Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Eur. Phys. J.

B (2019) 92: 5
https://doi.org/10.1140/epjb/e2018-90524-7 THE EUROPEAN
PHYSICAL JOURNAL B
Regular Article

Thermal resonance and energy transport in a biharmonically


driven Frenkel–Kontorova lattice
Mauricio Romero-Bastida a and Santiago Guerrero-Gonzalez
SEPI ESIME-Culhuacán, Instituto Politécnico Nacional, Av. Santa Ana No. 1000, Col. San Francisco Culhuacán,
Delegación Coyoacan, Distrito Federal 04430, México, Mexico

Received 30 August 2018 / Received in final form 12 November 2018


Published online 9 January 2019
c EDP Sciences / Società Italiana di Fisica / Springer-Verlag GmbH Germany, part of Springer Nature,
2019

Abstract. In this work, we study the heat conduction properties of a one-dimensional Frenkel–Kontorova
lattice driven by an external, time-periodic biharmonic force applied locally at one boundary and in contact
with two heat reservoirs operating at different temperature by means of molecular dynamics simulations. In
the single-frequency externally driven case already studied it was observed that there is a value of the driving
frequency at which the heat flux takes its maximum value, a phenomenon termed as thermal resonance.
It was also determined that it is possible to direct the heat flow against the imposed temperature bias by
adjusting the frequency of the single harmonic driving force. With the implementation of the biharmonic
forcing we have explored the temperature range at which thermal resonance effect is present. Furthermore,
we have determined that by changing the relative amplitude of both harmonic components as well as the
frequency of the second, taken always as a multiple, not necessarily integer, of the first one, we can adjust
the frequency at which the studied effect is present in the proposed model.

1 Introduction desired pumping effect. Later this effect was obtained,


without employing an external non-equilibrium bias, by
The desire to design and build micromechanical machines time-periodically modulated heat reservoirs with average
[1], and in particular heat controlling devices [2–4], has zero-temperature difference in contact with a molecu-
motivated a lot of interest in the last decade or so. lar wire system wherein the heat flow also generates an
Recently biologically inspired molecular machines that electric current [18]. Previous studies employed classi-
can walk [5], rotate [6], and pump [7], stochastic heat cal nonlinear junctions [19] as well as various models of
engines [8], a mechanical analog of a laser [9] as well as nonlinear oscillator lattices [20] coupled to the aforemen-
nanopower generators [10] have been proposed, among tioned reservoirs. Afterwards, in reference [21] heat flux
others. In particular, and within the realm of heat control inversion was also observed in a Frenkel–Kontorova (FK)
in low-dimensional systems, some of the thermal devices anharmonic lattice under the influence of a finite thermal
that have been envisioned recently are the heat rectifier gradient and acted upon by an external time-periodical,
[11–13], a thermal transistor [14], thermal logic gates [15], single-frequency force.
and a thermal memory as well [16]. All this recurring Motivated by some results in the field of Brownian
theoretical interest partly lies in the rapid progress in motors [22,23] we have studied the model proposed in ref-
probing and manipulating thermal properties of nanoscale erence [21] when acted by an external forcing constituted
systems, which unveils the possibility of designing thermal by two harmonic components. This technique has been
devices with optimized performance at the atomic scale. applied in the context of nanoparticle transport as well as
Among the proposed thermal devices is the heat pump, to design ring gyroscopes and annular Josephson junctions
which directs the heat flow against a temperature bias. (see, e.g. Ref. [23] and references therein). It has also been
A quantum heat pump model was proposed in reference previously employed to obtain a finite heat flux with a har-
[17], wherein a molecule with two allowed energy levels monic mixing drive on the heat reservoirs connected to
interacts with two heat reservoirs kept at different tem- various nonlinear oscillator lattice models [20]. However,
peratures. An external periodic modulation is applied to the resulting ratchet heat flux obeys no simple adiabatic
the energy levels and an asymmetry is incorporated by estimate so that even the direction of the ratcheting heat
taking reservoirs with different spectral properties and flux cannot be predicted a priori; also the aforementioned
different couplings to the molecule, thus leading to the effect was observed for a single temperature value. Apply-
ing this external biharmonic driving to the model studied
a in reference [21] we obtain a well defined heat flux value
e-mail: mromerob@ipn.mx
Page 2 of 8 Eur. Phys. J. B (2019) 92: 5

