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

Role of bias and tunneling asymmetries in nonlinear Fermi-liquid transport
through an SU(NN) quantum dot

Kazuhiko Tsutsumi Department of Physics, Osaka City University, Sumiyoshi-ku, Osaka 558-8585, Japan NITEP, Osaka Metropolitan University , Sumiyoshi-ku, Osaka 558-8585, Japan    Yoshimichi Teratani Department of Physics, Osaka City University, Sumiyoshi-ku, Osaka 558-8585, Japan NITEP, Osaka Metropolitan University , Sumiyoshi-ku, Osaka 558-8585, Japan    Kaiji Motoyama Department of Physics, Osaka City University, Sumiyoshi-ku, Osaka 558-8585, Japan    Rui Sakano Department of Physics, Keio University, 3-14-1 Hiyoshi, Kohoku-ku, Yokohama, Kanagawa 223-8522, Japan    Akira Oguri Department of Physics, Osaka City University, Sumiyoshi-ku, Osaka 558-8585, Japan NITEP, Osaka Metropolitan University , Sumiyoshi-ku, Osaka 558-8585, Japan
Abstract

We study how bias and tunneling asymmetries affect nonlinear current through a quantum dot with NN discrete levels in the Fermi liquid regime, using an exact low-energy expansion of the current derived up to terms of order V3V^{3} with respect to the bias voltage. The expansion coefficients are described in terms of the phase shift, the linear susceptibilities, and the three-body correlation functions, defined with respect to the equilibrium ground state of the Anderson impurity model. In particular, the three-body correlations play an essential role in the order V3V^{3} term, and their coupling to the nonlinear current depends crucially on the bias and tunnel asymmetries. The number of independent components of the three-body correlation functions increases with NN the internal degrees of the quantum dots, and it gives a variety in the low-energy transport. We calculate the correlation functions over a wide range of electron fillings of the Anderson impurity model with the SU(NN) internal symmetry, using the numerical renormalization group. We find that the order V3V^{3} nonlinear current through the SU(NN) Kondo state, which occurs at electron fillings of 11 and N1N-1 for strong Coulomb interactions, significantly varies with the three-body contributions as tunnel asymmetries increase. Furthermore, in the valence fluctuation regime toward the empty or fully occupied impurity state, a sharp peak emerges in the coefficient of V3V^{3} current in the case at which bias and tunneling asymmetries cooperatively enhance the charge transfer from one of the electrodes.

I Introduction

The ground state and low-lying excited states of quantum-impurity systems [1, 2] such as dilute magnetic alloys and quantum dots can be described as a local Fermi liquid [3, 4, 5, 6, 7], in which localized electrons with discrete energies are strongly coupled to conductions electrons in host metals or electrodes. These low-energy states continuously evolve as the occupation number of the discretized states varies, and various interesting phenomena such as the Kondo effects and the valence fluctuations occur, depending on electron fillings and the configurations [8, 9, 10].

After early observations of the Kondo effect in quantum dots [11, 12, 13, 14, 15, 16, 17], universal Fermi-liquid behaviors were explored through highly sensitive measurements [18, 19, 20, 21, 22, 23] and precise calculations [24, 25, 26, 27]. Furthermore, various kinds of internal degrees of freedom bring an interesting variety to the Kondo effects in quantum dots, such as the one with the SU(44) symmetry that can be realized in multiorbital dots and in carbon-nanotube (CNT) dots [28, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 44]. External magnetic fields also induce interesting crossover phenomena, such as the one between the SU(4) and SU(2) Kondo states observed in a CNT quantum dot [45, 46, 47, 48].

Recent development in the Fermi-liquid theory also reveals that the three-body correlations of electrons passing through the quantum dot play an essential role in the next-leading order terms of the transport coefficients [49, 50, 51, 52, 53] when the system does not have both electron-hole and time-reversal symmetries. This is because, in addition to the well-studied damping of order ω2\omega^{2}, T2T^{2}, and (eV)2(eV)^{2}, quasiparticles of the local Fermi liquid capture the energy shift of the same quadratic order which is induced by the three-body correlations, at low but finite frequencies ω\omega, temperatures TT, and bias voltages eVeV. The three-body contributions have also been confirmed experimentally in magnetoconductance and nonlinear thermocurrent spectroscopy measurements very recently [54, 55].

Here we focus on the effects of bias and tunneling asymmetries on the nonlinear current II at low energies. Specifically, we examine the bias asymmetry that can be described by the parameter (μL+μR)/2EF(\mu_{L}+\mu_{R})/2-E_{F}, with μL\mu_{L} and μR\mu_{R} the chemical potentials of the source (LL) and drain (RR) electrodes, applied such that eVμLμReV\equiv\mu_{L}-\mu_{R}, and EFE_{F} the Fermi level at thermal equilibrium eV=0eV=0. The other one, tunneling asymmetry, occurs through the difference between the tunnel couplings ΓL\Gamma_{L} and ΓR\Gamma_{R} for the source and drain electrodes, respectively. In the previous paper, we demonstrated how these asymmetries affect the transport through a quantum dot described by a spin-1/2 Anderson impurity model with no orbital degrees of freedom [56], and showed that the three-body correlations give significant contributions to the order (eV)3(eV)^{3} nonlinear current in the valence fluctuation regime, using the numerical renormalization group (NRG) approach.

The purpose of this paper is to clarify how the conjunction of these asymmetries and the internal degrees of freedom, which increases the independent components of the three-body correlation functions, affects low-energy transport in the Fermi-liquid regime. The low-bias expansion of the conductance up to terms order (eV)2(eV)^{2} can be described by the Fermi-liquid theory with the three-body correlations, which at zero temperature T=0T=0 takes the following form for a quantum dot with NN discrete levels which include the spin components,

dIdV=g0σ=1N[sin2δσ+cV,σ(2)eVcV,σ(3)(eV)2+].\displaystyle\frac{dI}{dV}\,=\,g_{0}\sum_{\sigma=1}^{N}\left[\,\sin^{2}\delta_{\sigma}+c_{V,\sigma}^{(2)}\,eV-c_{V,\sigma}^{(3)}\,(eV)^{2}\,+\,\cdots\right]. (1)

Here, g0=e2h 4ΓLΓR/(ΓL+ΓR)2g_{0}=\frac{e^{2}}{h}\,4\Gamma_{L}\Gamma_{R}/(\Gamma_{L}+\Gamma_{R})^{2}, which depends on tunneling asymmetries. We present the exact formulas for the coefficients cV,σ(2)c_{V,\sigma}^{(2)} and cV,σ(3)c_{V,\sigma}^{(3)} of the multilevel Anderson impurity model which are applicable to arbitrary impurity electron fillings and arbitrary level structures ϵdσ\epsilon_{d\sigma}. These coefficients are determined by the phase shift δσ\delta_{\sigma} and the other renormalized parameters, including the three-body correlation functions, and depend crucially on the bias and tunneling asymmetries. We show that, for N3N\geq 3, the three-body correlations between electrons in three different levels also couple to order (eV)3(eV)^{3} nonlinear current as well as the other components when there are some extents of bias and/or tunneling asymmetries. Our formula includes the previous results of the other group as some special limiting cases: cV,σ(2)c_{V,\sigma}^{(2)} derived by Aligia for the Anderson model [57, 58], and cV,σ(3)c_{V,\sigma}^{(3)} derived by Mora et al.  for the SU(NN) Kondo model [40].

We also calculate cV,σ(2)c_{V,\,\sigma}^{(2)} and cV,σ(3)c_{V,\,\sigma}^{(3)} for the SU(NN) symmetric quantum dots with N=4N=4 and 66, using the NRG in a wide range of electron fillings, varying the parameter corresponding to the gate voltage. There emerge (N1)(N-1) different Kondo states in the SU(NN) symmetric case, which can be classified according to the occupation number nd=1\langle n_{d}\rangle=1, 2,2, \ldots, N1N-1. We find that large Coulomb interaction suppresses charge fluctuations throughout the region of 1ndN11\lesssim\langle n_{d}\rangle\lesssim N-1, and it makes the coefficients cV,σ(2)c_{V,\sigma}^{(2)} and cV,σ(3)c_{V,\sigma}^{(3)} less sensitive to the bias asymmetry, whereas the tunneling asymmetry affects these coefficients. In particular, in the SU(NN) Kondo states at electron fillings of nd1\langle n_{d}\rangle\simeq 1 and N1N-1 for N3N\geq 3, the three-body contributions become sensitive to tunnel asymmetries, and it significantly varies the behavior of the order (eV)3(eV)^{3} nonlinear current.

In contrast, in the valence fluctuation regime toward the empty or fully occupied impurity state, i.e., at 0nd10\lesssim\langle n_{d}\rangle\lesssim 1 or N1ndNN-1\lesssim\langle n_{d}\rangle\lesssim N, both the bias and tunneling asymmetries affect the nonlinear transport. We find that, in these regions, a sharp peak emerges in the coefficient cV,σ(3)c_{V,\sigma}^{(3)}, in the case at which the bias and tunneling asymmetries cooperatively enhance the charge transfer from one of the electrodes.

This paper is organized as follows. In Sec. II, we describe an outline of the microscopic Fermi-liquid theory for the multilevel Anderson model, and derive the low-energy asymptotic form of the differential conductance. Section III shows the NRG results of quasiparticle parameters for SU(44) and SU(66) quantum dots. In Secs.  IV and V, we show the NRG results for the coefficients cV,σ(2)c_{V,\sigma}^{(2)} and cV,σ(3)c_{V,\sigma}^{(3)}, respectively. Summary is given in Sec. VI.

II Formulation

We consider a quantum dot with NN-level coupled to two noninteracting leads by using the Anderson Hamiltonian:

H=\displaystyle H\,= Hd+Hc+HT\displaystyle\,H_{d}\,+\,H_{c}\,+\,H_{T}\, (2)
Hd=\displaystyle H_{d}\,= σ=1Nϵdσndσ+U2σσndσndσ,\displaystyle\ \sum_{\sigma=1}^{N}\epsilon_{d\sigma}n_{d\sigma}+\frac{U}{2}\sum_{\sigma\neq\sigma^{\prime}}\,n_{d\sigma}n_{d\sigma^{\prime}}\,,\qquad (3)
Hc=\displaystyle H_{c}\,= ν=L,Rσ=1NDD𝑑ϵϵcϵνσcϵνσ,\displaystyle\sum_{\nu=L,\,R}\,\sum_{\sigma=1}^{N}\int_{-D}^{D}d\epsilon\,\,\epsilon\,c_{\epsilon\nu\sigma}^{\dagger}c_{\epsilon\nu\sigma}^{\,}\,, (4)
HT=\displaystyle H_{T}\,= ν=L,Rσ=1Nvν(ψν,σdσ+dσψν,σ),\displaystyle\ \sum_{\nu=L,\,R}\,\sum_{\sigma=1}^{N}v_{\nu}\,(\psi_{\nu,\,\sigma}^{\dagger}d_{\sigma}+d_{\sigma}^{\dagger}\,\psi_{\nu,\,\sigma}^{\,})\,, (5)
ψν,σ\displaystyle\psi_{\nu,\,\sigma}\equiv DD𝑑ϵρccϵνσ.\displaystyle\int_{-D}^{D}d\epsilon\,\sqrt{\rho_{c}}\,c_{\epsilon\nu\sigma}\,. (6)

Here, dσd_{\sigma}^{\dagger} for σ=1,2,,N\sigma=1,2,\ldots,N creates an impurity electron with energy ϵdσ\epsilon_{d\sigma}, ndσdσdσn_{d\sigma}\equiv d_{\sigma}^{\dagger}d_{\sigma}, and UU is the Coulomb interaction between electrons in the quantum dot. cϵνσc_{\epsilon\nu\sigma}^{\dagger} creates an electron with energy ϵ\epsilon in the lead on the left or right (ν=L,R\nu=L,R), and it is normalized as {cϵνσ,cϵνσ}=δ(ϵϵ)δννδσσ\bigl{\{}c_{\epsilon\nu\sigma}\,,\,c_{\epsilon^{\prime}\nu^{\prime}\sigma^{\prime}}^{\dagger}\bigr{\}}=\delta(\epsilon-\epsilon^{\prime})\delta_{\nu\nu^{\prime}}\delta_{\sigma\sigma^{\prime}}. The linear combination of the conduction electron ψν,σ\psi_{\nu,\sigma} couples to the dot via the tunneling matrix element vνv_{\nu}. It determines the resonance width of the impurity level Δ=ΓL+ΓR\Delta=\Gamma_{L}+\Gamma_{R}, with Γνπρcvν2\Gamma_{\nu}\equiv\pi\rho_{c}v_{\nu}^{2} the tunnel energy scale due to the lead ν\nu, and ρc=1/(2D)\rho_{c}=1/(2D) the density of state of the conduction band with a half-width DD. We set kB=1k_{B}=1 throughout this paper, and consider the parameter region where the half bandwidth DD is much greater than the other energy scales, Dmax(U,Δ,|ϵdσ|,T,|eV|)D\gg\max(U,\Delta,|\epsilon_{d\sigma}|,T,|eV|).

II.1 Fermi-liquid parameters

We describe here the definition of the correlation functions that play an essential role in the microscopic Fermi-liquid theory.

The occupation number and the linear susceptibilities of the impurity level can be derived from the free energy:

ndσ=\displaystyle\bigl{\langle}n_{d\sigma}\bigr{\rangle}\,= Ωϵdσ,ΩTlog[TreβH],\displaystyle\ \frac{\partial\Omega}{\partial\epsilon_{d\sigma}}\,,\qquad\quad\Omega\,\equiv\,-T\log\,\bigl{[}\,\mathrm{Tr}\,e^{-\beta H}\,\bigr{]}\,, (7)
χσσ\displaystyle\chi_{\sigma\sigma^{\prime}}\,\equiv 2Ωϵdσϵdσ=0β𝑑τδndσ(τ)δndσ,\displaystyle\ -\frac{\partial^{2}\Omega}{\partial\epsilon_{d\sigma}\partial\epsilon_{d\sigma^{\prime}}}\,=\,\int_{0}^{\beta}\!d\tau\,\bigl{\langle}\delta n_{d\sigma}(\tau)\,\delta n_{d\sigma^{\prime}}\bigr{\rangle}\,, (8)

where δndσ=ndσndσ\delta n_{d\sigma}=n_{d\sigma}-\langle n_{d\sigma}\rangle, and \langle\cdots\rangle represents an equilibrium average, with β=1/T\beta=1/T the inverse temperature.

In addition to linear susceptibilities, the nonlinear ones χσ1σ2σ3[3]\chi_{\sigma_{1}\sigma_{2}\sigma_{3}}^{[3]} also play an important role away from half filling:

χσ1σ2σ3[3]3Ωϵdσ1ϵdσ2ϵdσ3=χσ1σ2ϵdσ3\displaystyle\chi_{\sigma_{1}\sigma_{2}\sigma_{3}}^{[3]}\,\equiv\,-\frac{\partial^{3}\Omega}{\partial\epsilon_{d\sigma_{1}}\partial\epsilon_{d\sigma_{2}}\partial\epsilon_{d\sigma_{3}}}\,=\,\frac{\partial\chi_{\sigma_{1}\sigma_{2}}}{\partial\epsilon_{d\sigma_{3}}}
=\displaystyle= 0β𝑑τ10β𝑑τ2Tτδndσ1(τ1)δndσ2(τ2)δndσ3.\displaystyle-\int_{0}^{\beta}\!d\tau_{1}\!\int_{0}^{\beta}\!d\tau_{2}\,\bigl{\langle}T_{\tau}\delta n_{d\sigma_{1}}(\tau_{1})\,\delta n_{d\sigma_{2}}(\tau_{2})\,\delta n_{d\sigma_{3}}\bigr{\rangle}\,. (9)

Here, TτT_{\tau} is the imaginary-time ordering operator. This correlation function has the permutation symmetry: χσ1σ2σ3[3]=χσ2σ1σ3[3]=χσ3σ2σ1[3]=χσ1σ3σ2[3]=.\chi_{\sigma_{1}\sigma_{2}\sigma_{3}}^{[3]}=\chi_{\sigma_{2}\sigma_{1}\sigma_{3}}^{[3]}=\chi_{\sigma_{3}\sigma_{2}\sigma_{1}}^{[3]}=\chi_{\sigma_{1}\sigma_{3}\sigma_{2}}^{[3]}=\cdots. We will use hereafter the ground-state values for ndσ\langle n_{d\sigma}\rangle, χσσ\chi_{\sigma\sigma^{\prime}} and χσ1σ2σ3[3]\chi_{\sigma_{1}\sigma_{2}\sigma_{3}}^{[3]}, and thus the occupation number can be deduced from the phase shift δσ\delta_{\sigma} through the Friedel sum rule: ndσT0δσ/π\langle n_{d\sigma}\rangle\xrightarrow{\,T\to 0\,}\delta_{\sigma}/\pi [6].

