisting of two electrons and a nucleus; for example,
helium.
Although the extension from hydrogen to helium seems simple it is actually extremely complicated. In fact, it is so complicated that it cant be solved exactly.
The helium atom is an example of the three-body-problemdicult to handle
even in classical mechanicsone can not get a closed form solution.
The Hamiltonian for helium is
2
= ~ 21
H
2m
| {ze }

~2 2

2me 2
| {z }

Ze2

4 r
| {z 0 1}

Ze2

+
4 0 r2
| {z }

K.E of electron 1 K.E of electron 2 P.E of electron 1 P.E of eletcron 2

e2
4 0 r12
| {z
}

, (7.1)

elec.elec. repulsion

where r12 = |r1 r2 | is the distance between the electrons.

The electronelectron repulsion term is responsible for the diculty of the problem. It makes a closed form solution impossible.
The problem must be solved by one of the following methods
Numerical solutions (we will not discuss this)
55
55

Perturbation theory (next semester)


Variational theory (next semester)
Ignore the electronelectron repulsion (good for qualitative work only)

7.2. The Pauli Exclusion Principle


Electron are fundamentally indistinguishable. They can not truly be labelled.
All physical properties of a system where we have labelled the electrons as, say, 1
and 2 must be exactly the same as when the electrons are labelled 2 and 1.
Now, only ||2 is directly measurablenot itself.
All this implies that

+(2, 1),
(1, 2) =
or

(2, 1)

symmetric
(7.2)
antisymmetric

The Pauli exclusion principle states: The total wavefunctions for fermions
(e.g., electrons) must be antisymmetric under the exchange of indistinguishable
fermions.
Note: a similar statement exists for bosons (e.g., photons): The total wavefunction
for bosons must be symmetric under exchange of indistinguishable bosons.
Let us consider the two electron atom, helium

56

The total wavefunction is


= (1, 2)(1, 2)

(7.3)

Since a complete solution for helium is not possible we must use approximate
wavefunctions. Since we are doing this, we may as well simplify matters and use
product state wavefunctions (products of the hydrogenic wavefunctions).
= (1)(2)(1)(2),
| {z }| {z }

(7.4)

spatial part spin part

where the single particle wavefunctions are that of the hydrogenic system.
The Pauli exclusion principle implies that if the spatial part is even with respect
to exchange then the spin part must be odd. Likewise if the spatial part is odd
then the spin part must be even.
Now lets blindly list all possibilities for the ground state wave function of helium
a = 1s (1)(1)1s (2)(2)

(7.5)

b = 1s (1)(1)1s (2)(2)
c = 1s (1)(1)1s (2)(2)
d = 1s (1)(1)1s (2)(2)
These appear to be four reasonable ground state wavefunctions which would imply a four-fold degeneracy. However considering the symmetry with respect to
exchange we see the following
a has symmetric spatial and spin parts and is there for symmetric. It must
be excluded.
Similarly for d .
b and c have symmetric spatial parts, but the spin part is neither symmetric or antisymmetric. So, one must make an antisymmetric linear combination of the spin parts.

57

The appropriate linear combination is


(1)(2) (2)(1).

(7.6)

So the ground state wave function for helium is


g = 1s (1)1s (2) [(1)(2) (2)(1)] .

(7.7)

Consequences of the Pauli exclusion principle


No two electrons can have the same five quantum numbers
Electrons occupying that same subshell must have opposite spins

7.3. Many Electron Atoms


The remaining atoms on the periodic table are handled in a manner similar to
helium.
Namely the wavefunction is product state that must be antisymmeterized in accordance with the Pauli exclusion principle.
The product wavefunction for the ground state is determined by applying the
aufbau principle. The aufbau principle states that the ground state wavefunction
is built-up of hydrogenic wavefunctions
To arrive at an antisymmetric wavefunction we construct the Slater determinant:

(1)(1)
1s (1)(1)
n (1)(1)
n (1)(1)
1s

1s (2)(2)
1s (2)(2)
n (2)(2)
n (2)(2)

=
(7.8)
..
..
..
..
..

.
.
.
.
.

(N)(N) (N)(N)

(N
)(N)

(N)(N)
1s
1s
n
n

58

The reason one can be sure that this wavefunction is the antisymmeterized is that
we know from linear algebra that the determinant is antisymmetric under exchange
of rows (corresponds to exchanging two electrons). It is also antisymmetric under
exchange of columns.
Another property of the determinant is that if two rows are the same (corresponds
to two electrons in the same state) the determinant is zero. This agrees with the
Puli exclusion principle.
As an example consider lithium:
There are three electrons so we need three hydrogenic wavefunctions: 1s ,
1s , and 2s (or 2s ).
We construct the Slater determinant as

(1)(1) (1)(1) (1)(1)


1s
1s
2s

1 = 1s (2)(2) 1s (2)(2) 2s (2)(2)

(3)(3) (3)(3) (3)(3)


1s
1s
2s
or

(1)(1) (1)(1) (1)(1)


1s
2s

1s

2 = 1s (2)(2) 1s (2)(2) 2s (2)(2)

(3)(3) (3)(3) (3)(3)


1s
1s
2s

(7.9)

(7.10)

The short hand notation for these states is (1s)2 (2s)1


7.3.1. The Total Hamiltonian

The total Hamiltonian for a many electron (ignoring spin-orbit coupling which
will be discussed next semester) atom is
"
#
N
2
2
X
X e2
Ze
~
=
2i
+
(7.11)
H
2m
4 0 ri j>i 4 0 rij
e
i=1

59

8. Diatomic Molecules and the Born


Oppenheimer Approximation
Now that we have applied quantum mechanics to atoms, we are able to begin the
discussion of molecules.
This chapter will be limited to diatomic molecules.

8.1. Molecular Energy


A diatomic molecule with n electrons requires that 3n+6 coordinates be specified.
Three of these describe the center of mass position.
3n of these describe the position of the n electrons.
This leaves three degrees of freedom (R, , ) which describe the position of the
nuclei relative to the center of mass. R determines the internuclear separation
and and determine the orientation.

60
60

8.1.1. The Hamiltonian


In the center of mass coordinates the Hamiltonian for a diatomic molecule is
= TN + Te + VN N + VNe + Vee .
H

(8.1)

TN is the nuclear kinetic energy operator and is given by


~2 2
~2 2
~2 2

+ J ,
R
TN = N =
2
2R2 R R 2
where J is angular momentum operator for molecular rotation and =
the reduced mass of the diatomic molecule.

(8.2)
m1 m2
m1 +m2

is

P
~2 2
Te = i 2m
ei is the kinetic energy operator for the electrons.
e
VNN =

ZA ZBe e2
4 0 R

P
VNe = i
P
Vee = i>j

is the nuclearnuclear potential energy operator.

ZA e2
4 0 rAi

e2
4 0 rji

ZB e2
4 0 rBi

is the nuclearelectron potential energy operator.

is the electronelectron potential energy operator.

61

8.1.2. The BornOppenheimer Approximation


The BornOppenheimer approximation: The nuclei move much slower than
the electrons. (classical picture)
We put the BornOppenheimer approximation to work by first defining an eective Hamiltonian
ef f = Te + VN N + VNe + Vee .
H
(8.3)
The approximation comes in by treating R as a parameter rather than an operator
(or variable). So one writes
ef f e (R, {ri }) = Ee (R) e (R, {ri }).
H

(8.4)

e is the so-called electronic wavefunction.


Now the Schrdinger equation for the diatomic molecule is

ef f (R, {ri }) = E(R, {ri }).


TN + H

(8.5)

Since the Hamiltonian is a sum of two terms, one can write the wavefunction
(R, {ri }) as a product wavefunction
= N e ,

(8.6)

where N is the so-called nuclear wavefunction.


Substituting the product wavefunction into the Schrdinger equation gives

ef f N e = EN e
(8.7)
TN + H

TN + Ee (R) N /e = EN /e

TN + Ee (R) N = EN .

62

The last equation is exactly like a Schrdinger equation with a potential equal to
Ee (R).
One now models Ee (R) or determines it experimentally.

8.2. Molecular Vibrations


As stated earlier R is the internuclear separation and and determine the
orientation. Consequently, R is the variable involved with vibration whereas
and are involved with rotation.
Considering only the R part of the Hamiltonian (under the BornOppenheimer
approximation), we have
2 2

+ Ee (R) vib = Evib vib .


(8.8)
2 R2
It is convenient at this point to expand Ee (R) in a Taylor series about the equilibrium position, Req :

E
1 2E
0
(R Req ) +
(R Req )2 + . (8.9)
Ee (R) = E +
2
R Req
2! R Req
Now E 0 is just a constant which, by choice of the zero of energy, can be set to an
arbitrary value.
Since we are at a minimum,
One defines

2E
R2

Req

R Req

must be zero, so the linear term vanishes.

ke as the force constant.

The remaining terms in the expansion can collective be defined as O[(RReq )3 ]


Vanh , the anharmonic potential.

63

As a first approximation we can neglect the anharmonicity. With this, the Schrdinger
equation becomes
2 2

~
1
2

+ ke (R Req ) vib = Evib vib .


(8.10)
2 R2 2
If we let x = (R Req ) this becomes

2 2
1
~
2
+ ke x vib = Evib vib ,

2 x2 2

(8.11)

which is exactly the harmonic oscillator equation. Hence

where

(8.12)

1
Evib,n = hc
e (n + ),
2

(8.13)

ke
.
~

And

where
e

2
vib,n = An Hn ( x)ex /2 ,

1
2

ke
.

8.2.1. The Morse Oscillator


Neglecting anharmonicity and using the harmonic oscillator approximation works
well for low energies. However, it is a poor model for high energies.
For high energies we need a more realistic potentialone that will allow of bond
dissociation.
The Morse potential
Ee (R) = De [1 e(RRe q ) ]2 ,

(8.14)

64

where De is the well depth and = 2c


e 2D
is the Morse parameter. Note:
e
this expression for the Morse potential has the zero of energy at the bottom of
the well (i.e. R = Req , ;Ee (Req ) = 0).

The Morse Potential can also be written as


Ee (R) = De [e2(RReq ) 2e(RRe q ) ].

(8.15)

Now the zero of energy is the dissociated state (i.e. R , ;Ee (R ) = 0).
We approach this quantum mechanical problem exactly like all the other.
The Schrdinger equation is

2 2
~
(RReq ) 2
+ De [1 e
] vib = Evib vib

2 R2

(8.16)

This is another dierential equation that is dicult to solve.

As it turns out, this Schrdinger equation can be transformed into a one of a broad
class of known dierential equations called confluent hypergeometric equations
the solutions of which are the confluent hypergeometric functions, 1 F1 .
Doing this yields the wavefunctions of the form
vib,n (z) = z Apn ez 1 F1 (n, 1 + 2Apn , 2z),

2De x
e ,
z =
h

2
,
A =
h
p
1 n
pn =
De + 2
A
and energy levels of the form
1
1
Evib,n = De + hc
e xe (n + )2 ,
e (n + ) hc
2
2

(8.17)

(8.18)

65

where
e xe together is the anharmonicity constant, with xe =

hc
e
.
4De

See Handout
8.2.2. Vibrational Spectroscopy
Infrared (IR) and Raman spectroscopy are the two most widely used techniques
to probe vibrational levels.
The spectral peaks appear at v =

4E
hc

(in units of wavenumbers, cm1 ).

The transition from the n = 0 to the n = 1 state is called the fundamental


transition.
Transitions from n = 0 to n = 2, 3, 4 are called overtone transitions.
Transitions from n = 1 to 2, 3, 4 , n = 2 to 3, 4, 5 , etc. are called hot
transitions (or hot bands)
Since the energy levels depend on mass, isotopes will have a dierent transition
energy and hence appear in a dierent place in the spectrum. Heavier isotopes
have lower transition energies.

66

9. Molecular Orbital Theory and


Symmetry
9.1. Molecular Orbital Theory
One of the most important concepts in all of chemistry is the chemical bond.
In freshman chemistry we learn of one model for chemical bondingVSEPR (valence shell electron-pair repulsion) theory, where hybridized atomic orbitals determine the bonding geometry of a given molecule.
We are now prepared to discuss a bonding theory that is more rigorously based
in quantum mechanics.
Basically we will treat the molecules in the same way as all our other quantum
mechanical problems (e.g., particle in a box, harmonic oscillator, etc.)
As you might expect, it is not possible to obtain the exact wavefunctions and
energy levels so, we must settle for approximate solutions.
As a first example, let us consider the molecular hydrogen ion H+
2.
The Hamiltonianfor H+
2 is
= TN + Tel + VNel + VNN
H

(9.1)

67
67

We use the Born-Oppenheimer approximation and treat the nuclear coordinates


as a parameters rather than as variables. So we only worry about parts of the
Hamiltonian that deal with the electron.
The eective Hamiltonian becomes
= Tel + VNel
H
e2
e2
~2 2

.
=
2me
4 0 rA 4 0 rB
The eigenfunctions of this Hamiltonian are called molecular orbitals.

(9.2)

The molecular orbitals are the analogues of the atomic orbitals.


Atomic orbitals: Hydrogen is the prototype and all other atomic orbitals
are built from the hydrogen atomic orbitals.
Molecular orbitals: The hydrogen molecular ion is the prototype and all
other molecular orbitals are built from the hydrogen molecular ion molecular
orbitals.

There is one significant dierence between the above, which is the hydrogen atomic
orbitals are exact whereas the hydrogen molecular ion molecular orbitals are not
exact.
In fact, we shall see that these molecular orbitals are constructed as linear combinations of atomic orbitals.

9.2. Symmetry
Let the atoms of the hydrogen molecular ion lie on the z-axis of the center of mass
coordinate system.

68

Inversion symmetry
The potential field of the hydrogen molecular ion is cylindrically symmetric
about the z-axis.
Because of the symmetry the electron density at (x, y, z) must equal the
electron density at (x, y, z).
The above symmetry therefore requires that the molecular orbitals be eigenfunctions of the inversion operator, . That is
(x, y, z) = (x, y, z) = a(x, y, z).

(9.3)

Moreover the eigenvalue a can be either +1 or 1.


If a = +1 the molecular wavefunction is even with respect to inversion and
is called gerade and labelled with a g: g = g
If a = 1 the molecular wavefunction is odd with respect to inversion and
is called ungerade and labelled with a u: u = u
The terms gerade and ungerade apply only to systems that posses inversion
symmetry.

Cylindrical symmetry

69

The cylindrical symmetry implies that the potential energy can not depend
on the .
The molecular wavefunction is described by an eigenvalue = 0, 1, 2, . . .
We use to label the molecular orbitals as shown in the table

0
label

1 2

Mirror plane symmetry

70

There is also a symmetry about the x-y plane called horizontal mirror plane
symmetry: operator
h.
Thus the molecular wavefunction must be an eigenfunction of
h with eigenvalue 1.
If the eigenvalue is +1 (even with respect to
h ) the molecular orbital
is called a bonding orbital.
If the eigenvalue is 1 (odd with respect to
h ) the molecular orbital
is called an antibonding orbital.
There are also vertical mirror plane symmetries, but we will put that discussion o for the time being.

71

10. Molecular Orbital Diagrams


10.1. LCAOLinear Combinations of Atomic Orbitals
Now that we know what symmetry the molecular orbitals must posses, we need
to find some useful approximations for them.
Useful can mean qualitatively useful or quantitatively useful.
Unfortunately we cant have both.
We will discuss the approximation which models the molecular orbitals as linear
combinations of atomic orbitals (LCAO).
LCAO is qualitatively very useful but it lacks quantitative precision.
Let us again consider the hydrogen molecular ion H+
2 : let one H atom be labelled
A and the other labelled B.
Linear combination of the 1s atomic orbital from each H atom is used for the
molecular orbital of H+
2:
(1sA ) = kerA /a0
(10.1)
and
(1sB ) = kerB /a0

(10.2)

72
72

We construct two molecular orbitals as


+ = C+ (1sA + 1sB )

(10.3)

= C (1sA 1sB )

(10.4)

and
The normalization condition is

d = 1

(10.5)

As can be seen from the above figure, + represents a situation in which the
electron density is concentrated between the nuclei and thus represents a bonding
orbital.
Conversely represents a situation in which the electron density is very low
between the nuclei and thus represents an antibonding orbital
10.1.1. Classification of Molecular Orbitals
With atoms we classified atomic orbitals according to angular momentum.
For molecular orbitals we shall also classify them according to angular momentum.
But we shall also classify them according to their inversion symmetry and wether
or not they are bonding or antibonding.

73

The classification according to angular momentum is as follows.

0
orbital symbol

1 2

Atomic orbitals with m = 0 form type molecular orbitals, e.g., s , pz .


Those with m = 1 form type molecular orbitals, e.g., px etc.
The classification according to inversion symmetry is simply a subscript g or
u. For example, g or u etc.
The classification according to bonding or antibonding is an asterisk is used to
denote antibonding. For example, g is a bonding orbital and u is an antibonding
orbital.

10.2. The Hydrogen Molecule


Let us now consider the hydrogen molecule. This molecules is a homonuclear
diatomic with two electrons.
If the two atoms are infinitely far apart. The ground state of the system would
consist of two separate hydrogen molecules in their ground atomic states: (1s)1

74

As the atom are brought closer together, their respective s orbitals begin to overlap.

It is now more appropriate to speak in terms of molecular orbitals, so one forms


linear combinations of the atomic orbitals.
There are two acceptable linear combinations. These are
g = 1sA + 1sB

(10.6)

u = 1sA 1sB .

(10.7)

and

75

It can be shown mathematically that the energy level associated with g is lower
than u .
We can intuit this qualitatively however since the u orbital must have a node
whereas the g does not.
It is also to be expected since we know H2 is a stable molecule.

10.3. Molecular Orbital Diagrams


The energy levels associated with the molecular orbitals are drawn schematically
is what is called a molecular orbital diagram.
The molecular orbital diagram for H2 is shown below

Molecular orbital diagrams can be drawn for any molecule. Some get very complicated. We will focus on the second row homonuclear diatomics and some simple
heteronuclear diatomics.

76

The molecular orbital diagrams for the second row homonuclear diatomics are
rather simple.
See Supplement
The supplement that follows this section contains examples for each of the second
row diatomics.
Heteronuclear diatomics are some what more complicated since there is a disparity
in the energy levels of the atomic orbitals for the separated atoms. This disparity
is not present for homonuclear diatomics.
A consequence of this energy level disparity is that molecular orbitals may be
formed from nonidentical atomic orbitals. For example a high lying 1s orbital
may combine with a low lying 2s orbital to form a molecular orbital.
The supplement that follows this section contains some examples of heteronuclear
diatomics.
Bond order
One important property that can be predicted from the molecular orbital
diagrams is bond order.
Bond order is defined as
BO =

1
(# of bonding electrons # of antibonding electrons)
2

(10.8)

Examples follow in the supplement.

77

10.4. The Complete Molecular Hamiltonian and Wavefunction


We have discussed molecular vibrations which under the Born-Oppenheimer approximation are governed by the vibrational Hamiltonian and described by the
vibrational wavefunction.
Likewise we have discussed molecular orbitals which are the electronic wavefunctions.
Next semester we will discuss molecular rotations and just like for vibrations
and electronic transitions they are governed by the rotational Hamiltonian and
described by the rotational wavefunction.
We can succinctly express the Schrdinger equation for a molecule as follows.
(Next semester will we look at the details of this for polyatomic molecules)
mol mol = Emol mol
H
(10.9)

vib + H
rot ele vib rot = (Eele + Evib + Erot ) ele vib rot
ele + H
H

78

11. An Aside: Light ScatteringWhy


the Sky is Blue
This chapter addresses the topic of light scattering from two dierent perspectives.
Classical electrodynamics
Classical statistical mechanics
Since this is not a course on electrodynamics, we have to take several key results
from that theory on faith.

11.1. The Classical Electrodynamics Treatment of Light Scattering


As usual we work under the electric dipole approximation and only focus on the
interaction of the electric field part of light with a dipole.
When the light interacts with the molecule an electric dipole is induced according
to
= E,
(11.1)
where is the polarizability of the molecule describing the flexibility of its
electron cloud.

79
79

For light, the electric field part is


(11.2)

E(t) = E0 cos t.

The polarizability also depends on the positions of nuclei to some degree. That
is, there is a vibrational (and rotational) contribution to the polarizability:
(11.3)

(t) = 0 + 1 cos v t
(here for simplicity we assume only one vibrational mode).
Thus the lightmatter interaction is described as
(t) = (t)E(t) = (0 + 1 cos v t) E0 cos t

(11.4)

= 0 E0 cos t + 1 E0 cos v t cos t

1 E0
cos( v )t + cos( + v )t
= 0 E0 cos
| {zt} + 2
{z
} |
{z
}
|
Rayleigh

Stokes Raman

AntiStokes Raman

where a trig identity was used in the last step.

According to classical electrodynamics an oscillating dipole emits an electromagnetic field at the oscillation frequency.
In this case we see the dipole oscillates at three distinct frequencies: , v
and + v as part of three terms in the above expression.
The first term corresponds to Rayleigh scattering where the scattered light is at
the same frequency as the incident light.
The second term corresponds to Stokes Raman scattering where the scattered
light is shifted to the red of the incident frequency.

80

The third term corresponds to anti-Stokes Raman scattering where the scattered
light is shifted to the blue of the incident frequency.
Classical electrodynamics can describe exactly how the oscillating electric dipole
emits electromagnetic radiation. It can be shown that the emitted intensity is
I=

4 2
,
3c3 0

(11.5)

where 0 = 0 E0 for the case of Rayleigh scattering and 0 = 1 E0 /2 for the case
of Raman scattering.
To explicitly derive this expression we would need a fair bit of electrodynamics
and so the derivation is not shown here.
The important point to note is that I 4 or alternatively I 1/4 . There is a
very strong dependence on frequency (or wavelength).
This quartic scattering dependence is, in fact, the reason why the sky is blue (from
the point of view of classical electrodynamics) and is called the Rayleigh scattering
law.

11.2. The Blue Sky


The spectrum of visible light from the sun incident on the outer atmosphere is
essentially flat as shown below.

81

We just learned that light scatters as it traverses the atmosphere according to


Rayleighs scattering law: I() 1/4 .

The following figures illustrate why Rayleigh scattering implies that the sky is
blue.
11.2.1. Sunsets
We have focused on a blue sky, but red sunsets occur for the same reason
Rayleigh scattering.

82

If we look directly at the sun during a sunset (or sunrise) it appears red because
most of the blue light has scattered in other directions.
This more pronounced at dawn or dusk since the light must traverse more of the
atmosphere at those times then at noonday at which time the sun appears yellow
in color.

11.2.2. White Clouds


We might expect that clouds should be highly colored since they consist of droplets
of water which scatter light very eectively.

83

The key dierence between light scattering by clouds versus by the atmosphere is
the size of the scatterer.
The water droplets are much larger than the wavelenght of the lightquite the
opposite case as above.
In this limit an entirely dierent analysis is madeone does not have Rayleigh
scattering but instead has a process called Mie scattering.
In some contexts, particularly in liquid suspensions, Mie scattering is referred to
as Tyndall scattering

84

Key Equations for Exam 2


Listed here are some of the key equations for Exam 2. This section should not
substitute for your studying of the rest of this material.
The equations listed here are out of context and it would help you very little to
memorize this section without understanding the context of these equations.
The equations are collected here simply for handy reference for you while working
the problem sets.

Equations
The wavefunctions for the hydrogenic system are
(11.6)

nlm (r, , ) = Rnl (r)Ylm (, )


The radial part is.
Rnl () = Anl

2
n

/n

2l+1
Ln+1

2
n

(11.7)

where the normalization constant, Anl , depends on the n and l quantum


numbers as
s
3
2Z
(n l 1)!
Anl =
(11.8)
na0
2n[(n + l)!]3
85
85

The energy levels for the hydrogenic system are given by


En =

Z 2R
n2

(11.9)

The wavefunctions for the harmonic oscillator are


1

2
km 4
1
y2
n (y) = An Hn (y)e , y =
x, An = p
,
2
n
~
2 n!

(11.10)

where An is the normalization constant for the nth eigenfunction and Hn (y)
are the Hermite polynomials.

The energy levels are


1
En = (n + )~, =
2

k
m

(11.11)

The Morse potential is


Ee (R) = De [1 e(RRe q ) ]2 ,
(11.12)
q

where De is the well depth and = 2c


e 2D
is the Morse parameter.
e
Note: this expression for the Morse potential has the zero of energy at the
bottom of the well (i.e. R = Req , ;Ee (Req ) = 0).
The Morse Potential can also be written as
Ee (R) = De [e2(RReq ) 2e(RReq ) ].

(11.13)

Now the zero of energy is the dissociated state (i.e. R , ;Ee (R ) =


0).
The energy levels for the Morse oscillator are of the form
1
1
e xe (n + )2 ,
e (n + ) hc
Evib,n = De + hc
2
2
where
e xe together is the anharmonicity constant, with xe =

(11.14)
hc
e
.
4De

86

Bond order is defined as


BO =

1
(# of bonding electrons # of antibonding electrons)
2

(11.15)

The Rayleigh scattering law is


I() 1/4 4

(11.16)

87

Part III
Statistical Mechanics and The
Laws of Thermodynamics

88
88

12. Rudiments of Statistical


Mechanics
When we study simple systems like a single molecule, we use a very detailed
theory, quantum mechanics.
However, most of the time in the real world we are dealing with macroscopic
systems, say, at least 100 million molecules.
It is simply impossible, even with the fastest computers, to write down and solve
the Schrdinger equation for those 100 million molecules, but often Avogadros
number of molecules.
So we need a less detailed theory called statistical mechanics, which allows one to
handle macroscopic sized systems without losing to much of the rigor.

12.1. Statistics and Entropy


Probability and statistics is at the heart of statistical mechanics.
We will need some definitions
Ensemble: A large collection of equivalent macroscopic systems. The systems are the same except that each one is in a dierent so-called microstate.

89
89

Microstate: The single particular state of one member of the ensemble given
by listing the individual states of each of the microscopic systems in the
macroscopic state.
Configuration: The collection of all equivalent microstates. The number of
possible configurations is defined as W.

Boltzmann developed an equation to connect the microscopic properties of an


ensemble to the macroscopic properties. The Boltzmann equation is
S = k ln W

(12.1)

Where S is entropy and k is Boltzmanns constant.


12.1.1. Combinations and Permutations
Consider a random system that when measured can appear in one of two outcomes
(e.g., flipping coins).
One valuable piece of statistical information about system is knowing how many
dierent ways the system appears p times in, say, outcome 1 after N measurements.
This is given by the mathematical formula for combinations
C(N, p) =

N!
.
p!(N p)!

(12.2)

The number C(N, p) is also called the binomial coecient because it gives the
coecient for the pth order term in the expansion
N

(1 + x) =

N
X

C(N, p)xp .

(12.3)

p=0

90

This formula will allow us to derive a normalization constant so that we can obtain
the probability of obtaining p measurements of state 1.
Set x = 1 in the above. This gives
N

(1 + 1)

N
X

C(N, p)(1)p

(12.4)

p=0

2N =

N
X

C(N, p).

p=0

So the probability of any one outcome of N measurements is


P (N, p) =

1
1
N!
C(N, p) = N
N
2
2 p!(N p)!

(12.5)

For combinations we did not care what order the results of the measurements
occurred.
Sometimes the order is important.
So rather than a particular combination, we are interested in a particular permutation. This is given by
W (N, {Ni }) =

N!
N1 !N2 !N3 !

(12.6)

where N is the total number of measurements and Ni is the number of indistinguishable results of type i.
See Examples on Handout
For both combinations and permutations we need to evaluated factorials.