at the same temperature value employed when only a a timestep of ∆t = 0.005 for 4 × 107 time units after a
single-frequency forcing is applied, as well as for higher transient of 6 × 107 time units.
temperatures. Furthermore, by varying the magnitude of Once the non-equilibrium stationary state is attained,
the frequency of the second harmonic component, taken as the local heat flux is computed as Ji = khq̇i (qi − qi−1 )i
a multiple of the frequency of the first one and thus hav- with i ∈ [2, N − 1] and the local temperature as Ti =
ing a single frequency as a parameter, we have been able hp2i /mi i; in both instances h· · · i indicates a time aver-
to modify the frequency at which the heat flux inversion age that, under the considered temperature bias, can be
effect occurs; this control can also be achieved, although to safely assumed to be equivalent to an ensemble one. In
a lesser degree, by manipulating the relative amplitudes of the stationary state, the total heat flux J is calculated as
both components. Thus, our modification affords a way to the algebraic average of Ji over the number N of unther-
control the aforementioned effect by means of the parame- mostatted oscillators. In the following, the behavior of
ters of the external forcing, notwithstanding the fact that this quantity will be studied as a function of both the
this effect is stronger and more robust against ambient temperature difference ∆T = TL − TR and the average
changes than the original one. temperature T0 ≡ (TL + TR )/2 of the system.
This paper is organized as follows: in Section 2, the
model system and methodology are presented. Our results
on the dependence of the current reversal on the extra 3 Results
force term as well as its dependence on the average tem-
perature wherewith the device operates are reported in 3.1 External forcing with a single harmonic
Section 3. The discussion of the results, as well as our
conclusions, are presented in Section 4. In Figure 1a, we present the results of the dependence
of the heat flux J as a function of the driving frequency
ω1 ≡ ω with A1 = 1, A2 = 0, TL = 0.4 and various values
2 System description of TR for a lattice with N = 64 oscillators. In all instances
TL < TR and thus, in the absence of the external drive, the
The equations of motion (EOMs) for each lattice oscillator heat flows from oscillator i = N to i = 1, which renders
in the considered 1D model can be written as q̇i = pi /mi a negative sign for J. The results for TR = 0.6 reproduce
and those already reported in reference [21]. In the adiabatic
limit ω → 0 heat transport is mainly dominated by the
V
ṗi = k(qi+1 + qi−1 − 2qi ) − sin(2πqi /a) + δi1 F (t) temperature gradient, being negligible the effects of the
2π external force. Furthermore, in this frequency range the
+(ξL − λL pi ) δ1i + (ξR − λR pi ) δiN , (1)
oscillators are confined near their equilibrium positions
within the valley of the nonlinear on-site potential [12].
where N is the system size, k is the harmonic spring For an intermediate frequency value the first oscillator
constant, a is the lattice constant, and V is the ampli- has enough kinetic energy to overcome the potential well,
tude of the on-site potential. Here, we consider only the the on-site potential becomes negligible and the FK√lat-
commensurate case where the on-site potential assumes
the same spatial periodicity as the lattice constant. tice degenerates to a harmonic one with a 0 < ω < 4k
{mi , qi , pi }N phonon band composed mainly of noninteracting phonons.
i=1 are the mass, displacement, and momen- Thus, the driving frequency can resonate with the sys-
tum of the ith oscillator. For convenience of numerical
calculations we use dimensionless parameters by measur- tem’s characteristic frequencies, i.e. those contained in
ing positions in units of [a], spring constants in units the aforementioned phonon band, leading to a heat flux
increase that attains a maximum value Jmax > 0 – that
of [k], time in units of [(m/k)1/2 ], momenta in units of is, a reversal in the heat flux direction – at a critical value
[a(mk)1/2 ], and temperatures in unit of [a2 k/kB ]; see the of ω1 ≡ ωmax = 1.25; therefore, the cooperation between
Appendix of reference [24] for a detailed procedure on the intrinsic response time of the system and the external
how to construct such dimensionless units. Fixed bound- driving frequency leads to the occurring of thermal reso-
ary conditions are assumed (q0 = qN +1 = 0). Henceforth nance. In the limit ω → ∞, the first oscillator experiences
we will consider a homogeneous system, i.e. mi = 1 ∀ i. a time averaged zero constant force as if there is no driv-
The Gaussian white noise ξL/R has zero mean and correla- ing and the contribution due to the temperature gradient
tion hξL,R (t)ξL,R (t0 )i = 2λL,R kB TL,R mi (δ1i + δN i )δ(t − t0 ), becomes again dominant. Therefore, the J value in the
being λL,R (taken as 1 in all computations hereafter absence of the external driving is the one attained in either
reported) the coupling strength between the system and one of the above limits. Next we notice that, as the value
the left and right thermal reservoirs operating at temper- of TR increases while keeping TL fixed, thermal resonance
atures TL and TR , respectively. The external driving force is still present, but the current reversal disappears: the
on the first oscillator will be taken as J values in the whole frequency range steadily decrease
as TR increases, which results in a negative Jmax value
F (t) = A1 sin(ω1 t) + A2 sin(ω2 t), (2) in the whole frequency range. In Figure 1b, we plot the
corresponding temperature profiles for each TR value con-
being A1,2 and ω1,2 the amplitude and frequency of sidered. The observed change in slope corresponds to the
each component. The aforementioned EOMs were inte- loss of the current reversal effect. Thus in these instances,
grated with a stochastic velocity-Verlet integrator with the resonance effect renders a moderate increase of the
Eur. Phys. J. B (2019) 92: 5 Page 3 of 8

