This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Report

Aniket Singha Department of Electronics and Electrical Communication Engineering,
Indian Institute of Technology Kharagpur, Kharagpur-721302, India

A realistic non-local heat engine based on Coulomb coupled systems

Aniket Singha Department of Electronics and Electrical Communication Engineering,
Indian Institute of Technology Kharagpur, Kharagpur-721302, India
Abstract

Optimal non-local heat-engines, based on Coulomb-coupled systems, demand a sharp step-like change in the energy resolved system-to-reservoir coupling around the ground state of quantum-dots Walldorf et al. [2017]; Daré [2019]; Zhang and Chen [2019]; Daré and Lombardo [2017]; Sánchez and Büttiker [2011]; Erdman et al. [2018]. Such a sharp step-like transition in the system-to-reservoir coupling cannot be achieved in a realistic scenario. Here, I propose realistic design for non-local heat engine based on Coulomb-coupled system, which circumvents the need for any change in the system-to-reservoir coupling, demanded by the optimal set-ups discussed in literature. I demonstrate that an intentionally introduced asymmetry (or energy difference) in the ground state configuration between adjacent tunnel coupled quantum dots, in conjugation with Coulomb coupling, is sufficient to convert the stochastic fluctuations from a non-local heat source into a directed flow of thermoelectric current. The performance, along with the regime of operation, of the proposed heat engine is then theoretically investigated using quantum-master-equation (QME) approach. It is demonstrated that the theoretical maximum power output for the proposed set-up is limited to about 50%50\% of the optimal design. Despite a lower performance compared to the optimal set-up, the novelty of the proposed design lies in the conjunction of fabrication simplicity along with reasonable power output. At the end, the sequential transport processes leading to a performance deterioration of the proposed set-up are analyzed and a method to alleviate such transport processes is discussed. The set-up proposed in this paper can be used to design and fabricate high-performance non-local cryogenic heat engines.

I Introduction

Refer to caption
Figure 1: Schematic diagram of the proposed non-local heat-engine based on Coulomb-coupled quantum dots. The entire system consists of three dots S1,S2S_{1},~{}S_{2} and G1G_{1} which are electrically coupled to the reservoirs L,RL,~{}R and GG respectively. The dots S1S_{1} and S2S_{2} are tunnel coupled, while S1S_{1} and G1G_{1} are Coulomb coupled. The ground states of S1S_{1} and S2S_{2} form a staircase configuration with εs2=εs1+Δε\varepsilon_{s}^{2}=\varepsilon_{s}^{1}+\Delta\varepsilon. In the proposed arrangement, current can be driven between the cold reservoirs LL and RR by absorbing thermal energy from the hot reservoir GG.

With the progress in fabrication and scaling technology, efficient heat harvesting in lower dimensional systems has gained a lot of attention Singha et al. [2015]; Singha [2020a]; Singha and Muralidharan [2017]; Mi et al. [2008]; Li et al. [2010]; Sumithra et al. [2012]; Bahk et al. [2013]; Agarwal and Muralidharan [2014]; Kim and Lundstrom [2012, 2011]; Faleev and Léonard [2008]; Neophytou and Kosina [2013]; Singha [2018]; Singha and Muralidharan [2018]. One of the major issues affecting heat harvesting performance in nano-systems is the drastic lattice heat flux resulting from a small spatial separation between the heat reservoir and the heat sink. The large lattice heat flux severely limits the overall efficiency of the heat engine and poses a major performance issue in cases where supply of heat energy is limited. Tailoring the lattice thermal conductance, in an attempt to gain enhanced harvesting efficiency generally affects the current path, thereby deteriorating the peak harvested power. As such, one of the crucial focus of the modern thermoelectric community is to facilitate an independent optimization of the electron transport path and lattice heat conduction path, by introducing a spatial separation between the current path and the heat source Uchida et al. [2016]; Sinova et al. [2015]; Saitoh et al. [2006]. This phenomenon of harvesting heat from a reservoir, which is spatially separated from the current conduction path, is known as non-local heat harvesting Uchida et al. [2016]; Sinova et al. [2015]; Saitoh et al. [2006]; Walldorf et al. [2017]; Daré [2019]; Zhang and Chen [2019]; Daré and Lombardo [2017]; Zhang et al. [2016]; Sánchez and Büttiker [2011]. Thus, non-local heat engines are three terminal systems where power is delivered between two terminals while extracting heat from the third terminal which is spatially separated from the path of current flow. \colorblack In this case, tailoring the lattice heat transport path, in an attempt to gain enhanced efficiency, can be accomplished without altering the current conduction path. Recently designs and concepts of non-local heat engines and refrigerators using Coulomb coupled quantum dots have been proposed and explored theoretically Walldorf et al. [2017]; Daré [2019]; Zhang and Chen [2019]; Daré and Lombardo [2017]; Zhang et al. [2016]; Sánchez and Büttiker [2011] as well as experimentally Thierschmann et al. [2015] in literature. However, the operation of such non-local heat engines demand a sharp step-like change in the system-to-reservoir coupling around the ground state energy Walldorf et al. [2017]; Daré [2019]; Zhang and Chen [2019]; Daré and Lombardo [2017]; Zhang et al. [2016]; Sánchez and Büttiker [2011], which is impossible to achieve in a practical scenario Thierschmann et al. [2015]. In this paper, I propose a realistic design strategy to accomplish non-local heat harvesting using Coulomb coupled quantum dots. Unlike the optimal non-local heat-engine based on Coulomb coupled systems Walldorf et al. [2017]; Daré [2019]; Zhang and Chen [2019]; Daré and Lombardo [2017]; Zhang et al. [2016]; Sánchez and Büttiker [2011], the proposed design doesn’t demand a change in the system-to-reservoir coupling near the ground state. The performance proposed heat-engine is then evaluated and compared with the optimal set-up. It is demonstrated that the performance of the proposed heat engine hovers around 50%50\% of the optimal set-up. However, the novelty of the proposed set-up is the conjugation of fabrication simplicity along with a reasonable power output. At the end, the processes leading to a performance deterioration of the proposed set-up are discussed and analyzed. This paper is organized as follows. In Sec. II, I illustrate the proposed design strategy and elaborate the transport formulation employed to analyze the thermoelectric performance of the same. Next, Sec III elaborates a detailed analysis on the heat harvesting performance and regime of operation of the proposed heat-engine. A performance comparison between the proposed heat engine and the optimal set-up is also conducted along with a brief discussion on the transport processes leading to a performance deterioration of the proposed heat engine. I conclude this paper briefly in Sec. IV. The Appendix Sec. elaborates some intuitive and conceptual understanding of the proposed heat engine operation, as well as details the derivations of the equations employed to investigate the performance of the heat engine.

II Proposed design and transport formulation

The proposed heat engine, schematically demonstrated in Fig. 1, consists of three dots S1,S2S_{1},~{}S_{2} and G1G_{1} which are electrically coupled to the reservoirs LL, RR and GG respectively. I will now discuss the ground state configuration and other features of the system consisting of the three dots S1,S2S_{1},~{}S_{2} and G1G_{1}. S1S_{1} and S2S_{2} are tunnel coupled to each other, while G1G_{1} is capacitively coupled to S1S_{1}. The ground states of S1S_{1} and S2S_{2} form a stair-case configuration with εs2=εs1+Δε\varepsilon_{s}^{2}=\varepsilon_{s}^{1}+\Delta\varepsilon. Any electronic tunneling between the dots S1S_{1} and G1G_{1} is suppressed via suitable fabrication techniques. Energy exchange between the two dots is, however, possible via Coulomb coupling Hübel et al. [2007a]; Chan et al. [2002]; Molenkamp et al. [1995]; Hübel et al. [2007b]; Ruzin et al. [1992]. In the optimal Coulomb-coupled system based heat-engine discussed in literature, an asymmetric system-to-reservoir coupling is required for heat harvesting Sánchez and Büttiker [2011]; Erdman et al. [2018]. In the proposed set-up, instead of the asymmetric system-to-reservoir coupling, the system itself is made asymmetric with respect to the reservoir LL and RR by choosing an energy difference between the ground states of the adjacent quantum dots S1S_{1} and S2S_{2}, with εs2=εs1+Δε\varepsilon_{s}^{2}=\varepsilon_{s}^{1}+\Delta\varepsilon. The purpose of such arrangement is to deliver power to an external load connected between LL and RR while extracting heat from reservoir GG. Another equivalent realistic set-up, based on Coulomb coupled systems, that can be employed for efficient non-local heat harvesting is demonstrated in Fig. 7 and discussed briefly in Appendix A. \colorblack\colorblack Due to random electron exchange between the dot G1G_{1} and the reservoir GG, the total energy of the system fluctuates stochastically with time. For a Coulomb coupled system with constant energy resolved system-to-reservoir coupling and with symmetric ground state configuration with respect to reservoir LL and RR, the stochastic fluctuations result in the same probability of electron transfer from LL to RR and from RR to LL, making the average current zero, although there is a non-zero component of electronic heat flow between GG and L(R)L(R) (See Appendix B). However, for a Coulomb coupled system with asymmetric ground state configuration, the same fluctuation can be converted to a directed current flow when TGTL(R)T_{G}\neq T_{L(R)} (Appendix C). I will demonstrate via numerical calculations and theoretical arguments (given in Appendix C) that in the system detailed above, thermoelectric power can be delivered between the terminals LL and RR (with a net electronic flow from LL to RR) by extracting heat energy from the reservoir GG when TG>TL(R)T_{G}>T_{L}(R). Thus, power can be delivered between terminals non-local to the heat source. The excess energy Δε=εs2εs1\Delta\varepsilon=\varepsilon_{s}^{2}-\varepsilon_{s}^{1}, required for the electrons to tunnel from S1S_{1} to S2S_{2} is supplied from the reservoir GG via Coulomb coupling. Coming to the realistic fabrication possibility of such a system, due to the recent advancement in solid-state nano-fabrication technology, triple and quad quantum dot systems with and without Coulomb coupling have already been realized experimentally Eng et al. [2015]; Flentje et al. [2017]; Froning et al. [2018]; Noiri et al. [2017]; Hong et al. [2018]; Takakura et al. [2014]. In addition, it has been experimentally demonstrated that quantum dots that are far from each other in space, may be bridged to obtain strong Coulomb coupling, along with excellent thermal isolation between the hot and cold reservoirs Hübel et al. [2007a]; Chan et al. [2002]; Molenkamp et al. [1995]; Hübel et al. [2007b]; Ruzin et al. [1992]. Also, the bridge may be fabricated between two specific quantum dots to drastically enhance their mutual Coulomb coupling, without affecting the electrostatic energy of the other quantum dots Hübel et al. [2007a]; Chan et al. [2002]; Molenkamp et al. [1995]; Hübel et al. [2007b]; Ruzin et al. [1992]. \colorblack Thus, the change in electron number nS1(nG1)n_{S_{1}}~{}(n_{G_{1}}) of the dot S1S_{1} (G1G_{1}) influences the electrostatic energy of the dot G1G_{1} (S1S_{1}). In general, the total increase in electrostatic energy UU of the configuration demonstrated in Fig. 1 (a), consisting of three dots, due to fluctuation in electron number can be given by (Refer to Appendix D for detailed derivation):

U(nS1,nG1,nS2)=xUxself(nxtotnxeq)2+(x1,x2)x1x2Ux1,x2m(nx1totnx1eq)(nx2totnx2eq)U(n_{S_{1}},n_{G_{1}},n_{S_{2}})=\sum_{x}U^{self}_{x}\left(n_{x}^{tot}-n_{x}^{eq}\right)^{2}+\sum_{(x_{1},x_{2})}^{x_{1}\neq x_{2}}U^{m}_{x_{1},x_{2}}\left(n_{x_{1}}^{tot}-n_{x_{1}}^{eq}\right)\left(n_{x_{2}}^{tot}-n_{x_{2}}^{eq}\right) (1)

where nxtotn_{x}^{tot} is the total electron number, and Uxself=q2CxselfU^{self}_{x}=\frac{q^{2}}{C^{self}_{x}} is the electrostatic energy due to self-capacitance CxselfC^{self}_{x} (with the surrounding leads) of quantum dot ‘xx’ (Refer to Appendix D for details). Ux1,x2mU^{m}_{x_{1},x_{2}} is the electrostatic energy arising out of Coulomb coupling between two different quantum dots that are separated in space (Appendix D) and nxeqn_{x}^{eq} is the total number of electrons present in dot xx in equilibrium at 0K0K and is determined by the minimum possible electrostatic energy of the system. nx=nxtotnxeqn_{x}=n_{x}^{tot}-n_{x}^{eq} is the total number of excess electrons added in the ground state of the dot xx due to stochastic fluctuations from the reservoirs (Refer to Appendix D for details). Here, a minimal physics based model is used to investigate the heat engine performance under the assumption that the change in potential due self-capacitance is much greater than than the average thermal voltage kT/qkT/q or the applied bias voltage VV, that is Uxself=q2Cxself>>(kT,qV)U^{self}_{x}=\frac{q^{2}}{C^{self}_{x}}>>(kT,~{}qV). Hence, electron occupation probability or transfer rate via the Coulomb blocked energy level, due to self-capacitance, is negligibly small. The analysis of the entire system of dots may hence be approximated by limiting the maximum number of electrons in each dot to one. Thus the analysis of the entire system may be limited to eight multi-electron levels, which I denote by the electron occupation number in the ground state of each quantum dot. Hence, a possible state of interest in the system may be denoted as |nS1,nG1,nS2=|nS1|nG1|nS2\ket{n_{S_{1}},n_{G_{1}},n_{S_{2}}}=\ket{n_{S_{1}}}\mathbin{\mathop{\otimes}\limits}\ket{n_{G_{1}}}\mathbin{\mathop{\otimes}\limits}\ket{n_{S_{2}}}, where nS1,nG1,nS2(0,1)n_{S_{1}},n_{G_{1}},n_{S_{2}}\in(0,1), are used to denote the number of electrons present in the ground-states of S1,G1S_{1},~{}G_{1} and S2S_{2} respectively. I also assume that the electrostatic coupling between S1,S2S_{1},~{}S_{2} and between S2,G1S_{2},~{}G_{1} are negligible, such that, for all practical purposes under consideration, US1,S2m0U^{m}_{S_{1},S_{2}}\approx 0 and UG1,S2m0U^{m}_{G_{1},S_{2}}\approx 0. Since, the electronic transport and ground states in S1S_{1} and G1G_{1} are mutually coupled, I treat the pair of dots S1S_{1} and G1G_{1} as a sub-system (ς1\varsigma_{1}), S2S_{2} being the complementary sub-system (ς2\varsigma_{2}) of the entire system consisting of three dots Singha [2020b]. The state probability of ς1\varsigma_{1} is denoted by Pi,jς1P_{i,j}^{\varsigma_{1}}, ii and jj being the number of electrons in the ground state of dot S1S_{1} and G1G_{1} respectively. Pkς2P_{k}^{\varsigma_{2}}, on the other hand, denotes the probability of occupancy of the dot S2S_{2} in the sub-system ς2\varsigma_{2}. It can be shown that if Δε\Delta\varepsilon is much greater than the ground state broadening due to system-to-reservoir coupling, then the interdot tunneling rate between S1S_{1} and S2S_{2} is optimized when εs1+US1,G1m=εs2\varepsilon_{s}^{1}+U^{m}_{S_{1},G_{1}}=\varepsilon_{s}^{2}, that is when Δε=US1,G1m\Delta\varepsilon=U^{m}_{S_{1},G_{1}} Singha [2020b] (See Appendix D). To evaluate the optimal performance of the proposed heat-engine, I hence assume Δε=US1,G1m\Delta\varepsilon=U^{m}_{S_{1},G_{1}} Singha [2020b]. Henceforth, I would simply represent US1,G1mU^{m}_{S_{1},G_{1}} as UmU_{m}. Under the assumption stated above, the equations governing sub-system state probabilities in steady state can be derived as Singha [2020b] (See Appendix D for detailed derivation):

