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

Pairing by a dynamical interaction in a metal

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

We consider pairing of itinerant fermions in a metal near a quantum-critical point (QCP) towards some form of particle-hole order (nematic, spin-density-wave, charge-density-wave, etc). At a QCP, the dominant interaction between fermions comes from exchanging massless fluctuations of a critical order parameter. At low energies, this physics can be described by an effective model with the dynamical electron-electron interaction V(Ωm)1/|Ωm|γV(\Omega_{m})\propto 1/|\Omega_{m}|^{\gamma}, up to some upper cutoff Λ\Lambda. The case γ=0\gamma=0 corresponds to BCS theory, and can be solved by summing up geometric series of Cooper logarithms. We show that for a finite γ\gamma, the pairing problem is still marginal (i.e., perturbation series are logarithmic), but one needs to go beyond logarithmic approximation to find the pairing instability. We discuss specifics of the pairing at γ>0\gamma>0 in some detail and also analyze the marginal case γ=0+\gamma=0+, when V(Ωm)=λlog(Λ/|Ωm|)V(\Omega_{m})=\lambda\log{(\Lambda/|\Omega_{m}|)}. We show that in this case the summation of Cooper logarithms does yield the pairing instability at λlog2(Λ/Tc)=O(1)\lambda\log^{2}{(\Lambda/T_{c})}=O(1), but the logarithmic series are not geometrical. We reformulate the pairing problem in terms of a renormalization group (RG) flow of the coupling, and show that the RG equation is different in the cases γ=0\gamma=0, γ=0+\gamma=0+, and γ>0\gamma>0.

I Preface

It is our great pleasure to present this paper for the special issue of JETP devoted to 90th birthday of Igor Ekhielevich Dzyaloshinskii. His works on correlated electron systems are of highest scientific quality. He made seminal contributions to quantum magnetism, superconductivity, Fermi-liquid theory, and to other subfields of modern condensed matter physics. In a number of works, including the famous ones on the interplay between d-wave superconductivity and antiferromagnetism at the beginning of “high -TcT_{c} era” (Refs.  Zheleznyak et al. (1997); Dzyaloshinskii (1987); Dzyaloshinskii and Yakovenko (1988)), Igor Ekhielevich used the renormalization group technique to obtain the flow of the couplings upon integrating out fermions with higher energies. In this study, we apply the same RG technique to analyze the pairing instability in systems with critical dynamical pairing interaction. We consider this as a natural extension of his works. Happy birthday, Igor Ekhielevich, and the very best wishes.

II Introduction.

The pairing near a quantum-critical point (QCP) in a metal and its interplay with non-Fermi-liquid behavior in the normal state, is a fascinating subject, which attracted substantial attention in the correlated electron community after the discovery of superconductivity (SC) in the cuprates, Fe-based systems, heavy-fermion materials, organic materials, and, most recently, twisted bilayer graphene 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); Cao et al. (2018). Itinerant QC models, analyzed in recent years, include models of fermions in spatial dimensions D3D\leq 3, various two-dimensional models near zero-momentum spin and charge nematic instabilities, and instabilities towards spin and charge density-wave order with either real or imaginary (current) order parameter, 2D fermions at a half-filled Landau level, Sachdev-Ye-Kitaev (SYK) and SYK-Yukawa models, strong coupling limit of electron-phonon superconductivity, and even color superconductivity of quarks, mediated by gluon exchange. These problems have also been studied analytically and using various numerical techniques. We refer a reader to Ref. Abanov and Chubukov (2020), where the extensive list of references to these works has been presented.

From theory perspective, pairing near a QCP is a fundamentally novel phenomenon because an effective dynamic electron-electron interaction, V(q,Ω)V(q,\Omega), mediated by a critical collective boson, which condenses at a QCP, provides a strong attraction in one or more pairing channels and, at the same time, gives rise to a non-Fermi liquid (NFL) behavior in the normal state. The two tendencies compete with each other: fermionic incoherence, associated with the NFL behavior, destroys the Cooper logarithm and reduces the tendency to pairing, while an opening of a SC gap eliminates the scattering at low energies and reduces the tendency to a NFL. To find the winner of this competition (SC or NFL), one needs to analyze the set of integral equations for the fermionic self-energy, Σ¯(𝐤,ω)\bar{\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 on the Fermi surface (FS) and computed within the one-loop approximation. The self-energy on the FS, Σ¯(𝐤,ω)\bar{\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. E.g., near an antiferromagnetic QCP in a system with a single FS, the strongest attraction is in the dd-wave channel. In each particular case, one has to project the pairing interaction into the irreducible channels, find the strongest one, and solve for the pairing vertex for a given pairing symmetry.

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, and the equation for Δ(ω)\Delta(\omega) is similar to that in a conventional Eliashberg theory for phonon-mediated superconductivity; 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 is qualitatively different because the effective interaction V(Ω)V(\Omega), mediated by a critical massless boson, is a singular function of frequency. Quite generally, such interaction behaves at the smallest Ωm\Omega_{m} on the Matsubara axis as V(Ωm)1/|Ωm|γV(\Omega_{m})\propto 1/|\Omega_{m}|^{\gamma}, where γ>0\gamma>0 is some exponent. (Fig. 1). This holds at frequencies below some upper cutoff Λ\Lambda. At larger Ωm>Λ\Omega_{m}>\Lambda, the interaction drops even further, and can be safely neglected.

In these notations, BCS pairing corresponds to γ=0\gamma=0. The superconducting transition temperature TcT_{c} inthe BCS case can be most straightforwardly obtained by computing the pairing susceptibility χpp\chi_{pp} in the order-by-order expansion in the coupling and identifying the temperature at which it diverges. The series contain the powers of the product of the Cooper logarithm logΛ/Tc\log{\Lambda/T_{c}} and the dimensionless coupling λ\lambda, defined such that 1+λ1+\lambda is the ratio of the renormalized and the bare electron masses. The series are geometrical, i.e., χpp=χ0(1+λlogΛ/Tc+(λlogΛ/Tc)2+)=χ0/(1λlogΛ/Tc)\chi_{pp}=\chi_{0}(1+\lambda\log{\Lambda/T_{c}}+(\lambda\log{\Lambda/T_{c}})^{2}+...)=\chi_{0}/(1-\lambda\log{\Lambda/T_{c}}). We will see that for γ>0\gamma>0, the pairing kernel still has a marginal 1/|ω|1/|\omega| form, and, as a result, the series for χpp\chi_{pp} still contain logarithms. However, contrary to the BCS case, where 1/|ω|1/|\omega| comes from the fermionic propagator, the marginal exponent 1-1 for γ>0\gamma>0 is the sum of the exponent γ\gamma from the interaction and 1γ1-\gamma from the fermionic self-energy (see below). As a result, the logarithmic series are not geometric, and we will see that χpp\chi_{pp} does not diverge down to T=0T=0. We show that the pairing instability still exists, however one needs to go beyond the logarithmic approximation to obtain it.

We show that there exists a special case, which falls in between γ=0\gamma=0 and γ>0\gamma>0. This is the limit γ0+\gamma\to 0+, when V(Ωm)logΛ/|Ωm|V(\Omega_{m})\propto\log{\Lambda/|\Omega_{m}|}. In this limit, the pairing instability still can be detected by summing up the series of logarithms for χpp\chi_{pp}. However, the series are not geometric, and some extra efforts are needed to sum them up to find the form of χpp\chi_{pp} and detect the pairing instability at TΛeπ/(2λ)T\sim\Lambda e^{-\pi/(2\sqrt{\lambda})}.

The model with the interaction V(Ωm)1/|Ωm|γV(\Omega_{m})\propto 1/|\Omega_{m}|^{\gamma} displays a very rich physics, and our group has studied it over the last few years. This physics is particularly interesting for γ2\gamma\leq 2, where a wide range of the pseudogap (preformed pair) behavior emerges, and for γ>2\gamma>2, when the new, non-superconducting ground state emerges Wu et al. (2020a). In this communication, we will not discuss these values of γ\gamma but instead focus on small γ\gamma and discuss in some detail the difference between BCS pairing at γ=0\gamma=0 and the pairing at γ=0+\gamma=0+ and at γ>0\gamma>0. We derive the Eliashberg equation for the pairing vertex and analyze it within logarithmic perturbation theory and beyond it. We then convert the integral Eliashberg equation into the approximate differential equation and obtain and solve the corresponding renormalization group (RG) equation.

The pairing problem at small γ\gamma attracted a lot of attention in the last few years from various physics sub-communities of physicists Mross et al. (2010); Metlitski et al. (2015); Raghu et al. (2015); Mahajan et al. (2013); *raghu2; *raghu3; *raghu4; *raghu5; Wang et al. (2017, 2018); Fitzpatrick et al. (2015b); Wang and Chubukov (2013); Wu et al. (2019a); Son (1999); Chubukov and Schmalian (2005); Schäfer and Wilczek (1999); Pisarski and Rischke (2000); Wang and Rischke (2002); Damia et al. (2020) . The pairing interaction V(Ω)1/|Ω|γV(\Omega)\sim 1/|\Omega|^{\gamma} with γ1\gamma\ll 1 emerges for fermions near a generic particle-hole instability in a weakly anisotropic 3D system (more explicitly, in dimension D=3ϵD=3-\epsilon, where ϵγ\epsilon\sim\gamma, Refs.  Mross et al. (2010); Metlitski et al. (2015); Raghu et al. (2015) ). A similar gap equation with small γ\gamma holds for the pairing in graphene Khveshchenko (2009). The model with γ=0+\gamma=0+ describes the pairing in 3D systems and color superconductivity of quarks due to gluon exchange Son (1999); Schäfer and Wilczek (1999); Pisarski and Rischke (2000); Wang and Rischke (2002). The γ=0+\gamma=0+ model yields a marginal Fermi liquid form of the fermionic self-energy in the normal state and was argued Varma et al. (1989); *Littlewood_92; Varma (2020) to be relevant to pairing in the cuprates and, possibly, in Fe-based superconductors.

