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

Interplay between superconductivity and non-Fermi liquid at a quantum-critical point in a metal.
I: The γ\gamma model and its phase diagram at T=0T=0. The case 0<γ<10<\gamma<1.

Artem Abanov Department of Physics, Texas A&M University, College Station, USA    Andrey V. Chubukov School of Physics and Astronomy and William I. Fine Theoretical Physics Institute, University of Minnesota, Minneapolis, MN 55455, USA
Abstract

Near a quantum-critical point in a metal a strong fermion-fermion interaction, mediated by a soft boson, acts in two different directions: it destroys fermionic coherence and it gives rise to an attraction in one or more pairing channels. The two tendencies compete with each other. We analyze a class of quantum-critical models, in which momentum integration and the selection of a particular pairing symmetry can be done explicitly, and the competition between non-Fermi liquid and pairing can be analyzed within an effective model with dynamical electron-electron interaction V(Ωm)1/|Ωm|γV(\Omega_{m})\propto 1/|\Omega_{m}|^{\gamma} (the γ\gamma-model). In this paper, the first in the series, we consider the case T=0T=0 and 0<γ<10<\gamma<1. We argue that tendency to pairing is stronger, and the ground state is a superconductor. We argue, however, that a superconducting state is highly non-trivial as there exists a discrete set of topologically distinct solutions for the pairing gap Δn(ωm)\Delta_{n}(\omega_{m}) (n=0,1,2,n=0,1,2...,\infty). All solutions have the same spatial pairing symmetry, but differ in the time domain: Δn(ωm)\Delta_{n}(\omega_{m}) changes sign nn times as a function of Matsubara frequency ωm\omega_{m}. The n=0n=0 solution Δ0(ωm)\Delta_{0}(\omega_{m}) is sign-preserving and tends to a finite value at ωm=0\omega_{m}=0, like in BCS theory. The n=n=\infty solution corresponds to an infinitesimally small Δ(ωm)\Delta(\omega_{m}), which oscillates down to the lowest frequencies as Δ(ωm)|ωm|γ/2cos[2βlog(|ωm|/ω0)]\Delta(\omega_{m})\propto|\omega_{m}|^{\gamma/2}\cos\left[2\beta\log(|\omega_{m}|/\omega_{0})\right], where β=O(1)\beta=O(1) and ω0\omega_{0} is of order of fermion-boson coupling. As a proof, we obtain the exact solution of the linearized gap equation at T=0T=0 on the entire frequency axis for all 0<γ<10<\gamma<1, and an approximate solution of the non-linear gap equation. We argue that the presence of an infinite set of solutions opens up a new channel of gap fluctuations. We extend the analysis to the case where the pairing component of the interaction has additional factor 1/N1/N and show that there exists a critical Ncr>1N_{cr}>1, above which superconductivity disappears, and the ground state becomes a non-Fermi liquid. We show that all solutions develop simultaneously once NN gets smaller than NcrN_{cr}.

I Introduction.

The interplay between superconductivity and pairing near a quantum-critical point (QCP) in a metal is a fascinating subject, which attracted substantial attention in the correlated electron community after the discovery of superconductivity (SC) in the cuprates, heavy-fermion and organic materials, and, more recently, Fe-pnictides and Fe-chalcogenides. (see, e.g., Refs. Monthoux et al. (2007); Scalapino (2012); Norman (2014); Maiti and Chubukov (2014); Keimer et al. (2015); Shibauchi et al. (2014); Fernandes and Chubukov (2016); Fradkin et al. (2010); Yang et al. (2006); Fratino et al. (2016); Sachdev (2018); Coleman (2015)). Itinerant QC models, analyzed analytically in recent years include, e.g., models of fermions in spatial dimensions D3D\leq 3 (Refs. Mross et al. (2010); Metlitski et al. (2015); Mahajan et al. (2013); *raghu2; *raghu3; *raghu4; *raghu5; Wang et al. (2018, 2017a); Fitzpatrick et al. (2015b)), two-dimensional (2D) models near a spin-density-wave (SDW) and charge-density-wave (CDW) instabilities (Refs.  Millis (1992); Altshuler et al. (1995a); Sachdev et al. (1995); Abanov et al. (2001a, 2003); *acs2; *finger_2001; *acn; Sachdev et al. (2009); *Subir2; *Sachdev2019; Vojta and Sachdev (1999); Wang and Chubukov (2013a); Efetov et al. (2013); *efetov2; *efetov3; Tsvelik (2017); Metlitski and Sachdev (2010a); Dalidovich and Lee (2013); *sslee_2018; Bauer and Sachdev (2015); Castellani et al. (1995); *ital2; *ital3; Wang and Chubukov (2013a); *wang23; *wang_22; Khodas et al. (2010); *tsvelik; *tsvelik_1; *tsvelik_2; *tsvelik_3; Metzner et al. (2003); *DellAnna2006; Vilardi et al. (2018); *metzner_1; *metzner_2; Sordi et al. (2012); Chubukov and Wölfle (2014)), 2kF2k_{F} density-wave instability (Refs.Altshuler et al. (1995b); *2kf2; *2kf3), pair-density-wave instability  Kim et al. (2008); Lawler and Fradkin (2007); *Fradkin2009; *Fradkin2014, q=0q=0 instabilities towards a Pomeranchuk order instabilityBonesteel et al. (1996); *Oganesyan2001; *q=01; *q=02; Metzner et al. (2003); *DellAnna2006; Rech et al. (2006); Lee (2009); Metlitski and Sachdev (2010b); Lederer et al. (2015); Maslov and Chubukov (2012); *maslov2; *maslov3; *maslov4; You et al. (2016); Wang et al. (2001); *triplet2; *triplet3; Lederer et al. (2017); Klein et al. (2019); *avi_2 and towards circulating currents Bok et al. (2016), 2D fermions at a half-filled Landau level Lee (1989); *Monien1993; *Nayak1994; *Altshuler1994; *Kim_1994, generic QC models with different critical exponents Raghu et al. (2015); Moon and Chubukov (2010); Khveshchenko and Shively (2006); She and Zaanen (2009); *She2010; Wang et al. (2016); Lee et al. (2018); Wu et al. (2019a, b); *Abanov_19; Yuzbashyan et al. (a), Sachdev-Ye-Kitaev (SYK) and SYK-Yukawa models Patel and Sachdev (2019); Esterlis and Schmalian (2019); Wang (2020); Hauck et al. (2019); Chowdhury and Berg (2020), strong coupling limit of electron-phonon superconductivity, Combescot (1995); Bergmann and Rainer (1973); *Bergmann2; *ad; Marsiglio et al. (1988); *Marsiglio_91; Karakozov et al. (1991); Chubukov et al. (2020a), and even color superconductivity of quarks, mediated by gluon exchangeSon (1999); *son2. These problems have also been studied using various numerical techniques Yang et al. (2011); Gerlach et al. (2017); *berg_2; *berg_3; *berg_4; Haule and Kotliar (2007); *kotliar2; Sordi et al. (2012); Georges et al. (2013); *georges2.

From theory perspective, the key interest in the pairing near a QCP is due to the fact that an effective dynamic electron-electron interaction, V(q,Ω)V(q,\Omega), mediated by a critical collective boson, which condenses at a QCP, provides strong attraction in one or more pairing channels and, at the same time, gives rise to non-Fermi liquid (NFL) behavior in the normal state. The two tendencies compete with each other: fermionic incoherence, associated with NFL behavior, destroys Cooper logarithm and reduces the tendency to pairing, while the opening of a SC gap eliminates the scattering at low energies and reduces the tendency to NFL. To find the winner (SC or NFL), one needs to analyze the set of integral equations for the fermionic self-energy, Σ(𝐤,ω)\Sigma({\bf k},\omega), and the gap function, Δ(𝐤,ω)\Delta({\bf k},\omega), for fermions with momentum/frequency (𝐤,ω)({\bf k},\omega) and (𝐤,ω)(-{\bf k},-\omega).

We consider the subset of models in which collective bosons are slow modes compared to dressed fermions, for one reason or the other. In this situation, which bears parallels with Eliashberg theory for electron-phonon interaction Eliashberg (1960), the self-energy and the pairing vertex can be approximated by their values at the Fermi surface (FS) and computed within one-loop approximation. The self-energy on the FS, Σ(𝐤,ω)\Sigma({\bf k},\omega) is invariant under rotations from the point group of the underlying lattice. The rotational symmetry of the gap function Δ(𝐤F,ω)\Delta({\bf k}_{F},\omega) and the relation between the phases of Δ(𝐤F,ω)\Delta({\bf k}_{F},\omega) on different FS’s in multi-band systems are model specific. Near a ferromagnetic QCP, the pairing interaction mediated by a soft boson is attractive in the p-wave channel. Near an antiferromagnetic QCP, the strongest attraction is in dd-wave channel for the case of a single FS and fermionic density near half-filling. For a nearly compensated metal with hole and electron pockets, as in Fe-based superconductors, the two attractive channels near an antiferromagnetic QCP are s+s^{+-} and dd-wave. Near a q=0q=0 nematic QCP, the pairing interaction, mediated by soft nematic fluctuations, is attractive in all channels: ss-wave, pp-wave, dd-wave, etc. In each particular case, one has to project the pairing interaction into the proper irreducible channel, find the strongest one, and solve for the pairing vertex within a given pairing symmetry. In principle, even after projection, one has to solve an infinite set of coupled equations in momentum space as in a lattice system each irreducible representation contains an infinite set of eigenfunctions. However, near a QCP the pairing is often confined to a narrow range on the FS around special ”hot spots”. In this situation, the momentum integration near the FS can be carried out exactly, and the set of coupled equations for the self-energy and the gap function reduce to two one-dimensional (1D) equations for Σ(ω)\Sigma(\omega) and Δ(ω)\Delta(\omega), each with frequency-dependent effective ’local” interaction V(Ω)V(\Omega) (Ref. Abanov et al. (2001a). The same holds for the cases of s-wave pairing by a soft optical phonon, when momentum-dependencies of Σ\Sigma and Δ\Delta are not crucial and can be neglected, of non-s-wave pairing, when one eigenfunction gives the dominant contribution to the gap (e.g., coskxcosky\cos k_{x}-\cos k_{y} for dd-wave pairing in the cuprates), and for the case when the fermionic density of states is peaked at particular 𝐤F{\bf k}_{F} due to, e.g. van-Hove singularities ( see e.g., Maiti and Chubukov, 2014 and references therein).

Away from a QCP, the effective V(Ω)V(\Omega) tends to a finite value at Ω=0\Omega=0. In this situation, the fermionic self-energy has a FL form at the smallest frequencies, the equation for Δ(ω)\Delta(\omega) is similar to that in a conventional Eliashberg theory for phonon-mediated superconductivity, and the only qualitative distinction for electronically-mediated pairing is that V(Ω)V(\Omega) by itself changes below TcT_{c} due to feedback from fermionic pairing on collective modes.

At a QCP, the situation becomes qualitatively different because the effective interaction V(Ω)V(\Omega), mediated by a critical massless boson, becomes a singular function of frequency: on Matsubara axis V(Ωm)1/|Ωm|γV(\Omega_{m})\propto 1/|\Omega_{m}|^{\gamma} (Fig. 1). The exponent γ>0\gamma>0 depends on the model, ranging from small γ=0(ϵ)\gamma=0(\epsilon) in models in D=3ϵD=3-\epsilon to γ1\gamma\leq 1 in 2D models at SDW, CDW, and nematic QCP and in Yukawa-SYK model Chubukov et al. (2020b); Maiti and Chubukov (2014); Chubukov et al. (2020a); Esterlis and Schmalian (2019); Wang (2020). The case γ=2\gamma=2 corresponds to fermions, interacting with a critical Einstein phonon. The set of models with V(Ωm)1/|Ωm|γV(\Omega_{m})\propto 1/|\Omega_{m}|^{\gamma} has been nicknamed the γ\gamma-model, and we will use this notation. We present an (incomplete) set of models in Tables 1- 3.

Refer to caption
Figure 1: Frequency dependence of the effective interaction V(Ωm)V(\Omega_{m}), mediated by a soft boson, at T=0T=0. Away from a QCP, V(Ωm)V(\Omega_{m}) tends to a finite value at Ωm=0\Omega_{m}=0. Right at a QCP, the boson becomes massless, and V(Ωm)V(\Omega_{m}) diverges as 1/|Ωm|γ1/|\Omega_{m}|^{\gamma}.
Refer to caption
Figure 2: The phase diagram of the γ\gamma-model. In this and subsequent papers we argue that for the pairing near a QCP the onset temperatures for the pairing and for long-range superconducting order, TpT_{p} and TcT_{c}, differ. In between TpT_{p} and TcT_{c} the system displays pseudogap behavior. The width of the pseudogap region increases with increasing γ\gamma and extends to T=0T=0 for γ2\gamma\geq 2. The behavior at larger γ\gamma requires separate consideration Yuzbashyan et al. (b).

In this and subsequent papers, we present comprehensive analysis of the competition between NFL and SC within the γ\gamma-model. We define the dimensionless V(Ωm)V(\Omega_{m}) as V(Ωm)=(g¯/|Ωm|)γV(\Omega_{m})=({\bar{g}}/|\Omega_{m}|)^{\gamma}, where g¯{\bar{g}} is the effective fermion-boson coupling. This g¯{\bar{g}} is the only parameter in the model with the dimension of energy, and we will see that it determines both the magnitude of the gap function Δ(ωm)\Delta(\omega_{m}) and the upper limit of NFL behavior in the normal state, i.e., the scale below which Σ(ωm)>ωm\Sigma(\omega_{m})>\omega_{m}. We show that the competition between NFL and SC holds for all values of γ\gamma, but the physics and the computational analysis are different for the models with γ<1\gamma<1, γ=1\gamma=1, 1<γ<21<\gamma<2, γ=2\gamma=2, and γ>2\gamma>2, which we consider separately. We show that for all γ\gamma, a NFL self-energy in the normal state does not prevent the formation of bound pairs of fermions at a non-zero onset temperature TpT_{p}. However, we argue that TpT_{p} is generally higher than the actual superconducting TcT_{c}, and at Tc<T<TpT_{c}<T<T_{p} bound pairs remain incoherent. In this TT range various observables, e.g., the spectral function and the density of states, display pseudogap behavior. This holds for all γ\gamma, but the width of the pseudogap phase increases with γ\gamma. We relate the pseudogap behavior to strong gap fluctuations, but argue that these fluctuations can be viewed as long-wavelength phase fluctuations only in some range T¯p<T<Tp{\bar{T}}_{p}<T<T_{p}. At smaller Tc<T<T¯pT_{c}<T<{\bar{T}}_{p}, pseudogap behavior is predominantly due to the existence of low-energy ’longitudinal” gap fluctuations, which change the functional form of Δ(ω)\Delta(\omega) and cause phase slips. We argue that longitudinal gap fluctuations develop a zero mode at γ=2\gamma=2 in which case SC order gets destroyed already at T=0T=0. Specifically, the presence of a zero mode implies that there exists an infinite number of solutions for Δ(ω)\Delta(\omega), all with the same condensation energy. In this situation, the ground state is a mixture of different Δ(ω)\Delta(\omega), each with its own phase. We will argue that TcT_{c} gradually vanishes at γ2\gamma\to 2 and remains zero at larger γ\gamma, while TpT_{p} and T¯p{\bar{T}}_{p} stay finite. We show the phase diagram in Fig. 2. Our results present the scenario for the pseudogap, which holds even if for a given solution for Δ(ω)\Delta(\omega) SC stiffness is larger than TcT_{c}, i.e., conventional phase fluctuations are weak. This scenario is complementary to the one in which the smallness of the superfluid stiffness is caused by the closeness to a Mott transition (see, e.g., Ref. Simard et al. (2019) and references therein).

I.1 A brief summary of the results of this work

In this paper, the first in the series, we analyze the γ\gamma-model at T=0T=0 at 0<γ<10<\gamma<1. For definiteness we focus on spin-singlet pairing. The analysis of spin-triplet pairing is more involved because of different spin factors in the self-energy and the pairing vertex Wang et al. (2001); Roussev and Millis (2001); Chubukov et al. (2003) In this paper we discuss even frequency pairing. The analysis of odd-frequency pairing Linder and Balatsky (2019) is more involved and requires a separate consideration.

Our key goal is to understand the interplay between the competing tendencies toward pairing and toward NFL behavior. The first comes from the interaction in the particle-particle channel, the second from the interaction in the particle-hole channel. In the original γ\gamma-model, both interactions are given by the same V(Ωm)V(\Omega_{m}). In order to separate the two tendencies we extend the model and introduce the knob to vary the relative strength of the interaction in the two channels. Specifically, we multiply the pairing interaction by 1/N1/N and treat NN as a parameter. For N>1N>1 (N<1N<1) the tendency towards pairing decreases (increases) compared to the one towards NFL ground state. The extension to integer N>1N>1 can be formally justified by extending the model from SU(1)SU(1) to an SU(N)SU(N) global symmetry Raghu et al. (2015). In our analysis, we use the extension to arbitrary N1N\neq 1 just as a computational trick to better understand what happens in the physical case of N=1N=1.

Refer to caption
Figure 3: The T=0T=0 phase diagram of the γ\gamma-model for 0<γ<10<\gamma<1 as a function of NN. At smaller N<NcrN<N_{cr} the ground state is a superconductor. At larger N>NcrN>N_{cr} it is a NFL with no superconducting order. The value of NcrN_{cr} is larger than one.

The T=0T=0 phase diagram of the γ\gamma-model for N1N\geq 1 has been studied before Raghu et al. (2015); Wang et al. (2016). It was argued that for any γ<1\gamma<1 there exists a critical Ncr>1N_{cr}>1, separating a SC ground state at N<NcrN<N_{cr} and a normal, NFL ground state at N>NcrN>N_{cr} (see Fig. 3). The physical case N=1N=1 falls into the SC region. A similar result has been recently found in the study of the pairing in the SYK-type model Hauck et al. (2019).

The conventional wisdom holds that Δ(ω)=0\Delta(\omega)=0 for N>NcrN>N_{cr}, is infinitesimally small for N=NcrN=N_{cr}, and has a finite value for N<NcrN<N_{cr}, i.e., that the linearized gap equation has the solution at N=NcrN=N_{cr}, and the non-linear gap equation has the solution for N<NcrN<N_{cr}. We argue that in our case the conventional wisdom fails. Namely, we prove that the linearized equation has a solution not only for N=NcrN=N_{cr}, but also for all N<NcrN<N_{cr}, including the physical case of N=1N=1. We obtain the exact solution of the linearized gap equation, Δ(ωm)\Delta(\omega_{m}), for all N<NcrN<N_{cr} and all 0<γ10<\gamma\leq 1. This Δ(ωm)\Delta(\omega_{m}) oscillates at small ωmg¯\omega_{m}\ll{\bar{g}} as Δ(ωm)|ωm|γ/2cos(βNlog((|ωm|/g¯)γ+ϕN))\Delta(\omega_{m})\propto|\omega_{m}|^{\gamma/2}\cos\left(\beta_{N}\log((|\omega_{m}|/{\bar{g}})^{\gamma}+\phi_{N})\right), where βN\beta_{N} and ϕN\phi_{N} are particular functions of NN (for a given γ\gamma), e.g., βN(NcrN)1/2\beta_{N}\propto(N_{cr}-N)^{1/2} for NNcrN\lesssim N_{cr}. At large ωmg¯\omega_{m}\gg{\bar{g}}, Δ(ωm)\Delta(\omega_{m}) decreases as 1/|ωm|γ1/|\omega_{m}|^{\gamma} (see Fig. 11).

At small γ\gamma, when Δ(ωm)\Delta(\omega_{m}) is a smooth function of frequency, the integral equation for Δ(ωm)\Delta(\omega_{m}) can be approximated by second-order differential equation, whose solution, Δdiff(ωm)\Delta_{\text{diff}}(\omega_{m}), also exists for all NNcrN\leq N_{cr} and can be expressed analytically as a combination of two complex-conjugated hypergeometric functions. This Δdiff(ωm)\Delta_{\text{diff}}(\omega_{m}) also oscillates at small ωm\omega_{m} and decays as 1/|ωm|γ1/|\omega_{m}|^{\gamma} at large frequencies (see Fig. 9). We show that Δdiff(ωm)\Delta_{\text{diff}}(\omega_{m}) coincides with the exact Δ(ωm)\Delta(\omega_{m}) to leading order in γ\gamma.

We then analyze the non-linear gap equation. We argue that it has an infinite, discrete set of solutions, specified by integer nn, which ranges between 0 and \infty. All solutions have the same spatial gap symmetry (i.e., ss-wave, dd-wave, etc). A solution Δn(ωm)\Delta_{n}(\omega_{m}) changes sign nn times as a function of frequency. Each Δn(ωm)\Delta_{n}(\omega_{m}) tends to a finite value at ωm=0\omega_{m}=0, but the magnitude of Δn(0)\Delta_{n}(0) progressively decreases with increasing nn. At n=n=\infty, Δ(ωm)\Delta_{\infty}(\omega_{m}) is the solution of the linearized gap equation, which has an infinite number of sign changes due to cos(βNlog(|ωm|/g¯)γ+ϕN)\cos(\beta_{N}\log{(|\omega_{m}|/{\bar{g}})^{\gamma}}+\phi_{N}) oscillations running down to the smallest ωm\omega_{m}. The n=0n=0 solution yields sign-preserving Δ0(ω)\Delta_{0}(\omega).

The existence of an infinite set of Δn(ω)\Delta_{n}(\omega), each representing a local minimum of the Luttinger-Ward (LW) functional, opens up a new channel of longitudinal gap fluctuations, which cause phase slips. As long as the set is discrete, there is a single global minimum of the LW functional. In our case, it corresponds to n=0n=0. Then, at T=0T=0 the system should display a SC order. Still, the existence of the infinite set of Δn(ωm)\Delta_{n}(\omega_{m}) is a non-trivial aspect of the pairing at a QCP. In the next paper, we analyze the γ\gamma-model for 0<γ<10<\gamma<1 at a finite TT and show explicitly that for any N<NcrN<N_{cr} there exists a discrete, infinite set of onset temperatures for the pairing, Tp,nT_{p,n}, and the corresponding eigenfunctions Δn(ωm)\Delta_{n}(\omega_{m}) change sign nn times as a function of Matsubara frequency. We argue that at T0T\to 0, each Δn(ωm)\Delta_{n}(\omega_{m}) approaches the nthn-th solution of the non-linear integral gap equation at T=0T=0. We show that away from a QCP, only solutions with n<nmaxn<n_{max} remain, where nmaxn_{max} decreases with the deviation from a QCP.

We note in passing that a discrete set of solutions for the gap function at T=0T=0, with different momentum dependencies along the FS, has been detected in the analysis of Δ(𝐤F)\Delta({\bf k}_{F}) near a nematic QCP, Refs. Yang and Sondhi (2000); Klein et al. (2019) There, however, the physics is different, and the number of solutions is finite at a QCP. This is similar to our case away from a critical point.

Refer to caption
Figure 4: A phase diagram away from a QCP (left) and at a QCP (right). Away from a QCP, the gap equation has a solution for infinitesimally small Δ(ωm)\Delta(\omega_{m}) at N=NcrN=N_{cr}, and for a finite Δ(ωm)\Delta(\omega_{m}) at N<NcrN<N_{cr}, At a QCP, a solution with infinitesimally small gap function exists for all NNcrN\leq N_{cr}, and for N<NcrN<N_{cr} there is a discrete set of solutions for a finite Δn(ωm)\Delta_{n}(\omega_{m}). A gap Δn(ωm)\Delta_{n}(\omega_{m}) changes sign nn times as a function of ωm\omega_{m}.

I.2 Relevance of the exact solution of the linearized gap equation

The existence of an infinite set of solutions of the non-linear gap equation is a direct consequence of a highly unusual result that at a QCP the linearized gap equation has a solution not only at N=NcrN=N_{cr}, but also for any N<NcrN<N_{cr} (see Fig. 4). This does not away from a QCP, where the linearized gap equation as a solution only at N=NcrN=N_{cr}. The proof of the existence of the solution for infinitesimally small Δ(ωm)\Delta(\omega_{m}) for all NNcrN\leq N_{cr}, including the original N=1N=1, is, therefore, the central element of our analysis.

Refer to caption
Figure 5: The gap equation for infinitesimally small Δ(ωm)\Delta(\omega_{m}). The solution can be found separately at large and small frequencies. There is no guarantee, however, that for any NNcrN\leq N_{cr} there exists the solution, which interpolates between the two limiting forms.

The linearized gap equation can be analyzed at frequencies much smaller and much larger than g¯{\bar{g}}. In these two limits one can truncate the gap equation by keeping only leading terms either in ωm/g¯\omega_{m}/{\bar{g}} or in g¯/ωm{\bar{g}}/\omega_{m}. The truncated equation can be solved and yields Δ(ωm)1/|ωm|γ\Delta(\omega_{m})\propto 1/|\omega_{m}|^{\gamma} at large ωn\omega_{n} and Δ(ωm)|ωm|γ/2cos(βNlog((|ωm|/g¯)γ+ϕ)\Delta(\omega_{m})\propto|\omega_{m}|^{\gamma/2}\cos\left(\beta_{N}\log((|\omega_{m}|/{\bar{g}})^{\gamma}+\phi\right) at small ωm\omega_{m}, where the phase factor ϕ\phi is a free parameter. The generic task is then to verify whether by fixing ϕ\phi one can find Δ(ωm)\Delta(\omega_{m}), which smoothly interpolates between small and large ωm\omega_{m} (see Fig. 5). The verification would be straightforward if there was a finite frequency range where both forms were valid and could be made equal by fixing ϕ\phi. In our case, however, the low-frequency and the high-frequency regimes do not overlap, so the only option is to solve the full equation. We present the exact solution Δex(ωm)\Delta_{\text{ex}}(\omega_{m}) and show that at small and large frequencies it reduces to known forms. The analytic expression for Δex(ωm)\Delta_{\text{ex}}(\omega_{m}) is rather complex, but one can straightforwardly plot Δex(ωm)\Delta_{\text{ex}}(\omega_{m}) for any input parameters (see Fig. 11).

At small γ\gamma, where the interaction V(Ωm)1/|Ωm|γV(\Omega_{m})\propto 1/|\Omega_{m}|^{\gamma} is a slowly varying function of frequency, the actual, integral gap equation can be approximated by a differential equation. The latter can be solved analytically, and the solution, Δdiff(ωm)\Delta_{\text{diff}}(\omega_{m}), is a hypergeometric function of ω\omega. Using the properties of a hypergeometric function at small and large values of the argument one can explicitly verify that Δdiff(ωm)\Delta_{\text{diff}}(\omega_{m}) does interpolate between the known forms at small and large ωm\omega_{m} if one properly chooses the value of the phase ϕ\phi.

We show that Δex(ωm)\Delta_{\text{ex}}(\omega_{m}) and Δdiff(ωm)\Delta_{\text{diff}}(\omega_{m}) coincide at small γ\gamma and use the analytic form of Δdiff(ωm)\Delta_{\text{diff}}(\omega_{m}) to analyze the structure of the gap function at different NNcrN\leq N_{cr}. We demonstrate by the direct comparison that at larger γ\gamma, Δex(ωm)\Delta_{\text{ex}}(\omega_{m}) and Δdiff(ωm)\Delta_{\text{diff}}(\omega_{m}) differ qualitatively. We argue that the series for the exact Δex(ωm)\Delta_{\text{ex}}(\omega_{m}) contain non-local terms, not present for Δdiff(ωm)\Delta_{\text{diff}}(\omega_{m}). These non-local terms are negligible at small γ\gamma, but must be kept for γ=O(1)\gamma=O(1), particularly near ωm=ωmax\omega_{m}=\omega_{max}, which separates oscillating behavior at smaller ωm\omega_{m} and 1/|ωm|γ1/|\omega_{m}|^{\gamma} decay at larger ωm\omega_{m}.

I.3 Structure of the paper

The paper is organized as follows. In Sec. II we briefly review the γ\gamma-model and extend it to N1N\neq 1. We present the set of coupled Eliashberg equations for the pairing vertex Φ(ωm)\Phi(\omega_{m}) and the fermionic self-energy Σ(ωm)\Sigma(\omega_{m}) and combine them into the equations for the gap function Δ(ωm)\Delta(\omega_{m}) and the inverse quasiparticle residue Z(ωm)Z(\omega_{m}).

In Secs. III and IV we study the linearized gap equation with infinitesimally small Δ(ωm)\Delta(\omega_{m}). In Sec. III we analyze the truncated gap equation at small and large ωm\omega_{m} and obtain the solutions, valid in the corresponding limits. Here we identify NcrN_{cr} and show that the solution at small frequencies changes qualitatively between N>NcrN>N_{cr} and N<NcrN<N_{cr}. In Sec. IV we consider the limit of small γ\gamma and approximate the actual integral gap equation by second-order differential equation. We solve the differential equation for all ωm\omega_{m} and explicitly show how one can match low-frequency and high-frequency forms by fixing a single phase factor. We show that this only holds for NNcrN\leq N_{cr}, while for N>NcrN>N_{cr} the potential solution does not satisfy the condition of normalizability. In Sec. V we obtain the exact solution Δex(ωm)\Delta_{\text{ex}}(\omega_{m}) of the full linearized gap equation. We show that it exists for all N<NcrN<N_{cr}. The solution, Δex(ωm)\Delta_{\text{ex}}(\omega_{m}), oscillates at small frequencies and decays as 1/|ωm|γ1/|\omega_{m}|^{\gamma} at large |ωm||\omega_{m}|. In Sec. V.1 we analyze the structure of Δex(ωm)\Delta_{\text{ex}}(\omega_{m}) and compare it with Δdiff(ωm)\Delta_{\text{diff}}(\omega_{m}). We show that the two coincide at the smallest γ\gamma, but differ at γ=O(1)\gamma=O(1). We argue that the difference is due to the fact that Δex(ωm)\Delta_{\text{ex}}(\omega_{m}) contain non-local terms, not present in Δdiff(ωm)\Delta_{\text{diff}}(\omega_{m}). In Sec. VI we consider the non-linear gap equation. We first analyze the non-linear differential equation and argue that it has an infinite, discrete set of solutions Δdiff,n(ωm)\Delta_{{\text{diff}},n}(\omega_{m}) for any N<NcrN<N_{cr}. The index n=0,1,2,n=0,1,2,... specifies the number of times Δdiff,n(ωm)\Delta_{{\text{diff}},n}(\omega_{m}) changes sign as a function of frequency. The solution with n=0n=0 is sign-preserving. The solution nn\to\infty coincides with the solution of the linearized differential equation. We conjecture that the actual, non-linear integral gap equation also possesses an infinite, discrete set of topologically distinct solutions Δn(ωm)\Delta_{n}(\omega_{m}). Sec. VII presents the summary of our results.