The retarded Green’s function also plays a central role in the microscopic description of the Fermi-liquid transport:

Gσr(ω)=\displaystyle G_{\sigma}^{r}(\omega)= i0𝑑tei(ω+i0+)t{dσ(t),dσ}eV,\displaystyle\ -i\int_{0}^{\infty}dt\,e^{i(\omega+i0^{+})t}\,\Bigl{\langle}\,\Bigl{\{}d_{\sigma}(t),\,d_{\sigma}^{\dagger}\Bigr{\}}\Bigr{\rangle}_{eV}, (10)
Aσ(ω)\displaystyle A_{\sigma}(\omega)\,\equiv 1πImGσr(ω).\displaystyle\ -\frac{1}{\pi}\,\mathrm{Im}\,G_{\sigma}^{r}(\omega)\;. (11)

represents a nonequilibrium steady-state average taken with the statistical density matrix, which is constructed at finite bias voltages eVeV and temperatures TT, using the Keldysh formalism [59, 60].

The phase shift δσ\delta_{\sigma} is related to the value of the Green’s function at the equilibrium ground state T=eV=0T=eV=0 as Gσr(0)=|Gσr(0)|eiδσG_{\sigma}^{r}(0)=-\left|G_{\sigma}^{r}(0)\right|e^{i\delta_{\sigma}}. It also determines the behavior of the equilibrium spectral function ρdσ(ω)Aσ(ω)|T=eV=0\rho_{d\sigma}(\omega)\equiv\left.A_{\sigma}(\omega)\right|_{T=eV=0} in the low-frequency limit. At the Fermi level ω=0\omega=0, it takes the form

ρdσ\displaystyle\rho_{d\sigma}\,\equiv ρdσ(0)=sin2δσπΔ.\displaystyle\ \rho_{d\sigma}(0)\,=\,\frac{\sin^{2}\delta_{\sigma}}{\pi\Delta}\,. (12)

Furthermore, the first derivative of ρdσ(ω)\rho_{d\sigma}(\omega) is related to the diagonal susceptibility χσσ\chi_{\sigma\sigma}, as

ρdσρdσ(ω)ω|ω=0=χσσΔsin2δσ.\displaystyle\rho_{d\sigma}^{\prime}\equiv\,\left.\frac{\partial\rho_{d\sigma}(\omega)}{\partial\omega}\right|_{\omega=0}\ =\ \frac{\chi_{\sigma\sigma}}{\Delta}\,\sin 2\delta_{\sigma}\,. (13)

This is a result of a series of exact Fermi-liquid relations, obtained by Yamada-Yosida [see Appendix A].

II.2 Low-energy expansion of nonlinear current

Nonlinear current through quantum dots can be calculated using a Landauer-type formula [59, 60]:

I=\displaystyle I\,= eh4ΓLΓR(ΓL+ΓR)2\displaystyle\ \frac{e}{h}\frac{4\Gamma_{L}\Gamma_{R}}{(\Gamma_{L}+\Gamma_{R})^{2}}
×σ=1Ndω[fL(ω)fR(ω)]πΔAσ(ω).\displaystyle\times\sum_{\sigma=1}^{N}\int_{-\infty}^{\infty}d\omega\,\bigl{[}f_{L}(\omega)-f_{R}(\omega)\bigr{]}\,\pi\Delta A_{\sigma}(\omega). (14)

Here, fν(ω)=[eβ(ωμν)+1]1f_{\nu}(\omega)=[e^{\beta(\omega-\mu_{\nu})}+1]^{-1} is the Fermi distribution function for the conduction band on ν=L,R\nu=L,\,R. The chemical potentials of the left and right leads are driven from the Fermi energy at equilibrium EF=0E_{F}=0 by the bias voltage: μL=αLeV\mu_{L}=\alpha_{L}\,eV and μR=αReV\mu_{R}=-\alpha_{R}\,eV, with αL\alpha_{L} and αR\alpha_{R} the parameters which satisfy αL+αR=1\alpha_{L}+\alpha_{R}=1, i.e., μLμReV\mu_{L}-\mu_{R}\equiv eV. Asymmetries in tunnel couplings and that in bias voltages can be described, respectively, by the following parameters,

γdif\displaystyle\gamma_{\mathrm{dif}}\,\equiv ΓLΓRΓL+ΓR,\displaystyle\ \frac{\Gamma_{L}-\Gamma_{R}}{\Gamma_{L}+\Gamma_{R}}\,, (15)
αdif\displaystyle\alpha_{\mathrm{dif}}\,\equiv μL+μRμLμR=αLαR.\displaystyle\ \frac{\mu_{L}+\mu_{R}}{\mu_{L}-\mu_{R}}\,=\,\alpha_{L}-\alpha_{R}\,. (16)

In this work, we have derived the explicit expressions of the coefficients cV,σ(2)c_{V,\sigma}^{(2)} and cV,σ(3)c_{V,\sigma}^{(3)} for the first two nonlinear-response terms of dI/dVdI/dV in Eq. (1), using the exact low-energy asymptotic form of the spectral function Aσ(ω)A_{\sigma}(\omega) of quantum dots embedded in asymmetric junctions obtained up to terms of order ω2\omega^{2}, (eV)2(eV)^{2}, and T2T^{2}. The derivation is given in Appendix B. Specifically, we will use Eqs. (55) and (56) to clarify the roles of the tunnel and bias asymmetries, which enter through the parameters γdif\gamma_{\mathrm{dif}} and αdif\alpha_{\mathrm{dif}}, in the nonlinear Fermi-liquid transport.

II.2.1 dI/dVdI/dV of an SU(NN) dot with tunnel and bias asymmetries

In this paper, we consider the SU(NN) symmetric case, at which the impurity level has the NN-fold degeneracy: ϵdσϵd\epsilon_{d\sigma}\equiv\epsilon_{d} for σ=1,2,,N\sigma=1,2,\ldots,N. For convenience, we use a shifted impurity energy level, defined by ξdϵd+(N1)U/2\xi_{d}\equiv\epsilon_{d}+(N-1)U/2 in the following. Note that the system additionally has an electron-hole symmetry at ξd=0\xi_{d}=0.

The linear susceptibilities have two independent components in the SU(NN) symmetric case, i.e., the diagonal component χσσ\chi_{\sigma\sigma} and the off-diagonal one χσσ\chi_{\sigma\sigma^{\prime}} for σσ.\sigma\neq\sigma^{\prime}. These two parameters determine the essential properties of quasiparticles:

T14χσσ,R 1χσσχσσ.\displaystyle T^{\ast}\,\equiv\,\frac{1}{4\chi_{\sigma\sigma}}\,,\qquad\quad R\,\equiv\,1-\frac{\chi_{\sigma\sigma^{\prime}}}{\chi_{\sigma\sigma}}\,. (17)

Here, TT^{\ast} is a characteristic energy scale of the SU(NN) Fermi liquid, for instance, the TT-linear specific heat of impurity electrons is scaled in the form 𝒞imp=Nπ212T/T\mathcal{C}_{\mathrm{imp}}=\frac{N\pi^{2}}{12}T/T^{*}. The Wilson ratio RR corresponds to a dimensionless residual interaction between quasiparticles [61]: we will use the following rescaled Wilson ratio K~\widetilde{K} which is bounded in the range 0K~10\leq\widetilde{K}\leq 1,

K~(N1)(R1).\displaystyle\widetilde{K}\,\equiv\,(N-1)(R-1)\,. (18)

The differential conductance for SU(NN) symmetric quantum dots can be expressed in the following form, scaling the bias voltage eVeV by TT^{\ast},

dIdV\displaystyle\frac{dI}{dV}\, =Ng0[sin2δ+CV(2)eVTCV(3)(eVT)2+],\displaystyle=\ Ng_{0}\left[\,\sin^{2}\delta\,+\,C_{V}^{(2)}\,\frac{eV}{T^{\ast}}\,-\,C_{V}^{(3)}\left(\frac{eV}{T^{\ast}}\right)^{2}\,+\,\cdots\right], (19)
g0\displaystyle g_{0}\, e2h4ΓLΓR(ΓL+ΓR)2=e2h(1γdif2).\displaystyle\equiv\ \frac{e^{2}}{h}\,\frac{4\Gamma_{L}\Gamma_{R}}{(\Gamma_{L}+\Gamma_{R})^{2}}\ =\,\frac{e^{2}}{h}\bigl{(}1-\gamma_{\mathrm{dif}}^{2}\bigr{)}\,. (20)

The dimensionless coefficients CV(2)C_{V}^{(2)} and CV(3)C_{V}^{(3)}, can be deduced from the general formulas given in Eqs. (55) and (56), taking into account the SU(NN) symmetry:

CV(2)=\displaystyle C_{V}^{(2)}\,= π4[αdif(1K~)γdifK~]sin2δ.\displaystyle\ \frac{\pi}{4}\left[\,\alpha_{\mathrm{dif}}\left(1-\widetilde{K}\right)\,-\,\gamma_{\mathrm{dif}}\,\widetilde{K}\,\right]\,\sin 2\delta\,. (21)

For N=2N=2, this reproduces the previous result obtained by Aligia for the spin-1/2 Anderson model [57, 58]. CV(2)C_{V}^{(2)} consists of a linear combination of αdif\alpha_{\mathrm{dif}} and γdif\gamma_{\mathrm{dif}}, and it identically vanishes CV(2)0C_{V}^{(2)}\to 0 when both the tunnel couplings and the bias voltages are symmetrical αdif=γdif=0\alpha_{\mathrm{dif}}=\gamma_{\mathrm{dif}}=0. The magnitude is determined also by K~\widetilde{K} and sin2δ\sin 2\delta which can be related to the derivative of the spectral function ρd\rho_{d}^{\prime}, given in Eq. (13): sin2δ/T=4Δρd\sin 2\delta/T^{\ast}=4\Delta\rho_{d}^{\prime}. Furthermore, CV(2)C_{V}^{(2)} is an odd function of ξd\xi_{d}. Note that CV(2)C_{V}^{(2)} depends on the Coulomb interaction only through the real part of the self-energy.

In contrast, CV(3)C_{V}^{(3)} given in the following depends on both the real part that determines the high-order energy shifts and imaginary part that destroy phase coherence [62, 63], specifically on the order ω2\omega^{2} and (eV)2(eV)^{2} terms. The exact formula of the coefficient CV(3)C_{V}^{(3)} for order (eV)3(eV)^{3} nonlinear current is composed of two parts: WVW_{V} and ΘV\Theta_{V} which represent the two-body contribution and three-body one, respectively,

CV(3)=\displaystyle C_{V}^{(3)}\,= π264(WV+ΘV),\displaystyle\ \frac{\pi^{2}}{64}\,\bigl{(}W_{V}\,+\,\Theta_{V}\bigr{)}\,, (22)
WV\displaystyle W_{V}\equiv cos2δ[1+3αdif26(αdif2+αdifγdif)K~\displaystyle\ -\cos 2\delta\Biggl{[}1+3\alpha_{\mathrm{dif}}^{2}-6\Bigl{(}\alpha_{\mathrm{dif}}^{2}+\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}}\Bigr{)}\,\widetilde{K}
+{5N1+3αdif2+6αdifγdif+3(N2)N1γdif2}K~2],\displaystyle\!\!\!\!\!\!\!\!\!\!\!\!\!\!+\left\{\frac{5}{N-1}+3\alpha_{\mathrm{dif}}^{2}+6\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}}+\frac{3(N-2)}{N-1}\,\gamma_{\mathrm{dif}}^{2}\right\}\,\widetilde{K}^{2}\Biggr{]}, (23)
ΘV\displaystyle\Theta_{V}\equiv [1+3αdif2]ΘI+3[1+3αdif2+4αdifγdif]Θ~II\displaystyle\Bigl{[}1+3\alpha_{\mathrm{dif}}^{2}\Bigr{]}\Theta_{\mathrm{I}}+3\Bigl{[}1+3\alpha_{\mathrm{dif}}^{2}+4\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}}\Bigr{]}\,\widetilde{\Theta}_{\mathrm{II}}
+6[αdif2+2αdifγdif+γdif2]Θ~III.\displaystyle\ +6\Bigl{[}\alpha_{\mathrm{dif}}^{2}+2\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}}+\gamma_{\mathrm{dif}}^{2}\Bigr{]}\,\widetilde{\Theta}_{\mathrm{III}}\,. (24)

CV(3)C_{V}^{(3)} depends on the asymmetry parameters through the quadratic terms, i.e., αdif2\,\alpha_{\mathrm{dif}}^{2}, αdifγdif\,\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}}, and γdif2\,\gamma_{\mathrm{dif}}^{2}. Effects of the three-body corrections enter through the dimensionless parameters:

ΘI\displaystyle\Theta_{\mathrm{I}}\,\equiv sin2δ2πχσσσ[3]χσσ2,\displaystyle\ \frac{\sin 2\delta}{2\pi}\,\frac{\chi_{\sigma\sigma\sigma}^{[3]}}{\chi_{\sigma\sigma}^{2}}\,, (25)
Θ~II\displaystyle\widetilde{\Theta}_{\mathrm{II}}\,\equiv (N1)sin2δ2πχσσσ[3]χσσ2,\displaystyle\ (N-1)\,\frac{\sin 2\delta}{2\pi}\,\frac{\chi_{\sigma\sigma^{\prime}\sigma^{\prime}}^{[3]}}{\chi_{\sigma\sigma}^{2}}\,, (26)
Θ~III\displaystyle\widetilde{\Theta}_{\mathrm{III}}\,\equiv (N1)(N2)2sin2δ2πχσσσ′′[3]χσσ2,\displaystyle\ \frac{(N-1)(N-2)}{2}\,\frac{\sin 2\delta}{2\pi}\,\frac{\chi_{\sigma\sigma^{\prime}\sigma^{\prime\prime}}^{[3]}}{\chi_{\sigma\sigma}^{2}}\,, (27)

for σσσ′′σ\sigma\neq\sigma^{\prime}\neq\sigma^{\prime\prime}\neq\sigma. In particular, Θ~III\widetilde{\Theta}_{\mathrm{III}} represents the correlation between three different levels which appears for N3N\geq 3. This component Θ~III\widetilde{\Theta}_{\mathrm{III}} will give no contribution to CV(3)C_{V}^{(3)} through Eq. (24) if there is no tunnel or bias asymmetry [44, 64]: WVW_{V} and ΘV\Theta_{V} take the following form at αdif=γdif=0\alpha_{\mathrm{dif}}=\gamma_{\mathrm{dif}}=0,

WVαdif=γdif=0\displaystyle W_{V}\,\xrightarrow{\,\alpha_{\mathrm{dif}}=\gamma_{\mathrm{dif}}=0\,} cos2δ(1+5K~2N1),\displaystyle\ -\cos 2\delta\,\left(1\,+\,\frac{5\widetilde{K}^{2}}{N-1}\right), (28)
ΘVαdif=γdif=0\displaystyle\Theta_{V}\,\xrightarrow{\,\alpha_{\mathrm{dif}}=\gamma_{\mathrm{dif}}=0\,} ΘI+ 3Θ~II.\displaystyle\ \,\Theta_{\mathrm{I}}\,+\,3\,\widetilde{\Theta}_{\mathrm{II}}\,. (29)

When both the chemical potentials and the tunnel couplings are inverted such that (αdif(\alpha_{\mathrm{dif}}, γdif)(αdif\gamma_{\mathrm{dif}})\to(-\alpha_{\mathrm{dif}}, γdif)-\gamma_{\mathrm{dif}}), the coefficients CV(2)C_{V}^{(2)} and CV(3)C_{V}^{(3)} exhibit odd and even properties, respectively: CV(2)CV(2)C_{V}^{(2)}\to-C_{V}^{(2)} and CV(3)CV(3)C_{V}^{(3)}\to C_{V}^{(3)} as shown in Appendix B. These formulas of CV(2)C_{V}^{(2)} and CV(3)C_{V}^{(3)} of the Anderson model for arbitrary level structures ϵdσ\epsilon_{d\sigma} are consistent, in the limit of strong interaction UU\to\infty, with the corresponding formulas for the SU(NN) Kondo model obtained by Mora et al. [40] at integer-filling points nd=1,2,,N1\bigl{\langle}n_{d}\bigr{\rangle}=1,2,\ldots,N-1, with ndσndσn_{d}\equiv\sum_{\sigma}n_{d\sigma}.