The structure of the paper is at follows. In the next Section we present the set of coupled Eliashberg equations for the pairing vertex Φ(ωm)\Phi(\omega_{m}) and the fermionic self-energy Σ¯(ωm)\bar{\Sigma}(\omega_{m}) and combine them into the equation for the gap function Δ(ωm)\Delta(\omega_{m}). In Sec. IV we analyze the structure of the logarithmic perturbation theory for γ=0+\gamma=0+ and γ>0\gamma>0. In Sec. V we go beyond perturbation theory and re-express the integral Eliashberg equation as an approximate differential equation for the pairing vertex and solve it. We show that for γ=0+\gamma=0+, the solution coincides with the result of summation of logarithmic series. For γ>0\gamma>0, we show that the system does become unstable towards pairing if the interaction exceeds a certain threshold. In Sec. VI we analyze the pairing at small γ\gamma from RG perspective and reproduce the results of the previous Section. We present our conclusions in Sec. VII.

III The model

We consider itinerant fermions at the onset of a long-range particle-hole order in either spin or charge channel. At a critical point, the propagator of a soft boson becomes massless and mediates singular interaction between fermions. A series of earlier works on spin-density-wave order, charge-density-wave order, Ising-nematic order, etc (see Ref.Abanov and Chubukov (2020) for references) have found that this interaction is attractive in at least one pairing channel. We take this as in input and project boson-mediated interaction into the channel with the strongest attraction. As we said, at small frequencies, the interaction scales as V(Ωm)1/|Ωm|γV(\Omega_{m})\propto 1/|\Omega_{m}|^{\gamma}. We incorporate dimension-full factors, like the density of states, into V(Ωm)V(\Omega_{m}) and treat it as dimensionless. We assume that the power-law form holds up to the scale Λ\Lambda, and set V(Ωm)=0V(\Omega_{m})=0 above this scale. To keep V(Ωm)V(\Omega_{m}) continuous at Λ\Lambda and also to transform gradually between γ=0+\gamma=0+ and γ>0\gamma>0, we use the following form for V(Ωm)V(\Omega_{m}) (Fig.1):

Refer to caption
Figure 1: Frequency dependence of the effective interaction V(Ω¯m)V(\bar{\Omega}_{m}), mediated by a soft boson, at T=0T=0. Away from a QCP, V(Ω¯m)V(\bar{\Omega}_{m}) tends to a finite value at Ω¯m=0\bar{\Omega}_{m}=0. Right at a QCP, the boson becomes massless, and V(Ω¯m)V(\bar{\Omega}_{m}) diverges as 1/|Ω¯m|γ1/|\bar{\Omega}_{m}|^{\gamma}.
V(Ωm)=λγ(g¯|Ωm|)γ(1(|Ωm|Λ)γ)V(\Omega_{m})=\frac{\lambda}{\gamma}\left(\frac{{\bar{g}}}{|\Omega_{m}|}\right)^{\gamma}\left(1-\left(\frac{|\Omega_{m}|}{\Lambda}\right)^{\gamma}\right) (1)

Here λ\lambda is a dimensionless coupling, and g¯{\bar{g}} has units of energy. We assume that λ1\lambda\ll 1, but the ratio λ/γ\lambda/\gamma can be arbitrary. At γ=0+\gamma=0+, the last term in (1) can be expanded in γ\gamma and yields the logarithmic interaction

V(Ωm)=λlogΛ|Ωm|V(\Omega_{m})=\lambda\log{\frac{\Lambda}{|\Omega_{m}|}} (2)

The conventional BCS/Eliashberg case in this notations is obtained by extending (1) to the case when a pairing boson has a finite mass. In this situation, Ωm\Omega_{m} is replaced by a constant below a certain scale. In BCS theory, this scale is assumed to be comparable to Λ\Lambda, such that V(Ωm)λBCSV(\Omega_{m})\approx\lambda_{BCS} up to |Ωm|Λ|\Omega_{m}|\sim\Lambda, and zero at larger |Ωm||\Omega_{m}|.

Below we measure all quantities with the dimension of energy, i.e., T,Ωm,ΛT,\Omega_{m},\Lambda, in units of g¯{\bar{g}}, i.e., introduce T¯=T/g¯,Ω¯m=Ωm/g¯{\bar{T}}=T/{\bar{g}},{\bar{\Omega}}_{m}={\Omega_{m}}/{\bar{g}}, and Λ¯=Λ/g¯{\bar{\Lambda}}=\Lambda/{\bar{g}}. Throughout this paper we assume that Λ¯1\bar{\Lambda}\gg 1.

Earlier works on the pairing mediated by a soft collective boson found that a boson is overdamped due to Landau damping into a particle-hole pair and can be treated as slow mode compared to a fermion, i.e., at the same momentum qq, a typical fermionic frequency is much larger than a typical bosonic frequency. This is the same property that justified Eliashberg theory of phonon-mediated superconductivity. By analogy, the theory of electronic superconductivity, mediated by soft collective bosonic excitations in spin or charge channel, is often referred to as Eliashberg theory, and we will be using this convention.

In Eliashberg theory, one can explicitly integrate first over the momentum component perpendicular to the Fermi surface and then over the component(s) along the Fermi surface, and reduce the pairing problem to a set of coupled integral equations for frequency dependent self-energy Σ¯(ωm)\bar{\Sigma}(\omega_{m}) and the pairing vertex Φ(ωm)\Phi(\omega_{m}).

At T=0T=0, the coupled Eliashberg equations are

Φ¯(ω¯m)\displaystyle\bar{\Phi}(\bar{\omega}_{m}) =\displaystyle= 12𝑑ω¯mΦ(ω¯m)Σ¯~2(ω¯m)+Φ¯2(ω¯m)V(|ω¯mω¯m|),\displaystyle\frac{1}{2}\int d\bar{\omega}^{\prime}_{m}\frac{\Phi(\bar{\omega}^{\prime}_{m})}{\sqrt{{\tilde{\bar{\Sigma}}}^{2}(\bar{\omega}^{\prime}_{m})+\bar{\Phi}^{2}(\bar{\omega}^{\prime}_{m})}}{V}(|\bar{\omega}_{m}-\bar{\omega}^{\prime}_{m}|),
Σ¯~(ω¯m)\displaystyle{\tilde{\bar{\Sigma}}}(\bar{\omega}_{m}) =\displaystyle= ω¯m+12𝑑ω¯mΣ¯~(ω¯m)Σ¯~2(ω¯m)+Φ¯2(ω¯m)V(|ω¯mω¯m|)\displaystyle\bar{\omega}_{m}+\frac{1}{2}\int d\bar{\omega}^{\prime}_{m}\frac{{\tilde{\bar{\Sigma}}}(\bar{\omega}^{\prime}_{m})}{\sqrt{{\tilde{\bar{\Sigma}}}^{2}(\bar{\omega}^{\prime}_{m})+\bar{\Phi}^{2}(\bar{\omega}^{\prime}_{m})}}{V}(|\bar{\omega}_{m}-\bar{\omega}^{\prime}_{m}|) (3)

where Σ¯~(ω¯m)=ω¯m+Σ¯(ω¯m){\tilde{\bar{\Sigma}}}(\bar{\omega}_{m})=\bar{\omega}_{m}+\bar{\Sigma}(\bar{\omega}_{m}). In these equations, both Σ¯(ωm)\bar{\Sigma}(\omega_{m}) and Φ¯(ωm)\bar{\Phi}(\omega_{m}) are real functions. Observe that we define Σ¯(ωm)\bar{\Sigma}(\omega_{m}) with the overall plus sign and without the overall factor of ii. In the normal state (Φ¯0\bar{\Phi}\equiv 0), we have at γ>0\gamma>0, and |ω¯m|1|\bar{\omega}_{m}|\ll 1

Σ¯~(ω¯m)=ω¯m+|ω¯m|1γ1γsgnω¯m,{\tilde{\bar{\Sigma}}}(\bar{\omega}_{m})=\bar{\omega}_{m}+\frac{|\bar{\omega}_{m}|^{1-\gamma}}{1-\gamma}\operatorname{sgn}{\bar{\omega}_{m}}, (4)

At γ=0+\gamma=0+,

Σ¯~(ωm)=ω¯m(1+λlogΛ¯|ω¯m|),{\tilde{\bar{\Sigma}}}(\omega_{m})=\bar{\omega}_{m}\left(1+\lambda\log{\frac{\bar{\Lambda}}{|\bar{\omega}_{m}|}}\right), (5)

The superconducting gap function Δ¯(ω¯m)\bar{\Delta}(\bar{\omega}_{m}) is defined as the ratio

