19
Solvation dynamics for an ion pair in a polar solvent: Time-dependent fluorescence and photochemical charge transfer Emily A. Carte@ and James T. Hynes Department of Chemistry and Biochemistry, University of Colorado, Boulder, Colorado 80309-0215 (Received 12 December 1990;accepted 15 January 199 1) The results of a molecular dynamics (MD) computer simulation are presented for the solvation dynamics of an ion pair instanteously produced from a neutral pair, in a model polar aprotic solvent. These time-dependent fluorescence dynamics are analyzed theoretically to examine the validity of severallinear response theory approaches, as well as of various theoretical descriptions (e.g., Langevin equation) for the solvent dynamics per se. It is found that thesedynamics are dominated for short times by a simple inertial Gaussianbehavior, a feature which is absentin many current theoretical treatments, and which is related to the approximate validity of linear response theory. Nonlinear aspects, such as an overall spectral narrowing, but a transient initial spectral broadening, are also discussed. A model photochemical charge transfer process is also briefly consideredto elucidate aspects of the connection betweensolvation dynamics and chemical kinetic population evolution. I. INTRODUCTION The theoretical and experimental study of solvation dy- namics has seenan explosive growth in recent years.’ This intense interest stems both from the accessibility of direct experimental scrutiny via picosecond and subpicosecond time-dependent fluorescence’” and the potential impor- tance of these dynamics in heavy particle charge transfer reactions5 and electron6and proton transfers’ in solution. In this paper, we present a molecular dynamics (MD) simulation study of the solvation dynamics of an ion pair instantaneously created from a neutral pair, in a dipolar aprotic solvent. Such simulationss-‘4 are a useful tool in as- sessing the validity of approximate theoretical descrip- tions,‘5-20 as well as a source of microscopic insight for the interpretation of experiment (and theory). Our attention is focused on the time-dependent fluores- cence spectral frequency shift and emission spectrum as probes of solvation dynamics, and we attempt to place the MD results in theoretical perspective,with an eye toward experimental consequences. In particular, we establish the major role played by a simple inertial Gaussiantime behav- ior,” point out its absence in many popular theoretical treat- ments, and indicate its assistance in the approximate validity of a linear response treatment. We also discuss somenonlin- ear aspects of the problem, including an overall spectral nar- rowing-due to solute-dependent solvent force constants,as well as transient spectral broadening and solute-dependent solvent friction. We also briefly consider a closely related model photo- chemical reaction. Our emphasis here is on the relation between the solvation dynamics and the chemical kinetic aspects of the reaction. ” Permanent address: Department of Chemistry and Biochemistry, Uni- versity of California, Las Angeles, CA 90024-I 569. The outline of this paper is as follows. In Sec. II, we describethe model and simulation procedures, and present the simulation results. In Sec.III, we interpret the results in terms of linear response descriptions and examinethe appli- cability of a Langevin equation for the solvent coordinate. Here we also demonstrate that a significant fraction of the relaxation can be simply accounted for by short time Gaus- sian behavior. Section IV is devoted to examination of some details of the molecular trajectories, while Sec.V concludes. Some particular results of our study have been published previously. II. MODEL DESCRIPTION AND SIMULATION RESULTS A. Solute and solvent description We will considera donor-acceptor solute pair (SP)DA, where D = A, immersed in a solvent of rigid dipolar mole- cules. This SP will alternately be assigned point chargesof zero or plus and minus e, respectively, at the atomic D and A site centers to represent a ground electronic state neutral pair (NP)DA and an electronically excited ion pair ( IP ) D + A - . The constituent members of the SP eachhave mass 40 amu; their centersare rigidly separated by 3.0 A, but the SP is free to translate and rotate without vibration. The solvent is composed of 342 rigid dipolar molecules with constituent atoms of mass40 amu separated from each other by a fixed distance of 2.0 A and with partial charges f q such that the dipole moment is 2.4 D. The number den- sity is 0.012 A-’ and the temperature is 250 K. This sol- vent,21-23 which is very roughly similar to methyl chloride, is similar to members of the class of dipolar aprotic solvents currently under experimental investigation2’b)T4f(a)*4(b) for solvation dynamics and charge transfer reactions. The total potential energy of interaction U consists of Lennard-Jones and Coulomb potentials U(r,> = ULJ(rij) +ZiZjry’ (2.1) between each atomicsite,with Zi the change on sitei. The LJ J. Chem. Phys. 94 (9), 1 May 1991 0021-9606/91/095961-19$03.00 0 1991 American Institute of Physics 5961 Downloaded 19 Mar 2007 to 128.112.35.7. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

The Carter Group - Solvation dynamics for an ion pair in a ......5962 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent parameters are dks = 200 K and IJ = 3.5 b; for each

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: The Carter Group - Solvation dynamics for an ion pair in a ......5962 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent parameters are dks = 200 K and IJ = 3.5 b; for each

Solvation dynamics for an ion pair in a polar solvent: Time-dependent fluorescence and photochemical charge transfer

Emily A. Carte@ and James T. Hynes Department of Chemistry and Biochemistry, University of Colorado, Boulder, Colorado 80309-0215

(Received 12 December 1990; accepted 15 January 199 1)

The results of a molecular dynamics (MD) computer simulation are presented for the solvation dynamics of an ion pair instanteously produced from a neutral pair, in a model polar aprotic solvent. These time-dependent fluorescence dynamics are analyzed theoretically to examine the validity of several linear response theory approaches, as well as of various theoretical descriptions (e.g., Langevin equation) for the solvent dynamics per se. It is found that these dynamics are dominated for short times by a simple inertial Gaussian behavior, a feature which is absent in many current theoretical treatments, and which is related to the approximate validity of linear response theory. Nonlinear aspects, such as an overall spectral narrowing, but a transient initial spectral broadening, are also discussed. A model photochemical charge transfer process is also briefly considered to elucidate aspects of the connection between solvation dynamics and chemical kinetic population evolution.

I. INTRODUCTION

The theoretical and experimental study of solvation dy- namics has seen an explosive growth in recent years.’ This intense interest stems both from the accessibility of direct experimental scrutiny via picosecond and subpicosecond time-dependent fluorescence’” and the potential impor- tance of these dynamics in heavy particle charge transfer reactions5 and electron6 and proton transfers’ in solution.

In this paper, we present a molecular dynamics (MD) simulation study of the solvation dynamics of an ion pair instantaneously created from a neutral pair, in a dipolar aprotic solvent. Such simulationss-‘4 are a useful tool in as- sessing the validity of approximate theoretical descrip- tions,‘5-20 as well as a source of microscopic insight for the interpretation of experiment (and theory).

Our attention is focused on the time-dependent fluores- cence spectral frequency shift and emission spectrum as probes of solvation dynamics, and we attempt to place the MD results in theoretical perspective, with an eye toward experimental consequences. In particular, we establish the major role played by a simple inertial Gaussian time behav- ior,” point out its absence in many popular theoretical treat- ments, and indicate its assistance in the approximate validity of a linear response treatment. We also discuss some nonlin- ear aspects of the problem, including an overall spectral nar- rowing-due to solute-dependent solvent force constants, as well as transient spectral broadening and solute-dependent solvent friction.

We also briefly consider a closely related model photo- chemical reaction. Our emphasis here is on the relation between the solvation dynamics and the chemical kinetic aspects of the reaction.

” Permanent address: Department of Chemistry and Biochemistry, Uni- versity of California, Las Angeles, CA 90024-I 569.

The outline of this paper is as follows. In Sec. II, we describe the model and simulation procedures, and present the simulation results. In Sec. III, we interpret the results in terms of linear response descriptions and examine the appli- cability of a Langevin equation for the solvent coordinate. Here we also demonstrate that a significant fraction of the relaxation can be simply accounted for by short time Gaus- sian behavior. Section IV is devoted to examination of some details of the molecular trajectories, while Sec. V concludes. Some particular results of our study have been published previously. ’ ’

II. MODEL DESCRIPTION AND SIMULATION RESULTS

A. Solute and solvent description

We will consider a donor-acceptor solute pair (SP)DA, where D = A, immersed in a solvent of rigid dipolar mole- cules. This SP will alternately be assigned point charges of zero or plus and minus e, respectively, at the atomic D and A site centers to represent a ground electronic state neutral pair (NP)DA and an electronically excited ion pair ( IP ) D + A - . The constituent members of the SP each have mass 40 amu; their centers are rigidly separated by 3.0 A, but the SP is free to translate and rotate without vibration.

The solvent is composed of 342 rigid dipolar molecules with constituent atoms of mass 40 amu separated from each other by a fixed distance of 2.0 A and with partial charges f q such that the dipole moment is 2.4 D. The number den-

sity is 0.012 A-’ and the temperature is 250 K. This sol- vent,21-23 which is very roughly similar to methyl chloride, is similar to members of the class of dipolar aprotic solvents currently under experimental investigation2’b)T4f(a)*4(b) for solvation dynamics and charge transfer reactions.

The total potential energy of interaction U consists of Lennard-Jones and Coulomb potentials

U(r,> = ULJ(rij) +ZiZjry’ (2.1) between each atomic site, with Zi the change on site i. The LJ

J. Chem. Phys. 94 (9), 1 May 1991 0021-9606/91/095961-19$03.00 0 1991 American Institute of Physics 5961 Downloaded 19 Mar 2007 to 128.112.35.7. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 2: The Carter Group - Solvation dynamics for an ion pair in a ......5962 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent parameters are dks = 200 K and IJ = 3.5 b; for each

5962 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent

parameters are dks = 200 K and IJ = 3.5 b; for each site in the SP and the solvent molecules.