I.4 Brief outline of subsequent papers

In the next paper in the series (Paper II, Ref. Wu et al. (2020)) we consider the linearized gap equation for the same range 0<γ<10<\gamma<1 at a finite TT and show that there exists an infinite discrete set of critical temperatures Tp,nT_{p,n} for the pairing instability, all within the same pairing symmetry. The corresponding eigenfunction Δn(ωm)\Delta_{n}(\omega_{m}) changes sign nn times as a function of discrete Matsubara frequency ωm=πT(2m+1)\omega_{m}=\pi T(2m+1). We argue that at T0T\to 0 these finite TT solutions become Δn(ωm)\Delta_{n}(\omega_{m}), which we find here in the T=0T=0 analysis.

For γ<1\gamma<1, the sign-preserving solution Δ0(ωm)\Delta_{0}(\omega_{m}) has the largest condensation energy and the largest Tp,0T_{p,0}. Still, the presence of an infinite set of Δn(ωm)\Delta_{n}(\omega_{m}) at T=0T=0 is not only highly unusual, but creates a new channel of ”longitudinal” gap fluctuations. In subsequent papers we will focus on the physical case N=1N=1 and extend the analysis at T=0T=0 to γ>1\gamma>1. We will show that the set Δn(ω)\Delta_{n}(\omega) becomes more dense with increasing γ\gamma and eventually becomes continuous at γ=2\gamma=2. For this special γ\gamma, all solutions of the non-linear gap equation with finite nn have equal condensation energy, and long-range superconducting order at T=0T=0 gets destroyed upon averaging over these solutions, each with its own phase. We will corroborate this by the analysis of the gap equation for the same γ\gamma at finite TT and obtain the phase diagram, shown in Fig. 2. Later, we will show Yuzbashyan et al. (b) that the system behavior is rather special for larger γ\gamma, particularly for γ>3\gamma>3.

II γ\gamma-model.

We consider itinerant fermions at the onset of a long-range order in either spin or charge channel. At the critical point, the propagator of a soft boson becomes massless and mediates singular interaction between fermions. We follow earlier works Abanov et al. (2001a, 2003); Moon and Chubukov (2010); Metlitski and Sachdev (2010b); Mross et al. (2010); Monthoux et al. (2007); Efetov et al. (2013); Metlitski et al. (2015); Raghu et al. (2015); Haslinger and Chubukov (2003); Wang et al. (2016); Lee et al. (2018) and assume that this interaction is attractive in at least one pairing channel and that a pairing boson can be treated as slow mode compared to a fermion, i.e., at a given momentum qq, typical fermionic frequency is much larger than typical bosonic frequency. This is the case for a conventional phonon-mediated superconductivity, where for qkFq\sim k_{F} a typical fermionic frequency is of order EFE_{F}, while typical bosonic frequency is of order Debye frequency ωD\omega_{D}. The ratio δE=ωD/EF\delta_{E}=\omega_{D}/E_{F} is the small parameter for Eliashberg theory of phonon-mediated superconductivity. This theory allows one to obtain a set of coupled integral equations for frequency dependent fermionic self-energy and the pairing vertex. By analogy, the theory of electronic superconductivity, mediated by soft collective bosonic excitations in spin or charge channel, is also often called Eliashberg theory. We will use this convention.

Within the Eliashberg approximation, one can explicitly integrate over the momentum component perpendicular to the Fermi surface (for a given pairing symmetry) and reduce the pairing problem to a set of coupled integral equations for frequency dependent self-energy Σ(ωm)\Sigma(\omega_{m}) and the pairing vertex Φ(ωm)\Phi(\omega_{m}) with effective frequency-dependent dimensionless interaction V(Ω)=(g¯/|Ω|)γV(\Omega)=({\bar{g}}/|\Omega|)^{\gamma}. This interaction gives rise to NFL form of the self-energy in the normal state and, simultaneously, gives rise to the pairing.

At T=0T=0, the coupled Eliashberg equations for the pairing vertex and the fermionic self-energy are, in Matsubara formalism,

Φ(ωm)\displaystyle\Phi(\omega_{m}) =\displaystyle= g¯γ2𝑑ωmΦ(ωm)Σ~2(ωm)+Φ2(ωm)1|ωmωm|γ,\displaystyle\frac{{\bar{g}}^{\gamma}}{2}\int d\omega^{\prime}_{m}\frac{\Phi(\omega^{\prime}_{m})}{\sqrt{{\tilde{\Sigma}}^{2}(\omega^{\prime}_{m})+\Phi^{2}(\omega^{\prime}_{m})}}~{}\frac{1}{|\omega_{m}-\omega^{\prime}_{m}|^{\gamma}},
Σ~(ωm)\displaystyle{\tilde{\Sigma}}(\omega_{m}) =\displaystyle= ωm+g¯γ2𝑑ωmΣ~(ωm)Σ~2(ωm)+Φ2(ωm)1|ωmωm|γ\displaystyle\omega_{m}+\frac{{\bar{g}}^{\gamma}}{2}\int d\omega^{\prime}_{m}\frac{{\tilde{\Sigma}}(\omega^{\prime}_{m})}{\sqrt{{\tilde{\Sigma}}^{2}(\omega^{\prime}_{m})+\Phi^{2}(\omega^{\prime}_{m})}}~{}\frac{1}{|\omega_{m}-\omega^{\prime}_{m}|^{\gamma}} (1)

where Σ~(ωm)=ωm+Σ(ωm){\tilde{\Sigma}}(\omega_{m})=\omega_{m}+\Sigma(\omega_{m}). In these equations, both Σ(ωm)\Sigma(\omega_{m}) and Φ(ωm)\Phi(\omega_{m}) are real functions. Observe that we define Σ(ωm)\Sigma(\omega_{m}) with the overall plus sign and without the overall factor of ii. 111In obtaining the Eliashberg equations we assumed that states away from the Fermi surface are irrelevant for the pairing and NFL behavior, and extended the integration over momenta to infinite limits. In this situation, fermionic self-energy Σ(ωm)\Sigma(\omega_{m}) is real. If the momentum integration over kkFk-k_{F} holds in finite and non-symmetric limits, Σ(ωm)\Sigma(\omega_{m}) has both real and imaginary components. The same holds in SYK-type models at a non-zero chemical potential (see Ref. Gu_2020 and references therein). In the normal state (Φ0\Phi\equiv 0),

Σ~(ωm)=ωm+ω0γ|ωm|1γsgnωm,{\tilde{\Sigma}}(\omega_{m})=\omega_{m}+\omega_{0}^{\gamma}|\omega_{m}|^{1-\gamma}\operatorname{sgn}{\omega_{m}}, (2)

where ω0=g¯/(1γ)1/γ\omega_{0}={\bar{g}}/(1-\gamma)^{1/\gamma}. At small γ\gamma, ω0=g¯e\omega_{0}={\bar{g}}e.

The superconducting gap function Δ(ωm)\Delta(\omega_{m}) is defined as a real function

Δ(ωm)=ωmΦ(ωm)Σ~(ωm)=Φ(ωm)1+Σ(ωm)/ωm.\Delta(\omega_{m})=\omega_{m}\frac{\Phi(\omega_{m})}{{\tilde{\Sigma}}(\omega_{m})}=\frac{\Phi(\omega_{m})}{1+\Sigma(\omega_{m})/\omega_{m}}. (3)

The equation for Δ(ωm)\Delta(\omega_{m}) is readily obtained from (II):

Δ(ωm)=g¯γ2𝑑ωmΔ(ωm)Δ(ωm)ωmωm(ωm)2+Δ2(ωm)1|ωmωm|γ.\Delta(\omega_{m})=\frac{{\bar{g}}^{\gamma}}{2}\int d\omega^{\prime}_{m}\frac{\Delta(\omega^{\prime}_{m})-\Delta(\omega_{m})\frac{\omega^{\prime}_{m}}{\omega_{m}}}{\sqrt{(\omega^{\prime}_{m})^{2}+\Delta^{2}(\omega^{\prime}_{m})}}~{}\frac{1}{|\omega_{m}-\omega^{\prime}_{m}|^{\gamma}}. (4)

This equation contains a single function Δ(ωm)\Delta(\omega_{m}), but for the cost that Δ(ωm)\Delta(\omega_{m}) appears also in the r.h.s., which makes Eq. (4) less convenient for the analysis than Eqs. (II).

Eqs. (II)-(4) describe color superconductivity Son (1999); *son2 and pairing in 3D (γ=0+\gamma=0_{+}, V(Ωm)log|ωm|V(\Omega_{m})\propto\log{|\omega_{m}|}), spin- and charge-mediated pairing in D=3ϵD=3-\epsilon dimensions Mross et al. (2010); Metlitski et al. (2015); Raghu et al. (2015) and superconductivity in graphene Khveshchenko (2009) (γ=O(ϵ)1\gamma=O(\epsilon)\ll 1), a 2D pairing  Altshuler et al. (1995b) with interaction peaked at 2kF2k_{F} (γ=1/4\gamma=1/4), pairing at a 2D nematic/Ising-ferromagnetic QCP Bonesteel et al. (1996); Lederer et al. (2015); Wang et al. (2001); *triplet2; *triplet3 (γ=1/3\gamma=1/3), pairing at a 2D (π,π)(\pi,\pi) SDW QCP Millis (1992); Abanov et al. (2001a, 2003); Wang and Chubukov (2013a) and an incommensurate CDW QCP Castellani et al. (1995); *ital2; *ital3; Chowdhury and Sachdev (2014b); *wang_22; *wang23 (γ=1/2\gamma=1/2), dispersionless fermions randomly interacting with an Einstein phonon Esterlis and Schmalian (2019); Wang (2020); Hauck et al. (2019) and a spin-liquid model for the cuprates Tsvelik (2017) (γ=0.7\gamma=0.7) a 2D pairing mediated by an undamped propagating boson (γ=1\gamma=1), pairing in several Fe-based superconductors Haule and Kotliar (2007) (γ=1.2\gamma=1.2) and even the strong coupling limit of phonon-mediated superconductivity for either dispersion-full Combescot (1995); Bergmann and Rainer (1973); *Bergmann2; *ad; Marsiglio et al. (1988); *Marsiglio_91; Karakozov et al. (1991) or dispersion-less Esterlis and Schmalian (2019); Wang (2020) fermions (γ=2\gamma=2). The pairing models with parameter-dependent γ\gamma have been analyzed as well (Refs. Sachdev et al., 2009; Moon and Chubukov, 2010). The case γ=0\gamma=0 describes a BCS superconductor. We list some of the model in Tables 1 - 3.

A justification of Eliashberg theory for electronically mediated superconductivity (e.g., the reasoning to neglect vertex corrections) is case specific. For Ising-nematic fluctuations (the case γ=1/3\gamma=1/3), vertex corrections are small in g¯/EF{\bar{g}}/E_{F}, which is a small parameter of the theory Metzner et al. (2003); *DellAnna2006; Maslov and Chubukov (2010); Klein et al. (2020); *avi_last_1. In other cases, e,g., in SYK models, a small parameter for Eliashberg approximation is 1/N1/N, where NN is the number of of fermionic flavors Patel and Sachdev (2019); Esterlis and Schmalian (2019); Wang (2020); Hauck et al. (2019); Chowdhury and Berg (2020). For several 2D models, the corrections to Eliashberg approximation for the self-energy in the normal state are logarithmically singular and in the absence of the pairing would change the system behavior at the smallest frequencies Abanov et al. (2003); Metlitski and Sachdev (2010a); Dalidovich and Lee (2013); *sslee_2018. One-loop logarithmic corrections just change γ\gamma but keep the model intact (Ref. Abanov et al. (2003)), but higher-order corrections go beyond the γ\gamma model (Refs. Metlitski and Sachdev (2010a); Dalidovich and Lee (2013); *sslee_2018). Here, we assume that the onset temperature for the pairing, TpT_{p}, is larger, at least numerically, than the scale at which corrections to Eliashberg approximation become relevant, and stick with the Eliashberg theory. Recent Quantum Monte Carlo (QMC) calculations for superconductivity, mediated by antiferromagnetic spin fluctuations Wang et al. (2017b) and Ising-nematic fluctuations Klein et al. (2020), have found the onset of superconductivity at a temperature, almost identical to the one in the γ\gamma model with γ=1/2\gamma=1/2 and 1/31/3, respectively, without vertex corrections. Normal state QMC calculations for the models with antiferromagnetic and Ising-nematic fluctuations also found Klein et al. (2020); *avi_last_1 very good agreement with the behavior of the γ\gamma models with γ=1/2\gamma=1/2 and 1/31/3. A good agreement has been also found between QMC calculations of superconducting TcT_{c} for electron-phonon interaction and the γ\gamma-model with γ=2\gamma=2 (Ref. Chubukov et al. (2020a)), for couplings smaller than the Fermi energy.

Table 1: Examples of pairing near quantum-critical point in D=3D=3 and D=3ϵD=3-\epsilon.
Model/Order
V(q,Ωm)V(q,\Omega_{m})
V(Ωm)=d2qV(q,Ωm)V(\Omega_{m})=\int d^{2}qV(q,\Omega_{m})
γ\gamma
Ising-nematic order,
Ising FM,
1q2+Γ|Ωm|q\frac{1}{q^{2}+\Gamma\frac{|\Omega_{m}|}{q}}
log1|Ωm|\log\frac{1}{|\Omega_{m}|}
0+0+
SDW/CDW order,
hot-spot models
1q2+Γ|Ωm|\frac{1}{q^{2}+\Gamma|\Omega_{m}|}
log1|Ωm|\log\frac{1}{|\Omega_{m}|}
0+0+
Anisotropic Ising-nematic
1q2+ϵ+Γ|Ωm||q|\frac{1}{q^{2+\epsilon}+\Gamma\frac{|\Omega_{m}|}{|q|}}
1|Ωm|ϵ3+ϵ\frac{1}{|\Omega_{m}|^{\frac{\epsilon}{3+\epsilon}}}
ϵ3+ϵ\frac{\epsilon}{3+\epsilon}
Anisotropic SDW/CDW
1q2+ϵ+Γ|Ωm|\frac{1}{q^{2+\epsilon}+\Gamma|\Omega_{m}|}
1|Ωm|ϵ2+ϵ\frac{1}{|\Omega_{m}|^{\frac{\epsilon}{2+\epsilon}}}
ϵ2+ϵ\frac{\epsilon}{2+\epsilon}
Table 2: Examples of pairing near quantum-critical point in 2D2D.
Model/Order
V(q,Ωm)V(q,\Omega_{m})
V(Ωm)=𝑑qV(q,Ωm)V(\Omega_{m})=\int dqV(q,\Omega_{m})
γ\gamma
Ising-nematic order,
Ising FM,
fermions at 1/21/2 filled Landau level
1q2+Γ|Ωm|q\frac{1}{q^{2}+\Gamma\frac{|\Omega_{m}|}{q}}
1|Ωm|2/3\frac{1}{|\Omega_{m}|^{2/3}}
2/32/3
SDW/CDW order,
hot-spot models
1q2+Γ|Ωm|\frac{1}{q^{2}+\Gamma|\Omega_{m}|}
1|Ωm|1/2\frac{1}{|\Omega_{m}|^{1/2}}
1/21/2
Undamped fermions
1q2+Ωm2\frac{1}{q^{2}+\Omega_{m}^{2}}
1|Ωm|\frac{1}{|\Omega_{m}|}
11
FeFe-based superconductors
1|Ωm|1.2\frac{1}{|\Omega_{m}|^{1.2}}
1.21.2
Table 3: Examples of pairing due to dispersionless phonons.
Model/Order
V(q,Ωm)V(q,\Omega_{m})
V(Ωm)=𝑑qV(q,Ωm)V(\Omega_{m})=\int dqV(q,\Omega_{m})
γ\gamma
Pairing by a soft
Einstein phonon
1Ωm2\frac{1}{\Omega_{m}^{2}}
1Ωm2\frac{1}{\Omega_{m}^{2}}
22
SYK-model with phonons.
Weak coupling
1|Ωm|0.6\frac{1}{|\Omega_{m}|^{0.6}}
1|Ωm|0.6\frac{1}{|\Omega_{m}|^{0.6}}
0.60.6

In this paper, we consider the set of γ\gamma-models with 0<γ<10<\gamma<1. The analysis for γ1\gamma\geq 1 requires separate consideration because of divergencies in the r.h.s. of both equations in (II) (but not in (4)) and will be presented in subsequent papers.

Notice that for any γ>0\gamma>0 the pairing interaction V(Ω)V(\Omega) decays at large frequencies as 1/|Ω|γ1/|\Omega|^{\gamma}. A simple experimentation shows that the integrals in (II) and (4) are then convergent in the ultraviolet. The only exceptions are the BCS case γ=0\gamma=0, when V(Ω)V(\Omega) is just a constant, and the case V(Ωm)log|ωm|V(\Omega_{m})\propto\log{|\omega_{m}|}, i.e., γ=0+\gamma=0_{+} (color superconductivity/pairing in 3D, Refs. Son (1999); *son2; Metlitski et al. (2015)). For these two cases one needs to set the upper cutoff of frequency integration at some Λ\Lambda to cut ultra-violet divergence. We discuss the limit γ0\gamma\to 0 in Appendix A.

The full set of Eliashberg equations for electron-mediated pairing contains also the equation describing the feedback from the pairing on the bosonic propagator. This feedback is small by δE\delta_{E} in the case of electron-phonon interaction, but is generally not small when the pairing is mediated by a collective mode because the dispersion of a collective mode may change qualitatively below TpT_{p}. The most known example of this kind is the transformation of Landau overdamped spin collective mode in the normal state to a propagating mode (often called a resonance mode) below the onset of dd-wave pairing mediated by antiferromagnetic spin fluctuations (see, e.g., Refs.Millis and Monien (1996); Eschrig (2006); Abanov et al. (2002); Abanov and Chubukov (2000)). To avoid additional complications, we do not include this feedback explicitly into our consideration. In general, the feedback from the pairing makes bosons less incoherent and can be modeled by assuming that γ\gamma moves towards a larger value as TT decreases ( e.g., from γ=1/2\gamma=1/2 to γ=1\gamma=1 for the case of antiferromagnetic fluctuations and from γ=1/3\gamma=1/3 to γ=2/3\gamma=2/3 for the case of Ising-nematic fluctuations).

The coupled equations for Φ\Phi and Σ~{\tilde{\Sigma}}, Eq. (II), describe the interplay between the two competing tendencies – one towards superconductivity, specified by Φ\Phi, and the other towards incoherent non-Fermi liquid behavior, specified by Σ~{\tilde{\Sigma}}. The competition between the two tendencies is encoded in the fact that Σ~{\tilde{\Sigma}} appears in the denominator of the equation for Φ\Phi and Φ\Phi appears in the denominator of the equation for Σ~{\tilde{\Sigma}}. In more physical terms a self-energy Σ~{\tilde{\Sigma}} is an obstacle to Cooper pairing, while if Φ\Phi is non-zero, it reduces the strength of the self-energy and moves the system back into a FL regime.

In Eq. (II) the couplings in the particle-particle and particle-hole channels have the same magnitude g¯{\bar{g}}. To study the interplay, it is convenient to have a parameter, which would increase either the tendency towards NFL or towards pairing. With this in mind, we multiply the coupling in the particle-particle channel by a factor 1/N1/N, i.e., set it to be g¯γ/N{\bar{g}}^{\gamma}/N instead of g¯γ\bar{g}^{\gamma}, and keep the coupling in the particle-hole channel intact. For N<1N<1, the tendency towards pairing is enhanced, for N>1N>1 the tendency towards NFL effectively gets larger. We will treat NN as a free parameter, but use the extension as just a way to see the relevant physics more clearly, as our ultimate goal is to understand system behavior in the physical case of N=1N=1. We note in passing that the extension to integer N>1N>1 can be formalized by extending the original model to matrix SU(N)SU(N) model Raghu et al. (2015).

The modified equations for Φ(ωm)\Phi(\omega_{m}) and Σ~(ωm){\tilde{\Sigma}}(\omega_{m}) are

Φ(ωm)=g¯γ2N𝑑ωmΦ(ωm)Σ~2(ωm)+Φ2(ωm)1|ωmωm|γ,\displaystyle\Phi(\omega_{m})=\frac{{\bar{g}}^{\gamma}}{2N}\int d\omega^{\prime}_{m}\frac{\Phi(\omega^{\prime}_{m})}{\sqrt{{\tilde{\Sigma}}^{2}(\omega^{\prime}_{m})+\Phi^{2}(\omega^{\prime}_{m})}}~{}\frac{1}{|\omega_{m}-\omega^{\prime}_{m}|^{\gamma}}, (5)
Σ~(ωm)=ωm+g¯γ2𝑑ωmΣ~(ωm)Σ~2(ωm)+Φ2(ωm)1|ωmωm|γ,\displaystyle{\tilde{\Sigma}}(\omega_{m})=\omega_{m}+\frac{{\bar{g}}^{\gamma}}{2}\int d\omega^{\prime}_{m}\frac{{\tilde{\Sigma}}(\omega^{\prime}_{m})}{\sqrt{{\tilde{\Sigma}}^{2}(\omega^{\prime}_{m})+\Phi^{2}(\omega^{\prime}_{m})}}~{}\frac{1}{|\omega_{m}-\omega^{\prime}_{m}|^{\gamma}}, (6)

and the equation for Δ(ωm)\Delta(\omega_{m}) becomes

Δ(ωm)=g¯γ2N𝑑ωmΔ(ωm)NΔ(ωm)ωmωm(ωm)2+Δ2(ωm)1|ωmωm|γ.\Delta(\omega_{m})=\frac{{\bar{g}}^{\gamma}}{2N}\int d\omega^{\prime}_{m}\frac{\Delta(\omega^{\prime}_{m})-N\Delta(\omega_{m})\frac{\omega^{\prime}_{m}}{\omega_{m}}}{\sqrt{(\omega^{\prime}_{m})^{2}+\Delta^{2}(\omega^{\prime}_{m})}}~{}\frac{1}{|\omega_{m}-\omega^{\prime}_{m}|^{\gamma}}. (7)

Below we will occasionally refer to the equation for Φ(ωm)\Phi(\omega_{m}) as the gap equation, notwithstanding that the gap equation is given by Eqs. (7). Indeed, once we know Φ(ωm)\Phi(\omega_{m}) and Σ~(ωm){\tilde{\Sigma}}(\omega_{m}), we also know Δ(ωm)=Φ(ωm)ωm/Σ~(ωm)\Delta(\omega_{m})=\Phi(\omega_{m})\omega_{m}/{\tilde{\Sigma}}(\omega_{m}).

We will be searching for normalized solutions for Φ(ωm)\Phi(\omega_{m}) and Δ(ωm)\Delta(\omega_{m}). Physically, the normalizability of a solution follows from the requirement that the Free energy of a superconductor should be free from divergencies. The Free energy of a superconductor, FscF_{sc}, can be obtained by either applying Hubbard-Stratonovich formalism  Yuzbashyan et al. (a) or by using a generic Luttinger-Ward-Eliashberg expression Luttinger and Ward (1960); Eliashberg (1960), and explicitly integrating over momentum, approximating the density of states by its value at the Fermi level Haslinger and Chubukov (2003); Wu et al. (2019b). The result is

FscN0=𝑑ωmωm2ωm2+Δ2(ωm)g¯γ4𝑑ωm𝑑ωmωmωm+1NΔ(ωm)Δ(ωm)ωm2+Δ2(ωm)(ωm)2+Δ2(ωm)1|ωmωm|γ.\frac{F_{sc}}{N_{0}}=-\int d\omega_{m}\frac{\omega^{2}_{m}}{\sqrt{\omega^{2}_{m}+\Delta^{2}(\omega_{m})}}-\frac{{\bar{g}}^{\gamma}}{4}\int d\omega_{m}d\omega^{\prime}_{m}\frac{\omega_{m}\omega^{\prime}_{m}+\frac{1}{N}\Delta(\omega_{m})\Delta(\omega^{\prime}_{m})}{\sqrt{\omega^{2}_{m}+\Delta^{2}(\omega_{m})}\sqrt{(\omega^{\prime}_{m})^{2}+\Delta^{2}(\omega^{\prime}_{m})}}\frac{1}{|\omega_{m}-\omega^{\prime}_{m}|^{\gamma}}. (8)

The gap equation (7) is obtained from the condition δFsc/δΔ(ωm)=0\delta F_{sc}/\delta\Delta(\omega_{m})=0.

The condensation energy EcE_{c} is the difference between FscF_{sc}, with Δ(ωm)\Delta(\omega_{m}) satisfying the Eliashberg equation (7), and FnF_{n} (the free energy for Δ=0\Delta=0). Using Eq. (8) and the gap equation (7) we obtain Wada (1964); Bardeen and Stephen (1964); Yuzbashyan et al. (a)

EcN0=𝑑ωm|ωm|1+D2(ωm)(1+D2(ωm)1)2g¯γ4dωmdωm1+D2(ωm)1+D2(ωm)\displaystyle\frac{E_{c}}{N_{0}}=-\int d\omega_{m}\frac{|\omega_{m}|}{\sqrt{1+D^{2}(\omega_{m})}}\left(\sqrt{1+D^{2}(\omega_{m})}-1\right)^{2}-\frac{{\bar{g}}^{\gamma}}{4}\int\frac{d\omega_{m}d\omega^{\prime}_{m}}{\sqrt{1+D^{2}(\omega_{m})}\sqrt{1+D^{2}(\omega^{\prime}_{m})}}
×(1+D2(ωm)1+D2(ωm))2(1|ωmωm|γ1|ωm+ωm|γ)\displaystyle\times\left(\sqrt{1+D^{2}(\omega_{m})}-\sqrt{1+D^{2}(\omega^{\prime}_{m})}\right)^{2}\left(\frac{1}{|\omega_{m}-\omega^{\prime}_{m}|^{\gamma}}-\frac{1}{|\omega_{m}+\omega^{\prime}_{m}|^{\gamma}}\right) (9)

where D(ωm)=Δ(ωm)/ωmD(\omega_{m})=\Delta(\omega_{m})/\omega_{m}. Both terms in the r.h.s. of (9) are negative, hence, the condensation energy is negative for any solution of the gap equation Δ(ωm)\Delta(\omega_{m}). Note that the factor NN is not present in (9).

To understand whether or not the ground state at T=0T=0 is a NFL state or a superconducting state, we first consider infinitesimally small Φ(ωm)\Phi(\omega_{m}). The corresponding equation is obtained by neglecting Φ\Phi in the denominator of Eq. (5) and using (2) for Σ~{\tilde{\Sigma}}:

NΦ(ωm)=1γ2𝑑ωmΦ(ωm)|ωm|1γ|ωmωm|γ11+(|ωm|ω0)γN\Phi(\omega_{m})=\frac{1-\gamma}{2}\int d\omega^{\prime}_{m}\frac{\Phi(\omega^{\prime}_{m})}{|\omega^{\prime}_{m}|^{1-\gamma}|\omega_{m}-\omega^{\prime}_{m}|^{\gamma}}\frac{1}{1+\left(\frac{|\omega^{\prime}_{m}|}{\omega_{0}}\right)^{\gamma}} (10)

This is an equation for an eigenfunction of a linear operator, in which NN plays a role of the eigenvalue. Observe that the fermion-boson coupling g¯{\bar{g}} appears only in the last term in the denominator, via ω0g¯\omega_{0}\propto{\bar{g}}. Without this term, the r.h.s. of (10) is marginal by power counting (the total exponent in the denominator is 1γ+γ=11-\gamma+\gamma=1). Then, once we rescale frequency to ω¯m=ωm/ω0{\bar{\omega}}_{m}=\omega_{m}/\omega_{0}, the equation for Φ(ω¯m)\Phi({\bar{\omega}}_{m}) becomes fully universal:

Φ(ω¯m)=1γ2N𝑑ω¯mΦ(ω¯m)|ω¯m|1γ|ω¯mω¯m|γ11+|ω¯m|γ\Phi({\bar{\omega}}_{m})=\frac{1-\gamma}{2N}\int d{\bar{\omega}}^{\prime}_{m}\frac{\Phi({\bar{\omega}}^{\prime}_{m})}{|{\bar{\omega}}^{\prime}_{m}|^{1-\gamma}|{\bar{\omega}}_{m}-{\bar{\omega}}^{\prime}_{m}|^{\gamma}}\frac{1}{1+|{\bar{\omega}}^{\prime}_{m}|^{\gamma}} (11)

The same is true for the linearized equation for Δ(ωm)\Delta(\omega_{m}):

Δ(ωm)=g¯γ2N𝑑ωmΔ(ωm)NΔ(ωm)ωmωm|ωm|1|ωmωm|γ.\Delta(\omega_{m})=\frac{{\bar{g}}^{\gamma}}{2N}\int d\omega^{\prime}_{m}\frac{\Delta(\omega^{\prime}_{m})-N\Delta(\omega_{m})\frac{\omega^{\prime}_{m}}{\omega_{m}}}{|\omega^{\prime}_{m}|}~{}\frac{1}{|\omega_{m}-\omega^{\prime}_{m}|^{\gamma}}. (12)

Using g¯γ=ω0γ(1γ){\bar{g}}^{\gamma}=\omega^{\gamma}_{0}(1-\gamma) and introducing again ω¯m=ωm/ω0{\bar{\omega}}_{m}=\omega_{m}/\omega_{0}, we re-express (12) as

Δ(ω¯m)=1γ2N𝑑ω¯mΔ(ω¯m)NΔ(ω¯m)ω¯mω¯m|ω¯m|1|ω¯mω¯m|γ.\Delta({\bar{\omega}}_{m})=\frac{1-\gamma}{2N}\int d{\bar{\omega}}^{\prime}_{m}\frac{\Delta({\bar{\omega}}^{\prime}_{m})-N\Delta({\bar{\omega}}_{m})\frac{{\bar{\omega}}^{\prime}_{m}}{{\bar{\omega}}_{m}}}{|{\bar{\omega}}^{\prime}_{m}|}~{}\frac{1}{|{\bar{\omega}}_{m}-{\bar{\omega}}^{\prime}_{m}|^{\gamma}}. (13)

For infinitesimal Φ\Phi and Δ\Delta, Δ(ωm)=Φ(ωm)/(1+(ω0/|ωm|)γ)\Delta(\omega_{m})=\Phi(\omega_{m})/\left(1+(\omega_{0}/|\omega_{m}|)^{\gamma}\right), or, equivalently, Δ(ω¯m)=Φ(ω¯m)/(1+|ω¯m|γ)\Delta({\bar{\omega}}_{m})=\Phi({\bar{\omega}}_{m})/(1+|{\bar{\omega}}_{m}|^{-\gamma}).