Δ¯(ω¯m)=ω¯mΦ¯(ω¯m)Σ¯~(ω¯m)=Φ¯(ω¯m)1+Σ¯(ω¯m)/ω¯m.\bar{\Delta}(\bar{\omega}_{m})=\bar{\omega}_{m}\frac{\bar{\Phi}(\bar{\omega}_{m})}{{\tilde{\bar{\Sigma}}}(\bar{\omega}_{m})}=\frac{\bar{\Phi}(\bar{\omega}_{m})}{1+\bar{\Sigma}(\bar{\omega}_{m})/\bar{\omega}_{m}}. (6)

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

Δ¯(ω¯m)=12𝑑ω¯mΔ¯(ω¯m)Δ¯(ω¯m)ω¯mω¯m(ω¯m)2+Δ¯2(ω¯m)1|ω¯mω¯m|γ.\bar{\Delta}(\bar{\omega}_{m})=\frac{1}{2}\int d\bar{\omega}^{\prime}_{m}\frac{\bar{\Delta}(\bar{\omega}^{\prime}_{m})-\bar{\Delta}(\bar{\omega}_{m})\frac{\bar{\omega}^{\prime}_{m}}{\bar{\omega}_{m}}}{\sqrt{(\bar{\omega}^{\prime}_{m})^{2}+\bar{\Delta}^{2}(\bar{\omega}^{\prime}_{m})}}~{}\frac{1}{|\bar{\omega}_{m}-\bar{\omega}^{\prime}_{m}|^{\gamma}}. (7)

This equation contains a single function Δ¯(ω¯m)\bar{\Delta}(\bar{\omega}_{m}), but for the cost that Δ¯(ω¯m)\bar{\Delta}(\bar{\omega}_{m}) appears also in the r.h.s., which makes Eq. (7) less convenient for the analysis than Eqs. (III).

For a generic γ>0\gamma>0, the coupled equations (III) for Φ¯\bar{\Phi} and Σ¯~{\tilde{\bar{\Sigma}}} describe the interplay between the two competing tendencies – one towards superconductivity, specified by Φ¯\bar{\Phi}, and the other towards incoherent NFL normal-state behavior, specified by Σ¯~{\tilde{\bar{\Sigma}}}. The competition between the two tendencies is encoded in the fact that Σ¯~{\tilde{\bar{\Sigma}}} appears in the denominator of the equation for Φ¯\bar{\Phi} and Φ¯\bar{\Phi} appears in the denominator of the equation for Σ¯~{\tilde{\bar{\Sigma}}}. In more physical terms, a self-energy Σ¯~{\tilde{\bar{\Sigma}}} is an obstacle to Cooper pairing, while when Φ¯\bar{\Phi} is non-zero, it reduces the strength of the self-energy and renders fermionic coherence.

The full set of the equations for electron-mediated pairing generally must contain the third equation, describing the feedback from the pairing on the bosonic propagator. This feedback is small 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 changes qualitatively below TcT_{c} (Refs. Abanov et al. (2003, 2001a)). In this work, we only consider the computation of TcT_{c} and will not discuss system behavior below TcT_{c}. For this computation, it is sufficient to restrict with the two equations (III). Moreover, we can (i) treat Φ¯(ω¯m)\bar{\Phi}(\bar{\omega}_{m}) as infinitesimally small and neglect it in the denominator of Eq. (III), and (ii) use Eqs. (4) or (5) for Σ¯~{\tilde{\bar{\Sigma}}}. For a small, but finite γ>0\gamma>0, the linearized equation for the pairing vertex is

Φ¯(ω¯m)=12𝑑ω¯mΦ¯(ω¯m)|ω¯m|1γ|ω¯mω¯m|γ(1+(γ/λ)|ω¯m|γ)(1(|ω¯mω¯m|Λ¯)γ)\bar{\Phi}(\bar{\omega}_{m})=\frac{1}{2}\int d\bar{\omega}^{\prime}_{m}\frac{\bar{\Phi}(\bar{\omega}^{\prime}_{m})}{|\bar{\omega}^{\prime}_{m}|^{1-\gamma}|\bar{\omega}_{m}-\bar{\omega}^{\prime}_{m}|^{\gamma}(1+(\gamma/\lambda)|\bar{\omega}^{\prime}_{m}|^{\gamma})}\left(1-\left(\frac{|\bar{\omega}_{m}-\bar{\omega}^{\prime}_{m}|}{\bar{\Lambda}}\right)^{\gamma}\right) (8)

Observe that the overall coupling is just a number, equal to one, because the factor λ/γ\lambda/\gamma in V(Ω¯m)V(\bar{\Omega}_{m}) cancels out with the same factor in the self-energy. This factor is still present in the denominator, but in the term, which contains |ω¯m|γ|\bar{\omega}^{\prime}_{m}|^{\gamma} and becomes irrelevant at the smallest frequencies

At γ=0+\gamma=0+, the linearized equation for Φ¯(ω¯m)\bar{\Phi}(\bar{\omega}_{m}) is

Φ¯(ω¯m)=λ2𝑑ω¯mΦ¯(ω¯m)|ω¯m|logΛ¯|ω¯mω¯m|1+λlogΛ¯|ω¯mω¯m|\bar{\Phi}(\bar{\omega}_{m})=\frac{\lambda}{2}\int d\bar{\omega}^{\prime}_{m}\frac{\bar{\Phi}(\bar{\omega}^{\prime}_{m})}{|\bar{\omega}^{\prime}_{m}|}\frac{\log{\frac{\bar{\Lambda}}{|\bar{\omega}_{m}-\bar{\omega}^{\prime}_{m}|}}}{1+\lambda\log{\frac{\bar{\Lambda}}{|\bar{\omega}_{m}-\bar{\omega}^{\prime}_{m}|}}} (9)

IV Summing up the logarithms

As a first try, we analyze the linearized equation for the pairing vertex perturbatively, by adding up a trial, infinitesimally small Φ¯0\bar{\Phi}_{0} to the r.h.s of Eqs. (8) and (9) and computing the pairing susceptibility perturbatively, order-by-order. Such approach has been commonly used for BCS pairing. The perturbation series there contain singular Cooper logarithms. The logarithmic singularity can be cut either by a finite TT or at T=0T=0, by a finite total incoming bosonic frequency, Ω¯tot\bar{\Omega}_{tot}. For consistency with other cases, we set T=0T=0 and keep Ω¯tot\bar{\Omega}_{tot} finite. The result of order-by-order analysis for a BCS pairing is well known:

Φ¯(Ω¯tot)=Φ¯0(1+λBCSlogΛ¯|Ω¯tot|+λBCS2log2Λ¯|Ω¯tot|+)=Φ¯01λBCSlogΛ¯|Ω¯tot|\bar{\Phi}(\bar{\Omega}_{tot})=\bar{\Phi}_{0}\left(1+\lambda_{BCS}\log{\frac{\bar{\Lambda}}{|{\bar{\Omega}}_{tot}|}}+\lambda^{2}_{BCS}\log^{2}{\frac{\bar{\Lambda}}{|{\bar{\Omega}}_{tot}|}}+...\right)=\frac{\bar{\Phi}_{0}}{1-\lambda_{BCS}\log{\frac{\bar{\Lambda}}{|\bar{\Omega}_{tot}|}}} (10)

The ratio Φ¯(Ω¯tot)/Φ¯0\bar{\Phi}(\bar{\Omega}_{tot})/\bar{\Phi}_{0} (the pairing susceptibility) diverges at |Ω¯tot|=Λ¯e1/λBCS|\bar{\Omega}_{tot}|=\bar{\Lambda}e^{-1/\lambda_{BCS}} and becomes negative at smaller |Ω¯tot||\bar{\Omega}_{tot}|, indicating that the normal state is unstable towards pairing. (In a more accurate description, the pole in Φ¯(Ω¯tot)\bar{\Phi}(\bar{\Omega}_{tot}) moves from the lower to the upper half-plane of complex frequency Abrikosov et al. (1965)).

We now do the same calculation for γ>0\gamma>0. The kernel in Eq. (8) is still marginal at ω¯m,ω¯m1\bar{\omega}^{\prime}_{m},\bar{\omega}_{m}\ll 1, but, as we said, now the scaling dimension 1-1 is the sum of γ-\gamma, coming from the interaction, and 1+γ-1+\gamma, coming from the self-energy. This implies that perturbation theory still contains logarithms, but in distinction to BCS, 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|ω¯max=(λ/γ)1/γ|\bar{\omega}^{\prime}_{m}|\sim\bar{\omega}_{max}=(\lambda/\gamma)^{1/\gamma} and the lower cutoff at |ω¯m||ω¯m||\bar{\omega}^{\prime}_{m}|\sim|\bar{\omega}_{m}|. Because the lower cutoff is finite, we can set Ω¯tot=0\bar{\Omega}_{tot}=0. Summing up the logarithmical series, we then obtain Abanov et al. (2001b)

Φ¯(ω¯m)=Φ¯0(1+logω¯max|ω¯m|+12log2ω¯max|ω¯m|+.)=Φ¯0elogω¯max/|ω¯m|=Φ¯0ω¯max|ω¯m|\bar{\Phi}(\bar{\omega}_{m})=\bar{\Phi}_{0}\left(1+\log{\frac{\bar{\omega}_{max}}{|\bar{\omega}_{m}|}}+\frac{1}{2}\log^{2}{\frac{\bar{\omega}_{max}}{|\bar{\omega}_{m}|}}+....\right)=\bar{\Phi}_{0}e^{\log{\bar{\omega}_{max}/|\bar{\omega}_{m}|}}=\bar{\Phi}_{0}\frac{\bar{\omega}_{max}}{|\bar{\omega}_{m}|} (11)

