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

Planckian superconductor

Y. Cheipesh Instituut-Lorentz, Universiteit Leiden, P.O. Box 9506, NL-2300 RA Leiden, The Netherlands    A. I. Pavlov The Abdus Salam International Centre for Theoretical Physics (ICTP) Strada Costiera 11, I-34151 Trieste, Italy Institute for Theoretical Solid State Physics, Leibniz-Institut für Festkörper und Werkstoffforschung IFW-Dresden, Helmholtzstrasse 20, D-01169 Dresden, Germany    V. Scopelliti Instituut-Lorentz, Universiteit Leiden, P.O. Box 9506, NL-2300 RA Leiden, The Netherlands    J. Tworzydło Institute of Theoretical Physics, Faculty of Physics, University of Warsaw, ul. Pasteura 5, PL-02-093 Warszawa, Poland    N. V. Gnezdilov nikolay.gnezdilov@yale.edu Department of Physics, Yale University, New Haven, CT 06520, USA
(September 19, 2025)
Abstract

The Planckian relaxation rate /tP=2πkBT\hbar/t_{\mathrm{P}}=2\pi k_{\mathrm{B}}T sets a characteristic timescale for both the equilibration of quantum critical systems and maximal quantum chaos. In this note, we show that at the critical coupling between a superconducting dot and the complex Sachdev-Ye-Kitaev model, known to be maximally chaotic, the pairing gap Δ\Delta behaves as η/tP\eta\,\hbar/t_{\mathrm{P}} at low temperatures, where η\eta is an order one constant. The lower critical temperature emerges with a further increase of the coupling strength so that the finite Δ\Delta domain is settled between the two critical temperatures.

The Bardeen-Cooper-Schrieffer mechanism of conventional superconductivity Bardeen et al. (1957) requires two species of fermions coupled by an attractive two-body interaction. Altland and Simons (2010) The mean-field analysis of such a model results in the gapped quasiparticle excitation spectrum below the critical temperature. Meanwhile, the absence of long-living quasiparticles in high-temperature superconducting materials above the critical temperature is an immutable characteristic of the so-called strange metal state. Senthil (2008); Keimer et al. (2015) In contrast to the quasiparticle nature of superconductors, strange metals exhibit a power-law behavior in the spectral function, Varma et al. (1989) similarly to quantum critical systems. Sachdev (2011) A lack of quasiparticles manifests itself in fast equilibration at low temperature on a timescale set by the Planckian relaxation time tP=/(2πkBT)t_{\mathrm{P}}=\hbar/\left(2\pi k_{\mathrm{B}}T\right). Zaanen (2004); Sachdev (2011) The same timescale appears as an upper bound on quantum chaos setting the maximal rate of information scrambling. Maldacena et al. (2016) It is usually formulated Maldacena et al. (2016); Shenker and Stanford (2015); Roberts and Swingle (2016) in terms of the out-of-time ordered correlator Larkin and Ovchinnikov (1969) (OTOC): In quantum many-body systems the OTOC grows no faster than exponentially et/tLe^{t/t_{\mathrm{L}}} with the Lyapunov time tLt_{\mathrm{L}} bounded from below as tLtPt_{\mathrm{L}}\geq t_{\mathrm{P}}. Maldacena et al. (2016)

The widely known Sachdev-Ye-Kitaev (SYK) model, Sachdev and Ye (1993); Kitaev (2015) describing strongly interacting Majorana zero modes in 0+10+1 dimensions, saturates the chaos bound tL=tPt_{\mathrm{L}}=t_{\mathrm{P}}. Kitaev (2015); Maldacena and Stanford (2016) It does not possess an underlying quasiparticle description while being solvable in the infrared, with a spectral function that scales as a power law of frequency. These properties do not change upon replacing Majoranas with conventional fermions (complex SYK model). Sachdev (2015); Bulycheva (2017) The extensions of this model to the cSYK coupled clusters predict thermal diffusivity Davison et al. (2017) tP\propto t_{\mathrm{P}} and reproduce the linear in temperature resistivity, Song et al. (2017) observed in strange metals. Takagi et al. (1992); Taillefer (2010) Recently, a proposed theory of a Planckian metal, Patel and Sachdev (2019) based on the destruction of a Fermi surface by the cSYK-like interactions, shows that the universal scattering time equals the Planckian time tPt_{\mathrm{P}}. The latter one characterizes the linear in temperature resistivity property Bruin et al. (2013) and was detected in cuprates, Legros et al. (2019) pnictides, Nakajima et al. (2019) and twisted bilayer graphene, Cao et al. (2019) regardless of their different microscopic nature.

The success in applying the SYK model to qualitative studies of strange metals and the minimalistic structure of the model itself fostered the effort to find a mechanism by which the superconducting state is formed out of an incoherent SYK metal. Patel et al. (2018); Esterlis and Schmalian (2019); Wang (2019); Chowdhury and Berg (2019) Driven by the same curiosity, we consider a (0+1)(0+1)-dimensional toy model which consists of a superconducting quantum dot Kouwenhoven and Marcus (1998) coupled to the complex-valued SYK model. Sachdev (2015) At the critical coupling the pairing gap turns out to be proportional to the Planckian relaxation rate at low temperatures,

ΔηtP,\Delta\approx\eta\,\frac{\hbar}{t_{\mathrm{P}}}, (1)

where η\eta is a number close to one. This theoretical finding that we refer to as a Planckian superconductor draws parallels to the phenomenon of reentrant superconductivity Maple et al. (1972); Simons et al. (2012) in Kondo superconductors Müller-Hartmann and Zittartz (1971); Riblet and Winzer (1971); Müller-Hartmann et al. (1976) and the physics of Andreev billiards. Melsen et al. (1997); Schomerus and Beenakker (1999); Lodder and Nazarov (1998); Adagideli and Beenakker (2002); Vavilov and Larkin (2003)

We start with a superconducting Hamiltonian HSCH_{\rm SC} that contains 2M2M modes described by the Richardson Hamiltonian Richardson (1963); von Delft et al. (1996); Matveev and Larkin (1997) without single-particle energies coupled to the SYK model HSYKH_{\rm SYK} with NN fermions through a random tunneling term HtunH_{\rm tun},