III NRG results for Fermi-liquid parameters

In this section, we summarize the basic properties of quasiparticles in the SU(44) and SU(66) symmetric cases. Specifically, we have calculated the Fermi-liquid parameters δ\delta, χσσ\chi_{\sigma\sigma^{\prime}}, and χσσσ′′[3]\chi_{\sigma\sigma^{\prime}\sigma^{\prime\prime}}^{[3]} with the NRG approach, dividing NN channels into N/2N/2 pairs, and using the SU(22) spin and U(11) charge symmetries for each pair, i.e., totally k=1N/2[SU(2)U(1)]k\prod_{k=1}^{N/2}\left[\mathrm{SU}(2)\otimes\mathrm{U}(1)\right]_{k} symmetries. The discretization parameter Λ\Lambda and the number of low-lying energy states NtruncN_{\mathrm{trunc}} are chosen such that (Λ,Ntrunc)=(6,10000)(\Lambda,N_{\mathrm{trunc}})=(6,10000) for N=4N=4, and (20,30000)(20,30000) for N=6N=6 [44]. The phase shift and renormalized parameters have been deduced from the energy flow of NRG [65, 66, 67, 68, 48].

  Refer to captionRefer to caption Refer to captionRefer to caption Refer to captionRefer to caption

Figure 1: Fermi-liquid parameters of an SU(4) dot are plotted as a function of ξd/U\xi_{d}/U for U/(πΔ)=1/3U/(\pi\Delta)=1/3, 2/32/3, 5/35/3, 10/310/3, 5.05.0: (a) nd\langle n_{d}\rangle,  (b) sin2δ\sin^{2}\delta,  (c) rescaled Wilson ratio K~(N1)(R1)\widetilde{K}\equiv(N-1)(R-1),  (d) renormalized factor zz,  (e) TK/TT_{K}/T^{\ast},   and   (f) sin2δ\sin 2\delta. Here, TKT_{K} is defined as the value of T1/(4χσσ)T^{\ast}\equiv 1/(4\chi_{\sigma\sigma}) at half filling ξd=0\xi_{d}=0.

III.1 SU(4) Fermi-liquid properties

III.1.1 Two-body correlation functions of an SU(4) dot

The Fermi-liquid parameters of an SU(4) quantum dot, which can be deduced from the phase shift and the linear susceptibilities, are shown in Fig. 1 as a function of ξd\xi_{d} for several different values of U/(πΔ)=1/3U/(\pi\Delta)=1/3, 2/32/3, 5/35/3, 10/310/3, 55. For large U/(πΔ)2.0U/(\pi\Delta)\gtrsim 2.0, the occupation number in Fig. 1(a) shows a Coulomb-staircase behavior with the plateaus of integer height nd1.0\langle n_{d}\rangle\simeq 1.0, 2.02.0, 3.03.0 and the steps at ξdU\xi_{d}\simeq-U, 0, UU: the structure becomes clearer for stronger interactions. Figure 1(b) shows sin2δ\sin^{2}\delta corresponding to the linear term of the differential conductance in Eq. (1). For strong interactions U/(πΔ)2.0U/(\pi\Delta)\gtrsim 2.0, the Kondo ridges which reflect the step structure of the occupation number of the values nd1,2,3\langle n_{d}\rangle\simeq 1,2,3 evolve at ξdU,0,U\xi_{d}\simeq U,0,-U, respectively, as UU increases.

Figure 1(c) shows the rescaled Wilson ratio K~(N1)(R1)\widetilde{K}\equiv(N-1)(R-1). It has a wide plateau that reaches the strong-coupling limit value K~1\widetilde{K}\simeq 1 in the region 1.5Uξd1.5U-1.5U\lesssim\xi_{d}\lesssim 1.5U, for large interactions U/(πΔ)3.0U/(\pi\Delta)\gtrsim 3.0. This is caused by the fact that charge fluctuations are suppressed in this region and it makes the charge susceptibility vanish: χcχσσ+(N1)χσσ0\chi_{c}\propto\chi_{\sigma\sigma}+(N-1)\chi_{\sigma\sigma^{\prime}}\to 0. The shallow dips of K~\widetilde{K} at ξd±0.5U\xi_{d}\simeq\pm 0.5U is caused by the charge fluctuations at the steps of the Coulomb staircase structure of nd\langle n_{d}\rangle.

Correspondingly, the renormalization factor zz in Fig. 1(d) exhibits a broad valley structure at 1.5Uξd1.5U-1.5U\lesssim\xi_{d}\lesssim 1.5U, which becomes deeper as UU increases. It has a clear local minimum for U/(πΔ)2.0U/(\pi\Delta)\gtrsim 2.0 at ξd\xi_{d}\simeq 0 and ±U\pm U, where nd\langle n_{d}\rangle takes integer values: it also has a local maximum at intermediate valence states in between the two adjacent minima. Figure 1(e) shows the gate voltage ξd\xi_{d} dependence of the inverse of the characteristic energy scale, 1/T1/T^{\ast}, scaled by TKT_{K} that is defined as the value of TT^{\ast} at the elelctron-hole symmetric point ξd=0\xi_{d}=0. It shows an oscillatory behavior for strong interactions U/(πΔ)3.0U/(\pi\Delta)\gtrsim 3.0, reflecting the dependence of zz on ξd\xi_{d}. In particular, 1/T1/T^{\ast} reaches a local maximum at the integer-filling points where the SU(44) Kondo effect occurs.

Some of the two-body correlation functions contribute to the low-energy transport through the derivative of the impurity density of state: ρd=(χσσ/Δ)sin2δ\rho_{d}^{\prime}=(\chi_{\sigma\sigma}/\Delta)\sin 2\delta given in Eq. (13). For instance, the coefficient CV(2)C_{V}^{(2)} is proportional to ρd\rho_{d}^{\prime}. Figure 1(f) shows sin2δ\sin 2\delta. It is an odd function of ξd\xi_{d} and vanishes at ξd=0\xi_{d}=0 where the phase shift takes the value δ=π/2\delta=\pi/2. For strong interactions U/(πΔ)3.0U/(\pi\Delta)\gtrsim 3.0, it has a wide maximum (minimum) at ξdU\xi_{d}\simeq U (U-U), where the 1/41/4-filling (3/43/4-filling) SU(44) Kondo occurs.

  Refer to captionRefer to caption Refer to captionRefer to caption

Figure 2: Three-body correlation functions of an SU(4) dot plotted vs ξd/U\xi_{d}/U for U/πΔ=1/3U/\pi\Delta=1/3, 2/32/3, 5/35/3, 10/310/3, 5.05.0: (a) ΘI\Theta_{\mathrm{I}}, (b) Θ~II\widetilde{\Theta}_{\mathrm{II}}, (c) Θ~III\widetilde{\Theta}_{\mathrm{III}}. (d) comparison of ΘI\Theta_{\mathrm{I}}, Θ~II-\widetilde{\Theta}_{\mathrm{II}}, and Θ~III\widetilde{\Theta}_{\mathrm{III}} at a large interaction U/(πΔ)=5.0U/(\pi\Delta)=5.0. It indicates that ΘIΘ~IIΘ~III\Theta_{\mathrm{I}}\simeq-\widetilde{\Theta}_{\mathrm{II}}\simeq\widetilde{\Theta}_{\mathrm{III}} over a wide range of UξdU-U\lesssim\xi_{d}\lesssim U.

III.1.2 Three-body correlation functions of an SU(4) dot

We next consider the three-body correlations between electrons passing through an SU(4) symmetric dot. Figure 2 shows ΘI\Theta_{\mathrm{I}}, Θ~II\widetilde{\Theta}_{\mathrm{II}}, and Θ~III\widetilde{\Theta}_{\mathrm{III}} as a function of ξd/U\xi_{d}/U for several different values of UU. These dimensionless three-body correlation functions vanish in the electron-hole symmetric case ξd=0\xi_{d}=0. For strong interactions U/(πΔ)3.0U/(\pi\Delta)\gtrsim 3.0, a plateau of the width UU emerges at ξd0\xi_{d}\simeq 0 and ±U\pm U, i.e., at integer filling points nd1,\langle n_{d}\rangle\simeq 1, 22, 33. The plateau structure evolves further as interaction UU increases. Specifically, the correlation function between three different levels Θ~III\widetilde{\Theta}_{\mathrm{III}} appears for SU(NN) quantum dots with N3N\geq 3, and contributes to the nonlinear conductance when there are some asymmetries in the tunnel couplings or the bias voltages.

A comparison of the three independent components ΘI\Theta_{\mathrm{I}}, Θ~II-\widetilde{\Theta}_{\mathrm{II}}, and Θ~III\widetilde{\Theta}_{\mathrm{III}} is made in Fig. 2(d) for a large interaction U/(πΔ)=5.0U/(\pi\Delta)=5.0. It shows that these three components approach each other very closely over a wide range of gate voltages 1.5Uξd1.5U-1.5U\lesssim\xi_{d}\lesssim 1.5U,

ΘIΘ~IIΘ~III.\displaystyle\Theta_{\mathrm{I}}\,\simeq\,-\widetilde{\Theta}_{\mathrm{II}}\,\simeq\,\widetilde{\Theta}_{\mathrm{III}}\,. (30)

This is caused by the fact that the derivatives of the two independent components of linear susceptibilities, |χσσϵd||\frac{\partial\chi_{\sigma\sigma}}{\partial\epsilon_{d}}| and |χσσϵd||\frac{\partial\chi_{\sigma\sigma^{\prime}}}{\partial\epsilon_{d}}|, become much smaller than an inverse of the characteristic energy scale (T)2(T^{\ast})^{-2} in a wide range of electron fillings 1ndN11\lesssim\langle n_{d}\rangle\lesssim N-1, in addition to the suppression of the charge fluctuations χc0\chi_{c}\simeq 0, mentioned above [see Appendix C] [44].

  Refer to captionRefer to caption Refer to captionRefer to caption Refer to captionRefer to caption

Figure 3: Fermi-liquid parameters of an SU(6) dot are plotted as a function of ξd/U\xi_{d}/U for U/(πΔ)=2/5U/(\pi\Delta)=2/5, 1.01.0, 2.02.0, 5.05.0. (a) nd\langle n_{d}\rangle,  (b) sin2δ\sin^{2}\delta,  (c) rescaled Wilson ratio K~(N1)(R1)\widetilde{K}\equiv(N-1)(R-1),  (d) renormalized factor zz,  (e) TK/TT_{K}/T^{\ast},  and   (f) sin2δ\sin 2\delta. Here, TKT_{K} is defined as the value of T1/(4χσσ)T^{\ast}\equiv 1/(4\chi_{\sigma\sigma}) at half filling ξd=0\xi_{d}=0.

  Refer to captionRefer to caption Refer to captionRefer to caption

Figure 4: Three-body correlation functions of an SU(6) dot plotted vs ξd/U\xi_{d}/U for U/πΔ=2/5U/\pi\Delta=2/5, 1.01.0, 2.02.0, 5.05.0: (a) ΘI\Theta_{\mathrm{I}}, (b) Θ~II\widetilde{\Theta}_{\mathrm{II}}, (c) Θ~III\widetilde{\Theta}_{\mathrm{III}}. (d) comparison of ΘI\Theta_{\mathrm{I}}, Θ~II-\widetilde{\Theta}_{\mathrm{II}}, and Θ~III\widetilde{\Theta}_{\mathrm{III}} at a large interaction U/(πΔ)=5.0U/(\pi\Delta)=5.0. It indicates that ΘIΘ~IIΘ~III\Theta_{\mathrm{I}}\simeq-\widetilde{\Theta}_{\mathrm{II}}\simeq\widetilde{\Theta}_{\mathrm{III}} over a wide range of (N1)/2Uξd(N1)/2-(N-1)/2U\lesssim\xi_{d}\lesssim(N-1)/2 for N=6N=6.

III.2 SU(6) Fermi-liquid properties

III.2.1 Two-body correlation functions for an SU(6) dot

We next summarize the low-energy Fermi-liquid properties of SU(66) quantum dots. The phase shift and the renormalized parameters that can be deduced from the two-body correlations are plotted as a function of the gate voltage ξd/U\xi_{d}/U in Fig. 3, for several different values of U/(πΔ)=2/5U/(\pi\Delta)=2/5, 1.01.0, 2.02.0, 5.05.0, i.e., from weak to strong interactions. The behaviors of these parameters are similar to those of the SU(44) quantum dots. However, the number of different SU(NN) Kondo states occurring at integer fillings increases with NN, i.e., nd=1\langle n_{d}\rangle=1, 22, \ldots, N1N-1. It takes place at ξd0.0\xi_{d}\simeq 0.0, ±U\pm U, …, ±N22U\pm\frac{N-2}{2}U, and gives an interesting variety in low-energy transport. As NN increases, quantum fluctuations caused by the Coulomb interaction UU is suppressed, and in particular, the mean-field theory becomes exact in the large NN limit of the finite-UU Anderson impurity model [69]. Hence for larger NN, electron-correlation effects occur for stronger interactions UU.

In Fig. 3(a), we can see that the Coulomb staircase behavior of the occupation number nd\langle n_{d}\rangle emerges for a large interaction U/(πΔ)=5.0U/(\pi\Delta)=5.0, which has steps at ξd0.0\xi_{d}\simeq 0.0, ±U\pm U, ±2U\pm 2U for SU(66) quantum dots. The transmission probability sin2δ\sin^{2}\delta, shown in Fig. 3(b), reaches the unitary limit value sin2δ=1.0\sin^{2}\delta=1.0 at half filling ξd=0\xi_{d}=0. We can see that the plateau structure develops as UU increases, and it becomes visible at a strong interaction U/(πΔ)=5.0U/(\pi\Delta)=5.0.

Figure 3(c) shows that the rescaled Wilson ratio K~\widetilde{K} for N=6N=6 has a wide flat peak at (N1)U/2ξd(N1)U/2-(N-1)U/2\lesssim\xi_{d}\lesssim(N-1)U/2, the height of which increases with UU. In particular, it approaches saturation value K~1\widetilde{K}\simeq 1 at U/(πΔ)=5.0U/(\pi\Delta)=5.0 due to the suppression of charge fluctuations, as mentioned.

The renormalization factor zz for SU(66) quantum dots, shown in Fig.  3(d), exhibits a broad valley structure similar to the one for the SU(4) symmetric case. The valley becomes deeper as UU increases, and the local minima emerge at the integer-filling points, reflecting the electron correlations due to the SU(6) Kondo effects. Figure 3(e) shows that the inverse of the characteristic energy 1/T1/T^{\ast}, which is scaled by the value TKT_{K} defined at half filling ξd=0\xi_{d}=0. It has peaks situated at ξd0\xi_{d}\simeq 0, ±U\pm U, ±2U\pm 2U for a strong interaction U/(πΔ)=5.0U/(\pi\Delta)=5.0, although the ones at ±2U\pm 2U are still developing. These peaks correspond to the SU(6) Kondo temperature at each of the integer-filling points.

Figure 3(f) shows sin2δ\sin 2\delta for N=6N=6, which is proportional to the derivative of the density of states ρd=(χσσ/Δ)sin2δ\rho_{d}^{\prime}=(\chi_{\sigma\sigma}/\Delta)\sin 2\delta, and determines the magnitude of the coefficient CV(2)C_{V}^{(2)} of the nonlinear current, as mentioned. This factor sin2δ\sin 2\delta is an odd function of ξd\xi_{d}, and at a strong interaction U/(πΔ)=5.0U/(\pi\Delta)=5.0 it takes a broad peak (dip) at δ=π/4\delta=\pi/4 (3π/43\pi/4) corresponding to the half-integer filling nd=1.5\langle n_{d}\rangle=1.5 (4.54.5) where charge fluctuations are not fully suppressed.

III.2.2 Three-body correlation functions of an SU(6) dot

Figure 4 shows three-body correlations between electrons passing through an SU(6) quantum dot, calculated as a function of ξd\xi_{d} for several values of interactions U/πΔ=2/5U/\pi\Delta=2/5, 1.01.0, 2.02.0, 5.05.0. These dimensionless correlation functions ΘI\Theta_{\mathrm{I}}, Θ~II\widetilde{\Theta}_{\mathrm{II}}, and Θ~III\widetilde{\Theta}_{\mathrm{III}} vanish at half filling ξd=0\xi_{d}=0, and away from half filling, they are enhanced as UU increases. We can see that for a strong interaction U/(πΔ)=5.0U/(\pi\Delta)=5.0 there emerges either a wide peak, a wide dip, or a plateau at ξd±U\xi_{d}\simeq\pm U, ±2U\pm 2U, i.e., at integer filling points. The component between three different levels Θ~III\widetilde{\Theta}_{\mathrm{III}} also contributes to the order (eV)3(eV)^{3} term of nonlinear current through an SU(66) quantum dot when there are some asymmetries in the tunnel couplings or the bias voltages.

