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

The effect of superconducting fluctuations on the ac conductivity of a 2D electron system in the diffusive regime

I. S. Burmistrov L. D. Landau Institute for Theoretical Physics, Semenova 1a, 142432, Chernogolovka, Russia
Abstract

We report a complete analytical expression for the one-loop correction to the ac conductivity σ(ω)\sigma(\omega) of a disordered two-dimensional electron system in the diffusive regime. The obtained expression includes the weak localization and Altshuler–Aronov corrections as well as the corrections due to superconducting fluctuations above superconducting transition temperature. The derived expression has no 1/(iω)1/(i\omega) divergency in the static limit, ω0\omega\to 0, in agreement with general expectations for the normal state conductivity of a disordered electron system.

keywords:
electron transport , superconducting fluctuations , quantum corrections
journal: Annals of Physics

1 Introduction

The corrections to the physical observables of an electron system due to the superconducting fluctuations are the subject of research with more than 50 years old history (see Refs. [1, 2] for a review). Recently the study of superconducting fluctuations has gained a significance as a tool to elucidate the fundamental aspects of a superconducting state. The conductivity in the normal state is among physical observables which are affected significantly by superconducting fluctuations. Near the superconducting transition temperature TcT_{c}, the most substantial contributions to the dc conductivity are due to Aslamazov–Larkin [5, 6] and Maki–Thompson [3, 4] processes. While the dc conductivity is sensitive to the position of TcT_{c} only, the ac conductivity contains information about the energy and time scales involved. The experimental studies of the microwave conductivity near the superconducting transition in thin films were pioneered in Refs. [7, 8, 9]. Recently, the ac conductivity measurements have been used to elucidate physics behind superconductor-insulator transitions in thin films [10, 11, 12, 13, 14, 15].

It is well understood theoretically [16] that the conductivity corrections due to superconducting fluctuations in disordered electron systems in the diffusive regime stand in the same row as the weak localization [17] and Altshuler–Aronov corrections [18]. The contributions to the conductivity due to pairing fluctuations are nothing but quantum corrections due to interaction in the Cooper channel dressed by the scattering off a random potential. Although the final expressions for the weak localization and Altshuler–Aronov corrections to the ac conductivity σ(ω)\sigma(\omega) are well established [19, 20], the corresponding expression for the contribution due to superconducting fluctuations is still absent in the literature. The point is that in the diagrammatic approach the pairing conductivity is given by the sum of ten diagrams [1, 2]. Some diagrams produce contributions proportional to 1/(iω)1/(i\omega) in the static limit, ω0\omega\to 0. However, the sum of all ten diagrams is expected to give a finite contribution to the dc conductivity in the normal state. Only recently this problem has been finally solved and the general expression for the superconducting pairing contribution to the dc conductivity has been established. For the diffusive regime it was derived with the help of the Keldysh path integral and Usadel equation [21]. In the ballistic regime the fluctuation corrections to the dc conductivity were computed by means of a standard diagrammatic approach [22]. An attempt to obtain a general expression for the fluctuation correction to the ac conductivity, σ(ω)\sigma(\omega), was performed in Ref. [23] with the help of the Keldysh nonlinear sigma model (see Ref. [24] for a review). However, the expression derived in Ref. [23] diverges as 1/(iω)1/(i\omega) in the static limit, ω0\omega\to 0.

In this paper we report the general analytical expression for the quantum correction to the ac conductivity of a disordered electron system in the diffusive regime which includes the weak localization and Altshuler–Aronov contributions and contributions due to superconducting fluctuations above the transition temperature. We derived our results with the help of the replica Finkel’stein nonlinear sigma model (NLσ\sigmaM) (see Refs. [25, 26] for a review). In order to find σ(ω)\sigma(\omega) we performed the analytic continuation from Matsubara to real frequencies. We emphasize that our result for the contributions to σ(ω)\sigma(\omega) due to superconducting fluctuations has no 1/(iω)1/(i\omega) divergence as ω0\omega\to 0. In the static limit, ω0\omega\to 0, our expression reproduces the results reported for the dc conductivity in Refs. [21, 22].

The outline of the paper is as follows. In Sec. 2 we introduce the formalism of the Finkel’stein NLσ\sigmaM. The results of the one-loop computation of the ac conductivity are given in Sec. 3. In Sec. 4 the behavior of different contributions to the ac conductivity due to superconducting fluctuations is analysed. We finish the paper with conclusion (Sec. 5). Some technical details are given in Appendices.

2 Formalism

2.1 Finkel’stein NLσ\sigmaM action

The action of the Finkel’stein NLσ\sigmaM is given as the sum of the non-interacting NLσ\sigmaM, SσS_{\sigma}, and contributions due to electron-electron interactions, Sint(ρ)S_{\rm int}^{(\rho)} (the particle-hole singlet channel), Sint(σ)S_{\rm int}^{(\sigma)} (the particle-hole triplet channel), and Sint(c)S_{\rm int}^{(c)} (the particle-particle channel) (see Refs. [25, 26, 27] for a review):

