Sina Yeganeh and Mark A. Ratner- Effects of anharmonicity on nonadiabatic electron transfer: A model

download Sina Yeganeh and Mark A. Ratner- Effects of anharmonicity on nonadiabatic electron transfer: A model

of 9

Transcript of Sina Yeganeh and Mark A. Ratner- Effects of anharmonicity on nonadiabatic electron transfer: A model

  • 8/3/2019 Sina Yeganeh and Mark A. Ratner- Effects of anharmonicity on nonadiabatic electron transfer: A model

    1/9

    Effects of anharmonicity on nonadiabatic electron transfer: A modelSina Yeganeha and Mark A. Ratnerb

    Department of Chemistry and Center for Nanofabrication and Molecular Self Assembly, NorthwesternUniversity, Evanston, Illinois 60208-3113

    Received 28 September 2005; accepted 6 December 2005; published online 27 January 2006

    The effect of anharmonicity in the intramolecular modes of a model system for exothermic

    intramolecular nonadiabatic electron transfer is probed by examining the dependence of thetransition probability on the exoergicity. The Franck-Condon factor for the Morse potential is

    written in terms of the Gauss hypergeometric function both for a ground initial state and for the

    general case, and comparisons are made between the first-order perturbation theory results for

    transition probability for harmonic and Morse oscillators. These results are verified with quantum

    dynamical simulations using wave-packet propagations on a numerical grid. The

    transition-probability expression incorporating a high-frequency quantum mode and low-frequency

    medium mode is compared for Morse and harmonic oscillators in different temperature ranges and

    with various coarse-graining treatments of the delta function from the Fermi golden rule expression.

    We find that significant deviations from the harmonic approximation are expected for even

    moderately anharmonic quantum modes at large values of exoergicity. The addition of a second

    quantum mode of opposite displacement negates the anharmonic effect at small energy change, but

    in the inverted regime a significantly flatter dependence on exoergicity is predicted for anharmonic

    modes. 2006 American Institute of Physics. DOI: 10.1063/1.2162172

    I. INTRODUCTION

    The classical and semiclassical theories of electron

    transfer ET developed by Marcus and others13 have beensuccessful in describing and predicting many aspects of the

    ET rate constant including its dependence on driving force

    the negative free-energy change, G0 Ref. 4 and ontemperature. However, the neglect of quantum-mechanical

    tunneling causes the classical theory to disagree with experi-

    mental results in significant ways. The Marcus expression

    gives a Gaussian dependence on free energy, resulting in too

    steep a drop-off in the inverted region. In addition, the theory

    predicts that the rate constant approaches zero as temperature

    falls to zero while experimental results have shown a nonva-

    nishing rate constant at low temperatures.5

    These discrepan-

    cies can be dealt with by treating the problem quantum me-

    chanically, proceeding from the golden rule result of first-

    order perturbation theory and in the simplest approach

    assuming a single coupled vibrational mode represented by

    displaced harmonic potentials with same frequency.6

    Several

    authors have extended this treatment to include multiple

    modes of differing frequency,7,8

    and the physically important

    case of change in both displacement and frequency has beenaddressed.

    9Starting from the Kubo-Lax generating function

    for multiple harmonic oscillators with displacement but no

    frequency change, Jortner and co-workers1012

    formulated

    the ET rate constant in terms of a two-mode model with a

    high-frequency mode, representing an average of the relevant

    quantum modes, and a lower-frequency solvent mode. A very

    important early result came in applying the single-mode limit

    to the experimental data of DeVault and Chance5

    in chroma-

    tium, providing an explanation for the temperature depen-

    dence of the ET rate.

    While the harmonic approximation is generally appropri-

    ate for describing the low-frequency solvent modes,13

    it is

    significantly unrealistic for the quantum modes at large val-

    ues of as significantly anharmonic behavior is expected.

    Several authors have accounted for anharmonicity using

    Morse potentials14

    in treatments of nonradiative decay in

    crystals,15

    proton transfer reactions,16,17

    as well as nonadia-

    batic ET.11,1821

    In Sec. II of this paper we derive a more

    compact form of the transition probability for nonadiabatic

    ET with anharmonic quantum modes, making use of a finite

    sum form of the Gauss hypergeometric functions.22

    In Sec.

    III the results of computations with this closed form are

    compared with the corresponding harmonic results, and the

    golden rule results are compared with grid-based wave-

    packet propagations using the methods of Kosloff and co-

    workers. Finally, the two-mode case of anharmonic quantum

    mode and a harmonic low-frequency phonon mode is con-

    sidered in Sec. IV, and comparisons are made with the har-

    monic results from the treatment by Jortner.12

    Finally, the

    expected effects of significantly anharmonic modes on theET rate constant are discussed for parameters of interest in

    experimental systems such as the exoergicity in the optimal

    regime. Regions are identified in which the anharmonic ef-

    fect on the rate constant is expected to be largest.

    II. GOLDEN RULE FORMULATION: MORSEPOTENTIALS

    The starting point for the nonadiabatic quantum-

    mechanical treatments is the Fermi golden rule result,23

    aElectronic mail: [email protected]

    bElectronic mail: [email protected]

    THE JOURNAL OF CHEMICAL PHYSICS 124, 044108 2006

    0021-9606/2006/1244 /044108/9/$23.00 2006 American Institute of Physics124, 044108-1

    Downloaded 28 May 2009 to 129.105.215.213. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp

    http://dx.doi.org/10.1063/1.2162172http://dx.doi.org/10.1063/1.2162172http://dx.doi.org/10.1063/1.2162172http://dx.doi.org/10.1063/1.2162172
  • 8/3/2019 Sina Yeganeh and Mark A. Ratner- Effects of anharmonicity on nonadiabatic electron transfer: A model

    2/9

    WRP =2

    PR

    2EP , 1

    where WRP, the transition probability per unit time from re-

    actants to products, is given in terms of the product density

    of states DOS, EP, and the matrix element over the re-actant R and product P wave functions. Making theCondon approximation

    24that the electronic coupling is inde-

    pendent of the nuclear motion, the transition rate can be writ-ten as

    25

    WRP =2

    HRP

    2rp2EP, 2

    where the electronic coupling matrix element HRP has been

    factored out and is assumed to be independent of separation,

    leaving the Franck-Condon FC overlap integral over reac-tant and product vibrational states, r and p, respectively.In this single-mode expression, energy conservation requires

    that the product vibrational state be given by Enp =+Enr, where En is the energy of the nth vibrational

    eigenstate. The energy conservation requirement can be re-laxed in the single-mode case by introducing coarse graining,

    assuming a vibrational coupling to environmental modes.12,26

    For the multimode case, a more general form is used where

    the delta function allows contributions only from energy-

    conserving transitions primed and unprimed variables referto the final and initial states, respectively,

    WRP =2

    HRP

    2k

    nk=0

    n

    k=0

    Ink,nk

    k2

    eEnkk

    /kBT

    Qk

    k

    En

    k

    k Enk

    k . 3Here the product is over the k different modes, Qk is the

    partition function for the kth mode, Ink,nk

    kis the FC overlap

    integral given by nkk

    nk

    k , and Enkk and E

    nk

    kare the vibra-

    tional energies. The summations over nk and nk in Eq. 3represent the possible transitions between initial and final

    vibrational levels weighted by the fractional population in

    each reactant initial vibrational level, as given by

    eEnkk

    /kBT/Qk. The delta function then picks out combinations

    of nk and nk which result in energy-conserving transitions.

    As noted by Jortner,12

    this delta function should be replaced

    by a Lorentzian to account for energy-time uncertainty

    broadening resulting from finite lifetimes. Kubo7

    and Lax8

    were able to reduce Eq. 3 to give a generating function forthe FC factors for multiple harmonic oscillators.

    For our treatment of anharmonicity in the quantum

    mode, we consider the eigenfunctions of two Morse poten-

    tials centered at r0 and r0,

    Vr = D1 e rr02, Vr = D1 e rr02, 4

    and proceeding in the usual fashion, we take the ranges = and strengths D =D of the potentials to be the same.Unlike the harmonic case, however, in which analytic forms

    can be written for the FC factor of oscillators of different

    frequencies, the Morse result can only be calculated analyti-

    cally when =. The eigenfunctions of this anharmonic

    potential14

    can be written as27

    nN,;r r0 = Mn,NNw + wN/2n

    exp N+ 12

    wLnN2nNw + w , 5where Ln

    k is the generalized Laguerre polynomial and

    w = err0, Mn,N = j=0nN 2n + j

    j! 1/2

    ,

    6

    N= 22D

    1 .

    The parameter N is physically important because N/ 2 gives

    the highest bound level in the Morse potential. The FC over-

    lap integral,

    In,n = nN,;r r0nN,;r r0dr, 7was first given for Morse eigenfunctions by Fraser and

    Jarmain.28,29

    The expression simplifies as N=N for our case

    of identical strengths and ranges, and we can write in the

    notation of Iachello and Ibrahim,27

    In,n = Mn,NMn,NN/2n 2

    1 + Nnn

    m=0

    n

    m=0

    n 1m+mm ! m!

    N nn m

    N nn m

    m

    2

    1 + m+m

    N n n

    + m + m

    , 8

    with = er0r0.

    We now proceed to simplify Eq. 8 in the low-temperature limit where the initial state is entirely in its

    ground vibrational level. In keeping with our notation we set

    n =0, yielding

    I0,n = M0,NMn,NN/2 2

    1 + Nn

    m=0

    n 1mm!

    N nn m

    2

    1 +

    m

    N n + m

    . 9

    The binomial coefficient can be expanded in its gamma func-

    tion representation, and some grouping produces

    I0,n = M0,NMn,NN/2 2

    1 + NnN n + 1

    m=0

    n 2/1 + mm!

    N n + m

    n m + 1N 2n + m + 1 . 10

    044108-2 S. Yeganeh and M. A. Ratner J. Chem. Phys. 124, 044108 2006

    Downloaded 28 May 2009 to 129.105.215.213. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp

  • 8/3/2019 Sina Yeganeh and Mark A. Ratner- Effects of anharmonicity on nonadiabatic electron transfer: A model

    3/9

    The summation can be rewritten using a finite sum form of

    the Gauss 2F1 hypergeometric series that holds when n is a

    positive integer,30

    2F1 n,b;c ;z = m=0

    n nmbm

    cm

    zm

    m!, 11

    where the Pochhammer symbol is used,

    ak=a + k/a.31 After some manipulations of Eq. 10to obtain an expression in the form of Eq. 11, the summa-tion can be written as

    m=0

    n 2/1 + m

    m!

    N n + m

    n m + 1N 2n + m + 1

    =N n

    n + 1N 2n + 1

    2F1 n,N n;N 2n + 1 ; 21 +

    . 12The normalization constants M0,N and Mn,N can be simpli-

    fied as well, with the former given by N1/2, and the latterby

    32

    Mn,N = N n + 1N 2nn + 11/2

    . 13

    Combining terms, we write the FC overlap integral for

    eigenfunctions of two displaced Morses with identical

    strengths and ranges as

    I0,n

    = N/2

    2

    1 +

    Nn N n

    N 2n + 1

    N 2nN n + 1n + 1N

    1/2

    2F1 n,N n;N 2n + 1 ; 21 +

    . 14Equation 14 was compared for a number of parameterswith the numerical method of Lopez et al.,

    33and the values

    obtained agreed.

    For the multimode case at sufficiently low temperatures

    such that the quantum mode is not thermally populated, all

    that remains is to substitute I0,n in Eq. 3 for the appropriatemode. For the single-mode case, the simpler form of Eq. 2can be used after finding an expression for n and a suitable

    form for the product DOS. From the energy eigenvalues of

    the Morse potential, it can be shown that for a transition

    from the n = 0 initial state, n is given by energy conservation

    as

    n = 1

    2 2D

    D +

    1

    4, 15

    where =2D/ , with the reduced mass for the os-cillator. We use the classical form of the DOS Ref. 34 by

    taking n continuous and differentiating with respect to the

    energy E,35,36

    E =dn

    dE= 2

    D E1 . 16

    Thus, for a single anharmonic mode in the low-temperature

    regime, the transition-probability rate is given by

    WRP = 2

    HRP2I0,n2

    2

    D 1

    . 17

    III. NUMERICAL SOLUTION OF THE SCHRDINGEREQUATION

    In order to investigate the results of the previous section,

    one-dimensional wavefunctions were propagated on a nu-

    merical Cartesian grid with equally spaced points. The meth-

    ods are only outlined here as numerous reviews have been

    written on the subject.3739

    The Schrdinger equation given

    by

    i

    r

    ,tt

    = Hr,t 18

    has the formal solution

    r,t = eiH

    t/r,0 , 19

    where the Hamiltonian is written in the usual way, H = T+ V.

    The potential operator V is local in coordinate space, and its

    action is given simply by Vxx. The operation in the

    kinetic term T= 2/2m2/x2 can be calculated with theFourier method

    40,41by a discrete fast Fourier transform

    FFT of the wave function into momentum space followedby multiplication by K

    i

    2, where Ki is the wave number in

    the frequency domain, and finally an inverse FFT back into

    coordinate space. With the use of the FFT algorithm, the

    calculation of T scales efficiently as ONlog N, where N isthe number of grid points.

    39The exponential in Eq. 19 is

    then expanded using a global propagator through the New-

    tonian interpolation method,38,42

    where the sampling points

    are taken to be the zeros of the Chebychev polynomials.

    The initial wave function was prepared as the ground

    eigenstate of the initial potential as obtained through propa-

    gation in imaginary time.43

    As shown in Fig. 1, two states,

    harmonic or Morse, were used and the dependence of the

    transition-probability rate on the exoergicity was probed by

    lowering the final potential in small increments of . Thepopulation in the initial state was monitored over the course

    of the propagation and then fit to an exponential, i2t

    =Aect, where c was taken to be the transition-probability

    rate. Since the rate expressions arise from the first-order per-

    turbation theory result, the population at short times was

    used. By applying the single-mode Morse treatment from

    Sec. II and the analogous result for harmonic oscillators,44,45

    the perturbation theory results are compared with the nu-

    merical propagations Fig. 2. The numerical propagationpredicts a smaller transition probability than the golden rule

    result at small values of . This numerical calculation is

    somewhat artificial since a dephasing process is implicit in

    044108-3 Nonadiabatic electron transfer J. Chem. Phys. 124, 044108 2006

    Downloaded 28 May 2009 to 129.105.215.213. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp

  • 8/3/2019 Sina Yeganeh and Mark A. Ratner- Effects of anharmonicity on nonadiabatic electron transfer: A model

    4/9

    smoothing the delta function from Eq. 3. When becomessubstantially greater than c, the Morse results from these

    two treatments agree as the density of states becomes large.

    The chosen potentials for this quantum mode correspond

    to a stretch with the final state at larger displacement than the

    initial state. While the symmetry of the harmonic potential

    makes the distinction between a stretching or compression

    mode irrelevant in calculating the FC factor and transition

    probability, this distinction has an important effect for a

    Morse potential.11

    The plot for the Morse oscillator in this

    case exhibits a flatter dependence on exoergicity and is lower

    valued than in the harmonic case, as is expected from the FC

    factor with r0r0. For a compression mode with r0r0 the

    opposite would be true as the plot would be more sharply

    peaked and higher valued than the corresponding harmonic,

    as seen in Fig. 3. In addition, the computed transition prob-

    ability stops abruptly at D for the Morse potential, as

    energy conservation requires a transition from a bound to

    unbound state when approaches D. The expression in Eq.

    14 holds only for bound states, when n N/ 2. Beyondthis point, the FC factor for a bound/unbound state of the

    anharmonic oscillator is required.

    IV. TWO-MODE TREATMENT

    Although the single-mode treatment was found to be suf-ficient in treating some situations such as the chromatium

    data,12

    it is often the case that the relevant modes in a system

    can be separated into high-frequency quantum modes and a

    collection of low-frequency phonon modes. The low-

    frequency modes can be treated accurately as an average

    FIG. 1. Initial V and final V potentials are shown for a harmonic and b Morse cases with parameters shown. All values are in a.u. unless otherwisestated. The electronic coupling matrix element HRP was taken as 0.02 eV. The reduced mass c was 2000 a .u. The potentials were constructed such that the

    frequencies, Morse2D/c and Harmonic =kH/c, would be the same and equal to c 1870 cm1. The Morse potential has 26 bound states. Thereorganization energies are i

    H=0.11 eV and iM=0.18 eV. The ground-state eigenfunction for each case is also shown - - -. The exoergicity is

    represented as the bottom-bottom vertical separation of the potentials.

    FIG. 2. Golden rule results and wave-packet propagations for the transition

    probability as a function of for a single-quantum mode from Fig. 1.

    Harmonic potentials: golden rule result and wave packet ; Morsepotentials: golden rule result - - - and wave packet . Propagationparameters are given: nr is the number of spatial grid points, dr is the

    distance between grid points, nt is the number of time steps taken for each

    propagation, dt is the length of a time step, and nint is the number of inter-

    polation points taken in the Newton scheme.

    FIG. 3. Transition probability for single harmonic mode , Morse stretch-ing mode - - - with r0 r0 =1.5, and Morse compression mode withr0 r0 =1.5 potentials from Fig. 1 are plotted.

    044108-4 S. Yeganeh and M. A. Ratner J. Chem. Phys. 124, 044108 2006

    Downloaded 28 May 2009 to 129.105.215.213. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp

  • 8/3/2019 Sina Yeganeh and Mark A. Ratner- Effects of anharmonicity on nonadiabatic electron transfer: A model

    5/9

    over many modes of similar frequencies, but incorporating

    the effect of many quantum modes explicitly changes the

    transition probability significantly.46

    We focus our attention

    on a system with only one relevant quantum mode with fre-

    quency c and one collective solvent phonon mode with fre-

    quency s. For harmonic oscillators in both modes, Jortner12

    has identified several limiting cases in which simpler forms

    for the ET transition probability can be written, and we pro-

    ceed analogously. Ulstrup and Jortner11

    have shown that an-harmonicity and frequency change can be treated by simply

    substituting the appropriate FC overlap integrals and energy

    eigenvalues into the harmonic form.

    a For the zero-temperature limit, when kTsc,

    12

    WRPHarmonic =

    2

    HRP2

    sm=0

    ScmeScm!

    SspmeSspm!

    , 20where pm is taken as mc/s and Sc and Ss arethe Huang-Rhys factors for the quantum and solvent modes,

    respectively, given in terms of the inner- and outer-sphere

    reorganization energies,

    i and

    o, by Sc =

    i/c and Ss=o/s with i =cc

    2r0 r02 / 2 and o =ss

    2z0z0

    2 /2. In the summation, the two groups of terms in pa-

    rentheses are the squared FC overlap integrals for the quan-

    tum and solvent modes, respectively. We can incorporate the

    effect of anharmonicity in the quantum mode by

    substituting11

    I0,m from Eq. 14 to get

    WRPMorse =

    2

    HRP2

    sm=0

    N/2

    I0,m2SspmeSs

    pm! , 21

    where pm is Emm +Em0/s and Emn is thenth Morse energy eigenvalue. This result can be compared

    with that obtained from the zero-temperature limit of Eq. 3with two modes,

    WRPMorse =

    2

    HRP

    2m=0

    N/2

    I0,m2

    n=0

    Ssn

    eSs

    n!

    LEmm Em0 + n s , 22

    where the delta function has been replaced by the Lorentzian,

    xLx =//x2 +2.12,47 Note that for the Morse thesummation over m is only over the bound levels formally,

    the full spectrum of possible transitions should be repre-

    sented by the sum over bound levels and an integral over the

    dissociative states, but the latter will typically be small in

    comparison and is ignored throughout. The results of bothexpressions are shown in Fig. 4. The substitution of Morse

    FC factors and eigenvalues into Eq. 20 Jortners12 Eq. 3gives the same result as the more general form of Eq. 22;the similarity of the two results can be understood by rewrit-

    ing the delta function,

    Emm Em0 + n s

    =1

    s Emm + Em0

    s n , 23

    where the delta function is nonzero only when n =pm, andso the summation over m and n is subsumed into a single

    sum over coupled values of m and pm. The rounding ofpm or pm to the nearest integer gives width to the deltafunction,

    48as does the substitution of the Lorentzian, and so

    the results are nearly identical.

    The oscillatory structure in Fig. 4 is an interesting fea-

    ture that will always be present at zero temperature when

    there are two modes of significantly different frequencies and

    io.11

    For the harmonic oscillator, the peaks occur at in-

    teger multiples of c shifted by o. The Morse oscillator

    plot shows similar spacing at small values of , but for

    larger values of exoergicity the peaks become more closely

    spaced together, corresponding to the DOS approaching a

    continuum at the dissociative limit. In addition, as expected

    from the stretching nature of the quantum mode, the anhar-

    monic system has a smaller transition rate near the optimal

    regime and broader dependence on the exoergicity. At higher

    temperatures, broadening of the peaks is expected as excited

    states in the solvent contribute more allowed combinations

    of m and n which are energy conserving.

    b For the low-temperature regime with skTc, the solvent mode is treated classically, resulting in

    the following expression for harmonic oscillators in both

    modes:12

    WRPHarmonic =

    2

    HRP

    2 m=0

    exp o m c24okT

    eScScm

    m!4okT1/2. 24

    The corresponding expression for a Morse oscillator can be

    written once again by a simple substitution of the appropriate

    FC overlap integrals and energies,

    FIG. 4. The zero-temperature transition probability for two modes with

    values from Fig. 1 for harmonic and Morse oscillators in the quantum mode

    and with the solvent mode parameters shown. The solvent frequency s is

    approximately 30 cm1 and o =0.006 eV. The results of Eq. 20 andEq. 21 - - - as well as with a Lorentzian, Eq. 22, for a harmonic and Morse mode are plotted. For the Lorentzian, =1105 hartree.

    044108-5 Nonadiabatic electron transfer J. Chem. Phys. 124, 044108 2006

    Downloaded 28 May 2009 to 129.105.215.213. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp

  • 8/3/2019 Sina Yeganeh and Mark A. Ratner- Effects of anharmonicity on nonadiabatic electron transfer: A model

    6/9

    WRPMorse =

    2

    HRP

    2

    m=0

    N/2

    exp o Emm + Em024okT

    I0,m

    24okT1/2. 25

    For the Morse result with a Lorentzian and the exact form for

    the solvent modes, we make use of Eq. 3:

    WRPMorse =

    2

    HRP

    2

    m=0

    N/2

    I0,m2

    n=0

    n=0

    In,nH 2

    1 es/kT

    esn/kT

    LEmm Em0 + n n s , 26

    where In,n

    His the FC overlap integral for the harmonic oscil-

    lators in the solvent mode given by45,49

    In,n

    H 2 = n ! n ! SsnneSs

    j=0

    n Ss

    j

    j ! n j ! n n + j!2 . 27

    Figure 5 shows the result for the transition probability at

    150 K. The quantum and solvent modes have characteristic

    temperatures, Tc = /k, of 2700 and 43 K, respectively, so

    in this regime the low-temperature approximation for quan-

    tum modes and classical approximation for the solvent

    modes are appropriate. Although broader than in Fig. 4, os-

    cillations are still visible; Ulstrup and Jortner placed the con-

    dition on the presence of these peaks as 2okTc,which is satisfied for the parameters in Fig. 5 with the small

    value of o chosen. The low-temperature regime described

    here is valid for many temperatures of interest in biological

    systems.

    c For higher temperatures such that skT c,the full form of the Morse FC overlap integral In,n is re-

    quired as the molecular modes become thermally excited.

    This general equation can be obtained by similar manipula-

    tions of Eq. 8 to give

    In,n =N/2n

    N 2n + 1 2

    1 + Nnn N 2nN 2nn + 1N n + 1N n + 1

    n + 11/2

    m=0

    n 21 +

    m N n n + mm + 1n m + 1N 2n + m + 1 2

    F1 n,N n n + m;N 2n + 1 ; 21 +

    . 28

    Equipped with this result, transition rates at any temperature

    can be calculated within the classical approximation for thesolvent modes. First, for comparison, the harmonic oscillator

    expression can be reduced to12,49

    WRPHarmonic =

    2

    HRP

    2exp Sc2v+ 1

    4okT

    m=

    ecm/2kTIm2Scvv+ 1

    exp o m c24okT

    . 29

    where v= expc/kT 11 and Im is the modified Bessel

    function. We write the Morse ET transition probability,

    WRPMorse =

    2

    HRP

    24okT1/2

    m=0

    N/2

    m=0

    N/2

    exp o Emm + Emm24okT

    Im,m

    2eEmm/kT

    Qm, 30

    where Qm is the Morse partition function, which we take as

    FIG. 5. The transition probability in the low-temperature regime. The same

    potentials as in Fig. 4, at a temperature of 150 K. The classical approxima-

    tion for the solvent modes with Eq. 24 for the harmonic quantummode or Eq. 25 - - - for a Morse quantum mode as well as the full sumover both modes using Eq. 3 for the harmonic and Morse mode are plotted. =2.5105 hartree.

    044108-6 S. Yeganeh and M. A. Ratner J. Chem. Phys. 124, 044108 2006

    Downloaded 28 May 2009 to 129.105.215.213. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp

  • 8/3/2019 Sina Yeganeh and Mark A. Ratner- Effects of anharmonicity on nonadiabatic electron transfer: A model

    7/9

    j=0N/2expEmj/kT. For the Lorentzian result, we use the

    full two-mode form of Eq. 3,

    WRPMorse =

    2

    HRP

    2m=0

    N/2

    m=0

    N/2

    Im,m2

    eEmm/kT

    Qm

    n=0

    n=0

    In,n

    H 21 es/kT

    esn/kT

    LEmm Emm + n n s . 31

    The results of these expressions at 300 K are given in Fig. 6.

    There is significantly more broadening at this temperature as

    is expected from the 1/T temperature dependence in the

    Gaussian.

    Examining Figs. 5 and 6, we note the following things.

    In Fig. 5, there is good agreement between the classical ap-

    proximation for the solvent modes and the fully quantum-

    mechanical treatment of the solvent modes. In Fig. 6, there is

    also good agreement between the two forms; at this higher

    temperature, the numerical calculation of Eq. 27 for thesolvent modes is avoided for large values of n and n by

    using the recursion relations for the harmonic oscillator FC

    factors.45,50

    V. CONCLUSIONS

    Applying the theory of nonradiative multiphonon transi-

    tions to nonadiabatic ET has yielded forms for the transition

    probability that incorporate anharmonicity in the molecular

    mode. The anharmonicity is treated by substituting the

    Morse FC overlap integral and energies into the analogous

    relations for harmonic oscillators. At low temperatures and

    small values of o, interesting oscillatory behavior in the

    transition probability as a function of exoergicity is exhib-ited. Regardless of whether the mode corresponds to a stretch

    or compression, significant deviations are seen from the har-

    monic result for a relatively anharmonic mode. For increas-

    ing values of, as the density of states for the Morse oscil-

    lator rapidly increases, the peaks will no longer be equally

    separated by approximately c but will move closer to each

    other. As other authors have pointed out,11,18

    in a system with

    two anharmonic quantum modes with opposite displace-

    ments, a stretching and compression mode, the changes in

    the FC factors from the harmonic values will roughly cancel

    and anharmonic effects will disappear. However, at low tem-

    peratures and small o such that the oscillatory structure dis-cussed in Sec. IV is present, anharmonic effects will be vis-

    ible through the nonuniform spacing of peaks. We are not

    aware, however, of a condensed-phase experimental system

    which possesses these characteristics.

    Turning our attention to a more realistic system with

    o =0.16 eV at 150 K, the transition probability for a har-

    monic quantum mode is contrasted with stretching and com-

    pression anharmonic modes in Fig. 7. One significant feature

    is in the inverted regime: the transition probability for the

    Morse compression mode falls off steeply as a Gaussian aswould be predicted from a classical harmonic oscillator treat-

    ment, while the harmonic oscillator and Morse stretchingmode are described by exponential decay. The sharp drop

    leads to an ET rate that is several orders of magnitude

    smaller than for a harmonic or stretching mode, even at

    =0.5 eV where the inverted regime is just beginning. In ad-

    dition, the stretching mode has a significantly broader opti-

    mal regime than the harmonic case, and in the inverted re-

    gime this results in a rate increase of an order of magnitude

    relative to the harmonic case. At this larger value of o, all

    oscillatory structure is smoothed outfor distinct peaks a

    temperature of 10 K is required. For this single anharmonic

    mode, we see that the optimal exoergicity opt can be or-

    dered as: optcompression

    optstretch.

    51We cannot in general predict

    the position of

    opt

    harmonic

    relative to the Morse optimal exoer-gicities because this relation depends on the magnitude of the

    horizontal displacement r0 r0. For harmonic oscillators athigh temperatures for which Jortners two-mode form con-

    verges to the Marcus expression, the optimal regime is given

    FIG. 6. The transition probability with excited initial states in the quantum

    modes at 300 K same values as Fig. 4. The full form of Eq. 3 using theappropriate partition functions is plotted for harmonic and Morse potentials =1104 hartree as well as Eq. 29 for the harmonic and Eq. 30 for the Morse - - - mode.

    FIG. 7. Logarithmic plot of transition probability in the low-temperature

    regime at 150 K. Same as Fig. 5 with o =0.16 eV. Using Eqs. 24 and25, a harmonic mode , Morse stretching mode - - -, and Morsecompression mode are plotted.

    044108-7 Nonadiabatic electron transfer J. Chem. Phys. 124, 044108 2006

    Downloaded 28 May 2009 to 129.105.215.213. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp

  • 8/3/2019 Sina Yeganeh and Mark A. Ratner- Effects of anharmonicity on nonadiabatic electron transfer: A model

    8/9

    by optharmonic =o +i

    H with iH= kHr0 r0

    2 /2. This relation-

    ship can be used to approximate the optimal regime at lower

    temperatures. The Morse reorganization energy is given by

    iM=D1 er0r02, but opt

    Morse o +iM is a poor approxi-

    mation to the actual optimal region. This is due to the fact

    that the governing factor for the optimal region is the exoer-

    gicity at which energy conservation requires the maximum

    FC factor within o, and this will not necessarily peak nearthe activationless exoergicity given by i. While the har-

    monic oscillator FC factor does peak close to iH, the maxi-

    mum in the Morse FC overlap integral is not well approxi-

    mated by

    i

    M

    . An alternative prediction of the location of theoptimal region is not possible without specific parameters.

    We can predict that the ET transition probability at the opti-

    mal regime for the Morse compression mode will decrease

    much more gradually than the harmonic and the stretching

    mode as a function of increasing displacement r0 r0, sug-gesting a possible role for anharmonic compression modes in

    nonadiabatic ET reactions with large inner-sphere reorgani-

    zation.

    In Fig. 8 the transition probability for a system with two

    quantum modes and one solvent mode is plotted; the two

    Morse quantum modes are chosen to be stretching and com-

    pression modes, while the two harmonic modes are identical.

    The combination of compression and stretch causes theMorse and harmonic results to be nearly identical at low

    values of, but at higher exoergicities the Morse result is

    dominated by the stretching mode, and a more gradual de-

    crease in transition rate is seen. Even at the higher o, anhar-

    monic effects are noticeable. This effect is even more pro-

    nounced at larger values of i.

    It is interesting to consider the case of displaced-

    distorted harmonic oscillators in the quantum mode where

    both displacement and frequency change are considered. Us-

    ing the appropriate FC factors,52

    the effect of frequency

    change can be compared to the anharmonic effect, but while

    frequency change does slightly shift the W vs plot to

    higher or lower exoergicities depending on the sign of RP, the shape of the relation is generally the same. Thus,displaced-distorted harmonic oscillators cannot account for

    the anharmonic deviations of broadening or narrowing for

    Morse stretching or compression modes discussed earlier.

    In this paper, we have written the FC factors for Morse

    oscillators in condensed form using the hypergeometric 2F1function, and the expected deviations from harmonic behav-

    ior for anharmonic stretching and compression modes havebeen discussed. We predict that significant differences in the

    electron transfer rate constant are expected, especially in the

    inverted regime. As seen in the logarithmic plot in Fig. 7, the

    Morse stretching and compression modes will differ by many

    orders of magnitude from the harmonic result. In addition, at

    large values ofi the Morse compression mode is expected

    to have a much larger transition rate than the harmonic and

    Morse stretching modes. It is expected that these anharmonic

    effects will be significant and experimentally observable in a

    properly constructed system.

    ACKNOWLEDGMENT

    The authors would like to thank Abraham Nitzan for

    helpful discussions. This work was supported by the Chem-

    istry Division of the NSF and the MURI/DURINT program

    of the DoD.

    1R. A. Marcus, J. Chem. Phys. 24, 966 1956.

    2R. A. Marcus, Discuss. Faraday Soc. 29, 21 1960.

    3R. A. Marcus, Annu. Rev. Phys. Chem. 15, 155 1964.

    4For the harmonic case with same frequencies the entropy change van-

    ishes, and G = TS = . This, however, is not true for the case of

    anharmonic oscillators with a nonuniform density of states. We therefore

    consider only the transition probability dependence on exoergicity and

    not the free-energy change.5

    D. DeVault and B. Chance, Biophys. J. 6, 825 1966.6 K. Huang and R. Rhys, Proc. R. Soc. London, Ser. A 204, 406 1950.7

    R. Kubo, Phys. Rev. 86, 929 1952.8

    M. Lax, J. Chem. Phys. 20, 1752 1952.9

    T. Kakitani and H. Kakitani, Biochim. Biophys. Acta 635, 498 1981.10

    N. R. Kestner, J. Logan, and J. Jortner, J. Phys. Chem. 78, 2148 1974.11

    J. Ulstrup and J. Jortner, J. Chem. Phys. 63, 4358 1975.12

    J. Jortner, J. Chem. Phys. 64, 4860 1976.13

    M. Marchi, J. N. Gehlen, D. Chandler, and M. Newton, J. Am. Chem.

    Soc. 115, 4178 1993.14

    P. M. Morse, Phys. Rev. 34, 57 1929.15

    M. D. Sturge, Phys. Rev. B 8, 6 1973.16

    N. Brniche-Olsen and J. Ulstrup, J. Chem. Soc., Faraday Trans. 2 74,

    1690 1978.17

    N. Brniche-Olsen and J. Ulstrup, J. Chem. Soc., Faraday Trans. 1 75,

    205 1979.18

    N. C. Sndergaard, J. Ulstrup, and J. Jortner, Chem. Phys. 17, 417

    1976.19J. Ulstrup, Charge Transfer Processes in Condensed Media, Lecture

    Notes in Chemistry, Vol. 10 Springer-Verlag, Berlin, 1979.20

    R. Islampour and S. H. Lin, Chem. Phys. Lett. 179, 147 1991.21

    R. Islampour and S. H. Lin, J. Phys. Chem. 95, 10261 1991.22

    The hypergeometric function appears often when dealing with matrix

    elements of the Morse eigenstates. See, for example, N. M. Avram and G.