A conventional wisdom, borrowed from BCS theory, would imply that Eqs. (11) and (13) must have solutions at a single critical NcrN_{cr}, separating NFL and superconducting states, assuming that such critical NcrN_{cr} exists. By the same logic, the full non-linear gap equation has no solution at N>NcrN>N_{cr}, while at N<NcrN<N_{cr}, the solution Φ(ωm)\Phi(\omega_{m}) (or Δ(ωm)\Delta(\omega_{m})) has a finite magnitude, which increases with NcrNN_{cr}-N.

We show below that the situation in the γ\gamma-model is different. Specifically, we show that critical NcrN_{cr} exists, but the linearized gap equation has a solution not only at N=NcrN=N_{cr}, but for all N<NcrN<N_{cr}. Furthermore, N=NcrN=N_{cr} turns out to be a multi-critical point of the γ\gamma-model (for 0<γ<10<\gamma<1), below which there exists an infinite number of solutions of the full non-linear gap equation, Δn(ωm)\Delta_{n}(\omega_{m}). The magnitude of the nnth solution initially increases as eAn/NcrNe^{-A_{n}/\sqrt{N_{cr}-N}}, where the factor AnA_{n} depends on the number nn of the solution.

In the next four sections we discuss the linearized gap equation, Eqs. (11) and (13). We return to the non-linear gap equation in Sec. VI.

To verify that a potential solution of the linearized gap equation is normalizable, we will need to analyze the Free energy (8) order Δ2\Delta^{2}. Expanding in (8) we obtain

Fsc\displaystyle F_{sc} =\displaystyle= Fnorm+N02NJ(Δ,N)\displaystyle F_{norm}+\frac{N_{0}}{2N}J(\Delta,N) (14)
J(Δ,N)\displaystyle J(\Delta,N) =\displaystyle= N𝑑ω¯mΔ2(ω¯m)(1+|ω¯m|γ)|ω¯m|1+γ1γ2𝑑ωm𝑑ωmΔ(ωm)Δ(ωm)|ω¯m||ω¯m|1|ω¯mω¯m|γ.\displaystyle N\int d{\bar{\omega}}_{m}\frac{\Delta^{2}(\bar{\omega}_{m})(1+|\bar{\omega}_{m}|^{\gamma})}{|\bar{\omega}_{m}|^{1+\gamma}}-\frac{1-\gamma}{2}\int d\omega_{m}d\omega^{\prime}_{m}\frac{\Delta(\omega_{m})\Delta(\omega^{\prime}_{m})}{|\bar{\omega}_{m}||\bar{\omega}^{\prime}_{m}|}\frac{1}{|\bar{\omega}_{m}-\bar{\omega}^{\prime}_{m}|^{\gamma}}.

The linearized gap equation (13) is obtained by varying this FscF_{sc} over Δ(ωm)\Delta(\omega_{m}). A solution is normalizable if the variation δJ(Δ,N)/δN\delta J(\Delta,N)/\delta N is finite, i.e., if

𝑑ω¯mΔ2(ω¯m)(1+|ω¯m|γ)|ω¯m|1+γ\int d\bar{\omega}_{m}\frac{\Delta^{2}(\bar{\omega}_{m})(1+|\bar{\omega}_{m}|^{\gamma})}{|\bar{\omega}_{m}|^{1+\gamma}} (15)

is non-divergent. In terms of Φ(ω¯m)\Phi(\bar{\omega}_{m}), the same integral is

𝑑ω¯mΦ2(ω¯m)|ω¯m|1γ(1+|ω¯m|γ)\int d\bar{\omega}_{m}\frac{\Phi^{2}(\bar{\omega}_{m})}{|\bar{\omega}_{m}|^{1-\gamma}(1+|\bar{\omega}_{m}|^{\gamma})} (16)

III The limits of small and large ωm\omega_{m}.

We begin with the analysis of the truncated gap equation at small and large frequencies. The analysis can be done most straightforwardly for Eq. (11) for the pairing vertex. At large |ω¯m|1|\bar{\omega}_{m}|\gg 1, we can pull out the external ω¯m\bar{\omega}_{m} from the integral in the r.h.s. of (10) and obtain

Φ(ω¯m)=1|ω¯m|γ1γN0O(|ω¯m|)𝑑ω¯mΦ(ω¯m)|ω¯m|1γ11+|ω¯m|γ.\Phi({\bar{\omega}}_{m})=\frac{1}{|\bar{\omega}_{m}|^{\gamma}}\frac{1-\gamma}{N}\int_{0}^{O(|\bar{\omega}_{m}|)}d{\bar{\omega}}^{\prime}_{m}\frac{\Phi({\bar{\omega}}^{\prime}_{m})}{|{\bar{\omega}}^{\prime}_{m}|^{1-\gamma}}\frac{1}{1+|{\bar{\omega}}^{\prime}_{m}|^{\gamma}}. (17)

Substituting Φ(ω¯m)1/|ω¯m|γ\Phi({\bar{\omega}}^{\prime}_{m})\propto 1/|\bar{\omega}_{m}|^{\gamma} into the r.h.s of (17) we find that the integral converges at |ω¯m|=O(1)|{\bar{\omega}}^{\prime}_{m}|=O(1). This shows that pulling out ω¯m\bar{\omega}_{m} from the integral is justified when |ω¯m|1|\bar{\omega}_{m}|\gg 1. The outcome is that at large frequencies

Φ(ω¯m)=C|ω¯m|γ\Phi({\bar{\omega}}_{m})=\frac{C_{\infty}}{|\bar{\omega}_{m}|^{\gamma}} (18)

In this limit, Φ(ω¯m)=Δ(ω¯m)\Phi({\bar{\omega}}_{m})=\Delta(\bar{\omega}_{m}), hence, we also have

Δ(ω¯m)=C|ω¯m|γ\Delta({\bar{\omega}}_{m})=\frac{C_{\infty}}{|\bar{\omega}_{m}|^{\gamma}} (19)

In the opposite limit of small ω¯m\bar{\omega}_{m} we assume and then verify that one can truncate (11) by replacing 1/(1+|ω¯m|γ)1/(1+|\bar{\omega}^{\prime}_{m}|^{\gamma}) in the r.h.s. by the upper cutoff at |ω¯m|=O(1)|{\bar{\omega}}^{\prime}_{m}|=O(1). The precise value of the cutoff frequency will play no role, and we set the cutoff at |ω¯m|=1|{\bar{\omega}}^{\prime}_{m}|=1. The equation for the pairing vertex then becomes

Φ(ω¯m)=1γ2N11𝑑ω¯mΦ(ω¯m)|ω¯m|1γ|ω¯mω¯m|γ\Phi(\bar{\omega}_{m})=\frac{1-\gamma}{2N}\int_{-1}^{1}d\bar{\omega}^{\prime}_{m}\frac{\Phi(\bar{\omega}^{\prime}_{m})}{|\bar{\omega}^{\prime}_{m}|^{1-\gamma}|\bar{\omega}_{m}-\bar{\omega}^{\prime}_{m}|^{\gamma}} (20)

Below we analyze this equation for different NN.

III.1 Large NN.

We first consider the limit of large NN. The effective coupling constant in (20) scales as 1/N1/N, hence, the solution with a non-zero Φ(ωm)\Phi(\omega_{m}) emerges only if the smallness of the coupling is compensated by a large value of the frequency integral in the r.h.s. of (20). This is what happens in a BCS superconductor (the case γ=0\gamma=0). There, the pairing kernel scales as 1/|ω¯m|1/|\bar{\omega}_{m}|, Φ(ω¯m)=Φ\Phi(\bar{\omega}_{m})=\Phi is independent of the running fermionic frequency, and the integral 11𝑑ω¯mΦ/|ω¯m|\int_{-1}^{1}d\bar{\omega}^{\prime}_{m}\Phi/|\bar{\omega}^{\prime}_{m}| is logarithmically singular. The logarithm compensates for the smallness of the coupling 1/N1/N, and superconductivity emerges already for arbitrary weak attraction. A way to see this is to compute the pairing vertex in the presence of an infinitesimally small initial Φ0\Phi_{0}. At T=0T=0 this has to be done at a non-zero total incoming bosonic frequency Ωtot=Ω¯tot/ω0\Omega_{tot}={\bar{\Omega}}_{tot}/\omega_{0} to avoid divergencies. The result for a BCS superconductor is well known: to logarithmic accuracy

Φ(Ω¯tot)=Φ0(1+1Nlog1|Ω¯tot|+1N2log21|Ω¯tot|+)=Φ011Nlog1|Ω¯tot|\Phi({\bar{\Omega}}_{tot})=\Phi_{0}\left(1+\frac{1}{N}\log{\frac{1}{|{\bar{\Omega}}_{tot}|}}+\frac{1}{N^{2}}\log^{2}{\frac{1}{|{\bar{\Omega}}_{tot}|}}+...\right)=\frac{\Phi_{0}}{1-\frac{1}{N}\log{\frac{1}{|{\bar{\Omega}}_{tot}|}}} (21)

The ratio Φ(Ω¯tot)/Φ0\Phi({\bar{\Omega}}_{tot})/\Phi_{0} (the pairing susceptibility) diverges at |Ωtot|=ω0eN|\Omega_{tot}|=\omega_{0}e^{-N} and becomes negative at smaller |Ωtot||\Omega_{tot}|, indicating that the normal state is unstable towards pairing. (In a more accurate description, the pole in Φ(Ωtot)\Phi(\Omega_{tot}) moves from the lower to the upper half-plane of complex frequency Abrikosov et al. (1965)).

For a non-zero γ\gamma, the pairing kernel is the function of both internal ω¯m{\bar{\omega}}^{\prime}_{m} and external ω¯m{\bar{\omega}}_{m}

K(ω¯,ω¯m)=1|ω¯m|1γ|ω¯mω¯m|γK({\bar{\omega}},{\bar{\omega}}^{\prime}_{m})=\frac{1}{|\bar{\omega}^{\prime}_{m}|^{1-\gamma}|\bar{\omega}_{m}-\bar{\omega}^{\prime}_{m}|^{\gamma}} (22)

If we set the external ω¯m\bar{\omega}_{m} to zero, we find that K(0,ω¯m)=1/|ω¯m|K(0,\bar{\omega}^{\prime}_{m})=1/|\bar{\omega}^{\prime}_{m}| is marginal, like in BCS theory. Then, if we add Φ0\Phi_{0} and compute Φ(ω¯m)\Phi(\bar{\omega}_{m}) perturbatively, the series will be logarithmical, like in the BCS case. In distinction to BCS, however, each logarithmical integral 𝑑ω¯m/|ω¯m|\int d\bar{\omega}^{\prime}_{m}/|\bar{\omega}^{\prime}_{m}| runs between the upper cutoff at |ω¯m|=1|\bar{\omega}^{\prime}_{m}|=1 and the lower cutoff at |ω¯m||ω¯m||\bar{\omega}^{\prime}_{m}|\sim|\bar{\omega}_{m}|. Because the lower cutoff is finite at a finite ω¯m\bar{\omega}_{m}, we can safely set Ωtot=0\Omega_{tot}=0. Summing up logarithmical series we then obtain

Φ(ω¯m)=Φ0(1+1γNlog1|ω¯m|+(1γ)22N2log21|ω¯m|+.)=Φ0(1|ω¯m|)1γN\Phi(\bar{\omega}_{m})=\Phi_{0}\left(1+\frac{1-\gamma}{N}\log{\frac{1}{|\bar{\omega}_{m}|}}+\frac{(1-\gamma)^{2}}{2N^{2}}\log^{2}{\frac{1}{|\bar{\omega}_{m}|}}+....\right)=\Phi_{0}\left(\frac{1}{|\bar{\omega}_{m}|}\right)^{\frac{1-\gamma}{N}} (23)

We see that the pairing susceptibility remains finite and positive for all finite ωm\omega_{m}, even when Ωtot=0\Omega_{tot}=0. Re-doing calculations at a finite Ωtot\Omega_{tot} we find the same result as in (23), but with |ω¯m||\bar{\omega}_{m}| replaced by max(|ω¯m|,Ω¯tot)(|\bar{\omega}_{m}|,{\bar{\Omega}}_{tot}). The implication is that at a finite γ\gamma, summing up logarithms does not give rise to the divergence of the pairing susceptibility, i.e., within a logarithmic approximation, the system at T=0T=0 remains in the normal state.

The difference between logarithmic approximation at γ=0\gamma=0 and at a finite γ\gamma can be also understood by expressing the flow of Φ(ω¯m,Ω¯tot)\Phi(\bar{\omega}_{m},{\bar{\Omega}}_{tot}) under external Φ0\Phi_{0} in terms of RG-type differential equation. In BCS theory one can re-sum logarithmical ladder series by selecting a cross-section in the middle with the smallest running frequency ω¯m\bar{\omega}_{m} and integrate over frequencies larger than ω¯m\bar{\omega}_{m} in the cross-sections on both sides of the selected one. One such integration gives Φ(ω¯m)\Phi(\bar{\omega}_{m}), another gives the pairing susceptibility Φ(ω¯m)/Φ0\Phi(\bar{\omega}_{m})/\Phi_{0}. Combining and differentiating over L=log1/Ω¯totL=\log{1/{\bar{\Omega}}_{tot}}, we obtain the RG equation dΦ(L)/dL=Φ2(L)/NΦ0d\Phi(L)/dL=\Phi^{2}(L)/N\Phi_{0}, with the boundary condition Φ(0)=Φ0\Phi(0)=\Phi_{0}. The solution of this equation is Eq. (21). For a non-zero γ\gamma, the logarithmical integral in a given cross-section is cut not by an external bosonic Ω¯tot{\bar{\Omega}}_{tot}, but by the external fermionic frequency in the neighboring cross-section. A simple experimentation shows that in this case the corresponding RG equation is dΦ(L)/dL=Φ(L)/Nd\Phi(L)/dL=\Phi(L)/N, where now L=log(1/|ω¯m|)L=\log(1/|\bar{\omega}_{m}|). The solution of this equation is Eq. (23).

We now go beyond perturbation theory and analyze Eq. (20) without the Φ0\Phi_{0} term. Our first observation is that Φ(ω¯m)(1/|ω¯m|)(1γ)/N\Phi(\bar{\omega}_{m})\propto\left(1/|\bar{\omega}_{m}|\right)^{(1-\gamma)/N}, which we found by summing up logarithms, does satisfy Eq. (20) at |ω¯m|1|\bar{\omega}_{m}|\ll 1. Indeed, substituting this form into (20) we find that it is satisfied, because to leading order in 1/N1/N,

0dx|x|(1γ)(N+1)/N1|1x|γ=N1γ\int_{0}^{\infty}\frac{dx}{|x|^{(1-\gamma)(N+1)/N}}\frac{1}{|1-x|^{\gamma}}=\frac{N}{1-\gamma} (24)

We next observe that there is another possibility to compensate for the 1/N1/N smallness of the coupling constant in (20), by choosing Φ(ω¯m)(1/|ω¯m|)γ(1γ)/N\Phi(\bar{\omega}_{m})\propto\left(1/|\bar{\omega}_{m}|\right)^{\gamma-(1-\gamma)/N}, such that the integral over ω¯m\bar{\omega}^{\prime}_{m} in the r.h.s. of (20) almost diverges at small ω¯m\bar{\omega}^{\prime}_{m}. Indeed, substituting this form into (20) we find that the equation is satisfied because to leading order in 1/N1/N,

0dx|x|1(1γ)/N1|1x|γ=N1γ\int_{0}^{\infty}\frac{dx}{|x|^{1-(1-\gamma)/N}}\frac{1}{|1-x|^{\gamma}}=\frac{N}{1-\gamma} (25)

Note that the factor NN now comes from small |x|1|x|\ll 1. This implies that this solution could not be obtained within a conventional logarithmic approximation, or, equivalently, from the RG equation, as the latter assumes that the logarithms, which sum up into the anomalous power-law form, come from internal frequencies, which are much larger than the external one.

The solution of (20) at large NN is then the combination of the two power-law forms

Φ(ω¯m)=CA|ω¯m|(1γ)/N+CB|ω¯m|γ(1γ)/N.\Phi(\bar{\omega}_{m})=\frac{C_{A}}{|\bar{\omega}_{m}|^{(1-\gamma)/N}}+\frac{C_{B}}{|\bar{\omega}_{m}|^{\gamma-(1-\gamma)/N}}. (26)

The overall factor doesn’t matter because Φ(ωm)\Phi(\omega_{m}) is defined up to a constant factor, but the ratio CA/CBC_{A}/C_{B} is a free parameter at this moment.

We now argue that while both terms in (26) satisfy (20), only one component represents the normalized solution and should be kept. Indeed, substituting Φ(ω¯m)\Phi(\bar{\omega}_{m}) into (16) and using the fact find that for large NN, (1γ)/Nγ/2(1-\gamma)/N\ll\gamma/2, we find that the integral in (16) is finite for the first term in (26) but diverges for the second one. We then have to drop this term and keep only the CAC_{A} term in (26).

The full gap equation (11) has the solution if Φ(ω¯m)=CA/|ω¯m|(1γ)/N\Phi(\bar{\omega}_{m})=C_{A}/|\bar{\omega}_{m}|^{(1-\gamma)/N} at small ω¯m\bar{\omega}_{m} and Φ(ω¯m)=C/|ω¯m|γ\Phi(\bar{\omega}_{m})=C_{\infty}/|\bar{\omega}_{m}|^{\gamma} at large ω¯m\bar{\omega}_{m} can be smoothly connected (see Fig. 5). We show below that these two limiting forms cannot be connected, i.e., the linearized equation for the pairing vertex doesn’t have a solution.

Refer to caption
Refer to caption
Figure 6: The function ϵ(b)\epsilon(b) from Eq. 28 for real bb (left panel) and imaginary bb (right panel) for γ=0.6\gamma=0.6. The solution of the truncated gap equation at small ω¯\bar{\omega} exists if the equation ϵ(b)=N\epsilon(b)=N has a solution.

III.2 Arbitrary NN.

The analysis of the truncated equation for the pairing vertex can be straightforwardly extended to arbitrary NN. Like before, we use the fact that the kernel in (20) is marginal and search for power-law solution of (20) in the form Φ(ω¯m)1/|ω¯m|γ(1/2+b)\Phi(\bar{\omega}_{m})\propto 1/|\bar{\omega}_{m}|^{\gamma(1/2+b)}, where bb needs to be obtained self-consistently (we pulled out γ\gamma in the exponent for future convenience). Substituting into (20) and evaluating the integral we find that bb is the solution of

ϵb=N\epsilon_{b}=N (27)

where

ϵb=1γ2Γ(γ/2γb)Γ(γ/2+γb)Γ(γ)(1+cos(πγb)cos(πγ/2))\epsilon_{b}=\frac{1-\gamma}{2}\frac{\Gamma(\gamma/2-\gamma b)\Gamma(\gamma/2+\gamma b)}{\Gamma(\gamma)}\left(1+\frac{\cos(\pi\gamma b)}{\cos(\pi\gamma/2)}\right) (28)

The integral in the r.h.s. of (20) is evaluated using the identities

0dyya(1+y)γ\displaystyle\int_{0}^{\infty}\frac{dy}{y^{a}(1+y)^{\gamma}} =\displaystyle= Γ(1a)Γ(γ+a1)Γ(γ)\displaystyle\frac{\Gamma(1-a)\Gamma(\gamma+a-1)}{\Gamma(\gamma)}
01dyya(1y)γ\displaystyle\int_{0}^{1}\frac{dy}{y^{a}(1-y)^{\gamma}} =\displaystyle= Γ(1a)Γ(1γ)Γ(2aγ)\displaystyle\frac{\Gamma(1-a)\Gamma(1-\gamma)}{\Gamma(2-a-\gamma)}
1dyya(y1)γ\displaystyle\int_{1}^{\infty}\frac{dy}{y^{a}(y-1)^{\gamma}} =\displaystyle= Γ(a+γ1)Γ(1γ)Γ(a)\displaystyle\frac{\Gamma(a+\gamma-1)\Gamma(1-\gamma)}{\Gamma(a)} (29)

We plot ϵb\epsilon_{b} in Fig. 6a. At small γ\gamma,

ϵb4γ114b2\epsilon_{b}\approx\frac{4}{\gamma}\frac{1}{1-4b^{2}} (30)

We see that ϵb\epsilon_{b} is an even function of bb, i.e., if bb is a solution, then b-b is also a solution. This implies that

Φ(ω¯m)=Φ(z)=CAz1/2b+CBz1/2+b,\Phi(\bar{\omega}_{m})=\Phi(z)=\frac{C_{A}}{z^{1/2-b}}+\frac{C_{B}}{z^{1/2+b}}, (31)

where

z=|ω¯m|γ=(|ωm|ω0)γz=|\bar{\omega}_{m}|^{\gamma}=\left(\frac{|\omega_{m}|}{\omega_{0}}\right)^{\gamma} (32)

At large NN, b1/2(1γ)/(Nγ)b\approx 1/2-(1-\gamma)/(N\gamma) and (31) coincides with (26). As NN increases, bb gets smaller. As long as it remains positive, the second term in the r.h.s. of (31) has to be dropped as it does not satisfy the normalization condition (the integral in (16) diverges for Φ(ω¯m)1/(z1/2+b\Phi(\bar{\omega}_{m})\propto 1/(z^{1/2+b} ). Then at small zz,

Φ(z)=Cz1/2b,Δ(z)=Cz1/2+b.\Phi(z)=\frac{C}{z^{1/2-b}},~{}~{}~{}\Delta(z)=Cz^{1/2+b}. (33)

This is very similar to the case of large NN.

Refer to caption
Figure 7: The critical NcrN_{cr} (for which Eq. (28) has a solution for b=0b=0) as a function of γ\gamma. At small γ\gamma, Ncr4/γN_{cr}\approx 4/\gamma. At γ1\gamma\to 1, Ncr1N_{cr}\to 1.

We next observe that for any 0<γ10<\gamma\leq 1, there exists a critical NcrN_{cr}, for which b=0b=0. It is given by

Ncr=ϵ0\displaystyle N_{cr}=\epsilon_{0} =\displaystyle= π2(1γ)sinπ2(1γ)πΓ(γ)(1cosπγ2)1Γ2(1γ/2)=1γ2Γ2(γ/2)Γ(γ)(1+1cos(πγ/2)),\displaystyle\frac{\frac{\pi}{2}(1-\gamma)}{\sin{\frac{\pi}{2}(1-\gamma)}}\frac{\pi}{\Gamma(\gamma)}\frac{\left(1-\cos{\frac{\pi\gamma}{2}}\right)^{-1}}{\Gamma^{2}\left(1-\gamma/2\right)}=\frac{1-\gamma}{2}\frac{\Gamma^{2}(\gamma/2)}{\Gamma(\gamma)}\left(1+\frac{1}{\cos(\pi\gamma/2)}\right), (34)

We plot NcrN_{cr} vs γ\gamma in Fig. 7

We see that Ncr>1N_{cr}>1 for all γ<1\gamma<1. At small γ\gamma, we have from (30), Ncr4/γN_{cr}\approx 4/\gamma. At γ1\gamma\to 1, Ncr1N_{cr}\to 1.

Right at N=NcrN=N_{cr}, the two terms in (31) become equal, i.e., one solution of (20) is Φ(z)1/z\Phi(z)\propto 1/\sqrt{z}. On a more careful look, we find that Φ(z)logz/z\Phi(z)\propto\log{z}/\sqrt{z} is also the solution. Indeed, substituting this form into (20) and using 𝑑y|y|γ/21|y1|γlog|y|=0\int dy|y|^{\gamma/2-1}|y-1|^{-\gamma}\log|y|=0 and ((1γ)/2Ncr)𝑑y|y|γ/21|y1|γ=1((1-\gamma)/2N_{cr})\int dy|y|^{\gamma/2-1}|y-1|^{-\gamma}=1, we find that this equation is satisfied, i.e.,

1γ2Ncr𝑑ωmlog|ωm||ωm|1γ/2|ωmωm|γ=log|ωm||ωm|γ/2\frac{1-\gamma}{2N_{cr}}\int d\omega^{\prime}_{m}\frac{\log{|\omega^{\prime}_{m}|}}{|\omega^{\prime}_{m}|^{1-\gamma/2}|\omega_{m}-\omega^{\prime}_{m}|^{\gamma}}=\frac{\log{|\omega_{m}|}}{|\omega_{m}|^{\gamma/2}} (35)

The full solution of (20) at N=NcrN=N_{cr} is then

Φ(z)=1z(CAlogz+CB)\Phi(z)=\frac{1}{\sqrt{z}}\left(C_{A}\log{z}+C_{B}\right) (36)

or, equivalently,

Φ(z)=Cz(logz+ϕ),Δ(z)=Cz(logz+ϕ).\Phi(z)=\frac{C}{\sqrt{z}}\left(\log{z}+\phi\right),~{}~{}~{}\Delta(z)=C\sqrt{z}\left(\log{z}+\phi\right). (37)

This Φ(z)\Phi(z) contains two free parameters: CC and ϕ\phi. The overall factor CC is irrelevant for the solution of the linearized equation, but ϕ\phi is a free parameter, which we can vary in a hope that at a particular value of ϕ\phi, Φ(z)\Phi(z) and Δ(z)\Delta(z) interpolate smoothly between small zz and large zz limits. This gives an indication that N=NcrN=N_{cr} may be the onset for superconductivity.

We now move to N<NcrN<N_{cr}. There is no solution of (27) for real bb, consistent with the conventional wisdom that the linearized gap equation has no solution inside the superconducting phase. However, it turns out that the solution of (20) still exists, but for imaginary b=iβNb=i\beta_{N}, i.e., with complex exponents 1/2±iβN1/2\pm i\beta_{N}. Indeed, for a generic b=iβb=i\beta,

ϵiβ\displaystyle\epsilon_{i\beta} =\displaystyle= π2(1γ)sinπ2(1γ)πΓ(γ)(coshπγβcosπγ/2)1Γ(1γ/2(1+2iβ))Γ(1γ/2(12iβ))\displaystyle\frac{\frac{\pi}{2}(1-\gamma)}{\sin{\frac{\pi}{2}(1-\gamma)}}\frac{\pi}{\Gamma(\gamma)}\frac{\left(\cosh{\pi\gamma\beta}-\cos{\pi\gamma/2}\right)^{-1}}{\Gamma\left(1-\gamma/2(1+2i\beta)\right)\Gamma\left(1-\gamma/2(1-2i\beta)\right)} (38)
=\displaystyle= 1γ2|Γ(γ/2(1+2iβ))|2Γ(γ)(1+cosh(πγβ)cos(πγ/2))\displaystyle\frac{1-\gamma}{2}\frac{|\Gamma(\gamma/2(1+2i\beta))|^{2}}{\Gamma(\gamma)}\left(1+\frac{\cosh(\pi\gamma\beta)}{\cos(\pi\gamma/2)}\right)

is real, even, and monotonically decreases with increasing β\beta from its maximum value ϵ0=Ncr>1\epsilon_{0}=N_{cr}>1 to 0, i.e., Eq. (27) has a solution for N<NcrN<N_{cr} at some non-zero β=βN\beta=\beta_{N}. We plot ϵiβ\epsilon_{i\beta} in Fig. 6b. At small γ\gamma,

ϵiβ4γ11+4β2\epsilon_{i\beta}\approx\frac{4}{\gamma}\frac{1}{1+4\beta^{2}} (39)

and

βN=12NcrNN\beta_{N}=\frac{1}{2}\sqrt{\frac{N_{cr}-N}{N}} (40)

up to corrections of order γ\gamma. Solutions with complex exponents have been recently found in several other physics problems Liu et al. (2011); Cortez et al. (2011); Kim et al. (2019). For superconductivity, the solution with complex exponents has been found in Ref. Abanov et al. (2001a) for the γ\gamma model with γ=1/2\gamma=1/2.

At NN slightly below NcrN_{cr}, βNNcrN\beta_{N}\propto\sqrt{N_{cr}-N} for all γ<1\gamma<1. For N=1N=1, βN=1γ1/2(1((15π2)/24)γ+O(γ2))\beta_{N=1}\approx\gamma^{-1/2}(1-((15-\pi^{2})/24)\gamma+O(\gamma^{2})) for small γ\gamma, βN=11.268\beta_{N=1}\approx 1.268 for γ=1/2\gamma=1/2, and βN=10.792\beta_{N=1}\to 0.792 for γ1\gamma\to 1. For N=O(1)N=O(1), but N1N\neq 1, the behavior at γ<1\gamma<1 is very similar, but for γ1\gamma\to 1, βN\beta_{N} diverges as βN0.561(1/N)1/(1γ)\beta_{N}\approx 0.561(1/N)^{1/(1-\gamma)}.

Combining the contributions with βN\beta_{N} and βN-\beta_{N}, we obtain

Φ(z)=C21z(eiϕziβN+eiϕziβN)=Czcos(βNlogz+ϕ)\Phi(z)=\frac{C}{2}\frac{1}{\sqrt{z}}\left(\frac{e^{i\phi}}{z^{i\beta_{N}}}+\frac{e^{-i\phi}}{z^{-i\beta_{N}}}\right)=\frac{C}{\sqrt{z}}\cos{\left(\beta_{N}\log{z}+\phi\right)} (41)

where ϕ\phi is a free parameter. We see that Φ(z)\Phi(z) oscillates on a logarithmical scale down to the lowest frequencies. Such an oscillating solution could not be obtained perturbatively, starting from Φ(z)=Φ0\Phi(z)=\Phi_{0}, because within perturbation theory Φ(z)\Phi(z) remains of the same sign as Φ0\Phi_{0}.

