You are on page 1of 11

JOURNAL OF CHEMICAL PHYSICS

VOLUME 111, NUMBER 9

1 SEPTEMBER 1999

The semirigid vibrating rotor target model for quantum polyatomic reaction dynamics
J. Z. H. Zhang
Department of Chemistry, New York University, New York, New York 10003

Received 27 May 1999; accepted 10 June 1999 In this paper, we present detailed quantum treatment of the semirigid vibrating rotor target SVRT model for reaction dynamics involving polyatomic molecules. In the SVRT model, the reacting target molecule is treated as a semirigid vibrating rotor which can be considered as a three-dimensional generalization of the diatomic molecule. This model provides a realistic framework to treat reaction dynamics of polyatomic systems. Using the SVRT model, it becomes computationally practical to carry out quantitatively accurate quantum dynamics calculation for a variety of dynamics problems in which the reacting molecule is a polyatomic or complex molecule. In this work, specic theoretical treatment and mathematical formulation of the SVRT model are presented for three general classes of reaction systems: 1 reaction of an atom with a polyatomic molecule atompolyatom reaction , 2 reaction between two polyatomic molecules polyatom polyatom reaction , and 3 polyatomic reaction with a rigid surface polyatomsurface reaction . Since the number of dynamical degrees of freedom in the SVRT model for the above three classes of dynamical problems is limited, accurate quantum both ab initio and dynamical calculations are possible for many reactions of practical chemical interest. In this paper, a time-dependent wave packet approach is employed to implement the SVRT model for dynamics calculation of polyatomic reactions. 1999 American Institute of Physics. S0021-9606 99 03133-5

I. INTRODUCTION

Currently, a major challenge in the eld of quantum reaction dynamics is to develop quantitatively accurate methods for practical computational study of chemical reactions involving polyatomic molecules. At present, rigorous quantum dynamics calculations are limited only to systems involving no more than four atoms.1 In order to carry out quantitatively accurate quantum dynamics study for vast majority of chemical reactions that are of chemical or biological interest, it is necessary to develop practical computational methods to treat reaction dynamics of polyatomic molecules. To this end, some reduced dimensionality methods have been proposed to treat polyatomic systems tetraatomic systems in particular by reducing the dynamical degrees-offreedom DOF from six to three.2,3 By proper counting, some useful results of these reduced dimensionality calculations for the benchmark H2 OH system have been reported.47 However, it is desirable to develop a more systematic model to treat reaction dynamics for more general polyatomic molecules. In this paper we propose a semirigid vibrating rotor target SVRT model to treat reactive scattering involving polyatomic molecules. Additionally, we provide specic theoretical formulations for atompolyatom, polyatompolyatom, and polyatomsurface reactions using a time-dependent wave packet approach. Since the description of the reacting polyatomic molecule by the SVRT model is quite general and applies to essentially any polyatomic or complex molecule that reacts during collision, it becomes possible to employ this model to treat other similar types of reaction processes involving polyatomic molecules. For example, the
0021-9606/99/111(9)/3929/11/$15.00 3929

SVRT model for the atompolyatom reaction is a natural generalization of the exact atomdiatom reaction and it permits realistic quantum dynamics study for a general atom polyatom reaction with just four mathematical dimensions 4D . Similarly, the SVRT model for a polyatompolyatom reaction is a generalization of the exact diatomdiatom reaction and is described by seven mathematical degrees-offreedom 7D . In particular, the SVRT model can be applied to treat the dynamical process of dissociative adsorption of polyatomic molecules on a solid surface polyatomsurface reaction which can also be considered as a generalization of the exact treatment for the diatomsurface reaction.810 As a result, the dynamical SVRT model for the reaction of a polyatomic molecule with a rigid surface is completely described in 7D. However, it simplies to 5D if the xed site approximation1113 is employed, and to 4D if the surface is further assumed to be at.1421 In this paper, we present a detailed theoretical development and the time-dependent wave packet implementation of the SVRT model to treat reaction dynamics of three classes of polyatomic reactions; atompolyatom, polyatom polyatom, and polyatomsurface reactions. Although the basic SVRT model for the reacting polyatomic molecule is the same in all three classes of reactions, detailed mathematical formulation as well as the computational aspect is quite different in each case. For example, the SVRT model for an atompolyatom reaction involves only four DOF which can be accurately solved numerically for many practical reaction systems based on current computer powers. On the other hand, the exact numerical computation of the 7D SVRT model for the polyatompolyatom reaction is still quite chal 1999 American Institute of Physics

3930

J. Chem. Phys., Vol. 111, No. 9, 1 September 1999

J. Z. H. Zhang

lenging at present. However, such 7D numerical calculation is expected to be common in the near future as the computer power increases, as evidenced by recent exact 6D quantum dynamics calculations for tetraatomic reactions.2230 Similarly, with the recent report of 6D quantum dynamics calculations for the diatomsurface reaction,810 we expect the 7D SVRT model calculation for the polyatomsurface reaction to be commonplace in the near future. Thus, the computational prospect of the SVRT model for studying polyatomic reaction dynamics is very promising. In the basic SVRT model, the internal structure of the molecule is xed which may not be reasonable for a molecule whose structure changes from that in the reactant conguration to that in the product conguration during reactive collision. In order to account for this type of internal adjustment of the molecule, we need to relax the rigidity condition assumed in the basic SVRT model. A simple but often good approach is to apply adiabatic correction for the neglected internal DOF. Since the adiabatic approach accounts for the zero point energy quite accurately, it often gives a good reaction threshold energy without adding extra dynamical degrees-of-freedom in the coupled dynamics calculation.2 The implementation of this adiabatic correction to the basic SVRT model for all three types of polyatomic reactions are provided in the paper. The paper is organized as follows: Sec. II presents the basic SVRT model to describe a reacting polyatomic molecule that breaks into two fragments in reactive collision. Section III presents the time-dependent quantum dynamical formulation to treat the atompolyatom reaction for the basic SVRT model in four DOF. In addition, the adiabatic version of the SVRT model ASVRT model is also discussed and the mathematical formulation is given to include an adiabatic correction due to neglected internal degrees of freedom in the SVRT model. Section IV gives the mathematical formulation of the SVRT model for the polyatompolyatom reaction in seven degrees-of-freedom. Similarly, the timedependent formulation of the SVRT model for the polyatomsurface reaction is given in detail in Sec. V. Finally, discussion of the SVRT model and the future prospects of its application are given in Sec. VI.
II. THE SVRT MODEL FOR THE REACTING MOLECULE: Hamiltonian of the SVRT molecule