In Fig. 4(d) the three independent components ΘI\Theta_{\mathrm{I}}, Θ~II-\widetilde{\Theta}_{\mathrm{II}}, and Θ~III\widetilde{\Theta}_{\mathrm{III}} are compared in a strong interaction case U/(πΔ)=5.0U/(\pi\Delta)=5.0. It shows that these three components approach each other very closely ΘIΘ~IIΘ~III\Theta_{\mathrm{I}}\simeq-\widetilde{\Theta}_{\mathrm{II}}\simeq\widetilde{\Theta}_{\mathrm{III}} over a wide range of the gate voltage (N1)U/2ξd(N1)U/2-(N-1)U/2\lesssim\xi_{d}\lesssim(N-1)U/2 with N=6N=6. This is due to the fact that the diagonal component χσσσ[3]\chi_{\sigma\sigma\sigma}^{[3]} dominates the three-body correlation, and the derivatives |χσσϵd||\frac{\partial\chi_{\sigma\sigma}}{\partial\epsilon_{d}}| and |χσσϵd||\frac{\partial\chi_{\sigma\sigma^{\prime}}}{\partial\epsilon_{d}}| become much smaller than (T)2(T^{\ast})^{-2} in a wide range of electron fillings 1ndN11\lesssim\langle n_{d}\rangle\lesssim N-1 [see Appendix C] [44].

IV Order (eV)2(eV)^{2} nonlinear current for SU(4) and SU(6) cases

The coefficient CV(2)C_{V}^{(2)}, defined in Eqs. (19) and (21), emerges when the tunnel coupling or the bias voltage is not symmetric. Its magnitude is determined by the Wilson ratio K~\widetilde{K} and the derivative of the density of states: ρd=(χσσ/Δ)sin2δ\rho_{d}^{\prime}=(\chi_{\sigma\sigma}/\Delta)\sin 2\delta.

In this section, we first of all describe behavior of CV(2)C_{V}^{(2)} in some limiting cases, and then discuss the NRG results to show how the coefficient evolves with the tunnel and bias asymmetries.

IV.1 Behavior of CV(2)C_{V}^{(2)} in some limiting cases

We have seen in Figs. 1 and 3 that the Wilson ratio reaches the saturation value K~1\widetilde{K}\simeq 1 over a wide range of gate voltages (N1)U/2ξd(N1)U/2-(N-1)U/2\lesssim\xi_{d}\lesssim(N-1)U/2 for large UU, where the quantum dot is partially filled 1ndN11\lesssim\langle n_{d}\rangle\lesssim N-1. This is caused by the fact that the charge fluctuations are significantly suppressed in this region, as mentioned. In the limit of strong interactions, CV(2)C_{V}^{(2)} becomes independent of the bias asymmetry αdif\alpha_{\mathrm{dif}}:

CV(2)K~1\displaystyle C_{V}^{(2)}\xrightarrow{\,\widetilde{K}\to 1\,} π4γdifsin2δ,\displaystyle\,-\frac{\pi}{4}\,\gamma_{\mathrm{dif}}\,\sin 2\delta\,, (31)

In contrast, in the limit of |ξd||\xi_{d}|\to\infty where nd\langle n_{d}\rangle approaches 0 or NN, the Wilson ratio approaches the noninteracting value K~0\widetilde{K}\to 0, and CV(2)C_{V}^{(2)} becomes independent of tunnel asymmetry γdif\gamma_{\mathrm{dif}}:

CV(2)K~0π4αdifsin2δ.\displaystyle C_{V}^{(2)}\xrightarrow{\,\widetilde{K}\to 0\,}\,\frac{\pi}{4}\,\alpha_{\mathrm{dif}}\,\sin 2\delta\,. (32)
Refer to captionRefer to caption
Figure 5: Effects of tunnel asymmetry γdif\gamma_{\mathrm{dif}} on order (eV)2(eV)^{2} nonlinear current: CV(2)/γdifC_{V}^{(2)}/\gamma_{\mathrm{dif}} given by Eq. (33) for symmetrical bias voltages αdif=0.0\alpha_{\mathrm{dif}}=0.0 is plotted vs ξd/U\xi_{d}/U, varying interactions from weak to strong. (a) for SU(4) quantum dots with U/(πΔ)=1/3U/(\pi\Delta)=1/3, 2/32/3, 5/35/3, 10/310/3, 5.05.0. (b) for SU(6) quantum dots with U/(πΔ)=2/5U/(\pi\Delta)=2/5, 1.01.0, 2.02.0, 5.05.0.

IV.2 Effects of tunnel asymmetry γdif0\gamma_{\mathrm{dif}}\neq 0 on CV(2)C_{V}^{(2)}
for symmetric bias voltages αdif=0\alpha_{\mathrm{dif}}=0

We first of all consider effects of tunneling asymmetries γdif\gamma_{\mathrm{dif}} on CV(2)C_{V}^{(2)}, taking bias voltages to be symmetric αdif=0\alpha_{\mathrm{dif}}=0:

CV(2)αdif=0π4γdifK~sin2δ.\displaystyle C_{V}^{(2)}\xrightarrow{\,\alpha_{\mathrm{dif}}=0\,}\,-\frac{\pi}{4}\,\gamma_{\mathrm{dif}}\,\widetilde{K}\,\sin 2\delta\,. (33)

In this case, CV(2)C_{V}^{(2)} is proportional to γdif\gamma_{\mathrm{dif}}, and is determined by K~\widetilde{K} and sin2δ\sin 2\delta. Figure 5 shows CV(2)/γdifC_{V}^{(2)}/\gamma_{\mathrm{dif}} as a function of ξd/U\xi_{d}/U for symmetric bias voltages αdif=0.0\alpha_{\mathrm{dif}}=0.0. We have examined the behaviors from weak to strong interactions: (a) for SU(4) quantum dots with U/(πΔ)=1/3U/(\pi\Delta)=1/3, 2/32/3, 5/35/3, 10/310/3, 5.05.0, and (b) for SU(6) quantum dots with U/(πΔ)=2/5U/(\pi\Delta)=2/5, 1.01.0, 2.02.0, 5.05.0.

As UU increases, a wide peak and a wide dip of CV(2)C_{V}^{(2)} evolve at ξdU/4\xi_{d}\simeq U/4 and U/4-U/4, where the occupation number reaches the value of ndN/4\langle n_{d}\rangle\simeq N/4 and 3N/43N/4, respectively. It corresponds to phase shifts of δπ/4\delta\simeq\pi/4 and 3π/43\pi/4, at which sin2δ\sin 2\delta takes an extreme value as seen in Figs. 1(f) and 3(f). The peak and dip structures also reflect the fact that the Wilson ratio is almost saturated K~1.0\widetilde{K}\simeq 1.0 in a wider range of |ξd|(N1)U/2|\xi_{d}|\lesssim(N-1)U/2 for large UU in Figs. 1(c) and 3(c). The Kondo effect of an integer filling occurs at the flat peak and the flat dip for N=4N=4 quantum dots. In contrast, for N=6N=6, it is an intermediate valence state that occurs at the peak and dip, and thus the structures become round rather than flat since charge fluctuations remain active.

Outside the correlated region, the absolute value of CV(2)C_{V}^{(2)} decreases as UU increases, for both N=4N=4 and 66 in Figs. 5 (a) and (b), respectively. In particular, at |ξd|(N1)U/2|\xi_{d}|\gg(N-1)U/2, it vanishes asymptotically CV(2)0C_{V}^{(2)}\to 0 as the occupation number approaches nd0\langle n_{d}\rangle\to 0 or NN.

Refer to caption
Refer to caption
Figure 6: Effects of bias-voltage asymmetry αdif\alpha_{\mathrm{dif}} on order (eV)2(eV)^{2} nonlinear current: CV(2)/αdifC_{V}^{(2)}/\alpha_{\mathrm{dif}} given by Eq. (34) for symmetric junctions with γdif=0.0\gamma_{\mathrm{dif}}=0.0 is plotted vs ξd/U\xi_{d}/U, varying interactions from weak to strong. (a) for SU(4) quantum dots with U/(πΔ)=1/3U/(\pi\Delta)=1/3, 2/32/3, 5/35/3, 10/310/3, 5.05.0. (b) for SU(6) quantum dots with U/(πΔ)=2/5U/(\pi\Delta)=2/5, 1.01.0, 2.02.0, 5.05.0.

IV.3 Effects of bias asymmetry αdif0\alpha_{\mathrm{dif}}\neq 0 on CV(2)C_{V}^{(2)}
for symmetric tunnel junctions γdif=0\gamma_{\mathrm{dif}}=0

We next consider the effects of bias asymmetries αdif\alpha_{\mathrm{dif}} on CV(2)C_{V}^{(2)}, setting tunnel junctions to be symmetric γdif=0\gamma_{\mathrm{dif}}=0. In this case, CV(2)C_{V}^{(2)} becomes proportional to αdif\alpha_{\mathrm{dif}}, as

CV(2)γdif=0π4αdif(1K~)sin2δ.\displaystyle C_{V}^{(2)}\xrightarrow{\,\gamma_{\mathrm{dif}}=0\,}\,\frac{\pi}{4}\,\alpha_{\mathrm{dif}}\,\Bigl{(}1-\widetilde{K}\Bigr{)}\,\sin 2\delta\,. (34)

The ratio CV(2)/αdifC_{V}^{(2)}/\alpha_{\mathrm{dif}} for this case is plotted vs ξd/U\xi_{d}/U in Fig. 6. We have also examined from weak to strong interactions for symmetric tunnel couplings γdif=0.0\gamma_{\mathrm{dif}}=0.0: (a) for SU(4) quantum dots with U/(πΔ)=1/3U/(\pi\Delta)=1/3, 2/32/3, 5/35/3, 10/310/3, 5.05.0, and (b) for SU(6) quantum dots with U/(πΔ)=2/5U/(\pi\Delta)=2/5, 1.01.0, 2.02.0, 5.05.0.

At |ξd|(N1)U/2|\xi_{d}|\lesssim(N-1)\,U/2, where the localized levels of quantum dots are partially filled 1ndN11\lesssim\langle n_{d}\rangle\lesssim N-1, the factor 1K~1-\widetilde{K} in Eq. (34) becomes very small as UU increases since the Wilson ratio approaches the saturation value K~1.0\widetilde{K}\to 1.0. Therefore, CV(2)C_{V}^{(2)} almost vanishes in this region for strong interactions U/(πΔ)3.0U/(\pi\Delta)\gtrsim 3.0. However, weak oscillatory behavior survives at ξd/U±(N12m)/2\xi_{d}/U\simeq\pm(N-1-2m)/2 for m=1,,N/21m=1,\ldots,N/2-1, and it is caused by charge fluctuations in the intermediate valence states between two adjacent integer filling points.

In contrast outside the correlated region, at |ξd|(N1)U/2|\xi_{d}|\gg(N-1)\,U/2, the rescaled Wilson ratio decreases and it eventually vanishes in the limit of |ξd||\xi_{d}|\to\infty as shown in Figs. 1(c) and 3(c), so that 1K~1.01-\widetilde{K}\to 1.0 in Eq. (34). Therefore, the behavior of CV(2)C_{V}^{(2)} in this region is mainly determined by the other factor sin2δ\sin 2\delta. The localized levels of quantum dots become almost empty or fully occupied in the limit of |ξd||\xi_{d}|\to\infty. Correspondingly, at the crossover region |ξd|(N1)U/2|\xi_{d}|\simeq(N-1)\,U/2, the phase shift takes the value around δπ/N\delta\simeq\pi/N at ξd(N1)U/2\xi_{d}\simeq(N-1)U/2, and δπ(N1)/N\delta\simeq\pi(N-1)/N at ξd(N1)U/2\xi_{d}\simeq-(N-1)U/2. Thus, the peak height, or the dip depth, of CV(2)C_{V}^{(2)} decreases as NN increases: the amplitude becomes larger for the SU(4) quantum dots than the SU(6) in Fig. 6.

For weak interactions, effects of bias asymmetries αdif\alpha_{\mathrm{dif}} appear in the whole region of ξd\xi_{d}, particularly CV(2)C_{V}^{(2)} takes finite values in the region of |ξd|(N1)U/2|\xi_{d}|\simeq(N-1)U/2.

IV.4 CV(2)C_{V}^{(2)} under both asymmetries αdif0\alpha_{\mathrm{dif}}\neq 0 and γdif0\gamma_{\mathrm{dif}}\neq 0

We next consider the behavior of CV(2)C_{V}^{(2)} in the presence of both asymmetries, i.e., αdif0\alpha_{\mathrm{dif}}\neq 0 and γdif0\gamma_{\mathrm{dif}}\neq 0. We have seen above that in the strongly correlated region at |ξd|(N1)U/2|\xi_{d}|\lesssim(N-1)U/2 for large UU, the coefficient CV(2)C_{V}^{(2)} is given by Eq. (31) and it becomes almost independent of bias asymmetries αdif\alpha_{\mathrm{dif}}, since the Wilson ratio approaches K~1.0\widetilde{K}\simeq 1.0. Conversely, CV(2)C_{V}^{(2)} is given by Eq. (32), which does not depend on tunnel asymmetries γdif\gamma_{\mathrm{dif}}, for small UU or outside the correlated region as K~0.0\widetilde{K}\simeq 0.0. These properties provide the key to clarify overall characteristics of CV(2)C_{V}^{(2)} in a wide parameter range.

IV.4.1 Effects of bias asymmetry αdif0\alpha_{\mathrm{dif}}\neq 0 on CV(2)C_{V}^{(2)}
at large tunnel asymmetry γdif\gamma_{\mathrm{dif}} (ΓLΓR\Gamma_{L}\gg\Gamma_{R})

The coefficient CV(2)C_{V}^{(2)} for a large fixed tunnel asymmetry γdif=0.8\gamma_{\mathrm{dif}}=0.8, is plotted vs ξd\xi_{d} in Fig. 7, varying bias asymmetries, αdif=0.0\alpha_{\mathrm{dif}}=0.0, ±0.5\pm 0.5, ±1.0\pm 1.0. The upper panels describe the behavior for a strong interaction U/(πΔ)=5.0U/(\pi\Delta)=5.0, and the lower ones describe that for weak interactions, i.e., U/(πΔ)=1/3U/(\pi\Delta)=1/3 for SU(44), and U/(πΔ)=2/5U/(\pi\Delta)=2/5 for SU(66).

The results in Figs. 7 (a) and (b) show that the coefficient CV(2)C_{V}^{(2)} becomes almost independent of bias asymmetries in the strong coupling region |ξd|(N1)U/2|\xi_{d}|\lesssim(N-1)U/2 for large UU since the Wilson ratio approaches the saturated value K~1.0\widetilde{K}\simeq 1.0. In this region, CV(2)C_{V}^{(2)} is proportional to γdif\gamma_{\mathrm{dif}}, and shows the same behavior as that in Figs. 5(a) and (b). For N4N\geq 4, CV(2)C_{V}^{(2)} has a peak and a dip which correspond to the points sin2δ±1.0\sin 2\delta\simeq\pm 1.0 inside the correlated region, and thus the coefficient takes the value CV(2)±(π/4)γdifC_{V}^{(2)}\simeq\pm(\pi/4)\gamma_{\mathrm{dif}} at the extreme points. This kind of extreme points in the correlated region do not take place for SU(2) quantum dots since, for N=2N=2, the phase shift takes the values δ=π/4\delta=\pi/4 and 3π/43\pi/4 in the valence fluctuation regime, instead of the Kondo regime [56]. Around the extreme points, CV(2)C_{V}^{(2)} takes a typical flat structure of the Kondo state for N=4N=4, while it takes a round structure typical to the intermediate valence state for N=6N=6, as mentioned above.