S=Sσ+Sint(ρ)+Sint(σ)+Sint(c),\displaystyle S=S_{\sigma}+S_{\rm int}^{(\rho)}+S_{\rm int}^{(\sigma)}+S_{\rm int}^{(c)}, (1)

where

Sσ\displaystyle S_{\sigma} =g32d𝒓Tr(Q)2+4πTZωd𝒓TrηQ,\displaystyle=-\frac{g}{32}\int d\bm{r}\operatorname{Tr}(\nabla Q)^{2}+4\pi TZ_{\omega}\int d\bm{r}\operatorname{Tr}\eta Q, (2a)
Sint(ρ)\displaystyle S_{\rm int}^{(\rho)} =πT4Γsα,nr=0,3𝑑𝒓TrInαtr0QTrInαtr0Q,\displaystyle=-\frac{\pi T}{4}\Gamma_{s}\!\sum_{\alpha,n}\!\sum_{r=0,3}\!\int\!\!d\bm{r}\operatorname{Tr}I_{n}^{\alpha}t_{r0}Q\operatorname{Tr}I_{-n}^{\alpha}t_{r0}Q, (2b)
Sint(σ)\displaystyle S_{\rm int}^{(\sigma)} =πT4Γtα,nr=0,3𝑑𝒓TrInα𝒕𝒓QTrInα𝒕𝒓Q,\displaystyle=-\frac{\pi T}{4}\Gamma_{t}\!\sum_{\alpha,n}\!\sum_{r=0,3}\int\!\!d\bm{r}\operatorname{Tr}I_{n}^{\alpha}\bm{t_{r}}Q\operatorname{Tr}I_{-n}^{\alpha}\bm{t_{r}}Q, (2c)
Sint(c)\displaystyle S_{\rm int}^{(c)} =πT4Γcα,nr=1,2𝑑𝒓Trtr0LnαQTrtr0LnαQ.\displaystyle=-\frac{\pi T}{4}\Gamma_{c}\!\sum_{\alpha,n}\!\sum_{r=1,2}\int\!\!d\bm{r}\operatorname{Tr}t_{r0}L_{n}^{\alpha}Q\operatorname{Tr}t_{r0}L_{n}^{\alpha}Q. (2d)

Here the matrix field Q(𝒓)Q(\bm{r}) (as well as the trace Tr\operatorname{Tr}) acts in the replica, Matsubara, spin, and particle-hole spaces. The matrix field obeys the following constraints:

Q2=1,TrQ=0,Q=CTQTC,\displaystyle Q^{2}=1,\quad\operatorname{Tr}Q=0,\qquad Q^{\dagger}=C^{T}Q^{T}C, (3)

where the charge-conjugation is realized by the matrix C=it12C=it_{12}. The action of the NLσ\sigmaM involves four constant matrices:

Λnmαβ\displaystyle\Lambda_{nm}^{\alpha\beta} =sgnnδnmδαβt00,(Ikγ)nmαβ=δnm,kδαβδαγt00,\displaystyle=\operatorname{sgn}n\,\delta_{nm}\delta^{\alpha\beta}t_{00},\,(I_{k}^{\gamma})_{nm}^{\alpha\beta}=\delta_{n-m,k}\delta^{\alpha\beta}\delta^{\alpha\gamma}t_{00},
(Lkγ)nmαβ\displaystyle(L_{k}^{\gamma})_{nm}^{\alpha\beta} =δn+m,kδαβδαγt00,ηnmαβ=nδnmδαβt00,\displaystyle=\delta_{n+m,k}\delta^{\alpha\beta}\delta^{\alpha\gamma}t_{00},\quad\eta_{nm}^{\alpha\beta}=n\,\delta_{nm}\delta^{\alpha\beta}t_{00}, (4)

where α,β=1,,Nr\alpha,\beta=1,\dots,N_{r} stand for replica indices and integers n,mn,m correspond to the Matsubara fermionic frequencies εn=πT(2n+1)\varepsilon_{n}=\pi T(2n+1). The sixteen matrices,

trj=τrsj,r,j=0,1,2,3,t_{rj}=\tau_{r}\otimes s_{j},\qquad r,j=0,1,2,3, (5)

operate in the particle-hole (subscript rr) and spin (subscript jj) spaces. The matrices τ0,τ1,τ2,τ3\tau_{0},\tau_{1},\tau_{2},\tau_{3} and s0,s1,s2,s3s_{0},s_{1},s_{2},s_{3} are the standard sets of the Pauli matrices. Also we introduced the vector 𝒕𝒓={tr1,tr2,tr3}\bm{t_{r}}=\{t_{r1},t_{r2},t_{r3}\} for convenience.

The bare value of the total conductivity (in units e2/he^{2}/h and including spin) is denoted as gg. The interaction amplitude Γs\Gamma_{s} (Γt\Gamma_{t}) encodes interaction in the singlet (triplet) particle-hole channel. The interaction in the Cooper channel is expressed by Γc\Gamma_{c}. Its negative magnitude, Γc<0\Gamma_{c}<0, corresponds to an attraction in the particle-particle channel. The parameter ZωZ_{\omega} describes the frequency renormalization. If Coulomb interaction is present the following relation holds, Γs=Zω\Gamma_{s}=-Z_{\omega}. This condition remains intact under action of the renormalization group flow [25, 28].