The Hamiltonian of the ground-state NP and the excit- ed-state IP systems are, respectively,

H, =&.a +HO,, +Hs + Urw,s; (2.2)

H, = K,p + H;p -t Hs + Up,, . (2.3) Here KNp and KIp are the Hamiltonian, comprising the translational and rotational kinetic energies, of the rigid iso- lated NP and IP solutes, while Hip and H$ are corre- sponding constant electronic energies, which include all in- ternal potential energies in the solute pairs. We take

+ia”=H&-HHO,, (2.4) to be a positive number, consistent with the excited elec- tronic character of the IP state. The solvent Hamiltonian, including kinetic and potential energy, is Hs . Finally, U,,, is composed of Lennard-Jones interactions between the sol- vent molecules and the members of the NP, while

u w,s = UN,, + AK (2.5) is the total interaction potential energy between the solvent molecules and the charged members of the IP; H, - H& thus differs from Hg - H& by the Coulombic potential en- ergy

AE = 4,s - UN,, = V cod NP,S (2.6)

of interaction between the charged ions in the IP and the point charges embedded in each of the finite dipolar solvent molecules. It is this energy difference or gap which concerns us in all that follows.

In the point dipole approximation for solvent molecules, AE would be given by

AE= - s

drP(r)*[Ez(r-r*)+E$(r-rB)],

where P(r) is the solvent orientational polarization at posi- tion r in the solvent and Ep(r - ri) is the vacuum electric field at r due to ionic sites in the IP. This is in fact the conven- tional dielectric continuum limit solvent coordinate em- ployed in a number of analytic studies,15-20 which are there- by restricted to accounting for the solvent solely via its point polarization field. (The difficulties with this approximation for water are clearly apparent in Ref. 8.) In the present work, however, we employ the full microscopic definition Eq. (2.6) of U. Nonetheless, this indicates that AE is the mi- croscopic solvent coordinate, a perspective also exploited in other studies.8~‘0-14~2294

It is worth mentioning here that there are no electronic polarization effects included in either the model solute or solvent. In real systems, such effects would contribute, in a static way, to solvent shifts.‘5-‘7 These are completely ig- nored in the present work.25

1. Time-dependent fluorescence

According to the Franck-Condon principle, the SP and solvent nuclei remain fixed in a fluorescence transition. The instantaneous fluorescent frequency, averaged over the ini-

tial conditions, is, from Eqs. (2.2)-( 2.6), given by (cf. Fig. 1)

+2-z(t) = H,(t) - H,(l)

=fiw’+ AE (t). (2.7) Thus the dynamics of &Z(t) are governed by the average solvent-excited state IP Coulomb interaction AE (t); this will become more negative as the solvent equilibrates to the IP by transiting downward in the IP well. Note that by the Franck-Condon principle, the average energy difference between the (nonequilibrium) free energies of the excited and ground states; there is no entropy change in a transition occuring at fixed nuclear coordinates.

It is convenient to define two measures of this TDF fre- quency which are independent of the gas-phase frequency w”. The first is the dynamic average TDF red shift

S+iiG(t) = +iZ(t) -tie = AE (t), (2.8) and the second is the normalized TDF shift

s(t) = S=(t) - a%( CcJ 1 S%(O) - SG( Co)

= AE(t)- AE(co) A??(O) - AE( co) ’ (2.9)

which characterizes the equilibration in the excited state as AE (r) relaxes from its initial value of AE (0) to its final

equilibrium value AE ( CO ). [This quantity is termed “C(t)” in a number of studies.] Both these measures have been examined in experimental TDF studies.14

2. Photochemical charge transfer

Our discussion above was appropriate for the TDF problem. We now shift to the perspective of a model photo-

FIG. 1. Time-dependent fluorescence observed after initial excitation from the NP state to the IP state, followed by fluorescent decay [with frequency fiw( t) ] back to the NP state. Free energy ( AG) curves are shown as func- tion of the many-body solvent coordinate he = VS!‘A ; see Fq. (2.6). As the solvent relaxes in the presence of the IP, the fluorescence frequency de- creases.

J. Chem. Phys., Vol. 94, No. 9,1 May 1991

Downloaded 19 Mar 2007 to 128.112.35.7. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 3: The Carter Group - Solvation dynamics for an ion pair in a ......5962 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent parameters are dks = 200 K and IJ = 3.5 b; for each

E. A. Carter and J. T. Hynes: An ion pair in a polar solvent 5963

chemical charge transfer reaction, illustrated in Fig. 2. We hasten to stress that such a model is ultrasimplified. A transi- tion from a nonpolar ground state leading to a chemically distinct charge transfer (IP) state will typically occur by initial transition to a locally excited (LE) nonpolar state which is electronically coupled to the IP state. Figure 2 cor- responds to the very idealized case where the LErIP state electronic coupling is very strong and there is no barrier on the upper surface. Our sole purpose for considering the situ- ation depicted in Fig. 2 in the photochemical context is to explore, in the simplest case, the question of the connection of the solvation dynamics per se to the chemical kinetic aspects of the problem, a topic of considerable current inter- est. l.lO,i 1.26-29

In this case, we can view the role of the frequency tie in the TDF problem to be played by the intrinsic free energy difference

AGyNT = Hyp -H&s - TASvN’,, (2.10) is the charge shift reaction problem. This internal difference accounts for electronic energy and entropy differences in the isolated IP and NP species, exclusive of solvation free ener- getic terms. In this case, the essential point is that Eqs. (2.8) and (2.9) provide descriptions of the solvent coordinate time dependence as the nascent IP becomes solvent-equili- brated product after laser excitation from the NP reactant. As discussed in more detail in Sec. III B, this variable gauges the product relaxation insofar as the IP curve illustrated in Fig. 2 is dynamically followed; i.e., no electronically nona- diabatic transition IP + NP occur at the crossing point of the two curves to regenerate the reactant NP.

8. Simulation results

” Constant temperature MD simulations were carried out in a periodically replicated cubic box with sides of length 30.52 A. The integration of the equations of motion was ef- fected via the Verlet algorithm3’ and a time step of lo-’ ps. The long range forces were treated by using the Ewald sum- mation method,32 and the bond constraints for the SP and solvent molecules were implemented with the SHAKE algo- rithm.33

FIG. 2. Schematic of photochemically induced charge transfer from NP-IP, where the free energy curves AG are again functions of the many- body solvent coordinate he. The free energy curve for the NP is more shal- low and has a higher energy minimum than the IP curve, since Coulombic interactions with the solvent are absent in the former case.

300

$@e)

200

Ae(crri’)

(a)

“““1

Tp We) 200

-25050 cAE> IP -9050

Ae(ctri’)

0)) FIG. 3. Equilibrium probability distributions of he values for (a) the NP and (b) the IP, determined via constant temperature (250 K) MD simula- tion runs of 410 ps each. The he sampling was performed every 0.1 ps. (he),, = - 81.5 cm-’ with a width of a,,,,, = f 1 804 cm-‘, while (he),, = - 17 659 cm- ’ with a width of o - + 1456cm-‘. AC.lP - -

The TDF characteristics Eqs. (2.8) and (2.9) are ob- tained win a nonequilibrium MD simulation as follows. We focus first on the average dynamic red shift S+iZ( t), Eq. (2.8). The NP-solvent system is first equilibrated for 10 ps. The probability distribution of AE values in the presence of the NP (determined in a 410 ps simulation) is shown in Fig. 3 (a); it is, to a good approximation, a Gaussian, and is cen- tered at the equilibrium average ( AE ) NP = 0. [Since there are no solute charges present to bias the solvent configura- tions, positive and negative values of the Coulomb interac- tion energy AE of the solvent are equally likely.] Initial con- ditions are sampled from the trajectories that comprise this distribution, and for each trajectory, the charges in the IP are turned on instantaneously, h la Franck-Condon. At this initial instant, the solvent configurations are those originally present for the solvent in equilibrium with the uncharged NP, so that the initial value of the average red shift Eq. (2.8) is, of course, zero:

6%(O) = AE (0) = (AE)NP = 0. (2.11)

J. Chem. Phys., Vol. 94, No. 9,l May 1991 Downloaded 19 Mar 2007 to 128.112.35.7. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 4: The Carter Group - Solvation dynamics for an ion pair in a ......5962 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent parameters are dks = 200 K and IJ = 3.5 b; for each

5964 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent

-4000 - 6A5(t), cm-’ -8000 -

-16000-

-18000 . . . I . . . . . . . . .m 0 1 2 4 5 6

0.8

0.6

(b) 0.4

0.2

0.0 0 1 2 4 5 6

FIG. 4. (a) Time-dependent averaged red shift [Eq. (2.8)], Gfiw(t) = (he(t)) computed every 0.1 ps from 198 trajectories. (b) Nor- malized time-dependent red shift, S(t); see Eq. (2.9). Uncertainties in the values of Mw(r) ranged from o,,,, = -h 1 293 to * 2 296 cm - ’ , with the largest uncertainties occurring at early times.

10

u- &oaO Ae(cm-’ )

+wo Ae(cm-t )

The time-dependent red shift Eq. (2.8) is generated by aver- aging AE over 198 trajectories at different subsequent times. The result is shown in Fig. 4(a)

The overall dynamics exhibited by the TDF red shift are fairly rapid, and are distinctly bimodal in time. In particular, there is an extensive initial rapid shift of z 12 000 cm - t in -0.3 ps, followed by a much slower relaxation of some further E; 6 000 cm - i , as the solvent ultimately equilibrates to the IP.34 In a separate simulation of 410 ps duration (after a 10 ps equilibration), the equilibrium distribution of AE for the solvent in the presence of the IP was determined and is displayed in Fig. 3 (b) . Again, as for the NP case, the distri- bution is fairly Gaussian, but this time is centered about the IP equilibrium average (AE ),r = - 17 660 cm - ’ ; this highly negative value reflects the strongly attractive IP-sol- vent Coulomb interactions. The distribution is also more narrow than in the NP case, indicating a higher degree of order for the solvent molecules imposed by the Coulomb forces (uide infia). Finally, comparison of this ( AE ) tp val- ue with the long-time average red shift in Fig. 4(a) indicates that equilibration has indeed occurred in the nonequilibrium dynamics.