The coefficient CV(2)C_{V}^{(2)} has an additional zero point other than the one at ξd=0\xi_{d}=0 in the case αdifγdif>0\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}}>0, at which bias and tunneling asymmetries cooperatively enhance the charge transfer from one of the electrodes, and the Wilson ratio satisfies the following condition,

K~=αdifαdif+γdif.\displaystyle\widetilde{K}=\frac{\alpha_{\mathrm{dif}}}{\alpha_{\mathrm{dif}}+\gamma_{\mathrm{dif}}}\,. (35)

It represents the condition that the first and second terms in Eq. (21) cancel each other out. It takes place valence fluctuation regime |ξd|(N1)U/2|\xi_{d}|\gtrsim(N-1)U/2 in Fig. 7. Conversely, effects of tunnel and bias asymmetries become constructive for αdifγdif<0.0\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}}<0.0. The behavior at |ξd|(N1)U/2|\xi_{d}|\gg(N-1)U/2, is determined by the factor αdifsin2δ\alpha_{\mathrm{dif}}\sin 2\delta.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 7: CV(2)C_{V}^{(2)} for a junction with large tunnel asymmetry γdif=0.8\gamma_{\mathrm{dif}}=0.8 is plotted vs ξd/U\xi_{d}/U, varying bias-voltage asymmetries, as αdif=1.0\alpha_{\mathrm{dif}}=-1.0 (\bullet), 0.5-0.5 (×\times), 0.00.0 (\blacksquare), 0.50.5 (\square), 1.01.0 (\triangle). (Left panels) for SU(4) quantum dots and (Right panels) for SU(6) ones, with (top panels) a strong interaction U/(πΔ)=5.0U/(\pi\Delta)=5.0 and with (bottom panels) weak interactions U/(πΔ)=1/3U/(\pi\Delta)=1/3 for SU(44) and U/(πΔ)=2/5U/(\pi\Delta)=2/5 for SU(66).

IV.4.2 Effects of tunnel asymmetry γdif0\gamma_{\mathrm{dif}}\neq 0 on CV(2)C_{V}^{(2)}
at large bias asymmetry αdif=1\alpha_{\mathrm{dif}}=1

We next consider the effects of tunnel asymmetry γdif\gamma_{\mathrm{dif}} on CV(2)C_{V}^{(2)} at large bias asymmetry αdif=1.0\alpha_{\mathrm{dif}}=1.0, which describes the situation that the right lead is grounded. Figure 8 shows the result of CV(2)C_{V}^{(2)}, plotted as a function ξd\xi_{d} varying tunnel asymmetries, as γdif=0.0\gamma_{\mathrm{dif}}=0.0, ±0.2\pm 0.2, ±0.5\pm 0.5, ±0.8\pm 0.8: (top panels) for a strong U/(πΔ)=5.0U/(\pi\Delta)=5.0, and (bottom panels) for weak interactions with (left panels) U/(πΔ)=1/3U/(\pi\Delta)=1/3 for SU(44) quantum dots and (right panels) U/(πΔ)=2/5U/(\pi\Delta)=2/5 for SU(66) ones.

In the strong-coupling limit region |ξd|(N1)U/2|\xi_{d}|\lesssim(N-1)U/2, the results in Figs. 8 (a) and (b) show almost the same behavior as that for symmetric bias voltage αdif=0.0\alpha_{\mathrm{dif}}=0.0 given in Figs. 5 (a) and (b), respectively, except for the ones for γdif=0.0\gamma_{\mathrm{dif}}=0.0. This is because the Wilson ratio reaches saturated K~1.0\widetilde{K}\simeq 1.0 and CV(2)C_{V}^{(2)} becomes independent of αdif\alpha_{\mathrm{dif}}, as shown in Eq. (31).

In the valence fluctuation region |ξd|(N1)U/2|\xi_{d}|\simeq(N-1)U/2, CV(2)C_{V}^{(2)} has an extra zero point other than the one at ξd=0\xi_{d}=0 in the case at which αdifγdif>0\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}}>0, mentioned above at Eq. (35). As the impurity level deviates further away from the electron-hole symmetric point, i.e., at |ξd|(N1)U/2|\xi_{d}|\gtrsim(N-1)U/2, the coefficient CV(2)C_{V}^{(2)} becomes less sensitive to γdif\gamma_{\mathrm{dif}} and its behavior is described by Eq. (32) as the Wilson ratio approaches K~0.0\widetilde{K}\simeq 0.0. We also see in Fig. 8(c) that CV(2)C_{V}^{(2)} does not have an extra zero point for SU(4) symmetric quantum dots with a weak interaction. This is because the Wilson ratio in this case takes values in the range K~0.54\widetilde{K}\lesssim 0.54 as shown in Fig. 1(c) which does not satisfy the condition Eq. (35) as αdif/(αdif+γdif)0.56\alpha_{\mathrm{dif}}/(\alpha_{\mathrm{dif}}+\gamma_{\mathrm{dif}})\simeq 0.56 for αdif=1.0\alpha_{\mathrm{dif}}=1.0 and γdif=0.8\gamma_{\mathrm{dif}}=0.8. Nevertheless, extra zero points will emerge even in this situation if tunnel asymmetries are slightly larger, i.e., γdif0.85\gamma_{\mathrm{dif}}\gtrsim 0.85. An example is shown in Fig. 8(d) for SU(66) quantum dots: CV(2)C_{V}^{(2)} has an extra zero point clearly for γdif=0.8\gamma_{\mathrm{dif}}=0.8, whereas it does not for γdif0.5\gamma_{\mathrm{dif}}\lesssim 0.5. In this case, the Wilson ratio is bounded in the range of K~0.72\widetilde{K}\lesssim 0.72 as shown in Fig. 3(c), and thus the condition Eq. (35) is satisfied for γdif=0.8\gamma_{\mathrm{dif}}=0.8 at which αdif/(αdif+γdif)0.56\alpha_{\mathrm{dif}}/(\alpha_{\mathrm{dif}}+\gamma_{\mathrm{dif}})\simeq 0.56, whereas it is not, for instance, for γdif=0.2\gamma_{\mathrm{dif}}=0.2 as αdif/(αdif+γdif)0.83\alpha_{\mathrm{dif}}/(\alpha_{\mathrm{dif}}+\gamma_{\mathrm{dif}})\simeq 0.83.

Refer to captionRefer to caption
Refer to captionRefer to caption
Figure 8: CV(2)C_{V}^{(2)} for a large bias-voltage asymmetry αdif=1.0\alpha_{\mathrm{dif}}=1.0 is plotted vs ξd\xi_{d}, varying tunnel asymmetries, as γdif=0.8\gamma_{\mathrm{dif}}=-0.8 (\circ), 0.5-0.5 (×\times), 0.2-0.2 (\diamond), 0.00.0 ( ), 0.20.2 (\triangledown), 0.50.5 (\square) and 0.80.8 (\triangle). (Left panels) for SU(4) quantum dots and (Right panels) for SU(6) ones, with (top panels) a strong interaction U/(πΔ)=5.0U/(\pi\Delta)=5.0 and (bottom panels) weak interactions U/(πΔ)=1/3U/(\pi\Delta)=1/3 for SU(44) and U/(πΔ)=2/5U/(\pi\Delta)=2/5 for SU(66).

V Order (eV)3(eV)^{3} nonlinear current for SU(4) and SU(6) cases

The coefficient CV(3)C_{V}^{(3)} of the order (eV)2(eV)^{2} term of the nonlinear conductance, defined in Eqs. (22)–(24) has a quadratic dependence on the bias and tunnel asymmetries of the form, αdif2\alpha_{\mathrm{dif}}^{2}, αdifγdif\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}}, and γdif2\gamma_{\mathrm{dif}}^{2}. In particular, the order γdif2\gamma_{\mathrm{dif}}^{2} term, which is absent in the SU(2) case [70], emerges for multilevel quantum dots with N3N\geq 3 and plays an important role in the Kondo states with no elelctron-hole symmetry. In this section, we describe some limiting cases of CV(3)C_{V}^{(3)}, and then discuss the NRG results for SU(4) and SU(6) quantum dots.

V.1 Behavior of CV(3)C_{V}^{(3)} in some limiting cases

The two-body and three-body parts of CV(3)C_{V}^{(3)} defined in Eqs. (22)–(24) takes the following values in the strong-coupling case where the Wilson ratio reaches the saturated value K~1.0\widetilde{K}\to 1.0 and the three-body correlations functions show the properties ΘIΘ~IIΘ~III\Theta_{\mathrm{I}}\simeq-\widetilde{\Theta}_{\mathrm{II}}\simeq\widetilde{\Theta}_{\mathrm{III}} shown in Figs. 14,

WV\displaystyle W_{V} K~1cos2δ[1+5N1+3(N2)N1γdif2],\displaystyle\xrightarrow{\,\widetilde{K}\to 1\,}\,-\cos 2\delta\left[1+\frac{5}{N-1}+\frac{3(N-2)}{N-1}\gamma_{\mathrm{dif}}^{2}\right]\,, (36)
ΘV\displaystyle\Theta_{V} ΘIΘ~IIΘ~III2[13γdif2]ΘI.\displaystyle\xrightarrow{\,\Theta_{\mathrm{I}}\simeq-\widetilde{\Theta}_{\mathrm{II}}\simeq\widetilde{\Theta}_{\mathrm{III}}\,}\,-2\,\Bigl{[}1-3\gamma_{\mathrm{dif}}^{2}\Bigr{]}\,\Theta_{\mathrm{I}}\,. (37)

Our result in this limit is consistent with the corresponding result for the SU(NN) Kondo model, given in Eqs. (22) and (32) of Ref. 40. Note that their notation and our one in the strong-interaction limit correspond to each other such that α1/(πTK)χσσ\alpha_{1}/(\pi T_{K})\Leftrightarrow\chi_{\sigma\sigma}, and α2/(πTK2)12χσσσ[3]\alpha_{2}/(\pi T_{K}^{2})\Leftrightarrow-\frac{1}{2}\,\chi_{\sigma\sigma\sigma}^{[3]}.

In the limit of |ξd||\xi_{d}|\to\infty, the occupation number reaches nd0\langle n_{d}\rangle\to 0 or NN, and the correlation functions take the noninteracting values, and thus

WV\displaystyle W_{V} |ξd|(1+3αdif2),\displaystyle\xrightarrow{\,|\xi_{d}|\to\infty\,}\,-\bigl{(}1+3\alpha_{\mathrm{dif}}^{2}\bigr{)}\,, (38)
ΘV\displaystyle\Theta_{V} |ξd|2(1+3αdif2),\displaystyle\xrightarrow{\,|\xi_{d}|\to\infty\,}\,-2\bigl{(}1+3\alpha_{\mathrm{dif}}^{2}\bigr{)}\,, (39)
CV(3)\displaystyle C_{V}^{(3)} |ξd|3π264(1+3αdif2).\displaystyle\xrightarrow{\,|\xi_{d}|\to\infty\,}\,-\frac{3\pi^{2}}{64}\,\bigl{(}1+3\alpha_{\mathrm{dif}}^{2}\bigr{)}\,. (40)

Note that in this limit, the three-body correlations are given by ΘI2\Theta_{\mathrm{I}}\to-2, Θ~II0\widetilde{\Theta}_{\mathrm{II}}\to 0, and Θ~III0\widetilde{\Theta}_{\mathrm{III}}\to 0 [see Appendix B in Ref. 70].

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 9: Coefficient CV(3)C_{V}^{(3)} is plotted vs ξd/U\xi_{d}/U for the case where both tunnel couplings and bias voltages are symmetric, γdif=0.0\gamma_{\mathrm{dif}}=0.0 and αdif=0.0\alpha_{\mathrm{dif}}=0.0. Upper panels show the total value of CV(3)=(π2/64)(WV+ΘV)C_{V}^{(3)}=(\pi^{2}/64)(W_{V}+\Theta_{V}) together with the two-body WVW_{V} and three-body ΘV\Theta_{V} components, (a) for SU(4) and (b) for SU(6) quantum dots choosing a large interaction U/(πΔ)=5.0U/(\pi\Delta)=5.0. Bottom panels show UU dependence of CV(3)C_{V}^{(3)}: (c) U/(πΔ)=1/3U/(\pi\Delta)=1/3, 2/32/3, 5/35/3, 10/310/3, 5.05.0 for SU(4), and (d) U/(πΔ)=2/5U/(\pi\Delta)=2/5, 1.01.0, 2.02.0,5.05.0 for SU(6).

V.2 CV(3)C_{V}^{(3)} for symmetric tunnel coupling and symmetric bias voltage γdif=αdif=0\gamma_{\mathrm{dif}}=\alpha_{\mathrm{dif}}=0

We describe here previous results obtained for symmetric tunnel coupling and bias voltage γdif=αdif=0\gamma_{\mathrm{dif}}=\alpha_{\mathrm{dif}}=0 [53] in order to clarify how the breaking of these symmetries affects the CV(3)C_{V}^{(3)}.

Figures 9(a) and (b) show that CV(3)C_{V}^{(3)}, WVW_{V} and ΘV\Theta_{V} for αdif=γdif=0.0\alpha_{\mathrm{dif}}=\gamma_{\mathrm{dif}}=0.0 as functions of the gate voltage ξd\xi_{d} for strong interaction U/(πΔ)=5.0U/(\pi\Delta)=5.0 for SU(4) and SU(6) quantum dots. In the strong-coupling limit region |ξd|(N2)U/2|\xi_{d}|\lesssim(N-2)U/2, CV(3)C_{V}^{(3)} takes plateau structures of the height (64/π2)CV(3)2.67(64/\pi^{2})\,C_{V}^{(3)}\simeq 2.67 for SU(4), and 2.02.0 for SU(6). In particular, the plateau at half filling, i.e., at ξd=0\xi_{d}=0, is caused by the two-body correlation WVW_{V}. In contrast, the plateau around the Kondo state with the fillings of nd1.0\langle n_{d}\rangle\simeq 1.0 and N1N-1, i.e., corresponding to the one electron and one hole occupancies, are caused by the three-body correlation ΘV\Theta_{V}. Specifically, among three independent components defined in Eqs. (25)–(27), Θ~II\widetilde{\Theta}_{\mathrm{II}} determines the peak structure of ΘV\Theta_{V} since ΘV2Θ~II\Theta_{V}\simeq-2\widetilde{\Theta}_{\mathrm{II}} owing to the property ΘIΘII\Theta_{\mathrm{I}}\simeq-\Theta_{\mathrm{II}} in the strongly correlated region. In the limit of |ξd||\xi_{d}|\to\infty outside the plateau, the coefficient approaches the noninteracting value (64/π2)CV(3)3(64/\pi^{2})\,C_{V}^{(3)}\to-3, WV1.0W_{V}\to-1.0, and ΘV2.0\Theta_{V}\to-2.0, given by Eqs. (38)–(40).

In Figs. 9(c) and (d), CV(3)C_{V}^{(3)} for αdif=γdif=0.0\alpha_{\mathrm{dif}}=\gamma_{\mathrm{dif}}=0.0 is plotted as a function of ξd\xi_{d}, varying interaction strengths (c) U/(πΔ)=1/3U/(\pi\Delta)=1/3, 2/32/3, 5/35/3, 10/310/3, 5.05.0 for SU(4), and (d) U/(πΔ)=2/5U/(\pi\Delta)=2/5, 1.01.0, 2.02.0, 5.05.0 for SU(6). The plateau structure and the peak at the edge of the plateau of CV(3)C_{V}^{(3)} as shown above appear as UU increases in the strong-coupling limit |ξd|(N2)U/2|\xi_{d}|\lesssim(N-2)U/2.

V.3 Effects of tunnel asymmetries γdif0\gamma_{\mathrm{dif}}\neq 0 on CV(3)C_{V}^{(3)} for symmetric bias voltages αdif=0\alpha_{\mathrm{dif}}=0

We next consider the effect of tunnel asymmetry on CV(3)C_{V}^{(3)}, setting bias voltages to be symmetric αdif=0\alpha_{\mathrm{dif}}=0. In this case, among the three types of quadratic terms αdif2\alpha_{\mathrm{dif}}^{2}, αdifγdif\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}}, and γdif2\gamma_{\mathrm{dif}}^{2} in Eqs. (22)–(24), the only γdif2\gamma_{\mathrm{dif}}^{2} term remains finite and contributes to CV(3)C_{V}^{(3)}:

WV\displaystyle W_{V} αdif=0cos2δ[1+{5N1+3(N2)N1γdif2}K~2],\displaystyle\xrightarrow{\alpha_{\mathrm{dif}}=0\,}\,-\cos 2\delta\biggl{[}1+\biggl{\{}\frac{5}{N-1}+\frac{3(N-2)}{N-1}\,\gamma_{\mathrm{dif}}^{2}\biggr{\}}\widetilde{K}^{2}\biggr{]}, (41)
ΘV\displaystyle\Theta_{V} αdif=0ΘI+3Θ~II+6γdif2Θ~III.\displaystyle\xrightarrow{\alpha_{\mathrm{dif}}=0\,}\,\Theta_{\mathrm{I}}\,+3\,\widetilde{\Theta}_{\mathrm{II}}\,+6\,\gamma_{\mathrm{dif}}^{2}\,\widetilde{\Theta}_{\mathrm{III}}\,. (42)

The γdif2\gamma_{\mathrm{dif}}^{2} term emerges for N3N\geq 3. In particular, it couples to Θ~III\widetilde{\Theta}_{\mathrm{III}}, the three-body correlation between electrons in three different levels which does not contribute to CV(3)C_{V}^{(3)} for symmetric tunnel couplings.