We see that the pairing susceptibility increases as ω¯m\bar{\omega}_{m} decreases, but remains finite and positive for all finite ω¯m\bar{\omega}_{m}, even when Ω¯tot=0\bar{\Omega}_{tot}=0. Re-doing calculations at a finite Ω¯tot\bar{\Omega}_{tot} we find the same result as in (11), with |ω¯m||\bar{\omega}_{m}| replaced by max(|ω¯m|,Ω¯¯tot)(|\bar{\omega}_{m}|,{\bar{\bar{\Omega}}}_{tot}). This implies that for γ>0\gamma>0, there is no signature of a pairing instability within the logarithmic approximation. The absence of a pairing instability within the logarithmic approximation generally implies that pairing does not develop at weak coupling and, if exists, is a threshold phenomenon. In our case, the dimensionless coupling in the series in Eq. (11) is a number equal to one, i.e., the problem we consider is not weak-coupling. A weak-coupling limit can be reached if we extend the model and make the pairing interaction parametrically smaller than the one in the particle-hole channel. In practice, this is done by multiplying the interaction in the pairing channel by 1/N1/N, where N>1N>1 (Refs. Raghu et al. (2015); Wang et al. (2016); Abanov et al. (2019); *Wu_19_1; Abanov and Chubukov (2020); Damia et al. (2020)), while the interaction in the particle-hole channel is left intact  111The extension to N>1N>1 was originally rationalized by extending the original model to matrix SU(N)SU(N), with integer NN, hence the notation. We treat NN as a continuous parameter. In the extended model, logω¯max/|ω¯m|\log{\bar{\omega}_{max}/|\bar{\omega}_{m}|} gets multiplied by 1/N1/N, and at large NN the problem becomes weak-coupling. We show below that indeed there is no pairing instability at large NN.

The case γ=0+\gamma=0+ falls in between BCS and γ>0\gamma>0 cases. Namely, we show below that the series of logarithms are not geometric, like the ones for γ>0\gamma>0, however by summing up the series one does find the scale of a pairing instability, like in BCS. We assume and than verify that for relevant ω¯m\bar{\omega}^{\prime}_{m}, λlog(Λ¯/|ω¯m|)\lambda\log{(\bar{\Lambda}/|\bar{\omega}^{\prime}_{m}|)} is small for λ1\lambda\ll 1, and neglect this term in the denominator of (9). We keep Ω¯tot\bar{\Omega}_{tot} non-zero to avoid divergencies and for simplicity set ω¯m\bar{\omega}_{m} and Ω¯tot\bar{\Omega}_{tot} to be comparable. To logarithmic accuracy, we can then view Φ¯\bar{\Phi} as a function of a single parameter Ω¯tot\bar{\Omega}_{tot}.

Because the interaction is logarithmic, perturbation series hold in λlog2(Λ¯/|ω¯m|)\lambda\log^{2}{(\bar{\Lambda}/|\bar{\omega}_{m}|)}. Evaluating the pairing vertex in order-by-order calculations, we find (see Appendix for details)

Φ¯(Ω¯tot)=Φ¯0(1+12λlog2Λ¯|Ω¯tot|+524λ2log4Λ¯|Ω¯tot|+61720λ3log6Λ¯|Ω¯tot|+)\bar{\Phi}(\bar{\Omega}_{tot})=\bar{\Phi}_{0}\left(1+\frac{1}{2}\lambda\log^{2}{\frac{\bar{\Lambda}}{|\bar{\Omega}_{tot}|}}+\frac{5}{24}\lambda^{2}\log^{4}{\frac{\bar{\Lambda}}{|\bar{\Omega}_{tot}|}}+\frac{61}{720}\lambda^{3}\log^{6}{\frac{\bar{\Lambda}}{|\bar{\Omega}_{tot}|}}+...\right) (12)

At a first glance, the coefficients in (12) are just some uncorrelated numbers. On a more careful look, we realize that the series fall into

Φ¯(Ω¯tot)=Φ¯01cos(λlog(Λ¯/|Ω¯tot|))\bar{\Phi}(\bar{\Omega}_{tot})=\bar{\Phi}_{0}\frac{1}{\cos{\left(\sqrt{\lambda}\log{(\bar{\Lambda}/|\bar{\Omega}_{tot}|)}\right)}} (13)

Accordingly, the pairing susceptibility diverges when the argument of cos\cos becomes π/2\pi/2, i.e., at Ω¯tot=Λ¯eπ/2λ\bar{\Omega}_{tot}=\bar{\Lambda}e^{-\pi/2\sqrt{\lambda}}. It is natural to associate this scale with TcT_{c} (Ref. Son (1999)). The susceptibility also diverges at a set of smaller Ω¯tot,n=Λ¯eπ(1+2n)/2λ\bar{\Omega}_{tot,n}=\bar{\Lambda}e^{-\pi(1+2n)/2\sqrt{\lambda}}, but here we focus only on the highest onset temperature. We see that TcT_{c} has exponential dependence on the coupling, like in BCS theory, however the argument of the exponent contains λ\sqrt{\lambda} rather than λ\lambda. This in turn justifies the neglect of the self-energy, because for relevant frequencies, λlog(Λ¯/|ω¯m|)λ1\lambda\log{(\bar{\Lambda}/|\bar{\omega}^{\prime}_{m}|)}\sim\sqrt{\lambda}\ll 1 (the corrections due to self-energy have been analyzed in Ref. Wang and Rischke (2002)).

To summarize, the case γ=0+\gamma=0+ is similar to BCS in the sense that pairing occurs for arbitrary weak coupling (TcT_{c} remains finite even if we replace λ\lambda by λ/N\lambda/N and set NN to be large). However, for and finite γ>0\gamma>0, there is no indication of the pairing instability within the logarithmic approximation.

V The differential equation for Φ¯(ω¯m)\bar{\Phi}(\bar{\omega}_{m})

We now analyze the linearized equation for the pairing vertex for γ>0\gamma>0 beyond the logarithmic approximation. To do this, we convert integral equation (8) into to a differential equation with certain boundary conditions and solve it non-perturbatively.

V.1 The case Λ¯=\bar{\Lambda}=\infty

We first consider the case Λ=\Lambda=\infty. We keep N1N\geq 1 as a parameter, as we need to verify our earlier conjecture that there is no pairing instability for large enough NN. To convert (8) into differential equation, we use the fact that at small γ\gamma, the integral in the r.h.s. of (8) predominantly comes from internal ω¯m\bar{\omega}^{\prime}_{m}, which are either substantially larger or substantially smaller than the external ω¯m\bar{\omega}_{m}. We then split the integral over ω¯m\bar{\omega}^{\prime}_{m} into two parts: in one we approximate |ω¯mω¯m||\bar{\omega}_{m}-\bar{\omega}^{\prime}_{m}| by |ω¯m||\bar{\omega}^{\prime}_{m}|, in the other by |ω¯m||\bar{\omega}_{m}|. Introducing z=|ω¯m|γz=|\bar{\omega}_{m}|^{\gamma}, we then simplify (8) to

Φ¯(z)=1Nγ[z𝑑yΦ¯(y)y(1+αy)+1z0z𝑑yΦ¯(y)1+αy]\bar{\Phi}(z)=\frac{1}{N\gamma}\left[\int^{\infty}_{z}dy\frac{\bar{\Phi}(y)}{y(1+\alpha y)}+\frac{1}{z}\int_{0}^{z}dy\frac{\bar{\Phi}(y)}{1+\alpha y}\right] (14)

where α=γ/λ\alpha=\gamma/\lambda (ω¯max=(1/α)1/γ\bar{\omega}_{max}=(1/\alpha)^{1/\gamma}). Differentiating this equation twice over zz and replacing Φ¯(z)\bar{\Phi}(z) by Δ¯(z)=αΦ¯(z)z/(1+αz)\bar{\Delta}(z)=\alpha\bar{\Phi}(z)z/(1+\alpha z), we obtain the second order differential gap equation in the form (Refs.   Wang et al. (2017, 2018, 2016); Abanov and Chubukov (2020))

(Δ¯diff(z)(1+αz))′′=(1/4bN2)Δ¯diff(z)z2,(\bar{\Delta}_{{\text{diff}}}(z)(1+\alpha z))^{{}^{\prime\prime}}=-\left(1/4-b_{N}^{2}\right)\frac{\bar{\Delta}_{{\text{diff}}}(z)}{z^{2}}, (15)

where ()′′=d2()/dz2(...)^{{}^{\prime\prime}}=d^{2}(...)/dz^{2} and

bN=141γNb_{N}=\sqrt{\frac{1}{4}-\frac{1}{\gamma N}} (16)

This Δ¯diff(z)\bar{\Delta}_{{\text{diff}}}(z) has to be real and satisfy the boundary conditions at large zz and at z=0z=0. At large zz, we expect perturbation theory to work, hence limzΔ¯diff(z)=Φ¯0\lim\limits_{z\rightarrow\infty}\bar{\Delta}_{{\text{diff}}}(z)=\bar{\Phi}_{0}. The boundary condition at z=0z=0 is set by the requirement that Δ¯diff(z)\bar{\Delta}_{{\text{diff}}}(z) is normalizable, i.e., that the ground state energy for a non-zero Δ¯diff(z)\bar{\Delta}_{{\text{diff}}}(z) must be free from divergencies. This requirement imposes the condition that dz(Δ¯diff2(z)/z2\int dz(\bar{\Delta}^{2}_{{\text{diff}}}(z)/z^{2} should be infra-red convergent (Refs. Abanov and Chubukov (2020); Yuzbashyan et al. ).

