THE JOURNAL OF CHEMICAL PHYSICS 131, 094109 共2009兲

Dynamics of electron transfer reactions in the presence of mode mixing: Comparison of a generalized master equation approach with the numerically exact simulation Kirill A. Velizhanina兲 and Haobin Wangb兲 Department of Chemistry and Biochemistry, New Mexico State University, MSC 3C, Las Cruces, New Mexico 88003, USA

共Received 8 July 2009; accepted 8 August 2009; published online 4 September 2009兲 A generalized master equation approach is developed to describe electron transfer 共ET兲 dynamics in the presence of mode mixing. Results from this approximate approach are compared to the numerically exact simulations using the multilayer multiconfiguration time-dependent Hartree theory. The generalized master equation approach is found to work well for nonadiabatic resonant ET. Depending on the specific situation, it is found that the introduction of mode mixing may either increase or decrease the ET time scale. The master equation fails in the adiabatic ET regime, where the introduction of mode mixing may lead to electron trapping. From both the approximate theory and the numerically exact simulation it is shown how neglecting mode mixing in practical calculations may lead to inaccurate predictions of the ET dynamics. © 2009 American Institute of Physics. 关doi:10.1063/1.3213435兴 I. INTRODUCTION

Electron transfer 共ET兲 reactions in the condensed phase have received much attention due to their importance in chemistry and biology.1 The spin-boson Hamiltonian,2–4 where two electronic states are linearly coupled to a harmonic bath, is one of the most widely used models for studying the donor-acceptor ET processes. Extensive theoretical work has been devoted to solving the dynamics of this model, both approximately and numerically exactly.2,5–17 These studies serve as the basis for interpreting timeresolved experiments on ET reactions as well as other related processes. To model the effect of vibrations on the ET dynamics the nuclear environment is usually divided into two parts: The inner sphere and the outer sphere. The former includes the strongly coupled intramolecular modes, whereas the latter represents the contribution from the further, more weakly coupled 共in an average sense兲 degrees of freedom. In the standard spin-boson Hamiltonian both parts are modeled by a set of harmonic oscillators with linear electronic-nuclear couplings. The difference is that for outer sphere environment there is an infinite number of bath modes, representing a condensed phase continuum. The usual justification why such a harmonic representation is appropriate for modeling the outer sphere environment comes from the linear response theory in statistical mechanics. There is a good indication that for ET reactions in some condensed phase media, e.g., polar solvents, the linear response mapping is a good approximation for such a purpose. This, however, should not be confused with the instantaneous normal mode approximation because the mapped Present address: Center for Nonlinear Studies 共CNLS兲 and Theoretical Division, Los Alamos National Laboratory, Los Alamos, New Mexico 87545. b兲 Electronic mail: [email protected]. a兲

0021-9606/2009/131共9兲/094109/9/$25.00

bath modes are fictitious and the only equivalent property to the original nuclear environment is the force-force correlation function18–20 at a certain temperature 共thus the mapping is temperature dependent兲. The actual nuclear potential energy function could be quite anharmonic,18–20 which renders a local harmonic approximation invalid. Things are different for the inner sphere degrees of freedom. They are not in the condensed phase and, in general, one cannot apply the linear response mapping. The harmonic oscillator/linear electronic-nuclear coupling model thus comes from the approximation to the potential energy surface. At a stationary point of one electronic state 共usually the ground state兲 the nuclear potential energy is expanded to the second order. A normal mode analysis is used to identify the set of independent harmonic oscillators. Then it is assumed that the nuclear potential function for another 共excited兲 electronic state is described by the same set of harmonic oscillators, only shifted linearly according to the gradient at the vertical excitation point. Such an approximation may be very crude for describing intramolecular modes and, in situations where the dynamics of the ET reaction is dominated by those modes, may lead to serious errors. There are two immediate corrections to this harmonic potential/linear coupling approximation, both in the dynamical studies and the electronic structure theory calculations to obtain the potential surfaces. First, one may take into account the diagonal anharmonicity by making a higher-order expansion of the potential energy along each normal mode of each electronic state. Second, within the harmonic expansion one may take into account the rotation of normal mode axes, which is often referred to as the Duschinsky rotation effect or mode mixing.21 While both of these effects are important to model the intramolecular modes for realistic ET processes, for the purpose of obtaining a clear physical picture it is

131, 094109-1

© 2009 American Institute of Physics

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

094109-2

J. Chem. Phys. 131, 094109 共2009兲

K. A. Velizhanin and H. Wang

advantageous to study each effect separately. One may then combine these two corrections to have a more comprehensive model. The effect of including the diagonal anharmonic corrections has been reported in the previous publication20 using the numerically exact multilayer multiconfiguration timedependent Hartree 共ML-MCTDH兲 simulation method. Theoretical work has also been done to include mode-mixing effect, both numerically exactly22 and approximately.23–31 However, in this case the numerically exact methodology 共the original single layer MCTDH兲 is restricted to relatively small systems. In turn, the approximate theories are mainly based on the assumption of a rate process in the limit of weak coupling between the electronic states, i.e., Fermi’s golden rule. The purpose of this paper is to extend the generalized master equation approach to study the ultrafast ET reactions with the presence of mode mixing, and treat processes that cannot be characterized by single rate constants. For comparison we will use the recently developed ML-MCTDH theory32 to investigate the validity of our approximate theory in different physical regimes. This is important because it has been shown that in certain situations mode mixing can strongly affect ET rate constants or time scales.27,30 Experimental observations of spectroscopic signatures for Duschinsky rotations have also been reported.33,34 Our theory takes ingredients from and is complementary to the previous work30 in that it provides time-dependent electronic dynamics information. It is also important to note that the results obtained from our approximate theory have been compared to those from the numerically exact ML-MCTDH method. This provides a rigorous test on the range of the applicability of the theory. To fully explore this range and to study possible implications of the Duschinsky rotation on the ET reactions our study spans a certain physical parameter space ranging from nonadiabatic to adiabatic regimes. The application of the thus developed approximate theory to realistic systems will be reported in our future work. The remaining part of this paper is organized as follows. In Sec. II we describe the ET model and the corresponding notations. Section III discusses the generalized master equation approach to studying the time-dependent ET dynamics in the presence of mode mixing, and Sec. IV summarizes the ML-MCTDH theory for numerically exact simulations of the same processes. Section V presents numerical results obtained from the quantum master equation approach and their comparison with the ML-MCTDH simulations. Section VI concludes.

II. SPIN-BOSON MODEL WITH MODE MIXING

With the Duschinsky rotation/mode-mixing effect included the spin-boson Hamiltonian is generalized to 共2.1a兲

H = H0 + V,

H0 =



k=g,d,a

兩k典Hk具k兩;

V = ⌬共兩a典具d兩 + 兩d典具a兩兲,

共2.1b兲

˜ 2Q . Hk = ⑀k + 21 pTp + 21 QTk ⍀ k k

共2.1c兲

In this expression Hg, Hd, and Ha define the vibrational diabatic potential energy surfaces 共DPESs兲 for the ground, the donor, and the acceptor electronic states, respectively. The strength of the nonadiabatic coupling between the donor and the acceptor states is given by ⌬. The electronic energy for each state is denoted by ⑀k, e.g., the energy gap between the donor and acceptor DPES minima is ⑀ad = ⑀a − ⑀d. Diagonal ˜ , k = g , d , a, contain frequencies of the vibramatrices ⍀ k tional normal modes. It should be noted that in Eq. 共2.1c兲 a different set of normal mode 共mass-weighted兲 nuclear coordinates Qk is used for each electronic state k. The coordinates of the donor state Qd and the acceptor state Qa are related to those of the ground state Qg by Qg = ˜SkQk + Dk ;

k = d,a,

共2.2兲

where ˜Sk is the Duschinsky rotation matrix21 and Dk represents the coordinate displacement vector between the DPES minima of the ground and kth electronic state. The convention of our notation is that the column vectors are denoted by boldface, row vectors by an additional superscript “T” 共transpose兲, and the matrices are denoted by boldface with a tilde. The nonadiabatically coupled part of the Hamiltonian in Eq. 共2.1兲 reduces to the original spin-boson model when ˜Sa ˜ =⍀ ˜ , i.e., when there is neither Duschinsky ro= ˜Sd and ⍀ a d tation nor frequency change of the normal mode coordinates upon donor-to-acceptor ET. In the following we refer to the ˜ ⫽⍀ ˜ as “frequency shift” and the case ˜S ⫽ ˜S as case ⍀ a d a d mode-mixing or Duschinsky rotation. In the absence of both Duschinsky rotation and frequency shift, the reorganization energy ␭ for the transition between donor and acceptor electronic states is given by ˜ 2共D − D 兲, ␭ = 21 共Dd − Da兲T⍀ d a

共2.3兲

˜ =⍀ ˜ =⍀ ˜ . The reorganization energy is a classical where ⍀ a d concept in Marcus theory35,36 and is a useful measure of the coupling strength between the electronic states and the nuclear degrees of freedom. However, Eq. 共2.3兲 does not apply if mode mixing and/or frequency shift is introduced. For mere referencing purpose we denote the “reorganization energy” in this paper as ˜ 2共D − D 兲. ␭ = 21 共Dd − Da兲T⍀ d a d

共2.4兲

It still reflects the electronic-nuclear coupling strength but is no longer relevant to the real physical concept. The primary observables of interest are populations of the donor and acceptor electronic states Pk共t兲 = Tr关兩k典具k兩␳共t兲兴;

k = d,a,

共2.5兲

where ␳共t兲 is the time-dependent density matrix 共throughout the paper we use atomic units where ប = 1兲

␳共t兲 = e−iHt␳eiHt , which is the solution of the Liouville equation

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

共2.6兲

094109-3

J. Chem. Phys. 131, 094109 共2009兲

Electron transfer with mode mixing

FIG. 1. The ground, donor, and acceptor potential energy surfaces for two ˜ 共1 , 2 ; 15°兲 vibrational normal modes. The Duschinsky matrices are ˜Sd = G ˜ ˜ and Sa = G共1 , 2 ; −45°兲 for the donor and acceptor state, respectively.

i

⳵ ␳共t兲 = 关H, ␳共t兲兴. ⳵t

共2.7兲

共2.8兲

where the equilibrium vibrational density matrix ␳g 共␳d , ␳a兲 for the ground 共donor, acceptor兲 state DPES is introduced as

␳k =

e −␤Hk ; Tr共e−␤Hk兲

1 T 1 T˜ vert T Hvert k 共Qg兲 = ⑀k + 2 p p + ck Qg + 2 Qg KkQg ;

k = d,a, 共2.11兲

The initial density matrix is assumed to be in the factorized form

␳ ⬅ ␳共0兲 = 兩d典具d兩␳g ,

cal calculations it is more convenient to define the DPESs with respect to the vertical excitation point, i.e., the minimum energy point of the ground electronic state located at Qg = 0. This is particularly relevant in the context of photoinduced ET processes upon vertical excitation. It is also useful in practice for obtaining parameters of the Hamiltonian 共2.1兲 from electronic structure theory calculations, by paying attention to the important regions. In the limit that neither mode mixing nor frequency shift is present, the energy gradient 共force兲 of a mode at an electronic state defines the linear electronic-nuclear coupling in the original spin-boson model. Thus, the DPES representation at the vertical excitation reference point is

where the vector ck denotes the gradient of the kth DPES at the vertical excitation point. The ground state DPES remains ˜ is defined as the same as in Eq. 共2.1兲. The Hessian K k ˜ = ˜S ⍀ ˜ 2˜ T K k k k Sk .

共2.12兲

The two representations are related via k = g,d,a,

共2.9兲

where ␤ = 1 / kT. In this paper the ground electronic state 兩g典 is assumed to be decoupled from both the donor and the acceptor states of the ET process. This is typical in many photoinduced ET reactions where the ground state is only coupled to the excited states through the electromagnetic field. Thus, the nuclear Hamiltonian at the ground electronic state, Hg, is not directly involved in the ET dynamics. Instead, it defines the initial vibrational wave packet for the bright donor state upon photoexcitation 共within the approximation of an ultrashort laser pulse.兲 In the following we consider a simple model by defining the Duschinsky rotation matrix ˜S as a consecutive multipli˜ 共k , l ; ␪兲, cation of a number of primitive rotation matrices G each of which defines the rotation of two particular normal mode axes in the corresponding plane. Such a rotation matrix differs from the identity matrix only by four elements ˜ 共k,l; ␪兲兴 = 关G ˜ 共k,l; ␪兲兴 = cos ␪ , 关G k,k l,l 共2.10兲 ˜ 共k,l; ␪兲兴 = − 关G ˜ 共k,l; ␪兲兴 = sin ␪ , 关G k,l l,k and defines the rotation 共␪兲 of two orthogonal axes 共k , l兲 from their initial position to the final position in the 共k , l兲 plane. We adopt the convention that the rotations are defined for the donor/acceptor DPESs with respect to the ground state configuration 共rather than rotations of the corresponding coordinate frames.兲 Assuming k ⬍ l, a negative angle ␪ describes the counterclockwise rotation of a corresponding potential surface. This is illustrated in Fig. 1 for the case of two normal modes. The coordinate transformation in Eq. 共2.2兲 maps the DPESs according to their minima. However, in some practi-

1 T˜ ⑀vert k = ⑀ k + 2 D k K kD k ,

共2.13兲 ˜ D . ck = − K k k In our model the vibrational modes are divided into the inner sphere and the outer sphere modes. The first group consists of a finite number of intramolecular modes that are in general subject to the Duschinsky rotation. The other group comprises a continuous bath 共B兲 mapped from the linear response theory, which is characterized by a spectral density2,3 J B共 ␻ 兲 =

␲ 兺 ␻3关Dd − Da兴2j ␦共␻ − ␻ j兲. 8 j苸B j

共2.14兲

In this work we employ an Ohmic form with exponential cutoff J B共 ␻ 兲 = ␭ B

␲␻ −␻/␻ B. e 4␻B

共2.15兲

Here, ␭B and ␻B are the reorganization energy and the characteristic frequency of the bath, respectively. For practical calculations the continuous spectral density can be discretized following the procedure described previously.15,32 III. A GENERALIZED MASTER EQUATION APPROACH A. General discussion

As an approximate approach for describing ET dynamics in the presence of mode-mixing, we extend the generalized master equation 共GME兲 approach introduced by Evans and Coalson37,38 and later rederived by Golosov and Reichman39 using the Nakajima–Zwanzig projection operator technique.40–42 This approach treats nonadiabatic coupling perturbatively 共to the second order兲 and the electronic-

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

094109-4

J. Chem. Phys. 131, 094109 共2009兲

K. A. Velizhanin and H. Wang

vibrational coupling exactly 共to all orders兲. Another frequently used approach to describe ET dynamics, i.e., the Redfield theory,43,44 does not allow one to treat strong electronic-vibrational couplings for cases where modemixing effects are known to be important.27 Thus, we do not consider it in this paper. In the remaining part of the section we follow closely the derivation of GME by Golosov and Reichman39 but include the more sophisticated mode-mixing treatment. For convenience we transform Eqs. 共2.1兲 and 共2.7兲 to the interaction picture

⳵ ␳ˆ 共t兲 = 关Vˆ共t兲, ␳ˆ 共t兲兴, ⳵t

共3.1a兲

␳ˆ 共t兲 = eiH0t␳共t兲e−iH0t ,

共3.1b兲

i

Vˆ共t兲 = eiH0tVe−iH0t = ⌬兩a典eiHate−iHdt具d兩 + ⌬兩d典eiHdte−iHat具a兩, 共3.1c兲 in which the operators are denoted with a hat. It is seen from Eq. 共2.1兲 that the time evolution operator eiH0t commutes with the electronic population operator. Therefore, observables of interest remain formally the same as in Eq. 共2.5兲 Pk共t兲 = Tr关兩k典具k兩␳ˆ 共t兲兴;

k = d,a.

共3.2兲

To carry out perturbation expansions of the density matrix we adopt the Liouville space notation.45,46 An arbitrary quantum mechanical operator A becomes a double ket vector 具兩A典典 in Liouville space. The inner product 具具A 兩 B典典 is defined as 具具A兩B典典 = TrA艚B共A†B兲,

共3.3兲

where TrA艚B indicates that the trace is evaluated only over degrees of freedom which are simultaneously present in both operators A and B. Furthermore, the Liouville space superoperator Vˆ 共t兲 is introduced so that 关Vˆ共t兲 , ␳共t兲兴 becomes Vˆ 共t兲兩␳ˆ 共t兲典典. Using this notation the equation of motion 共3.1a兲 and the observables 共3.2兲 can be recast in the Liouville space form, yielding

⳵ i 具兩␳ˆ 共t兲典典 = Vˆ 共t兲具兩␳ˆ 共t兲典典, ⳵t

choice of the reference operator ␳c = ␳a was reported in an earlier work of Sparpaglione and Mukamel.47 The projection operator in Eq. 共3.5兲, together with the factorized initial density operator in Eq. 共2.8兲, leads to a closed form equation of motion for P具兩␳ˆ 共t兲典典

⳵ P具兩␳ˆ 共t兲典典 = − ⳵t



t

K共t, ␶兲P具兩␳ˆ 共␶兲典典d␶ ,

共3.6兲

0

where the memory kernel superoperator K共t , ␶兲 is given by K共t, ␶兲 = PVˆ 共t兲G共t, ␶兲Vˆ 共␶兲P,

共3.7兲

and the time propagator G共t , ␶兲 satisfies the differential equation i

⳵ G共t, ␶兲 = 共1 − P兲Vˆ 共t兲共1 − P兲G共t, ␶兲, ⳵t

共3.8兲

with the initial condition G共␶ , ␶兲 = 1. Equation 共3.6兲 is to be multiplied by 具具1v兩典具具d , d兩典 to obtain the equation of motion for donor state population Pd共t兲

⳵ Pd共t兲 = − ⳵t



t

Kd共t, ␶兲Pd共␶兲d␶ +

0



t

Ka共t, ␶兲Pa共␶兲d␶ ,

0

共3.9兲 with the exact memory kernels Kd共t, ␶兲 = 具具d,d兩典具具1v兩典Vˆ 共t兲G共t, ␶兲Vˆ 共␶兲具兩d,d典典具兩␳g典典, 共3.10a兲 Ka共t, ␶兲 = − 具具d,d兩典具具1v兩典Vˆ 共t兲G共t, ␶兲Vˆ 共␶兲具兩a,a典典具兩␳c典典. 共3.10b兲 The equation of motion for Pa共t兲 is obtained accordingly. Since the exact memory kernels usually cannot be evaluated analytically or numerically, approximations need to be introduced. In the nonadiabatic ET approximation all but the first nonvanishing term are neglected in the perturbation expansion of the memory kernel in powers of ⌬, i.e., ˆ ˆ Kd共t, ␶兲 ⬇ K共2兲 d 共t, ␶兲 = 具具d,d兩典具具1v兩典V共t兲V共␶兲具兩d,d典典具兩␳g典典,

共3.4a兲

共3.11a兲

共3.4b兲

ˆ ˆ Ka共t, ␶兲 ⬇ K共2兲 a 共t, ␶兲 = − 具具d,d兩典具具1v兩典V共t兲V共␶兲具兩a,a典典具兩␳c典典,

where 具兩k , k典典 is the Liouville space representation for 兩k典具k兩, and the identity operator acting on the vibrational degrees of freedom is denoted by 1v. The key point is that the Liouville space projection operator can be defined as39

共3.11b兲

Pk共t兲 = 具具k,k兩典具具1v兩␳ˆ 共t兲典典;

k = d,a,

P = 具兩d,d典典具兩␳g典典具具1v兩典具具d,d兩典 + 具兩a,a典典具兩␳c典典具具1v兩典具具a,a兩典, 共3.5兲 where ␳g has been defined previously 关Eq. 共2.9兲兴 and ␳c is an arbitrary reference density operator for the vibrational degrees of freedom. The use of this projection operator in the case of equilibrium initial preparation 共␳d = ␳g兲 and the

where the superscript “共2兲” stands for the second order perturbation. One can rewrite these memory kernels using the standard Hilbert space notation 2 iHdt −iHa共t−␶兲 −iHd␶ e e ␳g兲其, K共2兲 d 共t, ␶兲 = 2⌬ Re兵Trv共e

共3.12a兲

2 iHat −iHd共t−␶兲 −iHa␶ e e ␳c兲其, K共2兲 a 共t, ␶兲 = 2⌬ Re兵Trv共e

共3.12b兲

where Trv denotes the trace over the vibrational degrees of freedom. Although ␳c is, in principle, an arbitrary reference operator, Evans and Coalson suggested that the choice ␳c

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

094109-5

J. Chem. Phys. 131, 094109 共2009兲

Electron transfer with mode mixing

= ␳a, i.e., thermal equilibrium on acceptor DPES, is reasonable. This choice is adopted for the rest of this paper. B. Vibration correlation function

Memory kernels in Eq. 共3.12兲 can be expressed in terms of the vibration correlation function C共␶1 , ␶2 , . . . , ␶n兲 C共␶1, ␶2, . . . , ␶n兲 = Trv共e−␶1H1e−␶2H2 ¯ e−␶nHn兲, Hj =

1 T 2p p

+ V j共q兲,

j = 1,2, . . . ,n,

共3.13a兲 共3.13b兲

where V j共q兲 defines a generic quadratic potential surface 共assumed to be positive definite兲 and Re共␶ j兲 ⱖ 0. Following Ianconescu and Pollak,29 the quantum mechanical trace in Eq. 共3.13a兲 is evaluated as C共␶1, ␶2, . . . , ␶n兲 =

冕冕 冕 ⬁



−⬁

−⬁

⫻ 具q1兩e

¯

−␶1H1



−⬁

共3.14兲

dQ⬘dQ具q⬘兩Q⬘典具Q⬘兩e−␶ jH j兩Q典具Q兩q典.

Q = ˜STj 共q − D j兲, 共3.16兲 Q⬘ = ˜STj 共q⬘ − D j兲, where translation vector D j is chosen so that Q = 0 corresponds to the minimum of potential surface V j. Hamiltonian ˜ H j in the normal mode coordinates is 共⍀ j = diag兵␻1 , ␻2 , . . .其兲 +

具Q兩q典 = ␦关Q − ˜STj 共q − D j兲兴, 共3.21兲 具Q⬘兩q⬘典 = ␦关Q⬘ − ˜STj 共q⬘ − D j兲兴. Combining all the expressions, Eq. 共3.15兲 is given as 具q⬘兩e−␶ jH j兩q典 =

冑det共a˜ 兲 共2␲兲 f/2



1 ˜ q ⬘ + q TA ˜ q兲 exp − ␶ j⑀ j − 共q⬘TA 2



共3.22兲

where ˜ = ˜S ˜a˜ST , A j j ˜ = ˜S ˜b˜ST , B j j 共3.23兲 ˜ TD , ˜ − ˜b兲S C = ˜S j共a j j ˜ TD . ˜ − ˜b兲S E = DTj ˜S j共a j j

The coordinates are related through

Hj = ⑀j +

According to the coordinate transformation 共3.16兲, we have

兩q2典

共3.15兲

1 T 2P P

共3.20兲

␻1 ␻2 ˜b = diag , ,... . sinh共␶ j␻1兲 sinh共␶ j␻2兲

˜ q + CT共q⬘ + q兲 + E , + q ⬘TB

Each matrix element of type 具q⬘兩e−␶ jH j兩q典 can be evaluated by introducing the normal mode coordinates Q

冕冕

冎 冎

␻1 ␻2 , ,... , tanh共␶ j␻1兲 tanh共␶ j␻2兲

dq1dq2 ¯ dqn

⫻具q2兩e−␶2H2兩q3典 ¯ 具qn兩e−␶nHn兩q1典.

具q⬘兩e−␶ jH j兩q典 =

再 再

˜a = diag

1 T˜ 2 2 Q ⍀ j Q,

共3.17兲

According to Eq. 共3.22兲, expression 共3.14兲 becomes an n-fold Gaussian integral along the multidimensional coordinates qm 共m = 1 , 2 , . . . , n兲. This can be sequentially evaluated for each qm using the Gaussian integral identity 1 共2␲兲 f/2





冋 冋

1 ˜ dx exp − xTM x + N Tx 2 −⬁







˜ 兲exp 1 NTM ˜ −1N . = det共M 2

共3.24兲

where P = ˜STj p. Using the well-known expression48 for the one-dimensional harmonic oscillator HHO = P2 / 2 + ␻2Q2 / 2

冑a共␶兲

冋 册

+ b共␶兲Q⬘Q , a共␶兲 =

IV. NUMERICALLY EXACT METHODOLOGY

1 2 2 1/2 exp − a共␶兲共Q⬘ + Q 兲 2 共2␲兲

具Q⬘兩e−␶HHO兩Q典 =

␻ , tanh共␶␻兲

b共␶兲 =

A. Quantum trace evaluation

共3.18a兲

␻ , sinh共␶␻兲

共3.18b兲

the matrix element 具Q⬘兩e−␶ jH j兩Q典 can be easily expressed as 具Q⬘兩e

−␶ jH j

冑det共a˜ 兲



1 T˜ 兩Q典 = f/2 exp − ␶ j⑀ j − 共Q⬘ aQ⬘ 2 共2␲兲



+ QT˜aQ兲 + Q⬘T˜bQ , where f is the number of degrees of freedom and

The simulation of population dynamics in Eq. 共2.5兲 involves the evaluation of the quantum mechanical trace, including a thermal Boltzmann average over the initial state 共2.8兲, as well as the calculation of the time propagation. The quantum mechanical trace for evaluating the physical observables, as described above, is of the general form 具A共t兲典 =

1 Tr关␳eiHtAe−iHt兴. Tr关␳兴

共4.1兲

If the initial density matrix ␳ can be diagonalized, i.e., 共3.19兲

␳ = 兺 pn兩n典具n兩, n

then Eq. 共4.1兲 reduces to a simple summation

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

共4.2兲

094109-6

J. Chem. Phys. 131, 094109 共2009兲

K. A. Velizhanin and H. Wang

具A共t兲典 =

1 兺 pn具n兩eiHtAe−iHt兩n典, 兺n pn n

共4.3兲

which can be carried out by Monte Carlo importance sampling methods.49–51 On the other hand, even if ␳ cannot be trivially diagonalized, more sophisticated methods are available to cast Eq. 共4.1兲 into a sum of wave functions with proper weights.51

B. Multilayer multiconfiguration time-dependent Hartree theory

V. RESULTS AND DISCUSSION A. Resonant nonadiabatic ET

The time evolution for the wave function of each sample, e−iHt兩n典, is achieved by employing the recently developed multilayer multiconfiguration time-dependent Hartree 共ML-MCTDH兲 theory.32 This is a variational approach to describing quantum dynamics in systems with many degrees of freedom. It extends the capability of the original MCTDH method52–55 for application to significantly larger systems. In the original MCTDH method, the overall wave function is expanded as 兩⌿共t兲典 = 兺 AJ共t兲兩⌽J共t兲典 J

M

⬅ 兺 兺 ¯ 兺 A j1 j2,. . .,jM 共t兲 兿 兩␾kj 共t兲典, j1

j2

jM

k=1

共4.4兲

k

where 兩␾kj 共t兲典 is the “single-particle” 共SP兲 function for the k kth SP degree of freedom and is expanded in terms of timeindependent basis functions, 兩␾kn共t兲典 = 兺 BIk,n共t兲兩uIk典.

共4.5兲

I

In contrast to that, the ML-MCTDH method employs a dynamic contraction of the basis functions that constitute the SP functions, where a time-dependent multiconfigurational expansion of the SP functions is used, 兩␾kn共t兲典 = 兺 BIk,n共t兲兩uIk共t兲典 I

⬅兺兺¯兺 i1

iQ共k兲

i2

Q共k兲

Bik,ni . . .i 共k兲共t兲 12 Q

兿 兩vik,q共t兲典.

q=1

q

共4.6兲

Here, Q共k兲 denotes the number of level two 共L2兲-SP degrees of freedom in the kth level one 共L1兲-SP group and 兩vik,q共t兲典 is q the L2-SP function for the qth L2-SP degree of freedom. Employing two dynamical layers, the expansion of the overall wave function can thus be written in the form 兩⌿共t兲典 = 兺 兺 ¯ 兺 A j1 j2. . .jM 共t兲 j1

j2

M

⫻兿

k=1

冋兺 兺 jM

i1

i2

¯兺

iQ共k兲

Q共k兲

Bik,ji k. . .i 共t兲 12 Q共k兲



q=1

straightforward. For the calculations presented in this paper up to four dynamical layers are employed. The inclusion of several dynamically optimized layers in the ML-MCTDH method provides more flexibility in the variational functional, which significantly advances the capabilities of performing wave packet propagations in complex system. This has been demonstrated by several applications to quantum dynamics in the condensed phase including many degrees of freedom.49–51,57,58

兩vik,q共t兲典 q



In this work we parametrized the frequencies and the linear couplings of the intramolecular modes 共I兲 by the following expression:

冉 冑

˜ 兴 = ␻ ln 关⍀ i,j I

关Da − Dd兴i =



N ␦ij, N − i − 1/2

i, j 苸 I, 共5.1兲

2␭I N␻2i

,

where N is the number of intramolecular modes, ␭I is the total reorganization energy of intramolecular modes defined by Eq. 共2.4兲, and ␻I is a characteristic frequency. It should be noted that Eq. 共5.1兲 is used only for the convenience of model studies. In reality those frequencies and displacement vectors should be determined from electronic structure theory calculations, which may not be cast into a simple ˜ =⍀ ˜ mathematical expression. We further set Dd = 0, ⍀ d a ˜ , and ˜S = ˜1—the unit matrix. This set of parameters cor=⍀ g d responds to no frequency shift and an equilibrium initial preparation, i.e., the initial density matrix 共2.8兲 would describe the equilibrium state of the system if the nonadiabatic coupling parameter ⌬ were set to zero. Figure 2 shows the population dynamics of the acceptor state for the case of twelve intramolecular modes, in which the donor and acceptor states have the same electronic energy, i.e., ⑀ad = 0. Panels 共a兲 and 共b兲 represent population dynamics for different nonadiabatic coupling parameters, ⌬ / ␻I = 0.012 and ⌬ / ␻I = 0.12, respectively. Results are shown for cases with and without Duschinsky rotation. To model the former case the twelve intramolecular modes are divided into four groups. Each of the three modes in the first group is mixed with the corresponding one in the second group, and each of the three modes in the third group is mixed with the corresponding one in the fourth group, this yields the parameterization of the Duschinsky rotation matrix as ˜S = a



˜ 共n,n + 3;− 30°兲. G

共5.2兲

n=1–3,7–9

.

共4.7兲 The equations of motion within the ML-MCTDH approach32 can be obtained from the Dirac–Frenkel variational principle.56 The extension to more dynamical layers is

Figure 2 illustrates that the inclusion of mode mixing makes the ET dynamics proceed more slowly. Both the approximate GME approach and the numerically exact ML-MCTDH method predict this trend qualitatively correctly. At smaller ⌬ 关panel 共a兲兴 the GME results and the ML-MCTDH results are in excellent agreement. This is expected since the GME ap-

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

094109-7

J. Chem. Phys. 131, 094109 共2009兲

Electron transfer with mode mixing 0.06 ML-MCTDH, no mode-mixing ML-MCTDH, mode-mixing GME, no mode-mixing GME, mode-mixing

0.2

ML-MCTDH, no mode-mixing ML-MCTDH, no mode-mixing GME, no mode-mxing GME, mode-mixing

Pa(t)

Pa(t)

0.04

0.3

0.02

0 0

0.1

10

(a)

20

10

tωI

20

20

30

ML-MCTDH, no mode-mixing ML-MCTDH, mode-mixing GME, no mode-mixing GME, mode-mixing

0.5

0 0

30

FIG. 2. The dynamics of the acceptor state population for 12 intramolecular vibrational modes. The Duschinsky rotation matrix for the mode-mixing ˜ 共n , n + 3 ; −30°兲. The total reorganization energy case is taken ˜Sa = 兿n=1–3,7–9G of the intramolecular modes is ␭I / ␻I = 0.1. Shown are results for different coupling strengths: 共a兲 ⌬ / ␻I = 0.012, 共b兲 ⌬ / ␻I = 0.12. Other parameters are ⑀ad = 0 and kT / ␻I = 0.4.

tωI

1

ML-MCTDH, no mode-mixing ML-MCTDH, mode-mixing GME, no mode-mixing GME, mode-mixing

(b)

10

(a)

0.5

0 0

0 0

30

Pa(t)

Pa(t)

1

tωI

(b)

10

tωI

20

30

FIG. 3. Same as Fig. 2 except the bias ⑀ad / ␻I = −3.3 and the nonadiabatic coupling parameter ⌬ / ␻I = 0.12 共a兲 and 1.2 共b兲.

Figure 3 shows the population dynamics for a nonresonant ET. The parameters are the same as in Fig. 2 except for the electronic bias ⑀ad / ␻I = −3.3. High frequency coherent oscillations observed in both panels are associated with the Rabi oscillations of the bare two-level system 共兩d典 ↔ 兩a典兲 2 + 4⌬2. In this parameter regime, the incluwith ␻Rabi = 冑⑀ad sion of mode mixing accelerates the ET dynamics, as it was demonstrated earlier by a number of groups.25,27,30 GME predicts this trend qualitatively correctly. In panel 共a兲 the GME results are in good agreement with the ML-MCTDH simulation for ⌬ / ␻I = 0.12. However, as shown in Fig. 3共b兲 for a larger nonadiabatic coupling 共⌬ / ␻I = 1.2兲 the GME population deviates from the numerically exact one after a short time, and predicts an incorrect long time plateau of Pa共t兲. A similar problem exists in the noninteracting blip approximation based treatment of the original spin-boson model with a large electronic energy bias.3,5 Our results show that this problem remains in the GME approximation when mode mixing is present.

ergy surface of the traditional spin-boson model2,3 is known to be monostable at Ea / ⌬ ⬍ 41 and bistable at Ea / ⌬ ⬎ 41 , where Ea is the Marcus activation energy 共Ea = 41 ␭ in the resonant regime with no mode mixing兲.3 In the latter case if Ea is large, then the energy barrier between the two potential energy minima is high and population trapping in one of the minima may occur. Since the value of Ea can be affected by the degree of mode mixing, it is natural to expect population trapping to turn on/off by introducing mode mixing at certain physical regimes. Figure 4 shows the population dynamics for an example of resonant adiabatic ET with twelve intramolecular modes. The parameters are ␭I = ⌬ = 10␻I. At zero temperature 关panel 共a兲兴, large amplitude coherent motions are observed for Pa共t兲 in the case of no mode mixing. This remnant of the Rabi oscillation is quenched at a higher temperature 关panel 共b兲兴. The introduction of mode mixing by the Duschinsky rotation ˜ 共n , n + 6 ; −45°兲 leads to a qualitative matrix ˜Sa = 兿n=2–6G change: The electron is trapped in the donor state and thus the population of the acceptor state remains low at long times. In this regime the GME results are in poor agreement with the ML-MCTDH simulations. At zero temperature 关panel 共a兲兴 it does not even preserve positivity, whereas at finite temperature 关panel 共b兲兴 it overestimates decoherence for the case of no mode mixing and predicts faster ET when mode mixing is introduced. These shortcomings are consistent with the fact that the GME approach assumes nonadiabatic ET dynamics, which is invalid for the parameter regime in Fig. 4.

C. Resonant adiabatic ET

D. Intramolecular modes+ bath

This is the parameter regime where ⌬ / ␻I Ⰷ 1, i.e., the vibrational environment is slow compared to the electronic transition. In this regime the lowest adiabatic potential en-

Here we consider the situation where both the inner and outer sphere 共bath兲 modes are present. Figure 5 shows the dynamics of the acceptor state population in the nonadiabatic

proach is a better approximation if the perturbation, i.e., the nonadiabatic coupling, is sufficiently weak. For a larger ⌬ 关panel 共b兲兴 the GME result still agrees with the exact simulation at short times but deviates at a longer time scale. B. Nonresonant nonadiabatic ET „⑀d > ⑀a…

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

094109-8

J. Chem. Phys. 131, 094109 共2009兲

K. A. Velizhanin and H. Wang

Pa(t)

1

ML-MCTDH, no mode-mixing ML-MCTDH, mode-mixing GME, no mode-mixing GME, mode-mixing

0.5

0 0 1

Pa(t)

1

0.5 tωI

(a)

ML-MCTDH, no mode-mixing ML-MCTDH, mode-mixing GME, no mode-mixing GME, mode-mixing

0.5

0 0

1

0.5 tωI

(b)

FIG. 4. The dynamics of the acceptor state population for 12 intramolecular vibrational modes. The Duschinsky rotation matrix in the mode-mixing case ˜ 共n , n + 6 ; −45°兲. The total reorganization energy of the inis ˜Sa = 兿n=2–6G tramolecular modes is ␭I / ␻I = 10. Other parameters are ⌬ / ␻I = 10 and ⑀ad = 0. Temperature is kT / ␻I = 0 共a兲 and kT / ␻I = 5 共b兲.

regime 共⌬ / ␻I = 0.12, ␻I = ␻B兲 with four intramolecular modes and an Ohmic 共with exponential cutoff兲 bath. The strength of the electronic-vibrational interaction is given by ␭I / ␻I = 0.1 and ␭B / ␻B = 0.4.

Pa(t)

1

0.5 ML-MCTDH, no mode-mixing ML-MCTDH, mode-mixing GME, no mode-mixing GME, mode-mixing

0 0

50

(a)

Pa(t)

1

100

150

VI. CONCLUDING REMARKS

ML-MCTDH, no mode-mixing ML-MCTDH, mode-mixing GME, no mode-mixing GME, mode-mixing

0.5

0 0

(b)

tωI

50

tωI

100

For practical purposes the bath was discretized to the form of Eq. 共2.14兲 using the previously established procedure.15,32 For dynamics in Fig. 5 the convergence with respect to the discretization is reached with 50–100 bath modes in both the zero temperature 关panel 共a兲兴 and the finite temperature 关panel 共b兲兴 cases. Different from the previous subsections, the mode mixing is introduced at the vertical excitation point, Eqs. 共2.11兲 and 共2.12兲. As has been previously discussed, this is more convenient for the purpose of parametrizing the Hamiltonian 共2.1兲 from electronic structure theory calculations. It should be noted that upon the introduction of Duschinsky rotation the effective energy difference between the donor state and the acceptor state may be changed in this case. Thus, in practical modeling 共e.g., from electronic structure theory calculations兲 of ET processes at Frank–Condon regime, the neglect of mode mixing may affect both the ET dynamics and the equilibrium populations. This is what is observed in Fig. 5. Thereby both the short time dynamics and the long time limit of Pa共t兲 are changed once mode mixing is taken into account. The difference is larger for the zero temperature case in panel 共a兲, and is not as pronounced at a finite temperature, panel 共b兲. The GME results are in reasonable agreement with the numerically exact ML-MCTDH simulations. At zero temperature 关Fig. 5共a兲兴, the GME results are qualitatively correct except sometimes not preserving positivity in the presence of mode-mixing. The performance of GME becomes better at a finite temperature, Fig. 5共b兲. Finally, we briefly comment on the computational effort in carrying out both GME and ML-MCTDH computations. GME is relatively inexpensive and takes at most 2–3 CPU hours for the examples considered in this paper. The numerically exact ML-MCTDH simulation is more expensive, where each wave function may take 4–10 hours. At zero temperature this is all it takes. However, at a finite temperature an ensemble of 500–1000 wave functions need to be propagated to evaluate the quantum mechanical trace, Eq. 共4.3兲. This is achieved by running a parallel version of the program on a computer cluster.

150

FIG. 5. The dynamics of the acceptor state population for four intramolecular vibrational modes plus the bath. The Duschinsky rotation matrix in the ˜ 共1 , 3 ; −20°兲G ˜ 共2 , 4 ; −15°兲. The total reorganizamode-mixing case is ˜Sa = G tion energies for the intramolecular and the bath modes are taken to be ␭I / ␻I = 0.1 and ␭B / ␻B = 0.4, respectively, with ␻I = ␻B. Temperature is grad kT / ␻I = 0 共a兲 and kT / ␻I = 0.5. Other parameters are ⌬ / ␻I = 0.12 and ⑀ad = 0.

In this work we have investigated two theoretical methods for simulating ET dynamics in the presence of mode mixing. The first one is a generalized master equation approach, which is based on the approximation that treats the nonadiabatic coupling strength ⌬ as the perturbation parameter. The second is the ML-MCTDH method, which is a nonperturbative and numerically exact approach. Comparison of the numerical results of these two methodologies shows that GME becomes accurate if ⌬ is significantly smaller than the characteristic frequency of the vibrational environment. When necessary, the accuracy of the GME approach may be substantially improved by performing a partial resummation of the perturbation expansion terms using the Padé extrapolation technique.39,59 This will be investigated in future studies.