The relaxation dynamics are presented in the more com- pact format of the TDF shift function S( t), Eq. (2.9)) in Fig. 4(b). Again (and of course), the bimodal time-decay char- acter is quite apparent.

More details of the relaxation can be gleaned from other probes not often examined. Inspection of Fig. 3 (a) and 3 (b) shows that the initial BE probability distribution-appro-

0 -!awa +mso

Ae(cm “1

40 M

t- 0.5ps

10) 20

10 E- .-

Ae(cm-’ ) Ae(cm“) 4Bso

Ae(cm-’ )

FIG. 5. Simulated time-dependent fluorescence spectrum I( t,Ae), computed from sampling he every 0.1 ps from 200 trajectories, for t = O.O,O. 1,0.2,0.5, 1.0, and 3.0 ps. Notice the continuous shift to the red and the initial broadening of the population distribution, followed by a narrowing of the distribution due to equilibration of the solvent to the excited state (IP).

J. Chem. Phys., Vol. 94, No. 9.1 May 1991

Downloaded 19 Mar 2007 to 128.112.35.7. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 5: The Carter Group - Solvation dynamics for an ion pair in a ......5962 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent parameters are dks = 200 K and IJ = 3.5 b; for each

E. A. Carter and J. T. Hynes: An ion pair in a polar solvent 5965

priate for solvent in equilibrium with the NP-is wider than the final distribution-appropriate for equilibrium with the IP. This suggests that the TDF spectrum should exhibit an overall narrowing with time as solvent equilibration to the IP proceeds. We now examine this.

We can formally define the time-dependent red shift spectrum for the distribution Gfiw( t) by

I(t,Ae) = S[AE(t) - he] , (2.12) where he now represents the numerical value of the red shift 6% We can approximately determine this spectrum by con- structing histograms of the observed numerical values he of the microscopic dynamical variable AE at various stages of the relaxation. A number of these are exhibited in Fig. 5. While an overall narrowing trend with time is indeed ob- served (and this is the major point), there is a discernible initial broadening, and subsequent narrowing, of the red shift spectrum at early times. This is exposed more clearly in Fig. 6 where we display the time dependent square “width” function

w(t) = AE’(t) - [hE( (2.13) for the spectrum. Here the transient broadening followed by a subsequent narrowing is clearly visible. As we will see in Sec. III, the dynamics of W(t) is entirely governed by non- linear effects, while the major aspects of the TDF shift S(t) can be comprehended quite well within a linear framework.

III. THEORETICAL ANALYSIS

In this Section, we provide some theoretical analyses of the simulation results of Sec. II. We discuss in turn the TDF nonequilibrium solvation dynamics in terms of solvation dy- namics in equilibrium, and the special issues arising in the photochemical charge transfer case. We then address a Lan- gevin equation description, and finally a short time Gaussian approximation.

A. TDF shift and spectral width

1. Equilibrium time correlation functions and TDF shift

Two forms of what we will call linear response theory may be employed to attempt to connect the nonequilibrium

W(t), lo6 cm-*

l! . . . . . . . . . . . . . . . . . . . . . . . I 0 1 2 3 4 5 6

t (PS)

FIG. 6. Simulated time-dependent square width function W(t) [Eq. (2.13)), computed every 0.1 ps from 198 trajectories.

normalized TDF shift S(t) to equilibrium time correlation functions. In the early linear response treatments of TDF based on a solvent continuum approach,‘5-*7 such a distinc- tion was unnecessary. But as we will see, the distinction is highly relevant for the interpretation of the MD results.

In the first linear response treatment, the system initial- ly in equilibrium is subject to a step perturbation at t = 0, here given by

H’(t) = (fix0 i- AE)8(t), (3.1) where 8(t) is the step function. This represents the instanta- neous Franck-Condon transition to the electronically excit- ed IP state. The average of any observable is then calculated to first order in this perturbation and, as a consequence, is expressed in terms of an equilibrium tcf in the absence of the perturbation. In the present case, this standard procedure35 yields’ the relation

S?&(t) -S%(co) =p(AEAE(t)),, (3.2) involving the tcf of AE in the ground, NP state, with fl = (k, 7’) - ‘. The corresponding result for S(t) is then

(3.3)

in terms of the normalized tcf of AE in the ground, NP state. In the TDF context, Eqs. (3.2) and (3.3), by virtue of

their focus on averages in the initial NP state, correctly treat the distribution of the initial conditions, which by the Franck-Condon principle, is given by the NP ground state equilibrium distribution

e-“H” fess =

Sdre - BHg ’ (3.4)

where I’ denotes the totality of phase space coordinates and momenta. But on the other hand, Eqs. (3.2) and (3.3) refer to dynamics, e.g., of the solvent, in the presence of the un- charged NP, whereas in fact the solvent dynamics actually occur in the presence of the charged IP. A shift to this latter perspective on the dynamics gives a second and in principle different “linear response” theory approach, which we now describe.

To begin, we write the nonequilibrium average of AE for t>O as

AE (t) = s

drp(t)AE = s

dT(e-‘L’p,,,8)AE

= dQxx,,gAE(th s (3.5)

in which iL is the Liouville operator iL = {H,. } appropriate for dynamics in the presence of the IP; AE( t) is the value of AE at time t calculated with those dynamics. Next, since by Eq. (2.6)) the difference of excited- and ground-state Hamil- tonian is H, - H, = fio” + AE, we can write the NP state equilibrium distribution peq,g, Eq. (3.4), in terms of the IP equilibrium distribution peq,e,

e-PH<

P eq,e = Jdre - RH, ’

to give the nonequilibrium average of AE in the form

(3.6)

J. Chem. Phys., Vol. 94, No. 9,1 May 1991 Downloaded 19 Mar 2007 to 128.112.35.7. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 6: The Carter Group - Solvation dynamics for an ion pair in a ......5962 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent parameters are dks = 200 K and IJ = 3.5 b; for each

5966 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent

m(t) = (@AEAE(t))w VA”)IP ’

(3.7)

where ( ( * * * ) ) Ip denotes the IP excited-state equilibrium average, i.e., with distributionp,,,,. Expansion of exp(PAE) through first order in /3AE then gives for the average TDF shift the relation

S%(t) -SG(o3) =fl(SAESAE(t)),,; SAE= AE- (AE),,,

and for the normalized TDF dhift (3.8)

S(t) = @uaAE(t) )IP ((SAE)‘),, =“‘(‘) (3.9)

involving the normalized equilibrium tcf of SAE in the excit- ed, IP state [in contrast to Eqs. (3.2) and (3.3) I. We can view Eqs. (3.8) and (3.9) as correctly incorporating the fea- ture that the dynamics occur under the influence of the excit- ed state IP Hamiltonian, but with the approximation that the distribution of initial conditions--which should in fact be those appropriate to the NP-is perturbatively described.

We now have two separate linear response theory results for the TDF shifts s(t). In a linear dielectric continuum theory,‘5-17 there is no distinction between them. But at a molecular level, they can differ, as first empirically estab- lished by Maroncelli and Fleming8 for simulated ion solva- tion in water. We note in passing that the relation Eq. (3.9) has been termed an “Onsager linear regression hypothe- sis,“” and indeed this perspective follows naturally from a focus on relaxation in a nonequilibrium ensemble.35 We have emphasized here the (previously unstated) critical physical issues of the different approximations in Eqs. (3.3) and (3.9) for the initial conditions and the dynamics, in order to provide an interpretation below of the simulation results now described.

Figure 7 shows the calculated normalized nonequilibri- urn TDF shifts compared to the two calculated equilibrium tcf s Asp (t) . (The latter are determined in separate equilib- rium simulations of 410 ps duration after a 10 ps equilibrium period.) Both the ground NP state-based tcf A,, (t) and the excited IP state-based tcf A,, (t) are in close agreement with S(t) at short times, while AIp (t) is clearly in better conson- ance with the nonequilibrium shift dynamics at longer times. [A similar phenomenon was observed by Maroncelli and Fleming,* who, however, did not discuss the basis for their analog of Eq. (3.9).] This tells us first that the feature that the NP initial conditions are correctly incorporated in ANP (t)-but not in ALIp (t&appears to have surprisingly little impact on the short time dynamics. It also tells us that a central feature of the longer time relaxation dynamics is its dependence on the presence of the charged IP solute.

From the broader perspective, it is somewhat remark- able that the agreement between S(t) and, e.g., the linear response approximation A,, (t) is so good, given the fact that AE( t) is changing by some 18 000 cm - ’ . Similar appli- cability of linear response approximations has been observed in other simulations.*T’0-‘4 Our discussion in Sec. III D be- low sheds some light on this phenomenon.

We will fasten our attention in Sets. III C and III D

below on examination of the TDF shift, exploiting this ap- proximate linearity. But first we will explore the nonlinear characteristics of the phenomenon. The deviation from lin- ear behavior is much more clearly exposed in the spectrum, Fig. 5, and in particular, in the square width function W(t), Eq. (2.12) and Fig. 6, This is worth some inspection, now undertaken.

2. Dynamic spectral width To begin, we note that the Gaussian character of the AE

distributions in Figs. 3 (a) and 3(b) indicates that there are, to a good approximation, harmonic free energy curves gov- erning fluctuations in AE for both the NP and IP cases. In particular, the formal definition for the probability Psp (he) that AE assumes the numerical value he in a solvent tluctu- ation in the presence of a given solute pair (SP) is

Psp (Ae) = GRAB - Ae) jsp, (3.10) where S denotes the delta function. By Fourier representa- tion of the delta function and a second-order cumulant ex- pansion, the harmonic free energy becomes explicit as”

PSP(Ae) aexp[ -flG(Ae)]

(he- W),P)* 1 2((AE- (AE),,)2),, * (3.11)

(a) 0 1 2 3 4 5 6

t (PSI

..- 0.8 -

(b)

0 1 2 4 5 6

FIG. 7. Normalized nonequilibrium TDF shift function S(t) (solid lines) and simple linear response theory approximations (dashed lines) to S(t): the normalized equilibrium time correlation functions (a) AIP (r) and (b) A,,(t); see Eqs. (3.3) and (3.9). The ALsp (t)‘s were each computed every 0.1 ps over 410 ps equilibrium trajectories in separate calculations.