(a) (a) 0.2


0
0.1

Jmax
-0.05
J

0
A1 = 0.5
-0.1 -0.1
A1 = 1.0
A1 = 1.5
A1 = 2.0

-0.15 -0.2
0.01 0.1 1 10 0.5 1 1.5
TR = 0.6 ω ∆T
2 TR = 1.0
(b)
(b) TR = 1.4
TR = 1.6
TR = 2.0 0
1.5

Jmax
Ti

1 -0.1 ∆T = 0.2
∆T = 0.6
∆T = 1.0
∆T = 1.4
0.5
0 10 20 30 40 50 60 0.5 1 1.5 2
i To
Fig. 1. (a) Heat flux J vs. frequency ω for an external force Fig. 2. (a) Maximum heat flux Jmax vs. temperature difference
with A1 = 1 and A2 = 0 with V = 5, k = 1, TL = 0.4, and ∆T for various values of the amplitude A1 , with A2 = 0 and a
N = 64. Each curve corresponds to a value of TR and increases fixed TL = 0.4 value. (b) Maximum heat flux Jmax vs. average
from top to bottom. Maximum heat flux is at ωmax = 1.25 in temperature T0 for various values of the temperature difference
all instances. (b) Temperature profiles corresponding to the ∆T increasing from top to bottom with A1 = 1 and A2 = 0.
ωmax value in panel (a) for different values of TR , from low to Same V , k, and N values as in Figure 1. Continuous lines are
high values increasing from bottom to top. a guide to the eye.

heat flux at ωmax with respect to its value when there is zero values of ∆T are taken, the so obtained results (not
no external driving, but with no inversion of the direc- shown) are indistinguishable from those corresponding to
tion of the heat flux. Our results seem to indicate that ∆T = 0.2, the lowest employed value in both figures.
this later effect is extremely weak in all instances and
highly dependent on the temperature differences as readily 3.2 Biharmonic external forcing
shown.
Now, in Figure 2a we present the dependence of In order to assess the influence of the second external forc-
Jmax with respect to the temperature difference ∆T at ing term in both the thermal resonance and change in
fixed TL , A2 = 0, and increasing A1 values. It is readily direction of the heat flux, we begin to perform simulations
observed that, for the lowest A1 values depicted, the with various values of the parameters A1 , A2 , ω1 ≡ ω,
current reversal effect rapidly disappears and the heat and taking the quotient ω2 /ω as an irrational number.
flux takes increasingly negative values as ∆T increases, This particular choice stems from the desire to test the
thus enhancing the heat flow in the expected direction thermal resonance phenomenon in a setup in which the
when there is no external driving. For larger A1 values the second frequency is incommensurable with the first one,
current reversal effect is maintained and the magnitude thus avoiding any trivial summation of both contributions
of the heat flux increases, becoming almost independent that could occur at integer multiples of a given period.
of ∆T for large enough values of the amplitude A1 (not As can be readily observed in Figure 3 in both depicted
shown). In Figure 2b, we again plot Jmax , but now as a instances an increase in A1 with A2 = 0 renders a signifi-
function of the average temperature T0 for increasing ∆T cant increment of Jmax . When A1 = A2 = 1 an increment
values. Under these conditions only for ∆T = 0.2 there is of lower magnitude than the former is obtained, but with
a well defined current reversal effect that slightly increases an important new feature: there is a shift in ωmax to
as the T0 value does so. Already for ∆T = 0.6 this effect lower frequencies,
√ which corresponds to ωmax = 0.9 for
is greatly diminished, with a Jmax magnitude almost ω2 /ω = 2, Figure 3a, and to ωmax = 0.6 for ω2 /ω = e,
independent of the T0 value. As the temperature differ- Figure 3b, being e the Euler constant. This effect can
ence further increases it completely disappears, i.e. Jmax be understood if we recall that, as more external driv-
remains negative, with its absolute magnitude increasing ing frequencies are involved, the power supply (∼ ω 2 )
in magnitude as T0 takes larger values at fixed tempera- into the lattice is increased and thus more low frequency
ture difference. Thus, our remarks are the same as those phonons associated with heat conduction are involved
at the end of the previous paragraph. Finally, if close to in the transport process. Therefore, thermal resonance
Page 4 of 8 Eur. Phys. J. B (2019) 92: 5

