forked from digitalmoleculardesign/bill_goddard_chem_120
-
Notifications
You must be signed in to change notification settings - Fork 0
/
Copy pathchap10.tex
3833 lines (3452 loc) · 145 KB
/
chap10.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{Homonuclear Diatomic Molecules}
\section{Introduction}
In this chapter, we will start with the Aufbau diagram for atoms
and construct an energy level scheme for predicting states of
homonuclear molecules. The idea is to consider two limits:
\begin{enumerate}
\item $R = \infty$ or the \emph{separated atoms limit} where the energy
levels are just those of two atoms having nuclear charge $Z$.
\item $R = 0$ or the \emph{united atom limit}, where the energy levels
are those of the united atoms with nuclear charge $2Z$.
\end{enumerate}
Using symmetry considerations to convert the levels, allows one to
predict the occupation of various MOs for finite distances.
First, we digress to consider names of symmetries for diatomic
molecules.
\section{Symmetry Names}
We saw, in Chapter 2, that for a molecule with inversion symmetry,
each wavefunction of the system was either $g$ or $u$ with respect to
inversion,
\begin{eqnarray}
{\hat{I}}\varphi_g &=& \varphi_g\\
{\hat{I}} \varphi_u &=& - \varphi_u.
\end{eqnarray}
We derived this result by noting that
\begin{equation}
H(I\Phi) = E(I \Phi),
\end{equation}
if $H$ is invariant under inversion. Thus, $\Phi$ and $I \Phi$ must
both be eigenfunctions of $H$ with the same energy.
It is generally the case that symmetries in the molecular
framework result in symmetries of the wavefunctions, and for
this reason, names, being $g$ or $u$, are given to the
wavefunctions to indicate these symmetries. The general study
of such symmetries, called group theory, provides powerful
theorems that can be used to analyze wavefunctions.
However, we will not have time to pursue this here. Rather, we will
merely examine the naming conventions so that. we can use the names
commonly associated with various wavefunctions.
\subsection{Atoms}
With only one nucleus, the potential energy of an atom is
unchanged by any rotation about the nucleus, and the atom is
said to be spherically symmetric. Thus, if
\begin{equation}
H \Phi = E \Phi,
\end{equation}
then
\begin{equation}
H(R \Phi) = E(R \Phi),
\end{equation}
where $R$ is any rotation of the atom. Consequently, starting with
any eigenstate $\Phi$ or $H$, then all new wavefunctions $R \Phi$ that
can be obtained from $\Phi$ by rotating about some axis, must also be
an eigenstate of $H$ with the same energy. Thus, starting with
\begin{equation}
\Phi_z ( r , v , \varphi ) = {z \over r} f(r) = {1 \over R} f (r)
\cos \theta
\label{chap10-eqno1}
\end{equation}
that is, an eigenstate of $H$, and rotating by 90$^{\circ}$ about
the $x$ and $y$ axes, we obtain
\begin{eqnarray}
\Phi_x &= {x \over r} f(r) = {1 \over r} f(r) \sin\theta\cos\varphi\cr
\Phi_y &= {y \over r} f(r) = {1 \over r} f(r) \sin\theta\sin\varphi,
\label{chap10-eqno2}
\end{eqnarray}
both of which must be eigenstates of $H$ and degenerate with $\Phi_z$. Of
course, there are many other (an infinite number of) rotations $R$
that we could apply to $\Phi_z$. However, every one of these leads to
a linear combination of $\Phi_x$, $\Phi_y$, and $\Phi_z$. Thus, for one
atom, any eigenstate having the angular wavefunction
\begin{equation}
Z_{10} (\theta , \varphi ) = {z \over r} = \cos\theta
\end{equation}
is necessarily triply degenerate with additional eigenstates having the
same radial wavefunction and the same energy, but angular terms of the form
\begin{eqnarray}
Z_{11}(\theta , \varphi ) &=& {x \over r} = \sin\theta\cos \varphi\cr
Z_{1{\bar{1}}}(\theta , \varphi ) &=& {y \over r} = \sin\theta\cos
\varphi.
\end{eqnarray}
Such symmetry analyses lead to the result that for a one-electron
system, the exact wavefunctions all have the form
\begin{equation}
\Phi_{nlm} ( r ,\theta , \varphi ) = f_{nl} (r) Z_{lm} (\theta , \varphi ) ,
\end{equation}
where the angular wavefunctions must have the form of spherical
harmonics $Z_{lm}(\theta,\varphi)$. Here $l$ is a non-negative
integer, $l = 0 , 1 , 2 , \dots ,$ and the $2l + 1$ values of $m$,
associated with a given $l$, are all degenerate, same energy.
Thus, for $l = 2$, we obtained the five states, ignoring normalization,
\begin{equation}
\begin{array}{cccc}
\varphi_{20} & f(r) (3 \cos^2\theta - 1) & d_{x^2} & m = 0\\
\varphi_{21} & f(r) (\cos\theta \sin\theta)\cos\varphi & d_{zz} & m = 1 \\
\varphi_{1{\bar{1}}} & f(r)(\cos\theta\sin\theta)\sin \varphi &d_{yz}&m = -1\\
\varphi_{22} & f(r)(\sin^2\theta)\cos2\varphi & d_{x^2-y^2} & m = 2\\
\varphi_{2{\bar{2}}} & f(r)(\sin^2\theta)\sin 2 \varphi& d_{xy}& m =-2\\
\end{array}
\label{chap10-eqno3}
\end{equation}
Starting with any one of these functions, and rotating a proper amount
about appropriate axies, will lead to the other four. But applying
all the possible rotations, will just lead to a linear combination of
these functions.
For a one-electron system, the $l$ correspond to the total number of
angular nodal planes and $|m|$ is the number of nodal lines in the $xy$
plane. For reasons we will not dwell upon, $l$ is generally referred
to as the \emph{angular momentum quantum number}, and $m$ is related
to as the projection of $l$ along the $z$ axis. The names associated
with $l$ are summarized in Table \ref{chap10-tab1}. One-electron
cases are distinguished from many-electron cases by use of lower case
letters, in the table.
\begin{table}
\caption{Symmetry names for atoms.}
\label{chap10-tab1}
\begin{tabular}{cccccccccccccc}\\ \hline
$l$ & 0 & 1 & 2 & 3 & 4 & 5 & 6 & 7 & 8 & 9 & 10 & 11 & 12\cr
$(2l+1)^a$ & 1 & 3 & 5 & 7 & 9 & 11 & 13 & 15 & 17 & 19 & 21 & 23 &
25\cr
one-electron & s & p & d & f & g & h & i & k & I & m & n & o & q\cr
many-electron & S & P & D & F & G & H & I & K & L & M & N & 0 & Q\cr
\hline
\end{tabular}\\
$^a$degeneracy
\end{table}
An atom also has \emph{inversion} symmetry. For a one-electron
wavefunction, it is not necessary to indicate the inversion symmetry
because one-electron wavefunctions with even $l$ are always $g$, and
those with odd $l$ are always $u$. For many-electron wavefunctions,
we will sometimes append a $g$ or $u$ to indicate overall inversion
symmetry.