H=\displaystyle H\!= HSC+HSYK+Htun,\displaystyle{}H_{\rm SC}+H_{\rm SYK}+H_{\rm tun}\,, (2)
HSC=\displaystyle H_{\rm SC}\!= UMi,j=1Mψiψiψjψjμi=1Mσ=,ψσiψσi,\displaystyle{}-\frac{U}{M}\!\sum_{i,j=1}^{M}\!\!\psi^{\dagger}_{\uparrow i}\psi^{\dagger}_{\downarrow i}\psi_{\downarrow j}\psi_{\uparrow j}-\mu\!\sum_{i=1}^{M}\!\sum_{\sigma=\uparrow,\downarrow}\!\!\psi^{\dagger}_{\sigma i}\psi_{\sigma i}, (3)
HSYK=\displaystyle H_{\rm SYK}\!= 1(2N)3/2i,j,k,l=1NJij;klcicjckcl,\displaystyle{}\frac{1}{(2N)^{3/2}}\!\sum_{i,j,k,l=1}^{N}\!\!J_{ij;kl}c^{\dagger}_{i}c^{\dagger}_{j}c_{k}c_{l}, (4)
Htun=\displaystyle H_{\rm tun}\!= 1(MN)1/4i=1Nj=1Mσ=,(tijσciψσj+h.c.).\displaystyle{}\frac{1}{(MN)^{1/4}}\!\sum_{i=1}^{N}\sum_{j=1}^{M}\!\sum_{\sigma=\uparrow,\downarrow}\!\!\left(t^{\sigma}_{ij}c^{\dagger}_{i}\psi_{\sigma j}+\mathrm{h}.\mathrm{c}.\right). (5)

The couplings tijσt^{\sigma}_{ij} and Jij;klJ_{ij;kl} are assumed to be independent Gaussian random variables with finite variances tσijtijσ¯=t2δσσ\overline{{t^{\sigma}}^{*}_{ij}t^{\sigma^{\prime}}_{ij}}=t^{2}\delta_{\sigma\sigma^{\prime}}, |Jij;kl|2¯=J2\overline{|J_{ij;kl}|^{2}}=J^{2} (Jij;kl=Jji;kl=Jij;lk=Jkl;ijJ_{ij;kl}=-J_{ji;kl}=-J_{ij;lk}=J^{*}_{kl;ij}), and zero means.

The interaction terms in the Hamiltonian (2) are decoupled within the Hubbard-Stratonovich transformations, Altland and Simons (2010); Sachdev (2015) so that in the large M,NM,N limit the self-consistent saddle-point equations are app (a)

Σc(τ)\displaystyle\Sigma_{c}(\tau) =J2Gc(τ)3+2pt2G+(τ),\displaystyle=J^{2}G_{c}(\tau)^{3}+2\sqrt{p}\,t^{2}G_{+}(\tau)\,, (6)
Gc(iωn)1\displaystyle G_{c}(\mathrm{i}\omega_{n})^{-1} =iωnΣc(iωn),\displaystyle=\mathrm{i}\omega_{n}-\Sigma_{c}(\mathrm{i}\omega_{n}), (7)
G+(iωn)\displaystyle G_{+}(\mathrm{i}\omega_{n}) =iωnt2pGc(iωn)(iωnt2pGc(iωn))2|Δ|2,\displaystyle=\frac{\mathrm{i}\omega_{n}-\frac{t^{2}}{\sqrt{p}}G_{c}(\mathrm{i}\omega_{n})}{\left(\mathrm{i}\omega_{n}-\frac{t^{2}}{\sqrt{p}}G_{c}(\mathrm{i}\omega_{n})\right)^{\!2}\!\!-|\Delta|^{2}}, (8)
1U\displaystyle\frac{1}{U} =Tn=+1(ωn+it2pGc(iωn))2+|Δ|2,\displaystyle=T\!\!\sum_{n=-\infty}^{+\infty}\!\frac{1}{\left(\omega_{n}+\frac{\mathrm{i}t^{2}}{\sqrt{p}}G_{c}(\mathrm{i}\omega_{n})\!\right)^{\!2}\ \!\!+|\Delta|^{2}}, (9)

where ωn=πT(2n+1)\omega_{n}=\pi T(2n+1) are Matsubara frequencies and p=M/Np=M/N controls the ratio between the “sites” Banerjee and Altman (2017); Chen et al. (2017); Jian and Yao (2017) in the superconductor/SYK sector. The self-energy of the SYK fermions appears in the equations (6,7) as Σc(τ)\Sigma_{c}(\tau), while Gc(τ)G_{c}(\tau) denotes the corresponding Green’s function N1i=1NTτci(τ)c¯i(0)-N^{-1}\sum_{i=1}^{N}\left\langle\mathrm{T}_{\tau}c_{i}(\tau)\bar{c}_{i}(0)\right\rangle. The Green’s functions of the \uparrow,\downarrow fermions in the superconductor Gσ(τ)=M1i=1MTτψiσ(τ)ψ¯iσ(0)G_{\sigma}(\tau)=-M^{-1}\sum_{i=1}^{M}\left\langle\mathrm{T}_{\tau}\psi_{i\sigma}(\tau)\bar{\psi}_{i\sigma}(0)\right\rangle enter the equation (8) as a half trace of the Gor’kov’s function Gor’kov (1958) G+(τ)=12(G+G)(τ)G_{+}(\tau)=\tfrac{1}{2}(G_{\uparrow}+G_{\downarrow})(\tau). Finally, relation (9) is a modified gap equation, Altland and Simons (2010) which accounts for the amount of the SYK impurity in the superconductor through Gc(τ)G_{c}(\tau) under the assumption of frequency independent pairing Δ\Delta. The chemical potential μ\mu can be accounted in the equations (6-9) by the shift |Δ|2|Δ|2+μ2|\Delta|^{2}\to|\Delta|^{2}+\mu^{2}. Below, we set μ=0\mu=0.