We also see that Φ¯=Φ(z)zei(βNlogz+ϕ)+c.c{\bar{\Phi}}=\Phi(z)\sqrt{z}\propto e^{i(\beta_{N}\log{z}+\phi)}+c.c has the form of a wave function of a free fermion if we associate x=logzx=\log{z} with the coordinate. In terms of xx, the integral in (16) is expressed as 𝑑xΦ¯(x)2\int dx{\bar{\Phi}}(x)^{2}, and is nothing but the norm of the eigenfunction of continuum spectrum. The norm can be made finite by adding infinitesimal eδ|x|e^{-\delta|x|} to the integral, after which it converges. The addition of eδ|x|e^{-\delta|x|} also makes the integral in (16) convergent for Φ(z)\Phi(z) at N=NcrN=N_{cr}, however for N>NcrN>N_{cr}, the integral in (16) still diverges for one term in (31). We discuss the normalizability for N>NcrN>N_{cr} and NNcrN\leq N_{cr} in more detail in Appendix C.

The gap function Δ(z)\Delta(z) behaves as

Δ(z)=Czcos(βNlogz+ϕ)\Delta(z)=C\sqrt{z}\cos{\left(\beta_{N}\log{z}+\phi\right)} (42)

We see that the solution of the truncated equation for the pairing vertex at small zz again has two free parameters. In Eq. (42) these are the overall factor CC and the phase ϕ\phi. The overall factor is irrelevant, but the phase ϕ\phi is a free parameter, which we can vary to verify whether at some particular ϕ\phi, Δ(z)\Delta(z) interpolates between (42) and (19).

We emphasize that a free parameter for Δ(z)\Delta(z) at small zz (or, equivalently, for Φ(z)\Phi(z)) exists both at N=NcrN=N_{cr} and N<NcrN<N_{cr}. By conventional wisdom, one would expect that the linearized gap equation has the solution only for one value of NN, which in our case is expectedly N=NcrN=N_{cr} (see Fig. 4). However, the presence of a free parameter for all NNcrN\leq N_{cr} hints that our case may go against the conventional wisdom.

Matching the limiting forms of an integral equation is a non-trivial procedure as in general one should find a finite frequency range where the two forms coincide. This cannot be implemented in our case because the regions, where the integral equation can be truncated, do not overlap. Because of this, we will be searching for the solution of (11) without specifically looking at the limits of small and large ω¯m\bar{\omega}_{m}.

IV The case γ1\gamma\ll 1: The linearized gap equation as the differential equation.

We first consider the limit γ1\gamma\ll 1, when the pairing interaction (g¯/|Ωm|)γ({\bar{g}}/|\Omega_{m}|)^{\gamma} is a shallow function of frequency, and reduce the linearized integral gap equation to the differential equation, for which we obtain the exact analytical solution, Δdiff(ωm)\Delta_{{\text{diff}}}(\omega_{m}), which for convenience of notations we present below as a function of zz, defined in (32). We show that a generic normalized solution exists for NNcrN\leq N_{cr}. At small ωm\omega_{m}, Δdiff(z)\Delta_{{\text{diff}}}(z) coincides with the solution of the truncated gap equation. It oscillates as a function of βNlogz+ϕ\beta_{N}\log{z}+\phi. For arbitrary ϕ\phi, Δdiff(z)\Delta_{{\text{diff}}}(z) tends to a constant at large zz, instead of decaying as 1/z1/z. However, for a certain value of ϕ\phi the constant term cancels out, and the gap function has the correct asymptotic behavior at high frequencies. In other words, by fixing the phase in the oscillations of Δdiff(z)\Delta_{{\text{diff}}}(z) at small zz one recovers the correct form of sign-preserving Δdiff(z)\Delta_{{\text{diff}}}(z) at large frequencies.

To obtain the differential equation, we return to the linearized equation for the pairing vertex Φ(ωm)\Phi(\omega_{m}), Eq. (11), and use the fact that for small γ\gamma, the integral in the r.h.s. of (10) comes from internal ωm\omega^{\prime}_{m}, which are either substantially larger or substantially smaller than external ωm\omega_{m}. We then split the integral over ωm\omega^{\prime}_{m} into two parts: in one we approximate |ωmωm||\omega_{m}-\omega^{\prime}_{m}| by |ωm||\omega^{\prime}_{m}|, in the other by |ωm||\omega_{m}|. The equation for Φ(ωm)=Φ(z)\Phi(\omega_{m})=\Phi(z) then simplifies to

Φ(z)=1γNγ[z𝑑yΦ(y)y(1+y)+1z0z𝑑yΦ(y)1+y].\Phi(z)=\frac{1-\gamma}{N\gamma}\left[\int^{\infty}_{z}dy\frac{\Phi(y)}{y(1+y)}+\frac{1}{z}\int_{0}^{z}dy\frac{\Phi(y)}{1+y}\right]. (43)

Differentiating this equation twice over zz and replacing Φ(z)\Phi(z) by Δ(z)=Φ(z)z/(1+z)\Delta(z)=\Phi(z)z/(1+z), we obtain second order differential gap equation

(Δdiff(z)(1+z))′′=Ncr4NΔdiff(z)z2,(\Delta_{{\text{diff}}}(z)(1+z))^{{}^{\prime\prime}}=-\frac{N_{cr}}{4N}\frac{\Delta_{{\text{diff}}}(z)}{z^{2}}, (44)

where ()′′=d2()/dz2(...)^{{}^{\prime\prime}}=d^{2}(...)/dz^{2} and we used the fact that for small γ\gamma, Ncr4/γN_{cr}\approx 4/\gamma. This Δdiff(z)\Delta_{{\text{diff}}}(z) has to be real and satisfy the boundary conditions at small and large zz. At large zz we have from (19),

limzΔdiff(z)=C/z,\qquad\lim\limits_{z\rightarrow\infty}\Delta_{{\text{diff}}}(z)=C_{\infty}/z, (45)

At small zz we have from (33), (37) and (42)

limz0Δdiff(z)={CAz1/2+b,for N>NcrCz1/2(logz+ϕ),for N=NcrCz1/2cos(βNlogz+ϕ),for N<Ncr\lim\limits_{z\rightarrow 0}\Delta_{\text{diff}}(z)=\left\{\begin{array}[]{ll}C_{A}z^{1/2+b},&\quad\mbox{for $N>N_{cr}$}\\ Cz^{1/2}\left(\log{z}+\phi\right),&\quad\mbox{for $N=N_{cr}$}\\ Cz^{1/2}\cos\left(\beta_{N}\log z+\phi\right),&\quad\mbox{for $N<N_{cr}$}\end{array}\right. (46)

We recall that Δdiff(z)\Delta_{{\text{diff}}}(z), which satisfies the boundary condition (46), is normalizable in the sense that the corresponding condensation energy is finite.

A similar differential gap equation has been obtained in Ref. Wang et al. (2017a) for γ1\gamma\ll 1 and NNcrN\sim N_{cr}. These authors, however, set by hand a UW cutoff at some z=O(1)z=O(1). In our case, the differential equation (44) is valid for all zz and the solution must recover 1/z1/z behavior at high frequencies, see (45).

Before we proceed with the analysis of Eqs. (44), (45), and (46), a remark is in order. The integral equation (43) is “non-local” in the sense that the integrals in the r.h.s. are determined by all yy, not only yzy\approx z. The differential equation (44), on the other hand, is local – both r.h.s. and l.h.s of (44) contain Δdiff(z)\Delta_{\text{diff}}(z) with the same zz. However, the boundary conditions (45) and (46) imply that once we fix the asymptotic form of Δdiff(z)\Delta_{\text{diff}}(z) at z0z\rightarrow 0, we also determine the constant CC_{\infty} in the asymptotic form at zz\rightarrow\infty. Let us take zz large enough such that the asymptotic form of Δdiff(z)\Delta_{\text{diff}}(z) holds. From Eq. (44) we obtain for such zz, C=Ncr4N0Δdiff(y)𝑑y/yC_{\infty}=\frac{N_{cr}}{4N}\int_{0}^{\infty}\Delta_{\text{diff}}(y)dy/y. One can verify that the main contribution to this integral comes from yzy\ll z, i.e., the prefactor in Δdiff(z)\Delta_{\text{diff}}(z) for zz\to\infty is determined by the form of Δdiff(z)\Delta_{\text{diff}}(z) for much smaller zz, roughly by z1z\leq 1, for which Δdiff(z)\Delta_{\text{diff}}(z) is close to its form at small zz. In this sense the non-locality of the initial integral equation (44) reflects itself in the boundary conditions (45), (46) for the local differential gap equation. We also emphasize that (44) is a second order differential equation, hence, Δdiff(z)\Delta_{\text{diff}}(z) is fully determined by the two boundary conditions (45) and (46)

We now analyze Eqs. (44), (45), (46) separately for N<NcrN<N_{cr}, N=NcrN=N_{cr}, and N>NcrN>N_{cr}. We will show that the solution exists for all NNcrN\leq N_{cr}.

IV.1 𝐍<𝐍𝐜𝐫{\bf{N<N_{cr}}}

We use Eq. (40) and re-express Ncr/(4N)N_{cr}/(4N) as Ncr/(4N)=βN2+1/4N_{cr}/(4N)=\beta^{2}_{N}+1/4. Eq. (44) then becomes

(Δdiff(z)(1+z))′′=(βN2+14)Δdiff(z)z2,(\Delta_{{\text{diff}}}(z)(1+z))^{{}^{\prime\prime}}=-\left(\beta^{2}_{N}+\frac{1}{4}\right)\frac{\Delta_{{\text{diff}}}(z)}{z^{2}}, (47)

At z1z\ll 1 (i.e., at ωω0\omega\ll\omega_{0}) Eq. (44) simplifies to

(Δdiff(z))′′=(βN2+14)Δdiff(z)z2,(\Delta_{{\text{diff}}}(z))^{{}^{\prime\prime}}=-\left(\beta^{2}_{N}+\frac{1}{4}\right)\frac{\Delta_{{\text{diff}}}(z)}{z^{2}}, (48)

The solution of this equation is the combination of two power-law functions Δdiff(z)z1/2±iβN\Delta_{{\text{diff}}}(z)\propto z^{-1/2\pm i\beta_{N}}. Combining the two, we obtain the expected result

Δdiff(z)=Cz1/2cos(βNlogz+ϕ)\Delta_{{\text{diff}}}(z)=Cz^{1/2}\cos\left(\beta_{N}\log{z}+\phi\right) (49)

At z1z\gg 1, we have

(Δdiff(z)z)′′=(βN2+14)Δdiff(z)z2(\Delta_{{\text{diff}}}(z)z)^{{}^{\prime\prime}}=-\left(\beta^{2}_{N}+\frac{1}{4}\right)\frac{\Delta_{{\text{diff}}}(z)}{z^{2}} (50)

The solution of (50) is

Δdiff(z)=1z[A1J1(4βN2+1z)+A2Y1(4βN2+1z)],\Delta_{{\text{diff}}}(z)=\frac{1}{\sqrt{z}}\left[A_{1}J_{1}\left(\sqrt{\frac{4\beta^{2}_{N}+1}{z}}\right)+A_{2}Y_{1}\left(\sqrt{\frac{4\beta^{2}_{N}+1}{z}}\right)\right], (51)

where J1J_{1} and Y1Y_{1} are Bessel and Neumann functions. At small value of the argument (or large zz) J1(p)pJ_{1}(p)\approx p and Y1(p)1/pY_{1}(p)\approx 1/p. Substituting into (51) we find that to satisfy the boundary condition (45) we must set A2=0A_{2}=0. Then

Δdiff(z)=A1zJ1(4βN2+1z).\Delta_{{\text{diff}}}(z)=\frac{A_{1}}{\sqrt{z}}J_{1}\left(\sqrt{\frac{4\beta^{2}_{N}+1}{z}}\right). (52)

A similar result has been obtained in Ref. Wang et al. (2017a).

Refer to caption
Figure 8: The solution of the differential gap equation, Eq. (53) for arbitrary ϕ\phi (left panel) and for particular ϕ=ϕN\phi=\phi_{N} given by Eq. (55) (right panel). In the last case Δdiff(z)\Delta_{\text{diff}}(z) satisfies both boundary conditions at small and at large zz.

At arbitrary zz, the solution of (47) is expressed via hypergeometric function

Δdiff(z)=Cz×Re(eiϕziβNF2[12+iβN,32+iβN,1+2iβN,z]1)\Delta_{{\text{diff}}}(z)=C\sqrt{z}\times Re\left(e^{i\phi}z^{i\beta_{N}}{{}_{2}}F{{}_{1}}\left[\frac{1}{2}+i\beta_{N},\frac{3}{2}+i\beta_{N},1+2i\beta_{N},-z\right]\right) (53)

At small zz, F2[,z]11{{}_{2}}F{{}_{1}}\left[...,-z\right]\approx 1, and Δdiff(z)\Delta_{{\text{diff}}}(z) is the same as in (49). At z1z\gg 1, we use large zz asymptotic of the hypergeometric function

F2[1/2+iβN,3/2+iβN,1+2βN,z]1=\displaystyle{{}_{2}}F{{}_{1}}\left[1/2+i\beta_{N},3/2+i\beta_{N},1+2\beta_{N},-z\right]=
z1/2iβN[Γ(1+2iβN)Γ(3/2+iβN)Γ(1/2+iβN)(1+(14+βN2)logzz)+O(1z)]\displaystyle z^{-1/2-i\beta_{N}}\left[\frac{\Gamma(1+2i\beta_{N})}{\Gamma(3/2+i\beta_{N})\Gamma(1/2+i\beta_{N})}\left(1+\left(\frac{1}{4}+\beta^{2}_{N}\right)\frac{\log{z}}{z}\right)+O\left(\frac{1}{z}\right)\right] (54)

and obtain that for arbitrary ϕ\phi, Δdiff(z)\Delta_{{\text{diff}}}(z) tends to a constant, which is inconsistent with the boundary condition (45) (see Fig. 8 left panel). However, for a particular ϕ=ϕN\phi=\phi_{N}, where

ϕN=arctanLL,whereL=Γ(1+2iβN)Γ(1/2+iβN)Γ(3/2+iβN),\phi_{N}=\arctan{\frac{\Re L}{\Im L}},~{}~{}\text{where}~{}L=\frac{\Gamma(1+2i\beta_{N})}{\Gamma(1/2+i\beta_{N})\Gamma(3/2+i\beta_{N})}, (55)

the constant term in Δ(z)\Delta(z) vanishes (along with logz/z\log{z}/z correction), and Δdiff(z)\Delta_{{\text{diff}}}(z) does scale as 1/z1/z at large zz, consistent with (45) (see Fig. 8 right panel).

Refer to caption
Refer to caption
Figure 9: The solution of the differential gap equation, Eq. (53) for various βN\beta_{N}. For βN1\beta_{N}\ll 1 (panel a) Δdiff(z)\Delta_{{\text{diff}}}(z) oscillates as a function of logz\log{z} up to ze1/βNz\sim e^{-1/\beta_{N}} and decays as 1/z1/z at z>1z>1. For βN=O(1)\beta_{N}=O(1) (panel b) logarithmic oscillations extend to z=O(1)z=O(1), and at larger zz, Δdiff(z)1/z\Delta_{{\text{diff}}}(z)\propto 1/z. For βN1\beta_{N}\gg 1 (panel c) logarithmic oscillations persist up to z=O(1)z=O(1), and at larger zz Δdiff(z)\Delta_{{\text{diff}}}(z) again oscillates as Δdiff(z)cos(3π/42βN/z)/z1/4\Delta_{{\text{diff}}}(z)\propto\cos(3\pi/4-2\beta_{N}/\sqrt{z})/z^{1/4}. These oscillations persist up to z(βN)21z\sim(\beta_{N})^{2}\gg 1. At larger zz, Δdiff(z)1/z\Delta_{{\text{diff}}}(z)\propto 1/z. Panel d – A closer look at oscillations at z1z\ll 1 and 1zβN21\ll z\ll\beta^{2}_{N} for large βN\beta_{N} (small γ\gamma).

At small βN\beta_{N}, the solution of (55) is ϕN=π/20.77259βN\phi_{N}=\pi/2-0.77259\beta_{N}. At βN=1\beta_{N}=1, ϕN=π/20.93251\phi_{N}=\pi/2-0.93251, and at large βN\beta_{N}, ϕN3π/4βNlog4\phi_{N}\approx 3\pi/4-\beta_{N}\log{4}. Once ϕ=ϕN\phi=\phi_{N} is fixed, one can also express the prefactor A1A_{1} in (51) in terms of CC in (49), e.g., A1=C(πβN)1/2A_{1}=C(\pi\beta_{N})^{1/2} for βN1\beta_{N}\gg 1.

Analyzing Eq. (53) further, we note that both the location and the width of the crossover between small zz and large zz behavior of Δdiff(z)\Delta_{{\text{diff}}}(z) depend on NN (Fig. 9). For NNcrN\leq N_{cr}, βN(NcrN)1/2\beta_{N}\propto(N_{cr}-N)^{1/2} is small. In this situation Δdiff(z)\Delta_{{\text{diff}}}(z) oscillates as a function of logz\log{z} up to ze1/βNz\sim e^{-1/\beta_{N}}, then monotonically increases up to z0.2z\sim 0.2, passes through a maximum, and decays as 1/z1/z at larger frequencies (Fig. 9a). For NN such that βN=O(1)\beta_{N}=O(1), logarithmic oscillations extend to z=O(1)z=O(1), and at larger zz, Δdiff(z)1/z\Delta_{{\text{diff}}}(z)\propto 1/z (Fig. 9b). For N=O(1)N=O(1), βN0.5(Ncr/N)1/2\beta_{N}\approx 0.5(N_{cr}/N)^{1/2} is large because Ncr4/γ1N_{cr}\approx 4/\gamma\gg 1. In this situation, logarithmic oscillations of Δ(z)=Czcos(βNlogz/4+3π/4)\Delta(z)=C\sqrt{z}\cos(\beta_{N}\log{z/4}+3\pi/4), persist up to z=O(1)z=O(1), and at larger zz there is a range 1<z<βN21<z<\beta^{2}_{N}, where Δdiff(z)\Delta_{{\text{diff}}}(z) again oscillates as Δdiff(z)Ccos(3π/42βN/z)/z1/4\Delta_{{\text{diff}}}(z)\approx C\cos(3\pi/4-2\beta_{N}/\sqrt{z})/z^{1/4}. These oscillations persist up to z(βN)21/γz\sim(\beta_{N})^{2}\sim{1/\gamma}, i.e., up to ω=ωmaxg¯e|logγ|/γ\omega=\omega_{max}\sim{\bar{g}}e^{|\log\gamma|/\gamma}. At larger zz, Δdiff(z)1/z\Delta_{{\text{diff}}}(z)\propto 1/z (Fig.9c). Oscillations at small zz are best seen if we use the logarithmical variable

x=logz=log(|ωm|g¯)γx=\log{z}=\log{\left(\frac{|\omega_{m}|}{{\bar{g}}}\right)^{\gamma}} (56)

while oscillations at z>1z>1 are best seen if we plot Δdiff(z)\Delta_{\text{diff}}(z) as a function of 1/z1/\sqrt{z}. We present both plots in Fig. 9(d).

The existence of a large intermediate range 1<z<βN21<z<\beta^{2}_{N} for βN21\beta^{2}_{N}\gg 1 can be inferred already from Eq. (51). Indeed, in this range the argument of the Bessel function y=2βN/zy=2\beta_{N}/\sqrt{z} is large, J1(y)y1/2cos(y3π/4)J_{1}(y)\propto y^{-1/2}\cos(y-3\pi/4), and Δ(z)\Delta(z) displays an oscillating behavior with the period set by 1/z1/\sqrt{z} rather than by logz\log{z}.

The scale zβN2z\sim\beta^{2}_{N} also shows up in the expansion of Δdiff(z)\Delta_{{\text{diff}}}(z) in both 1/z1/z and zz. Expanding in 1/z1/z, we find

Δdiff(z)1z(1βN22z+O(1z2))\Delta_{{\text{diff}}}(z)\propto\frac{1}{z}\left(1-\frac{\beta^{2}_{N}}{2z}+O\left(\frac{1}{z^{2}}\right)\right) (57)

The expansion in zz yields

Δdiff(z)Cz(1+z)3/4cos(3π4+βNQ(z))f(zβN2),\displaystyle\Delta_{{\text{diff}}}(z)\approx C\frac{\sqrt{z}}{(1+z)^{3/4}}\cos{\left(\frac{3\pi}{4}+\beta_{N}Q(z)\right)}f\left(\frac{z}{\beta^{2}_{N}}\right), (58)

where f(0)=1f(0)=1 and

Q(z)=logz4z2+3z2165z348+19z4256+,Q(z)=\log{\frac{z}{4}}-\frac{z}{2}+\frac{3z^{2}}{16}-\frac{5z^{3}}{48}+\frac{19z^{4}}{256}+..., (59)

We see that oscillations extend to z>1z>1. At 1zβN21\ll z\ll\beta_{N}^{2}, the form of Q(z)Q(z) is determined by by comparison with (51). We have in this range Q(z)2/zQ(z)\approx-2/\sqrt{z}. Then βNQ(z)\beta_{N}Q(z) becomes of order one at zβN2z\sim\beta^{2}_{N}, where the argument of f(z/βN2)f(z/\beta^{2}_{N}) also becomes of order one. This sets zβN2z\sim\beta^{2}_{N} as the scale where Eqs. (58) and (57) match.

The scale zβN2z\sim\beta^{2}_{N} (or, equivalently, ωmωmaxg(c/γ)1/γ\omega_{m}\sim\omega_{max}\sim g(c/\gamma)^{1/\gamma}, c=O(1)c=O(1)) is also the scale at which the boundary condition at zz\to\infty becomes relevant and the phase ϕ\phi gets locked at ϕ=ϕN\phi=\phi_{N}. To see this, we substituted Φdiff(z)=Δdiff(z)(1+z)/z\Phi_{\text{diff}}(z)=\Delta_{\text{diff}}(z)(1+z)/z back into the r.h.s. of (43) and evaluated the integrals. We found that Φdiff(z)z1/2cos(βNlogz+ϕ)\Phi_{\text{diff}}(z)\propto z^{-1/2}\cos{(\beta_{N}\log{z}+\phi)} at z<1z<1 and Φdiff(z)z1/4cos(2βN/z+ϕ)\Phi_{\text{diff}}(z)\propto z^{-1/4}\cos{(2\beta_{N}/\sqrt{z}+\phi)} at 1<z<βN21<z<\beta^{2}_{N} are reproduced for any ϕ\phi, and the corresponding integrals are confined to internal yzy\sim z in (43) (for 1<z<βN21<z<\beta^{2}_{N} the integrals are expressed via Fresnel C-function, which have to be expanded to appropriate order). However, for z>βN2z>\beta^{2}_{N}, Eq. (43) is reproduced only if we set ϕ=ϕN\phi=\phi_{N}.

Refer to caption
Figure 10: The gap function Δdiff(z)\Delta_{\text{diff}}(z) for N=Ncr0N=N_{cr}-0 (βN0\beta_{N}\to 0). The function is sign-preserving. It scales as logzz\log{z}\sqrt{z} at small zz, passes through a maximum at z0.2z\sim 0.2, and decreases as 1/z1/z at z>1z>1.

IV.2 𝐍=𝐍𝐜𝐫{\bf{N=N_{cr}}}

We now analyze the form of Δdiff(z)\Delta_{\text{diff}}(z) at N=NcrN=N_{cr}, where, we expect, the normalized solution for Δdiff\Delta_{\text{diff}} first to appear. This analysis requires the same extra care as we exercised in Sec. III.2, when we analyzed the truncated gap equation. Namely, we need to take the limit βN0\beta_{N}\to 0 rather than just set βN=0\beta_{N}=0. To take the limit properly, we re-express Eq. (53) as

Δdiff(z)=C1S(βN,z)+C2S(βN,z)\Delta_{{\text{diff}}}(z)=C_{1}S(\beta_{N},z)+C_{2}S(-\beta_{N},z) (60)

where

S(βN,z)=ziβN1/2F2[1/2+iβN,3/2+iβN,1+2iβN,z]1,S(\beta_{N},z)=z^{i\beta_{N}-1/2}{{}_{2}}F{{}_{1}}[1/2+i\beta_{N},3/2+i\beta_{N},1+2i\beta_{N},-z], (61)

and C1C_{1} and C2C_{2} are two independent free parameters. Expanding to linear order in βN\beta_{N}, we obtain

Δdiff(z)=C1S(0,z)+C2S(0,z)\Delta_{{\text{diff}}}(z)=C^{*}_{1}S(0,z)+C^{*}_{2}S^{\prime}(0,z) (62)

where the derivative is with respect to iβNi\beta_{N}. As long as βN\beta_{N} is non-zero, C1C^{*}_{1} and C2C^{*}_{2} are two independent parameters. The functions S(0,z)S(0,z) and S(0,z)S^{\prime}(0,z) are expressed in terms of Meijer G-function and tend to finite values at large zz, with subleading term of order 1/z1/z. Then Δdiff(z)=a+b/z\Delta_{{\text{diff}}}(z)=a+b/z at large zz. A constant aa vanishes once we choose C2/C1=1.294C^{*}_{2}/C^{*}_{1}=1.294. For this particular C2/C1C^{*}_{2}/C^{*}_{1}, Δdiff(z)\Delta_{{\text{diff}}}(z) from (60) satisfies the boundary condition at zz\to\infty. This implies that the solution does indeed exist at N=Ncr0N=N_{cr}-0. At small zz, S(0,z)zS(0,z)\approx\sqrt{z}, S(0,z)logzzS^{\prime}(0,z)\approx\log{z}\sqrt{z}, hence, Δ(z)logzz\Delta(z)\propto\log{z}\sqrt{z}, consistent with (46). Computing the subleading terms and using C2/C1=1.294C^{*}_{2}/C^{*}_{1}=1.294, we obtain at z<1z<1,

Δdiff(z)=Cz(13z4)(logz2.165z4+O(z2)).\Delta_{{\text{diff}}}(z)=C\sqrt{z}\left(1-\frac{3z}{4}\right)\left(\log{\frac{z}{2.165}}-\frac{z}{4}+O(z^{2})\right). (63)

In Fig. 10 we show Δdiff(z)\Delta_{{\text{diff}}}(z) over the whole range of zz. We see that Δdiff(z)\Delta_{{\text{diff}}}(z) does not oscillate. It monotonically increases with zz at small zz, passes through a maximum at z0.2z\sim 0.2, and then decreases as 1/z1/z. We will study the consequences of this behavior in Sec. VI, where we analyze the non-linear gap equation. A similar behavior has been obtained in Ref. Wang et al. (2017a). These authors, however, put a hard UV cutoff Δ(z)=0\Delta^{\prime}(z)=0 at the maximum (at z=0.2z=0.2 in our case), and only analyzed the range z<0.2z<0.2.

IV.3 𝐍>𝐍𝐜𝐫{\bf{N>N_{cr}}}

At N>NcrN>N_{cr}, the exponent bN=((NNcr)/4N)1/2b_{N}=((N-N_{cr})/4N)^{1/2} is a real number, and the solution of (50) is

Δdiff(z)=C1S[bN,z]+C2S[bN,z]\Delta_{{\text{diff}}}(z)=C_{1}S[b_{N},z]+C_{2}S[-b_{N},z] (64)

where

S[bB,z]=xbN+1/2F2[12+bN,32+bN,1+2bN,z]1.S[b_{B},z]=x^{b_{N}+1/2}{{}_{2}}F{{}_{1}}\left[\frac{1}{2}+b_{N},\frac{3}{2}+b_{N},1+2b_{N},-z\right]. (65)

The boundary conditions at z1z\gg 1 and z1z\ll 1, Eqs. (45) and (46), set two conditions on the prefactors:

C2=0andC2C1=Γ(3/2bN)Γ(1/2bN)Γ(1+2bN)Γ(3/2+bN)Γ(1/2+bN)Γ(12bN).C_{2}=0~{}~{}{\text{and}}~{}~{}\frac{C_{2}}{C_{1}}=-\frac{\Gamma(3/2-b_{N})\Gamma(1/2-b_{N})\Gamma(1+2b_{N})}{\Gamma(3/2+b_{N})\Gamma(1/2+b_{N})\Gamma(1-2b_{N})}. (66)

These two conditions are incompatible for any bNb_{N}, hence, there is no normalized solution of the gap equation for N>NcrN>N_{cr}.

IV.4 Larger γ\gamma

At larger γ\gamma, the approximations used to reduce the integral equation for the pairing vertex to Eq. (43) are not justified. If we formally extend (43) to γ1\gamma\leq 1, we still obtain Eq. (44), but with Ncrdiff=4(1γ)/γN^{{\text{diff}}}_{cr}=4(1-\gamma)/\gamma instead of the actual NcrN_{cr} given by (34). At small γ\gamma, NcrdiffNcr4/γN^{{\text{diff}}}_{cr}\approx N_{cr}\approx 4/\gamma, but for larger γ\gamma, they differ. The difference becomes crucial for γ1\gamma\to 1, where NcrdiffN^{{\text{diff}}}_{cr} tends to zero, while the actual NcrN_{cr} approaches 11. This can be partly corrected by (i) expressing the differential equation in terms of βN\beta_{N}, as in (47), and using the correct expression for βN\beta_{N} in terms of γ\gamma and NN, and (ii) modifying the derivation of the differential equation by adding the contribution from internal ωω\omega^{\prime}\approx\omega, as we show in Appendix (B). With this additional contribution, the differential equation for NNcrN\leq N_{cr} remains the same as Eq. (47), but zz is rescaled to z¯=z/d{\bar{z}}=z/d, where d=4(1γ)/(Nγ(4βN2+1))d=4(1-\gamma)/(N\gamma(4\beta^{2}_{N}+1)). For small γ\gamma, d1d\approx 1, but for γ1\gamma\to 1 and N1N\to 1, d=(1γ)/(βN2+1/4)d=(1-\gamma)/(\beta^{2}_{N}+1/4). In this limit z=(|ωm|/g)(1γ)z=(|\omega_{m}|/g)(1-\gamma) vanishes, but z¯(|ωm|/g)(βN2+1/4){\bar{z}}\approx(|\omega_{m}|/g)(\beta^{2}_{N}+1/4) remains finite.