In Chapter \ref{chap05} we considered some examples of many-electron
wavefunctions, such as the ${^3P}_g$ wavefunction of C(2p)$^2$, the
${^3S}_u$ wavefunction of N(2p)$^3$, and the ${^3P}_g$ wavefunction of
O(2p)$^4$. In these cases, we were able to deduce the $L$ by
examining the total $M$ values for the various equivalent
wavefunctions. This simple approach is generally sufficient for the
ground state and we will not dwell upon the subject here.
\subsection{Linear Molecules}
\subsubsection{Rotational Symmetry}
\begin{figure}
% fig1 (ch 10 p 4) not found
%\includegraphics[scale=0.75]{fg10-}
\caption{}
\label{chap10-fig1}
\end{figure}
Consider a linear molecule with z as the molecular axis, e.g., CO,
\begin{equation}
% figure missing ch 10 p 4
\end{equation}
In this case, the molecule is symmetric with respect to all rotations
about the $z$ axis, but not about any other axis. The explicit form
of the changes in $x$ and $y$ are given in Figure \ref{chap10-fig1}.
In Figure \ref{chap10-fig1}, we see that $\gamma$ can have any value
of 0 to $2\pi$, other values would be equivalent to one of these. We
will denote a general rotation by $R_z(\gamma$), but if the angle
$\gamma$ can be written as $2 \pi / n$, where $n$ is an integer, the
rotation is customarily denoted as
\begin{equation}
C_n = R \left( {2 \pi \over n} \right).
\end{equation}
Thus, C$_2$ is a 180$^{\circ}$ rotation, and C$_3$ is a 120$^{\circ}$ rotation.
Just as for atoms, starting with $\Phi$ (that is, our eigenstate of
$H$) then any $R_z \Phi$ obtained by rotating $\Phi$ about the z axis
must also be an eigenstate with the same energy. Since only rotations
about one axis are allowed, there will be fewer degenerate states for
molecules. Thus, rotations of the $p_z$ function in equation
(\ref{chap10-eqno1}) never change it, whereas rotations of the $p_x$
and $p_y$ functions of equation (\ref{chap10-eqno2}) merely change
these functions into each other. Consequently, for a molecule the
$p_z$ function need not be degenerate with the $p_x$ and $p_y$
functions, but $p_x$ and $p_y$ are still degenerate.
Similarly, starting with the $d$ functions of equation
(\ref{chap10-eqno3}), we find that $d_{zz}$ and $d_{yz}$ must be
degenerate, and $d_{xy}$ and $d_{x^2-y^2}$ must be degenerate.
However, there need not be any relation between the energy of these
two sets of wavefunctions, nor need there be any between the $d_{z^2}$
functions and the others.
The conclusion from such symmetry analyses is that the total
wavefunction of a one-electron linear molecule will have the form
\begin{eqnarray}
\psi_{+\lambda} \left( \rho, \varphi , z \right) &=& f ( \rho , z )
\cos \lambda \varphi\\
\psi_{- \lambda} ( \rho , \varphi , z ) &=& f ( \rho , z ) \sin
\lambda \varphi
\label{chap10-eqno8}
\end{eqnarray}
for $\lambda \not= 0$. Here, $\lambda$
is the number of nodal lines in the $xy$ plane, $\varphi$ is the angle in the
$xy$ plane ($\varphi = 0$ is along the $x$ axis) and
\begin{equation}
\rho = \sqrt{x^2 + y^2}.
\end{equation}
Thus, for linear
molecules, the symmetry states are never more than twofold degenerate,
whereas, for an atom the degeneracy is $(2l + 1)$.
Rotating the wavefunctions in equation (\ref{chap10-eqno8}), by an
angle $\gamma$, leads to
\begin{eqnarray}
R( \gamma ) \psi_{+ \gamma} ( \rho , \varphi , z ) &=& \psi_{+ \gamma}
\left( \rho , \varphi + \gamma , z \right) = f ( \rho , z ) \cos
\lambda ( \varphi + \gamma )\cr
&=& f ( \rho , z ) \left[ \cos \lambda \varphi \cos \lambda \gamma -
\sin \lambda \gamma \sin \lambda \gamma \right]\cr
&=& \left( \cos \lambda \gamma \right) \psi_{+ \lambda} - \left( \sin
\lambda \gamma \right) \psi_{+ \lambda} .
\label{chap10-eqno10}
\end{eqnarray}
\begin{eqnarray}
R( \gamma ) \psi_{- \gamma} ( \rho , \varphi , z ) &=& \psi_{- \gamma}
\left( \rho , \varphi + \gamma , z \right) = f ( \rho , z ) \cos
\lambda ( \varphi + \gamma )\cr
&=& f ( \rho , z ) \left[ \sin \lambda \varphi \cos \lambda \gamma +
\cos \lambda \gamma \sin \lambda \gamma \right]\cr
&=& \left( \cos \lambda \gamma \right) \psi_{- \lambda} + \left( \sin
\lambda \gamma \right) \psi_{+ \lambda} .
\label{chap10-eqno11}
\end{eqnarray}
so that, as alleged, the two functions $\psi_{+ \lambda}$ and $\psi_{-
\lambda}$ change only into
each other upon various rotations about the intermolecular axis.
For a many-electron atom, the only rotations leaving the total
Hamiltonian invariant are those in which all electrons are
rotated simultaneously, because of the $1/r_{ij}$ terms in $H$. The
result is that the eigenstates of $H$ have the form $\Phi_{+
\Lambda}$ and $\Phi_{- \Lambda}$, and the rotation of all electrons
through an angle $\gamma$ leads to
\begin{equation}
R ( \gamma ) \Phi_{+ \Lambda} = \left( \cos \Lambda \gamma \right)
\Phi_{+ \Lambda} - \left( \sin \Lambda \gamma \right) \Phi_{- \Lambda}
\end{equation}
and
\begin{equation}
R ( \gamma ) \Phi_{- \Lambda} = \left( \cos \Lambda \gamma \right)
\Phi_{- \Lambda} + \left( \sin \Lambda \gamma \right) \Phi_{+ \Lambda},
\end{equation}
just as for the one-electron cases in equation (\ref{chap10-eqno10})
and equation (\ref{chap10-eqno11}). Analogous to the $\lambda = 0$
state in equation (\ref{chap10-eqno10}), the many-electron
wavefunction with $\Lambda = 0$ is nondegenerate and does not change
upon rotation.
