In the present work, we theoretically study thermoelectric transport and heat transfer in a junction including a double quantum dot in a serial configuration coupled to nonferromagnetic electrodes. We focus on the electron transport within the Coulomb blockade regime in the limit of strong intradot interactions between electrons. It is shown that under these conditions, characteristics of thermoelectric transport in such systems strongly depend on electron occupation on the dots and on interdot Coulomb interactions. We demonstrate that these factors may lead to a heat current rectification and analyze potentialities of a double-dot in a serial configuration as a heat diod.

Studies of thermoelectric transport in tailored nanoscale systems attract significant attention due to its fundamental and applied perspectives.1–6 In general, thermoelectric charge and heat transport is governed by simultaneous action of electric and thermal driving forces. The combined effect of these forces depends on several factors, including coupling of the “bridge” part of the system (most commonly, a molecule, single or multiple quantum dots, or carbon-based nanostructure) to electrodes7–15 and Coulomb interactions between electrons. Besides other things, the latter lead to oscillations of electron conductance and thermal conductance,16–19 violations of the Wiedemann–Franz law,20,21 and quantum interference effects.22–29 Also, thermoelectric transport may be strongly affected by electron–phonon interactions30–40 and by the interplay between charge/energy transfer and spin transport occurring when “the bridge” is sandwiched between ferromagnetic electrodes.22,41 Advances in control and the measurements of heat transferred through mesoscopic and nanoscale systems42–44 open up possibilities for manufacturing novel nanoelectronic devices such as heat-to-electric energy converters,45–48 cooling systems,2,3,49,50 and heat rectifiers.51–56 

In the present work, we concentrate on thermoelectric properties of a double quantum dot placed in between nonferromagnetic electrodes. More specifically, we theoretically analyze the effect of interdot Coulomb interactions on thermoelectric transport in such systems. Methods of quantum dots fabrication are sufficiently well advanced.57,58 Usually, sets including several dots are manufactured by applying a suitable electrostatic potential to GaAs/AlGaAs superstructures or heavy molecules.59–66 In such sets, the dots may be arranged in a serial configuration59–63 with hard59–61 or soft62,63 separating barriers or in a parallel geometry.64–66 

Transport properties of double quantum dots were (still are) rather intensely studied. Specifically, electron transport through a double quantum dot in a serial configuration was discussed in several works. Various manifestations of Kondo effect in such systems were analyzed67–70 along with the effects of ferromagnetic69,71 and superconducting72 electrodes, the effects of spin–orbit interactions and those of the charge and spin pumping,73 and the influence of Coulomb interactions between electrons on the efficiency of heat-to-electric energy conversion.74 Also, it was shown that double dots in a serial configuration might undergo quantum dynamical phase transitions.75 

Nevertheless, in studies of transport properties of double and/or triple quantum dots, more attention so far was given to dots in parallel geometry. Here, we study thermoelectric transport through a double dot in a serial configuration schematically shown in Fig. 1. In the considered system, single-level dot 1 is coupled to the left electrode and another single-level dot (dot 2) to the right one. Electron tunneling between the dots is characterized by a coupling parameter β. Also, we take into account Coulomb repulsion between electrons on the same dot, as well as interdot Coulomb interactions.

FIG. 1.

Schematics of the considered system. Single-level dot 1 with the electron energy on the dot E01 and the intradot charging energy U1 is coupled to the left electrode with the coupling strength ΓL. Dot 2 with the electron energy on the dot E02 and the intradot charging energy U2 is coupled to the right electrode with the coupling strength ΓR. The charging energy U and the parameter β characterize interdot Coulomb interactions and the coupling between the dots, respectively.

FIG. 1.

Schematics of the considered system. Single-level dot 1 with the electron energy on the dot E01 and the intradot charging energy U1 is coupled to the left electrode with the coupling strength ΓL. Dot 2 with the electron energy on the dot E02 and the intradot charging energy U2 is coupled to the right electrode with the coupling strength ΓR. The charging energy U and the parameter β characterize interdot Coulomb interactions and the coupling between the dots, respectively.

Close modal

This paper is organized as follows: In Sec. II, we describe the employed model and the formalism. We apply obtained results to briefly analyze the effect of interdot Coulomb interactions on the Seebeck effect and on thermally induced charge current in the considered system in Secs. III and IV. In Sec. V, we concentrate on the heat transport. It is shown that interdot Coulomb interactions may result in a significant heat current rectification. Conclusions are given in Sec. VI. In  Appendixes A and  B, we describe derivation of the relevant Green’s functions.

Studies of electron transport through quantum dots are simplified by the fact that phonon contributions to their electrical and thermal conductances are often small due to mismatch between frequencies of phonon modes associated with the dots embedded in their matrices and modes associated with the environment, preventing an overlap between the two.76,77 On these grounds, we omit electron–phonon interactions and assume that the coherent electron transmission is the predominant mechanism governing electron transport. Within Anderson model, the pair of quantum dots coupled in series and put between electrodes is described by the Hamiltonian

(1)

Here, the contribution from the electrodes has the form

(2)

where α = {L, R} labels the left and right electrodes, ϵα,k,σ are single electron energies in the electrodes, and Cα,k,σ,Cα,k,σ are creation and annihilation operators for electron states k, σ (the index σ labeling spin-up and spin-down electrons). The term Hd is associated with the dots,

(3)

In this expression, i = 1, 2, E0i are electron energies on the dots active in the transport process, Ui are intradot charging energies, U describes the interdot Coulomb interactions, ni,σ=di,σdi,σ, and di,σ,di,σ are creation and annihilation operators for electrons on the dots.