In the strongly correlated region for large UU, the three-body part of CV(3)C_{V}^{(3)} takes a simplified form ΘV13γdif2\Theta_{V}\propto 1-3\gamma_{\mathrm{dif}}^{2} which does not depend on αdif\alpha_{\mathrm{dif}}, as shown in Eq. (37). Therefore, ΘV\Theta_{V} decreases as tunnel asymmetry which enters through γdif2\gamma_{\mathrm{dif}}^{2} increases, and vanishes ΘV0.0\Theta_{V}\simeq 0.0 at γdif=±1/3\gamma_{\mathrm{dif}}=\pm 1/\sqrt{3} (±0.577\simeq\pm 0.577). In particular, Θ~II\widetilde{\Theta}_{\mathrm{II}} has a wide plateau at the fillings nd1\langle n_{d}\rangle\simeq 1 and N1N-1 for large UU in Fig. 2(b) and Fig. 4(b), the contribution of the three-body correlations ΘV\Theta_{V} becomes significant at these fillings.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 10: Effects of tunnel asymmetry γdif\gamma_{\mathrm{dif}} on order (eV)3(eV)^{3} nonlinear current: (Top panels) CV(3)C_{V}^{(3)}, and the components (Upper middle panels) WVW_{V} and (Lower middle panels) ΘV\Theta_{V} defined in Eqs. (41) and (42) for symmetrical bias voltages αdif=0.0\alpha_{\mathrm{dif}}=0.0 are plotted vs ξd/U\xi_{d}/U, varying tunnel asymmetries, as for γdif=0.0\gamma_{\mathrm{dif}}=0.0 ( ), 0.20.2 (\triangledown), 0.50.5 (\square), 0.580.58 (×\times), 0.80.8 (\triangle). Interaction strength is chosen to be U/(πΔ)=5.0U/(\pi\Delta)=5.0 for (a)–(f): (Left panels) for SU(4) quantum dots, and (Right panels) for SU(6) quantum dots. Bottom panels: UU dependence of CV(3)C_{V}^{(3)}, calculated for a fixed large tunnel asymmetry γdif=0.8\gamma_{\mathrm{dif}}=0.8, (g) U/(πΔ)=1/3U/(\pi\Delta)=1/3 ( ), 2/32/3 (\diamond), 5/35/3 (\circ), 10/310/3 (\triangle), 5.05.0 (×\times) for SU(4), and (h) U/(πΔ)=2/5U/(\pi\Delta)=2/5 ( ), 1.01.0 (\diamond), 2.02.0 (\circ), 5.05.0 (\triangle) for SU(6).

The NRG results of CV(3)C_{V}^{(3)}, WVW_{V}, and ΘV\Theta_{V}, for the bias symmetric case αdif=0.0\alpha_{\mathrm{dif}}=0.0 are plotted vs ξd\xi_{d} in the top, upper-middle, and lower-middle panels of Fig. 10, for (left panels) SU(4) and (right panels) SU(6) quantum dots, choosing interaction strength to be U/(πΔ)=5.0U/(\pi\Delta)=5.0. Specifically, Fig. 10 (a)–(f) are obtained, varying tunnel asymmetries, as γdif=0.0\gamma_{\mathrm{dif}}=0.0, 0.20.2, 0.50.5, 0.580.58 and 0.80.8. The plateau of CV(3)C_{V}^{(3)} emerging at |ξd|U/2|\xi_{d}|\lesssim U/2 is due to the half-filling Kondo state ndN/2\langle n_{d}\rangle\simeq N/2. The height of this plateau increases with γdif\gamma_{\mathrm{dif}}, and approaches the upper bound (64/π2)CV(3)4+2/(N1)(64/\pi^{2})\,C_{V}^{(3)}\to 4+2/(N-1) in the limit γdif21\gamma_{\mathrm{dif}}^{2}\to 1. This structure is determined by the two-body part WVW_{V} as the three-body part ΘV\Theta_{V} vanishes around the elelctron-hole symmetric point ξd=0\xi_{d}=0.

We see that the plateau structure of CV(3)C_{V}^{(3)} at the Kondo states away from half filling depend sensitively on tunnel asymmetry. The positive plateau emerging at |ξd|(N2)U/2|\xi_{d}|\simeq(N-2)U/2 are due to the Kondo state of the filling nd1\langle n_{d}\rangle\simeq 1 and N1N-1. It disappears as γdif\gamma_{\mathrm{dif}} increases, and a negative dip develops for γdif21/3\gamma_{\mathrm{dif}}^{2}\geq 1/3. This behavior of CV(3)C_{V}^{(3)} reflects the evolution of the three-body part ΘV\Theta_{V}, which is determined by Eq. (37) in the strongly correlated region. Therefore, if CV(3)C_{V}^{(3)} is measured varying tunneling asymmetries γdif\gamma_{\mathrm{dif}}, the three-body correlation function ΘV\Theta_{V} and Θ~III\widetilde{\Theta}_{\mathrm{III}} can experimentally be deduced, using Eqs. (37) and (42).

In the limit of |ξd||\xi_{d}|\to\infty, the coefficients approach the noninteracting value: WV1W_{V}\to-1, Θ2\Theta\to-2, and (64/π2)CV(3)3(64/\pi^{2})\,C_{V}^{(3)}\to-3 for symmetric bias αdif=0\alpha_{\mathrm{dif}}=0.

Figures 10(g) and (h) compare CV(3)C_{V}^{(3)} for several different interaction strengths: (g) U/(πΔ)=1/3U/(\pi\Delta)=1/3, 2/32/3, 5/35/3, 10/310/3, 5.05.0 for SU(4), and (h) U/(πΔ)=2/5U/(\pi\Delta)=2/5, 1.01.0, 2.02.0, 5.05.0 for SU(6), choosing a large tunnel asymmetry γdif=0.8\gamma_{\mathrm{dif}}=0.8. The dip structure due to the Kondo state at the filling nd=1\langle n_{d}\rangle=1 and N1N-1 becomes clear as UU increases, as well as the plateau due to the half-filling Kondo state.

Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 11: Behavior of CV(3)C_{V}^{(3)} for two different tunnel couplings, (top panels) γdif=0.0\gamma_{\mathrm{dif}}=0.0 and (bottom panels) γdif=0.8\gamma_{\mathrm{dif}}=0.8, is plotted as a function of ξd\xi_{d}, varying bias asymmetries αdif=1.0\alpha_{\mathrm{dif}}=-1.0 (\bullet), 0.8-0.8 (\circ), 0.5-0.5 (×\times), 0.2-0.2 (\diamond), 0.00.0 (\blacksquare), 0.20.2 (\triangledown), 0.50.5 (\square), 0.80.8 (\blacktriangle) and 1.01.0 (\triangle). Interaction strength is fixed at U/(πΔ)=5.0U/(\pi\Delta)=5.0, for both (left panels) SU(4) and (right panels) SU(6) quantum dots.

V.4 CV(3)C_{V}^{(3)} under asymmetric bias voltages αdif0\alpha_{\mathrm{dif}}\neq 0

The coefficient CV(3)C_{V}^{(3)} depends on the bias and tunnel asymmetries through the terms of order αdif2\alpha_{\mathrm{dif}}^{2}, αdifγdif\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}}, and γdif2\gamma_{\mathrm{dif}}^{2}, in Eqs. (23) and (24). We have discussed in the role of the γdif2\gamma_{\mathrm{dif}}^{2} term of the tunnel asymmetry, setting the bias voltages to be symmetric αdif=0\alpha_{\mathrm{dif}}=0. In this section, we examine the contributions of the other two terms, αdif2\alpha_{\mathrm{dif}}^{2} and αdifγdif\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}}.

V.4.1 Effects of bias asymmetry at small γdif=0\gamma_{\mathrm{dif}}=0 and large γdif=0.8\gamma_{\mathrm{dif}}=0.8 tunnel asymmetries

In Fig. 11, the coefficient CV(3)C_{V}^{(3)} is plotted vs ξd\xi_{d}, for (left panels) SU(4) and (right panels) SU(6) quantum dots, varying bias asymmetries αdif=0.0\alpha_{\mathrm{dif}}=0.0, ±0.2\pm 0.2, ±0.5\pm 0.5, ±0.8\pm 0.8 and ±1.0\pm 1.0. Specifically, two different tunnel couplings are examined. One (top panels) is the symmetric coupling γdif=0.0\gamma_{\mathrm{dif}}=0.0, at which the role of the αdif2\alpha_{\mathrm{dif}}^{2} term can be clarified. The other one (bottom panels) is a large asymmetric coupling γdif=0.8\gamma_{\mathrm{dif}}=0.8, for which all the quadratic terms, αdif2\alpha_{\mathrm{dif}}^{2}, αdifγdif\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}}, and γdif2\gamma_{\mathrm{dif}}^{2} contribute to CV(3)C_{V}^{(3)}.

As a large interaction U/(πΔ)=5.0U/(\pi\Delta)=5.0 is chosen for each of the panels, the coefficient CV(3)C_{V}^{(3)} does not depend on bias asymmetry αdif\alpha_{\mathrm{dif}} in the region of |ξd|(N1)U/2|\xi_{d}|\lesssim(N-1)U/2, and it takes the values determined by γdif2\gamma_{\mathrm{dif}}^{2} with Eqs. (36) and (37).

In contrast, the coefficient CV(3)C_{V}^{(3)} varies with αdifγdif\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}} in the valence fluctuation regime |ξd|(N1)U/2|\xi_{d}|\gtrsim(N-1)U/2. It approaches the asymptotic value which is given by Eq. (40) in the limit of |ξd||\xi_{d}|\to\infty. Since the effect of the bias asymmetry in the tunnel symmetric case γdif=0.0\gamma_{\mathrm{dif}}=0.0 is determined by the terms of αdif2\alpha_{\mathrm{dif}}^{2}, the results shown in Figs. 11(a) and (b) do not vary with the sign of the parameter αdif\alpha_{\mathrm{dif}}.

In Figs. 11(c) and (d), the coefficient CV(3)C_{V}^{(3)} has a sharp peak in the valence fluctuation region at ξd±(N1)U/2\xi_{d}\simeq\pm(N-1)U/2, which grows as bias asymmetry αdif\alpha_{\mathrm{dif}} increases. This is caused by the cross term with a positive sign αdifγdif>0\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}}>0: we will examine the contributions of this term to the peak structures more precisely in the following.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 12: Behavior of CV(3)C_{V}^{(3)} at large bias asymmetry αdif=1\alpha_{\mathrm{dif}}=1 is plotted vs ξd\xi_{d} for (left panels) SU(4) and (right panels) SU(6) quantum dots. Top panels: tunnel asymmetry is varied, as γdif=0.8\gamma_{\mathrm{dif}}=-0.8 (\circ), 0.5-0.5 (×\times), 0.2-0.2 (\diamond), 0.00.0 ( ), 0.20.2 (\triangledown), 0.50.5 (\square), 0.80.8 (\triangle). In addition, two-body part WVW_{V} (\square) and three-body part ΘV\Theta_{V} (\circ) are plotted together with CV(3)C_{V}^{(3)} (\blacktriangle) for two large opposite tunnel asymmetries: (upper middle panels) γdif=0.8\gamma_{\mathrm{dif}}=0.8, and (lower middle panels) γdif=0.8\gamma_{\mathrm{dif}}=-0.8. Interaction strength is chosen to be U/(πΔ)=5.0U/(\pi\Delta)=5.0 in (a)–(f). Bottom panels show the UU dependence of CV(3)C_{V}^{(3)} for γdif=0.8\gamma_{\mathrm{dif}}=0.8: (g) U/(πΔ)=1/3U/(\pi\Delta)=1/3, 2/32/3, 5/35/3, 10/310/3, 5.05.0 for SU(4), and (h) U/(πΔ)=2/5U/(\pi\Delta)=2/5, 1.01.0, 2.02.0, 5.05.0 for SU(6).

V.4.2 Effects of tunnel asymmetry γdif0\gamma_{\mathrm{dif}}\neq 0 on CV(3)C_{V}^{(3)} at large bias asymmetry αdif=1\alpha_{\mathrm{dif}}=1

The bias and tunnel asymmetries affect the coefficient CV(3)C_{V}^{(3)} through thus quadratic terms αdif2\alpha_{\mathrm{dif}}^{2}, αdifγdif\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}}, and γdif2\gamma_{\mathrm{dif}}^{2} in Eqs. (22)–(24). In order to clarify the contribution of the cross term, we set the bias asymmetry to be the upper-bound value αdif=1\alpha_{\mathrm{dif}}=1, which describes the situation where the bias asymmetry is maximized by grounding the right leads. Therefore, in this case WVW_{V} and ΘV\Theta_{V} can be expressed in the following form,

WVαdif=1\displaystyle W_{V}\xrightarrow{\,\alpha_{\mathrm{dif}}=1\,} cos2δ[46(1+γdif)K~\displaystyle\,-\cos 2\delta\Biggl{[}4-6\bigl{(}1+\gamma_{\mathrm{dif}}\bigr{)}\widetilde{K}
+{3N+2N1+6γdif+3(N2)N1γdif2}K~2],\displaystyle+\Biggl{\{}\frac{3N+2}{N-1}+6\,\gamma_{\mathrm{dif}}+\frac{3(N-2)}{N-1}\gamma_{\mathrm{dif}}^{2}\Biggr{\}}\widetilde{K}^{2}\Biggr{]}\,, (43)
ΘVαdif=1\displaystyle\Theta_{V}\xrightarrow{\,\alpha_{\mathrm{dif}}=1\,}   4[ΘI+6(1+γdif2)Θ~II+6(1+γdif2)2Θ~III].\displaystyle\,\,4\Biggl{[}\Theta_{\mathrm{I}}+6\biggl{(}\frac{1+\gamma_{\mathrm{dif}}}{2}\biggr{)}\widetilde{\Theta}_{\mathrm{II}}+6\biggl{(}\frac{1+\gamma_{\mathrm{dif}}}{2}\biggr{)}^{2}\widetilde{\Theta}_{\mathrm{III}}\Biggr{]}\,. (44)

Here, the cross term αdifγdif\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}} appears as linear order terms with respect to γdif\gamma_{\mathrm{dif}}.

Our discussion here is based on Fig. 12, in which the NRG results of CV(3)C_{V}^{(3)} for SU(4) quantum dots and those for SU(6) quantum dots are presented in the left and right panels, respectively. Top panels of Fig. 12 show CV(3)C_{V}^{(3)} as a function of the gate voltage ξd\xi_{d}, varying tunneling asymmetries γdif=0.0\gamma_{\mathrm{dif}}=0.0, ±0.2\pm 0.2, ±0.5\pm 0.5, ±0.8\pm 0.8, for a strong interaction strength U/(πΔ)=5.0U/(\pi\Delta)=5.0. The upper middle panels show CV(3)C_{V}^{(3)}, WVW_{V}, ΘV\Theta_{V} for large positive tunnel asymmetries γdif=0.8\gamma_{\mathrm{dif}}=0.8, and correspondingly lower middle panels show the ones for large negative tunnel asymmetries γdif=0.8\gamma_{\mathrm{dif}}=-0.8. Bottom panels of Fig. 12 show the results, calculated for several interaction strengths U/(πΔ)=1/3U/(\pi\Delta)=1/3, 2/32/3, 5/35/3, 10/310/3, 5.05.0 for SU(4), and U/(πΔ)=2/5U/(\pi\Delta)=2/5, 1.01.0, 2.02.0, 5.05.0 for SU(6), taking asymmetric tunnel coupling to be γdif=0.8\gamma_{\mathrm{dif}}=0.8.

In the strongly correlated region |ξd|(N1)U/2|\xi_{d}|\lesssim(N-1)U/2 for large UU, the result of CV(3)C_{V}^{(3)} in Figs. 12 (a) and (b) almost does not depend on whether or not αdif=1\alpha_{\mathrm{dif}}=1, and the behavior in this region is determined essentially by the γdif2\gamma_{\mathrm{dif}}^{2} term in Eqs. (36) and (37). The cross term αdifγdif\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}} becomes important at |ξd|(N1)U/2|\xi_{d}|\gtrsim(N-1)U/2, outside the correlated region.