2.2 Kubo formula for the ac conductivity

Within the Finkel’stein NLσ\sigmaM approach, the physical observables, associated with the mean-field parameters of the action (1), can be written as correlation functions of the matrix field QQ. The ac conductivity σ(ω)\sigma(\omega) can be obtained after the analytic continuation to the real frequencies, iωnω+i0+i\omega_{n}\to\omega+i0^{+}, of the following Matsubara response function (ωn=2πTn\omega_{n}=2\pi Tn):

σ(iωn)=\displaystyle\sigma(i\omega_{n})= g16nTr[Jnα,Q(𝒓)][Jnα,Q(𝒓)]+g264dn𝑑𝒓TrJnαQ(𝒓)Q(𝒓)TrJnαQ(𝒓)Q(𝒓).\displaystyle-\frac{g}{16n}\Bigl{\langle}\operatorname{Tr}[J_{n}^{\alpha},Q(\bm{r})][J_{-n}^{\alpha},Q(\bm{r})]\Bigr{\rangle}+\frac{g^{2}}{64dn}\int d\bm{r}^{\prime}\Bigl{\langle}\operatorname{Tr}J_{n}^{\alpha}Q(\bm{r})\nabla Q(\bm{r})\operatorname{Tr}J_{-n}^{\alpha}Q(\bm{r}^{\prime})\nabla Q(\bm{r}^{\prime})\Bigr{\rangle}. (6)

Here the expectation values \langle\dots\rangle are taken with respect to the action (1), dd stands for the spatial dimensionality, and the matrix JnαJ_{n}^{\alpha} is defined as follows

Jnα=t30t002Inα+t30+t002Inα.J_{n}^{\alpha}=\frac{t_{30}-t_{00}}{2}I_{n}^{\alpha}+\frac{t_{30}+t_{00}}{2}I_{-n}^{\alpha}. (7)

At the classical level, Q=ΛQ=\Lambda, the conductivity is independent of the frequency, σ(ω)=g\sigma(\omega)=g.

3 One-loop corrections to the ac conductivity

3.1 Perturbative expansion

Our aim is to compute correction to σ(ω)\sigma(\omega) in the lowest order in 1/g1/g. For this purpose we shall use the square-root parametrization of the matrix field QQ:

Q=W+Λ1W2,W=(0ww¯0).\displaystyle Q=W+\Lambda\sqrt{1-W^{2}},\qquad W=\begin{pmatrix}0&w\\ \overline{w}&0\end{pmatrix}. (8)

We adopt the following notations: Wn1n2=wn1n2W_{n_{1}n_{2}}=w_{n_{1}n_{2}} and Wn2n1=w¯n2n1W_{n_{2}n_{1}}=\overline{w}_{n_{2}n_{1}} where n10n_{1}\geqslant 0 and n2<0n_{2}<0. The blocks ww and w¯\overline{w} satisfy the charge-conjugation constraints:

w¯=CwTC,w=CwC.\displaystyle\overline{w}=-Cw^{T}C,\qquad w=-Cw^{*}C. (9)

These constraints imply that some elements (wn1n2αβ)rj(w^{\alpha\beta}_{n_{1}n_{2}})_{rj} in the expansion, wn1n2αβ=rj(wn1n2αβ)rjtrjw^{\alpha\beta}_{n_{1}n_{2}}=\sum_{rj}(w^{\alpha\beta}_{n_{1}n_{2}})_{rj}t_{rj}, are purely real and the others are purely imaginary.

The part of the action (1), which is quadratic in WW, determines the following propagators for diffusive modes in the theory. The propagators of diffusons (modes with r=0,3r=0,3 and j=0,1,2,3j=0,1,2,3) read

[wrj(𝒑)]n1n2α1β1[w¯rj(𝒑)]n4n3β2α2=2gδα1α2δβ1β2δn12,n34𝒟p(iΩ12ε)[δn1n332πTΓjgδα1β1𝒟p(j)(iΩ12ε)],\displaystyle\Bigl{\langle}[w_{rj}(\bm{p})]^{\alpha_{1}\beta_{1}}_{n_{1}n_{2}}[\bar{w}_{rj}(-\bm{p})]^{\beta_{2}\alpha_{2}}_{n_{4}n_{3}}\Bigr{\rangle}=\frac{2}{g}\delta^{\alpha_{1}\alpha_{2}}\delta^{\beta_{1}\beta_{2}}\delta_{n_{12},n_{34}}\mathcal{D}_{p}(i\Omega_{12}^{\varepsilon})\Bigl{[}\delta_{n_{1}n_{3}}-\frac{32\pi T\Gamma_{j}}{g}\delta^{\alpha_{1}\beta_{1}}\mathcal{D}_{p}^{(j)}(i\Omega_{12}^{\varepsilon})\Bigr{]}, (10)

where Ω12ε=εn1εn2=2πTn12=2πT(n1n2)\Omega_{12}^{\varepsilon}=\varepsilon_{n_{1}}-\varepsilon_{n_{2}}=2\pi Tn_{12}=2\pi T(n_{1}-n_{2}), Γ0Γs\Gamma_{0}\equiv\Gamma_{s}, and Γ1=Γ2=Γ3Γt\Gamma_{1}=\Gamma_{2}=\Gamma_{3}\equiv\Gamma_{t}. The diffuson in the absence of interaction is given as

𝒟p1(iωn)=p2+16Zω|ωn|/g.\mathcal{D}^{-1}_{p}(i\omega_{n})=p^{2}+{16Z_{\omega}|\omega_{n}|}/{g}. (11)

The diffusons renormalized by a ladder resummation of interaction in the singlet and triplet particle-hole channels have the following form, respectively,