The last term

(4)

describes the couplings of the dots to the electrodes where factors tα,k,σ represent coupling strengths.

Basing on the Hamiltonian (1), we derive expressions for the retarded [Gr(E)] and advanced [Ga(E)] electron Green’s functions for the double-dot within the Coulomb blockade regime where charging energies significantly exceed thermal energies kTα (Tα being the electrodes temperatures and k being the Boltzmann constant) as well as energies characterizing couplings of the dots to electrodes. We use the equations of motion (EOM) method, as described in  Appendix A. The resulting expression for Gr agrees with the corresponding result of Ref. 78. Note that in a nonmagnetic system, Gr,a(E) do not depend on the electron spin orientation provided that electron transport is spin conserving.

In further analysis, we restrict ourselves to the limit of strong intradot Coulomb interactions, assuming that UiU, Ei, β and to moderately biased systems (|eV| < Ui). Under such conditions, each dot in a double-dot system may contain only a single electron that can participate in transport. Energy levels that could receive a second electron become shifted beyond the conduction window and remain empty for electrons tunnel from these levels to the electrodes. Therefore, strong intradot Coulomb interactions do not affect electron transport through the double dot, and only much weaker interactions between electrons on different dots remain relevant. Thus, the double dot may be treated as a system where moderate/weak Coulomb interactions occur, and the expression for Gr(E) for a double-dot symmetrically coupled to electrodes may be reduced to the form

(5)

Here, ij,

(6)

and the parameter Γ describes the coupling of the double-dot to the electrodes in a symmetrically coupled system (ΓL=ΓR=12Γ). Equations (5) and (6) were derived applying the wide band approximation for electrodes and treating Γ as a constant independent of the tunnel energy. Note that dependencies of coupling parameters Γα on the tunnel energy E may bring changes into the Green’s functions energy dependencies and the characteristics of electron transport, but analysis of these changes is beyond the scope of the present work.

Electron occupation numbers on the dots ni are determined by the following equation:

(7)

where the lesser Green’s function G<(E). The approximation for G<(E) is derived in  Appendix B. Under the considered conditions, it may be approximated as G<(E) = i∑αfαGr(E)ΓαGa(E) and fα (α = {L, R}) are Fermi distribution functions for the electrodes with chemical potentials μα and temperatures Tα, and 2 × 2 matrices Γα for a symmetrically coupled double-dot have the form

(8)

We compute ni by self-consistently solving Eq. (7). The occupation numbers are controlled by several factors including the electrodes temperatures and chemical potentials, electron energies on the dots, the interdot charging energy U, coupling strengths Γ and β, and magnitudes and polarities of the gate and bias voltages [the latter originates from the difference in the electrodes chemical potentials in the Fermi functions included in the expression for G<(E)]. Electron occupation on the dots strongly depends on the interplay between interdot electron tunnelings and Coulomb interactions. When electron tunneling between the dots is weak, the dots occupations are mostly determined by positions of the on-dot electron energies on the energy scale and are considerably affected by Coulomb interactions between electrons on different dots. For strong interdot tunneling, the effect of Coulomb interactions is much less pronounced.

In further analysis, we focus on the case of weakly coupled double-dots to better understand effects originating from Coulomb interactions.

For the past three decades, a significant effort was applied to study thermoelectric transport in nanoscale systems.19,45–47,49,51,78,79 To study these properties, we assume that the system is biased by both the voltage ΔV = μLμR and the temperature gradient ΔT = TLTR. Within the linear response regime, the charge current I and the energy current J may be presented as80 

(9)

and

(10)

Here, coefficients Ln have the form

(11)

where μ is the chemical potential of electrodes in the absence of bias, f is the Fermi function with the chemical potential μ and temperature T, and τ(E) = Tr[Ga(E)ΓRGr(E)ΓL] is the electron transmission function. When a double-dot in a serial configuration is symmetrically coupled to electrodes, the transmission accepts the form

(12)

Equating the charge current to zero, one finds the expression for thermopower (Seebeck coefficient),

(13)

The efficiency of the energy conversion is characterized by a dimensionless thermoelectric figure of merit ZT, which is commonly defined as

(14)

Thus, ZT is expressed in terms of thermopower, electronic electric conductance G, and thermal conductance κ, which includes both electronic and phonon contributions. In the considered case of a double quantum dot, we can omit the contribution from phonons to the thermal conductance, assuming that electron–phonon interactions are weak because of the above-mentioned mismatch between phonon modes frequencies. Then, the figure of merit may be presented in the form

(15)

Both thermopower and ZT depend on the positions of the energy levels on the dots, on the coupling of the dots to the electrodes, and on electron–electron interactions. In general, these transport characteristics may be strongly affected by quantum interference phenomena, but we do not expect these effects to be manifested in the presently analyzed double-dot in a serial configuration.

Now, we consider electron transport solely driven by a difference between TL and TR assuming a symmetrical distribution of the temperature difference between and electrodes: TL,R=T±12ΔT [T=12(TL+TR)]. The transport occurs when an energy level of a considered system linking the electrodes is situated near their Fermi level μ. If this level is below μ, it works as a transport channel for holes, and if it is situated above μ, charge carriers mainly participating in transport are electrons. The Seebeck coefficient becomes zero when a certain level crosses E = μ, and electron and hole fluxes counterbalance each other. It changes sign when the type of charge carriers (electrons/holes) changes.

