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:
16
Добавлен:
15.08.2013
Размер:
4.29 Mб
Скачать

182

 

 

 

 

 

 

satrajit adhikari and gert due billing

Fx, and so on, we obtain

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Fx ¼ FðX þ x tg f=d12Þ

ð102Þ

 

 

 

 

 

 

 

 

 

Fy ¼ FðY þ y tg f=d12Þ

ð103Þ

 

 

 

 

 

 

 

 

 

Fz ¼ FðZ þ z tg f=d12Þ

ð104Þ

 

 

 

 

 

 

 

 

 

FX ¼ Fðx X tg fd12Þ

ð105Þ

 

 

 

 

 

 

 

 

 

FY ¼ Fðy Y tg fd12Þ

ð106Þ

 

 

 

 

 

 

 

 

 

FZ ¼ Fðz Z tg fd12Þ

ð107Þ

where the

function F is given as

F ¼ ðh=2mÞcos f sin fr RÞ and d12 ¼

m

1ð

1

 

m

 

M

Þ=m

with M

¼

m

m

2 þ

m

 

[72,73]. In

six dimensions, the

 

 

1

=

 

 

1 þ

 

 

3

 

6

amplitudes diðtÞ [in Eq. (97)] will be of dimension N ¼ i¼1Ni. Here, in mass scaled coordinates we have used [48]

r2=d12

¼

1

r2ð1

þ sin y cos fÞ

 

2

R2d12

¼

1

r2ð1

sin y cos fÞ

 

2

r R ¼

1

r2sin y sin f

 

2

ð108Þ

ð109Þ

ð110Þ

Since the geometric phase effect is related to the angle f we express f as

 

 

tg

 

r R

 

 

 

 

f ¼ r2=d1 d1R2

 

 

 

 

 

 

 

 

and obtain

 

 

 

 

 

 

qf

cos f sin f

 

2

 

 

 

¼

 

 

ðX þ x tg f=d1

Þ

 

qx

 

r R

qf cos f sin f

2

 

 

 

 

¼

 

 

ðx X tg fd1

Þ

 

 

qX

 

r R

 

ð111Þ

ð112Þ

ð113Þ

plus similar expressions for the y and z components.

Note that in this TDGH–DVR formulation of quantum dynamics, the inclusion of the geometric phase effects through the addition of a vector potential is very simple and the calculations can be carried out with about the same effort as what is involved in the ordinary scattering case.

Figure 3 shows the results with and without including the geometric phase effect for the D þ H2 reaction. The basis set is taken as 1,1,1,15,15,15, that is,

non-adiabatic effects in chemical reactions

183

Figure 3. Cross-sections obtained with a (1,1,1,15,15,15) basis set and the TDGH–DVR method for the D þ H2 ðv ¼ 1; j ¼ 1Þ ! DH ðv0 ¼ 1; j0Þ þ H reaction at 1.8-eV total energy. The solid line indicates the values obtained without the vector potential and the dashed with a vector potential. The dashed line indicates the experimental results [49–52].

the X; Y; Z variables are treated classically. Altogether 200 trajectories were calculated. We notice that the branching ration, that is, the total reactive crosssection is obtained from the trajectories but the distribution is obtained by a projection of the DVR points on final rotational–vibrational states of the product. The maximum of the distribution is now j0 ¼ 9 (in better agreement with full quantum calculations). It is shifted to j0 ¼ 8 if the geometric phase is included. The agreement with experimental data is good for j0 values <8 but overestimated at higher values. Since part of the system is still treated classically, we attribute this discrepancy to the lacking ability of classical trajectories to yield proper state-resolved reaction cross-sections (see also Fig. 1).

V.CONCLUSION

In this chapter, we discussed the significance of the GP effect in chemical reactions, that is, the influence of the upper electronic state(s) on the reactive and nonreactive transition probabilities of the ground adiabatic state. In order to include this effect, the ordinary BO equations are extended either by using a HLH phase or by deriving them from first principles. Considering the HLH phase due to the presence of a conical intersection between the ground and the first excited state, the general form of the vector potential, hence the effective

184

satrajit adhikari and gert due billing

kinetic energy operator, for a quasi-JT model and for an A þ B2 type reactive system were presented.

