You are on page 1of 5

9298

J. Phys. Chem. 1991, 95, 9298-9302

Transition-State Swltchlngs for Single Potential Well Ionic Dissociationst


C. Lifshitz,***F. Louage, V. Aviyente,i
Department of Physical Chemistry and The Fritz Haber Research Center for Molecular Dynamics, The Hebrew University of Jerusalem, Jerusalem 91 904, Israel

and K. SongL
Department of Chemistry, Texas Tech University, Lubbock, Texas 79409 (Received: January 17, 1991: In Final Form: June 24, 1991)

Microcanonical variational transition state theory calculations were applied to the reaction C6HsBr'' C6H5++ Br' in bromobenzene. Calculations were compared with experimental results for C6HS+ time-resolved photoionization efficiency curves and available k ( E ) data for this system. Calculations demonstrate, in agreement with previous results for CHI'', that multiple transition states are possible for single-well potintials of ionic systems. However, comparison with experiments indicates the dominance of the orbiting transition state over a wide energy range above threshold. Further extension of k ( E ) measurements to higher energies is recommended, in order to be able to confidently rule out transition-state switching to a tight transition state as E increases.

Introduction Many ionic dissociations have negligible reverse activation energies. As a result, there is no pronounced maximum in the potential energy surface between reactants and products (save for a rotational barrier) and the location of the transition state is not well-defined. Lifshitz suggested' that a variational criterion for the location of the transition state may be particularly well suited for these systems. In VTST (variational transition state theory) one chooses the transition state at the dividing surface between reactants and products that produces the smallest transition state theory rate c o n ~ t a n t . ~ This * ~ corresponds to a local minimum in the sum of quantum-mechanical states Wof the s - 1 degrees of freedom perpendicular to the reaction coordinate r at energy less than or equal to E - V ( r ) ,where V ( r )is the potential energy at
r

dWv,(E-V(r);r)/t3r= 0 (1) This criterion corresponds to the "bottleneck" for the quasiequilibrium flow of systems from a configuration of reactants to one of products. If it does not occur at the top of an actual potential energy barrier, it is a so-called "entropy" bottleneck. It has been demonstrated4 that the minimum-sum-of-states criterion I and not a minimum-density-of-states criterion is the correct one for microcanonical VTST (pVTST) and that in canonical VTST (CVTST) this corresponds to a maximum free energy criterion. We set out to study the importance of this concept in ionic systems. In order to decide whether entropy bottlenecks are important in ionic systems, one has to search for them in systems where there is definite knowledge that the reaction proceeds via a single-well potential surface. This is almost certainly the case for

duced dipole interactions) than for other potential surfaces. However, TSS has not yet been clearly observed experimentally for reactions proceeding on single-well potential surfaces. TSS may be understood in the following way: As the reactants move along the reaction coordinate, a compromise is struck between two factors:'~~ (a) The energy E - V ( r )decreases as r is extended; this causes a decrease in W. (b) The transitional mode frequencies decrease as r is extended; this causes an increase in W . At low energies the first factor is the dominant one; the transition state occurs at very large r and is the OTS. At higher E's, the two factors may offset each other and TSS is expected. Theoretical microcanonical studies have shown for neutral system^^^,^ that transition states tighten with increasing energy. Marcus and co-workers have recently appliedg-" unimolecular reaction rate theory for highly flexible transition states in neutral systems. The tendency for r+, the fragment-fragment separation distance in the transition state, to decrease with increasing energy E has been noticed, e.g., for C2H6 2CH3 and NCNO N C + NO. When ketene dissociates to CH2 + CO, the transition state may also move in, along the reaction coordinate, to a point closer than complete separation of the reaction products, as the energy increases.12 The role of TSS in ionic systems has been reviewed recently.') Bowers and co-workers have successfully applied7~14~15 a TSS model

in halobenzenes,s which are simple bond cleavage reactions. Transition-state switching (TSS) from a loose, orbiting transition state (OTS) at low internal energies of the ion to a tight transition state (TTS) at high internal energies is expected. Hu and Hase have demonstrated in recent CVTST calculations6 that these two types of transition states are expected to coexist over a wider range of temperatures for potential surfaces with long-range radial potentials proportional to r4 (which are appropriate for ion-in'Dedicated to the memory of Walter J . Chesnavich with whom this research has been initiated. t Archie and Marjorie Sherman Professor of Chemistry. 8 Permanent address: Chemistry Department, Bokazisi Universitesi Bebek, Istanbul, Turkey. Permanent address: Department of Chemistry, Korea National University of Education, Chungwon, Chungbuk 363-791 Korea.

(1) Lifshitz, C. Adu. Mass Spectrom. 1978, 7A, 3. (2) Hase, W. L. Acc. Chem. Res. 1983, 16, 258. (3) Chesnavich, W. J.; Bowers, M. T. In Gas Phase ton Chemistry; Bowers, M. T., Ed.; Academic Press: New York, 1979; Vol. l , p 119. (4) (a) Marcus, R. A. J . Chem. Phys. 1966, 45,2630. (b) Bunker, D. L.; Pattengill, M. J . Chem. Phys. 1968, 48, 772. (c) Wong, W. H.; Marcus, R. A. J . Chem. Phys. 1971, 55, 5625. (d) Hase, W. L. J . Chem. Phys. 1972, 57,730. (e) Quack, M.; Troe, J. Ber. Bumen-Ges. Phys. Chem. 1977,81,329. (f) Garrett, B. C.; Truhlar, D. G. J . Chem. Phys. 1979, 70, 1593. (8) Truhlar, D. G.; Garrett, B. C. Annu. Reo. Phys. Chem. 1984, 35, 159. (h) Wardlaw, D. M.; Marcus, R. A. Adu. Chem. Phys. 1988, 70, 231. (5) Malinovich, Y.; Arakawa, R.; Haase, G.; Lifshitz, C. J . Phys. Chem. 1985, 89, 2253. ( 6 ) Hu, X.; Hase, W. L. J . Chem. Phys. 1989, 93, 6029. (7) Bowers. M. T.; Jarrold. M. J.; Wagner-Redeker. W.; Kemper, P. R.; Bas's, L. M. Faraday Discuss. Chem. So:. 1983, 75, 57. (8) Miller, W. H. J . Chem. Phys. 1976, 65, 2216. (9) Wardlaw, D. M.; Marcus, R. A . J . Phys. Chem. 1986, 90, 5383. ( I O ) Klippenstein, S.J.; Marcus, R. A. J . Phys. Chem. 1988, 92, 3105. ( I I ) Klippenstein, S. J.; Khundkar, L. R.; Zewail, A. H.; Marcus, R. A. J Chem. Phys. 1988,89, 4761. (12) Baum, R. M. Science 1988, 66, 31. ( I 3) Lifshitz, C . Ado. Mass. Specrrom. 1989, I I , 7 1 3 . (14) Chesnavich, W. J.; Bass, L.; Su, T.; Bowers, M . T. J . Chem. Phys. 1981. 74, 2228. (15) Jarrold, M. F.; Wagner-Redeker, W.; Illies, A . J.; Kirchner, N . J.; Bowers, M . T. I n t . J . Mass Spectrom. Ion Phys. 1984, 58, 63.

0022-3654/91/2095-9298$02.50/0

0 1991 American Chemical Society

Single Potential Well Ionic Dissociations for unimolecular and bimolecular reactions in the C4H6.+,C4H8*+, and C6H;+ systems. For example, the reaction between ethylene ion and ethylene molecule proceeds via an OTS at low energies; however, the unimolecular dissociations of C4H8*+ which occur at slightly higher energies proceed via TTS. All of these systems possess double- or multiple-well potential surfaces. There has been further discussion of the applicability of tight transition states representative of entropy bottlenecks and of TSS to ionic syst e m ~ . ~However, ~ ' ~ until now there has been no realistic comparison of a VTST calculation on a single-well potential surface with an experiment for an ionic system. In addition to proceeding on single-well potential surfaces, reactions 1 in halobenzenes have additional advantages that make them good candidates for theoretical modeling. The heat of formation of the phenyl cation AH&c6H5+) and the critical energies Eo for the reactions are quite well-known.20 There are several independent determinations of the microcanonical rate coefficients k ( E ) as a function of albeit Over limited energy ranges. These reactions have also been studied experimentally by threshold ion photodissociation26and by time-resolved photoionization mass ~pectrometry.~*~' In the present study we have concentrated on reaction 1 in bromobenzene. We have applied the microcanonical VTST a p proach developed originally by Chesnavich2" for H' loss in the methane radical cation.

The Journal of Physical Chemistry, Vol. 95, No. 23, 1991 9299 where Eo is the critical energy of reaction 1. Summation is done for the reaction coordinate (670 cm-I) and two bending frequencies of bromine (254 and 181 cm-I). Thus, for Eo = 2.76 eV, De = 2.83 eV. The equilibrium C-Br bond length is re = 2.00 A. The remaining constants in eq I1 were chosen to yield the correct long-range ion-induced dipole force and an appropriate value for hoc, the C-Br stretching frequency in the harmonic oscillator limit. We have taken a,the polarizability of Br' to be 3.5 A3 and, as noted above, hoc = 670 cm-I. The parameter C2is calculated from the charge-induced dipole interaction
C2 = aq2/2r$De

( W

to be C2 = 0.557. The parameter CI is obtained by the condition V"(r,) = mw,2 (VI where hw, is the reaction coordinate frequency of 670 cm-I. This gives

C I = [-(IlC2-g-21)
where

+ [(11C2-g-21)24(3 - C2)(6g - 24c2)1"21/[2(3 - C2)l (VI)


g=m r : (
h~,)~/2D,h~ (VW

The Computational Model There are 30 vibrational frequencies in the bromobenzene cation radical but only 27 vibrations in the phenyl cation. The reaction coordinate, r, correlates with a C-Br stretching vibration of the reagent and with the radial translational degree of freedom of the products. Two additional modes are the so-called "transitional" modes which change from bending vibrations of the reagent to hindered rotational modes and finally to free rotors of the products. The exact position along r of the loosening of these transitional modes determines whether there is TSS. Early loosening refers to the C6H5Br'+region of configuration space and late loosening to the C6H5++ Br' region. The following empirical was used for the reaction coordinate potential
VC-dr) = [D,/(CI - 6 ) l M 3 - C2) exp[Cl(l - x)l - (4c2 C i G + C I )- ~ (CI - 6 ) C P l (11) where X = r/rc and where De and re are the C-Br dissociation energy and equilibrium bond length. The dissociation energy contains a zero-point energy correction De

and CI= 19.34. Previous experience2"has shown that choosing several different empirical power series expressions for the reaction coordinate potential gave variational results that were virtually identical with those obtained with eq 11, provided the empirical parameters were chosen to give the same values for re, De, hoe, and a. The two C-Br bending vibrations of C&$r+, with frequencies of 254 and 181 cm-I, were considered as transitional modes. The bending mode potential is calculated by Vb(6) = (V0(r)/2)(1 -cos 26) (VIII) Each of the two transitional modes was assumed to be a hindered rotor with the rotor barrier height Vo(r) taken to be a smoothly decreasing function of distance along the reaction coordinate Vo(r) = Vc exp[-a(r -

(IX)

Eo

+ f/ZChw,

(IW

(16) Dodd, J. A,; Golden, D. M.; Brauman, J. L. J . Chem. Phys. 1984,80, 1894. (17) Truhlar, D. C. J . Chem. Phys. 1985, 82, 2166. (18) Chesnavich, W. J.; Bowers, M. T. J . Chem. Phys. 1985,82, 2168. (19) Dodd, J. A.; Golden, D. M.; Brauman, J. L. J . Chem. Phys. 1985,82, 2 169. (20) Lias, S.G.;Bartmess, J. E.; Liebman, J. F.; Holmes, J. L.; Levin, R. D.; Mallard, W. G. J . Phys. Chem. Ref Data 1988, 17 (Suppl. 1). (21) Baer, T.: Tsai, B. P.; Smith, D.; Murray, P. T. J . Chem. Phys. 1976, 64, 2460. (22) (a) Rosenstock, H. M.; Stockbauer, R.; Parr, A. C. J . Chem. Phys. 1979,71,3708. (b) Rosenstock, H. M.; Stockbauer, R.;Parr, A. C. J . Chem. Phys. 1980, 73, 733. (c) Dannacher, J.; Rosenstock, H. M.; Buff, R.; Parr, A. C.; Stockbauer, R.L.; Bombach, R.;Stadelmann, J.-P. Chem. Phys. 1983, 75. 23. (23) Pratt, S .T.; Chupka, W. A. Chem. Phys. 1981,62, 153. (24) Durant, J . L.; Rider, D. M.; Anderson, S.L.; Proch, F. D.; Zare, R. N. J . Chem. Phys. 1984,80, 1817. (25) Stanley, R. J.; Cook, M.; Castleman, Jr., A. W. J. Phys. Chem. 1990, 94, 3668. (26) (a) Dunbar, R.C.; Honovich, J. P. Inr. J . Mass Spectrom. Ion Processes 1984,58,25. (b) Dunbar, R.C. J . Am. Chem. Soc. 1989. f l f , 5572. (27) Malinovich, Y.; Lifshitz, C. J . Phys. Chem. 1986, 90, 2200. (28) Chesnavich, W. J. J . Chem. Phys. 1986,84, 2615. (29) Equation I1 follows directly from the general expression V(r) = a exp[-b(l x)] - c x 6 - dx4 by requiring that V(rJ = -0, and V'(r,) = 0.

where V, is the equilibrium barrier height and a is a parameter. Smaller a (a = 1) makes loosening of the transitional modes occur late, while larger a (a = 2) makes it occur early, Le., at low r values. V, in eq IX was chosen so that at r = re each hindered rotor becomes, in the low-energy harmonic oscillator limit, the corresponding bending frequency. V, is calculated from the geometric mean of the two bending frequencies to be 7.66 eV. The variational sums of states were calculated as a function of energy, reaction coordinate (C-Br distance), and parameter a. The microcanonical k ( E ) rate coefficients were calculated by using the minimum sum of states at a given a and energy. Twenty-seven of the C6H5Br*+ vibrations are "spectatorwmodes in that they undergo only minor changes during the decomposition process. In the terminology of Klippenstein et al." these are the so-called "conserved" modes. Vibrational frequencies and rotational constants of C6H5Br'+ and C6H5+were adopted as previously.s In the zero angular momentum limit the sum of states as a function of E and r is given by W(E,r) = Cglg2Wspec[E- VCBr(r) - El - E21
nlc2

(x)

where Wspec is the sum of states of the spectator modes, ni the quantum number of the ith transitional mode, g, the degeneracy of the ith transitional mode, and E, the energy level of the ith transitional mode. A direct count algorithmMwas used for WsF The frequencies given by Klots" for the phenyl cation were used as the frequencies for the spectator modes. The hindered rotor energy levels Ei and degeneracies gi were calculated from the
(30) Stein, S.E.; Rabinovich, B. S.J . Chem. Phys. 1972. 58, 2438. (31) Klots, C. E. Z . Narurforsch. 1972. 27A, 553.

9300 The Journal of Physical Chemistry, Vol. 95, No. 23, I991

Lifshitz et al.
TABLE I: Microcanonical VTST Sums of States W ( E , r ) and Rate Coefficients k(E) at a Critical Energy of Activation, Eo = 2.83 eV, for C6H5Br'+ C6H5+ Br'

E , eV 2.93 3.03 3.13 3.23 3.33 3.43 3.53 3.63 3.73 3.83 3.93 4.03 4.13 4.23 4.33 4.43 4.53 4.63 4.73 4.83 2.93 3.03 3.1 3 3.23 3.33 3.43 3.53 3.63 3.73 3.83 3.93 4.03 4.13 4.23 4.33 4.43 4.53 4.63 4.73 4.83 2.93 3.03 3.13 3.23 3.33 3.43 3.53 3.63 3.73 3.83 3.93 4.03 4.13 4.23 4.33 4.43 4.53 4.63 4.73 4.83

W(E A
1 .20106E+04C

r+(E),"%,
50b 50 50 3.7 3.7 3.6 3.6 3.6 3.6 3.5 3.5 3.5 3.5 3.5 3.5 3.4 3.4 3.4 3.4 3.4 50 50 50 50 50 50 50 50 50 50 50 50 50 50 3.3 3.2 3.2 3.2 3.2 3.2 50 50 50 50 50 50 50 50 50 50 50 50 50 50
50

k ( E ) , s-I 9.5 1760E+01 6.58441E+02 3.24665E+03 1.13759E+04 2.8701 1E+04 6.49244E+04 1.33437E+05 2.587398+05 4.75571E+05 8.305678+05 1.38031E+06 2.21536E+06 3.456318+06 5.24775E+06 7.7831 9E+06 1.120878+07 1.57717E+07 2.18 106E+07 2.969998+07 3.985258+07 9.51760E+OI 6.58441 E+02 3.246658+03 1.1547 1E+04 3.471848+04 9.07 I948+04 2.1 1985E+05 4.574858+05 9.12875E+05 1.72736E+06 3.090868+06 5.318428+06 8.76819E+06 1.40235E+07 2.07703E+07 2.917858+07 4.016 l6E+07 5.44238E+07 7.270568+07 9.58570E+07 9.51760E+OI 6.58441E+02 3.24665E+03 1.1547I E+04 3.47184E+04 9.07 194E+04 2.13985E+05 4.574858+05 9.128758+05 1.727368+06 3.09086E+06 5.3 1842E+06 8.768 19E+06 1.402358+07 2.16988E+07 3.27698E+07 4.82110E+07 6.95099E+07 9.80990E+07 1.36118E+OS

Loosening Parameter, a = 1 .O
1.509828+05 1.33493E+06 8.28205E+06 3.65551E+07 1.43006E+08 5.027258+08 1.64986E+09 5.08084E+09 1.47236E+IO 4.022388+10 1.05178E+lI 2.65045E+lI 6.44609E+ll 1.51921E+12 3.44980E+12 7.59724E+12 1.63250E+l3 3.43021E+13 7.054668+13
I .20 106E+04 I .50982E+05 1.33493 E+06 8.406758+06 4.42 I9 1E+07 1.998248+08 7.98659E+08 2.917178+09 9.75286E+09 3.06210E+10 9.00713E+10 2.525028+11 6.72383E+I 1 3.72258E+12 4.0541 8E+ 12 8.980508+12 1.93459E+13 4.073558+ 13 8.39717E+13 1.69685E+14

4
r

(A)

Figure I . Variational sum of states W(E,r)values for bromobenzene with Eo = 2.76 eV.

harmonic oscillator or Pitzer rotor expression, depending on ni, as previously described.28 The values of Eiwere calculated as follows: First, the smallest even integer n* 1 ( I r V o / 2 h Z ) 'was /Z determined, where I, is the reduced moment of inertia
= Iph+-l I/cc(r

Loosening Parameter, a = 1.5

+ rb)Z

Here, rb = 1.54 A is the distance between the center of the benzene is the moment of inertia of the ring and a carbon atom, and IPh+ phenyl cation. Ei is then calculated, when ni I n* - 2 by the harmonic oscillator expression

Ei = (ni + f/2)(2V,h2/Z,)1/z
with g, = 1. If ni

(XIU

> n* - 2, the Pitzer rotor expression is used


(XIII)

with gi = 0 for odd n, gi = 2 for even n, and gi = 1 for n = 0.


Results and Discussion

Two transition states were observed for each energy, one at a small internuclear distance and one at a large internuclear distance; i.e., the simultaneous existence of multiple transition states is possible (see Figure I). Late loosening of the transitional modes leads to TSS. Thus, for a loosening parameter a = 2 there is no TSS and the OTS is the dominant transition state (Table I). For late loosening of the transitional modes ( a = I ) , TSS occurs at an internal energy of 0.4 eV (Table I); below this energy the OTS is rate determining, while above this energy the TTS is rate determining. The a = 2 result is essentially the phase space theory (PST) result, and the rotational angular momentum J distribution appropriate for 300 K has practically no effect on the calculated microcanonical rate coefficients. The existence of multiple transition states as we have observed here (Figure 1 and Table I ) warrants further discussion. McAd00'~has suggested that ion-neutral complexes may reside in "entropic wells" between the TTS and the OTS. Indeed, for systems with two transition states, a maximum must exist on the W(E,r)curve between the transition states.33 Two transition states were discovered33in the HeHz+ HeH' + H half-reaction with switching between them as a function of energy. The lifetime of the pseudocomplex trapped between the two transition states was estimated to be about 200 fs, suggesting that femtosecond spectroscopy may be able to be used as a probe for multiple transition states. The suggestion of ion-neutral complexes between the ?TS

Loosening Parameter, a = 2.0


1.20106E+04 1.50982E+05 1.33493E+06 8.40675E+06 4.42191E+07 1.99824E+08 7.98659Et08 2.917178+09 9.75286E+09 3.06210Et10 9.00713E+10 2.52502Et 11 6.72383E+Il 1.72258E+ 12 4.23540E+12 1.00858&+12 2.322338+13 5.20273E+l3 1.13300E+14 2.40955E+14

50 50 50 50 50

a r+ is the fragment-fragment separation distance for which W ( E , r ) is a minimum. bCalculationswere carried up to r = 50 %, so that rt L 50 8, for all the entries of 50 A. 'Read as 1.20106 X IO4.

and OTS has been borne out by ab initio calculations on the Icetone ion system.34 The ab initio calculation was carried out
(34) Heinrich, N.; Louage, F.; Lifshitz, C.; Schwarz, SOC. 1988, 110, 8183.

(32) McAdoo, D. J . Mass Spectrom. Reo. 1988, 7, 363. (33) Song, K.: Chesnavich, W. J. J . Chem. Phys. 1989, 91, 4664

H.J . Am. Chem,

Single Potential Well Ionic Dissociations for a-cleavage in acetone.34 When the C-C bond was elongated systematically, there was initially a steep increase in the potential energy, reflecting the loss of covalent interaction, following which the methyl moiety was found to migrate along the positively charged rod for the acylium ion attracted by purely electrostatic interactions about 5 kcal/mol below the combined energies of the separated products. The surface in this region turned out to be extremely flat, and significant geometrical changes exerted practically negligible effects on the total energy. The transition structure for the methyl migration, separating the steeply rising part of the surface from the flat part, was found to have one extremely low imaginary frequency of 25 cm-I. The CH3CO+/CH3*separating pair thus forms an ion-neutral complex between an inner saddle point transition state, which does not lead to any pronounced barrier, and an outer OTS which exists at large separations. There are cases for which multiple transition states on ab initio potential energy surfaces with no pronounced barriers have been demonstrated for neutral reactions. For some of these, the potential surfaces lead to peculiar shapes of the minimumenergy path-similar to the one for the acetone surface. It has been demonstrated by Song and Chesnavich that the peculiar curvature changes in the potential are the origin of the inner transition state (TTS) found by Rai and T r ~ h l a rfor , ~ the ~ dissociation H 0 2 0 OH. The empirical potential for the bromobenzene reaction (eq 11) demonstrates no such peculiarities, yet our calculations demonstrate multiple transition states and TSS a t a = 1. Hase, who has performed calculations for model ionic systems,%has come to a contradictory conclusion. According to Hase, long-range potentials -c/rc' have only one transition state at each energy. As the energy E increases, the transition state moves to lower r (Le., becomes tighter), but there is no TSS. What is the reason for the disagreement with the present results as well as previous results by Chesna~ich.***~~ We feel that the disagreement stems from the fact that Hase employs the classical harmonic oscillator expression for the sum of states

The Journal of Physical Chemistry, Vol. 95, No. 23, I991 9301

Eo=63.65 kcal /mol

11.6

11.8

12.0 12.2 E N E R G Y , eV

12.4

12.6

Figure 2. Experimental (open circles and squares, ref 5) and calculated (continuous curves) time-resolved photoionization efficiency curves for C6H5+ from C,H5Br. The loosening parameter a = 1 corresponds to late loosening of the transitional modes.
0.2 E,=65.26

60

40

20

W(E,r) = [ E - V(r)ls/s!hhvi(r)
i= I

0
11.4 11:6

which is physically unrealistic for the long-range behavior. In addition, Hase's model assumes different reaction coordinate and transitional mode potentials than we have used. Hase and cow o r k e r ~ ~ .have ~~,~ occasionally ' employed the exponential attenuation of vibrational frequencies upon bond elongation, following J o h n ~ t o n 'and ~ Quack and T r ~ e , " . namely ~~

11.8

12.0

12.2

12.4

12.6

E N E R G Y , eV

Figure 3. As in Figure 2 except that the calculations are for early


loosening of the transitional modes and higher critical energy Eo.
IO0
a=I
v

44

exp(-air)

(XV)

If no restrictions are imposed upon the parameters a and ai in eqs IX and XV, it would not be hard to come up with parameters that would give similar results. However, Hase's value ai = 0.10 A-l for ionic systems36implies that the transitional mode potential is still within 40% of the equilibrium value even when the ion and neutral are separated by IO A. This restriction on ai would lead to a rather different behavior of the two expressions employed for the transitional modes, and the exp(-a,r) curve will, in fact, be dominated by a TTS even near threshold. One expects the presence of multiple transition states to depend on the form and parameters of the analytic potential. However, given a certain potential, exact counting of states may be preferable. Hase's model does not include the fact that at large r the transitional modes become free rotations and orbiting motions. At large r, eqs XIV and XV predict W(E,r) t o change as exp(+xa,r). In reality, though, at long range the orbiting motion causes W(E,r)to have an 9 behavior.28 Hase does not count the orbiting transition state
(35) Rai, S.N.; Truhlar, D. G. J . Chem. Phys. 1983, 79,6046. (36) Hase, W.L. Chem. Phys. Lett. 1987, 139, 389. (37) Hase. W.L.:Hu, X. Chem. Phys. Lett. 1989, 156, 115. (38) Johnston, H. S. Gas Phase Reaction Rate Theory; Ronald: New York, 1966; p 82. (39) Quack, M.; Troe, J. Ber. Bunsen-Ges. Phys. Chem. 1974, 78, 240.

8oi

Eo=65.26 kcol/mole

2ms

11.4

11.6

11.8

12.0

12.2

12.4

12.6

E N E R G Y , eV

Figure 4. As in Figure 3 but for late loosening of the transitional modes.

as a minimum in the sum of states; consequently, his conclusion that long-range potentials -e/# have only one transition state at each energy is not inconsistent with the present results. We next compared the results of VTST calculations with experiment. The range of J values populated in the unimolecular reaction 1 is limited, and the comparison of the PST calculations with the VTST data for early loosening of the transitional modes

9302
IO '

J . Phys. Chem. 1991, 95, 9302-93 I O

*:

IO6

I 2 I -

z
W

0 0

t-

a:

lo5

lo4 1 2

1 24

1 28

13.2

E N E R G Y , eV

Figure 5. Comparison of calculated VTST k ( E ) curves (dashed) with experiment (continuouscurve). The calculations are for a critical energy 2.83 eV (= 65.26 kcal/mol). The loosening parameters a = 2 and a = 1 correspond to early and late loosening of the transitional modes, respectively. Early loosening is basically the phase space theory (PST) result. The experimental curve RSP is from coincidence experimentszzb by Rosenstock, Stockbauer, and Parr. demonstrates that the present calculations for the J = 0 approximation may be compared with experiment. We compared calculated VTST time-resolved photoionization efficiency curves with experimental ones from ref 5 . The method for calculating PIE curves has been de~cribed.~ We employed a critical energy Eo = 2.76 eV (= 63.65 kcal/mol) which originally gave agreement,5on the basis of a fixed RRKM transition state with a 1000 K activation entropy AS* = +8.07 eu. The result for late loosening with a = 1 and Eo = 2.76 eV is seen to be in disagreement with experimental results (Figure 2). A slightly higher critical energy, Eo = 2.83 eV (= 65.26 kcal/mol), and early loosening ( a = 2), which is the PST result, are in excellent agreement with experiment (Figure 3). Keeping Eo constant at 2.83 eV and assuming late loosening of the transitional modes ( a = 1) has a very minor

effect on the agreement with experiment (Figure 4). We next compared the calculated microcanonical VTST rate coefficients with the photoelectron photoion coincidence (PEPICO) results of Rosenstock, Stockbauer, and Parr22b (RSP, Figure 5 ) . The coincidence data are the ones that were fittedub by an RRKM model with Eo = 2.76 eV and AS* = 8.07 eu. Figure 5 shows that the RSP results are in good agreement with early loosening PST results ( a = 2, Eo = 2.83 eV) between IO5and IO6 s-l above and below which the curves diverge slightly, while the late loosening curve ( a = 1, Eo = 2.83 eV) is in fair agreement with the RSP result between IO4 and IO5 s-I and diverges considerably at energies higher than 12.4 eV ( E > 3.43 eV). Comparison of the VTST results with experiments is still incomplete. More accurate experiments giving k ( E ) over wider energy ranges, particularly in the range k ( E ) IIO7 s-', are desirable. The present data indicate best agreement for a totally loose OTS, Le., early loosening of the transitional modes. Pratt and C h ~ p k were a ~ ~the first to suggest the use of VTST for the halobenzene ions but have not actually performed the calculations. In agreement with their results, we find the OTS to fit experimental data and to give a slightly higher critical energy than a fixed TTS (Eo(OTS) = 2.83 eV versus Eo(TTS) = 2.76 eV) and as a result a slightly higher heat of formation for the = 1148 kJ mol-'. phenyl cation,27 AHfoo(C,H5+) The comparison of VTST calculations with microcanonical rate coefficients performed here demonstrates the necessity of more refined ab initio potential energy surfaces in addition to extension of rate-energy measurements to wider energy ranges. Particularly desirable would be detailed studies of energy deposition in ionic systems, since these have been found to be very sensitive measures for the degree of tightness in the transition state in the case of neutral systems such as C H 2 C 0 CH, + C0.40

Conclusions Microcanonical VTST calculations for ionic systems demonstrate the existence of multiple transition states. Comparison of calculated k ( E ) data with experimental ones indicates that early loosening of the transitional modes prevails.
Acknowledgment. This research was supported by a grant from the United States-Israel Binational Science Foundation (BSF), Jerusalem, Israel. Prof. M. T. Bowers served as the American Cooperative Investigator for this grant. Registry No. C,H5Brt, 55450-33-4.
(40) Klippenstein. S. J.: Marcus, R. A. J . Chem. Phys. 1990, 93, 2418.

Role of Sllylene In the Deposition of Hydrogenated Amorphous Silicon


Thomas R. Dietrich, Stefan0 Chiussi, Michael Marek, Angelika Roth, and Franz J. Comes*
Institut fur Physikalische und Theoretische Chemie, J . W. Goethe- Uniuersitat, Niederurseler Hang, W-6000 Frankfurt am Main 50, Germany (Received:January 22, 1991; In Final Form: June 19, 1991)

The role of silylene in the laser deposition of hydrogenated amorphous silicon has been studied with laser-induced fluorescence and deposition rate measurements. The rate constants of the reactions of silylene with silane and disilane and of the reverse reactions have been determined. The results show that silylene is rapidly consumed, exhibiting only a small effective lifetime. It proves that generally silylene will hardly be able to reach the surface to form amorphous silicon. The comparison of the kinetic data with the deposition rates shows that in 1R laser CVD silylene starts the gas-phase chemistry and that disilene is the main film-forming molecule. The UV laser process starts with a different primary dissociation leading to silylsilylene, which also rearranges to the film-forming disilene.

Introduction With the aid of photovoltaic cells, solar energy can be transformed into electric energy. For about 15 years hydrogenated amorphous silicon (a-Si:H) has been studied successfully, mainly for applications in solar cells.'4 Due to its high capability in
0022-3654/9 I /2095-9302$02.50/0

absorbing visible radiation, film thicknesses of 0.5 pm are sufficient for photovoltaic devices,' whereas crystalline material thicknesses
( I ) Heywang, W.; Plattner, R. D. Merall (Berlin) 1983, 37, 49. (2) Winterling, G.Phys. unserer Zeir 1985, 16. 50.

0 199 1 American Chemical Society

You might also like