Still, the assumption that the approximation of the actual integral gap equation by the differential one can be rigorously justified only for small γ\gamma, when the pairing interaction is a weak function of frequency. At larger γ=O(1)\gamma=O(1) one must analyze the actual, integral gap equation. This is what we will do next.

V The exact solution of the linearized gap equation

In this section we prove that for any γ\gamma in the interval 0<γ<10<\gamma<1, the linearized gap equation has the solution for any NNcrN\leq N_{cr}. We obtain the exact analytical solution that satisfies Eq. (11) and the normalization condition, Eq. (15). The analysis is somewhat involved, so we discuss the details in Appendix C and here list the computational steps and present the final result.

We start by re-writing Eq. (11) as eigen-value/eigen-function equation of a linear integral operator

NΦ(ω¯m)=1γ2Φ(ω¯m)dω¯|ω¯mω¯m|γ|ω¯m|1γ11+|ω¯m|γN\Phi(\bar{\omega}_{m})=\frac{1-\gamma}{2}\int_{-\infty}^{\infty}\frac{\Phi(\bar{\omega}^{\prime}_{m})d\bar{\omega}^{\prime}}{|\bar{\omega}_{m}-\bar{\omega}^{\prime}_{m}|^{\gamma}|\bar{\omega}^{\prime}_{m}|^{1-\gamma}}\frac{1}{1+|\bar{\omega}^{\prime}_{m}|^{\gamma}} (67)

We then introduce a set of functions

Φβ(Ωm)|Ωm|iβγ+δΩm|Ωm|γ/2,\Phi_{\beta}(\Omega_{m})\equiv\frac{|\Omega_{m}|^{i\beta\gamma+\delta_{\Omega_{m}}}}{|\Omega_{m}|^{\gamma/2}}, (68)

which obey the orthogonality condition:

dΩm|Ωm|1γΦβΦβ=𝑑yeiγ(ββ)y=2πγδ(ββ),\int\frac{d\Omega_{m}}{|\Omega_{m}|^{1-\gamma}}\Phi^{*}_{\beta}\Phi_{\beta^{\prime}}=\int dye^{-i\gamma(\beta-\beta^{\prime})y}=\frac{2\pi}{\gamma}\delta(\beta-\beta^{\prime}), (69)

and define the function bβb_{\beta} as

bβ=12ϵβdΩm|Ωm|1γΦβ(Ωm)Φ(Ωm).b_{\beta}=\frac{1}{2\epsilon_{\beta}}\int\frac{d\Omega_{m}}{|\Omega_{m}|^{1-\gamma}}\Phi_{\beta}(\Omega_{m})\Phi(\Omega_{m}). (70)
Refer to caption
Figure 11: The exact solution of the linearized gap equation, Δex(x)\Delta_{\text{ex}}(x), as a function of x=log(|ωm|/ω0)γx=\log{(|\omega_{m}|/\omega_{0})^{\gamma}} for N=1N=1 and various γ\gamma from the interval 0<γ<10<\gamma<1. For all γ\gamma, Δex(x)\Delta_{\text{ex}}(x) oscillates at small frequencies with the period set by xx, and decays as (ω0/|ωm|)γ=ex(\omega_{0}/|\omega_{m}|)^{\gamma}=e^{-x} at large xx.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 12: The exact Δex(x)\Delta_{\text{ex}}(x) and the solution of the differential gap equation Δdiff(x)\Delta_{\text{diff}}(x) (Eq. (53) as functions of logz=log((|ωm|/ω0)γ)\log z=\log((|\omega_{m}|/\omega_{0})^{\gamma}). For Δdiff\Delta_{\text{diff}}, we rescaled zz to z¯=z/d{\bar{z}}=z/d, where d=4(1γ)/(Nγ(4βN2+1))d=4(1-\gamma)/(N\gamma(4\beta^{2}_{N}+1)), see Appendix B.

Multiplying both sides of Eq. (67) by Φβ(Ωm)\Phi_{\beta}(\Omega_{m}) and integrating over dΩm/|Ωm|1γd\Omega_{m}/|\Omega_{m}|^{1-\gamma}, we obtain the equation for bβb_{\beta}:

Nbβ=iϵiβsinh(π(ββ+i0))bβ𝑑βNb_{\beta}=i\int_{-\infty}^{\infty}\frac{\epsilon_{i\beta^{\prime}}}{\sinh\left(\pi(\beta^{\prime}-\beta+i0)\right)}b_{\beta^{\prime}}d\beta^{\prime} (71)

where ϵiβ\epsilon_{i\beta} is defined in (38). The normalizability condition for Φ(ω¯m)\Phi(\bar{\omega}_{m}) for given NN and γ\gamma is expressed in terms of bβb_{\beta} as the non-divergence of the integral

[11Nϵβ]ϵβbβbβ𝑑β.\int_{-\infty}^{\infty}\left[1-\frac{1}{N}\epsilon_{\beta^{\prime}}\right]\epsilon_{\beta^{\prime}}b_{\beta^{\prime}}b^{\ast}_{\beta^{\prime}}d\beta^{\prime}. (72)

We solve Eq. (71) subject to (72) in Appendix C and use (70) and (69) to extract Φ(ω¯m)\Phi(\bar{\omega}_{m}) and Δ(ω¯m)=Φ(ω¯m)/(1+(ω¯m)γ)\Delta(\bar{\omega}_{m})=\Phi(\bar{\omega}_{m})/\left(1+(\bar{\omega}_{m})^{-\gamma}\right). We find that there is no normalizable solution for N>NcrN>N_{cr}, but for NNcrN\leq N_{cr}, the normalizable solution does exist. The solution Δex(z)\Delta_{\text{ex}}(z) is expressed as

Δex(z)=sinh(πβN)eiβlogzi2I(β)dβcosh(π(ββN)cosh(π(β+βN)).\Delta_{\text{ex}}(z)=\int_{-\infty}^{\infty}\frac{\sinh(\pi\beta_{N})e^{-i\beta{\log z}-\frac{i}{2}I(\beta)}d\beta}{\sqrt{\cosh(\pi(\beta-\beta_{N})\cosh(\pi(\beta+\beta_{N}))}}. (73)

Here βN\beta_{N} is the same as before (the solution of ϵβN=N\epsilon_{\beta_{N}}=N), and the function I(β)I(\beta) is given by

I(β)=𝑑βlog|11Nϵβ|[tanh(π(ββ))tanh(πβ)].I(\beta)=\int_{-\infty}^{\infty}d\beta^{{}^{\prime}}\log\left|1-\frac{1}{N}\epsilon_{\beta^{{}^{\prime}}}\right|\left[\tanh\left(\pi(\beta^{{}^{\prime}}-\beta)\right)-\tanh\left(\pi\beta^{{}^{\prime}}\right)\right]. (74)

The gap function Δex(z)\Delta_{\text{ex}}(z) likely cannot be represented by a simple analytical formula, but can be straightforwardly computed numerically. We plot Δex(z)\Delta_{\text{ex}}(z) for different γ\gamma and N=1N=1 in Fig. 11. To show both the oscillations at small zz and the power-law behavior at larger zz, we use the logarithmical variable x=log(ω¯m)γ=logzx=\log(\bar{\omega}_{m})^{\gamma}=\log{z}, defined in (56).

Refer to caption
Figure 13: The function ϵb/γ\epsilon_{b/\gamma} from Eq. (28) for γ=0.5\gamma=0.5.

V.1 The exact Δex(z)\Delta_{\text{ex}}(z) vs Δdiff(z)\Delta_{\text{diff}}(z)

In Fig. 12 we compare the exact Δex(z)\Delta_{\text{ex}}(z) for N=1N=1 and various γ\gamma with the solution of the differential equation Δdiff(z)\Delta_{\text{diff}}(z) (both are expressed in terms of x=logzx=\log{z}). We see that the two essentially coincide for small γ\gamma and remain close to each other for γ=O(1)\gamma=O(1), but the discrepancy increases with increasing γ\gamma. To understand where the discrepancy comes from, we expanded the exact Δex(z)\Delta_{\text{ex}}(z) in powers of zz at small zz and in powers of 1/z1/z at large zz and compared with the expansion of Δdiff\Delta_{\text{diff}}. We present the details of the calculations in Appendix D and here list the results. The expansion of Δex(z)\Delta_{ex}(z) in zz holds in

Δex(z)=11+zRen=0e(iβNlogz+ϕN)Cn<zn+1/2+n,m=0Dn,m<z(n+bm/γ+1/2).\Delta_{\text{ex}}(z)=\frac{1}{1+z}{\text{R}e}~{}\sum_{n=0}^{\infty}e^{(i\beta_{N}\log{z}+\phi_{N})}C^{<}_{n}~{}z^{n+1/2}+\sum_{n,m=0}^{\infty}D^{<}_{n,m}z^{(n+b_{m}/\gamma+1/2)}. (75)

The expansion of Δex(z)\Delta_{\text{ex}}(z) in 1/z1/z holds in

Δex(z)=n=0Cn>z(n+1)+n,m=0Dn,m>z(n+2(m+1)/γ+1)\Delta_{\text{ex}}(z)=\sum_{n=0}^{\infty}C^{>}_{n}~{}z^{-(n+1)}+\sum_{n,m=0}^{\infty}D^{>}_{n,m}z^{-(n+2(m+1)/\gamma+1)} (76)

The real bmb_{m} in (75) satisfy ϵbm/γ=N\epsilon_{b_{m}/\gamma}=N. There are no solutions for |bm|<γ/2|b_{m}|<\gamma/2, but there are solutions for larger bb, see Fig. 13. A given bmb_{m} is in the interval γ/2+2m<bm<γ/2+2(m+1)\gamma/2+2m<b_{m}<\gamma/2+2(m+1). The crossover between the two regimes is at z=zmax=(ωmax/ω0)γz=z_{max}=(\omega_{max}/\omega_{0})^{\gamma}. This is the scale at which oscillating behavior of Δex(z)\Delta_{\text{ex}}(z) crosses over to 1/z1/z behavior.

The expansion of Δdiff(z)\Delta_{\text{diff}}(z) yields

Δdiff(z)=11+zRen=0e(iβNlogz+ϕN)C~n<zn+1/2\Delta_{\text{diff}}(z)=\frac{1}{1+z}{\text{R}e}~{}\sum_{n=0}^{\infty}e^{(i\beta_{N}\log{z}+\phi_{N})}{\tilde{C}}^{<}_{n}~{}z^{n+1/2} (77)

for the expansion in zz, and

Δdiff(z)=n=0C~n>z(n+1)\Delta_{\text{diff}}(z)=\sum_{n=0}^{\infty}{\tilde{C}}^{>}_{n}~{}z^{-(n+1)} (78)

for the expansion in 1/z1/z.

We see that the expansion of Δdiff(z)\Delta_{\text{diff}}(z) is a regular expansion in powers of either zz or 1/z1/z. The coefficients C~n<{\tilde{C}}^{<}_{n} and C~n>{\tilde{C}}^{>}_{n} can be easily obtained by expanding the hypergeometric function either in zz or in 1/z1/z, see Eqs. (58) and (59).

The expansion of Δex(z)\Delta_{\text{ex}}(z) is more complex and contains two types of terms: the CnC_{n} series and Dn,mD_{n,m} series. The Cn>C^{>}_{n} series in (75) are local terms in the sense that in the direct perturbative expansion of the gap equation (7) they come from ωmωm\omega^{\prime}_{m}\sim\omega_{m} in the integral in (7). The Dn,m<D^{<}_{n,m} series describe non-local corrections, for which the integral over ω\omega^{\prime} is determined by ωmaxωm\omega_{max}\gg\omega_{m}. Examining perturbation series, we find that the coefficients Cn<C^{<}_{n} can be obtained analytically and are given by

Cn<=C0<[Inm=1n1Im1]C^{<}_{n}=C^{<}_{0}~{}\left[I_{n}\displaystyle\prod_{m=1}^{n}\frac{1}{I_{m}-1}\right] (79)

where

Im=1γ2NQ(m,γ,βN),\displaystyle I_{m}=\frac{1-\gamma}{2N}Q(m,\gamma,\beta_{N}),
Q(m,γ,βN)=Γ((m+1/2)γ+iβNγ)Γ((1/2m)γiβNγ)Γ(γ)+\displaystyle Q(m,\gamma,\beta_{N})=\frac{\Gamma((m+1/2)\gamma+i{\beta_{N}}\gamma)\Gamma((1/2-m)\gamma-i{\beta_{N}}\gamma)}{\Gamma(\gamma)}+
Γ(1γ)(Γ((m+1/2)γ+iβNγ)Γ(1(1/2m)γ+iβNγ)+Γ((1/2m)γiβNγ)Γ(1(m+1/2)γiβNγ))\displaystyle\Gamma(1-\gamma)\left(\frac{\Gamma((m+1/2)\gamma+i{\beta_{N}}\gamma)}{\Gamma(1-(1/2-m)\gamma+i{\beta_{N}}\gamma)}+\frac{\Gamma((1/2-m)\gamma-i{\beta_{N}}\gamma)}{\Gamma(1-(m+1/2)\gamma-i{\beta_{N}}\gamma)}\right) (80)

The series do not converge absolutely because at large nn, Cn<1/nC^{<}_{n}\propto 1/n, but each term in the series can be easily calculated. At small γ\gamma and N=1N=1, when βN(Ncr/4N)1\beta_{N}\approx(N_{cr}/4N)\gg 1, the Cn<C^{<}_{n} series in (75) reproduce, order by order, the expansion of Δdiff(z)\Delta_{\text{diff}}(z), Eqs. (58) and (59). The same holds for the expansion in 1/z1/z – the Cn>C^{>}_{n} series in (76) reproduce the expansion in (57). The DD terms are small for γ1\gamma\ll 1 by a combination of two factors: (i) the prefactors are relatively small, e.g., D0,0>C0>γ/4D^{>}_{0,0}\approx-C^{>}_{0}~{}\gamma/4, and (ii) bm2(m+1)b_{m}\approx 2(m+1), hence, the non-local terms contain additional small factors z2m/γz^{2m/\gamma} in the expansion in zz and and (1/z)2m/γ(1/z)^{-2m/\gamma} in the expansion in 1/z1/z. As the consequence, for small γ\gamma, Δex(z)\Delta_{\text{ex}}(z) and Δdiff(z)\Delta_{\text{diff}}(z) coincide, up to small corrections, as Fig. 12(a) demonstrates.

At larger γ=O(1)\gamma=O(1), the non-local terms become relevant, particularly near zmaxz_{max}, where Δex(z)\Delta_{\text{ex}}(z) crosses over to 1/z1/z behavior. Numbers-wise, Δdiff(z)\Delta_{\text{diff}}(z) still agrees reasonably well with Δex(z)\Delta_{\text{ex}}(z), as Fig. 12 shows, but qualitatively, the exact Δex(z)\Delta_{\text{ex}}(z) is different from Δdiff(z)\Delta_{\text{diff}}(z).

VI Non-linear gap equation

We now analyze the full non-linear gap equation, Eq. (7). We argue that it has an infinite, discrete set of solutions, which can be specified by topological index nn, and the limit nn\to\infty corresponds to the solution of the linearized gap equation.

VI.1 Qualitative reasoning

We begin with qualitative reasoning. We assume that for any finite nn, the gap function Δn(ωm)\Delta_{n}(\omega_{m}) tends to a finite value at ωm=0\omega_{m}=0, while at ωm>ωn\omega_{m}>\omega^{*}_{n}, Δn(ωm)\Delta_{n}(\omega_{m}) is small and has the same dependence on ωm\omega_{m} as Δ(ωm)=Δex(ωm)\Delta_{\infty}(\omega_{m})=\Delta_{\text{ex}}(\omega_{m}). We further assume that ωn\omega^{*}_{n} is the only relevant scale for Δn(ωm)\Delta_{n}(\omega_{m}), i.e., Δn(ωm)\Delta_{n}(\omega_{m}) saturates at ωm<ωn\omega_{m}<\omega^{*}_{n}. This reasoning has two implications. First, Δn(ωn)ωn\Delta_{n}(\omega^{*}_{n})\sim\omega^{*}_{n}. Second, Δn(ωm)\Delta_{n}(\omega_{m}) must satisfy the modified linearized gap equation, in which the lower limit of integration over positive and negative ωm\omega^{\prime}_{m} is set at a frequency of order ωn\omega^{*}_{n}. Because Δex(ωm)\Delta_{\text{ex}}(\omega_{m}) is the solution for ωn=0\omega^{*}_{n}=0, the modified linearized gap equation has a solution if

0ωn𝑑ωmΔex(ωm)ωm=0.\int_{0}^{\omega^{*}_{n}}d\omega_{m}^{\prime}\frac{\Delta_{ex}(\omega^{\prime}_{m})}{\omega^{\prime}_{m}}=0. (81)

These two conditions determine ωn\omega^{*}_{n} and Δn(ωn)\Delta_{n}(\omega^{*}_{n}). Below we use the variable zz instead of ωm\omega_{m} and define zn=(ωn/ω0)γz^{*}_{n}=(\omega^{*}_{n}/\omega_{0})^{\gamma}. In these variables, Δn(0)Δn(zn)ω0(zn)1/γ\Delta_{n}(0)\sim\Delta_{n}(z^{*}_{n})\sim\omega_{0}(z_{n})^{1/\gamma}.

For NNcrN\leq N_{cr}, Δex(z)\Delta_{\text{ex}}(z) behaves at z<1z<1 as Czcos(βNlogz+ϕN)C\sqrt{z}\cos(\beta_{N}\log{z}+\phi_{N}), where βN(NcrN)1/2\beta_{N}\propto(N_{cr}-N)^{1/2} is small and ϕN=π/20.773βN\phi_{N}=\pi/2-0.773\beta_{N}. Evaluating the integral in (81) and re-expressing the result in terms of zz, we obtain

sin(βN(logz2.77))=0,\sin\left(\beta_{N}\left(\log{z^{*}}-2.77\right)\right)=0, (82)

This equation has a discrete set of solutions for z1z^{*}\ll 1: logz=2.77π(n+1)/βN\log{z^{*}}=2.77-\pi(n+1)/\beta_{N}, where n=0,1,2n=0,1,2..., i.e,

zneπ(n+1)/βN,z^{*}_{n}\propto e^{-\pi(n+1)/\beta_{N}}, (83)

All znz^{*}_{n}, including z0z^{*}_{0}, are exponentially small at small βN\beta_{N}. The gap functions Δn(0)ω0(zn)1/γ\Delta_{n}(0)\sim\omega_{0}(z^{*}_{n})^{1/\gamma} are also exponentially small:

Δn(0)ω0eπ(n+1)/(γβN).\Delta_{n}(0)\propto\omega_{0}e^{-\pi(n+1)/(\gamma\beta_{N})}. (84)

One can easily verify that Δn(z)\Delta_{n}(z) changes sign nn times between z=0z=0 and z=z=\infty.

The result zneπn/βNz^{*}_{n}\propto e^{-\pi n/\beta_{N}} at n1n\gg 1 also follows from a generic reasoning that Δex(z)\Delta_{\text{ex}}(z) should have an extremum at zznz\sim z^{*}_{n} to match with a constant Δn(z)Δn(0)\Delta_{n}(z)\approx\Delta_{n}(0) for smaller zz. We note, however, that the presence of an extremum in Δex(z)\Delta_{\text{ex}}(z) by itself does not guarantee that the solution of the non-linear gap equation exists. Indeed, Δex(z)\Delta_{\text{ex}}(z) for N=NcrN=N_{cr} has a maximum at z=O(1)z=O(1), but there is no solution with a finite Δ(z)\Delta(z) for this NN.

When βN1\beta_{N}\gg 1 (e.g., for γ1\gamma\ll 1 and N=O(1)N=O(1)), we can still apply the same reasoning for large enough nn, when znz^{*}_{n} is exponentially small in nn. However, for n=O(1)n=O(1), zn1z^{*}_{n}\geq 1, and we should use appropriate expression for Δex(z)\Delta_{\text{ex}}(z) in Eq. (81). Replacing Δex(z)\Delta_{\text{ex}}(z) by Δdiff(z)\Delta_{{\text{diff}}}(z) as the two are very close, and using the asymptotic form of Δdiff(z)\Delta_{{\text{diff}}}(z) for z>1z>1: Δdiff(z)(1/z)J1(2βN/z)\Delta_{{\text{diff}}}(z)\sim(1/\sqrt{z})J_{1}(2\beta_{N}/\sqrt{z}), we obtain the condition on znz^{*}_{n} in the form

J0(2βNzn)=O(1).J_{0}\left(\frac{2\beta_{N}}{\sqrt{z^{*}_{n}}}\right)=O\left(1\right). (85)

This equation has a discrete set of solutions zn=4βN2snz^{*}_{n}=4\beta^{2}_{N}s_{n}, where where sns_{n} is a decreasing function of nn. The corresponding Δn(zn)Δn(0)ω0(βN2)1/γ\Delta_{n}(z^{*}_{n})\sim\Delta_{n}(0)\sim\omega_{0}(\beta^{2}_{N})^{1/\gamma}. For small γ\gamma, βN21/γ\beta^{2}_{N}\approx 1/\gamma, hence, Δn(0)ω0(1/γ)1/γ\Delta_{n}(0)\sim\omega_{0}(1/\gamma)^{1/\gamma} (Refs. Metlitski et al. (2015); Wang et al. (2016); Wu et al. (2019a); Yuzbashyan et al. (a)). This analysis holds up to n=nmaxn=n_{max}, for which znmax=O(1)z^{*}_{n_{max}}=O(1) For larger nn, znz^{*}_{n} are given by Eq. (83). Still, Δn(z)\Delta_{n}(z) changes sign nn times between z=0z=0 and z=z=\infty.

Note that at small γ\gamma, Δn(0)ω0(zn)1/γ\Delta_{n}(0)\propto\omega_{0}(z^{*}_{n})^{1/\gamma} rapidly evolves from Δn(0)ω0\Delta_{n}(0)\ll\omega_{0} when zn<1z^{*}_{n}<1 to Δn(0)ω0\Delta_{n}(0)\gg\omega_{0} when zn>1z^{*}_{n}>1. In Ref. Wu et al. (2019a) we solved the non-linear gap equation at small γ\gamma and N=1N=1 for sign-preserving Δ0(z)\Delta_{0}(z). We obtained Δ0(0)=0.885g¯(1/1.4458γ)1/γ0.326ω0(1/1.4458γ)1/γ\Delta_{0}(0)=0.885{\bar{g}}(1/1.4458\gamma)^{1/\gamma}\equiv 0.326\omega_{0}(1/1.4458\gamma)^{1/\gamma}, consistent with our estimate. The number 1.44581.4458 comes from the fact that at small γ\gamma, the more explicit form of (85) is J0(2βN/z0)=O(γ)J_{0}(2\beta_{N}/\sqrt{z^{*}_{0}})=O(\gamma) (Refs. Metlitski et al. (2015); Wang et al. (2016); Wu et al. (2019a); Yuzbashyan et al. (a)).

VI.2 Non-linear differential equation

We now corroborate qualitative reasoning by more direct analysis in which we derive and solve the non-linear differential gap equation. To derive this equation, we again depart from the original integral equations for the pairing vertex and the self-energy, Eqs. (5) and (6), or, equivalently, from the non-linear gap equation (7) and approximate the integral over ωm\omega_{m}^{\prime} in the l.h.s. of (5) by the integrals over ωmωm\omega^{\prime}_{m}\gg\omega_{m} and ωm<ωm\omega^{\prime}_{m}<\omega_{m}, each time neglecting either ωm\omega^{\prime}_{m} or ωm\omega_{m} in the pairing interaction. However, for a finite Δ(ωm)\Delta(\omega_{m}) we have to take into account the fact that the self-energy has a more complex form in the presence of Δ\Delta. In particular, one can easily verify that if Δ(ωm)\Delta(\omega_{m}) tends to a finite value at ωm=0\omega_{m}=0, as we assume to be the case, Σ(ωm)\Sigma(\omega_{m}) acquires a Fermi liquid form Σ(ωm)ωm(ω0/Δ(0))γ\Sigma(\omega_{m})\sim\omega_{m}(\omega_{0}/\Delta(0))^{\gamma} at the smallest ωm\omega_{m}. The non-Fermi liquid Σ(ωm)=|ωm|1γω0γsignωm=ωm/z\Sigma(\omega_{m})=|\omega_{m}|^{1-\gamma}\omega^{\gamma}_{0}{\text{sign}\omega_{m}}=\omega_{m}/z is recovered at frequencies for which |ωm|Δ(ωm)|\omega_{m}|\gg\Delta(\omega_{m}). These two limiting forms of the self-energy can be combined into the interpolation formula

Σ(ωm)=ωmω0γ(ωm2+Δ2(ωm))γ/2\Sigma(\omega_{m})=\omega_{m}\frac{\omega^{\gamma}_{0}}{(\omega^{2}_{m}+\Delta^{2}(\omega_{m}))^{\gamma/2}} (86)

If we include this self-energy and then follow the same derivation for the linearized differential gap equation, we would obtain

[Δ¯(z)(z+1(1+D2(z))γ/2)]′′=Ncr4Nz2Δ¯(z)1+D2(z)\left[{\bar{\Delta}}(z)\left(z+\frac{1}{(1+D^{2}(z))^{\gamma/2}}\right)\right]^{{}^{\prime\prime}}=-\frac{N_{cr}}{4Nz^{2}}\frac{{\bar{\Delta}}(z)}{\sqrt{1+D^{2}(z)}} (87)

where Δ¯(z)=Δ(z)/ω0{\bar{\Delta}}(z)=\Delta(z)/\omega_{0} and D(z)=Δ(z)/z1/γD(z)=\Delta(z)/z^{1/\gamma}. There is a second complication, however. One can verify that a at non-zero Δ(0)\Delta(0), the r.h.s of the actual, non-linear integral gap equation (7) (with the self-energy contribution kept in the l.h.s) has a regular expansion in powers of ωm2\omega^{2}_{m}. Indeed, a direct expansion in ωm\omega_{m} yields

g¯γ2N𝑑ωmΔ(ωm)(ωm)2+Δ2(ωm)1|ωmωm|γ=A+Bωm2\frac{{\bar{g}}^{\gamma}}{2N}\int d\omega^{\prime}_{m}\frac{\Delta(\omega^{\prime}_{m})}{\sqrt{(\omega^{\prime}_{m})^{2}+\Delta^{2}(\omega^{\prime}_{m})}}~{}\frac{1}{|\omega_{m}-\omega^{\prime}_{m}|^{\gamma}}=A+B\omega^{2}_{m} (88)

where AA and BB are given by convergent integrals:

A\displaystyle A =\displaystyle= g¯γ2N𝑑ωmΔ(ωm)(ωm)2+Δ2(ωm)1|ωm|γ\displaystyle\frac{{\bar{g}}^{\gamma}}{2N}\int d\omega^{\prime}_{m}\frac{\Delta(\omega^{\prime}_{m})}{\sqrt{(\omega^{\prime}_{m})^{2}+\Delta^{2}(\omega^{\prime}_{m})}}\frac{1}{|\omega^{\prime}_{m}|^{\gamma}}
B\displaystyle B =\displaystyle= γ(γ+1)2g¯γ2Ndωm(ωm)2+Δ2(ωm)(Δ(ωm)+(ωm)2+Δ2(ωm))1|ωm|γ.\displaystyle-\frac{\gamma(\gamma+1)}{2}\frac{{\bar{g}}^{\gamma}}{2N}\int\frac{d\omega^{\prime}_{m}}{\sqrt{(\omega^{\prime}_{m})^{2}+\Delta^{2}(\omega^{\prime}_{m})}\left(\Delta(\omega_{m})+\sqrt{(\omega^{\prime}_{m})^{2}+\Delta^{2}(\omega^{\prime}_{m})}\right)}\frac{1}{|\omega^{\prime}_{m}|^{\gamma}}. (89)

The combination of Eqs. (86) and (88) then implies that at small ωm\omega_{m}, Δ(ωm)\Delta(\omega_{m}) is a regular function of frequency Δ(ωm)=Δ0+Cωm2\Delta(\omega_{m})=\Delta_{0}+C\omega^{2}_{m}, as is expected in a Fermi liquid regime. Equivalently, Δ¯(z)=Δ¯0+C¯z2/γ{\bar{\Delta}}(z)={\bar{\Delta}}_{0}+{\bar{C}}z^{2/\gamma}. This regular expansion of Δ¯(z){\bar{\Delta}}(z) is not reproduced by Eq. (87). The reason can be easily understood by analyzing how the ωm2\omega^{2}_{m} term appears in the r.h.s. of (88): to get it one has to keep both ωm\omega^{\prime}_{m} and ωm\omega_{m} in the interaction, i.e., go beyond the approximation used to derive (87). A simple experimentation shows that to reproduce regular behavior of the gap function at small frequencies, one has to multiply the r.h.s. of (87) by the factor (1+D2(z)|D(z)|)1+γ\left(\sqrt{1+D^{2}(z)}-|D(z)|\right)^{1+\gamma}. This factor reduces to 11 when D(z)D(z) is small, but gives additional z(1+γ)/γz^{(1+\gamma)/\gamma} when D(z)D(z) is large. With this extra factor, the non-linear differential gap equation takes the form

[Δ¯(z)(z+1(1+D2(z))γ/2)]′′=Ncr4Nz2Δ¯(z)1+D2(z)(1+D2(z)|D(z)|)1+γ\left[{\bar{\Delta}}(z)\left(z+\frac{1}{(1+D^{2}(z))^{\gamma/2}}\right)\right]^{{}^{\prime\prime}}=-\frac{N_{cr}}{4Nz^{2}}\frac{{\bar{\Delta}}(z)}{\sqrt{1+D^{2}(z)}}\left(\sqrt{1+D^{2}(z)}-|D(z)|\right)^{1+\gamma} (90)

One can verify that at small zz, the solution of this equation, if it exists, has an asymptotic form Δ¯(z)=Δ¯(0)+C¯z2/γ{\bar{\Delta}}(z)={\bar{\Delta}}(0)+{\bar{C}}z^{2/\gamma}, consistent with asymptotic form of the solution of the actual integral equation.

