DataPDF Available

KT13JPCB

Authors:

Figures

Content may be subject to copyright.
Quantum Suppression of Ratchet Rectication in a Brownian System
Driven by a Biharmonic Force
Akihito Kato
and Yoshitaka Tanimura*
,,
Department of Chemistry, Graduate School of Science, Kyoto University, Kyoto606-8502, Japan
Universität Augsburg, Institut für Physik, Universitätsstrasse 1, 86135 Augsburg, Germany
ABSTRACT: We rigorously investigate the quantum dis-
sipative dynamics of a ratchet system described by a periodic
potential model based on the CaldeiraLeggett Hamiltonian
with a biharmonic force. In this model, we use the reduced
hierarchy equations of motion in the Wigner space
representation. These equations represent a generalization of
the GaussianMarkovian quantum FokkerPlanck equation
introduced by Tanimura and Wolynes (1991), which was
formulated to study non-Markovian and nonperturbative
thermal eects at nite temperature. This formalism allows
us to treat both the classical limit and the tunneling regimes,
and it is helpful for identifying purely quantum mechanical
eects through the time evolution of the Wigner distribution.
We carried out extensive calculations of the classical and quantum currents for various temperatures, coupling strengths, and
barrier heights. Our results reveal that at low temperature, while the quantum current is larger than the classical current in the
case of a high barrier, the opposite is true in the case of a low barrier. We nd that this behavior results from the fact that the
tunneling enhances the current in the case of a high barrier, while it suppresses the current in the case of a low barrier. This is
because the eect of the ratchet potential is weak in the case of a low barrier due to the large dispersion of the distribution
introduced by tunneling. This causes the spatiotemporal asymmetry, which is necessary for ratchet current, to be weak, and as a
result, the net current is suppressed.
1. INTRODUCTION
A system that is able to rectify thermal or mechanical
uctuations through a periodic potential is called a ratchet.
18
With recent advances in microscopy and nanotechnology,
several new mechanisms of rectication of uctuations resulting
in transport have been identied through theoretical works
919
and experimental works
2030
in biology, physics, and chemistry.
Well-known examples include asymmetric quantum dots,
2022
vortices in superconductors,
2325
cold atoms in asymmetric
optical lattices,
2628
and molecular motors in biological
systems.
2931
A ratchet system can produce a directed current
from uctuations only when the system is in a nonequilibrium
state because the extraction of work from unbiased uctuations
is not allowed by the second law of thermodynamics. Moreover,
to realize the ratchet eects, it is necessary to break the
spatiotemporal symmetries because otherwise the contributions
from positive and negative currents would cancel.
3236
Quantum mechanical eects may also play a role because
ratchet systems must be of microscopic size to rectify the
uctuating motion.
3747
A FokkerPlanck approach
48
and a Langevin approach
49
have been used in classical studies of ratchets, while varieties of
equation of motion approaches have been employed for
quantum mechanical studies.
3747
In the quantum mechanical
case, dissipative systems are commonly modeled as potential
systems coupled to heat-bath degrees of freedom at nite
temperature. This coupling gives rise to thermal uctuations
and dissipation that drive the systems toward the thermal
equilibrium state. The heat-bath degrees of freedom are then
reduced using such methods as the projection operator method
and the path integral method, for example. The quantum
Langevin equation
37
and the quantum FokkerPlanck
equation
42
have been used for the purpose of understanding
the quantum aspects of ratchet dynamics in a heat bath.
Although these equations are analogous to the classical kinetic
equations, which have proved to be useful in the treatment of
classical transport problems, such equations cannot be derived
in a quantum mechanical framework without signicant
approximations or assumptions. For example, in dealing with
the quantum Langevin equation expressed in operator form, it
is generally assumed that the antisymmetric correlation
function of the noise is positive, but this is valid only for
slow non-Markovian modulation at high temperature, as we
demonstrate in Section 2. Similarly, the quantum Fokker
Planck equation can be derived from the CaldeiraLeggett
Special Issue: Peter G. Wolynes Festschrift
Received: March 28, 2013
Revised: April 23, 2013
Published: May 2, 2013
Article
pubs.acs.org/JPCB
© 2013 American Chemical Society 13132 dx.doi.org/10.1021/jp403056h |J. Phys. Chem. B 2013, 117, 1313213144
Hamiltonian under a Markovian approximation,
5155
but for
this to be possible, the heat bath must be at a suciently high
temperature, in which case quantum tunneling processes play a
minor role. The quantum master equation expressed in terms
of Floquet states provides a more rigorous treatment of
quantum dissipative dynamics than the methods mentioned
above,
46
but it can only be applied to systems possessing weak
interactions.
50
The methods discussed to this point represent
the main approaches used in the treatment of ratchet systems
carried out to this time. However, as we have pointed out, none
of them provide fully quantum mechanical descriptions of
broad validity.
We employ the reduced hierarchy equations of motion
(HEOM) in the Wigner space representation,
56
which are a
generalization of the GaussianMarkovian quantum Fokker
Planck equation introduced by Tanimura and Wolynes
57,58
to
study non-Markovian and nonperturbative thermal eects at
nite temperature.
5963
Because the HEOM are derived from
the full system-bath Hamiltonian with no approximation, the
entire system approaches a thermal equilibrium state at nite
temperature when no external perturbation is applied.
56,64
This
means that the second law of thermodynamics holds within the
HEOM approach, and hence there can be no nite current in
the absence of a driving force, which is the essential behavior to
investigate the ratchet rectication of thermal uctuations. The
HEOM have been used to study a variety of phenomena and
systems, including exciton transfer,
6569
electron transfer,
7073
quantum information,
74,75
and resonant tunneling diodes.
62,63
The HEOM are ideal for studying quantum transport systems
when implemented using the Wigner representation because
they allow the treatment of continuous systems utilizing open
boundary conditions and periodic boundary conditions.
76
Elucidation of the dynamical behavior of the system through
the time evolution of the Wigner distribution functions is also
possible. The classical hierarchy equations of motion can be
obtained easily by taking the classical limit of the HEOM, and
utilizing the Wigner representation, we can easily compare
quantum and classical distribution functions.
58,61
Considering
both classical and quantum mechanical cases, we report the
results of numerical calculations of the ratchet current, which is
the induced net current resulting from ratchet eects, in a
dissipative environment for various temperatures and system-
bath coupling strengths. We then clarify the roles of
uctuations, dissipation, and the biharmonic force in both the
classical and quantum cases.
The organization of the paper is as follows. In Section 2, we
introduce the HEOM approach in the Wigner representation.
In Section 3, we discuss ratchet systems from the point of view
of space-time symmetries. In Section 4, the numerical results
obtained for the ratchet current in classical and quantum
mechanical cases are presented. Section 5 is devoted to
conclusions.
2. REDUCED HIERARCHY EQUATIONS OF MOTION
FORMALISM
Ratchet systems are often modeled by a Brownian particle in a
periodic potential under an ac or dc driving force. We take this
approach and employ a model based on the CaldeiraLeggett
(or Brownian) Hamiltonian
77
ω
ω
̂=̂+̂
+̂+̂̂
H
p
mUq t
p
m
mxaV q
m
2(; )
22
()
j
j
j
jj
j
j
jj
tot
2
22
2
2
(1)
Here m,p̂,q̂, and U(q̂,t) are the mass, momentum, position,
and potential of the particle, and mj,p̂j,x̂j, and ωjare the mass,
momentum, position, and frequency variables of the jth bath
oscillator mode. The quantities ajare coecients that depend
on the nature of the system-bath coupling. From eq 1, it is seen
that the interaction part of the Hamiltonian is assumed to take
the form ĤI=V(q̂)X̂, where V(q̂) is an arbitrary function of q̂
and X̂Σjajx̂jis the interaction coordinate. We introduced the
counter term Σjaj
2V(q̂)2/2mjωj
2to maintain the translational
symmetry of the Hamiltonian for U(q̂;0) = 0. Here we consider
areection-symmetric potential system driven by a biharmonic
force.
1419,4549
This force prevents the system from reaching
equilibrium and breaks the spatiotemporal symmetry. We write
the potential as U(q̂;t)=U(q̂)q̂F(t), where U(q̂) is static,
reection-symmetric potential, and q̂F(t) is the dynamic
potential. Note that the time average of F(t) is zero. The total
system obeys the von-Neumann equation (the quantum
Liouville equation).
After the bath degrees of freedom are traced out, the reduced
density matrix elements of the system are obtained in the path
integral form as
∫∫
ρ
ρ
ρ
′= ′ ′
×′′ ℏ
×−
qq t q q Dq Dq q q
qq t q q iS qt Fqq t
iS q t
(, ; ) d d ( , )
( , , ; , )exp( [ , ]/ ) [ , ; ]
exp( [ , ]/ )
0000
CS 00A
A(2)
where SA[q;t] is the action corresponding to the systems
Hamiltonian, ĤA=q̂2/2m+U(q̂;t), ρ(q0,q0) is the initial state
of the system at time t0, and ρCS(q,q,t;q0,q0) is the initial
correlation function between the system and the heat bath.
78
The eect of the bath is incorporated into the Feynman
Vernon inuence functional, which is written
′= −
×
Ψ− °
+−
×
×
Fq q t sV s
suisuVu
uCs uV u
[, ; ] exp 1d()
d2()()
d( )()
t
t
t
s
t
s
2
0
0
0(3)
where V×(t)V(q(t)) V(q(t)) and V°(t)V(q(t)) +
V(q(t)). Although the method we employ here can be used
with any form of V(q),
5961
in this paper, we consider the
linearlinear system bath coupling case, dened by V(q)=q.
The interaction coordinate X̂Σjajx̂jis regarded as a driving
force through the interaction
̂̂
q
X
. The canonical and
symmetrized correlation functions, respectively, are then
expressed as Ψ(t)βX̂;X̂(t)Band C(t)(1/
2){X̂(t),X̂(0)}B, where β=1/kBTis the inverse temperature,
X̂(t)isX̂in the Heisenberg representation, and ···Brepresents
the thermal average over the bath modes.
56,64
Using the
spectral density J(ω)=Σjaj
2δ(ωωj)/2mjωj, we can rewrite
these functions as
The Journal of Physical Chemistry B Article
dx.doi.org/10.1021/jp403056h |J. Phys. Chem. B 2013, 117, 131321314413133
ωω
ωω
Ψ
=
tJt() 2 d ()
cos( )
0(4)
and
ωω ω βω
=ℏ
⎜⎟
Ct J t() d ( )cos( )coth 2
0(5)
The function C(t) is analogous to the classical correlation
function of X(t) and corresponds to the correlation function of
the bath-induced noise (uctuations), whereas Ψ(t) corre-
sponds to dissipation. The function C(t) is related to Ψ(t)
through the quantum version of the uctuationdissipation
theorem, C[ω]=ωcoth(βω/2)/2Ψ[ω], which ensures that
the system evolves toward the thermal equilibrium state for
nite temperatures, trB{exp[βĤtot]}, in the case that there is
no driving force.
79
As shown in Figure 1, the noise correlation C(t) becomes
negative at low temperature due to the contribution of the
Matsubara frequency terms with ν=2π/βin the region of
small t. This behavior is characteristic of quantum noise. The
fact that the noise correlation takes negative values introduces
problems when the quantum Langevin equation is applied to
quantum tunneling at low temperature. We note that the
characteristic time scale over which we have C(t)<0is
determined by the temperature and is not inuenced by the
spectral distribution J(ω). Thus, the validity of the Markovian
(or δ(t)-correlated) noise assumption is limited in the quantum
case to the high-temperature regime. Approaches employing
the Markovian master equation and the Redeld equation,
which are usually applied to systems possessing discretized
energy states, ignore or simplify such tunneling contributions,
often through use of the rotating wave approximation (RWA).
The main problem created by the RWA is that a system treated
under this approximation will not satisfy the uctuation
dissipation theorem, and thus it may introduce signicant error
in the evolution of the system toward equilibrium. In the
classical limit, with tending to zero, C(t) is always positive.
We assume that spectral density J(ω) has an Ohmic form
with a Lorentzian cuto, that is
ωζ
π
γω
γω
=+
Jm
()
2
22 (6)
where ζis the system-bath coupling strength, which represents
the magnitude of damping, and γis the width of the spectral
density of the bath mode. The canonical and symmetrized
correlation functions then become
ζγ
Ψ
=γ−||
tm() e t(7)
and
ζγ β γ
β
ν
γν
=ℏℏ
γν−||
=
−||
⎜⎟
Ct m
() 2cot 2e14e
t
k
k
k
t
2
1
22
k
(8)
where νk=2πk/βis the kth Matsubara frequency.
80
Note that
in the high- temperature limit, βγ1, the noise correlation
function reduces to C(t)mζγeγ|t|/β. This indicates that the
heat bath oscillators interact with the system in the form of
GaussianMarkovian noise.
The reduced HEOM can be obtained by considering the
time derivative of the reduced density matrix elements given in
eq 2 in the Wigner representation,
56
as a generalization of the
GaussianMarkovian quantum FokkerPlanck equation in-
troduced by Tanimura and Wolynes.
57,58
If we choose Kso as
to satisfy νK=2πK/(β)ωc, where ωcis the characteristic
frequency of the system, the factor eνk|t|in eq 8 can be replaced
with the Dirac delta function, using the approximation νkeνk|t|/
2δ(t) (for kK+ 1), with negligible error. For the kernel
Figure 1. Symmetric correlation C(t)dened by eq 8 is depicted as a function of the dimensionless time tfor (a) the slow modulation case, γ=1,
and (b) the fast modulation case, γ= 10. The constants mζγ2/2 and are set to unity. Note that γcorresponds to the Ohmic (Markovian)
limit, as can be seen from eq 6. In each of the Figures, the inverse temperatures are, from top to bottom, β= 0.2, 0.5, 1.0, and 5. The noise
correlation C(t) becomes negative for (b), the fast modulation case, at low temperature (large β) due to the contribution of Matsubara frequency
terms.
The Journal of Physical Chemistry B Article
dx.doi.org/10.1021/jp403056h |J. Phys. Chem. B 2013, 117, 131321314413134
eqs 7 and 8, the number of hierarchy elements for γis denoted
by n, and the number of kth Matsubara frequencies is denoted
by jk.
56,81
Then, the HEOM can be expressed as
6163
γν
γ
ν
=− +Ξ
̂++
×+Φ
̂
++Θ
̂
̂
···
=
··· ···
+
=
··· + ··· ···
=
··· − ···
tWqpt n j
Wqpt Wqpt
WqptnWqpt
jW qpt
(, ; ) [ ]
(, ; ) [ (, ; )
(, ; )] (, ; )
(, ; ),
jj
n
k
K
kk
jj
n
jj
n
k
K
jj j
n
jj
n
k
K
kkk j j j
n
,,
() QM
1
,,
() ,,
(1)
1
,, 1,,
() 0,,
(1)
1
,, 1,,
()
K
KK
kK K
kK
1
11
11
1
3
(9)
where the quantum Liouvillian in the Wigner representation is
expressed as
76
π
−=
−′ ′
−∞
Wqp p
mq
Wqp
pUqp ptWqp
(, ) (, )
d
2(, ; ) (, )
W
QM
2
3
(10)
with UW(q,p;t)=20
drsin(pr/)[U(q+r/2;t)U(qr/
2;t)]. The other operators are dened as
Φ
̂
p(11)
ζγβγ
Θ̂≡+
ℏℏ
⎜⎟
pm
p2cot 2
0(12)
γζ
βν γ
Θ̂
m
p
2
()
k
k
2
22
(13)
for k> 0 and
ζ
β
βγ βγ γ
νγ
Ξ
̂≡− ℏℏ
=
⎜⎟
m
p
12cot 2
2
k
K
k
1
2
22
2
2
(14)
Note that the zeroth element is identical to the Wigner
function W0,0,···,0
(0) (q,p;t)W(q,p;t), and the other elements are
introduced in the numerical calculations in order to treat the
nonperturbative, non-Markovian system-bath interaction.
Although these elements do not have direct physical meaning,
they allow us to take into account the quantum coherence and
entanglement between the system and the bath.
74,75
The
importance of the system-bath coherence was pointed out in
the context of correlated initial conditions,
78
and it was shown
with a nonlinear response theory approach that the system-bath
coherence plays an essential role in the case that the system is
driven by a time-dependent external force.
56
The HEOM consist of an innite number of equations, but
they can be truncated at nite order with negligible error.
57,58
Essentially, the condition necessary for the error introduced by
the truncation to be negligibly small is that the total number of
hierarchy elements or the total number of Matsubara
frequencies retained be suciently large. Explicitly, it can be
shown that the condition Nn+Σk=1
Kjkωc/min(γ,ν1)is
sucient for this purpose.
81
For the CaldeiraLeggett
Hamiltonian, the equations of motion are then truncated by
using the terminators, expressed in the Wigner representation
as
6163
=− +Ξ
̂
··· ···
tWqpt Wqpt(, ; ) ( ) (, ; )
jj
n
jj
n
,,
() QM , ,
()
KK11
3
(15)
with n+Σk=1
Kjk=N. The HEOM formalism can be used to
treat a strong system-bath coupling nonperturbatively. It is ideal
for studying quantum transport systems when employing the
Wigner representation because it allows the treatment of
continuous systems utilizing open boundary conditions and
periodic boundary conditions.
76
In addition, the formalism can
accommodate the inclusion of an arbitrary time-dependent
external eld while still accounting for the system-bath
coherence. Note that such coherence cannot be described if
we assume that the state of the total system takes a factorized
system-bath form.
56
Such system-bath coupling features are
necessary to properly treat quantum ratchet systems. In the
white noise (or Markovian) limit, γ, which is taken after
imposing the high-temperature approximation, βγ1, the
quantum FokkerPlanck equation can be derived in a similar
form as the Kramers equation,
51,52
which is identical to the
quantum master equation without RWA.
82
Because we assume
βγ1 with γ, this equation cannot be applied to low-
temperature systems, where quantum eects play a major role.
The Wigner distribution function reduces to the classical one
in the limit ℏ→0, and hence we can directly compare the
quantum results with the classical results. The classical HEOM,
which is derived from eq 9 by setting ℏ→0, is given by
57,58
γ
γζ β
=− +
+
++
+
tWqpt nWqpt
pWqpt
np m
pWqpt
(, ; ) ( ) (, ; )
(, ; )
(, ; )
nn
n
n
() CL ()
(1)
(1)
3
(16)
and the classical terminator is
γ
ζβ
γζ β
=− +
+
+
++
tWqpt NWqpt
ppm
pWqpt
Np m
pWqpt
(, ; ) ( ) (, ; )
(, ; )
(, ; )
NN
N
N
() CL ()
()
(1)
3
(17)
The classical Liouvillian is dened by
−=
+
Wqp p
mq
Wqp Uq t
qp
Wqp(, ) (, ) (; ) (, )
CL
3
(18)
The classical equation of motion is helpful because knowing the
classical limit allows us to identify the purely quantum
mechanical eects.
58,61
Note that the above equations of
motion can also be derived from the classical Langevin
equation, where the classical uctuation dissipation theorem
is hold.
57
Then, in the white noise (or Markovian) limit, the
above equations reduce to the Kramers equation.
In Section 4, we use eqs 915 and eqs 1618 to calculate
quantum and classical ratchet currents, respectively. Because we
are dealing with the distribution function for both quantum and
classical cases, there is no need to sample kinetic trajectories,
which is necessary in approaches employing the Langevin
equation and in some kinetic approaches. This is convenient if
The Journal of Physical Chemistry B Article
dx.doi.org/10.1021/jp403056h |J. Phys. Chem. B 2013, 117, 131321314413135
we wish to calculate the current at low temperature, even in the
classical case, because this sampling becomes inecient due to
the localization of the particle trajectories trapped in the
potential. Another important aspect of the present method-
ology is that it allows elucidation of the dynamical behavior of
the system through the time evolution of the distribution
function in the phase space. In addition to the dependence of
the current on the temperature, the barrier height, and the
system-bath coupling, we analyze the mechanism of ratchet
rectication using the Wigner and classical distribution
functions.
3. POTENTIAL, FORCE, AND SYMMETRY
In the present study, we consider the 1D symmetric potential
(see Figure 2)
κ=
U
qU q() cos( )
02(19)
and the biharmonic force
θ+ Ω+Ft F t F t( ) cos( ) cos(2 )
12
(20)
where U0is the barrier height, κis the wavenumber, and F1,F2,
Ω, and θdenote the amplitudes, frequency, and the phase
dierence of the biharmonic forces, respectively.
1419,4549
Experimentally, such a situation has been realized for uxons in
long Josephson junctions
25
and for cold rubidium and cesium
atoms in optical lattices.
2628
Before reporting the results of our calculations of the ratchet
current, it is important to mention two symmetry conditions
under which the current vanishes.
3236
The rst one is the
time-shift symmetry dened by the transformation (q,p,t)
(q,p,t+T/2), where Tis the period of the external force,
and the second one is the time-reversal symmetry dened by
the transformation (q,p,t)(q,p,t). When the equation of
motion of the system is invariant under either of these two
symmetry transformations, the current vanishes because each of
these transformations reverses the sign of the momentum. For
the external force given by eq 20 and for the quantum and
classical HEOM appearing in eqs 915 and eqs 1618,
respectively, time-shift symmetry exists if F1=0orF2=0.
Time-reversal symmetry is broken if θ0orπin the
Hamiltonian case, that is, if ζ= 0, while, in the presence of
dissipation, this symmetry is always broken, which implies that
directed motion can exist for arbitrary θ. For the classical
overdamped Langevin equation with white noise, the system
possesses time-reversal symmetry if θ=π/2 or 3π/2. The
overdamped (Smoluchowski) case can be studied within the
HEOM formalism by choosing a large value for ζwith γΩ
in the classical case and by choosing a large value of ζwith γ
Ωand βγ1 (i.e., the high-temperature limit) in the
quantum case. In such cases, we expect the calculated current to
vanish if θ=π/2 or 3π/2.
4. NUMERICAL RESULTS
We numerically integrated the HEOM in the form of nite
dierence equations using the fourth-order RungeKutta
method. We also imposed the periodical boundary condition
W(p,q)=W(p,q+2π/κ). The spatial derivative of the kinetic
term in the Liouville operator, (p/m)W(p,q)/q,was
approximated using a third-order left-handed or right-handed
dierence scheme, depending on the sign of the momentum,
while other derivatives with respect to pwere approximated
using a third-order center dierence scheme.
63
We employed
the following units of the length, momentum, and frequency: qr
=1/κ,pr=κ, and ωr=κ2/m. Note that with these units we
have m==κ= 1. The dimensionless position and
momentum then are given by q̅=q/qrand p̅=p/pr. The mesh
sizes for the position and the momentum were chosen in the
ranges 0.0524 < Δq̅< 0.1047 and 0.04 < Δp̅< 0.10. The depth
of the hierarchy and the number of Matsubara frequencies were
chosen so as to satisfy N{5 10} and K{1 2}. In this
study, we xed the inverse of the noise correlation time to γ=
1.0ωrand the amplitude and frequency of the biharmonic force
to F1=F2= 0.20ωr/qrand Ω= 1.0ωr, respectively. Three
values were used for the height of potential barrier, U0=ωr,
2ωr, and 4ωr(see Figure 2).
Figure 2. U0cos 2(q̅) potential for the cases of (a) a low barrier, U0=
1, (b) an intermediate barrier, U0= 2, and (c) a high barrier, U0=4.In
each case, eigenenergies and eigenfunctions with periodic boundary
conditions are plotted as the black dashed lines and the red curves,
respectively.
The Journal of Physical Chemistry B Article
dx.doi.org/10.1021/jp403056h |J. Phys. Chem. B 2013, 117, 131321314413136
Employing the biharmonic force given in eq 20, we
integrated the classical and quantum HEOM until the
distributions reached the steady state. Then, the classical and
quantum directed currents (or the ratchet current) were
evaluated from the distribution function as
≡′
→∞
+
JTtptlim 1d()
tt
tT
(21)
where
⟩=pt q ppWq p t() d d ( , ; ) (22)
We repeated the calculation for xed physical conditions, while
varying θin the range 0 < θ<2π. From these calculations, we
nd that as a function of θthe current roughly takes the
following form:
26,48
θθθ=−JJ() sin( )
max 0(23)
where Jmaxand θ0are the maximum value of the current and the
phase lag evaluated from the numerical simulations, respec-
tively. Below we investigate the dependence of the current on
the temperature, the barrier height, and the bath coupling
strength through the dependences of Jmax and θ0on these
quantities. To gain further insight into the mechanism of
ratchet rectication, we also present plots of the Wigner and
classical distribution functions where helpful.
4.1. Temperature Eects: the Classical Case. Figure 3
displays (a) the maximum value, Jmax, and (b) the phase lag, θ0,
of the current as functions of the inverse temperature, β, for
three values of the barrier height, U0, evaluated using the
classical HEOM given in eqs 1618. Here we chose the weak
coupling strength ζ= 0.10ωr. In all three cases depicted in
Figure 3a, it is seen that the maximum current realizes a
maximum value at some intermediate value of the inverse
temperature. The value of βat which this maximum is realized
is found to increase as U0decreases. As shown in Figure 3b, a
similar prole is found for the phase lag θ0as a function of β,
although the peak positions are dierent.
The behavior depicted in Figure 3 can be understood in
terms of the thermal activation of a particle. To illustrate this, in
Figure 4, we plot snapshots of the classical distribution function
for (a) low-, (b) intermediate-, and (c) high-temperature cases
at ve dierent values of the time for one cycle. In Figure 4a-i,b-
i, it is seen that when the temperature takes a low or
intermediate value the distribution function is concentrated in
the potential wells. The distributions are periodically
accelerated by the biharmonic force and rotate clockwise
around the bottom of each potential well. Figure 4aii,b-ii
reveals that when the distribution approaches the left side of
the barrier a small part of the distribution with positive
momentum is transferred to the right potential well by crossing
the barrier. This can be seen as the elongation of the
distribution on the right side. In Figures 4a-iii,b-iii, it is seen
that while the distribution begins to ow in the positive
direction the sign of the biharmonic force changes, and the
main part of the distribution moves away from the barrier. The
part of the distribution that crosses the barrier ows into the
right potential and rotates clockwise in a spiral form while
losing energy due to dissipation. Figure 4a-iv,b-iv shows the
ow stops shortly after the biharmonic force change direction,
with a small time delay. In Figure 4a-v,b-v, it is seen that when
the distribution approaches the left-hand side of the barrier, a
part of the distribution with negative momentum also ows in
the opposite direction, but the eect of the acceleration caused
by the biharmonic force is smaller in this case than in the
positive case due to the asymmetric inuence of the biharmonic
force on the system dynamics with the phase lag θ0.
Figure 3. Maximum current in (a) the classical and (a) the quantum case and the corresponding phase lag in (b) the classical and (b) the quantum
case as functions of the inverse temperature, β, for three values of the barrier height U0. The coupling strength here is ζ= 0.10ωr.
The Journal of Physical Chemistry B Article
dx.doi.org/10.1021/jp403056h |J. Phys. Chem. B 2013, 117, 131321314413137
In the intermediate temperature case depicted in Figure 4b,
the distribution is broadened compared with Figure 4a due to
thermal eects. The current is greater than in the low-
temperature case because the weight of the distribution at
energies above the barrier under the biharmonic force is
greater. The response of the distribution to the external force is
also delayed due to the broadening of the distribution in the p
direction. Because the phase lag, θ0, changes in accordance with
the timing of the excitation created by F1cos(Ωt)qand F2
cos(2Ωt+θθ0)qthrough the gradients of the potential, this
delay may be the reason that θ0increases as βdecreases up to
the maximum point.
In the high-temperature case considered in Figure 4c, the
distribution is populated even at the top of barrier. This causes
the spatiotemporal asymmetry induced by biharmonic
perturbation, which is necessary for ratchet current to be
weak, and as a result, the current decreases as the temperature
increases. Because the activation energy increases as the barrier
increases, the maxima of the peaks shift toward higher
temperature as U0increases. The phase lag, θ0, decreases as
the temperature increases because the distribution can ow to
the neighboring potentials in both the positive and negative
directions with only a small delay. This can be understood from
Figure 4c-ii, where it is seen that the current ows quickly
following the movement of the distribution.
4.2. Temperature Eects: the Quantum Case. Next, we
consider the quantum mechanical case under the same physical
conditions as in the classical case considered above. Figure 3a
displays the maximum value, Jmax, and Figure 3bdisplays the
phase lags, θ0, of the current as functions of the inverse
Figure 4. Snapshots of the classical distribution for θ= 0.7πin the case of intermediate barrier height, U0= 2, for three values of the inverse
temperature: (a) low temperature, β= 3; (b) intermediate temperature, β= 2; and (c) high temperature, β= 1. The snapshots correspond to the
following times: t= (i) 0, (ii) 0.4π, (iii) 0.8π, (iv) 1.2π, and (v) 1.6π(1/Ω).
The Journal of Physical Chemistry B Article
dx.doi.org/10.1021/jp403056h |J. Phys. Chem. B 2013, 117, 131321314413138
temperature, β, for three values of the barrier height, U0,
calculated using the quantum mechanical HEOM given in eqs
915. It is seen that the proles of Jmax and θ0dier
signicantly from those in the classical case. In the intermediate
and high barrier cases, The Figure also shows that the quantum
mechanical Jmax is signicantly larger than the classical one for
the cases of an intermediate and high barrier in the low-
temperature regime (β> 2), while they exhibit similar behavior
in the high-temperature regime (β< 1).
To see a role of quantum eects, we consider the crossover
temperature at which the classical thermal-activated regime
crosses over to the quantum tunneling regime with regard to
chemical reaction rates
83
for the bath spectral density, J(ω),
given in eq 6 with the potential heights U0= 1, 2, and 4. The
evaluated crossover temperatures are at βc= 4.5, 3.2, and 2.2,
respectively. These values of βcare much smaller than those at
which we observed the dierence between the quantum and
classical cases. This suggests that our analysis based on the
reaction rate may not apply directly, because here we
considered Jmax under the external driving force while altering
the phase.
A distinctive feature of the quantum mechanical results is
found in the low barrier case (the red curve) in Figure 3a. One
may expect that the ratchet current in the quantum mechanical
case is larger than that in the classical case, especially at low
temperatures due to tunneling. However, our results in the low
barrier case presented in Figure 3areveal that in fact this is not
the case.
Figure 5. Snapshots of the quantum distribution for θ= 0.4πin the case of low barrier height, U0= 1, calculated using the HEOM for three values of
the inverse temperature: (a) low temperature, β= 3; (b) intermediate temperature, β= 2; and (c) high temperature, β= 1. These snapshots
correspond to the following times: t= (i) 0, (ii) 0.4π, (iii) 0.8π, (iv) 1.2π, and (v) 1.6π(1/Ω).
The Journal of Physical Chemistry B Article
dx.doi.org/10.1021/jp403056h |J. Phys. Chem. B 2013, 117, 131321314413139
To elucidate the reason for the behavior depicted in Figure
3afor the case U0= 1, we consider the Wigner distribution in
the low and intermediate barrier cases for three values of the
temperature in Figures 5 and 6, respectively. In both cases, the
Wigner distribution function is dispersed over the potential,
and it is negative in some regions (the regions inside the white
and black loops) at low temperature (β2). The Wigner
function for the quantum system is not denitely positive, and
in cases that the tunneling process is important, the region in
which it takes negative values is larger.
84
This indicates that
delocalization of the distribution in the low barrier case
depicted in Figure 5a arises from tunneling. This can be
understood from Figure 2a, where the ground state in the low
potential case is seen to have a large population under the
barrier. The tunneling contribution induced by biharmonic
perturbation may be suppressed at high temperature, because
the temperature enters the theory through β, and hence the
classical limit, ℏ→0, is equivalent to the high-temperature
limit, β0. In this case, however, thermal activation causing
ow over the potential barrier is signicant. For this reason, the
distribution is dispersed for all temperatures. As shown in the
high-temperature classical case depicted in Figure 4b, the eect
of the ratchet potential is weak when the distribution function
is dispersed. This is the reason that the ratchet current is
smaller in the quantum case than the classical case if the barrier
is low.