The solution of (15) is expressed via a Hypergeometric function, and the form of Δ¯diff(z)\bar{\Delta}_{{\text{diff}}}(z) depends on whether bNb_{N} is real or imaginary, i.e., whether NN is larger or smaller than Ncr=4/γN_{cr}=4/\gamma. When N>NcrN>N_{cr}, bNb_{N} is real, and the solution is

Δ¯diff(z)=C1S[bN,z]+C2S[bN,z]\bar{\Delta}_{{\text{diff}}}(z)=C_{1}S[b_{N},z]+C_{2}S[-b_{N},z] (17)

where

S[bN,z]=zbN+1/2F2[12+bN,32+bN,1+2bN,αz]1.S[b_{N},z]=z^{b_{N}+1/2}{{}_{2}}F{{}_{1}}\left[\frac{1}{2}+b_{N},\frac{3}{2}+b_{N},1+2b_{N},-\alpha z\right]. (18)

Both S[bN,z]S[b_{N},z] and S[bN,z]S[-b_{N},z] are sign-preserving functions. At small zz, S[±bN,z]z1/2±bNS[\pm b_{N},z]\approx z^{1/2\pm b_{N}}, and at large zz, S[±bN,z]S[\pm b_{N},z] tend to finite values 4±bNΓ(1±bN)πΓ(32±bN)\frac{4^{\pm b_{N}}\Gamma(1\pm b_{N})}{\sqrt{\pi}\Gamma\left(\frac{3}{2}\pm b_{N}\right)}. The boundary condition at z=0z=0 sets C2=0C_{2}=0, and the one at z1z\gg 1 sets the linear relation between C1C_{1} and Φ¯0\bar{\Phi}_{0}. We plot Δ¯diff(z)\bar{\Delta}_{{\text{diff}}}(z) for bN=0.2b_{N}=0.2 and α=1\alpha=1 in Fig. 2. We see that Δ¯diff(z)\bar{\Delta}_{{\text{diff}}}(z) remains positive for all zz, i.e., the normal state remains stable with respect to pairing. At γN1\gamma N\gg 1, bN1/21/(γN)b_{N}\approx 1/2-1/(\gamma N), and Δ¯diff(z)/Φ¯01/|z|1/(Nγ)\bar{\Delta}_{{\text{diff}}}(z)/\bar{\Phi}_{0}\propto 1/|z|^{1/(N\gamma)}. This is the same result that we obtained by summing up the logarithms. We see that in this parameter range the non-logarithmic terms just change the exponent from 1/N1/N to (γ/2)14/(Nγ)(\gamma/2)\sqrt{1-4/(N\gamma)}.

Refer to caption
Refer to caption
Figure 2: Left: The function Δ¯diff(z)\bar{\Delta}_{\text{diff}}(z), Eq. (17), for real bN=0.2b_{N}=0.2 (N>NcrN>N_{cr}) and α=1\alpha=1. Right: the same function for imaginary bN=2ib_{N}=2i (N<NcrN<N_{cr}). For real bNb_{N} perturbation theory around the normal state is regular, which indicates that there is no pairing instability. For imaginary bN=iβNb_{N}=i\beta_{N}, Δ¯diff(z)\bar{\Delta}_{\text{diff}}(z) oscillates at small zz. We argue in the text that this implies that the normal state becomes unstable towards pairing.

This behavior holds as long as bNb_{N} is real, i.e., N>Ncr=4/γN>N_{cr}=4/\gamma. For models with smaller N<NcrN<N_{cr}, including our original model with N=1N=1, bNb_{N} is imaginary, bN=iβNb_{N}=i\beta_{N}, were βN=1/(Nγ)1/4=(1/Nγ)1/2(1N/Ncr)1/2\beta_{N}=\sqrt{1/(N\gamma)-1/4}=(1/N\gamma)^{1/2}(1-N/N_{cr})^{1/2}. In this situation, the solution of the differential equation is

Δ¯diff(z)=Cz×Re(eiψziβNF2[12+iβN,32+iβN,1+2iβN,αz]1)\bar{\Delta}_{{\text{diff}}}(z)=C\sqrt{z}\times Re\left(e^{i\psi}z^{i\beta_{N}}{{}_{2}}F{{}_{1}}\left[\frac{1}{2}+i\beta_{N},\frac{3}{2}+i\beta_{N},1+2i\beta_{N},-\alpha z\right]\right) (19)

The boundary condition at z=0z=0 is satisfied as at small zz, Φ¯diff(z)\bar{\Phi}_{{\text{diff}}}(z) acquires a form of a free quantum particle with a ”coordinate” x=logzx=\log{z}. At z1z\gg 1, Δ¯diff(z)\bar{\Delta}_{{\text{diff}}}(z) tends to a constant for a generic ψ\psi, and the boundary condition sets up a linear relation between CC and Φ¯0\bar{\Phi}_{0}.

We plot Δ¯diff(z)\bar{\Delta}_{{\text{diff}}}(z) in Fig. (2). We see that the gap function now passes through a minimum at some zz^{*} and oscillates at smaller zz. The oscillating behavior at z<zz<z^{*} could not be obtained within perturbation theory starting from Φ¯(z)=Φ¯0\bar{\Phi}(z)=\bar{\Phi}_{0} as the kernel in Eq. (8) is entirely positive. This strongly suggests that the normal state is now unstable towards pairing, and (z)1/γ(z^{*})^{1/\gamma} sets the value of T¯c\bar{T}_{c}. To verify this, in Ref.  Wu et al. (2020b) we solved the gap equation at a finite TT. This calculation is a bit more tricky due to special behavior of fermions with Matsubara frequencies ωm=±πT\omega_{m}=\pm\pi T, but nevertheless it confirms the two key points of our T=0T=0 analysis that superconducting T¯c\bar{T}_{c} is finite and is generally of order zz^{*}.

Note that the value of zz^{*} depends on the magnitude of βN\beta_{N}, i.e., on the ratio Ncr/NN_{cr}/N. For small βN\beta_{N}, which holds when NN is only slightly below NcrN_{cr}, the first sign change of χpp(z)Δ¯diff/Φ¯0\chi_{pp}(z)\equiv\bar{\Delta}_{\text{diff}}/\bar{\Phi}_{0} occurs at small zea/βNz^{*}\sim e^{-a/\beta_{N}}, where a=O(1)a=O(1). For larger βN\beta_{N}, the first sign change occurs at αz1/(Nγ)1\alpha z^{*}\sim 1/(N\gamma)\geq 1, i.e., at zλ/Nγ2z^{*}\sim\lambda/N\gamma^{2}, or, equivalently, at ω¯mω¯max(1/(Nγ)1/γ\bar{\omega}_{m}\sim\bar{\omega}_{max}(1/(N\gamma)^{1/\gamma}. This holds for the original model with N=1N=1.

We note in passing that Δ¯diff(z)\bar{\Delta}_{{\text{diff}}}(z) contains two parameters: the overall factor CC and the phase factor ψ\psi. The boundary condition at z1z\gg 1 then still leaves the freedom of, e.g., choosing ψ\psi for a given CΦ¯0C\sim\bar{\Phi}_{0}. This extra freedom comes about because it turns out Abanov and Chubukov (2020) that the homogeneous gap equation (the one without Φ¯0=0\bar{\Phi}_{0}=0) by itself has a non-zero solution at T=0T=0, such that Δ¯diff(z)\bar{\Delta}_{{\text{diff}}}(z) is the sum of the induced and the homogeneous solutions of (8). To demonstrate this more explicitly, we set z>1/αz>1/\alpha, use the fact that in this range a hypergeometric function reduces to a combination of Bessel and Neumann functions, and re-express (19) as

Δ¯diff(z)1z[A1J1(4βN2+1αz)+A2Y1(4βN2+1αz)],\bar{\Delta}_{{\text{diff}}}(z)\sim\frac{1}{\sqrt{z}}\left[A_{1}J_{1}\left(\sqrt{\frac{4\beta^{2}_{N}+1}{\alpha z}}\right)+A_{2}Y_{1}\left(\sqrt{\frac{4\beta^{2}_{N}+1}{\alpha z}}\right)\right], (20)

where J1J_{1} and Y1Y_{1} are Bessel and Neumann functions, respectively. At small value of the argument, J1(p)pJ_{1}(p)\approx p and Y1(p)1/pY_{1}(p)\approx 1/p, i.e., the first term vanishes at large zz, while the second one tends to a constant. Then A2A_{2} is uniquely determined by the boundary condition set by Φ¯0\bar{\Phi}_{0}, while the A1A_{1} piece is the solution of the equation without Φ¯0\bar{\Phi}_{0}.

The existence of the solution of the linearized gap equation at T=0T=0 (well below TcT_{c}) is a highly non-trivial feature of pairing at a QCP, that affects fluctuation corrections to superconducting order parameter. For our purposes, however, we only need to analyze only the induced solution to get an estimate of TcT_{c}. We can set A1=0A_{1}=0 in (20) and determine A2A_{2} from the boundary condition at zz\to\infty.

