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

Nonlocal Andreev transport through a quantum dot in a magnetic field:
Interplay between Kondo, Zeeman, and Cooper-pair correlations

Masashi Hashimoto Department of Physics, Osaka City University, Sumiyoshi-ku, Osaka, 558-8585, Japan    Yasuhiro Yamada NTT Basic Research Laboratories, NTT Corporation, Atsugi, Kanagawa, 243-0198, Japan    Yoichi Tanaka Advanced Simulation Technology Of Mechanics R&\&D Co., Ltd., Bunkyo-ku, Tokyo, 112-0002, 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    Takuro Kemi Department of Physics, Osaka City University, Sumiyoshi-ku, Osaka, 558-8585, Japan    Norio Kawakami Department of Physics, Ritsumeikan University, Kusatsu, Shiga, 525-8577, Japan Department of Materials Engineering Science, Osaka University, Toyonaka, Osaka, 560-8531, 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 the nonlocal magnetotransport through a strongly correlated quantum dot, connected to multiple terminals consisting of two normal and one superconducting (SC) leads. Specifically, we present a comprehensive view on the interplay between the crossed Andreev reflection (CAR), the Kondo effect, and the Zeeman splitting at zero temperature in the large SC gap limit. The ground state of this network shows an interesting variety, which varies continuously with the system parameters, such as the coupling strength ΓS\Gamma_{S} between the SC lead and the quantum dot, the Coulomb repulsion UU, the impurity level εd\varepsilon_{d}, and the magnetic field bb. We show, using the many-body optical theorem which is derived from the Fermi-liquid theory, that the nonlocal conductance is determined by the transmission rate of the Cooper pairs 𝒯CP=14sin2Θsin2(δ+δ)\mathcal{T}_{\mathrm{CP}}=\frac{1}{4}\sin^{2}\Theta\,\sin^{2}\bigl{(}\delta_{\uparrow}+\delta_{\downarrow}) and that of the Bogoliubov particles 𝒯BG=12σsin2δσ\mathcal{T}_{\mathrm{BG}}=\frac{1}{2}\sum_{\sigma}\sin^{2}\delta_{\sigma}. Here, δσ\delta_{\sigma} is the phase shift of the renormalized Bogoliubov particles, and Θcot1(ξd/ΓS)\Theta\equiv\cot^{-1}(\xi_{d}/\Gamma_{S}) is the Bogoliubov-rotation angle in the Nambu pseudo spin space, with ξd=εd+U/2\xi_{d}=\varepsilon_{d}+U/2. It is also demonstrated, using Wilson’s numerical renormalization group approach, that the CAR is enhanced in the crossover region between the Kondo regime and the SC-proximity-dominated regime at zero magnetic field. The magnetic fields induce another crossover between the Zeeman-dominated regime and the SC-dominated regime, which occurs when the renormalized Andreev resonance level of majority spin crosses the Fermi level. We find that the CAR is enhanced and becomes less sensitive to magnetic fields in the SC-dominated regime close to the crossover region spreading over the angular range of π/4Θ3π/4\pi/4\lesssim\Theta\lesssim 3\pi/4. At the level crossing point, a spin-polarized current flows between the two normal leads, and it is significantly enhanced in the directions of Θ0\Theta\simeq 0 and Θπ\Theta\simeq\pi where the SC proximity effect is suppressed.

I Introduction

Quantum dots (QD) connected to multi terminal networks consisting of normal and superconducting (SC) leads is one of the active fields of current research. In such networks, the quantum coherence and entanglements can be probed through the Andreev reflection Hofstetter et al. (2009); Schindele et al. (2012); Das et al. (2012); Schindele et al. (2014); Fülöp et al. (2014); Tan et al. (2015); Borzenets et al. (2016); Tan et al. (2021); Golubev and Zaikin (2007); Eldridge et al. (2010); Sánchez et al. (2018); Wrześniewski and Weymann (2017); Rech et al. (2012); Walldorf et al. (2018); Hussein et al. (2019); Wójcik and Weymann (2019); Walldorf et al. (2020); Lee et al. (2014); Bordoloi et al. (2022) and Josephson effect. Choi et al. (2000); Wang and Hu (2011); Deacon et al. (2015); Padurariu et al. (2017); Yokoyama and Nazarov (2015); Ueda et al. (2019)

In particular, the crossed Andreev reflection (CAR) is one of the most interesting processes caused by a Cooper-pair tunneling in which an incident electron entering from a normal lead forms a Cooper pair with another electron from the other normal leads to tunnel into the SC leads, leaving a hole in the normal lead where the second electron came from. The time-reversal process of the CAR corresponds to a splitting of a Cooper pair that is emitted from the SC lead into two entangled electrons penetrating the different normal leads. The CAR and the Cooper-pair splitting have also been studied in the multi-terminal systems without quantum dots. Cadden-Zimansky and Chandrasekhar (2006); Lesovik, G. B. et al. (2001); Russo et al. (2005); Yeyati et al. (2007); Cadden-Zimansky et al. (2009); Wei and Chandrasekhar (2010); Golubev and Zaikin (2019); Kirsanov et al. (2019); Ranni et al. (2021)

Quantum dots give a variety to the transport properties of multi-terminal systems, through the tunable parameters such as electron correlations, resonant-level positions, and local magnetic fields which can polarize the spins of electrons. The strong electron correlations induce an interesting crossover between the Kondo singlet and the Cooper-pair singlet. Yoshioka and Ohashi (2000); Tanaka et al. (2007a); Buizert et al. (2007); Deacon et al. (2010a, b); Yamada et al. (2011); Governale et al. (2008); Oguri et al. (2013); Koga (2013); Domański et al. (2017); Vecino et al. (2003); Oguri et al. (2004); Tanaka et al. (2007b); Lee et al. (2022); Pillet et al. (2013); Kadlecová et al. (2017); Pokorný and Žonda (2023) Furthermore, the magnetic field induces a crossover occurring between the Kondo singlet state and the spin-polarized state due to the Zeeman splitting of discrete energy levels of quantum dots, which has recently been revisited to find that the three-body Fermi-liquid corrections play an essential role in the crossover region. Filippone et al. (2018); Oguri and Hewson (2018)

The CAR contributions can be probed through the nonlocal conductance for the current flowing from the QD towards one of the normal drain electrode when the bias voltage is applied to the source electrode. Hofstetter et al. (2009); Schindele et al. (2014); Fülöp et al. (2014); Tan et al. (2015) However, the nonlocal current also includes the contributions of the single-electron-tunneling process, in which an incident electron from the source electrode transmits directly towards the drain electrode through the QD. In order to observe the CAR contributions, it is important to find some sweet spots in the parameter space, at which the superconducting proximity effect dominates the nonlocal current and enhances the Cooper-pair tunneling by reconciling it with the other effects from electron correlations and magnetic fields.

The CAR in a single correlated quantum dot has theoretically been studied over a decade, particularly for a three-terminal QD connected to two normal and one superconducting leads. In the early stage, Futterer et al.Futterer et al. (2009) and Michałek et al.Michałek et al. (2013, 2015) demonstrated some behaviors of the nonlocal transport conductance typical to this three-terminal configuration,Michałek et al. (2015) taking also into account the Coulomb interaction with a generalized master equationFutterer et al. (2009) or the equation-of-motion method.Michałek et al. (2013) It has been extended to the configuration in which the normal leads are replaced by ferromagnetic metals and has been investigated intensively, using also the methods such as the real-time diagrammatic method and the numerical renormalization group (NRG). Futterer et al. (2009); Trocha and Wrześniewski (2018); Weymann and Trocha (2014); Wójcik and Weymann (2014); Weymann and Wójcik (2015)

Effects of the Zeeman splitting induced by the external magnetic field applied to quantum dots have also been theoretically investigated, mainly for two-terminal systems in which a quantum dot is connected to a single paramagnetic normal and a SC lead so far.Zhao and Wang (2001); Górski et al. (2020); Yamada et al. (2007); Domański et al. (2008, 2016); Žitko et al. (2015) Specifically, these theories addressed such subjects as the field dependence of the Andreev transport,Zhao and Wang (2001); Górski et al. (2020) the role of the Coulomb interaction in this configuration, Yamada et al. (2007); Domański et al. (2008, 2016) and the quantum phase transition between the spin-singlet and -doublet ground states.Žitko et al. (2015) However, it is still not fully clarified how the CAR contributions evolve at low energies in a wide parameter space of the multi-terminal networks, with and without magnetic fields.

Refer to caption
Figure 1: Single quantum dot (QD) connected to two normal leads (N) and one superconductor lead (SC). ΓL\Gamma_{L}, ΓR\Gamma_{R}, and ΓS\Gamma_{S} represent the coupling strengths of the QD with the left (LL), the right (RR), and the SC leads, respectively. The contributions of the normal tunnelings are given by ΓN=ΓL+ΓR\Gamma_{N}=\Gamma_{L}+\Gamma_{R}.
Refer to caption
Figure 2: Parameter space of HeffH_{\mathrm{eff}} at zero magnetic field b=0b=0, defined in Eqs. (50) and (8). The semicircle represents the line along which the energy of the Andreev level EAξd2+ΓS2E_{A}\equiv\sqrt{\xi_{d}^{2}+\Gamma_{S}^{2}} coincides with one-half of the Coulomb interaction U/2U/2, where ξdεd+U/2\xi_{d}\equiv\varepsilon_{d}+U/2. In the atomic limit ΓN=0\Gamma_{N}=0, the ground state is a magnetic spin doublet inside the semicircle, which eventually is screened by conduction electrons to form the Kondo singlet when the tunnel coupling ΓN\Gamma_{N} is switched on, whereas the ground state is a spin singlet due to the Cooper paring outside the circle. Θ\Theta is the Bogoliubov-rotation angle, which parametrizes the contributions of the Andreev scattering on the transport coefficients.

The purpose of this paper is to provide a comprehensive view of the Andreev transport through a strongly-correlated quantum state, the characteristics of which vary due to the interplay between the Kondo, Zeeman, and Cooper-pair correlations. To this end, we calculate the transport coefficients, using the Fermi-liquid theory Nozières (1974); Yamada (1975a, b); Shiba (1975); Yoshimori (1976) in conjunction with Wilson’s numerical renormalization group (NRG). Specifically, we consider a three-terminal quantum dot connected to two normal and one superconducting leads, as illustrated in Fig. 1, in the large SC gap limit.Tanaka et al. (2007a) We first of all derive the optical theorem for the CAR at zero temperature, using the Fermi-liquid theory that describes the interacting Bogoliubov particles moving throughout the entire system. It elucidates the fact that the nonlocal conductance is determined by the transmission rate of Cooper pairs 𝒯CP=14sin2Θsin2(δ+δ)\mathcal{T}_{\mathrm{CP}}=\frac{1}{4}\sin^{2}\Theta\,\sin^{2}\bigl{(}\delta_{\uparrow}+\delta_{\downarrow}) and that of the Bogoliubov particles 𝒯BG=12σsin2δσ\mathcal{T}_{\mathrm{BG}}=\frac{1}{2}\sum_{\sigma}\sin^{2}\delta_{\sigma}. Here, Θcot1(ξd/ΓS)\Theta\equiv\cot^{-1}(\xi_{d}/\Gamma_{S}) is the angular coordinate in the ξdεd+U/2\xi_{d}\equiv\varepsilon_{d}+U/2 vs ΓS\Gamma_{S} plane shown in Fig. 2, and εd\varepsilon_{d} and UU are the discrete level and the Coulomb interaction of electrons in the QD, respectively, and ΓS\Gamma_{S} is the coupling strength between the QD and SC lead. In this case, the phase shift δσ\delta_{\sigma} of the interacting Bogoliubov particles does not depend on the angle Θ\Theta but varies along the radial coordinate EA=ξd2+ΓS2E_{A}=\sqrt{\xi_{d}^{2}+\Gamma_{S}^{2}}.

We also calculate the transmission rate 𝒯CP\mathcal{T}_{\mathrm{CP}} and 𝒯BG\mathcal{T}_{\mathrm{BG}} with the NRG in a wide range of the parameter space. It is demonstrated that, at zero magnetic field, the CAR contributions are significantly enhanced near the crossover region between the Kondo regime and the SC-proximity-dominated regime. Specifically, it takes place in a crescent-shaped region spreading over the range of U/2EAU/2+ΓNU/2\lesssim E_{A}\lesssim U/2+\Gamma_{N} in the radial direction and π/4Θ3π/4\pi/4\lesssim\Theta\lesssim 3\pi/4 in the angular direction: ΓN\Gamma_{N} is the resonance width due to the tunneling between the QD and normal leads. When a magnetic field is applied, another crossover occurs between the Zeeman-dominated regime and the SC-proximity-dominated regime when the spin-polarized Andreev level crosses the Fermi level. We find that the CAR-dominated transport taking place in the crescent region is less sensitive to magnetic fields, and it emerges as a flat valley structure in the magnetic-field dependence of the nonlocal conductance. This parameter region provides an optimal condition for observing the Cooper-pair tunneling, i.e., a sweet spot, especially in the direction of Θπ/2\Theta\simeq\pi/2 where the Cooper pairs are most entangled and become equal-weight linear combinations of an electron and a hole.

This paper is organized as follows. In Sec. II, we introduce an Anderson impurity model for quantum dots connected to SC and normal leads, and rewrite the Hamiltonian and the Green’s function in terms of interacting Bogoliubov particles. Then, the optical theorem and the formula for the nonlocal conductance are derived at zero temperature using the Fermi-liquid description for the interacting Bogoliubov particles in Sec. III. We investigate the CAR contributions to the nonlocal conductance, using the NRG, at zero and finite magnetic fields in Secs. IV and V, respectively. Summary and discussion are given in Sec. VI.

II Fermi-liquid description for interacting Bogoliubov particles

In this section, we show how the contributions of the CAR to the nonlocal conductance of the multi-terminal network can be described in the context of the Fermi-liquid theory for the interacting Bogoliubov particles at zero temperature.Tanaka et al. (2007a)

II.1 Anderson impurity model for the CAR

We start with an Anderson impurity model for a single quantum dot (QD) connected to two normal (N) and one superconducting (SC) leads, as shown in Fig. 1:

H=\displaystyle H\,= Hdot+HN+HTN+HS+HTS,\displaystyle\ \,H_{\mathrm{dot}}\,+\,H_{\text{N}}\,+\,H_{\text{TN}}\,+\,H_{\text{S}}+H_{\text{TS}}\,, (1)
Hdot=\displaystyle H_{\mathrm{dot}}= ξd(nd1)b(nd,nd,)+U2(nd1)2,\displaystyle\ \xi_{d}\bigl{(}n_{d}-1\bigr{)}-b\,\bigl{(}n_{d,\uparrow}-n_{d,\downarrow}\bigr{)}\,+\frac{U}{2}\bigl{(}n_{d}-1\bigr{)}^{2}, (2)
HN=\displaystyle H_{\text{N}}= ν=L,RσDD𝑑εεcε,ν,σcε,ν,σ,\displaystyle\ \sum_{\nu=L,R}\sum_{\sigma}\int_{-D}^{D}\!d\varepsilon\,\varepsilon\,c^{\dagger}_{\varepsilon,\nu,\sigma}c_{\varepsilon,\nu,\sigma}, (3)
HTN=\displaystyle H_{\text{TN}}= ν=L,RvνσDDdερc(cε,ν,σdσ+H.c.),\displaystyle\ \sum_{\nu=L,R}v_{\nu}\sum_{\sigma}\int_{-D}^{D}\!d\varepsilon\,\sqrt{\rho_{c}}\,\Bigl{(}c^{\dagger}_{\varepsilon,\nu,\sigma}d_{\sigma}+\mathrm{H.c.}\Bigr{)}, (4)
HS=\displaystyle H_{\text{S}}= σDSDS𝑑εεsε,σsε,σ\displaystyle\ \sum_{\sigma}\!\int_{-D_{S}}^{D_{S}}\!\!d\varepsilon\,\varepsilon\,s^{\dagger}_{\varepsilon,\sigma}s_{\varepsilon,\sigma}
+DSDSdε(ΔSsε,sε,+H.c.),\displaystyle\quad\ +\int_{-D_{S}}^{D_{S}}\!\!d\varepsilon\left(\Delta_{S}\,s^{\dagger}_{\varepsilon,\uparrow}s^{\dagger}_{\varepsilon,\downarrow}+\mathrm{H.c.}\right), (5)
HTS=\displaystyle H_{\text{TS}}= vSσDSDSdερS(sε,σdσ+H.c.).\displaystyle\ v_{\text{S}}\sum_{\sigma}\!\int_{-D_{S}}^{D_{S}}\!\!d\varepsilon\,\sqrt{\rho_{S}}\,\Bigl{(}s^{\dagger}_{\varepsilon,\sigma}d_{\sigma}+\mathrm{H.c.}\Bigr{)}. (6)

Here, HdotH_{\mathrm{dot}} describes the QD part: ξdεd+U/2\xi_{d}\equiv\varepsilon_{d}+{U}/{2}, with εd\varepsilon_{d} the discrete energy level, UU the Coulomb interaction, and bb (μBB)\equiv\mu_{B}B) the Zeeman energy due to the magnetic field BB applied to the QD, with μB\mu_{B} the Bohr magneton. dσd^{\dagger}_{\sigma} is the creation operator for an electron with spin σ\sigma, and ndnd,+nd,n_{d}\equiv n_{d,\uparrow}+n_{d,\downarrow} is the number operator with nd,σdσdσn_{d,\sigma}\equiv d^{\dagger}_{\sigma}d_{\sigma}. A constant energy shift, which does not affect the physics, is included in Eq. (2) in order to describe clearly that the system has the electron-hole symmetry at ξd=0\xi_{d}=0.

HNH_{\text{N}} describes the conduction electrons in the normal leads, the density of states of which is assumed to be flat ρc=1/(2D)\rho_{c}=1/(2D), with DD the half-width of the bands. cε,ν,σc^{\dagger}_{\varepsilon,\nu,\sigma} is the creation operator for conduction electrons with spin σ\sigma and energy ε\varepsilon. The operators for conduction electrons satisfy the following anti-commutation relation that is normalized by the Dirac delta function: {cε,ν,σ,cε,ν,σ}=δννδσσδ(εε)\{c_{\varepsilon,\nu,\sigma},c^{\dagger}_{\varepsilon^{\prime},\nu^{\prime},\sigma^{\prime}}\}=\delta_{\nu\nu^{\prime}}\,\delta_{\sigma\sigma^{\prime}}\delta(\varepsilon-\varepsilon^{\prime}). HTNH_{\text{TN}} describes the tunnel couplings between the QD and the normal leads. The level broadening of the discrete energy level in the QD is given by ΓNΓL+ΓR\Gamma_{N}\equiv\Gamma_{L}+\Gamma_{R}, with Γνπρcvν2\Gamma_{\nu}\equiv\pi\rho_{c}v_{\nu}^{2} the contributions of the two normal leads on the left ν=L\nu=L and right ν=R\nu=R.