P0,0ς1{fL(εs1)+fG(εg)}+P0,1ς1{1fG(εg)}+P1,0ς1{1fL(εs1)}=0\displaystyle-P_{0,0}^{\varsigma_{1}}\{f_{L}(\varepsilon_{s}^{1})+f_{G}(\varepsilon_{g})\}+P_{0,1}^{\varsigma_{1}}\{1-f_{G}(\varepsilon_{g})\}+P_{1,0}^{\varsigma_{1}}\{1-f_{L}(\varepsilon_{s}^{1})\}=0
P1,0ς1{1fL(εs1)+fG(εg+Um)}+P1,1ς1{1fG(εg+Um)}+P0,0ς1fL(εs1)\displaystyle-P_{1,0}^{\varsigma_{1}}\left\{1-f_{L}(\varepsilon_{s}^{1})+f_{G}(\varepsilon_{g}+U_{m})\right\}+P_{1,1}^{\varsigma_{1}}\left\{1-f_{G}(\varepsilon_{g}+U_{m})\right\}+P_{0,0}^{\varsigma_{1}}f_{L}(\varepsilon_{s}^{1})
P0,1ς1{1fg(εg1)+fL(εs1+Um)+γγcP1ς2}+P0,0ς1fG(εg)+P1,1ς1{1fL(εs1+Um)+γγcP0ς2}=0\displaystyle-P_{0,1}^{\varsigma_{1}}\left\{1-f_{g}(\varepsilon_{g}^{1})+f_{L}(\varepsilon_{s}^{1}+U_{m})+\frac{\gamma}{\gamma_{c}}P^{\varsigma_{2}}_{1}\right\}+P_{0,0}^{\varsigma_{1}}f_{G}(\varepsilon_{g})+P_{1,1}^{\varsigma_{1}}\left\{1-f_{L}(\varepsilon_{s}^{1}+U_{m})+\frac{\gamma}{\gamma_{c}}P^{\varsigma_{2}}_{0}\right\}=0
P1,1ς1{[1fg(εg1+Um)]+[1fL(εs1+Um)]+γγCP0ς2}+P1,0ς1fG(εg+Um)+P0,1ς1{fL(εs1+Um)+γγcP1ς2}=0\displaystyle-P_{1,1}^{\varsigma_{1}}\left\{[1-f_{g}(\varepsilon_{g}^{1}+U_{m})]+[1-f_{L}(\varepsilon_{s}^{1}+U_{m})]+\frac{\gamma}{\gamma_{C}}P^{\varsigma_{2}}_{0}\right\}+P_{1,0}^{\varsigma_{1}}f_{G}(\varepsilon_{g}+U_{m})+P_{0,1}^{\varsigma_{1}}\left\{f_{L}(\varepsilon_{s}^{1}+U_{m})+\frac{\gamma}{\gamma_{c}}P^{\varsigma_{2}}_{1}\right\}=0 (2)
P0ς2{fR(εs2)+γγcP1,1ς1}+P1ς2{1fR(εs2)+γγcP0,1ς1}=0\displaystyle-P_{0}^{\varsigma_{2}}\{f_{R}(\varepsilon_{s}^{2})+\frac{\gamma}{\gamma_{c}}P_{1,1}^{\varsigma_{1}}\}+P_{1}^{\varsigma_{2}}\left\{1-f_{R}(\varepsilon_{s}^{2})+\frac{\gamma}{\gamma_{c}}P^{\varsigma_{1}}_{0,1}\right\}=0
P1ς2{1fR(εs2)+γγcP0,1ς1}+P0ς2{fR(εs2)+γγcP1,1ς1}=0,\displaystyle-P_{1}^{\varsigma_{2}}\{1-f_{R}(\varepsilon_{s}^{2})+\frac{\gamma}{\gamma_{c}}P^{\varsigma_{1}}_{0,1}\}+P_{0}^{\varsigma_{2}}\left\{f_{R}(\varepsilon_{s}^{2})+\frac{\gamma}{\gamma_{c}}P_{1,1}^{\varsigma_{1}}\right\}=0, (3)

where γc=γl(ε)=γr(ε)=γg(ε)\gamma_{c}=\gamma_{l}(\varepsilon)=\gamma_{r}(\varepsilon)=\gamma_{g}(\varepsilon) and γ\gamma are related to the reservoir-to-system tunnel coupling and the inter-dot tunnel coupling respectively Singha [2020b]; Datta [2005], ε\varepsilon being the independent energy variable. In the above set of equations, fλ(ε)f_{\lambda}(\varepsilon) denotes the probability of occupancy of the reservoir λ\lambda at energy ε\varepsilon. For the purpose of calculations in this paper, I assume an equilibrium Fermi-Dirac statistics at the reservoirs. Hence, fλ(ε)f_{\lambda}(\varepsilon) is given by:

fλ(ε)=(1+exp{εμλkTλ})1,f_{\lambda}(\varepsilon)=\left(1+exp\left\{\frac{\varepsilon-\mu_{\lambda}}{kT_{\lambda}}\right\}\right)^{-1}, (4)

where μλ\mu_{\lambda} and TλT_{\lambda} respectively denote the quasi-Fermi energy and temperature of the reservoir λ\lambda. From the set of Eqns. (2) and (3), it is clear that an electron in S1S_{1} can tunnel into S2S_{2} only when the ground state in the dot G1G_{1} is occupied with an electron. The set of Eqns. (2) and (3) are coupled to each other and may be solved using any iterative method. Here, I use Newton-Raphson iterative method to solve the steady-state values of sub-system probabilities. On calculation of the sub-system state probabilities Pi,jς1P_{i,j}^{\varsigma_{1}} and Pkς2P_{k}^{\varsigma_{2}}, the electron current flow into (out of) the system from the reservoirs L(R)L(R) can be given as:

Refer to caption
Figure 2: Variation of the proposed heat engine performance with variation in the the ground states εg\varepsilon_{g} and εs1\varepsilon_{s}^{1} for Um=3.9meV(6kTq)U_{m}=3.9meV~{}(\approx 6\frac{kT}{q}) and V=1.3meV(2kTq)V=1.3meV~{}(\approx 2\frac{kT}{q}). Colour plot demonstrating the variation in (a) generated power (PP) and (b) efficiency η/ηc\eta/\eta_{c}. T=TL(R)+TG2=7.5KT=\frac{T_{L(R)}+T_{G}}{2}=7.5K is the average temperature between the heat source and the heat sink. The efficiency of generation is measured with respect to the Carnot efficiency ηc=1TL(R)/TG\eta_{c}=1-T_{L(R)}/T_{G}
IL=\displaystyle I_{L}= qγc×{P0,0ς1fL(εs1)+P0,1ς1fL(εs1+Um)}\displaystyle q\gamma_{c}\times\left\{P^{\varsigma_{1}}_{0,0}f_{L}(\varepsilon_{s}^{1})+P^{\varsigma_{1}}_{0,1}f_{L}(\varepsilon_{s}^{1}+U_{m})\right\}
qγcP1,0ς1{1fL(εs1)}qγcP1,1ςs1{1fL(εs1+Um)}\displaystyle-q\gamma_{c}P^{\varsigma_{1}}_{1,0}\{1-f_{L}(\varepsilon_{s}^{1})\}-q\gamma_{c}P^{\varsigma_{s}^{1}}_{1,1}\{1-f_{L}(\varepsilon_{s}^{1}+U_{m})\}
IR=\displaystyle I_{R}= qγc×{P0ς2fR(εs1)P1ς2{1fR(εs1)}},\displaystyle-q\gamma_{c}\times\left\{P^{\varsigma_{2}}_{0}f_{R}(\varepsilon_{s}^{1})-P^{\varsigma_{2}}_{1}\{1-f_{R}(\varepsilon_{s}^{1})\}\right\}, (5)

In addition, the electronic component of heat flow from the reservoir GG can be given by:

IQe=Umγc{P10ς1fG(εg+Um)P11ς1{1fG(εg+Um)}}I_{Qe}=U_{m}\gamma_{c}\left\{P^{\varsigma_{1}}_{10}f_{G}(\varepsilon_{g}+U_{m})-P^{\varsigma_{1}}_{11}\{1-f_{G}(\varepsilon_{g}+U_{m})\}\right\} (6)

Interestingly, we note that Eqn. (6) is not directly dependent on εg\varepsilon_{g}. This is due to the fact that the net electronic current into or out of the reservoir GG is zero (See Appendix D for details). Next, I use the set of Eqns. (2), (3), (5) and (6), to evaluate the thermoelectric generation performance of the the set-up demonstrated in Fig. 1. To analyze the performance of the heat engine, I use a voltage-controlled set-up demonstrated in literature Leijnse et al. [2010]; Sothmann et al. [2014]; Nakpathomkun et al. [2010], where a bias voltage VV is applied between LL and RR, with the positive and negative terminals being connected to LL and RR respectively, to emulate the voltage drop due to thermoelectric current flow across an external load. Thus the bias voltage VV mimics the voltage drop between the terminals across which power is to be delivered. Assuming the equilibrium electrochemical potential across the entire set-up is μ0\mu_{0}, and a voltage drop VV across the external load distributed symmetrically between the left and right system-to-reservoir interfaces, the quasi-Fermi levels at the reservoirs LL, RR and GG may be written as μL=μ0V2,μR=μ0+V2\mu_{L}=\mu_{0}-\frac{V}{2},~{}\mu_{R}=\mu_{0}+\frac{V}{2} and μG=μ0\mu_{G}=\mu_{0} respectively. The generated power (PP) and efficiency (η\eta) can be defined as:

P\displaystyle P =IL(R)×V,\displaystyle=I_{L(R)}\times V,
η\displaystyle\eta =PIQ,\displaystyle=\frac{P}{I_{Q}}, (7)

where VV is the applied potential bias in the voltage controlled model and IQI_{Q} is the sum of electronic and lattice heat flux from the heat source (GG). In the non-local heat engine the lattice heat flux can be favourably engineered Androulakis et al. [2006]; Hsu et al. [2004]; Pan et al. [2015]; Feser et al. [2012]; Davis and Hussein [2014] without affecting the current conduction path. In addition the lattice heat flux is generally independent of the the electronic heat conductivity and doesn’t change with the system energy configuration Mingo and Broido [2004]; Mingo [2004]; Zhou et al. [2007, 2010]; Boukai et al. [2008]; Hochbaum et al. [2008]; Balandin et al. [1999]; Chen [1998]. Hence, to simplify the calculation, I assume ideal condition by neglecting the lattice heat flux, as done in recent literature Walldorf et al. [2017]; Daré and Lombardo [2017]; Zhang et al. [2016]; Sánchez and Büttiker [2011]; Singha et al. [2015]; Singha and Muralidharan [2017]; Whitney [2014, 2015]. The generation efficiency for our case, can hence be defined as:

η=PIQe,\eta=\frac{P}{I_{Q_{e}}}, (8)

where IQeI_{Qe} is the electronic component of the heat current extracted from the heat source GG (defined in Eq. 6). It should be noted that the efficiency given by Eq. (8) is the maximum achievable efficiency under ideal conditions and is limited by thermodynamic considerations of heat to work conversion for the given system energy configuration. \colorblack
To understand the operation of the proposed heat engine, let us assume that the system is initially in the vacuum state |0,0,0\ket{0,0,0}, where the ground state of all the dots are empty. Now, let us consider a complete cycle F|0,0,0|1,0,0|1,1,0|0,1,1|0,1,0|0,0,0F\Rightarrow\ket{0,0,0}\rightarrow\ket{1,0,0}\rightarrow\ket{1,1,0}\rightarrow\ket{0,1,1}\rightarrow\ket{0,1,0}\rightarrow\ket{0,0,0}. In this cycle, the system starts from the initial vacuum state |0,0,0\ket{0,0,0}. Next an electron enters into S1S_{1} from LL, with an energy εs1\varepsilon_{s}^{1}, followed by an electron tunneling into G1G_{1} from GG with an energy εg+Um\varepsilon_{g}+U_{m}. Next, the electron in S1S_{1} tunnels into S2S_{2} with an energy εs2=εs1+Um\varepsilon_{s}^{2}=\varepsilon_{s}^{1}+U_{m} and finally tunnels out of S2S_{2} into reservoir RR. The system, at last returns to its initial state when the electron in G1G_{1}, finally tunnels out into GG with an energy εg\varepsilon_{g}. In this cycle, an electron is transferred from reservoir LL to RR while absorbing a heat packet UmU_{m} from GG. In the reverse cycle, corresponding to the cycle FF described above, that is R|0,0,0|0,1,0|0,1,1|1,1,0|1,0,0|0,0,0R\Rightarrow\ket{0,0,0}\rightarrow\ket{0,1,0}\rightarrow\ket{0,1,1}\rightarrow\ket{1,1,0}\rightarrow\ket{1,0,0}\rightarrow\ket{0,0,0}, an electron is transferred from reservoir RR to LL, while dumping a heat packet UmU_{m} into reservoir GG. It can be shown that when TG>TL(R)T_{G}>T_{L(R)}, under short-circuited condition between LL and RR, the rate of the forward cycle FF is greater than the reverse cycle RR (See Appendix D), resulting in a net unidirectional flow of electrons from LL to RR while absorbing heat from reservoir GG. Also for all other equivalent cycles, resulting in transfer of electrons between LL and RR under short-circuited condition, it can be shown that a rate of forward sequence that results in electron transfer from LL to RR while absorbing heat from GG is greater than the reverse sequence when TG>TL(R)T_{G}>T_{L(R)}. Thus for TG>TL(R)T_{G}>T_{L(R)}, a net thermoelectric electronic flow occurs from LL to RR by extracting heat from the spatially separated reservoir GG and hence the set-up proposed in this paper constitutes a non-local heat engine. Without loss of generality, I assume that γc=105qh\gamma_{c}=10^{-5}\frac{q}{h} and γ=104qh\gamma=10^{-4}\frac{q}{h} (The effect of the ratio γ/γc\gamma/\gamma_{c} on the heat harvesting performance of the set-up is discussed in Appendix E). Such values of γ\gamma and γc\gamma_{c} limits electronic transport phenomena in the weak coupling regime, where the effects of cotunneling can be neglected and the transport through the entire system can be described in terms of the quantum master equation (QME) approach, given by the set of Eqns. (2)-(3). This simplifies the analysis to a great extent and facilitates an intuitive understanding of the computed results. \colorblack The temperature of the reservoirs L(R)L(R) and GG are assumed to be TL(R)=5KT_{L(R)}=5K and TG=10KT_{G}=10K. The average temperature between the hot and the cold reservoirs is, hence, given by T=TL(R)+TG2=7.5KT=\frac{T_{L(R)}+T_{G}}{2}=7.5K.

Refer to caption
Figure 3: Variation in the peak performance of the heat engine with variation in Coulomb coupling energy UmU_{m} and applied bias VV. Colour plot depicting the (a) peak generated power PMP_{M} and (b) efficiency at the peak generated power for a range of values of VV and UmU_{m}. To find out the maximum power PMP_{M} for a given value of VV and UmU_{m}, the ground states of the dots are tuned to optimal position. T=TL(R)+TG2=7.5KT=\frac{T_{L(R)}+T_{G}}{2}=7.5K is the average temperature between the heat source and the heat sink. The efficiency of maximum power (PM)P_{M}) generation is normalized with respect to the Carnot efficiency ηc=1TL(R)/TG\eta_{c}=1-T_{L(R)}/T_{G}

III Results

Refer to caption
Figure 4: Performance comparison of the proposed non-local heat engine (solid lines) with the optimal set-up Walldorf et al. [2017]; Daré [2019]; Zhang and Chen [2019]; Daré and Lombardo [2017] (dashed lines) for different values of the Coulomb coupling energy. Plot of (a) maximum generated power PMP_{M} vs bias voltage VV, (b) efficiency (with respect to Carnot efficiency) at the maximum generated power vs bias voltage, (c) maximum power PMP_{M} vs efficiency at the maximum power for the optimal set-up (d) maximum power PMP_{M} vs efficiency at the maximum power for the proposed design.