J. Chem. Phys., Vol. 94, No. 9,1 May 1991

Downloaded 19 Mar 2007 to 128.112.35.7. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 7: The Carter Group - Solvation dynamics for an ion pair in a ......5962 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent parameters are dks = 200 K and IJ = 3.5 b; for each

We can use this in turn to define** a solvent force constant

k, = [P(@W2)s,] -‘; (3.12)

SAEz AE - (AE jsp, associated with the harmonic free energies. Figures 8 (a) and 8(b) display these free energy curves. It is especially noteworthy that k,, > kNP, i.e., the solvent force constant depends on the solute. The fluctuations in AE about the average value are more difficult in the presence of the charged IP compared to the uncharged NP. This nonlinear aspect, which we have described in more detail elsewhere,** is consistent with the picture that, for this system, the nearby polar solvent molecules are more highly ordered in the pres- ence of the IP charges.22,36*37 In earlier work,22 we have ex- plored the consequences of this solute dependence of the sol- vent force constant for the activation energy patterns for electron transfer reactions, and contrasted the estimated magnitudes of such effects with the very much larger values previously suggested by Kakitani, Mataga, and co- workers.36

The function W(t) governs the square width of the time evolving spectrum I( t,Ae), Eq. (2.11); indeed, application to Eq. (2.11) of the procedures used to derive Eq. (3.11) yields the Gaussian approximation

I(t,Ae) = [27rW(t)] -“*exp - [Ae ~~~(““) .

(3.13)

E. A. Carter and J. T. Hynes: An ion pair in a polar solvent 5967

If we develop the averages in Eq. (2.12) up to second order in /3AE using Eq. (3.7) for AE (t) and its analog for AE2(t), we obtain

w(t) = W( co 1 +kq,(t) + $ [G*(t) - 2c:, (t>

+ WhC,,W] 9 (3.14) where the final equilibrium square width is

W( co) = (AE*& - (AE)fp

= Whp) -’ (3.15)

with our IP solvent force constant definition (3.12). Here we have defined the assorted time correlation functions

C,,(t) = ((hW2),,A,,(t) = @hEShE(t)) C,,(t) = (SAE [SAE(t)l*),,; (3.16)

C22tf) = ((SAE)*[SAE(O12),p - WAE)2%, all of which vanish as t-t CO.

The first nonlinear aspect of the spectrum Fig. 5 is its overall narrowing with time. From Eqs. (3.14)-( 3.16) and (3.12), the degree of narrowing in the square width is W(O) - W(m) = @kNp)-‘-- @‘klp)-’

= $- [ ((6AE)4h, - 3WW2):,] , -

(3.17)

be (cm-')

(a)

200 -

o-

-6OO-

.soo-

.lOOOl . . , 1 * . f I . * * I , * ’ 1 . . . .lOOOO -6000 -2000 2000 6000 10000

Ae(cti')

0)

FIG. 8. Simulated solvent free energy curves AG,, (Ae) for SP = (a) IP and (b) NP, versus he. Parabolic fits to the curves are shown by the thin solid lines.

where we have assumed that the IP-solvent free energy is symmetric, even though it is anharmonic to some degree.** There are several points we need to make. First, the overall narrowing arises from the larger solvent force constant in the IP state compared to the NP state-the equilibrium distribu- tion is more narrow in the former than in the latter. This narrowing amounts to a change of several hundred cm - ’ out of several thousand cm - ’ in the width. (Our inference** that the NP-IP force constant size could be reversed in wa- ter suggests that an overall spectral broadening could occur in that solvent.) It is interesting to note that spectral narrow- ing has been experimentally observed by Maroncelli and Fleming’ and by Simon and SU~‘~’ (though it is not clear what role might be played in those systems by vibrational contributions’7929 ) . Second, this narrowing is a completely nonlinear effect: for if the free energy G,, (he) for the IP were perfectly harmonic and the corresponding equilibrium distribution PIP (he) were perfectly Gaussian, then k,, = k,, and Eq. (3.17) shows that there would be no narrowing. [The same statement applies to G,, (he) and PNp (he) 1. Third, and finally, the entire time dependence of W(t) arises from nonlinear effects. This follows of course from the second point just made, and in more detail from Eq. (3.14). It is easy to show that for a linear Gaussian system, the tcf relations C,,(t) = C22(t) - 2C:, (t) = 0 hold, and there is no dynamic narrowing and indeed no change at all in the width. The transient initial broadening prior to ultimate narrowing observed in Fig. 6 (and Fig. 5) arises from the nonlinear tcfs in Eq. (3.14); each term in the series dis- played rises from zero3* and ultimately decays to zero. We have attempted to find a simple connection between the ini-

J. Chem. Phys., Vol. 94, No. 9,l May 1991 Downloaded 19 Mar 2007 to 128.112.35.7. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 8: The Carter Group - Solvation dynamics for an ion pair in a ......5962 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent parameters are dks = 200 K and IJ = 3.5 b; for each

tial transient broadening reflected in W( t)34 and the devi- ation of the TDF shift S(t) from its linear approximation, An, (t) [most apparent in Fig. 7(a) in the 0.25-0.5 ps range], involving only the leading order nonlinear AE tcf C,,(t) in Eqs. (3.14): w(t) - W(O) = w co 1 - W(O) +B [C,*(t) - C,,(O)l.

(3.18) Unfortunately this attempt fails, strongly indicating that the short time transient broadening phenomenon in Fig. 6 is not at all simple in character. We return this question in Sec. IV.

B. Photochemical reaction

In the photochemical reaction case (Fig. 2), a natural focus of attention would be on the rate of appearance of the product IP population in the IP well, subsequent to initial excitation from the NP well. Here we assume that any elec- tronic coupling between the NP and IP curves is sufficiently small, and that the passage through the curve crossing re- gion is sufficiently fast; a Landau-Zener perspective39 then indicates that passage into the IP product well will occur without any electronically nonadiabatic back passage to the NP well.

In the initial Franck-Condon excitation, the initial nearly Gaussian AE distribution in the equilibrated NP well will be transferred up to the repulsive IP curve. This distri- bution, which will call a “thermal packet,” will then evolve in AE under the influence of the electrostatic forces on the solvent originating from the charged IP solute and will des- cend into the IP well. We can couch the description of the dynamics of the reaction process in chemical kinetic terms by suitable definitions ofthe thermal packets to be associated with the reactant and product. The latter clearly should re- flect the population near the IP well minimum. The former can be defined by imagining that a suitable laser dump pulse arrangement could transfer, at any time t, any residual popu- lation on the upper IP surface in the initially FC populated IP region back down to the NP well. In such a (highly ideal- ized) conception, the “reactant” population would be de- fined by the thermal packet still residing at time ton the IP curve in the neighborhood of the AE values appropriate to the thermal population of the NP well. In effect, this reac- tant population is the overlap of the thermal packet on the IP surface evolving from the initial FC packet, translated up from the thermal NP distribution, with that thermal NP distribution.40v41

We will extract these product and reactant thermal packets as follows. Rather than use the “exact” MD-genera- ted time-dependent spectrum for this purpose, we employ the time-dependent Gaussian approximation I( t,Ae), Eq. (3.13), in terms of the MD-simulated average AE (t) and time dependent square width W(t) . To define the “product” population Pip ( t), we integrate I( t,Ae) over the region in he containing most of the equilibrium product distribution Pi: ( Ae), centered at (he),, , with a width wrr = 2,/m:

[J

W - 1 A& PIP (t) = dAel( 03 ,Ae) dAeI( t,Ae);

4, I s AeG A4 = (AEhp + wIp. (3.19)

5968 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent

This gives us the formula

PIP(f) = [2erf(w,,/[2W(oo)])“*] -’

ww + @E),,%t) [2W(t)]“* 1

- wxp + (AEM I) [2W(t)]“* ’

(3.20)

where erfis the error function, S(t) is the MD shift, W(t) is the MD square width function, and W( co ) = (@rip ) - ‘.

The corresponding expression for the initial “reactant” distribution is

P,,(t) = [2erf(w,,/[2W(O)]“*)] --I

- (AE),,{l -s(t)) [2IY(f)]“* II

- wNP - (AE hp{l - S'(t)} - erf II [2W(t)]“* ’ (3.21)

where wNp = [2W(O) I”* with W(0) = (PkNP ) - ’ gauges the width of the equilibrium distribution of the reactant dis- tribution Pgp ( Ae).

Figure 9 displays our results, which illustrate several important features. First, the reactant population very rap- idly decreases to nearly zero in consequence of the underly- ing dynamics propelling the FC thermal packet away from its initial high free energy side down into the GIp well. By contrast, the product population builds up much more slow- ly, as significant population growth in the product well re- quires extensive (and slower) relaxation into the thermal IP region. Both these trends mirror the bimodal relaxation be- havior of S( t) [and AE (t) ] . The first trend illustrates in an MD context the point4’ that passage away from a high free energy point with steep slope (as in a high barrier acti- vated electron transfer23) is governed by the short time dy- namics. This aspect plays a central role in the Hynes theo- retical description of electron transfer reactions,26 and has received some experimental support in the studies of Weaver and co-workers.27 The second trend indicates that in a strongly exothermic process the time dependence of the ap-

d 1 2 3 4 5 6

t (PSI

FIG. 9. Relative populations Psp of the IP and NP as a function of time computed via Eqs. (3.20) and (3.21). Note the fast rate ofdisappearanceof the NP compared with the slower appearance of the IP.