HSH_{\text{S}} and HTSH_{\text{TS}} describe the contributions of the superconducting lead with an ss-wave SC gap ΔS|ΔS|eiϕS\Delta_{S}\equiv\left|\Delta_{S}\right|\,e^{i\phi_{S}}: sε,σs^{\dagger}_{\varepsilon,\sigma} is the creation operator for electrons in the SC lead, with DSD_{S} the half-band width and ρS=1/(2DS)\rho_{S}=1/(2D_{S}). One of the key parameters for the SC proximity effects is ΓSπρSvS2\Gamma_{S}\equiv\pi\rho_{S}v_{\text{S}}^{2}, i.e., the coupling strength between the QD and the SC lead.

In this paper, we study the crossed Andreev reflection occurring at low energies, much lower than the SC energy gap. To this end, we consider the large gap limit |ΔS|\left|\Delta_{S}\right|\to\infty, which is taken at |ΔS|DS\left|\Delta_{S}\right|\ll D_{S} keeping ΓS\Gamma_{S} constant.Tanaka et al. (2007a) In this case, the superconducting proximity effects can be described by the pair potential penetrating into the QD:

ΔdΓSeiϕS.\displaystyle\Delta_{d}\equiv\Gamma_{S}\,e^{i\phi_{S}}\,. (7)

The Coulomb interaction UU induces the correlation effects for electrons in the QD and the symmetrized linear combination of the conduction bands defined in Eq.  (46), which can be described by an effective Hamiltonian HeffH_{\mathrm{eff}} given in Eq. (50) (see Appendix A). Furthermore, carrying out the Bogoliubov rotation defined in Eq. (53), it can be transformed further into a system of interacting Bogoliubov particles described by a standard Anderson model:

Heff=\displaystyle H_{\mathrm{eff}}= EA(σγd,σγd,σ1)b(γd,γd,γd,γd,)\displaystyle\ E_{A}\left(\sum_{\sigma}\gamma^{\dagger}_{d,\sigma}\gamma_{d,\sigma}-1\right)-b\,\Bigl{(}\gamma^{\dagger}_{d,\uparrow}\gamma_{d,\uparrow}-\gamma^{\dagger}_{d,\downarrow}\gamma_{d,\downarrow}\Bigr{)}
+U2(σγd,σγd,σ1)2+σDD𝑑εεγε,σγε,σ\displaystyle+\frac{U}{2}\left(\sum_{\sigma}\gamma^{\dagger}_{d,\sigma}\gamma_{d,\sigma}-1\right)^{2}+\sum_{\sigma}\int_{-D}^{D}\!d\varepsilon\,\varepsilon\ \gamma^{\dagger}_{\varepsilon,\sigma}\gamma_{\varepsilon,\sigma}
+vNσDDdερc(γε,σγd,σ+H.c.),\displaystyle+v_{N}\sum_{\sigma}\int_{-D}^{D}\!d\varepsilon\,\sqrt{\rho_{c}}\,\left(\gamma^{\dagger}_{\varepsilon,\sigma}\gamma_{d,\sigma}+\mathrm{H.c.}\right)\,, (8)
Nγ=\displaystyle N_{\gamma}= σγd,σγd,σ+σDD𝑑εγε,σγε,σ.\displaystyle\ \sum_{\sigma}\gamma^{\dagger}_{d,\sigma}\gamma_{d,\sigma}+\sum_{\sigma}\int_{-D}^{D}\!d\varepsilon\,\gamma^{\dagger}_{\varepsilon,\sigma}\gamma_{\varepsilon,\sigma}\,. (9)

Here, EAξd2+ΓS2E_{A}\equiv\sqrt{\xi_{d}^{2}+\Gamma_{S}^{2}} is the effective impurity level, and vNvL2+vR2v_{N}\equiv\sqrt{v_{L}^{2}+v_{R}^{2}}. The operators γd,σ\gamma_{d,\sigma} and γε,σ\gamma_{\varepsilon,\sigma} describe the Bogoliubov particles in the dot and the symmetrized part of the conduction band, respectively. The effective Hamiltonian conserves the total number of the Bogoliubov particles NγN_{\gamma}, reflecting the U(1)U(1) symmetry along the principal axis in the Nambu pseudo-spin space.

Figure 2 illustrates the parameter space of HeffH_{\mathrm{eff}} at zero magnetic field b=0b=0. For finite ΓN\Gamma_{N}, the Kondo screening due to the normal conduction electrons occurs inside the semicircle region, at which the impurity level is occupied by a single Bogoliubov particle: Q1.0Q\simeq 1.0 with

Q\displaystyle Q\, Q+Q,Qσγd,σγd,σ.\displaystyle\equiv\,Q_{\uparrow}+Q_{\downarrow},\qquad Q_{\sigma}\,\equiv\,\left\langle\gamma^{\dagger}_{d,\sigma}\gamma_{d,\sigma}\right\rangle. (10)

Bogoliubov particles show also the valence-fluctuation behavior near EAU/2E_{A}\simeq U/2, at which the crossover between the Kondo singlet and the superconducting singlet occurs. The Bogoliubov rotation angle corresponds to Θ=cot1(ξd/ΓS)\Theta=\cot^{-1}(\xi_{d}/\Gamma_{S}) shown in Fig. 2. In particular, the crossed Andreev scattering is enhanced in the angular range of π/4Θ3π/4\pi/4\lesssim\Theta\lesssim 3\pi/4 outside the semicircle EAU/2E_{A}\gtrsim U/2, as discussed later in Secs. IV and V.

II.2 Renormalized Bogoliubov quasiparticles

In this work, we calculate the nonlocal conductance for the current flowing into the drain electrode, using the retarded Green’s function for electrons in the QD:

𝑮ddr(ω)i0𝑑tei(ω+i0+)t\displaystyle\!\!\!\bm{G}^{r}_{dd}(\omega)\equiv-i\int_{0}^{\infty}\!dt\,e^{i\left(\omega+i0^{+}\right)t}
×({d(t),d}{d(t),d}{d(t),d}{d(t),d}).\displaystyle\qquad\qquad\ \ \times\begin{pmatrix}\left\langle\left\{d_{\uparrow}(t),\,d^{\dagger}_{\uparrow}\right\}\right\rangle&\left\langle\left\{d_{\uparrow}(t),\,d_{\downarrow}\right\}\right\rangle\\ \left\langle\left\{d^{\dagger}_{\downarrow}(t),\,d^{\dagger}_{\uparrow}\right\}\right\rangle&\left\langle\left\{d^{\dagger}_{\downarrow}(t),\,d_{\downarrow}\right\}\right\rangle\rule{0.0pt}{17.07182pt}\end{pmatrix}. (11)

Here, \langle\cdots\rangle denotes the thermal average at equilibrium. This matrix Green’s function can be diagonalized with the Bogoliubov transformation:

𝓤𝑮ddr(ω)𝓤\displaystyle\bm{\mathcal{U}}^{\dagger}\,\bm{G}^{r}_{dd}(\omega)\ \bm{\mathcal{U}}\, =(Gγ,r(ω)00Gγ,a(ω)).\displaystyle=\,\begin{pmatrix}G^{r}_{\gamma,\uparrow}(\omega)&0\\ 0&-G^{a}_{\gamma,\downarrow}(-\omega)\rule{0.0pt}{11.38092pt}\end{pmatrix}. (12)

We will choose the Josephson phase of the pair potential to be ϕS=0\phi_{S}=0 in the following, so that 𝓤\bm{\mathcal{U}} is determined solely by a pseudo-spinor rotation with the angle Θ/2\Theta/2:

𝓤\displaystyle\bm{\mathcal{U}}\, =(cosΘ2sinΘ2sinΘ2cosΘ2).\displaystyle=\,\begin{pmatrix}\cos\frac{\Theta}{2}&\,-\sin\frac{\Theta}{2}\\ \sin\frac{\Theta}{2}&\rule{0.0pt}{14.22636pt}\quad\cos\frac{\Theta}{2}\end{pmatrix}. (13)

The matrix elements of 𝓤\bm{\mathcal{U}} determine the behaviors of transport coefficients as the superconducting coherence factors,

cosΘ2=12(1+ξdEA),sinΘ2=12(1ξdEA).\displaystyle\cos\frac{\Theta}{2}=\sqrt{\frac{1}{2}\left(1+\frac{\xi_{d}}{E_{A}}\right)},\quad\sin\frac{\Theta}{2}=\sqrt{\frac{1}{2}\left(1-\frac{\xi_{d}}{E_{A}}\right)}.\rule{0.0pt}{28.45274pt}

The diagonal elements Gγ,σrG^{r}_{\gamma,\sigma} and Gγ,σaG^{a}_{\gamma,\sigma} on the right-hand side of Eq. (12) are the retarded and advanced Green’s functions for the interacting Bogoliubov particles, described by HeffH_{\mathrm{eff}}. These diagonal elements can be expressed in the form, using Eq. (53),

Gγ,σr(ω)\displaystyle G^{r}_{\gamma,\sigma}(\omega)\equiv i0𝑑tei(ω+i0+)t{γd,σ(t),γd,σ}\displaystyle\,-i\int_{0}^{\infty}\!dt\,e^{i\left(\omega+i0^{+}\right)t}\left\langle\left\{\gamma_{d,\sigma}(t),\,\gamma^{\dagger}_{d,\sigma}\right\}\right\rangle
=\displaystyle= 1ωEA,σΣγ,σU(ω)+iΓN,\displaystyle\ \frac{1}{\omega\,-E_{A,\sigma}-\Sigma^{U}_{\gamma,\sigma}(\omega)+i\Gamma_{N}}, (14)

and Gγ,σa(ω)={Gγ,σr(ω)}G^{a}_{\gamma,\sigma}(\omega)=\left\{G^{r}_{\gamma,\sigma}(\omega)\right\}^{*}. Here, EA,σEAσbE_{A,\sigma}\equiv E_{A}-\sigma\,b, and Σγ,σU(ω)\Sigma^{U}_{\gamma,\sigma}(\omega) represents the self-energy corrections due to the Coulomb interaction term, (U/2)(nd1)2({U}/{2})\left(n_{d}-1\right)^{2}, defined in Eq. (2). The unperturbed part of the denominator describes the Andreev resonance level with the width ΓN\Gamma_{N} situated at ω=EA,σ\omega=E_{A,\sigma}.

At low energies, effects of the electron correlations on the transport properties can be deduced from the behavior of the self-energy near ω0\omega\simeq 0 at zero temperature T=0T=0:

Gγ,σr(ω)ZσωE~A,σ+iΓ~N,σ.\displaystyle G^{r}_{\gamma,\sigma}(\omega)\,\simeq\,\frac{Z_{\sigma}}{\omega-\widetilde{E}_{A,\sigma}+i\widetilde{\Gamma}_{N,\sigma}}\,. (15)

The asymptotic form of the Green’s function defines a renormalized resonance level of quasiparticles in the Fermi liquid, the position E~A,σ\widetilde{E}_{A,\sigma} and the width Γ~N,σ\widetilde{\Gamma}_{N,\sigma} of which are given byNozières (1974); Yamada (1975a, b); Shiba (1975); Yoshimori (1976)

Γ~N,σ=\displaystyle\widetilde{\Gamma}_{N,\sigma}= ZσΓN,1Zσ= 1Σγ,σU(ω)ω|ω=0,\displaystyle\ Z_{\sigma}\,\Gamma_{N},\qquad\frac{1}{Z_{\sigma}}\,=\,1-\left.\frac{\partial\Sigma^{U}_{\gamma,\sigma}(\omega)}{\partial\omega}\right|_{\omega=0}, (16)
E~A,σ=\displaystyle\widetilde{E}_{A,\sigma}\,= Zσ[EA,σ+Σγ,σU(0)].\displaystyle\ Z_{\sigma}\Bigl{[}\,E_{A,\sigma}+\Sigma^{U}_{\gamma,\sigma}(0)\,\Bigr{]}\,. (17)

Furthermore, the phase shift δσ\delta_{\sigma} of the interacting Bogoliubov particles is defined by Gγ,σr(0)=|Gγ,σr(0)|eiδσG^{r}_{\gamma,\sigma}(0)=-\left|G^{r}_{\gamma,\sigma}(0)\right|e^{i\delta_{\sigma}}, i.e.,

δσ=\displaystyle\delta_{\sigma}\,= π2tan1(E~A,σΓ~N,σ),\displaystyle\ \frac{\pi}{2}-\tan^{-1}\left(\frac{\widetilde{E}_{A,\sigma}}{\widetilde{\Gamma}_{N,\sigma}}\right)\,, (18)

plays a primary role in the ground-state properties. These renormalized parameters can be calculated, for instance, using the NRG approach described in the next section.

The Friedel sum rule also holds for the interacting Bogoliubov particles, and thus the average number of the Bogoliubov particles in the QD is determined by the phase shift,

Qσ\displaystyle Q_{\sigma} T0δσπ.\displaystyle\xrightarrow{\,T\to 0\,}\frac{\delta_{\sigma}}{\pi}. (19)

The phase shift varies in the range of 0δσπ/20\leq\delta_{\sigma}\leq\pi/2 along the radial coordinate EAE_{A} in the ξd\xi_{d} vs ΓS\Gamma_{S} plane but is independent of the angle Θ\Theta.

The ground-state properties, such as the occupation number of electrons nd\left\langle n_{d}\right\rangle and the pair correlation function dd+dd\bigl{\langle}d^{\dagger}_{\uparrow}\,d^{\dagger}_{\downarrow}+d_{\downarrow}\,d_{\uparrow}\bigr{\rangle}, can be deduced from QQ:

nd1=\displaystyle\left\langle n_{d}\right\rangle-1\ = (Q1)cosΘ,\displaystyle\,\left(Q-1\right)\,\cos\Theta\,, (20)
dd+dd=\displaystyle\left\langle d^{\dagger}_{\uparrow}\,d^{\dagger}_{\downarrow}+d_{\downarrow}\,d_{\uparrow}\right\rangle\ = (Q1)sinΘ.\displaystyle\,\left(Q-1\right)\,\sin\Theta\,. (21)

These two averages correspond to the projection of a vector of magnitude Q1Q-1 directed along the principal axis onto the zz axis and the xx axis of the Nambu space, respectively. Furthermore, a local magnetization mdm_{d} is induced in the quantum dot at finite magnetic fields,

md\displaystyle m_{d}\,\equiv\ nd,nd,=QQ.\displaystyle\ \left\langle n_{d,\uparrow}\right\rangle-\left\langle n_{d,\downarrow}\right\rangle\ =\,Q_{\uparrow}-Q_{\downarrow}\,. (22)

Therefore, the occupation number of electrons with spin σ\sigma is given by

nd,σ=1+cosΘ2δσπ+1cosΘ2(1δσ¯π),\displaystyle\left\langle n_{d,\sigma}\right\rangle\,=\,\frac{1+\cos\Theta}{2}\,\frac{\delta_{\sigma}}{\pi}\,+\,\frac{1-\cos\Theta}{2}\,\left(1-\frac{\delta_{\overline{\sigma}}}{\pi}\right)\,, (23)

where σ¯\overline{\sigma} represents an opposite-spin component of σ\sigma.

III Linear-response theory for CAR

III.1 Cooper-pair transmission in a local Fermi liquid

We consider the linear-response current IRI_{R} flowing from the QD to the normal lead on the right, induced by bias voltages VLV_{L} and VRV_{R} applied to the left and right leads, respectively. It can be expressed in the following form at T=0T=0 (see Appendix B):

IR=\displaystyle I_{R}\,= IRET+IRCP,\displaystyle\ \,I_{R}^{\mathrm{ET}}\,+\,I_{R}^{\mathrm{CP}}, (24)
IRET=\displaystyle I_{R}^{\mathrm{ET}}= 2e2h𝒯ET4ΓRΓLΓN2(VLVR),\displaystyle\ \frac{2e^{2}}{h}\mathcal{T}_{\mathrm{ET}}\ \frac{4\Gamma_{R}\Gamma_{L}}{\Gamma_{N}^{2}}\,\bigl{(}V_{L}-V_{R}\bigr{)},\rule{0.0pt}{17.07182pt}
IRCP=\displaystyle I_{R}^{\mathrm{CP}}= 2e2h𝒯CP[4ΓRΓLΓN2(VL+VR)+4ΓR2ΓN2 2VR].\displaystyle\ -\frac{2e^{2}}{h}\mathcal{T}_{\mathrm{CP}}\left[\frac{4\Gamma_{R}\Gamma_{L}}{\Gamma_{N}^{2}}\,\bigl{(}V_{L}+V_{R}\bigr{)}+\frac{4\Gamma_{R}^{2}}{\Gamma_{N}^{2}}\,2V_{R}\right].