The boundary conditions for (90) are a finite Δ(0)\Delta(0) and Δ(z)1/z\Delta(z)\propto 1/z at zz\to\infty. The last condition is the same as for the linearized gap equation. According to our qualitative reasoning, this equation should have an infinite, discrete set of solutions Δn(z)\Delta_{n}(z).

Refer to caption
Figure 14: The ratio of Δ(ω)/Δ(ω0)\Delta(\omega\to\infty)/\Delta(\omega\rightarrow 0) vs logΔ(ω0)\log\Delta(\omega\rightarrow 0).

We solve Eq. (90) numerically. We set Δ(0)\Delta(0) to some value, which we vary. For a generic Δ(0)\Delta(0), Δ(z)\Delta(z) evolves with zz towards some constant at zz\to\infty. We verify whether for some particular Δ(0)\Delta(0), a constant vanishes and Δ(z)\Delta(z) decays as 1/z1/z. We show the result in Fig. (14) for γ=0.3\gamma=0.3 and N=1N=1 (Ncr/N13N_{cr}/N\approx 13, βn1.75\beta_{n}\approx 1.75). We plot the ratio of Δ(z)/Δ(0)\Delta(z\to\infty)/\Delta(0) vs Δ(0)\Delta(0). We see that for a generic Δ(0)\Delta(0), Δ()\Delta(\infty) remains finite, but there is a discrete set of Δn(0)\Delta_{n}(0), for which Δ()\Delta(\infty) vanishes. With our numerical accuracy we can clearly identify three such Δn(0)\Delta_{n}(0). In three inserts in Fig. (14) we show Δ(z)\Delta(z) for these three Δn(0)\Delta_{n}(0). We see that the gap function changes sign nn times as a function of zz (n=0,1,2n=0,1,2), precisely as we anticipated.

In the top right insert in Fig. (14) we plot Δn(0)\Delta_{n}(0) vs nn for N=1N=1 and γ=0.3\gamma=0.3. We clearly see that Δn(0)\Delta_{n}(0) has exponential dependence on nn. We fitted the results to Eq. (83). The best fit yields βN1.51\beta_{N}\approx 1.51. This is quite consistent with the actual value βN1.75\beta_{N}\approx 1.75 (we note that Eq. (83) is by itself only valid for βN<1\beta_{N}<1).

Refer to caption
Refer to caption
Figure 15: Left: the ratio of Δ(ω)/Δ(ω0)\Delta(\omega\to\infty)/\Delta(\omega\rightarrow 0) vs logΔ(ω0)\log\Delta(\omega\rightarrow 0). for different NN. The inset shows Δ0\Delta_{0} vs πγβN\frac{\pi}{\gamma\beta_{N}}, the best fit has the slope of 0.95-0.95. Right Δ0\Delta_{0} vs zz for γ=0.3\gamma=0.3 and N=6Ncr/2N=6\approx N_{cr}/2.

In the left panel of Fig. (15) we collect the results for γ=0.3\gamma=0.3 and different NN between N=1N=1 and NcrN_{cr}, the inset shows a comparison of logΔ0\log\Delta_{0} and πγβn\frac{\pi}{\gamma\beta_{n}}, where βN=0.5(Ncr/N1)1/2\beta_{N}=0.5(N_{cr}/N-1)^{1/2}. We see that the agreement is quite good: the theoretical slope is 1-1 (see Eq. (84)) and the fitting of numerical data yields the slope 0.95-0.95.

Finally, in the right panel of Fig. (15) we plot Δ0\Delta_{0} vs zz for NNcr/2N\approx N_{cr}/2, where Δ0(0)\Delta_{0}(0) is already very small. We see that the gap function flattens at z(Δ0(0))γz^{*}\sim(\Delta_{0}(0))^{\gamma}, which is much smaller than the scale at which Δ0(z)\Delta_{0}(z) has a broad maximum. This is again consistent with our qualitative reasoning.

In the subsequent publication (Paper II, Ref.Wu et al. (2020)) we collaborate the T=0T=0 analysis with the analysis of the linearized gap equation at a finite TT. We show that for N<NcrN<N_{cr}, there exists a discrete set of onset pairing temperatures Tp,nT_{p,n}, and the eigenfunction Δn(ωm)\Delta_{n}(\omega_{m}) changes sign nn times as a function of Matsubara frequency. We argue that these Δn(ωm)\Delta_{n}(\omega_{m}) grow in magnitude below Tp,nT_{p,n}, but preserve the number of sign changes, and at T=0T=0 coincide with Δn(ωm)\Delta_{n}(\omega_{m}), which we found here.

We note in passing that Δn(ωm)\Delta_{n}(\omega_{m}) has a maximum at ωω0g¯\omega\sim\omega_{0}\sim{\bar{g}}. The reason for the maximum is that at T=0T=0, a non-zero Δn(ωm)\Delta_{n}(\omega_{m}) emerges only at N<NcrN<N_{cr}, and Δn(0)\Delta_{n}(0), including n=0n=0, scale with NcrNN_{cr}-N. Meanwhile, the onset temperature for the pairing into the n=0n=0 state, Tp,0T_{p,0}, is of order ω0\omega_{0}. The gap structure reflects this discrepancy: Δn(ωm)\Delta_{n}(\omega_{m}) at the lowest frequencies scales with NcrNN_{cr}-N, but the gap at ωmω0\omega_{m}\sim\omega_{0} scales with ω0\omega_{0}. This gives rise to a maximum at ωmω0\omega_{m}\sim\omega_{0}.

VII Conclusions

In this paper we analyzed the competition between two opposing trends in the behavior of interacting fermions near a quantum critical point in a metal: non-Fermi-liquid physics and pairing. Both trends are captured within the model of fermions interacting by exchanging soft bosonic fluctuations of an order parameter, which condenses at the critical point. The non-Fermi liquid behavior is the result of strong, non-analytic self-energy due to boson-mediated scattering near the Fermi surface, the pairing is due to strong attraction in at least one pairing channel, provided by the same soft boson exchange. We considered a class of quantum-critical models with an effective dynamical electron-electron interaction V(Ωm)1/|Ωm|γV(\Omega_{m})\propto 1/|\Omega_{m}|^{\gamma} (the γ\gamma-model). In this paper, the first in the series, we considered the case 0<γ<10<\gamma<1 and restricted the analysis to T=0T=0. The limit γ=0\gamma=0 corresponds to BCS theory without the upper cutoff, but for all finite γ\gamma the interaction drops off at large Ω\Omega and the pairing problem is ultra-violet convergent. To parametrically separate the tendencies towards non-Fermi liquid and pairing, we extended the model and introduced the parameter 1/N1/N as additional overall factor in the pairing interaction only. At large NN, the tendency towards non-Fermi liquid normal state at T=0T=0 is stronger by NN, at small NN the tendency towards pairing is stronger by 1/N1/N.

Our analysis brings two conclusions. First, we found that there indeed exists a critical N=NcrN=N_{cr}, separating non-Fermi liquid normal state at larger NN and superconducting state at smaller NN. The critical Ncr>1N_{cr}>1 for all γ<1\gamma<1, such that the original N=1N=1 model is superconducting at T=0T=0. Second, we found that the system behavior for N<NcrN<N_{cr} is rather unconventional in the sense that there exists an infinite set of solutions of the non-linear gap equation, Δn(ωm)\Delta_{n}(\omega_{m}), n=0,1,2n=0,1,2..., all within the same pairing symmetry. The solutions are topologically distinct: Δn(ωm)\Delta_{n}(\omega_{m}) changes sign nn times as a function of Matsubara frequency ωm\omega_{m}. All Δn\Delta_{n} emerge at N=NcrN=N_{cr}, and for NNcrN\leq N_{cr} their magnitudes scale as Δneπ(n+1)/γβN\Delta_{n}\propto e^{-\pi(n+1)/\gamma\beta_{N}}, where βN(NcrN)1/2\beta_{N}\propto(N_{cr}-N)^{1/2}. The end point of the set, Δ(ωm)\Delta_{\infty}(\omega_{m}), is the solution of the linearized gap equation at T=0T=0. We obtained the exact analytical solution of the linearized gap equation at T=0T=0 and also a highly accurate simple approximate solution for all NNcrN\leq N_{cr}. In the subsequent paper we analyze the linearized gap equation at a finite TT and show that there is an infinite set of the onset temperatures for the pairing, Tp,nT_{p,n}, and the corresponding eigenfunctions change sign nn times as functions of ωm\omega_{m}. We argue that these topologically distinct gap functions grow in magnitude below the corresponding Tp,nT_{p,n}, but largely preserve the functional forms and at T=0T=0 become Δn\Delta_{n} that we found in this work.

The sign-preserving solution with n=0n=0 has the largest condensation energy and is the global minimum of the condensation energy EcE_{c}. All other solutions are local minima. Still, the presence of the infinite set of Δn(ωm)\Delta_{n}(\omega_{m}), including the solution of the linear gap equation for all NNcrN\leq N_{cr} is a highly non-trivial feature of the pairing at a QCP. In subsequent publications we extend the T=0T=0 analysis to γ>1\gamma>1 and argue that the local minima become more closely spaced with increasing γ\gamma form a continuous set for γ=2\gamma=2. We argue that in this case all solutions have equal condensation energy, and superconducting order gets destroyed already at T=0T=0. As the consequence, the true superconducting TcT_{c} terminates at γ=2\gamma=2, while the onset temperature for the pairing Tp,0=TpT_{p,0}=T_{p} remains finite (see Fig. 2). The two temperatures then must be different for all 0<γ20<\gamma\leq 2, including 0<γ<10<\gamma<1, which we considered here. In between TcT_{c} and TpT_{p} the system displays pseudogap behavior. We emphasize again that the root to this highly unusual behavior is the existence of the solution of the linearized gap equation for all NN, rather than only at N=NcrN=N_{cr}.

Acknowledgements.
We thank I. Aleiner, B. Altshuler, E. Berg, D. Chowdhury, L. Classen, K. Efetov, R. Fernandes, A. Finkelstein, E. Fradkin, A. Georges, S. Hartnol, S. Karchu, S. Kivelson, I. Klebanov, A. Klein, R. Laughlin, S-S. Lee, G. Lonzarich, D. Maslov, F. Marsiglio, M. Metlitski, W. Metzner, A. Millis, D. Mozyrsky, C. Pepan, V. Pokrovsky, N. Prokofiev, S. Raghu, S. Sachdev, T. Senthil, D. Scalapino, Y. Schattner, J. Schmalian, D. Son, G. Tarnopolsky, A-M Tremblay, A. Tsvelik, G. Torroba, E. Yuzbashyan, J. Zaanen, and particularly R. Combescot, Y. Wang and Y. Wu for useful discussions. The work by AVC was supported by the NSF DMR-1834856.

Appendix A The case of small γ\gamma and a finite frequency cutoff

The limit γ0\gamma\to 0 of the γ\gamma model attracted a lot of attention from various sub-communities in physicsRaghu et al. (2015); Wu et al. (2019a); Wang et al. (2017a, 2018); Fitzpatrick et al. (2015b); Metlitski et al. (2015); Mross et al. (2010); Son (1999); *son2 and has been analyzed in both Eliashberg-type and renormalization group approaches. This limit requires extra care because strictly at γ=0\gamma=0, the interaction V(Ω)|Ω|γV(\Omega)\propto|\Omega|^{-\gamma} becomes frequency independent, and the pairing problem becomes equivalent to BCS, but without the upper cutoff for the interaction. Meanwhile for any non-zero γ\gamma, the gap equation converges at ω>ωmaxge|logγ|/γ\omega>\omega_{max}\propto ge^{|\log\gamma|/\gamma}. More explicit calculation yields ωmaxg(1/1.446)1/γe|logγ|/γ\omega_{max}\sim g(1/1.446)^{1/\gamma}e^{|\log\gamma|/\gamma} (Refs. Metlitski et al. (2015); Wang et al. (2016); Wu et al. (2019a); Yuzbashyan et al. (a)). This ωmax\omega_{max} remains finite at any non-zero γ\gamma, but grows with decreasing γ\gamma faster than the exponent.

In this Appendix we consider how the gap equation gets modified if we introduce the upper cutoff for V(Ωm)V(\Omega_{m}) at Ωm=Λ\Omega_{m}=\Lambda, i.e., modify V(Ωm)=(g¯/|Ωm|)γV(\Omega_{m})=({\bar{g}}/|\Omega_{m}|)^{\gamma} to

V(Ωm)=(g¯|Ωm|)γ[1(|Ωm|Λ)γ]V(\Omega_{m})=\left(\frac{{\bar{g}}}{|\Omega_{m}|}\right)^{\gamma}\left[1-\left(\frac{|\Omega_{m}|}{\Lambda}\right)^{\gamma}\right] (91)

for |Ωm|<Λ|\Omega_{m}|<\Lambda, and V(Ωm)=0V(\Omega_{m})=0 for |Ωm|>Λ|\Omega_{m}|>\Lambda. We consider γ\gamma for which Λωmax\Lambda\ll\omega_{max}. In this case we can expand V(Ω)V(\Omega) to first order in γ\gamma and express it for Ωm<Λ\Omega_{m}<\Lambda as V(Ω)=γlog(Λ/|Ω|)V(\Omega)=\gamma\log{(\Lambda/|\Omega|)}. The normal state self-energy is Σ(ωm)=γωmlogΛ/|ωm|\Sigma(\omega_{m})=\gamma\omega_{m}\log{\Lambda/|\omega_{m}|}. Substituting these V(Ω)V(\Omega) and Σ(ωm)\Sigma(\omega_{m}) into the equation for the pairing vertex, we obtain

Φ(ωm)=γ2NdωmlogΛ|ωmωm||ωm|(1+γlogΛ|ωm|)\Phi(\omega_{m})=\frac{\gamma}{2N}\int\frac{d\omega^{\prime}_{m}\log{\frac{\Lambda}{|\omega_{m}-\omega^{\prime}_{m}|}}}{|\omega^{\prime}_{m}|(1+\gamma\log{\frac{\Lambda}{|\omega^{\prime}_{m}|}})} (92)

Introducting logarithmic variables x=logΛ/|ωm|x=\log{\Lambda/|\omega_{m}|}, x=logΛ/|ωm|x^{\prime}=\log{\Lambda/|\omega^{\prime}_{m}|} and restricting with the contributions from ωmωm\omega^{\prime}_{m}\gg\omega_{m} and ωmωm\omega^{\prime}_{m}\ll\omega_{m}, as we did in the derivation of (43), we reduce (92) to

Φ(x)=γN[0x𝑑xxΦ(x)1+γx+x1+γxx𝑑xΦ(x)]\Phi(x)=\frac{\gamma}{N}\left[\int_{0}^{x}dx^{\prime}\frac{x^{\prime}\Phi(x^{\prime})}{1+\gamma x^{\prime}}+\frac{x}{1+\gamma x}\int_{x}^{\infty}dx^{\prime}\Phi(x^{\prime})\right] (93)

Differentiating twice over xx and introducing y=γxy=\gamma x, we obtain

(Φ(y)(1+y)2)=Φ(y)Nγ\left(\Phi^{\prime}(y)(1+y)^{2}\right)^{\prime}=-\frac{\Phi(y)}{N\gamma} (94)

The boundary conditions are Φ(y=0)=0\Phi(y=0)=0 (Φ(ωm)\Phi(\omega_{m}) must vanish at ωm=Λ\omega_{m}=\Lambda) and Φ(y=)=0\Phi(y=\infty)=0, i.e., Φ(ωm=0)=0\Phi(\omega_{m}=0)=0. The last condition is set to avoid the divergence in (92) at vanishing ωm\omega^{\prime}_{m}. The solution of (94) is

Φ(y)=C1+ycos(ϕ+1Nγ14log(1+y))\Phi(y)=\frac{C}{\sqrt{1+y}}\cos\left(\phi+\sqrt{{\frac{1}{N\gamma}-\frac{1}{4}}}\log{(1+y)}\right) (95)

This Φ(y)\Phi(y) satisfies the boundary condition Φ(y=)=0\Phi(y=\infty)=0 provided 1/Nγ>1/41/N\gamma>1/4, i.e., N<Ncr=4/γN<N_{cr}=4/\gamma. To satisfy the other boundary condition we choose ϕ=π/2\phi=\pi/2.

We plot Φ(y)\Phi(y) in Fig.16 for N=1N=1 and γ=0.01\gamma=0.01. We see that Φ(y)\Phi(y) is an oscillating function of yy for y1y\ll 1 and an oscillating function of logy\log{y} for y1y\gg 1. For y1y\ll 1 we have, restoring the units of frequency,

Φ(ωm)=Csin(γ1Nγ14logΛ|ωm|)\Phi(\omega_{m})=-C\sin\left(\gamma\sqrt{{\frac{1}{N\gamma}-\frac{1}{4}}}\log{\frac{\Lambda}{|\omega_{m}|}}\right) (96)
Refer to caption
Figure 16: The function Φ(y)\Phi(y) from Eq. (95).

The largest frequency for oscillations is ω0Λeγ1/(Nγ)1/4\omega^{*}_{0}\sim\Lambda e^{-\gamma\sqrt{1/(N\gamma)-1/4}}. This scale determines the magnitude of the gap Δ0(0)\Delta_{0}(0). For Nγ1N\gamma\ll 1, ω0Λeπ2N/γ\omega^{*}_{0}\sim\Lambda e^{-\frac{\pi}{2}\sqrt{N/\gamma}}. The ratio γ/N\gamma/N plays the role of a dimensionless coupling λ\lambda, and Δ0\Delta_{0} can be cast into Δ0(0)Λeπ2λ\Delta_{0}(0)\sim\Lambda e^{-\frac{\pi}{2\sqrt{\lambda}}}. This agrees with Refs. Son (1999); *son2; Metlitski et al. (2015). Comparing Δ0(0)\Delta_{0}(0) obtained with and without the cutoff at Λ\Lambda we see that the analysis of the γ\gamma model without the upper cutoff is valid as long as g(1/1.446γ)1/γΛeγ1/Nγ1/4g(1/1.446\gamma)^{1/\gamma}\ll\Lambda e^{-\gamma\sqrt{1/N\gamma-1/4}}. At smaller Λ\Lambda, Δ0(0)Λeγ1/Nγ1/4\Delta_{0}(0)\sim\Lambda e^{-\gamma\sqrt{1/N\gamma-1/4}}. We emphasize that even in this case there are an infinite number of solutions

Δn(0)Λeπ2λenπλ.\Delta_{n}(0)\sim\Lambda e^{-\frac{\pi}{2\sqrt{\lambda}}}e^{-\frac{n\pi}{\sqrt{\lambda}}}. (97)

This is valid up to n1/Nγn\sim 1/\sqrt{N\gamma}. For larger nn, relevant yy in (94) become larger than one, and the formula for Δn(0)\Delta_{n}(0) changes.

At small γ\gamma and N=O(1)N=O(1), the ratio Σ(ω)/ω\Sigma(\omega)/\omega is small at ωω0\omega\sim\omega^{*}_{0}, both when Λ\Lambda is finite and when it is set to infinity. In the first case, the ratio is of order γ\sqrt{\gamma}, in the second it is of order γ\gamma. This implies that for the computations Δ0\Delta_{0}, the self-energy can be safely neglected. Once self-energy is neglected, one can easily obtain the differential gap equation for the full V(Ωm)V(\Omega_{m}) in (91), without expanding it to first order in γ\gamma. This allows us to study the crossover between the behavior at finite and infinite Λ\Lambda.

Introducing, as before, z=(|ωm|/ω0)γz=(|\omega_{m}|/\omega_{0})^{\gamma} and anticipating that relevant ωmω0\omega_{m}\gg\omega_{0}, we find that the equation for Δdiff(z)\Delta_{{\text{d}iff}}(z) for z1z\gg 1 and V(Ωm)V(\Omega_{m}) given by (91), is the same as when Λ=\Lambda=\infty, i.e.,

(Δdiff(z)z)′′=1NγΔdiffz2,\left(\Delta_{{\text{d}iff}}(z)z\right)^{\prime\prime}=-\frac{1}{N\gamma}\frac{\Delta_{{\text{d}iff}}}{z^{2}}, (98)

However, now Δdiff(z)\Delta_{{\text{d}iff}}(z) has to satisfy the boundary condition Δdiff(Λ)=0\Delta_{{\text{d}iff}}(\Lambda^{*})=0, where Λ=(Λ/ω0)γ\Lambda^{*}=(\Lambda/\omega_{0})^{\gamma}. The proper solution of (98) is

Δdiff(z)1z×\displaystyle\Delta_{{\text{d}iff}}(z)\propto\frac{1}{\sqrt{z}}\times
[J1(2(Nγz)1/2)Y1(2(NγΛ)1/2)Y1(2(Nγz)1/2)J1(2(NγΛ)1/2)].\displaystyle\left[J_{1}\left(\frac{2}{(N\gamma z)^{1/2}}\right)Y_{1}\left(\frac{2}{(N\gamma\Lambda^{*})^{1/2}}\right)-Y_{1}\left(\frac{2}{(N\gamma z)^{1/2}}\right)J_{1}\left(\frac{2}{(N\gamma\Lambda^{*})^{1/2}}\right)\right]. (99)

In the limit NγΛ1N\gamma\Lambda^{*}\ll 1, we use

J1(x)2πxcos(x3π/4),\displaystyle J_{1}(x)\approx\sqrt{\frac{2}{\pi x}}\cos{(x-3\pi/4)},
Y1(x)2πxsin(x3π/4)\displaystyle Y_{1}(x)\approx\sqrt{\frac{2}{\pi x}}\sin{(x-3\pi/4)} (100)

and obtain

Δdiff(z)1z1/4sin(2(NγΛ)1/22(Nγz)1/2)\Delta_{{\text{d}iff}}(z)\propto\frac{1}{z^{1/4}}\sin{\left(\frac{2}{(N\gamma\Lambda^{*})^{1/2}-\frac{2}{(N\gamma z)^{1/2}}}\right)} (101)

In original variables, ωm\omega_{m} and Λ\Lambda, this reduces, to the leading order in γ\gamma, to

Δdiff(ωm)C|ωm|γ/4sin(γNlog|ωm|Λ).\Delta_{{\text{d}iff}}(\omega_{m})\propto\frac{C}{|\omega_{m}|^{\gamma/4}}\sin{\left(\sqrt{\frac{\gamma}{N}}\log{\frac{|\omega_{m}|}{\Lambda}}\right)}. (102)

This coincides with (96), up to subleading terms.

In the opposite limit, NγΛ1N\gamma\Lambda^{*}\gg 1, we use J1(x1)xJ_{1}(x\ll 1)\sim x, Y1(x1)1/xY_{1}(x\ll 1)\sim 1/x, where x=2/NγΛx=2/\sqrt{N\gamma\Lambda^{*}}. Keeping only Y1(x)Y_{1}(x), we obtain from (99) that the dependence on Λ\Lambda disappears, and

Δdiff(z)1zJ1(2(Nγz)1/2)\Delta_{{\text{d}iff}}(z)\propto\frac{1}{\sqrt{z}}J_{1}\left(\frac{2}{(N\gamma z)^{1/2}}\right) (103)

This coincides with Eq. (52).

The same result can be obtained from the RG equations Metlitski et al. (2015); Wang et al. (2017a). One has to write a set of two RG equations for the two-fermion pairing vertex gg and the effective 4-fermion pairing interaction uu. Both depend on the logarithmic scale Ln=logΛ/|ωn|L_{n}=\log{\Lambda}/|\omega_{n}|. To leading order in γ\gamma, the equations for the running g(L)g(L) and u(L)u(L) are

g=γg,u=u2+gg^{\prime}=\gamma g,~{}~{}u^{\prime}=u^{2}+g (104)

The initial condition is

g(0)=(g¯Λ)γγN,u(0)=0.g(0)=\left(\frac{\bar{g}}{\Lambda}\right)^{\gamma}\frac{\gamma}{N},~{}~{}u(0)=0. (105)

The solution of (104) is g(L)=g(0)eγLg(L)=g(0)e^{\gamma L} and

u(L)=g1/2(0)eγL/2J1(2g1/2(0)γeγL/2)Y1(2g1/2(0)γ)Y1(2g1/2(0)γeγL/2)J1(2g1/2(0)γ)J0(2g1/2(0)γeγL/2)Y1(2g1/2(0)γ)Y0(2g1/2(0)γeγL/2)J1(2g1/2(0)γ)u(L)=g^{1/2}(0)e^{\gamma L/2}\frac{J_{1}\left(\frac{2g^{1/2}(0)}{\gamma}e{\gamma L/2}\right)Y_{1}\left(\frac{2g^{1/2}(0)}{\gamma}\right)-Y_{1}\left(\frac{2g^{1/2}(0)}{\gamma}e{\gamma L/2}\right)J_{1}\left(\frac{2g^{1/2}(0)}{\gamma}\right)}{J_{0}\left(\frac{2g^{1/2}(0)}{\gamma}e{\gamma L/2}\right)Y_{1}\left(\frac{2g^{1/2}(0)}{\gamma}\right)-Y_{0}\left(\frac{2g^{1/2}(0)}{\gamma}e{\gamma L/2}\right)J_{1}\left(\frac{2g^{1/2}(0)}{\gamma}\right)} (106)

For g(0)/γ2=(g¯/Λ)γ(1/Nγ)1g(0)/\gamma^{2}=({\bar{g}}/\Lambda)^{\gamma}(1/N\gamma)\gg 1 (i.e., for small γ\gamma and finite Λ\Lambda), relevant values of the arguments of Bessel and Neumann functions are large, and using (100) we find that γ\gamma cancels out, and

u(L)=g1/2(0)tan(g1/2(0)L)u(L)=g^{1/2}(0)\tan{(g^{1/2}(0)L)} (107)

The 4-fermion vertex formally diverges at the set of ωn=Λeπ/(2g1/2(0))enπ/g1/2(0)\omega^{*}_{n}=\Lambda e^{-\pi/(2g^{1/2}(0))}e^{-n\pi/g^{1/2}(0)}. Using g(0)λg(0)\approx\lambda, valid when Λ\Lambda is finite and γ0\gamma\to 0, and associating the corresponding ωn\omega^{*}_{n} with Δn(0)\Delta_{n}(0), we reproduce (97).

In the opposite limit, g(0)/γ21g(0)/\gamma^{2}\ll 1, valid when γ\gamma is kept finite and Λ\Lambda is set to be sufficiently large, we again use that at small xx, Y1(x)J1(x)Y_{1}(x)\gg J_{1}(x). Keeping only Y1(2g01/2/γ)Y_{1}(2g^{1/2}_{0}/\gamma) in (106), we obtain

u(L)=g1/2(0)eγL/2J1(2g1/2(0)γeγL/2)J0(2g1/2(0)γeγL/2)u(L)=g^{1/2}(0)e^{\gamma L/2}\frac{J_{1}\left(\frac{2g^{1/2}(0)}{\gamma}e{\gamma L/2}\right)}{J_{0}\left(\frac{2g^{1/2}(0)}{\gamma}e{\gamma L/2}\right)} (108)

The 4-fermion vertex now diverges at the set of zeros of J0(p)J_{0}(p), where

p=2g1/2(0)γ(Λ|ωm|)γ/2=2(Nγz)1/2p=\frac{2g^{1/2}(0)}{\gamma}\left(\frac{\Lambda}{|\omega_{m}|}\right)^{\gamma/2}=\frac{2}{(N\gamma z)^{1/2}} (109)

where, as before, z=(ω0/|ωm|)γz=(\omega_{0}/|\omega_{m}|)^{\gamma} (we used that for small γ\gamma, g¯γ=ω0γ(1γ)ω0γ{\bar{g}}^{\gamma}=\omega^{\gamma}_{0}(1-\gamma)\approx\omega^{\gamma}_{0}. We see that now Λ\Lambda cancels out, and the condition J0[2/(Nγz)1/2]=0J_{0}[2/(N\gamma z)^{1/2}]=0 gives the same set of znz^{*}_{n} as for the case when Λ\Lambda is infinite.

Appendix B Differential gap equation for arbitrary γ\gamma.

The derivation of the differential gap equation, Eq. (43), from the original integral equation is only justified for small γ\gamma. In this appendix we modify the derivation and show that for N=1N=1 the modified gap function Δdiff(ωm)\Delta_{\text{diff}}(\omega_{m}) shows qualitatively the same behavior as the exact Δ(ωm)\Delta(\omega_{m}) even for γ1\gamma\leq 1.

Specifically, we add to the r.h.s of (43) the additional contribution from ωω\omega^{\prime}\sim\omega, where the interaction is strongly peaked at γ1\gamma\to 1. There is some uncertainty with this contribution as we have to specify the range of integration around ω=ω\omega^{\prime}=\omega. We choose this range to be |ωω|aγω|\omega^{\prime}-\omega|\leq a_{\gamma}\omega and keep aγa_{\gamma} as a parameter. We assume that for small γ\gamma, aγ0a_{\gamma}\to 0, but for γ1\gamma\leq 1, aγ=O(1)a_{\gamma}=O(1). Adding this contribution to the r.h.s. of Eq. (43) we find that it becomes

Φ(z)d+z1+z=1γNγ[z𝑑yΦ(y)y(1+y)+1z0z𝑑yΦ(y)1+y]\Phi(z)\frac{d+z}{1+z}=\frac{1-\gamma}{N\gamma}\left[\int^{\infty}_{z}dy\frac{\Phi(y)}{y(1+y)}+\frac{1}{z}\int_{0}^{z}dy\frac{\Phi(y)}{1+y}\right] (110)

where d=1aγ1γ/Nd=1-a^{1-\gamma}_{\gamma}/N. Introducing Δ\Delta instead of Φ\Phi, rescaling zz to z¯=z/d{\bar{z}}=z/d and differentiating twice over z¯{\bar{z}} we obtain differential gap equation in the same form as in (47)

(Δdiff(z¯)(1+z¯))′′=(βN2+14)Δdiff(z¯)z¯2,(\Delta_{{\text{diff}}}({\bar{z}})(1+{\bar{z}}))^{{}^{\prime\prime}}=-\left(\beta^{2}_{N}+\frac{1}{4}\right)\frac{\Delta_{{\text{diff}}}({\bar{z}})}{{\bar{z}}^{2}}, (111)

once we identify

1γNγaγ1γ=βN2+14\frac{1-\gamma}{N\gamma-a_{\gamma}^{1-\gamma}}=\beta^{2}_{N}+\frac{1}{4} (112)

Hence, the solution is given by Eq. (53) with z¯{\bar{z}} instead of zz.