J. Chem. Phys., Vol. 94, No. 9,1 May 1991

Downloaded 19 Mar 2007 to 128.112.35.7. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 9: The Carter Group - Solvation dynamics for an ion pair in a ......5962 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent parameters are dks = 200 K and IJ = 3.5 b; for each

pearance of the lower free energy product species should be sensitive to the longer time relaxation dynamics; this point has been stressed for exothermic electron transfer reactions by Fonseca.28

C. Langevin equation description

The excited IP state based tcf A,, (t) was seen in Sec. III A 1 to reasonably well describe the nonequilibrium sol- vation dynamics in the TDF shift S(t) at all times. It is of interest to explore here whether simple models can shed any light on the behavior of this tcf.

One possible model for A,, (t) would be a dielectric con- tinuum description. In this characterization, A,, (t) [and A,, (t) for that matter] would, in the simplest instance, be exponential in time with a decay time proportional to the solvent longitudinal relaxation time rr. .‘*15-” A test of this would involve simulation determinations of the Debye relax- ation time r,, and the static dielectric constant E for the pure solvent (since rr. = E, r, /E, and E, = 1 for this nonpolari- zable solvent). We have not followed this course here for several reasons. First, the accurate simulation determination of e is quite difficult. But more importantly, the lack of valid- ity of a dielectric continuum description for small solutes in solvents of comparable size molecules is easily anticipated (see, e.g., Ref. 17) and is already adequately established in other studies (see, e.g., Ref. 8) Instead, we explore the valid- ity of a more general, yet still simple, description: a Langevin equation (LE) approach.

We note at the outset that the LE description is bound to fail in important ways; this is not unanticipated, given the observed bimodal relaxation character of the solvation dy- namics and the very short time scale for the dominant sol- vent relaxation. Nonetheless, we persist in pursuing this ave- nue for several reasons. First, the LE approach is already moregeneral than most theoretical descriptions used in dis- cussing TDF dynamics, *‘s-2o these typically ignore any sol- vent inertia and describe the solvent as highly overdamped, i.e., diffusive. The LE in contrast includes solvent inertia.” Second, we will see that the nature of the failure of a LE description is itself instructive.

A LE for the dynamics of AE in either of the NP or IP harmonic free energy wells can be established43 with the aid of standard (two variable) projection operator techniques.35 When applied to the variable of interest, 6AE = AE - (AE )sp, one obtains the LE:”

SA&t) = - w;,SAE(t) - &,SA&t) + R(t). (3.22)

Here w& is the square frequency of the SP we11,22*23

w;p = (@AJ%~)s, b ((6AE)*),, =-’ msp

E. A. Carter and J. T. Hynes: An ion pair in a polar solvent 5969

Finally, the friction constant <sp is defined by

SSP = -df (RR(O),,;

R = 6A% t w&SAE, (3.25) which involves the tcf of the fluctuating “random” force (with the systematic harmonic force - wgp AE subtracted out). The evolution of R (t) is governed, not by conventional Hamiltonian dynamics, but rather by projection-operator modified dynamics. 35,43 We note, for later reference, that the LE (3.22) is an approximation to an exact generalized Lan- gevin equation (GLE),43 which involves the time nonlocal friction term

SAE= -o&GAE(t) - s

‘dr&(t-r)SA&) +R(t), 0

(3.26) first introduced in an approximate form for solvation dy- namics in Ref. 26 and previously used in connection with the MD simulations of activated electron transfer reactions.23 A reduction to the LE assumes that the random force correla- tions (RR ( t ) ) sp decay very rapidly compared to the time- scale of the speed SAk of the 6AE fluctuations.

The frequency and friction parameters w&, and csp can be obtained via simulation as follows. First, from Eqs. ( 3.12) and (3.23)) wgp can be determined from separate equilibri- um MD simulations of ((SAl?)*),, and ((SAE)2),p for each of the two SP cases. The results are displayed in Table I. The IP and NP frequencies are quite close to each other, largely as a result of the mild square root dependence of&., on the solute-dependent force constant k,, .45

The second quantities to be determined are the friction constants gsp. One can readily show from the LE (3.22) (and in fact also from the GLE23V4’ ) that the time areas of the normalized tcfs A,, (t), Eqs. (3.3) and (3.9), are given by the average relaxation time,

s cv s

TSP = dt A,, (t) = 2 ; 0 &P

(3.27)

such an average time is often used to characterize experi- mental solvation dynamics.‘-“ Thus, with known wzp val- ues, csp can be extracted from the MD results for the tcf’s A,, (t) and A,, (t) . The results displayed in Table I indicate the striking feature that the friction for SAE in the presence of the charged IP is about twice as large as in the presence of the NP solute. Since the solvent frequencies are comparable, this translates to an average relaxation time which is longer by about a factor of 1.5 for the IP relaxation. The solvent is definitely “slower” in the presence of the field of the

(3.23)

with k,, the force constant and msp the effective mass asso- ciated with the fluctuations in 6AE. Comparison of Eqs. (3.12) and (3.23) shows that in effect msp is defined by the fluctuations in the “velocity” of 6AE via the equipartition theorem:22*23+”

TABLE I. Simulation results for solute pair properties.

4,~ (lo-‘cm-‘) %P cps-‘) &sp (Ps-‘) 7sp (ps)

l (SA-b2),, = (Pmsp ) - ‘. (3.24)

NP 4.9 6.2 13.3 0.35 IP 8.0 1.4 29.3 0.54

J. Chem. Phys., Vol. 94, No. 9,1 May 1991 Downloaded 19 Mar 2007 to 128.112.35.7. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 10: The Carter Group - Solvation dynamics for an ion pair in a ......5962 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent parameters are dks = 200 K and IJ = 3.5 b; for each

charges-the solvent relaxation time is solute-dependent; this feature is clearly associated with the differing longer time “tails” of the tcf s, and not with the initial time behav- ior. We will return to this point.

We can examine how an LE description, even provided with the correct frequency and friction parameters, fares for the IP tcf [which gives a reasonable approximation to the nonequilibrium shift S(t) in Fig. 3 (b) 1. The story is merci- fully brief: Fig. lO( a) shows that the LE picture begins to fail right from the earliest times and, of course, completely misses the bimodal decay character of A,, (t). These same failures are apparent in the LE comparison with A,, (t) in Fig. 10(b). It is worth stressing that the LE description does not even reproduce the longer time tails for either time corre- lation function. Obviously a different description on all time scales is required.

D. Short time Gaussian approximation

A quite different tack is suggested by the Gaussian time character of both A,, (t) and ANP (t) evident in Fig. 7 at short times. If we write

ASP(r) = 1 -w;,t2/2+ ~~~~=e-“g”“2~A~p(t) (3.28)

for suitably short times, then Fig. 10 shows that in each case the short time behavior of the correlation function is quite well reproduced over a significant fraction of the decay. This striking agreement,” which also holds for the shift S(t) in Fig. IO(c), has a number of implications, which we now discuss.

First, we have seen in Fig. 10 that both the equilibrium tcfs AIp (t) and A,, (t) agree well with the nonequilibrium shift S(t) in the initial decay period. That the former should work well is at first glance a cause for surprise since, as dis- cussed in Sec. III A, the initial solvent distribution relevant for the shift S(t) [and A,, (t) ] is that appropriate for the uncharged NP and not the charged IP: the IP solute charges are not present to induce any average solvent dipolar orien- tation. Now we can see that (a) the initial dynamics for A,, (t) are clearly governed by the frequency wu, , and (b) that on is nearly equal to w .,-which governs the initial decay of A,, (t). Equivalently stated, the extensive initial dynamics of S( t), A,, ( t) and A,, (I) are all approximately governed by the solvent frequency, and this is only mildly sensitive to the initial solvent distribution, be it that for the NP or for the IP. This observation goes some way to account for the validity of a linear response treatment. [It also sug- gests that the approximate validity of linear response is by no means generally guaranteed.]

The physical interpretation of this Gaussian time behav- ior is the following. In all cases, we can picture a distribution of solvent configurations responsible for the initial AE val- ues. For very short times, the change in those configurations is governed, not by the intermolecular forces and torques, but rather by the free streaming, inertial motion of the sol- vent molecules, subject to a Maxwellian velocity distribu- tion. These inertial changes in solvent configurations change AE, and their average net effect is described by the frequency WSP.

5970 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent

1 (4

A IP 0)

0 1 2 4 5 6

lb)

0 1 2 4 5 6

1.0

0.8

0.6

0.4

0.2

0.0 0 1 2

t Al

4 5 6

FIG. 10. Langevin equation (LE; dashed lines) and short time Gaussian (A$ (t); dotted lines) approximations to Asp (t) are compared to the exact time-correlation functions (solid lines) for the (a) IP and (b) NP. (c) Com- parison of A$ (I) to the normalized TDF shift function S(t). Frequency and friction parameters used to calculate the LE and Gaussian tcf approxi- mations are: o,r = 7.4 ps-‘, oNp = 6.2 ps-‘, {n, = 29.3 ps-‘, and &, = 13.3 ps - ‘.

Second, the extensive initial Gaussian time relaxation forS( I), A,, (t), and AIp ( t) dependent upon w& highlights in a particularly stark and instructive fashion the particular inadequacies of a simple LE description. Figures 10 (a ) and 10(b) indicate that the constant friction assumption, with its requirement that the frictional forces begin to act very quick- ly [at order t 3 for Asp (t) 1, causes a very rapid departure from the simple correct exp( - &t 2/2) behavior which is itself completely independent of the frictional forces. Allowance for the finite time scale of the friction forces, i.e., the use of a time-dependent friction kernel csp (t) in a GLE,

J. Chem. Phys., Vol. 94, No. 9,1 May 1991

Downloaded 19 Mar 2007 to 128.112.35.7. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 11: The Carter Group - Solvation dynamics for an ion pair in a ......5962 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent parameters are dks = 200 K and IJ = 3.5 b; for each

will remedy this failing by recognizing that these forces do not act instantly. As shown elsewhere,42*43 it is in fact possi- ble to extract csp (t) from A,, (t) via Fourier transform techniques; the result for SIP (t) is displayed in Fig. 1 1.46 Comparison of this with Fig. 7(a) shows that the LE as- sumption that cIp (t) decays rapidly compared to the decay of AE fluctuations is simply untenable. Not only does this approximation largely obliterate the important short time Gaussian behavior in A,, ( t) and S(t) as discussed above, it can be also shown43 that in fact explicit retention of the full time dependence of cIp (t) is necessary to account for the long time tail behavior in An, (t) [and S(t) 1. [Similar re- marks apply to A,, (t). ]

Unfortunately, provision of a theory for the full time dependence of JIp (t) [or for cNP (t) ] is no mean feat; in particular, the physical origin of the bimodal time-depen- dent structure will need to be understood before the GLE provides a useful nonempirical, nonsimulation route to un- derstanding the tail behavior in the TDF shift S(t). For the present, we content ourselves with the emphasis that a GLE level of description is necessary.