Correspondingly, the current ILI_{L} flowing from the left normal lead towards the QD takes the form, IL=ILET+ILCPI_{L}=I_{L}^{\mathrm{ET}}+I_{L}^{\mathrm{CP}}, with ILET=IRETI_{L}^{\mathrm{ET}}=I_{R}^{\mathrm{ET}} and

ILCP=\displaystyle I_{L}^{\mathrm{CP}}= 2e2h𝒯CP[4ΓRΓLΓN2(VL+VR)+4ΓL2ΓN2 2VL].\displaystyle\ \frac{2e^{2}}{h}\mathcal{T}_{\mathrm{CP}}\left[\frac{4\Gamma_{R}\Gamma_{L}}{\Gamma_{N}^{2}}\,\bigl{(}V_{L}+V_{R}\bigr{)}+\frac{4\Gamma_{L}^{2}}{\Gamma_{N}^{2}}\,2V_{L}\right]. (25)

The two components of the current IνETI_{\nu}^{\mathrm{ET}} and IνCPI_{\nu}^{\mathrm{CP}} represent the contribution of the single-electron tunneling and that of the Cooper-pair tunneling, respectively. The transmission probabilities 𝒯ET\mathcal{T}_{\mathrm{ET}} and 𝒯CP\mathcal{T}_{\mathrm{CP}} are determined by the equilibrium Green’s functions at the Fermi level ω=0\omega=0, and can be expressed in terms of the phase shifts and the Bogoliubov angle (see Appendix B):

𝒯ET\displaystyle\!\!\!\mathcal{T}_{\mathrm{ET}}\equiv ΓN22[|{𝑮ddr(0)}11|2+|{𝑮ddr(0)}22|2]\displaystyle\ \frac{\Gamma_{N}^{2}}{2}\left[\ \Bigl{|}\bigl{\{}\bm{G}_{dd}^{r}(0)\bigl{\}}_{11}\Bigr{|}^{2}+\Bigl{|}\bigl{\{}\bm{G}_{dd}^{r}(0)\bigl{\}}_{22}\Bigr{|}^{2}\ \right]
=\displaystyle= 12σsin2δσ14sin2Θsin2(δ+δ),\displaystyle\ \frac{1}{2}\sum_{\sigma}\sin^{2}\delta_{\sigma}-\frac{1}{4}\,\sin^{2}\Theta\,\sin^{2}\bigl{(}\delta_{\uparrow}+\delta_{\downarrow})\,, (26)
𝒯CP\displaystyle\!\!\!\mathcal{T}_{\mathrm{CP}}\equiv ΓN22[|{𝑮ddr(0)}12|2+|{𝑮ddr(0)}21|2]\displaystyle\ \frac{\Gamma_{N}^{2}}{2}\left[\ \Bigl{|}\bigl{\{}\bm{G}_{dd}^{r}(0)\bigl{\}}_{12}\Bigr{|}^{2}+\Bigl{|}\bigl{\{}\bm{G}_{dd}^{r}(0)\bigl{\}}_{21}\Bigr{|}^{2}\ \right]
=\displaystyle= 14sin2Θsin2(δ+δ).\displaystyle\,\frac{1}{4}\,\sin^{2}\Theta\,\sin^{2}\bigl{(}\delta_{\uparrow}+\delta_{\downarrow})\,. (27)

These two are bounded in the range, 0𝒯ET10\leq\mathcal{T}_{\mathrm{ET}}\leq 1 and 0𝒯CP1/40\leq\mathcal{T}_{\mathrm{CP}}\leq 1/4, and are related to each other through the optical theorem (see Appendix C):

𝒯ET+𝒯CP=𝒯BG,𝒯BG12σsin2δσ.\displaystyle\mathcal{T}_{\mathrm{ET}}+\mathcal{T}_{\mathrm{CP}}\,=\,\mathcal{T}_{\mathrm{BG}}\,,\qquad\mathcal{T}_{\mathrm{BG}}\,\equiv\,\frac{1}{2}\,\sum_{\sigma}\sin^{2}\delta_{\sigma}\,. (28)

Here, 𝒯BG\mathcal{T}_{\mathrm{BG}} can be regarded as a transmission probability of the Bogoliubov particles.

The linear-response coefficients, given in the above for the large gap limit |ΔS||\Delta_{S}|\to\infty, are determined by δσ\delta_{\sigma} and Θ\Theta. Therefore, the Cooper-pair contributions, which vary depending on the parameter regions shown in Fig. 2, can systematically be explored by using the polar coordinate (EAE_{A}, Θ\Theta) since the phase shift δσ\delta_{\sigma} through which the many-body effects enter is independent of the angle Θ\Theta that determines the superconducting coherence factor sin2Θ\sin^{2}\Theta for the transmission probability 𝒯CP\mathcal{T}_{\mathrm{CP}}.

III.1.1 Nonlocal conductance for IRI_{R} at VL0V_{L}\neq 0 and VR=0V_{R}=0

Equations (24)–(27) provide a set of formulas that describe how the single-electron and the Cooper-pair tunneling parts, IRETI_{R}^{\mathrm{ET}} and IRCPI_{R}^{\mathrm{CP}}, contribute to the total current IRI_{R} for arbitrary bias voltages VLV_{L} and VRV_{R}. We next consider the situation, at which the right lead is grounded VR=0V_{R}=0 in order to clarify the contributions of the CAR to the nonlocal conductance gRLg_{\mathrm{RL}} for the current IRI_{R},

gRLIRVL=\displaystyle g_{\mathrm{RL}}\,\equiv\,\frac{\partial I_{R}}{\partial V_{L}}\,= 2g0(𝒯ET𝒯CP)\displaystyle\ 2\,g_{0}\,\bigl{(}\mathcal{T}_{\mathrm{ET}}-\mathcal{T}_{\mathrm{CP}}\bigr{)}
=\displaystyle= 2g0(𝒯BG2𝒯CP),\displaystyle\ 2\,g_{0}\,\bigl{(}\mathcal{T}_{\mathrm{BG}}-2\mathcal{T}_{\mathrm{CP}}\bigr{)}, (29)

where g0=e2h 4ΓRΓL/ΓN2g_{0}=\frac{e^{2}}{h}\,{4\Gamma_{R}\Gamma_{L}}/{\Gamma_{N}^{2}}. In the last line, the Bogoliubov angle Θ\Theta enters gRLg_{\mathrm{RL}} solely through 𝒯CP\mathcal{T}_{\mathrm{CP}} since 𝒯BG\mathcal{T}_{\mathrm{BG}} does not depend on it. The contribution of Cooper-pair tunnelings in gRLg_{\mathrm{RL}} is negative as it induces the current flowing from the right lead towards the QD at the center.

The CAR efficiency ηCAR\eta_{\text{CAR}} is one of the useful parameters for measuring the CAR contribution to the nonlocal conductance gRLg_{\mathrm{RL}}:

ηCAR\displaystyle\eta_{\text{CAR}}\ |IRCP||IRET|+|IRCP|=𝒯CP𝒯BG\displaystyle\equiv\,\frac{|I_{R}^{\mathrm{CP}}|}{|I_{R}^{\mathrm{ET}}|+|I_{R}^{\mathrm{CP}}|}=\frac{\mathcal{T}_{\mathrm{CP}}}{\mathcal{T}_{\mathrm{BG}}}
=sin2(δ+δ)sin2δ+sin2δsin2Θ2.\displaystyle=\,\frac{\sin^{2}\left(\delta_{\uparrow}+\delta_{\downarrow}\right)}{\,\sin^{2}\delta_{\uparrow}+\sin^{2}\delta_{\downarrow}}\,\frac{\sin^{2}\Theta}{2}\,. (30)

Alternatively, the nonlocal conductance can also be expressed in terms of the efficiency:

gRL=\displaystyle g_{\mathrm{RL}}\,=  2g0𝒯BG(1 2ηCAR).\displaystyle\ \,2g_{0}\,\mathcal{T}_{\mathrm{BG}}\Bigl{(}1\,-\,2\,\eta_{\text{CAR}}\Bigr{)}\,. (31)

Here, the Θ\Theta dependence of gRLg_{\mathrm{RL}} arises from the efficiency ηCAR\eta_{\text{CAR}}. The efficiency ηCAR\eta_{\text{CAR}} is enhanced by the coupling between the QD and the SC lead. In the limit ΓS0\Gamma_{S}\to 0 where the SC lead is disconnected, Eq. (31) reproduces the usual Landauer formula with the single-electron tunneling probability 𝒯BG\mathcal{T}_{\mathrm{BG}}.

Similarly, the local conductance for the current from the left lead ILI_{L} can also be expressed in the following form,

ILVL= 2g0𝒯BG(1+2ΓLΓRηCAR).\displaystyle\frac{\partial I_{L}}{\partial V_{L}}\,=\,2g_{0}\,\mathcal{T}_{\mathrm{BG}}\left(1\,+\,\frac{2\Gamma_{L}}{\Gamma_{R}}\,\eta_{\text{CAR}}\right)\,. (32)

Here, the second term on the right-hand side represents the contribution of the direct Andreev reflection (DAR), inducing the current component ILDAR4g0𝒯CP(ΓL/ΓR)VLI_{L}^{\mathrm{DAR}}\equiv 4g_{0}\,\mathcal{T}_{\mathrm{CP}}\,(\Gamma_{L}/\Gamma_{R})\,V_{L} ΓL2/ΓN2\propto\Gamma_{L}^{2}/\Gamma_{N}^{2} for VR=0V_{R}=0. Therefore, the ratio of the DAR contribution to ILI_{L} is determined by δ\delta and Θ\Theta through the efficiency ηCAR\eta_{\text{CAR}},

ILDARIL=2ΓLηCARΓR+ 2ΓLηCAR.\displaystyle\frac{I_{L}^{\mathrm{DAR}}}{I_{L}}\,=\,\frac{2\Gamma_{L}\,\eta_{\text{CAR}}}{\Gamma_{R}\,+\,2\Gamma_{L}\eta_{\text{CAR}}}\;. (33)

III.1.2 Andreev transport for VL=VRV_{L}=V_{R}

Here we briefly discuss another setting, in which bias voltages are applied in a symmetrical way VL=VRV_{L}=V_{R} (V\equiv V). In this case, the contribution of single-electron process vanishes IRET=ILET=0I_{R}^{\mathrm{ET}}=I_{L}^{\mathrm{ET}}=0, and the Cooper-pair tunnelings determine both IRI_{R} and ILI_{L}, as

IRVL=VR=V\displaystyle I_{R}\,\xrightarrow{\,V_{L}=V_{R}=V\,} 4e2h𝒯CP[4ΓRΓLΓN2+4ΓR2ΓN2]V,\displaystyle\ -\frac{4e^{2}}{h}\mathcal{T}_{\mathrm{CP}}\left[\,\frac{4\Gamma_{R}\Gamma_{L}}{\Gamma_{N}^{2}}\,+\,\frac{4\Gamma_{R}^{2}}{\Gamma_{N}^{2}}\,\right]V\,, (34)
ILVL=VR=V\displaystyle I_{L}\,\xrightarrow{\,V_{L}=V_{R}=V\,} +4e2h𝒯CP[4ΓRΓLΓN2+4ΓL2ΓN2]V.\displaystyle\ +\frac{4e^{2}}{h}\mathcal{T}_{\mathrm{CP}}\left[\,\frac{4\Gamma_{R}\Gamma_{L}}{\Gamma_{N}^{2}}\,+\,\frac{4\Gamma_{L}^{2}}{\Gamma_{N}^{2}}\,\right]V\,. (35)

For both IRI_{R} and ILI_{L}, the first and the second terms in the square brackets on the right-hand side represent the contributions of the crossed Andreev reflection and the direct Andreev reflection, respectively. These terms depend sensitively on the asymmetry of the tunnel couplings. For instance, the CAR dominates IRI_{R} for ΓRΓL\Gamma_{R}\ll\Gamma_{L}, as the direct Andreev scattering occurring in the right lead is suppressed.

The current flowing into the SC lead through the QD is given by ILIRI_{L}-I_{R}. It reaches the maximum value 4e2V/h4e^{2}V/h in the case at which 𝒯CP=14\mathcal{T}_{\mathrm{CP}}=\frac{1}{4} for symmetric junctions ΓL=ΓR\Gamma_{L}=\Gamma_{R} (ΓN/2\equiv\Gamma_{N}/2). Note that the behavior of this current ILIRI_{L}-I_{R} into the SC lead is equivalent to the one flowing through an N-QD-SC junction, which was investigated in the previous work.Tanaka et al. (2007a)

III.2 NRG approach to the CAR

In the following two sections, we numerically investigate the contribution of the CAR over a wide range of the parameter space. To this end, we have calculated the phase shift δσ\delta_{\sigma} and the other correlation functions of Bogoliubov quasiparticles, applying the NRG approachKrishna-murthy et al. (1980a, b); Hewson et al. (2004) to the effective Hamiltonian HeffH_{\mathrm{eff}} given in Eq. (8),Tanaka et al. (2007a) choosing the discretization parameter to be Λ=2.0\Lambda=2.0 and ΓN/D=1/(100π)\Gamma_{N}/D=1/(100\pi). We have also constructed the interpolating functions for the phase shift δσ\delta_{\sigma} from a discrete set of the NRG data obtained along the radial-EAE_{A} direction in the parameter space, described in Fig. 2. The dependence of the transport coefficients on the Bogoliubov-rotation angle Θ\Theta of the polar coordinate has been determined by using the exact formulas presented in the above.

We will discuss the CAR contribution to the nonlocal transport at zero field in Sec. IV, and then consider magnetic-field dependence in Sec. V.

IV Crossed-Andreev transport at zero field b=0b=0

In this section, we show the NRG results for the nonlocal conductance and renormalized parameters calculated at zero magnetic field b=0b=0, extending the previous results obtained for a two terminal N-QD-S system.Tanaka et al. (2007a) Before going into the details, we describe some general features which can be deduced from the transport formulas presented above.

At zero magnetic field b=0b=0, the phase shift becomes independent of spin component δ=δ\delta_{\uparrow}=\delta_{\downarrow} (δ\equiv\delta), and thus the transport coefficients are determined by two angular parameters δ\delta and Θ\Theta. The average occupation number of the Andreev level in this case is given by the phase shift Q=2δ/πQ=2\delta/\pi. It decreases from the unitary limit value Q=1Q=1 as EAE_{A} deviates from the origin, EA=0E_{A}=0, of the parameter space illustrated in Fig. 2. In contrast, the SC pair correlation function dd+dd\bigl{\langle}d^{\dagger}_{\uparrow}\,d^{\dagger}_{\downarrow}+d_{\downarrow}\,d_{\uparrow}\bigr{\rangle}, defined in Eq. (21), depends not only on the phase shift δ\delta but also the coherence factor, sinΘ\sin\Theta, which takes a maximum at Θ=π/2\Theta=\pi/2.

Similarly, at zero magnetic field, the transmission probabilities defined in Eqs.  (26) and (27) can be simplified, as

𝒯ET=\displaystyle\mathcal{T}_{\mathrm{ET}}= sin2δ𝒯CP,𝒯CP=14sin2Θsin22δ,\displaystyle\ \sin^{2}\delta-\mathcal{T}_{\mathrm{CP}}\,,\qquad\mathcal{T}_{\mathrm{CP}}=\frac{1}{4}\sin^{2}\Theta\,\sin^{2}2\delta\,, (36)

and 𝒯BG=sin2δ\mathcal{T}_{\mathrm{BG}}=\sin^{2}\delta. Therefore, the Cooper-pairing part 𝒯CP\mathcal{T}_{\mathrm{CP}} takes a maximum at Θ=π/2\Theta=\pi/2 and δ=π/4\delta=\pi/4, where the Andreev level for Bogliubov particles is quarter-filling Qσ=14Q_{\sigma}=\frac{1}{4}. Correspondingly, the nonlocal conductance and the CAR efficiency defined in Eqs. (29)–(31) can be expressed in the following forms, at b=0b=0:

gRL=\displaystyle g_{\mathrm{RL}}\,= 2g0sin2δ(12sin2Θcos2δ),\displaystyle\ 2\,g_{0}\sin^{2}\delta\,\Bigl{(}1-2\,\sin^{2}\Theta\,\cos^{2}\delta\Bigr{)}, (37)
ηCAR=\displaystyle\eta_{\text{CAR}}\,= sin2Θcos2δ.\displaystyle\ \sin^{2}\Theta\,\cos^{2}\delta\,. (38)

Thus, for the CAR to dominate the nonlocal conductance, taking a negative value gRL<0g_{\mathrm{RL}}<0, the Bogoliubov angle must be in the range π/4<Θ<3π/4\pi/4<\Theta<3\pi/4, i.e., 2sin2Θ>12\sin^{2}\Theta>1.

In particular, Cooper pairs are most entangled at Θ=π/2\Theta=\pi/2, and in this case the transport coefficients take the form,

𝒯ETΘ=π2sin4δ,gRLΘ=π2 2g0sin2δcos2δ.\displaystyle\mathcal{T}_{\mathrm{ET}}\,\xrightarrow{\,\Theta=\frac{\pi}{2}\,}\,\,\sin^{4}\delta,\qquad g_{\mathrm{RL}}\xrightarrow{\,\Theta=\frac{\pi}{2}\,}\,-\,2\,g_{0}\,\sin^{2}\delta\,\cos 2\delta. (39)

Hence, along the ΓS\Gamma_{S}-axis in Fig. 2, the nonlocal conductance gRLg_{\mathrm{RL}} becomes negative for 0<δπ/40<\delta\lesssim\pi/4, where the ground state of HeffH_{\mathrm{eff}} is in the valence-fluctuation regime or the empty-orbital regime of the Bogoliubov particles. It takes a minimum of the depth gRL/g0=1/4g_{\mathrm{RL}}/g_{0}=-1/4 at δ=π/6\delta=\pi/6. As the phase shift approaches δπ/2\delta\simeq\pi/2, the Kondo effect dominates and the transmission probability of the Bogoliubov-particle shows a Kondo-ridge structure, at which 𝒯BG1.0\mathcal{T}_{\mathrm{BG}}\simeq 1.0, as we will demonstrate later.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 3: Renormalized parameters plotted vs EA/UE_{A}/U for U/(πΓN)=1.0U/(\pi\Gamma_{N})=1.0, 2.02.0, 3.03.0, 5.05.0 at b=0b=0. (a): Occupation number of Bogoliubov particles QQ (=2δ/π=2\delta/\pi). (b): Pair correlation |dd+dd||\bigl{\langle}d^{\dagger}_{\uparrow}d^{\dagger}_{\downarrow}+d_{\downarrow}d_{\uparrow}\bigr{\rangle}|. (c): Renormalized Andreev-resonance energy E~A\widetilde{E}_{A}. (d): Renormalization factor Z=Γ~N/ΓNZ=\widetilde{\Gamma}_{N}/\Gamma_{N}, which at EA=0E_{A}=0 takes the values 0.6290.629, 0.2390.239, 0.0800.080, 0.0080.008, respectively, for the above four values of UU. Note that nd=1.0\langle n_{d}\rangle=1.0 and |dd+dd|=1Q|\bigl{\langle}d^{\dagger}_{\uparrow}d^{\dagger}_{\downarrow}+d_{\downarrow}d_{\uparrow}\bigr{\rangle}|=1-Q, at Θ=π/2\Theta=\pi/2 along the ΓS\Gamma_{S} axis of Fig. 2. The dashed line in (c) denotes the Hartree-Fock energy shift EAHFEAU/2EAU/2E_{A}^{\mathrm{HF}}\xrightarrow{E_{A}\gg U/2}E_{A}-U/2, given in Eq. (41).

IV.1 Ground state properties at Θ=π/2\Theta=\pi/2

We next consider how the ground state of HeffH_{\mathrm{eff}} evolves as EAE_{A} varies along the radial direction in the ξd\xi_{d} vs ΓS\Gamma_{S} plane, shown in Fig. 2. Note that the eigenstates and eigenvalues of the effective Hamiltonian defined in Eq. (8) do not depend on the angular coordinate Θ\Theta.