V.2 The case of a finite Λ¯\bar{\Lambda}

At this stage, we have two different energy scales, which we identified with TcT_{c}. Namely, at a finite Λ¯\bar{\Lambda} and γ0\gamma\to 0, we found T¯cΛ¯eπ/(2λ)\bar{T}_{c}\sim\bar{\Lambda}e^{-\pi/(2\sqrt{\lambda})}. At Λ¯\bar{\Lambda}\to\infty and finite γ\gamma, we found T¯cω¯max(1/Nγ)1/γ\bar{T}_{c}\sim\bar{\omega}_{max}(1/N\gamma)^{1/\gamma}. We now analyze the crossover between the two energies. For definiteness, we consider the original model with N=1N=1.

We argued earlier that for the pairing at γ=0+\gamma=0+, fermions can be treated as free quasiparticles, because for relevant fermions the ratio Σ¯(ω¯m)/ω¯mλ1\bar{\Sigma}(\bar{\omega}_{m})/\bar{\omega}_{m}\sim\sqrt{\lambda}\ll 1. The same holds for the case Λ¯=\bar{\Lambda}=\infty and γ>0\gamma>0. Here, the ratio Σ¯(ω¯m)/ω¯m1/(αz)\bar{\Sigma}(\bar{\omega}_{m})/\bar{\omega}_{m}\sim 1/(\alpha z), and for zzz\sim z^{*}, Σ¯(ω¯m)/ω¯mγ1\bar{\Sigma}(\bar{\omega}_{m})/\bar{\omega}_{m}\sim\gamma\ll 1. We now use this simplification and re-analyze the differential equation for Δ¯(z)\bar{\Delta}(z) at a finite Λ¯\bar{\Lambda}.

One can verify that the differential equation retains the same for as for Λ¯=\bar{\Lambda}=\infty:

(Δ¯diff(z)z)′′=λγ2Δ¯diff(z)z2,(\bar{\Delta}_{{\text{diff}}}(z)z)^{{}^{\prime\prime}}=-\frac{\lambda}{\gamma^{2}}\frac{\bar{\Delta}_{{\text{diff}}}(z)}{z^{2}}, (21)

and its solution is still a linear combination of Bessel and Neumann functions, Eq. (20). However, we have an extra requirement

Δ¯diff(Λ¯)=0,\bar{\Delta}_{{\text{diff}}}(\bar{\Lambda}^{*})=0, (22)

where Λ¯=(Λ¯)γ\bar{\Lambda}^{*}=(\bar{\Lambda})^{\gamma}. There is no solution of the homogeneous equation at a finite Λ¯\bar{\Lambda}, and Eq. (22) together with the boundary condition Δ¯diff()=Φ¯0\bar{\Delta}_{{\text{diff}}}(\infty)=\bar{\Phi}_{0} uniquely specify the coefficients A1A_{1} and A2A_{2} in (20):

Δ¯diff(z)=Cz×\displaystyle\bar{\Delta}_{\text{diff}}(z)=\frac{C}{\sqrt{z}}\times
[J1(2(γαz)1/2)Y1(2(γαΛ¯)1/2)Y1(2(γαz)1/2)J1(2(γαΛ¯)1/2)],\displaystyle\left[J_{1}\left(\frac{2}{(\gamma\alpha z)^{1/2}}\right)Y_{1}\left(\frac{2}{(\gamma\alpha\bar{\Lambda}^{*})^{1/2}}\right)-Y_{1}\left(\frac{2}{(\gamma\alpha z)^{1/2}}\right)J_{1}\left(\frac{2}{(\gamma\alpha\bar{\Lambda}^{*})^{1/2}}\right)\right], (23)

where CΦ¯0C\sim\bar{\Phi}_{0}. It is convenient to introduce a dimensionless parameter

B=γαΛ¯=γ2λ(Λ¯)γ.B=\gamma\alpha\bar{\Lambda}^{*}=\frac{\gamma^{2}}{\lambda}(\bar{\Lambda})^{\gamma}. (24)

In the left panel of Fig. 3 we plot Δ¯ diff(ω¯m)\bar{\Delta}_{\text{ diff}}(\bar{\omega}_{m}) for a representative B=0.3B=0.3. We see that the gap function is regular at |ω¯m|Λ¯|\bar{\omega}_{m}|\leq\bar{\Lambda}, but passes through an extremum and oscillates at smaller ω¯m\bar{\omega}_{m}. The position of the first extremum depends on BB. We now show that the two forms of T¯c\bar{T}_{c}, which we found earlier, correspond to the limits B1B\ll 1 and B1B\gg 1.

When B1B\ll 1, one can use the forms of Bessel and Neumann functions at large argument,

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)}, (25)

and obtain

Δ¯ diff(z)Φ¯0z1/4sin(2B1/22(αγz)1/2)=Cz1/4sin2B1/2(1(Λ¯z)1/2)\bar{\Delta}_{\text{ diff}}(z)\propto\frac{\bar{\Phi}_{0}}{z^{1/4}}\sin{\left(\frac{2}{B^{1/2}}-\frac{2}{(\alpha\gamma z)^{1/2}}\right)}=\frac{C}{z^{1/4}}\sin{\frac{2}{B^{1/2}}\left(1-\left(\frac{\bar{\Lambda}^{*}}{z}\right)^{1/2}\right)} (26)

In the original variables ω¯m\bar{\omega}_{m} and Λ\Lambda, this reduces, to the leading order in γ\gamma, to

Δ¯diff(ω¯m)Φ¯0|ω¯m|γ/4sin(λlog|ω¯m|Λ¯).\bar{\Delta}_{{\text{diff}}}(\bar{\omega}_{m})\propto\frac{\bar{\Phi}_{0}}{|\bar{\omega}_{m}|^{\gamma/4}}\sin{\left(\sqrt{\lambda}\log{\frac{|\bar{\omega}_{m}|}{\bar{\Lambda}}}\right)}. (27)
Refer to caption
Refer to caption
Figure 3: Left: Δ¯diff(z)\bar{\Delta}_{\text{diff}}(z) from (23) for B=0.3B=0.3, and Λ¯=10\bar{\Lambda}^{\ast}=10. Right: the sketch of the evolution of TcT_{c} as a function of BB.

Associating T¯c\bar{T}_{c} with the position of the first extremum of Δ¯ diff(ω¯m)\bar{\Delta}_{\text{ diff}}(\bar{\omega}_{m}), we find T¯cΛ¯eπ/(2λ)\bar{T}_{c}\sim\bar{\Lambda}e^{-\pi/(2\sqrt{\lambda})}. This coincides with the estimate of T¯c\bar{T}_{c} from the analysis of the pairing susceptibility as a function of the total frequency of two fermions in a pair (see Eq. (13)).

In the opposite limit B1B\gg 1, we use J1(2/B1/2)1/B1/2J_{1}(2/B^{1/2})\sim 1/B^{1/2} and Y1(2/B1/2)B1/2Y_{1}(2/B^{1/2})\sim B^{1/2} and keep only Y1(2/B1/2)Y_{1}(2/B^{1/2}). We then obtain from (23) that the dependence on Λ¯\bar{\Lambda} appears only in the overall factor:

Δ¯diff(z)1zJ1(2(γαz)1/2),\bar{\Delta}_{\text{diff}}(z)\propto\frac{1}{\sqrt{z}}J_{1}\left(\frac{2}{(\gamma\alpha z)^{1/2}}\right), (28)

or, in terms of ω¯m\bar{\omega}_{m},

Δ¯diff(ω¯m)1|ω¯m|γ/2J1(2B1/2(Λ¯|ω¯m|)γ/2)\bar{\Delta}_{\text{diff}}(\bar{\omega}_{m})\propto\frac{1}{|\bar{\omega}_{m}|^{\gamma/2}}J_{1}\left(\frac{2}{B^{1/2}}\left(\frac{\bar{\Lambda}}{|\bar{\omega}_{m}|}\right)^{\gamma/2}\right) (29)

The first extremum is now located at ω¯m=(z)1/γT¯cΛ¯/B1/γ\bar{\omega}^{*}_{m}=(z^{*})^{1/\gamma}\sim\bar{T}_{c}\sim\bar{\Lambda}/B^{1/\gamma}, i.e., T¯c(λ/γ)1/γ(1/γ)1/γ\bar{T}_{c}\sim(\lambda/\gamma)^{1/\gamma}(1/\gamma)^{1/\gamma}. This agrees with the analysis at Λ¯=\bar{\Lambda}=\infty.

In the right panel in Fig. 3 we sketch the evolution of T¯c\bar{T}_{c} as a function of BB. In terms of BB, T¯c\bar{T}_{c} scales as Λ¯/B1/γ\bar{\Lambda}/B^{1/\gamma} for B1B\gg 1 (the limit of large Λ¯\bar{\Lambda} and finite γ\gamma), and saturates at Λeπ/2λ\Lambda e^{-\pi/2\sqrt{\lambda}} for B1B\ll 1 (the limit γ=0+\gamma=0+ and finite Λ¯\bar{\Lambda}). The crossover between the two regimes occurs at Beπγ/(2λ)1B\sim e^{\pi\gamma/(2\sqrt{\lambda})}\approx 1.