In particular, CV(3)C_{V}^{(3)} takes a sharp peak near ξd(N1)U/2\xi_{d}\simeq(N-1)U/2 in the valence fluctuation regime for large positive γdif\gamma_{\mathrm{dif}}, i.e., αdifγdif>0\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}}>0 at which bias and tunneling asymmetries cooperatively enhance the charge transfer from one of the electrodes. The results of WVW_{V} and ΘV\Theta_{V}, plotted for γdif=0.8\gamma_{\mathrm{dif}}=0.8 in Figs. 12 (c)-(d), show the sharp peak structure is mainly due to the three-body part ΘV\Theta_{V} and two-body part has a small constructive peak at the same position. In contrast, for tunnel asymmetry in the opposite direction γdif=0.8\gamma_{\mathrm{dif}}=-0.8, each of the two components, WVW_{V} and ΘV\Theta_{V}, does not have a peak in Figs. 12 (e)-(f).

We consider more precisely the peak that emerged in the three-body part ΘV\Theta_{V}, taking the limit of γdif1\gamma_{\mathrm{dif}}\to 1 in Eq. (44) 111 This limit of strong tunnel asymmetry γdif±1\gamma_{\mathrm{dif}}\to\pm 1 is meaningful for investigating asymptotic behavior of CV(3)C_{V}^{(3)} although it represents the situation where one of the leads is disconnected and the current is determined by Eq. (20).,

ΘVαdif=1,γdif1 4(ΘI+6Θ~II+6Θ~III).\displaystyle\Theta_{V}\xrightarrow{\,\alpha_{\mathrm{dif}}=1,\,\gamma_{\mathrm{dif}}\to 1\,}\,4\,\Bigl{(}\Theta_{\mathrm{I}}+6\,\widetilde{\Theta}_{\mathrm{II}}+6\,\widetilde{\Theta}_{\mathrm{III}}\Bigr{)}\,. (45)

Note that each of the three-body correlation functions of the SU(NN) symmetric quantum dots has a definite sign: ΘI<0\Theta_{\mathrm{I}}<0, Θ~II>0\widetilde{\Theta}_{\mathrm{II}}>0, and Θ~III<0\widetilde{\Theta}_{\mathrm{III}}<0, as shown in Figs. 2 and 4. Specifically, in the valence fluctuation regime, the positive contribution of 6Θ~II6\,\widetilde{\Theta}_{\mathrm{II}} in Eq. (45) becomes greater than the negative contribution of ΘI+6Θ~III\Theta_{\mathrm{I}}+6\,\widetilde{\Theta}_{\mathrm{III}}, and the difference between these components yields a sharp peak of ΘV\Theta_{V} at |ξd|(N1)U/2|\xi_{d}|\gtrsim(N-1)U/2. The height measured from the base value of ΘV\Theta_{V} that is defined at the Kondo state in the close vicinity of the valence fluctuation region increases with NN. Note that in the Kondo state, the three-body correlation functions have a property Θ~IΘ~IIΘ~III,\widetilde{\Theta}_{\mathrm{I}}\simeq-\widetilde{\Theta}_{\mathrm{II}}\simeq\widetilde{\Theta}_{\mathrm{III}}, and ΘV\Theta_{V} in this limit γdif1\gamma_{\mathrm{dif}}\to 1 takes a negative value ΘV4ΘI<0\Theta_{V}\to 4\Theta_{\mathrm{I}}<0. In the opposite limit of the tunnel asymmetries γdif1\gamma_{\mathrm{dif}}\to-1, the cross term is negative αdifγdif<0\alpha_{\mathrm{dif}}\,\gamma_{\mathrm{dif}}<0. Therefore, the three-body contribution also becomes negative ΘV4ΘI<0\Theta_{V}\to 4\Theta_{\mathrm{I}}<0, and CV(3)C_{V}^{(3)} monotonically decreases at |ξd|(N1)U/2|\xi_{d}|\gtrsim(N-1)U/2.

We also examine how the peak structure evolves with UU in Figs. 12(g) and (h). The results show that the dip structure in the SU(NN) Kondo regime near |ξd|(N1)U/2|\xi_{d}|\simeq(N-1)U/2 and the sharp peak structure in the valence fluctuation regime |ξd|(N1)U/2|\xi_{d}|\gtrsim(N-1)U/2 clearer as interaction UU increases.

VI Conclusion

We have derived the exact low-bias expansion formula of the differential conductance dI/dVdI/dV through a multilevel Anderson impurity up to terms of order (eV)2(eV)^{2}, without assuming symmetries in tunnel couplings or bias voltages. It is applicable to a wide class of quantum dots with arbitrary energy level ϵdσ\epsilon_{d\sigma}. The expansion coefficients are expressed in terms of the phase shift δσ\delta_{\sigma}, linear susceptibilities χσσ\chi_{\sigma\sigma^{\prime}}, and three-body correlation functions χσσσ′′[3]\chi_{\sigma\sigma^{\prime}\sigma^{\prime\prime}}^{[3]}, defined with respect to the equilibrium ground state.

The tunnel and bias asymmetries enter the transport coefficients through the parameters γdif=(ΓLΓR)/(ΓL+ΓR)\gamma_{\mathrm{dif}}=(\Gamma_{L}-\Gamma_{R})/(\Gamma_{L}+\Gamma_{R}) and αdif=(μL+μR2EF)/(μLμR)\alpha_{\mathrm{dif}}=(\mu_{L}+\mu_{R}-2E_{F})/(\mu_{L}-\mu_{R}). In contrast to the linear conductance which depends only on the tunnel asymmetry through the prefactor g0=e2h 4ΓLΓR/(ΓL+ΓR)2g_{0}=\frac{e^{2}}{h}\,4\Gamma_{L}\Gamma_{R}/(\Gamma_{L}+\Gamma_{R})^{2}, the nonlinear terms cV,σ(2)c_{V,\sigma}^{(2)} and cV,σ(3)c_{V,\sigma}^{(3)}, given in Eqs.  (B.5) and (B.6), depend on both asymmetries. The γdif\gamma_{\mathrm{dif}}-dependence enters additionally through the self-energy corrections due to the Coulomb interactions, and the αdif\alpha_{\mathrm{dif}}-dependence arises also through a shift of bias window. In particular, cV,σ(2)c_{V,\sigma}^{(2)} emerges when tunnel couplings and/or bias voltages are asymmetrical, and is given by a linear combination of γdif\gamma_{\mathrm{dif}} and αdif\alpha_{\mathrm{dif}}.

We have explored the behaviors of these coefficients of SU(NN) quantum dots of N=4N=4 and 66 with the NRG, in a wide range of parameter space, the varying gate voltage ξd\xi_{d} and interaction UU as well as γdif\gamma_{\mathrm{dif}} and αdif\alpha_{\mathrm{dif}}. In particular, for large UU, transport exhibits quite different behaviors depending on electron fillings, especially in the two regions: one is the strongly-correlated region 1ndN11\lesssim\langle n_{d}\rangle\lesssim N-1, and the other is the valence fluctuation region in which 0nd10\lesssim\langle n_{d}\rangle\lesssim 1 or N1ndNN-1\lesssim\langle n_{d}\rangle\lesssim N. The coefficient CV(2)C_{V}^{(2)} of the first nonlinear term of dI/dVdI/dV for SU(NN) quantum dots becomes almost independent of bias asymmetries in the strongly-correlated region as charge fluctuation is suppressed in this region. Conversely, it becomes less sensitive to tunnel asymmetries in the valence fluctuation region since interaction effects are suppressed as the filling approaches nd0\langle n_{d}\rangle\to 0 or NN.

The three-body correlation functions contribute to the order (eV)2(eV)^{2} nonlinear term of dI/dVdI/dV, especially in the SU(NN) Kondo regime other than the half-filled one occurring in electron-hole asymmetric cases, and also in the valence fluctuation region. The coefficient CV(3)C_{V}^{(3)} of the order (eV)2(eV)^{2} term shows the quadratic αdif2\alpha_{\mathrm{dif}}^{2}, αdifγdif\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}} and γdif2\gamma_{\mathrm{dif}}^{2} dependences on the bias and tunnel asymmetries.

In particular, the γdif2\gamma_{\mathrm{dif}}^{2} term, which is absent in the SU(2) case, emerges for multilevel quantum dots with N3N\geq 3, and it couples to a three-body correlation between electrons occupying three different local levels: χσσσ′′[3]\chi_{\sigma\sigma^{\prime}\sigma^{\prime\prime}}^{[3]} for σσσ′′σ\sigma\neq\sigma^{\prime}\neq\sigma^{\prime\prime}\neq\sigma. We have found that, as γdif2\gamma_{\mathrm{dif}}^{2} increases, the structure of SU(NN) Kondo plateaus of CV(3)C_{V}^{(3)} at nd1\langle n_{d}\rangle\simeq 1 and N1N-1 fillings vary significantly from the one for symmetric junctions with γdif=αdif=0\gamma_{\mathrm{dif}}=\alpha_{\mathrm{dif}}=0. It suggests that the tunnel asymmetries could be used as a sensitive probe for observing three-body correlations in the SU(NN) Kondo states.

The cross term αdifγdif\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}} plays an important role, especially in the valence fluctuation region. It yields a sharp peak of CV(3)C_{V}^{(3)} when αdifγdif>0\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}}>0, i.e., in the case at which the tunneling and bias asymmetries cooperatively enhance the charge transfer from one of the electrodes. This is caused by a constructive enhancement of the three-independent components of the three-body correlation function of SU(NN) quantum dots: χσσσ[3]\chi_{\sigma\sigma\sigma}^{[3]}, χσσσ[3]\chi_{\sigma\sigma^{\prime}\sigma^{\prime}}^{[3]}, and χσσσ′′[3]\chi_{\sigma\sigma^{\prime}\sigma^{\prime\prime}}^{[3]}. Our results indicate that these three-independent components can separately be deduced if CV(3)C_{V}^{(3)} is measured varying tunneling asymmetries γdif\gamma_{\mathrm{dif}} and bias asymmetries αdif\alpha_{\mathrm{dif}}.

Acknowledgements.
This work was supported by JSPS KAKENHI Grant Nos. JP18K03495, JP18J10205, JP21K03415, and JP23K03284, and by JST CREST Grant No. JPMJCR1876. KM was supported by JST, the establishment of university fellowships towards the creation of science technology innovation, Grant Number JPMJFS2138.

Appendix A Fermi-liquid relations

The retarded Green’s function defined in Eq. (10) can be expressed in the form,

Gσr(ω)=\displaystyle G_{\sigma}^{r}(\omega)\,= 1ωϵdσ+iΔΣσr(ω).\displaystyle\ \frac{1}{\omega-\epsilon_{d\sigma}+i\Delta-\Sigma_{\sigma}^{r}(\omega)}\,. (46)

Specifically, the ground-state properties and the leading Fermi-liquid corrections are determined by the low-frequency behavior of the equilibrium self-energy Σeq,σr(ω)Σσr(ω)|T=eV=0\Sigma_{\mathrm{eq},\sigma}^{r}(\omega)\equiv\left.\Sigma_{\sigma}^{r}(\omega)\right|_{T=eV=0}, or the Green’s function:

Gσr(ω)zσωϵ~dσ+iΔ~σ.\displaystyle G_{\sigma}^{r}(\omega)\,\simeq\,\frac{z_{\sigma}}{\omega-\widetilde{\epsilon}_{d\sigma}+i\widetilde{\Delta}_{\sigma}}\,. (47)

Here, the renormalized parameters are defined by

ϵ~dσ\displaystyle\widetilde{\epsilon}_{d\sigma}\,\equiv zσ[ϵdσ+Σeq,σr(0)]=Δ~σcotδσ,\displaystyle\ z_{\sigma}\left[\epsilon_{d\sigma}+\Sigma_{\mathrm{eq},\sigma}^{r}(0)\right]\,=\,\widetilde{\Delta}_{\sigma}\cot\delta_{\sigma}\,,
Δ~σ\displaystyle\widetilde{\Delta}_{\sigma}\,\equiv zσΔ,1zσ 1Σeq,σr(ω)ω|ω=0.\displaystyle\ z_{\sigma}\Delta,\qquad\frac{1}{z_{\sigma}}\,\equiv\ 1-\left.\frac{\partial\Sigma_{\mathrm{eq},\sigma}^{r}(\omega)}{\partial\omega}\right|_{\omega=0}.\rule{0.0pt}{17.07182pt} (48)

Furthermore, the renormalization factor zσz_{\sigma} and the derivative of Σeq,σr(0)\Sigma_{\mathrm{eq},\sigma}^{r}(0) with respect to the impurity level ϵdσ\epsilon_{d\sigma^{\prime}} are related to each other through the Ward identity [4, 7]:

1zσ=χ~σσ,χ~σσδσσ+Σeq,σr(0)ϵdσ.\displaystyle\frac{1}{z_{\sigma}}\,=\,\widetilde{\chi}_{\sigma\sigma}\,,\qquad\quad\widetilde{\chi}_{\sigma\sigma^{\prime}}\,\equiv\,\delta_{\sigma\sigma^{\prime}}+\frac{\partial\Sigma_{\mathrm{eq},\sigma}^{r}(0)}{\partial\epsilon_{d\sigma^{\prime}}}\,. (49)

Note that χ~σσ\widetilde{\chi}_{\sigma\sigma^{\prime}} corresponds to an enhancement factor for the linear susceptibilities defined in Eq. (8), i.e., χσσ=ndσ/ϵdσ=ρdσχ~σσ\chi_{\sigma\sigma^{\prime}}=-\partial\bigl{\langle}n_{d\sigma}\bigr{\rangle}/\partial\epsilon_{d\sigma^{\prime}}=\rho_{d\sigma}\widetilde{\chi}_{\sigma\sigma^{\prime}} at T=0T=0.

Recently, the Ward identity for the order ω2\omega^{2} real part of the self-energy has also been obtained, as [50, 51, 52, 53]

2ω2ReΣeq,σr(ω)|ω0=2Σeq,σr(0)ϵdσ2=χ~σσϵdσ.\displaystyle\left.\frac{\partial^{2}}{\partial\omega^{2}}\mathrm{Re}\,\Sigma_{\mathrm{eq},\sigma}^{r}(\omega)\right|_{\omega\to 0}\,=\,\frac{\partial^{2}\Sigma_{\mathrm{eq},\sigma}^{r}(0)}{\partial\epsilon_{d\sigma}^{2}}\,\ =\ \frac{\partial\widetilde{\chi}_{\sigma\sigma}}{\partial\epsilon_{d\sigma}}\,\,. (50)

This identity shows that the ω2\omega^{2} real part is determined by the intra-level component of the three-body correlation function χσσσ[3]\chi_{\sigma\sigma\sigma}^{[3]}. Physically, this term of the self-energy induces higher-order energy shifts for single-quasiparticle excitations.