In the normal phase (Δ=0\Delta=0) the equations (6-8) can be written as

Σ(τ)\displaystyle\Sigma(\tau) =J2Gc(τ)3,\displaystyle=J^{2}G_{c}(\tau)^{3}, (10)
(iωnΣ(iωn))Gc(iωn)\displaystyle\left(\mathrm{i}\omega_{n}-\Sigma(\mathrm{i}\omega_{n})\right)G_{c}(\mathrm{i}\omega_{n}) =iωnt2(12p)pGc(iωn)iωnt2pGc(iωn),\displaystyle=\frac{\mathrm{i}\omega_{n}-\frac{t^{2}\left(1-2p\right)}{\sqrt{p}}G_{c}(\mathrm{i}\omega_{n})}{\mathrm{i}\omega_{n}-\frac{t^{2}}{\sqrt{p}}G_{c}(\mathrm{i}\omega_{n})}, (11)

ensuring a convenient self-energy translation ΣΣc2pt2G+\Sigma\equiv\Sigma_{c}-2\sqrt{p}\,t^{2}G_{+}. If p1/2p\ll 1/2 (2MN2M\ll N), the bare SYK Green’s function GSYK(iωn)=iπ1/4sgn(ωn)/|Jωn|G_{\rm SYK}(\mathrm{i}\omega_{n})=-\mathrm{i}\pi^{1/4}\mathrm{sgn}\left(\omega_{n}\right)/\sqrt{|J\omega_{n}|} solves the equations (10,11) in the infrared ωnJ\omega_{n}\ll J. In this regime, the Green’s function of the ψ\psi fermions G+(iωn)G_{+}(\mathrm{i}\omega_{n}) scales as ωn\sqrt{\omega_{n}} for ωn/Jp1/3(t/J)4/3\omega_{n}/J\ll p^{-1/3}(t/J)^{4/3}. In the equal sites case 2M=N2M=N, which corresponds to p=1/2p=1/2, the bare SYK Green’s function survives for (t/J)4/3ωn/J1\left(t/J\right)^{4/3}\ll\omega_{n}/J\ll 1. Another solution appears at p=1/2p=1/2 if one supposes ωn{t2|Gc|,|Σ|}\omega_{n}\ll\left\{t^{2}\left|G_{c}\right|,\left|\Sigma\right|\right\}. Then the equation (11) shortens to

Σ(iωn)\displaystyle\Sigma(\mathrm{i}\omega_{n}) =iωn2t2Gc(iωn)2.\displaystyle=\frac{\mathrm{i}\omega_{n}}{\sqrt{2}t^{2}}\,G_{c}(\mathrm{i}\omega_{n})^{\!-2}. (12)

The Green’s function that satisfies the equations (10,12) is Gc(iω)isgn(ω)/(J2t2|ωn|)1/5G_{c}(\mathrm{i}\omega)\propto-\mathrm{i}\,\mathrm{sgn}(\omega)/\left(J^{2}t^{2}|\omega_{n}|\right)^{\!-1/5} for the frequencies (t/J)3ωn/J(t/J)4/3\left(t/J\right)^{3}\ll\omega_{n}/J\ll\left(t/J\right)^{4/3}, that are achievable in the weak tunneling limit tJt\ll J. Note that the frequency window strictly depends on the coupling tt. For p1/2p\gg 1/2, the Green’s function of the cc fermions in the low-frequency limit is Gc(iωn)iωnG_{c}(\mathrm{i}\omega_{n})\propto-\mathrm{i}\omega_{n}, Jian and Yao (2017) which leads to the density of states π1ImGc(iωnω+i0+)0-\pi^{-1}\mathrm{Im}G_{c}(\mathrm{i}\omega_{n}\to\omega+\mathrm{i}0^{+})\simeq 0 vanishing in the SYK sector. Therefore, at large pp, the normal phase is given by the free fermions in the ψ\psi–dot, whose Green’s function is G+(iωn)=i/ωnG_{+}(\mathrm{i}\omega_{n})=-\mathrm{i}/\omega_{n}. To follow the frequency scaling of the Green’s function Gc(iωn)G_{c}(\mathrm{i}\omega_{n}) while changing pp, we introduce the logarithmic derivative ν=lnGc/lnωn\nu=\partial\ln G_{c}/\partial\ln\omega_{n} plotted in Figure 1 at low temperatures. Summarizing, the normal phase in the infrared limit is described by the inverse Green’s function of the SYK model at small pp, whereas it crosses over to free fermions for large pp values.

Refer to caption
Figure 1: Scaling of the Green’s function GcG_{c} in the normal phase. We plot ν=lnGc/lnωn\nu=\partial\ln G_{c}/\partial\ln\omega_{n} as a function of pp at given frequencies and finite coupling t=0.475Jt=0.475J. At low frequencies, ν\nu close to 1/2-1/2 is robust against pp increase for p<1/2p<1/2. The frequency rise moves ν\nu towards 1-1 (free fermion limit), while ν\nu crosses over to 11 for large pp. The temperature is T=104JT=10^{-4}J.
Refer to caption
Refer to caption
Figure 2: Left panel: Critical temperature as a function of the coupling strength to the SYK dot. The curves for p<0.5p<0.5 are bent at low temperatures. This illustrates the presence of two critical temperatures. At p=0.5p=0.5 the bend disappears, whereas for the values of p>0.5p>0.5 a single critical temperature decays to zero asymptotically. Right panel: The pairing gap as a function of temperature at p=0.02p=0.02. The critical coupling value is tc0.127Jt_{c}\approx 0.127J. UU is set equal to JJ in both panels.

The gap equation (9) at Δ=0\Delta=0 makes a boundary in between the normal phase and the superconducting one by setting the critical temperature TcT_{c} as a function of the coupling rate tt. Let us notice that the SYK model (4) does not have a spin degree of freedom after disorder averaging. app (a) Thus, it may be thought of as spin polarized. It suppresses superconductivity similar to magnetic impurities: Increase of the coupling to the SYK subsystem decreases the critical temperature. De Gennes (1966) There exists a critical coupling tct_{c},

1U=+dω2π(ω+itc2pGc(ω))2,\frac{1}{U}=\!\int_{-\infty}^{+\infty}\frac{d\omega}{2\pi}\!\left(\omega+\frac{\mathrm{i}t_{c}^{2}}{\sqrt{p}}G_{c}(\omega)\!\right)^{\!-2}, (13)

such as to abolish superconductivity at zero temperature. The constraint (13) follows from the gap equation (9) when Δ,T=0\Delta,T=0.