91

This is no problem for small numbers, but when we consider macroscopic systems
(1020 or so molecules) no calculator can handle factorials of such large numbers.
Sterlings Approximation:
In place of evaluating factorials of large number one can use Sterlings approximation to approximate the value of the factorial.
Sterlings approximation is
ln(N!) ' N ln N N

(12.7)

12.2. Fluctuations
When we list the macroscopic properties of a material such as a beaker of benzene
or the air of the atmosphere, we speak of the average value of the property.
Macroscopic equilibrium is a dynamic rather than static equilibrium. Consequently, the value of a certain property fluctuates about the average value. Often
this fluctuation is not important, but sometimes it is important.
The fluctuation about an average value for any observable property O is described
by the variance which is defined as
2.
2O O2 O

(12.8)

O is consider the range of the observable property.


It can be shown that

O
1

O
N
where N is the number of particles. So for example if N = 1024 then

(12.9)
1
N

= 1012

92

For ensembles having large numbers of particles measured values of a property are
extremely sharply peaked about the average value.

93

13. The Boltzmann Distribution


Consider a isolated system of N molecules that has the set { i } energy levels
associated with it.
Since the system is isolated the total energy, E, and the total number of particles
will be constant.
The total energy is given by
E=

Ni i ,

(13.1)

where Ni is the number of particles in energy state i.


The total number of particles is, of course,
N=

Ni

(13.2)

The number of configurations for the system is then given by the number of
distinct permutations of the system
W =

N!
.
N1 !N2 !

(13.3)

A system in equilibrium always tries to maximize entropy and minimize energy


and so the equilibrium configuration is a compromise between these two cases.
94
94

For the moment let us relax the isolation constraint.


Maximizing entropy corresponds to maximizing W (via S = k ln W ). This would
be the situation in which every particle was in a dierent energy state. That is
all Ni = 1 or 0.
Minimizing energy would be the case where all the particles are in the ground
state (say 1 ).
These two situations are contradictory and some compromise must be obtained.
We start by considering our original systemthat being one with constant energy,
E and number of particles N
To determine the equilibrium configuration we must find the maximum W subject
to the constraint of constant energy and constant number of particles.
This is done using the mathematical technique of Lagrange multipliers (page 951
of your calc book).
We will not discuss this method in detail and consequently we cannot derive the
equilibrium configuration.
The derivation using Lagrange multipliers arrives at the configuration in which
the
gi e i
Ni = N P
,
(13.4)
j
j gj e
| {z }
pi

where

1
kT

and gj denotes the degeneracy of states having energy

j.

95

The pi represents the probability of finding the a randomly chosen particle or


system which has energy i . This is the Boltzmann distribution
gi e i
Pi = P

j gj e

(13.5)

Since we started with a isolated system, and hence T are constants. A given
energy E will correspond to a unique temperature T.
The analysis readily generalizes to variable energy i.e., nonisolated systems by
considering T as a variable.

13.1. Partition Functions


We have already come across both the partition functions that we will use in this
class.
The first is W the number of configurations. This is called the microcanonical
partition function.
This partition function is not very useful to us so we will not discuss it further.
The second partition function is
Q=

gj eEj

(13.6)

and is called the canonical partition function.

96

This was first encountered as the denominator of the Boltzmann distribution and
it is extremely important in statistical mechanics. (Note: the symbol Z is also
often used for the canonical partition function.)
The partition function is to statistical mechanics as the wavefunction is to quantum mechanics. That is, the partition function contains all that can be known
about the ensemble.
We shall see in the next chapter that the partition function will provide a link
between the microscopic (quantum mechanics or classical mechanics) and the
macroscopic (thermodynamics).
In fact we have already seen this in the S = k ln W. But this an inconvenient
connection because, for among other reasons, energy levels and temperature do
not explicitly appear.
There are other partition functions that are useful in dierent situations but we
will do nothing more than list two important ones here: i) the grand canonical
partition function and ii) the isothermalisobaric partition function
13.1.1. Relation between the Q and W
When we get to connecting quantum mechanics with thermodynamics it will prove
convenient to use Boltzmanns equation (S = k ln W ) but as was stated earlier it
is not convenient to use the microcanonical partition function (W ).
In the following we give an argument which provides a relation between the partition functions. It is not an exact relation as we derive it, but it is a very good
approximation for large numbers of particles.

97

The microcanonical partition function describes a system at fixed energy E. In


fact W is the number of available states of the ensemble at the particular energy
E. This is essentially the same as the degeneracy of the ensemble gE .
Conversely the canonical partition function describes a system with variable energy.
However, based on our previous discussion of fluctuations, even though the energy
of the ensemble is allowed to vary, the number of states with energy equal to the
is overwhelmingly large. That is, almost every state available
average energy E

to the ensemble has energy E.


We can express these ideas mathematically to come up with a relation between
W and Q.
The canonical partition function is
Q=

gj e j ,

(13.7)

but to a good approximation

Q ' gE e E .

(13.8)

Now since the degeneracy is essentially the microcanonical partition function we


have

(13.9)
Q ' W e E .
So the canonical partition function is a Boltzmann weighted version of the microcanonical partition function.
We will soon make use of the Boltmanns equation in terms of the canonical

98

partition function:

ln Q ' ln(W e E ) = ln W + ln(e E )

= ln
W
| {z } kT .

(13.10)

S/k

so,

S = k ln Q +

E
T

(13.11)

13.2. The Molecular Partition Function


We ended the previous chapter by stating the total molecular energy (about the
center of mass) as
= ele + vib + rot .
(13.12)
This is a consequence of the Born Oppenheimer approximation
If we include the center of mass translational motion this is
=

ele

vib

rot

vib,v

rot,J

trans

(13.13)

The ith total energy level is


i

ele,n

trans,m .

(13.14)

Now if we have a collection of molecules in a macroscopic system. A given configuration (say, configuration j) of that system has total energy Ej .
So the canonical partition function is
Q=

gj eEj

(13.15)

99

But, each Ej is made up of the contributions of all of the molecules:


a
l

Ej =

b
m

c
n

(13.16)

The partition function for the molecule is written as


Q =

gj eEj =

|l

{z

l,m,n

gla e

qm o l,a

a
l

}|m

a
gm
e

{z

qm o l,b

b c
(gla gm
gn )e( l + m + n + )
a
m

(13.17)

gna e n

}| n {z

qm o l,c

where the qmol,i are the molecular partition functions.

The total canonical partition function is the product of the molecular partition
functions.
For the case where the molecules are the same then all the qmol,i are the same:
qmol,i = qmol thus
qN
(13.18)
Q = mol .
N!
This allows us to focus only on a single molecule:
qmol =

gi e i =

X
|n

n s,m )
gele,n gvib,v grot,J gtrans,m e ( ele,n + v ib ,v + ro t,J + tra(13.19)

n,v,J,m

gele,n e
{z

qele

ele ,n

}| v

gvib,v e
{z

qv ib

v ib ,v

}| J

grot,J e
{z

qro t

ro t,J

}|m

gtrans,m e
{z

qtra n s

tra n s,m

We now collect below the expression for each of these partition functions. You
will get the chance to derive each of these for your home work

100

The Translational Partition Function


qtrans =

V
3

(13.20)

where

h

2mkT
is the thermal de Broglie wavelength.

(13.21)

The Rotational Partition Function (linear molecules)


We will discuss rotations next semester.
However, the high temperature limit, which works for all gases (of linear molecules)
except H2 is
T
(13.22)
qrot
r
2

where r 8h2 Ik (I is the moment of inertia) and is the so-called symmetry


number in which = 1 for unsymmetrical molecules and = 2 for symmetrical
molecules.
The Vibrational Partition Function
1

qvib

e 2 ~
1
=
=
~
1e
2 sinh 12 ~

(13.23)

Note this is for the harmonic oscillator. At temperatures well below the dissociation energy this is a very good approximation. (You will derive this as a homework
problem.)
The Electronic Partition Function
There is usually only a very few electronic states of interest. Only at exceedingly
high temperatures does any state other that the ground state(s) become important

101

so
qele =

X
i

gele,i e

tele ,i

gele,ground

(13.24)

102

14. Statistical Thermodynamics


The partition function allows one to calculate ensemble averages which correspond
to macroscopically measurable properties such as internal energy, free energy,
entropy etc.
In this chapter we will obtain expressions for internal energy, U, pressure, P,
entropy, S, and Helmholtz free energy, A. With these quantities in hand we will,
in the subsequent chapters, formally develop thermodynamics with no need to
refer back to the partition function.
Ensemble averages
The ensemble average of any property is given by
X
= 1
O
Oi gi e i .
Q i

(14.1)

Internal energy
One critical property of an ensemble is the average (internal) energy U.
1 X
i
.
U E =
i gi e
Q i

(14.2)

Let us look closer at the above expression. Recall that


Q=

gi e i .

(14.3)

103
103

Now take the derivative of Q with respect to gives

#!
"

X e i
X i
Q
=
gi e
=
gi
n,V
i

n,V
i
n,V
X
=
gi i e i

(14.4)

By comparing this to the expression for U, we see

1 Q
ln Q
U =
=
,
Q n,V

n,V
where we used the identity

1 y
y x

(14.5)

ln y
.
x

Pressure
Another important property is pressure.
When the ensemble is in the particular state i, d
temperature and number of particles

i
pi =
V n,

= pi dV . So at constant
(14.6)

Thus the ensemble average pressure is given by

1 X
i
P = p =
gi
e i .
Q i
V n,

(14.7)

Multiplying by / we get
1 X
P =
gi
Q i

i
V

e i .

(14.8)

n,

Using the chain rule in reverse, i.e.,


e

e
V

z }| {

i
e i
i
=
=
e
i
V
V

(14.9)

104

we proceed as
!

i
e
1
1 X
X i
P =
gi
=
gi e
Q i
V
Q V i
n,
n,

Q
1
1 ln Q
=
=
.
Q V n,
V
n,

(14.10)

Entropy
We have already obtained the expression for entropy. It is
U
+ k ln Q
T

ln Q
= k
+ k ln Q

n,V

S =

(14.11)

105

Helmholtz Free Energy


Free energy is the energy contained in the system which is available to do work.
That is, it is the energy of the system minus the energy that is tied-up in the
random (unusable) thermal motion of the particle in the system: A U T S
Free energy is probably the key concept in thermodynamics and so we will discuss
it in much greater detail later. We will make the distinction between the Helmholtz
free energy and the more familiar Gibbs free energy (G) later as well.
The Helmholtz free energy has the most direct relation to the partition function
as can be seen from

ln Q
ln Q
A U TS =
+ kT
kT ln Q (14.12)

n,V
n,V
= kT ln Q
Any thermodynamic property can now be obtained from the above functions as
we shall see in the following chapters.

106

15. Work
We now begin the study of thermodynamics.
Thermodynamics is a theory describing the most general properties of macroscopic
systems at equilibrium and the process of transferring between equilibrium states.
Thermodynamics is completely independent of the microscopic structure of the
system.

15.1. Properties of Partial Derivatives


Of critical importance in mastering thermodynamics is to become proficient with
partial derivatives.
See Handout
15.1.1. Summary of Relations
1. The total derivative of z(x, y):


z
z
dz =
dx +
dy
x y
y x
2. The chain rule for partial derivatives:


u
z
z
=
x y
u y x y

(15.1)

(15.2)

107
107

3. The reciprocal rule:

4. The cyclic rule:

5. Finally

z
x

z
x

z
x

x
z

z
=
y

z
y

z
x

(15.3)

=1

y
x

(15.4)

y
x

(15.5)

15.2. Definitions
System: a collection of particles
Macroscopic systems: Systems containing a large number of particles.
Microscopic systems: Systems containing a small number of particles.
Environment: Everything not included in the system (or set of systems)
Note that the distinction between the system and the environment is arbitrary
and is chosen as a matter of convenience.
15.2.1. Types of Systems
Isolated system: A system that cannot exchange matter or energy with its environment.
Closed system: A system that cannot exchange matter with its environment but
may exchange energy.

108

Open system: A system that may exchange matter and energy with its environment.
Adiabatic system: A closed system that also can not exchange heat energy with
its environment.
15.2.2. System Parameters

Extensive parameters (or properties): properties that depend on the amount of


matter.
For example, volume, mass, heat capacity.
Intensive parameters (or properties): properties that are independent of the
amount of matter.
For example, temperature, pressure, density.
Extensive properties can be converted to intensive properties through ratios:
Extensive property
Intensive property.
Extensive property
For example

mass
volume

= density,

volume
moles

= molar volume,

heat capacity
mass

(15.6)
= specific heat.

15.3. Work and Heat


A system may exchange energy with its environment or another system in the
form of work or heat.
Work is exchanged if external parameters are changed during the process.
Heat is exchanged if only internal parameters are changed during the process.

109

Convention
Work, w, is positive (w > 0) if work is done on the system.
Work is negative (w < 0) if work is done by the system.
Heat, q, is positive (q > 0) if heat is absorbed by the system.
Heat is negative (q < 0) if heat is released from the system.
15.3.1. Generalized Forces and Displacements
In physics you learned that an infinitesimal change in work is given by the product
of force, F , times and infinitesimal change in position, dx:
dw = F dx.

(15.7)

For thermodynamics, we need a more general definition if infinitesimal work.


Any given external parameter, A may be considered as a generalized force which
is coupled to a particular internal parameter, a, which acts as generalized displacement.
Note that the generalized force need not have units of force (e.g., Newtons) and
the generalized displacement need not have units of position (e.g., meters), but
the product of the two must have units of energy (e.g., Joules).
The infinitesimal amount of work done on the system is then given by
dw = Ada,

(15.8)

(15.9)

or more generally as
dw =

Ai dai

110

if more than one set of parameters change.


The following table gives some examples of generalized forces and displacements
Generalized Force, A Generalized Displacement, a Contribution to dw
Pressure, P
Stress,
Surface tension,
Voltage, E
Magnetic Field, H
Chemical Potential,
Gravity, mg

Volume, dV
Strain, d
Surface area, dA
Charge, dQ
Magnetization, dM
Moles, dn
Height, dh

P dV
d
dA
EdQ
HdM
dn
mgdh

15.3.2. P V work
In principle all work is interchangeable so that without loss of generality we will
develop the formal aspects of thermodynamics assuming all work is due to changes
in volume under a given pressure. That is
dw = P dV,

(15.10)

this is called P V work.


When we get to applications of thermodynamics we will then be concerned with
the various forms of work like those shown in the table above.

Expanding Gases
Consider the work done by a gas expanding in piston from volume V1 to V2 against
some constant external pressure P = Pex (see figure)

111

The force exerted on a gas by a piston is equal to the external pressure times the
area of the piston: F = Pex A Pex = F/A.
Rx
Recall from physics that work is the (path) integral over force: w = x12 F dx.
This can be manipulated as
Z x2
Z x2
Z V2
F
w=
F dx =
Adx =
Pex dV
(15.11)
A |{z}
x1
x1 |{z}
V1
dV
Pex

If Pex is independent of V then


Z
Z V2
Pex dV = Pex
w=
V1

V2

V1

dV = Pex 4V

(15.12)

112

16. Maximum Work and Reversible


changes
Now that we have learned about PV work we will consider the situation where
the system does the maximum amount of work possible.

16.1. Maximal Work: Reversible versus Irreversible changes


The value of w depends on Pex during the entire expansion.

In the figure
wA =

V2

V1

Patm dV = Patm (V2 V1 )

(16.1)

113
113

and
wB = w1 + w2 ,
where
w1 =
and

(16.2)

Vi

V1

w2 =

Patm+2W dV = Patm+2W (Vi V1 )

(16.3)

(16.4)

V2

Vi

Patm dV = Patm (V2 Vi )

Hence it is clear that |wB | > |wA | .


Now consider case in the figure below

The expansion is reversible. That is, there is always an intermediate equilibrium


throughout the expansion. Namely Pgas = Pex . So,
wrev =

V2

Pgas dV

(16.5)

V1

This is the limiting case of path B in the previous figure. Thus wrev is the maximum possible work that can be done in an expansion. wrev = wmax .

114

16.2. Heat Capacity


Temperature and heat are dierent.
Temperature is not the amount of heat.
Temperature is an intensive property and heat is an extensive property.
However, heat is related to temperature through the heat capacity
dq
dT
n.b., heat capacity is a function of T ; it is not a constant.
C(T ) =

(16.6)

From this equation


(16.7)

dq = C(t)dT,

That is, when the temperature of a substance having a heat capacity C(t) is
changed by dT, dq amount of heat energy is transferred.
The heat capacity also depends on the conditions during the temperature change,
dq
dq
and CP (T ) = dT
are not the same
e.g., CV (T ) = dT
V
P
Heat capacity is an extensive property. To make an intensive property
1. divide by the number of moles to get molar heat capacity

1 dq
CV m (T ) =
n dT V
2. divide by mass to get specific heat
1
cV =
m

dq
dT

(16.8)

(16.9)

We will discuss heat capacity more later.

115

16.3. Equations of State


The macroscopic properties of matter are related to one another via a phenomenological equation of state.
The state of a pure, homogeneous material (in the absence of external fields) is
given by the values of any two intensive properties.
(More complicated systems require more than two independent variables, but
behave in the same way as the more simple pure system, so we will focus our
development of thermodynamics on simple systems.)
The functional dependence of any property on the two independent variables is
an equation of state. e.g., T , P independent then heat capacity is a function of T
and P , C(T, P ).
16.3.1. Example 1: The Ideal Gas Law
The equation of state for volume of an ideal gas is
P V = nRT ,

(16.10)

where R is the gas constant (8.315 J K1 mol1 ) and n is the number of moles.
The ideal gas equation of state can be expressed in terms of intensive variables
only
P Vm = RT ,
(16.11)
where Vm =

V
.
n

The equation of state can also be expressed in terms of density =


mass m/n)
MP
mP
=
.
=
nRT
RT

m
V

(and molar
(16.12)

116

16.3.2. Example 2: The van der Waals Equation of State


A more realistic equation of state was presented by van der Waals:
P =

nRT
n2 a
2.
V nb
V

(16.13)

The parameter a attempts to account for the attractive forces among the particles
The parameter b attempts to account for the repulsive forces among the particles
b originates from hard sphere collisions (see figure):

117

In term of intensive variables


P =

a
RT
2.
Vm b Vm

(16.14)

16.3.3. Other Equations of State


The van der Waals equation of state is not the only one that has been proposed.
Some other equations of state are
Berthelot

P =

Dieterici

n2 a
a
nRT
RT

=
2
V nb T V
Vm b T Vm2
an

(16.15)

RT e RT Vm
nRT e RT V
=
P =
V nb
Vm b

(16.16)

Redlich-Kwang
P =

nRT
n2 a
a
RT

=
V nb
Vm b
T V (V nb)
T Vm (Vm b)

(16.17)

118

17. The Zeroth and First Laws of


Thermodynamics
Over the course of the next two lectures we will discuss the four core laws of
thermodyanmics.
Today we will cover the zeroth and first laws, which deal with temperature and
total energy respectively.
Next time we will cover the second and third laws which both deal with entropy.

17.1. Temperature and the Zeroth Law of Thermodynamics


Temperature tells us the direction of thermal energy (heat) flow.
Heat flows from high T to Low T.
Temperature scales
Celsius: A relative scale based on water (T = 0 C for melting ice and
T = 100 C for boiling water)
Kelvin: An absolute temperature scale based on the ideal gas law. The
temperature at which (for fixed V and n) the pressure is zero is defined as
T =0K
T (Kelvin) = T (Celsius) + 273.15
119
119

Standard conditions
standard temperature and pressure (STP): T = 273.15 K and P = 1 atm.
(Vm (STP) = 22.414 L/mol)
standard ambient temperature and pressure (SATP): T = 298.15 K and
P = 1 bar. (Vm (SATP) = 24.789 L/mol)

Diathermic wall: A wall that allows heat to flow through it.


Adiabatic wall: A wall the does not allow heat to flow through it.
Thermal equilibrium: If two systems are in contact along a diathermic wall and
no heat flows across the wall, then the systems are in thermal equilibrium.
The zeroth law of thermodynamics
Mathematical statement:
If TA = TB and TB = TC , then TA = TC

(17.1)

This the mathematical statement of transitivity


Verbal statement: If system A is in thermal equilibrium with system B
and system B is in thermal equilibrium system C then system A is also in
thermal equilibrium with system C.

The zeroth law implies that if an arbitrary system, C, is chosen as a thermometer


then it will read the same temperature when it is in thermal contact along a
diathermic wall with system A as when it is in thermal contact along a diathermic
wall with system B.

120

17.2. The First Law of Thermodynamics


Definitions:
State: the state of a system is defined by specifying a minimum number in
intensive variables
State Function: A function of the chosen independent variables that describes a property of the state (e.g., V (T, P )). The value of the state function depends only on that given state and on no other possible state of the
system.
17.2.1. The internal energy state function
For characterizing the change in energy of a system, one is concerned with the
work done on the system (w) and the heat supplied to the system (q). The energy
of a system is called the internal energy (U) of the system.
The first law of thermodynamics:
Mathematical statement:

4U = q + w

(17.2)

dU = dq + dw

(17.3)

or in dierential form

Verbal statement: The change in internal energy of a system is equal to the


amount of work done on the system plus the amount of heat provided to the
system.

So for a system where all the work is P V work the first law becomes
Z V2
4U = q
Pex dV

(17.4)

V1

121

in dierential form this is


dU = dq Pex dV

(17.5)

Although U can be expressed as a function of any two state variables, the most
convenient at this time are V and T. U U(T, V ).
The total dierential of U (T, V ) is

U
U
dU =
dT +
dV
T V
V T
Consider adding heat at a constant volume then

U
U
dT +
dV = dq Pex dV.
dU =
T V
V T
So,

U
U
dq
= CV
dT = dq =
=
T V
T V
dT

is the heat capacity.
Hence the slope U
T V
The other slope,

U
V

(17.6)

(17.7)

(17.8)

, is called the internal pressure (it has no standard symbol).

A useful relation (derivation to come) is

U
P
=T
P
V T
T V

(17.9)

Example: A van der Waals gas


n2 a
nRT
2
P =
V nb
V

P
T

nR
V nb

(17.10)

122

so the useful relation becomes

nRT
nRT
n2 a
nR
U
P =

+ 2
= T
V T
V nb
V nb V nb
V
2
na
= + 2
V

(17.11)

The equation of state for U : Express U in terms of T, V, and P.


Start with the total dierential of U

U
U
dT +
dV
dU =
T V
V T
U


= CV and V
= T P
P (useful relation). Hence
but U
T V
T V
T

P
P dV
dU = CV dT + T
T V

(17.12)

(17.13)

is the equation of state for U.


A useful approximation is 4U = CV 4T which is valid for
i) heat capacity nearly constant over 4T and with no phase transitions.
ii) ideal gas or at constant volume.

123

18. The Second and Third Laws of


Thermodynamics
18.1. Entropy and the Second Law of Thermodynamics
We learned from statistical mechanics that entropy, S, is a measure of the disorder
of the system and is expressed via Boltzmanns equation S = k ln W (where W is
the micocanonical partition function)
We expressed Boltzmanns law in terms of the more convenient canonical partition
function as

E
S = + k ln Q.
(18.1)
T
Now, the average energy of the system E is in fact what we call internal energy:

U E.
Furthermore we derived the simple relation between the Helmholtz free energy
and the canonical partition function as A = kT ln Q.
Hence,
U
A
1
= (U A).
T
T
T
Since U, A, and T are state functions, S is also a state function .
S=

So we may write
dS =

1
(dU dA)
T

(18.2)

(18.3)

124
124

for an isothermal process.


Recall the definition of Helmholtz free energythe energy of the system available
to do work.
We learned previously that the maximum amount of work one can extract from
the system is the work done during a reversible process. Hence dA = dwrev .
For now let us limit the discussion to reversible processes. Then
1
1
(dU dwrev ) = (dqrev + dw
/ rev dw
/ rev )
T
T
dqrev
. (Reversible process)
=
T

dS =

(18.4)

Note: An alternative approach to thermodynamics which makes no reference to


molecules or statistical mechanics is to simply begin by defining entropy as dS
dqrev
T

The principle of Clausius


The entropy of an isolated system will always increase in a spontaneous
process
Mathematical statement: (dS)U,V 0
For a general process: dU = dq Pex dV
For a reversible process Pex = P and dq = T dS so dU = T dS P dV

125

Since U, S, T, P, and V are state functions, dU = T dS P dV holds for any


process, but in general, T dS is not heat and P dV is not work. (see figure)
T dS is heat and P dV is work only for reversible processes.

For some dU,


dq Pex dV = T dS P dV T dS = dq Pex dV + P dV

(18.5)

T dS = dq + (P Pex ) dV
Case i) Pex > P then (spontaneous) dV is negative so (P Pex )dV is positive.
Case ii) P > Pex then (spontaneous) dV is positive so (P Pex )dV is positive.
Case iii) P = Pex then (spontaneous) dV is zero so (P Pex )dV is zero.
Thus for any spontaneous process T dS dq.
This is a mathematical statement of the second law of thermodynamics

126

18.1.1. Statements of the Second Law


Unlike the first law, the second law has a number of equivalent statements
1. A cyclic process must transfer heat from a hot to cold reservoir if it is to
convert heat into work.
2. Work must be done to transfer heat from a cold to a hot reservoir.
3. A useful perpetual motion machine does not exist.
4. The entropy of the universe is increasing
5. Spontaneous processes are irreversible in character.
6. The entropy of an isolated system will always increase in a spontaneous
process (the principle of Clausius)

18.2. The Third Law of Thermodynamics


Consider the first law for a reversible change at constant volume.
dU = dq + dw = dq Pex dV

(18.6)

From our earlier discussion of heat capacity dq = CV dT (CV since constant volume). So,
(18.7)

dU = CV dT
but also dU = T dS. So
CV dT
= 4S =
dS =
T

T2

T1

CV
dT.
T

(18.8)

127

A very similar derivation can be done for a reversible change at constant pressure
(we can not do it quite yet) to yield
Z T2
CP
dT
(18.9)
4S =
T
T1
18.2.1. The Third Law
Verbal statement
The third law of thermodynamics permits the absolute measurement of entropy.
To derive the mathematical statement of the third laws we starting with
Z T2
CP
dT
(18.10)
4S =
T
T1
now let T1 0
4S = S2 S0 =

T2

CP
dT
T

Hence the mathematical statement of the third law is


Z T2
CP
S(T2 ) =
dT + S0
T
0

(18.11)

(18.12)

From a macroscopic point of view S0 is arbitrary. However, a microscopic point of


view suggests S0 = 0 for perfect crystals of atoms or of totally symmetric molecules
(e.g., Ar, O2 etc.). S0 6= 0 for imperfect crystals and crystals of asymmetric
molecules (e.g., CO).
Alternative statement of the third law: Absolute zero is unattainable.
Consider the heat capacity near T 0.
For S0 to have significance

CP
T

must be finite (not infinite) as T 0. Thus CP 0.

128

But CP =

dq
dT

0 implies

dT
dq

In other words, an infinitesimal amount of heat causes an infinite change in temperature.

In view of what we have learned about fluctuations, the ever present random
fluctuations in energy provide the infinitesimal amount of heat and so you can
never reach absolute zero corresponding to an average energy of zero.

18.2.2. Debyes Law


Heat capacity data only goes down so far. So one needs a theoretical extrapolation
down to T = 0. (Debye)
Postulate: CP m = aT 3 . That is at low temperatures heat capacity goes as the
cube of the temperature.
CP m , T are the lowest temperature data points. So, a = CP m /T 3 .

129

The molar entropy is

Sm (T ) =
=

CP
C
=aT 3 CP m
dT P m=
T
T 3
0

T
CP m
CP m T 3
=
.
T 3 3
3

T 2 dT

(18.13)

18.3. Times Arrow

Entropy and the second law give a direction to time.


For example, if we see a picture of your PChem book in mint condition and we see
a picture of your PChem book all battered and beaten. We know which picture
was taken first.
The interesting thing is that each molecule in a macroscopic system obeys time
invariant dynamics. Both Newtons laws and Quantum dynamics (next semester)
are the same if you replace t with t.
Yet, the behavior of the macrosystem definitely changes if you replace t with t.
Thus the simple fact that you have an enormous number of particles induces a
perceived asymmetry in time.