E. Implications for current theories

We have already noted in Sec. III B and III C that even a Langevin equation description including solvent inertia is unable to account for our MD results; instead a Generalized Langevin Equation description is necessary. Much current theoretical effort and experimental attention have been de- voted to a different approach which incorporates some mi- croscopic solvent aspects, the dynamical MSA Theory, in- troduced by Wolynesi8 and since extended in various ways. I9 Here we comment briefly on the implications of our results for such interesting MSA theories.

It is important to first point out that the bimodal struc- tureofS( t) and Asp (t) that we have observed is unrelated to the (approximately) bimodal time structure found in var- ious MSA studies. As discussed extensively above, the initial time behavior is governed by solvent inertial effects, and sol-

vent inertia has not been included in all the MSA studies to date. (The dielectric continuum non-Debye several relaxa- tion time solvent models studied in the literature2”28 also ignore solvent inertia, as have continuum treatments includ- ing solvent translational contributions’(c’*‘7 and dielectric saturatiom4’ again, any bimodal time structure arising in these descriptions is unrelated to the present inertial effects.)

In the MSA, the solvent dynamics enters via the dynam- ic dielectric response E(O) of the solvent in the absence of the solute. At the very least, inclusion of solvent inertial effects via E(W) 17,26 seems necessary. Even then, however, current MSA models must fail to account for the observation that the longer time tails in the saturation dynamics depend on the charge distribution in the solute pair (cf. Fig. 10 and Table I). This points clearly to the need to account for solva- tion dynamics more directly in terms of solute-solvent inter- actions, rather than via a purely solvent dynamical property [E(W) ] as in the MSA approach.

IV. NONEQUlLlBRlUM SOLVATION TRAJECTORIES

Immediately after the Franck-Condon transition creat- ing the IP from the initial NP solute, the surrounding polar solvent molecules find themselves out of equilibrium with the IP and subject to the Coulomb interaction with the site charges in the IP. They must then reposition and orient themselves ultimately to achieve a new average equilibrium structure about the IP. Judging from Fig. 4, this accommo- dation is largely completed within about 3 ps, but with con- tinuing minor adjustments up to times of order 6 ps. Here we attempt to gain some insight on the microscopic details and mechanism of this nonequilibrium solvation process.

The long-time equilibrium spatial and orientational pat- terns of the solvent about the IP have been previously deter- mined in an MD simulation study by Ciccotti et al.*’ The results are shown schematically in Fig. 12. The main struc- tural feature is a cylindrical sheath of four solvent molecules whose dipoles are oriented antiparallel to the IP axis, at a radius of approximately 4 A on the average. There is also a much weaker solvation pattern comprising two solvent mol- ecules each flanking the IP axis and on average parallel to it.

200.0

<,,( 0 ‘OO.0

0.0

\

E. A. Carter and J. T. Hynes: An ion pair in a polar solvent 5971

w-w

e-m >

0.0 0.2 0.4 0.6 0.6 1.0

t (PS)

FIG. II. Time-dependent friction, <,,, (t), in the presence of the IP. Units FIG. 12. Schematic illustration of the equilibrium solvation patterns around areps-‘. the IP.

J. Chem. Phys., Vol. 94, No. 9,l May 1991 Downloaded 19 Mar 2007 to 128.112.35.7. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 12: The Carter Group - Solvation dynamics for an ion pair in a ......5962 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent parameters are dks = 200 K and IJ = 3.5 b; for each

5972 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent

0 1 2 3 4 5 6 t, PS

0 1 2 3 4 5 6

t, P=

0 1 2 3 4 5 s

t, PS 0 1 2 3 4 6 6

t, PS

0.6 0.4 I 0.4

0.2 -0.0 -0.2 -0.4 -0.6 -0.6 -1 .o

z 0.2 = a -0.0 b ::

-02 -0.4

d i i i s t,PS

0 i i 3 i 5

ZPS

0.6 - .- 0.6 0.6

4 -0.6 -I -0.6 1 0 1 2 3 4 5 6

SPS

(4

-1.0-j . ..(. ‘.,...I. ..I. .-,..r 0 1 2 3 4 5

t, P=

FIG. 13. Representative trajectories ofthe first solvent shell molecules subsequent to the excitation NP + IP. The characteristic quantities (z and the parallel and perpendicular components ofp) are described in the text. The panels refer, in sequence, to molecules A-E.

J. Chem. Phys., Vol. 94, No. 9,1 May 1991

Downloaded 19 Mar 2007 to 128.112.35.7. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 13: The Carter Group - Solvation dynamics for an ion pair in a ......5962 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent parameters are dks = 200 K and IJ = 3.5 b; for each

E. A. Carter and J. T. Hynes: An ion pair in a polar solvent

0 1 2 3 4 5 6

6.0

4.5

4.0

3.6

3 3.0

2.5

2.0

1 .s

1 .o

s_ ‘i ;; 9.

f

E P t n

t, PS

P % 4$$@?F qf&! &c?” O % -88

, . . , . . . , . . . . . . . . 0 1 2 3 4 5

t, PS

4 6

0.6 0.6 OJ- 02- %

-0.0 - -0.2 -

0 i i i ii SPS

-l.o:...,...,...,...,...,...~ 0 1 2 3 4 5 6

f PS w

t, PS

0 1 2 3 4 5

4 PS

0.6 ,0.6

02 i

1 -0.0 A ;j,+*{*yw 1 2 3 4 5 6

I 6

SPS

0 1 2 3 4 5 6

t,PS 0)

FIG. 13. (continued)

J. Chem. Phys., Vol. 94, No. 9,1 May 1991 Downloaded 19 Mar 2007 to 128.112.35.7. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 14: The Carter Group - Solvation dynamics for an ion pair in a ......5962 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent parameters are dks = 200 K and IJ = 3.5 b; for each

2 0 1

N -I -2 -3 -4 -5?...,...1.-.I...I..-‘...’

0 1 2 3 4 5 6

t, P=

0 1 2 3 4 5 6

t, P=

Z = 2 x

2 z u s E x

02 Aa -A

-0.0 A A a I

-0.2 -0.4 A

-0.6 -0.6 3 -1 .o 31

0 I 2 3 4 5 6

t, P=

0 1

t, P=

(E)

FIG. 13 (continued).

E. A. Carter and J. T. Hynes: An ion pair in a polar solvent

In the absence of the IP electrical charges, there will be no net antiparallel solvent orientation in any cylindrical sheath about the NP, and thus no such net polarization ex- ists initially after the IP has been created in the Franck- Condon transition. In order to examine the ensuing solva- tion dynamics, we construct a cylindrical coordinate system, whose z axis is anchored to the IP axis and whose zero is located at the midpoint between the two IP members. The spatial position of the center of mass of a solvent molecule is then located by the spatial cylindrical coordinates z (the cyl- inder height) andp (the cylinder radius) and an azimuthal angle 4 that is referenced to some convenient value (see be- low). The orientation of a solvent molecule can be specified in terms relevant to the solvation as follows. A unit vector P, is imagined to be attached to the dipolar axis of a solvent molecule, pointing from the negatively charged site to the positively charged site. A corresponding unit vector 2,, is rigidly attached to the IP cylindrical axis pointing from the negative site to the positive site. The signed parallel and per- pendicular components of the solvent molecule orientation with respect to the IP axis are then defined by

parallel = Zs (t) *Pi, (t); perpendicular = P, ( t ) -&, ( t), (4.1)

where &, ( t) is a unit vector perpendicular to P,, ( t), defined to have a positive projection along the intermolecular axis from the center of mass of the IP to that of the solvent mole- cule.

With the above coordinate system, we have followed the fate of (typically) six solvent molecules closest to the newly created IP in a number of trajectories subsequent to the Franck-Condon transition. We have found that there is con- siderable variation in the detailed dynamics of these mole- cules between trajectories, but that certain tendencies emerge, involve both reorientational and translational mo- tion. We now discuss two sets of these in detail, but stress that these trajectories should more properly be regarded as representative (and instructive), rather than as necessarily completely typical.

I*“I--*I”-I--. 2 3 4 5 6

Figures 13 (a)-1 3 (c) illustrate the histories of three sol- vent molecules, labeled A, B and C, which are initially favor- ably oriented: at t = 0 they are approximately oriented anti- parallel to the IP axis. On a time scale of 5 2 ps, there is a tendency to remain antiparallel or to become more so. On this same time scale (and longer), the solvent molecular centers of mass stay close to the IP bond midpoint z = 0; an initial motion towards z = 0 is especially noticeable for B and C. Molecules A and B are typically located nearp = 4 A over the 6 ps period examined; in contrast, after tz3 ps, molecule C wanders to the periphery-it drifts to somewhat largerp away from the z = 0 midpoint, and loses its antipar- allel character; it has exited the inner solvation shell. Solvent molecules D and E [Figs. 13 (d) and (e) ] have a different behavior than A-C; they illustrate IP-induced antiparallel alignment from initially unfavorable orientations. This oc- curs fairly rapidly in the case of D, but is rather delayed for E, whose slower alignment coincides with its center of mass positioning near the IP midpoint z = 0. The interference between all these solvent molecules attempting to accommo-

J. Chem. Phys., Vol. 94, No. 9,1 May 1991

Downloaded 19 Mar 2007 to 128.112.35.7. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 15: The Carter Group - Solvation dynamics for an ion pair in a ......5962 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent parameters are dks = 200 K and IJ = 3.5 b; for each

Azimuthal Anale of Solvent Molecules Relative to Molecule D

3 120 f! EI E s? So l B 3 D A iZ 84 l c

0

30

t, PS

FIG. 14. Dynamics in me azimuthal angle for solvent molecules followed in Fig. 13, referenced to molecule D.

date to the IP can result in strong disturbances of previously favorable positionings, e.g., molecule C past t- 2 ps. Finally, Fig. 14 illustrates the azimuthal solvation pattern referenced to molecule D. Within rather large fluctuations, there is a basic pattern of two solvent molecules (A and E) located near 90” (i.e., * 90”) with respect to D, while the fourth (B) is near 180” (recall that molecule C drifts out of the inner solvation shell).