The symmetry names for a linear molecule are given in Table
\ref{chap10-tab2}. Bear in mind that the $\delta$
or $\Delta$ state of a linear molecule is always twofold degenerate,
having $M = \pm 2$, whereas the $d$ or $D$ state of an atom is
fivefold degenerate, having $m = 0 ,
\pm 1 , \pm 2$.
\begin{table}
\caption{Names for linear molecules.}
\label{chap10-tab2}
\begin{tabular}{cccccccc}\\ \hline
$\Lambda$ & 0 & 1 & 2 & 3 & 4 & 5 & 6\cr
degeneracy & 1 & 2 & 2 & 2 & 2 & 2 & 2\cr
angular&1&cos$\varphi$&
cos2$\varphi$&cos3$\varphi$&cos4$\varphi$&cos5$\varphi$&cos6$\varphi$\cr
wavefunctions &&sin$\varphi$&sin2$\varphi$&sin3$\varphi$&sin4$\varphi$&sin5$\varphi$&sin6$\varphi$\cr
one-electron & $\sigma$ & $\pi$ & $\delta$ & $\phi$ & $\gamma$ &
$\eta$ & i\cr
many-electron & $\Sigma$ & $\Pi$ & $\Delta$ & $\Phi$ & $\Lambda$ &
h & I\cr
\hline
\end{tabular}
\end{table}
As an example, consider a linear molecule in which we put a B
atom, ${^2P}$ state, at the origin, and a pair of positive charges
along the $z$ axis at $z= \pm \infty$. The three states of B are
${^2P}_z$ for $m = 0 \rightarrow {^2\Sigma} , {^2P}_x$ for $m = + 1
\rightarrow {^2\Pi}$, and ${^2P}_y$ for $m = - 1 \rightarrow
{^2\Pi}$, so, in terms of molecular names, these states are
${^2\Sigma}$ and ${^2\Pi}$, as
indicated. As the charges are moved to finite distance, they will
stabilize the ${^2P}_z$ state more than ${^2P}_x$ or ${^2P}_y$, so
that the ${^2\Sigma}$ and ${^2\Pi}$ states will have different energies,
as indicated in Figure \ref{chap10-fig2}(a).
As a second example, starting with Sc in the ${^2D}$ state, we would
obtain the results in Figure \ref{chap10-fig2}(b) where the fivefold
degenerate $D$ state has split into a nondegenerate $\Sigma$ state,
plus two twofold degenerate states, $\Pi$ and $\Delta$.
\begin{figure}
% fig 2 not found (ch10 p 7)
%\includegraphics[scale=0.75]{fg10-}
\caption{}
\label{chap10-fig2}
\end{figure}
\subsubsection{Angular Momentum States}
Consider the pair of states (\ref{chap10-eqno8}), corresponding to
$\lambda \not= 0$, and combine these states as
\begin{equation}
\psi_{\lambda +} = \psi_{+ \lambda} + i \psi_{-\lambda} =
f ( \rho , z ) e^{i \lambda \varphi}
\label{chap10-eqno12a}
\end{equation}
\begin{equation}
\psi_{\lambda -} = \psi_{+ \lambda} - i \psi_{- \lambda} =
f ( \rho , z ) e^{- i \lambda \varphi}
\label{chap10-eqno12b}
\end{equation}
Since the angular momentum projection operator is
\begin{equation}
{\hat{\ell}}_z = {\hbar \over i} {\partial \over \partial \varphi} =
x p_y - yp_x = {\hbar \over i} \left[ x {\partial \over \partial
y} - y {\partial \over \partial x} \right],
\end{equation}
we see that
\begin{equation}
{\hat{\ell}}_z \psi_{\lambda +} = + \lambda \psi_{\lambda +}
\end{equation}
\begin{equation}
{\hat{\ell}}_z \psi_{\lambda -} = - \lambda \psi_{\lambda -}
\end{equation}
Thus, the states (\ref{chap10-eqno12a})--(\ref{chap10-eqno12b} are
eigenstates of the angular momentum projection operator with
eigenvalues $\pm \lambda$. We prefer to use the real form of these
wavefunctions (\ref{chap10-eqno8}). However, the two states in
(\ref{chap10-eqno12a}) and (\ref{chap10-eqno12b}) also provide a
correct description of these states.
\subsubsection{Reflection Symmetry}
In addition to the rotational symmetry, discussed above, a linear
molecule is always invariant under reflection in the xz plane, that is
\begin{equation}
{{x \atop y} \atop z} \rightarrow {{x \atop -y} \atop z}
\label{chap10-eqno13a}
\end{equation}
or equivalently
\begin{equation}
{{\rho \atop \varphi} \atop z} \rightarrow
{{\rho \atop - \varphi} \atop z}
\label{chap10-eqno13b}
\end{equation}
This reflection operation is denoted as $\sigma_v$, where $\sigma$ derives from
Spiegel, meaning mirror in German, and $v$ derives from vertical,
visualizing the molecule to be standing up. Since application
of $\sigma_v$ twice leads to the identity $(x,y,x) \rightarrow
(x,y,z)$, just as
for inversion and just as for transposition of electrons. Then,
from the analysis of previous chapters, we know that each
eigenstate of $H$ can be taken as symmetric, denoted as $+$, or as
antisymmetric, denoted as $-$, under $\sigma_v$. Since
\begin{eqnarray}
\sigma_v \cos \lambda \varphi &=& \cos \left( - \lambda \varphi
\right) = \cos \lambda \varphi\cr
\sigma_v \sin \lambda \varphi &=& \sin \left( - \lambda \varphi
\right) = -\sin \lambda \varphi,
\end{eqnarray}
we see that for $\lambda \not= 0$, there is always one $+$ and one $-$
state. These can be denoted as $\pi^+$ for cos $\varphi$, $\pi^-$ for sin
$\varphi$, $\delta^+$ for cos $2 \varphi$, and $\delta^-$ for sin $2
\varphi$, etc., but usually the $\sigma_v$ symmetry is deleted so
that the symbol $\pi$ or $\delta$
refers to both states, just as $d$ refers to all five $m$ states. For a
one-electron state with $\lambda = 0$, i.e., $\sigma$, the wavefunction
must have $+$ symmetry for $\sigma_v$, i.e., $\sigma^+$. Hence,
the $+$ is suppressed. For example, the
bonding state of H$_2$ is ${^1\Sigma}^+_g$, whether we use the valence
bond (VB) or the molecular orbital (MO) wavefunction,
\begin{equation}
\Phi^{VB} = \left( \chi_l \chi_r + \chi_r \chi_l \right) \left(
\alpha \beta - \beta \alpha \right)
\end{equation}
or
\begin{equation}
\Phi^{MO} = \left( \varphi_g \varphi_g \right) \left(
\alpha \beta - \beta \alpha \right)
\end{equation}
and the antibonding state is ${^3\Sigma}^+_u$.
An example of $\Sigma^-$ is the wavefunction
\begin{equation}
\Psi (1,2) = \left[ \varphi_{px} (1) \varphi_{py} (2) - \varphi_{py}
(1) \varphi_{px} (2) \right] \alpha (1) \alpha (2).
\end{equation}
This is seen most easily by expanding
\begin{eqnarray}
\varphi_{px} &= f_p ( \rho , z ) \cos \varphi\cr
\varphi_{py} &= f_p ( \rho , z ) \sin \varphi
\end{eqnarray}
we obtain
\begin{equation}
\Psi(1,2) = f_p(1) f_p(2) \underbrace{[\cos \varphi_1 \sin
\varphi_2 - \sin \varphi_1 \cos \varphi_2]}_{\sin( \varphi_2 -
\varphi_1 )} \alpha (1) \alpha (2)
\end{equation}
Since rotating both electrons through an angle $\gamma$, leads to
\begin{equation}
\sin \left[ \left( \varphi_2 + \gamma \right) - \left( \varphi_1 +
\gamma \right) \right] = \sin \left( \varphi_2 - \varphi_1 \right),
\end{equation}
this wavefunction is $\Sigma$, but
\begin{equation}
\sigma_v \varphi_{px} = \varphi_{px}
\end{equation}
\begin{equation}
\sigma_v \varphi_{py} = - \varphi_{py},
\end{equation}
and hence,
\begin{equation}
\sigma_v \Psi = - \Psi,
\end{equation}
so that this wavefunction has ${^3\Sigma}^-$ symmetry.
\subsubsection{Other Reflections}
Of course, the $\sigma_v$ in (\ref{chap10-eqno13a}) is not the only reflection symmetry a
linear molecule, there is an infinite number of such reflection
planes corresponding to all possible planes containing the $z$
axis. However, if $\sigma$ denotes
the reflection at an angle $\gamma$, with respect to the $xz$ plane
it is easy to show that $\sigma$ can be expressed in terms of
$R_z(\gamma)$ and $\sigma_v$ as
\begin{equation}
\sigma = R_z ( \gamma ) \sigma_v R_z (- \gamma ),
\end{equation}
so that a knowledge of the rotational and $\sigma_v$ symmetry
suffices to specify the symmetry for all other reflections $\sigma$.
For a $\Sigma$ state, the wavefunction will have the same reflection
symmetry for all such reflection planes. For $\Lambda \not= 0$, the
wavefunctions can be combined to have $+$ or $-$ symmetry, with
respect to only one $\sigma$ plane at a time, and by convention, we
choose this plane to be the $xz$ plane.
\subsubsection{Inversion Symmetry}
A general linear molecule, e.g., CO, NiH, HCN, has only the
rotational and reflection symmetries discussed above. However,
some molecules, e.g., N$_2$, O$_2$, HC$=$CH, also have inversion
symmetry. Just as discussed earlier, this leads to states
with either $g$ or $u$ symmetry, and this will be indicated by
subscripts, as always, e.g., ${^1\Sigma}^+_g$, ${^3\Sigma}^-_u$, and
${^5\Delta}_u$.
\subsubsection{Horizontal Reflections}
For a linear molecule with inversion symmetry, there is a
second type of reflection symmetry, referred to as the
horizontal reflection, and denoted as $\sigma_h$ (remember that
the molecule is standing up). This takes $x, y, z$ into $x, y, -z$
\begin{equation}
{{x \atop y} \atop z} \stackrel{\sigma_h}{\rightarrow} {{x \atop y} \atop -z}.
\end{equation}
That is, $\sigma_h$ interchanges the two atoms. Since C$_{2z}$ takes $x, y, z$
into $-x, -y, +z$, and $i$ takes $-x, -y, z$ into $x, y, -z$, we see
that $\sigma_h$ is equivalent to first applying C$_2$ and then $i$,
$\sigma_h = iC_2$.
Consequently, we need not consider the $\sigma_h$ symmetry separately, it
is determined by the rotational and inversion symmetry.
\subsubsection{Molecular Symmetry for Atomic Wavefunctions}
Summarizing the discussions, from the above sections, we see that the
one-electron atom functions, centered at the bond midpoint, have the
following symmetries as shown in Table \ref{chap10-tab3}.
\begin{table}
\caption{}
\label{chap10-tab3}
\begin{tabular}{ccc}\\ \hline
Atomic Symmetry & Function & Molecular Symmetry\cr
s & 1 & $\sigma_g = s \sigma$\cr
p & $x$ & $\pi^+_u = p \pi$\cr
& $y$ & $\pi^-_u = p \pi$\cr
& $z$ & $\sigma_u = p \sigma$\cr
d & $xy$ & $\delta^-_g = d \delta$\cr
& $x^2 - y^2$ & $\delta^+_g = d \delta$\cr
& $xz$ & $\pi^+_g = d \pi$\cr
& $yz$ & $\pi^-_g = d \pi$\cr
& $2z^2-(x^2+y^2)$ & $\sigma_g = d \sigma$\cr
\hline
\end{tabular}
\end{table}
\subsection{Examples of Many-Electron Wavefunctions}
\subsubsection{The $\pi^2$ Configuration}
Since we will have many occasions to consider wavefunctions in which
two electrons are distributed over two equivalent $\pi$ orbitals,
$\pi_x$ and $\pi_y$, we will work out the wavefunctions and symmetries
for this case. Considering only the spatial wavefunctions, there are
four possible two-electron wavefunctions, $\pi_x (1) \pi_x (2)$,
$\pi_y (1) \pi_y (2)$, $\pi_x (1) \pi_y(2)$, and $\pi_y(1) \pi_x(2)$,
which we will combine as
\begin{eqnarray}
x(1) y(2) + y(1) x(2) &=& F \sin ( \varphi_1 + \varphi_2 )\cr
x(1) y(2) - y(1) x(2) &=& F \sin ( \varphi_2 - \varphi_1 )\cr
x(1) x(2) + y(1) y(2) &=& F \cos ( \varphi_2 - \varphi_1 )\cr
x(1) x(2) - y(1) y(2) &=& F \cos ( \varphi_2 + \varphi_1 )
\label{chap10-eqno14}
\end{eqnarray}
where $F$ is independent of $\varphi_1$ and $\varphi_2$, and where
$\pi_x$ and $\pi_y$ are
denoted as $x$ and $y$, respectively. Since $\pi_x = f ( \rho , z )
\cos \varphi$ and $\pi_y = f ( \rho , z )$, then $F = f ( \rho_1 ,
z_1 ) f ( \rho_2 , z_2 )$. Rotating the coordinates by an
angle $\gamma$ changes both $\varphi_1$ and $\varphi_2$ by this angle,
leading to
\begin{eqnarray}
\sin ( \varphi_1 + \varphi_2 + 2 \gamma ) &=& \sin ( \varphi_1 +
\varphi_2 ) \cos 2 \gamma + \cos ( \varphi_1 + \varphi_2 ) \sin 2
\gamma\cr
\sin ( \varphi_2 - \varphi_1 ) &&\cr
\cos ( \varphi_2 - \varphi_1 ) &&\cr
\cos ( \varphi_2 + \varphi_1 + 2 \gamma ) &=& \cos ( \varphi_2 +
\varphi_1 ) \cos 2 \gamma - \sin ( \varphi_2 + \varphi_1 ) \sin 2
\gamma
\label{chap10-eqno15}
\end{eqnarray}
for the four wavefunctions in equation (\ref{chap10-eqno14}). Thus,
the rotational symmetries in equation (\ref{chap10-eqno14}) are
$\Delta$, $\Sigma$, $\Sigma$, and $\Delta$, respectively. The first
two functions are odd under $\sigma_v$, and the other two are even.
From equations (\ref{chap10-eqno13a})--(\ref{chap10-eqno13b}) we see
that $\sigma_v$ changes the signal of all $\varphi$. Thus,
\begin{eqnarray}
\pi_x (1) \pi_y (2) + \pi_y (1) \pi_x (2) &=& \Delta^-\cr
\pi_x (1) \pi_y (2) - \pi_y (1) \pi_x (2) &=& \Sigma^-\cr
\pi_x (1) \pi_x (2) + \pi_y (1) \pi_y (2) &=& \Sigma^+\cr
\pi_x (1) \pi_x (2) - \pi_y (1) \pi_y (2) &=& \Delta^+
\label{chap10-eqno16}
\end{eqnarray}
From equation (\ref{chap10-eqno15}) we see that applying a rotation to
either of the $\Delta$ states, leads to a mixture of these two states.
Consequently, these states are degenerate, same energy, and are
generally considered together as $\Delta$, it being understood that
both $+$ and $-$ are included.
\begin{figure}
% fig 3 not found (ch10 p 14)
%\includegraphics[scale=0.75]{fg10-}
\caption{}
\label{chap10-fig3}
\end{figure}
Now we consider the spin functions. There are two sets, a triplet, $S
= 1 , \alpha \alpha , \alpha \beta + \beta \alpha , \beta \beta$, and
a singlet $S = 0 , ( \alpha \beta - \beta \alpha )$. However, because
of the Pauli principle, each spatial function of equation
(\ref{chap10-eqno16}) goes with a specific spin function
\begin{eqnarray}
( \pi_x \pi_y + \pi_y \pi_x ) ( \alpha \beta - \beta \alpha ) &=&
{^1\Delta}^-\cr
( \pi_x \pi_y - \pi_y \pi_x )( \alpha \alpha )&=& {^3\Sigma}^-\cr
( \pi_x \pi_y - \pi_y \pi_x )( \alpha \beta + \beta \alpha )
&=& {^3\Sigma}^-\cr
( \pi_x \pi_y - \pi_y \pi_x )( \beta \beta ) &=& {^3\Sigma}^-\cr
(\pi_x \pi_x + \pi_y \pi_y ) ( \alpha \beta - \beta \alpha ) &=&
{^1\Sigma}^+\cr
(\pi_x \pi_x - \pi_y \pi_y ) ( \alpha \beta - \beta \alpha ) &=&
{^1\Delta}^+
\label{chap10-eqno17}
\end{eqnarray}
From equation (\ref{chap10-eqno14}) we can even estimate the energies
of these states. All four states are equivalent in $r$ and $\theta$,
differing only in their dependence upon $\varphi$. Since electrons
repel each other, it is favorable for the wavefunction to be small
when the electrons are close to each other. From equation
(\ref{chap10-eqno14}) we saw that the ${^3\Sigma}^-_g$ wavefunction
goes to zero as $\varphi_2
\rightarrow
\varphi_1$, and hence, the electrons are never
on top of each other, good. On the other hand, the ${^1\Sigma}^+$
wavefunction is a maximum when $\varphi_2 \rightarrow \varphi_1$, bad.
The ${^1\Delta}_g$ wavefunctions are intermediate. This suggests that
the ordering of states is as in Figure \ref{chap10-fig3}. Indeed,
evaluation of the energy expression for equation
(\ref{chap10-eqno17}), leads to
\begin{eqnarray}
E \left( {^1\Sigma}^+_g \right) &= 2 h_{\pi} + J_{xx} + K_{xy}\cr
E \left( {^1\Delta}_g \right) &= 2h_{\pi} + J_{xx} - K_{xy}\cr
E \left( {^1\Delta}^-_g \right) &= 2 h_{\pi} + J_{xy} + K_{xy}\cr
E \left( {^3\Sigma}^-_g \right) &= 2 h_{\pi} + J_{xy} - K_{xy}
\end{eqnarray}
where $h_{\pi}$ is the one-electron energy for a $\pi$ orbital, and
\begin{equation}
E \left( {^1\Delta}^+_g \right) = E \left( {^1\Delta}^-_g \right) .
\end{equation}
Thus, the state separations are $2K_{xy}$ as indicated in Figure
\ref{chap10-fig3}.
\subsubsection{The $\pi^4$ Configuration}
With four electrons distributed among two $\pi$ orbitals, $\pi_x$
and $\pi_y$, there is only one possible, Pauli allowed, wavefunction
\begin{equation}
\Psi = {\cal A} \left\{ \left[ x (1) \alpha (1) \right] \left[ y (2)
\alpha (2) \right] \left[ x (3) \beta (3) \right] \left[ y (4) \beta
(4) \right] \right\}.
\label{chap10-eqno18}
\end{equation}
In order to determine the symmetry of equation (18), consider first the symmetry of
\begin{equation}
{\cal A} \left[ x (1) \alpha (1) \right] \left[ y (2) \alpha (2)
\right] = \left( xy - yx \right) \alpha \alpha .
\end{equation}
Rotating through an angle $\gamma$ changes this wavefunction to
\begin{equation}
{\cal A} \left[ {\bar{x}} (1) \alpha (1) \right] \left[ {\bar{y}} (2)
\alpha (2) \right],
\label{chap10-eqno19}
\end{equation}
where ${\bar{x}} = x \cos \gamma - y \sin \gamma$ and where
${\bar{y}} = x \sin \gamma + y \cos \gamma$. However,
because both orbitals have the same spin, and because of the
antisymmetrizer, the new wavefunction is equal to the old one
\begin{eqnarray}
{\cal A} \left[ {\bar{x}} \alpha \right] \left[ {\bar{y}} \alpha
\right] &=& \cos^2 {\cal A} [ x \alpha ] [ y \alpha ] - \sin^2 {\cal
A} [ y \alpha ] [ x \alpha ]\cr
&=& \left( \cos^2 \gamma + \sin^2 \gamma \right) {\cal A} [ x
\alpha ] [ y \alpha ] {\cal A} [ x \alpha ] [ y \alpha ].
\label{chap10-eqno20}
\end{eqnarray}
Thus, the wavefunction is a $\Sigma$ state. Since there is one electron
in the $y$ orbital, it is of $\Sigma^-$ symmetry. Similarly, rotating the
wavefunction equation (\ref{chap10-eqno18}) through an angle $\gamma$, leads to
\begin{eqnarray}
{\bar{\Psi}}_{20} = R ( \gamma ) \Psi &=& {\cal A} \left[ {\bar{x}}
\alpha \right] \left[ {\bar{y}} \alpha \right] \left[ {\bar{x}} \beta
\right] \left[ {\bar{y}} \beta \right]\cr
&=& {\cal A} \left[ x \alpha \right] \left[ y \alpha \right] \left[ x
\beta \right] \left[ y \beta \right] = \Psi_{20}.
\end{eqnarray}
That is, $\Psi_{20}$ is a $\Sigma$ state. Because there are now two
electrons in $y$ orbitals $\sigma_v \Psi_{20} = + \Psi_{20}$. Thus,
the $\pi^4$ configuration equation (\ref{chap10-eqno18}) leads to a
${^1\Sigma}^+_g$ state. Similarly,
\begin{equation}
\left( \sigma_g \right) ^2 , \left( \sigma_u \right)^2 , \left(
\pi_g \right)^4 , \left( \pi_u \right)^4 , \left( \delta_g
\right)^4 , \left( \delta_u \right)^4 , \dots
\end{equation}
configurations all lead to ${^1\Sigma}^+_g$. Such configurations
are referred to as closed shells.
\subsubsection{The $\pi^3$ Configuration}
With one electron in a $\pi$ orbital, there are four possible wavefunctions
\begin{eqnarray}
\phi_{x \beta} &= \pi_x \beta ,\cr
\phi_{x \alpha} &= \pi_x \alpha ,\cr
\phi_{y \beta} &= \pi_y \beta ,\cr
\phi_{x \beta} &= \pi_y \alpha ,
\label{chap10-eqno21}
\end{eqnarray}
comprising the four degenerate wavefunctions for a ${^2\Pi}$ state,
${^2\Pi}_g$ or ${^2\Pi}_u$.
In addition, with three electrons distributed among the
two $\pi_x$ and $\pi_y$ orbitals, we find four Pauli allowed wavefunctions
\begin{eqnarray}
\psi_{x \beta} &= {\cal A} \left( \pi_x \alpha \right) \left( \pi_y
\alpha \right) \left( \pi_x \beta \right)\cr
\psi_{y \beta} &= {\cal A} \left( \pi_x \alpha \right) \left( \pi_yu
\alpha \right) \left( \pi_x \beta \right)\cr
\psi_{x \alpha} &= {\cal A} \left( \phi_x \beta \right) \left( \pi_y
\beta \right) \left( \pi_x \alpha \right)\cr
\psi_{y \alpha} &= {\cal A} \left( \pi_x \beta \right) \left( \pi_y
\beta \right) \left( \pi_y \alpha \right)
\label{chap10-eqno22}
\end{eqnarray}
comprising a ${^2\Pi}$ state (${^2\Pi}_g$ or ${^2\Pi}_u$ depending on
the inversion symmetry of the orbitals). Using equation
(\ref{chap10-eqno19}) and equation (\ref{chap10-eqno20}), we find
\begin{eqnarray}
R ( \gamma ) \psi_{x \beta} &=& {\cal A}| ( {\bar{x}} \alpha ) (
{\bar{y}} \alpha ) ( {\bar{x}} \beta )\cr
&=& {\cal A} ( x \alpha ) ( y \alpha ) ( {\bar{x}} \beta ) = \psi_{x
\beta} \cos \gamma - \psi_{y \beta} \sin \gamma
\end{eqnarray}
whereas
\begin{equation}
R ( \gamma ) \phi_{x \beta} = \phi_{x \beta} \cos \gamma - \phi_{y
\beta} \sin \gamma .
\end{equation}
Thus, the four wavefunctions in equation (\ref{chap10-eqno21}) and the
four wavefunctions in equation (\ref{chap10-eqno22}) transform in
exactly the same way, both corresponding to ${^2\Pi}$ states.
In general, the symmetry for a shell with one hole (i.e., $(\pi)^3 ,
(\delta)^3,(\sigma)^2$) is the same as the symmetry of the orbital.
\subsubsection{Inverted Configurations}
The spin-orbit coupling term of the Hamiltonian is
\begin{equation}
h_{so} = \xi (r) l \cdot s .
\end{equation}
For an orbital having angular momentum projections of $m_l$ and $m_s$,
the spin orbit energy is
\begin{equation}
E_{so} = \langle | \xi (r) l \cdot s | \rangle = \gamma m_l m_s ,
\end{equation}
where $\gamma > 0$. Thus, the lowest energy spin orbital has $m_l$
and $m_s$ with opposite signs. For a one-electron atom the spin-orbit
coupling splits the state of orbital angular momentum $l$, and
spin angular momentum $s = 1/2$, into two states with total angular momentum
\begin{equation}
j = l + {1 \over 2} ~~~~~\mathrm{higher}
\end{equation}
\begin{equation}
j = l - {1 \over 2} ~~~~~\mathrm{lower}.
\end{equation}
For a many-electron atom, with orbital angular momentum $L$ and
spin angular momentum $S$, the spin-orbit coupling leads to a
series of states characterized by different total angular momentum
\begin{equation}
J = L + S, L + S - 1, \cdots , | L - S |.
\end{equation}
For heavy atoms, e.g., the Pt row, the spin-orbit coupling is
sufficiently large that this description, in terms of $LS$ states,
is not very accurate, only the $J$ is a good quantum number.
For molecules, the situation is similar, except that the $L$
states are replaced by $\Lambda$ states, orbital angular momentum
about the molecular axis, and the total angular momentum
$J$ is replaced by $\Omega$, the total angular
momentum about the molecular axis. We assume, here, that the
separation between various $\Lambda$ states is large compared to
spin-orbit coupling, when those separations are small, e.g., for
Rydberg states, the coupling between spin, orbital, and rotations
angular momenta becomes complicated. Thus,
\begin{eqnarray}
{^2\Pi} &\rightarrow& {^2\Pi}_{{3 \over 2}} = {^2\Pi}_{{1 \over 2}}\cr
{^3\Sigma} &\rightarrow& {^3\Sigma}_0 = {^3\Sigma}_1\cr
{^2\Phi} &\rightarrow& {^2\Phi}_{{7 \over 2}} = {^2\Phi}_{{5 \over 2}}\cr
{^3\Delta} &\rightarrow& {^3\Delta}_1 , {^3\Delta}_2 = {^3\Delta}_3\cr
{^3\Pi} &\rightarrow& {^3\Pi}_0 , {^3\Pi}_1 , {^3\Pi}_2
\end{eqnarray}
Consider, for example, a ${^2\Pi}$ state. The spatial wavefunction
leads to possible projections along the axis of $+1$ and $-1$, while
the spin leads to $+1/2$ and $-1/2$. Thus, the possible total angular
momentum projections, $\Omega$, are $+3/2$ and $-3/2$, leading to
$^2\Pi_{3/2}$, and $+1/2$ and $-1/2$, leading to $^2\Pi_{3/2}$. Just
as for atoms, one electron in the shell stabilizes lowest $\Omega$
leading to a ${^2\Pi}_{1/2}$ ground state (the regular case), while
three electrons stabilizes highest $\Omega$, leading to a
${^2\Pi}_{3/2}$ ground state (the inverted case). Thus an
experimental determination of which $\Omega$ is lower, can lead to an
assignment for the orbital configuration. When available, the
$\Omega$ ordering is indicated by $r$ or $i$, for regular or inverted.
\section{Homonuclear Molecules}
Starting with the ground state configurations of the atoms Na-Ar
\begin{equation}
\includegraphics[scale=0.75]{fig6-06c}
\end{equation}
and combining to form wavefunctions for homonuclear diatomic
molecules, leads to the group state configurations
\begin{equation}
%figure missing here, p 8.2-2
\end{equation}
In discussing the molecular states, we will often find it convenient
to utilize MOs constructed as symmetry combinations of
atomic orbitals
\begin{equation}
%figure missing here, p 8.2-4
\end{equation}
which will be analyzed in more detail in the balance of this chapter.
The ground state configurations from Na$_2$ to Ar$_2$, are listed in
Tables \ref{chap10-tab4a}--\ref{chap10-tab4b}.
\begin{table}
\caption{Properties of third row homonuclear diatomic
molecules.$^a$ The first column if the number of valence electrons.}
\label{chap10-tab4a}
\begin{tabular}{cccccccc}\\ \hline
& Molecule & $R_e$ (\AA)& $\omega_e$ (cm$^{-1}$)&\multicolumn{2}{c}{$D_0$}
&\multicolumn{2}{c}{IP (eV)}\cr
& & & &eV &kcal & molecule & atom$^b$\cr
1 & Na$^+_2$ & (3.54) & (126) & 0.96 & 22.14\cr
2 & Na$_2$ & 3.0789 & 159.1 & 0.720 & 16.60 & 4.90 & 5.138\cr
3 \cr
4 & Mg$_2$ & 3.890$^5$ & 51.1 & 0.0501 & 1.16 & & 7.644\cr
5 \cr
6 &Al$_2$ & 2.466 & 350.0 & 1.5$^5$ & 35.74 & & 5.984\cr
7 \cr
8 & Si$_2$ & 2.246 & 511.0 & 3.21 & 74.02 & & 8.149\cr
9 & P$^+_2$ & 1.985$^9$ & 672.2 & 4.99 & 115.07\cr
10 & P$_2$ & 1.8934 & 780.8 & 5.033 & 116.06 & 10.53 & 10.484\cr
11 & S$^+_2$ & (1.82$^5$) & 790 & 5.37 & 123.83\cr
12 & S$_2$ & 1.8892 & 725.7 & 4.3693 & 100.76 & 9.36 & 10.357\cr
13 & S$^-_2$ & - & 600.8 & 3.95$^5$ & 91.20 & 1.66$^3$ & 2.0772\cr
& Cl$^+_2$ & 1.8915 & 645.6 & 3.95 & 91.09\cr
14 & Cl$_2$ & 1.987$^9$ & 559.7 & 2.4794 & 57.17 & 11.50 & 13.01\cr
15 & Cl$_2$ & & & 1.26 & 29.06 & 2.39 & 3.615\cr
& Ar$^+_2$ & (2.43) & & 1.33 & 30.67\cr
16 & Ar$_2$ & 3.758 & [25.7] & 0.0105 & 0.24 & 14.54 & 15.755\cr
\hline
\end{tabular}
\end{table}
\begin{table}
\caption{MO configurations.}
\label{chap10-tab4b}
\begin{tabular}{ccccccccc}\\ \hline
& Molecule & Symmetry & $\sigma_g3s$ & $\sigma_u3s$ & $\sigma_g
3p$ & $\pi_u3p$ & $\pi_g3p$ & $\sigma_u3p$\cr
1 & Na$^+_2$ & ${^2\Sigma}^+_g$ & 1 & 0 & 0 & 0 & 0 & 0\cr
2 & Na$_2$ & ${^1\Sigma}^+_g$ & 2 & 0 & 0 & 0 & 0 & 0\cr
3 &\cr
4 & Mg$_2$ & ${^1\Sigma}^+_g$ & 2 & 2 & 0 & 0 & 0 & 0\cr
5 &\cr
6 &Al$_2$ & ${^3\Sigma}^-_g$ & 2 & 2 & 0 & 2 & 0 & 0\cr
7 &\cr
8 & Si$_2$ & ${^3\Sigma}^-_g$ & 2 & 2 & 2 & 2 & 0 & 0\cr
9 & P$^+_2$ & ${^1\Pi}_{ui}$ & 2 & 2 & 2 & 2 & 3 & 0\cr
10 & P$_2$ & ${^1\Sigma}^+_g$ & 2 & 2 & 2 & 4 & 0 & 0\cr
11 & S$^+_2$ & ${^2\Pi}_{gr}$ & 2 & 2 & 2 & 4 & 1 & 0\cr
12 & S$_2$ & ${^3\Sigma}^-_g$ & 2 & 2 & 2 & 4 & 2 & 0\cr
13 & S$^-_2$ & ${^2\Pi}_{gi}$ & 2 & 2 & 2 & 4 & 3 & 0\cr
& Cl$^+_2$ & ${^2\Pi}_{gi}$ & 2 & 2 & 2 & 4 & 3 & 0\cr
14 & Cl$_2$ & ${^1\Sigma}^+_g$ & 2 & 2 & 2 & 4 & 3 & 0\cr
15 & Cl$_2$ & ${^2\Sigma}^+_u$ & 2 & 2 & 2 & 4 & 4 & 1\cr
& Ar$^+_2$ & & 2 & 2 & 2 & 4 & 4 & 1\cr
16 & Ar$_2$ & ${^1\Sigma}^+_g$ & 2 & 2 & 2 & 4 & 4 & 2\cr
\hline
\end{tabular}\\
$^a$ See reference 1. Unless stated otherwise, all experimental data
on diatomic molecules in this chapter are from this reference.
$^b$ Table \ref{chap5-table5} of Chapter 5.
$^c$ The subscript $r$ or $i$ indicates a less than
half-filled, $r$ for regular, or more than half-filled,
$i$ for inverted, shell with $\lambda \not= 0$.
$^d$ $D_0(AB)$ is the energy difference between separated atoms A
and B, and the lowest, $v$, vibrational level of the molecule, $v = 0$.
\end{table}
\subsection{Na$_2$}
Except for the presence of the two Ne cores, the wavefunctions for
Na$_2$
\begin{equation}
\left( \phi_{3sl} \phi_{3sr} + \phi_{3sr} \phi_{3sl} \right) \left(
\alpha \beta - \beta \alpha \right)
\end{equation}
and Na$^+_2$
\begin{equation}
\left( \phi_{3sl} + \phi_{3sr} \right) \alpha
\end{equation}
are quite analogous to those of H$_2$ and H$_2^+$. These core
orbitals will be ignored in all wavefunctions. In MO
language, the wavefunction for Na$_2$ is
\begin{equation}
\phi_{\sigma_g 3s} \phi_{\sigma_g3s} ( \alpha \beta - \beta \alpha )
\end{equation}
and for Na$_2^+$ is
\begin{equation}
\phi_{\sigma_g3s}\alpha
\end{equation}
where
\begin{equation}
\phi_{\sigma_g3s} = \phi_{3sl} + \phi_{3sr}.
\end{equation}
Since H$_2$ is considerably more strongly bound than H$^+_2$ (there being two
electrons in the bonding orbital rather than one) it is surprising
to find that Na$_2$ is more \emph{weakly} bound than Na$^+_2$, and the same for
other alkali diatomics. The origin of this effect will be explained
later.
\subsection{Mg$_2$}
In order to make a bond between Mg atoms, we must unpair the lobe
pairs of orbitals on each atom
\begin{equation}
% figure missing p 8.2-5
%\includegraphics
\end{equation}
The net result is a very weak bond,
\begin{equation}
D_0 ( {\rm Mg}_2) = 0.05 {\rm eV} = 1.15~{\rm kcal}.
\end{equation}
At large distances, the overlap of the lobe orbitals is too small to unpair
the orbitals on each atom. The result is a potential curve with a small
hump, and a small minimum, as shown in Figure \ref{chap10-fig4}.
\begin{figure}
% fig 4 not foudn (p 8.2-6)
%\includegraphics[scale=0.75]{fg10-}
\caption{}
\label{chap10-fig4}
\end{figure}
We saw a bonding and a nonbonding pair:
\begin{equation}
\includegraphics{fg10-4a}
\end{equation}
Since the net bond is almost negligible, we can consider the second
pair as an antibond equal in strength to the bond.
In the MO description, the configuration for Mg$_2$ is
\begin{equation}
\left( \sigma_g 3s \right)^2 \left( \sigma_u 3s \right)^2.
\end{equation}
Thus, with two electrons in the bonding orbital, $\sigma_g 3s$, and two
electrons in the antibonding orbital, $\sigma_u 3s$ it is expected that
Mg$_2$ is not
bound. On the other hand, Mg$_2^+$, with only one electron in the
$\sigma_u 3s$ orbital should be strongly bound, as indicated in Figure
\ref{chap10-fig4}.
\subsection{P$_2$}
We will skip Al$_2$ and Si$_2$ for now, and consider P$_2$. As indicated in
\begin{equation}
\includegraphics{fg10-4b}
\end{equation}
we can form three two-electron bonds, one $\sigma$ bond, and two $\pi$
bonds. As a result, the bond energy of P$_2$ is quite high,
\begin{equation}
D_0 ({\rm P}_2) = 5.03 {\rm eV} = 116.1 {\rm kcal}.
\end{equation}
In MO language, the three bonding orbitals are $\sigma_g3p$ (top
pair), $\pi_u 3p_x$ (bottom pair), and $\pi_u 3p_y$ (bottom pair, but
rotated out of the plane):
\begin{equation}
\includegraphics{fg10-4c}
\end{equation}
so that the total valence configuration is
\begin{equation}\left( \sigma_g 3s \right)(^2 \left( \sigma_u 3s
\right)^2 \left(\sigma_g 3p \right)^2 \left( \sigma_u 3p \right)^2.
\end{equation}
Since all occupied orbital shells are completely filled, closed shell,
the total symmetry is ${^1\Sigma}^+_g$.
The $\pi$ bond is expected to be weaker than the $\sigma$ bond, lower overlap,
and hence, the lowest ionization of P$_2$ should be from the $\pi_u
3p$ orbital, leading
to a ${^2\Pi}_u$ state for P$_2^+$. Since the $\pi_u3p$ orbital is bonding, we
expect for P$^+_2$ a larger $R_e$, a smaller
vibrational frequency, $\omega_e$, and a weaker bond.
These expectations are correct,
\begin{equation}
\begin{array}{lcc} \\
& \mathrm{P}_2 & \mathrm{P}_2^+\\
R_e ( {\rm \AA} ) & 1.8934& 1.9859\cr
\omega_e ( {\rm cm}^{-1} ) & 780.8 & 672.2\cr
D_0 ( {\rm kcal} ) & 116.1& 115.1
\end{array}
\end{equation}
although the bond is not weakened much.
\subsection{S$_2$}
For S$_2$, the favorable bonding configurations are
\begin{equation}
\includegraphics{fg10-4d}
\label{chap10-eqno24}
\end{equation}
all of which correspond to the MO configuration of
\begin{equation}
\left( \sigma_g 3s \right)^2 \left( \sigma_u 3s \right)^2 \left( \sigma_g
3sp \right)^2 \left( \pi_u 3p \right)^4 \left( \pi_g 3p \right)^2
.
\label{chap10-eqno25}
\end{equation}