In the absence of electron–electron interactions, a double-dot including two single-level dots possesses two energy levels E1 and E2, which differ from original electron energy levels on the decoupled dots E0i due to the interdot coupling. In a weakly coupled double-dot, Ei are close to E0i. Intradot electron–electron interactions split each level in two. In the considered limit of very strong intradot electron–electron interactions (U1,2U), the levels at E = Ei + Ui are shifted beyond the conduction window. Thus, within this limit, actual transport channels may be associated with levels whose energies equal Ei and/or Ei + U. As follows from Eq. (6), two energy levels of a double-dot system affected by interdot Coulomb interactions may be approximated as ẼiEi when both dots are nearly empty and as ẼiEi+U when they are nearly full. When one dot (dot i) is nearly full and another one (dot j) is nearly empty, we obtain ẼiEi and ẼjEj+U.

In Fig. 2, we show dependencies of S and ZT on the gate voltage Vg, which shifts energy levels on the dots and drags them through the vicinity of the electrodes’ Fermi energy where (due to the presence of a thermal gradient) they become transport channels. The figure is plotted assuming that E01 is located on the energy scale above the chemical potential of electrodes μ and E02 is placed below μ. Therefore, we may expect dot 1 to be nearly empty (Ẽ1E1+U) and dot 2 to be nearly full (Ẽ2E2). Accordingly, the thermopower sign changes three times. This happens at Vg ≈ −E2, at Vg ≈ −(E1 + U), and at the point Vg12(E1+E2). At this value of the gate potential, the energy levels Ẽ1 and Ẽ2 are symmetrically arranged with respect to the Fermi level of electrodes and simultaneously serve as transport channels for electrons (E=Ẽ1) and for holes (E=Ẽ2). Both channels equally contribute to the transport, and therefore, electron and hole currents are counterbalanced and S becomes zero.

FIG. 2.

Thermopower S and thermoelectric figure of merit ZT (inset) as functions of gate voltage applied to the double-dot. Curves are plotted assuming kT = 5 meV, Γ = 4 meV, β = 10 meV, E01 = 30 meV, E02 = −10 meV, and μL = μR = μ = 0.

FIG. 2.

Thermopower S and thermoelectric figure of merit ZT (inset) as functions of gate voltage applied to the double-dot. Curves are plotted assuming kT = 5 meV, Γ = 4 meV, β = 10 meV, E01 = 30 meV, E02 = −10 meV, and μL = μR = μ = 0.

Close modal

Two doubled peaks appear at each ZT vs Vg curve (see inset) at VgẼi in agreement with the thermopower behavior. Peak values of ZT are considerably greater than 1, indicating a potential efficiency of the considered system as a heat-to-electric energy convertor. Note that the peaks heights do not depend on the value of charging constant U. The strength of interdot Coulomb interactions determines a position of the peak emerging at Vg ≈ −(E1 + U) but not its magnitude.

When the temperature difference across the system increases, it may switch to a nonlinear regime of operation. Then, a linear relationship between ΔT and thermally induced voltage is violated and thermopower becomes a function of ΔT.9,22,31,34,36,72 Here, we focus on the thermocurrent Ith, which also is an important quantity characterizing thermoelectric charge transport through nanoscale systems. It could be defined as a difference between the charge current driven by both thermal and electric biases and the current solely driven by the bias voltage at ΔT = 0,81 

(16)

where the charge current is given by the following standard Landauer expression:5,73

(17)

According to this definition, the thermocurrent Ith represents a contribution to the tunnel current originating from the thermally excited flow of charge carriers. We first consider Ith in an unbiased double-dot, which is solely driven by a thermal gradient ΔT. The effect of interdot Coulomb interactions is better pronounced when both E1 and E2 are situated below the electrodes Fermi energy μ = 0 and both dots are nearly full. Then, the thermally induced current reaches maxima at UẼi(Ei+U) when corresponding energy levels appear in a close vicinity of E = μ and may serve as transport channels for charge carriers, as shown in Fig. 3. Note that in the considered case, the charge carriers involved in transport are holes, so the current takes on positive values at positive ΔT (that is, when TL > TR). If Ei < μ (dot i is nearly full) and Ej > μ (dot j is nearly empty; ij), a single peak located at U ≈ −Ei should appear at each Ith vs U curve. Finally, if both dots are nearly empty (Ei > μ; i = 1, 2), Ith vs U curves become featureless, as demonstrated in the inset. The thermocurrent magnitude slowly decreases as interdot Coulomb interactions strengthen.

FIG. 3.

Thermally induced current Ith flowing through an unbiased double-dot vs the strength of interdot Coulomb interactions U. Curves are plotted assuming that kT = 5 meV, Γ = 4 meV, β = 10 meV, E01 = −40 meV, E02 = −20 meV, and μ = Vg = 0. Inset: E01 = 40 meV and E02 = 20 meV. Remaining parameters are taken at the same values as in the main figure.

FIG. 3.

Thermally induced current Ith flowing through an unbiased double-dot vs the strength of interdot Coulomb interactions U. Curves are plotted assuming that kT = 5 meV, Γ = 4 meV, β = 10 meV, E01 = −40 meV, E02 = −20 meV, and μ = Vg = 0. Inset: E01 = 40 meV and E02 = 20 meV. Remaining parameters are taken at the same values as in the main figure.