An impression of the complexity of the solvation trajec- tories and the frequent mutual interference between them is given in another set of trajectories shown in Fig. 15. Solvent molecule A is initially positioned and oriented favorably with respect to the IP axis but is subsequently (at - 1 ps) severely perturbed. Molecules B and C rapidly become aligned approximately antiparallel to the IP axis starting from unfavorable orientations (in fact B flips!). Molecule D at first succeeds in the induced alignment, but subsequently drifts away from it, eventually exiting the shell at -6 ps. Molecules E and F, which are initially flanked away from the IP m idpoint z = 0 meander toward partial alignment but even wander out of and back into the primary solvation shell (Eat -3psandFat -2psand4ps).

From these two sets of trajectories (and others not shown), we have the impression that initially favorable sol- vent molecule positionings and alignments tend to strengthen and to be preserved for at least a few picoseconds. Initially unfavorable solvent molecule positionings and alignments can be quickly converted to favorable ones. However, the competition within and between these classes can create considerable complexity, thwarting the attain- ment of favorable solvation by some molecules and disrupt- ing the pre-existing favorable solvation ofothers. We think it likely that the initial broadening in the spectrum (Figs. 5 and 6) is associated with the rather large range of solvent mole- cule trajectories induced by the sudden exertion of the Cou- lomb forces and torques on them by the newly created IP. Actually, in view of the complexity of the trajectories,it must be counted as remarkable that the spectral shift S(t) can be accounted for so well overall by the linear response approxi- mation A,, (t) (Fig. 7) and that the simple Gaussian ap-

proximations As”p (t) (Fig. 10) are so accurate at short times.

V. CONCLUDING REMARKS

The present simulation and theoretical studies of time dependent fluorescence have shed some light on the approxi- mate validity of an appropriate linear response treatment, on the importance for the solvation dynamics of the inertial Gaussian time behavior, and on the necessity for a general- ized Langevin equation level of description. The impact of these dynamics on a model photochemical reaction high- lights the importance of short time solvation dynamics in the kinetic evolution. In addition, the nonlinear aspects of initial spectral broadening and ultimate spectral broadening have been examined.

The results of the present study suggest a number of avenues for further exploration. The critical role of the iner- tial Gaussian time behavior of the solvation dynamics, for example, indicates that such a component should be searched for in femtosecond experiments, and be incorporat- ed in future analytical treatments.

An important role should be played by short time Gaus- sian dynamics in the early time solvation behavior for other solvents, including, e.g., water,* and in other environments, e.g., proteins.48 (These have now also been found for aceto- nitrile13 and methanol14). A novel feature in aqueous and other hydrogen-bonded solutions should be a relatively early and marked departure from Gaussian behavior associated with, e.g., the high frequency librational dynamics which can contribute at short times. Indeed, Fonseca and La- danyi14 have very recently demonstrated this for simulated methanol. They have also found marked departure from lin- ear response theory expectations, and been able to explain this in terms of the methanol dynamics.

The governance of the very extensive initial Gaussian relaxation in the TDF shift S(t) by the solvent frequency *IP =*NP suggests that considerable progress m ight be made in the description of general TDF systems by construc- tion of theories for wgp. This is an equilibrium time-indepen- dent average quantity [cf. Eq. (3.20) 1, and the application of modern equilibrium statistical mechanical methods for solutes in polar solvents49 to its calculation could prove use- ful in providing estimates for more general solute and sol- vent systems. One can imagine, for example, that such a the- ory could shed light on the relative importance of solvent translational, librational and reorientational free streaming motion on the initial Gaussian decay, and on the relative importance of solvent shell molecules for this decay. It m ight also reveal interesting initial relaxation dynamics for a solute in a m ixture of solvents of differing polarity,arising from “solvent-sorting” effects. 5o Much remains to be learned here.

In addition, further investigation of the probe depen- dence of the solvation dynamics and of the overall and tran- sient spectral narrowing and broadening could prove to be useful experimental and theoretical indicators for nonlinear effects. The possibility that these m ight differ qualitatively in dipolar aprotic versus hydroxylic solvents could provide ad- ditional illumination on these solvation nonlinearities.

E. A. Carter and J. T. Hynes: An ion pair in a polar solvent 5975

J. Chem. Phys., Vol. 94, No. 9,1 May 1991 Downloaded 19 Mar 2007 to 128.112.35.7. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 16: The Carter Group - Solvation dynamics for an ion pair in a ......5962 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent parameters are dks = 200 K and IJ = 3.5 b; for each

5976 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent

-- I 4 3

200 i 1

!a@@@

E -1

%

T?Y

-2 2 & #

-3 -4 “1 -s!...,...,...l...I..-I...~

0 1 2 3 4 5 6

G PS 0 1 2 3 4 5 6

t, P=

2.0

1.6 I

0 1 2 3 4 6 6

t, P*

0 1 2 3 4 5 6

t, PS

0.3 0.6 A 0.4 A

A

0 i i s ts w

= -04 A

= OS Y

-0.6

0 1 2 3 4 5 6

%P

(4 @I

0 i i i i 6

1, I#

0.6 - 0.6: 4

Oe4 ’ A A A 02- A f -0.0 :k :h

-02-

‘=‘r’\jA4A

-0.4 - A A r4

-0.6 - -0.3- -1.0 .-.~.-.1--.1...‘.--‘.--

0 1 2 3 4 5 6

fP

FIG. 15. A second set of representative trajectories of the first solvent shell molecules subsequent to the excitation NP -+ IP. The description of various coordinates is as in Fig. 13. The panels refer, in sequence, to molecules A-F.

J. Chem. Phys., Vol. 94, No. 9,1 May 1991

Downloaded 19 Mar 2007 to 128.112.35.7. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 17: The Carter Group - Solvation dynamics for an ion pair in a ......5962 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent parameters are dks = 200 K and IJ = 3.5 b; for each

E. A. Carter and J. T. Hynes: An ion pair in a polar solvent 5977

0 1 2 3 4 6 6

t, PS

0 1 2 3 4 5 6

t PS

5 7s r’ P

..- 0.3 0.6 0.4 0.2

0 1 2 3 4 5 6

t, w

A 4

-0.6 -0.6 I -1.0 ]I

0 1 2 3 4 5 6

% v

(C)

-5t...,...,.-.I.-.I-*-I-.~ 0 1 2 3 4 6

t, PS

0 1 2 3 4 6

t, PS

-0.0 -0.2 Aa -0.4

&,“.a A.. UP -

AA -0.6 . %h

-l.O!. ..I. .., .-- I... I.. .I.. . 0 1 2 3 4 5

t, P

..- 0.6 0.6 4

0.4 AA 02-

-0.0- A A -0.2 - -0.4 - -0.6 - -0.6 - -1.0-I . ..I -.- I...,.. -,-.-I-.-

0 1 2 3 4 5

t, PS

0)

FIG. IS. lcontinuecf.)

J. Chem. Phys., Vol. 94, No. 9,1 May 1991 Downloaded 19 Mar 2007 to 128.112.35.7. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 18: The Carter Group - Solvation dynamics for an ion pair in a ......5962 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent parameters are dks = 200 K and IJ = 3.5 b; for each

5978 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent

ia? q rn& cp %I fl$!!- 0

I---l---,*--l---l--C 1 2 3 4 5

t, PS

6.0 , I 4.5

b 0

4.0 rcI “0” 0 o@@?ip 0% ,"%

z 3.5 O %I 3.0 I 0

0 2.5 -i 2.0 1.5

1 .o 1 =o 0 1 2 3 4 5 6

t, PS

02- -0.0 - A -02. -0.4 - -0.6 - -0.6 - -1.0 .-- ,...,.,.(...(...,...

0 1 2 3 4 5 6

t, P

Ir A A

AAAr A Ir-

0 1 2 3 4 6 6

t, P

(El

4 3

1 2

g l N -1 ~ 2&o:

-2 $i#!!ip

-3 a 5!? 0

0

-4 -5: . . . . . . . . . ..I...... .I..

0 1 2 3 4 5

t PS

0 1 2 3 4 6 6 t PS

I 0.6- AA 0.6 - A+

A "g +i@

A = 2 -o.o- 4a 116 %a E -02- A

-0.4 - A -0.6 - A -0.6 - -1.0,. . . , . . . , . . . , . . . , . . . , . . -1

0 1 2 3 4 6 6

1, yu,

0.6 $ 0.4

z 02 u r -0.0

*. . , . . *. . . . . . . . . . . A A 1

0 1 i i i i

t, ps

F)

FIG. 15. (continued.)

J. Chem. Phys., Vol. 94, No. 9,i May 1991

Downloaded 19 Mar 2007 to 128.112.35.7. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 19: The Carter Group - Solvation dynamics for an ion pair in a ......5962 E. A. Carter and J. T. Hynes: An ion pair in a polar solvent parameters are dks = 200 K and IJ = 3.5 b; for each

E. A. Carter and J. T. Hynes: An ion pair in a polar solvent 5979

ACKNOWLEDGMENTS

This work was supported in part by NSF Grants No. CHE 84.19830 and No. CHE 88.07852.