FIG. 1. The SVRT molecule T is composed of two rigid parts B and C. The centers-of-mass of B and C are separate by the reactive coordinate r which describes the vibrational motion responsible for breaking the BC bond. The angle is the rotational angle of the molecule about its molecular z-axis.

The exact classical Hamiltonian for a vibrating rotor target molecule can be written31 H ex 1 2
i i

ij

Gij

P2 VT , k

where i and i are, respectively, the angular momentum and vibrational angular momenta, P k is the momentum conjugate to the normal coordinate q k (k 1,..,3N 6) nonlinear molecule , and V T is the molecular potential. The G i j is the inverse of the shifted moment of inertia tensor I i j . 31 The corresponding quantum Hamiltonian to Eq. 3 for a nonlinear molecule was given by Wilson and Howard,32 Darling and Dennison,33 and later simplied by Watson34 to have the form H ex 1 2
ij i

i Gij
2 2

j 2

j G ii V T , 4

q2 k

To avoid confusion, we use the letter T to denote the target reacting molecule and its associated quantities, and letter R to denote the reactant molecule for the gas-phase reaction. In the case of the molecular reaction with a solid surface, R denotes the surface. In the SVRT model, the target molecule T is composed of two rigid fragments B and C as depicted in Fig. 1. A general reactive collision between molecules T and R can be expressed as R TR B C R C B, 1

with T B C. In the case of dissociative adsorption with R being a rigid surface, the reaction formula becomes R TR B R C. 2

where i and i are, respectively, the angular momentum and vibrational angular momentum operators. Since the solution of the exact Hamiltonian of Eq. 4 to include all vibrational degrees of freedom is impractical for polyatomic molecules, it is necessary to simplify the model in order to be computationally tractable for polyatomic molecules. In the SVRT model, we treat the target molecule T as a semirigid rotor composed of two rigid parts B and C as shown in Fig. 1. The two rigid parts B and C can exercise only one dimensional relative motion along the coordinate r connecting the center-of-mass CMS of B and C. Except for this relative motion between B and C, the molecule T behaves like a rigid rotor in three-dimensional space for any given value of r. Treating the effect of other neglected internal motions of T will be discussed later in the paper by using an adiabatic correction. Using this simplied SVRT model for molecule T, it is easy to show that the exact classical kinetic energy in Eq. 3 simplies to HT 1 2
iG i j 2 pr j

ij

VT r ,
r

J. Chem. Phys., Vol. 111, No. 9, 1 September 1999

SVRT model for reaction dynamics

3931

where the rst term is the rotational energy of molecule T, the second term is the relative translational energy between fragments B and C, and V T (r) is a one-dimensional interaction potential between fragments B and C. In Eq. 5 , T is the reduced mass dened by
T

M BM C MB MC

and G i j is the inverse of I, Gij I


1 i j,

7
FIG. 2. The SVRT model for the atompolyatom reaction. The angle is the rotational angle of the target molecule T about its molecular z-axis.

where the moment of inertia tensor I is dened as Iij m r2


ij

x ix

8
III. SVRT MODEL FOR THE ATOMPOLYATOM REACTION A. The basic SVRT model

for (i, j 1,2,3) corresponding to (x,y,z) axes in the molecular frame and the summation of in Eq. 8 is over all the atoms in molecule T. Equation 5 can be derived as follows. The classical kinetic energy of a vibrating rotor in the CMS frame can be evaluated31 K 1 m 2 1 m 2 r
2

In this section we treat the case when the reactant molecule R is a single atom or point mass A which represents a general case for the atompolyatom reaction. In the basic SVRT model, the full interaction Hamiltonian for the reactive collision between the atomic projectile A and the SVRT molecule T takes the form
2 2

r v 9

H ap

L2 2 R2 L2 2 R
2 2

2
2

R2
2

HT V 1 2 V, G i ij

1 m v2 , 2

2
2

ij

where is the angular velocity of the rotating frame, v is the velocity of atom relative to the rotating frame, and the summation of is over all the atoms in molecule T. Since in the SVRT model we have the relation v vB for atoms in fragment B and v vC for atoms in fragment C, Eq. 9 becomes K
ij
1 2

r2

VT r

13

where R is the relative radial distance between the CMS of A and T, L is the orbital angular momentum operator, V is the interaction potential, and is the reduced translational mass M AM T MA MT 14

iI i j 2 M B vB

M B rB vB M C rC vC
2 M C vC

10 with M T M B M C . In Eq. 13 , the CMS motion of the collision system has been set to zero and the standard substitution of the wave function by 1/R is assumed. It is important to note that the interaction potential V depends on four internal coordinates only as shown in Fig. 2. A natural choice of these four coordinates are (R,r, , ), where the polar angle species the angle between the vecspecies the rotation of the tors R and r and the angle molecule T along its own molecular z axis coordinate r cf. Fig. 2 . Thus, the SVRT model for the atompolyatom reaction is of four mathematical dimensions which is well within the current computational reach. The numerical solution for the atompolyatom reactive collision with the above described SVRT model can be implemented using the time-dependent TD wave packet approach for the reaction.46 In the TD approach, one solves the TD Schrodinger equation,

Since the second term vanishes and the CMS kinetic energy is zero, the above equation simplies to K
1 2

iG i j

1 2

2 T vT ,

11

which gives Eq. 5 after the addition of the interaction potential V T (r). Finally, the quantum Hamiltonian corresponding to the classical SVRT Hamiltonian of Eq. 5 can be written straightforwardly HT 1 2
2 ij 2

G i ij

r2

VT r .