130

Key Equations for Exam 3


Listed here are some of the key equations for Exam 3. This section should not
substitute for your studying of the rest of this material.
The equations listed here are out of context and it would help you very little to
memorize this section without understanding the context of these equations.
The equations are collected here simply for handy reference for you while working
the problem sets.

Equations
The Boltzmann equation is

S = k ln W.

(18.14)

The Boltzmann distribution :


g e i
P i
j gj e

(18.15)

The canonical partition function is


Q=

gj eEj

(18.16)

131
131

The relation between the partition function and the molecular partition
function is
qN
Q = mol .
(18.17)
N!
The Translational Partition Function
V
qtrans = 3
(18.18)

where
h

(18.19)
2mkT
is the thermal de Broglie wavelength.
The Rotational Partition Function (linear molecules) is
T
qrot
,
r

(18.20)

where r 8h2 Ik (I is the moment of inertia) and is the so-called symmetry number in which = 1 for unsymmetrical molecules and = 2 for
symmetrical molecules
The Vibrational Partition Function is
qvib =

1
.
2 sinh 12 ~

(18.21)

The ensemble average of any property is given by


X
= 1
Oi gi e i .
O
Q i

(18.22)

The relations between the canonical partition function and the thermodynamics variables are
Helmholtz Free Energy A = kTln Q

Internal energy

U = Q1

Entropy

Pressure

n,V

ln Q
= k
n,V
Q
1
= Q V n, = 1

ln Q

+ k ln Q
ln Q
V

n,V

n,

132

P V work is

dw = P dV.

(18.23)

dq = C(t)dT.

(18.24)

4U = q + w,

(18.25)

dU = dq Pex dV.

(18.26)

dU = T dS P dV.

(18.27)

T dS dq.

(18.28)

Heat capacity:
General forms of the first law:

in dierential form this is

Also,

The second law


The third law

S(T2 ) =

T2

CP
dT + S0
T

(18.29)

Debyes law for entropy at very low temperatures


Sm (T ) =

CP m
,
3

(18.30)

where CP m is the molar heat capacity at the lowest temperature for which
there is data.

133

Part IV
Basics of Thermodynamics

134
134

19. Auxillary Functions and Maxwell


Relations
We have stated that thermodynamics as we are studying it deals with states in
equilibrium or transitions between equilibrium states.
Consequently, the concept of equilibrium plays a key role in much of what we will
discuss for the remainder of the year.
The equilibrium constant for a thermodynamic process, K, (which you are familiar
with from general chemistry) serves are a common point which connects thermodynamics, electrochemistry, and kineticstopics we will encounter throughout
the year.

19.1. The Other Important State Functions of Thermodynamics


As was the case in quantum mechanics, here too is energy the key property with
which to work.
So far we have encountered two state functions which characterize the energy of a
macroscopic systemthe internal energy and, briefly the Helmholtz free energy.

135
135

From the first law as stated as


dU = T dS P dV

(19.1)

we say that the natural (most convenient) variables for the equation of state for
U are S and V . This is U = U(S, V )
Unfortunately S can not be directly measured and most often P is a more convenient variable than V
Because of this fact, it is handy to define state functions which have dierent pairs
of natural variables, so that no mater what situation arises we have convenient
equations of state to work with.
The other pairs of natural variables being (S and P ), (T and V ) and (T and P )
The table below lists these state functions
State function
Internal Energy
Enthalpy
Helmholtz free energy
Gibbs free energy

Symbol Natural variables


U
H
A
G

S
S
T
T

and V
and P
and V
and P

Definition
H U + PV
A U TS
G H TS

Units
energy
energy
energy
energy

We consider each of these functions in turn

19.2. Enthalpy
We want a state function whose natural variables are S and P
Let us try the definition H U + P V.

136

Now formally
dH = dU + d(P V ) = dU + P dV + V dP,

(19.2)

but dU = T dS P dV, so

/ + P dV/ + V dP
dH = T dS P dV

(19.3)

= T dS + V dP.
Hence Enthalpy does indeed have the desired natural variables.
19.2.1. Heuristic definition:
Enthalpy is the total energy of the system minus the pressure volume energy. So
a change in enthalpy is the change in internal energy adjusted for the P V work
done. If the process occurs at constant pressure then the enthalpy change is the
heat given o or taken in.
For example, consider an reversibly expanding gas under constant pressure (dP =
0) and adiabatic (dq = 0) conditions.
The system does work during the expansion; in doing so it must lose energy. Since
the process is adiabatic no heat energy can flow in to compensate for the work
done and the gas cools.
The total internal energy decreases. The enthalpy of the system on the other hand
does not changeit is the internal energy adjusted by an amount of energy equal
to the P V work done by the system. As Freshmen we learn this as 4H = qp .

19.3. Helmholtz Free Energy


Now we want a state function whose natural variables are T and V

137

Let us try the definition A U T S.


Formally
dA = dU d(T S) = dU T dS SdT,

(19.4)

but dU = T dS P dV, so

/ P dV T dS
/ SdT
dA = T dS

(19.5)

= P dV SdT.
Hence Helmholtz free energy does indeed have the desired natural variables.
19.3.1. Heuristic definition:
As we have said before Helmholtz free energy is the energy of the system which is
available to do workIt is the internal energy minus that energy which is used
up by the random thermal motion of the molecules.

19.4. Gibbs Free Energy


Finally we want a state function whose natural variables are T and P
Let us try the definition G H T S.
Now formally
dG = dH + d(T S) = dH T dS SdT,

(19.6)

but from above dH = T dS + V dP, so


dG = T dS
/ + V dP T dS
/ SdT

(19.7)

= V dP SdT.
Hence Gibbs free energy does indeed have the desired natural variables.

138

19.4.1. Heuristic definition:


Gibbs free energy is the energy of the system which is available to do non P V
workIt is the internal minus both that energy which is used up by the random
thermal motion of the molecules and used up in doing the P V work.

19.5. Heat Capacity of Gases


19.5.1. The Relationship Between CP and CV
To find how CP and CV are related we begin with
(19.8)

dH = T dS + V dP
at constant pressure and reversible conditions

(19.9)

dH = T dS
dH = dq
but

(19.10)

dq = CP dT

The constant pressure heat capcity can then be expressed in terms of enthalpy as

H
.
(19.11)
CP =
T P
So,

(U + P V )
U
V
CP =
=
+P
(19.12)
T
T P
T P
P
U

is
not
C
we
need
. Use an identity of partial derivatives
note U
V
T P
T V

U
T

U
T

U
V

V
T

(19.13)

139

thus

V
U
V
+
+P
(19.14)
CP =
V T T P
T P
V

U
V
= CV +
+P .
T P
V T
P
U
=
T
P . Then
Recall the expression for internal pressure V
T V
T

U
T

CP = CV +

V
T

P
T
P
/ + P/
T V
P

Finally
CP = CV + T

V
T

P
T

(19.15)

(19.16)

Example: Ideal gases


1. Ideal gas (equation of state: P V = nRT ): This equation is easily made
explicit in either P or V so we dont need any of the above replacements

P
V
CP = CV + T
(19.17)
T P T V
nRT
nR nR
=
nR
= CV + T
P V
PV
Thus CP = CV + nR or
CP m = CV m + R

(19.18)

19.6. The Maxwell Relations


Summary of thermodynamic relations weve seen so far
Definitions and relations:
H = U + PV

140

A = U TS
G = H TS
CV =

U
T

basic equations
dU = T dS P dV
dH = T dS + V dP
dA = P dV SdT
dG = V dP SdT

CP =

H
T

Maxwell relations
T

= P
V S
T
VS V
=
SP S
SP P
= + T V
V
T
V
S
=

P T
T P

working equations

dU = CV dT + T P
P dV
T V

dH = CP dT T V
V dP
T P

dS = CTV dT + P
dV
T V

V
CP
dS = T dT T P dP

We will get plenty of practice with derivations based on these equations and on
the properties of partial derivatives. (See handout and Homework)

141

20. Chemical Potential


20.1. Spontaneity of processes
Two factors drive spontaneous processes
1. The tendency to minimize energy
2. The tendency to maximize entropy

Let us begin with Helmholtz free energy


The total dierential of A is (A = U T S)
dA = dU T dS SdT = dq Pex dV T dS SdT

(20.1)

For constant T and V, (dA)T,V = dq T dS


From the second law, T dS dq for a spontaneous process, (dA)T,V 0 for a
spontaneous process.
Hence at equilibrium (dA)T,V = 0.
For chemistry it is most often more convenient to use Gibbs free energy
The total dierential of G is
dG = dH T dS SdT = dq Pex dV + P dV + V dP T dS SdT

(20.2)

142
142

For constant T and P = Pex , (dG)T,P = dq T dS


Again from the second law, T dS dq for a spontaneous process, (dG)T,P 0 for
a spontaneous process.
Hence at equilibrium (dG)T,P = 0.
So free energy provides a measure of the thermodynamic driving force towards
equilibrium.
Note free energy provides no information about how fast a process proceeds to
equilibrium.
The free energy functions are the workhorses of applied thermodynamics so we
want to get a feel for them.
Returning to the total dierentials of free energy,
dA = dU T dS SdT

(20.3)

dG = dH T dS SdT.

(20.4)

and
Expressing dU and dH generally as dU = T dS P dV and dH = T dS + V dP
(remember that in general T dS cannot be identified with dq and P dV cannot be
identified with w).
Plugging these into the total dierentials of free energy gives
dA = SdT P dV

(20.5)

dG = SdT + V dP

(20.6)

and

143

These expressions are quite general, but i) only P V work and ii) closed systems.
The total dierential of A is also
dA = dq + dw T dS SdT.

(20.7)

For a reversible process dq = T dS and work is maximal.


Hence (dA)T = dwmax = (4A)T = wmax . As we have stated in words a number
of times before.
The total dierential of G is also
dG = dq + dw + P dV + V dP T dS SdT.

(20.8)

In general dw = dw0 Pex dV where dw0 is the non-P V work.


The total dierential of G becomes
dG = dq + dw0 Pex dV + P dV + V dP T dS SdT.

(20.9)

For constant T and P = Pex , (dG)T,P = dq + w0 T dS.


For reversible processes q = T dS and this becomes
0
0
(dG)T,P = dwmax
= (4G)T,P = wmax

(20.10)

So, as stated earlier, the Gibbs free energy is the energy of the system available
to do non-P V work.

20.2. Chemical potential


What if the amount of substance can change?

144

Extensive properties depend on the amount of stu


For example A(T, V ) now becomes A(T, V, n) and the total dierential becomes


A
A
A
dA(T, V, n) =
dT +
dV +
dn
(20.11)
T V,n
V T,n
n V,T
Lets focus on the slope

n V,T

This is a measure of the change in Helmholtz free energy of a system (at


constant T and V ) with the change in the amount of material.
Physically, this is a measure of the potential to change the amount of material.
It defines the chemical potential

n V,T

So we can also write


dA = SdT P dV + dn

(20.12)

What about the relation of the chemical potential to Gibbs free energy?
G = H TS = U
|
{zT S} + P V = A + P V so,
=A

(20.13)

dG = dA + P dV + V dP

= SdT P
/ dV/ + P/ dV/ + V dP + dn
= SdT + V dP + dn,
but from
dG =

G
T

P,n

dT +

G
P

T,n

dP +

G
n

dn

(20.14)

P,T

145

we see that
=

G
n

(20.15)

P,T

So, is also a measure of the change in Gibbs free energy of a system (at constant
T and P ) with the change in the amount of material and it still has the same
physical meaning.
The Gibbs free energy per mole (Gm ) for a pure substance is equal to the chemical
potential. (Gm = )

20.3. Activity and the Activity coecient


When, for example, a solute is dissolved in a solvent, there exist complicated
interactions which cause deviations from ideal behavior.
To account for this one must introduce the concept of activity and the activity
coecient.
Activity is hard to define in words and indeed it has an awkward mathematical
definition as we will soon see.
The activity coecient has a more convenient definition which is that it is the
measure of how a particular real system deviates from some reference system
which is usually taken to be ideal.
The mathematical definition of activity ai of some species i is implicitly stated as
lim

ai
=1
g()

(20.16)

where g() is any reference function (e.g., pressure, mole fraction, concentration
etc.), and is the value of at the reference state.

146

This implicit definition is awkward so for convenience one defines the activity
coecient as the argument of the above limit,
i

ai
g()

(20.17)

which we can rearrange as


ai = i g().

(20.18)

The definition of activity implies that i = 1 at g( ) (the reference state)


20.3.1. Reference States
Thermodynamics is founded on the concept of energy which we know to have an
arbitrary scale. That is, we can define are zero of energy any where we want.
Because of this it is always necessary to specify a reference state to which our real
state can be compared.
The choice of this state is completely up to us, but it is often the case that the
reference state is chosen to be some ideal state.
For example, if we are talking about a gas we will mostly likely choose the ideal
gas law in terms of pressure (P = nRT /V ) as our reference function and the
reference state being when P = 0 since we know all gases behave ideally in the
limit of zero pressure.
Let us consider the activity of a real gas for the above reference function and
reference state. Note: the activity of gases as referenced to pressure has the
special name fugacity (fugacity is a special case of activity).

147

Our reference function is very simple: g() = = P , so


=

a
a = P.
P

(20.19)

Thus the activity of our real gas is given by the activity coecient times the
pressure of an ideal gas under the same conditions.
Based on the condition that 1 as we approach the reference state (P = 0
in this case) we see that the activity (or fugacity) of a real gas becomes equal to
pressure for low pressures

20.3.2. Activity and the Chemical Potential


One cannot measure absolute chemical potentials, only relative potentials can be
measured. By convention we chose a standard state and measure relative to that
state.
The deviation of the chemical potential at the state of interest versus at the
reference state is determined by the activity at the current state (the activity at
the reference state is unity by definition).
i
i = RT ln ai .

(20.20)

Rather than referencing to the standard state one can also reference to any convenient ideal state. This ideal state is in turn referenced to the standard state.
For the state of interest
i =
(20.21)
i + RT ln ai
and for the ideal state

id

id
id
id
i = i + RT ln ai i = i RT ln ai .

(20.22)

148

Thus,
id
i = id
i RT ln ai + RT ln ai

(20.23)

i id
= +RT ln ai RT ln aid
i
i
ai
= RT ln id
ai

Example: Real and ideal gases at constant temperature, but any pressure.
Starting from the begining
=0

id

z}|{
= dGm = Sm dT + Vm dP

(20.24)

did = Vm dP
RT
dP.
did =
P

Now we integrate from the reference state to the current state of interest
Z
Z
RT
id
dP.
(20.25)
d =

P P
This gives
P
.
P
The usual standard state is the ideal gas at P = 1, so
id = RT ln

id = + RT ln P.

(20.26)

(20.27)

(Note that as P 1, id ).
Lets say our gas is not ideal, then at a given pressure
= + RT ln a.

(20.28)

For gases activity is usually called fugacity and given the symbol f , so a = f for
real gasses. Thus
= + RT ln f.
(20.29)

149

Lets say that instead of referencing to the ideal gas at P = 1, we want to reference
to the ideal gas at the current pressure P.
This is easily done by using = id RT ln P in the above equation for ,
= id RT ln P + RT ln f
f
= id + RT ln .
P
Example: The barometric equation for an ideal gas.
We have an ideal gas so,
id = + RT ln P

(20.30)

where we will take the reference state to be at sea level, i.e. P = 1 atm.
So at sea level

=0

z }| {
(0) = + RT ln 1 =

(20.31)

id (h) = + RT ln Ph

(20.32)

id

and at elevation h

The gas fields the gravitational force which gives it a potential energy per mole
of Mgh at height h. We add this energy per mole term to the chemical potential
(which is free energy per mole) thus at equilibrium
id (0) = id (h) + Mgh

(20.33)

Referencing to the reference state we get

= /
+ RT ln Ph + Mgh
/

(20.34)

RT ln Ph = Mgh
Ph = e

M gh
RT

The last line is the barometric equation and it shows that pressure is exponentially
decreasing function of altitude.

150

21. Equilibrium
First let us consider the equilibrium A B.
Since A and B are in equilibrium their chemical potentials must be equal
A = B

(21.1)

A =
A + RT ln aA

(21.2)

Now,

and
B =
B + RT ln aB
So the equilibrium condition becomes

A + RT ln aA = B + RT ln aB

(21.3)

4 =
A B = RT ln aB RT ln aA
aB
4 = RT ln
aA

Since chemical potential is free energy per mole, if we multiply the above by n
moles we have
aB
4G = nRT ln
aA
as a consequence of the equilibrium condition.
The quantity aaBA defines the equilibrium constant, Ka , for this process.

151
151

Say the system A B is not in equilibrium then we can not write


A = B
but we can write

z }| {
A + B A = B

Proceeding as above we get

A + RT ln aA + 4 = B + RT ln aB

(21.4)

(21.5)

4 =
B A + RT ln aB +RT ln aA
aB
4 = 4 + RT ln .
aA

Again multiplying by n gives


4G = 4G + nRT ln

aB
.
aA

If the 4G < 0 then the transition A B proceeds spontaneously as written.


Consider a more complicated equilibrium
aA + bB cC + dD.

(21.6)

aA + bB = cC + dD .

(21.7)

The equilibrium condition is

In a manner similar to the above

a
A + aRT ln aA + bB + bRT ln aB = cC + cRT ln aC + dD + dRT ln aD (21.8)

Rearranging gives
4rx n G

z
}|
{
acC adD

+
b

d
=
RT
ln
a
A
B
C
D
aaA abB

(21.9)

152

the equilibrium constant is


acC adD
Ka = a b = 4G = RT ln Ka
aA aB

(21.10)

Note: n is absent in the above since the molar values are implied by the stoichiometry.
21.0.3. Equilibrium constants in terms of KC
Equilibrium constant in terms of a condensed phase concentration:
KC0 =

[C]c [D]d
[A]a [B]b

which is related to Ka by
Ka =

KC0

cC dD
aA bB

(21.11)

(21.12)

If the reactants are solutes then as the solution is diluted all the activity coecients
go to unity and KC0 Ka .

21.0.4. The Partition Coecient


Up to now we have only considered miscible solutions.
We now consider the problem of determining the equilibrium concentrations of a
solute A in both phases of an immiscible mixture.

153

The equilibrium equation is


A A

(21.13)

The equilibrium expression for this process is


4G = 0 = 4G
nRT ln Ka ,

(21.14)

where, 4G
G G . The equilibrium constant for this process has a special
/
name; it is called the partition coecient, P / Kpart , for species A in the
mixture.

We can solve for the partition coecient to yield


P

4G

a
nRT
= A
=
e
.

aA

For low concentrations


P / '

[A]
.
[A]

(21.15)

(21.16)

Knowledge of the partition function is important on the delivery of drugs because,


to enter the body, the drugs must transfer between an aqueous phase and a oil
phase.

154

For most drugs


o/w

0 < Ppart < 4

Partition coecient

(21.17)

Delivery mechanism

o/w

low Ppart (likes water)


injection
o/w
medium Ppart
oral
o/w
high Ppart (likes oil) skin patch/ointment

Factors other than the partition coecient influence the drug delivery choice. For
example, can the drug handle the acidic environment of the stomach?

155

22. Chemical Reactions


Up to now we have only been considering systems in the absence of chemical
reactions. After chemical reactions take place the system is in a final product
thermodynamic state that is in general dierent from the initial reactant state.
For any extensive property
4rxn (Property) = property of products property of reactants
Example
Reaction: aA+bB= cC+dD
4rxn S = cSm,C + dSm,D aSm,A bSm,B

22.1. Heats of Reactions


Exothermic reaction: heat is given o to the surroundings
Endothermic reaction: heat is given taken in from the surroundings
At constant pressure (Pex = P
q = 4rxn U w = 4rxn U P 4rxn V = 4rxn H

(22.1)

4rxn H < 0 for Exothermic reactions.


4rxn H > 0 for Endothermic reactions.
156
156

22.1.1. Heats of Formation


Hesss Law of heat summation: 4rxn H is independent of chemical pathway
Example: C2 H2 +H2 = C2 H4 .
This direct reaction is not easy but it can be done in steps
C2 H2 + 52 O2 2CO2 +H2 O(liq) 4rxn H = 1299.63 kJ
2CO2 +2H2 O(liq)C2 H4 + 3O2 4rxn H = +1410.97 kJ
H2 + 12 O2 H2 O(liq)
4rxn H = 285.83 kJ
C2 H2 +H2 = C2 H4

4rxn H = 174.49 kJ

The heat of formation 4f H is the 4rxn H at STP in forming a compound from


its constituent atoms in their natural states.
O2 , H2 , C(graphite) are examples of atoms in their natural state.
Example: Formation of water
H2 + 12 O2 = H2 O not 2H2 +O2 = 2H2 O
4rxn H =
ponent.

i 4f H(i), where i is the stoichiometric factor of the ith com-

Example: H2 O(liq)H2 O(gas) at SATP


H2 + 12 O2 = H2 O(gas)
4f H = 241.818 kJ
H2 + 12 O2 = H2 O(liq)
4f H = 285.830 kJ
H2 O(liq)H2 O(gas)

4rxn H = 241.818 (285.830) = 44.012 kJ

22.1.2. Temperature dependence of the heat of reaction


Z T2
4rxn CP dT
4rxn H(T2 ) = 4rxn H(T1 ) +

(22.2)

T1

157

22.2. Reversible reactions


Recall the requirement for a spontaneous change: 4G < 0 for constant T and P.
4rxn G = G(products) G(reactants) =

i i ,

(22.3)

(remember i = Gm,i for pure substance i).


As we saw before i can be defined in terms of activity
i =
i + RT ln ai .

(22.4)

So,
4rx n G

zX}| {
X
i
+
RT
i ln ai .
4rxn G =
i
i

(22.5)

Using the property of logarithms: a ln x + b ln y = ln(xa y b ) the above expression


becomes
4rxn G = 4rxn G + RT ln

ai i

(22.6)

4rxn G = 4rxn G + RT ln Q,
where Q

ai i is the activity quotient.

At equilibrium, 4rxn G = 0 and Q = Ka (Thermodynamic equilibrium constant).


Ka depends on T but is independent of P.
For the reaction aA + bB = cC + dD
Ka =

acC adD
aaA abB

(22.7)

Note that the activity of any pure solid or liquid is for all practical purposes
equal to 1.

158

For ideal gases, ai =


useful relation
KP =

Pi
P

Xi P
P

(P = 1 bar) This leads to the sometimes

c+dab
PCc PDd
(P aC )c (P aD )d
=
=
K
,
P
a
PAa PBb
(P aA )a (P aB )b
S

or more generally KP = Ka (P )

(22.8)

So at equilibrium, 4rxn G = 4rxn G + RT ln Q becomes


0 = 4rxn G + RT ln Ka 4rxn G = RT ln Ka .

(22.9)

22.3. Temperature Dependence of Ka


Starting with
G =H T S or G/T = H/T S.
)
From this (G/T
= H.
(1/T )
P
Applying this to
4rxn H
4rxn G
=
4rxn S
T
T
gives

(4rxn G /T )
= 4rxn H
(1/T )
P

Using 4rxn G = RT ln Ka , we get

4rxn H
ln Ka
ind. d ln Ka
=
=
(1/T ) P of P d(1/T )
R
or (using

d
d(1/T )

dT
d
d(1/T ) dT

(22.10)

(22.11)

(22.12)

d
= T 2 dT
)

4rxn H
d ln Ka
=
dT
RT 2

(22.13)

Integration gives
1
ln Ka (T2 ) = ln Ka (T1 ) +
R

T2

T1

4rxn Hm
T2

For a reasonably small range T2 T1 this is well approximated by

1
4rxn Hm
1

ln Ka (T2 ) = ln Ka (T1 )
R
T2 T1

(22.14)

(22.15)

159

22.4. Extent of Reaction


There are other equilibrium constants that are used in the literature.
From Pj = Xj P , KX = KP P 4g
V
From nj = Pj RT
(ideal gas approximation), Kn = KP

From concentration Cj =

nj
V

Pj
,
RT

Equilibrium constants
constants
expression
KP

activity(products)
activity(reactants)
partial pressure(products)
partial pressure(reactants)

KX

mole fraction(products)
mole fraction(reactants)

Kn

moles(products)
moles(reactants)

Ka

KC

concentration(products)
concentration(reactants)

RT 4g
V

KC = KP (RT )4g

relation to Ka

situation used

when an exact answer is needed

Ka
4 g
K P

Ka
4 g
K P

Ka
4 g
K P

Ka
4 g
K P

gas reactions
4 g

RT 4g
V

(RT )4g

when eq. P is known


when V is known and constant
when concentration known

160

23. Ionics
Many chemical processes involve electrolytes and or acids and bases.
To understand these processes we must know something about how ions behave
in solution.

23.1. Ionic Activities


Consider a salt in solution
Mv+ Xv v+ M z+ (aq) + v X z (aq),

(23.1)

where v+ (v ) is the number of cations (anions) and z+ (z ) is the charge on the


cation (anion).
The chemical potential for the salt may be written in terms of the chemical potential for each of the ions:
salt = v+ + + v

(23.2)

To determine the activity we start with


j
j
,
ln aj =
RT
and
ln asalt =

j = + or

salt
salt
.
RT

(23.3)

(23.4)

161
161

Substituting the expression for salt into this gives


ln asalt

v+ + v + v+
+ v
=
RT
v v
v+ + v+
+

+
=
RT
RT
|
{z
} |
{z
}
v+ ln a+

So,

(23.5)

v ln a

ln asalt = v+ ln a+ + v ln a

(23.6)

asalt = av+ av

(23.7)

or, alternatively,

It is the case that 1 mole of salt behaves like v = v+ + v moles of nonelectrolytes


in terms of the colligative properties. This suggests that the interesting quantity
is salt
:
v

salt
1/v
= salt + RT ln asalt .
(23.8)
v
v
We see that
1/v
asalt = (av+ av )1/v a .
(23.9)
The quantity a is the mean ionic activity.
23.1.1. Ionic activity coecients
The activity coecients for ionic solutions can also be defined via
a+ = + m+ , a = m ,

(23.10)

where m+ = v+ m and m = v m.
The mean ionic activity coecient is
v

1/v

= ( ++ )

(23.11)

162

The quantities a+ , a , + and cannot be measured individually.


One can use the colligative properties to measure the ionic activity coecients.
It is convenient to redefine the osmotic coecient as
=

1000 g/kg
ln a1 ,
vmM1

(23.12)

where the subscript 1 refers to the solvent.


Similarly freezing point depression is redefined as
= vKf m.

(23.13)

So, v corresponds to the empirical factor i discussed earlier.


Recall how was calculated from the Gibbs-Duhem equation:
Z m
j
ln = j
dm0 ,
0
m
0

(23.14)

where j = 1 .

23.2. Theory of Electrolytic Solutions


Ionic strength is defined as
I=

1X 2
z mi ,
2 i i

(23.15)

where z is the charge of the ion and m its concentration.


Results from DebyeHckel theory: point charge in a continuum
The DebyeHckel equation:

|z+ z | I
,
ln =
1 Ba0 I

(23.16)

163

where
=

e3
(kT )3/2

2 L
1000

1/2

(23.17)

8Le2
,
(23.18)
1000kT
a0 is the radius of closest approach, e is the charge on the electron, is the
density of the pure solvent, is the dielectric constant for the pure solvent and L
is Avogadros number.
B=

Notice that the parameters and B depend only on the solvent.


One important approximation to this equation is to neglect the B term to get the
DebyeHukel limiting Law (DHLL):

ln = |z+ z | I.
(23.19)
This gives the dependence of ln for dilute solutions (m 0). It is seen that

the DHLL correctly predicts the m dependence of ln , which is observed exP


perimentally (recall I = 12 i zi2 mi ).
A useful empirical approximation is to set Ba0 = 1 and to add an empirical
correction to get the :

2
v+ + v
|z+ z | I

ln =
+ 2m
.
(23.20)
v+ + v
1 I
This equation works well to ionic strengths of about I = 0.1

23.3. Ion Mobility


Current, I is given by the rate of change (in time) of charge, Q:
I=

dQ
dt

(23.21)

164

(Electrical) work, w, is required to move a change through a potential (or voltage),


:
w = Q
(23.22)
Power is given by the product of the voltage and the current:
p = I

(23.23)

Resistance is given by the ratio of the voltage to current:


R=

Conductance is the inverse of the resistance (R1 ).


Some relevant constants
charge of an electron e = 1.602177 1019 C.
Faradays constant F = Le = 96485 C/mol (Avogadros number of electrons)
23.3.1. Ion mobility

165

The total current passing through an ionic solution is determined by the sum of
the current carried by the cations and by the anions

Now
Ii =

I = I+ + I

(23.24)

dQi
dNi
= |zi | e
,
dt
dt

(23.25)

where i = +, .
For uniform ion velocity (vi ) the number of ions arriving at the electrode during
any given time interval 4t is
4Ni =

Ni
Ni
dNi
Avi 4t =
=
Avi
V
dt
V

so
Ii = |zi | e

(23.26)

Ni
Avi
V

(23.27)

(in vacuum)

(23.28)

dvi
= zi eE.
dt

(23.29)

Recall Coulombs law


Fi = zi eE,
where E is the electric field, E =
Also recall Newtons law

d
.
dx

Fi = mai = m

The moving ions experience a viscous drag f that is proportional to their velocities.
So the total force on the ions is a sum of the Coulomb force and the viscous drag
Fi = zi eE fvi

(in solution).

(23.30)

The ions quickly reach terminal velocity, i.e., the viscous drag equals the Coulomb
force. Hence Fi = 0.
zi eE
zi eE = f vi = vi =
.
(23.31)
f
The drag f has three basic origins.

166

1. Stokes Law type force


spherical ion moving through a continuous medium
this contribution is independent of the other ions
2. Electrophoretic eect.
oppositely charged ions pull at each other

3. Relaxation eects
solvation shell must re-adjust as ion moves. a dressed ion.

167

A more fundamental quantity than ion velocity is the ion mobility, ui which is the
ions velocity per field,
vi
ui = .
(23.32)
E
For the case for parallel plate capacitors E = l , where l is the separation of the
plates. So,
vi l
.
(23.33)
ui =

Here the current carried by ion i is


Ii = |zi | e

Ni ui
A .
V
l

(23.34)

Suppose a salt has a degree of dissociation ( = 1 for strong electrolytes) to


produce + cations and anions, then each mole of salt gives: N+ = + Ln
and N = Ln.
The current then becomes
Ii = |zi | e

i Ln ui F=Le

A
= i n |zi | ui AF
V
l
Vl

(23.35)

It is of interest to determine the ratio of the current carried by the cation versus
the anion.
=1
z }| {

/
n |z+ | u+ A
/ /F V/
+/
I+
u+
+ |z+ | u+
l
=
=
=
(23.36)

I
|z | u
u
/
/
n |z | u A
/ /F V/l
Thus the ratio of the currents is determined by simply the ratio of the mobilities.

168

24. Thermodynamics of Solvation


An extremely important application of thermodynamics is to that of ion solvation.
Solvation describes how a solute dissolves in a solvent.
We will focus on ions in solution.
As a basic treatment of solvation we shall consider the solvent as a non-structural
continuum and the ion as a charged particle.
Of course this is an approximation and numerous statistical mechanical models
for solvents which incorporate a more realistic structure can be used, but we will
stick with this simple thermodynamic model.
The way to investigate the ionsolvent interaction upon solvation from a thermodynamics point of view is to consider the change in the properties of the ion in a
vacuum versus the ion in solution.
Primarily we will determine 4Gvs Gion in solv. Gion in vac .
Since Gibbs free energy corresponds to non-P V work, 4Gvs can be determined
by calculating the reversible work done in transferring an ion into the bulk of the
solvent.

169
169

24.1. The Born Model


The Born model is a simple solvation model in which the ions are taken to be
charged spheres and the solvent is take to be a continuum with dielectric constant
s

170

4Gvs for the Born model is obtained by considering the following contribution
to the work of ion transfer from the vacuum state to the solvated state (see figure)
Begin with the state in which the charged sphere (the ion) is in a vacuum.
Determine the work, wdis , done in discharging the sphere.
Assume the uncharged sphere can pass from the (neutral) vacuum to the
neutral solvent without doing any work, wtr = 0. (This is an approximation).
Determine the work, wch , done in charging the sphere which is now in the
solvent.

171

So,
4Gvs = wdis + wtr + wch = wdis + wch

(24.1)

Work done in discharging the sphere:


The act of discharging a sphere involves bringing out to infinity from the surface
infinitesimal amounts of charge.
The work done is discharging is some what complicated since as one removes the
charge the work done in removing more charge changes according to the amount
of charge currently on the sphere.

172

This is expressed mathematically as


Z 0Z

wdis =
drd
2
ze ri 4 0 r
Z 0

d
=
ze 4 0 ri
(ze)2
,
=
8 0 ri

(24.2)

where z is the oxidation state of the ion, e is the charge of the electron, ri is the
radius of the sphere (ion) and 0 is the permittivity of free space.
Work done in charging the sphere:
The only dierence in charging the sphere is that the sign of the work will be dierent and that since we are charging in a solvent we must multiply the permittivity
of free space by the dielectric constant of the solvent.
So,
wch = +

(ze)2
8 0 s ri

(24.3)

24.1.1. Free Energy of Solvation for the Born Model


Combining the above two expression for work gives
(ze)2
(ze)2
+
8 0 ri 8 0 s ri

(ze)2
1
=
1
8 0 ri s

4Gvs =

(24.4)

The above expression is 4Gvs /ion. For n moles of ions (nL = N)


4Gvs

N (ze)2
=
8 0 ri

1
1
s

(24.5)

173

The dielectric constant of any solvent is always greater than unity so 1s 1 is


always negative hence 4Gvs < 0. Thus ions always exist more stably in solution
than in a vacuum.

24.1.2. Ion Transfer Between Phases


We can quickly generalize the Born model to describe ion transfer between phases
in a solution of two immiscible phases
Consider an immiscible solution of two phases and having dielectric constants
and .
Since Gibbs free energy is a state function we can write the change in free energy
for transfer of an ion form the phase to the phase as
=4Gv

z }| {
4Gv + 4Gv

N (ze)2 1
N (ze)2 1
=
1 +
1
8 0 ri
8 0 ri

1
N (ze)2 1

=
8 0 ri

4G =

(24.6)

The Partition Coecient


We can now write the partition coecient for the Born model as
/
Pi

=e

4G

nRT

L(ze)2
i 0 RT

8r

=e

1
1

(24.7)

24.1.3. Enthalpy and Entropy of Solvation


We may employ the standard thermodynamic relations which we have derived
earlier to obtain the entropy and enthalpy for the Born model.

174

From

G
T

= S

=
T

"

4Gvs
T

= 4Svs ,

(24.8)

we find entropy to be
4Svs

N (ze)2
8 0 ri

#
1
1 .
s

(24.9)

The only variable in the above equation that has a temperature dependence is the
dielectric constant of the solvent so,

1
N (ze)2 s
N (ze)2
.
(24.10)
=
4Svs =
8 0 ri T s
8 0 ri 2s T
Enthalpy is obtained via the relation:
4Hvs = 4Gvs + T 4Svs

N (ze)2 1
N (ze)2 T s
=
1 +
8 0 ri s
8 0 ri 2s T

N (ze)2 1
T s
1
=
+
8 0 ri s 2s T

(24.11)

24.2. Corrections to the Born Model


The Born model is very valuable because of its simplicityqualitative statements
about solvation and ion transfer between phases can be made.
Unfortunately however, the Born model does not make quantitatively correct predictions in many cases.
We simply list here several phenomena that more sophisticated theories of solvation must consider

175

1. The solvophobic eect: a cavity must form in the solvent to accommodate


the ion.
2. Changes in solvent structure: the local environment of the ion has a dierent
arrangement of solvent molecules than that of the bulk solvent, so the initial
structure of the solvent must breakdown and the new structure must form.
3. Specific interactions: any interaction energy specific to the particular ionsolvent pair: Hydrogen bonding being the prime example.
4. Annihilation of defects: A small ion may be captured in a micro-cavity
within the solvent releasing the energy of the micro-cavity defect.

176

25. Key Equations for Exam 4


Listed here are some of the key equations for Exam 4. This section should not
substitute for your studying of the rest of this material.
The equations listed here are out of context and it would help you very little to
memorize this section without understanding the context of these equations.
The equations are collected here simply for handy reference for you while working
the problem sets.

Equations
Some thermodynamic relations
H = U + PV
A = U TS
G = H TS

dH = T dS + V dP
dA = SdT P dV
dG = SdT + V dP

The chemical potential equation


i =
i + RT ln ai

(25.1)

The 4G equation (this should be posted on your refrigerator)


4G = 4G + RT ln Q.

(25.2)

177
177

At equilibrium 4G = 0 and
4G = RT ln Ka

(25.3)

CP m = Cvm + R

(25.4)

For an ideal gas

The DebyeHukel limiting Law (DHLL):

ln = |z+ z | I.

(25.5)

The ratio of the current carried by the cation versus the anion in terms of
ion mobility is
I+
u+
=
(25.6)
I
u
The chemical potential equation
i =
i + RT ln ai

(25.7)

The 4G equation (this should be posted on your refrigerator)


4G = 4G + RT ln Q.

(25.8)

At equilibrium 4G = 0 and
4G = RT ln Ka
4G for the Born model:
4Gvs

N (ze)2
=
8 0 rs

1
1
s

4G for transfer of an ion form the phase to the phase,

N (ze)2 1
1

4G =
8 0 ri

(25.9)

(25.10)

(25.11)

178

Chemistry 352: Physical


Chemistry II

179
179

Part V
Quantum Mechanics and
Dynamics

180
180

26. Particle in a 3D Box


We now return to quantum mechanics and investigate some of the important
models that we omitted from the first semester.
In particular we will look at the particle in a box in more than one dimension.
We will also solve models which deal with rotations.

26.1. Particle in a Box


Recall that the important ideas from the 1D particle in a box problem were
The potential, V (x), is given by

V (x) =

x0
0<x<a .
xa

(26.1)

Because of the infinities at x = 0 and x = a, we need to partition the x-axis into


the three regions shown in the figure.

181
181

Now, in region I and III, where the potential is infinite, the particle can never
exist so, must equal zero in these regions.
The particle must be found only in region II.
The Schrdinger equation in region II is (V (x) = 0)
2 2
= E = ~ d (x) = E,
H
2m dx2

(26.2)

The general solution of this dierential equation is


(x) = A sin kx + B cos kx,
where k =

(26.3)

2mE
.
~2

Now must be continuous for all x. Therefore it must satisfy the boundary
conditions (b.c.): (0) = 0 and (a) = 0.
From the (0) = 0 b.c. we see that the constant B must be zero because
cos kx|x=0 = 1.
So we are left with (x) = A sin kx for our wavefunction.

182

The second b.c., (a) = 0, places certain restrictions on k.


In particular,

n
,
n = 1, 2, 3, .
a
The values of k are quantized. So, now we have
kn =

n (x) = A sin

nx
.
a

(26.4)

(26.5)

The constant A is the normalization constant.


Solving for A gives
A=

2
.
a

(26.6)

Thus our normalized wavefunctions for a particle in a box are (in region II)
r
nx
2
sin
.
(26.7)
n (x) =
a
a
We found the energy levels to be
En =

h
n2 2 ~2 ~= 2
n2 h2
=
.
2ma2
8ma2

(26.8)

26.2. The 3D Particle in a Box Problem


We now consider the three dimensional version of the problem.
The potential is now
V (x, y, z) =

0,

0 < x < a, 0 < y < b, 0 < z < c


.
,
else

(26.9)

183

Now the Schrdinger equation is


2
= E ~ 2 = E
H
2m

~2 2 2 2
+ 2 + 2 = E.

2m x2
y
z

(26.10)

It is generally true that when the Hamiltonian is a sum of independent terms, we


can write the wavefunction as a product of wavefunctions
(x, y, z) = x (x) y (y) z (z).

(26.11)

This lets us perform a mathematical trick which is sometimes useful in solving


partial dierential equations.
Subbing the product wavefunction into the Schrdinger equation we get

~2 2 x y z 2 x y z 2 x y z
= E x y z
+
+
(26.12)
2m
x2
y 2
z 2

~2 y z 2 x x z 2 y x y 2 z
= E x y z .
+
+
2m
x2
y 2
z 2

We now divide both sides by x y z to get

1 2y
1 2z
~2 1 2 x
= E.
+
+
2m x x2
y y 2
z z 2

(26.13)

This equation is now of the form


f (x) + g(y) + h(z) = C,

(26.14)

where C is a constant.
If we take the derivative with respect to x we get
d
f (x) + g(y) + h(z) = C,
dx
dC
df (x) dg(y) dh(z)
+
+
=
,
dx
dx
dx
dx
df (x)
= 0,
dx

(26.15)

184

So, f (x) is a constant. Similarly for g(y) and h(z)


Applying this to our Schrdinger equation means that we have converted our
partial dierential equation into three independent ordinary dierential equations,
~2 d2 x
~2 1 d2 x
=
E
=
= Ex x
x
2m x dx2
2m dx2
~2 d2 y
~2 1 d2 y
=
E
=
= Ey y
y
2m y dy 2
2m dy 2
~2 d2 z
~2 1 d2 z
=
E
=
= Ez z
z
2m z dz 2
2m dz 2

(26.16)

which we recognize as the 1D particle in a box equations.


Hence we immediately have
nx x
2
sin
,
a
a
r
ny y
2
sin
,
=
b
b
r
nz z
2
sin
=
c
c

x =
y
z

(26.17)

and
n2x h2
,
8ma2
n2y h2
=
,
8mb2
n2z h2
=
.
8mc2

Ex,nx =
Ey,ny
Ez,nz

(26.18)

The total wavefunction is

ny y
nz z
2 2
nx x
sin
sin
=
sin
a
b
c
abc

(26.19)

and the total energy is


E = Ex,nx + Ey,ny + Ez,nz .

(26.20)

185

Degeneracy
The 3D particle in a box model brings up the concept of degeneracy.
When n(> 1) states have the same total energy they are said to be n-fold degenerate.
Let the 3D box be a cube (a = b = c) then the states
(nx = 2, ny = 1, nz = 1),

(26.21)

(nx = 1, ny = 2, nz = 1),
(nx = 1, ny = 1, nz = 2)
have the same total energy and thus are degenerate.

186

27. Operators
27.1. Operator Algebra
We now take a mathematical excursion and discuss the algebra of operators.
Definitions
Function: A function, say f, describes how a dependent variable, say y, is
related to an independent variable, say x: y = f(x)
e.g., y = x2 , y = sin x, etc.
transforms a function, say f , into another
Operator: An operator, say O,

function, say g: Of(x)


= g(x).
Algebra: An algebra is a specific collection of rules applied to a set of objects
and a particular operation
Rules
Transitivity
Associativity
Existence of an identity
Existence of an inverse
e.g., Addition on the set of real numbers, Multiplication on the set of
real numbers
187
187

Note: Commutivity is not a requirement of an algebra


example 1: multiplication on the set of real number is commutive:
ab = ba
example 2: multiplication on the set of n n matrices is not commutive: ab 6= ba in general. e.g.,
"
#"
# "
#
1 0
3 1
3 1
=
(27.1)
2 1
1 1
7 3
but

"

3 1
1 1

#"

1 0
2 1

"

5 1
3 1

6=

"

3 1
7 3

(27.2)

Algebraic rules for operators


1. Equality:
then
(x)
if
= ,
f (x) = g(x) = f

(27.3)

(x) = h(x),
if
f (x) = g(x) and f
(x) =
(x) = g(x) + h(x)
f (x) + f
then (
+ )f

(27.4)

2. Addition:

3. Multiplication:

f (x) =
f (x)

(27.5)


f (x) = (
f (x)) ,
(x) 6=
but in general
f
f (x).
4. Inverse:

if
f (x) = g(x) and g(x)
= f (x)
inverse
then =
1 and is said to be

(27.6)

188

Linear operators:
A special and important class of operators
They obey all of the above properties in addition to

(f (x) + g(x)) =
f (x) +
g(x), and

(f (x)) =
f (x), where is a complex number.
Hermitian operators:
A special class of linear operators
All observables in quantum mechanics are associated with Hermitian operators
The eigenvalues of Hermitian operators are real
Some important operators
1.

x: xf(x) = xf(x)

df
(x) =
d:

d
f (x)
dx

(x) = d d f (x) =
d2 : d2 f (x) = d df
dx

d
dx

dx

f (x) =

d2
f (x)
dx2

: f (x, y, z) = f (x, y, z)

: f(x, y, z) = x ex + y ey + z ez f(x, y, z)

2:
2 f (x, y, z) = 22 + 22 + 22 f (x, y, z)

x
y
z

Commutators:
We have seen that in general
6=
. This leads to the construction of the
commutator, [, ]:
h
i

,

.
(27.7)

189

h
i

If
=
, then
, = 0 and
and are said to commute with one another.
The eigenvalue equation:
If
f (x) = g(x) and g(x) = af(x), then the operator equation,
f (x) = g(x)
becomes the eigenvalue equation

f (x) = af (x).

(27.8)

The eigenvalue equation is of fundamental importance in quantum theory. We


shall see that eigenvalues of certain operator can be identified as experimental
observables.
Commuting operators and simultaneous sets of eigenfunctions.
(x) = bf (x).
If
f (x) = af (x) and and
commute, then f
The proof goes as follows: On the one hand,

(
f ) = (af) = af

(27.9)

because f is an eigenfunction of
.
On the other hand,

because and
commute.
Thus

(
f ) =
f

(27.10)

f
= a f

(27.11)

is an eigenfunction of
which states that f
with eigenvalue a. The only way for

this to be true is if f = bf.

190

27.2. Orthogonality, Completeness, and the Superposition


Principle
Theorem 1: The eigenfunctions of a Hermitian operator corresponding to dierent eigenvalues are orthogonal:
Z
j k = 0,
j 6= k.
(27.12)
space

Theorem 2: The eigenfunctions of a Hermitian operator form a complete set


Corollary (the superposition principle): Any arbitrary function in the
space of eigenfunctions {i } can be written as a superposition of these eigenfunctions:
X
=
ai i
(27.13)
i

191

28. Angular Momentum


We will encounter several dierent types of angular momenta, but fortunately
they are all described by a single theory
Before starting with the quantum mechanical treatment of angular momentum,
we first review the classical treatment.

28.1. Classical Theory of Angular Momentum


The classical angular momentum, L, is given by
L=xp

(28.1)

The vector cross-product can be computed by finding the following determinant:

ex ey ez
Ly
Lx
L
z }|

{
z
}|
{
z
}|z
{

(28.2)
L = x y z = (ypz zpy )ex + (zpx xpz )ey + (xpy ypx )ez

px py pz

Hence,

Lx = (ypz zpy ) ,

(28.3)

Ly = (zpx xpz ) ,

(28.4)

Lz = (xpy ypx ) .

(28.5)

Another quantity that we will find useful is


L2 = L L = L2x + L2y + L2z

(28.6)

192
192

28.2. Quantum theory of Angular Momentum


So, in accordance with postulate II, we replace the classical variables with their
operators. That is,

y
z
,
(28.7)
Lx = (
y pz zpy ) =
i
z
y

~
y = (
z
x
,
(28.8)
z px xpz ) =
L
i
x
z

~
z = (
x y
.
(28.9)
xpy ypx ) =
L
i
y
x
Recall the basic commutators.

, u = 1,
u

, v = 0,
u

(28.10)

where u, v = x, y, or z and u 6= v.
From these basic commutators one can derive
h
i
i
h

x,

Lx , Ly = i~Lz ,
Ly , Lz = i~L

and

i
h
y

Lz , Lx = i~L

i h
i h
i
h
2
2
2

L , Lx = L , Ly = L , Lz = 0

(28.11)

(28.12)

It is often convenient to express the angular momentum operators in spherical


polar coordinates as follows.

x = i~ sin + cot cos


L
,
(28.13)

,
(28.14)
Ly = i~ cos cot sin

193

z = i~
L

2
1

2
2
= ~
+ cot +
L
sin2 2
2

(28.15)
(28.16)

28.3. Particle on a Ring


Consider a particle of mass confined to move on a ring of radius R.
The moment of inertia is I = R2
The Hamiltonian is given by
2
2
2
= Lz = ~ d
H
2I
2I d2

(28.17)

(note that we use d rather than since the problem is one-dimensional).


The Schrdinger equation becomes
~2 d2
= E
2I d2

(28.18)

Notice that this Schrdinger equation is exactly the same form as the particle in
a box. The only dierence is the boundary conditions.
The boundary condition for the particle in a box were was zero outside the box.
Now the boundary condition is that () = ( + 2). The wavefunction must
by 2 periodic.
The allowable wavefunctions are

A cos m
m () =
A sin m ,

Aeim

(28.19)

194

m = 0, 1, 2, 3, . . .
These wavefunctions are really the same. It will be most convenient to use
m () = Aeim as our wave functions.
Plugging m () = Aeim into the Schrdinger equation gives
~2 d2 Aeim
= Em Aeim
2I
d2
~2 m2 im
Ae
= Em Aeim
2I
Therefore the energy levels (the eigenvalues) for a particle in a ring are
m2 h2
~2 m2
=
.
2I
82 I
Next we need to find the normalization constant, A.
Z 2
d
1 =
Z0 2
1 =
A2 eim eim d
0
Z 2
2
1 = A
d = 2A2 ,
Em =

(28.20)

(28.21)

(28.22)

thus

1
.
2
Hence the normalized wavefunctions for a particle on a ring are
1
= eim .
2
A=

(28.23)

(28.24)

28.4. General Theory of Angular Momentum


To discuss angular momentum in a more general way it is convenient to define
two so-called ladder operators
+ L
x + iL
y
L

(28.25)

195

and
x iL
y
L
L

:
+ and L
We collect here the commutators of L
h
i
+ L
+ = L
+L
zL
+
z = L
+ L
z, L
L
i
h
L
= L
L
zL

z = L
+ L
z, L
L

(28.26)

(28.27)
(28.28)

2 commute there must exist a set of simultaneous eigenfunc z and L


Now, since L
tions {i }
z i = m i
(28.29)
L
and
2 i = k2 i
L

(28.30)

Physically, k~ represents the length of the angular momentum vector and m~


represents the projection onto the z-axis. (Note: for simplicity in writing we are
hiding the ~ in the wavefunctions.)
On these physical grounds we conclude |m| k, i.e., k sets an upper and lower
limit on m.
Lets define the maximum value of m to be a new quantum number l mmax .
(Thus l k).
And lets define the minium value of m to be a new quantum number l0 mmin .
(Thus l0 k)
Now, at least one of the eigenfunctions in the set { i } yields the eigenvalue mmax
z . Lets call that eigenfunction l ;
(or l) when operated on by L
z l = l l .
L

(28.31)

:
Now we can operate on both sides of this equation with L
l l
z l = L
L
L

(28.32)

196

L
zL
we get
z = L
+ L
Using the commutator relation L

l = lL
l
+ L
zL
L

(28.33)

l = lL
l
l + L
zL
L

Bringing the second term on the left hand side over to the right hand side gives
zL
l L
l
l = lL
L
zL
l
l = (l 1)L
L
| {z
}
| {z
}
l1

(28.34)

l1

l l1 is in fact an eigenfunction of L
z (with associated eigenWe see that L
value (l 1)) and is thus a member of {i } .
The eigenfunction l1 has an associated eigenvalue that is one unit less then the
maximum value.
n l = ln provided n
The above procedure can be repeated n times so that L
does not exceed l l0 .
The eigenfunction ln has an associated eigenvalue that is n units less then the
maximum value, i.e.,
z ln = (l n) ln .
L
(28.35)
The largest value of n is l l0 . For that case,
z l0 = (l l + l0 )l0 = l0 l0 .
L

(28.36)

+ , except in the opposite directionthe


Similar behavior is seen for the operator L
+ . For example
eigenvalue is increased by one unit for each action of L
+L
+ l 0 l0
z l0 = L
L

+ l0 = l 0 L
+ L
+ l0
zL
L

(28.37)

+ l0 .
zL
+ l0 = (l0 + 1)L
L

197

+ and L
is why they are called ladder
The raising and lowering nature of L
operators.
+ and L
indefinitely since we are limited by lwe reach
We can not act with L
the ends of the ladder. This requires that
l0 = 0
L

(28.38)

(we cant go lower than the lowest step) and


+l = 0
L

(28.39)

(we cant go higher than the highest step).


+L
+ or L
so it
L
Often times the ladder operators appear in tandem either as L
is useful list some identities for these products
L
2 L
2z L
z
+ = L
L

(28.40)

+L
2 L
2z + L
z
= L
L

(28.41)

and

We can use these identities to derive a relation between the quantum numbers k
and l.
We begin with

L L+ l = L L+ l = 0,

but from the first of the above identities

2z L
z l = (k2 l2 l) l
L
+l = L
2 L
L

Therefore

k2 l2 l = 0 k =

p
l(l + 1).

(28.42)

(28.43)

(28.44)

198

We we can also consider

and

+ L
l0 = L
l0 = 0
+L
L

(28.45)

+L
2z + L
z l0 = (k 2 l02 + l0 )l0 .
l0 = L
2 L
L

(28.46)

l(l + 1) l02 + l0 = 0;

(28.47)

l = l0

(28.48)

substituting in the relation we just found for k gives

simplifying gives

Thus mmax = l, mmin = l and so m = l, l 1, l 2, . . . , l + 1 , l.


This also implies that the number of rungs is 2l + 1 and that l must be either an
integer or a half-integer.

28.5. Quantum Properties of Angular Momentum


The eigenfunctions of angular momentum are entirely specified by two quantum
numbers l and m: lm .
2 lm = l(l + 1) lm
L

z lm = m lm
L

(28.49)

If we write out the first of these explicitly in spherical polar coordinates as a


partial dierential equation we obtain
1 2 lm
2 lm
lm
+
+ cot
+ l(l + 1) lm = 0

sin2 2
2

(28.50)

The solutions to this partial dierential equation are known to be the spherical
harmonic functions
lm = Ylm (, ).
(28.51)

199

The spherical harmonics are functions of two variables, but they are a product of
a function only of and a function only of ,
|m|

lm = Ylm (, ) = APl ()eim ,

(28.52)

|m|

where the Pl () are the Legendra polynomials and A is normalization constant.


Both the spherical harmonics and the Legendra polynomials are tabulated. They
are also built-in functions of Mathematica.
The spherical harmonics (and hence the angular momentum wavefunctions) are
orthonormal; meaning,
(
Z 2 Z
1
l0 = l and m0 = m
(28.53)
Yl0 m0 (, )Ylm (, ) sin dd =
0
l0 6= l or m0 6= m
0
0
28.5.1. The rigid rotor
Rotational energy
For general rotation in three dimensions the is
2
= ~ L
2,
H
2I

(28.54)

so the Schrdinger equation is


2
2
2 lm = Elm lm ~ l(l + 1)lm = Elm lm .
lm = Elm lm ~ L
H
2I
2I

(28.55)

Thus

l(l + 1)h2
l(l + 1)~2
=
= El .
(28.56)
2I
82 I
There is no m dependence for the energy. In other words, the energy levels are
determined only by the value of l.
Elm =

We know that there are 2l + 1 dierent m values for a particular l value. All 2l + 1
of these wavefunctions correspond to the same energy. We say the there is a 2l + 1
degeneracy of the energy levels.

200

29. Addition of Angular Momentum


29.1. Spin Angular Momentum
We learned above that l may take on integer or half-integer values.
Systems in which l takes on half-integer values are peculiar.
These systems have no classical analogs.
One example of such a system is the spin of an electron, l = s = 1/2. The values
of m = ms are limited to +1/2 and 1/2.
One peculiarity of this system is that the wavefunctions are 4 periodic (and 2
antiperiodic):
s () = s ( + 2)
(29.1)
and
s () = s ( + 4).

(29.2)

That means that the system has to rotate twice (in spin space not coordinate
space) to get back to its original state.
See in-class demonstration: the belt trick

201
201

29.2. Addition of Angular Momentum


In atoms the are a number of sources of angular momentum: The ls and ss of
each of the electrons.
One measures, however, the total angular momentum, J.
The electrons in many electron atoms couple. The are two main coupling schemes
which account for the total angular momentum of the atom.
1. LS coupling (also called Russell-Saunders coupling)
works well for low atomic weight atoms (first couple of rows of the
periodic table)
P
find the total spin angular momentum S = Ms,max , (Ms = i msi )
P
find the total orbital angular momentum L = Mmax , (M = i mi )
then J = L + S

2. jj coupling
applies to higher atomic weight atoms
find subtotal angular momentum for each electron ji = li + si
P
then find total angular momentum by J = i ji .
we will not use this method.

29.2.1. The Addition of Angular Momentum: General Theory


Consider two sources of angular momentum for a system represented by the op or S angular momentum; we use J when
erators J1 and J2 (J1 and J2 could be L
we speak generally.)

202

The total angular momentum is JT = J1 + J2 .


The total z-component of the angular momentum is JzT = Jz1 + Jz2
The last statement implies that the orientation quantum number of the total
system is simple the sum of that for the components
M = m1 + m2

(29.3)

We need to determine the allowed values of the total angular momentum quantum
number J.
The maximum value of J is determined by the maximum value of M by
Jmax = Mmax = m1max + m2max = j1 + j2

(29.4)

This corresponds to a situation in which component angular momentums add in


the most favorable manner
The minimum value of J is determined by the case when the components add in
the least favorable manner. That is,
Jmin = |j1 j2 | .

(29.5)

The total angular momentum is quantized is exactly the same manner as any
other angular momentum. Thus the allowed values of J are
J = j1 + j2 , j1 + j2 1, . . . , |j1 j2 | + 1, |j1 j2 | .

(29.6)

29.2.2. An Example: Two Electrons


The table below shows the total spin angular momentum S for a two electron
system

203

spin state

ms1

ms1

MS

(1)(2)
(1)(2)
(1)(2) + (1)(2)
(1)(2) (1)(2)

1
2

1
2

12

12

1
1
0
0

1
1
1
0

0
0

0
0

Counting states:
The spin degeneracy, gS , of the states is given by 2S + 1. In the above example
the degeneracy is gS = 3 for the S = 1 states and gS = 1 for the S = 0 states.
29.2.3. Term Symbols
We have already seen several term symbols, those being 1 S and 3 S during our
discussion of helium.
Term symbols are simply shorthand notion used to identify states. Term symbols
are useful for predict and understanding spectroscopic data. So, it is worthwhile
to briefly discuss them.
In general the term symbol is simply notates the total orbital angular momentum
and spin degeneracies of a particular set of states (or a state in the case of a singlet
state).
The orbital degeneracy is given by gL = 2L + 1.
For historical reasons L values are associated with a letter like the l values of a
hydrogenic system are.
L
0
symbol S

1
P

2
D

3
F

4 5
.
G H

204

The term symbol for a particular states is constructed from the following general
template
gS
LJ .
Many electron atoms have term symbols associated with their states.
Rules:
1. All closed shells have zero spin and orbital angular momentums: L = 0,
S = 0. These states are all singlet S states, notated by 1 S
2. An electron and a hole lead to equivalent term symbols.
E.g., p1 and p5 have the same term symbol.
3. Hunds Rule for the ground state only.
1. The ground state will have maximum multiplicity.
2. If several terms have the same multiplicity then ground state will be
that of the largest L.
3. Lowest J value (regular) electron, Highest J value (inverted) hole
29.2.4. Spin Orbit Coupling
A charge possessing angular momentum has a magnetic dipole associated with it.
An electron has orbital and spin magnetic dipoles.
These dipoles interact with a certain spinorbit interaction energy ESO .
The spinorbit Hamiltonian is
[
SO = hcAL
S
H

hcA 2 2 2

J L S ,
HSO =
2

(29.7)

205

where A is the spinorbit coupling constant.


From the Hamiltonian the spinorbit interaction energy is
ESO =

hcA
[J(J + 1) L(L + 1) S(S + 1)]
2

(29.8)

206

30. Approximation Techniques


As we learned last semester, there are very few models for which we can obtain
an exact solution.
Consequently we must be satisfied with using approximation methods.
Last semester, we always took the simplest approximation to give the qualitative
properties of the unsolvable system.
Now we will consider two important quantitative approximation methods: (i)
perturbation theory and (ii) variational theory

30.1. Perturbation Theory


The basic procedure of perturbation theory
Find a solvable system that is similar to the system at hand.
Treat the dierence between the two systems as a perturbation to the solvable system
Use the solvable systems wavefunctions as a zeroth order approximation to
the wavefunctions for the unsolvable system.
These wavefunctions are used to find a first order correction to the energy.

207
207

The first order energy is then used to make a first order approximation to
the wavefunction.
The procedure is repeated to get higher and higher order approximations.
This process get algebraically intensive so we will only go as far as listing the first
order energy correction.
The nth state energy in perturbation theory:
En = En(0) + En(1) + . . . ,

(30.1)

(0)

(1)

where En is the nth state energy for the unperturbed (solvable) system and En
is the first order correction. This is given by
Z
(1)
(1) (0)
En =
(0)
(30.2)
n H
n dx,
all
space

(1) is the first order correction to the Hamiltonianthe perturbation.


where H
Example: the quartic oscillator
Consider the quartic oscillator described by the potential V (x) = 12 kx2 +ax4
where a is very small and can be treated as a perturbation.
The obvious solvable system is the harmonic oscillator:
2
2
= ~ d + 1 kx2 .
H
2m dx2 2

(30.3)

2
This has energy levels En = ~(n+ 12 ) and wavefunctions An Hn ( x)ex /2 ,
q
where = km
~

The perturbative part of the Hamiltonian is


(1) = ax4 .
H

(30.4)

208

For example, the ground state energy correction is then calculation from
Z
(1)
(0) (1) (0)
0 H
0 dx
(30.5)
E0 =

Z
2
2
A0 ex /2 ax4 A0 ex /2 dx
=

Z
2
2
x4 ex dx
= aA0

3 aA20
=
,
5
4 2
so the first order ground state energy for a quartic oscillator is

~ 3 aA20
+
E0 '
.
5
2
4 2

30.2. Variational method


The basic idea behind the variational method is to use a trial wavefunction with
an adjustable parameter. The value of the parameter which minimizes the energy,
Etrial , gives a trial wavefunction which is closest to the real wavefunction.
The basis for this is the variation theorem which states
Etrial E.
We will not prove this theorem here.
The trial energy is calculated by
R

all
space

Etrial = R

trial dx
trial H

all
space

trial trial dx

(30.6)

The trial energy is now a function of the adjustable parameter, p, that we use to
minimize the trial energy by setting
dEtrial
=0
dp

(30.7)

209

and solving for p. (Strictly speaking we should check that we have a minimum
and not a maximum or inflection point, but with reasonably good trial functions
one is pretty safe in having a minimum.)

210

31. The Two Level System and


Quantum Dynamics
Our entire discussion of quantum mechanics thus far had dealt only with time
independent quantum mechanics.
The time variable never appears in any expression.
Obviously there are cases where quantum objects move with time. For example,
firing an electron down a particle accelerator.
We shall finally get to quantum dynamics in this chapter, but first we will discuss
the very important model of the two level system.

31.1. The Two Level System


If the harmonic oscillator is the most important model in all a physics, the two
level system is a close second.
The spin system discussed above is an example of a two level system.
The two level system is inherently quantum mechanical in nature. Unlike the
harmonic oscillator it has no classical analogue.

211
211

Consequently, we can not use our usual procedure of writing down the classical
Hamiltonian and then replacing the variables with their corresponding operators.
The two level system consists of two states 1 and 2 separated by energy 4 =
2 1 as shown below

The states 1 and 2 are orthonormal:


Z

j k d =

TLS

1
0

j=k
,
j 6= k

(31.1)

R
where TLS d means integration over the two level space (which is really just the
P
sum 2i=1 ).
The states 1 and 2 are eigenfunctions of the two level Hamiltonian,
=
H

1 1,

+ 2 2, ,

(31.2)

where j, projects out the j th state of the wavefunction being acted on.

212

For example let some arbitrary wavefunction = a 1 + b2 , then

H
= ( 1 1, +
=

1 1,

2 2, ) (a 1

(a1 + b2 ) +

+ b2 )
2 2,

(31.3)

(a1 + b2 )

= a 1 1 + b 22
Another orthonormal set of wavefunctions are the so-called left

and right

1
1
L = 1 + 2
2
2

(31.4)

1
1
R = 1 2
2
2

(31.5)

states.

We can invert above equations and solve for 1 and 2 in terms of L and R

and

1
1
1 = L + R
2
2

(31.6)

1
1
2 = L R .
2
2

(31.7)

213

31.2. Quantum Dynamics


So far we have been concerned with the eigenfunctions and eigenvalues (energy
levels) of the various quantum systems that we have discussed.
What has been kept hidden up to now is the fact that the eigenfunctions are really
multiplied by a phase factor of the form .
i

n (x, t) n (x)e ~ En t

(31.8)

We can verify this by obtaining the time independent Schrdinger equation from
the more general time dependent
n (x, t)
t
i
n (x)e ~ En t
i~
t
i
e ~ En t
i~ n (x)
t

i
i
i~ n (x) En e ~ En t
~
i~

n (x, t)
= H

(31.9)

n (x)e ~i En t
= H
n (x)e ~i En t
= H
n (x)e ~i En t
= H
i

n (x)
En n (x)e ~ En t = e ~ En t H
n (x)
En n (x) = H

(31.10)

Does this mean the eigenstates are not stationary states? To determine this we
need to calculate the probability of finding the particle in the same eigenstate at
some future time. This is given by
Z
2

P (x, t) = n (x, 0)n (x, t)dx


(31.11)
2
Z

E
t
n
= n (x) n (x)e ~ dx

2
Z
iE t

n (x) n (x)dx
= e ~
i

~ En t 2
= e
(1) = 1,
214

so no matter what time t we check we will always find the system in the same
eigenstate. Thus the eigenstates are stationary states.
In general the state of the system need not be in one particular eigenstate; it may
be in a superposition of any number of eigenstates.
The left and right wavefunctions that we saw in the discussion of the two
level system are examples of superposition states.
The phase factor does become important for superposition states.
As an example consider the state
1
1
(x, t) = 1 (x, t) + 2 (x, t)
2
2

(31.12)

exposing the phase factors we get


i
i
1
1
(x, t) = 1 (x)e ~ E1 t + 2 (x)e ~ E2 t
2
2

(31.13)

Lets now track the probability of finding the particle in the same superposition
state. Similar to before we calculate
2
Z

P (x, t) = (x, 0)(x, t)


Z
2

1
1
1
1
~i E1 t
~i E2 t

1 (x) + 2 (x)
1 (x)e
=
+ 2 (x)e

2
2
2
2

Z
!
i
i
2
1
1 (x) 1 (x)e ~ E1 t + 1 (x) 2 (x)e ~ E2 t

dx .
=
(31.14)

~i E1 t
~i E2 t

2
+2 (x) 1 (x)e
+ 2 (x) 2 (x)e
The cross-terms (those of the form 1 (x) 2 (x) and 2 (x) 1 (x)) are zero when

215

integrated because the eigenfunctions are orthogonal. This leaves


2
Z

(31.15)
P (x, t) = (x, 0)(x, t)
Z
2
1
i
i

E
t

E
t
1 (x) 1 (x)e ~ 1 + 2 (x) 2 (x)e ~ 2 dx
=
2

2
Z
Z
1 i E t

E
t
1
2

e ~
1 (x) 1 (x)dx + e ~
2 (x) 2 (x)dx
=
2

2 1 i
i

1 i E t
~i E2 t
+ ~ E1 t
+ ~i E2 t
~ E1 t
~i E2 t
1

~
e
e
= e
+e
=
+
e
+
e

2
4

1
(E1 E2 )
1
+ ~i (E1 E2 )t
~i (E1 E2 )t
1+e
1 + cos
t .
+e
+1 =
=
4
2
~
The probability of find in the system in its original superposition states is not one
for all times t.

216

Key Equations for Exam 1


Listed here are some of the key equations for Exam 1. This section should not
substitute for your studying of the rest of this material.
The equations listed here are out of context and it would help you very little to
memorize this section without understanding the context of these equations.
The equations are collected here simply for handy reference for you while working
the problem sets.

Equations
The short cut for getting the normalization constant .
sZ
N=

space

|unnorm (x, y, z)|2 dxdydz.

(31.16)

The normalized wavefunction:


norm =

.
N unnorm

How to get the average value for some property,


Z

dxdydz.
h
i =

(31.17)

(31.18)

space

217
217

The Laplacian

2
2
2
+
+
x2 y 2 z 2

Normalized wavefunctions for the 3D particle in a box,

ny y
nz z
nx x
2 2
sin
sin
.
sin
n (x) =
a
b
c
abc

(31.19)

(31.20)

The energy levels for the 3D particle in a box,


Enx ,ny ,nz
Orthonormality:

n2y h2
n2x h2
n2z h2
=
+
+
.
8ma2 8mb2 8mc2

j k

space

Superpostion:

1,
0,

j=k
.
j 6= k

ai i

(31.21)

(31.22)

(31.23)

Commonly used comutators of the angular momentum operators are


h
i
i
i
h
h

(31.24)
Lx , Ly = i~Lz ,
Ly , Lz = i~Lx ,
Lz , Lx = i~L
and

h
i h
i h
i
2
2
2

L , Lx = L , Ly = L , Lz = 0.

(31.25)

The energy levels for a particle in a ring are


Em =

m2 h2
~2 m2
=
.
2I
82 I

(31.26)

The normalized wavefunctions for a particle on a ring are


1
= eim .
2

(31.27)

218

The eigenfunctions of angular momentum are entirely specified by two quantum numbers l and m: lm .
2 lm = l(l + 1) lm
L

z lm = m lm
L

(31.28)

The energy levels for the rigid rotor are


El =

l(l + 1)~2
.
2I

(31.29)

Degeneracy for general angular momentum is


(31.30)

gJ = 2J + 1.
The first order energy correction in pertubation theory is
Z
(1)
(1) (0)
(0)
En =
n H
n dx,
all
space

The trial energy in variation theory is calculated by


R

all trial H trial dx


space
Etrial = R

all trial trial dx

(31.31)

(31.32)

space

In general

n (x, t) n (x)e ~ En t

(31.33)

The left and right superposition states are

and

1
1
L = 1 + 2
2
2

(31.34)

1
1
R = 1 2
2
2

(31.35)

219

Part VI
Symmetry and Spectroscopy

220
220

32. Symmetry and Group Theory


We now take a short break from physical chemistry to discuss ideas from the
mathematical field of group theory.
Inherent to group theory is symmetry.
As far as we are concerned, we will
determine the symmetry of a particular molecule.
The types of symmetry it has will determine to which symmetry group it
belongs.
The mathematical properties of all the possible groups have been worked
out
These mathematical properties translate into a wide variety of variety of
physical properties including
Bonding
Properties of wavefunctions
Vibrational modes
Many more applications

221
221

32.1. Symmetry Operators


Any operator that leaves ||2 invariant are symmetry operators for that particular
system:
||2 = ||2 .
O
(32.1)
This implies
= .
O

(32.2)

That is, the eigenvalues for the particular symmetry operator are 1 or 1.
For molecules we will be dealing with point group symmetry operators. These
operators deal with symmetry about the center of mass.
We have seen two such operators in and
h.
An example of symmetry operator that is not a point group symmetry operator
would be an operator that performed some sort of translation in space. This type
of operator arrises in the treatment of extended crystal structures.
See Handout on Symmetry Elements

32.2. Mathematical Groups


In mathematics the term group has special meaning. It is a set of objects and
a single operation, which has the following properties.
1. The group is associative (but not necessarily communative) with respect to
the operation.
2. An identity element exits and is a member of the group

222

3. The product of any two members of the group yield a member of the
group.
4. The inverse of every member of the group is also in the group. In other
words, for any member of the group one can find another member of the
group which, upon multiplication, yields the identity element.

See Handout on Naming Point Groups


See Handout on Assigning Point Groups
Associated with a given group is a multiplication table.
32.2.1. Example: The C2v Group
C2 ,
The C2v group consists of the symmetry elements E,
v (in-plane) and
0v
(transverse).
Water is an example of a molecule described by this point group.
The multiplication table for the C2v group is
C2v

E
C2

E
E
C2

C2
C2
E

0v

0v

0v

0v

0v

E
C2

0v

v
C2

32.3. Symmetry of Functions


In the absence of degeneracy, the wavefunctions must be symmetric or antisymmetric with respect to all elements of the group.

223

Connecting with the C2v group example lets consider the wavefunctions for water.
In this case one can collect the eigenvalues (either +1 or 1) for each of the four
symmetry operators as a four component vector. As it turns out there is four
possible sets of eigenvalueshence four dierent vectors:
A1 = (1, 1, 1, 1)
A2 = (1, 1, 1, 1)
B1 = (1, 1, 1, 1)
B2 = (1, 1, 1, 1).
To see where these four vectors come from, consider the following.
is 1
The first value has to be +1 since the only eigenvalue of E
The eigenvalue of C2 can be +1 or 1
When it is +1 the vectors are labelled A
When it is 1 the vectors are labelled B
The eigenvalue of
v can be either +1 or 1
When it is +1 the vectors are labelled with a subscript 1
When it is 1 the vectors are labelled with a subscript 2
The eigenvalue of
0v can be either +1 or 1
Finally there is a restriction do to the fact that the eigenvalues must obey
the group multiplication table.
0v to be the same for
This restriction forces the eigenvalues of
v and
the A type vectors and opposite for the B type vectors.

224

The above considerations leave four vectors.


In fact, there will always be the same number of vectors as symmetry elements.
Altogether, the vectors represent what is call an irreducible representation of the
group.
These vectors make up the :
C2v

C2

A1
A2
B1
B2

1
1
1
1
1
1 1 1
1 1
1 1
1 1 1
1

0v

See Handout on Character Tables


32.3.1. Direct Products
The direct product of a two vectors is defined as
(x1 , x2 , x3 , . . .) (y1 , y2 , y3 , . . .) = (x1 y1 , x2 y2 , x3 y3 , . . .)

(32.3)

For the example of the C2v group consider


B1 B2 = (1, 1, 1, 1) (1, 1, 1, 1)
= (1, 1, 1, 1) = A2

(32.4)

32.4. Symmetry Breaking and Crystal Field Splitting


We shall investigate how degeneracies of energy levels are broken as one reduces
the overall symmetry of the system.

225

In doing this we will, for simplicity, consider only proper rotations (Cn ). Mirror
symmetry will not be considered (although in real applications one must consider
all symmetry).
First consider a free atom. In this case there is complete rotational symmetry.
Thus the symmetry group is the spherical group (see character table handout.)
This is the group associated with the particle on a sphere model and the angular
part of the hydrogen atom. The vectors are the labeled according to the angular
momentum quantum numbers S, P, D, F, etc.
The degeneracies of these vectors are 1 for S, 3 for P, 5 for D and so on as is
familiar to us already.
Now consider the free atom being placed in a crystal lattice of octahedral symmetry. For example placed at the center of a cube which has other atoms at the
centers of each face of the cube.
When moving to octahedral symmetry we now must look at the character table for
such a casethe O group (remember we are considering only proper rotations).
The S vector has the symmetry of a sphere (x2 + y 2 + z 2 ) and hence is totally
symmetric. It is also nondegenerate so it will be, of course, nondegenerate in
the octahedral case. It remains totally symmetric so it is now represented by the
vector A1 .
The P vector is triply degenerate and has the symmetry of x, y and z as we see
from the character table for the spherical group. In the octahedral crystal the
degeneracy remains in tact and these states are represented by the T1 group.

226

The D vector has a degeneracy of five and the symmetry of 2z 2 x2 y 2 , xz, yz,
xy, x2 y 2 . Looking at the table for the O group we see the degeneracy splits:
two states become E type and the remaining three become T2 type.
The F states have a degeneracy of 7 and the symmetry of z 3 , xz 2 , yz 2 , xyz,
z(x2 y 2 ), x(x2 3y 2 ) and y(3x2 y 2 ). In an octahedral environment the states
split with one becoming A2 , three becoming T1 and three becoming T2 . This is
not readily apparent from the character tables so one needs to inspect a little
harder to see it (see homework).
The octahedral group is still highly symmetric. Lets say that two atoms on opposite sides of the cube are moved slightly inward. The remaining four atoms remain
in place.
This breaks the octahedral symmetry and the system now assumes D4 symmetry.
Now the A1 vector of the O group becomes the A1 vector of the D4 group. The
triply degenerate T1 vector splits into a A2 state and a doubly degenerate E state.
The E states from the O group become a A1 type state and a B1 type state.
The T1 states from the O group become a A2 type state and a E type state.
The T2 states from the O group become a B2 and a E type state.

227

33. Molecules and Symmetry


From our chapter on diatomic molecules last semester we have learned a great
deal which caries over directly to polyatomic molecules.
So, in this chapter we simply investigate some of the specific details regarding
polyatomic molecules.

33.1. Molecular Vibrations


As for diatomic molecules, it is convenient to work with center of mass coordinates.
With polyatomic molecules one needs to specify the coordinates of N nuclei rather
than just two nuclei.
To do so we begin with the 3N nuclear degrees of freedom.
As for the diatomic case 3 degrees of freedom determine the center of mass motion.
That leaves us with 3N 3 coordinates to specify.
One must now consider two dierent types of polyatomic molecules: Linear and
Nonlinear.
For linear molecules there are 2 rotational degrees of freedom
For nonlinear molecules there are 3 rotational degrees of freedom
This now leaves one with 3N 5 vibrational degrees of freedom for linear polyatomic molecules and 3N 6 vibrational degrees of freedom for nonlinear molecules.

228
228

33.1.1. Normal Modes


Polyatomic molecules can undergo very complicated vibrational motion.
Regardless of what type of vibrational motion is taking place, however, that motion is some linear combination of fundamental vibrational motions called normal
modes.
This is analogous to writing an arbitrary wavefunction as a linear combination of
eigenfunctions. One example was the left and right states of the two level
system.
The number of normal modes equals the number of vibrational degrees of freedom.
At low energies the normal modes are well approximated as harmonic oscillators.

33.1.2. Normal Modes and Group Theory


The symmetry of the normal modes are associated with entries in the character
table of the point group of any particular polyatomic molecule.
Example: Water
The point group symmetry of the water molecule is C2v . The character table is
C2v

C2

A1
A2
B1
B2

1
1
1
1
1
1 1 1
1 1
1 1
1 1 1
1

0v

Water has three nuclei and it is nonlinear so it has 3(3) 6 = 3 normal modes.
The three modes are the bending vibration, the symmetric stretching vibration
and the asymmetric stretch.

229

The normal modes are associated with a particular vector (row) of the character
table by considering the action of the each of the symmetry elements on the normal
mode.
For the bending mode, the vibration is complete unchanged by any of the symmetry elements. Consequently the bending mode is associated with A1
The same is true for the symmetric stretching mode. It too is associated with A1 .
The asymmetric stretch, however, is associated with B1 since C2 and
0v transform
the mode into its opposite and
v leaves it unchanged.

230

34. Vibrational Spectroscopy and


Group Theory
We now investigate how group theory and, in particular, the character tables can
be used to determine IR and Raman spectra and selection rules for polyatomic
molecules

34.1. IR Spectroscopy
IR absorption is exactly the same as regular electronic absorption except the
frequency of the electromagnetic radiation is much less.
The typical energies for IR absorption are from 400 to 4000 cm1 . This is in
the Infrared region of the electromagnetic spectrum.
As for electronic absorption one typically employs the electric dipole approximation.
The electric dipole approximation
Molecule is viewed as a collection of charges
Multipole expansion
monopole + dipole + quadrapole+

(34.1)

231
231

Lightmatter interaction is dominated by the lightdipole coupling so the


other interactions are ignored.
In order for absorption of the electromagnetic radiation to take place, it must be
able to couple to a changing (oscillating) electric dipole.
The electric dipole is
= x ex + y ey + z ez

(34.2)

where x = qx, y = qy, z = qz.


The upshot of all this is as far as group theory is concerned is the following
selection rule:
The vibrational coordinates for an IR active transition must have the same
symmetry as either x, y, or z for the particular group.
Example: Water
Recall that the point group symmetry of the water molecule is C2v .
We now need a column of the character table which we have ignored up to this
point.
The character table is
C2v

C2

A1
A2
B1
B2

1
1
1
1 z, x2 , y 2 , z 2
1
1 1 1 xy
1 1
1 1 x, xz
1 1 1
1 y, yz

0v

Functions

The last column describes the symmetry of several important functions for the
point group.

232

Among these functions are x, y, and z.


So we can see immediately that the IR active modes of any molecule having this
point group will be A1 , B1 , and B2 .
The A2 mode is IR forbidden and any vibrations having this symmetry will not
appear in the IR spectrum (or it may appear as a very weak line).
From before we know the modes of water have A1, and B1 symmetry and hence
are all IR active and appear in the IR spectrum

34.2. Raman Spectroscopy


Raman spectroscopy is somewhat dierent than IR spectroscopy in that vibrational frequencies are measured by way of inelastic scattering of high frequency
(usually visible) light.
The light loses energy to the material in an amount equal to the vibrational energy
of the molecules is the sample.
This lose of energy shows up in the scattered light as a new down shifted frequency
from that of the original input light frequency.
Unlike IR absorption which is based on the electric dipole, Raman scattering is
based on the polarizability of the molecule
Roughly speaking the polarizability of a molecule determines how the electron
density is distorted through interaction with an electromagnetic field.

233

The molecular quantity of interest is the polarizability tensor, .


We will not get into tensors in this course except to say the polarizability tensor
elements are proportional to the quadratic functions, x2 , y 2 , z 2 , xy, xz, yz, (or
any combinations thereof).
One can now inspect the character table to determine which modes will be Raman
active.
For the example of water, all modes are Raman active
Rule of Mutual exclusion
Vibrational mode can be both IR and Raman active or inactive
If, however, the molecule has inversion symmetry (contains as a symmetry
element) then no modes will be both IR and Raman active.

234

35. Molecular Rotations


Recall that the three degrees of freedom that described the position of the nuclei
about the center of mass were (R, , ). The R was involved in vibrations. We
now turn our attention to the angular components to describe rotations.
Recall also the Kinetic energy operator for the nuclei in the center of mass coordinates
~2 2
~2 2
~2 2

TN = N =
+ J .
(35.1)
R
2
2R2 R R 2
We will now be concerned only with the angular part,
~2 2
(35.2)
J .
2I
Now, under the Born-Oppenheimer approximation, R is a parameter. For constant
R the rotational energy is given by

Erot =

J(J + 1)~2
J(J + 1)h2
.
=
2R2
8 2 I

(35.3)

This is the so-called rigid rotor energy.


It is common to define
Be

h
8 2 I

(35.4)

as the rotational constant. Then


Erot = J(J + 1)hBe

(35.5)

gJ = 2J + 1

(35.6)

with a degeneracy of

235
235

35.1. Relaxing the rigid rotor


Of course the rigid rotor is not a perfectly correct model for a diatomic molecule.
There are two corrections we will now make
1. Vibrational state dependence:
The R value is dependent on the particular vibrational level.
One defines a rotational interaction constant that depends on the vibrational level, n.

1
Bn Be n +
e ,
(35.7)
2
where e is an empirical rotationalvibrational interaction constant.

2. Centrifugal stretching:
Rotation tends to stretch the diatomic distance R.
This is corrected for by the term
J 2 (J + 1)2 Dc ,
where
Dc

4Be3

2e

(35.8)

(35.9)

is the centrifugal stretching constant.

35.2. Rotational Spectroscopy


A rotational transition can occur in the same vibrational level n. This is called a
pure rotational transition. Alternatively, a rotational transition can accompany a
vibrational transition.
In either case the selection rule for the transition is 4J = 1.

236

It turns out that typical rotational energy gaps are on the order of a few wavenumbers or less.
Thermal energy, kT, at room temperature is about 200 cm1 . This means that
at room temperature the many excited rotational states are populated.
See Handout
The selection rules and the thermalized states combine to yield a multi-peaked
ro-vibrational spectrum.
See Handout

35.3. Rotation of Polyatomic Molecules


There are a few additional details regarding rotations for polyatomic molecules as
compared to diatomics
Of course one could set-up an arbitrary center of mass coordinate system. But
one system is specialthe principle axes coordinate system.
The principle axes coordinate system is the one in which the z-axis is taken to be
along the principle symmetry axis.
The total moment of inertia, I = Ixx + Iyy + Izz
The Hamiltonian in the principle axes system is
#
"
2
x2
z2
Jy2
~
J
J
=
+
+
H
2 Ixx Iyy Izz

(35.10)

237

There are four classes of polyatomic molecules regarding rotations


1. Linear (e.g., carbon dioxide)
Izz = 0, Ixx = Iyy
J2 = Jx2 + Jy2

The Hamiltonian is

2
= ~ J2
H
2Ixx

(35.11)

Erot = hBJ(J + 1),

(35.12)

The rotational energy is

where
B=

h
82 Ixx

(35.13)

2. Symmetric tops (e.g., benzene)


Ixx = Iyy

J2 = Jx2 + Jy2 + Jz2


The Hamiltonian is

#
"
2 J2 + J2
2

~
J
x
y
=
+ z
H
2
Ixx
Izz

(35.14)

The rotational energy is


Erot = hBJ(J + 1) + h(A B)K 2 ,

(35.15)

where

h
,
(35.16)
8 2 Izz
h
(35.17)
B= 2
8 Ixx
and K is the quantum number describing the projection of the angular
momentum onto the z-axis
A=

238

3. Spherical tops (e.g., methane)


Ixx = Iyy = Izz

J2 = Jx2 + Jy2 + Jz2


The Hamiltonian is

2
= ~ J2
H
2Ixx

(35.18)

Erot = hBJ(J + 1),

(35.19)

The rotational energy is

where
B=

h
82 Ixx

(35.20)

4. Asymmetric tops
Ixx 6= Iyy 6= Izz
These are more complicated and we will not discuss them in detail

239

36. Electronic Spectroscopy of


Molecules
The electronic spectra of molecules are quite dierent than that of atoms.
Atomic spectra consist of single sharp lines due to transitions between energy
levels.
Molecular spectra, on the other hand, have numerous lines (bands) due to the
fact that electronic transitions are accompanied by vibrational and rotational
transitions.

36.1. The Structure of the Electronic State


Last semester we saw that under the BornOppenheimer approximation we were
able to write the molecular wavefunction as a product of an electronic part and a
nuclear part.
We found that in doing so the electronic energy level, Ee , was parameterized by
the internuclear distance, R.
Ee as a function of R describe the eective potential for the nuclei.
It had a qualitative shape similar to the Morse potential.

240
240

In the figure below the ground and first excited electronic levels (as a function of
R) are shown.
Note: The potential minima are not at the same value of R for each of the
electronic states.

36.1.1. Absorption Spectra


In absorption spectroscopy, light promotes an electron from the ground electronic
state (and usually from the ground vibrational state too) to the excited electronic
state and any of the excited vibrational states of the excited electronic state.
See Spectroscopy Supplement p1
36.1.2. Emission Spectra
In emission spectroscopy, light demotes an electron from the ground vibrational
state of the excited electronic state to any one of a number of excited vibrational
levels in the ground electronic state.

241

See Spectroscopy Supplement p2


36.1.3. Fluorescence Spectra
All during the process of absorption, the process of is taking place.
See Spectroscopy Supplement p3
As seen in the supplement the fluorescence spectrum is shifted to lower energies
(red shifted) from the absorption spectrum.
This is known as the Stokes shift.
The main stream explanation for the stokes shift is as follows
Light promotes the system from the ground vibrational and ground electronic state to excited vibrational levels in the excited electronic state.
The system then very rapidly (on the order of tens to hundreds of femtoseconds) relaxes to the ground vibrational state of the excited electronic
state.
This process is called
The molecule than emits a photon to drop back down into an excited vibrational state of the ground electronic state.
This requires a lower energy (or more red) photon. Hence the Stokes shift.

242

36.2. FranckCondon activity


We have seen than an electronic tranistion involves not only a change in the
electronic state but also in the vibrational state in general (and in the rotaitonal
state as well, but we will ingore this).
Assuming the electronic transition is allowed one must calculate the probability of
the vibrational transistion as well. This is down by evaulating the FranckCondon
integral.

36.2.1. The FranckCondon principle


When the BornOppenheimer approximation is applied to spectroscopic transitions, one obtains the FranckCondon principle.
The FranckCondon principle states that the nuclei do not move during an electronic transition.
Physically this means that for a particular transition to be FranckCondon active there must be good overlap of the vibrational wavefunctions involved in the
transition.
Mathematically this means that the strength of a transition from i = el,i vib,i
f = el,f vib,f is given by
Z

all
space

2 Z

f
el i =

el
space

vib
space

el,f vib,f
el el,i vib,i ,

(36.1)

243

where
el is the electronic transition dipole. We can separate the integrals as
2 Z
2
Z

el,f
el el,i
vib,f vib,i ,
(36.2)

vib

el
space
space
|
{z
}|
{z
}
if 6=0, allowed

FranckCondon

244

37. Fourier Transforms


As a spectroscopist it is imperative to have a deep understanding of the relationship between time and frequency.
Spectroscopic data is obtained either in the time domain or in the frequency
domain and one should readily be able to look at data in one domain and know
what is happening in the other domain.
One should be familiar with qualitative aspects of this timefrequency relation,
such as if a signal oscillates in time it will have a peak in it frequency spectrum
at the frequency with which it is oscillating.
Furthermore, if the signal decays rapidly it will have a broad spectrum and, conversely, if the signal decays slowly it will have a narrow spectrum. The mathematics which governs these qualitative statements is Fourier transform theory which
we now review.

37.1. The Fourier transformation


The Fourier transformation, =, of a function f (t) will, in this work, by denoted
by a tilde, f(), and is given by
Z

= [f (t)] = f () =
f(t)eit dt.
(37.1)

245
245

The Fourier transformation is unique and it has a unique inverse, =1 , which is


given by
Z
h
i
1
1
it

=
d.
(37.2)
f()e
f () = f (t) =
2
The above two relations form the convention used throughout this work.
Other authors use dierent conventions, so one must take care to know exactly
which convention is being used.
For simplicity the symbol = will be used to represent the Fourier transformation
1
operation, i.e., = [f (t)] = f(). Whereas
h
i the symbol = will represent the inverse

Fourier transformation, i.e., =1 f()


= f(t).

246

Key Equations for Exam 2


Listed here are some of the key equations for Exam 2. This section should not
substitute for your studying of the rest of this material.
The equations listed here are out of context and it would help you very little to
memorize this section without understanding the context of these equations.
The equations are collected here simply for handy reference for you while working
the problem sets.

Equations
Vibrational degrees of freedom
linear: 3N 5
not linear: 3N 6
The so-called rigid rotor energy is
Erot = J(J + 1)hBe .
where
Be

h
8 2 I

(37.3)

(37.4)

is the rotational constant.


247
247

The degeneracy of the rigid rotor is


(37.5)

gJ = 2J + 1
FranckCondon Factor:

vib
space

The Fourier transformation is

vib,f vib,i

= [f (t)] = f() =

f (t)eit dt.

(37.6)

(37.7)

The inverse Fourier transformation is


Z
h
i
1
1
it

d.
=
f()e
f () = f (t) =
2

(37.8)

248

Part VII
Kinetics and Gases

249
249

38. Physical Kinetics


We now turn our attention to the molecular level and in particular to molecular
motion.

38.1. kinetic theory of gases


A microscopic view of gases
Consider a gas of point mass (m), m is the molecular (or atomic) mass
Each particle of mass m has velocity v, hence a momentum of p = mv and
a kinetic energy of KE = 12 mv v = 12 mv 2 .
A sample of N molecules is characterized by its number density n =
From the ideal gas law P V = nRT =
N
PL
= RT
= n
V

N
RT
L

N
.
V

(L is Avogadros number):

Consider the ith particle at position xi = (x, y, z) in coordinate (position) space.


dxi dyi dzi
i
Its velocity is vi = dx
=
, dt , dt . This can represented in velocity space by
dt
dt
the vector vi = (vxi , vyi , vzi ).
The velocities of the particles are characterized by a probability distribution function for velocities F (vx , vy , vz , t) which is in general a function of time, t.
250
250

The number of particles, NVv , having velocities in a macroscopic volume, Vv , in


velocity space is
Z
Z Z Z
NVv = N F (v, t)dv = N
F (vx , vy , vz , t)dvx dvy dvz
(38.1)
Vv

Vv

It is more convenient to switch to spherical polar coordinates in velocity space


(v, , ); n.b., v is simply a magnitude (not a vector)it is the speed.

The probability distribution function then becomes F (v, , , t)


If we choose the origin of our coordinate system to be at the center of mass of the
gas, then for many cases the velocity distribution will be isotropicindependent
of and .
F (v, , , t) = F (v, t).
(38.2)
Furthermore, stationary distributionsthose independent of timeare often encountered.
F (v, , , t) = F (v, , ).
(38.3)

251

We shall consider stationary isotropic distributions F (v). So F (v) represents a


distribution of speeds.
It can be shown from first principles that

32
mv 2
m
F (v) = 4
e 2kb T v 2
(38.4)
2kb T
where kb = 1.380658 1023 is Boltzmanns constant. This is the Maxwells
distribution (of speeds).

38.2. Molecular Collisions


The average speed of a particle can calculated from Maxwells distribution:
Z
Z
m 3 mv2
) 2 e 2kT dv
hvi = v =
vF (v)dv =
v 3 4(
(38.5)
2kT
0
0
s
r

8kT L Lk=R
8RT
=
=
m L Lm=M
M
It will be convenient to define number density as n N
where N is the number
V
N
LP
P
of particles. For an ideal gas (V = nRT
), n =
= RT
.
P
RT
n
|{z}
=L

A simple model for molecular collisions:

252

Particles are hard spheres of radius .


A Particle moving at v sweeps out a cylinder of radius and length v4t =
V = 2 v4t.
See Handout
The number of collisions equals the number of particles with their centers
in V :
number of collisions = n 2 v4t
(38.6)
The collision frequency = n 2 v
For the above model we need to find the average collision frequency. Since the
molecules are moving relative to one another we must find the average relative
velocity, v12 = h|v1 v2 |i
It can be shown that
r
v12 =

collision
16RT
= 2
v =
= 2n 2 v.
M
frequency

(38.7)

From the above expression one defines the mean free path to be

/v
=
2n 2/v

LP
n = RT

RT

2P L 2

(38.8)

Example: Ar at SATP (T = 298 K, P = 1 bar):


v = 380.48

m
,
s

collision
= 5.25 109 s1 ,
frequency
= 72.5 nm

253

39. The Rate Laws of Chemical


Kinetics
Thermodynamics described chemical systems in equilibrium. For the study of
chemical reactions it is important understand systems that can be very far from
equilibrium. For this we turn to the field of chemical kinetics.
We can, from thermodynamics, address the question; Will the reaction occur?
We need kinetics, however, answer the question: How fast will the reaction occur?

39.1. Rate Laws


Consider a general four component reaction
aA + bB = cC + dD

(39.1)

The time dependence of this reaction can be observed by following the disappearance of either of the reactants or appearance of either of the products. That
is,
d[B] d[C] d[D]
d[A]
or
or
or
(39.2)

dt
dt
dt
dt
BUT this is ambiguous because a moles of A reacts with b moles of B and a does
not, in general, equal b. We must account for the stoichiometry.

254
254

We define the reaction velocity as


v=

1 d[I]
vi dt

(39.3)

where vi = a, b, c or d and I = A, B, C, or D.
This definition is useful but must be used with caution since for complicated
reactions all the vs may not be equal. An example of this is

bB cC + dD
aA +
(39.4)
0
b B0 c0 C0 + d0 D0
A rate law is the mathematical statement of how the reaction velocity depends
on concentration.
v = f (conc.)
(39.5)
For the most part, rate laws are empirical.
Many, but certainly not all, rate laws are of the form
v = k[A1 ]xA 1 [A2 ]xA 2 [An ]xA n .

(39.6)

The reaction is said to be of order xAi in species Ai and it is of overall order


P
i xA i .
In general an overall reaction is made up of so called elementary reactions
Reactant

Product

overall rxn

Reactant

Intermediates Product

(39.7)

Note that we shall use an equal sign when talking about the overall reaction and
arrows when talking about the elementary reactions
Example

255

Let
2A + 2B = C + D

(39.8)

be the overall reaction. One possible set of elementary steps could be


elementary rxn

molecularity

A + A A0

Bimolecular

A0 A00

Unimolecular

A00 + 2B C + D

Trimolecular

The rate laws for elementary reactions can be determined from the stoichiometry
molecularity

elementary rxn

rate law

Unimolecular
A Product
v = k[A]
Bimolecular
A + A Product
v = k[A]2
Bimolecular
A + B Product
v = k[A][B] .
Trimolecular A + A + A Product
v = k[A]3
Trimolecular A + A + B Product v = k[A]2 [B]
Trimolecular A + B + C Product v = k[A][B][C]
Conversely, rate laws for overall reactions can not be determined by stoichiometry.
Connection to thermodynamics
Consider the overall or elementary reaction
kf

aA + bB cC + dD
kr

(39.9)

where kf is the rate constant for the reaction to proceed in the forward direction
and kr is the rate constant for the reaction to proceed in the reverse direction.
Now, at equilibrium vf = vb which implies
kf [A]a [B]b = kr [C]c [D]d

(39.10)

256

bringing kr to the LHS and [A][B] to the RHS we get


[C]c [D]d
kf
=
= Kc0
a
b
kr
[A] [B]

(39.11)

where Kc0 is the thermodynamic equilibrium constant.


So, we have succeeded in connecting thermodynamics to kinetics BUT we have
done so through the ratio of rate constants. The velocity of a reaction is lost
in this ratio and hence we still can not determine the speed of a reaction from
thermodynamics.
Examples of rate laws
Consider the (overall) reaction between molecular hydrogen and molecular iodine,
H2 + I2 = 2HI.

(39.12)

The observed rate laws are vf = kf [H2 ][I2 ] and vr = kr [HI]2 . This suggests that the
reaction is elementary. In fact, the reaction is not elementary. Moral: Kinetics
is very much an empirical science.
Next consider the reaction between molecular hydrogen and molecular bromine,
H2 + Br2 = 2HBr.

(39.13)

The observed rate law for this reaction is very complicated,


v=

k[H2 ][Br2 ]1/2


1+

k0 [HBr]
[Br2 ]

this does not obey any common form.


The above two example are seemingly very similar but they have very dierent
observed rate laws. Moral: Kinetics is very much an empirical science.
Objectives of chemical kinetics

257

To establish empirical rate laws


To determine mechanisms of overall reactions
To empirically study elementary reactions
To establish theoretical links to statistical mechanics and quantum mechanics
This involve nonequilibrium thermodynamicsmore dicult
To study chemical reaction dynamics
the dynamics of molecular collisions that result in reactions

39.2. Determination of Rate Laws


Concentrations c(t) are measured not rates. To obtain the rate from the concentration we must take its time derivative dc(t)
. That is we must measure c(t) as a
dt
function of time and find the rate of change of this concentration curve.
The rates of chemical reactions vary enormously from sub-seconds to years. Consequently no one experimental technique can be used.
For slow reactions (hrs/days) almost any technique for measuring the concentration can be used.
For medium reactions (min) either a continuous monitoring technique or a
stopping technique can be used
A stopping technique used rapid cooling or destruction of the catalysts
to stop a reaction at a given point.
Very fast (sec/subsec) reactions cause problems because the reaction goes
faster than one can mix the reactants.

258

39.2.1. Dierential methods based on the rate law


Methods based directly on the rate law rely on the determination of the time
derivative of the concentration.
The main problem with such a method is that randomness in the concentration
measurements gets amplified when taking the derivative.
1. Method of initial velocities
for v = k[A]x [B]y rate laws.
initially v0 = kax by where a and b are the initial concentrations of A
and B respectively
taking the log of both sides gives lnv0 = ln[kax by ] = ln k + x ln a + y ln b
a and b can be varied independently so both x and y can be determined.
problems
1. if the concentration drops very sharply
2. if there is an induction period
2. Method of isolation
for v = k[A]x [B]y rate laws
start with initial concentrations a and b equal to the stoichometry; this
gives the overall order of x + y
flood with, say, A so v kax [B]y
39.2.2. Integrated rate laws
The above dierential methods look directly at the rate law which is a dierential
equation. The dierential equation is not solved.

259

We now solve the dierential equations to yield what are called the integrated
rate law.
The dierential equations (rate law) and their solutions (integrated rate law) are
simply listed here for a few rate laws.
type
rate lawa)
integrated rate lawa)
1st order
2nd order
nth orderb)

1
vi
1
vi
1
vi
1
vi

d[I]
dt
d[I]
dt
d[I]
dt
d[I]
dt

= k[I]
= k[I]2
= k[I]n

[I] = [I0 ]evi kt


1
= [I10 ] vi kt
[I]
1
[I]n1

1
[I0 ]n1

(n 1)vi kt

enyzme
= kmk[I]
km ln [I[I]0 ] + ([I0 ] [I]) = vi kt
+[I]
a)
[I] is the concentration of one of the reactants in an elementary reaction and
vi is the stoichiometric factor for [I] (n.b., vi is a negative number).
b)
The order need not be an integer. For example n = 3/2 is a three-halves
order rate law.

260

40. Temperature and Chemical


Kinetics
40.1. Temperature Eects on Rate Constants
An empirical rate constant was proposed by Arrhenious:
Ea
d ln k
=
or
dT
RT 2
Ea
d ln k
=
,
d(1/T )
R

(40.1)
(40.2)

where Ea is the Arrhenious activation energy.


Integration of the above yields
ln k = ln A

Ea
Ea
= k = Ae RT
RT

(40.3)

(A is the constant of integration). This is the Arrhenious equation


Recall the equilibrium constant can also be obtained from kinetics
Kc0 =

kf
' Ka .
kr

(40.4)

Now, take the log of this:

kf
= ln kf ln kr .
ln Ka = ln
kr

(40.5)

261
261

Substituting the Arrhenious equation for the rate constants gives

i
h
Ea
Ear
RTf
ln Ar e RT
ln Ka = ln Af e

Ear Eaf
Af
= ln
+
Ar
RT

(40.6)

40.1.1. Temperature corrections to the Arrhenious parameters


The Arrhenious parameters A and Ea are constants.
Theoretical approaches to reaction rates predict rate constants of the form
0

k = aT j eE /RT .

(40.7)

Forcing this to coincide with the Arrhenious implies


Ea = E 0 + jRT

(40.8)

A = aT j ej

(40.9)

and

We can verify this by starting with the Arrhenious equation and substituting the
above expressions,
Ea

k = Ae RT = aT j ej e

E 0 +jRT
RT

E 0

E 0

j
= aT j e/j e/
e RT = aT j e RT

(40.10)

40.2. Theory of Reaction Rates


Simple collision theory (SCT)
Bimolecular reactions (A,B)
Reaction rate determined by molecular collisions

262

Collision frequency for AB collisions


s
8RT
zAB = AB L2
[A][B]
L
where

mA mB
mA +mB

(40.11)

is the reduced mass and AB is the collision diameter.

zA B
L

The maximum reaction velocity is vmax =


reaction velocity will be less because

, but intuitively the actual

the ability to react depends on orientation = a steric factor p


a minimum amount of collisional energy is required= eEm in /RT
The actual reaction velocity is
pzAB e
v=
L

Em in
RT

(40.12)

The rate constant for a bimolecular reaction is


v
k=
[A][B]

(40.13)

so SCT predicts
k=

pzA B e
L

Em in
RT

[A][B]

= p AB L

8RT Em in
e RT
L

(40.14)

263

Comparison to the (temperature corrected) Arrhenious equation suggests


s
8RT 1
e2
(40.15)
A = p AB L
L
and

1
Ea = Emin + RT
2

(40.16)

Activated complex theory (ACT)


An intermediate active complex is formed during the reaction, e.g.,
A + B (AB) products.

(40.17)

ACT is not limited to bimolecular reactions.

The active complex is a state in the thermodynamic sense, thus we can apply
thermodynamics to it.
For the above example, the equilibrium constant is defined as
Ka =

a low []
'
aA aB conc. [A][B]

(40.18)

264

Definition: transmission factor, f


accounts for the fraction of activated complex that becomes product.
From statistical mechanics, it can be shown that f = kb T /h where kb
is Boltzmanns constant and h is Plancks constant.
The reaction rate constant for reactants going to products for ACT is
kb T
K
h a

(40.19)

4G = RT ln Ka

(40.20)

k = f Ka =
Thermodynamics tells us that

which can be written as


Ka = e

4G
RT

= e

4H
RT

4S
R

(40.21)

where 4G = 4H T 4S .
The ACT reaction rate constant now becomes
k=

kb T 4H 4S
e RT e R .
h

(40.22)

This is Eyrings equation

40.3. Multistep Reactions


Up to now, the reactions we have studied have been single step reactions.
In general, there is many steps from initial reactants to final products.
Reactions may occur in series or in parallel or both, in what is called a reaction
network.
Parallel reactions:

265

Parallel reactions are of the form, for example,


k

A + B1 1 C

(40.23)

k2

A + B2 D
The rate constant for the disappearance of [A] is simply the sum of the two
rate constants: k = k1 + k2

Series reactions:
Series reactions necessarily include and intermediate product. They are of
the form
k
k
A 1 B 2 C
(40.24)
The concentrations of A, B and C are determined by the system of dierential equations:

d[A]
= k1 [A]
dt
d[B]
= k1 [A] k2 [B]
dt
d[C]
= k2 [B],
dt

which, when solved yields


[A] = [A0 ]ek1 t

k1 [A0 ] k1 t
[B] =
ek2 t
e
k2 k1

k2 ek1 t k1 ek2 t
[C] = [A0 ] [A] [B] = [A0 ] 1
k2 k1

See in class animation

266

40.4. Chain Reactions


Chain reactions are reactions which have at least one step that is repeated indefinitely. The simplest chain reactions have three distinct steps (discussed below)
Chain reactions are extremely important in polymer chemistry
Steps of a chain reaction
1. Initiation: Typically a molecule M reacts to form some highly reactive radical
M R.

267

2. Propagation: The radical formed in the initiation step reacts with some so
molecule M0 to form another molecule M00 and another radical R0 . This step
repeats an indefinite number of times.
R+M0 M00 + R0 .
3. Termination: The radicals interact with each other or with the walls of the
container to forma stable molecule
R0 +R0 M000
or
R0 +wall removed

268

41. Gases and the Virial Series


Unlike liquids and solids, a particular particle has much less significant interactions
with the other particles.
This simplifies the theoretical treatment of gases.
We will now look in detail at the gases.

41.1. Equations of State


Recall from last semester several of the equations of states for gases.
The ideal gas equation of state
(41.1)

P V = nRT.
The equation of state can also be expressed in term of density =
=

mP
.
nRT

m
V

(41.2)

The van der Waals gas equation of state


P =
or

n2 a
nRT
2
V nb
V

(41.3)

a
RT
2,
(41.4)
Vm b Vm
where the parameter a accounts for the attractive forces among the particles
and parameter b accounts for the repulsive forces among the particles
P =

269
269

Berthelot

nRT
n2 a
a
RT

=
2
V nb T V
Vm b T Vm2

P =

Dieterici

(41.5)

an

RT e RT Vm
nRT e RT V
=
P =
V nb
Vm b

(41.6)

nRT
n2 a
a
RT

=
V nb
Vm b
T V (V nb)
T Vm (Vm b)

(41.7)

Redlich-Kwang
P =

41.2. The Virial Series


Definition: Compressibility Factor: z =

PV
nRT

P Vm
.
RT

z is unity for an ideal gas because for such a gas P V = nRT.


For a real gas z must approach unity upon dilution ( Vn 0).
z can be expended in a power series called the virial series.
The virial series in powers of
z = 1 + B(T )
or

1
z = 1 + B(T )
Vm

n
V

is
+ C(T )

n 2

1
+ C(T )
Vm

+ D(T )

n 3

+ ,

(41.8)

+ .

(41.9)

1
+ D(T )
Vm

B(T ), C(T ), etc. are called the virial coecients.


Conceptually B(T ) represents pair-wise interaction of the particles, C(T ) represents triplet interactions, etc.

270

41.2.1. Relation to the van der Waals Equation of State


Recall the van der Waals equation
P =
multiply both sides by

Vm
RT

a
RT
2
Vm b Vm

(41.10)

to get

V
/ a
//
T
Vm R
P Vm
=
m /2
RT
R
//
T Vm b RT Vm
a
Vm

=
Vm b RT Vm
1
a
=

b
RT Vm
1 Vm
but

P Vm
RT

(41.11)

= z so
z=

The first term is of the form

1
1x

1
a

.
b
RT Vm
1 Vm

(41.12)

which has the power series expansion

1
= 1 + x + x2 + .
1x
Therefore
b
a
+1+
+
z=
RT Vm
Vm
the first term is proportional to
in the series expansion, hence

1
Vm

b
Vm

(41.13)

+ .

and so it can be combined with the

a 1
b
z =1+ b
+
+ .
RT Vm
Vm

(41.14)
1
Vm

term

(41.15)

This series can now be compared term by term to the virial series to give expression
for the virial coecients:

a
, C(t) = b2 , D(T ) = b3 , etc.
(41.16)
B(T ) = b
RT

271

41.2.2. The Boyle Temperature


The temperature at which B(T ) = 0 is called the Boyle temperature, Tb .
The virial series at Tb becomes
2

3
1
1
1
z(T = Tb ) = 1 + 0
+ C(T )
+ D(T )
+
Vm
Vm
Vm
3
2
1
1
+ D(T )
+ .
(41.17)
= 1 + C(T )
Vm
Vm
2
The lowest order correction are now V1m . The gas behaves more like an ideal
gas at Tb then for other temperatures.
41.2.3. The Virial Series in Pressure
One can also expand the compressibility factor in pressure
z = 1 + B 0 (T )P + C 0 (T )P 2 + D0 (T )P 3 + .

(41.18)

The relation of this expansion to the one in V1m can be obtained. One finds (see
homework)
B(T )
B 0 (T ) =
,
(41.19)
RT

272

C(T ) B(T )2
(RT )2

(41.20)

D(T ) 3B(T )C(T ) 2B(T )3


(RT )3

(41.21)

C 0 (T ) =
and
D0 (T ) =

41.2.4. Estimation of Virial Coecients


The virial coecients can be estimated using empirical equations and tabulated
parameters.

Estimates based on Beattie-Bridgeman constants:


c
A0
3,
RT
T
A0 a
B0 c
B0 b 3 ,
C(T ) =
RT
T
B0 bc
D(T ) =
.
T3
B(T ) = B0

(41.22)
(41.23)
(41.24)

where A0 , B0 , a, b, c are tabulated constants


Estimates based on critical values (we will discuss critical values shortly, for
now treat them as empirical parameters):

9RTc
6Tc2
B(T ) =
1 2 .
(41.25)
128Pc
T

273

42. Behavior of Gases


42.1. P, V and T behavior
We shall briefly consider the P, V and T behavior of dense fluids (e.g., liquids).
Taking volume as a function of P and T, we consider the total derivative

V
V
dT +
dP.
(42.1)
dV (T, P ) =
T P
P T
We can change this from a extensive property equation to an intensive property
equation by dividing by V :

1 V
1 V
dV
=
dT +
dP.
V
V T P
V P T
| {z }
| {z }

is the coecient of thermal expansion.

At a given pressure, describes the change in volume with temperature.


Positive means the volume of the fluid increases with increasing temperature.

T is the isothermal compressibility


At a given temperature, T describes the change in volume with pressure.
Positive T means the volume of the fluid decreases with increasing pressure.
T is dierent from z, the compressibility factor.
274
274

42.1.1. and T for an ideal gas


As an exercise we shall calculate and T using the ideal gas equation of state
(n.b., it is, of course, absurd to treat a liquid as an ideal gas). Starting with the
ideal gas law: V = nRT
.
P

!

nRT
1 V
1 nRT
1
P
2
T =
(42.2)
=
=
V
P T
V
P
V
P
T

nR
//
T
1 /
1
1 nRT
=
=
(P V ) P
P
P
/n R
//
T
| {z }
=nRT

and

1
=
V

V
T

1
=
V


!
nRT
1
1
P
=
/n R
/ =
T
VP
T
|{z}
P

(42.3)

=n
/R
/T

42.1.2. and T for liquids and solids

In general, the compressibility and expansion of liquids (and solids) are very small.
So one can expand the volume in a Taylor series about a known pressure, P0 .
At constant T

V
V
(P P0 ) +
(P P0 )2 +
V (P ) = V0 +
P
P T
| {z T}

(42.4)

V0 T

so,

V (P ) V0 [1 T (P P0 )] .

(42.5)

This approximation is quite good even over a rather large pressure range (P P0 =
100 atm or so).

275

Likewise at constant P

V
V
(T T0 ) +
(T T0 )2 +
V (T ) = V0 +
T P
T T
| {z }

(42.6)

V0

so,

V (T ) V0 [1 + (T T0 )] .

(42.7)

As one final point, we can apply the cyclic rule for partial derivatives to determine
the ratio T :
V

P
T P cyclic
= V =
(42.8)
T
P T rule T V

42.2. Heat Capacity of Gases Revisited


This section is a review from the first semester with an additional example beyond
the ideal gas.
42.2.1. The Relationship Between CP and CV
To find how CP and CV are related we begin with

H
,H = U + PV
CP =
T P
so

(42.9)

(U + P V )
U
V
=
+P
(42.10)
CP =
T
T P
T P
P


is not CV we need U
. Use an identity of partial derivatives
note U
T P
T V

U
T

U
T

U
V

V
T

(42.11)

276

thus

V
U
V
+
+P
(42.12)
CP =
V T T P
T P
V

U
V
= CV +
+P .
T P
V T
P
U
=
T
P . Then
Recall the expression for internal pressure V
T V
T

U
T

CP = CV +

V
T

P
T
P
/ + P/
T V
P

Finally

CP = CV + T

V
T

P
T

(42.13)

(42.14)

For solids and liquids:

V
T

= V ,

P
T

(42.15)

so

2 T V
(42.16)
T
For gases we need the equation of state which often is conveniently explicit in P
or V but not both
CP = CV +

1. Explicit in P : Replace

V
T

P
T

2. Explicit in V : Replace

V
with T
P
V

(42.17)

T P

with V

(42.18)

P T

277

Examples
1. Ideal gas (equation of state: P V = nRT ): This equation is easily made
explicit in either P or V so we dont need any of the above replacements

P
V
(42.19)
CP = CV + T
T P T V
nRT
nR nR
=
= nR
= CV + T
P V
PV
Thus CP = CV + nR or CP m = CV m + R
2. One term viral equation (equation of state: V = nRT
+ nB). This is explicit
P
in V so use case 2 above

P
V
V
T P
V

CP = CV + T
= CV T
(42.20)
T P T V
T P P T
The partial derivatives are

V
nR
+ nB 0 ,
=
T P
P

so

T P
=
V
P T

Thus
CP

V
P

nRT
,
P2

nR
P

+ nB 0 /
n P (R + P B 0 )
=
.
nRT
/n RT
P2

0
nR
P
(R
+
P
B
)
+ nB 0
= CV + /
T
P
RT
/

2
P B0
= CV + nR 1 +
R

or
CP m

P B0
= CV m + R 1 +
R

(42.21)

(42.22)

(42.23)

(42.24)

278

42.3. Expansion of Gases


Expanding gases do work:
w =

V2

Pex dV

(42.25)

V1

As we learned last semester the value of w depends on Pex during the expansion.
Recall that if the expansion is reversible, there is always an intermediate equilibrium throughout the expansion. Namely Pgas = Pex . So,

nRT
)
V

V2

Pgas dV

(42.26)


V2
nRT
dV = nRT ln
V
V1

(42.27)

wrev =
For an ideal gas (P =

V1

this becomes

wrev =

V2

V1

Also recall that wrev is the maximum possible work that can be done in an
expansion. wrev = wmax .
42.3.1. Isothermal and Adiabatic expansions
We shall consider two limits for the expansion of gases
1. Isothermal expansion T is constant
2. Adiabatic expansion q = 0.

Isothermal expansion
For the case of a ideal gas, U (T, V ) = U(T ) (independent of V ). So for
isothermal expansion 4U = 0 = q + w = q = w.

279

Adiabatic expansion
Since q = 0, dU = dw = Pex dV = P dV (reversible).
For an ideal gas

dU = P dV =

nRT
dV
V

(42.28)

42.3.2. Heat capacity CV for adiabatic expansions


Considering an ideal gas going adiabatically from (T1 , V1 ) to (T2 , V2 ).
Recall

U
CV =
= dU = CV dT
T V

(42.29)

So from above

CV dT
nRdV
nRT
dV =
=
V
T
V
Going from (T1 , V1 ) to (T2 , V2 ):
CV dT =

T2

T1

CV
dT =
T

V2

V1

nR
dV.
V

(42.30)

(42.31)

If CV (T ) is reasonably constant over the internal T1 to T2 then this is approximately




V2
T
2
= nR ln
(42.32)
CV ln
T1
V1
where CV = 1 (CV (T1 ) + CV (T2 )) . Or, in terms of molar heat capacity
2

T2
CV m ln
T1

V2
= R ln
V1

(42.33)

280

42.3.3. When P is the more convenient variable


What if P is the more convenient variable? Then use H instead of U
Let us still consider an adiabatic expansion
H = U + P V, dH = dU + P dV + V dP (because both P and V can, in general,
change)

/ dV/ + V dP
dH = dq + dw + P

(42.34)

dH = V dP.
Now,
CP =

H
T

For an ideal gas this becomes

= dH = Cp dT = V dP

Cp dT =
Going from (T1 , P1 ) to (T2 , P2 ):
Z T2
T1

(42.35)

nRT
dP
P

CP
dT =
T

P2

P1

(42.36)

nR
dP.
P

(42.37)

If CP (T ) is reasonably constant over the internal T1 to T2 then this is approximately




P2
T
2
= nR ln
(42.38)
CP ln
T1
P1
where CP = 1 (CP (T1 ) + CP (T2 )) . Or, in terms of molar heat capacity
2

T2
CP m ln
T1

P2
= R ln
P1

From the above two cases





R
R
P2
V2
T2
= ln
= ln
ln
T1
P1
V1
CP m
CV m

(42.39)

(42.40)

281

So

P2
ln
P1


V2
CP m
= ln
V1
CV m
| {z
}

(42.41)

hence

Thus

P2
ln
P1

V2
= ln
V1

P2
P1

V1
V2

V1
= ln
V2

P2 V2 = P1 V1 ,

V1
= ln
V2

(42.42)

(42.43)

but Pi Vi are arbitrary so this implies P V = constant (** NOTE: The axes
should be reversed **)

42.3.4. Joule expansion


Consider a gas expanding adiabatically against a vacuum (Pex = 0). In this case
q = 0 (adiabatic) and w = 0 (since dw = Pex dV ).

282

This implies
4U = q + w = 0.

(42.44)

Internal energy is constant.


We want to find
Identity:

T
V

T
V

U
U
T
1
=
=
U V V T
CV V T
| {z }

(42.45)

1/CV

For an ideal gas V T = 0 (since U(T, V ) = U(T )). Thus in as much as the
T
gas can be considered ideal V
= 0. That is, for Joule type expansion the
U
temperature of the gas does not change. For real gases this is not strictly equal
to zero.
42.3.5. Joule-Thomson expansion
Consider the adiabatic expansion as illustrated by the figure below

283

The work done on the left is


wL = P1 4V = P1 (0 V1 ) = P1 V1 .

(42.46)

The work done on the right is


wR = P2 4V = P2 (V2 0) = P2 V2 .

(42.47)

4U = U2 U1 = wL + wR = P1 V1 P2 V2

(42.48)

U2 + P2 V2 = U1 + P1 V1 H2 = H1

(42.49)

Now,

Thus

For Joule-Thomson expansion the enthalpy is constant.


We want to find
Identity:

T
V

T
P

. (the Joule-Thomson coecient).


H
T
1 H
=
=
=
H P P T
CP P T
| {z }

(42.50)

1/CP

284

Recall the useful identity

Thus

H
P

=V T

V
T

(42.51)


V + T V
T P
=
CP

(42.52)

Example: The one term virial equation: (equation of state P V = nRT + nB)

nRT
1 nRT
0
nB +
+ nT B
=
CP
P
P
B + T B 0
.
=
CP m

(42.53)

Limts:
Low T : B 0 is positive and B is negative, so is positivethe gas cools upon
expansion
High T : B 0 is nearly zero and B is positive, so is negativethe gas warms
upon expansion
The Joule-Thomson inversion temperature is the temperature where = 0.

285

43. Entropy of Gases


43.1. Calculation of Entropy
Entropy must be calculated along reversible paths. This is not a problem though
since entropy is a state function.
Entropy change for changes in temperature.
At constant V :
dU = dq + dw

dq=CV dT

dU = CV dT, but also dU = T dS. So


Z T2
CV
CV
dT = 4S =
dT.
dS =
T
T
T1

(43.1)

At constant P : (use H = U + P V instead of U)


dq=C dT

P
dH = dU +P dV +V dP = dqP dV +P dV +V dP . So dH = dq =
dq=T dS
dH = CP dT, but also dH = T dS. So
Z T2
CP
CP
dT = 4S =
dT.
(43.2)
dS =
T
T
T1

Isothermal expansion of an ideal gas (P V = nRT ):


Recall that for isothermal expansion of an ideal gas dU = 0 = T dS P dV
dS = P TdV .
286
286

Using the equation of state


nRdV
= 4S =
dS =
V

V2

V1

V2
nR
dV = nR ln .
V
V1

(43.3)

Using the equation of state to express V1 and V2 in terms of P1 and P2 .

/
n R
//
T

V2
P2
dS = nR ln
= nR ln P2
= nR ln .
V1
P1
/
n R
//
T

(43.4)

P1

If two variables change in going from the initial to final states break the path into
two paths in which only one variable changes at a time.
Entropy of Mixing of an ideal gas

Since the gas is ideal, there are simply two separate equations:
4SA = nA R ln

VA + VB
,
VA

4SB = nB R ln

VB + VA
VB

(43.5)

and
4Smix = 4SA + 4SB

(43.6)

287

Recall Avogadros principle: n V for an ideal gas. So.

nA + nB
nB + nA

+ nB ln
4Smix = R nA ln
= R (nA ln XA + nB ln XB )
nA
nB

| {z }
| {z }
1/XA

1/XB

(43.7)

43.1.1. Entropy of Real Gases


Consider the question: How does S S ideal as P 0 ?
Use Maxwell relation
nRT
+ nB.
P
So

Hence

P T

S
P

V
T

V
=
T

and single term viral equation, V =

nR
nB 0
P

U
nR
P2

0
nB dP =
dS =
S2 S1 = nR ln
nB 0 (P2 P1 )
P
P1

(43.8)

(43.9)

For an ideal gas B 0 = 0, so


S2ideal S1ideal = nR ln

P2
P1

(43.10)

Thus
S2 S1 = S2ideal S1ideal nB 0 (P2 P1 )

(43.11)

Letting P1 0 and P2 P (Standard pressure 1 bar), this becomes


ideal

S2 /
S1

ideal

= S2ideal /
S1

nB 0 (P2 P1 )

(43.12)

Defining S2ideal , P2 P as S . So,


S(P ) = S nB 0 P

(43.13)

288

The entropy at any P and T can be obtained expresses as


S(T, P ) = S ideal (T, P ) nB 0 P
Thus
S(T, P ) = S (T ) nR ln

P
nB 0 P
P

(43.14)

(43.15)

289

Key Equations for Exam 3


Listed here are some of the key equations for Exam 3. This section should not
substitute for your studying of the rest of this material.
The equations listed here are out of context and it would help you very little to
memorize this section without understanding the context of these equations.
The equations are collected here simply for handy reference for you while working
the problem sets.

Equations
The Maxwells distribution of speeds is
F (v) = 4

m
2kb T

The average speed of a particle is


hvi =
The mean free path is

32

mv2

e 2kb T v2 .

8RT
M

RT
=
2P L 2

(43.16)

(43.17)

(43.18)

290
290

The reaction velocity is


v=

1 d[I]
vi dt

(43.19)

The relation between the rate constant and the thermodynamic equilibrium
constant is
kf
Kc =
(43.20)
kr
The Arrhenious equation

Ea

k = Ae RT

(43.21)

Important thermodynamic relation:


4G = 4H T 4S

(43.22)

Eyrings equation is
k=

kb T 4H 4S
kb T 4G
e RT =
e RT e R
h
h

(43.23)

The van der Waals gas equation of state:

Compressibility Factor:

P =

a
RT
2.
Vm b Vm

(43.24)

z=

P Vm
PV
=
.
nRT
RT

(43.25)

The virial series is



2
3
1
1
1
z = 1 + B(T )
+ C(T )
+ D(T )
+ .
Vm
Vm
Vm

(43.26)

Relation between heat capacities for an ideal gas:


CP m = CV m + R

(43.27)

291

Part VIII
More Thermodyanmics

292
292

44. Critical Phenomena


44.1. Critical Behavior of fluids
The point on the top of the coexistence curve is called the critical point. It is
characterized by a critical temperature, Tc , and a critical density c .

Law of rectilinear diameters: The average density [ave = 12 (liq + vap )] is


linear in temperature.

293
293

44.1.1. Gas Laws in the Critical Region

The vapor pressure of a substance is taken from the gas laws as the pressures
where A1 = A2 in the above figure.
Simple gas laws do not work well near critical points.

294

44.1.2. Gas Constants from Critical Data


Consider the van der Waals equation at the critical point (Pc , Tc , Vmc )
Pc =

a
RTc
2 .
Vmc b Vmc

(44.1)

dP
d P
There is an inflection point ( dV
= 0, dV
2 = 0) at the critical point. So, setting
m
m
the first and second derivatives at the critical point equal to zero we get

RTc
2a
dP
=
+ 3 =0
(44.2)

2
dVm c (Vmc b)
Vmc

2RTc
6a
d2 P
=
4 =0

2
3
dVm c (Vmc b)
Vmc

and

(44.3)

solving these three equations for Pc , Tc and Vmc gives

(44.4)

Vmc = 3b,
8a
,
Tc =
27bR
a
.
Pc =
27b2

(44.5)
(44.6)

These values can be used to find the compressibility factor, z, at the critical point
zc =

Pc Vmc
3
= = 0.375.
RTc
8

(44.7)

Notice that both a and b whose values depend on the particular gas have dropped
out. That is (for the van der Waals Equation) zc = 0.375 for all gases.
The other equations of state give similar results
van der Waals Berthelot
Dieterici
Redlich-Kwong
zc

3/8 = 0.375

3/8 = 0.375 2/e2 ' 0.27

0.33

295

44.2. The Law of Corresponding States


We have found that zc is predicted by the equations of state to be independent of
the particular gas. This is actually not too far from the truth experimentally.
One can define unitless reduced variables Tr = T /Tc , Pr = P/Pc , and Vr = V /Vc .
r Vr
Then zr = PRT
.
r
zr is a universal functionit is nearly the same for all gasses.
See Fig. 1.18 Laidler&Meiser

44.3. Phase Equilibrium


Consider a homogeneous substance consisting of two phases and at a constant
T and V.
Suppose some amount of material, dn, goes from
(dA )T = P dV dn
(dA )T = P dV + dn
= 0 since V is constant

(dA)T,V = P

z
}|
{
(dV + dV )

+ dn

For a spontaneous process A deceases (dA < 0)

At equilibrium dA = 0. This implies = is the condition for equilibrium.


When , denote liquid (or solid) and vapor phases, then for a given T , the
pressure of the system when = is the called the vapor pressure of the
material at temperature T.

296

For phase changes at constant T and P then (dG)T,P = dn. So again


= is the condition for equilibrium.
44.3.1. The chemical potential and T and P
How does vary with T and P ?
Generally for homogeneous substances,
dG = SdT + V dP + dn
Now,

(44.9)

G
=
=
=
.
n T
T n
T P

(44.10)

G
S=
T
So,

S
n

But S = nSm (T, P ) so,

P,T

(44.8)

Similarly,

P,n

= Sm .

(44.11)

= Vm .

(44.12)

Now the total dierential of is


S

m
z }|m {
z }|
{

dT +
dP
d(T, P ) =
T P
P T
d(T, P ) = Sm dT + Vm dP

(44.13)

297

44.3.2. The Clapeyron Equation


At equilibrium = so,
Sm dT + Vm dP = Sm dT + Vm dP
Now

Sm Sm
4 Sm
dP
=
=
dT
Vm Vm
4 Vm

4S= 4H
T

This is the Clapeyron Equation

4 Hm
T 4 Vm

4 Hm
dP
=
dT
T 4 Vm

(44.14)

(44.15)

(44.16)

44.3.3. Vapor Equilibrium and the Clausius-Clapeyron Equation


The above Clapeyron equation applies to any phase transition; consider the liquidvapor phase transition.
Now
4v V = Vm,vap Vm,liq ' Vm,vap

(44.17)

Assuming the vapor phase obeys the ideal gas equation of state,
4v V =

RT
P

(44.18)

Substituting this into the Clapeyron equation gives


4v Hm
4v Hm P
dP
=
=
RT
dT
RT 2
T P

(44.19)

Collecting the T s on one side of the equation and the P s on the other we get
4v Hm dT
dP
=
P
R T2
Now we identify

dP
P

= d(ln P ) and

dT
T2

(44.20)

= d(1/T ) so this becomes

d(ln P ) =

4v Hm
d(1/T )
R

(44.21)

298

Rearranging again leads to


4v Hm
d(ln P )
=
d(1/T )
R

(44.22)

This is the Clausius-Clapeyron equation.

44.4. Equilibria of condensed phases


Examples
Solidliquid
icewater, most other common liquids
Solidsolid
rhombic sulfurmonoclinic sulfur
grey tinwhite tin
graphitediamond

For example a diamond at STP is metastable with respect to graphite.


A diamond is not forever!
At equilibrium = this implies (for incompressible liquids and solids)

+ Vm (P P ) = + Vm (P P )

(44.23)

This can be rearranged so that terms independent of pressure (the standard chemical potentials) are one side and the terms that depend of pressure are on the other
side

(44.24)
= (Vm Vm ) (P P )

299

Thus for any given T only one P allows for equilibrium.


Recall the Clapeyron equation
4f Hm
Hm Hm
dP
=
=
dT
T 4f Vm
T (Vm Vm )

(44.25)

We make the good approximation that 4f Hm is independent of T and solve the


Clapeyron equation
Z
4f Vm dP
Tf
4f Vm (P P )
dT
=
ln =
(44.26)
4f Hm
4f Hm
Tf
T
where Tf is the freezing temperature at standard pressure (1 bar).

44.5. Triple Point and Phase Diagrams


Definitions
Phase Diagram: A graph of P vs. T for a system which shows the lines
of equal chemical potential
Critical Point: The terminal point of the liquid-vapor line. At temperatures above the critical point there is no distinction between vapor and
liquid.
Triple Point: The point where all three phases coexist in equilibrium:
solid = liq = vap

(44.27)

300

45. Transport Properties of Fluids


Transport properties of matter deal with the flow (or flux) of some property along
a gradient of some other property.
Flux: movement of something through a unit area.
We now consider three transport properties of fluids:
1. Diusion: The flux of material down a concentration gradient
2. Viscosity: The flux of momentum down a velocity gradient
3. Thermal Conductivity: The flux of energy down a temperature gradient

See Transport Phenomena handout

45.1. Diusion
At equilibrium concentration on a bulk solution will be uniform.
So if there exists a concentration gradient there will be a net flux, J, of material
from high concentration to low concentration so as to establish an equilibrium.
J=

1 dn
A dt

(45.1)

301
301

The flux of material through a plane depends on the concentration dierence


J = D

dC
1 dn
dC
=
= D
dx
A dt
dx

where D is the diusion constant


dC
1 dn
= D
A dt
dx

(45.2)

This is Ficks first law of diusion (in one dimension).


The change in concentration in a lamina between x and dx with time is given by
the flux in minus the flux out of the lamina:
J(x) J(x + dx)
J
C
=
=
t
dx
x

(45.3)

Using Ficks first law for J


C
C
=
D
.
t
x x
If D is truly constant we get Ficks second law of diusion:
2C
C
=D 2.
t
x

(45.4)

(45.5)

302

The solution of this partial dierential equation depends on the boundary conditions. Numerous methods of solution exist for this equation but they are beyond
the scope of the course.
The solution for two special boundary conditions are of interest and will simply
be presented here without derivation
1. Point source solution

x2
C0
C(x, t) =
e 4Dt
2 Dt

(45.6)

2. Step function solution


#
Z x
4Dt
1
1
2

ey dy
C(x, t) = C0
2
0

x
1
C0 1 erf
=
2
4Dt

1
x
=
C0 erfc
2
4Dt
"

(45.7)

where erf and erfc are tabulated functions respectively called the error function and complementary error function.

45.2. Viscosity
Viscosity, , is the resistance to dierential fluid flow, i.e., The tendency of a
liquid to flow at the same velocity throughout.

303

dv
The frictional (viscous) force is F = A dx
. (The units of are
g
= cms .)

mass
.
lenghttime

1 poise

Poiseuilles Formula
Applies to Laminar (nonturbulent) flow
For a liquid flowing trough a tube (radius r, length l), the volume of flow
4V in time 4t is
r4 4P
4V
=
(45.8)
4t
8l
where 4P is the driving pressure, i.e., the dierence in pressure on either
side of the tube.
For a gas

4V
r2
=
4t
16l

Pi2 Pf2
P0

(45.9)

where Pi is the inlet pressure, Pf is the outlet pressure and P0 is the pressure
at which the volume is read.

Stokes law: spheres falling through fluids

304

The frictional force (exerted upwards) is proportional to velocity: Ff = fv.


Stokes showed f = 6r
Gravitational force (exerted downwards): Fg =
gravitational acceleration (9.8 m/s2 ).

4r3
( 0 )g,
3

where g is the

Terminal velocity is reached when Ff + Fg = 0 giving


f vterm +
vterm =

4r3
( 0 )g = 0
3

(45.10)

4r3 ( 0 )g
3f

using f = 6r
vterm

4
/ r/3 ( 0 )g 2r2 ( 0 )g

=
=
9
3 6
/ r
/

(45.11)

Related to diusion constant:

D=

kT
f

f =6r

kT
6r

(45.12)

45.3. Thermal conductivity


(This section closely follows parts of chapter 8 in Transport Phenomena by R.B.
Bird, W.E. Stewart and E. N. Lightfoot Wiley New York 1960)
The thermal conductivity, , of a material is a measure of the tendency of energy
in the form of heat to flow through the material.
Consider a slab of solid material of area A between two large parallel plates a
distance D apart. The plates are held at constant but dierent temperatures T1
and T2 (T1 > T2 ) for a suciently long time that a steady state exists.

305

Under such conditions, a linear steady state temperature distribution across the
material is established. And a constant rate of heat flow dq
is needed to maintain
dt
the temperature dierence 4T = (T1 T2 )
4T
1 dq
=
.
(45.13)
A dt
D
If we take the limit where D becomes infinitesimally small (D dx) we obtain a
dierential form of this equation:
1 dq
dT
= Qf = ,
(45.14)
A dt
dx
where Qf is the heat flux. This is called Fouriers law of heat conduction
(one-dimensional version).
Thermal conductivities are positive quantities so Fouriers law says that heat flow
down a temperature gradient, i.e., from hot to cold.
45.3.1. Thermal Conductivity of Gases and Liquids
See Reduced thermal conductivity handout
From this handout we see that typically the thermal conductivity of gases at low
densities increases with increasing temperature, whereas the thermal conductivity
of most liquids decrease with increasing temperature.

306

45.3.2. Thermal Conductivity of Solids


For the most part, the thermal conductivity of solids have to be determined experimentally because many factors contributing to the thermal conductivity are
dicult to predict.
In general metals are better heat conductors than nonmetals and crystals are
better heat conductors than amorphous materials.
Dry porous materials are poor heat conductors
Rule of Thumb: Thermal conductivity and electrical conductivity go hand in
hand.
The Wiedemann, Frantz and Lorenz equation relates the thermal conductivity to
electrical conductivity, el for pure metals:

= L = const.
el T

(45.15)

where L is the Lorenz number (typically 22 to 29 109 V2 /K2 ).


The Lorenz number is taken as constant because it is only a very weak function
of temperature with a change of 10 to 20% per 1000 degrees being typical.
The Wiedemann, Frantz and Lorenz equation breaks down at low temperature
because metals become superconductive. There is no analog to superconductivity
for thermal conductivity.

307

46. Solutions
Solutions are mixtures of two or more pure substances. So, in addition to the
parameters needed to characterize a pure substance, one also needs to keep track
of the amount of individual species in solution

46.1. Measures of Composition


There are several measures of composition of solutions
mole ratio r =

n1
n2

mole fraction X2 =

n2
,
n1 +n2

molality m =

1000X2
,
M1 X1

Molarity c2 =

n2
L solution

X1 = 1 X2

where M1 is the molecular weight of species 1

46.2. Partial Molar Quantities


Thermodynamic properties, in general change upon mixing
X
4mix = properties of soln
properties of pure.

(46.1)

For example,

4mix V = Vsoln Vsolute Vsolvent

(46.2)

Consider a thermodynamic quantity, say, volume.


308
308

In general, it is a function of T, P, n1 and n2 : V (T, P, n1 , n2 ). So, the total


derivative is

V
V
V
V
dV =
dT +
dP +
dn1 +
dn2 ,
T P,n1 ,n2
P T,n1 ,n2
n1 T,P,n2
n2 T,P,n1
(46.3)

V
Vi , the partial molar volume.
ni
T,P,nj

Similarly

G
G
G
G
dT +
dP +
dn1 +
dn2 ,
dG =
T P,n1 ,n2
P T,n1 ,n2
n1 T,P,n2
n2 T,P,n1
(46.4)

G
i .
ni
T,P,nj

So now for the more general case of mixtures the chemical potential of a species
of the partial molar free energy for that species, rather than simply the molar free
energy as it was earlier.
46.2.1. Notation
The study of solutions brings with it
here for future reference.
Material
Pure liquid i
Vi

Pure liquid i per mole


Vmi
Whole solution
V
Solution/(total moles)
Vm
Partial molar of i in solution Vi
Apparent molar (of solute) V
Reference state
Vi

a large number of symbols which we collect

Hi

Hmi
H
Hm
i
H

H
Hi

Si

Smi
S
Sm
Si

Gi
i
G
Gm
i

Si

309

46.2.2. Partial Molar Volumes


Consider the partial molar volume
For constant T and P
dV = V1 dn1 + V2 dn2

(46.5)

Now, Vi depends on concentration, so change each amount of substance proportional to the amount substance present,
dn1 = n1 d,
So,

dn2 = n2 d.

dV = V1 n1 + V2 n2 d = V = V1 n1 + V2 n2

(46.6)

(46.7)

That is, the total volume of the solution is equal to the sum of the partial molar
volumes each weighted by their respective number of moles.
The total volume, however, is not necessarily the mole weighted sum of the volumes of each component in its pure (unmixed) state. More specifically
4mix V

= V (Vm1
n1 + Vm2
n2 )

= V1 n1 + V2 n2 (Vm1
n1 + Vm2
n2 )

= V1 Vm1
n1 + V2 Vm2
n2

(46.8)

4mix V can be positive, negative or zero.


For example,

1. one unit of baseballs are mixed with one unit of basketballs. 4mix V < 0.
2. one unit of baseballs are mixed with one unit of books. 4mix V > 0.

310

46.3. Reference states for liquids


For liquids there are two more convenient ideal states
1. neat (pure) solvent limit
1. all neighboring molecules are same as the given molecule
2. the ideal state for Raoults law
2. infinite dilution limit
1. all neighboring molecules are dierent than the given molecule
2. the ideal state for Henrys law

Raoults law limit

Henrys law limit

46.3.1. Activity (a brief review)


Recall that activity gives a measure of the deviation of the real state from some
reference state

311

Also recall that the mathematical definition of activity ai of some species i is


implicitly stated as
ai
lim
=1
(46.9)
g()
where g() is any reference function (e.g., pressure, mole fraction, concentration
etc.), and is the value of at the reference state.
This implicit definition is awkward so for convenience one defines the activity
coecient as the argument of the above limit,
i

ai
g()

(46.10)

which we can rearrange as


ai = i g().

(46.11)

The definition of activity implies that i = 1 at g( ) (the reference state)


That is i 1 as the real system approaches the reference state.
Connecting with the chemical potential we saw last semester that the deviation
of the chemical potential at the state of interest versus at the reference state is
determined by the activity at the current state (the activity at the reference state
is unity by definition).
i
(46.12)
i = RT ln ai .

46.3.2. Raoults Law


In discussing both Raoults law and Henrys law, we are describing the behavior
of a liquid solution by measuring the vapor (partial) pressures of the components

312

For simplicity we consider here only a two component solution.


dG = 1 dn1 + 2 dn2 .

(46.13)

Take dierential change along a line of constant concentration, so


dG = (1 n1 + 2 n2 ) d

(46.14)

G = 1 n1 + 2 n2 .

(46.15)

4mix G = G(soln) G(pure components)

(46.16)

4mix G = 1 n1 + 2 n2 1 n1 2 n2

(46.17)

then

Recall that

Hence,

= (1 1 ) n1 + (2 2 ) n2 .

Now,
1 1 = RT ln

ai low P
Pi
' RT ln ,

ai
Pi

(46.18)

where Pi is the vapor pressure of the ith component above the solution.

313

Thus

or at low P

a1
a2
4mix G = RT n1 ln + n2 ln
a1
a2

P1
P2
4mix G = RT n1 ln + n2 ln
P1
P2

(46.19)

(46.20)

46.3.3. Ideal Solutions (RL)


Raoults Law:
Pi = Xi Pi

(46.21)

That is, the vapor partial pressure of a component of a mixture is equal to the
mole fraction of the component times the vapor pressure that the component
would have if it were pure.
The change in free energy upon mixing for solutions ideally obeying Raoults law
is

!
X1 P/1
X2 P/2
id(RL)
4mix G = RT n1 ln
+ n2 ln
(46.22)
P/1
P/2
id(RL)

4mix G = RT (n1 ln X1 + n2 ln X2 )

(46.23)

Again, this is for an ideal solution in the Raoults Law sense.


From

G
(G/T )
S=
and H =
,
T P
(1/T ) P
the entropy of mixing for an ideal Raoult solution is

id(RL)

4mix S = R (n1 ln X1 + n2 ln X2 )

(46.24)

(46.25)

and the enthalpy of mixing is


id(RL)

4mix H = 0

(46.26)

314

(since G/T is independent of 1/T ).


The Reference State (RL)
Let us apply the definition of activity for the Raoults law reference state.
The reference function is g() = = Xi . and the reference state is Xi = 1
So,
(RL)

ai
=1
Xi 1 Xi
lim

(46.27)

implies
(RL)

ai
(RL)

and i

(RL)

= i

Xi ,

(46.28)

1 as Xi 1

Deviations from Raoults Law


Raoults law is a purely statistical law. It does not require any kind of interaction
among the constituent particle making up the solution.
Since, in reality, there are specific interactions between particles, real solutions
generally deviate from Raoults law.
The physical interpretation of deviation from Raoults law is
positive deviation: the molecules prefer to be around themselves rather than
other types of molecules.
negative deviation: the molecules prefer to be around other types of molecules than themselves.
no deviation: the molecules have no preference.

315

It is very important to note that this deviation from Raoults law is a property of
the solution and NOT any given component.
For example, for a given component, mixing with one substance may lead to
a positive deviation but mixing with another substance may lead to a negative
deviation.

Positive deviation from Raoults lawNegative deviation from Raoults law


46.3.4. Henrys Law
Henrys Law:
(46.29)

Pi = kXi Xi ,
where kXi is the Henrys law constant,
kXi = lim

Xi 0

Pi
Xi

(46.30)

Henrys law applies to the solute not to the solvent and becomes more correct for
real solution as the concentration of solute goes to zero (Xi 0), i.e., at infinite
dilution.

316

The Reference State (HL)


Referring to the definition of activity again we see that the reference function is
g() = = Xi . and the reference state is now Xi = 0
So,
(HL)
ai
lim
=1
(46.31)
Xi 0 Xi
implies
(HL)
(HL)
= i Xi ,
(46.32)
ai
(HL)

1 as Xi 0
and i
If instead of mole fraction, molality or molarity is used then
(HL)

= (HL)
mi mi

(46.33)

(HL)

= Mi Mi

(HL)

(46.34)

ai
and
ai
respectively.

Comparison of Raoults Law and Henrys Law


Both Raoults law and Henrys law become better approximations for real solutions as the solution becomes pure. But, they apply to opposite species in the
solution. Raoults law applies to the dominant species, X1 1, whereas Henrys
law applies to the subdominant species X2 0. So, in summary
Raoults law: 1 1 as X1 1
Henrys law: 2 1 as X2 0

317

46.4. Colligative Properties


Colligative properties: Properties of dilute solutions that are independent of
the chemical nature of the solute
Examples
Freezing point depression
Boiling point elevation
Vapor pressure lowering
Osmotic pressure
We will consider the examples of freezing point depression and osmotic pressure
46.4.1. Freezing Point Depression
At Tf (freezing point), 1 (solid) = 1 (soln).
| {z }
s1

318

Using the Raoults law reference state (since we are interested in the behavior of
the dominant species), 1 (soln) = 1 + RT ln a1 :
s1 = 1 + RT ln a1
Rearranging this and taking the derivative with respect to T yields

1 s1 1
1
ln a1

s

( 1 ) =
=

ln a1 =
T
RT 1
T
RT 2 T
T
Now, using

= H and integrating we get

Z
1
4f H
s

d ln a1 =
(H1 H1 ) dT =
dT
2
RT
RT 2

Z Tf
4f H
dT
ln a1 =
2
Tf RT

(46.35)

(46.36)

(46.37)

For small changes in the freezing point we may approximate T by Tf in the


integrand. So,
Z Tf
4f H
4f H
ln a1 '
dT =
,
(46.38)
2
RTf2
Tf RTf
where Tf Tf . The freezing point depression is
=

RTf2 ln a1
4f H

46.4.2. Osmotic Pressure


We consider the osmotic pressure at a constant temperature, T. (so, dG = V dP ).

319

In the above figure 1 (left) = 1 (right), hence


1 = 1 + RT ln a1 + V1 ,

(46.39)

where V1 is the partial molar volume of the solvent in solution (dicult to measure)
and is the hydrostatic (osmotic) pressure.
From the above equation

V1
RT

Now we make the approximations V1 = Vm1 , a1 = X1 = 1 X2 :


ln a1 =

ln(1 X2 ) =

Vm1

RT

(46.40)

(46.41)

For dilute solutions X2 is small so ln(1 X2 ) may be expanded as


ln(1 X2 ) = X2 +
but X2 =

n2
n1 +n2

'

n2
n1

X22 X23

' X2 ,
2
3

(46.42)

for dilute solutions. Thus


V

1
z }|
{

Vm1
n1 Vm1

n2
= n2 '
,
'
n1
RT
RT

(46.43)

320

or,
=

n2
RT = cRT,
V1
|{z}

(46.44)

'c

where c is the concentration of the solute.

Note the similarity of this equation with the ideal gas equation: P = cRT. Thus
the solute in a very dilute solution behaves as if it were an ideal gas.

321

47. Entropy Production and


Irreverisble Thermodynamics
We have seen that thermodynamics tells us if a process will occur and kinetics
tells us how fast a process will occur.
These two areas of physical chemistry appear to be rather disjoint.
We now we consider thermodynamics of nonequilibrium states and investigate
how (and how fast) these state move towards equilibrium.
This allows us to make a stronger connection between thermodynamics and kinetics.
The main concept of this approach is the idea of entropy production and, ultimately, entropy production per unit timehow fast we are producing entropy.

47.1. Fundamentals
We know the dierence between reversible and irreversible processes from before.
However, we will state their respective definitions here in a manner best suited
for this chapter.

322
322

Reversible process: dynamical equations are invariant under time inversion


(t t).
e.g., the one dimensional wave equation,
1 2 u 2 u tt 1 2 u
2u
1 2u 2u
=
=
=
=
= 2,
c t2
x2
c (t)2
x2
c t2
x

(47.1)

is invariant under time reversal


Irreversible process: dynamical equations are not invariant under time inversion (t t).
e.g., the one dimensional heat equation,
2 T tt 1 T
2T
2T
1 T
1 T
=
=
=
=
=

,
t
x2
(t)
x2
t
x2

(47.2)

is not invariant under time reversal.

We will be concerned with the change in entropy, dS, which can be split into two
components dS = de S + di S.
Definitions
de S is the change in entropy due to interactions with the exterior environment.
di S is the change in entropy due to internal changes of the system
The quantity di S is called the entropy production.

323

Splitting up dS into these two parts permits an easy discussion of both open and
isolated systemsthe dierence between the two appearing only in de S.
General criteria for irreversibility:
di S = 0 (reversible change)
di S > 0 (irreversible change)
For isolated systems have di S = dS and the principle of Clausius, di S = dS 0,
holds.

47.2. The Second Law


As you might expect, the second law underlies all the concepts of this chapter.
We need a local formulation of the second law:
Absorption of entropy in one part of the system, compensated by a sucient
production in another part is prohibited
i.e., in every macroscopic region of the system the entropy production
due to irreversible processes is positive.
This is simply another in our long list of alternative statements of the second law.

324

I
II

Considering the above figure of an isolated system, we write the principle of Clausius as
dS = dS I + dS II 0.
(47.3)
The local formulation statement implies
di S I 0 and di S II 0

(47.4)

and the possibility of, for example, di S I < 0 and di S II > 0 such that di S I + S II >
0 is excluded.

47.3. Examples
The idea of entropy production can be applied to any of the processes we have
talked about; mixing, phase changes, heat flow, chemical reactions, etc. As example we now consider the last two of these: heat flow and chemical reactions.

325

47.3.1. Entropy Production due to Heat Flow

Recall from the lecture on transport phenomena that the heat flux Qf is given by
Qf =

4T
D

We are now interested in exposing the time dependence, so, using Qf =


A4T
q
=
4t
D

(47.5)
q
4t

we get
(47.6)

in dierential form this is

dT
dq
= A .
(47.7)
dt
dx
Example: Find the entropy production in a system consisting of two identical
connected blocks of metal (I and II), one of which is held at temperature T1 and
the other at T2 (take T1 > T > T2 ) where T is the temperature at the interface.

326

Considering the whole system


d S

dS =

dqI dqII
+
T1
T2

dS

e
i
z
}|
{ z
}|
{
de qI de qII di qI di qII
=
+
+
+
.
T1
T2
T1
T2

(47.8)

The quantity de qj is the amount of heat supplied by the environment to hold block
j at its fixed temperature.
Furthermore the heat going out of I through the connecting wall is equal to the
heat coming into II through the connecting wall:
di qI = di qII .

(47.9)

Using this we see that the entropy production is

1
1
,

di S = di qI
T1 T2

(47.10)

which we see is positive because di qI < 0 when T1 > T2 .


We have still not made a connection to kinetics.
To do so we must consider the entropy production per unit time

di S
.
dt

327

For this example


di qI
di S
=
dt
dt
From chapter 24 we know

1
1

T1 T2

A4T
di qI
=
.
dt
D
So,
A4T
di S
=
dt
D

(47.11)

1
1

T1 T2

(47.12)

To determine T we use the fact that the heat flow out of I is equal to the heat
flow into II:
di qI
di qII
=
.
(47.13)
dt
dt
Using the above expression for heat flow gives us T since,

/ A
/
/ A
/
T1 + T2
;
T1 T =
T T2 T =
2
D
/
D
/

(47.14)

a result we might have guessed.

47.3.2. Entropy Production due to Chemical Reactions


Definitions:
1. Chemical anity: a (4rxn G)T,P =
P
i vi i

vi i and a (4rxn A)T,V =

i
2. Extent of reaction: is defined by d = dn
, where ni is the number of moles
vi
th
of the i component and vi the stoichiometric factor of the ith component.

328

e.g., for the reaction N2 + 3H2 2NH3


d =

dnH2
dnNH3
dnN2
=
=
(1)
(3)
(2)

(47.15)

and
a = 2NH3 N2 3H2
The connection to kinetics: reaction rate v =

(47.16)

d
dt

The connection to thermodynamics:


(dA)T,V =

X
i

but


1
dni = ad
i dni =
vi i
v
i
| i {z }| {z }
X

ad

z }| {
z }| {
(dA)
dq
T,V

= (dU)T,V T dS dS =
T
T
dq

(dA)T,V
so

(47.17)

d S

(47.18)

dS

e
i
z}|{
z}|{
dq
ad
+
dS =
T
T

(47.19)

The entropy production per unit time for a chemical reaction is a function of both
the chemical anity and of the reaction rate
a d
a
di S
=
= v0
dt
T dt
T

(47.20)

We see that for a spontaneous process the entropy production per unit time is
positive. This is because a = (4rxn A)T,V is positive as is v.

329

Simultaneous Reactions
For N simultaneous chemical reactions, the entropy production per unit time
generalizes to
N
1X
di S
=
aj vj 0.
(47.21)
dt
T j=1

The second law requires that the total entropy production for simultaneous reactions is positive. It says nothing about the entropy production of the individual
component reactions other then the sum of all the component entropy productions
must be positive.

For example in a system of two coupled reactions we could have a1 v1 < 0, a2 v2 > 0
such that a1 v1 + a2 v2 > 0.

47.4. Thermodynamic Coupling


Processes may be what is called thermodynamically coupled such that a process
that normally is not thermodynamically favored can be coupled to another process
that is thermodynamically favored so as to allow for the unfavorable process to
proceed spontaneously.
We just saw an example of such a situation with the discussion of simultaneous
reactions.
Thermodynamic coupling need not be confined to coupling between the same
types of processes.
That is, diusion is the flux of matter down a concentration gradient. The socalled Soret eect is flux of matter down a temperature gradient. Conversely, the
so-called Dufour eect is heat flux down a concentration gradient

330

The following table lists a number of thermodynamically coupled phenomena


q

material

Q (charge)

Thermoconductivity

Soret eect

Seebeck eect

Mechanocaloric
eect

Thermomechanical
eect
Hydrodynamic
flow

Reverse
osmosis

Potential of flow

Dufour eect

Osmosis

Diusion

Nernst Potential

Peltier eect

Electrophoresis

Migration

Electoconductivity

Flux
Gradient

T
P
C

47.5. Echo Phenonmena


Consider an ensemble that is perturbed away from thermal equilibrium by some
means such as by applying a field.
If the perturbation is released the system will begin to evolve in time as it heads
back towards the thermalized equilibrium state.
The ensemble evolves in two ways
Reversibly
A second perturbation can undo or reverse the evolution.
Irreversibly
The evolution towards equilibrium cannot be undoneit is irreversible
Example: The spin echo in pulsed NMR
A radio frequency pulse prepares an ensemble of nuclear spins such that
they are all spinning coherently.

331

A strong signal is seen because all the spinning nuclei cooperate.


Each nucleus is in a slightly dierent environment so each spin frequency is
slightly dierent.
The dierent environment (spin frequencies) cause the ensemble spinning
nuclei to dephase
Dephasing causes a decrease in the observed signal because now not all nuclei
are cooperating.
Now a radio pulse with the opposite phase is applied to make the nuclei spin
in the opposite direction
This undoes or reverses the dephasing process and the signal regains strength
The full signal is not recovered however since all the while random thermalization is taking place to irreversibly destroy the coherence among the
nuclei.
This cannot be undone with the second radio pulse.

332

Key Equations for Exam 4


Listed here are some of the key equations for Exam 4. This section should not
substitute for your studying of the rest of this material.
The equations listed here are out of context and it would help you very little to
memorize this section without understanding the context of these equations.
The equations are collected here simply for handy reference for you while working
the problem sets.

Equations
The Clapeyron Equation is
4 Hm
dP
=
.
dT
T 4 Vm

(47.22)

The Clausius-Clapeyron equation is


4 Hm
d(ln P )
=
d(1/T )
R

(47.23)

Ficks first law of diusion is


dC
1 dn
= D
A dt
dx

(47.24)

333
333

Ficks second law of diusion:

2C
C
=D 2.
t
x

(47.25)

Relation between the viscosity and the diusion constant:


kT
f

kT
.
6r

(47.26)

dT
1 dq
= Qf = .
A dt
dx

(47.27)

D=

f =6r

Fouriers law of heat conduction is

Mixing
4mix = properties of soln
Chemical potential

properties of pure.

(47.28)

= + RT ln a

(47.29)

Pi = Xi Pi

(47.30)

Raoults Law:
Raoults law reference
(RL)

ai

(RL)

= i

Henrys Law:

Xi ,

(RL)

1 as Xi 1

(47.32)

Pi = kXi Xi .

where kXi is the Henrys law constant,


kXi = lim

Xi 0

Pi
Xi

(47.31)

(47.33)

Henrys law reference


(HL)

ai

(HL)

= i

Xi ,

(HL)

1 as Xi 0.

(47.34)

334

Index
absorption spectroscopy 241
activity 146, 311
mathematical definition of 146
activity coecient 146, 312
adiabatic expansion 280
and heat capacity 280
adiabatic wall 120
angular momentum
addition of 202
classical 192
eigenfunctions for 199, 219
jj coupling 202
LS coupling 202
quantum numbers 199, 219
spin 201
angular momentum quantum number 52
antibonding orbital 71
Arrhenious activation energy 261
Arrhenious equation 261, 291
temperature corrected 262
atomic orbitals 49
chemists picture 50
physicists picture 50
aufbau principle 58
average value theorem 29
Berthelot gas 13, 270
binominal coecient 90
blue sky 81
Bohr model 18

Bohr radius 19
Boltzmann distribution 10, 96, 131
Boltzmanns equation 90, 97, 124,
131
bond order 77
bonding orbital 71
Born model 170
corrections to 175
enthalpy of solvation 174
entropy of solvation 174
free energy of solvation 173, 178
partition coecient 174
BornOppenheimer approximation
62, 99, 235, 240
and the FranckCondon principle 243
bosons 56
Boyle temperature 272
chain rule
for partial derivatives 107
character table
for the C2v group 225
chemical anity 328
chemical potential 144
for a salt 161
relation to activity 148
relation to Gibbs free energy
145
relation to Helmhotz free energy 145
335
335

Clapeyron equation 298, 300, 333


Clausius-Clapeyron equation 299,
333
coecient of thermal expansion 274
coexistence curve 293
colligative properties 318
commutator 30, 189
completeness 191
complimentary variables 30
compressibility factor
at the critical point 295
compressibilty factor 270, 291
configuration 90
confluent hypergeometric functions
65
correspondence principle 41
critical point 300
cyclic rule 14, 108
cylindrical symmetry 69
DebyeHuckel limiting law 164, 178
DebyeHuckel theory 163
DebyeHuckelGuggenheim equation
164
Debyes law 129, 133
degeneracy 186
of the ensemble 98
diathermic wall 120
diatomic molecules
electron-electron potential energy operator for 61
electronic kinetric energy operator for 61
electronic wavefunction for 62
Hamiltonian for 61
nuclear kinetic energy operator
for 61
nuclear-electron potential energy
operator for 61

nuclear-nuclear potential energy


operator for 61
Schrodinger equation for 62
Dieterici gas 270
diusion 301
diusion constant 302
eigenfunction 5
eigenvalue 5
eigenvalue equation 190
electric dipole approximation 79, 231
electrolytes
strong 161
electrophoretic eect 167
elementary reactions 255
and stoichiometry 256
molecularity 256
emission spectroscopy 241
enemble 89
ensemble average 103, 132
enthalpy 136
entropy 105
change for changes in temperature 286
change for isothermal expansion
286
change for mixing 287
of real gases 288
entropy production 322, 323
due to chemical reactions 328
due to heat flow 326
equation of state 116
for a Berthelot gas 118
for a Dieterici gas 118
for a RedlichKwang gas 118
for a van der Waals gas 117
for an ideal gas 116
for gases 269
equilibrium constant 135

336

equlibrium constant 153


Eulers identity 4
expansion
of gases 111
reversible 114
extent of reaction 328
Eyrings equation 265, 291
fermions 56
Ficks first law 302, 333
Ficks second law 302, 334
first law of thermodynamics 121,
133
flipping coins 90
fluctuation 92
fluorescence 242
stokes shift 242
Fouriers law of heat conduction 306,
334
FranckCondon integral 243
FranckCondon principle 243
free energy
Gibbs 138
Helmholtz 137
fugacity 147
fundamental transistions 66
general equlibrium 151
generalized displacement 110
generalized force 110
gerade 69
Gibbs free energy 106
Gibbs-Duhem equation 163
good theory 16
group
mathematical definition of 222
multiplication table 223
group theory 221
Hamiltonian operator 27
Hamitonian

classical 27
harmonic oscillator 38
energy levels for 40, 44, 86
potential energy 39
Schrodinger equation for 39
heat 109
sign convention 110
heat capacity 115, 133
Heisenberg uncertainty principle 30
and the harmonic oscillator 41
helium 55
electron-electron repulsion term
55
Hamiltonian 55
Helmholtz free energy 106
Henrys law 311, 316, 334
Henrys law constant 316, 334
Hermite polynominals 40
hot bands 66
Hunds rule 205
hydrogen atom
ioniztion energy of 19
hydrogen molecule 74
hydrogenic systems 46
energy levels for 49, 86
Hamiltonian 47
normalization constant 49, 85
potential energy for 47
Schrodinger equation for 47
wavefunction (no spin) 49
wavefunction (with spin) 52
ideal solution
Raoults law 314
immiscible solutions 153
infrared spectroscopy 66
internal energy 103, 121
intramolecular vibrational relaxation
(IVR) 242

337

inversion symmetry 69
operator 69
ion mobility 166
and current 168
ion transfer 174
IR spectroscopy 231
and the character table 232
isothermal compressibility 274
isothermal expansion 279
Joule expansion 282
Joule-Thomson expansion 283
kinetic theory of gases 250
Lagrange multipliers 95
Laguerre polynominals 49
laminar flow 304
law of corresponding states 296
law of rectilinear diameters 293
Legendra polynomials 200
linear combinations of atomic orbitals (LCAO) 72
Lorenz number 307
many electron atom
Hamlitonian for 59
maximal work 113
Maxwell relations 140
Maxwells distribution of speeds 252,
290
mean free path 253, 290
mean ionic activity 162
mean ionic activity coecient 162
method of initial velocities 259
method of isolation 259
microstate 90
Mie scattering 84
mirror plane symmetry 70
molar heat capacity 115
molecular collisions
simple model for 252

molecular hydrogen ion 67


Hamiltonian for 67
molecular orbital diagram 76
molecular orbitals 68
molecular rotations 235
asymmetric tops 239
centrifugal stretching 236
linear tops 238
polyatomic molecules 237
spherical tops 239
symmetric tops 238
vibrational state dependence of
236
molecular vibrations 228
molecule
Scrodinger equation for 78
momentum operator 5
Morse oscillator 64
energy levels for 65, 86
Schrodinger equation for 65
wavefunction for 65
Morse potential 64, 86, 240
force constant associated with
9
Taylor series expansion of 8
normal modes 229
operator
Hermitian 189
ladder 195
linear 189
symmetry 222
operator algebra 187
orientation quantum number 53
orthogonality 191
overtone transitions 66
parameters
extensive 109
intensive 109

338

particle in a box 31, 181


energy levels 183
energy levels for 34, 44, 218
features of the energy levels 35
normalization constant for 33
potenial energy 31
Schrodinger equation for 32
three dimensional 183
three dimensional energy levels
185
three dimensional wavefunction
185
wavefunction for 183
wavefunctions for 34, 44, 218
particle on a ring 194
boundary conditions 194
energy levels for 195, 218
Hamitonian for 194
wavefunctions for 195, 218
partition coecient 154
and drug delivery 155
for the Born model 174
partition function
canonical 96, 131
electronic 101
grand canonical 97
isothermalisobaric 97
microcanonical 96
molecular 100
rotational 101, 132
translational 101, 132
vibrational 101, 132
Pauli exclusion principle 56
consquences of 58
perturbation theory 207
example of the quartic oscillator 208
phase diagram 300

Poiseuilles formula 304


polarizability 79
postulate I (of quantum mechanics) 22
postulate II (of quantum mechanics) 24
postulate III (of quantum mechanics) 25
pressure 104
principle of Clausius 125, 324
principle quantum number 52
probability amplitude 22
probability distribution 22
PV work 111, 133
Raman scattering 80
Raman spectroscopy 66, 233
and the character table 234
Raoults law 311, 312, 314, 334
deviations from 315
reference state 315
rate law 255
rate laws 254
determination of 258
integrated 259
Rayleigh scattering 80
Rayleigh scattering law 81, 82, 87
reaction velocity 255, 291
reciprocal rule 108
red sunsets 82
Redlich-Kwang gas 270
reference states 147
relationship between CP and CV
139, 276
relaxation eects 167
rigid rotor 200
degeneracy of 235, 248
energy 235, 247
rotational energy levels 200, 219

339

degeneracy of 200
rotational Hamiltonian 200
rule of mutual exclusion 234
Rydberg constant 20
SATP 120
Schrodinger equation
time dependent 214
time independent 27
second law
local formulation 324
second law of thermodynamics 126,
133
statements of 127
simple collision theory 262
Slater determinant 58
for lithium 59
solar system model 17
solvation 169
solvophobic eect 176
specific heat 115
spherical harmonic functions 48, 200
spin 201
quantum number 51, 53
wavefunction 51
spin orientation
quantum number 51, 53
spin-orbit
coupling 205
Hamiltonian 205
interaction energy 205
spontaneous process 142
state function 121
table of important ones 136
Sterlings approximation 92
Stokes law 167, 304
STP 120
superposition 191
systems

types of 108
temperature 115
term symbols 204
thermal conductivity 301
of gases 306
of liquids 306
thermal equilibrium 120
third law of thermodynamics 128,
133
tips for solving problems 2
total derivative 107
transfer matrix 11
triple point 300
two level system 211
left and right states 213, 219
Hamiltonian for 212
Tyndall scattering 84
ungerade 69
van der Waals equation

340

You might also like