In this section, I discuss the optimal operation regimes of the proposed heat engine. In addition, I conduct a performance comparison of the proposed heat engine with the optimal set-up discussed in literature and investigate the transport processes leading to a performance deterioration of the proposed set-up. Fig. 2 demonstrates the performance of the heat engine, in particular the generated power PP and efficiency of generated power (η/ηc\eta/\eta_{c}) over a range of the positions of the ground states εg\varepsilon_{g} and εs1\varepsilon_{s}^{1} for Um=3.9meV(6kTq)U_{m}=3.9meV~{}(\approx 6\frac{kT}{q}) and V=1.3meV(2kTq)V=1.3meV~{}(\approx 2\frac{kT}{q}). It can be noted that the regime of heat harvesting corresponds to εs1\varepsilon_{s}^{1} lying a few kTL(R)kT_{L(R)} below μ0\mu_{0}. Such a behaviour can be expected since net interdot electron flow is optimized when εs1\varepsilon_{s}^{1} lies a few kTL(R)kT_{L(R)} below the equilibrium Fermi-energy. This is because for TG>TL(R)T_{G}>T_{L(R)}, electrons must tunnel into S1S_{1} with energy εs1\varepsilon_{s}^{1} and out of S2S_{2} with energy εs2=εs1+Um\varepsilon_{s}^{2}=\varepsilon_{s}^{1}+U_{m} for finite power generation, which demands finite value of fL(εs1)f_{L}(\varepsilon_{s}^{1}) and 1fR(εs2)1-f_{R}(\varepsilon_{s}^{2}). Both fL(εs1)f_{L}(\varepsilon_{s}^{1}) and 1fR(εs2)1-f_{R}(\varepsilon_{s}^{2}) are finite and large only when εs1\varepsilon_{s}^{1} lies a few kTL(R)kT_{L(R)} below μ0\mu_{0}. In fact the product fL(εs1){1fR(εs2)}f_{L}(\varepsilon_{s}^{1})\{1-f_{R}(\varepsilon_{s}^{2})\} is maximized when εs1μ0Um2\varepsilon_{s}^{1}\approx\mu_{0}-\frac{U_{m}}{2} (for small bias VV). Since, Um6kTq=8kTL(R)2U_{m}\approx 6\frac{kT}{q}=8\frac{kT_{L(R)}}{2}, the generated power is maximized when εs1μ0Um2=μ04kTL(R)q\varepsilon_{s}^{1}\approx\mu_{0}-\frac{U_{m}}{2}=\mu_{0}-4\frac{kT_{L(R)}}{q}, as demonstrated in Fig. 2. \colorblack Similarly, it can also be noted that for optimal heat harvesting, the ground state εg\varepsilon_{g} must lie below a few kTGkT_{G} of the equilibrium Fermi energy μ0\mu_{0}. This can be understood by the following: if εgμ0<fewkTG\varepsilon_{g}-\mu_{0}<-few~{}kT_{G}, then the ground state εg\varepsilon_{g} is always occupied with an electron and so the asymmetry of the system ground states with respect to the reservoir LL and RR disappears. Hence a directional thermoelectric current flow is not possible Sánchez and Büttiker [2011] by rectifying the stochastic fluctuations of the heat source. On the other hand, when εgμ0>fewkTG\varepsilon_{g}-\mu_{0}>few~{}kT_{G}, the probability of an electron tunneling into the reservoir G1G_{1} with an energy εg+Um\varepsilon_{g}+U_{m} (provided that the ground state of S1S_{1} is occupied) is negligibly small, resulting in the deterioration of the unidirectional current flow and power generation. In other words, for the generation of finite thermoelectric power with TG>TL(R)T_{G}>T_{L(R)}, electrons must tunnel in and out of G1G_{1} with energy εg+Um\varepsilon_{g}+U_{m} and εg\varepsilon_{g} respectively for the absorption of heat energy from GG, which calls for a finite value of both fG(εg+Um)f_{G}(\varepsilon_{g}+U_{m}) and 1fG(εg)1-f_{G}(\varepsilon_{g}). Both these quantities are finite and large only if εg\varepsilon_{g} lies below a few kTGkT_{G} of the equilibrium Fermi energy μ0\mu_{0}. In fact, the product fG(εg+Um){1fG(εg)}f_{G}(\varepsilon_{g}+U_{m})\{1-f_{G}(\varepsilon_{g})\}, and hence the generated power, is maximized when εgμ0Um2\varepsilon_{g}\approx\mu_{0}-\frac{U_{m}}{2}. Since, in this case, Um=6kTq=4kTGqU_{m}=6\frac{kT}{q}=4\frac{kT_{G}}{q}, the generated power is maximized, when εgμ02kTq\varepsilon_{g}\approx\mu_{0}-2\frac{kT}{q}, as depicted in Fig. 2(a). \colorblack Fig. 2(b) demonstrates the heat engine efficiency as a function of the ground state energy levels. The efficiency of heat harvesting decreases monotonically with decrease in εgμ0\varepsilon_{g}-\mu_{0}. This is because, as εg\varepsilon_{g} decreases, the probability occupancy of the ground state of G1G_{1} increases, which increases the probability of reverse electron flux due to the bias voltage. An increase in the reverse electron flux decreases the net thermoelectric current (and hence power) and thus degrades the efficiency. It should be noted that an equivalent trend of decrease in efficiency can be found in bulk and lower dimensional thermoelectric engines as the equilibrium Fermi-energy gradually moves inside the energy-bands Singha et al. [2015]; Singha and Muralidharan [2017]; Agarwal and Muralidharan [2014]; Whitney [2014, 2015]. The variation in generation efficiency with εs1\varepsilon_{s}^{1} is non-monotonic. Initially as εs1\varepsilon_{s}^{1} decreases and moves away the Fermi energy (or electron transport window), the generated power increases (as discussed above) leading to an increase in efficiency. As εs1\varepsilon_{s}^{1} further decreases, such that εs1+Um\varepsilon_{s}^{1}+U_{m} enters the electron transport window, the probability of reverse electronic flow due to voltage bias increases, leading to a decrease in the total net thermoelectric current (discussed later). A decrease in the net thermoelectric current leads to a deterioration in the generated power and hence, efficiency.
\colorblack The variation of the optimal performance of the heat engine with variation in the Coulomb coupling energy UmU_{m} and applied bias VV is demonstrated in Fig. 3. In particular, Fig. 3 (a) demonstrates the maximum generated power (PMP_{M}), while Fig. 3 (b) demonstrates the efficiency at the maximum generated power for a range of values of the applied bias VV and the Coulomb coupling energy UmU_{m}. To calculate the the maximum generated power PMP_{M} for a given value of VV and UmU_{m}, the ground states of the dots are tuned to the optimal energy position. It should be noted that the maximum generated power is low for low values of UmU_{m}. This is due to the fact that the ground state of the dots approach symmetrical arrangement with respect to the reservoir LL and RR as UmU_{m} approaches towards zero. Hence, the directional flow of electrons decreases. As UmU_{m} increases, the asymmetry of the system increases resulting in an increase in directional electron flow, and hence, the maximum generated power Sánchez and Büttiker [2011]. With further increase in UmU_{m}, the maximum generated power reaches its peak and then decreases due to lower probability of an electron tunneling into G1G_{1} with an energy εg+Um\varepsilon_{g}+U_{m}, when the ground state of S1S_{1} is already occupied. Mathematically, the non-monotonic change in maximum generated power with UmU_{m} can be explained in terms of the Eqns. (21)-(23) developed in Appendix C (Refer to Appendix C for details). \colorblack For a fixed value of UmU_{m}, the maximum generated power first increases and then decreases with an increase in the bias voltage VV. Such a behaviour is indeed expected from heat engines as the regime of operation approaches from short-circuited (zero voltage drop across the external load) condition to the open-circuited condition (zero net current across the external load) Nakpathomkun et al. [2010]; Whitney [2014, 2015]. Fig. 3(b) demonstrates the efficiency at the maximum generated power with variation in applied bias VV and Coulomb coupling energy UmU_{m}. The efficiency varies non-monotonically with the applied bias VV, that is,the efficiency increases with an increase in VV as the regime of operation approaches the point of maximum power and then gradually decreases as the generated power decreases with the regime of operation approaching the open circuited condition. Such a trend can also be noted in bulk and lower dimensional heat engines Singha et al. [2015]; Singha and Muralidharan [2017]. On the other hand with an increase in UmU_{m}, the efficiency at the maximum power increases monotonically as εg+Um\varepsilon_{g}+U_{m} gradually surpasses the Fermi energy. An equivalent trend, again, can be noted in bulk and lower dimensional thermoelectric engines as the band-edge gradually surpasses the Fermi-energy Singha et al. [2015]; Leijnse et al. [2010]; Sothmann et al. [2014]; Nakpathomkun et al. [2010]; Choi and Jordan [2015]. Fig. 4 demonstrates a performance comparison of the proposed heat engine with the optimal non-local heat engine put forward in literature Walldorf et al. [2017]; Daré [2019]; Zhang and Chen [2019]; Daré and Lombardo [2017]. The optimal non-local heat engine, discussed in literature, consists of a pair of Coulomb coupled quantum dots (demonstrated in Appendix Fig. 8.a) with asymmetric system-to-reservoir coupling Walldorf et al. [2017]; Daré [2019]; Zhang and Chen [2019]; Daré and Lombardo [2017] given by γl(ε)=γcθ(εs1+δϵε),γl(ε)=γcθ(εδϵεs1)\gamma_{l}(\varepsilon)=\gamma_{c}\theta(\varepsilon_{s}^{1}+\delta\epsilon-\varepsilon),~{}\gamma_{l}(\varepsilon)=\gamma_{c}\theta(\varepsilon-\delta\epsilon-\varepsilon_{s}^{1}) and γg(ε)=γc\gamma_{g}(\varepsilon)=\gamma_{c}, where δε<Um\delta\varepsilon<U_{m} and θ\theta is Heaviside step function. \colorblack In particular, Fig. 4(a) and (b) demonstrate the variation in the maximum power PMP_{M} and efficiency at the maximum power respectively with applied load bias VV for different values of UmU_{m} for both the proposed set-up (solid lines) and the optimal set-up (dashed lines). We note that the overall maximum generated power for the proposed design PMAXpropP_{MAX}^{prop} is approximately 3.23fW3.23fW, which is about 50%50\% of the overall maximum power output of 6.04fW6.04fW for the optimal design PMAXoptP_{MAX}^{opt}. In both these cases the maximum power is generated around Um4meV(6kTq)U_{m}\approx 4meV(\approx 6\frac{kT}{q}). As already discussed in literature Sánchez and Büttiker [2011], the generation efficiency for the optimal set-up increases linearly with the applied bias for a given value of UmU_{m}. The efficiency at the overall maximum power for our proposed design and the optimal set-up are 24.5%24.5\% and 60%60\% of the Carnot efficiency respectively. From Fig. 4(a) and (b), we also note that the open-circuit voltage (finite voltage at which out-put power just becomes zero) for both the proposed design and the optimal set-up increases with an increase in UmU_{m}. In addition, it can also be noted that the open circuit voltage for the optimal set-up is slightly higher compared to the proposed design. Fig. 4 (c) and (d) demonstrate the the maximum power vs efficiency trade-off trend Singha et al. [2015] for the optimal design and the proposed design respectively for various values of UmU_{m}. We note that the power-efficiency trade-off for the optimal set-up is mild compared to the proposed design.

Refer to caption
Figure 5: Schematic diagram demonstrating the components of electron flow between the reservoir LL and the system through the energy level εs1\varepsilon_{s}^{1} and the Coulomb blockaded level εs1+Um\varepsilon_{s}^{1}+U_{m}. Three current components are shown in the figure. (1) Electrons flow from the reservoir LL to RR while absorbing a heat packet UmU_{m} (per electron) from GG. (2) Electrons enter into εs1\varepsilon_{s}^{1} from LL. Next these electrons tunnel out of the system into LL through εs1+Um\varepsilon_{s}^{1}+U_{m}. (3) These electrons flow in the direction of the voltage bias (opposite to the thermoelectric current) and thus reduce the net current through the system.
Refer to caption
Figure 6: Colour plots demonstrating the electron flow into the system from the reservoir LL with variation in the the ground states εg\varepsilon_{g} and εs1\varepsilon_{s}^{1}, for Um=3.9meV(6kTq)U_{m}=3.9meV~{}(\approx 6\frac{kT}{q}) and V=1.3meV(2kTq)V=1.3meV~{}(\approx 2\frac{kT}{q}), when the (a) ground state of the dot G1G_{1} is not occupied (b) ground state of the dot is occupied (c) total average current between the system and the reservoir LL. Interestingly, when the ground state of the dot G1G_{1} is occupied, the net electronic flow is directed from the system into the reservoir LL. Since, the net direction of current for thermoelectric generation should be from reservoir LL to RR, the effect of net electronic flow from the system into the reservoir LL is to reduce the generated power and efficiency.

I end the discussion with a brief description of the processes leading to a deterioration in generated power and efficiency for the proposed set-up. First, let us consider the cycles leading to electron transport from the reservoir LL to the reservoir RR against the applied bias while absorbing a heat packet UmU_{m} from reservoir GG. Consider the cycle |0,0,0|1,0,0)|1,1,0|0,1,1|0,1,0|0,0,0\ket{0,0,0}\rightarrow\ket{1,0,0})\rightarrow\ket{1,1,0}\rightarrow\ket{0,1,1}\rightarrow\ket{0,1,0}\rightarrow\ket{0,0,0}. In this cycle, the system starts with an initial vacuum state |0,0,0\ket{0,0,0}. An electron tunnels from LL into S1S_{1} at energy εs1\varepsilon_{s}^{1}, followed by an electron tunneling into G1G_{1} from GG at energy εg+Um\varepsilon_{g}+U_{m}. At the next instant, the electron in S1S_{1} tunnels into S2S_{2}, after which the electron in G1G_{1} tunnels out into GG with energy εg\varepsilon_{g}. The cycle is completed and the system returns to the initial state when the electron in S2S_{2} tunnels out into the reservoir RR with an energy εs2=εs1+Um\varepsilon_{s}^{2}=\varepsilon_{s}^{1}+U_{m}. It is clear that in this process an electron is transferred from LL to RR while absorbing a heat packet UmU_{m} from GG. Another cycle that again transfers electrons from LL to RR, while absorbing heat packet from GG can be given by |0,0,0|1,0,0)|1,1,0|0,1,1|0,0,1|0,0,0\ket{0,0,0}\rightarrow\ket{1,0,0})\rightarrow\ket{1,1,0}\rightarrow\ket{0,1,1}\rightarrow\ket{0,0,1}\rightarrow\ket{0,0,0}. These transport processes contribute to thermoelectric power generation while absorbing heat energy from GG and are demonstrated as (1)(1) in Fig. 5. \colorblack Next, consider the cycle |0,0,0|1,0,0|1,1,0|0,1,0|0,0,0\ket{0,0,0}\rightarrow\ket{1,0,0}\rightarrow\ket{1,1,0}\rightarrow\ket{0,1,0}\rightarrow\ket{0,0,0}. This cycle consists of an electron tunneling into S1S_{1} from LL, with an energy εs1\varepsilon_{s}^{1}, followed by an electron tunneling into G1G_{1} with an energy εg+Um\varepsilon_{g}+U_{m}. At the next step, the electron in S1S_{1} exits into reservoir LL with an energy εs1+Um\varepsilon_{s}^{1}+U_{m}. The cycle is completed with the electron in G1G_{1} tunnels out into GG with energy εg\varepsilon_{g}. It is evident that in this process, a packet of heat energy UmU_{m} is transmitted from reservoir GG to LL without any net flow of electrons between LL and RR. So, effectively the heat packet UmU_{m} is wasted without any power conversion. This electron flow component, depicted in Fig. 5 as (2)(2), doesn’t contribute to power generation, but transmits heat packets from GG and hence results in degradation of the efficiency. The third electron flow component, depicted in Fig. 5 as (3), results from a reverse flux of electrons caused by the voltage bias (voltage drop across the load impedance) and flows in the direction opposite to the thermoelectric current. Hence, this component results in the degradation of the net thermoelectric current and thus deteriorates the generated power as well as the efficiency. \colorblack Once the ground state of both S1S_{1} and G1G_{1} are occupied, the electron existing in S1S_{1} can either tunnel into S2S_{2} and subsequently into RR giving rise to directional electronic flow (component 1) or tunnel out to LL without any net electron flow (component 2). Hence, the overall maximum power output of the proposed set-up hovers around 50%50\% of the optimal design. The generation efficiency at the overall maximum power is lower than 50%50\% of the optimal design due to component (3) of the total electron current. It should be noted that component (3) of the current only flows when the ground state of dot G1G_{1} is already occupied with an electron and that of dot S1S_{1} is empty.
To further establish the points discussed above, in Fig. 6, I separate out the current flow into the system from the reservoir LL as:

IL=IL1+IL2,I_{L}=I_{L1}+I_{L2},

where

IL1=qγc×{P0,0ς1fL(εs1)P1,0ς1(1fL(εs1))}\displaystyle I_{L1}=q\gamma_{c}\times\left\{P^{\varsigma_{1}}_{0,0}f_{L}(\varepsilon_{s}^{1})-P^{\varsigma_{1}}_{1,0}(1-f_{L}(\varepsilon_{s}^{1})\right)\}
IL2=qγc{P0,1ς1fL(εs1+Um)P1,1ς1{1fL(εs1+Um)}}\displaystyle I_{L2}=q\gamma_{c}\{P^{\varsigma_{1}}_{0,1}f_{L}(\varepsilon_{s}^{1}+U_{m})-P^{\varsigma^{1}}_{1,1}\{1-f_{L}(\varepsilon_{s}^{1}+U_{m})\}\} (9)

In Eq. (9), IL1I_{L1} and IL2I_{L2} denote the total electronic current from reservoir LL into the energy level εs1\varepsilon_{s}^{1} and the Coulomb blockaded level εs1+Um\varepsilon_{s}^{1}+U_{m} respectively. Fig. 6(a) and (b) demonstrate the electron current flow into the system from reservoir LL through the energy level εs1\varepsilon_{s}^{1} (IL1I_{L1}) and the Coulomb blockaded level εs1+Um\varepsilon_{s}^{1}+U_{m} (IL2I_{L2}) respectively, while Fig. 6(c) demonstrates the total electronic current IL=IL1+IL2I_{L}=I_{L1}+I_{L2} from the reservoir LL into the system. It should be noted that the electronic current demonstrated in Fig. 6, is opposite to the direction of conventional current flow (since electron charge is negative). We find that the electron current flow IL1I_{L1} through εs1\varepsilon_{s}^{1}, into the system from LL, is positive or against the voltage bias, generating a net value of thermoelectric power. Interestingly, we also find that the current component IL2I_{L2} through the Coulomb blockaded energy level εs1+Um\varepsilon_{s}^{1}+U_{m} is negative, as already shown in Fig. 5. That is, the net electron current, through the Coulomb blockaded level εs1+Um\varepsilon_{s}^{1}+U_{m} flows into the reservoir LL from the system in the direction of voltage bias. This is already illustrated in Fig. 5, where it is shown that the components (2)(2) and (3)(3) flow out of the system into LL via the Coulomb blocked level εs1+Um\varepsilon_{s}^{1}+U_{m}. As discussed above, these components (2)(2) and (3)(3) of electron current flow may transmit heat packets without any power generation, or may reduce the net forward current flow against the voltage bias. Thus, they impact negatively on the generated power, as well as the efficiency. The deterioration in the heat engine performance, due to these current components (2)(2) and (3)(3) (discussed above), can be alleviated by adding an extra filter between LL and S1S_{1}, to restrict the current flow via the Coulomb blocked level εs1+Um\varepsilon_{s}^{1}+U_{m}, thereby nullifying the current components (2)(2) and (3)(3). However, doing so neutralizes the novelty of the proposed set-up in terms of fabrication simplicity. It should be noted that in Fig. 6(c), negative values of total electronic current corresponds to a net electronic flow in the direction of the applied bias and thus the zone with negative electronic current flow indicates the regime of zero net thermoelectric power generation.

IV Conclusion

To conclude, in this paper I have proposed a realistic design strategy for non-local heat engine based on Coulomb coupled systems. The performance of the proposed design was then theoretically analyzed and compared with the optimal set-up Sánchez and Büttiker [2011] using the QME approach. It was demonstrated that the proposed set-up outputs a maximum power of around 50%50\% of the optimal set-up. However, the crucial advantage of the proposed design strategy is that along with a reasonable output power, it also circumvents the demand for a sharp step-like change in reservoir-to-system coupling, which is required for proper operation of the optimal set-up proposed in literatureSánchez and Büttiker [2011]. In the above discussions, I have limited transport through the quantum dots in the weak coupling regime, so that the effects of co-tunneling can be neglected. The generated power in the proposed system can be increased by a few orders of magnitude by tuning electronic transport in the strong coupling regime, that is by increasing the system-to-reservoir, as well as, the interdot tunnel coupling. It will be interesting to study the effects of cotunneling on heat-harvesting performance of the proposed system as electronic transport is gradually tuned towards the strong coupling regime. \colorblack In addition, I have also assumed ideal conditions by neglecting the lattice thermal conductance. Understanding the impact of lattice heat flux on the generation efficiency and an investigation on the effect electron-phonon interaction on the performance of the proposed design also constitute interesting research directions. Although not shown here, the proposed system can also work as an efficient non-local heat engine when the reservoir GG acts as a heat sink (cold) with respect to the reservoirs LL and RR (hot). In such a case, the direction of thermoelectric current flow is reversed. The different possible design strategies for non-local heat engines and their performance is left for future exploration. Nevertheless, the set-up proposed in this paper can be employed to fabricate high performance non-local heat engines using Coulomb coupled systems.
Acknowledgments: The author would like to thank Sponsored Research and Industrial Consultancy (IIT Kharagpur) for their financial support via grant no. IIT/SRIC/EC/MWT/2019-20/162.
Data Availability: The data that supports the findings of this study are available within the article

Appendix A Another Coulomb coupled system based set-up for efficient non-local heat harvesting

In this section, I show another possible set-up for non-local heat engine using Coulomb coupled systems that circumvents the need of a sharp transition in system-to-reservoir coupling around the ground states of the quantum dots. The system, demonstrated in Fig. 7, consists of three quantum-dots S1,S2S_{1},~{}S_{2} and G1G_{1}, which are coupled to the reservoirs L,RL,~{}R and GG respectively. Identical to the set-up in Fig. 1, S1S_{1} and S2S_{2} are tunnel coupled to each other, while G1G_{1} is Coulomb coupled to S1S_{1} with electrostatic charging energy UmU_{m}. However, unlike the proposed set-up in Fig. 1, the ground-states of the two quantum dots S1S_{1} and S2S_{2} are identical, that is εs1=εs2\varepsilon_{s}^{1}=\varepsilon_{s}^{2}. The asymmetry of the system ground state configuration, required for non local heat harvesting, creeps in when an electron tunnels from GG into G1G_{1}, such that an electron tunneling into (or out of) S1S_{1} must now have an energy εs1+Um\varepsilon_{s}^{1}+U_{m} when the ground state of G1G_{1} is occupied. Although not detailed here, the system depicted in Fig. 7 demonstrates similar heat-harvesting performance to the set-up shown in Fig. 1 with slightly shifted regime of operation. In this case, for TG>TL(R)T_{G}>T_{L(R)}, the short-circuited electron flux, due to thermoelectric force, flows from reservoir RR to LL, that is in the direction opposite to that of the set-up in Fig. 1. The set-up, demonstrated in Fig. 7, thus can also be employed for efficient non-local heat harvesting.

Refer to caption
Figure 7: Schematic diagram for another realistic design strategy to accomplish non-local heat harvesting using Coulomb coupled systems. The heat harvesting performance of this system was found to be similar to the proposed set-up in Fig. 1 with slightly shifted regime of operation.

Appendix B Direction of heat flow in the Coulomb coupled systems.

In this section, I discuss the direction of electronic heat flow in a simple coulomb coupled system with constant energy-resolved system-to-reservoir coupling, for TG>TL(R)T_{G}>T_{L(R)}. Such a system is depicted in Fig. 8. Let us consider a sequential cycle that transfers a heat packet UmU_{m} from the reservoir GG to the reservoir L(R)L(R). Consider the cycle, demonstrated schematically in Fig. 8(b). With nxn_{x} denoting the total number of electrons in the ground state of dot xx at any given instant of time, I denote each state of the complete system as |nS1,nG1\ket{n_{S_{1}},n_{G_{1}}}, where nS1,nG1(0,1)n_{S_{1}},n_{G_{1}}\in(0,1). Consider the sequence of sequential electron transport ( Fig. 8b) given by F1|0,0|1,0)|1,1|0,1|0,0F_{1}\Rightarrow\ket{0,0}\rightarrow\ket{1,0)}\rightarrow\ket{1,1}\rightarrow\ket{0,1}\rightarrow\ket{0,0}. In this cycle, the system starts with the initial vacuum state |0,0\ket{0,0}. An electron then tunnels from L(R)L(R) into S1S_{1} at energy εs1\varepsilon_{s}^{1}, followed by an electron tunneling from GG into G1G_{1} with an energy εg+Um\varepsilon_{g}+U_{m}. Next, the electron in S1S_{1} tunnels out into L(R)L(R) with an energy εs1+Um\varepsilon_{s}^{1}+U_{m}. The system returns to its ground state when the electron in G1G_{1} tunnels out to GG with an energy εg\varepsilon_{g}. In this cycle, a heat packet is transferred from GG to L(R)L(R). The reverse cycle given by R1|0,0|0,1|1,1|1,0|0,0R_{1}\Rightarrow\ket{0,0}\rightarrow\ket{0,1}\rightarrow\ket{1,1}\rightarrow\ket{1,0}\rightarrow\ket{0,0}, on the other hand, transfers a heat energy UmU_{m} from the reservoir L(R)L(R) to GG. In what follows, I show that that when TG>TL(R)T_{G}>T_{L(R)}, the probability of forward cycle F1F_{1} (given in Fig. 8.b) is higher than than the probability of occurance of the reverse cycle, resulting in a net flow of heat from reservoir GG to RR. If the energy resolved coupling of the reservoirs LL or RR to the dot S1S_{1} is constant, then the electron in S1S_{1} can tunnel in(out) from LL or RR with equal probability, thereby producing zero net current. Let us assume that the system, with schematic shown in Fig. 8(a), is initially in the vacuum state at time t=0t=0 with no electrons in the levels εg\varepsilon_{g} and εs1\varepsilon_{s}^{1}. For simplicity, I also assume that there is no voltage drop between the reservoirs LL and RR. Let us assume that the probability an electron enters the dot S1S_{1} from the left contact between time tt and t+dtt+dt is given by Pin(t)dtP_{in}(t)dt. We can write Pin(t)dtP_{in}(t)dt as:

Refer to caption
Figure 8: (a) Schematic of a Coulomb coupled system consisting of a pair of quantum dots S1S_{1} and G1G_{1}. The dot S1S_{1} is electrically connected to the reservoirs LL and RR, while G1G_{1} is electrically connected to the reservoir GG. G1G_{1} and S1S_{1} are Coulomb coupled with a mutual charging energy UmU_{m}. For constant energy resolved coupling γl(ε)=γr(ε)=γg(ε)=γc\gamma_{l}(\varepsilon)=\gamma_{r}(\varepsilon)=\gamma_{g}(\varepsilon)=\gamma_{c}, the system produces no net thermoelectric current between LL and RR. However, the stochastic thermal fluctuation of GG can be converted into a directed thermoelectric current flow for asymmetric system-to-reservoir coupling given by γl(ε)=γcθ(εs1+δϵε),γr(ε)=γcθ(εδϵεs1)\gamma_{l}(\varepsilon)=\gamma_{c}\theta(\varepsilon_{s}^{1}+\delta\epsilon-\varepsilon),~{}\gamma_{r}(\varepsilon)=\gamma_{c}\theta(\varepsilon-\delta\epsilon-\varepsilon_{s}^{1}) and γg(ε)=γc\gamma_{g}(\varepsilon)=\gamma_{c}, for TGTL(R)T_{G}\neq T_{L(R)} (δϵ<Um(\delta\epsilon<{U_{m}}) Walldorf et al. [2017]; Daré [2019]; Zhang and Chen [2019]; Daré and Lombardo [2017]; Zhang et al. [2016]; Sánchez and Büttiker [2011]. (b) A sequential electron transport cycle that is responsible for heat packet flow from the reservoir GG to the reservoir L(R)L(R). (c) Schematic of the realistic set-up proposed in this paper for efficient non-local heat harvesting. In this case, the system ground state configuration is asymmetric with respect to the reservoirs LL and RR, with εs2=εs1+Δε\varepsilon_{s}^{2}=\varepsilon_{s}^{1}+\Delta\varepsilon. In such a system, a non-local thermoelectric current can be driven between the reservoirs LL and RR, by extracting heat from the reservoir GG, even for the case of constant energy resolved system-to-reservoir coupling given by γl(ε)=γr(ε)=γg(ε)=γc\gamma_{l}(\varepsilon)=\gamma_{r}(\varepsilon)=\gamma_{g}(\varepsilon)=\gamma_{c}. (d) Sequential transport processes leading to transfer of electrons from LL to RR while absorbing heat from the reservoir GG.
Pin(t)dt=γcdt×fL(εs1)×{10tPin(τ)𝑑τ}P_{in}(t)dt=\gamma_{c}dt\times f_{L}(\varepsilon_{s}^{1})\times\left\{1-\int_{0}^{t}P_{in}(\tau)d\tau\right\} (10)

Differentiating the above equation with respect to tt, we get

dPin(t)dt=γc×fL(εs1)×Pin(t)\frac{dP_{in}(t)}{dt}=-\gamma_{c}\times f_{L}(\varepsilon_{s}^{1})\times P_{in}(t) (11)

Subject to the condition 0Pin(t)𝑑t=1\int_{0}^{\infty}P_{in}(t)dt=1, the above equation has the solution

Pin(t)=γCfL(εs1)eγCfL(εs1)tP_{in}(t)={\gamma_{C}f_{L}(\varepsilon_{s}^{1})}e^{-\gamma_{C}f_{L}(\varepsilon_{s}^{1})t} (12)

Similarly, it can shown that an electron present in the energy level εs1\varepsilon_{s}^{1} at t=0t=0 can exit to the left contact with a probability given by

Pout(t)=γC{1fL(εs1)}eγC{1fL(εs1)}t.P_{out}(t)={\gamma_{C}\{1-f_{L}(\varepsilon_{s}^{1})}\}e^{-\gamma_{C}\{1-f_{L}(\varepsilon_{s}^{1})\}t}. (13)

With the help of equations (12) and (13), the probability that the forward cycle in Fig. 8(b) (marked with black arrows) completes within a time-span Δt\Delta t can be written as:

ϱf=0ΔtγC{fL(εs1)+fR(εs1)}eγC{fL(εs1)+fR(εs1)}t1t1ΔtγCfG(εg+Um)eγCfG(εg+Um)t2\displaystyle\varrho_{f}=\int_{0}^{\Delta t}{\gamma_{C}\{f_{L}(\varepsilon_{s}^{1})+f_{R}(\varepsilon_{s}^{1})}\}e^{-\gamma_{C}\{f_{L}(\varepsilon_{s}^{1})+f_{R}(\varepsilon_{s}^{1})\}t_{1}}\int_{t_{1}}^{\Delta t}{\gamma_{C}f_{G}(\varepsilon_{g}+U_{m})}e^{-\gamma_{C}f_{G}(\varepsilon_{g}+U_{m})t_{2}}
×t2ΔtγC{2fL(εs1+Um)fR(εs1+Um)}eγC{2fL(εs1+Um)fR(εs1+Um)}t3\displaystyle\times\int_{t_{2}}^{\Delta t}{\gamma_{C}\{2-f_{L}(\varepsilon_{s}^{1}+U_{m})-f_{R}(\varepsilon_{s}^{1}+U_{m})}\}e^{-\gamma_{C}\{2-f_{L}(\varepsilon_{s}^{1}+U_{m})-f_{R}(\varepsilon_{s}^{1}+U_{m})\}t_{3}}
×t3ΔtγC{1fG(εg)}eγC{1fG(εg)}t4dt4dt3dt3dt1\displaystyle\times\int_{t_{3}}^{\Delta t}{\gamma_{C}\{1-f_{G}(\varepsilon_{g})}\}e^{-\gamma_{C}\{1-f_{G}(\varepsilon_{g})\}t_{4}}dt_{4}dt_{3}dt_{3}dt_{1} (14)
ϱf=Πf0ΔteγC{fL(εs1)+fR(εs1)}t1t1ΔteγCfG(εg+Um)t2t2ΔteγC{2fL(εs1+Um)fR(εs1+Um)}t3\displaystyle\Rightarrow\varrho_{f}=\Pi_{f}\int_{0}^{\Delta t}e^{-\gamma_{C}\{f_{L}(\varepsilon_{s}^{1})+f_{R}(\varepsilon_{s}^{1})\}t_{1}}\int_{t_{1}}^{\Delta t}e^{-\gamma_{C}f_{G}(\varepsilon_{g}+U_{m})t_{2}}\int_{t_{2}}^{\Delta t}e^{-\gamma_{C}\{2-f_{L}(\varepsilon_{s}^{1}+U_{m})-f_{R}(\varepsilon_{s}^{1}+U_{m})\}t_{3}}
×t3ΔteγC{1fG(εg)}t4dt4dt3dt3dt1,\displaystyle\times\int_{t_{3}}^{\Delta t}e^{-\gamma_{C}\{1-f_{G}(\varepsilon_{g})\}t_{4}}dt_{4}dt_{3}dt_{3}dt_{1}, (15)

where Πf=γC4{fL(εs1)+fR(εs1)}×fG(εg+Um)×{2fL(εs1+Um)fR(εs1+Um)}×{1fG(εg)}\Pi_{f}={\gamma_{C}^{4}\{f_{L}(\varepsilon_{s}^{1})+f_{R}(\varepsilon_{s}^{1})}\}\times{f_{G}(\varepsilon_{g}+U_{m})}\times{\{2-f_{L}(\varepsilon_{s}^{1}+U_{m})-f_{R}(\varepsilon_{s}^{1}+U_{m})}\}\times{\{1-f_{G}(\varepsilon_{g})}\}. To simplify the above equation, I work in the limit of very small Δt\Delta t such that the exponential quantities in the above equations is approximately 1. Hence,