0.25
(a)
(a) A11 = A2 = 0 0.2
TR = 0.6
0.2 A11 = 1.0 , A2 = 0 TR = 1.0
A11 = 2.0 , A2 = 0 TR = 1.4
0.15
J

A1 = A2 = 1.0 0.1 TR = 1.6


0.1 A1 = 2.0 , A2 = 1.0
TR = 2.0

J
A1 = 1.0 , A2 = 2.0
0.05
0
0
0.25
-0.1
(b) 0.2

0.15 0.01 0.1 1 10


ω
J

0.1
(b)
0.05 2.5

0
2
0.01 0.1 1 10

Ti
ω
1.5
Fig. 3. Heat flux J vs. driving frequency ω for various
√ values
of the amplitudes A1 and A2 , ω1 ≡ ω, (a) ω2 /ω = 2, and
(b) ω2 /ω = e. Same V , k, and N values as in Figure 1, with 1
TL = 0.4 and TR = 0.6. Filled circles correspond to A1 = 1
and A2 = 0, filled squares to A1 = 2 and A2 = 0, void circles 0 10 20 30 40 50 60
to A1 = A2 = 1, void squares to A2 = 2 and A1 = 1, and i
diamonds to A1 = 1 and A2 = 2. Continuous lines are a guide
to the eye and dot-dashed line correspond to the case with no Fig. 4. (a) Heat flux J vs. driving frequency for an external
external forcing, i.e. A1 = A2 = 0. force with parameters A1 = 1, A2 = 2, ω1 = ω, and ω2 = 2ω
for various TR values. Same N and TL values as in Figure 1.
Maximum heat flux is at ωmax = 0.7. (b) Temperature profiles
occurs at the lower frequencies associated with corresponding to the ωmax value in panel (a) for different values
√ the afore- of TR , from low to high values increasing from bottom to top.
mentioned phonons. Now, for the ω2 /ω = 2 case we
notice that, for A1 = 2 and A2 = 1 the increment of the
Jmax value is substantial, from Jmax = 0.06 to 0.25; if
we now take A1 = 1 and A2 = 2 the peak in the plot J feature is that both the thermal resonance and current
versus ω is somewhat sharper, but with no other new fea- reversal effects are present in the whole temperature range
ture. However, the behavior is different for the ω2 /ω = e studied, as can be seen from the temperature profiles pre-
instance: with A1 = 2 and A2 = 1 the behavior of J as a sented in Figure 4b, all of them with a negative slope
function of ω resembles quite close that of the case A1 = 2 and thus consistent with a positive heat flux. The cur-
and A2 = 0, i.e. the effect of the second harmonic is neg- rent reversal effect decreases as TR increases, just as in
ligible, with no frequency shift altogether. Nevertheless, if the single-frequency case, but now each Jmax value corre-
we now take A1 = 1 and A2 = 2 we obtain not only an sponds to an inversion in the heat current direction, from
increment in the Jmax value with a more defined peak, negative to positive. The thermal resonance suffers a dra-
but also a larger frequency shift, and thus a lower value matic increase since the heat flux changes from −0.0076, in
of ωmax = 0.5, than that obtained with A1 = A2 = 1. We both the adiabatic and fast-driving limits, to Jmax = 0.24
thus see that, with the inclusion of a second forcing term for TR = 0.6 and from −0.135 to Jmax = 0.12 for TR = 2,
with larger amplitude than the first, a dramatic incre- both at a resonance frequency of ωmax = 0.7.
ment of the magnitude of the current reversal effect is Next in Figure 5 we present the same results that in
obtained; the observed frequency shift even hints to the Figure 2, but now for a biharmonic forcing. Although Jmax
possibility of controlling the response of the system by a decreases as ∆T increases, as can be observed in Figure 5a,
manipulation of the parameters defining the external forc- the current reversal effect still exists for moderate ampli-
ing, something that seems not to be feasible with a single tude values, which may be more readily implemented than
frequency forcing. Therefore, in the following we will take the large magnitudes needed for the singe-harmonic case
mainly the values A1 = 1 and A2 = 2 for the amplitudes to maintain this effect. As for the dependence of Jmax
of the components of the external forcing. on T0 depicted in Figure 5b the results are definitely
In Figure 4a, we present our results when the external better than the corresponding ones in Figure 2b, since,
driving has the parameters ω1 ≡ ω, ω2 = 2ω, A1 = 1, and although as ∆T increases the corresponding data in the
A2 = 2 for J versus ω. It is clear that, when the second curve decrease in value compared with the ones in the pre-
harmonic is present in the external drive, the Jmax values vious one, the maximum heat flux increases as T0 does so
corresponding to each curve have both a larger magnitude in all instances and the obtained values are larger than
than the corresponding single-harmonic case and a lower those corresponding to the single-harmonic case. There-
decrease rate when TR increases. But the most remarkable fore, so far we can conclude that the biharmonic forcing
Eur. Phys. J. B (2019) 92: 5 Page 5 of 8

(a) 0.3
T0 = 0.5
0.2 10
T0 = 1
0.1

Jmax
Jmax

0 5
A2 = 0
A2 = 0.5
-0.1 A2 = 1.0
A2 = 1.5
-0.2 A2 = 2.0 0
0.5 1 1.5 0 2 4 6 8 10
∆T A2
(b)
0.4 ∆T = 0.2
Fig. 6. Maximum heat flux Jmax vs. amplitude A2 with A1 ,
∆T = 0.6
∆T = 1.0 ∆T = 0.2 and two different T0 values. Same V , k, ω1 , ω2 , and
∆T = 1.4
0.3 N values as in Figure 4. Dashed and dotted lines correspond
Jmax

to the power-law fit to the data. See text for details.

0.2

0.1 than those in Figure 4b; nevertheless, the magnitude of


Jmax for each curve in Figure 4a is almost identical to
0.5 1 1.5 2 the corresponding ones obtained with the previous choice
T A1 = 1 and A2 = 2. As for the value of the critical fre-
Fig. 5. (a) Same as Figure 2a, but with the values A1 , A2 ,
quency, ωmax = 0.9, we notice that it is identical to the
ω1 , and ω2 employed in Figure 4. (b) Same information as in case depicted in Figure 3a, thus corroborating that only
Figure 2b and with the same A1 , A2 , ω1 , and ω2 values as in for small values of the quotient ω2 /ω there is any shift
Figure 4a. Continuous lines are a guide to the eye. of ωmax to lower frequencies at all; for larger values of
the aforementioned quotient ωmax is identical to the value
obtained when A2 = 0. Therefore, we have determined
significantly enhances the performance of the device under that the manipulation of the relative amplitudes of both
the considered temperature conditions compared to the forcing terms is of lesser importance than the value of the
single-harmonic forcing previously considered. quotient of both frequencies in obtaining a shift of the
Since the previously reported results for the biharmonic critical frequency in which the change of direction of the
forcing were obtained for A1 = 1 and A2 = 2, and in view heat flux occurs and that this phenomenology seems to be
of Figure 5a which shows an increase of the Jmax value largely temperature independent.
as the amplitude of the second harmonic grows at either Besides improving the current reversal effect for dif-
fixed ∆T or T0 , it seems compelling to explore the behav- ferent temperature conditions, the biharmonic forcing
ior of the aforementioned quantity for larger A2 values. reduces the magnitude of the critical frequency wherewith
In Figure 6, we present the dependence of Jmax with A2 the resonance effect appears; the reduction was reported
for two values of the temperature difference. A power-like in Figure 3 for very specific values of the quotient ω2 /ω.
dependence Jmax ∼ Aα 2 is obtained in both instances, with
Therefore, it seems compelling to explore the effect of
α = 2.4 for T0 = 0.5 and α = 2.3 for T0 = 1. This result the second forcing frequency on thermal resonance for
shows that almost arbitrary increments in the maximum a wider value range of the aforementioned quotient. In
heat flux can be obtained for an increase in amplitude, Figure 8a, we present our results for a biharmonic forc-
although maybe this would not be the best way to do so ing wherein the quotient ω2 /ω is varied with A1 = 1 and
because an increase in amplitude could compromise the A2 = 2 for values of ω2 /ω > 1. This shift of Jmax towards
structural stability of the lattice. lower frequencies can be explained by recalling that, as
Now we explore the temperature dependence of the higher harmonics are employed in the external driving
behavior reported in Figure 3 for A1 = 2 and A2 = 1, with force, more energy is provided to the FK lattice and its
ω1 = ω and ω2 = 2ω; the result is presented in Figure 7. dynamical behavior becomes closer to √ that of a harmonic
It can be noticed that, for the chosen ω2 /ω value, there one with a phonon band of 0 < ω < 4k, as previously
is a competition between the two forcing terms that is explained. Thus, as the energy in the system increases
reflected in a maximum value of the heat flux not as well more low-frequency phonons, which are mostly responsi-
defined as in Figure 4a; the observed shoulder after the ble of the heat transport process, are involved and the
Jmax value reflects the influence of the second driving resonance frequency ωmax is shifted towards lower val-
term. But except for this feature the behavior is identical ues, although the associated Jmax diminishes in magnitude
to that depicted in Figure 4a with A1 = 1 and A2 = 2, i.e., because, in this frequency range, the system approaches
a decreasing value of Jmax as TR increases, with the cur- the adiabatic limit. At ω2 /ω = 1 the resonant heat flux
rent reversal still present. Furthermore, the temperature is maximum, with a value of Jmax = 0.267, as can be
profiles depicted in Figure 7b present larger fluctuations observed in Figure 8b where the dependence of Jmax with
Page 6 of 8 Eur. Phys. J. B (2019) 92: 5

(a) TR = 0.6 0.25


TR = 1.0
k=1 (a)
0.2 0.2 k=2
TR = 1.4 k=3
0.15 k=4
TR = 1.6 k=5

J
0.1 TR = 2.0
0.1
J

0.05
0
0
-0.1 0.01 0.1 1 10
ω
2
0.01 0.1 1 10 0.25
ω (b)
0.2 1.5
(b)

ωmax
Jmax
2.5 0.15
0.1 1
2 (c)
0.05
Ti

0.5
1 2 3 4 5 1 2 3 4 5
1.5 k k
Fig. 9. (a) Heat flux J vs. driving frequency for an external
1
force with parameters TL = 0.4, TR = 0.6, A1 = 1, A2 = 2,
ω1 = ω, and ω2 = 2ω for various k values. (b) Jmax and
0 10 20 30 40 50 60
(c) ωmax vs. k. Continuous lines in (a) are a guide to the eye,
i whereas in (b) and (c) it represents the best fit to the data.
See text for details.
Fig. 7. (a) Heat flux J vs. driving frequency for an external
force with parameters A1 = 2, A2 = 1, ω1 = ω, and ω2 = 2ω
for various TR values. Same N and TL values as in Figure 1.
Maximum heat flux is at ωmax = 0.9. (b) Temperature profiles In Figure 9, we explore the influence of the harmonic
corresponding to the ωmax value in panel (a) for different values coupling constant on thermal resonance. As the former
of TR , from low to high values increasing from bottom to top. increases, the maximum heat flux shifts to higher frequen-
cies, just as in the single-forced case already reported [21].
For the maximum heat flux, we have a dependence on the
(b) harmonic constant of the form Jmax = 1/(αk + β) with
0.25
α = 0.4 and β = 0.3 that can be readily appreciated in
Figure 9b. From Figure 9c we obtain a linear dependence
Jmax

0.25
1/2
ω2 = 2 ω 1 (a) 0.2
ω2 = 2ω1
ω2 = eω1 of the form ωmax ∼ Ak with A = 0.27; this result is very
0.2 ω2 = 3ω1
ω2 = πω1 0.15 different to the single force instance, since in the latter one
ω2 = 4ω1

0.15 ω2 = 5ω1
0.1
ωmax has a power-law dependence on the harmonic con-
ω2 = 6ω1
stant. This result exemplifies the complicated interplay
J

6 (c)
0.1
5 between the contribution to the thermal resonance effect
0.05 4 of structural properties, such as k in this case, and the
ωmax

3 form of the external forcing, which makes highly nontriv-


0
2 ial to predict the outcome for a particular instance. At this
0.01 0.1 1 10 1 point it is important to notice that, for the single forced
ω 0
instance, the shift of ωmax to higher frequencies is accom-
0 1 2 3 4 5 6
ω2 / ω1
panied by a weakening in the thermal transport, i.e. the
value of Jmax steadily diminishes, which will eventually
Fig. 8. (a) Heat flux J vs. driving frequency for an external lead to the disappearance of thermal resonance altogether
force with parameters TL = 0.4, TR = 0.6, A1 = 1, A2 = 2, for large enough k values. The situation is completely dif-
ω1 = ω, and various values of the quotient ω2 /ω > 1. (c) Jmax ferent for the biharmonic forcing case, where even the
and (d) ωmax vs. ω2 /ω. Continuous lines are a guide to the eye curve which has the lowest maximum frequency, k = 5,
in (a) and (b), whereas in (c) they represent the best fit to the has a higher Jmax value compared to that obtained in the
data. original, single-forced case.
We also explored the effect of the magnitude of the non-
linear on-site potential on the thermal resonance present
ω2 /ω is presented. Next, in Figure 8c the decrease of the in this system. In Figure 10a, we readily notice that
resonant frequency ωmax with respect to the frequency the resonance peak decreases as the height of the on-site
quotient is observed to follow a decrement of the form potential increases in a monotonic way, as can be observed
ωmax = 1/(αx + β), with α = 0.8 and β = −0.008. This in Figure 10b. This phenomenology arises because, as V
power-like behavior indicates that, as ω2 /ω increases, the increases, the oscillators are more confined into the valleys
critical frequency approaches an asymptotic value. of the on-site potential, hindering the heat transport and
Eur. Phys. J. B (2019) 92: 5 Page 7 of 8

0.4 (a) 50
V=3 (a) N = 64
0.3 V=4 40 N = 512
V=5 N = 1024
V=6 30
0.2 V=7 N = 2048

JN
J

20
0.1
10
0
0
0.01 0.1 1 10
ω 0.01 0.1 1 10
0.4 3
ω
(b) (c)
(b) 0.8 (c) 2.5
3.5
0.3
ωmax
Jmax

2
0.7

κ
Ti
3.4
1.5
0.2
0.6 1 3.3
0.5
0.1
3 4 5 6 7 3 4 5 6 7 0 0.5 1 100 1000
V V i/N N

Fig. 10. (a) Heat flux J vs. driving frequency for an external Fig. 11. (a) Total heat flux JN vs. driving frequency for an
force with parameters TL = 0.4, TR = 0.6, A1 = 1, A2 = 2, external force with parameters TL = 0.4, TR = 0.6, V = 5,
ω1 = ω, and ω2 = 2ω for various V values. (b) Jmax and A1 = 1, A2 = 2, ω1 = ω, and ω2 = 2ω for various system
(c) ωmax vs. V . Continuous lines are a guide to the eye in sizes N . (b) Temperature profiles for all reported system sizes.
all instances. (c) Thermal conductivity vs. system size N . Continuous lines
are a guide to the eye in all instances.

thus decreasing the Jmax value. Therefore, the response motors. First we determined that the original single-
time of the system is increased, which entails a shift of frequency external forcing rendered very weak resonance
the resonant frequency ωmax to a lower value within the and heat flux inversion effects that rapidly disappear as
allowed phonon band as V increases in a non-monotonic either the mean temperature or the one of the hot reservoir
way, as can be appreciated in Figure 10c. The resonance is increased. When the second harmonic external forcing
regime is thus driven towards longer timescales, although is included a significant increase in the magnitude of both
very large amplitudes would be needed to obtain any fur- aforementioned effects is obtained, which are also present
ther decrements in the ωmax value. Also the frequency in a larger temperature range than the single-frequency
shift occurs in a narrower frequency range compared to case. We also determined that, if the second harmonic
the corresponding single-forced case. forcing is a higher harmonic than the first one, the fre-
The size effects on thermal resonance is explored for quency at which thermal resonance appears shifts to lower
various system sizes and is reported in Figure 11. The frequencies, with a very modest decrease in its magnitude
phenomenology is the same as in the single-frequency case: as higher second frequencies are considered. The maxi-
the total heat flux increases up to a saturation value, the mum heat flux is obtained when the frequencies of both
effect does not disappear in the large system-size limit, components are identical and equal to the resonance fre-
as can be readily appreciated in Figure 11a, and the res- quency of the single-frequency forcing. Thus, by varying
onance frequency is independent of the system size. It is parameters of the external forcing we can achieve an effec-
worth noticing that the maximum heat flux is an order of tive control of the frequency range in which we desire the
magnitude larger than the single-frequency case. The only thermal resonance to appear within the allowed phonon
significant differences are that the convergence to the sat- band, although with a somewhat diminished magnitude.
uration value is somewhat slower in the biharmonic case, Nevertheless, we believe that this is an adequate trade-off,
as can be appreciated in the temperature profiles for the since varying the strength of the harmonic coupling can
employed system sizes depicted in Figure 11b and in the shift the value of the critical frequency, but just to higher
behavior of the thermal conductivity κ(N ) = −JN/∆T values and with an important decrease of the magnitude
as a function of the system size presented in Figure 11c of the heat flux, whereas a change in the magnitude of the
compared to the single-frequency case [21]. on-site potential decreases the aforementioned magnitude
with only a very weak shift of the critical frequency.
Since our proposed modification entails an external
4 Concluding remarks mechanism independent of the structural details of the
employed model, the herein employed driving force could
In this work, we have modified the previously proposed be readily implemented experimentally by many different
setup of a FK anharmonic lattice subjected to a temper- means. This is specially important if a future technolog-
ature gradient by means of a biharmonic external forcing ical application could be envisioned, since the external
that was previously proposed in the context of Brownian driving can be tailored according to the proposed use
Page 8 of 8 Eur. Phys. J. B (2019) 92: 5