Figure 3(a) shows the occupation number QQ as a function of EAE_{A} for U/(πΓN)=1.0U/(\pi\Gamma_{N})=1.0, 2.02.0, 3.03.0, and 5.05.0. We see that QQ decreases as EAE_{A} increases, especially near EAU/2E_{A}\simeq U/2, where the crossover from the Kondo regime to the valence-fluctuation regime of Bogoliubov particles occurs for large interactions U/(πΓN)2.0U/(\pi\Gamma_{N})\gtrsim 2.0. Figure 3(b) shows the magnitude of the pair correlation function for Θ=π/2\Theta=\pi/2, where the absolute value is given by |dd+dd|=1Q|\bigl{\langle}d^{\dagger}_{\uparrow}d^{\dagger}_{\downarrow}+d_{\downarrow}d_{\uparrow}\bigr{\rangle}|=1-Q. It increases significantly at EAU/2E_{A}\simeq U/2, i.e., near the quarter-filling point Q=0.5Q=0.5 (δ=π/4\delta=\pi/4) of Bogoliubov particles, and it saturates to the upper-bound value 1.01.0 as EAE_{A} increases further towards the empty-orbital regime.

The Kondo behaviors of Bogoliubov particles are clearly seen for the renormalized Andreev level E~A\widetilde{E}_{A} and the wave-function renormalization factor Z=Γ~N/ΓNZ=\widetilde{\Gamma}_{N}/\Gamma_{N}, plotted in Figs. 3(c) and 3(d), respectively. The renormalized level is almost locked at the Fermi level E~A0.0\widetilde{E}_{A}\simeq 0.0, for large interactions U/(πΓN)2.0U/(\pi\Gamma_{N})\gtrsim 2.0, over a wide Kondo-dominated region 0EAU/20\leq E_{A}\lesssim U/2, taking place the inside of the semicircle in Fig. 2. Correspondingly, the renormalization factor ZZ is significantly suppressed in this region, and it indicates the fact that the Kondo energy scale TT^{*},

TZ4ρd,ρd=sin2δπΓN,\displaystyle T^{*}\,\equiv\,\frac{Z}{4\rho_{d}}\,,\qquad\qquad\rho_{d}\,=\,\frac{\sin^{2}\delta}{\pi\Gamma_{N}}\,, (40)

becomes much smaller than the bare tunneling energy scale ΓN\Gamma_{N}.

In contrast, at EAU/2E_{A}\gtrsim U/2, i.e., in the valence-fluctuation or empty-orbital regime for Bogoliubov particles, the effects of electron correlations become less important: the renormalization factor approaches Z1.0Z\simeq 1.0 and the renormalized level E~A,σ\widetilde{E}_{A,\sigma} approaches the Hartree-Fock (HF) energy shift:

EA,σHFEAσb+U(Qσ¯12)\displaystyle E_{A,\sigma}^{\mathrm{HF}}\,\equiv\,E_{A}-\sigma b\,+\,U\left(Q_{\overline{\sigma}}-\frac{1}{2}\right) (41)
EAU/2,b=0EAU2,\displaystyle\ \ \xrightarrow{\,E_{A}\gg U/2,\,b=0\,}\,E_{A}\,-\,\frac{U}{2}\,,

since Qσ¯0.0\,Q_{\overline{\sigma}}\simeq 0.0 at EAU/2E_{A}\gg U/2 and b=0b=0.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 4: Transport coefficients plotted vs EA/UE_{A}/U for U/(πΓN)=1.0U/(\pi\Gamma_{N})=1.0, 2.02.0, 3.03.0, 5.05.0 at b=0b=0, keeping the Bogoliubov angle fixed at Θ=π/2\Theta=\pi/2 (i.e., ξd=0\xi_{d}=0). (a): Single-electron transmission 𝒯ET\mathcal{T}_{\mathrm{ET}}, defined in Eq.  (26). (b): Cooper-pair contributions 2𝒯CP-2\mathcal{T}_{\mathrm{CP}} (<0<0) and the Bogoliubov-particle transmission 𝒯BG=sin2δ\mathcal{T}_{\mathrm{BG}}=\sin^{2}\delta, defined in Eqs.  (27) and (28), respectively. (c): Nonlocal conductance gRL/g0=2(𝒯BG2𝒯CP)g_{\mathrm{RL}}/g_{0}=2(\mathcal{T}_{\mathrm{BG}}-2\mathcal{T}_{\mathrm{CP}}) with g0=e2h4ΓRΓL/ΓN2g_{0}=\frac{e^{2}}{h}{4\Gamma_{R}\Gamma_{L}}/{\Gamma_{N}^{2}}. (d): CAR efficiency ηCAR\eta_{\text{CAR}} defined in Eq. (30).

IV.2 Transport properties at Θ=π/2\Theta=\pi/2

We next discuss the transport properties. Specifically, in this subsection, we consider the case Θ=π/2\Theta=\pi/2, where ξd=0\xi_{d}=0 and the occupation number of impurity electrons is fixed at nd=1\langle n_{d}\rangle=1, reflecting the electron-hole symmetry of HeffH_{\mathrm{eff}} defined in Eq. (50). In this case, the Andreev level takes the value EA=ΓSE_{A}=\Gamma_{S}, which is determined solely by the coupling strength between the QD and the SC lead and it breaks the particle-hole symmetry of the Bogoliubov particles even at ξd=0\xi_{d}=0.

The transmission probability 𝒯ET\mathcal{T}_{\mathrm{ET}} of the single-electron tunneling process is shown in Fig. 4(a). We see that the plateau of the unitary limit 𝒯ET1.0\mathcal{T}_{\mathrm{ET}}\simeq 1.0 evolves at 0EAU/20\leq E_{A}\lesssim U/2, for large UU. Since 𝒯ET=𝒯BG𝒯CP\mathcal{T}_{\mathrm{ET}}=\mathcal{T}_{\mathrm{BG}}-\mathcal{T}_{\mathrm{CP}} due to the optical theorem mentioned above, it is the Bogoliubov-particle part 𝒯BG=sin2δ\mathcal{T}_{\mathrm{BG}}=\sin^{2}\delta that shows the genuine Kondo ridge, as demonstrated in Fig. 4(b). The single-particle contribution 𝒯ET\mathcal{T}_{\mathrm{ET}} decreases outside of the Kondo regime EAU/2E_{A}\gtrsim U/2, at which the occupation number QQ of Bogoliubov particles rapidly decreases and the SC pair correlation increases, as demonstrated in Figs. 3(a) and 3(b).

The Cooper-pair contribution 𝒯CP\mathcal{T}_{\mathrm{CP}} is also plotted in Fig. 4(b), choosing the Bogoliubov angle to be Θ=π/2\Theta=\pi/2 and multiplying a factor of 2-2 which emerges for the nonlocal conductance gRL𝒯BG2𝒯CPg_{\mathrm{RL}}\propto\mathcal{T}_{\mathrm{BG}}-2\mathcal{T}_{\mathrm{CP}}: the negative sign represents the fact that the crossed Andreev reflection induces the counterflow, flowing from the right lead towards the QD. In this case, Eq. (36) can be rewritten further into a similar form to the current noise of normal electrons: 𝒯CP=sin2δ(1sin2δ)\mathcal{T}_{\mathrm{CP}}=\sin^{2}\delta\,(1-\sin^{2}\delta). Blanter and Büttiker (2000); Oguri et al. (2022) Thus, the contribution of 𝒯CP\mathcal{T}_{\mathrm{CP}} to the nonlocal conductance is maximized in the case at which the phase shift becomes δ=π/4\delta=\pi/4 and it reaches the value 2𝒯CP=12-2\mathcal{T}_{\mathrm{CP}}=-\frac{1}{2}. The corresponding dip emerges in Fig. 4(b) at the crossover region EAU/2E_{A}\simeq U/2, the width of which becomes of the order of ΓN\Gamma_{N}.

Figure 4(c) shows the nonlocal conductance, which takes the form gRL/g0=2sin2δcos2δg_{\mathrm{RL}}/g_{0}=-2\sin^{2}\delta\cos 2\delta at Θ=π/2\Theta=\pi/2, as mentioned. It decreases from the unitary-limit value gRL/g0=1g_{\mathrm{RL}}/g_{0}=1 as EAE_{A} deviates from EA=0E_{A}=0, and vanishes gRL=0g_{\mathrm{RL}}=0 at EAU/2E_{A}\simeq U/2 where the phase shift reaches δ=π/4\delta=\pi/4. The nonlocal conductance becomes negative gRL<0g_{\mathrm{RL}}<0 at EAU/2E_{A}\gtrsim U/2 as the CAR contributions dominate in this region. In particular, it has a dip of the depth gRL/g0=1/4g_{\mathrm{RL}}/g_{0}=-1/4 at the point where the phase shift takes the value δ=π/6\delta=\pi/6.

Similarly, the CAR efficiency takes a simplified form ηCAR=cos2δ\eta_{\text{CAR}}=\cos^{2}\delta at Θ=π/2\Theta=\pi/2, and the NRG results are plotted in Fig. 4(d). The efficiency ηCAR\eta_{\text{CAR}} increases with EAE_{A}, and reaches ηCAR=0.5\eta_{\text{CAR}}=0.5 at δ=π/4\delta=\pi/4 where 2𝒯CP-2\mathcal{T}_{\mathrm{CP}} has the dip of the depth 1/2-1/2 seen in Fig. 4(b). The transient region of ηCAR\eta_{\text{CAR}} varying from 0 to 1 is estimated to be of the order of ΓN\Gamma_{N}. Furthermore, at EAU/2E_{A}\gg U/2, the efficiency approaches the saturation value ηCAR=1.0\eta_{\text{CAR}}=1.0 although the conductance gRLg_{\mathrm{RL}} itself becomes very small.

Refer to caption
Refer to caption
Figure 5: Three-dimensional plots of (a) |dd+dd||\bigl{\langle}d^{\dagger}_{\uparrow}d^{\dagger}_{\downarrow}+d_{\downarrow}d_{\uparrow}\bigr{\rangle}| and (b) nd\left\langle n_{d}\right\rangle, described as functions of ξd\xi_{d} and ΓS\Gamma_{S}, choosing U/(πΓN)=5.0U/(\pi\Gamma_{N})=5.0. Mesh lines are drawn along the polar coordinate (EAE_{A}, Θ\Theta), with EA=ξd2+ΓS2E_{A}=\sqrt{\xi_{d}^{2}+\Gamma_{S}^{2}} and Θ=tan1(ΓS/ξd)\Theta=\tan^{-1}(\Gamma_{S}/\xi_{d}).

IV.3 The characteristics of CAR along the
polar coordinates EAE_{A} and Θ\Theta at b=0b=0

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 6: Transport coefficients (a) 𝒯BG\mathcal{T}_{\mathrm{BG}}, (b) 𝒯CP\mathcal{T}_{\mathrm{CP}}, (c) gRLg_{\mathrm{RL}}, and (e) the CAR efficiency ηCAR\eta_{\text{CAR}} plotted as functions of ξd\xi_{d} and ΓS\Gamma_{S}, for U/(πΓN)=5.0U/(\pi\Gamma_{N})=5.0 and b=0b=0. For these three-dimensional plots, mesh lines are drawn along the polar coordinate (EAE_{A}, Θ\Theta). Two-dimensional plot (d) is a contour map for the region where the nonlocal conductance gRLg_{\mathrm{RL}} becomes negative: contour lines are drawn with 0.050.05 increments. The CAR dominates gRLg_{\mathrm{RL}} over the parameter region ΓSU/2\Gamma_{S}\gtrsim U/2 and π/4<Θ<3π/4\pi/4<\Theta<3\pi/4, i.e., ΓS>|ξd|\Gamma_{S}>|\xi_{d}|.

So far, we have discussed the transport properties at Θ=π/2\Theta=\pi/2, along the vertical ΓS\Gamma_{S} axis in the ξd\xi_{d} vs ΓS\Gamma_{S} plane. As ξd\xi_{d} varies from the electron-hole symmetric point ξd=0\xi_{d}=0, the Bogoliubov angle Θ\Theta deviates from π/2\pi/2. Here we discuss how the ground-state and transport properties vary along the angular direction over the range 0Θπ0\leq\Theta\leq\pi.

Figures 5 and 6 show the NRG results of the renormalized parameters and the transport coefficients as functions of ξd\xi_{d} and ΓS\Gamma_{S} for a relatively large Coulomb interaction U/(πΓN)=5.0U/(\pi\Gamma_{N})=5.0. In these three-dimensional plots, mesh lines are drawn along the polar coordinates (EA,Θ)(E_{A},\Theta). Note that the superconducting coherence factors, cosΘ\cos\Theta and sinΘ\sin\Theta, vary in the angular direction: Cooper pairs are strongly entangled at π/2\pi/2 and the SC proximity effect becomes weak as Θ\Theta deviates towards 0 or π\pi. In contrast, along the radial direction, the crossover between the Kondo regime and valence fluctuation regime of the Bogoliubov particles occurs near the semicircle of radius EA=U/2E_{A}=U/2, as mentioned.

IV.3.1 Θ\Theta dependence of dd+dd\,\bigl{\langle}d^{\dagger}_{\uparrow}d^{\dagger}_{\downarrow}+d_{\downarrow}d_{\uparrow}\bigr{\rangle} and nd\,\langle n_{d}\rangle

Among the renormalized parameters plotted in Fig. 3, the following three, QQ, ZZ, and E~A\widetilde{E}_{A} do not depend on the Bogoliubov angle Θ\Theta, and thus Figs. 3(a), 3(c), and 3(d) remain unchanged as angle Θ\Theta varies. In contrast, the correlation functions which are defined with respect to electrons, such as |dd+dd|=(1Q)sinΘ|\bigl{\langle}d^{\dagger}_{\uparrow}d^{\dagger}_{\downarrow}+d_{\downarrow}d_{\uparrow}\bigr{\rangle}|=(1-Q)\sin\Theta and nd=1+(Q1)cosΘ\langle n_{d}\rangle=1+(Q-1)\cos\Theta, evolve with the Bogoliubov angle Θ\Theta.

We can see in Fig. 5 (a) that the pair correlation is suppressed due to the Kondo effect at EAU/2E_{A}\lesssim U/2, especially along the valley at Θ=π/2\Theta=\pi/2, inside the semicircle shown in Fig. 2. The slope from the valley bottom towards the direction parallel to the ξd\xi_{d}-axis is suppressed by the coherence factor sinΘ\sin\Theta. Correspondingly, in Fig. 5(b), the occupation number nd\langle n_{d}\rangle of electrons clearly shows a plateau of a semicircle shape which spreads around the origin EA=0.0E_{A}=0.0 of the ξd\xi_{d} vs ΓS\Gamma_{S} plane. Note that the occupation number is locked exactly at nd=1.0\langle n_{d}\rangle=1.0 along the ΓS\Gamma_{S} axis. Outside the plateau EAU/2E_{A}\gtrsim U/2, the superconducting proximity effects dominate over the angular range of π/4<Θ<3π/4\pi/4<\Theta<3\pi/4, or equivalently at ΓS>|ξd|\Gamma_{S}>|\xi_{d}|. In particular, the ridge of the pair correlation develops at Θ=π/2\Theta=\pi/2, along the ΓS\Gamma_{S}-axis in Fig. 5(a).

IV.3.2 Θ\Theta dependence of transport properties

Figure 6(a) shows the NRG results of transmission probability of Bogoliubov particles 𝒯BG=sin2δ\mathcal{T}_{\mathrm{BG}}=\sin^{2}\delta calculated for U/(πΓN)=5.0U/(\pi\Gamma_{N})=5.0. It has an isotropic structure independent of Θ\Theta. In particular, the semi-cylindrical elevation of the height 𝒯BG1.0\mathcal{T}_{\mathrm{BG}}\simeq 1.0 at EA0.5UE_{A}\lesssim 0.5U corresponds to a rotating body of the Kondo ridge shown in Fig. 4 (b). On the slopes of this semicylindrical hill at EA0.5UE_{A}\simeq 0.5U, it spreads over the valence fluctuation region of the Bogoliubov particles, at which the transmission probability 𝒯BG\mathcal{T}_{\mathrm{BG}} rapidly decreases.

Figure 6(b) shows the transmission probability of Cooper pairs 𝒯CP=(1/4)sin2Θsin22δ\mathcal{T}_{\mathrm{CP}}=(1/4)\sin^{2}\Theta\,\sin^{2}2\delta. It is enhanced along the ridge of a crescent shape that is spreading over the angular range of π/4<Θ<3π/4\pi/4<\Theta<3\pi/4 (at which ΓS>|ξd|\Gamma_{S}>|\xi_{d}|) on the arc of radius EAU/2E_{A}\simeq U/2, where the crossover between the Kondo-singlet and the superconducting-singlet states takes place. The ridge height of 𝒯CP\mathcal{T}_{\mathrm{CP}} decreases from the maximum value 0.250.25 as Θ\Theta deviates from Θ=π/2\Theta=\pi/2, showing the sin2Θ\sin^{2}\Theta dependence. The width of the crescent region in the radial direction is of the order of ΓN\Gamma_{N} (0.06U\simeq 0.06U in Fig. 6(b) ).

The nonlocal conductance gRL/g0=2(𝒯BG2𝒯CP)g_{\mathrm{RL}}/g_{0}=2(\mathcal{T}_{\mathrm{BG}}-2\mathcal{T}_{\mathrm{CP}}) is shown in Fig. 6(c). It also features a flat-topped semicylindrical elevation at EAU/2E_{A}\lesssim U/2, which is mainly due to the contributions of the Bogoliubov-particle part 𝒯BG\mathcal{T}_{\mathrm{BG}} seen in Fig. 6(a). The nonlocal conductance gRLg_{\mathrm{RL}} becomes negative at the foot of the hill, specifically at EAU/2E_{A}\gtrsim U/2 along the arc of the range π/4<Θ<3π/4\pi/4<\Theta<3\pi/4, where the CAR dominates the transport. In order to see more precisely the profile of the negative-conductance region, a contour plot of gRLg_{\mathrm{RL}} is shown in Fig. 6(d). The dip in the profile becomes deepest at Θ=π/2\Theta=\pi/2 and EA/U0.55E_{A}/U\simeq 0.55, as seen also in Fig. 4(c). The behavior of gRLg_{\mathrm{RL}} along the Θ\Theta direction is determined by the coherence factor sin2Θ\sin^{2}\Theta of the Cooper-pair part 𝒯CP\mathcal{T}_{\mathrm{CP}} in Eq. (36). It suppresses the CAR contributions to the nonlocal conductance as Θ\Theta deviates from π/2\pi/2. The crescent-shaped dip emerged for gRLg_{\mathrm{RL}} reflecting the corresponding one seen in Fig. 6(b) for 𝒯CP\mathcal{T}_{\mathrm{CP}}, and the dip spreads from EAU/2E_{A}\simeq U/2 to EAU/2+ΓNE_{A}\sim U/2+\Gamma_{N} in the direction of the ΓS\Gamma_{S}-axis. These results suggest that the crescent dip region will be a plausible target to probe the CAR contributions in experiments.