ϱf\displaystyle\varrho_{f} =Πf0Δtt1Δtt2Δtt3Δt𝑑t4𝑑t3𝑑t3𝑑t1\displaystyle=\Pi_{f}\int_{0}^{\Delta t}\int_{t_{1}}^{\Delta t}\int_{t_{2}}^{\Delta t}\int_{t_{3}}^{\Delta t}dt_{4}dt_{3}dt_{3}dt_{1}
=ΠfΔt424\displaystyle=\Pi_{f}\frac{\Delta t^{4}}{24} (16)

Similarly, assuming that the system is in the vacuum state |0,0\ket{0,0} at t=0t=0, the probability that the reverse process is completed within a very small time interval Δt\Delta t is

ϱb\displaystyle\varrho_{b} =Πb0Δtt1Δtt2Δtt3Δt𝑑t4𝑑t3𝑑t3𝑑t1\displaystyle=\Pi_{b}\int_{0}^{\Delta t}\int_{t_{1}}^{\Delta t}\int_{t_{2}}^{\Delta t}\int_{t_{3}}^{\Delta t}dt_{4}dt_{3}dt_{3}dt_{1}
=ΠbΔt424,\displaystyle=\Pi_{b}\frac{\Delta t^{4}}{24}, (17)

where Πb=γC4{2fL(εs1)fR(εs1)}×{1fG(εg+Um)}×{fL(εs1+Um)+fR(εs1+Um)}×fG(εg)\Pi_{b}={\gamma_{C}^{4}\{2-f_{L}(\varepsilon_{s}^{1})-f_{R}(\varepsilon_{s}^{1})}\}\times\{1-f_{G}(\varepsilon_{g}+U_{m})\}\times{\{f_{L}(\varepsilon_{s}^{1}+U_{m})+f_{R}(\varepsilon_{s}^{1}+U_{m})}\}\times f_{G}(\varepsilon_{g}). So, the ratio of the probability of occurance of the forward process to the reverse process for a very small instant of time Δt\Delta t is

ϱf/ϱb\displaystyle\varrho_{f}/\varrho_{b} ={fL(εs1)+fR(εs1)}{2fL(εs1)fR(εs1)}×fG(εg+Um){1fG(εg+Um)}×{2fL(εs1+Um)fR(εs1+Um)}{fL(εs1+Um)+fR(εs1+Um)}×{1fG(εg)}fG(εg)\displaystyle=\frac{\{f_{L}(\varepsilon_{s}^{1})+f_{R}(\varepsilon_{s}^{1})\}}{\{2-f_{L}(\varepsilon_{s}^{1})-f_{R}(\varepsilon_{s}^{1})\}}\times\frac{{f_{G}(\varepsilon_{g}+U_{m})}}{\{1-f_{G}(\varepsilon_{g}+U_{m})\}}\times\frac{{\{2-f_{L}(\varepsilon_{s}^{1}+U_{m})-f_{R}(\varepsilon_{s}^{1}+U_{m})}\}}{{\{f_{L}(\varepsilon_{s}^{1}+U_{m})+f_{R}(\varepsilon_{s}^{1}+U_{m})}\}}\times\frac{{\{1-f_{G}(\varepsilon_{g})}\}}{f_{G}(\varepsilon_{g})}
=exp{εs1μkTL(R)}exp{εg+UmμgkTG}exp{εs1+UmμkTL(R)}exp{εgμgkTG}\displaystyle=exp\left\{-\frac{\varepsilon_{s}^{1}-\mu}{kT_{L(R)}}\right\}exp\left\{-\frac{\varepsilon_{g}+U_{m}-\mu_{g}}{kT_{G}}\right\}exp\left\{\frac{\varepsilon_{s}^{1}+U_{m}-\mu}{kT_{L(R)}}\right\}exp\left\{\frac{\varepsilon_{g}-\mu_{g}}{kT_{G}}\right\}
=exp{UmkTL(R)}exp{UmkTG}\displaystyle=exp\left\{\frac{U_{m}}{kT_{L(R)}}\right\}exp\left\{-\frac{U_{m}}{kT_{G}}\right\}
=exp{Umk(1TL(R)1TG)}\displaystyle=exp\left\{\frac{U_{m}}{k}\left(\frac{1}{T_{L(R)}}-\frac{1}{T_{G}}\right)\right\} (18)

From Eq. (18), it is clear that ϱf=ϱb\varrho_{f}=\varrho_{b} for TL(R)=TGT_{L(R)}=T_{G}. However, for TG>TL(R)T_{G}>T_{L(R)}, we have ϱf>ϱb\varrho_{f}>\varrho_{b}, that is the probability of the forward process in Fig. 8(b) is greater than the probability of the reverse process. The same can be proved in the limit of finite and large Δt\Delta t by a analytical/numerical solution of Eq. (15). Hence, for TG>TL(R)T_{G}>T_{L(R)}, the direction of electronic heat flow via the Coulomb coupled system is from the reservoir GG to the reservoir(s) L(R)L(R).

Appendix C Direction of electron flow (for V=0) in the proposed set-up (demonstrated in Fig. 1 or 8c)

Here, I show that for the arrangement demonstrated in Fig. 1 or 8(c), we get a directed motion of electrons from the reservoir LL to RR under short-circuited condition for TG>TL(R)T_{G}>T_{L(R)}. Let us assume that the system is initially in the vacuum state |0,0,0\ket{0,0,0}. From Eqns. (12) and (13), the probability that the forward cycle, that transfers an electron from LL to RR (demonstrated in Fig. 8d) while absorbing a heat packet UmU_{m} from the reservoir GG, is completed in a time-span Δt\Delta t can be written as:

ϱLR\displaystyle\varrho_{L\rightarrow R} =0ΔtγCfL(εs1)eγCfL(εs1)t1t1ΔtγCfG(εg+Um)eγCfG(εg+Um)t2×t2Δtγeγt3\displaystyle=\int_{0}^{\Delta t}\gamma_{C}f_{L}(\varepsilon_{s}^{1})e^{-\gamma_{C}f_{L}(\varepsilon_{s}^{1})t_{1}}\int_{t_{1}}^{\Delta t}{\gamma_{C}f_{G}(\varepsilon_{g}+U_{m})}e^{-\gamma_{C}f_{G}(\varepsilon_{g}+U_{m})t_{2}}\times\int_{t_{2}}^{\Delta t}\gamma e^{-\gamma t_{3}}
×t3ΔtγC{1fR(εs2)}eγC{1fR(εs2)}t4×t3ΔtγC{1fG(εg)}eγC{1fG(εg)}t5dt5dt4dt3dt3dt1\displaystyle\times\int_{t_{3}}^{\Delta t}{\gamma_{C}\{1-f_{R}(\varepsilon_{s}^{2})}\}e^{-\gamma_{C}\{1-f_{R}(\varepsilon_{s}^{2})\}t_{4}}\times\int_{t_{3}}^{\Delta t}{\gamma_{C}\{1-f_{G}(\varepsilon_{g})}\}e^{-\gamma_{C}\{1-f_{G}(\varepsilon_{g})\}t_{5}}dt_{5}dt_{4}dt_{3}dt_{3}dt_{1} (19)
ϱLR=ΠLR0ΔteγCfL(εs1)t1t1ΔteγCfG(εg+Um)t2×t2Δteγt3×t3ΔteγC{1fR(εs2)}t4\displaystyle\Rightarrow\varrho_{L\rightarrow R}=\Pi_{L\rightarrow R}\int_{0}^{\Delta t}e^{-\gamma_{C}f_{L}(\varepsilon_{s}^{1})t_{1}}\int_{t_{1}}^{\Delta t}e^{-\gamma_{C}f_{G}(\varepsilon_{g}+U_{m})t_{2}}\times\int_{t_{2}}^{\Delta t}e^{-\gamma t_{3}}\times\int_{t_{3}}^{\Delta t}e^{-\gamma_{C}\{1-f_{R}(\varepsilon_{s}^{2})\}t_{4}}
×t3ΔteγC{1fG(εg)}t5dt5dt4dt3dt3dt1\displaystyle\times\int_{t_{3}}^{\Delta t}e^{-\gamma_{C}\{1-f_{G}(\varepsilon_{g})\}t_{5}}dt_{5}dt_{4}dt_{3}dt_{3}dt_{1} (20)

where ΠLR=γγC4fL(εs1)×fG(εg+Um)×{1fR(εs2)}×{1fG(εg)}\Pi_{L\rightarrow R}=\gamma\gamma_{C}^{4}f_{L}(\varepsilon_{s}^{1})\times{f_{G}(\varepsilon_{g}+U_{m})}\times\{1-f_{R}(\varepsilon_{s}^{2})\}\times{\{1-f_{G}(\varepsilon_{g})}\} To simplify the above equation, I work in the limit of very small Δt\Delta t such that the exponential quantities in the above equations is approximately 1. Hence,

ϱLR\displaystyle\varrho_{L\rightarrow R} =ΠLR0Δtt1Δtt2Δtt3Δtt3Δt𝑑t5𝑑t4𝑑t3𝑑t3𝑑t1\displaystyle=\Pi_{L\rightarrow R}\int_{0}^{\Delta t}\int_{t_{1}}^{\Delta t}\int_{t_{2}}^{\Delta t}\int_{t_{3}}^{\Delta t}\int_{t_{3}}^{\Delta t}dt_{5}dt_{4}dt_{3}dt_{3}dt_{1}
=ΠLRΔt560\displaystyle=\Pi_{L\rightarrow R}\frac{\Delta t^{5}}{60} (21)

Similarly, assuming that the system is in ground state at t=0t=0, the probability that the reverse cycle is completed within a very small time interval Δt\Delta t is

ϱRL\displaystyle\varrho_{R\rightarrow L} =ΠRL0Δtt1Δtt2Δtt3Δtt3Δt𝑑t5𝑑t4𝑑t3𝑑t3𝑑t1\displaystyle=\Pi_{R\rightarrow L}\int_{0}^{\Delta t}\int_{t_{1}}^{\Delta t}\int_{t_{2}}^{\Delta t}\int_{t_{3}}^{\Delta t}\int_{t_{3}}^{\Delta t}dt_{5}dt_{4}dt_{3}dt_{3}dt_{1}
=ΠRLΔt560,\displaystyle=\Pi_{R\rightarrow L}\frac{\Delta t^{5}}{60}, (22)

where ΠRL=γγC4{1fL(εs1)}×{1fG(εg+Um)}×fR(εs2)×fG(εg)\Pi_{R\rightarrow L}={\gamma\gamma_{C}^{4}\{1-f_{L}(\varepsilon_{s}^{1})}\}\times\{1-f_{G}(\varepsilon_{g}+U_{m})\}\times f_{R}(\varepsilon_{s}^{2})\times f_{G}(\varepsilon_{g}). So, the ratio of the probability of the forward process to the reverse cycle to get completed within a very small interval of time Δt\Delta t, is

ϱLR/ϱRL\displaystyle\varrho_{L\rightarrow R}/\varrho_{R\rightarrow L} ={fL(εs1)}{1fL(εs1)}×fG(εg+Um){1fG(εg+Um)}×{1fR(εs2)}{fR(εs2)}×{1fG(εg)}fG(εg)\displaystyle=\frac{\{f_{L}(\varepsilon_{s}^{1})\}}{\{1-f_{L}(\varepsilon_{s}^{1})\}}\times\frac{{f_{G}(\varepsilon_{g}+U_{m})}}{\{1-f_{G}(\varepsilon_{g}+U_{m})\}}\times\frac{{\{1-f_{R}(\varepsilon_{s}^{2})}\}}{{\{f_{R}(\varepsilon_{s}^{2})}\}}\times\frac{{\{1-f_{G}(\varepsilon_{g})}\}}{f_{G}(\varepsilon_{g})}
=exp{εs1μkTL(R)}exp{εg+UmμgkTG}exp{εs2μkTL(R)}exp{εgμgkTG}\displaystyle=exp\left\{-\frac{\varepsilon_{s}^{1}-\mu}{kT_{L(R)}}\right\}exp\left\{-\frac{\varepsilon_{g}+U_{m}-\mu_{g}}{kT_{G}}\right\}exp\left\{\frac{\varepsilon_{s}^{2}-\mu}{kT_{L(R)}}\right\}exp\left\{\frac{\varepsilon_{g}-\mu_{g}}{kT_{G}}\right\}
=exp{UmkTL(R)}exp{UmkTG}\displaystyle=exp\left\{\frac{U_{m}}{kT_{L(R)}}\right\}exp\left\{-\frac{U_{m}}{kT_{G}}\right\}
=exp{Umk(1TL(R)1TG)}\displaystyle=exp\left\{\frac{U_{m}}{k}\left(\frac{1}{T_{L(R)}}-\frac{1}{T_{G}}\right)\right\} (23)

For TG=TL(R)T_{G}=T_{L(R)}, ϱLR=ϱRL\varrho_{L\rightarrow R}=\varrho_{R\rightarrow L} and no electronic current flows between the reservoirs LL and RR. However, for TG>TL(R)T_{G}>T_{L(R)}, we note that ϱLR>ϱRL\varrho_{L\rightarrow R}>\varrho_{R\rightarrow L}, causing a net electronic current flow from the reservoir LL to resevoir RR. The current reverses its direction for TG<TL(R)T_{G}<T_{L(R)}.
The variation in the thermoelectric power generation of the proposed set-up with UmU_{m} (demonstrated in Fig. 3a) can be explained in terms of Eqns, 21-23. For low values of UmU_{m}, the ratio ϱLR/ϱRL\varrho_{L\rightarrow R}/\varrho_{R\rightarrow L} is low which results in low values of the generated power. With an increase in UmU_{m}, the ratio ϱLR/ϱRL\varrho_{L\rightarrow R}/\varrho_{R\rightarrow L} increases and the forward cycle in Fig. 8(d), that generates power while harvesting heat from the reservoir GG, dominates (over the reverse cycle) resulting in an overall increase in the net generated power. With further increase in UmU_{m} beyond a certain point, despite of an increase in ϱLR/ϱRL\varrho_{L\rightarrow R}/\varrho_{R\rightarrow L}, the absolute value of ϱLR\varrho_{L\rightarrow R} deteriorates due to decrease in the product fG(εg+Um){1fG(εg)}f_{G}(\varepsilon_{g}+U_{m})\{1-f_{G}(\varepsilon_{g})\}, resulting in a decrease in the factor ΠLR\Pi_{L\rightarrow R}. Hence, the generated power decreases with further increase in UmU_{m}.