Close modal

When a double-dot is biased by an applied voltage V, the effect of thermal gradient is important for moderately biased systems where the thermal and electric driving forced are comparable. The interplay between these forces may result in peaks and dips at the current–voltage curves, indicating that a certain energy level E=Ẽi crosses the boundaries of the conduction window. The specifics of these features are mostly determined by the levels’ positions, as illustrated in Fig. 4. One may come to this conclusion by comparing the curves displayed in the main body of this figure with those shown in the inset. Note that Ith is only a small part of a total charge current flowing through a system biased by both applied voltage and thermal gradient even for a moderate electric bias. At sufficiently strong bias, the effect of thermal driving forces is negligible and Ith becomes zero.

FIG. 4.

Thermally induced current Ith as a function of the bias voltage V. Curves are plotted assuming that kT = 5 meV, Γ = 4 meV, β = 10 meV, E01 = 20 meV, E02 = −10 meV, μ = Vg = 0, and ΔT/T = 1.2. Inset: E01 = 20 meV and E02 = 10 meV. Remaining parameters are taken at the same values as in the main figure.

FIG. 4.

Thermally induced current Ith as a function of the bias voltage V. Curves are plotted assuming that kT = 5 meV, Γ = 4 meV, β = 10 meV, E01 = 20 meV, E02 = −10 meV, μ = Vg = 0, and ΔT/T = 1.2. Inset: E01 = 20 meV and E02 = 10 meV. Remaining parameters are taken at the same values as in the main figure.

Close modal

In this section, we study heat currents flowing through the system when it is biased by the applied temperature gradient ΔT. Again, we assume a symmetric distribution of the temperature between the electrodes, so T remains constant. We compute the heat current using the general expression82 

(18)

Here, the difference in the chemical potentials of electrodes Δμ = μLμR equals eΔVth where the potential difference ΔVth originates from the Seebeck effect and could be found by assuming that I = 0 (the open circuit condition) and then solving Eq. (9) for a given temperature difference ΔT. As before, we consider a double-dot in a serial geometry, and the electron transport occurs within the Coulomb blockade regime. Using the expression Eq. (18), we omit from consideration phonon mechanisms of energy transfer and concentrate on the electron contribution to the heat current.

In the same way as the thermally induced charge current Ith, the heat current J at a fixed ΔT is mostly controlled by positions of the double-dot energy levels Ẽi on the energy scale. These, in turn, are controlled by the strength of interdot Coulomb interactions. This is illustrated in Fig. 5. As shown in this figure, interdot Coulomb interactions noticeably suppress the heat transfer through the considered system and may bring asymmetry into Ith vs ΔT curves, which indicates heat current rectification. The asymmetry becomes stronger as U increases.

FIG. 5.

Heat current J as a function of the temperature gradient ΔT plotted at kT = 5 meV, β = 5 meV, Γ = 4 meV, E01 = −6 meV, and E02 = 4 meV for several values of the charging energy U. Inset: J as a function of E1 plotted at ΔT/T = 1.6. Remaining parameters take on the same values as in the main figure.

FIG. 5.

Heat current J as a function of the temperature gradient ΔT plotted at kT = 5 meV, β = 5 meV, Γ = 4 meV, E01 = −6 meV, and E02 = 4 meV for several values of the charging energy U. Inset: J as a function of E1 plotted at ΔT/T = 1.6. Remaining parameters take on the same values as in the main figure.

Close modal

To better understand a mechanism controlling the heat current rectification in the considered system, we need to more thoroughly analyze the specifics of electron transport. At TL = TR and μ = 0 and for chosen values of E0i (E01 = −6 meV, E02 = 4 meV), dot 1 is probably nearly full (Ẽ1E1<0) and dot 2 is nearly empty [Ẽ2(E2+U)>0]. Suppose that the thermal gradient drives charge carriers to the right (TL > TR, that is, ΔT > 0). Then, the energy level E=Ẽ2 does not participate in transport, whereas the level E=Ẽ1 being close enough to the electrodes Fermi level serves as a transport channel for holes. When the thermal gradient is reversed (ΔT < 0), dot 2 may become occupied. In this case, both energy levels E=Ẽi(Ei+U) are situated beyond the conduction window created by the thermal gradient, and electron transport between the electrodes ceases to exist.

It was suggested in some earlier works83–86 that nanoscale rectifiers may operate as such due to broken mirror symmetry in a nanoscale junction whose terminals are linked by a chain of elements with different spectral densities. In the considered system, we have a chain-like linker including two elements, and a difference between the spectral densities originates from interdot Coulomb interactions and from difference in electron energies on the decoupled elements (dots), which give rise to different spectral densities associated with the dots in a weakly coupled double-dot.

The rectification ratio for heat currents RJ may be defined as

(19)

where J± are the heat currents corresponding to the forward (ΔT > 0) and reversed (ΔT < 0) bias created by the thermal gradient. The rectification becomes more pronounced at strong interactions between electrons on the dots. Also, RJ and that depends on the on-dot electron energies, as follows from the above discussion. Assuming that the dot attached to the hot electrode is occupied, RJ is determined by the occupation on the dot attached to the cool electrode. When this dot is also occupied, interdot Coulomb interactions affect the electron transport, and their effect vanishes when this dot remains empty. The dependence of the rectification ratio on the electron energies on the dots is displayed in Fig. 6. As shown in this figure, at ΔT > 0, a pronounced rectification appears at certain values of the on-dot energies and sufficiently strong Coulomb repulsion between electrons on different dots. The rectification ratio may reach values of the order of 10, thus satisfying the least rigorous condition for diod operations. Therefore, the considered system may operate as a heat diod.