The ordinary BO approximate equations failed to predict the proper symmetry allowed transitions in the quasi-JT model whereas the extended BO equation either by including a vector potential in the system Hamiltonian or by multiplying a phase factor onto the basis set can reproduce the so-called exact results obtained by the two-surface diabatic calculation. Thus, the calculated transition probabilities in the quasi-JT model using the extended BO equations clearly demonstrate the GP effect. The multiplication of a phase factor with the adiabatic nuclear wave function is an approximate treatment when the position of the conical intersection does not coincide with the origin of the coordinate axis, as shown by the results of [60]. Moreover, even if the total energy of the system is far below the conical intersection point, transition probabilities in the JT model clearly indicate the importance of the extended BO equation and its necessity.

The integral and differential cross-section obtained by using QCT calculations on the ground adiabatic surface of the D þ H2 system at a total energy of 1.8 eV, clearly indicates the GP effect where the ground state of this system has a conical intersection with its’ first excited state at a total energy of 2.7 eV. Similarly, semiclassical calculations on the same system with or without including a vector potential in the system Hamiltonian confirms this effect. Preliminary calculations with the new TDGH–DVR method also show a less dramatic effect. In the case of the H þ D2 reaction at total energy 2.4 eV, calculated rotational state and scattering angle distributions obtained from the QCT calculations on the LSTH surface demonstrate quantitative change due to the GP effect but the qualitative variation, at least in the integral cross-section, is not significant.

Formulation of the extended BO approximate equations using the HLH phase is based on the consideration of two coupled states. If the ground state of a system is coupled with more than one excited state, it has been demonstrated that the phase factor could be different from the HLH phase factor. In this formulation, we consider the BO coupled equations with the aim of deriving an approximate set of uncoupled equations that will contain the effect of nonadiabatic coupling terms. When the electronic states are degenerate, some of the non-adiabatic coupling terms may become infinite and affect the dynamics of the nuclei irrespective of how far it occurs from the point of the degeneracy. Hence, the importance of non-adiabatic coupling terms has been taken into account when deriving the uncoupled BO from the coupled ones. In this approch, the non-adiabatic coupling terms are not eliminated but shifted from the off-diagonal position to the diagonal one and the BO approximation has been introduced afterward. This shift has been done with the physical assumption that the non-adiabatic coupling matrix guarantees the continuous, single-valued diabatic potential matrix in the CS, that is, along a close path the

non-adiabatic effects in chemical reactions

185

non-adiabatic coupling matrix follows the Bohr–Sommerfeld type quantization rule. This quantization guarantees that all N decoupled equations obtained by deleting the potential coupling terms are invariant under gauge transformations and follow proper boundary conditions. These extended–approximated BO equations are tested for a tri-state system. First, we performed a so-called exact calculation in the diabatic representation to obtain reactive and nonreactive transition probabilities on the ground adiabatic surface and then the extended– approximated BO equations for the ground adiabatic surface are solved to get the relevant results. State-to-state transition probabilities obtained by both calculations indicate that the new approximate BO equations yield correct results for a tri-state system.

Hence, systems having conical intersections between two or more than two electronic states exhibit geometric phase effects. For two-states systems, the HLH phase factor is the same as that obtained by Baer et al. from first principles but the new phase factor appears to be different and depends on the number of electronic states coupled. Considering a conical intersection between the ground and first excited state of the D þ H2 reactive system, the extended BO equations are the same in both of the above-mentioned approaches and we found significant GP effect at a total energy of 1.8 eV. However, it has been possible to obtain good agreement between experiment and theory without including the effect for the H þ D2 system at a total energy 2.4 eV. At this point, it is worth noting that the calculations on the H þ D2 reaction were carried out on a different potential energy surface than the one we used in our calculations. May be the reactivity of one potential energy surface could hide the GP effect while another could expose it. At the same time, the importance of the GP effect is clearly understood in the quasi-JT model. The inclusion of a simple phase factor (HLH) or by using the extended BO equations can change the parity for vibrational transitions in the 2D two-surface model and give good agreement with results obtained by an exact two-state diabatic calculation. Again, calculations on a tri-state 2D quasi-JT model using the extended BO equations (N 2) derived by Baer et al. not only exhibit geometric phase effects but also the new phase factor that changes with the number of electronic states coupled.

APPENDIX A: THE JAHN–TELLER MODEL AND THE HERZBERG–LONGUET–HIGGINS PHASE

When two electronic states are degenerate at a particular point in configuration space, the elements of the diabatic potential energy matrix can be modeled as a linear function of the coordinates in the following form:

W ¼ k

y

x

ðA:1Þ

x

y

 

 

 

 

186

satrajit adhikari and gert due billing

where k is the force constant and (x; y) are the nuclear coordinates. The eigenvalues and eigenvectors of the above matrix represent the adiabatic potential energy surfaces and the columns of the ADT matrix, respectively. In order to carry out this diabatization, we use the following transformations between the Cartesian (x; y) and polar (q; f) coordinates: x ¼ q sin f and y ¼ q cos f.

The eigenvalues and eigenfunctions of this model are

u1; 2 ¼ kq

ðA:2Þ

where q ¼ 0; 1 and f ¼ 0; 2p, and

x1

1

1

ðA:3Þ

1

1

¼ pp cos f=2;

pp sin f=2

 

x2

 

 

 

¼ pp sin f=2;

pp cos f=2

 

respectively.

These adiabatic eigenfunctions depend only on the angular coordinate f and are not single valued in configuration space when f changes to f þ 2p—a rotation that brings the adiabatic wave functions back to their initial position. This multivaluedness of the adiabatic eigenfunctions was first revealed by Herzberg and Longuet-Higgins. In order to avoid multivalued electronic eigenfunctions they suggested that the corresponding nuclear wave function be treated with care. While solving the nuclear SE, this feature needs to be incorporated explicitly through specific boundary conditions. It is worth mentioning that in the HLH state realistic ab initio electronic wave functions posses the multivaluedness feature.

Longuet-Higgins corrected the multivaluedness of the electronic eigenfunctions by multiplying them with a phase factor, namely,

ZjðfÞ ¼ expðiaÞxjðfÞ j ¼ 1; 2

ðA:4Þ

where a ¼ f=2. It is important to note that ZjðfÞ, j ¼ 1; 2 are single-valued complex eigenfunctions.

APPENDIX B: THE BORN–OPPENHEIMER TREATMENT

The total electron–nuclear Hamiltonian of a molecular sytem is defined as

^ ^

^

ðB:1Þ

H ¼ Tn þ Heðe j nÞ

 

non-adiabatic effects in chemical reactions

 

187

^

 

 

^

ðe j nÞ is

the

where Tn

is the kinetic energy operator for the nuclei and He

electronic Hamiltonian and

 

 

 

 

 

^

^

^

ðB:2Þ

 

He ¼ Te þ Vðe j nÞ

^

 

 

 

^

the

with Te being the kinetic energy operator of the electrons and

Vðe j nÞ

potential energy operator as a function of electronic coordinates(e) with nuclear coordinates(n).

The BO expansion of the molecular wave function

XN

ðe; nÞ ¼ ciðnÞxiðe j n0Þ

ðB:3Þ

i¼1

 

where the functions ciðnÞ are the nuclear coordinate-dependent coefficients, later considered as the nuclear wave function, and the xiðe j n0Þs are the electronic eigenfunctions satisfying the equation

^

Þxiðe j n0Þ ¼ uiðn0Þxiðe j n0Þ

i ¼ 1; . . . ; N

ðB:4Þ

Heðn0

Here, the uiðn0Þs are the electronic eigenvalues dependent on the nuclear coordinate n0. Note that n0 n is defined as the adiabatic case and n0 ¼6 n is defined as the diabatic case.

Substituting Eqs. (B.1) and (B.3) into the time-independent Schro¨dinger

equation H ðe; nÞ ¼ E ðe; nÞ, one obtains

 

 

 

 

 

ðT^n þ H^e EÞ X ciðnÞxiðe j n0Þ ¼ 0

ðB:5Þ

Below, we apply the bra–ket notation to electronic coordinates only,

 

hxjðnÞjxiðn0

gji n; n0

Þ

; forn

¼6

n0

ðB:6Þ

Þi ¼ ( dji;ðforn

n0

 

 

 

¼

 

 

 

By returning back to Eq. (B.5), we have

 

 

 

 

 

X

X

 

 

 

 

 

N

N

 

 

 

 

 

TnciðnÞjxiðe j n0Þi þ ciðnÞðHe EÞjxiðe j n0Þi ¼ 0

ðB:7Þ

i¼1

i¼1

 

 

 

 

 

If we consider the ADIABATIC (n0 n) case, we get

 

 

X

X

 

 

 

 

 

N

N

 

 

 

 

ðB:8Þ

TnciðnÞjxiðe j nÞi þ

ciðnÞðuiðnÞ EÞjxiðe j nÞi ¼ 0

i¼1

i¼1

 

 

 

 

 

188

satrajit adhikari and gert due billing

 

Multiplying by hxjj and integrating over electronic coordinates yields

 

 

X

 

 

 

 

N

 

j ¼ 1; . . . ; N

ðB:9Þ

 

hxjjTncnðnÞjxii þ ðujðnÞ EÞcjðnÞ ¼ 0

 

i¼1

 

 

 

where r is the gradient operator and Tn ¼ ð1=2mÞr2.