Eq. (112) determines aγa_{\gamma}. At small γ\gamma, βN2+1/41/(Nγ)\beta^{2}_{N}+1/4\approx 1/(N\gamma), hence, aa is small and z¯z{\bar{z}}\approx z. However, for larger γ\gamma, aγ=O(1)a_{\gamma}=O(1), and the rescaling zz¯z\to{\bar{z}} is essential. Fig. 12 shows that it brings Δdiff(z)\Delta_{{\text{diff}}}(z) closer to the exact Δex(z)\Delta_{ex}(z).

We note in this regard that for N=1N=1 and γ1\gamma\to 1, aγ1γ/N1+(1γ)logaγa^{1-\gamma}_{\gamma}/N\approx 1+(1-\gamma)\log{a_{\gamma}}, and the r.h.s. of (112) becomes of order one if we choose a=O(1)a=O(1). This is consistent with the fact that βN=1\beta_{N=1} tends to a finite value βN=10.79\beta_{N=1}\approx 0.79 at γ1\gamma\to 1. Also, the rescaling factor d=1a1γ/N(1γ)logaγd=1-a^{1-\gamma}/N\approx-(1-\gamma)\log{a_{\gamma}}, hence, for a=O(1)a=O(1), z¯=z/d{\bar{z}}=z/d becomes of order |ωm|/g|\omega_{m}|/g. As the consequence, the modified Δdiff(ωm)\Delta_{\text{diff}}(\omega_{m}) evolves at a non-singular ωmg\omega_{m}\sim g. This is consistent with the behavior of Δex(ωm)\Delta_{\text{ex}}(\omega_{m}) at γ1\gamma\to 1 and N=1N=1 (see Fig. 12(e)).

Appendix C The exact solution of the linearized gap equation – the computational details.

In this appendix we present the full details of the computations, leading to the formula for the exact solution of the equation (11). The presentation is self-contained in the sense that it does not use specific notations, introduced in the main text.

C.1 Reformulation of Eq. (11).

We write Eq. (11) as

EΦ(Ω)=1γ2Φ(ω)dω|ωΩ|γ|ω|1γ11+|ω|γE\Phi(\Omega)=\frac{1-\gamma}{2}\int_{-\infty}^{\infty}\frac{\Phi(\omega)d\omega}{|\omega-\Omega|^{\gamma}|\omega|^{1-\gamma}}\frac{1}{1+|\omega|^{\gamma}} (113)

and use the latter EE instead of NN to emphasize that this is a continuous variable.222The method, presented in this appendix, will also work for a generalized version of Eq. (113) with 11+|ω|α\frac{1}{1+|\omega|^{\alpha}} instead of 11+|ω|γ\frac{1}{1+|\omega|^{\gamma}}.

We introduce a set of functions

Φβ(Ω)=|Ω|2iβ+δΩ|Ω|γ/2\Phi_{\beta}(\Omega)=\frac{|\Omega|^{2i\beta+\delta_{\Omega}}}{|\Omega|^{\gamma/2}} (114)

where β\beta changes from -\infty to ++\infty and δΩ=+0sign(1|Ω|)\delta_{\Omega}=+0\mbox{sign}(1-|\Omega|) is a convergence factor. It is clear that

Φβ(ω)Φβ(Ω)dβ2π=12|Ω|1γδ(|Ω||ω|).\int\Phi_{\beta}(\omega)\Phi_{\beta}^{\ast}(\Omega)\frac{d\beta}{2\pi}=\frac{1}{2}|\Omega|^{1-\gamma}\delta(|\Omega|-|\omega|).

We now denote

aβ=12dΩ|Ω|1γΦβ(Ω)Φ(Ω),a_{\beta}=\frac{1}{2}\int\frac{d\Omega}{|\Omega|^{1-\gamma}}\Phi_{\beta}(\Omega)\Phi(\Omega), (115)

multiply (113) by Φβ(Ω)\Phi_{\beta}(\Omega), and integrate it over dΩ|Ω|1γ\frac{d\Omega}{|\Omega|^{1-\gamma}}. We obtain

Eaβ=1γ212𝑑ωϕ(ω)|ω|1γ11+|ω|γΦβ(Ω)dΩ|Ω|1γ|Ωω|1γ=ϵβ2𝑑ωΦ(ω)Φβ(ω)|ω|1γ11+|ω|γ,Ea_{\beta}=\frac{1-\gamma}{2}\frac{1}{2}\int d\omega\frac{\phi(\omega)}{|\omega|^{1-\gamma}}\frac{1}{1+|\omega|^{\gamma}}\int\frac{\Phi_{\beta}(\Omega)d\Omega}{|\Omega|^{1-\gamma}|\Omega-\omega|^{1-\gamma}}=\frac{\epsilon_{\beta}}{2}\int d\omega\frac{\Phi(\omega)\Phi_{\beta}(\omega)}{|\omega|^{1-\gamma}}\frac{1}{1+|\omega|^{\gamma}},

where

ϵβ=1γ2|Γ(γ/2+2iβ)|2Γ(γ)(1+cosh(2πβ)cos(πγ/2)).\epsilon_{\beta}=\frac{1-\gamma}{2}\frac{|\Gamma(\gamma/2+2i\beta)|^{2}}{\Gamma(\gamma)}\left(1+\frac{\cosh(2\pi\beta)}{\cos(\pi\gamma/2)}\right). (116)

From the definition (115) (we also use the fact that Φ(ω)\Phi(\omega) is taken to be an even function of frequency) we find

aβΦβ(ω)dβ2π=12dΩ|Ω|1γΦ(Ω)Φβ(Ω)Φβ(ω)dβ(2π)=12Φ(ω).\int a_{\beta}\Phi_{\beta}^{\ast}(\omega)\frac{d\beta}{2\pi}=\frac{1}{2}\int\frac{d\Omega}{|\Omega|^{1-\gamma}}\Phi(\Omega)\int\Phi_{\beta}(\Omega)\Phi_{\beta}^{\ast}(\omega)\frac{d\beta}{(2\pi)}=\frac{1}{2}\Phi(\omega).

We then have

Eaβ=ϵβdβ2πaβ𝑑ωΦβ(ω)Φβ(ω)|ω|1γ11+|ω|γ.Ea_{\beta}=\epsilon_{\beta}\int\frac{d\beta^{\prime}}{2\pi}a_{\beta^{\prime}}\int d\omega\frac{\Phi_{\beta^{\prime}}^{\ast}(\omega)\Phi_{\beta}(\omega)}{|\omega|^{1-\gamma}}\frac{1}{1+|\omega|^{\gamma}}.

We now compute the integral

𝑑ωΦβ(ω)Φβ(ω)|ω|1γ11+|ω|α=20dωωω2i(ββ)+δω1+ωγ=2γ𝑑xeix2γ(ββ)+xδx1+ex\int d\omega\frac{\Phi_{\beta^{\prime}}^{\ast}(\omega)\Phi_{\beta}(\omega)}{|\omega|^{1-\gamma}}\frac{1}{1+|\omega|^{\alpha}}=2\int_{0}^{\infty}\frac{d\omega}{\omega}\frac{\omega^{2i(\beta-\beta^{\prime})+\delta_{\omega}}}{1+\omega^{\gamma}}=\frac{2}{\gamma}\int_{-\infty}^{\infty}dx\frac{e^{ix\frac{2}{\gamma}(\beta-\beta^{\prime})+x\delta_{x}}}{1+e^{x}}

where δx=0sign(x)\delta_{x}=-0\mbox{sign}(x). At xx\rightarrow\infty the integral perfectly converges with or without the convergence factor. At xx\rightarrow-\infty we do need the convergence factor, and we assume that δx\delta_{x} is a small positive number. Evaluating the integral we then obtain

2γ𝑑xeix2γ(ββi0)1+ex=2γiπsinh(2πγ(ββi0)),\frac{2}{\gamma}\int_{-\infty}^{\infty}dx\frac{e^{ix\frac{2}{\gamma}(\beta-\beta^{\prime}-i0)}}{1+e^{x}}=-\frac{2}{\gamma}\frac{i\pi}{\sinh\left(\frac{2\pi}{\gamma}(\beta-\beta^{\prime}-i0)\right)}, (117)

and

Eaβ=iϵβ1γaβdβsinh(2πγ(ββi0)).Ea_{\beta}=-i\epsilon_{\beta}\frac{1}{\gamma}\int_{-\infty}^{\infty}\frac{a_{\beta^{\prime}}d\beta^{\prime}}{\sinh\left(\frac{2\pi}{\gamma}(\beta-\beta^{\prime}-i0)\right)}.

We now substitute 2β/γ=k2\beta/\gamma=k, and use aka_{k} instead of aβa_{\beta}. Then,

Eak=i2ϵγk/2akdksinh(π(kki0))Ea_{k}=-\frac{i}{2}\epsilon_{\gamma k/2}\int_{-\infty}^{\infty}\frac{a_{k^{\prime}}dk^{\prime}}{\sinh\left(\pi(k-k^{\prime}-i0)\right)} (118)

Introducing the new function bkb_{k} via

ak=bkϵγk/2,a_{k}=b_{k}\epsilon_{\gamma k/2}, (119)

we obtain from (118) the equation for bkb_{k} in the form

Ebk=i2ϵγk/2sinh(π(kk+i0))bk𝑑kEb_{k}=\frac{i}{2}\int_{-\infty}^{\infty}\frac{\epsilon_{\gamma k^{\prime}/2}}{\sinh\left(\pi(k^{\prime}-k+i0)\right)}b_{k^{\prime}}dk^{\prime} (120)

C.2 The functional F[ϕ]F[\phi].

For an infinitesimally small Φ(ω)\Phi(\omega), the free energy difference between the states with Φ(ω)=0\Phi(\omega)=0 and Φ(ω)0\Phi(\omega)\neq 0 is given by

F[Φ]N0ω0=12Φω2|ω|1γ(1+|ω|γ)𝑑ω1γ4EΦωΦωdωdω|ω|1γ(1+|ω|γ)|ωω|γ|ω|1γ(1+|ω|γ),\frac{F[\Phi]}{N_{0}\omega_{0}}=\frac{1}{2}\int\frac{\Phi_{\omega}^{2}}{|\omega|^{1-\gamma}\left(1+|\omega|^{\gamma}\right)}d\omega-\frac{1-\gamma}{4E}\int\frac{\Phi_{\omega}\Phi_{\omega^{\prime}}d\omega d\omega^{\prime}}{|\omega|^{1-\gamma}\left(1+|\omega|^{\gamma}\right)|\omega-\omega^{\prime}|^{\gamma}|\omega^{\prime}|^{1-\gamma}\left(1+|\omega^{\prime}|^{\gamma}\right)}, (121)

where we defined Φ(ωm)Φω\Phi(\omega_{m})\equiv\Phi_{\omega} to shorten the notations. Taking the first variation of this functional we obtain equation (113) This functional defines our space of functions, namely we must only consider the functions for which F[ϕ]F[\phi] is finite.

C.3 Normalizability condition.

Now, let us take ΦωE\Phi_{\omega}^{E} to be the solution of (113) with E1E\not=1. Then

F[Φωϵ]=N0g¯2(1γ)2/γ1E2Φω2|ω|1γ(1+|ω|γ)𝑑ω.F[\Phi_{\omega}^{\epsilon}]=N_{0}\bar{g}^{2}(1-\gamma)^{-2/\gamma}\frac{1-E}{2}\int_{-\infty}^{\infty}\frac{\Phi_{\omega}^{2}}{|\omega|^{1-\gamma}\left(1+|\omega|^{\gamma}\right)}d\omega. (122)

We use

Φω=2aβΦβ(ω)dβ2π,\Phi_{\omega}=2\int a_{\beta}\Phi^{\ast}_{\beta}(\omega)\frac{d\beta}{2\pi},

where Φβ(ω)\Phi_{\beta}(\omega) is defined in (114). According to (117), we have

Φω2dω|ω|1γ(1+|ω|γ)=4iπγaβaβdβdβsinh[2πγ(ββi0)].\int\frac{\Phi_{\omega}^{2}d\omega}{|\omega|^{1-\gamma}\left(1+|\omega|^{\gamma}\right)}=-\frac{4i}{\pi\gamma}\int\frac{a_{\beta}a_{\beta^{\prime}}^{\ast}d\beta d\beta^{\prime}}{\sinh\left[\frac{2\pi}{\gamma}(\beta-\beta^{\prime}-i0)\right]}.

Substituting into (122) and using

k=2β/γ,k=2\beta/\gamma,

we obtain

F[akE]N0ω0=γ(1E)i2πakakdkdksinh[π(kki0)].\frac{F[a_{k}^{E}]}{N_{0}\omega_{0}}=-\gamma(1-E)\frac{i}{2\pi}\int\frac{a_{k}a_{k^{\prime}}^{\ast}dkdk^{\prime}}{\sinh\left[\pi(k-k^{\prime}-i0)\right]}. (123)

Using (119), we rewrite (123) in terms of bkb_{k} as

F[bkE]N0ω0=γ(1E)iπϵγk/2ϵγk/2sinh[π(kki0)]bkbk𝑑k𝑑k.\frac{F[b_{k}^{E}]}{N_{0}\omega_{0}}=-\gamma(1-E)\frac{i}{\pi}\int\frac{\epsilon_{\gamma k/2}\epsilon_{\gamma k^{\prime}/2}}{\sinh\left[\pi(k-k^{\prime}-i0)\right]}b_{k}b_{k^{\prime}}^{\ast}dkdk^{\prime}.

It can be re-expressed as

F[bkE]N0ω0=γ(1E)iπϵγk/2ϵγk/2[1sinh[π(kk+i0)]+2iδ(kk)]bkbk𝑑k𝑑k.\frac{F[b_{k}^{E}]}{N_{0}\omega_{0}}=-\gamma(1-E)\frac{i}{\pi}\int\epsilon_{\gamma k/2}\epsilon_{\gamma k^{\prime}/2}\left[\frac{1}{\sinh\left[\pi(k-k^{\prime}+i0)\right]}+2i\delta(k-k^{\prime})\right]b_{k}b_{k^{\prime}}^{\ast}dkdk^{\prime}.

Comparing to (120), we finally obtain

F[bkE]N0g¯2(1γ)2/γγ=E(1E)π[11Eϵγk/2]ϵγk/2bkbk𝑑k\frac{F[b_{k}^{E}]}{N_{0}\bar{g}^{2}(1-\gamma)^{-2/\gamma}\gamma}=-\frac{E(1-E)}{\pi}\int\left[1-\frac{1}{E}\epsilon_{\gamma k^{\prime}/2}\right]\epsilon_{\gamma k^{\prime}/2}b_{k^{\prime}}b_{k^{\prime}}^{\ast}dk^{\prime} (124)

Eq. (124) defines a norm for our solutions. If we manage to solve Eq. (120) and find bkb_{k} for a given EE, we will need to verify whether for these bkb_{k}, F[bkE]F[b_{k}^{E}] is finite. This will give us the spectrum.

C.4 Solution of Eq. (120)

We now show that the normalizable solution of Eq. (120) exists for all E<ϵ0E<\epsilon_{0}. We obtain bkb_{k} and use them to obtain the pairing vertex Φ(ωm)\Phi(\omega_{m}) and the gap function Δ(ωm)\Delta(\omega_{m}). For this purpose we need to first find symmetry properties of bkb_{k} in (120).

C.4.1 bkb_{k} vs bkb_{-k}.

The equation (113) is real. So, (i) if an eigenfunction is complex, its complex conjugated must also be an eigenfunction with the same eigenvalue. The eigen value then is double degenerate; (ii) If an eigenvalue is non-degenerate, then the eigenfunction must be real (more precisely, it may have a trivial ω\omega independent phase.) If the eigenfunction of Eq. (113) is real, then, according to Eqs. (114) and (115), we have

ak=aka_{k}=a^{\ast}_{-k}

As the function ϵγk/2\epsilon_{\gamma k/2} is real and symmetric under the change kkk\rightarrow-k we have

bk=bkb_{k}=b^{\ast}_{-k} (125)

C.4.2 Periodicity.

Equation (120) has an obvious, but important property. If we change kk+ik\rightarrow k+i the r.h.s will turn into itself, but with the opposite sign. So we must have

bk+in=(1)nbk.b_{k+in}=(-1)^{n}b_{k}. (126)

It means, in particular, that a solution of (120) is periodic with the period 2i2i.

C.4.3 Analytical properties.

Let us consider Eq. (120) for arbitrary complex kk, keeping kk^{\prime} to be real. This way we define the analytical function bkb_{k}. The integral perfectly converges as long as kin\Im k\not=in for integer nn. So the function bkb_{k} can only have singularities on the lines k=in\Im k=in. It cannot have poles on these lines, as in this case the integral in the r.h.s. of Eq. (120) would not converge.

Let us return to Eq. (120) and use it to define an analytic function BkB_{k}:

Bk=i2Eϵγk/2sinh(π(kk))bk𝑑k=i2πEn=(1)nϵγk/2bkkk+in𝑑kB_{k}=\frac{i}{2E}\int_{-\infty}^{\infty}\frac{\epsilon_{\gamma k^{\prime}/2}}{\sinh\left(\pi(k^{\prime}-k)\right)}b_{k^{\prime}}dk^{\prime}=\frac{i}{2\pi E}\sum_{n=-\infty}^{\infty}(-1)^{n}\int_{-\infty}^{\infty}\frac{\epsilon_{\gamma k^{\prime}/2}b_{k^{\prime}}}{k^{\prime}-k+in}dk^{\prime}

(we assume E0E\not=0). The function bkb_{k} is related to BkB_{k} as

bk=limδ0Bkiδ,k=0,δ>0.b_{k}=\lim_{\delta\rightarrow 0}B_{k-i\delta},\qquad\Im k=0,\quad\delta>0. (127)

Using the fact that both ϵγk/2\epsilon_{\gamma k/2} and 1/sinh(πk)1/\sinh(\pi k) are analytic in narrow strip below the real axis, we can express BkB_{k} as

Bk=i2Eiδ~iδ~ϵγk/2sinh(π(kk))Bk𝑑k,δ~>0B_{k}=\frac{i}{2E}\int_{-\infty-i\tilde{\delta}}^{\infty-i\tilde{\delta}}\frac{\epsilon_{\gamma k^{\prime}/2}}{\sinh\left(\pi(k^{\prime}-k)\right)}B_{k^{\prime}}dk^{\prime},\qquad\tilde{\delta}>0 (128)

This formula must be understood in the sense that to compute bk=limδ0Bkiδb_{k}=\lim_{\delta\rightarrow 0}B_{k-i\delta}, for real kk, we must first take the limit δ~0\tilde{\delta}\rightarrow 0 and then δ0\delta\rightarrow 0, in which case kk in the r.h.s. of (128) is below the integration path. The function BkB_{k} obviously satisfies the periodicity condition, Eq. (126). It is also clear from (128) that BkB_{k} has branch cuts along the lines k=in\Im k=in.

Let us now compute limδ0(Bk+in+iδBk+iniδ)\lim_{\delta\rightarrow 0}(B_{k+in+i\delta}-B_{k+in-i\delta}), for real kk and positive δ\delta. By Sokhotski-Plemelj theorem: 1xx0i0=P1xx0+iπδ(xx0)\frac{1}{x-x_{0}-i0}=\mbox{P}\frac{1}{x-x_{0}}+i\pi\delta(x-x_{0}). Hence

limδ0(Bk+in+iδBk+iniδ)=(1)n1Eϵλk/2bk=1Eϵλk/2Bk+iniδ,δ0,δ>0\lim_{\delta\rightarrow 0}(B_{k+in+i\delta}-B_{k+in-i\delta})=-(-1)^{n}\frac{1}{E}\epsilon_{\lambda k/2}b_{k}=-\frac{1}{E}\epsilon_{\lambda k/2}B_{k+in-i\delta},\qquad\delta\rightarrow 0,\quad\delta>0

which means that

Bk+in+i0=(11Eϵγk/2)Bk+ini0.B_{k+in+i0}=\left(1-\frac{1}{E}\epsilon_{\gamma k/2}\right)B_{k+in-i0}. (129)

This equation is a particular case of the Riemann-Hilbert problem.

Here, one has to distinguish between the following two cases

  • (i)] E>maxϵγk/2=ϵ0E>\mbox{max}\epsilon_{\gamma k/2}=\epsilon_{0} – in this case the expression 11Eϵγk/21-\frac{1}{E}\epsilon_{\gamma k/2} does not change sign, and

  • (ii)] 0<E<ϵ00<E<\epsilon_{0} – in this case the expression 11Eϵγk/21-\frac{1}{E}\epsilon_{\gamma k/2} changes sign twice at k=±kEk=\pm k_{E}, where

ϵγkE/2=E\epsilon_{\gamma k_{E}/2}=E (130)

We consider the two cases separately.

Case (i). E>ϵ0E>\epsilon_{0}: In this case we take the logarithms of the both sides of the equation (129), we then get

logBk+in+i0logBk+ini0=log(11Eϵγk/2)\log B_{k+in+i0}-\log B_{k+in-i0}=\log\left(1-\frac{1}{E}\epsilon_{\gamma k/2}\right)

So the analytic function Φk=logBk\Phi_{k}=\log B_{k} has branch cuts along the lines k=in\Im k=in with the jumps given by above formula. According to Sokhotski-Plemelj theorem we have

logBk=12πin=log(11Eϵγk/2)kk+in𝑑k+Fk=12ilog(11Eϵγk/2)coth(π(kk))𝑑k+Fk,\log B_{k}=\frac{1}{2\pi i}\sum_{n=-\infty}^{\infty}\int_{-\infty}^{\infty}\frac{\log(1-\frac{1}{E}\epsilon_{\gamma k^{\prime}/2})}{k^{\prime}-k+in}dk^{\prime}+F_{k}=\frac{1}{2i}\int_{-\infty}^{\infty}\log(1-\frac{1}{E}\epsilon_{\gamma k^{\prime}/2})\coth\left(\pi(k^{\prime}-k)\right)dk^{\prime}+F_{k},

where the analytic function FkF_{k} has no branch cuts or singularities, so it must be a polynomial. In order to find this polynomial, we notice, that the first term in the right hand side does not change if we shift kk+ink\rightarrow k+in, while BkB_{k} must acquire a factor of (1)n(-1)^{n} and logBk\log B_{k} must acquire extra ±iπn\pm i\pi n. There are only two polynomials that have this property, they are are ±πk\pm\pi k, so there are two solutions:

Bk=e±πk+12ilog(11Eϵγk/2)coth(π(kk))𝑑k,B_{k}=e^{\pm\pi k+\frac{1}{2i}\int_{-\infty}^{\infty}\log\left(1-\frac{1}{E}\epsilon_{\gamma k^{\prime}/2}\right)\coth\left(\pi(k^{\prime}-k)\right)dk^{\prime}}, (131)

Correspondingly, there are two solutions for bkb_{k}

bk±=Bki0=e±πk+12ilog(11Eϵγk/2)coth(π(kk+i0))𝑑k,b_{k}^{\pm}=B_{k-i0}=e^{\pm\pi k+\frac{1}{2i}\int_{-\infty}^{\infty}\log\left(1-\frac{1}{E}\epsilon_{\gamma k^{\prime}/2}\right)\coth\left(\pi(k^{\prime}-k+i0)\right)dk^{\prime}}, (132)

where kk is now real.

Notice, that we have two solutions. One can check, that they are indeed complex conjugated.

Now we need to check whether the solutions (132) are normalizable. From (132) we find

bkbk=e±2πke12ilog(11Eϵγk/2)[coth(π(kk+i0))coth(π(kki0))]𝑑k\displaystyle b_{k}b_{k}^{\ast}=e^{\pm 2\pi k}e^{\frac{1}{2i}\int_{-\infty}^{\infty}\log\left(1-\frac{1}{E}\epsilon_{\gamma k^{\prime}/2}\right)\left[\coth\left(\pi(k^{\prime}-k+i0)\right)-\coth\left(\pi(k^{\prime}-k-i0)\right)\right]dk^{\prime}}
=e±2πke12ilog(11Eϵγk/2)2iδ(kk)𝑑k=cosh2(πk)elog(11Eϵγk/2)\displaystyle=e^{\pm 2\pi k}e^{-\frac{1}{2i}\int_{-\infty}^{\infty}\log\left(1-\frac{1}{E}\epsilon_{\gamma k^{\prime}/2}\right)2i\delta(k-k^{\prime})dk^{\prime}}=\cosh^{2}(\pi k)e^{-\log\left(1-\frac{1}{E}\epsilon_{\gamma k/2}\right)}

As the consequence,

F[bkE]N0g¯2(1γ)2/γγ=E(1E)πe2πkϵγk/2𝑑k\frac{F[b_{k}^{E}]}{N_{0}\bar{g}^{2}(1-\gamma)^{-2/\gamma}\gamma}=-\frac{E(1-E)}{\pi}\int e^{2\pi k}\epsilon_{\gamma k/2}dk

The last integral diverges. The divergence implies that the solutions with E>ϵ0E>\epsilon_{0} are not normalizable and are not the part of the spectrum.

Case (ii). 0<E<ϵ00<E<\epsilon_{0}. In this case there are two points (130) where the r.h.s of (129) changes sign. To avoid ambiguity of log(1)\log(-1), we first rewrite (129) as

Bk+in+i0=Θ(kkE)Θ(k+kE)|11Eϵγk/2|Bk+ini0,B_{k+in+i0}=\Theta(k-k_{E})\Theta(k+k_{E})\left|1-\frac{1}{E}\epsilon_{\gamma k/2}\right|B_{k+in-i0},

where Θ\Theta is the step function (Θ(x<0)=1\Theta(x<0)=-1 and Θ(x>0)=1\Theta(x>0)=1) Now we take the log of both sides

logBk+in+i0logBk+ini0=log|11Eϵγk/2|+logΘ(kkE)+logΘ(k+kE).\log B_{k+in+i0}-\log B_{k+in-i0}=\log\left|1-\frac{1}{E}\epsilon_{\gamma k/2}\right|+\log\Theta(k-k_{E})+\log\Theta(k+k_{E}).

The logΘ(kkE)\log\Theta(k-k_{E}) is zero for k>kEk>k_{E} and either +iπ+i\pi or iπ-i\pi for k<kEk<k_{E}. Let us introduce two yet unknown functions χn\chi_{n} and ξn\xi_{n}, which have values ±1\pm 1, depending on nn, and rewrite the equation above as

logBk+in+i0logBk+ini0=log|11Eϵγk/2|+iπχn12(1Θ(kkE))+iπξn12(1Θ(k+kE)).\log B_{k+in+i0}-\log B_{k+in-i0}=\log\left|1-\frac{1}{E}\epsilon_{\gamma k/2}\right|+i\pi\chi_{n}\frac{1}{2}(1-\Theta(k-k_{E}))+i\pi\xi_{n}\frac{1}{2}(1-\Theta(k+k_{E})).

The Sokhotski-Plemelj theorem now gives

logBk=12πin=log|11Eϵγk/2|kk+in𝑑k\displaystyle\log B_{k}=\frac{1}{2\pi i}\sum_{n=-\infty}^{\infty}\int_{-\infty}^{\infty}\frac{\log|1-\frac{1}{E}\epsilon_{\gamma k^{\prime}/2}|}{k^{\prime}-k+in}dk^{\prime}
+12n=χn12(1Θ(kkE))kk+in𝑑k+12n=ξn12(1Θ(k+kE))kk+in𝑑k+Fk.\displaystyle+\frac{1}{2}\sum_{n=-\infty}^{\infty}\chi_{n}\int_{-\infty}^{\infty}\frac{\frac{1}{2}(1-\Theta(k-k_{E}))}{k^{\prime}-k+in}dk^{\prime}+\frac{1}{2}\sum_{n=-\infty}^{\infty}\xi_{n}\int_{-\infty}^{\infty}\frac{\frac{1}{2}(1-\Theta(k+k_{E}))}{k^{\prime}-k+in}dk^{\prime}+F_{k}.

This can be re-expressed as

logBk=12ilog|11Eϵγk/2|coth(π(kk))dk\displaystyle\log B_{k}=\frac{1}{2i}\int_{-\infty}^{\infty}\log|1-\frac{1}{E}\epsilon_{\gamma k^{\prime}/2}|\coth\left(\pi(k^{\prime}-k)\right)dk^{\prime}
+12kEn=χnkk+indk+12kEn=ξnkk+indk+Fk.\displaystyle+\frac{1}{2}\int_{-\infty}^{k_{E}}\sum_{n=-\infty}^{\infty}\frac{\chi_{n}}{k^{\prime}-k+in}dk^{\prime}+\frac{1}{2}\int_{-\infty}^{-k_{E}}\sum_{n=-\infty}^{\infty}\frac{\xi_{n}}{k^{\prime}-k+in}dk^{\prime}+F_{k}.

According to (126), the difference logBk+imlogBk\log B_{k+im}-\log B_{k} should be equal to iπmi\pi m independent on kk. For BkB_{k} from (C.4.3) we have

logBk+imlogBk=12kEn=χn+mχnkk+indk+12kEn=ξn+mξnkk+indk+Fk+imFk.\log B_{k+im}-\log B_{k}=\frac{1}{2}\int_{-\infty}^{k_{E}}\sum_{n=-\infty}^{\infty}\frac{\chi_{n+m}-\chi_{n}}{k^{\prime}-k+in}dk^{\prime}+\frac{1}{2}\int_{-\infty}^{-k_{E}}\sum_{n=-\infty}^{\infty}\frac{\xi_{n+m}-\xi_{n}}{k^{\prime}-k+in}dk^{\prime}+F_{k+im}-F_{k}.