𝒟p(0)(iωn)𝒟ps(iωn)\displaystyle\mathcal{D}^{(0)}_{p}(i\omega_{n})\equiv\mathcal{D}^{s}_{p}(i\omega_{n}) =[p2+16(Zω+Γs)|ωn|/g]1,\displaystyle=\Bigl{[}p^{2}+{16(Z_{\omega}+\Gamma_{s})|\omega_{n}|}/{g}\Bigr{]}^{-1},
𝒟p(1,2,3)(iωn)𝒟pt(iωn)\displaystyle\mathcal{D}^{(1,2,3)}_{p}(i\omega_{n})\equiv\mathcal{D}^{t}_{p}(i\omega_{n}) =[p2+16(Zω+Γt)|ωn|/g]1.\displaystyle=\Bigl{[}p^{2}+{16(Z_{\omega}+\Gamma_{t})|\omega_{n}|}/{g}\Bigr{]}^{-1}. (12)

The propagators of singlet cooperons (modes with r=1,2r=1,2 and j=0j=0) can be written as

[wr0(𝒑)]n1n2α1β1[w¯r0(𝒑)]n4n3β2α2=2gδα1α2δβ1β2δn14,n32𝒞p(iΩ12ε)[δn1n34πTDδα1β1𝒞p(iΩ34ε)p(i12)],\displaystyle\Bigl{\langle}[w_{r0}(\bm{p})]^{\alpha_{1}\beta_{1}}_{n_{1}n_{2}}[\bar{w}_{r0}(-\bm{p})]^{\beta_{2}\alpha_{2}}_{n_{4}n_{3}}\Bigr{\rangle}=\frac{2}{g}\delta^{\alpha_{1}\alpha_{2}}\delta^{\beta_{1}\beta_{2}}\delta_{n_{14},n_{32}}\mathcal{C}_{p}(i\Omega_{12}^{\varepsilon})\Bigl{[}\delta_{n_{1}n_{3}}-\frac{4\pi T}{D}\delta^{\alpha_{1}\beta_{1}}\mathcal{C}_{p}(i\Omega_{34}^{\varepsilon})\mathcal{L}_{p}(i\mathcal{E}_{12})\Bigr{]}, (13)

where 12=εn1+εn2\mathcal{E}_{12}=\varepsilon_{n_{1}}+\varepsilon_{n_{2}}, 𝒞p(iωn)𝒟p(iωn)\mathcal{C}_{p}(i\omega_{n})\equiv\mathcal{D}_{p}(i\omega_{n}). The diffusion coefficient is D=g/(16Zω)D=g/(16Z_{\omega}). The fluctuation propagator has the standard form,

p1(iωn)=γc1ln(2πTτ)ψ(𝒳p,i|ωn|)+ψ(1/2),\displaystyle\mathcal{L}^{-1}_{p}(i\omega_{n})=\gamma_{c}^{-1}-\ln(2\pi T\tau)-\psi\left(\mathcal{X}_{p,i|\omega_{n}|}\right)+\psi\left({1}/{2}\right), (14)

where γc=Γc/Zω\gamma_{c}=\Gamma_{c}/Z_{\omega} and ψ(z)\psi(z) denotes the di-gamma function. Also we introduced the following notation

𝒳q,ω=Dq2iω4πT+12.\mathcal{X}_{q,\omega}=\frac{Dq^{2}-i\omega}{4\pi T}+\frac{1}{2}. (15)

The triplet cooperons (modes with r=1,2r=1,2 and j=1,2,3j=1,2,3) are insensitive to the Cooper-channel interaction and coincide with the non-interacting cooperons:

[wrj(𝒑)]n1n2α1β1[w¯rj(𝒑)]n4n3β2α2\displaystyle\Bigl{\langle}[w_{rj}(\bm{p})]^{\alpha_{1}\beta_{1}}_{n_{1}n_{2}}[\bar{w}_{rj}(-\bm{p})]^{\beta_{2}\alpha_{2}}_{n_{4}n_{3}}\Bigr{\rangle} =2gδα1α2δβ1β2δn1n3δn2n4𝒞p(iΩ12ε).\displaystyle=\frac{2}{g}\delta^{\alpha_{1}\alpha_{2}}\delta^{\beta_{1}\beta_{2}}\delta_{n_{1}n_{3}}\delta_{n_{2}n_{4}}\mathcal{C}_{p}(i\Omega_{12}^{\varepsilon}). (16)

3.2 One-loop renormalization

Expanding the matrix QQ up to the second order in WW we obtain the following expression from Eq. (6),

σ(iωn)=gg64nTr[Jnα,ΛW2(𝒓)][Jnα,ΛW2(𝒓)]+g264dnd𝒓TrJnαW(𝒓)W(𝒓)\displaystyle\sigma(i\omega_{n})=g-\frac{g}{64n}\Bigl{\langle}\operatorname{Tr}[J_{n}^{\alpha},\Lambda W^{2}(\bm{r})][J_{-n}^{\alpha},\Lambda W^{2}(\bm{r})]\Bigr{\rangle}+\frac{g^{2}}{64dn}\int d\bm{r}^{\prime}\Bigl{\langle}\operatorname{Tr}J_{n}^{\alpha}W(\bm{r})\nabla W(\bm{r})
×TrJnαW(𝒓)W(𝒓).\displaystyle\times\operatorname{Tr}J_{-n}^{\alpha}W(\bm{r}^{\prime})\nabla W(\bm{r}^{\prime})\Bigr{\rangle}. (17)