The NRG result of the CAR efficiency at b=0b=0, ηCAR=sin2Θcos2δ\eta_{\text{CAR}}=\sin^{2}\Theta\,\cos^{2}\delta, is shown in Fig. 6(e). We can see that its behavior is similar to that of the pair correlation described in Fig. 5(a): the ridge of ηCAR\eta_{\text{CAR}} evolves at EAU/2E_{A}\gtrsim U/2 in the direction of Θ=π/2\Theta=\pi/2 along the ΓS\Gamma_{S} axis. In the valley region at EAU/2E_{A}\lesssim U/2, however, the slope of ηCAR\eta_{\text{CAR}} in the direction parallel to the ξd\xi_{d} axis becomes steeper as it is determined by the coherence factor sin2Θ\sin^{2}\Theta, whereas that for the pair correlation function is sinΘ\sin\Theta. There are also some quantitative differences between the profiles of the CAR efficiency and the pair correlation function in the radial direction: it is because ηCARcos2δ\eta_{\text{CAR}}\propto\cos^{2}\delta, whereas |dd+dd|12δ/π|\bigl{\langle}d^{\dagger}_{\uparrow}d^{\dagger}_{\downarrow}+d_{\downarrow}d_{\uparrow}\bigr{\rangle}|\propto 1-2\delta/\pi.

Refer to caption
Figure 7: Parameter space of HeffH_{\mathrm{eff}} at finite magnetic fields. Near the semicircle of radius EA=U/2+bE_{A}=U/2+b with EA=ξd2+ΓS2E_{A}=\sqrt{\xi_{d}^{2}+\Gamma_{S}^{2}}, the occupation number of the Bogoliubov particles in the Andreev level varies rapidly from Q1.0Q\simeq 1.0 to Q0.0Q\simeq 0.0 for large UU. Specifically, in the atomic limit ΓN0\Gamma_{N}\to 0, the ground state is spin polarized Q1.0Q_{\uparrow}\to 1.0 inside the semicircle at finite fields b0b\neq 0. The Andreev scattering can dominate the transport in the range of π/4<Θ<3π/4\pi/4<\Theta<3\pi/4 outside the semicircle which evolves with bb.

V Crossed-Andreev transport at finite magnetic fields b0b\neq 0

Both the Kondo effect and the superconducting proximity effect are sensitive to a magnetic field. Here we study precisely how the CAR contributions vary at finite magnetic fields.

Figure 7 shows the parameter space of the effective Hamiltonian HeffH_{\mathrm{eff}} at finite magnetic fields (b>0b>0). In the atomic limit ΓN0\Gamma_{N}\to 0, the phase boundary evolves with bb, and the ground state of the isolated QD is spin polarized inside the semicircle of radius EA=U/2+bE_{A}=U/2+b where the Andreev level is occupied by a single Bogoliubov particle with majority spin: Q1.0Q_{\uparrow}\to 1.0. In contrast, outside the semicircle, the Andreev level is empty Q=0Q=0 and the ground state is unpolarized. The transition is caused by the level crossing between the energy level EA,EAbE_{A,\uparrow}\equiv E_{A}-b of the singly occupied majority-spin state and the spin-singlet empty state of energy U/2U/2, and thus it takes place at the circumference of the semicircle EA,=U/2E_{A,\uparrow}=U/2.

The level crossing becomes a gradual crossover, the width of which is of the order of ΓN\Gamma_{N}, when normal leads are connected. The CAR contribution to the nonlocal conductance is enhanced also at finite bb near the crossover region: specifically along the circumference of radius EAU/2+bE_{A}\simeq U/2+b over the angular range of π/4Θ3π/4\pi/4\lesssim\Theta\lesssim 3\pi/4. We will consider magnetic-field dependence of the CAR contribution precisely in this section.

V.1 Ground-state and transport properties at Θ=π/2\Theta=\pi/2

At finite magnetic fields, the renormalized parameters become dependent on spin components σ\sigma and vary as Zeeman splitting increases. Our discussions in the following are based on the transport formulas for the ground state given in Eqs. (26)–(30). The NRG calculations have been carried out for a strong interaction U/(πΓN)=5.0U/(\pi\Gamma_{N})=5.0 in order to clarify how the electron correlations affect the crossed Andreev reflection in the multiterminal network.

Refer to caption Refer to caption
Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 8: Magnetic-field dependence of the renormalized parameters calculated at different Andreev-level positions EA/U=0.0E_{A}/U=0.0, 0.20.2, 0.40.4, 0.50.5, 0.60.6, and 0.80.8, for a fixed interaction U/(πΓN)=5.0U/(\pi\Gamma_{N})=5.0. (a), (b) Occupation number of Bogoliubov particles QQ_{\uparrow} and QQ_{\downarrow}. Inset in (b) is an enlarged plot of QQ_{\downarrow} vs b/Tb/T^{*}, with TT^{*} the characteristic energy scale of the Kondo regime defined at b=0b=0 in Eq. (40). For EA=0.0E_{A}=0.0, 0.2U0.2U, and 0.4U0.4U, it takes the value T/(πΓN)=0.002T^{*}/(\pi\Gamma_{N})=0.002, 0.0050.005, and 0.0970.097, respectively (or T/U=0.0004T^{*}/U=0.0004, 0.0010.001, and 0.0190.019). The values of Z=Γ~N/ΓNZ=\widetilde{\Gamma}_{N}/\Gamma_{N} at these three points of EAE_{A} are given by Z=0.008Z=0.008, 0.020.02, 0.310.31, respectively. (c) Magnetization md=ndndm_{d}=\langle n_{d\uparrow}\rangle-\langle n_{d\downarrow}\rangle, which does not depend on the Bogoliubov angle Θ\Theta. (d) Enlarged view of mdm_{d} for EA=0.0E_{A}=0.0, 0.2U0.2U, and 0.4U0.4U plotted vs b/Tb/T^{*} in the Kondo regime. (e) Pair correlation |dd+dd||\bigl{\langle}d^{\dagger}_{\uparrow}d^{\dagger}_{\downarrow}+d_{\downarrow}d_{\uparrow}\bigr{\rangle}| at Θ=π/2\Theta=\pi/2, which in this case is given by 1Q1-Q and varies with bb and EAE_{A}, in contrast to the electron filling nd1.0\langle n_{d}\rangle\equiv 1.0 that remains unchanged along the ΓS\Gamma_{S} axis at ξd=0\xi_{d}=0. (f) Renormalized Andreev levels E~A,\widetilde{E}_{A,\uparrow}. The dashed straight lines represent the Hartree-Fock (HF) result EAHFEAU/2EAU/2bE_{A}^{\mathrm{HF}}\xrightarrow{\,E_{A}\gg U/2\,}E_{A}-U/2-b.

V.1.1 bb dependence of renormalized parameters

Figures 8(a)–8(f) show the magnetic-field dependence of the renormalized parameters, calculated for several different positions of the Andreev level: EA/U=0.0E_{A}/U=0.0, 0.20.2, 0.40.4, 0.50.5, 0.60.6, and 0.80.8. The results commonly reflect the Fermi-liquid properties of the Bogoliubov particles, which evolve as EAE_{A} and bb vary.

For 0EAU/20\leq E_{A}\lesssim U/2, the renormalized parameters exhibit a universal b/Tb/T^{*} scaling behavior at small magnetic fields, with TT^{*} the Kondo energy scale defined at zero field b=0b=0 by Eq. (40). The magnitude of TT^{*} depends sensitively on the interaction strength and the level position EAE_{A}, and, for instance, for U/(πΓN)=5.0U/(\pi\Gamma_{N})=5.0, it is given by T/(πΓN)=0.002T^{*}/(\pi\Gamma_{N})=0.002, 0.0050.005, and 0.0970.097 for EA/U=0.0E_{A}/U=0.0, 0.20.2, and 0.40.4, respectively. At the magnetic field of order at bTb\sim T^{*}, the Kondo resonance of Bogoliubov particles splits into two as the Zeeman effect dominates. In contrast, in the parameter region of EAU/2E_{A}\gtrsim U/2 where the Andreev level deviates further from the Fermi level, the renormalization effects due to the electron correlations are suppressed, and the low-energy states depend significantly on whether U/2EAU/2+bU/2\lesssim E_{A}\lesssim U/2+b or U/2+bEAU/2+b\lesssim E_{A}. The magnetization mdm_{d} of quantum dot is almost fully polarized at U/2EAU/2+bU/2\lesssim E_{A}\lesssim U/2+b, where the Zeeman effect dominates. In contrast, the superconducting proximity effect dominates outside the semicircle of radius EAU/2+bE_{A}\gtrsim U/2+b in the angular range of π/4<Θ<3π/4\pi/4<\Theta<3\pi/4. We will discuss these of variations of the ground state properties in more detail in the following.

Figures 8(a)–8(d) describe the occupation number QσQ_{\sigma} and the magnetization mdndndm_{d}\equiv\langle n_{d\uparrow}\rangle-\langle n_{d\downarrow}\rangle as functions of magnetic fields. Note that two of them, Fig. 8(d) and the inset presented for QQ_{\downarrow} in Fig. 8(b), are plotted vs b/Tb/T^{*} for small magnetic fields. We see in Fig. 8(d) that the magnetization mdm_{d} for EA/U0.4E_{A}/U\lesssim 0.4 exhibits the universal Kondo scaling behavior at bTb\lesssim T^{*}. It can also be recognized that the three curves for QQ_{\downarrow} shown in the inset in Fig. 8(b) will collapse into a single universal curve if we introduce the offset values along the vertical axis which is determined at b=0b=0 for each EAE_{A}: note that the occupation number takes the value Qσ=0.5Q_{\sigma}=0.5 at EA=b=0E_{A}=b=0.

However, as seen in Figs. 8(a) and 8(c), the Zeeman effect dominates at strong magnetic fields. Note that the magnetization can also be written as md=QQm_{d}=Q_{\uparrow}-Q_{\downarrow} and does not depend on the Bogoliubov angle Θ\Theta. The Kondo behavior disappears for EA/U0.5E_{A}/U\gtrsim 0.5, at which the Bogoliubov particles are in the valence fluctuation or empty orbital regime at b=0b=0. In this region of EA/UE_{A}/U, the occupation number QQ_{\uparrow} of the majority-spin component and the magnetization mdm_{d} show a steep increase at magnetic fields of bEAU/2b\simeq E_{A}-U/2 where the level crossing occurs. As magnetic field increases further bEAU/2b\gtrsim E_{A}-U/2, the magnetization rapidly approaches the saturation value md1.0m_{d}\to 1.0.

Figure 8(e) shows the magnetic-field dependence of the SC pair correlation function |dd+dd||\bigl{\langle}d^{\dagger}_{\uparrow}d^{\dagger}_{\downarrow}+d_{\downarrow}d_{\uparrow}\bigr{\rangle}| which becomes equal to 1Q1-Q in the direction of Θ=π/2\Theta=\pi/2. While the pair correlation increases with EAE_{A}, it decreases as bb increases. We can see that the SC proximity effect dominates at small fields bEAU/2b\lesssim E_{A}-U/2 in the parameter region of EAU/2E_{A}\gtrsim U/2, i.e, the outside of the semicircle of radius EAU/2+bE_{A}\gtrsim U/2+b shown in Fig. 7. In this region, the SC pair correlation function can take the value of the order of 10% of the maximum possible value |dd+dd|=1|\bigl{\langle}d^{\dagger}_{\uparrow}d^{\dagger}_{\downarrow}+d_{\downarrow}d_{\uparrow}\bigr{\rangle}|=1, as seen in Fig.8(e) at magnetic fields of the order of b0.1Ub\sim 0.1U. However, as magnetic field approaches bEAU/2b\simeq E_{A}-U/2, the crossover to the Zeeman-dominated spin-polarized regime occurs, and the pair correlation rapidly decreases. The sum of the phase shifts takes the value δ+δπ/2\delta_{\uparrow}+\delta_{\downarrow}\simeq\pi/2 in the crossover region. Therefore, the Andreev scattering is most enhanced at this point since the factor sin2(δ+δ)\sin^{2}(\delta_{\uparrow}+\delta_{\downarrow}) that determines 𝒯CP\mathcal{T}_{\mathrm{CP}} takes the maximum value.

Figure 8(f) shows the results for the majority-spin component of the renormalized Andreev level E~A,\widetilde{E}_{A,\uparrow} which includes the Zeeman energy and the many-body corrections defined in Eq. (17). For EAU/2E_{A}\lesssim U/2, the slope of E~A,\widetilde{E}_{A,\uparrow} is steep at small magnetic fields b0b\simeq 0. This is because the spin susceptibility, χs=md/b\chi_{s}=m_{d}/b, is enhanced in this region by the Kondo effect as seen in Fig. 8(c). The slope becomes gradual, however, as bb increases and the crossover to the Zeeman-dominated regime occurs at bTb\sim T^{*}. In contrast, for EAU/2E_{A}\gtrsim U/2, the renormalized level E~A,\widetilde{E}_{A,\uparrow} at small magnetic fields b0b\simeq 0 shifts away from the Fermi level, and the occupation number of the Bogoliubov particles QQ decreases as EAE_{A} increases. However, as magnetic field increases further, the renormalized level E~A,\widetilde{E}_{A,\uparrow} crosses the Fermi level at bEAU/2b\simeq E_{A}-U/2, and the occupation number of the majority spin component QQ_{\uparrow} increases abruptly at the crossing point. The dashed straight lines in Fig. 8(f) represent the Hartree-Fock energy shift EA,HFE_{A,\uparrow}^{\mathrm{HF}}, which asymptotically takes the following form at EAU/2E_{A}\gg U/2:

EA,HFEAb+U(Q12)\displaystyle E_{A,\uparrow}^{\mathrm{HF}}\,\equiv\,E_{A}-b\,+\,U\left(Q_{\downarrow}-\frac{1}{2}\right)
EAU/2EAU2b.\displaystyle\ \ \xrightarrow{\,E_{A}\gg U/2\,}\,E_{A}-\frac{U}{2}-b\,. (42)

The NRG results for E~A,\widetilde{E}_{A,\uparrow} and the Hartree-Fock energy shifts show a close agreement for EAU/2E_{A}\gtrsim U/2. This is caused by the fact that the renormalization factor approaches Zσ1.0Z_{\sigma}\simeq 1.0 and it becomes less important at the crossover region between the Zeeman-dominated spin-polarized regime and the SC-dominated regime.

Refer to caption Refer to caption
Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 9: Magnetic-field dependence of the transport coefficients at Θ=π/2\Theta=\pi/2 for different positions of EA/U=0E_{A}/U=0, 0.20.2, 0.40.4, 0.50.5, 0.60.6, and 0.80.8, for a strong interaction U/(πΓN)=5.0U/(\pi\Gamma_{N})=5.0. Top panel describes 𝒯BG=σsin2δσ/2\mathcal{T}_{\mathrm{BG}}=\sum_{\sigma}\sin^{2}{\delta_{\sigma}}/2, and Θ\Theta dependent part 2𝒯CP-2\mathcal{T}_{\mathrm{CP}} for (a) small EA0.5UE_{A}\lesssim 0.5U and (b) large EA0.5UE_{A}\gtrsim 0.5U. (c), (d) Nonlocal conductance gRL/g0g_{\mathrm{RL}}/g_{0}. Inset in (d) is an enlarged view in the region around gRL/g00.0g_{\mathrm{RL}}/g_{0}\approx 0.0: the dashed lines represent the perturbation results obtained with Eq. (43). The characteristic energy is given by T/U=0.0004T^{*}/U=0.0004, 0.0010.001, 0.0190.019, and 0.097 for EA/U=0.0E_{A}/U=0.0, 0.20.2, 0.40.4, and 0.50.5, respectively. (e) CAR efficiency ηCAR\eta_{\text{CAR}}. (f) gRL/g0g_{\mathrm{RL}}/g_{0} at a fixed Andreev level position EA=0.8UE_{A}=0.8U for several different angles Θ=0\Theta=0, π/8\pi/8, π/4\pi/4, 3π/83\pi/8, π/2\pi/2: the dashed lines here also represent the perturbation results.

V.1.2 bb dependence of transport properties at Θ=π/2\Theta=\pi/2

We next discuss the magnetic-field dependence of transport coefficients in the direction of Θ=π/2\Theta=\pi/2, i.e., along the ΓS\Gamma_{S} axis (ξd=0\xi_{d}=0). The NRG results are shown in Fig. 9: the magnetic field bb is scaled by TT^{*} in two of the panels 9(a) and 9(c), whereas the other panels are plotted vs b/Ub/U.

We see in Fig. 9(a) that the transmission probabilities of Bogoliubov particles 𝒯BG\mathcal{T}_{\mathrm{BG}} for EA0.2UE_{A}\lesssim 0.2U collapse into a single curve at small magnetic fields bTb\lesssim T^{*}, showing a universal b/Tb/T^{*} Kondo scaling behavior, whereas the probability of the Cooper-pairs 𝒯CP\mathcal{T}_{\mathrm{CP}} is suppressed in this region. Correspondingly, the nonlocal conductance gRLg_{\mathrm{RL}} exhibits the scaling behavior for EA0.2UE_{A}\lesssim 0.2U, as shown in Fig. 9(c). The results for 𝒯BG\mathcal{T}_{\mathrm{BG}} and gRLg_{\mathrm{RL}} at EA=0.4UE_{A}=0.4U still show a similar monotonic decrease although they did not collapse into the universal curves. Therefore, the CAR efficiencies ηCAR=𝒯CP/𝒯BG\eta_{\mathrm{CAR}}=\mathcal{T}_{\mathrm{CP}}/\mathcal{T}_{\mathrm{BG}} for EA/U=0.2E_{A}/U=0.2 and 0.40.4, described in Fig. 9(e), increase clearly with bb at small magnetic fields near b0b\simeq 0. It shows that the Zeeman splitting suppresses the Kondo correlations and assists the contributions of the Cooper-pair tunneling.