We note in passing that Eq. (27) could also be obtained directly, by converting the gap equation for γ=0+\gamma=0+ into the differential equation Chubukov and Schmalian (2005). For this we depart from Eq. (9), neglect the self-energy term in the denominator, split the integral over ωm\omega^{\prime}_{m} into contributions from ωmωm\omega^{\prime}_{m}\gg\omega_{m} and ωmωm\omega^{\prime}_{m}\ll\omega_{m}, as we did in the derivation of (14), and introduce logarithmic variables x=logΛ/|ωm|=logΛ¯/|ω¯m|x=\log{\Lambda/|\omega_{m}|}=\log{\bar{\Lambda}/|\bar{\omega}_{m}|}, x=logΛ¯/|ω¯m|x^{\prime}=\log{\bar{\Lambda}/|\bar{\omega}^{\prime}_{m}|}. We then obtain

Φ(x)=λ[0x𝑑xxΦ(x)+xx𝑑xΦ(x)]\Phi(x)=\lambda\left[\int_{0}^{x}dx^{\prime}x^{\prime}\Phi(x^{\prime})+x\int_{x}^{\infty}dx^{\prime}\Phi(x^{\prime})\right] (30)

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

Φ′′(y)=λγ2Φ(y)\Phi^{{}^{\prime\prime}}(y)=-\frac{\lambda}{\gamma^{2}}\Phi(y) (31)

This equation is valid only for 0<y<γ2/λ0<y<\gamma^{2}/\lambda, and the two boundary conditions are Φ(y=0)=Φ0\Phi(y=0)=\Phi_{0} and Φ(y=γ2/λ)=0\Phi(y=\gamma^{2}/\lambda)=0. The solution of (31), which satisfies the boundary conditions, is, for small λ\lambda,

Φ(y)=Φ¯0λγsin(λγ(yγ2λ))\Phi(y)=-\bar{\Phi}_{0}\frac{\sqrt{\lambda}}{\gamma}\sin{\left(\frac{\sqrt{\lambda}}{\gamma}\left(y-\frac{\gamma^{2}}{\lambda}\right)\right)} (32)

In the original units, this becomes

Φ(ω¯m)=Φ¯0λγsin(λlogΛ¯|ω¯m|γλ)\Phi(\bar{\omega}_{m})=-\bar{\Phi}_{0}\frac{\sqrt{\lambda}}{\gamma}\sin{\left(\sqrt{\lambda}\log{\frac{\bar{\Lambda}}{|\bar{\omega}_{m}|}}-\frac{\gamma}{\sqrt{\lambda}}\right)} (33)

We plot Φ(ω¯m)\Phi(\bar{\omega}_{m}) in Fig. 4. We see that Φ(ω¯m)\Phi(\bar{\omega}_{m}) is sign-preserving when logΛ¯|ω¯m|\log{\frac{\bar{\Lambda}}{|\bar{\omega}_{m}|}} is small, but oscillates when logΛ¯|ω¯m|\log{\frac{\bar{\Lambda}}{|\bar{\omega}_{m}|}} gets larger. Associating the position of the first extremum of Φ(ω¯m)\Phi(\bar{\omega}_{m}) with T¯c\bar{T}_{c}, as we did before, we find the same T¯cΛ¯eπ/(2λ)\bar{T}_{c}\sim\bar{\Lambda}e^{-\pi/(2\sqrt{\lambda})} as at B1B\ll 1 in our earlier treatment.

Refer to caption
Figure 4: The function Φ(ω¯m)\Phi(\bar{\omega}_{m}) from Eq. (33). It passes through a maximum at ω¯mT¯cΛ¯eπ/(2λ)\bar{\omega}_{m}\sim\bar{T}_{c}\sim\bar{\Lambda}e^{-\pi/(2\sqrt{\lambda})} and oscillates at smaller ω¯m\bar{\omega}_{m} as a simple trigonometric function of logΛ¯|ω¯m|\log{\frac{\bar{\Lambda}}{|\bar{\omega}_{m}|}} (shown in the insert).

VI RG analysis

The difference between γ=0+\gamma=0+ and at a finite γ\gamma can be also understood by analyzing the flow of the running 4-fermion pairing vertex within the RG formalism. The RG equations for small γ>0\gamma>0 have been derived in Refs.  Metlitski et al. (2015); Wang et al. (2017) and for γ=0+\gamma=0+ in Ref. Son (1999) (see also Refs. Schäfer and Wilczek (1999); Pisarski and Rischke (2000)).

The key point for the RG analysis is that at small but finite γ\gamma, the interaction V(Ω¯m)V(\bar{\Omega}_{m}) in (1) can be approximated by

V(Ω¯m)=λ|Ω¯m|γlogΛ¯|Ω¯m|V(\bar{\Omega}_{m})=\frac{\lambda}{|\bar{\Omega}_{m}|^{\gamma}}\log{\frac{\bar{\Lambda}}{|\bar{\Omega}_{m}|}} (34)

and both the logarithmic and the power-law dependence have to be kept. This interaction can be further re-expressed in terms of L=logΛ¯|Ω¯m|L=\log{\frac{\bar{\Lambda}}{|\bar{\Omega}_{m}|}} as the product of LL and the running effective V~(L)=V~(0)eγL{\tilde{V}}(L)={\tilde{V}}(0)e^{\gamma L}, where V~(0)=λ(1/Λ¯)γ=γ2/B{\tilde{V}}(0)=\lambda(1/\bar{\Lambda})^{\gamma}=\gamma^{2}/B.

The interaction V(L)=V~(L)LV(L)={\tilde{V}}(L)L acts as the source for the running 4-fermion pairing vertex, which we label as VSC(L)V_{SC}(L). Without the source, VSC(L)V_{SC}(L) would obey a BCS RG equation V˙SC(L)=VSC2(L){\dot{V}}_{SC}(L)=V^{2}_{SC}(L). With the source, the equation becomes

V˙SC(L)=VSC2(L)+V~(L){\dot{V}}_{SC}(L)=V^{2}_{SC}(L)+{\tilde{V}}(L) (35)

The solution of this equation, subject to VSC(0)=0V_{SC}(0)=0 (no pairing without the source), is  Metlitski et al. (2015); Wang et al. (2017)

VSC(L)=γB1/2eγL/2J1(2B1/2eγL/2)Y1(2B1/2)Y1(2B1/2eγL/2)J1(2B1/2)J0(2B1/2eγL/2)Y1(2B1/2)Y0(2B1/2eγL/2)J1(2B1/2)V_{SC}(L)=\frac{\gamma}{B^{1/2}}e^{\gamma L/2}\frac{J_{1}\left(\frac{2}{B^{1/2}}e^{\gamma L/2}\right)Y_{1}\left(\frac{2}{B^{1/2}}\right)-Y_{1}\left(\frac{2}{B^{1/2}}e^{\gamma L/2}\right)J_{1}\left(\frac{2}{B^{1/2}}\right)}{J_{0}\left(\frac{2}{B^{1/2}}e^{\gamma L/2}\right)Y_{1}\left(\frac{2}{B^{1/2}}\right)-Y_{0}\left(\frac{2}{B^{1/2}}e^{\gamma L/2}\right)J_{1}\left(\frac{2}{B^{1/2}}\right)} (36)

For B1B\ll 1, relevant values of the arguments of Bessel and Neumann functions are large, and using using their asymptotic forms, Eq. (25), we find that, to leading order in γ\gamma,

VSC(L)tan(λL)V_{SC}(L)\propto\tan{(\sqrt{\lambda}L)} (37)

The 4-fermion vertex diverges at the same ω¯m=Λeπ/(2λ)\bar{\omega}^{*}_{m}=\Lambda e^{-\pi/(2\sqrt{\lambda})} that we obtained before.

In the opposite limit B1B\ll 1, we use that Y1(2/B1/2)J1(2/B1/2)Y_{1}(2/B^{1/2})\gg J_{1}(2/B^{1/2}). Keeping only Y1(2/B1/2)Y_{1}(2/B^{1/2}) in (36), we obtain

VSC(L)=γB1/2eγL/2J1(2B1/2eγL/2)J0(2B1/2eγL/2)V_{SC}(L)=\frac{\gamma}{B^{1/2}}e^{\gamma L/2}\frac{J_{1}\left(\frac{2}{B^{1/2}}e^{\gamma L/2}\right)}{J_{0}\left(\frac{2}{B^{1/2}}e^{\gamma L/2}\right)} (38)

The 4-fermion vertex now diverges at the first zero of J0(p)J_{0}(p), i.e., at

2B1/2(Λ¯|ω¯m|)γ/2=2(γλz)1/22.4\frac{2}{B^{1/2}}\left(\frac{\bar{\Lambda}}{|\bar{\omega}_{m}|}\right)^{\gamma/2}=\frac{2}{(\gamma\lambda z)^{1/2}}\approx 2.4 (39)

Solving for ω¯m\bar{\omega}_{m}, we obtain ω¯m=(z)1/γω¯max(1/γ)1/γ\bar{\omega}^{*}_{m}=(z^{*})^{1/\gamma}\sim\bar{\omega}_{max}(1/\gamma)^{1/\gamma}. This agrees the result of our earlier analysis of the case Λ¯\bar{\Lambda}\to\infty and γ>0\gamma>0.