FIG. 6.

Heat current rectification ratio RJ for the considered double-dot system as a function of respective positions of the electron energy levels on the dots at ΔT/T = 1.4, kT = 5 meV, β = 5 meV, Γ = 4 meV, and U = 50 meV.

FIG. 6.

Heat current rectification ratio RJ for the considered double-dot system as a function of respective positions of the electron energy levels on the dots at ΔT/T = 1.4, kT = 5 meV, β = 5 meV, Γ = 4 meV, and U = 50 meV.

Close modal

In the present work, we theoretically analyzed thermoelectric transport through a double-dot in a serial configuration placed between electrodes. Such systems are being manufactured and studied both theoretically and experimentally. We analyzed electron transport through such a junction within a Coulomb blockade regime when the coupling of the dots to the electrodes was weaker than the characteristic thermal energy kT associated with the system. Effects of electron–phonon interactions on the electron transport were omitted based on the assumption that phonon contributions to both electrical and thermal conductance of the system were small and coherent electron transport predominated. Transport characteristics were computed using the Green’s functions formalism within the limit of strong intradot Coulomb interactions where intradot charging energies Ui greatly exceed interdot charging energy U.

It was shown that interdot Coulomb interactions may significantly affect the thermopower of the considered system as well as thermally induced charge current flowing through the latter. The effect of these interactions strongly depends on electron energies on the dots and interdot coupling strength, which affect electron occupation numbers on the dots. Note that the thermoelectric figure of merit in the considered system may reach values noticeably greater than 1, thus indicating that it may serve as a heat-to-electric energy converter.

Also, we showed that electron–electron interactions between the dots may result in rectification of heat currents flowing through the considered system. By varying occupation numbers on the dots by means of the applied bias and gate voltages and/or thermal gradient, one may control the rectification ratio. A model of a heat diod based on a double-dot in a parallel geometry and operating due to interdot Coulomb interactions was suggested in some earlier works.51,52,87 Here, we demonstrate that a satisfactory operating heat rectifier may be manufactured basing on a double-dot in a serial configuration. For an appropriate choice of electron energies on the dots and sufficiently strong interdot Coulomb interactions, the rectification ratio may reach values of the order of 10.

We believe that the presented results provide a step toward further understanding and modeling of thermoelectric and heat transport through quantum dots and molecules and could be important for manufacturing nanoscale heat rectifiers.

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

This work was supported by NSF-DMR-PREM 1523463. The author thanks G. M. Zimbovsky for help with the manuscript preparation

Within the Coulomb blockade regime, the couplings between the dots and the electrodes are supposed to be weak, and the EOMs for Gji,σr(E)dj,σ;di,σ have the form

(A1)

and

(A2)

where {i, j} = {1, 2}, (ij). Assuming that dot 1 is coupled to the left electrode and dot 2 to the right one, the parameters Γi,σ are further labeled ΓσL,R and are given by

(A3)

Within a wide band approximation, parameters ΓσL,R may be treated as constants independent of energy E. Assuming that the double dot is symmetrically coupled to the electrodes and coupling strengths do not depend on the electron spin orientation, then ΓσL=ΓσR=12Γ. Also, we assume that intradot charging energies are equal: U1=U2=Ũ.

Two-particle Green’s functions obey EOMs,

(A4)
(A5)

A simplified form for the term βdj,σ;di,σ is obtained using the following approximation to decouple relevant four-operator averages,82 

(A6)

Three-particle Green’s functions that appear in Eqs. (A4) and (A5) are determined by EOMs,

(A7)
(A8)

Here, Ni=di,σdi,σdi,σdi,σ are two-particle occupation numbers and G(4)r(E) is a four-particle Green’s function approximated as

(A9)

In a nonmagnetic system, ni,σ=ni,σ=12ni, and Green’s functions are independent of electron spin orientations provided that spin-flip processes are neglected. Using Eqs. (A4)–(A9), Eq. (A1) may be transformed to the form

(A10)

where

(A11)

In these expressions, P1=1nj+Nj, P2=nj2Nj, P3=Nj, Π1 = 0, Π2 = U, Π3 = 2U, and the index j satisfies the conditions j = {1, 2} and ji.

We consider the limit of strong intradot Coulomb interactions when ŨU,E0i and a moderately biased system (|eV|<Ũ), and therefore, we may expect that the double dot energy levels shifted in the energy scale due to intradot interactions remain beyond the conduction window. Electrons from these elevated levels tunnel to the electrodes leaving them empty, so it is highly likely that each single-level dot contains no more than a single electron and two-particles occupation numbers may be set to zero. Also, all terms in Eq. (A11) including Ũ in their denominators remain small inside the conduction window and do not contribute to the electron transmission. Therefore, one can simplify the expressions for υi by omitting the terms proportional to Ni as well as those including the intradot charging energies. Then, one arrives at the following approximation for υi:

(A12)

We may transform Eq. (A2) in a similar way and get

(A13)

Solving Eqs. (A10) and (A13), we obtain

(A14)

where again i, j = 1, 2 and ij.