In contrast, when the Andreev level situates further away from the Fermi level EAU/2E_{A}\gtrsim U/2, the ground state evolves from the superconductivity-dominated regime to the Zeeman dominated spin-polarized regime, as magnetic field increases. In particular, at bEAU/2b\simeq E_{A}-U/2, i.e., the crossover region between these two regimes, the transmission probability of the Bogoliubov particles 𝒯BG=(sin2δ+sin2δ)/2\mathcal{T}_{\mathrm{BG}}=\left(\sin^{2}\delta_{\uparrow}+\sin^{2}\delta_{\downarrow}\right)/2 has a peak, which emerges in Fig. 9(b), as the phase shifts take the value δπ/2\delta_{\uparrow}\simeq\pi/2 and δ0\delta_{\downarrow}\simeq 0. Similarly, the Cooper-pair contribution 𝒯CP=(1/4)sin2Θsin2(δ+δ)\mathcal{T}_{\mathrm{CP}}=(1/4)\sin^{2}\Theta\,\sin^{2}\bigl{(}\delta_{\uparrow}+\delta_{\downarrow}) reaches the maximum value 1/41/4 at bEAU/2b\simeq E_{A}-U/2 in the crossover region. This is consistent with the previous work that examined an N-QD-SC system with the modified second-order perturbation theory. Yamada et al. (2007) It revealed the fact that the Andreev scattering is significantly enhanced under the condition that the renormalized parameters satisfy σ(E~A,σ/Γ~N,σ)=1\prod_{\sigma}(\widetilde{E}_{A,\sigma}/\widetilde{\Gamma}_{N,\sigma})=1: this can be rewritten into the form cotδcotδ=1\cot\delta_{\uparrow}\cot\delta_{\downarrow}=1 and is fulfilled at δ+δ=π/2\delta_{\uparrow}+\delta_{\downarrow}=\pi/2.

We can see in Fig. 9(d) that, for EAU/2E_{A}\gtrsim U/2, the nonlocal conductance gRL=𝒯BG2𝒯CPg_{\mathrm{RL}}=\mathcal{T}_{\mathrm{BG}}-2\mathcal{T}_{\mathrm{CP}} becomes negative in the SC-dominated regime bEAU/2b\lesssim E_{A}-U/2, whereas gRLg_{\mathrm{RL}} takes a positive value in the Zeeman-dominated regime bEAU/2b\gtrsim E_{A}-U/2. In particular, for EA0.6UE_{A}\gtrsim 0.6U, the nonlocal conductance forms a flat valley structure at 0bEAU/20\leq b\lesssim E_{A}-U/2 the bottom of which is negative and is less sensitive to bb. This is caused by the fact that the peak of 𝒯BG\mathcal{T}_{\mathrm{BG}} and the dip of 2𝒯CP-2\mathcal{T}_{\mathrm{CP}} move almost synchronously, in Fig. 9(b), as EAE_{A} increases from 0.5U0.5U. For observing the flat valley structure of gRLg_{\mathrm{RL}}, the depth of which should not be too shallow, and thus EAU/2E_{A}-U/2 should be of the order of ΓN\Gamma_{N}, or should not be too much larger than ΓN\Gamma_{N}. Note that, in this magnetic-field region 0bEAU/20\leq b\lesssim E_{A}-U/2, the occupation number of the Bogoliubov particles with the minority spin is almost empty Q0.0Q_{\downarrow}\simeq 0.0 and the transport coefficients are determined by the majority-spin component QQ_{\uparrow}. In order to verify this quantitatively, we expand the nonlocal conductance into the following form, keeping the terms up to the first order with respect to δ\delta_{\downarrow},

gRL\displaystyle g_{\mathrm{RL}}\,\approx g0[cos2Θsin2δδsin2Θsin2δ]+O(δ2)\displaystyle\ \,g_{0}\Bigl{[}\,\cos^{2}\Theta\,\sin^{2}\delta_{\uparrow}-\delta_{\downarrow}\sin^{2}\Theta\,\sin 2\delta_{\uparrow}\,\Bigr{]}\,+\,O\!\left(\delta_{\downarrow}^{2}\right)
Θ=π2\displaystyle\xrightarrow{\,\Theta=\frac{\pi}{2}\,} g0[δsin2δ+O(δ2)].\displaystyle\ g_{0}\,\left[\,-\delta_{\downarrow}\,\sin 2\delta_{\uparrow}\,+\,O\!\left(\delta_{\downarrow}^{2}\right)\,\right]\,. (43)

The dashed lines plotted in the inset of Fig. 9(d) are the results evaluated with Eq. (43), using the NRG results for δσ\delta_{\sigma}. These results show close agreement with the full NRG calculations of gRLg_{\mathrm{RL}} plotted with the symbols, i.e., for EA/U=0.6E_{A}/U=0.6 (\bullet) and 0.80.8 (\square).

So far, we have considered behaviors along the angular direction of Θ=π/2\Theta=\pi/2. Figure 9(f) compares the magnetic-field dependence of gRLg_{\mathrm{RL}} for several different angles Θ\Theta, keeping the Andreev-level position at EA=0.8UE_{A}=0.8U. The dashed lines, which also show nice agreement with the full NRG results (symbols) of gRLg_{\mathrm{RL}} in this figure, are the perturbation results obtained from Eq. (43). We find that the flat structure with negative gRLg_{\mathrm{RL}} remains for Θ=3π/8\Theta=3\pi/8, where the angle largely deviates from π/2\pi/2. As Θ\Theta derives further, however, gRLg_{\mathrm{RL}} becomes positive at 0<Θ<π/40<\Theta<\pi/4, or 3π/4<Θ<π3\pi/4<\Theta<\pi. Note that the Θ\Theta dependence enters the nonlocal conductance through the coherence factor sin2Θ\sin^{2}\Theta in 𝒯CP\mathcal{T}_{\mathrm{CP}}, and thus gRLg_{\mathrm{RL}} is symmetrical with respect to the ΓS\Gamma_{S}-axis in parameter space shown in Fig. 7.

In the SC-dominated regime 0bEAU/20\leq b\lesssim E_{A}-U/2, both components of the phase shift approach zero as EAE_{A} increases keeping bb unchanged, i.e., δ0\delta_{\uparrow}\simeq 0 and δ0\delta_{\downarrow}\simeq 0 as seen in Figs. 8(a) and 8(b). The CAR efficiency ηCAR\eta_{\mathrm{CAR}} is enhanced in this region although it decreases as bb increases, as seen in Fig. 9(e) for EA/U=0.6E_{A}/U=0.6 and 0.80.8. In particular, for EAUE_{A}\gg U, the efficiency approaches saturation value ηCAR1.0\eta_{\mathrm{CAR}}\to 1.0 at small magnetic fields near b0.0b\simeq 0.0.

V.2 The characteristics of CAR along the
polar coordinates EAE_{A} and Θ\Theta at b0b\neq 0

In this subsection, we consider the Θ\Theta dependence of the transport properties at finite magnetic fields in more detail. Specifically, in order to clarify in which situations the CAR contribution can dominate the nonlocal conductance by overcoming the disturbance of the SC proximity effects by the Coulomb interaction and magnetic field, we explore a wide range of the parameter space, i.e., the ξd\xi_{d} vs ΓS\Gamma_{S} plane. Our discussion here is based on the NRG results in Figs. 10 and 11, obtained for a relatively large interaction U/(πΓN)=5.0U/(\pi\Gamma_{N})=5.0: the Kondo temperature in this case is given by TK/U=0.0004T_{K}/U=0.0004, which is defined as the value of the characteristic energy scale TT^{\ast} at EA=b=0E_{A}=b=0. These results can be compared with those for zero field presented in Figs. 5 and 6.

Refer to caption Refer to caption Refer to caption
Figure 10: Contour maps of |dd+dd||\bigl{\langle}d^{\dagger}_{\uparrow}d^{\dagger}_{\downarrow}+d_{\downarrow}d_{\uparrow}\bigr{\rangle}| at finite magnetic fields: (a) b=TKb=T_{K} and (b) b=0.1Ub=0.1U, for U/(πΓN)=5.0U/(\pi\Gamma_{N})=5.0. Here, TK=0.0004UT_{K}=0.0004U is defined as the value of TT^{\ast} at EA=b=0E_{A}=b=0. The contours are drawn with 0.1 increments, and the dashed line represents the semicircle of radius EA=U/2+bE_{A}=U/2+b.
Refer to caption Refer to caption Refer to caption
Refer to caption Refer to caption Refer to caption
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 11: Bogoliubov-angle dependence of transport coefficients in a strong interaction case U/(πΓN)=5.0U/(\pi\Gamma_{N})=5.0 at small b=TKb=T_{K} and large b=0.1Ub=0.1U fields, with TK/U=0.0004T_{K}/U=0.0004 defined as the value of TT^{\ast} at EA=b=0E_{A}=b=0. In the top panels for (a) b=0b=0, (b) b=TKb=T_{K}, and (c) b=0.1Ub=0.1U, the coefficients 𝒯BG\mathcal{T}_{\mathrm{BG}}, 𝒯CP\mathcal{T}_{\mathrm{CP}}, and CAR efficiency ηCAR\eta_{\mathrm{CAR}} at Θ=π/2\Theta=\pi/2 are plotted vs EA/UE_{A}/U. (d), (e) Contour maps of 𝒯CP\mathcal{T}_{\mathrm{CP}}, drawn with 0.05 increments. (f), (g) Contour maps of ηCAR\eta_{\mathrm{CAR}}, drawn with 0.1 increments. (h)–(k) Nonlocal conductance gRL/g0g_{\mathrm{RL}}/g_{0}. In particular, (j) and (k) are the contour maps for negative conductance region gRL<0g_{\mathrm{RL}}<0, for which contours are drawn with 0.050.05 increments. The dashed semicircle of radius EA=b+U/2E_{A}=b+U/2 in (d)–(g) and (j)–(k) corresponds to the one in Fig. 7.

V.2.1 Θ\Theta dependence of dd+dd\bigl{\langle}d^{\dagger}_{\uparrow}d^{\dagger}_{\downarrow}+d_{\downarrow}d_{\uparrow}\bigr{\rangle} at b0b\neq 0

Figure 10 shows the pair correlation |dd+dd|=(1Q)sinΘ|\bigl{\langle}d^{\dagger}_{\uparrow}d^{\dagger}_{\downarrow}+d_{\downarrow}d_{\uparrow}\bigr{\rangle}|=(1-Q)\sin\Theta for two different magnetic-field strengths: (a) b=TKb=T_{K} and (b) b=0.1Ub=0.1U. Here, the occupation number QQ of Bogoliubov particles does not depend on Θ\Theta but varies with bb and EAE_{A}, as mentioned and shown in Figs. 8(a) and 8(b).

The pair correlation function for a small magnetic field b=TKb=T_{K}, shown in Fig. 10(a), is enhanced outside the semicircle of radius EAU/2+bE_{A}\gtrsim U/2+b in the angular range of π/4<Θ<3π/4\pi/4<\Theta<3\pi/4, especially along the ΓS\Gamma_{S}-axis (Θ=π/2\Theta=\pi/2), where the SC proximity effects dominate. In contrast, it is suppressed by the Kondo effect inside the semicircle EAU/2+bE_{A}\lesssim U/2+b. Note that bb is much smaller than UU in this case (TK/U=0.0004T_{K}/U=0.0004).

Figure 10(b) shows the pair correlation function for a large magnetic field b=0.1Ub=0.1U. Although it is qualitatively similar to Fig. 10(a), we can see that the slope just inside of the circumference becomes steeper than that for b=TKb=T_{K}. The radius of the dashed semicircle at the crossover region in this case is EAU/2+bE_{A}\simeq U/2+b (=0.6U=0.6U), and thus the expansion of the circumference due to bb becomes visible in Fig. 10(b).

V.2.2 Θ\Theta dependence of transport properties at b0b\neq 0

The top panels of Fig. 11 show 𝒯BG\mathcal{T}_{\mathrm{BG}}, 𝒯CP\mathcal{T}_{\mathrm{CP}}, and ηCAR\eta_{\mathrm{CAR}} as functions of EAE_{A} for three different magnetic fields: (a) b=0b=0, (b) b=TKb=T_{K}, and (c) b=0.1Ub=0.1U, taking the Bogoliubov angle to be Θ=π/2\Theta=\pi/2, i.e., the direction in which the SC proximity effect is most enhanced. We can see that, as bb increases, the peak of the Cooper-pair tunneling part 𝒯CP\mathcal{T}_{\mathrm{CP}}, emerging at EAU/2+bE_{A}\simeq U/2+b, moves with the crossover region towards the larger EAE_{A} side along the horizontal axis. The peak height is 1/41/4 and the width becomes of the order of ΓN\Gamma_{N} (0.06U\simeq 0.06U in this case). The Bogoliubov-particle part 𝒯BG\mathcal{T}_{\mathrm{BG}} exhibits the flat Kondo plateau at zero field, plotted in Fig. 11(a) for comparison. However, the Zeeman splitting dominates at magnetic fields of the order of bTKb\simeq T_{K} and the top of the Kondo plateau caves in around EA0.0E_{A}\simeq 0.0, as seen in Fig. 11(b). As magnetic field increases further, the peak of 𝒯BG\mathcal{T}_{\mathrm{BG}} in Fig. 11(c) becomes small and approaches the peak of 𝒯CP\mathcal{T}_{\mathrm{CP}} that situates close to the crossover region.

The CAR efficiency ηCAR=𝒯CP/𝒯BG\eta_{\mathrm{CAR}}=\mathcal{T}_{\mathrm{CP}}/\mathcal{T}_{\mathrm{BG}} plotted in Fig. 11(c) takes relatively large value 0.1ηCAR0.50.1\lesssim\eta_{\mathrm{CAR}}\lesssim 0.5 even at EAU/2+bE_{A}\lesssim U/2+b. Such a behavior is not seen for small bb in Figs. 11(a) and 11(b), and reflects the suppression of 𝒯BG\mathcal{T}_{\mathrm{BG}} caused by a large magnetic field b=0.1Ub=0.1U. Outside the crossover region EAU/2+bE_{A}\gtrsim U/2+b, however, ηCAR\eta_{\mathrm{CAR}} shows a similar behavior in Figs. 11(a)–11(c): it approaches the saturation value ηCAR1.0\eta_{\mathrm{CAR}}\to 1.0 as EAE_{A} increases.

The Bogoliubov angle Θ\Theta enters the nonlocal conductance gRLg_{\mathrm{RL}} through 𝒯CP\mathcal{T}_{\mathrm{CP}} since the Bogoliubov part 𝒯BG\mathcal{T}_{\mathrm{BG}} is independent of it. Figures 11(d) and 11(e) show the contour maps of 𝒯CP\mathcal{T}_{\mathrm{CP}} described on the ξd\xi_{d} vs ΓS\Gamma_{S} plane, for magnetic fields of (d) b=TKb=T_{K} and (e) b=0.1Ub=0.1U. The Cooper-pair tunneling part 𝒯CP\mathcal{T}_{\mathrm{CP}} is enhanced along in the crescent-shaped region on the arc of radius EA=U/2+bE_{A}=U/2+b in the angular range of π/4<Θ<3π/4\pi/4<\Theta<3\pi/4. The crescent region spreads over the direction of the ΓS\Gamma_{S}-axis with the width of the order of ΓN\Gamma_{N} (0.06U\simeq 0.06U in this case). As the magnetic field increases, the crescent region moves upward along the ΓS\Gamma_{S}-axis, together with the arc indicated as a dashed semicircle in Figs. 11(d) and 11(e). This evolution of the crescent region causes the CAR-dominated flat structure of nonlocal conductance gRLg_{\mathrm{RL}} that emerged in the magnetic-field dependence in Figs. 9(d) and 9(f).

Figures 11(f) and 11(g) show the contour maps of the CAR efficiency ηCAR\eta_{\mathrm{CAR}} for magnetic fields of (f) b=TKb=T_{K} and (g) b=0.1Ub=0.1U. Figure 11(f) captures the typical profile of ηCAR\eta_{\mathrm{CAR}} at small fields of order bTKb\simeq T_{K}: the CAR efficiency is enhanced in the SC-dominated regime EAU/2+bE_{A}\gtrsim U/2+b and π/4<Θ<3π/4\pi/4<\Theta<3\pi/4, whereas it decreases rapidly outside this region, especially just inside the semicircle, EAU/2+bE_{A}\lesssim U/2+b, in the edge of the Zeeman-dominated spin-polarized regime. It reflects the steep slope along the direction of Θ=π/2\Theta=\pi/2, seen in Fig. 11(b), at the crossover region EAU/2+bE_{A}\simeq U/2+b. In contrast, at large fields of order b0.1Ub\simeq 0.1U, the corresponding slope of ηCAR\eta_{\mathrm{CAR}} inside the semicircle shows a slow modest decline as seen in Fig. 11(g). This modest decline of ηCAR=𝒯CP/𝒯BG\eta_{\mathrm{CAR}}=\mathcal{T}_{\mathrm{CP}}/\mathcal{T}_{\mathrm{BG}} is caused by the behavior of the transmission probability of the Bogoliubov particles in the denominator that is suppressed in the Zeeman-dominated regime for large magnetic fields, as seen in Fig. 11(c).

Figures 11(h) and 11(i) describe three-dimensional plots of the nonlocal conductance gRL/g0g_{\mathrm{RL}}/g_{0} for magnetic fields of (h) b=TKb=T_{K} and (i) b=0.1Ub=0.1U, respectively. These two examples show quite different behaviors inside the semicircle of radius EAU/2+bE_{A}\lesssim U/2+b. While the Kondo plateau of gRLg_{\mathrm{RL}} starts to cave in around EA=0E_{A}=0 for magnetic fields of the order of TKT_{K}, it is significantly suppressed at large magnetic fields of order 0.1U0.1U, almost in the whole region inside the semicircle EAU/2+bE_{A}\lesssim U/2+b, except for the rim of the semicircle. However, in both cases, there spreads commonly a CAR-dominated region with negative nonlocal conductance, outside the semicircle EAU/2+bE_{A}\gtrsim U/2+b in the direction of the ΓS\Gamma_{S}-axis, which also emerges at zero magnetic field in Fig. 6(c).

Figures 11(j) and 11(k) are the contour maps of the region, at which the nonlinear conductance becomes negative gRL<0.0g_{\mathrm{RL}}<0.0, for magnetic fields of (j) b=TKb=T_{K} and (k) b=0.1Ub=0.1U. It spreads in the ξd\xi_{d} vs ΓS\Gamma_{S} plane, over the region of EAU/2+bE_{A}\gtrsim U/2+b and π/4<Θ<3π/4\pi/4<\Theta<3\pi/4. These plots clearly show that the CAR contribution is enhanced, particularly at the crescent region just outside the circumference of the dashed semicircle. The CAR dominates the nonlocal conductance in this region, and the dip structure of gRLg_{\mathrm{RL}} still remains for finite magnetic fields of order 0.1U0.1U although the depth decreases as bb increases. Furthermore, these results demonstrate how the flat structure can emerge in the magnetic-field-dependence of gRLg_{\mathrm{RL}}, seen in Figs. 9(d) and 9(f). For example, at the point (EA=0.6U,Θ=π/2)(E_{A}=0.6U,\Theta=\pi/2) in the ξd\xi_{d} vs ΓS\Gamma_{S} plane situates in the dip region of gRLg_{\mathrm{RL}} when the magnetic field bb varies from 0 to the order 0.1U0.1U.