of the thermal resonance. Further research is being per- B.J. Lopez, P.M.G. Curmi, N.R. Forde, D.N. Woolfson,
formed to assess if other structural modifications in the H. Linke, HFSP J. 3, 204 (2009)
oscillator lattice could improve the observed effect in the 6. H. Murakami, A. Kawabuchi, K. Kotoo, M. Kunitake,
studied temperature range. The present work is a start- N. Nakashima, J. Am. Chem. Soc. 119, 7605 (1997)
ing point in the study of external control of the heat flux 7. C. Cheng, P.R. McGonigal, S.T. Schneebeli, H. Li, N.A.
in the nanometric scale that could find application in the Vermeulen, C. Ke, J.F. Stoddart, Nat. Nanotechnol. 10,
management of heat flux through tailor-build channels in 547 (2015)
artificial membranes, just to mention one many possible 8. V. Blickle, C. Bechinger, Nat. Phys. 8, 143 (2012)
applications. 9. I. Bargatin, M.L. Roukes, Phys. Rev. Lett. 91, 138302
(2003)
10. J. Feng, M. Graf, K. Liu, D. Ovchinnikov, D. Dumcenco,
M.R.B. thanks CONACyT, México and S.G.G. thanks “Pro-
M. Heiranian, V. Nandigana, N.R. Aluru, A. Kis, A.
grama Institucional de Formación de Investigadores” I.P.N., Radenovic, Nature 536, 197 (2016)
México for financial support. The authors also thank Salvador 11. M. Terraneo, M. Peyrard, G. Casati, Phys. Rev. Lett. 88,
Malagón Leon, Nely Sasiova, and Juan M. López for useful 094302 (2002)
comments and discussions. 12. B. Li, L. Wang, G. Casati, Phys. Rev. Lett. 93, 184301
(2004)
Author contribution statement 13. D. Segal, A. Nitzan, Phys. Rev. Lett. 94, 034301 (2005)
14. B. Li, L. Wang, G. Casati, Appl. Phys. Lett. 88, 143501
(2006)
M.R.B. and S.G.G. performed numerical simulations.
15. L. Wang, B. Li, Phys. Rev. Lett. 99, 177208 (2007)
M.R.B. devised the study and wrote the paper. 16. L. Wang, B. Li, Phys. Rev. Lett. 101, 267203 (2008)
17. D. Segal, A. Nitzan, Phys. Rev. E 73, 026109 (2006)
References 18. F. Zhan, N. Li, S. Kohler, P. Hänggi, Phys. Rev. E 80,
061115 (2009)
1. M. Bencowe, Science 304, 56 (2004) 19. N. Li, P. Hänggi, B. Li, Europhys. Lett. 84, 40009
2. F. Gianzotto, T.T. Hikkilä, A. Luukanen, A.M. Savin, J.P. (2008)
Pekola, Rev. Mod. Phys. 78, 217 (2006) 20. N. Li, F. Zhan, P. Hänggi, B. Li, Phys. Rev. E 80, 011125
3. M. Maldovan, Nature 503, 209 (2013) (2009)
4. D.G. Cahill, P.V. Braun, G. Chen, D.R. Clarke, S. Fan, 21. B.Q. Ai, D. He, B. Hu, Phys. Rev. E 81, 031124 (2010)
K.E. Goodson, P. Keblinski, W.P. King, G.D. Mahan, 22. P. Reimann, Phys. Rep. 361, 57 (2002)
A. Majumdar, H.J. Maris, S.R. Phillpot, E. Pop, L. Shi, 23. P. Hänggi, F. Marchesoni, Rev. Mod. Phys. 81, 387
Appl. Phys. Rev. 1, 011305 (2014) (2009)
5. E.H.C. Bromley, N.J. Kuwada, M.J. Zuckermann, 24. N. Li, J. Ren, L. Wang, G. Zhang, P. Hänggi, B. Li, Rev.
R. Donadini, L. Samii, G.A. Blab, G.J. Gemmen, Mod. Phys. 84, 1045 (2012)

You might also like