Expressions for G<(E) and G>(E) may be derived using the EOM technique as well as the expression for Gr(E).78,88,89 The results are

(B1)
(B2)

Here, I is the identity matrix and

(B3)

where fα(E) and α = {L, R} are Fermi distribution functions for the electrodes. These results satisfy the identity G<(E) − G>(E) = Gr(E) − Ga(E).

As mentioned in the text, at UiU, |e|V, Ei, β, the electron transport is not affected by strong intradot Coulomb interactions because each dot may contain a single electron. Therefore, within the Coulomb blockade regime, we may treat the considered double dot as a system where only moderate/weak interactions occur between the dots and the electrodes and between the dots themselves. Thus, we may use the relationship82,89

(B4)

For a serial arrangement of the dots and within a wide band approximation, the matrices Γα have the form

(B5)

Note that beyond the Coulomb blockade regime and at sufficiently strong interdot Coulomb interactions, matrix elements Γα should not be approximated by constants. They become dependent on both tunnel energy E and charging energy U.89 

Substituting Eqs. (B4) and (B5) into Eq. (B1), we arrive at the Keldysh equation,

(B6)

where

(B7)

This result shows that in the case considered in the present work, the EOM technique leads to the same approximation for the lesser Green’s function as the diagrammatic method.

Charge current flowing through the double dot is given by a general expression,

(B8)

Using Eqs. (B4), (B6), and (B7), one may reduce this expression to the simple formula used in the text [see Eq. (17)].