There are three competing phases contributing to the denominator of the self-consistency relation (9): SYK non-Fermi liquid, free fermions, and superconducting condensate Δ\Delta. If there are enough of the SYK fermions (N>2MN>2M), Δ\Delta interplays with the non-Fermi liquid at zero temperature. The latter one falls off with an increase in temperature, making room for the superconducting phase beyond the critical coupling, which results in the growth of the critical temperature. Indeed, Figure 2 (left) shows the bend of the critical temperature in the vicinity of the critical coupling. com (a) This phenomenon resembles the reentrant superconductivity Maple et al. (1972); Simons et al. (2012) in superconductors with Kondo impurities. Müller-Hartmann and Zittartz (1971); Riblet and Winzer (1971); Müller-Hartmann et al. (1976) The pairing gap goes down at low temperatures with an increase in coupling as in Figure 2 (right). Achieving the critical coupling when Δ\Delta vanishes at zero temperature leads to the appearance of the lower critical temperature. In contrast, the reentrant superconducting regime is absent for N<2MN<2M, since the normal phase behaves as the conventional Fermi liquid at low temperatures and large pp, as was noticed earlier. In Figure 2 (left), we show com (a) that p=1/2p=1/2 (N=2MN=2M) separates the regions with one or two critical temperatures. Similarly, consideration of the random free fermion model ijJijcicj\sum_{ij}J_{ij}c^{\dagger}_{i}c_{j} instead of the SYK model does not give the reentrance effect. In this case, the self-energy equation (6) changes to Σc(iωn)=J2Gc(iωn)+2pt2G+(iωn)\Sigma_{c}(\mathrm{i}\omega_{n})=J^{2}G_{c}(\mathrm{i}\omega_{n})+2\sqrt{p}\,t^{2}G_{+}(\mathrm{i}\omega_{n}). The results for the critical temperature are presented in Figure 3. It is still possible to suppress the superconductivity at zero temperature providing sufficient impurities, but there is only a single critical temperature as the normal phase is always set by the free fermions. com (b)

Refer to caption
Figure 3: Critical temperature as a function of the coupling strength to the random free fermions model.
Refer to caption
Refer to caption
Figure 4: The gap to temperature ratio as a function of inverse temperature at the critical coupling depends on neither the mode ratio pp (fixed U=JU=J, left panel) nor the Richardson interaction strength UU (fixed p=0.02p=0.02, right panel). In both cases, Δ\Delta saturates 2πT2\pi T at low temperatures. com (c) In the right panel, we notice that a decrease of the interaction in the superconducting dot reduces the critical temperature as in the bare Richardson model (3).

From Figure 2 (right), one notices the pairing gap at the critical coupling is T\propto T at low temperatures. We numerically examine com (a) Δ\Delta in the reentrant phase p<1/2p<1/2 for several values of pp and UU (see Figure 4). The gap saturates 2πT2\pi T almost irrespective of parameters of the problem. Unit recovery brings us to the above-mentioned relation (1) so that the gap is set by the inverse Planckian time 1/tP1/t_{\mathrm{P}} multiplied by \hbar.

This observation seems to be reminiscent of quite a peculiar feature of an Andreev billiard: Beenakker (2005) In a clean chaotic cavity proximate to a superconductor, the induced gap equals /tE=/(tLlnpFl)\hbar/t_{\mathrm{E}}=\hbar/\!\left(t_{\mathrm{L}}\ln\frac{p_{\mathrm{F}}l}{\hbar}\right),Lodder and Nazarov (1998); Adagideli and Beenakker (2002); Vavilov and Larkin (2003) where tEt_{\mathrm{E}} is the Ehrenfest time (the typical timescale of quantum dynamics), tLt_{\mathrm{L}} is the Lyapunov time, pFp_{\mathrm{F}} is the Fermi momentum, and ll is the characteristic cavity length. The effect is predicted in the regime of the Ehrenfest time far exceeds τ\tau the typical lifetime of an electron/hole excitation in the cavity. Oppositely, if tEτt_{\mathrm{E}}\ll\tau, the gap behaves as /τ\hbar/\tau, where τ\tau does not depend on the Planck constant. Melsen et al. (1997); Schomerus and Beenakker (1999) In the SYK model the Lyapunov time coincides with the Planckian relaxation time tL=/(2πkBT)=tPt_{\mathrm{L}}=\hbar/\left(2\pi k_{\mathrm{B}}T\right)=t_{\mathrm{P}}, Kitaev (2015); Maldacena and Stanford (2016) although those are different physical quantities. com (d) However, the Ehrenfest time is tLlnN/(2πkBT)t_{\mathrm{L}}\ln N\gg\hbar/(2\pi k_{\mathrm{B}}T), which differs from tPt_{\mathrm{P}} predicted in the pairing gap (1) by lnN\ln N.

To estimate the gap behavior at the critical coupling we consider the equations (6-8) at finite Δ\Delta,

(iωnΣ(iωn))Gc(iωn)=(iωnt2pGc(iωn))(iωnt2(12p)pGc(iωn))|Δ|2(iωnt2pGc(iωn))2|Δ|2,\displaystyle\Big{(}\mathrm{i}\omega_{n}-\Sigma(\mathrm{i}\omega_{n})\Big{)}G_{c}(\mathrm{i}\omega_{n})=\frac{\left(\mathrm{i}\omega_{n}\!-\!\frac{t^{2}}{\sqrt{p}}G_{c}(\mathrm{i}\omega_{n})\!\right)\!\!\left(\mathrm{i}\omega_{n}\!-\!\frac{t^{2}\left(1-2p\right)}{\sqrt{p}}G_{c}(\mathrm{i}\omega_{n})\!\right)\!-\!|\Delta|^{2}}{\left(\mathrm{i}\omega_{n}\!-\!\frac{t^{2}}{\sqrt{p}}G_{c}(\mathrm{i}\omega_{n})\!\right)^{\!2}\!\!-\!|\Delta|^{2}}, (14)

whereas the self-energy equation (10) stays unchanged. The right-hand side of the equation (14) tends to unity for p1/2p\ll 1/2. Thus it is sufficient to substitute the SYK Green’s function in the gap equation (9) in this regime.