These results suggest that, in order to experimentally probe the CAR contributions in the nonlocal conductance gRLg_{\mathrm{RL}}, this crescent region will be a plausible target to be examined. The CAR-dominated transport occurs in the parameter region ΓSU/2+b\Gamma_{S}\gtrsim U/2+b, where the Cooper pairs can penetrate into quantum dots, overcoming the repulsive interaction and magnetic field. Although we have chosen a rather strong interaction U/(πΓN)=5.0U/(\pi\Gamma_{N})=5.0 in this section, the sweet spot for the measurements, at which gRLg_{\mathrm{RL}} exhibits a dip structure, emerges for any UU, as demonstrated in Fig. 4(c) for b=0b=0.

V.3 Spin-polarized current between normal leads

So far, we have mainly considered the charge transport. In particular, we have seen in Fig. 11(i) that for a magnetic field of b=0.1Ub=0.1U, the nonlocal conductance has a peak in the angular directions Θ0\Theta\simeq 0 and Θπ\Theta\simeq\pi, along the rim of the semicircle of radius EAU/2+bE_{A}\simeq U/2+b. Here we discuss the resonant spin-polarized current which is significantly enhanced in this region where the crossover takes place between the Zeeman-dominated regime and the SC-proximity-dominated regime.

The spin current IR,spinIR,IR,I_{R,\mathrm{spin}}\equiv I_{R,\uparrow}-I_{R,\downarrow} flowing from the quantum dot to the right lead can be expressed in the following form, as shown in Appendix D:

IR,spin=\displaystyle I_{R,\mathrm{spin}}\,= e2h4ΓLΓRΓN2𝒯spin(VLVR),\displaystyle\ \frac{e^{2}}{h}\,\frac{4\Gamma_{L}\Gamma_{R}}{\Gamma_{N}^{2}}\,\mathcal{T}_{\mathrm{spin}}\,\left(V_{L}-V_{R}\right)\,, (44)

where 𝒯spin=(sin2δsin2δ)cosΘ\mathcal{T}_{\mathrm{spin}}=\left(\sin^{2}\delta_{\uparrow}-\sin^{2}\delta_{\downarrow}\right)\cos\Theta. The magnetic-field dependence of the spin current is determined by the difference sin2δsin2δ\sin^{2}\delta_{\uparrow}-\sin^{2}\delta_{\downarrow} between the transmission probability of the \uparrow-spin and that of the \downarrow-spin Bogoliubov particles. Similarly, the normalized current polarization is defined by López and Sánchez (2003); Souza et al. (2008); Hitachi et al. (2006); Kiyama et al. (2015)

PR\displaystyle P_{R}\, IR,IR,IR,+IR,ΓL=ΓRsin2δsin2δsin2δ+sin2δcosΘ.\displaystyle\equiv\frac{I_{R,\uparrow}-I_{R,\downarrow}}{I_{R,\uparrow}+I_{R,\downarrow}}\xrightarrow{\,\Gamma_{L}=\Gamma_{R}\,}\frac{\sin^{2}\delta_{\uparrow}-\sin^{2}\delta_{\downarrow}}{\sin^{2}\delta_{\uparrow}+\sin^{2}\delta_{\downarrow}}\cos\Theta. (45)

Figure 12 shows the NRG result of the spin-resolved transport coefficients calculated at a magnetic field of b=0.1Ub=0.1U, for a strong interaction U/(πΓN)=5.0U/(\pi\Gamma_{N})=5.0. In this case, the renormalized Andreev level for the majority spin E~A,\widetilde{E}_{A,\uparrow} crosses the Fermi level at EAU/2+bE_{A}\simeq U/2+b since E~A,\widetilde{E}_{A,\uparrow} can be approximated by the Hartree-Fock energy shift, defined in Eq. (42), in the crossover region.

Figure 12(a) shows that the resonant tunneling of the unitary limit sin2δ=1\sin^{2}\delta_{\uparrow}=1 occurs for the majority \uparrow-spin Bogoliubov particles, whereas the minority one sin2δ\sin^{2}\delta_{\downarrow} is very small and does not give any significant contribution to the current. Note that the occupation number of electrons nd,σ\langle n_{d,\sigma}\rangle depends on the coherence factor cosΘ\cos\Theta, and is given by a linear combination of the phase shifts as shown in Eq. (23). Therefore, the occupation number of \downarrow-spin electrons fluctuates significantly at the crossover region in the direction of Θ=π\Theta=\pi in such a way that nd,Θ=π1δ/π\langle n_{d,\downarrow}\rangle\xrightarrow{\,\Theta=\pi\,}1-\delta_{\uparrow}/\pi, whereas the \uparrow-spin electrons fluctuate in the direction of Θ=0\Theta=0 as nd,Θ=0δ/π\langle n_{d,\uparrow}\rangle\xrightarrow{\,\Theta=0\,}\delta_{\uparrow}/\pi.

Figures 12(b) and 12(c) clearly show that 𝒯spin\mathcal{T}_{\mathrm{spin}} and PRP_{R} are enhanced at the level-crossing point EAb+U/2E_{A}\simeq b+U/2 near the ξd\xi_{d} axis. The Θ\Theta dependencies of 𝒯spin\mathcal{T}_{\mathrm{spin}} and PRP_{R} are determined by the coherence factor cosΘ\cos\Theta, as shown in Eqs. (44) and (45). Therefore, these coefficients become most significant in the directions of Θ=0\Theta=0 and π\pi, where the resonant tunneling occurs for the \uparrow-spin and \downarrow-spin electron components, respectively. As the Bogoliubov angle Θ\Theta deviates away from the ξd\xi_{d} axis, the spin polarization is suppressed, especially in the SC-proximity-dominated regime at π/4<Θ<3π/4\pi/4<\Theta<3\pi/4, and the spin current IR,spinI_{R,\mathrm{spin}} vanishes at Θ=π/2\Theta=\pi/2.

Refer to caption
Refer to caption
Refer to caption
Figure 12: Spin-dependent transport coefficients at finite magnetic field b=0.1Ub=0.1U, for U/(πΓN)=5.0U/(\pi\Gamma_{N})=5.0. Top panel (a) shows sin2δ\sin^{2}\delta_{\uparrow} and sin2δ\sin^{2}\delta_{\downarrow}, plotted vs EAE_{A}. Three-dimensional figures represent (b) spin transmission coefficient 𝒯spin\mathcal{T}_{\mathrm{spin}} and (c) current polarization PRP_{R}, plotted as functions of ξd\xi_{d} and ΓS\Gamma_{S}, for ΓL=ΓR\Gamma_{L}=\Gamma_{R}.

VI Summary

We have studied the interplay between the crossed Andreev reflection, Kondo effect, and Zeeman splitting, occurring in a multi-terminal quantum dot, consisting of two normal and one SC leads.

It has been shown that the linear-response currents flowing through quantum dot at zero temperature T=0T=0 are determined by two angular variables, i.e., the phase shift δσ\delta_{\sigma} of Bogoliubov particles and the Bogoliubov rotation angle Θ=cot1(ξd/ΓS)\Theta=\cot^{-1}(\xi_{d}/\Gamma_{S}) in the the limit of large SC gap |ΔS||\Delta_{S}|\to\infty. In this limit, the phase shift can be deduced from an effective Anderson model for interacting Bogoliubov particles, which has a global U(1) symmetry along the principal axis in the Nambu pseudo-spin space. The Bogoliubov angle Θ\Theta enters the transport coefficients through the SC coherence factors, and plays an essential role in the conductance, together with the position of the Andreev level EA=ξd2+ΓS2E_{A}=\sqrt{\xi_{d}^{2}+\Gamma_{S}^{2}}.

In the first half of the paper, we have described the role of the many-body optical theorem on the CAR, and have shown that the multi-terminal conductance at finite magnetic fields is determined by the transmission probability 𝒯BG=12σsin2δσ\mathcal{T}_{\mathrm{BG}}=\frac{1}{2}\sum_{\sigma}\sin^{2}\delta_{\sigma} of the Bogoliubov particles, which does not depend on Θ\Theta, and by the Cooper-pair tunneling part 𝒯CP=14sin2(δ+δ)sin2Θ\mathcal{T}_{\mathrm{CP}}=\frac{1}{4}\sin^{2}(\delta_{\uparrow}+\delta_{\downarrow})\sin^{2}\Theta. In the second half, we have discussed the behaviors of nonlocal conductance, obtained by using the NRG approach in a wide range of the parameter space which consists of ξd\xi_{d}, ΓS\Gamma_{S}, ΓN\Gamma_{N}, the Coulomb interaction UU, and the magnetic field bb.

At zero magnetic field, the nonlocal conductance gRLg_{\mathrm{RL}} becomes negative at EAU/2E_{A}\gtrsim U/2 and π/4<Θ<3π/4\pi/4<\Theta<3\pi/4, where the CAR dominates. In particular, the contribution of Cooper-pair tunnelings 𝒯CP\mathcal{T}_{\mathrm{CP}} is maximized at a crescent-shaped crossover region between the Kondo-dominated and the SC-dominated regimes, emerging at EAU/2E_{A}\simeq U/2 in the angular direction of Θπ/2\Theta\simeq\pi/2. The width of the crescent region along the ΓS\Gamma_{S}-axis is of the order of ΓN\Gamma_{N}. The enhanced CAR occurring in this region is caused by the valence fluctuation of the Bogoliubov particles, in the middle of which the occupation number takes the value Q=1/2Q=1/2 and the phase shift due to the Cooper-pair tunneling reaches the unitary limit δ+δ=π/2\delta_{\uparrow}+\delta_{\downarrow}=\pi/2.

Magnetic fields lift the spin degeneracy of the Andreev resonance level. In the strongly-correlated case where Q1.0Q\simeq 1.0 with EAU/2E_{A}\lesssim U/2 and UΓNU\gg\Gamma_{N}, the crossover between the Kondo regime and Zeeman-dominated regime occurs at a magnetic field bTb\sim T^{*} of the order of the Kondo energy scale TT^{*}. In contrast, at EAU/2E_{A}\gtrsim U/2 in the valence-fluctuation region of the Bogoliubov particles, magnetic fields induce a crossover between the SC-proximity dominated regime and the Zeeman-dominated regime at bEAU/2b\simeq E_{A}-U/2, where the renormalized Andreev level E~A,\widetilde{E}_{A,\uparrow} for the majority spin component (σ=\sigma=\uparrow) crosses the Fermi level. It induces the resonant tunneling of the Bogoliubov particles and the Cooper-pair tunneling, the transmission probabilities of which are determined by the phase shifts δπ/2\delta_{\uparrow}\simeq\pi/2 and δ+δπ/2\delta_{\uparrow}+\delta_{\downarrow}\simeq\pi/2, respectively. Note that δ0.0\delta_{\downarrow}\simeq 0.0 as the renormalized Andreev level for the minority-spin component becomes almost empty Q0.0Q_{\downarrow}\simeq 0.0 in this region. It has also been demonstrated that the resonant spin current is enhanced in the angular direction of Θ0\Theta\simeq 0 or Θπ\Theta\simeq\pi when the Andreev level of the majority spin crosses the Fermi level.

The nonlocal conductance gRLg_{\mathrm{RL}} becomes negative in the parameter region of EAU/2+bE_{A}\gtrsim U/2+b and π/4<Θ<3π/4\pi/4<\Theta<3\pi/4. In particular, the CAR contribution is maximized in the crescent-shaped region, which moves in the ξd\xi_{d} vs ΓS\Gamma_{S} plane, together with the semi-circular boundary of radius EAU/2+bE_{A}\simeq U/2+b as bb increases. The crescent region evolves with the magnetic field and yields a flat valley structure which emerges in the bb dependence of gRLg_{\mathrm{RL}}, at 0bEAU/20\leq b\lesssim E_{A}-U/2. These results suggest that, in order to experimentally probe the CAR contributions measuring the nonlocal conductance, the crescent parameter region will be a plausible target to be examined.

Acknowledgements.
This work was supported by JSPS KAKENHI Grants No. JP18K03495 and No. JP23K03284, and JST Moonshot R & D-MILLENNIA Program Grant No. JPMJMS2061. Y. Teratani was supported by the Sasakawa Scientific Research Grant from the Japan Science Society Grant No. 2021-2009.

Appendix A Effective Hamiltonian for |ΔS||\Delta_{S}|\to\infty

The Hamiltonian HH defined in Eq. (1) can be separated into two independent parts since only the symmetrized linear combination αε,σ\alpha_{\varepsilon,\sigma} of conduction electrons has a finite tunnel coupling to the QD, whereas the anti-symmetrized linear combination βε,σ\beta_{\varepsilon,\sigma} is decoupled from the rest of the system:

αε,σ\displaystyle\alpha_{\varepsilon,\sigma}\,\equiv vLcε,L,σ+vRcε,R,σvL2+vR2,\displaystyle\ \frac{v_{L}\,c_{\varepsilon,L,\sigma}+v_{R}\,c_{\varepsilon,R,\sigma}}{\sqrt{v_{L}^{2}+v_{R}^{2}}}\,, (46)
βε,σ\displaystyle\beta_{\varepsilon,\sigma}\,\equiv vRcε,L,σ+vLcε,R,σvL2+vR2.\displaystyle\ \frac{-v_{R}\,c_{\varepsilon,L,\sigma}+v_{L}\,c_{\varepsilon,R,\sigma}}{\sqrt{v_{L}^{2}+v_{R}^{2}}}\,. (47)

Correspondingly, the conduction-electron part and the normal-tunneling part of the Hamiltonian can be rewritten in the form

HN=\displaystyle H_{\text{N}}\,= σDD𝑑εε(αε,σαε,σ+βε,σβε,σ),\displaystyle\ \sum_{\sigma}\int_{-D}^{D}\!d\varepsilon\,\varepsilon\,\Bigl{(}\alpha^{\dagger}_{\varepsilon,\sigma}\alpha_{\varepsilon,\sigma}+\beta^{\dagger}_{\varepsilon,\sigma}\beta_{\varepsilon,\sigma}\Bigr{)}, (48)
HTN=\displaystyle H_{\text{TN}}= vNσDDdερc(αε,σdσ+H.c.),\displaystyle\ v_{N}\sum_{\sigma}\int_{-D}^{D}\!d\varepsilon\,\sqrt{\rho_{c}}\,\Bigl{(}\alpha^{\dagger}_{\varepsilon,\sigma}d_{\sigma}+\mathrm{H.c.}\Bigr{)}\,, (49)

where vNvL2+vR2v_{N}\equiv\sqrt{v_{L}^{2}+v_{R}^{2}}.

Furthermore, in the large gap limit |ΔS|\left|\Delta_{S}\right|\to\infty which is taken at |ΔS|DS\left|\Delta_{S}\right|\ll D_{S} keeping ρS\rho_{S} constant, the superconducting proximity effects can be described by the pair potential ΔdΓSeiϕS\Delta_{d}\equiv\Gamma_{S}\,e^{i\phi_{S}} penetrating into the QD. Tanaka et al. (2007b, a) Therefore, at low energies, the subspace to which the QD belongs can be described by the following effective Hamiltonian:

Heff\displaystyle H_{\mathrm{eff}}\,\equiv 𝝍d𝓗dotSC𝝍d+U2(nd1)2\displaystyle\ \ \bm{\psi}_{d}^{\dagger}\ \bm{\mathcal{H}}_{\mathrm{dot}}^{\mathrm{SC}}\ \bm{\psi}_{d}\,+\,\frac{U}{2}\left(n_{d}-1\right)^{2}
+vNDD𝑑ερc[𝝍α(ε)𝝍d+𝝍d𝝍α(ε)]\displaystyle+v_{N}\int_{-D}^{D}\!\!d\varepsilon\,\sqrt{\rho_{c}}\left[\,\bm{\psi}_{\alpha}^{\dagger}(\varepsilon)\,\bm{\psi}_{d}+\bm{\psi}_{d}^{\dagger}\,\bm{\psi}_{\alpha}(\varepsilon)\,\right]
+DD𝑑εε𝝍α(ε)𝝍α(ε).\displaystyle+\int_{-D}^{D}\!\!d\varepsilon\,\varepsilon\ \bm{\psi}_{\alpha}^{\dagger}(\varepsilon)\,\bm{\psi}_{\alpha}(\varepsilon)\,. (50)

Here, 𝓗dotSC\bm{\mathcal{H}}_{\mathrm{dot}}^{\mathrm{SC}} is the following matrix defined in the Nambu pseudo-spin space,

𝓗dotSC\displaystyle\bm{\mathcal{H}}_{\mathrm{dot}}^{\mathrm{SC}}\,\equiv (ξdΔdΔdξd)b 1,\displaystyle\ \begin{pmatrix}\xi_{d}&\Delta_{d}\cr\Delta_{d}^{*}&-\xi_{d}\rule{0.0pt}{9.95863pt}\end{pmatrix}\,-\,b\,\bm{1}\,, (51)

with 𝟏\bm{1} the 2×22\times 2 unit matrix, and

𝝍d\displaystyle\bm{\psi}_{d}\equiv (dd),𝝍α(ε)(αε,αε,).\displaystyle\ \begin{pmatrix}d_{\uparrow}\cr d_{\downarrow}^{\dagger}\rule{0.0pt}{14.22636pt}\cr\end{pmatrix},\qquad\bm{\psi}_{\alpha}(\varepsilon)\equiv\begin{pmatrix}\alpha_{\varepsilon,\uparrow}\cr-\alpha_{-\varepsilon,\downarrow}^{\dagger}\rule{0.0pt}{14.22636pt}\cr\end{pmatrix}.\rule{0.0pt}{22.76228pt} (52)

The effective Hamiltonian HeffH_{\mathrm{eff}} has a global U(1)(1) symmetry with respect to the principal axis along the three-dimensional vector 𝒏^(ReΔd,ImΔd,ξd)\widehat{\bm{n}}\propto(\mathrm{Re}\,\Delta_{d},\,-\mathrm{Im}\,\Delta_{d},\,\xi_{d}) in the Nambu space. The conserved charge associated with this U(1)(1) symmetry corresponds to the total number of Bogoliubov particles, the operators for which are given by

(γd,γd,)=\displaystyle\begin{pmatrix}\gamma_{d,\uparrow}^{\phantom{\dagger}}\cr\gamma_{d,\downarrow}^{\dagger}\rule{0.0pt}{17.07182pt}\cr\end{pmatrix}= 𝓤𝝍d,(γε,γε,)=𝓤𝝍α(ε).\displaystyle\ \bm{\mathcal{U}}^{\dagger}\bm{\psi}_{d},\quad\begin{pmatrix}\gamma_{\varepsilon,\uparrow}^{\phantom{\dagger}}\cr-\gamma_{-\varepsilon,\downarrow}^{\dagger}\rule{0.0pt}{17.07182pt}\cr\end{pmatrix}=\,\bm{\mathcal{U}}^{\dagger}\bm{\psi}_{\alpha}(\varepsilon)\,. (53)

Here, 𝓤\bm{\mathcal{U}} is the unitary matrix which diagonalizes 𝓗dotSC\bm{\mathcal{H}}_{\mathrm{dot}}^{\mathrm{SC}}:

𝓤𝓗dotSC𝓤=EA𝝉3b𝟏,𝝉3=(1 001),\displaystyle\!\!\bm{\mathcal{U}}^{\dagger}\,\bm{\mathcal{H}}_{\mathrm{dot}}^{\mathrm{SC}}\ \bm{\mathcal{U}}\,=\,E_{A}\bm{\tau}_{3}-b\bm{1},\qquad\bm{\tau}_{3}=\begin{pmatrix}1&\ 0\cr 0&-1\rule{0.0pt}{5.69046pt}\end{pmatrix}, (54)

with EAξd2+ΓS2E_{A}\equiv\sqrt{\xi_{d}^{2}+\Gamma_{S}^{2}}. For example, in the case where the Josephson phase ϕS=0\phi_{S}=0, the matrix 𝓤\bm{\mathcal{U}} is determined by a single Bogoliubov angle Θ\Theta, as shown in Eq. (13).