In order to derive the correction to σ(iωn)\sigma(i\omega_{n}) in the lowest order in 1/g1/g, it is enough to average the correlation functions in Eq. (17) with the Gaussian part of the NLσ\sigmaM action. Using Wick theorem and computing the averages with the help of Eqs. (10) - (16), we find lengthy expression

σ(iωn)=\displaystyle\sigma(i\omega_{n})= g4q𝒞q(iωn)16π2T2ωnDωn>εn1,εn2>0q𝒞q(iΩ12ε)𝒞q(2iωniΩ12ε)q(i12)\displaystyle g-4\int_{q}\mathcal{C}_{q}(i\omega_{n})-\frac{16\pi^{2}T^{2}}{\omega_{n}D}\sum_{\omega_{n}>\varepsilon_{n_{1}},-\varepsilon_{n_{2}}>0}\int_{q}\mathcal{C}_{q}(i\Omega_{12}^{\varepsilon})\mathcal{C}_{q}(2i\omega_{n}-i\Omega_{12}^{\varepsilon})\mathcal{L}_{q}(i\mathcal{E}_{12}) (18a)
+\displaystyle+ 256πTωngdj=03Γjωm>0qq2min{ωm,ωn}𝒟q(iωm)𝒟q(j)(iωm)𝒟q(iωm+iωn)\displaystyle\frac{256\pi T}{\omega_{n}gd}\sum_{j=0}^{3}\Gamma_{j}\sum_{\omega_{m}>0}\int_{q}\,q^{2}\min\{\omega_{m},\omega_{n}\}\mathcal{D}_{q}(i\omega_{m})\mathcal{D}^{(j)}_{q}(i\omega_{m})\mathcal{D}_{q}(i\omega_{m}+i\omega_{n}) (18b)
\displaystyle- 16π2T2ωnDεn1,εn2>0σ,σ=±q𝒞q(iΩ12ε+iωnζσσ2)𝒞q(iΩ12ε+iωn(2ζσσ2))\displaystyle\frac{16\pi^{2}T^{2}}{\omega_{n}D}\sum_{\varepsilon_{n_{1}},-\varepsilon_{n_{2}}>0}\sum_{\sigma,\sigma^{\prime}=\pm}\int_{q}\mathcal{C}_{q}\left(i\Omega_{12}^{\varepsilon}+i\omega_{n}\zeta^{2}_{\sigma\sigma^{\prime}}\right)\mathcal{C}_{q}\left(i\Omega_{12}^{\varepsilon}+i\omega_{n}(2-\zeta^{2}_{\sigma\sigma^{\prime}})\right)
×q(i12+iωnζσσ)\displaystyle\,{}\hskip 99.58464pt\times\mathcal{L}_{q}\left(i\mathcal{E}_{12}+i\omega_{n}\zeta_{\sigma\sigma^{\prime}}\right) (18c)
+\displaystyle+ 32π2T2dωnDεn1,εn2>0σ,σ=±qq2𝒞q(iΩ12ε)𝒞q(iΩ12ε+iωn)𝒞q(iΩ12ε+iωn(2ζσσ2))\displaystyle\frac{32\pi^{2}T^{2}}{d\omega_{n}D}\sum_{\varepsilon_{n_{1}},-\varepsilon_{n_{2}}>0}\sum_{\sigma,\sigma^{\prime}=\pm}\int_{q}q^{2}\mathcal{C}_{q}\left(i\Omega_{12}^{\varepsilon}\right)\mathcal{C}_{q}\left(i\Omega_{12}^{\varepsilon}+i\omega_{n}\right)\mathcal{C}_{q}\left(i\Omega_{12}^{\varepsilon}+i\omega_{n}(2-\zeta^{2}_{\sigma\sigma^{\prime}})\right)
×q(i12+iωnζσσ)\displaystyle\,{}\hskip 99.58464pt\times\mathcal{L}_{q}\left(i\mathcal{E}_{12}+i\omega_{n}\zeta_{\sigma\sigma^{\prime}}\right) (18d)
+\displaystyle+ 32π2T2dωnDεn1,εn2>0σ,σ=±qq2𝒞q(iΩ12ε)𝒞q(iΩ12ε+iωn)𝒞q(iΩ12ε+iωn(1+σ))\displaystyle\frac{32\pi^{2}T^{2}}{d\omega_{n}D}\sum_{\varepsilon_{n_{1}},-\varepsilon_{n_{2}}>0}\sum_{\sigma,\sigma^{\prime}=\pm}\int_{q}q^{2}\mathcal{C}_{q}\left(i\Omega_{12}^{\varepsilon}\right)\mathcal{C}_{q}\left(i\Omega_{12}^{\varepsilon}+i\omega_{n}\right)\mathcal{C}_{q}\left(i\Omega_{12}^{\varepsilon}+i\omega_{n}(1+\sigma)\right)
×q(i12)\displaystyle\,{}\hskip 99.58464pt\times\mathcal{L}_{q}\left(i\mathcal{E}_{12}\right) (18e)
\displaystyle- 128π3T3zdωnD2εn1,3,εn2,4>0σ,σ=±qq2𝒞q(iΩ12ε)𝒞q(iΩ12ε+iωn)𝒞q(iΩ34ε)𝒞q(iΩ34ε+iωn)\displaystyle\frac{128\pi^{3}T^{3}z}{d\omega_{n}D^{2}}\sum_{\varepsilon_{n_{1,3}},-\varepsilon_{n_{2,4}}>0}\sum_{\sigma,\sigma^{\prime}=\pm}\int_{q}q^{2}\mathcal{C}_{q}\left(i\Omega_{12}^{\varepsilon}\right)\mathcal{C}_{q}\left(i\Omega_{12}^{\varepsilon}+i\omega_{n}\right)\mathcal{C}_{q}\left(i\Omega_{34}^{\varepsilon}\right)\mathcal{C}_{q}\left(i\Omega_{34}^{\varepsilon}+i\omega_{n}\right)
×δ12,34+iωnμσσq(i12)q(i34+iωnμσσ+).\displaystyle\,{}\hskip 99.58464pt\times\delta_{\mathcal{E}_{12},\mathcal{E}_{34}+i\omega_{n}\mu^{-}_{\sigma\sigma^{\prime}}}\mathcal{L}_{q}\left(i\mathcal{E}_{12}\right)\mathcal{L}_{q}\left(i\mathcal{E}_{34}+i\omega_{n}\mu^{+}_{\sigma\sigma^{\prime}}\right). (18f)

Here we use the following short-hand notations, ζσσ=(σ+σ)/2\zeta_{\sigma\sigma^{\prime}}=(\sigma+\sigma^{\prime})/2, μσσ±=σ(1±σ)/2\mu_{\sigma\sigma^{\prime}}^{\pm}=\sigma(1\pm\sigma^{\prime})/2, and qdd𝒒/(2π)d\int_{q}\equiv\int{d^{d}\bm{q}}/{(2\pi)^{d}}. We note that the contributions (18a) and (18c) come from the term in Eq. (17) which has no gradients acting on WW matrices. All the other contributions result from the last term in the right hand side of Eq. (17).

Traditionally, the conductivity is split into several parts: weak localization or interference contribution δgWL\delta g^{WL}, Altshuler–Aronov or interaction contribution δgAA\delta g^{AA}, and fluctuation conductivity which stems from the interaction in the Cooper channel, δgCC\delta g^{CC}, i.e.

σ(ω)=g+δgWL(ω)+δgAA(ω)+δgCC(ω).\sigma(\omega)=g+\delta g^{\rm WL}(\omega)+\delta g^{\rm AA}(\omega)+\delta g^{\rm CC}(\omega). (19)

The contribution due to the Cooper channel interaction involves the fluctuation propagator q\mathcal{L}_{q}. This contribution, δgCC\delta g^{\rm CC}, can be written as a sum of four terms [1]:

δgCC=δgMT,an+δg~MT,reg+δg~DOS+δg~AL.\delta g^{\rm CC}=\delta g^{\rm MT,an}+\delta\tilde{g}^{\rm MT,reg}+\delta\tilde{g}^{\rm DOS}+\delta\tilde{g}^{\rm AL}. (20)

In what follows we shall consider each of these terms separately.

3.2.1 Weak localization and Althsuler–Aronov corrections

The weak localization and Althsuler–Aronov contributions are given by the second term in the right hand side of Eq.(18a) and by Eq. (18b). At first, we perform analytic continuation to the real frequencies iωnω+i0+i\omega_{n}\to\omega+i0^{+}. Then the interference correction is expressed in terms of the non-interacting cooperon [17]:

δgWL(ω)=4q𝒞qR(ω).\delta g^{\rm WL}(\omega)=-4\int_{q}\mathcal{C}^{R}_{q}(\omega). (21)

Here 𝒞qR(ω)\mathcal{C}^{R}_{q}(\omega) stands for the retarded propagator corresponding to the Matsubara propagator 𝒞q(iωn)\mathcal{C}_{q}(i\omega_{n}).

The interaction correction reads [18, 29, 30, 31, 32]

δgAA(ω)=64iωgdj=03Γjq,Ωq2[ΩΩ(Ωω)Ωω]𝒟qR(Ω)𝒟q(j),R(Ω)𝒟qR(Ω+ω).\displaystyle\delta g^{\rm AA}(\omega)=\frac{64}{i\omega gd}\sum_{j=0}^{3}\Gamma_{j}\int_{q,\Omega}\,q^{2}\Bigl{[}\Omega\mathcal{B}_{\Omega}-(\Omega-\omega)\mathcal{B}_{\Omega-\omega}\Bigr{]}\mathcal{D}^{R}_{q}(\Omega)\mathcal{D}^{(j),R}_{q}(\Omega)\mathcal{D}^{R}_{q}(\Omega+\omega). (22)

Here Ω=coth[Ω/(2T)]\mathcal{B}_{\Omega}=\coth[\Omega/(2T)] denotes the bosonic distribution function for the particle-hole excitations. The retarded diffuson propagators are denoted as 𝒟qR(ω)\mathcal{D}^{R}_{q}(\omega), 𝒟q(j),R(ω)\mathcal{D}^{(j),R}_{q}(\omega). Also we introduced the short-hand notation Ω𝑑Ω\int_{\Omega}\equiv\int_{-\infty}^{\infty}d\Omega.

3.2.2 Anomalous Maki-Thompson correction

The anomalous Maki-Thompson correction [3, 4] is given by the last term in the right hand side of Eq. (18a). It is convenient to rewrite it as follows

δgMT,an(iωn)=4q𝒞q(iωn)βq(iωn),\delta g^{\rm MT,an}(i\omega_{n})=4\int_{q}\mathcal{C}_{q}(i\omega_{n})\beta_{q}(i\omega_{n}), (23)

where [33, 34]

βq(iωn)=πTωn|ωm|<ωnq(iωm)[ψ(𝒳q,i|ωm|)ψ(𝒳q,2iωni|ωm|)].\displaystyle\beta_{q}(i\omega_{n})=\frac{\pi T}{\omega_{n}}\sum_{|\omega_{m}|<\omega_{n}}\mathcal{L}_{q}(i\omega_{m})\Bigl{[}\psi(\mathcal{X}_{q,i|\omega_{m}|})-\psi(\mathcal{X}_{q,2i\omega_{n}-i|\omega_{m}|})\Bigr{]}. (24)

Performing analytic continuation to the real frequencies we obtain the final form of the anomalous Maki-Thompson correction

δgMT,an(ω)=4q𝒞qR(ω)βqR(ω),\delta g^{\rm MT,an}(\omega)=4\int_{q}\mathcal{C}_{q}^{R}(\omega)\beta_{q}^{R}(\omega), (25)

where

βqR(ω)=ΩqR(Ω)ΩΩω2ω[ψ(𝒳q,Ω)ψ(𝒳q,2ωΩ)].\displaystyle\beta_{q}^{R}(\omega)=\int_{\Omega}\mathcal{L}_{q}^{R}(\Omega)\frac{\mathcal{B}_{\Omega}-\mathcal{B}_{\Omega-\omega}}{2\omega}\Bigl{[}\psi(\mathcal{X}_{q,\Omega})-\psi(\mathcal{X}_{q,2\omega-\Omega})\Bigr{]}. (26)

We mention that the anomalous Maki-Thompson correction (25) coincides with the sum σMT1+σMT2\sigma_{MT1}+\sigma_{MT2} computed in Ref. [23] (see Eqs. (A1) and (A2) there).

It is instructive to compare the above result with the other expressions existing in the literature. For this purpose we use the following relations

ε(ε+ωε)𝒞qR(2ε+Ω)=iD[ψ(𝒳q,Ω)ψ(𝒳q,Ω2ω)]\int_{\varepsilon}\Bigl{(}\mathcal{F}_{\varepsilon+\omega}-\mathcal{F}_{\varepsilon}\Bigr{)}\ \mathcal{C}_{q}^{R}(2\varepsilon+\Omega)=iD\bigl{[}\psi(\mathcal{X}_{q,\Omega})-\psi(\mathcal{X}_{q,\Omega-2\omega})\bigr{]} (27)

and

εε+Ω(ε+ωε)𝒞qR(2ε+Ω)=iD{Ω[ψ(𝒳q,Ω)ψ(𝒳q,Ω)]\displaystyle\int_{\varepsilon}\mathcal{F}_{\varepsilon+\Omega}\Bigl{(}\mathcal{F}_{\varepsilon+\omega}-\mathcal{F}_{\varepsilon}\Bigr{)}\ \mathcal{C}_{q}^{R}(2\varepsilon+\Omega)=iD\Bigl{\{}\mathcal{B}_{\Omega}\bigl{[}\psi(\mathcal{X}_{q,\Omega})-\psi(\mathcal{X}_{q,-\Omega})\bigr{]}
Ωω[ψ(𝒳q,Ω2ω)ψ(𝒳q,Ω)]}.\displaystyle-\mathcal{B}_{\Omega-\omega}\bigl{[}\psi(\mathcal{X}_{q,\Omega-2\omega})-\psi(\mathcal{X}_{q,-\Omega})\bigr{]}\Bigr{\}}. (28)

Here ε=tanh[ε/(2T)]\mathcal{F}_{\varepsilon}=\tanh[\varepsilon/(2T)] stands for the fermionic distribution function. Then it is possible to rewrite Eq. (25) as follows

δgMT,an(ω)=2iDωq,Ω,ε𝒞qR(ω)qR(Ω)[ε+ωε][Ωε+Ω]𝒞qR(2ε+Ω).\displaystyle\delta g^{\rm MT,an}(\omega)=\frac{2i}{D\omega}\int_{q,\Omega,\varepsilon}\mathcal{C}_{q}^{R}(\omega)\mathcal{L}_{q}^{R}(\Omega)\Bigl{[}\mathcal{F}_{\varepsilon+\omega}-\mathcal{F}_{\varepsilon}\Bigr{]}\Bigl{[}\mathcal{B}_{\Omega}-\mathcal{F}_{\varepsilon+\Omega}\Bigr{]}\mathcal{C}^{R}_{q}(2\varepsilon+\Omega). (29)

In the dc limit, ω0\omega\to 0, the expression (29) is similar to Eq. (384) of Ref. [24].