As we look for a low-temperature correction to zero Δ\Delta at the critical coupling, we expand the gap equation (9) in powers of Δ\Delta up to the second order,

1U\displaystyle\frac{1}{U}\!\simeq  2Tn=0+1(ωn+itc2pGc(ωn))2(1|Δ|2(ωn+itc2pGc(ωn))2).\displaystyle\,2T\!\sum_{n=0}^{+\infty}\!\frac{1}{\left(\!\omega_{n}\!+\!\frac{\mathrm{i}t_{c}^{2}}{\sqrt{p}}G_{c}(\omega_{n})\!\right)^{\!\!2}}\!\!\left(\!\!1\!-\!\frac{|\Delta|^{2}}{\left(\!\omega_{n}\!+\!\frac{\mathrm{i}t_{c}^{2}}{\sqrt{p}}G_{c}(\omega_{n})\!\right)^{\!\!2}}\!\!\right)\!\!. (15)

The SYK Green’s function diverges at low frequencies as 1/ωn1/\sqrt{\omega_{n}} and decays as 1/ωn1/\omega_{n} in the ultraviolet. Hence the principal contribution to the sum (15) from the high frequencies is given by the bare ωn\omega_{n} in the denominator. On the other hand, a divergent Green’s function is crucial at low frequencies. Assuming GcG_{c} decays fast enough in comparison to ωn\omega_{n}, we replace GcG_{c} with the infrared SYK Green’s function GSYK(iωn)=iπ1/4sgn(ωn)/|Jωn|G_{SY\!K}(\mathrm{i}\omega_{n})=-\mathrm{i}\pi^{1/4}\mathrm{sgn}\left(\omega_{n}\right)/\sqrt{|J\omega_{n}|} in expression (15).

The low-temperature version of relation (15) can be written by means of the Euler-Maclaurin formula, Abramowitz and Stegun (1964)

1U\displaystyle\frac{1}{U}\simeq 0+dωπ1(ω+itc2pGSYK(ω))2(1|Δ|2(ω+itc2pGSYK(ω))2)\displaystyle\!\int_{0}^{+\infty}\!\!\frac{d\omega}{\pi}\!\frac{1}{\left(\!\omega\!+\!\frac{\mathrm{i}t_{c}^{2}}{\sqrt{p}}G_{SY\!K}(\omega)\!\right)^{\!\!2}}\!\!\left(\!\!1\!-\!\frac{|\Delta|^{2}}{\left(\!\omega\!+\!\frac{\mathrm{i}t_{c}^{2}}{\sqrt{p}}G_{SY\!K}(\omega)\!\right)^{\!\!2}}\!\!\right)\!
pTtc4GSYK(πT)2(1+2πT3GSYK(πT)/ωGSYK(πT)),\displaystyle-\!\frac{pT}{t_{c}^{4}\,G_{SY\!K}(\pi T)^{2}}\!\left(\!1+\!\frac{2\pi T}{3}\frac{\partial G_{SY\!K}(\pi T)/\partial\omega}{G_{SY\!K}(\pi T)}\!\right)\!, (16)

where we expand up to T2T^{2} keeping in mind that ΔT\Delta\propto T at the critical coupling. com (e) Finally, one notices two terms in the top row of the equation (16) that match the critical coupling condition (13). Therefore, we obtain com (f)

Δ(T)6πT.\displaystyle\Delta(T)\simeq\,\sqrt{6}\pi T. (17)

Although this estimate gives η1.22\eta\approx 1.22 that exceeds the found numerical value η0.96\eta\approx 0.96 for the pairing gap Δη/tP\Delta\approx\eta\,\hbar/t_{\mathrm{P}}, the derived low-temperature gap behavior (17) is independent of the problem parameters as in Figure 4.

Conclusion.— In this manuscript, we considered the superconducting proximity effect for the Sachdev-Ye-Kitaev model. We have shown, that the superconducting dot coupled to the complex SYK model possesses reentrant superconductivity. At the critical coupling, which gives rise to the occurrence of a lower critical temperature, the pairing gap disappears at T=0T=0 and grows linearly with an increase in temperature. The linear–TT growth of the gap is given by /tP\hbar/t_{\mathrm{P}}, where tP=/(2πkBT)t_{\mathrm{P}}=\hbar/\left(2\pi k_{\mathrm{B}}T\right) is the Planckian relaxation time. The same timescale serves as an ultimate bound on many-body quantum chaos, Maldacena et al. (2016) saturated in strongly coupled systems without quasiparticle excitations. Thereby a natural question arises whether the pairing gap is an appropriate physical observable for the Lyapunov spectrum Romero-Bermúdez et al. (2019) of the SYK model. Accurate studies of the OTOC in the proposed system (2) might shed light on that. On its own, Δη/tP\Delta\approx\eta\,\hbar/t_{\mathrm{P}} may be used to characterize the cSYK quantum dots. Chen et al. (2018); Danshita et al. (2017) However, this requires consideration of a more realistic setup such as a superconducting lead attached to the particular realization of the complex SYK model.

Acknowledgements.
We are grateful to C. W. J. Beenakker for drawing our attention to this problem. The authors have benefited from inspiring discussions with D. V. Efremov, Yu. Malitsky, and K. E. Schalm. This research was supported by the Netherlands Organization for Scientific Research (NWO/OCW), by the European Research Council, and by the DOE Contract DEFG02-08ER46482 (NVG).