Appendix B Derivation of linear nonlocal current

In this appendix, we provide a brief derivation of the nonlocal conductance defined in Eqs. (24)–(27)

The current flowing from the quantum dot to the normal lead on the right is described by the operator,

I^R,σ=ievRDD𝑑ερc(cε,R,σdσdσcε,R,σ)\widehat{I}_{R,\sigma}\,=\,-i\,ev_{R}\int_{-D}^{D}\!\!d\varepsilon\,\sqrt{\rho_{c}}\,\Bigl{(}c^{\dagger}_{\varepsilon,R,\sigma}d_{\sigma}-d^{\dagger}_{\sigma}c_{\varepsilon,R,\sigma}\Bigr{)} (55)

for spin σ\sigma component. The steady-state average of the total current IRI^R,+I^R,I_{R}\equiv\langle\widehat{I}_{R,\uparrow}\rangle+\langle\widehat{I}_{R,\downarrow}\rangle with IR,σI^R,σI_{R,\sigma}\equiv\langle\widehat{I}_{R,\sigma}\rangle can be expressed in terms of the Green function in the Keldysh formalism, Tanaka et al. (2007a)

IR=\displaystyle I_{R}\,= ieh𝑑ωΓRTr[𝑮ddr(ω)𝓒R(ω)𝑮dda(ω)𝝉3],\displaystyle\ -i\,\frac{e}{h}\int_{-\infty}^{\infty}\!d\omega\,\Gamma_{R}\,\mathrm{Tr}\,\Bigl{[}\bm{G}_{dd}^{r}(\omega)\,\bm{\mathcal{C}}_{R}(\omega)\,\bm{G}_{dd}^{a}(\omega)\,\bm{\tau}_{3}\Bigr{]}\,,
𝓒R\displaystyle\bm{\mathcal{C}}_{R}\,\equiv 𝚺totK(12𝒇R)[𝚺tot+𝚺tot+].\displaystyle\ \,\bm{\Sigma}_{\mathrm{tot}}^{\mathrm{K}}-\bigl{(}1-2\bm{f}_{R}\bigr{)}\Bigr{[}\,\bm{\Sigma}_{\mathrm{tot}}^{-+}-\bm{\Sigma}_{\mathrm{tot}}^{+-}\,\Bigr{]}\,. (56)

Here, Tr\mathrm{Tr} denotes the trace of the 2×22\times 2 matrices in the Nambu pseudo-spin space. 𝚺tot+\bm{\Sigma}^{-+}_{\mathrm{tot}} and 𝚺tot+\bm{\Sigma}^{+-}_{\mathrm{tot}} are the lesser and greater self-energies, respectively, and 𝚺totK𝚺tot+𝚺tot+\bm{\Sigma}^{\mathrm{K}}_{\mathrm{tot}}\equiv-\bm{\Sigma}^{-+}_{\mathrm{tot}}-\bm{\Sigma}^{+-}_{\mathrm{tot}}. The matrix 𝒇ν\bm{f}_{\nu} is defined as

𝒇ν(ω)=[fν(ω) 00f¯ν(ω)],ν=L,R.\displaystyle\bm{f}_{\nu}(\omega)=\left[\,\begin{matrix}f_{\nu}(\omega)&\ 0\cr 0\rule{0.0pt}{14.22636pt}&\ \overline{f}_{\nu}(\omega)\cr\end{matrix}\,\right],\qquad\quad\nu=L,\,R. (57)

The bias voltage eVνeV_{\nu} is applied to the leads such that fν(ω)f(ωeVν)f_{\nu}(\omega)\equiv f(\omega-eV_{\nu}) and f¯ν(ω)f(ω+eVν)\overline{f}_{\nu}(\omega)\equiv f(\omega+eV_{\nu}) with f(ω)=1/[eω/T+1]f(\omega)=1/[e^{\omega/T}+1] the Fermi distribution function.

Each self-energy matrix can be separated into two parts, e.g.,

𝚺totK(ω)=𝚺0K(ω)+𝚺UK(ω).\bm{\Sigma}^{\mathrm{K}}_{\mathrm{tot}}(\omega)\,=\,\bm{\Sigma}^{\mathrm{K}}_{0}(\omega)+\bm{\Sigma}^{\mathrm{K}}_{U}(\omega)\;. (58)

Here, the first term on the right-hand side represents the tunnel contributions at U=0U=0,

𝚺0K(ω)=2iν=L,RΓν[ 12𝒇j(ω)],\displaystyle\!\!\!\!\bm{\Sigma}^{\mathrm{K}}_{0}(\omega)=-2i\sum_{\nu=L,R}\Gamma_{\nu}\Bigl{[}\,\bm{1}-2\bm{f}_{j}(\omega)\,\Bigr{]}, (59)
𝚺0+(ω)𝚺0+(ω)=2i(ΓL+ΓR) 1,\displaystyle\!\!\!\!\bm{\Sigma}_{0}^{-+}(\omega)-\bm{\Sigma}_{0}^{+-}(\omega)\,=\,-2i\bigl{(}\Gamma_{L}+\Gamma_{R}\bigr{)}\,\bm{1}\,, (60)

with 𝟏\bm{1} the 2×22\times 2 unit matrix in the pseudo-spin space. The second term on the right-hand side of Eq. (58), 𝚺UK(ω)\bm{\Sigma}^{\mathrm{K}}_{U}(\omega), represents the self-energy corrections due to the Coulomb interaction UU. This and the corresponding terms of the lesser and greater self-energies, 𝚺U+(ω)\bm{\Sigma}^{-+}_{U}(\omega) and 𝚺U+(ω)\bm{\Sigma}^{+-}_{U}(\omega) are also pure imaginary in the frequency domain, and represent the damping of quasiparticles due to the multiple collisions. These imaginary parts of the interacting self-energies vanish at T=0T=0, eVν=0eV_{\nu}=0, and ω=0\omega=0, and thus they do not contribute to the linear-response current at zero temperature. Furthermore, the function 𝓒R(ω)\bm{\mathcal{C}}_{R}(\omega) identically vanishes at eVν=0eV_{\nu}=0 since there is no steady current at equilibrium.

Therefore, at T=0T=0, the linear-response current can be calculated, keeping the noninteracting terms in Eqs. (59) and (60) for 𝓒R(ω)\bm{\mathcal{C}}_{R}(\omega) in the right-hand side of Eq.  (56):

IR=4e2h\displaystyle I_{R}=\,\frac{4e^{2}}{h} [(|{𝑮ddr(0)}11|2+|{𝑮ddr(0)}22|2)\displaystyle\Biggl{[}\,\left(\Bigl{|}\bigl{\{}\bm{G}_{dd}^{r}(0)\bigl{\}}_{11}\Bigr{|}^{2}+\Bigl{|}\bigl{\{}\bm{G}_{dd}^{r}(0)\bigl{\}}_{22}\Bigr{|}^{2}\right)
×ΓRΓL(VLVR)\displaystyle\qquad\times\Gamma_{R}\Gamma_{L}\,(V_{L}-V_{R})
(|{𝑮ddr(0)}12|2+|{𝑮ddr(0)}21|2)\displaystyle-\left(\Bigl{|}\bigl{\{}\bm{G}_{dd}^{r}(0)\bigl{\}}_{12}\Bigr{|}^{2}+\Bigl{|}\bigl{\{}\bm{G}_{dd}^{r}(0)\bigl{\}}_{21}\Bigr{|}^{2}\right)
×{ΓRΓL(VL+VR)+2ΓR2VR}].\displaystyle\qquad\times\Bigl{\{}\Gamma_{R}\Gamma_{L}\,(V_{L}+V_{R})+2\Gamma_{R}^{2}\,V_{R}\Bigr{\}}\,\,\Biggr{]}. (61)

Note that the anomalous Green’s functions are related to each other through {𝑮ddr(ω)}21={𝑮dda(ω)}12\bigl{\{}\bm{G}_{dd}^{r}(\omega)\bigl{\}}_{21}=\bigl{\{}\bm{G}_{dd}^{a}(\omega)\bigl{\}}_{12}^{*}. Equation (61) can be rewritten further in terms of the phase shifts δσ\delta_{\sigma} and the Bogoliubov angle Θ\Theta, by using Eq. (12) to obtain Eqs. (26) and (27).

Appendix C Optical theorem for Andreev scattering

We provide a derivation of the optical theorem, which emerges in the form

σsin2δσ= 2(𝒯ET+𝒯CP).\displaystyle\sum_{\sigma}\sin^{2}\delta_{\sigma}\,=\,2\bigl{(}\mathcal{T}_{\mathrm{ET}}+\mathcal{T}_{\mathrm{CP}}\bigr{)}\,. (62)

We start with the matrix identity for the impurity Green’s function in the Nambu form

𝑮ddr(ω)𝑮dda(ω)\displaystyle\bm{G}_{dd}^{r}(\omega)-\bm{G}_{dd}^{a}(\omega)
=𝑮ddr(ω)[{𝑮dda(ω)}1{𝑮ddr(ω)}1]𝑮dda(ω)\displaystyle\quad=\bm{G}_{dd}^{r}(\omega)\Bigl{[}\left\{\bm{G}_{dd}^{a}(\omega)\right\}^{-1}-\left\{\bm{G}_{dd}^{r}(\omega)\right\}^{-1}\Bigr{]}\bm{G}_{dd}^{a}(\omega)
=𝑮ddr(ω)[𝚺totr(ω)𝚺tota(ω)]𝑮dda(ω).\displaystyle\quad=\bm{G}_{dd}^{r}(\omega)\Bigl{[}\bm{\Sigma}_{\mathrm{tot}}^{r}(\omega)-\bm{\Sigma}_{\mathrm{tot}}^{a}(\omega)\Bigr{]}\bm{G}_{dd}^{a}(\omega)\,. (63)

At ω=T=eV=0\omega=T=eV=0, it can be rewritten further in the form

ΓN2i[𝑮ddr(0)𝑮dda(0)]=\displaystyle-\frac{\Gamma_{N}}{2i}\Bigl{[}\,\bm{G}_{dd}^{r}(0)-\bm{G}_{dd}^{a}(0)\,\Bigr{]}\,= ΓN2𝑮ddr(0)𝑮dda(0).\displaystyle\ \Gamma_{N}^{2}\,\bm{G}_{dd}^{r}(0)\,\bm{G}_{dd}^{a}(0)\,. (64)

Here, we have used the property that the imaginary part of the interacting self-energy vanishes Im𝚺Ur(0)=0\mathrm{Im}\,\bm{\Sigma}_{U}^{r}(0)=0 and the noninteracting one is given by 𝚺0r(ω)𝚺0a(ω)=2iΓN𝟏\bm{\Sigma}_{0}^{r}(\omega)-\bm{\Sigma}_{0}^{a}(\omega)=-2i\Gamma_{N}\bm{1}. Taking trace of the Nambu matrices, the left-hand side of Eq. (64) can be calculated as

ΓN2iTr[𝑮ddr(0)𝑮dda(0)]\displaystyle\!\!\!\!-\frac{\Gamma_{N}}{2i}\mathrm{Tr}\,\Bigl{[}\,\bm{G}_{dd}^{r}(0)-\bm{G}_{dd}^{a}(0)\,\Bigr{]}
=ΓN2iσ[Gγ,σr(0)Gγ,σa(0)]=σsin2δσ.\displaystyle=-\frac{\Gamma_{N}}{2i}\sum_{\sigma}\Bigl{[}G_{\gamma,\sigma}^{r}(0)-G_{\gamma,\sigma}^{a}(0)\Bigr{]}=\sum_{\sigma}\sin^{2}\delta_{\sigma}. (65)

Similarly, the right-hand side of Eq. (64) takes the form

ΓN2Tr[𝑮ddr(0)𝑮ddr(0)]\displaystyle\Gamma_{N}^{2}\mathrm{Tr}\,\Bigl{[}\,\bm{G}_{dd}^{r}(0)\,\bm{G}_{dd}^{r}(0)\,\Bigr{]}
=ΓN2[|{𝑮ddr(0)}11|2+|{𝑮ddr(0)}12|2\displaystyle=\ \Gamma_{N}^{2}\biggl{[}\Bigl{|}\left\{\bm{G}_{dd}^{r}(0)\right\}_{11}\Bigr{|}^{2}+\Bigl{|}\left\{\bm{G}_{dd}^{r}(0)\right\}_{12}\Bigr{|}^{2}
+|{𝑮ddr(0)}21|2+|{𝑮ddr(0)}22|2].\displaystyle\qquad\quad+\Bigl{|}\left\{\bm{G}_{dd}^{r}(0)\right\}_{21}\Bigr{|}^{2}+\Bigl{|}\left\{\bm{G}_{dd}^{r}(0)\right\}_{22}\Bigr{|}^{2}\,\biggr{]}. (66)

The last line corresponds to 2(𝒯ET+𝒯CP)2\bigl{(}\mathcal{T}_{\mathrm{ET}}+\mathcal{T}_{\mathrm{CP}}\bigr{)} defined in Eqs. (26) and (27), and from this Eq. (62) follows.

Appendix D Derivation of the spin-current formula

We briefly describe here the linear-response formula for the spin current following between two normal leads at finite magnetic fields. The current formula presented in Appendix B can be decomposed into the contributions of the \uparrow- and \downarrow- spin components, which can be rearranged as a spin current:

IR,spinIR,IR,\displaystyle I_{R,\mathrm{spin}}\equiv I_{R,\uparrow}-I_{R,\downarrow}
=ieh𝑑ωΓRTr[𝑮ddr(ω)𝓒R(ω)𝑮dda(ω)].\displaystyle=\ -i\,\frac{e}{h}\int_{-\infty}^{\infty}\!d\omega\,\Gamma_{R}\,\mathrm{Tr}\Bigl{[}\bm{G}_{dd}^{r}(\omega)\,\bm{\mathcal{C}}_{R}(\omega)\,\bm{G}_{dd}^{a}(\omega)\,\Bigr{]}\,. (67)

Specifically at T=0T=0, the linear-response spin current can be expressed in the following form,

IR,spin=\displaystyle I_{R,\mathrm{spin}}\,= 4e2hΓLΓR(|{𝑮ddr(0)}11|2|{𝑮ddr(0)}22|2)\displaystyle\ \frac{4e^{2}}{h}\,\Gamma_{L}\Gamma_{R}\left(\Bigl{|}\bigl{\{}\bm{G}_{dd}^{r}(0)\bigl{\}}_{11}\Bigr{|}^{2}-\Bigl{|}\bigl{\{}\bm{G}_{dd}^{r}(0)\bigl{\}}_{22}\Bigr{|}^{2}\right)
×(VLVR)\displaystyle\times\left(V_{L}-V_{R}\right)
=\displaystyle= 4e2hΓLΓRΓN2𝒯spin(VLVR),\displaystyle\ \frac{4e^{2}}{h}\,\frac{\Gamma_{L}\Gamma_{R}}{\Gamma_{N}^{2}}\,\mathcal{T}_{\mathrm{spin}}\,\left(V_{L}-V_{R}\right)\,, (68)
𝒯spin\displaystyle\mathcal{T}_{\mathrm{spin}}\,\equiv (sin2δsin2δ)cosΘ.\displaystyle\ \left(\sin^{2}\delta_{\uparrow}-\sin^{2}\delta_{\downarrow}\right)\cos\Theta\,. (69)

Note that IL,IL,=IR,IR,I_{L,\uparrow}-I_{L,\downarrow}=I_{R,\uparrow}-I_{R,\downarrow}.

Similarly, the current polarization PRP_{R}, defined with respect to symmetric voltages VL=VRV_{L}=-V_{R}, can be used as a measure of the spin current relative to the charge current López and Sánchez (2003); Souza et al. (2008); Hitachi et al. (2006); Kiyama et al. (2015):

PRIR,IR,IR,+IR,=ΓLΓR𝒯spin2[ΓLΓR𝒯ET+ΓR2𝒯CP]\displaystyle P_{R}\,\equiv\,\frac{I_{R,\uparrow}-I_{R,\downarrow}}{I_{R,\uparrow}+I_{R,\downarrow}}\ =\ \frac{\Gamma_{L}\Gamma_{R}\mathcal{T}_{\mathrm{spin}}}{2\left[\Gamma_{L}\Gamma_{R}\mathcal{T}_{\mathrm{ET}}+\Gamma_{R}^{2}\mathcal{T}_{\mathrm{CP}}\right]}
ΓL=ΓRsin2δsin2δsin2δ+sin2δcosΘ.\displaystyle\ \ \xrightarrow{\,\Gamma_{L}=\Gamma_{R}\,}\ \frac{\sin^{2}\delta_{\uparrow}-\sin^{2}\delta_{\downarrow}}{\sin^{2}\delta_{\uparrow}+\sin^{2}\delta_{\downarrow}}\,\cos\Theta\,. (70)

References