In order for this expression to be independent of kk, the functions χn\chi_{n} and ξn\xi_{n} must be independent of nn and FkF_{k} must be equal to ±πk+c\pm\pi k+c, where cc is an arbitrary complex constant. We then denote χn=χ\chi_{n}=\chi, ξn=ξ\xi_{n}=\xi, where χ=±1\chi=\pm 1, ξ=±1\xi=\pm 1. The signs of χ\chi and ξ\xi can be chosen independently. After this, we obtain

logBk=12ilog|11Eϵγk/2|coth(π(kk))dk\displaystyle\log B_{k}=\frac{1}{2i}\int_{-\infty}^{\infty}\log|1-\frac{1}{E}\epsilon_{\gamma k^{\prime}/2}|\coth\left(\pi(k^{\prime}-k)\right)dk^{\prime}
+χπ2kEcoth(π(kk))𝑑k+ξπ2kEcoth(π(kk))𝑑k±πk+c\displaystyle+\chi\frac{\pi}{2}\int_{-\infty}^{k_{E}}\coth(\pi(k^{\prime}-k))dk^{\prime}+\xi\frac{\pi}{2}\int_{-\infty}^{-k_{E}}\coth(\pi(k^{\prime}-k))dk^{\prime}\pm\pi k+c (133)

Now, we can find the function bkb_{k}

bk=Bki0=exp[12ilog|11Eϵγk/2|coth(π(kk+i0))dk±πk+c\displaystyle b_{k}=B_{k-i0}=\exp\Big{[}\frac{1}{2i}\int_{-\infty}^{\infty}\log|1-\frac{1}{E}\epsilon_{\gamma k^{\prime}/2}|\coth\left(\pi(k^{\prime}-k+i0)\right)dk^{\prime}\pm\pi k+c
+χπ2kE[coth(π(kk+i0))coth(π(k+i0))]𝑑k\displaystyle+\chi\frac{\pi}{2}\int_{-\infty}^{k_{E}}\left[\coth(\pi(k^{\prime}-k+i0))-\coth(\pi(k^{\prime}+i0))\right]dk^{\prime}
+ξπ2kE[coth(π(kk+i0))coth(πk)]dk],\displaystyle+\xi\frac{\pi}{2}\int_{-\infty}^{-k_{E}}\left[\coth(\pi(k^{\prime}-k+i0))-\coth(\pi k^{\prime})\right]dk^{\prime}\Big{]}, (134)

(we redefined the constant cc)

Now, we need to check if the solution (134) is normalizable. For this, we need to verify whether the integral in (124) converges at large |k||k|. This requires us to compute the real part of the exponent in (134) for |k|kE|k|\gg k_{E}. We do this on term by term basis

[12ilog|11Eϵγk/2|coth(π(kk+i0))dk±πk+c]=12log|11Eϵγk/2|±πk+c,\displaystyle\Re\left[\frac{1}{2i}\int_{-\infty}^{\infty}\log|1-\frac{1}{E}\epsilon_{\gamma k^{\prime}/2}|\coth\left(\pi(k^{\prime}-k+i0)\right)dk^{\prime}\pm\pi k+c\right]=-\frac{1}{2}\log|1-\frac{1}{E}\epsilon_{\gamma k/2}|\pm\pi k+\Re c,
[χπ2kE[coth(π(kk+i0))coth(π(k+i0))]𝑑k]=χ2log|sinh(π(kk))sinh(πk)|kE\displaystyle\Re\left[\chi\frac{\pi}{2}\int_{-\infty}^{k_{E}}\left[\coth(\pi(k^{\prime}-k+i0))-\coth(\pi(k^{\prime}+i0))\right]dk^{\prime}\right]=\frac{\chi}{2}\log\left|\frac{\sinh(\pi(k^{\prime}-k))}{\sinh(\pi k^{\prime})}\right|_{-\infty}^{k_{E}}
=χ2log|sinh(π(kEk))sinh(πkE)|χπ2kχπ2(|k|k),for |k|,\displaystyle=\frac{\chi}{2}\log\left|\frac{\sinh(\pi(k_{E}-k))}{\sinh(\pi k_{E})}\right|-\chi\frac{\pi}{2}k\rightarrow\chi\frac{\pi}{2}\left(|k|-k\right),\qquad\mbox{for $|k|\rightarrow\infty$},
[ξπ2kE[coth(π(kk+i0))coth(πk)]𝑑k]=ξ2log|sinh(π(kk))sinh(πk)|kE\displaystyle\Re\left[\xi\frac{\pi}{2}\int_{-\infty}^{-k_{E}}\left[\coth(\pi(k^{\prime}-k+i0))-\coth(\pi k^{\prime})\right]dk^{\prime}\right]=\frac{\xi}{2}\log\left|\frac{\sinh(\pi(k^{\prime}-k))}{\sinh(\pi k^{\prime})}\right|_{-\infty}^{-k_{E}}
=ξ2log|sinh(π(kE+k))sinh(πkE)|ξπ2kξπ2(|k|k),for |k|,\displaystyle=\frac{\xi}{2}\log\left|\frac{\sinh(\pi(k_{E}+k))}{\sinh(\pi k_{E})}\right|-\xi\frac{\pi}{2}k\rightarrow\xi\frac{\pi}{2}\left(|k|-k\right),\qquad\mbox{for $|k|\rightarrow\infty$},

Combining, we find that the real part of the exponent in (134) is

12log|11Eϵγk/2|±πk+πχ+ξ2(|k|k)-\frac{1}{2}\log|1-\frac{1}{E}\epsilon_{\gamma k/2}|\pm\pi k+\pi\frac{\chi+\xi}{2}\left(|k|-k\right)

We see that if we chose χ=ξ=1\chi=\xi=-1, and the minus sign for the term ±πk\pm\pi k, we get

12log|11Eϵγk/2|π|k|.-\frac{1}{2}\log|1-\frac{1}{E}\epsilon_{\gamma k/2}|-\pi|k|.

This guarantees the convergence of the integral in (124).

With this choice of the constants, the function bkb_{k} becomes

bk=exp[12ilog|11Eϵγk/2|coth(π(kk+i0))dk\displaystyle b_{k}=\exp\Big{[}\frac{1}{2i}\int_{-\infty}^{\infty}\log|1-\frac{1}{E}\epsilon_{\gamma k^{\prime}/2}|\coth\left(\pi(k^{\prime}-k+i0)\right)dk^{\prime}
πkπ2kEkE[coth(π(kk+i0))coth(π(k+i0))]𝑑k\displaystyle-\pi k-\frac{\pi}{2}\int_{-k_{E}}^{k_{E}}\left[\coth(\pi(k^{\prime}-k+i0))-\coth(\pi(k^{\prime}+i0))\right]dk^{\prime}
πkE[coth(π(kk+i0))coth(πk)]dk],\displaystyle-\pi\int_{-\infty}^{-k_{E}}\left[\coth(\pi(k^{\prime}-k+i0))-\coth(\pi k^{\prime})\right]dk^{\prime}\Big{]}, (135)

where, we remind, kEk_{E} is defined in (130).

Notice the following: (i) The normalizability of bkb_{k} is due to the eπ|k|e^{-\pi|k|} asymptotic. This absolute value |k||k| comes from the contribution from the branch cut of the log. This holds only for E<ϵ0E<\epsilon_{0}; (ii) There exists only one normalizable solution. This means that this bkb_{k} must satisfy the condition (125). One can check explicitly that it is indeed so.

C.5 Pairing vertex Φ(ωm)\Phi(\omega_{m})

We now use bkb_{k} from Eq. (135) to compute the function Φ(ωm)\Phi(\omega_{m}). We recall that Φ(ωm)Φ(ω)\Phi(\omega_{m})\equiv\Phi(\omega) is expressed via bkb_{k} as

Φ(ω)=2aβΦβ(ω)dβ2π=γ2πbkϵγk/2Φγk/2(ω)𝑑k.\Phi(\omega)=2\int_{-\infty}^{\infty}a_{\beta}\Phi_{\beta}^{\ast}(\omega)\frac{d\beta}{2\pi}=\frac{\gamma}{2\pi}\int_{-\infty}^{\infty}b_{k}\epsilon_{\gamma k/2}\Phi_{\gamma k/2}^{\ast}(\omega)dk. (136)

We show that this expression is equivalent to

Φ(ω)=1+|ω|γ|ω|γ/2f(log|ω|γ)\Phi(\omega)=\frac{1+|\omega|^{\gamma}}{|\omega|^{\gamma/2}}f(\log|\omega|^{\gamma}) (137)

where

f(x)=bkeikx𝑑k.f(x)=\int_{-\infty}^{\infty}b_{k}e^{-ikx}dk. (138)

To prove this, we employ the following trick: use

ϵβΦβ(ω)=21γΦβ(Ω)dΩ|Ω|1γ|Ωω|γ\epsilon_{\beta}\Phi_{\beta}(\omega)=\frac{2}{1-\gamma}\int_{-\infty}^{\infty}\frac{\Phi_{\beta}(\Omega)d\Omega}{|\Omega|^{1-\gamma}|\Omega-\omega|^{\gamma}}

and write

Φ(ω)=1πγ1γdΩ|Ω|1γ|Ωω|γbkΦγk/2(Ω)𝑑k\Phi(\omega)=\frac{1}{\pi}\frac{\gamma}{1-\gamma}\int_{-\infty}^{\infty}\frac{d\Omega}{|\Omega|^{1-\gamma}|\Omega-\omega|^{\gamma}}\int b_{k}\Phi_{\gamma k/2}^{\ast}(\Omega)dk

or

Φ(ω)=1πγ1γdΩ|Ω|1γ/2|Ωω|γf(log(|Ω|γ),wheref(x)=bkeikxdk.\Phi(\omega)=\frac{1}{\pi}\frac{\gamma}{1-\gamma}\int_{-\infty}^{\infty}\frac{d\Omega}{|\Omega|^{1-\gamma/2}|\Omega-\omega|^{\gamma}}f(\log(|\Omega|^{\gamma}),\quad\mbox{where}\quad f(x)=\int_{-\infty}^{\infty}b_{k}e^{-ikx}dk.

Comparing this expression with Eq. (113), we recover (137). Notice, that this proof does not use an explicit form of the function bkb_{k}. It does, however, explicitly uses that fact that Φ(ω)\Phi(\omega) is the solution of (67).333There is another way to prove (138) which does not use the fact that Φ(ω)\Phi(\omega) is a solution of (67), instead it uses the analytical properties of the function BkB_{k}.

The gap function Δ(ω)\Delta(\omega) is expressed as

Δ(ω)=|ω|γ/2f(log|ω|γ)\Delta(\omega)=|\omega|^{\gamma/2}f(\log|\omega|^{\gamma}) (139)

C.6 Summary: The steps to obtain the pairing vertex Φ(Ω)\Phi(\Omega) and the gap function Δ(ω)\Delta(\omega) for E<ϵ0E<\epsilon_{0}.

The strategy is the following:

  1. 1.

    Evaluate the integral

    I(k)=12log|11Eϵγk/2|tanh(π(kk))dkI(k)=\frac{1}{2}\int_{-\infty}^{\infty}\log|1-\frac{1}{E}\epsilon_{\gamma k^{\prime}/2}|\tanh\left(\pi(k^{\prime}-k)\right)dk^{\prime} (140)

    Notice, that it is real and antisymmetric.

  2. 2.

    Construct the function

    bk=sinh(πkE)eiI(k)cosh(π(kkE)cosh(π(k+kE)).b_{k}=\frac{\sinh(\pi k_{E})e^{-iI(k)}}{\sqrt{\cosh(\pi(k-k_{E})\cosh(\pi(k+k_{E}))}}. (141)
  3. 3.

    Compute

    f(x)=bkeikx𝑑k.f(x)=\int_{-\infty}^{\infty}b_{k}e^{-ikx}dk. (142)
  4. 4.

    The exact solutions of the linearized equations for the pairing vertex and the gap function are

    Φex(ωm)=(1+|ω|γ)f(log|Ω|γ)\displaystyle\Phi_{\text{ex}}(\omega_{m})=\left(1+|\omega|^{-\gamma}\right)f(\log|\Omega|^{\gamma})
    Δex(ωm)=|ω|γ/2f(log|Ω|γ)\displaystyle\Delta_{\text{ex}}(\omega_{m})=|\omega|^{\gamma/2}f(\log|\Omega|^{\gamma}) (143)

Appendix D Asymptotic expansion at small and large frequencies

To understand the structure of Δex(ωm)\Delta_{ex}(\omega_{m}) it is instructive to expand it at small and large ωm\omega_{m}. One can do this in two ways: expand in the formulas for the exact solution, or perturbatively expand the gap equation (10) order by order in ωm\omega_{m} or 1/ωm1/\omega_{m}.

For shortness, we present here the details of the direct perturbative expansion. The point of departure for the expansion in ωm\omega_{m} is the solution of (10) without the last term 1/(1+|ω¯m|γ)1/(1+|\bar{\omega}^{{}^{\prime}}_{m}|^{\gamma}): Φ(ωm)=\Phi(\omega_{m})= ReΦ0(ωm)\Phi_{0}(\omega_{m}), where

Φ0(ωm)=C0<|ω¯m|γeiϕ+βNlog|ω¯m|γ\Phi_{0}(\omega_{m})=\frac{C^{<}_{0}}{\sqrt{|\bar{\omega}_{m}|^{\gamma}}}e^{i\phi+\beta_{N}\log{|\bar{\omega}_{m}|^{\gamma}}} (144)

and, we remind, ω¯m=ωm/ω0\bar{\omega}_{m}=\omega_{m}/\omega_{0} and βN\beta_{N} is the solution of ϵ(β)=N\epsilon(\beta)=N, where ϵ(β)\epsilon(\beta) is given by (38).

Let us expand the r.h.s. of (10) in ωm\omega_{m} and express Φ(ωm)\Phi(\omega_{m}) as Re[Φ0(ωm)+δΦ(ωm)][\Phi_{0}(\omega_{m})+\delta\Phi(\omega_{m})]. The equation for δΦ(ωm)\delta\Phi(\omega_{m}) is

δΦ(ωm)1γ2N𝑑ωmδΦ(ωm)|ωm|1γ|ωmωm|γ=\displaystyle\delta\Phi(\omega_{m})-\frac{1-\gamma}{2N}\int d\omega^{\prime}_{m}\frac{\delta\Phi(\omega^{\prime}_{m})}{|\omega^{\prime}_{m}|^{1-\gamma}|\omega_{m}-\omega^{\prime}_{m}|^{\gamma}}=
1γ2Nω0γ𝑑ωmΦ0(ωm)|ωm|12γ(1|ωmωm|γ1|ωm|γ)+1γ2Nω0γ𝑑ωmΦ0(ωm)|ωm|1γ\displaystyle-\frac{1-\gamma}{2N\omega^{\gamma}_{0}}\int d\omega^{\prime}_{m}\frac{\Phi_{0}(\omega^{\prime}_{m})}{|\omega^{\prime}_{m}|^{1-2\gamma}}\left(\frac{1}{|\omega_{m}-\omega^{\prime}_{m}|^{\gamma}}-\frac{1}{|\omega^{\prime}_{m}|^{\gamma}}\right)+\frac{1-\gamma}{2N\omega^{\gamma}_{0}}\int d\omega^{\prime}_{m}\frac{\Phi_{0}(\omega^{\prime}_{m})}{|\omega^{\prime}_{m}|^{1-\gamma}} (145)

There are two types of terms in the r.h.s of (145). The first term is local in the sense that the integral over ωm\omega_{m}^{\prime} is ultra-violet and infra-red convergent and is determined by ωm\omega^{\prime}_{m} comparable to ωm\omega_{m}. Evaluating this term, we obtain

1γ2ω0γN𝑑ωmΦ0(ωm)|ωm|12γ(1|ωmωm|γ1|ωm|γ)=|ω¯|γΦ0(ωm)I1-\frac{1-\gamma}{2\omega^{\gamma}_{0}N}\int d\omega^{\prime}_{m}\frac{\Phi_{0}(\omega^{\prime}_{m})}{|\omega^{\prime}_{m}|^{1-2\gamma}}\left(\frac{1}{|\omega_{m}-\omega^{\prime}_{m}|^{\gamma}}-\frac{1}{|\omega^{\prime}_{m}|^{\gamma}}\right)=-|\bar{\omega}|^{\gamma}\Phi_{0}(\omega_{m})I_{1} (146)

where

I1=1γ2N𝑑yy(3γ/21+iβNγ)(1|1y|γ1|y|γ)=1γ2NΓ(3γ/2+iβNγ)Γ(γ/2iβNγ)Γ(γ)+\displaystyle I_{1}=\frac{1-\gamma}{2N}\int_{-\infty}^{\infty}dyy^{(3\gamma/2-1+i{\beta_{N}}\gamma)}\left(\frac{1}{|1-y|^{\gamma}}-\frac{1}{|y|^{\gamma}}\right)=\frac{1-\gamma}{2N}\frac{\Gamma(3\gamma/2+i{\beta_{N}}\gamma)\Gamma(-\gamma/2-i{\beta_{N}}\gamma)}{\Gamma(\gamma)}+
Γ(1γ)1γ2N(Γ(3γ/2+iβNγ)Γ(1+γ/2+iβNγ)+Γ(γ/2iβNγ)Γ(13γ/2iβNγ))\displaystyle\Gamma(1-\gamma)\frac{1-\gamma}{2N}\left(\frac{\Gamma(3\gamma/2+i{\beta_{N}}\gamma)}{\Gamma(1+\gamma/2+i{\beta_{N}}\gamma)}+\frac{\Gamma(-\gamma/2-i{\beta_{N}}\gamma)}{\Gamma(1-3\gamma/2-i{\beta_{N}}\gamma)}\right) (147)

and the integration variable y=ωm/ωmy=\omega^{\prime}_{m}/\omega_{m}.

D.1 Local corrections.

Let us momentarily keep only the local term in the r.h.s. of (145). We label the corresponding portion of δΦ(ωm)\delta\Phi(\omega_{m}) as δΦL(ωm)\delta\Phi_{L}(\omega_{m}). The functional form of δΦL(ωm)\delta\Phi_{L}(\omega_{m}) is determined by (146): δΦL(ωm)=|ω¯|γΦ0(ωm)C1\delta\Phi_{L}(\omega_{m})=|\bar{\omega}|^{\gamma}\Phi_{0}(\omega_{m})C_{1}. Substituting into (145), we find the relation between C1C_{1} and I1I_{1}:

C1<=C0<(I1I11)C^{<}_{1}=C^{<}_{0}\left(\frac{I_{1}}{I_{1}-1}\right) (148)

This analysis can be straightforwardly extended to higher orders in the expansion in |ω¯|γ|\bar{\omega}|^{\gamma}. At each other we extract infra-red convergent contribution in the source term by subtracting from 1/|ωmωm|γ1/|\omega^{\prime}_{m}-\omega_{m}|^{\gamma} a proper number of terms ( 1/|ωm|γ1/|\omega^{\prime}_{m}|^{\gamma}, ωm2/|ωm|γ+2\omega^{2}_{m}/|\omega^{\prime}_{m}|^{\gamma+2}, etc) to make the remaining integral over ωm\omega^{\prime}_{m} ultra-violet convergent, and solve for δΦL(ωm)\delta\Phi_{L}(\omega_{m}) induced by such source terms. The result is

δΦL(z)=Φ0(z)n=1Cn<zn\delta\Phi_{L}(z)=\Phi_{0}(z)~{}\sum_{n=1}^{\infty}C^{<}_{n}~{}z^{n} (149)

where, we remind, z=|ω¯m|γz=|\bar{\omega}_{m}|^{\gamma}, and

Cn<=C0<[Inm=1n1Im1]C^{<}_{n}=C^{<}_{0}~{}\left[I_{n}\displaystyle\prod_{m=1}^{n}\frac{1}{I_{m}-1}\right] (150)

where

Im=1γ2NQ(m,γ,βN),\displaystyle I_{m}=\frac{1-\gamma}{2N}Q(m,\gamma,\beta_{N}),
Q(m,γ,βN)=Γ((m+1/2)γ+iβNγ)Γ((1/2m)γiβNγ)Γ(γ)+\displaystyle Q(m,\gamma,\beta_{N})=\frac{\Gamma((m+1/2)\gamma+i{\beta_{N}}\gamma)\Gamma((1/2-m)\gamma-i{\beta_{N}}\gamma)}{\Gamma(\gamma)}+
Γ(1γ)(Γ((m+1/2)γ+iβNγ)Γ(1(1/2m)γ+iβNγ)+Γ((1/2m)γiβNγ)Γ(1(m+1/2)γiβNγ))\displaystyle\Gamma(1-\gamma)\left(\frac{\Gamma((m+1/2)\gamma+i{\beta_{N}}\gamma)}{\Gamma(1-(1/2-m)\gamma+i{\beta_{N}}\gamma)}+\frac{\Gamma((1/2-m)\gamma-i{\beta_{N}}\gamma)}{\Gamma(1-(m+1/2)\gamma-i{\beta_{N}}\gamma)}\right) (151)

Note than Cn<C^{<}_{n} are complex numbers. Combining ΔΦL\Delta\Phi_{L} and Φ0\Phi_{0}, we obtain the total local contribution to the pairing vertex ΦL=\Phi_{L}= Re [Φ0+δΦL][\Phi_{0}+\delta\Phi_{L}] as

ΦL(z)=Re[ei(ϕ+βNlogz)n=0Cn<zn1/2]\Phi_{L}(z)={\text{R}e}\left[e^{i(\phi+{\beta_{N}}\log{z})}\sum_{n=0}^{\infty}C^{<}_{n}z^{n-1/2}\right] (152)

Expressing (152) as the equation for the local contribution to the gap function ΔL(z)=zΦL(z)/(1+z)\Delta_{L}(z)=z\Phi_{L}(z)/(1+z) , we obtain the first term in (75).

At small γ\gamma and N=O(1)N=O(1), βN1Nγ(1Nγ/8)1\beta_{N}\approx\frac{1}{\sqrt{N\gamma}}(1-N\gamma/8)\gg 1. Evaluating InI_{n} in two orders in 1/βN1/\beta_{N}, we find a simple expression

In=1+2inβN3(nβN)2I_{n}=1+2i\frac{n}{\beta_{N}}-3\left(\frac{n}{\beta_{N}}\right)^{2} (153)

Substituting this InI_{n} into (150) and then into (149), using the relation between Φ\Phi and Δ\Delta, and exponentiating the series to order z3z^{3}, we obtain, up to corrections of order z/βN2z/\beta^{2}_{N},

ΔL(z)=C0<z1/2(1+z)3/4cos(ϕ+βNQ1(z))\Delta_{L}(z)=C^{<}_{0}~{}\frac{z^{1/2}}{(1+z)^{3/4}}\cos{\left(\phi+\beta_{N}Q_{1}(z)\right)} (154)

where Q1(z)=logzz/2+3z2/16+Q_{1}(z)=\log{z}-z/2+3z^{2}/16+.... Eq. (154) and the expansion of Q1(z)Q_{1}(z) coincide with those for Δdiff(z)\Delta_{\text{diff}}(z), Eqs. (58) and (59). This explicitly confirms that for γ1\gamma\ll 1, Δ(ωm)ΔL(ωm)Δdiff(ωm)\Delta(\omega_{m})\approx\Delta_{L}(\omega_{m})\approx\Delta_{\text{diff}}(\omega_{m}).

D.2 Non-local corrections.

We now look at the other terms, which we had to add and subtract at each order to make local contributions ultra-violet convergent. At the leading order in the expansion the additional term in the r.h.s. of Eq. (145) is

1γ2Nω0γ𝑑ωmΦ0(ωm)|ωm|1γ=1γNγC0<eiϕ0𝑑xxiβN1/2\frac{1-\gamma}{2N\omega^{\gamma}_{0}}\int d\omega^{\prime}_{m}\frac{\Phi_{0}(\omega^{\prime}_{m})}{|\omega^{\prime}_{m}|^{1-\gamma}}=\frac{1-\gamma}{N\gamma}C^{<}_{0}e^{i\phi}\int_{0}^{\infty}dxx^{i{\beta_{N}}-1/2} (155)

This integral is ultra-violet divergent, but the divergence is fictitious because the actual Φ(ωm)\Phi(\omega_{m}) must drop as 1/|ωm|γ1/|\omega_{m}|^{\gamma} at large |ωm||\omega_{m}|. Accordingly, we cut ultra-violet divergencies at some |ωm|=ωmax|\omega_{m}|=\omega_{max}, which, we argue below, is roughly where xx^{\prime} changes sign. The separation into local and non-local terms obviously holds only for |ωm|<ωmax|\omega_{m}|<\omega_{max}. The non-local term (155) induces another set of corrections to the bare Φ(ωm)=\Phi(\omega_{m})= ReΦ0(ωm)\Phi_{0}(\omega_{m}). We label the corresponding non-local part of the full Φ(ωm)\Phi(\omega_{m}) as ΦnL(ωm)\Phi_{nL}(\omega_{m}). By construction, ΦnL(ωm)\Phi_{nL}(\omega_{m}) is real. Substituting (155) into (145), adding the complex conjugated term, and adding and subtracting 1/|ωm|γ1/|\omega^{\prime}_{m}|^{\gamma} to/from 1/|ωmωm|γ1/|\omega_{m}-\omega^{\prime}_{m}|^{\gamma}, we obtain the equation for ΦnL(ωm)\Phi_{nL}(\omega_{m}) in the form

ΦnL(ωm)1γ2N𝑑ωmΦnL(ωm)|ωm|1γ(1|ωmωm|γ1|ωm|γ)=K\Phi_{nL}(\omega_{m})-\frac{1-\gamma}{2N}\int d\omega^{\prime}_{m}\frac{\Phi_{nL}(\omega^{\prime}_{m})}{|\omega^{\prime}_{m}|^{1-\gamma}}\left(\frac{1}{|\omega_{m}-\omega^{\prime}_{m}|^{\gamma}}-\frac{1}{|\omega^{\prime}_{m}|^{\gamma}}\right)=K (156)

where

K=1γN0ωmax𝑑ωm(ΦnL(ωm)|ωm|γReΦ0(ωm)ω0γ)K=\frac{1-\gamma}{N}\displaystyle\int_{0}^{\omega_{max}}d\omega_{m}\left(\frac{\Phi_{nL}(\omega_{m})}{|\omega_{m}|^{\gamma}}-\frac{{\text{R}e}\Phi_{0}(\omega_{m})}{\omega^{\gamma}_{0}}\right) (157)

We first notice that Eq. (156) without the KK term does have the solution ΦnL(ωm)=E0,0<|ωm|(b0γ/2)\Phi_{nL}(\omega_{m})=E^{<}_{0,0}|\omega_{m}|^{(b_{0}-\gamma/2)}, where b0b_{0} is the solution of ϵb0/γ=N\epsilon_{b_{0}/\gamma}=N in the interval [γ/2,γ/2+2]\left[\gamma/2,\gamma/2+2\right]. For b0b_{0} within this range, the integral in the l.h.s. of (156) does not diverge in the infra-red and ultra-violet limits. To satisfy (156) one needs to satisfy also the condition K=0K=0. Substituting our trial solution for ΦnL(ωm)\Phi_{nL}(\omega_{m}) and ReΦ0(ωm)\Phi_{0}(\omega_{m}) into (157), we find that the condition K=0K=0 is satisfied for some particular ratio E0,0</C0<E^{<}_{0,0}/C^{<}_{0}.

This analysis can be extended to higher orders in the perturbation theory. The non-local terms in the r.h.s. of the equation for the pairing vertex appear in the form A10+A11|ω¯m|γ+A12ω¯m2γ++A20ω¯m2+A21|ω¯m|2+γ+A22|ω¯m|2+2γ++A30ω¯m4+..A_{10}+A_{11}|\bar{\omega}_{m}|^{\gamma}+A_{12}\bar{\omega}_{m}^{2\gamma}+...+A_{20}\bar{\omega}_{m}^{2}+A_{21}|\bar{\omega}_{m}|^{2+\gamma}+A_{22}|\bar{\omega}_{m}|^{2+2\gamma}+...+A_{30}\bar{\omega}_{m}^{4}+.., where all AijA_{ij} contain C0<C^{<}_{0} as the overall factor. To cancel all these non-local terms, we choose ΦnL(ωm)\Phi_{nL}(\omega_{m}) in the form

ΦnL(ωm)=1|ω¯|γ/2n,m=0Dn,m|ω¯|bm+γn\Phi_{nL}(\omega_{m})=\frac{1}{|\bar{\omega}|^{\gamma/2}}\sum_{n,m=0}^{\infty}D_{n,m}~{}|\bar{\omega}|^{b_{m}+\gamma n} (158)

where γ/2+2m<bm<γ/2+2(m+1)\gamma/2+2m<b_{m}<\gamma/2+2(m+1). The terms with m>0m>0 are solutions of the same equation (157), but with the proper number of terms subtracted from 1/|ωmωm|γ1/|\omega_{m}-\omega^{\prime}_{m}|^{\gamma} and added to the KK term in the r.h.s, to make the integral over ωm\omega^{\prime}_{m} in the l.h.s. ultra-violet convergent. We see from Fig. 13 that there is one bnb_{n} in each interval that satisfies (157). Substituting this form into the equation for the pairing vertex and collecting non-local terms, we find a set of coupled algebraic equations for Dn,mD_{n,m} with AijA_{ij} playing the role of source terms. The precise forms of these equations cannot be obtained within perturbative expansion in ωm\omega_{m}, and we just have to assume that the solution Dn,mD_{n,m} exists. Reexpressing (158) as the equation for ΔnL\Delta_{nL} instead of ΦnL\Phi_{nL}, we reproduce Dn,m<D^{<}_{n,m} series in (75).

Combining (152) and (158) we see that the r.h.s. of Eq. (75) is the sum of local and non-local contributions

Δ(z)=ΔL(z)+ΔnL(z).\Delta(z)=\Delta_{L}(z)+\Delta_{nL}(z). (159)

At the largest ωm\omega_{m} one can expand in 1/z1/z and use Δ0(z)=C0>/z\Delta_{0}(z)=C^{>}_{0}/z as the zero-order solution. The expansion is straightforward and yields Eq. (76).

References