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

094109-9

Both methodologies have been used to study the effect of Duschinsky rotation on the efficiency of ET reactions in various regimes. It is demonstrated that the introduction of mode mixing may lead to either deceleration or acceleration of ET dynamics in the nonadiabatic regime. For the parameters considered in this paper the former has been observed for resonant ET processes, whereas the latter takes place if the acceptor state is energetically lower than the donor. For an adiabatic ET reaction the presence of mode-mixing may induce trapping. In this regime the GME approach was found to be inaccurate and predicts incorrectly a quick damping of the coherent oscillations at a finite temperature 关Fig. 4共b兲兴. Furthermore, it has been shown that neglecting mode-mixing in practical calculations may lead to an inaccurate prediction of the short time behavior or rate constants as well as the long time limit of the electronic population. These results suggest that an accurate description of the vibrational environment is essential for a correct and quantitative understanding of the condensed phase ET processes. Apart from the effect of mode mixing, anharmonicity may also play an important role for low frequency modes. Including both effects in the accurate quantum dynamical simulation is currently pursued by us. ACKNOWLEDGMENTS

The authors would like to thank Michael Thoss for stimulating discussions and critical comments on this manuscript. This work has been supported by the National Science Foundation 共NSF兲 CAREER under Award No. CHE0348956 and used resources of the National Energy Research Scientific Computing Center 共NERSC兲, which is supported by the Office of Science of the U.S. Department of Energy under Contract No. DE-AC02-05CH11231. 1

J. Chem. Phys. 131, 094109 共2009兲

Electron transfer with mode mixing

Electron Transfer—from isolated Molecules to Biomolecules, edited by M. Bixon and J. Jortner 共Wiley, New York, 1999兲. 2 A. J. Leggett, S. Chakravarty, A. T. Dorsey, M. P. A. Fisher, A. Garg, and W. Zwerger, Rev. Mod. Phys. 59, 1 共1987兲. 3 U. Weiss, Quantum Dissipative Systems 共World Scientific, Singapore, 1999兲. 4 H. Grabert and A. Nitzan, Chem. Phys. 296, 101 共2004兲. 5 M. Grifoni, E. Paladino, and U. Weiss, Eur. Phys. J. B 10, 719 共1999兲. 6 R. Bulla, N. Tong, and M. Vojta, Phys. Rev. Lett. 91, 170601 共2003兲. 7 F. K. Wilhelm, S. Kleff, and J. von Delft, Chem. Phys. 296, 345 共2004兲. 8 A. Garg, J. N. Onuchic, and V. Ambegaokar, J. Chem. Phys. 83, 4491 共1985兲. 9 J. Ankerhold and H. Lehle, J. Chem. Phys. 120, 1436 共2004兲. 10 H. Wang, X. Song, D. Chandler, and W. H. Miller, J. Chem. Phys. 110, 4828 共1999兲. 11 E. Martin-Fierro and E. Pollak, J. Chem. Phys. 126, 164108 共2007兲. 12 R. Egger and U. Weiss, Z. Phys. 89, 97 共1992兲. 13 R. Egger and C. H. Mak, Phys. Rev. B 50, 15210 共1994兲.

N. Makri and D. E. Makarov, J. Chem. Phys. 102, 4600 共1995兲. H. Wang, M. Thoss, and W. H. Miller, J. Chem. Phys. 115, 2979 共2001兲. 16 M. Thoss, H. Wang, and W. H. Miller, J. Chem. Phys. 115, 2991 共2001兲. 17 H. Wang and M. Thoss, New J. Phys. 10, 115005 共2008兲. 18 A. Suárez and R. Silbey, J. Chem. Phys. 95, 9115 共1991兲. 19 N. Makri, J. Phys. Chem. B 103, 2823 共1999兲. 20 H. Wang and M. Thoss, J. Phys. Chem. 111, 10369 共2007兲. 21 F. Duschinsky, Acta Physicochim. URSS 7, 551 共1937兲. 22 A. Raab, G. A. Worth, H.-D. Meyer, and L. S. Cederbaum, J. Chem. Phys. 110, 936 共1999兲. 23 R. Kubo and Y. Toyozawa, Prog. Theor. Phys. 13, 160 共1955兲. 24 Y. J. Yan and S. Mukamel, J. Chem. Phys. 85, 5908 共1986兲. 25 A. M. Mebel, M. Hayashi, K. K. Liang, and S. H. Lin, J. Phys. Chem. A 103, 10674 共1999兲. 26 S. A. Egorov, E. Rabani, and J. Berne, J. Chem. Phys. 110, 5238 共1999兲. 27 G. M. Sando, K. G. Spears, J. T. Hupp, and P. T. Ruhoff, J. Phys. Chem. A 105, 5317 共2001兲. 28 K. K. Liang, A. M. Mebel, S. H. Lin, M. Hayashi, H. L. Selzle, E. W. Schlag, and M. Tachiya, Phys. Chem. Chem. Phys. 5, 4656 共2003兲. 29 R. Ianconescu and E. Pollak, J. Phys. Chem. A 108, 7778 共2004兲. 30 Q. Peng, Y. Yi, Z. Shuai, and J. Shao, J. Chem. Phys. 126, 114302 共2007兲. 31 S. Banerjee and G. Gangopadhyay, J. Chem. Phys. 126, 034102 共2007兲. 32 H. Wang and M. Thoss, J. Chem. Phys. 119, 1289 共2003兲. 33 S. Nakashima, Y. Nagasawa, K. Seike, T. Okada, M. Sato, and T. Konzuma, Chem. Phys. Lett. 331, 396 共2000兲. 34 T. Cimei, A. R. Bizzarri, G. Cerullo, S. D. Silvestri, and S. Cannistraro, Biophys. Chem. 106, 221 共2003兲. 35 R. A. Marcus, J. Chem. Phys. 24, 966 共1956兲. 36 R. A. Marcus, J. Chem. Phys. 43, 679 共1965兲. 37 D. G. Evans and R. D. Coalson, J. Chem. Phys. 102, 5658 共1995兲. 38 D. G. Evans and R. D. Coalson, J. Chem. Phys. 104, 3598 共1996兲. 39 A. A. Golosov and D. R. Reichman, J. Chem. Phys. 115, 9848 共2001兲. 40 S. Nakajima, Prog. Theor. Phys. 20, 948 共1958兲. 41 R. Zwanzig, J. Chem. Phys. 33, 1338 共1960兲. 42 R. Zwanzig, Nonequilibrium Statistical Mechanics 共Oxford University Press, New York, 2001兲. 43 A. G. Redfield, IBM J. Res. Dev. 1, 19 共1957兲. 44 A. G. Redfield, Adv. Magn. Reson. 1, 1 共1965兲. 45 S. Mukamel, Principles on Nonlinear Optical Spectroscopy 共Wiley– Interscience, Oxford, 1995兲. 46 H.-P. Breuer and F. Pettruccione, The Theory of Open Quantum Systems 共Clarendon, Oxford, 2006兲. 47 M. Sparpaglione and S. Mukamel, J. Chem. Phys. 88, 3263 共1988兲. 48 L. S. Schulman, Techniques and Applications for Path Integration 共Wiley, New York, 1981兲. 49 H. Wang and M. Thoss, J. Phys. Chem. A 107, 2126 共2003兲. 50 M. Thoss and H. Wang, Chem. Phys. 322, 210 共2006兲. 51 H. Wang and M. Thoss, J. Chem. Phys. 124, 034114 共2006兲. 52 H.-D. Meyer, U. Manthe, and L. S. Cederbaum, Chem. Phys. Lett. 165, 73 共1990兲. 53 U. Manthe, H.-D. Meyer, and L. S. Cederbaum, J. Chem. Phys. 97, 3199 共1992兲. 54 M. H. Beck, A. Jäckle, G. A. Worth, and H.-D. Meyer, Phys. Rep. 324, 1 共2000兲. 55 H.-D. Meyer and G. A. Worth, Theor. Chem. Acc. 109, 251 共2003兲. 56 J. Frenkel, Wave Mechanics 共Clarendon, Oxford, 1934兲. 57 M. Thoss, I. Kondov, and H. Wang, Chem. Phys. 304, 169 共2004兲. 58 K. A. Velizhanin, H. Wang, and M. Thoss, Chem. Phys. Lett. 460, 325 共2008兲. 59 A. A. Golosov and D. R. Reichman, Chem. Phys. 296, 129 共2004兲. 14 15

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Dynamics of electron transfer reactions in the presence ...

Received 8 July 2009; accepted 8 August 2009; published online 4 September 2009. A generalized ... Things are different for the inner sphere degrees of free- dom. They are not in ...... Scientific Computing Center NERSC, which is supported.

335KB Sizes 1 Downloads 349 Views

Recommend Documents

Dynamics of electron transfer reactions in the presence ...
Received 8 July 2009; accepted 8 August 2009; published online 4 September 2009. A generalized master equation approach is developed to describe electron transfer ET dynamics in ... Things are different for the inner sphere degrees of free- dom. They

Kondo Effect in the Presence of Itinerant-Electron Ferromagnetism ...
Dec 10, 2003 - 3Hungarian Academy of Sciences, Institute of Physics, TU Budapest, H-1521, Hungary. 4Institute of Molecular Physics, Polish Academy of ...

Kondo Effect in the Presence of Itinerant-Electron ... - Semantic Scholar
Dec 10, 2003 - system's ground state can be tuned to have a fully com- pensated local spin, in which case the QD conductance is found to be the same for each spin channel, G" И G#. The model.—For ferromagnetic leads, electron- electron interaction

Collective chemotactic dynamics in the presence of self-generated ...
Oct 22, 2012 - [7] to study swimmer transport and rotation in a given background ... *Corresponding author: [email protected] and-tumble dynamics.

Collective chemotactic dynamics in the presence of self ... - NYU (Math)
Oct 22, 2012 - them around the domain. The dynamics is one of .... [22] S. Park, P. M. Wolanin, E. A. Yuzbashyan, P. Silberzan, J. B.. Stock, and R. H. Austin, ...

Photoinduced Intramolecular Electron Transfer of ...
photoinduced electron transfer between carbazole derivatives and fullerenes has been well .... giving kinetic data of the charge-separation processes. As shown.

Photoinduced Intramolecular Electron Transfer of Carbazole Trimer ...
Jan 9, 2008 - of the electron-transfer processes in the donor-C60 dyads depend ... photoinduced electron transfer between carbazole derivatives.

Catalytic Effects of Dioxygen on Intramolecular Electron Transfer in ...
radical ion pairs. The rate constant of BET increases linearly with increasing oxygen concentration without, however, forming reactive oxygen species, such as singlet oxygen or superoxide anion. When ferrocene (Fc) is used as a terminal electron dono

Photoinduced Electron Transfer in “Two-Point” Bound ...
Photosynthetic reaction centers of green plants and bacteria consist of complex ... the two components, which can act as a conduit for the energy or electron ...

Photoinduced electron transfer between fullerenes ...
free energy changes ( Get) for the electron transfer reactions on the basis of Rehm–Weller relation. Addition of pyridine (Py) enhanced significantly the rates and ...

Photoinduced Electron Transfer in “Two-Point” Bound ...
reorganization energy,5 and structure,6 only a limited number of noncovalently linked ..... available free of charge via the Internet at http://pubs.acs.org. References and Notes ... Commun. 2003, 226. (k) El-Khouly, M. E.; Rogers, L. M.; Zandler,.

Electron spin dynamics in impure quantum wells for ...
lated within the Boltzmann formalism for arbitrary couplings to a Rashba spin-orbit field. It is shown that .... electron occupation distribution function. Equation 6 is.

Synthesis and photoinduced electron transfer ... - Wiley Online Library
Electron Transfer Process. YU CHEN,1 YING LIN,1 MOHAMED E. EI-KHOULY,2,3 NAN HE,1 AIXIA YAN,4 YING LIU,1. LIANGZHEN CAI,1 OSAMU ITO2.

Photoinduced electron transfer between fullerenes (C60/C70) and ...
free energy changes ( Get) for the electron transfer reactions on the basis of Rehm–Weller relation. Addition of pyridine (Py) enhanced significantly the rates and ...

Efficient photoinduced electron transfer between C60 ...
70 were observed, accompanied by the appearance of C. БА. 60 and C. БА. 70. The electron transfer rate constants (ket) and the quantum yields (Φet) of C. БА.

Asymptotic Optimality of the Static Frequency Caching in the Presence ...
that improves the efficiency and scalability of multimedia content delivery, benefits of ... probability, i.e., the average number of misses during a long time period.

Capacity of Cooperative Fusion in the Presence of ...
Karlof and Wagner. [11] consider routing security in wireless sensor networks. ...... thesis, Massachusetts Institute of Technology, Cambridge, MA, August. 1988.

Proton Translocation Coupled to Electron Transfer ...
The most extensively studied examples of these oxidases are mitochondrial. CcO from bovine heart; ...... The unstructured protein medium itself cannot facilitate ...... environment in typical optical and electrometrical measurements. In one case ...

Control of Chemical Reactions by Convective Turbulence in the ...
Feb 15, 1998 - In the bulk of the boundary layer the chemical ... bulk of the convective boundary layer plays only a minor role. ... E-mail: [email protected].

New reactions and reactive intermediates in the pyrolysis of ... - Arkivoc
Aug 31, 2017 - spectra were recorded using a Varian Gemini 2000 instrument at 300 MHz for ... reference and coupling constants (J) are in Hz. Flash vacuum ...

Polymeric latexes prepared in the presence of 2-acrylamido-2 ...
Feb 6, 2001 - recovery of natural resources in the mining, petroleum and ..... ization of the acid monomer With an alkaline agent such as a ..... This data,.

Query Answering using Views in the Presence of ...
The problem of finding equivalent rewritings is formally defined as follows: We have a database schema R, a set of. CQ views V over schema R, a set of tgds and ...

Round-the-Clock presence of Central Excise Officers in Cigarette ...
There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. Main menu.

Electron correlation in 1D electron gas
Electron correlation in 1D electron gas. Yan Jun. July 26, 2008. 1 The equation of motion [1]. The equation of motion for the classical one-particle distribution ...