Appendix A Derivation of the gap equation

The imaginary time action averaged over disorder is

S=\displaystyle S= 0β𝑑τ[i=1Nc¯iτci+i=1Mσ=,ψ¯σi(τμ)ψσiUMi,j=1Mψ¯iψ¯iψjψj]\displaystyle{}\int_{0}^{\beta}\!\!d\tau\Bigg{[}\sum_{i=1}^{N}\bar{c}_{i}\partial_{\tau}c_{i}+\!\sum_{i=1}^{M}\!\sum_{\sigma=\uparrow,\downarrow}\!\!\bar{\psi}_{\sigma i}\left(\partial_{\tau}-\mu\right)\psi_{\sigma i}-\frac{U}{M}\!\sum_{i,j=1}^{M}\!\!\bar{\psi}_{\uparrow i}\bar{\psi}_{\downarrow i}\psi_{\downarrow j}\psi_{\uparrow j}\Bigg{]}
0β𝑑τ0β𝑑τ[t2NMi=1Nj=1Mσ=,c¯iψσj(τ)ψ¯σjci(τ)J24N3i,j,k,l=1Nc¯ic¯jckcl(τ)c¯lc¯kcjci(τ)],\displaystyle{}-\int_{0}^{\beta}\!\!d\tau\!\!\int_{0}^{\beta}\!\!d\tau^{\prime}\Bigg{[}\frac{t^{2}}{\sqrt{NM}}\!\sum_{i=1}^{N}\sum_{j=1}^{M}\!\sum_{\sigma=\uparrow,\downarrow}\!\!\bar{c}_{i}\psi_{\sigma j}(\tau)\bar{\psi}_{\sigma j}c_{i}(\tau^{\prime})-\frac{J^{2}}{4N^{3}}\!\sum_{i,j,k,l=1}^{N}\!\bar{c}_{i}\bar{c}_{j}c_{k}c_{l}(\tau)\bar{c}_{l}\bar{c}_{k}c_{j}c_{i}(\tau^{\prime})\Bigg{]}, (18)

where β\beta is the inverse temperature. Following Refs. Altland and Simons, 2010; Sachdev, 2015, we decouple the interaction term on the top line of the action (18) with the Hubbard–Stratonovich transformation and introduce three non-local fields Gσ(τ,τ)=M1i=1Mψiσ(τ)ψ¯iσ(τ)G_{\sigma}(\tau,\tau^{\prime})=-M^{-1}\sum_{i=1}^{M}\psi_{i\sigma}(\tau)\bar{\psi}_{i\sigma}(\tau^{\prime}), Gc(τ,τ)=N1i=1Nci(τ)c¯i(τ)G_{c}(\tau,\tau^{\prime})=-N^{-1}\sum_{i=1}^{N}c_{i}(\tau)\bar{c}_{i}(\tau^{\prime}) together with Σσ(τ,τ)\Sigma_{\sigma}(\tau,\tau^{\prime}), Σc(τ,τ)\Sigma_{c}(\tau,\tau^{\prime}) as the corresponding Lagrange multipliers:

S=\displaystyle S= 0βdτ0βdτ[MUδ(ττ)|Δ|2i=1MΨ¯i(τ)(δ(ττ)(τμ)Σ(τ,τ)δ(ττ)Δδ(ττ)Δ¯δ(ττ)(τ+μ)Σ(τ,τ))Ψi(τ)\displaystyle{}\int_{0}^{\beta}\!\!d\tau\!\!\int_{0}^{\beta}\!\!d\tau^{\prime}\Bigg{[}\frac{M}{U}\delta(\tau-\tau^{\prime})|\Delta|^{2}-\!\sum_{i=1}^{M}\!\bar{\Psi}_{i}(\tau)\!\begin{pmatrix}-\delta(\tau-\tau^{\prime})\left(\partial_{\tau}-\mu\right)-\Sigma_{\uparrow}(\tau,\tau^{\prime})&\delta(\tau-\tau^{\prime})\Delta\\ \delta(\tau-\tau^{\prime})\bar{\Delta}&-\delta(\tau-\tau^{\prime})\left(\partial_{\tau}+\mu\right)-\Sigma_{\downarrow}(\tau,\tau^{\prime})\end{pmatrix}\!\Psi_{i}(\tau^{\prime})
i=1Nc¯i(τ)(δ(ττ)τΣc(τ,τ))ci(τ)Mσ=,(Σσ(τ,τ)NMt2Gc(τ,τ))Gσ(τ,τ)\displaystyle{}-\sum_{i=1}^{N}\bar{c}_{i}(\tau)\big{(}-\delta(\tau-\tau^{\prime})\partial_{\tau}-\Sigma_{c}(\tau,\tau^{\prime})\big{)}c_{i}(\tau^{\prime})-M\!\sum_{\sigma=\uparrow,\downarrow}\!\!\left(\Sigma_{\sigma}(\tau,\tau^{\prime})-\sqrt{\frac{N}{M}}\,t^{2}G_{c}(\tau,\tau^{\prime})\right)G_{\sigma}(\tau^{\prime},\tau)
N(Σc(τ,τ)Gc(τ,τ)+J24Gc(τ,τ)4)],\displaystyle{}-N\left(\Sigma_{c}(\tau,\tau^{\prime})G_{c}(\tau^{\prime},\tau)+\frac{J^{2}}{4}G_{c}(\tau,\tau^{\prime})^{4}\right)\!\Bigg{]}, (19)

where Ψ¯i=(ψ¯iψi)\bar{\Psi}_{i}=\begin{pmatrix}\bar{\psi}_{\uparrow i}&\psi_{\downarrow i}\end{pmatrix} and Ψi=(ψiψ¯i)T\Psi_{i}=\begin{pmatrix}\psi_{\uparrow i}&\bar{\psi}_{\downarrow i}\end{pmatrix}^{T} are Nambu spinors. Integrating out fermions and assuming constant Δ\Delta, we get:

S=\displaystyle S= βMU|Δ|2Mn=+log[(iωnΣ(iωn)+μ)(iωnΣ(iωn)μ)|Δ|2]Nn=+log[iωnΣc(iωn)]\displaystyle{}\frac{\beta M}{U}|\Delta|^{2}-\!M\!\!\sum_{n=-\infty}^{+\infty}\!\!\!\log\bigg{[}\!\left(\mathrm{i}\omega_{n}-\Sigma_{\uparrow}(\mathrm{i}\omega_{n})+\mu\right)\left(\mathrm{i}\omega_{n}-\Sigma_{\downarrow}(\mathrm{i}\omega_{n})-\mu\right)-|\Delta|^{2}\bigg{]}-\!N\!\!\sum_{n=-\infty}^{+\infty}\!\!\!\log\bigg{[}\mathrm{i}\omega_{n}-\Sigma_{c}(\mathrm{i}\omega_{n})\bigg{]}
0β𝑑τ0β𝑑τ[Mσ=,(Σσ(τ,τ)NMt2Gc(τ,τ))Gσ(τ,τ)+N(Σc(τ,τ)Gc(τ,τ)+J24Gc(τ,τ)4)],\displaystyle{}-\int_{0}^{\beta}\!\!d\tau\!\!\int_{0}^{\beta}\!\!d\tau^{\prime}\Bigg{[}M\!\!\sum_{\sigma=\uparrow,\downarrow}\!\!\left(\Sigma_{\sigma}(\tau,\tau^{\prime})-\sqrt{\frac{N}{M}}\,t^{2}G_{c}(\tau,\tau^{\prime})\!\right)\!G_{\sigma}(\tau^{\prime},\tau)\!+N\!\left(\Sigma_{c}(\tau,\tau^{\prime})G_{c}(\tau^{\prime},\tau)+\frac{J^{2}}{4}G_{c}(\tau,\tau^{\prime})^{4}\right)\!\Bigg{]}, (20)

where ωn=π(2n+1)/β\omega_{n}=\pi(2n+1)/\beta are Matsubara frequencies. In the limit of MM, N1N\gg 1, the saddle-point equations are:

Σ(τ)\displaystyle\Sigma_{\uparrow}(\tau) =t2pGc(τ),Σ(τ)=t2pGc(τ),\displaystyle{}=\frac{t^{2}}{\sqrt{p}}G_{c}(\tau),\;\;\;\Sigma_{\downarrow}(\tau)=\frac{t^{2}}{\sqrt{p}}G_{c}(\tau), (21)
Σc(τ)\displaystyle\Sigma_{c}(\tau) =J2Gc(τ)3+pt2σ=,Gσ(τ),\displaystyle{}=J^{2}G_{c}(\tau)^{3}+\sqrt{p}\,t^{2}\!\!\sum_{\sigma=\uparrow,\downarrow}\!\!G_{\sigma}(\tau), (22)
G(iωn)\displaystyle G_{\uparrow}(\mathrm{i}\omega_{n}) =iωnμΣ(iωn)(iωnΣ(iωn)+μ)(iωnΣ(iωn)μ)|Δ|2,\displaystyle{}=\frac{\mathrm{i}\omega_{n}-\mu-\Sigma_{\downarrow}(\mathrm{i}\omega_{n})}{\left(\mathrm{i}\omega_{n}-\Sigma_{\uparrow}(\mathrm{i}\omega_{n})+\mu\right)\left(\mathrm{i}\omega_{n}-\Sigma_{\downarrow}(\mathrm{i}\omega_{n})-\mu\right)-|\Delta|^{2}}, (23)
G(iωn)\displaystyle G_{\downarrow}(\mathrm{i}\omega_{n}) =iωn+μΣ(iωn)(iωnΣ(iωn)+μ)(iωnΣ(iωn)μ)|Δ|2,\displaystyle{}=\frac{\mathrm{i}\omega_{n}+\mu-\Sigma_{\uparrow}(\mathrm{i}\omega_{n})}{\left(\mathrm{i}\omega_{n}-\Sigma_{\uparrow}(\mathrm{i}\omega_{n})+\mu\right)\left(\mathrm{i}\omega_{n}-\Sigma_{\downarrow}(\mathrm{i}\omega_{n})-\mu\right)-|\Delta|^{2}}, (24)
Gc(iωn)1\displaystyle G_{c}(\mathrm{i}\omega_{n})^{-1} =iωnΣc(iωn),\displaystyle{}=\mathrm{i}\omega_{n}-\Sigma_{c}(\mathrm{i}\omega_{n}), (25)
1U\displaystyle\frac{1}{U} =1βn=+1(ωn+iΣ(iωn)iμ)(ωn+iΣ(iωn)+iμ)+|Δ|2,\displaystyle{}=\!\frac{1}{\beta}\!\sum_{n=-\infty}^{+\infty}\!\frac{1}{\left(\omega_{n}+\mathrm{i}\Sigma_{\uparrow}(\mathrm{i}\omega_{n})-\mathrm{i}\mu\right)\left(\omega_{n}+\mathrm{i}\Sigma_{\downarrow}(\mathrm{i}\omega_{n})+\mathrm{i}\mu\right)+|\Delta|^{2}}, (26)

where we introduced the parameter p=M/Np=M/N representing the amount of the SYK “impurities” in the superconductor sector.

We exclude the self-energies Σσ\Sigma_{\sigma} (21), so that one obtains four Schwinger-Dyson equations:

Σc(τ)\displaystyle\Sigma_{c}(\tau) =J2Gc(τ)3+2pt2G+(τ),\displaystyle{}=J^{2}G_{c}(\tau)^{3}+2\sqrt{p}\,t^{2}G_{+}(\tau), (27)
G+(iωn)\displaystyle G_{+}(\mathrm{i}\omega_{n}) =iωnt2pGc(iωn)(iωnt2pGc(iωn))2μ2|Δ|2,\displaystyle{}=\frac{\mathrm{i}\omega_{n}-\frac{t^{2}}{\sqrt{p}}G_{c}(\mathrm{i}\omega_{n})}{\left(\mathrm{i}\omega_{n}-\frac{t^{2}}{\sqrt{p}}G_{c}(\mathrm{i}\omega_{n})\right)^{2}-\mu^{2}-|\Delta|^{2}}, (28)
Gc(iωn)1\displaystyle G_{c}(\mathrm{i}\omega_{n})^{-1} =iωnΣc(iωn),\displaystyle{}=\mathrm{i}\omega_{n}-\Sigma_{c}(\mathrm{i}\omega_{n}), (29)
1U\displaystyle\frac{1}{U} =1βn=+1(ωn+it2pGc(iωn))2+μ2+|Δ|2,\displaystyle{}=\!\frac{1}{\beta}\!\sum_{n=-\infty}^{+\infty}\!\frac{1}{\left(\omega_{n}+\frac{\mathrm{i}t^{2}}{\sqrt{p}}G_{c}(\mathrm{i}\omega_{n})\right)^{2}+\mu^{2}+|\Delta|^{2}}, (30)

where the latter one (30) is a modified BCS gap equation Altland and Simons (2010) and G+=12(G+G)G_{+}=\frac{1}{2}\left(G_{\uparrow}+G_{\downarrow}\right).