‘Some recent reviews which address some aspects of these issues include: (a) P. F. Barbaraand W. Jarzeba, Adv. Photochem. 15,l (1990); (b) M. Maroncelli, J. MacInnis, and G. R. Fleming, Science 243,1674 (1989); (c) B. Bagchi, Annu. Rev. Phys. Chem. 40,115 (1989); (d) J. D. Simon, Act. Chem. Ra. 21, 128 (1988); (e) E. M. Kosower and D. Huppert, Annu. Rev. Phys. Chem. 37, 127 (1986).

’ (a) M. Maroncelli and G. R. Fleming, J. Chem. Phys. 86,622l ( 1987); (b) E. W. Castner, Jr., M. Maroncelli, and G. R. Fleming ibid. 86, 1090 (1987).

’ (a) S.-G. Su and J. D. Simon, J. Phys. Chem. Q&2693 (1987); (b) Chem. Phys. Lett. 158,423 (1989).

’ (a) M. A. Kahlow, T. J. Kang, and P. F. Barbara, J. Chem. Phys. 86.3183 (1987); (b) M. A. Kahlow, W. Jarzeba, T. J. Kang, and P. F. Barbara, ibid. 90, 151 (1988).

‘G. van der Zwan and J. T. Hynes, J. Chem. Phys. 76, 2993 (1982); D. Zichi and J. T. Hynes, ibid. 88,2513 (1988); S. Leeand J. T. Hynes, ibid. 88,6863 ( 1988); B. J. Gertner, K. R. Wilson, and J. T. Hynes, ibid. 90, 3537 ( 1989); B. J. Gertner, R. Whitnell, K. R. Wilson, and J. T. Hynes, J. Am. Chem. Sot. 113,74 (1991).

‘For recent reviews, see G. E. McManis and M. J. Weaver, Act. Chem. Res. ( 1990) and Refs. 1 (a) and 1 (d).

‘D. Borgis, S. Lee, and J. T. Hynes, Chem. Phys. Lett. 162, 19 (1989); D. Borgis and J. T. Hynes, J. Chem. Phys. (in press).

z M. Maroncelli and G. R. Fleming, J. Chem. Phys. 89,5044 ( 1988). ‘M. Rao and B. J. Beme, J. Phys. Chem. 85,1498 (1981); 0. A. Karim, A.

D. J. Haymet, M. HJ. Banet, and J. D. Simon, ibid. 92, 3391 (1988). “J. S. Bader and D. Chandler, Chem. Phys. Lett. 157,501 ( 1989). “J. T. Hynes, E. A. Carter, G. Ciccotti, H. J. Kim, D. A. Zichi, M. Fer-

rario, and R. Kapral, in Perspectives in Photosynthesis, edited by J. Jortner and B. Pullman (Kluwer, Dordrecht, 1990), p. 133.

” R. M. Levy, D. B. Kitchen, J. T. Blair, and K. Krogh-Jespersen, J. Phys. Chem. Q4,4470 (1990); J. T. Blair, K. Krogh-Jespersen, and R. M. Levy, ibid. 111, 6948 (1989).

I’M. Maroncelli, J. Chem. Phys. (in press). “T. Fonseca and B. M. Ladanyi, J. Phys. Chem. (in press). “N. G. Bakshiev, Opt. Spectrosc. (USSR) 16,446 (1964); Y. T. Mazur-

enko, ibid. 36,283 (1974); 48,388 (1980). 16B. Bagchi, D. W. Oxtoby, and G. R. Fleming, Chem. Phys. 86, 257

(1984). “G. van der Zwan and J. T. Hynes, J. Phys. Chem. 89,418l (1985). ‘*P. G. Wolynes, J. Chem. Phys. 86, 5133 (1987). I9 I. Rips, J. Klafter, and J. Jortner, J. Chem. Phys. 88, 3246 (1988); 89,

4288 (1988); A. L. Nichols III and D. F. Calef, ibid. 89,3783 (1988); Y. Zhou, H. L. Friedman, and G. Stell, ibid. 91,4885 (1989); B. Bagchi and A. Chandra, ibid. QO,7338 (1989); J. Phys. Chem. 93,6996 (1989).

*‘Y. J. Yan and S. Mukamel, J. Chem. Phys. 86.6085 (1987); V. Friedrich and D. Kivelson, ibid. 87, 6425 (1987); N. Agmon, J. Phys. Chem. 94, 2959 (1990); R. A. Marcus, ibid. 93.3078 ( 1989).

*’ G Ciccotti, M. Ferrario, J. T. Hynes, and R. Kapral, Chem. Phys. 129, 241 (1989); J. Chem. Phys. 93,7137 (1990).

**E, A. Carter and J. T. Hynes, J. Phys. Chem. 93,2184 (1989). 2’D. A. Zichi. G. Ciccotti, J. T. Hvnes. and M. Ferrario, J. Phvs. Chem. 93,

6261 (1989). - *‘A. Warshel, J. Phys. Chem. 86, 2218 (1982); A. Warshel and J. K.

Hwang, J. Chem. Phys. 84,493s (1986); J. K. Hwang and A. Warshel, J. Am. Chem. Sot. 109,715 (1987); 3. W. Halley and J. Hautman, Phys. Rev. B 38, 11704 ( 1988); R. A. Kuharski, J. S. Bader, D. Chandler, M. Sprik, M. Klein, and R. Impey, J. Chem. Phys. 89,3248 (1988).

25Al~o not included here are dynamic effects associated with quantum ex- change terms associated with a mixing of states in the charge distribution of the excited state [H. J. Kim and J. T. Hynes, J. Chem. Phys. 93, 5211 ( 1990) ] or any solvent dependence of the transition moment [H. J. Kim and J. T. Hynes (to be submitted) 1.

26J. T. Hynes, J. Phys. Chem. 90,370l ( 1986). 27G. E. McManis and M. J. Weaver, J. Chem. Phys. 90,912 (1989); M. J.

Weaver, G. E. McManis, W. Jarzeba, and P. F. Barbara, J. Phys. Chem. 94,1715 (1990).

*‘T. Fonseca, J. Chem. Phys. 91,2869 ( 1989). 29 (a) P. F. Barbara, T. J. Kang, W. Jarzeba, and T. Fonseca, in Perspectives

in Photosynthesis, edited by J. Jortner and B. Pullman (Kluwer, Dor- drecht, 1990), p. 273; (b) J. Kang, W. Jarzeba, P. F. Barbara, and T. Fonseca, Chem. Phys. 149,81 (1990).

3oS. Nose, J. Chem. Phys. 81,511 (1984). 3’ L. Verlet, Phys. Rev. 159,98 (1967). 32 See, for example, J. P. Hansen, in Molecular Dynamics Simulation ofSta-

tistical Mechanical Systems, edited by G. Ciccotti and W. G. Hoover (North-Holland, Amsterdam, 1986).

“G. Ciccotti and J. P. Ryckaert, Comput. Phys. Rep. 4,345 (1986). 34 Within 0.3 ps after the charges appear in the IP, the average kinetic energy

per molecule (K) of the first shell of solvent molecules rises from the thermal value (K ) ,,, = 434 cm - ’ at 250 K, to 800 cm- ’ . It then rapidly declines, returning to the initial thermal average by about 1 ps. The peak in this average occurs at the same time as does the peak in the width func- tion Eq. (2.12) and shown in Fig. 6 below.

35 See, for example, J. T. Hynes and J. M. Deutch, in Physical Chemistry an Advanced Treatise, edited by H. Eyring, W. Jost,and D. Henderson, Vol. 1 lb (Academic, New York, 1975), p. 153.

36T Kakitani and N. Mataga, Chem. Phys. 93,381 (1985); J. Phys. Chem. f&8,4752 (1985); 99,993 1986); Q&6277 (1987).

“AS pointed out in Ref. 22, the situation in water solvent is apparently different.

‘*This is subject to the neglect of certain small anharmonic effects (Ref. 22).

39 See, for example, M. D. Newton and N. Sutin, Annu. Rev. Phys. Chem. 35,437 (1984). See also Ref. 10.

*There is a certain analogy here to pure quantum wave packet perspectives in isolated molecule spectroscopy. See, for example, E. J. Heller, Act. Chem. Res. 14,369 (1981); S. Mukamel, J. Phys. Chem. 88,832 (1984).

4’See also Ref. 29 in this connection. 42 R. F. Grote and J. T. Hynes, J. Chem. Phys. 73,2715 ( 1980). See also J.

T. Hynes, in The Theory of Chemical Reaction Dynamics, Vol. IV, edited M. Baer (CRC, Boca Raton, FL, 1985), p. 171.

43 D A. Zichi, H. J. Kim, E. A. Carter, and J. T. Hynes (to be submited). 44 D: F. Calef and P. G. Wolynes, J. Phys. Chem. 87,3387( 1983). 45 Due to inadequacies in the shortness ofthe time step used in the determin-

ation of velocities in Ref. 22, the solvent masses and frequencies presented there need to be revised to the values quoted in the present work.

46The result for tNP( t) is given in Ref. 23. 47 E. W. Castner, Jr., M. Maroncelh, and G. R. Fleming, J. Chem. Phys. 86,

1090 (1987). 48G. McLendon, Act. Chem. Res. 21, 160 (1988); A. Gafni, R. P. De-

Toma, R. E. Manrow, and L. Brand, Biophys. 3.17,155 (1977); M. Kar- plus (personal communication); M. Nonella and K. Shulten (preprint).

49 See, for example, S. E. Houston, P. J. Rossky, and D. Zichi, J. Am. Chem. Sot. 111, 5680 (1989).

“‘B. M. Ladanyi and J. T. Hynes (to be submitted).

J. Chem. Phys., Vol. 94, No. 9,1 May 1991 Downloaded 19 Mar 2007 to 128.112.35.7. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp