Добавил:
Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:

Baer M., Billing G.D. (eds.) - The role of degenerate states in chemistry (Adv.Chem.Phys. special issue, Wiley, 2002)

.pdf
Скачиваний:
17
Добавлен:
15.08.2013
Размер:
4.29 Mб
Скачать

726

a. j. c. varandas and z. r. xu

 

and, similarly,

 

 

 

 

 

 

 

T^^SyT^ 1 ¼ S^y

 

T^^SzT^ 1 ¼ S^z

ðC:35Þ

Clearly, the above equations and Eq. (C.20) prove Eq. (C.25).

 

Finally, we demonstrate that

 

 

 

 

 

 

^2

 

 

^ 2S

 

ðC:36Þ

 

T

¼ ð 1Þ

 

From Eqs. (C.8) and (C.23), we have for S ¼ 21

 

 

 

^2

^

2

2

^

ðC:37Þ

 

T1=2 ¼ ðisyKÞ

 

¼ sy

¼ 1

^

is the unit operator in a 2 2 vector space. Note that, for spinless

where 1

 

^

 

 

1 vector space (and

particles, we have chosen U to be the unit operator in a 1

^

^

 

 

 

hence T0

¼ K), which leads to

 

 

 

 

^2

^2

^

ðC:38Þ

 

T0

¼ K

¼ 1

and hence proves Eq. (C.36).

The above discussion is now generalized to arbitrary spin values. First, we note that twice application of the time-reversal operator leads the system back to

 

 

 

 

^2

c

 

 

 

^2

 

^

 

 

its original state c, that is, T

¼ cc. Thus, we have T

¼ c1. Next, consider

the following two relations

 

 

 

 

 

 

 

 

 

hT^fjT^2c ðhfjT^yÞðT^2jciÞ ¼ hfT^yT^2jcÞ

?¼ hfjT^ci

?

¼ hT^cjfi ðC:39Þ

h

T^f

T^2ci ¼ c

T^f c

i

 

 

 

 

 

ð

C:40

Þ

j

i ¼ h

j

 

 

 

 

 

 

 

Thus, we have

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

^

 

 

^

 

 

ðC:41Þ

 

 

 

 

hTcjfi ¼ chTfjci

 

Similarly, we can show that

 

 

 

 

 

 

 

 

 

 

 

 

 

^

^2

 

^

 

 

ðC:42Þ

 

 

 

 

hTcjT

fi ¼ hTfjci

 

 

 

 

 

 

 

^

 

^

 

 

ðC:43Þ

 

 

 

 

 

hTfjci ¼ chTcjfi

 

from Eqs. (C.41) and (C.43) we can obtain

 

 

 

 

 

 

 

 

 

 

^

 

2

^

 

 

ðC:44Þ

 

 

 

 

hTcjfi ¼ c

hTcjfi

 

and hence

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

c2 ¼ 1

 

 

ðC:45Þ

permutational symmetry and the role of nuclear spin

727

^2 ¼ ^ ^^^ 1 ^^^ 1 which proves that T 1. Now, by substituting Eq. (C.8) in TrT , TpT ,

^^^ 1 ^

and TST , we may show that U satisfies equations similar to Eqs. (C.24) and (C.25).

As a first application, consider the case of a single particle with spin quantum number S. The spin functions will then transform according to the IRREPs DðSÞðaÞ of the 3D rotational group SOð3Þ, where a is the rotational vector,

written in the operator form as [36]

 

 

 

 

i

 

D^ðSÞðaÞ ¼ exp

 

S^ a

ðC:46Þ

h

^

 

The spin operator S is an irreducible tensor of rank one with the following

transformational properties

 

D^ðSÞðaÞSD^^ðSÞðaÞ 1 ¼ ^gðaÞS^

ðC:47Þ

where ^gðaÞ is an operator of SOð3Þ. Let us then take ^gðaÞ to be a rotation by p around the y axis. Thus, from Eqs. (C.46) and (C.47), one gets

exp

 

i

S^x exp

i

¼ S^x

 

 

 

 

S^y

 

S^y

ðC:48Þ

h

h

exp

 

i

S^y exp

i

¼ S^y

 

 

 

 

S^y

 

S^y

ðC:49Þ

h

h

exp

 

i

S^z exp

i

¼ S^z

 

 

 

S^y

 

S^y

ðC:50Þ

h

h

Comparing Eqs. (C.48)–(C.50) with Eq. (C.24), one obtains

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

^

¼ exp

 

 

i

^

 

 

 

ðC:51Þ

 

U

 

h pSy

 

 

 

^T

^

 

 

 

 

 

 

 

 

 

 

 

 

 

Since Sy

¼ Sy, we then have

 

 

 

 

 

 

 

 

 

 

 

 

 

 

U^ y ¼ exp

i

 

 

 

 

 

 

 

 

 

 

pS^yT

 

 

 

 

 

ðC:52Þ

 

h

 

 

 

 

 

 

¼ exp

i

 

 

 

 

 

 

 

 

 

p S^y

 

 

 

 

 

ðC:53Þ

 

h

 

 

 

 

 

 

 

 

 

 

i

 

 

 

 

i

 

 

 

¼ exp

 

2pS^yT exp

 

pS^y

 

ðC:54Þ

 

h

h

 

¼ D^ðSÞð2pjÞU^

 

 

 

 

 

 

ðC:55Þ

 

¼ ð 1Þ

2S ^

 

 

 

 

 

 

 

ðC:56Þ

 

 

U

 

 

 

 

 

 

 

728

a. j. c. varandas and z. r. xu

where 2pj indicates a 2p rotation about the y axis. Thus, we have

^

¼ exp

 

i

^

^

ðC:57Þ

T

h pSy

K

and finally, by comparing with Eq. (C.51), one gets

^2

^

2S

ðC:58Þ

T

¼ ð 1Þ

 

Finally, for a system of n identical particles, the result is

 

¼ exp

i

exp

 

i

 

 

exp

i

 

 

U^

 

pS^y;1

 

 

pS^y;2

 

pS^y;n

ðC:59Þ

h

h

h

and hence

 

 

 

 

 

Pn

 

 

 

 

 

 

 

 

 

 

 

^2

^

2Si

 

^

2S

 

 

 

ðC:60Þ

 

 

 

 

i 1

 

 

 

 

 

 

 

T

 

¼ ð 1Þ

¼

 

 

¼ ð 1Þ

 

 

 

 

APPENDIX D: DEGENERATE AND NEAR-DEGENERATE

VIBRATIONAL LEVELS

Here, we discuss the motion of a system of three identical nuclei in the vicinity of the D3h configuration. The conventional coordinates for the in-plane motion are employed, as shown in Figure 5. The normal coordinates ðQx; Qy; QzÞ, the plane polar coordinates ðr; j; zÞ, and the Cartesian displacement coordinates ðxi; yi; ziÞ of the three nuclei ði ¼ 1; 2; 3Þ are related by [20,94]

Qx ¼ r cosj ¼ p3

x1 þ

 

2 x2 þ

 

 

2

y2

þ

2 x3 2

y3

 

ðD:1Þ

 

1

 

 

 

 

 

 

 

 

p3

 

 

1

 

 

 

p3

 

 

 

 

 

 

 

1

 

 

 

 

 

 

 

 

 

 

 

 

 

 

1

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

p3

 

1

 

 

 

 

 

p3

1

 

 

 

 

Qy ¼ r sinj ¼ p3

y1 þ

2

x2

 

 

y2

þ

 

2

x3

 

 

y3

 

ðD:2Þ

2

2

1

 

1

 

 

p3

 

 

 

 

1

 

 

p3

 

 

 

 

ðD:3Þ

Qz ¼ z ¼ p3

x1 þ 2 x2

2 y2

þ 2 x3 þ

 

2 y3

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

where the coordinates ðQx; QyÞ are the doubly degenerate modes belonging to the E0 IRREP in D3h, and Qz belongs to the A01 one. Note that Qx is symmetric with respect to the xz plane, while Qy is antisymmetric.

The coordinates of interest to us in the following discussion are Qx and Qy, which describe the distortion of the molecular triangle from D3h symmetry. In the harmonic-oscillator approximation, the factor in the vibrational wave

permutational symmetry and the role of nuclear spin

729

function due to the two degenerate modes is then (except for a normalization constant and dependence on Qz) given by

ð

 

Þ ’

 

2a ð

 

Þ

 

2b

 

½ a

ð

x þ

y Þ

 

&

ð

 

Þ

w

v1; v2a; v2b

 

Hv

 

pa2Qx

Hv

 

pa2Qy exp

2

 

Q2

Q2

=2

 

 

D:4

 

where Hv2a and Hv2b are Hermite polynomials of order v2a and v2b, respectively; v2a and v2b are the vibrational quantum numbers, and a2 ¼ 2pn2=h, with n2 being the frequency of the degenerate mode.

Let us then consider the case where the degenerate mode is doubly excited. In this case, v2 ¼ v2a þ v2b ¼ 2 and the corresponding vibrational energy level will be triply degenerate with the associated wave functions being given by

w1 ¼ wðv1; 2; 0Þ " 4a2Qx2 2

ðD:5Þ

w2 ¼ wðv1; 1; 1Þ " 4a2QxQy

ðD:6Þ

w3 ¼ wðv1; 0; 2Þ " 4a2Qy2 2

ðD:7Þ

Note that only the polynomial factors have been given, since the exponential parts are identical for all wave functions. Of course, any linear combination of the wave functions in Eqs. (D.5)–(D.7) will still be an eigenfunction of the vibrational Hamiltonian, and hence a possible state. There are three such linearly independent combinations which assume special importance, namely,

w10

¼ w1 w3 þ 2iw2 " 4a2ðQx2 Qy2 þ 2iQxQyÞ

ðD:8Þ

w20

¼ w1 þ w3 " 4a2ðQx2 þ Qy2Þ 4

ðD:9Þ

w30

¼ w1 w3 2iw2 " 4a2ðQx2 Qy2 2iQxQyÞ

ðD:10Þ

By using the plane polar coordinates defined in Eq. (D.1), one obtains

w10

4a2r2expð a2r2

=2Þexpð2ijÞ

ðD:11Þ

w20

4ða2r2 1Þexpð a2r2=2Þ

ðD:12Þ

w30

4a2r2expð a2r2

=2Þexpð 2ijÞ

ðD:13Þ

These new wave functions are eigenfunctions of the z component of the angular

momentum p^z ¼ ih qjq with eigenvalues mv2 ¼ þ2; 0; 2 in units of h. Thus, Eqs. (D.11)–(D.13) represent states in which the vibrational angular momentum

of the nuclei about the molecular axis has a definite value. When treating the vibrations as harmonic, there is no reason to prefer them to any other linear combinations that can be obtained from the original basis functions in

730 a. j. c. varandas and z. r. xu

Eqs. (D.5)–(D.7). However, when perturbations occur due to anharmonicity, the wave functions in Eqs. (D.11)–(D.13) will provide the correct zerothorder ones. The quantum numbers v2a and v2b are therefore not physically significant, while v2 and mv2 or v2 and l2 ¼ jmv2 j are. It should also be pointed out that the degeneracy in the vibrational levels will be split due to anharmonicity [28].

Now, consider the general case of a v2 multiply excited degenerate vibrational level where v2 > 2, which is dealt with by solving the Schro¨dinger equation for the isotropic 2D harmonic oscillator with the Hamiltonian assuming the form [95]

 

^

 

hpl

 

q2

 

q2

 

2

 

2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Hv

¼

 

2

 

qqx2 qqy2 þ

qx

þ

qy

 

ð

D:14

Þ

 

 

 

 

!

 

 

 

where we

have used

the

 

dimensionless

normal

coordinates q

i ¼

p

Q

i

 

 

 

 

 

 

 

p

 

 

 

 

 

 

 

a2

 

 

ði ¼ x; yÞ,

with a2 ¼ 2pn2=h ¼

l=h. The

transformation of such a

Hamil-

 

 

 

 

 

tonian into polar coordinates leads to

 

 

 

 

 

 

 

 

 

 

 

 

H^v ¼ 2

 

qr2 þ r qr

þ r2 qj2

r2

 

hpl

 

 

 

 

 

 

1 q2

 

 

 

q2 1 q

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Separation of variables can then be achieved by using

wv ¼ RðrÞ ðjÞ

ðD:15Þ

ðD:16Þ

where

 

RðrÞ ¼ FðrÞexpð r2=2Þ

ðD:17Þ

and

 

X

 

1

ðD:18Þ

FðrÞ ¼ rs anrn

n ¼ 0

 

Assuming now that the power series expansion in FðrÞ can be terminated to keep RðrÞ well behaved at large r values, it may be shown [95] that

ðjÞ ¼ ð2pÞ 1=2expðimv2 jÞ mv2

¼ v2; ðv2 2Þ; . . . ; 1 or 0

ðD:19Þ

RðrÞ ¼ Nv2l2 rl2 Lnl2 ðr2Þexpð r2=2Þ

l2 ¼jmv2 j n ¼ðv2 þ l2Þ=2

ðD:20Þ

permutational symmetry and the role of nuclear spin

731

where Lln2 ðr2Þ are the associated Laguerre polynomials of order normalization factor assumes the form

 

s

Nv2l2 ¼

2½ðv2 l2Þ=2&!

3

 

f½ð

v2

þ

l2

Þ

=2

&

g

 

 

 

 

 

!

 

 

n, and the

ðD:21Þ

Let us now examine the case of a 3D harmonic oscillator possessing three degenerate normal coordinates ðQ1; Q2; Q3Þ, with the degenerate mode being v multiply excited; v ¼ v1 þ v2 þ v3. There are then ðv þ 1Þðv þ 2Þ=2 degenerate vibrational wave functions and energy levels for each value of v, corresponding to the possible different combinations of v1, v2, and v3. It is now convenient to define the polar coordinates ðr; y; jÞ by the corresponding dimensionless normal coordinates ðq1; q2; q3Þ according to

 

 

 

 

 

 

 

 

 

 

 

 

 

q1

¼ r siny cosj

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

q2

¼ r sinysinj

 

 

 

 

 

 

 

 

 

 

ðD:22Þ

 

 

 

 

 

 

 

 

 

 

 

 

 

q3

¼ r cosy

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

In such coordinates, the Hamiltonian assumes the form [95]

 

 

 

 

 

 

 

 

 

Hv

¼

 

2 qq12 qq22

qq32

þ q1

þ q2

þ q3

 

 

 

ðD:23Þ

 

 

^

 

 

 

hpl

 

q2

 

 

 

q2

 

 

 

q2

2

2

2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Transformation of the Hamiltonian into polar coordinates then leads to

 

H^v ¼

2 r2 qr

r2 qr þ r2 siny qy siny qy þ r2 sin2 y qj2 r2

 

 

hpl

1

 

 

q

 

 

 

q

 

 

1 1

 

 

q

 

 

 

q

1 1

 

q2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

ðD:24Þ

while the vibrational wave equation assumes the form

 

 

 

 

 

 

 

wv ¼ 0

r2 qr r2 qr þr2 siny qy

siny qy þr2 sin2y qj2 þ hpl r2

 

1 q

 

q

 

 

 

1 1 q

 

 

 

 

q

1 1

 

 

q2

 

2E

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

ðD:25Þ

Separation of variables may now be obtained by using

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

wv ¼ RðrÞ ðyÞ ðjÞ

 

 

 

 

 

 

 

 

 

 

ðD:26Þ

732

 

 

 

 

 

a. j. c. varandas and z. r. xu

 

 

 

 

 

 

which upon insertion into Eq. (D.25) leads to

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

d2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

þ m2 ¼ 0

ðD:27Þ

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

d2j

siny dy siny dy þ lðl þ 1Þ sin

2y ¼ 0

ðD:28Þ

 

1

 

d

 

 

 

 

 

d

 

 

 

 

 

 

 

m2

 

 

 

 

 

 

r2 dr

r

 

dr

þ hpl r

 

 

 

ð r2

1Þ

 

¼

0

ð

D:29

Þ

 

1 d

 

 

 

2

 

d

 

 

2E

2

 

 

 

l l þ

R

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

These have as solutions

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

ðjÞ ¼ ð2pÞ 1=2expðimjÞ

 

 

 

 

 

 

 

 

ðD:30Þ

 

 

 

 

 

ðyÞ ¼ NljmjPljmjðcos yÞ

 

 

 

 

 

 

 

 

 

 

ðD:31Þ

 

 

 

 

 

RðrÞ ¼ NvlrlLtlþ1=2ðr2Þexpð r2=2Þ

 

 

ðD:32Þ

where

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

l ¼ v; v 2; v 4; . . . ; 1 or 0

 

 

 

ðD:33Þ

 

 

 

 

 

 

m ¼ 0; 1; 2; . . . ; l

 

 

 

 

 

 

 

 

ðD:34Þ

 

 

 

 

 

 

 

 

t ¼ ðv þ l þ 1Þ=2

 

 

 

 

 

 

 

 

 

 

 

ðD:35Þ

The functions Pjlmj are associated Legendre polynomials of order jmj and degree l, and Lltþ1=2ðr2Þ are associated Laguerre polynomials of degree ðv 1Þ=2 in r2. In turn, the normalization factors are found to be

 

 

¼

2 l!

 

s!

ð

 

Þ

N

ljmj

 

ð 1Þl

 

ð2l þ 1Þ

ðl jmjÞ!

 

D:36

 

 

 

l

 

 

 

 

2

 

ð

l

 

 

m

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

þ j jÞ

 

 

 

 

 

 

s

 

 

 

Nvl ¼

 

f½ð

2½ðv 1Þ=2&!

g

3

 

ðD:37Þ

 

 

 

 

v

þ

l

þ

1

Þ

=2

&

 

 

 

 

 

 

 

 

 

 

 

 

 

!

 

 

 

 

 

 

In the configuration space spanned by ðq1; q2; q3Þ, we may then define the vibrational angular momentum p through its classical components, that is,

p1 ¼ q2p3 q3p2 and its ð123Þ cyclic permutations ðD:38Þ

where pi are the conjugate momenta associated to qi ði ¼ 1; 2; 3Þ. The operators associated with p2 ¼ p21 þ p22 þ p23 and its projection pz (denoted M3 in [95])

permutational symmetry and the role of nuclear spin

along the z axis assume in polar coordinates the form

p^2

¼ h2

 

q2

cosy q

þ

1 q2

 

 

 

þ

 

 

 

 

 

 

qy2

siny

qy

sin2 y

qj2

p^z

¼ ih

q

 

 

 

 

 

 

 

 

 

qj

 

 

 

 

 

 

 

 

 

 

733

ðD:39Þ

ðD:40Þ

As for the 2D case, it can be shown that wv in Eq. (D.26) are eigenfunctions of both p^2 and p^z defined by

^ 2

wv ¼ lðl þ

1Þh

2

wv

ðD:41Þ

M

 

^

 

 

 

 

ðD:42Þ

Mzwv ¼ mhwv

 

 

 

Thus, l and m quantize the vibrational angular momentum and its z component. So far, we have considered interactions that are degenerate at the harmonicoscillator level of approximation. For two levels that are nearly degenerate by accident in such an approximation, large perturbations may arise due to anharmonicity that are known as Fermi resonances. It should be noted that Fermi resonances occur only between states of the same symmetry. Thus, they cannot occur between two levels with different values of the vibrational angular momentum quantum number l. As usual, Fermi resonances increase the energy of the upper level while decreasing that of the lower one (in common language, they repel each other). Thus, the spectrum of a specific symmetry tends to be

more irregular in the presence of Fermi resonances.

APPENDIX E: ADIABATIC STATES IN THE VICINITY OF A

CONICAL INTERSECTION

I.JAHN–TELLER THEOREM

Following Moffitt and Liehr [73], in this appendix we give a proof of the Jahn– Teller theorem for X3 molecules pertaining to the D3h point group. Let c1 and c2 be the two electronic eigenfunctions that belong to the degenerate electronic states of E0 symmetry (denoted eE0). The two degenerate normal coordinates are Qx and Qy, the former being symmetric and the latter antisymmetric with respect to the xz plane (see Appendix D). Defining complex normal coordinates and electronic eigenfunctions as

Qþ ¼ Qx þ iQy ¼ rexpðijÞ

ðE:1Þ

Q ¼ Qx iQy ¼ rexpð ijÞ

ðE:2Þ

734

a. j. c. varandas and z. r. xu

 

and

 

 

 

cþ ¼ c1 þ ic2

ðE:3Þ

 

c ¼ c1 ic2

ðE:4Þ

the electronic energy of the system is in degenerate-state perturbation theory obtained by solving the secular equation

 

Hþþ W Hþ

W

 

¼

0

ð

E:5

Þ

H þ

H

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

where the matrix elements are given by

 

 

^

^

ðE:6Þ

Hþþ ¼ hcþjHejcþi

H ¼ hc jHejc i

^

^

ðE:7Þ

Hþ ¼ hcþjHejc i

H þ ¼ hc jHejcþi

and the integrations are defined with respect to all the electronic coordinates.

^

Then, by developing He in a power series expansion of the normal coordinates, one gets

H^e ¼ h^0 þ h^1þQ þ h^1 Qþ þ h^2þQ2 þ h^2 Qþ2 þ

ðE:8Þ

where we have considered only the dependence on the degenerate complex normal coordinates Qþ and Q . Substitution of Eq. (E.8) in Eqs. (E.6) and (E.7) gives

 

 

 

 

 

 

Hþþ ¼ hcþjh^0jcþi þ hcþjh^1þjcþiQ þ hcþjh^1 jcþiQþ

 

 

 

 

 

 

 

 

 

 

þ hcþjh^2þjcþiQ2 þ hcþjh^2 jcþiQþ2 þ

 

 

 

 

ðE:9Þ

 

 

 

 

 

 

Hþ ¼ hcþjh^0jc i þ hcþjh^1þjc iQ þ hcþjh^1 jc iQþ

 

 

 

 

 

 

 

 

 

 

þ hcþjh^2þjc iQ2 þ hcþjh^2 jc iQþ2 þ

 

 

 

ðE:10Þ

with similarly expressions for H

 

and H

.

 

 

 

 

 

 

 

 

 

 

 

 

 

^

3

 

?

 

þ

 

cþ

 

?

 

 

þ

 

 

 

 

 

o ¼

ð

 

p =

Þ

 

 

For a C

 

 

,

, and

c

are multiplied by

2

while

 

 

rotation, Q

 

 

 

o

 

 

exp

i 3

 

Q

 

,

c

, and

cþ

are multiplied by

 

¼

exp

ð

2

p

Þ

 

 

 

 

 

 

 

 

 

 

 

 

 

i=3 . Since the Hamiltonian

^þ ^ ^þ ^

must be totally symmetric, it follows that h1 , h1 , h2 , and h2 are multiplied by o, o , o2, and o 2, respectively. The integrals in Eqs. (E.9) and (E.10) will then be different from zero only if the integrands are invariant under all symmetry

^

operations allowed by the symmetry point group, in particular under C3. It is readily seen that the linear terms in Qþ and Q vanish in Hþþ and H . In turn,

permutational symmetry and the role of nuclear spin

735

the first term in Hþ and Hþ vanishes while one of the linear terms (Qþ for Hþ, and Q for Hþ) does not vanish. Thus, neglecting quadratic (and higher order) terms, one obtains

Hþþ ¼ H ¼ W0 Hþ ¼ cQþ Hþ ¼ cQ

ðE:11Þ

Substitution of Eq. (E.11) into Eq. (E.5), leads to

 

p

ðE:12Þ

W ¼ W0 c QþQ ¼ W0 cr

Clearly, Eq. (E.12) shows that to a first approximation the electronic energy varies linearly with displacements in r, increasing for one component state while decreasing for the other. Thus, the potential minimum cannot be at r ¼ 0. This is the statement of the Jahn–Teller theorem for a X3 molecule belonging to the D3h point group.

II. INVARIANT OPERATORS

We follow Thompson and Mead [13] to discuss the behavior of the electronic Hamiltonian, potential energy, and derivative coupling between adiabatic states

in the vicinity of the D3h conical intersection. Let ^ be an operator that

A

^

transforms only the nuclear coordinates, and A be one that acts on the electronic degrees of freedom alone. Clearly, the electronic Hamiltonian satisfies

^ ^

^

^ ^

ðE:13Þ

ðAHeÞAc ¼ AHec

^

 

 

^

since, if c is an eigenfunction of He, Eq. (E.13) just expresses the fact that Ac is

an eigenfunction of the transformed Hamiltonian with the same eigenvalue (for

 

 

^

an arbitrary c, it also follows upon its expansion in eigenfunctions of He). Thus,

^ ^

^ ^ ^

ðE:14Þ

ðAHeÞA ¼ AHe

^

If Eq. (E.14) is satisfied for all elements of some point group G, A will be an invariant operator [13] (the Hermitian conjugate as well as the sum and/or product of two invariant operators are also invariant operators). Such an operator can be expanded in the form

^

X

 

^

ðE:15Þ

He ¼

h gsQ gs

 

gs

 

where Q gs is a nuclear coordinate transforming as the g th component of the IRREP of G, the index s refers to different occurrences of the same IRREP, and