12

It is important to note that the matrix G i j is a function of the coordinate r and commutes with the angular momentum operator i .

3932

J. Chem. Phys., Vol. 111, No. 9, 1 September 1999

J. Z. H. Zhang

The generalized potential operator U is dened as i t t H t , 15 U H rot V , where H rot is the rotation Hamiltonian H rot L2 2 R
2

24

with the Hamiltonian dened in Eq. 13 . With an appropriate choice of basis sets, the expansion of (t) can take the form t
p v jKn

uv p

Z JM jK(n)

1 2

ij

G . i ij j

25

, ,

r C p v jKn t , 16

where is the translational basis function whose denition is given in Ref. 23, v (r) is the vibrational basis function given by the solution of the one-dimensional eigenequation,
2 2

u v (R) p

In carrying out the numerical propagation in Eq. 22 , it is convenient to employ the split operator algorithm again to express the exponential operator of the general potential U by either formula e or e
iU iU

iV /2

iH rot

iV /2

26

r2

VT r

r ,

17

, , ) is the body-xed BF total angular moand mentum eigenfunction to be dened below. As discussed in the Appendix of Ref. 46, the eigenfunction of the total angular momentum coupled from two individual angular momenta in the BF representation is given by Z JM jK(n) , , MK D J* Q jKn , . 18

Z JM ( jK(n)

iH rot /2

iV

iH rot /2

27

Since the orbital angular momentum operator L commutes with , the exponential rotation operator can be written as a simple product of two exponential operators, exp iH rot exp i L2 2 R2 i 1 2 G i ij
j

M Here D J K ( ) is the normalized rotation matrix dened in Eq. B2 in the Appendix. The three Euler angles specify the spatial orientation of the whole collision system. The Q jKn ( , ) is the rotational basis function for the angular momentum of the target molecule with its own SF Z axis pointing along the R axis as shown in Fig. 1. To simplify the mathematical treatment, we choose Q jKn ( , ) to be the eigenfunction of the symmetric top, i.e., Q jKn , j* 2 D Kn 0, , Kn d j* 1 2 e in , 19

exp

28

ij

where Kn ( ) is the normalized reduced rotation matrix ded j* ned as Kn d j* 2J 1 j d Kn 2 . 20

The operation of the exponential operator 2 /2 R 2 ) ) on the wave function requires some exexp( i(L planation. The most important step is to utilize the result of the operation of the orbital angular momentum operator L2 on the BF total angular momentum eigenfunction. The result of such an operation is well known for the atomdiatom and diatomdiatom systems as is outlined in the Appendix of Ref. 46. Here, however, we give a brief derivation of the result for the current SVRT model. First we utilize the formal operator relation to express the orbital angular momentum operator as L2 J
2

J2

2 2J z

, 29

Thus, the normalized total angular momentum eigenfunction in the BF frame can be expressed as Z JM jK(n) , , MK D J* Kn d j* 1 2 e in . 21

Here the quantum number n carries the meaning of angular momentum quantum number of the projection of along the z-axis of molecular frame dened in Fig. 1 for a symmetric top. The TD wave function can be propagated by employing the split-operator method47 t e
iH 0 /2

iU

iH 0 /2

t ,

22

where J denotes the raising lowering operator of the total angular momentum in the BF frame. Secondly, we note that the total angular momentum operator J in Eq. 29 is dened in the BF frame of the collision system with the BF Z axis pointing along the R vector cf. Fig. 2 . Therefore J in Eq. 29 satises the anomalous commutation relation given in Eq. B9 . However, the BF frame of the collision system serves as the space-xed SF frame for the component angular momentum in Eq. 29 . Thus in Eq. 29 satises the normal commutation relation of angular momentum operator.43 As a result, the operation of L2 on the BF total angular momentum eigenfunction of Eq. 18 gives the following result:46 jK(n) L2 Z JM J J 1
JK

where the operator H 0 can be dened as


2 2 2 2

j j

2K 2 Z JM jK(n)
JK JM jK Z jK 1(n) ,

H0

JM jK Z jK 1(n)

30

R2

r2

VT r .

23

where the coefcient is dened as

J. Chem. Phys., Vol. 111, No. 9, 1 September 1999

SVRT model for reaction dynamics

3933

JK

J J 1

K K 1

31

and the same denition for jK . Thus the K-states of the BF total angular momentum eigenfunction are coupled through the centrifugal potential or orbital angular momentum for the total angular momentum quantum number J 0. We notice, however, that the quantum number n in Eq. 30 is unchanged. Consequently, the rst exponential operator on the right-hand side RHS of Eq. 28 couples the K states only and the coupling matrix can be diagonalized to yield the result of the exponential operation. If the centrifugal sudden CS approximation is employed for total angular momentum J 0, the off diagonal terms in Eq. 30 are simply neglected to yield the CS result,
CS

where the summation of k is over the number of internal vibrational normal mode coordinates q k of the target molecule that are strongly coupled to the reactive coordinate r. In this GSVRT generalized SVRT model, the full atom polyatom collision Hamiltonian takes the form
2 g H ap 2

L2 2 R2 L2 2 R
2

2
2

R2
2

HT V 1 2
2

2
2

ij 2

G i ij

r2

q2 k

V int V.

35

jK(n) L2 Z JM

J J 1

j j

2K 2 Z JM jK(n)

32

which signicantly simplies the numerical computation by treating K as a conserved quantum number. The second exponential operator on the RHS of Eq. 30 couples the n quantum number only as shown in the Appendix for the rigid rotor eigenfunction. The rest of the procedure for the TD wave packet calculation such as the evaluation of the reactive ux, etc., is essentially identical to those given before23,46 and is not repeated here.