VII Conclusions

The goal of this work was to analyze the crossover from a conventional BCS pairing by a massive boson to a pairing at a quantum-critical point towards some particle-hole order, when the pairing boson becomes massless. We considered a subset of quantum-critical models, in which the pairing boson is a slow mode compared to fermions, and an effective low-energy theory is purely dynamical, with an effective dynamical interaction V(Ωm)1/|Ωm|γV(\Omega_{m})\propto 1/|\Omega_{m}|^{\gamma}, up to some upper cutoff Λ\Lambda. The case γ=0\gamma=0 corresponds to BCS theory od pairing from a Fermi liquid. We considered the pairing at a small, but finite γ\gamma, when the normal state is a NFL, and the case γ=0+\gamma=0+, when V(Ωm)logΛ/|Ωm|V(\Omega_{m})\propto\log{\Lambda/|\Omega_{m}|}, and the normal state is marginal Fermi liquid. We demonstrated that for γ=0+\gamma=0+, the pairing instability can still be detected by summing up series of Cooper logarithms, but for a finite γ\gamma one needs to go beyond the leading logarithmic approximation to analyze the pairing. We argued that in this last case, the pairing occurs only if the interaction exceeds some finite threshold. We approximated the original integer gap equation by the differential one and solved it. This allowed us to identify the threshold at γ>0\gamma>0 and the frequency scale, associated with superconductivity, once the interaction exceeds the threshold, We found the crossover between the pairing at a finite γ\gamma and at γ=0+\gamma=0+ and identified the parameter responsible for the crossover. We obtained the same crossover by analyzing the non-BCS RG equation for the running 4-fermion pairing vertex.

Acknowledgements.
We thank B. Altshuler, A. Finkelstein, S. Karchu, S. Kivelson, I. Mazin, M. Metlitski, V. Pokrovsky, S. Raghu, S. Sachdev, T. Senthil, J. Schmalian, D. Son, G. Torroba, E. Yuzbashyan, C. Varma, Y. Wang, Y. Wu, and S-S Zhang for useful discussions. The work by AVC was supported by the NSF DMR-1834856. AVC acknowledges the hospitality of KITP at UCSB, where part of the work has been conducted. The research at KITP is supported by the National Science Foundation under Grant No. NSF PHY-1748958.

Appendix A Derivation of Eq. (12)

Refer to caption
Figure 5: Diagrammatic representation of one-loop and two-loop renormalization of the pairing vertex Φ0\Phi_{0} for the case γ=0+\gamma=0+. Solid lines are propagators of free fermions, wavy lines are V(Ω¯m)=λlogΛ¯/|Ω¯m|V(\bar{\Omega}_{m})=\lambda\log{\bar{\Lambda}/|\bar{\Omega}_{m}|}, and the two-fermion vertices are Φ0\Phi_{0}. For definiteness, we set the frequencies of external fermions to be Ωtot/2\Omega_{tot}/2.

In this Appendix we present some details of the derivation of order-by-order logarithmic renormalization of the pairing vertex Φ¯(Ω¯tot)\bar{\Phi}(\bar{\Omega}_{tot}) for the case γ=0+\gamma=0+. The Eliashberg equation for Φ¯(Ω¯tot)\bar{\Phi}(\bar{\Omega}_{tot}) is Eq. (9). Adding a constant Φ¯0\bar{\Phi}_{0} to the r.h.s. and expanding in powers of the coupling λ\lambda, we obtain the series in λlog2Λ¯|Ω¯tot|\lambda\log^{2}{\frac{\bar{\Lambda}}{|\bar{\Omega}_{tot}|}}, where, we remind, Λ¯\bar{\Lambda} in the upper cutoff, and all quantities are in units of our base energy g¯{\bar{g}}. As we discussed in the main text, fermionic self-energy (the log term in the denominator of (9)) is small at relevant logΛ¯|Ω¯tot|1/λ\log{\frac{\bar{\Lambda}}{|\bar{\Omega}_{tot}|}}\sim 1/\sqrt{\lambda} and can be safely neglected.

We show the calculations at one-loop and two-loop order. The corresponding renormalizations are graphically presented in Fig. 5. In this Figure, the two-fermion vertex is Φ¯0\bar{\Phi}_{0}, the wavy line is V(Ω¯m)=λlogΛ¯|Ω¯m|V(\bar{\Omega}_{m})=\lambda\log{\frac{\bar{\Lambda}}{|\bar{\Omega}_{m}|}}, and we set both external frequencies to be Ω¯tot/2\bar{\Omega}_{tot}/2. The one-loop result is obtained by integrating over a single internal frequency and is

λ|Ω¯tot|Λ¯dω¯mω¯mlogΛ¯|ω¯m|=λ2(logΛ¯|Ω¯tot|)2\lambda\int_{|\bar{\Omega}_{tot}|}^{\bar{\Lambda}}\frac{d\bar{\omega}_{m}}{\bar{\omega}_{m}}\log{\frac{\bar{\Lambda}}{|\bar{\omega}_{m}|}}=\frac{\lambda}{2}\left(\log\frac{\bar{\Lambda}}{|\bar{\Omega}_{tot}|}\right)^{2} (40)

At two-loop order we have to integrate over two internal frequencies ω¯m\bar{\omega}_{m} and ω¯m\bar{\omega}^{\prime}_{m}. We introduce ω¯m=Ω¯totx\bar{\omega}_{m}=\bar{\Omega}_{tot}x, ω¯m=Ω¯toty\bar{\omega}^{\prime}_{m}=\bar{\Omega}_{tot}y and use the fact that the leading logarithm comes from x,y1x,y\gg 1. Taking this limit, we obtain the two-loop correction to Φ¯0\bar{\Phi}_{0} in the form

λ21Λ¯/|Ω¯tot|dxx1Λ¯/|Ω¯tot|dyy[(logΛ¯|Ω¯tot|+log1x)(logΛ¯|Ω¯tot|+12log1|xy|+12log1x+y)]\lambda^{2}\int_{\sim 1}^{\bar{\Lambda}/|\bar{\Omega}_{tot}|}\frac{dx}{x}\int_{\sim 1}^{\bar{\Lambda}/|\bar{\Omega}_{tot}|}\frac{dy}{y}\left[\left(\log\frac{\bar{\Lambda}}{|\bar{\Omega}_{tot}|}+\log{\frac{1}{x}}\right)\left(\log\frac{\bar{\Lambda}}{|\bar{\Omega}_{tot}|}+\frac{1}{2}\log{\frac{1}{|x-y|}}+\frac{1}{2}\log{\frac{1}{x+y}}\right)\right] (41)

A simple analysis shows that the highest power of logΛ¯|Ω¯tot|\log\frac{\bar{\Lambda}}{|\bar{\Omega}_{tot}|} comes from the ranges xyx\gg y and yxy\gg x. Evaluating the contributions from these regions, we obtain

1Λ¯/|Ω¯tot|dxx1Λ¯/|Ω¯tot|dyy[log1x(12log1|xy|+12log1x+y)]=38(logΛ¯|Ω¯tot|)4\displaystyle\int_{\sim 1}^{\bar{\Lambda}/|\bar{\Omega}_{tot}|}\frac{dx}{x}\int_{\sim 1}^{\bar{\Lambda}/|\bar{\Omega}_{tot}|}\frac{dy}{y}\left[\log{\frac{1}{x}}\left(\frac{1}{2}\log{\frac{1}{|x-y|}}+\frac{1}{2}\log{\frac{1}{x+y}}\right)\right]=\frac{3}{8}\left(\log\frac{\bar{\Lambda}}{|\bar{\Omega}_{tot}|}\right)^{4}
12logΛ¯|Ω¯tot|×1Λ¯/|Ω¯tot|dxx1Λ¯/|Ω¯tot|dyy(log1|xy|+log1x+y)=23(logΛ¯|Ω¯tot|)4\displaystyle\frac{1}{2}\log\frac{\bar{\Lambda}}{|\bar{\Omega}_{tot}|}\times\int_{\sim 1}^{\bar{\Lambda}/|\bar{\Omega}_{tot}|}\frac{dx}{x}\int_{\sim 1}^{\bar{\Lambda}/|\bar{\Omega}_{tot}|}\frac{dy}{y}\left(\log{\frac{1}{|x-y|}}+\log{\frac{1}{x+y}}\right)=-\frac{2}{3}\left(\log\frac{\bar{\Lambda}}{|\bar{\Omega}_{tot}|}\right)^{4}
logΛ¯|Ω¯tot|×1Λ¯/|Ω¯tot|dxx1Λ¯/|Ω¯tot|dyylog1x=12(logΛ¯|Ω¯tot|)4\displaystyle\log\frac{\bar{\Lambda}}{|\bar{\Omega}_{tot}|}\times\int_{\sim 1}^{\bar{\Lambda}/|\bar{\Omega}_{tot}|}\frac{dx}{x}\int_{\sim 1}^{\bar{\Lambda}/|\bar{\Omega}_{tot}|}\frac{dy}{y}\log{\frac{1}{x}}=-\frac{1}{2}\left(\log\frac{\bar{\Lambda}}{|\bar{\Omega}_{tot}|}\right)^{4} (42)

Collecting all contributions, we obtain for the two-loop renormalization

Aλ2(logΛ¯|Ω¯tot|)4,A=11223+38=524A\lambda^{2}\left(\log\frac{\bar{\Lambda}}{|\bar{\Omega}_{tot}|}\right)^{4},~{}~{}A=1-\frac{1}{2}-\frac{2}{3}+\frac{3}{8}=\frac{5}{24} (43)

This leads to Eq. (12) in the main text.

References