In the cases of intermediate and high barriers, depicted in
Figures 3a,b, tunneling also plays a signicant role. As shown
in Figure 6, the Wigner distribution is not dispersed when the
barrier is high. As a result, the ratchet mechanism is eective
Figure 6. Snapshots of the quantum distribution for θ= 0.6πin the case of intermediate barrier height, U0= 2, for the following values of β: (a) low
temperature, β= 3; (b) intermediate temperature, β= 2; and (c) high temperature, β= 1. These snapshots correspond to the following times: t= (i)
0, (ii) 0.4π, (iii) 0.8π, (iv) 1.2π, and (v) 1.6π(1/Ω).
The Journal of Physical Chemistry B Article
dx.doi.org/10.1021/jp403056h |J. Phys. Chem. B 2013, 117, 131321314413140
with the help of tunneling, and thus the quantum mechanical
current is larger than the classical current. However, when the
temperature becomes very high, the quantum coherence is lost
due to the thermal noise, and as a result, the quantum value
approaches the classical one.
As shown in Figure 3b, the phase lag has a critical point in
the high-temperature region (β< 1.5) as in the classical case
depicted in Figure 3b. However, in the low-temperature region
(β> 2.5), the phase is seen to increase monotonically in the
quantum mechanical case, while it decreases monotonically in
the classical case. We carefully investigated the dynamics of the
Wigner distribution in the low-temperature region by studying
snapshots of it and the corresponding current using ne time
slices. We thus found that the transfer of the distribution
through tunneling takes a longer time at low temperature. This
is because thermal transitions between the ground and excited
states, which are the cause of tunneling transition induced by
biharmonic perturbation, may be suppressed at low temper-
ature. Because the net current ows slowly in the low-
temperature case, the phase lag that controls the timing of the
biharmonic force increases when the temperature decreases.
For higher barriers, the tunneling plays a minor role, and thus
the increase in the phase is smaller than in the low barrier case.
Note that as discussed in ref 85 quantum diusion may be
suppressed by quantum reection. In the cases we studied,
however, the curvature at the top of the barrier and in the
potential well are equal, and the force that creates the dierence
between the curvatures is small.
86
Thus, in the present case,
quantum reection is not the cause of the suppression.
4.3. Varying the Coupling Strength. We next investigate
the change of the ratchet current as the system-bath coupling
strength ζis varied by taking advantage of the fact that the
HEOM constitute a nonperturbative theory. In Figure 7, we
plot the classical (red curve) and quantum (green curve)
ratchet current for (a) the low barrier case, U0= 1, and (b) the
intermediate barrier case, U0= 2, with β= 1. In both cases, the
classical and quantum currents seem to approach the same
value for large ζ, although the quantum values reach the large ζ
limit more rapidly. To compare the two cases, we plot the
quantum and classical distribution functions in Figure 8. It is
seen that in the low barrier case the quantum distribution is
dispersed, as in the case depicted in Figure 5. Moreover, in the
quantum mechanical case, the distribution in the pdirection
becomes a symmetric Gaussian for ζ> 0.3, while there is some
deviation from Gaussian in the classical case, due to the external
perturbation. This Gaussian feature arises from quantum
tunneling, as can be observed more clearly in the low barrier
case. It is interesting that the situation in which there exists a
symmetric Gaussian distribution in momentum space is similar
to that described by the classical Smoluchowski equation, which
can be derived in the overdamped limit from the Kramers
equation. In the overdamped limit, the distribution of the
momentum direction is Gaussian. In the presently considered
case, we nd that such a situation arises due to the eect of
tunneling even if the system-bath coupling is weak. Therefore,
it is natural to call this situation a tunneling-induced
Smoluchowski limit.
The phase approaches π/2 as ζincreases, as depicted in the
inset of Figure 7i). This is similar to the situation for the
classical Smoluchowski equation. For U0= 2, the quantum and
classical cases exhibit similar ζdependence. This is due to the
fact that the dispersion of the distribution through tunneling
does not occur in the higher barrier case, as depicted in Figure
6. The quantum values approach the classical values for large ζ
because in that case the quantum coherence is destroyed by the
system-bath coupling.
5. CONCLUDING REMARKS
The classical and quantum mechanical ratchet currents in a
system driven by a biharmonic force were calculated in a
rigorous manner for the rst time using the reduced hierarchy
equations of motion (HEOM) in the Wigner representation
over a wide range of values of the temperature, the system-bath
coupling strength, and the potential height. The roles of
uctuations, dissipation, and the biharmonic force on the
ratchet current were investigated in classical and quantum
mechanical cases by studying the dependence of the ratchet
current on these parameters.
In the classical case, we found that over the temperature
range we studied the current realizes a maximum at an
intermediate value and falls ofor both high and low
temperatures. This decrease on the high temperature side is a
result of thermal activation, which induces dispersion of the
distribution over the barrier in the high-temperature region. In
the quantum mechanical case, that tunneling enhances the
current in the high barrier case, while it suppresses the current
in the low barrier case. This is because the eect of the ratchet
potential is weak in the case of a low barrier due to the
dispersion of the distribution that arises from tunneling.
Figure 7. Maximum ratchet current in the classical (red curve) and
quantum mechanical (green curve) cases for (a) a low barrier, U0=1,
and (b) an intermediate barrier, U0= 2, with β= 1. The insets depict
the phase lag θ0as a function of ζ.
The Journal of Physical Chemistry B Article
dx.doi.org/10.1021/jp403056h |J. Phys. Chem. B 2013, 117, 131321314413141
Moreover, in the quantum mechanical case, there is a classical
Smoluchowski-like regime induced by tunneling eects. We
also found that the phase lag of the biharmonic force is related
to the response time of the system to the external perturbation.
As we demonstrated, the dynamical behavior of the system is
clearly and readily elucidated by the time evolution of the
Wigner distribution function. Although we must introduce
hierarchies of Wigner functions described by a discretized
mesh, a modern personal computer is suciently powerful to
solve the equations of motion in a reasonable amount of time.
Numerical techniques, such as the optimization of the
hierarchy,
8790
the utilization of a graphic processing unit
(GPU),
91
and memory allocation for massive parallel
computing,
92
have been developed for the HEOM approach
to accelerate numerical calculations. As an approximated
method, it may be possible to adapt a multiconguration
time-dependent Hartree (MCTDH) approach, which is shown
to be an ecient method to solve Schrödinger equations of
motion.
9395
With these developments, the extension of the
present study to the variety of multidimensional potential
systems with non-Drude-type spectral distribution functions
may be possible.
AUTHOR INFORMATION
Corresponding Author
*E-mail: tanimura@kuchem.kyoto-u.ac.jp. Tel:+81-75-753-
4017.
Notes
The authors declare no competing nancial interest.
ACKNOWLEDGMENTS
Y.T. is grateful for stimulating discussions with Professor Peter
Hänggi and the hospitality of him and his group members
during his stay at the University of Augsburg, made possible by
the Humboldt Foundation. A.K. acknowledges a research
fellowship from Kyoto University. We are grateful for useful
comments on the nite dierence expressions of the HEOM
with Dr. Atsunori Sakurai. This research is supported by a
Grant-in-Aid for Scientic Research (B2235006) from the
Japan Society for the Promotion of Science.
REFERENCES
(1) Smoluchowski, M. V. Experimentell Nachweisbare, der üblichen
Thermodynamik Widersprechende Molekularphänomene. Phys. Z.
1912,13, 10691080.
Figure 8. Snapshots of the classical (upper panel) and quantum mechanical (lower panel) distribution functions at the xed time t=2π(1/Ω) for
four values of the coupling strength: ζ= (a) 0.1, (b) 0.3, (c) 0.5, and (d) 0.7. The upper panels, (i) and (ii), describe the classical cases, whereas the
lower panels, (i) and (ii), describe the quantum cases for the case of a low barrier, U0= 1, and an intermediate barrier, U0= 2, respectively.
The Journal of Physical Chemistry B Article
dx.doi.org/10.1021/jp403056h |J. Phys. Chem. B 2013, 117, 131321314413142
(2) Feynman, R. P.; Leighton, R. B.; Sands, M. The Feynman Lectures
on Physics; Addison Wesley: Reading, MA, 1966; Vol. 1, chapt. 46.