1.
N.
Agraït
,
A. L.
Yeati
, and
J. M.
Van Ruitenbeck
,
Phys. Rep.
377
,
81
273
(
2003
).
2.
F.
Giazotto
,
T. T.
Heikkilä
,
A.
Luukanen
,
A. M.
Savin
, and
J. P.
Pekola
,
Rev. Mod. Phys.
78
,
217
(
2006
).
3.
Y.
Dubi
and
M.
Di Ventra
,
Rev. Mod. Phys.
83
,
131
(
2011
).
4.
J. C.
Cuevas
and
E.
Scheer
,
Molecular Electronics: An Introduction to Theory and Experiment
(
World Scientific
,
Singapore
,
2010
).
5.
M.
Galperin
,
M. A.
Ratner
, and
A.
Nitzan
,
J. Phys.: Condens. Matter
19
,
103201
(
2007
).
6.
N. A.
Zimbovskaya
and
M. R.
Pederson
,
Phys. Rep.
509
,
1
89
(
2011
).
7.
F.
Pauli
,
J. K.
Viljas
, and
J. C.
Cuevas
,
Phys. Rev. B
78
,
035315
(
2008
).
8.
C. M.
Finch
,
V. M.
Garcia-Suarez
, and
C. J.
Lambert
,
Phys. Rev. B
79
,
033405
(
2009
).
9.
P.
Reddy
,
S.-Y.
Jang
,
R. A.
Segalman
, and
A.
Majumdar
,
Science
315
,
1568
(
2007
).
10.
J. A.
Malen
,
P.
Doak
,
K.
Baheti
,
T. D.
Tilley
,
R. A.
Segalman
, and
A.
Majumdar
,
Nano Lett.
9
,
1164
(
2009
).
11.
A.
Tan
,
J.
Balachandran
,
S.
Sadat
,
V.
Gavini
,
B. D.
Dunietz
,
S.-Y.
Jang
, and
P.
Reddy
,
J. Am. Chem. Soc.
133
,
8838
(
2011
).
12.
J. R.
Wídawski
,
W.
Chen
,
H.
Vásquez
,
T.
Kim
,
R.
Breslow
,
M. S.
Hybertsen
, and
L.
Venkataraman
,
Nano Lett.
13
,
2889
(
2013
).
13.
Z.
Gonsanamlou
,
S. T.
Vishkayi
,
M. B.
Tagani
, and
H. R.
Soleimani
,
Chem. Phys. Lett.
594
,
51
(
2014
).
14.
R.-X.
Li
,
Y.
Ni
,
H.-D.
Li
,
X.-L.
Tian
,
K.-L.
Yao
, and
H.-H.
Fu
,
Physica B
493
,
1
(
2016
).
15.
M.
Tsutsui
,
K.
Yokota
,
T.
Morikawa
, and
M.
Taniguchi
,
Sci. Rep.
7
,
44276
(
2017
).
16.
C. W. J.
Beenakker
and
A. A. M.
Staring
,
Phys. Rev. B
46
,
9667
(
1992
).
17.
J.
Koch
,
F.
von Oppen
,
Y.
Oreg
, and
E.
Sela
,
Phys. Rev. B
79
,
195107
(
2004
).
18.
19.
N. A.
Zimbovskaya
,
J. Phys.: Condens. Matter
28
,
183002
(
2016
).
20.
B.
Kubala
,
J.
König
, and
J.
Pekola
,
Phys. Rev. Lett.
100
,
066801
(
2008
).
21.
P. G.
Murphy
,
S.
Mukerjee
, and
J. E.
Moore
,
Phys. Rev. B
79
,
161406(R)
(
2008
).
22.
M.
Wíerzbicki
and
R.
Swírkovicz
,
Phys. Rev. B
82
,
165334
(
2010
).
23.
P.
Trócha
and
J.
Barnás
,
Phys. Rev. B
85
,
085408
(
2012
).
24.
M.
Yoshida
and
L. N.
Oliveira
,
Physica B
404
,
3312
(
2009
).
25.
D.
Nozaki
,
S. M.
Avdoshenko
,
H.
Sevinçli
, and
G.
Cuniberti
,
J. Appl. Phys.
116
,
074308
(
2014
).
26.
M. A.
Siérra
,
M.
Sáiz-Bretín
,
F.
Domínguez-Adame
, and
D.
Sánchez
,
Phys. Rev. B
93
,
235452
(
2016
).
27.
J. P.
Biergfield
and
C. A.
Stafford
,
Nano Lett.
9
,
3072
(
2009
).
28.
M.
Sáiz-Bretín
,
A. V.
Malyshev
,
P. A.
Orellana
, and
F.
Domínguez-Adame
,
Phys. Rev. B
91
,
085431
(
2015
).
29.
H. M.
Pastawski
,
L. E. F.
Foa Torres
, and
E.
Medina
,
Chem. Phys.
281
,
257
(
2002
).
30.
A.
Nitzan
,
Chemical Dynamics in Condensed Phases
(
Oxford University Press
,
Oxford
,
2006
).
31.
M.
Leijinse
,
M. R.
Wegewijs
, and
K.
Flensberg
,
Phys. Rev. B
82
,
045412
(
2010
).
32.
M.
Galperin
,
A.
Nitzan
, and
M. A.
Ratner
,
Mol. Phys.
106
,
397
(
2008
).
33.
R.
Hartle
and
M.
Thoss
,
Phys. Rev. B
83
,
115414
(
2011
).
34.
J.
Argüello-Luéngo
,
D.
Sánchez
, and
R.
López
,
Phys. Rev. B
91
,
165431
(
2015
).
35.
G. T.
Craven
and
A.
Nitzan
,
J. Chem. Phys.
146
,
092305
(
2017
).
36.
T.
Koch
,
J.
Loos
, and
H.
Fleshke
,
Phys. Rev. B
89
,
155133
(
2014
).
37.
A.
Erpenbeck
,
R.
Hartle
, and
M.
Thoss
,
Phys. Rev. B
91
,
195418
(
2015
).
38.
J. K.
Sowa
,
J. A.
Mol
, and
E. M.
Gauger
,
J. Phys. Chem. C
123
,
4103
(
2019
).
39.
J. K.
Sowa
,
J. A.
Mol
,
G. A. D.
Briggs
, and
E. M.
Gauger
,
Phys. Rev. B
95
,
085423
(
2017
).
40.
N. A.
Zimbovskaya
and
A.
Nitzan
,
J. Phys. Chem. B
124
,
2632
(
2020
).
41.
Y.
Dubi
and
M.
Di Ventra
,
Phys. Rev. B
79
,
081302
(
2009
).
42.
S.
Jezouin
,
F. D.
Parmentier
,
A.
Anthore
,
U.
Gennser
,
A.
Cavanna
,
Y.
Jin
, and
F.
Pierre
,
Science
342
,
601
(
2013
).
43.
C.
Riha
,
P.
Miechowski
,
S. S.
Buchholz
,
O.
Chiatti
,
A. D.
Wieck
,
D.
Reuter
, and
S. F.
Fischer
,
Appl. Phys. Lett.
106
,
083102
(
2015
).
44.
L.
Cui
,
W.
Jeong
,
S.
Hur
,
M.
Matt
,
J. C.
Klöckner
,
F.
Pauly
,
P.
Nielaba
,
J. C.
Cuevas
,
E.
Meyhofer
, and
P.
Reddy
,
Science
355
,
1192