    E. Draganescu, Int. J. Quantum Chem. 64, 655 1997; D. A. Morales, J.Math. Chem. 22, 255 1997; E. F. de Lima and J. E. M. Hornos, J. Phys.B 38, 815 2005.

    23See, for example, C. Cohen-Tannoudji, B. Diu, and F. Lalo, Quantum

    Mechanics Wiley, New York, 1977, Vol. 2, p. 1299.24

    E. Condon, Phys. Rev. 32, 858 1928.25

    P. F. Barbara, T. J. Meyer, and M. A. Ratner, J. Phys. Chem. 100, 13148

    1996.26

    A. M. Kuznetsov and J. Ulstrup, Electron Transfer in Chemistry and

    FIG. 8. Transition probability with two quantum modes and one solvent

    mode at 150 K. Using Eqs. 24 and 25 with an additional quantum modesame D and with r0 r0 =1.5, harmonic and Morse - - - witho =0.006 eV, and harmonic and Morse with o =0.16 eV areplotted.

    044108-8 S. Yeganeh and M. A. Ratner J. Chem. Phys. 124, 044108 2006

    Downloaded 28 May 2009 to 129.105.215.213. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp

  • 8/3/2019 Sina Yeganeh and Mark A. Ratner- Effects of anharmonicity on nonadiabatic electron transfer: A model

    9/9

    Biology: An Introduction to the Theory Wiley, Chicester, UK, 1999,Chap. 8.

    27F. Iachello and M. Ibrahim, J. Phys. Chem. A 102, 9427 1998.

    28P. A. Fraser and W. R. Jarmain, Proc. Phys. Soc., London, Sect. A 66,

    1145 1953.29

    W. R. Jarmain and P. A. Fraser, Proc. Phys. Soc., London, Sect. A 66,

    1153 1953.30

    Handbook of Mathematical Functions with Formulas, Graphs, and Math-

    ematical Tables, 9th ed., edited by M. Abramowitz and I. Stegun Dover,New York, 1972, Chap. 15.

    31For negative values as we have for n

    m, the Pochhammer symbol is

    represented by ak=a + 1/a k+ 1.32

    We assume that N is not precisely an even integer.33

    J. C. Lopez, A. L. Rivera, Y. F. Smirnov, and A. Frank, Int. J. Quantum

    Chem. 88, 280 2002.34

    Formally, the density of states is given by En = En+1 En1. To ob-

    tain a simpler form, we take the classical density of states since the

    results are nearly identical except very near the dissociative limit when

    EnD.35

    P. C. Haarhoff, Mol. Phys. 7, 101 1963.36

    J. P. K. Doye and D. J. Wales, J. Chem. Phys. 102, 9659 1995.37

    R. Kosloff, J. Phys. Chem. 92, 2087 1988.38

    R. Kosloff, Annu. Rev. Phys. Chem. 45, 145 1994.

    39R. Kosloff, in Dynamics of Molecules and Chemical Reactions, edited by

    R. Wyatt and J. Zhang Marcel Dekker, New York, 1996, pp. 185230.40

    M. D. Feit, J. A. Fleck, Jr., and A. Steiger, J. Comput. Phys. 47, 412

    1982.41

    D. Kosloff and R. Kosloff, J. Comput. Phys. 52, 35 1983.42

    G. Ashkenazi, R. Kosloff, S. Ruhman, and H. Tal-Ezer, J. Chem. Phys.

    103, 10005 1995.43

    R. Kosloff and H. Tal-Ezer, Chem. Phys. Lett. 127, 223 1986.44

    E. Hutchisson, Phys. Rev. 36, 410 1930.45

    C. Manneback, Physica Amsterdam 17, 1001 1951.46

    A. Sarai, Chem. Phys. Lett. 63, 360 1979.47 The parameter for the Lorentzian was found by taking the harmonic-

    oscillator result at different values of and comparing with the equiva-

    lent expression from Jortner 1976 until the results were nearly identical.48

    J. J. Markham, Rev. Mod. Phys. 31, 956 1959.49

    D. DeVault, Quantum-Mechanical Tunnelling in Biological Systems

    Cambridge University Press, Cambridge, 1984.50

    J. Lerm, Chem. Phys. 145, 67 1990.51

    This relation is true as long as r0 r0 is not too large, but for largeseparations the relation will switch and opt

    compressionopt

    stretchat this large

    displacement, however, the transition rate is nearly zero for the stretch

    mode, and the discussion of optimal region is meaningless.52

    W. Siebrand, J. Chem. Phys. 46, 440 1967.

    044108-9 Nonadiabatic electron transfer J. Chem. Phys. 124, 044108 2006