(3) Hänggi, P.; Bartussek, R. In Nonlinear Physics of Complex Systems;
Parisi, J., Muller, S. C., Zimmermann, W., Eds.; Lecture Notes in
Physics Vol. 476; Springer: Berlin, 1996; pp 294308.
(4) Jülicher, F.; Ajdari, A.; Prost, J. Modeling Molecular Motors. Rev.
Mod. Phys 1997,69, 12691281.
(5) Astumian, R. D. Thermodynamics and Kinetics of a Brownian
Motor. Science 1997,276, 917922.
(6) Reimann, P. Brownian Motors: Noisy Transport far from
Equilibrium. Phys. Rep. 2002,361,57265.
(7) Hänggi, P.; Marchesoni, F.; Nori, F. Brownian Motors. Ann. Phys.
(Leipzig, Ger.) 2005,14,5170.
(8) Hänggi, P.; Marchesoni, F. Artificial Brownian Motors:
Controlling Transport on the Nanoscale. Rev. Mod. Phys 2009,81,
387442.
(9) Magnasco, M. O. Forced Thermal Ratchets. Phys. Rev. Lett. 1993,
71, 14771481.
(10) Bartussek, R.; Hänggi, P.; Kissner, J. G. Periodically Rocked
Thermal Ratchets. Europhys. Lett. 1994,28, 459464.
(11) Schreier, M.; Reimann, P.; Hänggi, P.; Pollak, E. Giant
Enhancement of Diffusion and Particle Selection in Rocked Periodic
Potentials. Europhys. Lett. 1998,44, 416422.
(12) Lehmann, J.; Kohler, S.; Hänggi, P.; Nitzan, A. Molecular Wires
Acting as Coherent Quantum Ratchets. Phys. Rev. Lett. 2002,88,
228305.
(13) Scheidl, S.; Vinokur, V. M. Quantum Brownian Motion in
Ratchet Potentials. Phys. Rev. B. 2002,65, 195305.
(14) Schneider, W.; Seeger, K. Harmonic Mixing of Microwaves by
Warm Electrons in Germanium. Appl. Phys. Lett. 1966,8, 133135.
(15) Wonneberger, W.; Breymayer, H.-J. Asymptotics of Harmonic
Microwave Mixing in a Sinusoidal Potential. Z. Phys. B 1981,43, 329
334.
(16) Breymayer, H.-J.; Risken, H.; Vollmer, H. D.; Wonneberger, W.
Harmonic Mixing in a Cosine Potential for Large Damping and
Arbitrary Field Strengths. Appl. Phys. B: Laser Opt. 1982,28, 335339.
(17) Breymayer, H.-J. Harmonic Mixing in a Cosine Potential for
Arbitrary Damping. Appl. Phys. A: Mater. Sci. Process. 1984,33,17.
(18) Marchesoni, F. Harmonic Mixing Signal: Doubly Dithered Ring
Laser Gyroscope. Phys. Lett A 1986,119, 221224.
(19) Goychuk, I.; Hänggi, P. Quantum Rectifiers from Harmonic
Mixing. Europhys. Lett. 1998,43, 503509.
(20) Linke, H.; Sheng, W.; Löfgren, A.; Hongqi, Xu.; Omling, P.;
Lindelof, P. E. A Quantum Dot Ratchet: Experiment and Theory.
Europhys. Lett. 1998,44, 341347.
(21) Linke, H.; Humphrey, T. E.; Löfgren, A.; Sushkov, A. O.;
Newbury, R.; Taylor, R. P.; Omling, P. Experimental Tunneling
Ratchets. Science 1999,286, 23142317.
(22) Linke, H. Ratchets and Brownian Motors: Basics, Experiments
and Applications. Appl, Phys. 2002,75, 167167.
(23) Majer, J. B.; Peguiron, J.; Grifoni, M.; Tusveld, M.; Mooij, J. E.
Quantum Ratchet Effect for Vortices. Phys. Rev. Lett. 2003,90, 056802.
(24) Villegas, J. E.; Savelev, S.; Nori, F.; Gonzalez, E. M.; Anguita, J.
V.; Garcia, R.; Vicent, J. L. A Superconducting Reversible Rectifier that
Controls the Motion of Magnetic Flux Quanta. Science 2003,302,
11881191.
(25) Ustinov, A. V.; Coqui, C.; Kemp, A.; Zolotaryuk, Y.; Salerno, M.
Ratchetlike Dynamics of Fluxons in Annular Josephson Junctions
Driven by Biharmonic Microwave Fields. Phys. Rev. Lett. 2004,93,
087001.
(26) Shiavoni, M.; S-Palencia, L.; Renzoni, F.; Grynberg, G. Phase
Control of Directed Diffusion in a Symmetric Optical Lattice. Phys.
Rev. Lett. 2003,90, 094101.
(27) Gommers, R.; Bergamini, S.; Renzoni, F. Dissipation-Induced
Symmetry Breaking in a Driven Optical Lattice. Phys. Rev. Lett. 2005,
95, 073003.
(28) Salger, T.; Kling, S.; Hecking, T.; Geckeler, C.; Morales-Molina,
L.; Weitz, M. Directed Transport of Atoms in a Hamiltonian Quantum
Ratchet. Science 2009,326, 12411243.
(29) Astumian, R. D.; Derenyi, I. Fluctuation Driven Transport and
Models of Molecular Motors and Pumps. Eur. Biophys 1998,27, 474
489.
(30) Nishiyama, M.; Higuchi, H.; Ishii, Y.; Taniguchi, Y.; Yanagida, T.
Single Molecule Processes on the Stepwise Movement of ATP-Driven
Molecular Motors. BioSystems 2003,71, 145156.
(31) Kolomeisky, A. B.; Fisher, M. E. Molecular Motors; A Theorists
Perspective. Annu. Rev. Phys. Chem. 2007,58, 675695.
(32) Flach, S.; Yevtushenko, O.; Zolotaryuk, Y. Directed Current
Due to Broken Time-Space Symmetry. Phys. Rev. Lett. 2000,84,
23582361.
(33) Yevtushenko, O.; Flach, S.; Zolotaryuk, Y.; Ovchinnikov, A. A.
Rectification of Current in AC-Driven Nonlinear Systems and
Symmetry Properties of the Boltzmann Equation. Europhys. Lett.
2001,54, 141147.
(34) Denisov, S.; Flach, S.; Ovchinnikov, A. A.; Yevtushenko, O.;
Zolotaryuk, Y. Broken Space-Time Symmetries and Mechanisms of
Rectification of AC Fields by Nonlinear (non)Adiabatic Response.
Phys. Rev. E 2002,66, 041104.
(35) Reimann, P. Supersymmetric Ratchets. Phys. Rev. Lett. 2001,86,
49924995.
(36) Quintero, N. R.; Cuesta, J. A.; A-Nodarse, R. Symmetries Shape
the Current in Ratchets Induced by a Biharmonic Driving Force. Phys.
Rev. E 2010,81, 030102.
(37) Reimann, P.; Grifoni, M.; Hänggi, P. Quantum Ratchets. Phys.
Rev. Lett. 1997,79,1013.
(38) Reimann, P.; Hänggi, P. Quantum Features of Brownian Motors
and Stochastic Resonance. Chaos. 1998,8, 629642.
(39) Goychuk, I.; Grifoni, M.; Hänggi, P. Nonadiabatic Quantum
Brownian Rectifiers. Phys. Rev. Lett. 1998,81, 649652.
(40) Goychuk, I.; Grifoni, M.; Hänggi, P. Nonadiabatic Quantum
Brownian Rectifiers (Erratum). Phys. Rev. Lett. 1998,81, 2837.
(41) Grifoni, M.; Ferreira, M. S.; Peguiron, J.; Majer, J. B. Quantum
Ratchets with Few Bands below the Barrrier. Phys. Rev. Lett. 2002,89,
146801.
(42) Garcia-Palacios, J. L.; Zueco, D. The Caldeira-Leggett Quantum
Master Equation in Wigner Phase Space: Continued-Fraction Solution
and Application to Brownian Motion in Periodic Potentials. J. Phys. A:
Math. Gen. 2004,37, 1073510770.
(43) Maier, S. A.; Ankerhold, J. Low-Temperature Quantum
Fluctuations in Overdamped Ratchets. Phys. Rev. E 2010,82, 021104.