(
2017
).
45.
B.
Sothmann
,
R.
Sánchez
, and
A. N.
Jordan
,
Nanotechnology
26
,
032001
(
2015
).
46.
H.
Saleghi
,
S.
Sangtarash
, and
C. J.
Lambert
,
Nano Lett.
15
,
7467
(
2015
).
47.
M.
Noori
,
H.
Sadeghi
, and
C. J.
Lambert
,
Nanoscale
9
,
5299
(
2017
).
48.
C. A.
Perroni
,
D.
Ninno
, and
V.
Cataudella
,
J. Phys.: Condens. Matter
28
,
373001
(
2016
).
49.
J.
Lukkebo
,
G.
Romano
 et al,
J. Chem. Phys.
144
,
114310
(
2016
).
50.
M.
Kulguir
and
D.
Segal
,
Phys. Rev. E
98
,
012114
(
2018
).
51.
D. M.-T.
Kuo
and
Y.-C.
Chang
,
Phys. Rev. B
81
,
205301
(
2010
).
52.
A.
Marcos-Vicioso
,
C.
López-Jurádo
,
M.
Rúiz-García
, and
R.
Sánchez
,
Phys. Rev. B
98
,
035414
(
2018
).
53.
D.
Segal
and
A.
Nitzan
,
Phys. Rev. Lett.
94
,
034301
(
2004
).
54.
L.-A.
Wu
,
C.-X.
Yu
, and
D.
Segal
,
Phys. Rev. E
80
,
041103
(
2009
).
55.
56.
G. T.
Craven
,
D.
He
, and
A.
Nitzan
,
Phys. Rev. Lett.
121
,
247704
(
2018
).
57.
D.
Bera
,
L.
Qian
,
T.-k.
Tseng
, and
P. H.
Holloway
,
Materials
3
,
2260
2345
(
2010
).
58.
H.
Choi
,
C.
Nahm
,
J.
Kim
,
C.
Kim
,
S.
Kang
,
T.
Hwang
, and
B.
Park
,
Curr. Appl. Phys.
13
,
S2
S13
(
2013
).
59.
H.
Imamura
,
P. A.
Maksym
, and
H.
Aoki
,
Phys. Rev. B
59
,
5817
5826
(
1999
).
60.
X.
Chen
,
H.
Buhmann
, and
L. W.
Molenkamp
,
Phys. Rev. B
61
,
16801
(
2000
).
61.
E.
Paspalakis
,
Z.
Kis
,
E.
Voitsinas
, and
A. F.
Terzis
,
Phys. Rev. B
69
,
155316
(
2004
).
62.
C. D.
Simserides
,
A.
Zora
, and
G.
Tribertis
,
Phys. Rev. B
73
,
155313
(
2006
).
63.
C. D.
Simserides
,
U.
Hohenester
,
G.
Goldoni
, and
E.
Molinari
,
Phys. Rev. B
62
,
13657
(
2000
).
64.
A. W.
Holleitner
,
C. R.
Decker
,
H.
Kin
,
K.
Eberl
, and
R. H.
Blick
,
Phys. Rev. Lett.
87
,
256802
(
2001
).
65.
A. W.
Holleitner
,
R. H.
Blick
,
A. K.
Huttel
,
K.
Eberl
, and
J. P.
Kotthaus
,
Science
297
,
70
74
(
2002
).
66.
L.
Yan
,
L.
Wang
,
A. R.
Garrigues
,
L.
Jiang
,
H. V.
Annadata
,
M. A.
Antonana
,
E.
Barko
, and
C. A.
Nijhuis
,
Nat. Nanotechnol.
13
,
322
334
(
2018
).
67.
Y.
Tanaka
and
N.
Kawakami
,
Phys. Rev. B
72
,
085304
(
2005
).
68.
S.
Grap
,
S.
Andergassen
,
J.
Paaske
, and
V.
Meden
,
Phys. Rev. B
83
,
115115
(
2011
).
69.
R.
Zitko
,
J.-S.
Lim
,
R.
Lópes
,
J.
Martinek
, and
P.
Simon
,
Phys. Rev. Lett.
108
,
166605
(
2012
).
70.
Y.
Wang
,
J.
Wei
, and
Y.
Yan
,
Chem. Phys.
152
,
164113
(
2020
).
71.
P.
Trócha
,
I.
Weymann
, and
J.
Barnás
,
Phys. Rev. B
80
,
165333
(
2009
).
72.
J. C.
Estrada-Saldana
,
A.
Verkis
,
G.
Steffensen
,
R.
Zitko
,
P.
Krogstrup
,
J.
Paaske
,
K.
Grove-Rasmussen
, and
J.
Nygard
,
Phys. Rev. Lett.
121
,
257701
(
2018
).
73.
R.-P.
Riwar
and
S.
Splettstoesser
,
Phys. Rev. B
82
,
205308
(
2010
).
74.
D. M.-T.
Kuo
,
Jpn. J. Appl. Phys., Part 1
50
,
025003
(
2011
).
75.
E. P.
Danieli
,
G. A.
Álvarez
,
P. R.
Levstein
, and
H. M.
Pastawski
,
Sold State Commun.
141
,
422
(
2007
).
76.
M.
Tsaousidou
and
G. P.
Triberis
,
J. Phys.: Condens. Matter
22
,
355304
(
2010
).
77.
J. P.
Bergfield
,
M. A.
Solis
, and
C. A.
Stafford
,
ACS Nano
4
,
5314
(
2010
).
78.
Y.-C.
Chang
and
D. M.-T.
Kuo
,
Phys. Rev. B
77
,
245412
(
2008
).
79.
D.
Sánchez
and
L.
Sérra
,
Phys. Rev. B
84
,
201307(R)
(
2011
).
80.
G. D.
Mahan
,
Many-Particle Physics
(
Plenum
,
New York
,
2000
).
81.
S. F.
Svensson
,
E. A.
Hoffmann
,
N.
Nakpathomkun
,
P. M.
Wu
,
H. Q.
Xu
,
H. A.
Nilsson
,
D.
Sánchez
,
V.
Kashcheyevs
, and
H.
Linke
,
New J. Phys.
15
,
105011
(
2013
).
82.
S.
Datta
,
Quantum Transport: From Atom to Transistor
(
Cambridge University Press
,
Cambridge, UK
,
2005
).
83.
T.
Ruokola
,
T.
Ojanen
, and
A.-P.
Jauho
,
Phys. Rev. B
79
,
144306
(
2009
).
84.
M. J.
Martínez-Pérez
,
A.
Fornieri
, and
F.
Giazotto
,
Nat. Nanotechnol.
10
,
303
(
2015
).
85.
J.
Ren
and
J.-X.
Zhu
,
Phys. Rev. B
87
,
241412
(
2013
).
86.
G.
Schaller
,
G. G.
Giusteri
, and
G. L.
Gelardo
,
Phys. Rev. E
94
,
032135
(
2016
).
87.
G.
Roselló
,
R.
López
, and
R.
Sánchez
,
Phys. Rev. B
95
,
235404
(
2017
).
88.
C.
Niu
,
D. L.
Lin
, and
T.-H.
Lin
,
J. Phys.: Condens. Matter
11
,
1511
(
1999
).
89.
R.
Swirkowics
,
J.
Barnás
, and
M.
Wilczynski
,
Phys. Rev. B
68
,
195318
(
2003
).