forked from digitalmoleculardesign/bill_goddard_chem_120
-
Notifications
You must be signed in to change notification settings - Fork 0
/
Copy pathchap03.tex
3163 lines (2870 loc) · 120 KB
/
chap03.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
\chapter{More Exact Wavefunctions}
\section{Introduction}
In Chapter 2 we discussed approximate wavefunctions for H$^+_2$
$\chi_{\ell} \pm \chi_r$ and for H$_2$ $\chi_{\ell} \chi_r \pm \chi_r
\chi_{\ell}$. In this chapter we will consider some aspects of more
exact wavefunctions of these molecules and for the two-electron atom,
He. Our emphasis here will be on qualitative ideas.
We first examine the variational principle and later we use this principle to
examine some of the useful methods for calculating wavefunctions such as
HF, GVB, and CI.
Accurate wavefunctions for H$^+_2$, He, and H$_2$ are discussed in Sections
3.3, 3.7, and 3.8, respectively. In Section 3.4 we re-examine the nature of
the bond in H$^+_2$ and in Section 3.8 we re-examine the bond in H$_2$
The energy $\epsilon_0$ of any approximate wavefunction $\varphi_0$ is an
upper bound on the exact energy of the ground state $E_0$,
\begin{equation}
\epsilon_0 \geq E_0 ,
\end{equation}
leading to the variational condition. If an approximate wavefunction, and
hence the energy, is a function of some parameter $\lambda$, then the optimum
wavefunction satisfies the necessary condition
\begin{equation}
{\partial \epsilon \over \partial \lambda} = 0 .
\end{equation}
Expanding the unknown wavefunction $\varphi$ in terms of a basis
\begin{equation}
\varphi = \sum^{p}_{\mu = 1} C_{\mu} \chi_{\mu}
\end{equation}
and applying the variational condition, leads to a set of matrix equations,
\begin{equation}
{\bf H} {\bf C} = E {\bf S} {\bf C}
\end{equation}
for obtaining the optimum coefficients, i.e., wavefunction.
More exact wavefunctions of H$^+_2$ are also considered, but later we find
that the description of bonding, in terms of exchange energies is retained. In
Section 3.6 we present an overview of three useful methods for
wavefunctions. They are, first, the \emph{Hartree-Fock} (HF) method, a
generalization of the MO wavefunction in which the
wavefunction (ground state of a two- electron system) is taken as
\begin{equation}
\Phi^{HF} ( 1 , 2 ) = \varphi ( 1 ) \varphi ( 2 )
\end{equation}
and the orbitals $\varphi$ optimized by solving the differential equation
\begin{equation}
H^{HF} \varphi = \left( h + J_{\varphi} \right) \varphi = \epsilon
\varphi
\label{chap3-eqno1}
\end{equation}
or the matrix equation
\begin{equation}
\mathbf{H}^{HF} \mathbf{C} = \epsilon \mathbf{S}
\mathbf{C}.
\label{chap3-eqno2}
\end{equation}
These equations are nonlinear and must be solved iteratively.
Secondly, the \emph{generalized valence bond} (GVB) method, a
generalization of the VB method with the wavefunction taken of the
form
\begin{equation}
\Phi^{GVB} (1 , 2) = \varphi_a (1) \varphi_b (2) + \varphi_b (1) \varphi_a (2)
\end{equation}
and the orbitals $\varphi_a$ and $\varphi_b$ optimized. This leads to two
matrix equations
\begin{equation}
\mathbf{H}^a \mathbf{C}_a = \epsilon_a \mathbf{S} \mathbf{C}_a
\end{equation}
\begin{equation}
\mathbf{H}^b \mathbf{C}_b = \epsilon_b \mathbf{S} \mathbf{C}_b
\end{equation}
analogous to the HF equations (\ref{chap3-eqno2}) and to two
differential equations analogous to (\ref{chap3-eqno1}).
Third, the \emph{configuration interaction} (CI) method with the
wavefunction taken of the form
\begin{equation}
\Phi^{CI} (1, 2) = \sum_{\mu , \nu} C_{\mu \nu} \chi_{\mu} ( 1 )
\chi_{\nu} ( 2 )
\end{equation}
For the ground state this wavefunction can always be written in terms of
natural orbitals $\{ {\bar{\chi}}_{\mu} \}$ as
\begin{equation}
\Phi^{CI} (1, 2) = \sum_{\mu} {\bar{C}}_{\mu \nu} {\bar{\chi}}_{\mu} ( 1 )
{\bar{\chi}}_{\nu} ( 2 ) .
\end{equation}
Later in Section 3.6 we find that the ground state of any two electron
systems is nodeless and symmetric
\begin{equation}
\Phi (1 , 2) = \Phi (2 , 1).
\end{equation}
In Sections 3.7 and 3.8 we find that the HF wavefunction accounts for
all but about 1.1 eV of the energy for He and H$_2$, and that a CI
wavefunction with five natural orbitals accounts for all but about
0.15 eV. The four correlating natural orbitals for this wavefunction
all involve one nodal plane.
All of these methods involve expansions in terms of basis sets. For He
it is possible to obtain highly accurate HF and GVB wavefunctions with
only two, $s$-like, basis functions. The \emph{double valence} (DV)
basis and for H$_2$ similar quality wavefunctions can be obtained with
six basis functions, two $s$ and one $p$ on each center, the
\emph{double valence plus polarization}(DVP) basis.
\section{The Variational Principle}
The wavefunction for the ground state of the hydrogen atom has the
form $e^{-r}$, and the wavefunction for the ground state of the
harmonic oscillator is $e^{- \alpha r^2}$. However, there are very
few interesting systems for which the Schr\"odinger equation can be
solved exactly. Even so, there are general procedures allowing one to
obtain highly accurate, indeed arbitrarily accurate,
wavefunctions. The powerful tool needed is the \emph{variational
principle}. Here, we will outline some of the key results of this
principle.
\subsection{Upper Bound Theorem}
If $\psi_0$ and $E_0$ are the exact ground state wavefunction, and energy
of a system
\begin{equation}
H \psi_0 = E_0 \psi_0
\end{equation}
and if $\varphi_0$ is an approximate wavefunction with energy
\begin{equation}
\epsilon_0 = {\langle \varphi_0 \vert H \vert \varphi_0 \rangle
\over \langle \varphi_0 \vert \varphi_0 \rangle} ,
\end{equation}
then
\begin{equation}
\epsilon_0 \geq E_0 .
\label{chap3-eqno3}
\end{equation}
That is, the energy evaluated using any approximate wavefunction is an
upper bound in the exact energy of the ground state.
\subsection{Variational Condition}
Given an approximate wavefunction $\varphi_{\lambda}(r)$ depending upon
some parameter $\lambda$, e.g.,
\begin{equation}
\varphi_{\lambda} = e^{- \lambda r} ,
\end{equation}
then the optimum value of $\lambda$, for describing the system, must satisfy
the necessary condition
\begin{equation}
{\partial \epsilon ( \lambda ) \over \partial \lambda} = 0 ,
\label{chap3-eqno4}
\end{equation}
where
\begin{equation}
\epsilon ( \lambda ) = {\langle \varphi_{\lambda} \vert H \vert
\varphi_{\lambda} \rangle \over \langle \varphi_{\lambda} \vert
\varphi_{\lambda} \rangle} .
\end{equation}
Equation (\ref{chap3-eqno4}) is referred to as the \emph{variational
condition}.
\subsection{Basis Set Expansions}
Given a set of functions
\begin{equation}
\left\{ \chi_{\mu} ( r ) , \mu = 1 , 2 , \cdot \cdot \cdot , P
\right\} ,
\label{chap3-eqno5}
\end{equation}
the optimum wavefunction of the form
\begin{equation}
\varphi = \sum^{p}_{\mu = 1} C_{\mu} \chi_{\mu}
\label{chap3-eqno6}
\end{equation}
satisfies the condition
\begin{equation}
\sum_{\nu} H_{\mu \nu} C_{\nu} = E \sum_{\nu} S_{\mu \nu} C_{\nu},
\label{chap3-eqno7a}
\end{equation}
or in matrix notation
\begin{equation}
{\bf H} {\bf C} = E {\bf S} {\bf C},
\label{chap3-eqno7b}
\end{equation}
where
\begin{equation}
H_{\mu \nu} \equiv \langle \chi_{\mu} \vert H \vert \chi_{\nu}
\rangle
\label{chap3-eqno8a}
\end{equation}
and
\begin{equation}
S_{\mu \nu} \equiv \langle \chi_{\mu} \vert \chi_{\nu} \rangle
.
\label{chap3-eqno8b}
\end{equation}
The set of functions (\ref{chap3-eqno5}) used for expansion of the
unknown function $\varphi$, is called a basis set. The unknowns are
the coefficients
\begin{equation}
\left\{ C_{\mu} ; \mu = 1 , \cdot \cdot \cdot , P \right\}
\end{equation}
which are obtained by solving the matrix equations
\ref{chap3-eqno7a}--\ref{chap3-eqno7b}.
\subsection{Discussion of Upper Bound Theorem}
The upper bound theorem (\ref{chap3-eqno3}) is easy to
derive. Consider that the exact eigenstates $\{ \psi_i \}$ of the
Hamiltonian were known
\begin{equation}
H \psi_i = E_i \psi_i
\label{chap3-eqno9}
\end{equation}
with $i = 0$ as the ground state.
Since the set of functions $\{ \psi_i \}$ is complete, we can expand any
approximate wavefunction $\varphi_0$ as
\begin{equation}
\varphi_0 = \sum_{i} C_i \psi_i .
\label{chap3-eqno10}
\end{equation}
Of course, in a real problem we will not know the functions $\{ \psi_i
\}$, and hence, we will not be able to calculate the $\{ C_i \}$.
However, the analysis in this section will serve to establish a
relation between the approximate and exact solutions. From equations
(\ref{chap3-eqno9}) and (\ref{chap3-eqno10})
\begin{equation}
H \varphi_0 = \sum_i C_i H \psi_i = \sum_i C_i E_i
\psi_i
\end{equation}
and the energy of the approximate wavefunction $\psi_0$ is
\begin{equation}
\epsilon_0 \equiv \langle \varphi_0 \vert H \vert \varphi_0
\rangle = \sum_i \sum_j C^*_j C_i E_i \langle \psi_j \vert \psi_i
\rangle = \sum_i \vert C_i \vert^2 E_i .
\label{chap3-eqno11}
\end{equation}
Assuming $\varphi_0$ is normalized,
\begin{equation}
1 = \langle \varphi_0 \vert \varphi_0 \rangle = \sum_{ij} C^*_j C_i \langle
\psi_j \vert \psi_i \rangle = \sum_i \vert C_i \vert^2 .
\label{chap3-eqno12}
\end{equation}
Using (\ref{chap3-eqno12}) in (\ref{chap3-eqno11}), we obtain
\begin{equation}
\epsilon_0 - E_0 = \sum_i \vert C_i \vert^2 \left( E_i - E_0 \right)
.
\label{chap3-eqno13}
\end{equation}
Since $E_i \geq E_0$ and $\vert C_i \vert^2 \geq 0$, the right-hand side
of (\ref{chap3-eqno13}) is necessarily positive, and hence
\begin{equation}
\epsilon_0 - E_0 \geq 0
\end{equation}
or
\begin{equation}
\epsilon_0 \geq E_0 .
\label{chap3-eqno14}
\end{equation}
That is, the energy calculated for any wavefunction is never lower than the
exact energy for the ground state wavefunction. Thus, we say that
$\epsilon_0$ is an upper bound on $E_0$.
In deriving (\ref{chap3-eqno14}), we assumed that $\varphi_0$ could be
expanded in terms of the eigenfunctions of $H$. For
example, if the boundary conditions for the system described by
(\ref{chap3-eqno9}) were such that all wavefunctions are antisymmetric,
then we could not allow $\varphi_0$ to contain a symmetric part.
\subsection{Discussion of the Variational Principle}
Since any approximate wavefunction must yield an energy above the
energy of the exact, ground state, wavefunction, we have a useful
criterion for improving approximate wavefunctions. Namely, if you can
find some change in the wavefunction that leads to a lower energy,
then do it. Then reject any change that increase the energy.
Ultimately, if we consider all possible changes in the wavefunction,
this procedure must yield the exact wavefunction. Normally we lose
patience before considering all changes in the wavefunction, and
instead we consider functions of certain restricted classes. Here our
basic criterion for approximating the wavefunction will be to select
the function of our specific restricted class leading to the lowest
energy. If $\lambda$ is some variable parameter for the restricted set
of functions being considered, then the optimum wavefunction must
satisfy
\begin{equation}
{\partial E \over \partial \lambda} = 0
\label{chap3-eqno15}
\end{equation}
since otherwise a lower energy could be obtained by changing $\lambda$ a
bit, as illustrated in Figure \ref{fig3-01}.
\begin{figure}
\begin{center}
\includegraphics[scale=0.75]{fig3-01}
\end{center}
\caption{}
\label{fig3-01}
\end{figure}
\noindent
This criterion for optimizing a wavefunction is called the
\emph{variational principle} and forms the basis of all methods we
will consider for determining wavefunctions. It should be noted here
that (\ref{chap3-eqno15}) is not sufficient to guarantee a minimum, with
respect to variations in $\lambda$ (this requires
${\partial^2 E \over \partial \lambda^2} > 0$)
and even if a minimum is found, it need not, in general, be the minimum
leading to the lowest energy. Fortunately, for the types of problems we
deal with, these potential difficulties can usually be avoided.
\subsection{Parameter Optimization}
Consider, as an approximation to the ground state of the hydrogen atom, the
function
\begin{equation}
\varphi_0 ( \alpha ) = e^{- \alpha r^2}
\label{chap3-eqno16}
\end{equation}
where $\alpha$ is a parameter. To determine the value of $\alpha$
minimizing the energy, we first calculate the energy as a function of
$\alpha$,
\begin{equation}
E ( \alpha ) = {\langle \varphi_0 \vert - {1 \over 2} \nabla^2 - {1
\over r} \vert \varphi_0 \rangle \over \langle \varphi_0 \vert \varphi_0
\rangle} = {3 \over 2}\alpha - \left( {8 \over \pi \alpha}
\right)^{{1 \over 2}}
\label{chap3-eqno17}
\end{equation}
The optimum value of $\alpha$ is given by
\begin{equation}
{dE ( \alpha ) \over d \alpha} = 0 = {3 \over 2} - {1 \over 2} \left(
{8 \over \pi \alpha} \right)^{{1 \over 2}}
\end{equation}
or
\begin{equation}
\alpha_{opt} = {8 \over 9 \pi} = 0.283 ...
\end{equation}
Substituting this into (\ref{chap3-eqno17}), we obtain
\begin{equation}
E \left( \alpha_{opt} \right) = - {4 \over 3 \pi} = - 0.4244 ...
\end{equation}
recalling that the exact energy is $E = -0.5$. Thus, even though
(\ref{chap3-eqno16}) is considerably different from the exact
eigenfunction for the ground state of the hydrogen atom, by optimizing
$\alpha$ we are able to account for 84.9\% of the energy.
\subsection{Basis Set Expansions}
We will now use the variational principle to determine the best
representation of an approximate wavefunction as an expansion
(\ref{chap3-eqno6}), in terms of the functions of some finite basis
set (\ref{chap3-eqno5}). The energy is
\begin{equation}
E = {\langle \varphi \vert H \vert \varphi \rangle \over \langle
\varphi \vert \psi \rangle} = {N \over D} ,
\label{chap3-eqno18}
\end{equation}
where
\begin{equation}
N = \sum_{\mu , \nu} C^*_{\mu} H_{\mu \nu} C_{\nu} ,
\end{equation}
\begin{equation}
D = \sum _{\mu , \nu} C^*_{\mu} S_{\mu \nu} C_{\nu},
\end{equation}
and, $H_{\mu \nu}$ and $S_{\mu \nu}$ are given in
(\ref{chap3-eqno8a})--(\ref{chap3-eqno8b}). We do not assume here
that the basis functions are orthonormal; they must, of course, be
linearly independent.
The energy (\ref{chap3-eqno18}) depends on the $P$ parameters $\{
C_{\mu} \}$, and thus, from the variational principle we require that
\begin{equation}
{\partial E \over \partial C_{\mu}} = 0 ; \mu = 1 , 2 , \cdot \cdot
\cdot , P.
\end{equation}
From (\ref{chap3-eqno18}), this leads to
\begin{equation}
{\partial E \over \partial C_{\mu}} = {1 \over D} {\partial N \over
\partial C_{\mu}} - {N \over D^2} {\partial D \over \partial
C_{\mu}} = {1 \over D} \left[ {\partial N \over \partial C_{\mu}} -
E {\partial D \over \partial C_{\mu}} \right] = 0
\end{equation}
and hence,
\begin{equation}
{\partial N \over \partial C_{\mu}} - E {\partial D \over \partial
C_{\mu}} = 0 .
\end{equation}
Assuming that the basis function $\{ \chi_{\mu} \}$ and coefficients
$\{ C_{\mu} \}$ are all real, we obtain
\begin{equation}
2 \left[ \sum_{\nu} \left( H_{\mu \nu} - ES_{\mu \nu} \right) C_{\nu}
\right] = 0
\end{equation}
and hence,
\begin{equation}
\sum_{\nu} H_{\mu \nu} C_{\nu} E \sum_{\nu} S_{\mu \nu} C_{\nu} .
\label{chap3-eqno7-other}
\end{equation}
The more general case leads to the same equations. Note that if the basis
functions are real then $H_{\mu \nu} = H_{\nu \mu}$ and $S_{\mu \nu} =
S_{\nu \mu}$.
In matrix notation, (\ref{chap3-eqno7-other}) becomes
\begin{equation}
{\bf H} {\bf C} = E {\bf S} {\bf C}.
\label{chap3-eqno19}
\end{equation}
If the basis functions are orthonormal
\begin{equation}
S_{\mu \nu} = \delta_{\mu \nu} ,
\end{equation}
the variational condition (\ref{chap3-eqno19}) becomes
\begin{equation}
{\bf H} {\bf C} = E {\bf C}.
\label{chap3-eqno20}
\end{equation}
Thus, the variational principle leads to a finite matrix equation
directly analogous to the Schr\"odinger equation. Indeed, if a
complete set of basis functions is used, the solution of
(\ref{chap3-eqno15}) or (\ref{chap3-eqno20}) is the exact solution of the
Schr\"odinger equation. Although the wavefunction and basis functions
were written as one electron functions, this procedure applies
identically for many-electron wavefunctions.
\section{Accurate Wavefunctions for H$^+_2$}
The LCAO wavefunction of H$^+_2$ discussed in Chapter 2, is an
approximate wavefunction and does not provide a quantitatively
accurate description of H$^+_2$ near $R_e$. In this section, we will
discuss more accurate wavefunctions of H$^+_2$. First we consider a
useful intermediate level description, the MBS wavefunction.
\subsection{Scaled LCAO Wavefunctions}
We will describe the wavefunction of H$^+_2$ in terms of linear combinations
of two orbitals, $\chi_\ell$ and $\chi_r$, centered on each proton, but rather
than atomic orbitals, we will use scaled atomic-like orbitals
\begin{equation}
\chi_\ell = \sqrt{\left( {\zeta^3 \over \pi} \right)} e^{- \zeta
r_a}
\label{chap3-eqno21a}
\end{equation}
and
\begin{equation}
\chi_r = \sqrt{\left( {\zeta^3 \over \pi} \right)} e^{- \zeta
r_b}
\label{chap3-eqno21b}
\end{equation}
The scaling parameter $\zeta$ is referred to as an \emph{orbital
exponent}. Use of $\zeta = 1$ leads to the LCAO description of
Chapter 2, $\zeta > 1$ leads to more contracted orbitals, while $\zeta
< 1$ leads to more diffuse orbitals.
Using the basis set (\ref{chap3-eqno21a})--(\ref{chap3-eqno21b}), the
wavefunctions of H$^+_2$ have the form
\begin{equation}
\varphi_g = {( \chi_\ell + \chi_r) \over \sqrt{2(1+S)}}
\end{equation}
\begin{equation}
\varphi_u = {( - \chi_\ell + \chi_r) \over \sqrt{2(1-S)}}
\end{equation}
just as in Chapter 2. However, the energies of these wavefunctions
depend upon both $\zeta$ and $R$ (see Section \ref{appendix-a} in
Chapter 2 for the specific dependence of the integrals on $\zeta$).
At each $R$ we will use the $\zeta$ leading to the lowest energy.
Since the forms of $E_g$ and $E_u$ are different, the optimum $\zeta$
will be different for the $g$ and $u$ states, as shown in Figure
\ref{fig3-02}.
\begin{figure}
\includegraphics[scale=0.75]{fig3-01-1}
\caption{The optimal orbital exponents for the MBS descriptions of the
$g$ and $u$ states of H$_2^+$.}
\label{fig3-02}
\end{figure}
\begin{table}
\caption{Optimum bond length, $R_e$, and bond strength,
$D_e$, for the $g$ state of H$^+_2$. All quantities in atomic units.}
\label{table3-01}
\parindent=0.5truein
\begin{tabular}{cccccc}\\ \hline
&\multicolumn{3}{c}{Non-Relativistic} &
\multicolumn{2}{c}{Relativistic} \\
&\multicolumn{3}{c} {Neglect $T_{nuc}$} & Neglect $T_{nuc}$ &
Include $T_{nuc}$ \\
&LCAO$^a$ & MBS$^a$ & Exact$^b$ & Exact & Exact \\
$R_e$ & 2.493 & 2.00 & 2.00379$^c$ & 2.00376 $^c$ & 2.00562 $^c$ \\
$D_e$ & 0.065 & 0.08651 & 0.102635$^c$ & 0.10264$^c$ & 0.101785$^c$ \\
\hline
\end{tabular}\\
$^a$See reference 1. $^b$See reference 2. $^c$See reference 3.
$^d$See reference 4.
\end{table}
As shown in Figure \ref{fig3-02} and seen in Table \ref{table3-01},
the improvement in the energy for the $g$ state is quite remarkable,
leading to energies close to the exact answer. For the $u$ state, both
the LCAO and the MBS
energies are quite close to the exact answer.
\begin{figure}
\includegraphics[scale=0.75]{fig3-02}
\caption{The LCAO ($\zeta=1.0$) MBS (optimum $\zeta$) and the exact
energies for the $g$ and $u$ states of H$^+_2$. Note that (b)
does contain three different lines. The vertical scale of (b) is twice
that of (a).}
\label{fig3-03}
\end{figure}
In discussing such wavefunctions, we will use the following terminology.
First, linear combination of atomic orbitals, LCAO, denotes the use of a
linear combination of atomic orbitals using the orbital exponents of the
atoms. Second, minimal basis set, MBS, indicates the smallest set of
atomic-like functions that would describe the case of $R = \infty$. For
finite $R$, the orbital exponents will generally be
optimized. The result of MBS calculations will be
discussed further, after a discussion of the exact wavefunctions of H$^+_2$
\subsection{The Exact Wavefunction for H$^+_2$}
Previously, we considered approximate solutions of the Schr\"odinger equation
\begin{equation}
H \psi_R(r) = E ( R ) \psi_R(r) ,
\label{chap3-eqno22}
\end{equation}
where the Hamiltonian is
\begin{equation}
H = - {1 \over 2} \nabla^2 - {1 \over r_a} - {1 \over r_b} +
{1 \over R} .
\end{equation}
Exact solutions to (\ref{chap3-eqno22}) have also been obtained, as
will now be described.
We can obtain arbitrarily accurate wavefunctions of H$^+_2$ by
expanding the orbital in terms of a sufficiently general basis
\begin{equation}
\left\{ \chi_{\mu} : \mu = 1 , 2 , \cdot \cdot \cdot , P \right\} ,
\end{equation}
\begin{equation}
\varphi ( r ) = \sum^{P}_{\mu = 1} C_{\mu} \chi_{\mu} ( r ) ,
\end{equation}
where the expansion coefficients are obtained by solving the $P$ by
$P$ matrix equation
\begin{equation}
\mathbf{HC} = \mathbf{C}E,
\end{equation}
with
\begin{equation}
H_{\mu \nu} = \langle \chi_{\mu} \vert H \vert \chi_{\nu} \rangle
\end{equation}
assuming the basis to be orthonormal. As the basis set is made more
complete ($P \rightarrow \infty$) the wavefunction approaches the
exact wavefunction.
Although the above procedure is practical, it is possible for H$^+_2$ to solve
directly for the exact solutions. The procedure is examined in more detail in
Section \ref{chap3-app-d}.
\subsection{Comparison of Wavefunctions and Energies}
The various wavefunctions of the $g$ and $u$ states are compared in
Figure \ref{fig3-04} for $R = 2a_0$. For the $g$ state, we see that
the shape of the linear combination of atomic orbitals wavefunction in
the bond region, is in good agreement with the exact
wavefunction. However, the magnitude of the density in the bond region
is $\approx 25$ to 30\% low.
\begin{figure}
\includegraphics[scale=0.75]{fig3-03}
\caption{The wavefunctions of H$^+_2$ at $R = 2.0a_0$.}
\label{fig3-04}
\end{figure}
\noindent
The MBS description leads to reasonably good densities near the nuclei
but too low a density in the bond region. Thus, with MBS the shape of
the wavefunction is not well described.
In the $u$ state, the LCAO
wavefunction is in much better agreement with the exact wavefunction
than is the MBS wavefunction.
In Figure \ref{fig3-05}, we compare the LCAO and MBS wavefunctions as
a function of $R$, finding that the LCAO description does reasonably
well for $R > 4\ a_0$. Note the large difference in the behavior of the
$g$ and $u$ states for small $R$. These differences were also
manifest in the optimum exponents of Figure \ref{fig3-02}.
\begin{figure}
\includegraphics[scale=0.75]{fig3-04}
\caption{Amplitudes of the LCAO (dashed) and MBS (solid) wave
functions for (a) the $g$ state and (b) the $u$ state of H$^+_2$}
\label{fig3-05}
\end{figure}
\section{More on the Chemical Bond}
In Chapter 2 we analyzed the bond of H$^+_2$ in terms of the linear
combination of atomic orbitals description. Now we will re-examine
the bond using more accurate wavefunctions. With more accurate
wavefunctions, we still find that the exchange energy $E^x$ (more
specifically the exchange kinetic energy $T^x$ part of $E^x$) is
responsible for the bonding or antibonding of the $g$ and $u$ states
of H$^+_2$. On the other hand, partitioning the energy into the total
potential energy ($V$) and the total kinetic, energy ($T$) we find
that neither can be solely responsible for bonding.
\subsection{The Classical and Exchange Energies}
Defining the classical and exchange terms, just as in Chapter 2,
\begin{equation}
E = E^{cl} + E^x
\label{chap3-eqno23}
\end{equation}
and
\begin{equation}
E^{cl} = \langle \chi_\ell \vert H \vert \chi_\ell \rangle
\end{equation}
but using the MBS wavefunctions, we obtain the results
of Figure \ref{fig3-06}.
\begin{figure}
\includegraphics[scale=0.75]{fig3-05}
\caption{(a) The total energy, $E$, and the
components $E^{cl}$ and $E^x$ for the MBS wavefunction
of the $g$ state of H$^+_2$. (b) The $T^x$ and $V^x$ components of
$E^x$ is shown. All quantities are relative to $R = \infty$.}
\label{fig3-06}
\end{figure}
\noindent
Thus the exchange energy dominates the bonding just as for the linear
combination of atomic orbitals wavefunction. Partitioning the $E^x$
into potential and kinetic parts, $V^x$ and $T^x$,
\begin{equation}
E^x = V^x + T^x
\end{equation}
as in Figure \ref{fig3-06}, we see that $T^x$ favors bond formation,
while $V^x$ opposes it, just as for the LCAO wavefunction.
Thus, in terms of the classical and exchange quantities, the linear
combination of atomic orbitals and MBS descriptions are quite
similar. In both cases, it is the large decrease in $T^x$ that is
responsible for bond formation. Just as discussed in Chapter 2, $T^x$
is large and negative because the atomic orbitals are contragradient
in the region between the nuclei. In particular, the $T^x$ is similar
in character for the LCAO and MBS descriptions. With $\zeta > 1$ the
gradients get larger and favor a smaller $R$ so that the differences
in $T^x$ for the LCAO and MBS
descriptions are easily understood.
\subsubsection{An Ambiguity}
There is a flaw with this procedure of decomposing the energy into
classical and exchange parts. Adding a second basis function on each
center, say $\chi_{2\ell}$ and $\chi_{2r}$, and optimizing the coefficients,
leads to
\begin{equation}
\Phi_g = C_1 \left( \chi_{1\ell} + \chi_{1r} \right) + C_2 \left(
\chi_{2\ell} + \chi_{2r} \right)
\end{equation}
and adding additional functions, we ultimately obtain the exact
wavefunction in the form
\begin{equation}
\Phi_g = \sum^{\infty}_{k=1} C_k \left( \chi_{k\ell} + \chi_{kr} \right) .
\end{equation}
Thus, we can define optimum left and right orbitals as
\begin{equation}
\chi_\ell = \sum_{k} C_k \chi_{k\ell}
\label{chap3-eqno24a}
\end{equation}
\begin{equation}
\chi_r = \sum_{k} C_k \chi_{kr}
\label{chap3-eqno24b}
\end{equation}
and obtain an exchange energy for the exact wavefunction. The problem
is that for the exact wavefunction there is not a unique choice for
the left and right functions $\chi_\ell$ and $\chi_r$. As a result,
there is some ambiguity in the exchange energy for the exact
wavefunction. On the other hand, with optimized basis functions only
a few functions (say, two $s$ and one $p_z$ on each center) lead to
quite accurate descriptions but with no ambiguity in the decomposition
(\ref{chap3-eqno24a})--(\ref{chap3-eqno24b}).
\subsection{Potential and Kinetic Energies}
Rather than the partition (\ref{chap3-eqno23}) of the energy into
classical and exchange terms, it has been much more common to
partition the energy into total potential energy, $V$, and total
kinetic energy, $T$,
\begin{equation}
E = T + V
\label{chap3-eqno25}
\end{equation}
This partition mixes up the things characteristic of
bonding with other quantities that are nearly independent of bonding
with the result that neither quantity, $T$ or $V$, consistently
contains the bonding stuff. A good illustration of this is to compare
the quantities for the LCAO and MBS
wavefunctions of H$^+_2$. As shown earlier, the classical and
exchange energies behave very similarly for these two cases. However,
as shown in Figure \ref{fig3-07}, the behavior of $T$ and $V$ for the
LCAO and MBS wavefunctions is
markedly different.
\begin{figure}
\includegraphics[scale=0.75]{fig3-06}
\caption{The kinetic and total potential energies for (a,b) the $g$
state, and (c,d) the $u$ state of H$^+_2$. All quantities
are relative to the value for $R = \infty$.}
\label{fig3-07}
\end{figure}
Thus, for LCAO the $T(R)$ is always lower than $T(\infty)$, while
$V(R)$ is always higher than $V(\infty)$. This might suggest that it
is kinetic energy that is responsible for the bond. However, for the
MBS wavefunction $T(R)$ is below $T(\infty)$ only for $R > 2.7\ a_0$.
Thus, at $R_e = 2\ a_0$, $T(R) > T(\infty)$ and it would be ludicrous
to assume that the total kinetic energy is the quantity dominating
bonding. On the other hand, in the MBS wavefunction, $V(R) >
V(\infty)$ for $R = 3.5\ a_0$. Thus, although $V(R)$ dominates the
bond at $R_e$, it opposes bond formation for $R < 3.5\ a_0$.
Furthermore, for the LCAO, $V(R)$ opposes bonding for all $R$.
Such difficulties show that (\ref{chap3-eqno25}) is not a useful
partition of the energy. The key indication of this is that although
the total energy changes monotomically from $R = \infty$ to $R_e$, the
$V$ and $T$ for the MBS and exact wavefunctions are not monotonic,
each dominating the energy over different regions. Hence, neither can
be uniquely responsible for bonding.
Occasionally, usually in the analysis of rotational and conformational barriers
in polyatomic molecules, energy curves are analyzed by partitioning
the $V$ into various parts
\begin{equation}
V = V^{en} + V^{nn} + V^{ee}
\end{equation}
where $ee$ denotes electron-electron repulsion, $en$ denotes
electron-nuclear attraction, and $nn$ denotes nuclear-nuclear
repulsion terms, $ee$ is not present for H$^+_2$. As shown in Figure
\ref{fig3-08}, each term is monotomic, with $V^{en}$ decreasing with
$R$. One might conclude from this that it is $V^{en}$ that is
responsible for bond formation. However, as seen from Figure
\ref{fig3-08}(b), the $V^{en}$ and $V^{nn}$ are also monotonic for the
$u$ state and, again $V^{en}$ decreases with $R$, but this state is
repulsive. Thus, despite similar $V^{en}$ and $V^{nn}$ for $g$ and $u$
we obtain radically different potential curves. Thus, $V^{en}$ is
dominated by quantities other than those responsible for bond
formation.
\begin{figure}
\includegraphics[scale=0.75]{fig3-07}
\caption{The total potential energy, $V$, and the partition
into $V^{nn}$ and $V^{en}$ for (a) the $g$ state and
(b) the $u$ state of H$^+_2$, exact wavefunctions is shown. All
quantities are relative to the value for $R = \infty$.}
\label{fig3-08}
\end{figure}
\section{Overview of Theoretical Methods}
This course focuses primarily on \emph{qualitative} ideas of chemical
bonding rather than particular theoretical methods. However,
familiarity with the theoretical methods is important for discussing
qualitative ideas and hence, we will outline these methods.
\subsection{Basis Sets}
Several methods involve solving for the optimum shape of one-electron orbitals
$\varphi_i(r)$. The general procedure for carrying out such calculations,
involve selection of a basis set
\begin{equation}
\left\{ \chi_{\mu} ; \mu = 1 , \cdot \cdot \cdot , P
\right\},
\label{chap3-eqno26}
\end{equation}
suitable for describing the optimum orbitals
\begin{equation}
\varphi_i ( r ) = \sum^{P}_{\mu = 1} C_{\mu i} \chi_{\mu} ( r ) .
\end{equation}
Here, the basis functions are fixed and hence, selection of the optimum
coefficients
\begin{equation}
\left\{ C_{1i} , C_{2i} , \cdot \cdot \cdot , C_{Pi} \right\}
\end{equation}
serves to determine the orbital $\varphi_i(r)$. This procedure is
analogous to a Fourier expansion where harmonic functions (sines and
cosines) are used as basis functions in (\ref{chap3-eqno26}).
For an exact description of the optimum orbital it is generally necessary
to use an infinite number, a complete set, of basis functions. However,
for practical reasons we must use a finite set. Indeed, from numerous
studies of molecular wavefunctions, there are principles that can be used
to select rather small basis sets that yield quite accurate wavefunctions.
In evaluating the wavefunctions and energies using a basis set, as in
(\ref{chap3-eqno26}), we must evaluate integrals of the form
\begin{equation}
\langle \chi_{\mu} \vert h \vert \chi_{\nu} \rangle
\label{chap3-eqno27a}
\end{equation}
\begin{equation}
\langle \chi_{\mu} ( 1 ) \chi_{\nu} ( 2 ) \vert {1 \over r_{12}}
\vert \chi_{\sigma} ( 1 ) \chi_{\eta} ( 2 ) \rangle ,
\label{chap3-eqno27b}
\end{equation}
where the functions may be centered at various regions of space. Thus, an
important criterion in selecting the basis is that the molecular integrals be
practicable to evaluate. In order to obtain the best wavefunctions
with the fewest basis functions, we want to choose the basis functions
to have shapes characteristic of the eigenstates of the molecular systems.
For a Coulomb potential, i.e. the hydrogen atom, the eigenstates have the form
\begin{equation}
1s : e^{-Zr}
\end{equation}
\begin{equation}
2s : (r - \alpha ) e^{-{1 \over 2}Zr}
\end{equation}
\begin{equation}
2p_z : r \cos \theta e^{-{1 \over 2} Zr}
\end{equation}
\begin{equation}
2p_x : r \sin \theta \cos \varphi e^{-{1 \over 2} Zr}
\label{chap3-eqno28}
\end{equation}
\begin{equation}
2p_y : r \sin \theta \cos \varphi e^{-{1 \over 2} Zr}
\end{equation}
\begin{equation}
3s : \left( r^2 - \beta r + \alpha \right) e^{-{1 \over 2} Zr}
\end{equation}
\centerline{etc.}
\noindent
where normalization is ignored and the constants $\alpha$ and $\beta$
are unimportant to our considerations here. In order to describe,
with finite number of basis functions, singular characteristics such
as the cusps occurring near the various nuclei, we should include in
our basis set functions having similar singular characteristics. Thus,
for a molecular system, we should use atomic functions like
(\ref{chap3-eqno28}) centered upon the various nuclei of the molecule.
The radial parts of the functions in (\ref{chap3-eqno28}) all can be
built from functions of the form
\begin{equation}
r^n e^{- \zeta r},
\label{chap3-eqno29}
\end{equation}
where various values of n and of the orbital exponents, $\zeta$, must
be allowed. Functions of the form (\ref{chap3-eqno29}) are preferable
to the hydrogen atom orbitals (\ref{chap3-eqno28}), since
(\ref{chap3-eqno29}) is more convenient for evaluating the molecular
integrals. Combining functions of the form (\ref{chap3-eqno29}) with
appropriate angular functions, $Z_{lm}$ the real spherical harmonics,
leads to a convenient set of one-particle orbitals
\begin{equation}
r^n e^{- \zeta r} Z_{lm} \left( \theta , \varphi \right)
\label{chap3-eqno30}
\end{equation}
for use in atomic and molecular wavefunctions. These functions
(\ref{chap3-eqno30}) are referred to as \emph{Slater functions}, or
\emph{Slater-type orbitals} (STO) in honor of an early exponent
\cite{chap3-ref5} of such functions. We will use the term function
when referring to an arbitrary function as in a basis function, and
the term orbital when referring to a specific optimized orbitals as in
a HF or GVB orbital. They are denoted as $1s$, $2s$, $2p$, etc., just
as for hydrogen atom orbitals. The orbital exponent, $\zeta$, in
(\ref{chap3-eqno29}) is considered as an adjustable parameter and is
generally chosen as the optimum value for the particular molecule and
basis set of interest (rather than taken as $\zeta = Z/n$ as suggested
by (\ref{chap3-eqno28})).
For example, a good basis for describing the wavefunction for H$_2$ is
to use two ls Slater functions (denoted as $1s$ and $1s^{\prime}$) a
$2s$ Slater function, and a set of the $2P$ Slater functions, $2p_z$,
$2p_x$, and $2p_y$, on each center. The optimum exponents at $R =
1.4\ a_0$ are \cite{chap3-ref6}
\begin{eqnarray}
\zeta ( 1 s ) &= 0.965\\
\zeta ( 1 s^{\prime} ) &= 1.43\\
\zeta ( 2 s ) &= 1.16\\
\zeta ( 2 pz ) &= 1.87\\
\zeta (2 px ) &= \zeta ( 2py ) = 1.71
\end{eqnarray}
\noindent
where the molecular axis is along $z$. With this basis, the CI
wavefunction leads to an energy of $-1.16696$ h, at $R = 1.4a_0$,
99.4\% of the exact answer$^7$ $-1.17447$ h. Note that the optimum
orbital exponents are significantly different from the values for the
free atom
\begin{equation}
\zeta_{1s} = 1.0
\end{equation}
\begin{equation}
\zeta_{2s} = 0.5
\end{equation}
\begin{equation}
\zeta_{2p} = 0. 5.
\end{equation}
The second type of basis functions, commonly used in molecular
calculations, are \emph{Gaussian functions} where the $e^{-\zeta r}$
of (\ref{chap3-eqno30}) is replaced by $e^{-\alpha r^2}$ and $n$ is
taken as $l$,
\begin{equation}
r^l e^{- \alpha r^2} Z_{lm} \left( \theta , \varphi \right)
\end{equation}
Although Gaussian functions have the wrong behavior as $r \rightarrow
0$, and as $r \rightarrow \infty$, they serve just as well as Slater
functions in describing the valence orbitals and the bonds of
molecules. The major advantage of Gaussian functions is that the
molecular integrals (\ref{chap3-eqno27a})--(\ref{chap3-eqno27b})
required for large molecules are much simpler, and less time
consuming, than for Slater functions.
Generally, basis sets are optimized for atoms. If properly carried
out, the atomic basis sets supplemented by a few additional functions,
polarization functions, serve to provide very accurate descriptions of
the molecular wavefunctions.
\subsection{The HF Method}
\subsubsection{The Basic Equations}
In Chapter 2 we described the simple MO wavefunction of H$_2$ in
which the two-electron wavefunction is expressed as
\begin{equation}
\Phi ( 1 , 2 ) = \varphi ( 1 ) \chi ( 2 )
\label{chap3-eqno31}
\end{equation}
where $\varphi$ is the MO
\begin{equation}
\varphi = {( \chi_\ell + \chi_r ) \over \sqrt{2(1+S)}}
\end{equation}
and $\chi_\ell$ and $\chi_r$ are hydrogen orbitals centered on the two
nuclei. Now we will consider the case where $\varphi$ is allowed to be
completely general. Thus, if $\{ \chi_{\mu} \}$ is some basis set,
we write
\begin{equation}
\varphi = \sum_{\mu} C_{\mu} \chi_{\mu}
\label{chap3-eqno32}
\end{equation}
with the coefficients $\{ C_{\mu} \}$ chosen so that $\Phi$ in
(\ref{chap3-eqno31}) leads to the lowest possible energy.
The energy of (\ref{chap3-eqno31}) is
\begin{equation}
E = {\langle \Phi \vert H^{el} \vert \Phi \rangle \over \langle
\Phi \vert \Phi \rangle} = 2 \langle \varphi \vert h \vert \varphi
\rangle + J_{\varphi \varphi} + {1 \over R} ,
\label{chap3-eqno33}
\end{equation}
where
\begin{equation}
J_{\varphi \varphi} = ( \varphi \varphi \vert \varphi \varphi ) = \int
d^3 r_1 \varphi^* (1) \varphi (1) \int d^3 r_2 {\varphi^* (2)
\varphi(2) \over r_{12}}
\end{equation}
and
\begin{equation}
\langle \varphi \vert \varphi \rangle = 1 .
\label{chap3-eqno34}
\end{equation}
Applying the variational principle to (\ref{chap3-eqno33})
\begin{equation}
{\partial E \over \partial C_{\mu}} = 0
\end{equation}
with the constraint (\ref{chap3-eqno34}), leads to
\begin{equation}
\langle \chi_{\mu} \vert \left( h + J_{\varphi} - \epsilon \right) \vert
\varphi \rangle = 0
\label{chap3-eqno35}
\end{equation}
where $\epsilon$ is referred to as the orbital energy,
\begin{equation}
\epsilon = \langle \varphi \vert h \vert \varphi \rangle + J_{\varphi \varphi}
,
\label{chap3-eqno36}
\end{equation}
and
\begin{equation}
J_{\varphi} ( r_1 ) = \int d^3 r_2 {\varphi^* (r_2) \varphi (r_2) \over r_{12}}