Appendix B Derivation of cV,σ(2)c_{V,\sigma}^{(2)} and cV,σ(3)c_{V,\sigma}^{(3)}

We describe here the derivation of the coefficients cV,σ(2)c_{V,\sigma}^{(2)} and cV,σ(3)c_{V,\sigma}^{(3)} which appeared in Eq. (1). The low-energy asymptotic form of the retarded self-energy Σσr(ω)\Sigma_{\sigma}^{r}(\omega) for multi-orbital Anderson impurity model was derived up to terms of order ω2\omega^{2}, T2T^{2}, and (eV)2(eV)^{2} in previous work [53, 64],

ImΣσr(ω)=π21ρdσσ(σ)χσσ2\displaystyle\mathrm{Im}\,\Sigma_{\sigma}^{r}(\omega)\,=\,-\frac{\pi}{2}\frac{1}{\rho_{d\sigma}}\sum_{\sigma^{\prime}(\neq\sigma)}\chi_{\sigma\sigma^{\prime}}^{2}
×[(ωαeV)2+3ΓLΓR(ΓL+ΓR)2(eV)2+(πT)2]+,\displaystyle\qquad\times\Biggl{[}(\omega-\alpha eV)^{2}+\frac{3\Gamma_{L}\Gamma_{R}}{(\Gamma_{L}+\Gamma_{R})^{2}}(eV)^{2}+(\pi T)^{2}\Biggr{]}+\cdots\!, (51)
ReΣσr(ω)=Σeq,σr(0)σ(σ)χ~σσαeV+(1χ~σσ)ω\displaystyle\mathrm{Re}\,\Sigma_{\sigma}^{r}(\omega)\,=\ \Sigma_{\mathrm{eq},\sigma}^{r}(0)-\sum_{\sigma^{\prime}(\neq\sigma)}\widetilde{\chi}_{\sigma\sigma^{\prime}}\,\alpha\,eV+(1-\widetilde{\chi}_{\sigma\sigma})\,\omega
+161ρdσσ(σ)χσσϵdσ[3ΓLΓR(ΓL+ΓR)2(eV)2+(πT)2]\displaystyle\qquad\ +\frac{1}{6}\frac{1}{\rho_{d\sigma}}\sum_{\sigma^{\prime}(\neq\sigma)}\frac{\partial\chi_{\sigma\sigma^{\prime}}}{\partial\epsilon_{d\sigma^{\prime}}}\Biggl{[}\frac{3\Gamma_{L}\Gamma_{R}}{(\Gamma_{L}+\Gamma_{R})^{2}}(eV)^{2}+(\pi T)^{2}\Biggr{]}
+12χ~σσϵdσω2+σ(σ)χ~σσϵdσαeVω\displaystyle\qquad\ +\frac{1}{2}\frac{\partial\widetilde{\chi}_{\sigma\sigma}}{\partial\epsilon_{d\sigma}}\,\omega^{2}+\sum_{\sigma^{\prime}(\neq\sigma)}\frac{\partial\widetilde{\chi}_{\sigma\sigma^{\prime}}}{\partial\epsilon_{d\sigma}}\,\alpha\,eV\omega\,
+12σ(σ)σ′′(σ)χ~σσϵdσ′′α2(eV)2+.\displaystyle\qquad\ +\frac{1}{2}\sum_{\sigma^{\prime}(\neq\sigma)}\sum_{\sigma^{\prime\prime}(\neq\sigma)}\frac{\partial\widetilde{\chi}_{\sigma\sigma^{\prime}}}{\partial\epsilon_{d\sigma^{\prime\prime}}}\,\alpha^{2}(eV)^{2}+\cdots\,. (52)

Here, α\alpha is a parameter defined in a way such that αeV(ΓLμL+ΓRμR)/(ΓL+ΓR)\alpha\,eV\equiv(\Gamma_{L}\,\mu_{L}+\Gamma_{R}\,\mu_{R})/(\Gamma_{L}+\Gamma_{R}), and thus

α\displaystyle\alpha\,\equiv αLΓLαRΓRΓL+ΓR=12(αdif+γdif).\displaystyle\ \frac{\alpha_{L}\Gamma_{L}-\alpha_{R}\Gamma_{R}}{\Gamma_{L}+\Gamma_{R}}\,=\,\frac{1}{2}\Bigl{(}\alpha_{\mathrm{dif}}+\gamma_{\mathrm{dif}}\Bigr{)}\,. (53)

The spectral function can be deduced exactly up to terms of order ω2\omega^{2}, T2T^{2}, and (eV)2(eV)^{2} from the above results of Σσr(ω)\Sigma_{\sigma}^{r}(\omega) using Eq. (46):

πΔAσ(ω)=\displaystyle\pi\Delta A_{\sigma}(\omega)\,= sin2δσ+πsin2δσ[χσσω+σ(σ)χσσ12(αdif+γdif)eV]\displaystyle\ \sin^{2}\delta_{\sigma}+\pi\sin 2\delta_{\sigma}\,\biggl{[}\,\chi_{\sigma\sigma}\,\omega+\sum_{\sigma^{\prime}(\neq\sigma)}\chi_{\sigma\sigma^{\prime}}\ \frac{1}{2}\bigl{(}\alpha_{\mathrm{dif}}+\gamma_{\mathrm{dif}}\bigr{)}\,eV\,\biggr{]}
+π2[cos2δσ(χσσ2+12σ(σ)χσσ2)sin2δσ2πχσσσ[3]]ω2\displaystyle+\pi^{2}\biggl{[}\,\cos 2\delta_{\sigma}\biggl{(}\chi_{\sigma\sigma}^{2}+\frac{1}{2}\sum_{\sigma^{\prime}(\neq\sigma)}\chi_{\sigma\sigma^{\prime}}^{2}\biggr{)}-\frac{\sin 2\delta_{\sigma}}{2\pi}\chi_{\sigma\sigma\sigma}^{[3]}\,\biggr{]}\,\omega^{2}
+π2[cos2δσσ(σ)(χσσχσσ12χσσ2)sin2δσ2πσ(σ)χσσσ[3]](αdif+γdif)ωeV\displaystyle+\pi^{2}\biggl{[}\,\cos 2\delta_{\sigma}\sum_{\sigma^{\prime}(\neq\sigma)}\biggl{(}\chi_{\sigma\sigma}\chi_{\sigma\sigma^{\prime}}-\frac{1}{2}\chi_{\sigma\sigma^{\prime}}^{2}\biggr{)}-\frac{\sin 2\delta_{\sigma}}{2\pi}\sum_{\sigma^{\prime}(\neq\sigma)}\chi_{\sigma\sigma\sigma^{\prime}}^{[3]}\,\biggr{]}\bigl{(}\alpha_{\mathrm{dif}}+\gamma_{\mathrm{dif}}\bigr{)}\,\omega\,eV
+π23σ(σ)[32cos2δσχσσ2sin2δσ2πχσσσ[3]][34(1+2αdifγdif+αdif2)(eV)2+(πT)2]\displaystyle+\frac{\pi^{2}}{3}\sum_{\sigma^{\prime}(\neq\sigma)}\biggl{[}\,\frac{3}{2}\cos 2\delta_{\sigma}\chi_{\sigma\sigma^{\prime}}^{2}-\frac{\sin 2\delta_{\sigma}}{2\pi}\chi_{\sigma\sigma^{\prime}\sigma^{\prime}}^{[3]}\,\biggr{]}\biggl{[}\,\frac{3}{4}\biggl{(}1+2\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}}+\alpha_{\mathrm{dif}}^{2}\biggr{)}(eV)^{2}+(\pi T)^{2}\,\biggr{]}
+π23σ(σ)σ′′(σ,σ)[cos2δσχσσχσσ′′sin2δσ2πχσσσ′′[3]]34(γdif2+2αdifγdif+αdif2)(eV)2+.\displaystyle+\frac{\pi^{2}}{3}\sum_{\sigma^{\prime}(\neq\sigma)}\sum_{\sigma^{\prime\prime}(\neq\sigma,\sigma^{\prime})}\biggl{[}\,\cos 2\delta_{\sigma}\chi_{\sigma\sigma^{\prime}}\chi_{\sigma\sigma^{\prime\prime}}-\frac{\sin 2\delta_{\sigma}}{2\pi}\chi_{\sigma\sigma^{\prime}\sigma^{\prime\prime}}^{[3]}\,\biggr{]}\ \frac{3}{4}\biggl{(}\gamma_{\mathrm{dif}}^{2}\!+2\alpha_{\mathrm{dif}}\gamma_{\mathrm{dif}}+\alpha_{\mathrm{dif}}^{2}\biggr{)}(eV)^{2}\ +\ \cdots\,. (54)

Substituting this low-energy asymptotic form into the the Landauer-type formula in Eq. (14), we obtain the exact expression for the coefficients cV,σ(2)c_{V,\sigma}^{(2)} and cV,σ(3)c_{V,\sigma}^{(3)} for the differential conductance at T=0T=0 defined in Eq. (1):

cV,σ(2)=\displaystyle c_{V,\sigma}^{(2)}\,= πsin2δσ[αdifχσσ+(αdif+γdif)σ(σ)χσσ],\displaystyle\ \pi\sin 2\delta_{\sigma}\left[\,\alpha_{\mathrm{dif}}\,\chi_{\sigma\sigma}\,+\,\bigl{(}\alpha_{\mathrm{dif}}+\gamma_{\mathrm{dif}}\bigr{)}\sum_{\sigma^{\prime}(\neq\sigma)}\chi_{\sigma\sigma^{\prime}}\right]\,, (55)
cV,σ(3)=\displaystyle c_{V,\sigma}^{(3)}\,= π24[cos2δσ{(1+3αdif2)χσσ2+(53γdif2)σ(σ)χσσ2+ 6αdif(αdif+γdif)χσσσ(σ)χσσ\displaystyle\ \frac{\pi^{2}}{4}\Biggl{[}\,-\cos 2\delta_{\sigma}\Biggl{\{}\bigl{(}1+3\alpha_{\mathrm{dif}}^{2}\bigr{)}\,\chi_{\sigma\sigma}^{2}\,+\,\left(5-3\gamma_{\mathrm{dif}}^{2}\right)\sum_{\sigma^{\prime}(\neq\sigma)}\chi_{\sigma\sigma^{\prime}}^{2}\,+\,6\alpha_{\mathrm{dif}}\,\bigl{(}\alpha_{\mathrm{dif}}+\gamma_{\mathrm{dif}}\bigr{)}\,\chi_{\sigma\sigma}\!\sum_{\sigma^{\prime}(\neq\sigma)}\chi_{\sigma\sigma^{\prime}}
+3(αdif+γdif)2σ(σ)σ′′(σ)χσσχσσ′′}\displaystyle\qquad\qquad\qquad\quad+3\bigl{(}\alpha_{\mathrm{dif}}+\gamma_{\mathrm{dif}}\bigr{)}^{2}\sum_{\sigma^{\prime}(\neq\sigma)}\sum_{\sigma^{\prime\prime}(\neq\sigma)}\chi_{\sigma\sigma^{\prime}}\chi_{\sigma\sigma^{\prime\prime}}\Biggr{\}}
+sin2δσ2π{(1+3αdif2)χσσσ[3]+ 6αdif(αdif+γdif)σ(σ)χσσσ[3]+ 3(1γdif2)σ(σ)χσσσ[3]\displaystyle\qquad+\,\frac{\sin 2\delta_{\sigma}}{2\pi}\Biggl{\{}\bigl{(}1+3\alpha_{\mathrm{dif}}^{2}\bigr{)}\,\chi_{\sigma\sigma\sigma}^{[3]}\,+\,6\alpha_{\mathrm{dif}}\,\left(\alpha_{\mathrm{dif}}+\gamma_{\mathrm{dif}}\right)\,\sum_{\sigma^{\prime}(\neq\sigma)}\chi_{\sigma\sigma\,\sigma^{\prime}}^{[3]}\,+\,3\left(1-\gamma_{\mathrm{dif}}^{2}\right)\sum_{\sigma^{\prime}(\neq\sigma)}\chi_{\sigma\sigma^{\prime}\sigma^{\prime}}^{[3]}
+3(αdif+γdif)2σ(σ)σ′′(σ)χσσσ′′[3]}].\displaystyle\qquad\qquad\qquad\quad\ +3\left(\alpha_{\mathrm{dif}}+\gamma_{\mathrm{dif}}\right)^{2}\sum_{\sigma^{\prime}(\neq\sigma)}\sum_{\sigma^{\prime\prime}(\neq\sigma)}\chi_{\sigma\sigma^{\prime}\sigma^{\prime\prime}}^{[3]}\Biggr{\}}\,\Biggr{]}\,. (56)

Note that cV,σ(2)c_{V,\sigma}^{(2)} is proportional to the derivative of the density of state as sin2δσ=Δρdσ/χσσ\sin 2\delta_{\sigma}=\Delta\rho_{d\sigma}^{\prime}/\chi_{\sigma\sigma} from Eq. (13). The coefficients cV,σ(2)c_{V,\sigma}^{(2)} and cV,σ(3)c_{V,\sigma}^{(3)} have odd and even inversion-symmetrical properties, respectively. Namely, if the left and right leads are inverted together with their tunnel couplings and chemical potentials, sign of cV,σ(2)c_{V,\sigma}^{(2)} changes while cV,σ(3)c_{V,\sigma}^{(3)} does not:

cV,σ(2)(αdif,γdif)=cV,σ(2)(αdif,γdif),\displaystyle c_{V,\sigma}^{(2)}(\alpha_{\mathrm{dif}},\,\gamma_{\mathrm{dif}})\,=\,-c_{V,\sigma}^{(2)}(-\alpha_{\mathrm{dif}},\,-\gamma_{\mathrm{dif}})\,, (57)
cV,σ(3)(αdif,γdif)=cV,σ(3)(αdif,γdif).\displaystyle c_{V,\sigma}^{(3)}(\alpha_{\mathrm{dif}},\,\gamma_{\mathrm{dif}})\,=\,c_{V,\sigma}^{(3)}(-\alpha_{\mathrm{dif}},\,-\gamma_{\mathrm{dif}})\,. (58)

Appendix C Behavior of three-body correlation functions for large UU

  Refer to captionRefer to captionRefer to captionRefer to caption

Figure 13: Correlation functions emerging on the right-hand side of Eqs. (59) and (60) are plotted vs ξd/U\xi_{d}/U, for N=4N=4. (a)-(b): the diagonal component (πΔ)2χσσσ[3](\pi\Delta)^{2}\chi_{\sigma\sigma\sigma}^{[3]}. (c): (πΔ)2χσσϵd(\pi\Delta)^{2}\frac{\partial\chi_{\sigma\sigma}}{\partial\epsilon_{d}}. (d): (πΔ)2(N1)χσσϵd(\pi\Delta)^{2}\,(N-1)\,\frac{\partial\chi_{\sigma\sigma^{\prime}}}{\partial\epsilon_{d}}. (b) represents an enlarged view of (a). In each of these panels, interaction strength is chosen to be U/(πΔ)=1/3U/(\pi\Delta)=1/3, 2/32/3, 5/35/3, 10/310/3, 5.05.0.

In the SU(NN) symmetric case, the two different off-diagonal components χσσσ[3]\chi_{\sigma\sigma^{\prime}\sigma^{\prime}}^{[3]} and χσσσ′′[3]\chi_{\sigma\sigma^{\prime}\sigma^{\prime\prime}}^{[3]} of the three-body correlation functions for σσσ′′σ\sigma\neq\sigma^{\prime}\neq\sigma^{\prime\prime}\neq\sigma can be expressed as a linear combination of the diagonal one χσσσ[3]\chi_{\sigma\sigma\sigma}^{[3]} and the derivative of the linear susceptibilities [44]:

(N1)χσσσ[3]\displaystyle(N-1)\chi_{\sigma\sigma^{\prime}\sigma^{\prime}}^{[3]} =χσσσ[3]+χσσϵd,\displaystyle=-\chi_{\sigma\sigma\sigma}^{[3]}+\frac{\partial\chi_{\sigma\sigma}}{\partial\epsilon_{d}}\,, (59)
(N1)(N2)2χσσσ′′[3]\displaystyle\frac{(N-1)(N-2)}{2}\chi_{\sigma\sigma^{\prime}\sigma^{\prime\prime}}^{[3]} =χσσσ[3]χσσϵd+N12χσσϵd.\displaystyle=\chi_{\sigma\sigma\sigma}^{[3]}-\frac{\partial\chi_{\sigma\sigma}}{\partial\epsilon_{d}}+\frac{N-1}{2}\frac{\partial\chi_{\sigma\sigma^{\prime}}}{\partial\epsilon_{d}}\,. (60)

Figure 13 compares the components emerging on the right-hand side for N=4N=4, i.e., χσσσ[3]\chi_{\sigma\sigma\sigma}^{[3]}, χσσϵd\frac{\partial\chi_{\sigma\sigma}}{\partial\epsilon_{d}}, and (N1)χσσϵd(N-1)\frac{\partial\chi_{\sigma\sigma^{\prime}}}{\partial\epsilon_{d}}, varying interaction strength UU. We can see that the diagonal component |χσσσ[3]||\chi_{\sigma\sigma\sigma}^{[3]}| becomes much larger than the derivative terms |χσσϵd||\frac{\partial\chi_{\sigma\sigma}}{\partial\epsilon_{d}}| and (N1)|χσσϵd|(N-1)\,|\frac{\partial\chi_{\sigma\sigma^{\prime}}}{\partial\epsilon_{d}}|, for strong interactions U/(πΔ)2.0U/(\pi\Delta)\gtrsim 2.0, over a wide parameter range (N1)U/2ξd(N1)U/2-(N-1)U/2\lesssim\xi_{d}\lesssim(N-1)U/2. Therefore, χσσσ[3]\chi_{\sigma\sigma\sigma}^{[3]} dominates on the right hand side of Eq. (59) and also on Eq. (60), and thus the corresponding dimensionless parameters show the property ΘIΘ~IIΘ~III\Theta_{\mathrm{I}}\simeq-\widetilde{\Theta}_{\mathrm{II}}\simeq\widetilde{\Theta}_{\mathrm{III}} described in Eq. (30) for the strong interaction region. We have confirmed that the same behavior occurs in the SU(6) symmetric case.

References