B. The adiabatic SVRT model

The GSVRT model in Eq. 35 allows energy exchange between the internal vibrational coordinates q k and the external coordinates (R,r, , ) and is therefore more realistic. However, it is important to keep in mind that any addition of a vibrational coordinate will signicantly increase the level of computational difculty. Therefore the inclusion of additional vibrational coordinates should be kept at a minimum. In many practical situations, the internal structure of fragments B and C undergo approximately adiabatic change before, during, and after the reaction. To a good approximation, the change of these structure can be treated as a function of the reactive coordinate r only. It is then possible to make adiabatic approximation to treat the internal vibrations of the molecule. In this adiabatic SVRT ASVRT model, the potential V int is approximated as the sum of individual vibrational potential V int q 1 , . . . ,q N ,r
vk qk r

In many practical situations, the internal structure of molecule T or R changes during the reaction process. For example, a particular bond angle may change from an initial value in the parent molecule T to a nal value in fragment B or C. In such cases, it may be necessary to accommodate the internal structural change in the dynamical model in order to describe the energetics of the reaction correctly. It is possible to systematically improve the basic SVRT model by including more internal coordinates of the target molecule that are strongly coupled to the reaction coordinate. By including more internal degrees of freedom, the kinetic energy operator of the target molecule becomes more complicated due to Coriolis or vibrational angular momentum coupling as evidenced in Eq. 3 . However, if we neglect all the vibrational angular momenta in Eq. 3 , we obtain the simplied classical Hamiltonian for the target molecule which is the sum of kinetic energy and vibrational energy,
g HT

VT r ,

36

where v k (q k r) is the vibrational potential for the kth vibrational mode. The internal vibrational wave function is described by the adiabatic wave function k (q k r) which is given by the solution of the adiabatic vibrational eigenequation,
2

d2

2 dq 2 k

vk qk r

qk r

a k

qk r ,

37

where the adiabatic dependence of the internal vibration wave function k on the reactive coordinate r is explicitly shown. By employing adiabatic approximation, the ASVRT Hamiltonian is given by
2 2 a H ap 1 N

1 2

ij

iG i j

2 pr j

p2 k 33

L2 2 R2

N 2 2

2 r2

R2 VT r

V int r,q 1 , . . . ,q N . The corresponding quantum version is g HT 1 2


2 ij 2 2 k 2

1 2

ij

G i ij ,

R,r, , V where 34 Gij


1 N

38

G i ij

r2

q2 k

V int r,q 1 , . . . ,q N .

Gij

39

3934

J. Chem. Phys., Vol. 111, No. 9, 1 September 1999

J. Z. H. Zhang

and V
1 N

a k

r .

40

Equations 39 and 40 may be calculated approximately by Gij r G i j r,q k r , V R,r, , ,q k r , 41 42

R,r, , V

where k (r) is the r-dependent average value of the vibraq tional coordinate q k dened as k r q
k

qk

dq k q k

qk r

43

The dynamical calculation for the ASVRT model of Eq. 38 is essentially identical to that for the basic SVRT model of Eq. 13 . Thus the TD wave packet treatment described in the previous section for the basic SVRT model can be directly applied to the ASVRT model with essentially no modication. The ASVRT reaction model allows the internal vibrational motions of the fragments to adjust to the change of the reactive coordinate r and thus can describe the change of internal structure of the fragments B and C during reactive collision.
IV. THE SVRT MODEL FOR THE POLYATOMPOLYATOM REACTION A. Hamiltonian for the reactant molecule R

FIG. 3. The SVRT model for the polyatompolyatom reactive collision between the target reacting molecule T and the reactant molecule R. The reactant molecule R is treated as a rigid asymmetric rotor.

The corresponding rotational eigenfunction is given by solving the rigid rotor equation
2 x

2 y

2 z

2I xx

2I y y

2I zz

QR jkn

R R jn Q jkn

49

where k and n are, respectively, the angular momentum quantum number along the SF and molecular xed Z-axis. The method of solving Eq. 49 is given in the Appendix.
B. The basic SVRT model

In this section we treat the case in which the reactant molecule R is a complex or polyatomic molecule with mass M R . The simplest approach is to treat the reactant as a rigid asymmetric rotor whose quantum Hamiltonian with zero kinetic energy for the center-of-mass is given by HR 1 2 RG R R , i ij j 44

Now we are in a position to construct the full interaction Hamiltonian for reactive collision between two molecules T and R,
2 2

H pp

L2 2 R2

H int V,

50

ij

where R is the projection of the angular momentum operai tor of the reactant molecule R along the molecular axis i and G Rj is given by the inverse matrix element i G Rj i IR
1 ij

where the variable R is the relative radial distance between the CMS of molecules R and T, L is the orbital angular momentum operator, is the reduced translational mass of the collision system dened by M RM T , MR MT and the internal Hamiltonian H int is given by H int 1 2 RG R i ij
R j 2 ij 2

51

45

where the moment of inertia IR is dened as I Rj i m r2


ij

x ix

46

ij

1 2

TG T i ij

T j

r2 52

VT r . for which the summation of is over all atoms in the reactant molecule R. For the reactant molecule, it is convenient to choose the principal axis of rotation such that the rotational Hamiltonian takes the diagonal form HR with G Rj i 1 2I i j
ij 2 x

2 y

2 z

2I xx

2I y y

2I zz

47

48

It is important to note that the interaction potential V in Eq. 50 depends on seven internal coordinates. For convenience, we choose these coordinates to be (R,r, T , T , R , R , ). Here the angle is dened as R T , where R and are, respectively, the azimuthal angles dened in the EuT ler angles R ( R , R , R ) for molecule R and T ( T , T , T ) for molecule T as illustrated in Fig. 3. Thus the SVRT dynamical model for the polyatompolyatom reaction dened by the Hamiltonian of Eq. 50 involves seven degrees-of-freedom 7D . It is useful to note that theoretical treatment of molecular system consisting of two rigid rotors in six DOF has been

J. Chem. Phys., Vol. 111, No. 9, 1 September 1999

SVRT model for reaction dynamics

3935

presented before for bound state problem.35,36 In particular, benchmark six dimensional bound state calculations for water dimer treated as two rigid rotors have recently been reported by Leforestier et al.37 and Chen et al.38 These stateof-the-art quantum dynamics calculations for water dimer are very encouraging because they demonstrated the practicality of carrying out rigorous quantum calculations for systems consisting of two polyatomic molecules. Many of the mathematical equations used in these bound state calculations are quite similar to what will be employed to treat reactive scattering problems discussed below. However, due to the inclusion of the extra bond coordinate r in the SVRT Hamiltonian and the nature of the reactive scattering problem, the mathematical and computational aspect of the current dynamical model for the polyatompolyatom reaction is not a trivial extension of that for the bound state problem. In the following, we present the general TD mathematical treatment of the SVRT model for the polyatompolyatom reaction in seven degrees of freedom. In time-dependent implementation of the SVRT model, we expand the TD wave function (t) in basis set t
p v j 12K j 1 k 1 j 2 k 2

where the operator H 0 can be dened as


2 2 2 2

H0

R2

r2

VT r .

57

The generalized potential operator U is dened as U H rot V , where H rot is the rotation Hamiltonian H rot L2 2 R
2

58

1 2

ij

TG T r i ij

T j

1 2

ij

RG R T . i ij j

59

Since the orbital angular momentum operator L commutes T or R , the exponential rotation operator can be with written as simple product of three exponential operators exp iH rot exp i L2 2 R2 i 1 2 1 2 TG T i ij RG R i ij
T j

exp

ij

uv R p exp T , R ,R
v

Z JMK( j k j k ) j 12 1 1 2 2

r C p v j 12K j 1 k 1 j 2 k 2 , 53
C. The adiabatic SVRT model

ij

R j

60

where u v (R) is the translational basis function whose dep nition is given in Ref. 23, v (r) is the vibrational basis function dened in Eq. 17 and Z JMK( j k j k ) ( T , R ,R) is j 12 1 1 2 2 the BF body-xed total angular momentum eigenfunction to be dened below.
1. Total angular momentum eigenfunction

As discussed in Sec. III A, the total angular momentum eigenfunction in the BF representation can be written as Z JMK( j j
12 1k1 j 2k2)

MK D J*

Y j 12 j

j K , 1 2 (k 1 k 2 )

54

where are three Euler angles specifying the overall spatial j K orientation of the collision complex. The function Y j 12 (k k ) j
1 2 1 2

is the coupled angular momentum eigenfunction of 12 T R dened as Y j 12 j


j K 1 2 (k 1 k 2 )

j 1 m j 2 K m j 12K
1 j* D mk 0, 1

2 ,
R,

Since the basic SVRT model for the polyatompolyatom reaction already involves the coupled equation with seven degrees of freedom, it is not computationally prudent to introduce more coordinates into the dynamics model explicitly. To a good approximation, we may treat the change of the internal structure as a function of all seven coordinates adiabatically. By employing the adiabatic approximation, we avoid the explicit coupling of internal vibrational motions to that of the scattering coordinates. For computational reasons, however, we wish to minimize the number of coordinates that the internal structure depend on. The simplest approach is to make the internal structure depend adiabatically on only one coordinate which is characteristic of the evolution of the reaction process such as the reaction coordinate s. In the following we use s r to simplify the treatment. Following the discussions on adiabatic correction in Sec. III B for the SVRT model for the atompolyatom reaction, the generalized Hamiltonian for the polyatompolyatom reaction takes the form
2 2

T,

j* D K2

mk 2

55

pp Hg

L2 2 R
2

R2

g g HT HR V,

61

where j 1 m j 2 K m j 12K are ClebschGordon CG coef j* cients, D mk are the normalized rotation matrix dened in Eq. B2 , and the angle is dened by R T.
2. Propagation of the wave function

g where H T of the reacting molecule T is dened in Eq. 34 g and H R of the reactant molecule R is similarly given by g HR 1 2
2 ij 2

RG R i ij

R j

q2 k 62

The TD wave function can be propagated by employing the split-operator method47 t e


iH 0 /2

V int q 1 , . . . ,q N .

iU

iH 0 /2

t ,

56

The ASVRT Hamiltonian for the polyatompolyatom collision system is therefore obtained from

3936

J. Chem. Phys., Vol. 111, No. 9, 1 September 1999

J. Z. H. Zhang

pp Ha

T k 2

R k

pp H g
2

T k

R k R j

L2 2 R
T j

2 1 2 with i G Rj s and i G Tj V
k

1 2
2

ij 2

RG R s ij i VT r

ij

TG T ij i

r2

63

R k

G Rj i

R k

64

T k

G Tj i
R k

T k

65

The potential in Eq. 63 is given by the averaging


T k T k

V
R k

T k

R k

FIG. 4. The SVRT model for the polyatomsurface reaction. The three Euler angles are ( , , ), where is the azimuthal angle, the polar angle and the rotational angle of the target molecule T about the molecular z-axis. The (X,Y ,Z) are three Cartesian coordinates of the center-of-mass of molecule T.

r
k

r .

66

Again the TD wave function is propagated by the splitoperator method t e


iH 0 /2

V. SVRT MODEL FOR THE POLYATOMSURFACE REACTION A. The basic SVRT model

iU

iH 0 /2

70

where the operator H 0 is dened as


2 2 2 R 2

H0

1. Fully corrugated surface

2M

r2

VT r ,

71

The Hamiltonian for reactive collision between the SVRT molecule T and a rigid surface takes the form
2 2 2 R 2 2 2 2

and the generalized potential U is U 1 2 G i ij


j

V.