Appendix D Derivation of quantum master equations (QME) for the proposed system

Refer to caption
Figure 9: Schematic diagram demonstrating electrostatic interaction of the system with the adjacent electrodes and other dots. The voltages Vg1,Vg2,Vg3V_{g1},~{}V_{g2},~{}V_{g3} are the voltages at the gate terminals of the dots S1S_{1}, S2S_{2} and G1G_{1} respectively.

Here, starting from the basic physics of Coulomb coupled systems, I derive the quantum master equations (QME) for the proposed non-local heat engine. Fig. 9 demonstrates the equivalent model for electrostatic interaction of the system dots with the adjacent electrodes as well as the neighbouring dots. The voltages Vg1,Vg2,Vg3V_{g1},~{}V_{g2},~{}V_{g3} are the voltages at the gate terminals of the dots S1S_{1}, S2S_{2} and G1G_{1} respectively. The other symbols in Fig. 9 are self-explanatory. The potentials of the dots S1,S2S_{1},~{}S_{2} and G1G_{1} can be written in terms of the dot charge and the neighbouring potentials as cb :

VS1=QS1CS1Σ+1CS1Σ{Cg1,S1Vg1+CL,S1VL+CS1,S2VS2+CS1,G1VG1}\displaystyle V_{S_{1}}=\frac{Q_{S_{1}}}{C_{S_{1}}^{\Sigma}}+\frac{1}{C_{S_{1}}^{\Sigma}}\left\{C_{g1,S_{1}}V_{g1}+C_{L,S_{1}}V_{L}+C_{S_{1},S_{2}}V_{S_{2}}+C_{S_{1},G_{1}}V_{G_{1}}\right\}
VG1=QG1CG1Σ+1CG1Σ{Cg3,G1Vg1+CG,G1VG+CG1,S2VS2+CS1,G1VS1}\displaystyle V_{G_{1}}=\frac{Q_{G_{1}}}{C_{G_{1}}^{\Sigma}}+\frac{1}{C_{G_{1}}^{\Sigma}}\left\{C_{g3,G_{1}}V_{g1}+C_{G,G_{1}}V_{G}+C_{G_{1},S_{2}}V_{S_{2}}+C_{S_{1},G_{1}}V_{S_{1}}\right\}
VS2=QS2CS2Σ+1CS2Σ{Cg2,S2Vg2+CR,S2VR+CG1,S2VG1+CS1,S2VS1},\displaystyle V_{S_{2}}=\frac{Q_{S_{2}}}{C_{S_{2}}^{\Sigma}}+\frac{1}{C_{S_{2}}^{\Sigma}}\left\{C_{g2,S_{2}}V_{g2}+C_{R,S_{2}}V_{R}+C_{G_{1},S_{2}}V_{G_{1}}+C_{S_{1},S_{2}}V_{S_{1}}\right\}, (24)

where QxQ_{x} is the charge in dot xx and the terms CxΣC_{x}^{\Sigma} is the total capacitance seen by the dot xx with its adjacent environment.

CS1Σ=Cg1,S1+CS1,G1+CL,S1+CS1,S2\displaystyle C_{S_{1}}^{\Sigma}=C_{g1,S_{1}}+C_{S_{1},G_{1}}+C_{L,S_{1}}+C_{S_{1},S_{2}}
CS2Σ=Cg2,S2+CG1,S2+CR,S2+CS1,S2\displaystyle C_{S_{2}}^{\Sigma}=C_{g2,S_{2}}+C_{G_{1},S_{2}}+C_{R,S_{2}}+C_{S_{1},S_{2}}
CS1Σ=Cg13,G1+CS1,G1+CG,G1+CG1,S2\displaystyle C_{S_{1}}^{\Sigma}=C_{g13,G_{1}}+C_{S_{1},G_{1}}+C_{G,G_{1}}+C_{G_{1},S_{2}} (25)

In general, each dot is strongly coupled with its corresponding gate. Hence, from a practical view-point, the effective capacitance CL,S1,CR,S2C_{L,S_{1}},~{}C_{R,S_{2}} and CG,G1C_{G,G_{1}} between the dots and the tunnel-coupled electrodes can be considered negligible with respect to the gate coupling capacitances Cg1,S1,Cg2,S2C_{g1,S_{1}},~{}C_{g2,S_{2}} and Cg3,G1C_{g3,G_{1}}. In addition, the dots S1S_{1} and G1G_{1} are strongly coupled (intentionally) by suitable fabrication techniques Hübel et al. [2007a]; Chan et al. [2002]; Molenkamp et al. [1995]; Hübel et al. [2007b]; Ruzin et al. [1992]. The electrostatic coupling between S1S2S_{1}-S_{2} and G1S2G_{1}-S_{2} are again negligible, such that for all practical purposes CS1,G1>>(CS1,S2,CG1,S2)C_{S_{1},G_{1}}>>(C_{S_{1},S_{2}},C_{G_{1},S_{2}}). For all practical purposes relating to the following derivations, I hence neglect the capacitances CL,S1,CR,S2,CG,G1,CS1,S2,CG1,S2C_{L,S_{1}},~{}C_{R,S_{2}},~{}C_{G,G_{1}},~{}C_{S_{1},S_{2}},~{}C_{G_{1},S_{2}}. Under such considerations, the total electrostatic energy of the system can be written as cb :

Utot=x(S1,S2,G1)Qx22CxΣ+QS1CS1Σ{Cg1,S1Vg1+CS1,G1VG1}+QS2CS2ΣCg2,S2Vg2+QG1CG1Σ{Cg3,G1Vg3+CS1,G1VS1},\displaystyle U_{tot}=\sum_{x\in(S_{1},S_{2},G_{1})}\frac{Q_{x}^{2}}{2C_{x}^{\Sigma}}+\frac{Q_{S_{1}}}{C^{\Sigma}_{S_{1}}}\left\{C_{g1,S_{1}}V_{g1}+C_{S_{1},G_{1}}V_{G_{1}}\right\}+\frac{Q_{S_{2}}}{C^{\Sigma}_{S_{2}}}C_{g2,S_{2}}V_{g2}+\frac{Q_{G_{1}}}{C^{\Sigma}_{G_{1}}}\left\{C_{g3,G_{1}}V_{g3}+C_{S_{1},G_{1}}V_{S_{1}}\right\},

where it is assumed that capacitances CL,S1,CR,S2,CG,G1,CS1,S2,CG1,S20C_{L,S_{1}},~{}C_{R,S_{2}},~{}C_{G,G_{1}},~{}C_{S_{1},S_{2}},~{}C_{G_{1},S_{2}}\approx 0, compared to the other capacitances in the system. At 0K0K temperature, the system would reach equilibrium at the lowest possible value of UtotU_{tot}, which we call UeqU_{eq}. The charge QS1eq=qnS1eq,QS2eq=qnS2eqQ_{S_{1}}^{eq}=-qn_{S_{1}}^{eq},~{}Q_{S_{2}}^{eq}=-qn_{S_{2}}^{eq} and QG1eq=qnG1eqQ_{G_{1}}^{eq}=-qn_{G_{1}}^{eq} stored in the dot S1,S2S_{1},~{}S_{2} and G1G_{1} respectively in equilibrium (minimum energy) condition at 0K0K can be found by solving the equations:

UtotQS1=QS1CS1Σ+1CS1Σ(Cg1,S1Vg1+CS1,G1VG1)+(CS1,G1)2QS1CG1Σ(CS1Σ)2+CS1,G1QG1CS1ΣCG1Σ=0\displaystyle\frac{\partial U_{tot}}{\partial Q_{S_{1}}}=\frac{Q_{S_{1}}}{C^{\Sigma}_{S_{1}}}+\frac{1}{C_{S_{1}}^{\Sigma}}(C_{g1,S_{1}}V_{g1}+C_{S_{1},G_{1}}V_{G_{1}})+\frac{(C_{S_{1},G_{1}})^{2}Q_{S_{1}}}{C_{G_{1}}^{\Sigma}(C_{S_{1}}^{\Sigma})^{2}}+\frac{C_{S_{1},G_{1}}Q_{G_{1}}}{C_{S_{1}}^{\Sigma}C_{G_{1}}^{\Sigma}}=0
UtotQS2=QS2CS2Σ+Cg2,S2CS2ΣVg2=0\displaystyle\frac{\partial U_{tot}}{\partial Q_{S_{2}}}=\frac{Q_{S_{2}}}{C^{\Sigma}_{S_{2}}}+\frac{C_{g2,S_{2}}}{C_{S_{2}}^{\Sigma}}V_{g2}=0
UtotQG1=QG1CG1Σ+1CG1Σ(Cg3,G1Vg3+CS1,G1VG3)+(CS1,G1)2QG1CS1Σ(CG1Σ)2+CS1,G1QS1CG1ΣCS1Σ=0\displaystyle\frac{\partial U_{tot}}{\partial Q_{G_{1}}}=\frac{Q_{G_{1}}}{C^{\Sigma}_{G_{1}}}+\frac{1}{C_{G_{1}}^{\Sigma}}(C_{g3,G_{1}}V_{g3}+C_{S_{1},G_{1}}V_{G_{3}})+\frac{(C_{S_{1},G_{1}})^{2}Q_{G_{1}}}{C_{S_{1}}^{\Sigma}(C_{G_{1}}^{\Sigma})^{2}}+\frac{C_{S_{1},G_{1}}Q_{S_{1}}}{C_{G_{1}}^{\Sigma}C_{S_{1}}^{\Sigma}}=0 (27)

The above equations have been derived by partial differentiation of Eq. LABEL:eq:utot, in conjugation with replacing appropriate expressions by partially differentiating the set of Eqns. 24 along with some algebraic manipulation. At finite temperature, the number of electrons in the dots may vary stochastically due to thermal fluctuations from the reservoir. The small increase in the total electrostatic energy of the system due to thermal fluctuations from the reservoirs can be written by Taylor expanding Eq. (LABEL:eq:utot) around the equilibrium dot charges (qnS1eq,qnS2eq-qn_{S_{1}}^{eq},~{}-qn_{S_{2}}^{eq} and qnG1eq-qn_{G_{1}}^{eq}), in conjugation with the condition UtotQS1|QS1=qnS1eq=UtotQS2|QS2=qnS2eq=UtotQG1|QG1=qnG1eq=0\frac{\partial U_{tot}}{\partial Q_{S_{1}}}\Big{|}_{Q_{S_{1}}=-qn_{S_{1}}^{eq}}=\frac{\partial U_{tot}}{\partial Q_{S_{2}}}\Big{|}_{Q_{S_{2}}=-qn_{S_{2}}^{eq}}=\frac{\partial U_{tot}}{\partial Q_{G_{1}}}\Big{|}_{Q_{G_{1}}=-qn_{G_{1}}^{eq}}=0 as:

U(nS1,nG1,nS2)=UtotUeq=x(S1,G1,S2)q2Cxself(nxtotnxeq)2+(x1,x2)(S1,G1,S2)x1x2Ux1,x2(nx1totnx1eq)(nx2totnx2eq)\displaystyle U(n_{S_{1}},n_{G_{1}},n_{S_{2}})=U_{tot}-U_{eq}=\sum_{x\in(S_{1},G_{1},S_{2})}\frac{q^{2}}{C^{self}_{x}}\left(n_{x}^{tot}-n_{x}^{eq}\right)^{2}+\sum_{(x_{1},x_{2})\in(S_{1},G_{1},S_{2})}^{x_{1}\neq x_{2}}U_{x_{1},x_{2}}\left(n_{x1}^{tot}-n_{x1}^{eq}\right)\left(n_{x2}^{tot}-n_{x2}^{eq}\right)
(28)

where nxtotn_{x}^{tot} is the total number of electrons in the dot xx, and CxselfC_{x}^{self} is self capacitance of dot xx. Ux1,x2U_{x_{1},x_{2}} is the electrostatic energy arising out of mutual Coulomb coupling between two different quantum dots. These quantities can be derived from Eqns. (24), (LABEL:eq:utot) and (27), along with the assumption CL,S1=CR,S2=CG,G1=CS1,S2=CG1,S2=0C_{L,S_{1}}=C_{R,S_{2}}=C_{G,G_{1}}=C_{S_{1},S_{2}}=C_{G_{1},S_{2}}=0 as:

1CS1self=2UtotQS12=1CS1Σ+2(CS1,G1)2(CS1Σ)2CG1Σ\displaystyle\frac{1}{C_{S_{1}}^{self}}=\frac{\partial^{2}U_{tot}}{\partial Q_{S_{1}}^{2}}=\frac{1}{C_{S_{1}}^{\Sigma}}+2\frac{(C_{S_{1},G_{1}})^{2}}{(C_{S_{1}}^{\Sigma})^{2}C_{G_{1}}^{\Sigma}}
1CS2self=2UtotQS22=1CS2Σ\displaystyle\frac{1}{C_{S_{2}}^{self}}=\frac{\partial^{2}U_{tot}}{\partial Q_{S_{2}}^{2}}=\frac{1}{C_{S_{2}}^{\Sigma}}
1CG1self=2UtotQG12=1CG1Σ+2(CS1,G1)2(CG1Σ)2CS1Σ\displaystyle\frac{1}{C_{G_{1}}^{self}}=\frac{\partial^{2}U_{tot}}{\partial Q_{G_{1}}^{2}}=\frac{1}{C_{G_{1}}^{\Sigma}}+2\frac{(C_{S_{1},G_{1}})^{2}}{(C_{G_{1}}^{\Sigma})^{2}C_{S_{1}}^{\Sigma}}
US1,G1q2=2UtotQG1QS1=2UtotQS1QG1=2CS1,G1CS1ΣCG1Σ\displaystyle\frac{U_{S_{1},G_{1}}}{q^{2}}=\frac{\partial^{2}U_{tot}}{\partial Q_{G_{1}}\partial Q_{S_{1}}}=\frac{\partial^{2}U_{tot}}{\partial Q_{S_{1}}\partial Q_{G_{1}}}=2\frac{C_{S_{1},G_{1}}}{C_{S_{1}}^{\Sigma}C_{G_{1}}^{\Sigma}}
US1,S2q2=2UtotQS1QS2=2UtotQS2QS1=0\displaystyle\frac{U_{S_{1},S_{2}}}{q^{2}}=\frac{\partial^{2}U_{tot}}{\partial Q_{S_{1}}\partial Q_{S_{2}}}=\frac{\partial^{2}U_{tot}}{\partial Q_{S_{2}}\partial Q_{S_{1}}}=0
US2,G1q2=2UtotQG1QS2=2UtotQS2QG1=0\displaystyle\frac{U_{S_{2},G_{1}}}{q^{2}}=\frac{\partial^{2}U_{tot}}{\partial Q_{G_{1}}\partial Q_{S_{2}}}=\frac{\partial^{2}U_{tot}}{\partial Q_{S_{2}}\partial Q_{G_{1}}}=0 (29)

From Eq. (28), I proceed to derive the QME of the entire system. First, I consider the rate of inter-dot tunneling for the system proposed in this paper (schematically demonstrated in Fig. 1 or 8.c). In this arrangement, the additional quantum dot S2S_{2} is tunnel coupled to the quantum dot S1S_{1}, while G1G_{1} is Coulomb coupled to to S1S_{1}. I assume that the electrostatic energy arising out of self-capacitance is much greater than the average thermal energy or the applied bias voltage, that is Exself=q2Cxself>>(kT,qV)E^{self}_{x}=\frac{q^{2}}{C^{self}_{x}}>>(kT,~{}qV), such that transport through the Coulomb blocked energy level, due to self-capacitance, is negligible. Under these assumptions, the analysis may be restricted to eight multi-electron states, that can be denoted by the electron numbers in each quantum dot as |nS1,nG1,nS2=|nS1|nG1|nS2\ket{n_{S_{1}},n_{G_{1}},n_{S_{2}}}=\ket{n_{S_{1}}}\mathbin{\mathop{\otimes}\limits}\ket{n_{G_{1}}}\mathbin{\mathop{\otimes}\limits}\ket{n_{S_{2}}}, where (nS1,nG1,nS2)(0,1)(n_{S_{1}},n_{G_{1}},n_{S_{2}})\in(0,1). For simplifying the representation of these multi-electron states, I rename the states as |0,0,0|0\ket{0,0,0}\rightarrow\ket{0}, |0,0,1|1\ket{0,0,1}\rightarrow\ket{1}, |0,1,0|2\ket{0,1,0}\rightarrow\ket{2}, |0,1,1|3\ket{0,1,1}\rightarrow\ket{3}, |1,0,0|4\ket{1,0,0}\rightarrow\ket{4}, |1,0,1|5\ket{1,0,1}\rightarrow\ket{5}, |1,1,0|6\ket{1,1,0}\rightarrow\ket{6}, and |1,1,1|7\ket{1,1,1}\rightarrow\ket{7}
The Hamiltonian of the entire system consisting of these three quantum dots, hence, can be written as:

H=βϵβ|ββ|+t{|36|+|14|}\displaystyle H=\sum_{\beta}\epsilon_{\beta}\ket{\beta}\bra{\beta}+t\{\ket{3}\bra{6}+\ket{1}\bra{4}\}
+Um{|66|+|77|}+h.c.,\displaystyle+U_{m}\{\ket{6}\bra{6}+\ket{7}\bra{7}\}+h.c., (30)

where Um=US1,G1mU_{m}=U^{m}_{S_{1},G_{1}} is the electrostatic coupling energy between S1S_{1} and G1G_{1} in Fig. 1 (or 8c), tt is the interdot tunnel coupling element between S1S_{1} and S2S_{2} and ϵβ\epsilon_{\beta} is the total energy of the state |β\ket{\beta} with respect to the vacuum state |0\ket{0}. Under the assumption that the interdot coupling element tt or the reservoir to dot coupling are small, the time evolution of the system density matrix can be calculated by taking the partial trace over the total density matrix of the combined set-up consisting of the reservoirs and the dots Gurvitz [1998]; Hazelzet et al. [2001]; Dong et al. [2008, 2004]; Sztenkiel and Świrkowicz [2007]; Wegewijs and Nazarov [1999]. In particular, the diagonal and the non-diagonal terms of the density matrix ρ\rho of the system consisting only of the arrays of quantum dots may be written as a set of modified Liouville euqation Gurvitz [1998]; Hazelzet et al. [2001]; Dong et al. [2008, 2004]; Sztenkiel and Świrkowicz [2007]; Wegewijs and Nazarov [1999]:

ρηηt=i[H,ρ]ηηνΓηνρηη+δΓδηρδδ\displaystyle\frac{\partial\rho_{\eta\eta}}{\partial t}=-i[H,\rho]_{\eta\eta}-\sum_{\nu}\Gamma_{\eta\nu}\rho_{\eta\eta}+\sum_{\delta}\Gamma_{\delta\eta}\rho_{\delta\delta}
ρηβt=i[H,ρ]ηβ12ν(Γην+Γβν)ρηβ,\displaystyle\frac{\partial\rho_{\eta\beta}}{\partial t}=-i[H,\rho]_{\eta\beta}-\frac{1}{2}\sum_{\nu}\Big{(}\Gamma_{\eta\nu}+\Gamma_{\beta\nu}\Big{)}\rho_{\eta\beta},
(31)

where [x,y][x,y] denotes the commutator of the operators xx and yy and ρηβ=η|ρ|β\rho_{\eta\beta}=\bra{\eta}\rho\ket{\beta}. The elements ρηη\rho_{\eta\eta} and ρηβ\rho_{\eta\beta} in the above equation denote any diagonal and non-diagonal element of the system density matrix respectively. The off-diagonal elements ρηβ\rho_{\eta\beta} result from coherent inter-dot tunnelling and tunnelling of electrons between the dots and the reservoirs. The off-diagonal terms, hence, are only non-zero when electron tunneling can result in the transition between the states η\eta and β\beta. The parameters Γxy\Gamma_{xy} account for the transition between system states due to electronic tunneling between the system and the reservoirs and are only finite when the system state transition from |x\ket{x} to |y\ket{y} (or vice-versa) is possible due to electron transfer between the system and the reservoirs. Assuming a statistical quasi-equilibrium distribution of electrons inside the reservoirs, Γxy\Gamma_{xy} can be expressed as:

Γxy=γcfλ(ϵyϵx),\displaystyle\Gamma_{xy}=\gamma_{c}f_{\lambda}(\epsilon_{y}-\epsilon_{x}), (32)

where γc=γl=γr=γg{\gamma_{c}=\gamma_{l}=\gamma_{r}=\gamma_{g}} denotes the system to reservoir coupling corresponding to the state transition, fλ(ϵ)f_{\lambda}(\epsilon) denotes the probability of occupancy of an electron in the corresponding reservoir λ\lambda (driving the state transition) at energy ϵ\epsilon and ϵx(y)\epsilon_{x(y)} is the total electronic energy of the system in the state |x(y)\ket{x(y)} compared to the vacuum state. In the context of this discussion, I assume equilibrium Fermi-Dirac carrier statistics in the reservoir.
For this system, the tunneling of electrons between the quantum dots corresponds to the change of the system states from |4\ket{4} to |1\ket{1} (or vice-versa) and from |3\ket{3} to |6\ket{6} (or vice-versa). In steady state, the time-derivative of each element of the density matrix [ρ][\rho] vanishes. Hence, using the second equation of (31), we get,

ρ4,1=ρ1,4=ρ4,4ρ1,1ϵ4ϵ1iΥ4,12\rho_{4,1}=\rho^{*}_{1,4}=\frac{\rho_{4,4}-\rho_{1,1}}{\epsilon_{4}-\epsilon_{1}-i\frac{\Upsilon_{4,1}}{2}} (33)
ρ6,3=ρ3,6=ρ6,6ρ3,3ϵ6ϵ3iΥ6,32,\rho_{6,3}=\rho^{*}_{3,6}=\frac{\rho_{6,6}-\rho_{3,3}}{\epsilon_{6}-\epsilon_{3}-i\frac{\Upsilon_{6,3}}{2}}, (34)

where Υx,y\Upsilon_{x,y} is the sum of all process rates arising from the system to reservoir coupling that leads to the decay of the states |x\ket{x} and |y\ket{y}. In Eqns. (33) and (34), Υ4,1\Upsilon_{4,1} and Υ6,3\Upsilon_{6,3} can be written as:

Υ4,1=Γ|4,|0+Γ|4,|6+Γ|4,|5+Γ|1,|0+Γ|1,|6+Γ|1,|3\displaystyle\Upsilon_{4,1}=\Gamma_{\ket{4},\ket{0}}+\Gamma_{\ket{4},\ket{6}}+\Gamma_{\ket{4},\ket{5}}+\Gamma_{\ket{1},\ket{0}}+\Gamma_{\ket{1},\ket{6}}+\Gamma_{\ket{1},\ket{3}}
Υ6,3=Γ|6,|4+Γ|6,|2+Γ|6,|7+Γ|3,|1+Γ|3,|2+Γ|3,|7\displaystyle\Upsilon_{6,3}=\Gamma_{\ket{6},\ket{4}}+\Gamma_{\ket{6},\ket{2}}+\Gamma_{\ket{6},\ket{7}}+\Gamma_{\ket{3},\ket{1}}+\Gamma_{\ket{3},\ket{2}}+\Gamma_{\ket{3},\ket{7}} (35)

From the first equation of (31), the time derivative of the diagonal elements ρ6,6\rho_{6,6} and ρ3,3\rho_{3,3} of the density matrix can be written as:

ρ˙6,6=\displaystyle\dot{\rho}_{6,6}= it(ρ6,3ρ3,6)(Γ|6,|4Γ|6,|2+Γ|6,|7)ρ6,6\displaystyle it(\rho_{6,3}-\rho_{3,6})-\left(\Gamma_{\ket{6},\ket{4}}-\Gamma_{\ket{6},\ket{2}}+\Gamma_{\ket{6},\ket{7}}\right)\rho_{6,6}
+Γ|4,|6ρ4,4+Γ|2,|6ρ2,2+Γ|7,|6ρ7,7\displaystyle+\Gamma_{\ket{4},\ket{6}}\rho_{4,4}+\Gamma_{\ket{2},\ket{6}}\rho_{2,2}+\Gamma_{\ket{7},\ket{6}}\rho_{7,7}
ρ˙4,4=\displaystyle\dot{\rho}_{4,4}= it(ρ4,1ρ1,4)(Γ|4,|0Γ|4,|6+Γ|4,|5)ρ4,4\displaystyle it(\rho_{4,1}-\rho_{1,4})-\left(\Gamma_{\ket{4},\ket{0}}-\Gamma_{\ket{4},\ket{6}}+\Gamma_{\ket{4},\ket{5}}\right)\rho_{4,4}
+Γ|0,|4ρ0,0+Γ|6,|4ρ6,6+Γ|5,|4ρ5,5\displaystyle+\Gamma_{\ket{0},\ket{4}}\rho_{0,0}+\Gamma_{\ket{6},\ket{4}}\rho_{6,6}+\Gamma_{\ket{5},\ket{4}}\rho_{5,5}

Substituting the values of ρ6,3,ρ3,6,ρ4,1\rho_{6,3},~{}\rho_{3,6},~{}\rho_{4,1} and ρ1,4\rho_{1,4} in Eq. (LABEL:eq:b8) from Eq. (33) and (34), the probability of decay of the states |4\ket{4} and |6\ket{6} can be written as:

p6˙=ρ˙6,6=α=(Γ|6,|αp6+Γ|α,|6pα)\displaystyle\dot{p_{6}}=\dot{\rho}_{6,6}=\sum_{\alpha=}\left(-\Gamma_{\ket{6},\ket{\alpha}}p_{6}+\Gamma_{\ket{\alpha},\ket{6}}p_{\alpha}\right)
Λ|6,|3p6+Λ|3,|6p3\displaystyle-\Lambda_{\ket{6},\ket{3}}p_{6}+\Lambda_{\ket{3},\ket{6}}p_{3} (37)

p4˙=ρ˙4,4=α=(Γ|4,|αp3+Γ|α,|4pα)\displaystyle\dot{p_{4}}=\dot{\rho}_{4,4}=\sum_{\alpha=}\left(-\Gamma_{\ket{4},\ket{\alpha}}p_{3}+\Gamma_{\ket{\alpha},\ket{4}}p_{\alpha}\right)
Λ|4,|1p4+Λ|1,|4p1,\displaystyle-\Lambda_{\ket{4},\ket{1}}p_{4}+\Lambda_{\ket{1},\ket{4}}p_{1}, (38)

where pη=ρη,ηp_{\eta}=\rho_{\eta,\eta} and

Λ|6,|3=Λ|3,|6=t2Υ6,3(ϵ6ϵ3)2+Υ6,324\displaystyle\Lambda_{\ket{6},\ket{3}}=\Lambda_{\ket{3},\ket{6}}=t^{2}\frac{\Upsilon_{6,3}}{(\epsilon_{6}-\epsilon_{3})^{2}+\frac{\Upsilon_{6,3}^{2}}{4}}
Λ|4,|1=Λ|1,|4=t2Υ4,1(ϵ4ϵ1)2+Υ4,124\displaystyle\Lambda_{\ket{4},\ket{1}}=\Lambda_{\ket{1},\ket{4}}=t^{2}\frac{\Upsilon_{4,1}}{(\epsilon_{4}-\epsilon_{1})^{2}+\frac{\Upsilon_{4,1}^{2}}{4}}

In the above equation, Λ|4,|1\Lambda_{\ket{4},\ket{1}} and Λ|6,|3\Lambda_{\ket{6},\ket{3}} are the interdot tunneling rates when the ground state of G1G_{1} is unoccupied and occupied respectively. By a clever choice of the energy states, such that, ϵ6=εg+εs1+Um=εg+εs2=ϵ3\epsilon_{6}=\varepsilon_{g}+\varepsilon_{s}^{1}+U_{m}=\varepsilon_{g}+\varepsilon_{s}^{2}=\epsilon_{3}, that is by choosing εs2=εs1+Um\varepsilon_{s}^{2}=\varepsilon_{s}^{1}+U_{m}, we can arrive at a condition where Λ|6,|3>>Λ|4,|1\Lambda_{\ket{6},\ket{3}}>>\Lambda_{\ket{4},\ket{1}} (under the condition Um>>|Υ4,1|U_{m}>>|\Upsilon_{4,1}|). Such a condition implies that the tunneling probability between the dots is negligible in the absence of an electron in G1G_{1}, which initiates a unidirection flow of electrons when TG>TL(R)T_{G}>T_{L(R)}.
For the calculation of current, it is sufficient to know the probability of occupancy of the dot S1S_{1} or S2S_{2}. Since, the electronic transport in S1S_{1} and G1G_{1} are coupled to each other via Coulomb interaction, I will treat S1S_{1} and G1G_{1} as a separate sub-system (ς1\varsigma_{1}) of the entire system, S2S_{2} being another sub-system (ς2\varsigma_{2}) of the entire system consisting of the three dots. To simplify my calculations, I assume that Λ|4,|1<<Λ|6,|3\Lambda_{\ket{4},\ket{1}}<<\Lambda_{\ket{6},\ket{3}}, such that for all practical purposes relating to electron transport Λ|4,|10\Lambda_{\ket{4},\ket{1}}\approx 0. In what follows, I simply denote Λ|6,|3\Lambda_{\ket{6},\ket{3}} as γ\gamma. γ\gamma thus denotes the interdot tunnel coupling. I write the probability of occupancy of the subsystem ς1\varsigma_{1} as Pi,jς1P_{i,j}^{\varsigma_{1}}, where ii and jj denote the number of electrons in the ground state of the dot S1S_{1} and G1G_{1} respectively. Pzς2P_{z}^{\varsigma_{2}}, on the other hand, would be used to denote the probability of occupancy of the dot S2S_{2}. Note that breaking down the entire system into two sub-system in this fashion is permissible only in the limit of weak tunnel and Coulomb coupling between the two sub-systems so that the state of one sub-system doesn’t affect the state of the complementary sub-system. In such a limit, we can write ρ0,0=P0,0ς1P0ς2,ρ1,1=P0,0ς1P1ς2,ρ2,2=P0,1ς1P0ς2,ρ3,3=P0,1ς1P1ς2,ρ4,4=P1,0ς1P0ς2,ρ5,5=P1,0ς1P1ς2,ρ6,6=P1,1ς1P1ς2,ρ7,7=P1,1ς1P1ς2\rho_{0,0}=P^{\varsigma_{1}}_{0,0}P^{\varsigma_{2}}_{0},~{}\rho_{1,1}=P^{\varsigma_{1}}_{0,0}P^{\varsigma_{2}}_{1},~{}\rho_{2,2}=P^{\varsigma_{1}}_{0,1}P^{\varsigma_{2}}_{0},~{}\rho_{3,3}=P^{\varsigma_{1}}_{0,1}P^{\varsigma_{2}}_{1},~{}\rho_{4,4}=P^{\varsigma_{1}}_{1,0}P^{\varsigma_{2}}_{0},~{}\rho_{5,5}=P^{\varsigma_{1}}_{1,0}P^{\varsigma_{2}}_{1},~{}\rho_{6,6}=P^{\varsigma_{1}}_{1,1}P^{\varsigma_{2}}_{1},~{}\rho_{7,7}=P^{\varsigma_{1}}_{1,1}P^{\varsigma_{2}}_{1} The rate equations for the sub-system ς1\varsigma_{1} can be written in terms of two or more diagonal elements of the density matrix, in (31), as Singha [2020b]:

ddt(P0,0ς1)=ddt(ρ0,0+ρ1,1)=\displaystyle\frac{d}{dt}(P_{0,0}^{\varsigma_{1}})=\frac{d}{dt}\left(\rho_{0,0}+\rho_{1,1}\right)= γc×{P0,0ς1{fL(εs1)+fG(εg)}+P0,1ς1{1fG(εg)}+P1,0ς1{1fL(εs1)}}\displaystyle\gamma_{c}\times\left\{-P_{0,0}^{\varsigma_{1}}\{f_{L}(\varepsilon_{s}^{1})+f_{G}(\varepsilon_{g})\}+P_{0,1}^{\varsigma_{1}}\{1-f_{G}(\varepsilon_{g})\}+P_{1,0}^{\varsigma_{1}}\{1-f_{L}(\varepsilon_{s}^{1})\}\right\}
ddt(P1,0ς1)=ddt(ρ5,5+ρ4,4)=\displaystyle\frac{d}{dt}(P_{1,0}^{\varsigma_{1}})=\frac{d}{dt}\left(\rho_{5,5}+\rho_{4,4}\right)= γc×{P1,0ς1{1fL(εs1)+fG(εg+Um)}+P1,1ς1{1fG(εg+Um)}+P0,0ς1fG(εg)}\displaystyle\gamma_{c}\times\left\{-P_{1,0}^{\varsigma_{1}}\left\{1-f_{L}(\varepsilon_{s}^{1})+f_{G}(\varepsilon_{g}+U_{m})\right\}+P_{1,1}^{\varsigma_{1}}\left\{1-f_{G}(\varepsilon_{g}+U_{m})\right\}+P_{0,0}^{\varsigma_{1}}f_{G}(\varepsilon_{g})\right\}
ddt(P0,1ς1)=ddt(ρ2,2+ρ3,3)=\displaystyle\frac{d}{dt}(P_{0,1}^{\varsigma_{1}})=\frac{d}{dt}\left(\rho_{2,2}+\rho_{3,3}\right)= γc×{P0,1ς1{1fg(εg1)+fL(εs1+Um)+γγcP1ς2}}\displaystyle\gamma_{c}\times\left\{-P_{0,1}^{\varsigma_{1}}\left\{1-f_{g}(\varepsilon_{g}^{1})+f_{L}(\varepsilon_{s}^{1}+U_{m})+\frac{\gamma}{\gamma_{c}}P^{\varsigma_{2}}_{1}\right\}\right\}
+γc{P0,0ς1fG(εg)+P1,1ς1{1fL(εs1+Um)+γγcP0ς2}}\displaystyle+\gamma_{c}\left\{P_{0,0}^{\varsigma_{1}}f_{G}(\varepsilon_{g})+P_{1,1}^{\varsigma_{1}}\left\{1-f_{L}(\varepsilon_{s}^{1}+U_{m})+\frac{\gamma}{\gamma_{c}}P^{\varsigma_{2}}_{0}\right\}\right\}
ddt(P1,1ς1)=ddt(ρ7,7+ρ6,6)=\displaystyle\frac{d}{dt}(P_{1,1}^{\varsigma_{1}})=\frac{d}{dt}\left(\rho_{7,7}+\rho_{6,6}\right)= γc×{P1,1ς1{[1fg(εg1+Um)]+[1fL(εs1+Um)]+γγCP0ς2}}\displaystyle\gamma_{c}\times\left\{-P_{1,1}^{\varsigma_{1}}\left\{[1-f_{g}(\varepsilon_{g}^{1}+U_{m})]+[1-f_{L}(\varepsilon_{s}^{1}+U_{m})]+\frac{\gamma}{\gamma_{C}}P^{\varsigma_{2}}_{0}\right\}\right\}
+γc{P1,0ς1fG(εg+Um)+P0,1ς1{fL(εs1+Um)+γγcP1ς2}}\displaystyle+\gamma_{c}\left\{P_{1,0}^{\varsigma_{1}}f_{G}(\varepsilon_{g}+U_{m})+P_{0,1}^{\varsigma_{1}}\left\{f_{L}(\varepsilon_{s}^{1}+U_{m})+\frac{\gamma}{\gamma_{c}}P^{\varsigma_{2}}_{1}\right\}\right\}

where Λ|4,|1=Λ|1,|4=0\Lambda_{\ket{4},\ket{1}}=\Lambda_{\ket{1},\ket{4}}=0 and γ=Λ|6,|3=Λ|3,|6\gamma=\Lambda_{\ket{6},\ket{3}}=\Lambda_{\ket{3},\ket{6}}. I assume quasi Fermi-Dirac carrier statistics at the reservoirs such that fλ(ϵ)={1+exp(ϵμλkTλ)}1f_{\lambda}(\epsilon)=\left\{1+exp\left(\frac{\epsilon-\mu_{\lambda}}{kT_{\lambda}}\right)\right\}^{-1}, corresponding to the reservoir λ\lambda, and λ(L,R,G)\lambda\in(L,R,G). Similarly, the rate equations of the sub-system ς2\varsigma_{2} can be written as:

ddt(P0ς2)=ddt(ρ6,6+ρ4,4+ρ2,2+ρ0,0)=γc×{P0ς2{fR(εs2)+γγcP1,1ς1}+P1ς2{1fR(εs2)+γγcP0,1ς1}}\displaystyle\frac{d}{dt}(P_{0}^{\varsigma_{2}})=\frac{d}{dt}\left(\rho_{6,6}+\rho_{4,4}+\rho_{2,2}+\rho_{0,0}\right)=\gamma_{c}\times\left\{-P_{0}^{\varsigma_{2}}\{f_{R}(\varepsilon_{s}^{2})+\frac{\gamma}{\gamma_{c}}P_{1,1}^{\varsigma_{1}}\}+P_{1}^{\varsigma_{2}}\{1-f_{R}(\varepsilon_{s}^{2})+\frac{\gamma}{\gamma_{c}}P^{\varsigma_{1}}_{0,1}\}\right\}
ddt(P1ς2)=ddt(ρ7,7+ρ5,5+ρ3,3+ρ1,1)=γc×{P1ς2{1fR(εs2)+γγcP0,1ς1}+P0ς2{fR(εs2)+γγcP1,1ς1}}\displaystyle\frac{d}{dt}(P_{1}^{\varsigma_{2}})=\frac{d}{dt}\left(\rho_{7,7}+\rho_{5,5}+\rho_{3,3}+\rho_{1,1}\right)=\gamma_{c}\times\left\{-P_{1}^{\varsigma_{2}}\{1-f_{R}(\varepsilon_{s}^{2})+\frac{\gamma}{\gamma_{c}}P^{\varsigma_{1}}_{0,1}\}+P_{0}^{\varsigma_{2}}\{f_{R}(\varepsilon_{s}^{2})+\frac{\gamma}{\gamma_{c}}P_{1,1}^{\varsigma_{1}}\}\right\}

In steady state, the L.H.S of Eqns. (LABEL:eq:first_sys1) and (LABEL:eq:second_sys1) are zero. Both the sets of Eqns. (LABEL:eq:first_sys1) and (LABEL:eq:second_sys1) form dependent sets of equations. The dependency of the sets of Eqns. (LABEL:eq:first_sys1) and (LABEL:eq:second_sys1) is broken by introducing the probability conservation rules, that is, x,yPx,yς1=1\sum_{x,y}P_{x,y}^{\varsigma_{1}}=1 for  (LABEL:eq:first_sys1) and zPzς2=1\sum_{z}P_{z}^{\varsigma_{2}}=1 for (LABEL:eq:second_sys1). The set of Eqns. (LABEL:eq:first_sys1) and (LABEL:eq:second_sys1) form a coupled system of equations which, for the case of my study, were solved using iterative Newton-Raphson method. On solution of the state probabilities given by Eqns. (LABEL:eq:first_sys1) and (LABEL:eq:second_sys1), the charge current IL(R)I_{L(R)} through the system and the electronic heat current (IQeI_{Qe}) extracted from the reservoir GG can be calculated using the equations:

IL=qγc×{P0,0ς1fL(εs1)+P0,1ς1fL(εs1+Um)P1,0ς1{1fL(εs1)}P1,1ςs1{1fL(εs1+Um)}}I_{L}=q\gamma_{c}\times\left\{P^{\varsigma_{1}}_{0,0}f_{L}(\varepsilon_{s}^{1})+P^{\varsigma_{1}}_{0,1}f_{L}(\varepsilon_{s}^{1}+U_{m})-P^{\varsigma_{1}}_{1,0}\{1-f_{L}(\varepsilon_{s}^{1})\}-P^{\varsigma_{s}^{1}}_{1,1}\{1-f_{L}(\varepsilon_{s}^{1}+U_{m})\}\right\} (42)
IR=qγc×{P0ς2fR(εs1)P1ς2{1fR(εs1)}}I_{R}=-q\gamma_{c}\times\left\{P^{\varsigma_{2}}_{0}f_{R}(\varepsilon_{s}^{1})-P^{\varsigma_{2}}_{1}\{1-f_{R}(\varepsilon_{s}^{1})\}\right\} (43)
IQe=γc×{(εg+Umμg){P1,0ς1fG(εg+Um)P1,1ς1{1fG(εg+Um)}}}\displaystyle I_{Qe}=\gamma_{c}\times\left\{(\varepsilon_{g}+U_{m}-\mu_{g})\left\{P^{\varsigma_{1}}_{1,0}f_{G}(\varepsilon_{g}+U_{m})-P^{\varsigma_{1}}_{1,1}\{1-f_{G}(\varepsilon_{g}+U_{m})\}\right\}\right\}
+γc×{(εgμg)×{P0,0ς1fG(εg)P0,1ςn{1fG(εg)}}}\displaystyle+\gamma_{c}\times\left\{(\varepsilon_{g}-\mu_{g})\times\left\{P^{\varsigma_{1}}_{0,0}f_{G}(\varepsilon_{g})-P^{\varsigma_{n}}_{0,1}\{1-f_{G}(\varepsilon_{g})\}\right\}\right\} (44)

Since, no net current flows from the reservoir GG, we have

IG=qγc×{P1,0ς1fG(εg+Um)P1,1ς1{1fG(εg+Um)}+P0,0ς1fG(εg)P0,1ςn{1fG(εg)}}=0I_{G}=q\gamma_{c}\times\left\{P^{\varsigma_{1}}_{1,0}f_{G}(\varepsilon_{g}+U_{m})-P^{\varsigma_{1}}_{1,1}\{1-f_{G}(\varepsilon_{g}+U_{m})\}+P^{\varsigma_{1}}_{0,0}f_{G}(\varepsilon_{g})-P^{\varsigma_{n}}_{0,1}\{1-f_{G}(\varepsilon_{g})\}\right\}=0 (45)

Substituting Eq. (45) in Eq. (44), the equation for IQeI_{Qe} becomes modified as:

IQe=γc×Um{P1,0ς1fG(εg+Um)P1,1ς1{1fG(εg+Um)}}I_{Qe}=\gamma_{c}\times U_{m}\left\{P^{\varsigma_{1}}_{1,0}f_{G}(\varepsilon_{g}+U_{m})-P^{\varsigma_{1}}_{1,1}\{1-f_{G}(\varepsilon_{g}+U_{m})\}\right\} (46)

Appendix E Effect of the ratio r=γ/γcr=\gamma/\gamma_{c} on heat harvesting properties of the proposed set-up

Refer to caption
Figure 10: Variation in heat harvesting performance of the proposed system as the ratio r=γ/γcr=\gamma/\gamma_{c} is tuned, keeping the system-to-reservoir coupling constant at γc=105qh\gamma_{c}=10^{-5}\frac{q}{h}. Plot of (a) maximum generated power (PMP_{M}) and (b) Efficiency at the maximum generated power, for various values of rr, with variation in the applied voltage for Um=3.9meV(6kT/q)U_{m}=3.9meV(\approx 6kT/q) and TG=10K,TL(R)=5KT_{G}=10K,~{}T_{L(R)}=5K. T=TG+TL(R)2=7.5KT=\frac{T_{G}+T_{L(R)}}{2}=7.5K is the average temperature of the system.
Refer to caption
Figure 11: Variation in heat harvesting properties of the proposed system as the values of γc\gamma_{c} and γ\gamma are interchanged. The two cases demonstrated here are (i) γc=105qh\gamma_{c}=10^{-5}\frac{q}{h} and γ=104qh\gamma=10^{-4}\frac{q}{h}, resulting in r=10r=10 (shown in green) and (ii) γc=104qh\gamma_{c}=10^{-4}\frac{q}{h} and γ=105qh\gamma=10^{-5}\frac{q}{h}, resulting in r=0.1r=0.1 (shown in brown). Plot of (a) maximum generated power (PMP_{M}) and (b) Efficiency at the maximum generated power, with variation in the applied voltage for Um=3.9meV(6kT/q)U_{m}=3.9meV(\approx 6kT/q) and TG=10K,TL(R)=5KT_{G}=10K,~{}T_{L}(R)=5K. T=TG+TL(R)2=7.5KT=\frac{T_{G}+T_{L(R)}}{2}=7.5K is the average temperature of the system.

In this section, I demonstrate the variation in heat harvesting properties of the proposed set-up with a variation in r=γ/γcr=\gamma/\gamma_{c}. Throughout the main discussion in this paper, the values of the system-to-reservoir and interdot coupling were chosen to be γc=105qh\gamma_{c}=10^{-5}\frac{q}{h} and γ=104qh\gamma=10^{-4}\frac{q}{h}. Since, the conductance between the reservoirs LL and RR depends on both the interdot coupling and system-to-reservoir coupling , reducing either of the two deteriorates the conductance and hence, degrades the generated power. In the Fig. 10 below, I plot the variation in maximum generated power, as well as the efficiency at the maximum generated power for γc=105qh\gamma_{c}=10^{-5}\frac{q}{h}, while tuning the ratio r=γ/γcr=\gamma/\gamma_{c}. We note that for high values of rr or for low values of rr, the conductance of the system is determined by min(γc,γ)min(\gamma_{c},~{}\gamma). For r10r\geq 10, the conductance of the system is mainly determined by γc\gamma_{c} and hence increasing γ\gamma further has negligible effect on the heat harvesting properties of the system. However, as rr is gradually decreased from r=10r=10, the overall conductance of the system decreases, resulting in a decrease in the generated power. The decrease in generation efficiency with decrease in γ\gamma is a bit complicated to understand. Let us consider the cycle I1|0,0,0|1,0,0|1,1,0|0,1,1|0,1,0|0,0,0I_{1}\Rightarrow\ket{0,0,0}\rightarrow\ket{1,0,0}\rightarrow\ket{1,1,0}\rightarrow\ket{0,1,1}\rightarrow\ket{0,1,0}\rightarrow\ket{0,0,0}. This cycle (explained in the main text) transmits an electron from LL to RR while absorbing a heat packet UmU_{m} from GG. Note that in this cycle, an electron has to undergo interdot tunneling and hence the rate of this cycle depends on both the interdot coupling γ\gamma and the system-to-reservoir coupling. Next, let us consider the cycle I2|0,0,0|1,0,0|1,1,0|0,1,0|0,0,0I_{2}\Rightarrow\ket{0,0,0}\rightarrow\ket{1,0,0}\rightarrow\ket{1,1,0}\rightarrow\ket{0,1,0}\rightarrow\ket{0,0,0}. In this cycle, a heat packet UmU_{m} is absorbed from the reservoir GG without a net transport of electrons between the reservoirs. Hence, such processes lead to a deterioration of the overall efficiency. Note that in this cycle, an electron doesn’t undergo inter-dot tunneling and hence the rate of this process is dependent mainly on γc\gamma_{c}. As the ratio rr is decreased keeping γc\gamma_{c} constant, the rate of the cycle I1I_{1} decreases while I2I_{2} remains unaffected, that is a higher fraction of heat energy is lost without any transport of electrons between LL and RR. This leads to a deterioration in the generation efficiency with decrease in rr.
Fig. 11 demonstrates the variation in heat harvesting performance of the system when the values of γ\gamma and γc\gamma_{c} are interchanged. In Fig. 11, I demonstrate the heat harvesting performance for (i) γc=105qh\gamma_{c}=10^{-5}\frac{q}{h} and γ=104qh\gamma=10^{-4}\frac{q}{h}, resulting in r=10r=10 (shown in green) and (ii) γc=104qh\gamma_{c}=10^{-4}\frac{q}{h} and γ=105qh\gamma=10^{-5}\frac{q}{h}, resulting in r=0.1r=0.1 (shown in brown). Since the overall conductance of the system, for large and small rr, is determined by the min(γc,γ)min(\gamma_{c},~{}\gamma), which is equal to 105qh10^{-5}\frac{q}{h} in both the cases, there is negligible variation in the generated power between the two cases. The generation efficiency, however, decreases drastically for case (ii) when the values of γ\gamma and γc\gamma_{c} are interchanged. This is because a higher value of γc\gamma_{c} in case (ii) drastically increases the rate of the cycles I2I_{2} (as discussed above), leading to more heat packets being wasted without any net transmission of electrons between LL and RR. This causes a deterioration in the generation efficiency.

References