(44) Zhan, F.; Denisov, S.; Ponomarev, A.; Hänggi, P. Quantum
Ratchet Transport with Minimal Dispersion Rate. Phys. Rev. A 2011,
84, 043617.
(45) Denisov, S.; Morales-Molina, L.; Flach, S.; Hänggi, P.
Periodically Driven Quantum Ratchets: Symmetries and Resonances.
Phys. Rev. A 2007,75, 063424.
(46) Denisov, S.; Kohler, S.; Hänggi, P. Underdamped Quantum
Ratchets. Europhys. Lett. 2009,85, 40003.
(47) Rohling, N.; Grossmann, F. Optimization of Electron Pumping
by Harmonic Mixing. Phys. Rev. B 2011,83, 205310.
(48) Denisov, S.; Hänggi, P.; Mateos, J. L. AC-Driven Brownian
Motors: A Fokker-Planck Treatment. Am. J. Phys. 2009,77, 602606.
(49) Machura, L.; Luczka, J. Transport Driven by Biharmonic Forces:
Impact of Correlated Thermal Noise. Phys. Rev. E 2010,82, 031133.
(50) Kohler, S.; Dittrich, T.; Hänggi, P. Floquet-Markovian
Description of Parametrically Driven, Dissipative Harmonic Quantum
Oscillator. Phys. Rev. E 1997,85, 300313.
(51) Caldeira, A. O.; Leggett, A. J. Path Integral Approach to
Quantum Brownian Motion. Physica A 1983,121, 587616.
(52) Chang, L.-D.; Waxman, D. Quantum Fokker-Planck Equation. J.
Phys. C 1985,18, 58735879.
(53) Coffey, W. T.; Kalmykov, Yu. P.; Titov, S. V.; Mulligan, B. P.
Semiclassical Master Equation in Wigners Phase Space Applied to
Brownian Motion in a Periodic Potential. Phys. Rev. E 2007,75,
041117.
(54) Coffey, W. T.; Kalmykov, Yu. P.; Titov, S. V.; Mulligan, B. P.
Wigner Function Approach to the Quantum Brownian Motion of a
Particle in a Potential. Phys. Chem. Chem. Phys. 2007,9, 33613382.
The Journal of Physical Chemistry B Article
dx.doi.org/10.1021/jp403056h |J. Phys. Chem. B 2013, 117, 131321314413143
(55) Shit, A.; Chattopadhyay, S.; Chaudhuri, J. R. Towards an
Understanding of Escape Rate and State Dependent Diffusion for a
Quantum Dissipative System. Chem. Phys. 2011,386,5672.
(56) Tanimura, Y. Stochastic Liouville, Langevin, Fokker-Planck, and
Master Equation Approaches to Quantum Dissipative Systems. J. Phys.
Soc. Jpn. 2006,75, 082001.
(57) Tanimura, Y.; Wolynes, P. G. Quantum and Classical Fokker-
Planck Equations for a Gaussian-Markovian Noise Bath. Phys. Rev. A
1991,43, 41314142.
(58) Tanimura, Y.; Wolynes, P. G. The Interplay of Tunneling,
Resonance, and Dissipation in Quantum Barrier Crossing: A
Numerical Study. J. Chem. Phys. 1992,96, 84858496.
(59) Tanimura, Y.; Steffen, T. Two Dimensional Spectroscopy for
Harmonic Vibrational Modes with Nonlinear System-Bath Inter-
actions: II. Gaussian-Markovian Case. J. Phys. Soc. Jpn. 2000,69,
40954106.
(60) Kato, T.; Tanimura, Y. Vibrational Spectroscopy of a Harmonic
Oscillator System Nonlinearly Coupled to a Heat Bath. J. Chem. Phys.
2002,117, 62216234.
(61) Sakurai, A.; Tanimura, Y. Does Play a Role in Multidimen-
sional Spectroscopy? Reduced Hierarchy Equations of Motion
Approach to Molecular Vibrations. J. Phys. Chem. A 2011,115,
40094022.
(62) Sakurai, A.; Tanimura, Y. An Approach to Quantum Transport
Based on Reduced Hierarchy Equations of Motion: Application to a
Resonant Tunneling Diode. J. Phys. Soc. Jpn. 2013,82, 033707.
(63) Sakurai, A.; Tanimura, Y. Self-Excited Current Oscillations in a
Resonant Tunneling Diode Described by a Model Based on the
Caldeira-Leggett Hamiltonian. 2013, arXiv:1304.5596. arXiv.org e-
Print archive. http://arxiv.org/abs/1304.5596 (accessed April 20,
2013).
(64) Tanimura, Y.; Kubo, R. Time Evolution of a Quantum System
in Contact with a Nearly Gaussian-Markoffian Noise Bath. J. Phys. Soc.
Jpn. 1989,58, 101114.
(65) Ishizaki, A.; Fleming, G. R. Theoretical Examination of
Quantum Coherence in a Photosynthetic System at Physiological
Temperature. Proc. Natl. Acad. Sci. U.S.A. 2009,106, 1725517260.
(66) Chen, L.; Zheng, R.; Jing, Y.; Shi, Q. Simulation of the Two-
Dimensional Electronic Spectra of the Fenna-Matthews-Olson
Complex using the Hierarchical Equations of Motion Method. J.
Phys. Chem. 2011,134, 194508.
(67) Kreisbeck, C.; Kramer, T. Long-Lived Electronic Coherence in
Dissipative Exciton Dynamics of Light-Harvesting Complexes. J. Phys.
Chem. Lett. 2012,3, 28282833.
(68) Strümpfer, J.; Schulten, K. Excited State Dynamics in
Photosynthetic Reaction Center and Light Harvesting Complex 1. J.
Chem. Phys. 2012,137, 065101.
(69) Dijkstra, A. G.; Tanimura, Y. Correlated Fluctuations in the
Exciton Dynamics and Spectroscopy of DNA. New J. Phys. 2010,12,
055005.
(70) Xu, R.-X.; Chen, Y.; Cui, P.; Ke, H. W.; Yan, Y.-J. The Quantum
Solvation, Adiabatic versus Nonadiabatic, and Markovian versus Non-
Markovian Nature of Electron-Transfer Rate Processes. J. Phys. Chem.
A2007,111, 96189626.
(71) Tanaka, M.; Tanimura, Y. Quantum Dissipative Dynamics of
Electron Transfer Reaction System: Nonperturbative Hierarchy
Equation Approach. J. Phys. Soc. Jpn. 2009,78, 073802.
(72) Tanaka, M.; Tanimura, Y. Multistate Electron Transfer
Dynamics in the Condensed Phase: Exact Calculations from the
Reduced Hierarchy Equations of Motion Approach. J. Chem. Phys.
2010,132, 214502.
(73) Tanimura, Y. Reduced Hierarchy Equations of Motion
Approach with Drude Plus Brownian Spectral Distribution: Probing
Electron Transfer Processes by Means of Two-Dimensional
Correlation Spectroscopy. J. Chem. Phys. 2012,137, 22A550.
(74) Dijkstra, A. G.; Tanimura, Y. Non-Markovian Entanglement
Dynamics in the Presence of System-Bath Coherence. Phys. Rev. Lett.
2010,104, 250401.
(75) Yin, X.; Ma, J.; Wang, X.; Nori, F. Spin Squeezing under non-
Markovian Channels by the Hierarchy Equation Method. Phys. Rev. A
2012,86, 012308.
(76) Frensley, W. R. Boundary Conditions for Open Quantum
Systems Driven far from Equilibrium. Rev. Mod. Phys. 1990,62, 745
791.
(77) Caldeira, A. O.; Leggett, A. J. Quantum Tunneling in a
Dissipative System. Ann. Phys. 1983,149, 374456.
(78) Grabert, H.; Schramm, P.; Ingold, G. L. Quantum Brownian
Motion: The Functional Integral Approach. Phys. Rep. 1988,168,
115207.
(79) Kubo, R.; Toda; M.; Hashitsume, N. Statistical Physics; Springer-
Verlag: Berlin, 1985; Vol. 2.
(80) Tanimura, Y. Nonperturbative Expansion Method for a
Quantum System Coupled to a Harmonic-Oscillator Bath. Phys. Rev.
A1990,41, 66766687.
(81) Ishizaki, A.; Tanimura, Y. Quantum Dynamics of a System
Strongly Coupled to a Low Temperature Colored Noise Bath:
Reduced Hierarchy Equations Approach. J. Phys. Soc. Jpn. 2005,74,
31313134.
(82) Carmichael, H. J. Statistical Methods in Quantum Optics 1;
Springer: Berlin, 1999.
(83) Hänggi, P.; Grabert, H.; Ingold, G. L.; Weiss, U. Quantum
Theory of Activated Events in Presence of Long-Time Memory. Phys.
Rev. Lett. 1985,55, 761764.
(84) Hillery, M.; OConnell, R. F.; Scully, M. O.; Wigner, E. P.
Distribution Functions in Physics: Fundamentals. Phys. Rep. 1984,106,
121167.
(85) Griff, U.; Grabert, H.; Hänggi, P.; Riseborough, P. S. Possibility
of Quantum Effects Reducing the Rate of Escape from a Metastable
Well. Phys. Rev. B 1989,40, 72957297.
(86) Shushin, A. I.; Pollak, E. Quantum and Classical Aspects of
Activated Surface Diffusion. J. Chem. Phys. 2003,119, 1094110952.
(87) Shi, Q.; Chen, L. P.; Nan, G. J.; Xu, R. X.; Yan, Y.-J. Efficient
Hierarchical Liouville Space Propagator to Quantum Dissipative
Dynamics. J. Phys. Chem. 2009,130, 084105.
(88) Hu, J.; Xu, R. X.; Yan, Y.-J. Communication: PadéSpectrum
Decomposition of Fermi Function and Bose Function. J. Chem. Phys.
2010,133, 101106.
(89) Tian, B. L.; Ding, J. J.; Xu, R. X.; Yan, Y.-J. Biexponential Theory
of Drude Dissipation via Hierarchical Quantum Master Equation. J.
Chem. Phys. 2010,133, 114112.
(90) Zhu, J.; Kais, S.; Rebentros, P.; Aspuru-Guzik, A. Modified
Scaled Hierarchical Equation of Motion Approach for the Study of
Quantum Coherence in Photosynthetic Complex. J. Phys. Chem. B
2011,115, 15311537.
(91) Kreisbeck, C.; Kramer, T.; Rodriguez, M.; Hein, B. High-
Performance Solution of Hierarchical Equations of Motion for
Studying Energy Transfer in Light-Harvesting Complexes. J. Chem.
Theory Comput. 2011,7, 21662174.
(92) Strümpfer, J.; Schulten, K. Open Quantum Dynamics
Calculations with the Hierarchy Equations of Motion on Parallel
Computers. J. Chem. Theory Comput. 2012,8, 28082816.
(93) Meyer, H.-D.; Manthe, U.; Cederbaum, L. S. The Multi-
Configurational Time-Dependent Hartree Approach. Chem. Phys. Lett.
1990,165,7378.
(94) Wang, H.; Thoss, M. Multilayer Formulation of the Multi-
configuration Time-Dependent Hartree Theory. J. Chem. Phys. 2003,
119, 1580111.
(95) Burghardt, I.; Worth, G. A. Multimode Quantum Dynamics
using Gaussian Wavepackets: The Gaussian-Based Multiconfiguration
Time-Dependent Hartree (G-MCTDH) Method Applied to the
Absorption Spectrum of Pyrazine. J. Chem. Phys. 2008,129, 174104.
The Journal of Physical Chemistry B Article
dx.doi.org/10.1021/jp403056h |J. Phys. Chem. B 2013, 117, 131321314413144
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
A multistate displaced oscillator system strongly coupled to a heat bath is considered a model of an electron transfer (ET) reaction system. By performing canonical transformation, the model can be reduced to the multistate system coupled to the Brownian heat bath defined by a non-ohmic spectral distribution. For this system, we have derived the hierarchy equations of motion for a reduced density operator that can deal with any strength of the system bath coupling at any temperature. The present formalism is an extension of the hierarchy formalism for a two-state ET system introduced by Tanimura and Mukamel into a low temperature and multistate system. Its ability to handle a multistate system allows us to study a variety of problems in ET and nonlinear optical spectroscopy. To demonstrate the formalism, the time-dependent ET reaction rates for a three-state system are calculated for different energy gaps.
Article
Full-text available
A test system is assumed to interact with a heat bath consisting of harmonic oscillators or an equivalent bath with a proper frequency spectrum producing a Gaussian-Markoffian random perturbation. The effect of reaction of the test system to the bath is considered in the high temperature approximation. Elimination of the bath using the influence functional method of Feynman and Vernon yields a continuous fraction expression for the reduced density matrix of the test system. The result affords a basis to clarify the relationship between the stochastic and the dynamical approaches to treat the problem of partial destruction of quantum coherence of a system interacting with its environment.
Article
Full-text available
We study both dissipative and nondissipative quantum transport in discrete Brownian rectifiers being driven by nonthermal noise that is unbiased on average. In the absence of dissipation the current is always zero and the ballistic diffusion changes into normal diffusion. The dissipative quantum dynamics exhibits current (with distinctive reversals) as a result of the cooperative interplay between dissipative forces and external fluctuations. Considering the nonlinear current response to aperiodic, noisy forces we predict aperiodic quantum stochastic resonance. The nonthermal fluctuations can considerably enhance, as well as suppress, the thermal quantum diffusion.
Article
We investigate quantum Brownian motion in adiabatically rocked ratchet systems. Above a crossover temperature Tc tunneling events are rare, yet they already substantially enhance the classical particle current. Below Tc, quantum tunneling prevails and the classical predictions grossly underestimate the transport. Upon approaching T = 0 the quantum current exhibits a tunneling induced reversal, and tends to a finite limit.
Article
This is a study of simple kinetic models of open systems, in the sense of systems that can exchange conserved particles with their environment. The system is assumed to be one dimensional and situated between two particle reservoirs. Such a system is readily driven far from equilibrium if the chemical potentials of the reservoirs differ appreciably. The openness of the system modifies the spatial boundary conditions on the single-particle Liouville-von Neumann equation, leading to a non-Hermitian Liouville operator. If the open-system boundary conditions are time reversible, exponentially growing (unphysical) solutions are introduced into the time dependence of the density matrix. This problem is avoided by applying time-irreversible boundary conditions to the Wigner distribution function. These boundary conditions model the external environment as ideal particle reservoirs with properties analogous to those of a blackbody. This time-irreversible model may be numerically evaluated in a discrete approximation and has been applied to the study of a resonant-tunneling semiconductor diode. The physical and mathematical properties of the irreversible kinetic model, in both its discrete and its continuum formulations, are examined in detail. The model demonstrates the distinction in kinetic theory between commutator superoperators, which may become non-Hermitian to describe irreversible behavior, and anticommutator superoperators, which remain Hermitian and are used to evaluate physical observables.
Article
We address the stochastic dynamics of an open quantum system coupled to a heat reservoir that is driven out of thermal equilibrium by an external noise. By constructing Langevin and Fokker–Planck equations, we obtain the rate of decay from a metastable state of the system when the dissipation is state dependent. We discuss the effects and consequences of the non-linear interaction(s) stemming out of the system-bath coupling alongside the modulation of the bath by an external noise on the rate expression. We demonstrate that the temperature dependence of the escape rate is not only embedded in the so-called Arrhenius type factor, the second exponential factor also includes the temperature dependence. The last effect has a purely quantum origin. Interestingly, we also envisage that this quantum effect is entangled with dissipation. The results offer a basis for clarifying the relationship between the dissipation and exponential factor of the obtained rate expression.
Article
Harmonic mixing of two alternating electric fields due to a Brownian charged particle in a nonlinear one-dimensional potential of cosine shape is investigated. The dynamics of the system are described by a time dependent Fokker-Planck equation. The appropriate distribution function is obtained by a matrix continued fraction expansion method, which is treated numerically. The dc signal due to mixing is computed for strong thermal fluctuations in all relevant parameter ranges of the pinning potential strength, damping and frequency. The dc signal without fluctuations is discussed separately. Resonance effects are shown in the electric dc field and the additional phase shift, caused by intrinsic relaxation processes.