72

ij

H ps

2M
2

HT V 1 2

2M X 2
2

2M Y 2
2

2M Z 2

ij

G i ij

r2

VT r

V, 67

2. Fixed site approximation In the xed site model for the surface reaction,11 the lateral coordinates (X,Y ) of the CMS of molecule T is xed and the corresponding Hamiltonian is obtained by simply dropping the relevant kinetic energy operators in Eq. 67 ,
2 2

where M M B M C is the mass of molecule T, (X,Y ,Z) are the Cartesian coordinates of the CMS of molecule T, and V is the interaction potential. It should be noted here that the interaction potential V depends on three Cartesian coordinates (X,Y ,Z) and four molecular coordinates. The latter can be chosen to be (r, ), where ( , , ) are three Euler angles that specify the spatial orientation of molecule T as shown in Fig. 4. In the time-dependent implementation of the SVRT model for polyatomsurface reaction, one can expand the TD wave function as t
pv jmk
v j* u p R D mk v

fs H ps

2M Z 2

1 2

2 ij

G i ij

r2

VT r

V, 73

where the (X,Y ) coordinates of the interaction potential V are xed at the chosen site. Thus the Hamiltonian for the xed site SVRT model is of ve degrees-of-freedom. The expansion of the TD wave function takes the form t
p v jmk

j* u v Z D mk p

r C p v jmk t ,

74

r C pv jmk t ,

68

where only the Z-component kinetic energy operator of the CMS motion of molecule T is included.
3. Flat surface If we further assume that the dependence of the interaction potential V on the azimuthal angle can be neglected, we arrive at the so called at surface model in which the potential only depends on four coordinates, (Z,r, , ). The corresponding Hamiltonian appears exactly the same as given by Eq. 73 but the dependence of the interaction potential on is gone. As a result, the expansion of the TD wave function takes the form

v where u p(R) are three dimensional translational basis functions with three indices p (p x , p y , p z ) and can be taken as a direct product of the one-dimensional function v up R

u px X u py Y u v Z , p
z

69

where u p x (X) and u p y (Y ) are sine functions with the proper periodicity of the surface, and u v (Z) is dened in Ref. 23. pz The vibrational function v (r) is dened in Eq. 17 .

J. Chem. Phys., Vol. 111, No. 9, 1 September 1999

SVRT model for reaction dynamics

3937

t
p v jk

j* u v Z D mk p

r C p v jk t .

75

Here the quantum number m, which is the projection of the angular momentum of molecule T on the SF Z axis, becomes a conserved quantum number. As shown in early studies, the dissociation probability of hydrogen on copper shows a sensitive dependence on the quantum number m especially at low collision energies.11,15
B. The adiabatic SVRT model

In the same spirit as for gas-phase polyatomic reaction, we can employ the ASVRT model to include the adiabatic correction to the basic SVRT treatment for the polyatom surface reaction. By employing the adiabatic approximation, the polyatomsurface Hamiltonian is obtained from
2

ps Ha

H ps

1 2

N 2

2M VT r

2 R

1 2

ij

G i ij

r2

R,r, V

, 76

where Gij and V


1 N 1 N

Gij

77

a k

r ,

78

SVRT model, it becomes computationally realistic to study quantitatively chemical reaction dynamics involving polyatomic molecules using a limited number of degrees-offreedom. For the atompolyatom reaction, the SVRT model involves only four mathematical dimensions and is subject to rigorous numerical calculation at present. Both ab initio and dynamics calculations could be carried out at a high level of accuracy. The improved ASVRT model gives a more realistic description than the basic SVRT model by allowing the internal structure of the fragments to adjust to the change of the reactive coordinate. Thus the ASVRT model gives a more accurate representation of the transition state and the reaction barrier for the reaction. The dynamical calculation of the ASVRT model also involves four degrees of freedom and is subject to rigorous numerical computation as well. Computationally, there is still some room to include a number of dynamical degrees of freedom in the generalized SVRT treatment for atompolyatom reaction. The SVRT model for the polyatompolyatom reaction is a generalization of the exact diatomdiatom reaction model. Although it is currently still quite challenging to carry out rigorous numerical calculations in 7D, we expect such calculations to be much more accessible in the near future. Similarly, the SVRT model for the polyatomsurface reaction is a direct generalization of the diatomsurface reaction and also involves seven degrees-of-freedom for the fully corrugated surface. The xed-size approximation involves only ve mathematical dimensions and is subject to rigorous numerical computation at present.
ACKNOWLEDGMENTS

where the adiabatic function k is dened in Eq. 37 . Equations 77 and 78 may also be calculated approximately by Gij r and R,r, V V R,r, , , ,q k r 80 G i j r,q k r 79

Funding by the National Science Foundation through a Presidential Faculty Fellowship and the Petroleum Research Fund are acknowledged.
APPENDIX A: ROVIBRATIONAL EIGENFUNCTION OF THE SVRT HAMILTONIAN

where k (r) is calculated via Eq. 43 . q

VI. CONCLUSION

The proposed SVRT model for the polyatomic molecule is a natural generalization of the exact treatment for the diatomic molecule to three dimensions. As a result of this generalization, a wide range of reactive collision dynamics problems involving polyatomic molecules can now be treated quantitatively using only a limited number of degrees-offreedom as is described in detail in this paper for the atom polyatom, polyatompolyatom, and polyatomsurface reactions. Thus the SVRT model opens the door for carrying out quantitatively accurate quantum dynamics calculation for polyatomic reactions regardless of the number of atoms in the target molecule. It should be noted that since the SVRT model describes the rotational motion of the complex molecule correctly or realistically at the least , the steric dynamics or effect of the reaction can be correctly described which is of fundamental importance for studying reaction dynamics involving large complex molecules. Using the

The exact rovibrational eigenfunction of molecule T, dejm noted by v n ( T ,r) with three Euler angles T and the vibrational coordinate r, satises the Schrodinger equation 1 2
2 ij 2

G i ij

r2

VT r

jm vk

E v jk

jm vk

. A1

Here j and m are, respectively, quantum numbers of the angular momentum and its projection on the space-xed SF Z axis, k is an additional angular momentum label, and v is the label for the vibrational state. Equation A1 can be solved exactly by numerical methods to give the eigenvalues and eigenfunctions. However, in practice, one may wish to solve Eq. A1 by employing adiabatic approximation to separate rotation from vibration
jm vk T ,r

Q jmk

v jk

r ,
T

A2 r) is the solution of

where the rotational function Q jkm ( the rotational eigenequation,

3938

J. Chem. Phys., Vol. 111, No. 9, 1 September 1999

J. Z. H. Zhang

1 2

ij

G i ij

Q jmk

rot jk

r Q jmk

A3
k

Hk

Ck

rot j jk C k k ,

B7

and the vibrational function v jk (r) is given by the solution of the one-dimensional eigenequation,
2 2

where the rotation Hamiltonian matrix is dened as Hk


j k

r2

VT r

rot jk

* D mk

v jk

E v jk

v jk

r . A4

1 2

ij

G i ij

* D mk .

B8

In practical application, one may wish to approximate Q jkm ( T r) by Q jmk


T

Q jmk

, r

A5

In computing the rotation Hamiltonian matrix in Eq. B8 , we should note that i is the projection of the angular momentum operator along the molecular axis. Thus the angular momentum operator i satises the anomalous commutation relation4143 , i
j

where is some sort of vibration-averaged position. This is r equivalent to using an averaged momentum of inertia Gij r Gij r A6

i
k

i jk

, k

B9

in solving Eq. A3 . The detailed method of solving Eq. A3 will be given later.
APPENDIX B: ROTATIONAL EIGENFUNCTION

where i jk is the standard antisymmetric tensor. As a result the denition of raising and lowering operators are reversed,44 i.e.,
x x

i i

y y

B10

Since the method to solve rotational eigenfunction of either Eq. A3 for molecule T or Eq. 49 for molecule R is the same, we drop the specic reference to T or R in describing the method of solution in this section unless specied otherwise. First, we expand the angular momentum eigenfunction by Q jkm
k

which can be inverted to give


x

1 2 i 2

B11

j* D mk

Ck k ,

B1

j where D mk ( ) is the normalized rotation matrix dened as j D mk 2j 8 1


2 j D mk

B2

where denotes the raising lowering operator in the molecular frame. By utilizing Eq. B11 and the well known property of the raising and lowering operator of the angular momentum,45 the evaluation of the matrix element of Eq. B8 is straightforward. For example, we can obtain D j* x mk 1 2 1 2 j j j j
j * D mk

j where D mk is the Wigner rotation matrix dened as39,40 j D mk

, , e

jm e
im

i Z j j d mk

i Y j

i Z j

jk B3

ik

1 1

j * k k 1 D mk j * k k 1 D mk

with the reduced rotation matrix dened as


j d mk

jm e

i Y j

jk .

B4

B12

j The normalization condition for D mk is j Dm


k

and a similar result for the operation by y . Of course, the j * operation by z on D mk simply yields D j * kD j * . z mk mk B13

j D mk

2 0 0

d sin d d
0

*j Dm k

, ,

j D mk

, ,

k B5

j j k k m m,

where the normalized reduced rotation matrix is given by mk dj 2j 2 1


j d mk

B6

For the target molecule T, the molecular z axis is chosen to be along the r coordinate as shown in Fig. 1. With this choice of the molecular z axis, the Euler angles T that specify the spatial orientation of molecule T are independent of the coordinate r. The (x,y) axes, on the other hand, can be chosen such that the element of moment of inertia tensor I xy I yx vanishes.
APPENDIX C: SYMMETRIC TOP MOLECULE

Substituting the expansion of Eq. B1 in Eq. A3 or Eq. 49 and integrating out the Euler angles, we obtain the linear algebraic equation for the expansion coefcient

In case that the molecule is a symmetric top, the rotational Hamiltonian simplies to

J. Chem. Phys., Vol. 111, No. 9, 1 September 1999

SVRT model for reaction dynamics


13

3939

H A

C A

2 z

C1

with rotation constants given by A and C 1 . 2I zz C3 1 2I xx 1 2I y y C2

j In such an instance, the rotation eigenfunction Q jmk becomes j * identical to the rotation matrix D mk with eigenvalue rot jk

Aj j

C A k 2.

C4

APPENDIX D: LINEAR MOLECULE

In case that the symmetric top is also a linear molecule, the quantum number k must be zero and the rotation Hamiltonian in Eq. C1 simplies to H
2

2I

D1

where I is the moment of inertia of the linear molecule. The corresponding angular momentum eigenfunction for the linear rotor is Y jm which is the special case of the rotation j* matrix D mk with k 0. 40 If both T and R are linear molecules, the collision Hamiltonian of Eq. 50 becomes identical to the diatomdiatom Hamiltonian in Ref. 23 with the replacement of mr 2 by the moment of inertia I. Thus the SVRT model for polyatom polyatom reaction can be viewed as a generalization of the exact treatment for the diatomdiatom reaction.
J. Z. H. Zhang, J. Dai, and W. Zhu, J. Phys. Chem. A 101, 2746 1997 , and references therein. 2 Q. Sun and J. M. Bowman, J. Chem. Phys. 92, 5201 1990 . 3 A. N. Brook and D. C. Clary, J. Chem. Phys. 92, 4178 1990 . 4 a J. M. Bowman and D. Wang, J. Chem. Phys. 96, 7852 1992 ; b D. Wang and J. M. Bowman, ibid. 96, 8906 1992 . 5 a D. C. Clary, J. Chem. Phys. 95, 7298 1991 ; b 96, 3656 1992 . 6 H. Szichman, I. Last, A. Baram, and M. Baer, J. Phys. Chem. 97, 6436 1993 . 7 N. Balakrishnan and G. D. Billing, J. Chem. Phys. 101, 2785 1994 . 8 A. Gross, B. Hammer, M. Schefer, and W. Brenig, Phys. Rev. Lett. 73, 3121 1995 ; 75, 2718 1995 . 9 G. J. Kroes, E. J. Baerends, and R. C. Mowrey, J. Chem. Phys. 107, 3309 1997 ; Drew A. McCormack, G. J. Kroes, Roar A. Olsen, E. J. Baerends, and R. C. Mowrey, ibid. 110, 7008 1999 . 10 J. Dai and J. C. Zhang, J. Chem. Phys. 107, 1676 1997 ; 108, 7816 1998 . 11 J. Dai and J. Z. H. Zhang, Surf. Sci. 319, 193 1994 . 12 G. R. Darling and S. Holloway, J. Chem. Phys. 101, 3268 1994 .
1

R. C. Mowrey, G. J. Kroes, E. J. Baerends, and G. Wiesenekker, J. Chem. Phys. 106, 4248 1997 ; R. C. Mowrey, G. J. Kroes, and E. J. Baerends, ibid. 108, 6909 1998 . 14 J. Sheng and J. Z. H. Zhang, J. Chem. Phys. 97, 6784 1992 ; 99, 1373 1993 . 15 J. Dai, J. Sheng, and J. Z. H. Zhang, J. Chem. Phys. 101, 1555 1994 ; Surf. Sci. 319, 193 1994 . 16 J. Y. Ge and J. Z. H. Zhang, Chem. Phys. Lett. 292, 51 1998 . 17 U. Nielsen, D. Halstead, S. Holloway, and J. K. Norskov, J. Chem. Phys. 93, 2879 1990 . 18 S. Holloway and B. Jackson, Chem. Phys. Lett. 172, 40 1990 . 19 A. Cruz and B. Jackson, J. Chem. Phys. 84, 5715 1991 . 20 P. Saalfrank and W. H. Miller, J. Chem. Phys. 98, 9040 1993 . 21 R. C. Mowrey, J. Chem. Phys. 94, 7098 1991 ; 99, 7049 1993 . 22 D. H. Zhang and J. Z. H. Zhang, J. Chem. Phys. 99, 5615 1993 ; 100, 2697 1994 . 23 D. H. Zhang and J. Z. H. Zhang, J. Chem. Phys. 101, 1146 1994 . 24 a D. H. Zhang and J. Z. H. Zhang, Chem. Phys. Lett. 232, 370 1995 ; b D. H. Zhang, J. Z. H. Zhang, Y. C. Zhang, D. Wang, and Q. Zhang, J. Chem. Phys. 102, 7400 1995 ; c Y. Zhang, D. Zhang, W. Li, Q. Zhang, D. Wang, D. H. Zhang, and J. Z. H. Zhang, J. Phys. Chem. 99, 16824 1995 ; d D. H. Zhang and J. Z. H. Zhang, J. Chem. Phys. 103, 6512 1995 . 25 a W. Zhu, J. Dai, and J. Z. H. Zhang, J. Chem. Phys. 105, 4881 1996 ; b J. Dai, W. Zhu, and J. Z. H. Zhang, J. Phys. Chem. 100, 13901 1996 . 26 a W. Zhu, J. Z. H. Zhang, Y. C. Zhang, Y. B. Zhang, L. X. Zhan, and S. L. Zhang, J. Chem. Phys. 108, 3509 1998 ; b W. Zhu, J. Z. H. Zhang, and D. H. Zhang, Chem. Phys. Lett. 292, 46 1998 . 27 D. Neuhauser, J. Chem. Phys. 100, 9272 1994 . 28 U. Manthe, T. Seideman, and W. H. Miller, J. Chem. Phys. 99, 10078 1993 ; 101, 4759 1994 . 29 a F. Matzkies and U. Manthe, J. Chem. Phys. 108, 4828 1998 ; b U. Manthe and F. Matzkies, ibid. 282, 442 1998 . 30 D. H. Zhang and J. C. Light, J. Chem. Phys. 104, 4544 1996 ; 105, 1291 1996 ; 106, 551 1997 . 31 E. B. Wilson, Jr., J. C. Decius, and P. C. Cross, Molecular Vibrations Dover, New York, 1980 , p. 273. 32 E. B. Wilson, Jr. and J. B. Howard, J. Chem. Phys. 4, 260 1936 . 33 B. T. Darling and D. M. Dennison, Phys. Rev. 57, 128 1940 . 34 J. K. G. Watson, Mol. Phys. 15, 479 1968 . 35 G. Brooks, A. van der Avoid, B. T. Sutcliffe, and J. Tennyson, Mol. Phys. 50, 1025 1983 . 36 S. C. Althorpe and D. C. Clary, J. Chem. Phys. 101, 3603 1994 . 37 a C. Leforestier, L. B. Braly, K. Liu, M. J. Elrod, and R. J. Saykally, J. Chem. Phys. 106, 8527 1997 ; b R. S. Fellers, L. B. Braly, R. J. Saykally, and C. Leforestier, ibid. 110, 6306 1999 . 38 H. Chen, S. Liu, and J. C. Light, J. Chem. Phys. 110, 168 1999 . 39 E. P. Wigner, Group Theory Academic, New York, 1959 . 40 M. E. Rose, Elementary Theory of Angular Momentum Wiley, New York, 1957 . 41 O. Klein, Z. Phys. 58, 730 1929 . 42 H. B. G. Casimir, Rotation of a Rigid Body in Quantum Mechanics Woltjers, The Hague, 1931 . 43 J. H. Van Vleck, Rev. Mod. Phys. 23, 213 1951 . 44 R. N. Zare, Angular Momentum Wiley, New York, 1988 . 45 For example, see A. R. Edmonds, Angular Momentum in Quantum Mechanics Princeton University Press, Princeton, 1974 . 46 J. Z. H. Zhang, Theory and Application of Quantum Molecular Dynamics World Scientic, Singapore, 1998 . 47 J. A. Fleck, Jr., J. R. Morris, and M. D. Feit, Appl. Phys. 10, 129 1976 .

You might also like