 

Hence, the following matrix element becomes

 

 

 

1

fdijr2ci þ 2hxjjrxii þ hxjjr2xicig

 

 

hxjjTnciðnÞjxii ¼

 

ðB:10Þ

 

2m

and the non-adiabatic coupling matrix elements are defined as below,

 

 

tjið1Þ ¼ hxjjrxii

tjið2Þ ¼ hxjjr2xii

ðB:11Þ

For example, in the case of the x component of the nuclear coordinates we have

1

xj

 

q

2

xj

 

q2

 

ðB:12Þ

txjið Þ ¼

qx xi

txjið Þ ¼

qx2 xi

 

 

 

 

 

 

 

 

 

 

 

 

 

Therefore, Eq. (B.10) in terms

of this notation becomes

 

 

hxjjTnciðnÞjxii ¼

 

1

fdjir2ci þ 2tjið1Þ rci þ tjið2Þcig

ðB:13Þ

 

2m

It is important to note that the non-adiabatic coupling terms have a direct effect on the momentum of the nuclei, which is the reason it is called a dynamic coupling. By substituting Eq. (B.13) in Eq. (B.9), we get

 

1

 

1

 

X

 

 

 

N

 

2m

r2cj þ ðujðnÞ EÞcjðnÞ

2m

ð2tjið1Þ rci þ tjið2ÞciÞ ¼ 0 ðB:14Þ

 

 

 

 

 

i¼1

This is the electronic adiabatic Schro¨dinger equation and in the case of a single coordinate x Eq. (B.14) takes the following form:

 

1 d2

 

X

 

 

d

 

 

 

1 N

2txjið1Þ

 

¼ 0

 

 

 

 

 

cj þ ðujðnÞ EÞcjðnÞ

 

 

 

 

 

ci þ txjið2Þci

ðB:15Þ

2m

dx2

2m i

1

dx

 

 

 

 

 

¼

 

 

 

 

 

 

 

When the non-adiabatic coupling terms tð1Þ and tð2Þ are considered negligibly small and dropped from Eq. (B.15), we get the uncoupled approximate Schro¨dinger equation

 

1 d2

ðB:16Þ

2m dx2 cj þ ðujðnÞ EÞcjðnÞ ¼ 0

non-adiabatic effects in chemical reactions

189

or more general,

 

1

r2cjðnÞ þ ðujðnÞ EÞcjðnÞ ¼ 0

ðB:17Þ

 

 

2m

The approximation involved in Eq. (B.17) is known as the Born–Oppenheimer approximation and this equation is called the Born–Oppenheimer equation.

By assuming the Hilbert space of dimension N, one can easily establish the relation between coupling matrices sð1Þ and sð2Þ by considering the (ij)th matrix element of r tð1Þ,

rtðij1Þ ¼ rhxijrxji ¼ hrxijrxji þ hxijr2xji ¼ hrxijrxji þ tðij2Þ

We can resolve the unity operator in the following way:

XN

I ¼ jxkihxkj

k¼1

and obtain,

X

 

 

!jxji

hrxijrxji ¼ hrxijIjrxji ¼ hxij

kN

1 jxkihxk

X

¼

X

 

N

 

N

 

¼ hrxijxkihxkjrxji ¼ hxkjrxiihxkjrxji

k¼1

 

k¼1

 

XN

¼ tðki1Þtðkj1Þ ¼ ðtð1ÞÞ2ij

k¼1

Hence, the elements of tð1Þ and tð2Þ are related as below

tðij2Þ ¼ ðtð1ÞÞ2ij þ rtðij1Þ

and finally in matrix notation

sð2Þ ¼ ðtð1ÞÞ2 þ rtð1Þ

ðB:18Þ

Incorporating relation (B.18) in Eq. (B.14), we can write in matrix form,

 

1

r2c þ

u

1

tð1Þ2 E

c

1

 

ð2tð1Þ r þ rtð1ÞÞc ¼ 0

ðB:19Þ

 

 

 

 

2m

2m

2m

190

satrajit adhikari and gert due billing

 

which can be expressed in compact form as

 

 

1

ðr þ tÞ2c þ ðu EÞc ¼ 0

ðB:20Þ

 

 

 

 

2m

So far, we have treated the case n n0, which was termed the adiabatic representation. We will now consider the diabatic case where n is still a variable

but

n0 is constant as defined in Eq. (B.3). By multiplying Eq.

(B.7) by

hxjðe j n0Þj and integrating over the electronic coordinates, we get

 

 

1

r2 E cjðnÞ þ i

N

 

 

1 hxjðe j n0ÞjH^eðe j nÞjxiðe j n0ÞiciðnÞ¼0

ðB:21Þ

 

 

 

2m

¼

 

 

 

X

 

We can rewrite the electronic Hamiltonian in the following form:

 

 

 

 

Heðe j nÞ ¼ Te þ Vðe j nÞ

 

 

 

 

Heðe j n0Þ ¼ Te þ Vðe j n0Þ

ðB:22Þ

 

 

 

Heðe j nÞ ¼ Heðe j n0Þ þ fVðe j nÞ Vðe j n0Þg

 

and by using Eq. (B.22), we can calculate the following matrix element:

hxjðe j n0ÞjHeðe j nÞjxiðe j n0Þi ¼ ujðn0

Þdji þ nijðn j n0Þ

ðB:23Þ

 

 

~

 

where

 

 

 

njiðn j n0Þ ¼ hxjðe j n0ÞjVðe j nÞ Vðe j n0Þjxiðe j n0Þi

 

~

 

 

ðB:24Þ

~

Þ þ ujðn0Þdji

 

nijðn j n0Þ ¼ nijðn j n0

 

 

By substituting the expression for the matrix elements in Eq. (B.21), we get the final form of the Schro¨dinger equation within the diabatic representation

 

 

X

 

 

 

1

N

 

 

 

2m

r2 E cjðnÞ þ i

1

njiðn j n0ÞciðnÞ ¼ 0

ðB:25Þ

 

 

¼

 

 

 

where the coupling terms among the states are due to potential coupling. By substituting the following transformation

c ¼ A ðB:26Þ

into the adiabatic Schro¨dinger equation (B.20), we obtain the following expression,

 

1

 

fAr2 þ 2ðrA þ tAÞ r þ fðt þ rÞ ðrA þ tAÞg g þ ðu EÞA

 

2m

 

¼

0

ðB:27Þ

non-adiabatic effects in chemical reactions

191

If the transformation matrix A is chosen in such a way that rA þ tA ¼ 0,

Eq. (B.27) can be rearranged to the following form:

 

 

1

ðB:28Þ

2m r2 þ ðA 1uA EÞ ¼ 0

which is basically in ‘‘diabatic’’ representation and identical looking with Eq. (B.25) and A is the adiabatic–diabatic transformation matrix.

APPENDIX C: FORMULATION OF THE VECTOR POTENTIAL

The vector potential is derived in hyperspherical coordinates following the procedure in [54], where the connections between Jacobi and the hyperspherical coordinates have been considered as below (see [67])

 

 

 

 

y

 

 

y

 

 

 

 

 

f

 

 

r

 

 

 

 

 

 

 

 

 

r

 

y

 

 

 

y

 

 

f

 

 

rx ¼ p2 cos 2 þ sin 2

cos 2

 

rz ¼ 0

cos

 

 

 

 

 

sin

 

 

 

 

 

 

ry ¼ p2

2

sin

2

2

 

 

 

 

r

cos

y

þ sin

y

 

f

 

ðC:1Þ

r

y

y

 

f

 

Rx ¼ p2

2

2 sin

2

 

 

 

 

Rz ¼ 0

cos

 

 

 

 

 

cos

 

 

 

 

 

Ry ¼ p2

2

sin

2

2

 

 

The interatomic distances of the triangle ABC formed due to any A þ BC type reactive system are as follows:

 

RAB2

¼

r2

ð1

þ sin y cos fÞ

 

 

d12

 

2

 

 

RBC2

¼

r2

ð1

þ sin y cos ðf x2ÞÞ

ðC:2Þ

 

d22

 

2

 

RCA2

¼

r2

ð1

þ sin y cos ðf þ x3ÞÞ

 

 

d32

 

2

 

and these interatomic distances can also be expressed in terms

of Jacobi