Appendix B Saddle-point numerical analysis

B.1 The algorithm

To solve the equations (27-30), we use an iterative approach that is equivalent to finding the fixed point (the point to which the iterative procedure converges) of the operator T^\hat{T} representing the Schwinger-Dyson equations (27-29) set on a fixed grid of Matsubara frequencies. com (g) One starts with an empty seed G0G^{0} and applies iterations

Gk+1=T^Gk\displaystyle G^{k+1}=\hat{T}G^{k} (31)

until

Gk+1Gkε,\|G^{k+1}-G^{k}\|\leq\varepsilon, (32)

where we set the precision to ε=104\varepsilon=10^{-4} and \|\cdot\| denotes the euclidean norm of the vector.

The straightforward approach (31) converges rarely. One improves convergence modifying (31) as

Gk+1=λGk+(1λ)T^Gk,\displaystyle G^{k+1}=\lambda G^{k}+(1-\lambda)\hat{T}G^{k}, (33)

where 0<λ<10<\lambda<1 is a tunable parameter. This particular approach (33) has been used to compute the Green’s function of the SYK model. Maldacena and Stanford (2016) However, the convergence of the algorithm (33) may sufficiently slow down when one considering extra Schwinger-Dyson equations coupled to those of the bare SYK model or expands the parameter space. In our case, that happens due to coupling of the SYK model to a superconductor. To cope with this problem, we suggest using the adaptive golden ratio algorithm, Malitsky (2019) where the weight λ\lambda is not fixed but automatically adjusted to the local properties of the operator T^\hat{T}:

λk\displaystyle\lambda_{k} =min{109λk1,916λk2GkGk12GkT^GkGk1+T^Gk12},\displaystyle=\min\left\{\frac{10}{9}\lambda_{k-1},\ \frac{9}{16\lambda_{k-2}}\frac{\|G^{k}-G^{k-1}\|^{2}}{\|G^{k}-\hat{T}G^{k}-G^{k-1}+\hat{T}G^{k-1}\|^{2}}\right\}, (34)
G¯k\displaystyle\bar{G}^{k} =Gk+2G¯k13,\displaystyle=\frac{G^{k}+2\bar{G}^{k-1}}{3}, (35)
Gk+1\displaystyle G^{k+1} =G¯kλkGk+λkT^Gk.\displaystyle=\bar{G}^{k}-\lambda_{k}G^{k}+\lambda_{k}\hat{T}G^{k}. (36)

Above we introduce G¯\bar{G} as an auxiliary function that requires G¯0=G1\bar{G}^{0}=G^{1} and λ0=λ1>0\lambda_{0}=\lambda_{-1}>0. Computationally, the algorithm (34-36) is of the same complexity as (31) and (33), while the adaptive step allows for a significant speedup.

We treat the pairing gap Δ\Delta, the temperature TT, and the coupling strength tt that enter the equations (27-29) as an external set of parameters. Once the Green’s functions are found within the procedure (34-36), we choose the data that satisfies the self-consistency relation (30) to produce the finite-temperature phase diagrams.

B.2 Precision and grid

Matsubara frequencies ωn=πT(2n+1)\omega_{n}=\pi T(2n+1) define a natural discrete grid. We set the ultraviolet cut-off NN such that n[N,N+1,,N1,N]n\in\left[-N,-N+1,\ldots,N-1,N\right], where the reliable NN is of the order 10410^{4}10510^{5} with the accuracy criteria (32) ε=104\varepsilon=10^{-4}. The numerical analysis becomes more demanding as one enters the low-temperature regime in the vicinity of the critical coupling. We reach the lowest temperature of T103T\sim 10^{-3} using N=1.5×106N=1.5\times 10^{6}, with a main computational bottleneck coming from the computer memory. Also, the computation of the lowest critical temperatures requires an increase of the accuracy for the self-consistency condition (30) and ε\varepsilon (32) to 10510^{-5}10610^{-6}.

Refer to caption
Refer to caption
Figure 5: The pairing gap as a function of temperature at the critical coupling. Left panel: fixed U=JU=J. Right panel: fixed p=0.02p=0.02.
pp 0.0020.002 0.020.02 0.020.02 0.020.02 0.050.05
UU 1.01.0 1.01.0 0.750.75 0.50.5 1.01.0
tct_{c} 0.0710112J0.0710112J 0.126827J0.126827J 0.10057J0.10057J 0.07294J0.07294J 0.1607J0.1607J
η\eta 0.95880.9588 0.96210.9621 0.94870.9487 0.95330.9533 0.97210.9721
δ\delta 2.96×105-2.96\times 10^{-5} 5.47×104-5.47\times 10^{-4} 3.73×104-3.73\times 10^{-4} 1.54×103-1.54\times 10^{-3} 9.86×105-9.86\times 10^{-5}
Table 1: The values of the critical coupling and the interpolation parameters for given pp and UU.

One of the objectives of this manuscript is to study the pairing gap at the critical coupling and low temperatures. In this regime, the gap grows linearly in temperature as shown in Figure 5. The critical coupling tct_{c} is found as a condition when the off-set δ\delta of the interpolating function Δ=2πηT+δ\Delta=2\pi\eta\,T+\delta vanishes (see numerical values in Table 1). The system is sensitive to the coupling changes for small values of pp, therefore, the precision of tct_{c} reaches 10710^{-7} for p=0.002p=0.002.

References