3.2.3 Regular Maki-Thompson correction

The so-called regular part of the Maki-Thompson correction is determined by the contribution (18c). Performing summation over one of the fermionic energies, we obtain

δg~MT,reg(iωn)=Dωnqωm{2ψ(𝒳q,i|ωm+n|+iωn)+4πTωn[ψ(𝒳q,i|ωm|+2iωn)ψ(𝒳q,i|ωm|)]}q(iωm).\displaystyle\delta\tilde{g}^{\rm MT,reg}(i\omega_{n})=-\frac{D}{\omega_{n}}\int_{q}\sum_{\omega_{m}}\Biggl{\{}2\psi^{\prime}\left(\mathcal{X}_{q,i|\omega_{m+n}|+i\omega_{n}}\right)+\frac{4\pi T}{\omega_{n}}\Bigl{[}\psi\left(\mathcal{X}_{q,i|\omega_{m}|+2i\omega_{n}}\right)-\psi\left(\mathcal{X}_{q,i|\omega_{m}|}\right)\Bigr{]}\Biggr{\}}\mathcal{L}_{q}(i\omega_{m}). (30)

After the analytic continuation to the real frequency, iωnω+i0i\omega_{n}\to\omega+i0, we find

δg~MT,reg(ω)=2DπTωq,ΩΩqR(Ω)ψ(𝒳q,Ω)+D2πTωq,ΩqR(Ω){2ΩΦ2ω(Ω)\displaystyle\delta\tilde{g}^{\rm MT,reg}(\omega)=-\frac{2D}{\pi T\omega}\int_{q,\Omega}\mathcal{B}_{\Omega}\mathcal{L}_{q}^{R}(\Omega)\psi^{\prime}(\mathcal{X}_{q,\Omega})+\frac{D}{2\pi T\omega}\int_{q,\Omega}\mathcal{L}_{q}^{R}(\Omega)\Biggl{\{}2\mathcal{B}_{\Omega}\Phi_{-2\omega}(\Omega)
+Ω[ψ(𝒳q,Ω)ψ(𝒳q,Ω+2ω)]+[ΩΩω][ψ(𝒳q,Ω)ψ(𝒳q,2ωΩ)]},\displaystyle+\mathcal{B}_{\Omega}\Bigl{[}\psi^{\prime}(\mathcal{X}_{q,\Omega})-\psi^{\prime}(\mathcal{X}_{q,\Omega+2\omega})\Bigr{]}+\Bigl{[}\mathcal{B}_{\Omega}-\mathcal{B}_{\Omega-\omega}\Bigr{]}\Bigl{[}\psi^{\prime}(\mathcal{X}_{q,\Omega})-\psi^{\prime}(\mathcal{X}_{q,2\omega-\Omega})\Bigr{]}\Biggr{\}}, (31)

where

Φω(Ω)=ψ(𝒳q,Ω)+4πTiω[ψ(𝒳q,Ω)ψ(𝒳q,Ωω)].\Phi_{\omega}(\Omega)=\psi^{\prime}(\mathcal{X}_{q,\Omega})+\frac{4\pi T}{i\omega}\Bigl{[}\psi(\mathcal{X}_{q,\Omega})-\psi(\mathcal{X}_{q,\Omega-\omega})\Bigr{]}. (32)

We note that in the course of derivation of Eq. (31) we have also used the following symmetry properties: qA(Ω)=qR(Ω)\mathcal{L}_{q}^{A}(\Omega)=\mathcal{L}_{q}^{R}(-\Omega), and Ω=Ω\mathcal{B}_{-\Omega}=-\mathcal{B}_{\Omega}.

It is useful to relate the regular Maki-Thompson correction with the correction to the tunneling density of states due to interaction in the Cooper channel [35, 36]. The correction to the density of states can be written as [24]

δρCC(ε)=ρ0ReΥ(ε),\delta\rho^{\rm CC}(\varepsilon)=\rho_{0}\operatorname{Re}\Upsilon(\varepsilon), (33)

where

Υ(ε)=32Zωig2q,Ω𝒞qR2(2εΩ)[qK(Ω)+εΩqR(Ω)].\Upsilon(\varepsilon)=\frac{32Z_{\omega}}{ig^{2}}\int_{q,\Omega}\mathcal{C}^{R2}_{q}(2\varepsilon-\Omega)\Bigl{[}\mathcal{L}^{K}_{q}(\Omega)+\mathcal{F}_{\varepsilon-\Omega}\mathcal{L}^{R}_{q}(\Omega)\Bigr{]}. (34)

Here qK(Ω)=2iΩImqR(Ω)\mathcal{L}^{K}_{q}(\Omega)=2i\mathcal{B}_{\Omega}\operatorname{Im}\mathcal{L}^{R}_{q}(\Omega) stands for the Keldysh component of the fluctuation propagator.

We define the correction to the conductivity that is related with the correction to the density of states in the following way

δgDOS(ω)=gω𝑑ε[fF(εω)fF(ε)]Υ(ε),\delta g^{\rm DOS}(\omega)=\frac{g}{\omega}\int d\varepsilon\Bigl{[}f_{F}(\varepsilon-\omega)-f_{F}(\varepsilon)\Bigr{]}\Upsilon(\varepsilon), (35)

where fF(ε)=(1ε)/2f_{F}(\varepsilon)=(1-\mathcal{F}_{\varepsilon})/2 is the Fermi-Dirac distribution function. Then, using the identities (27) and (28), we obtain the following result

δgDOS(ω)=D4πTωq,ΩqR(Ω){[ΩΩω][ψ(𝒳q,Ω)ψ(𝒳q,Ω+2ω)]\displaystyle\delta g^{\rm DOS}(\omega)=\frac{D}{4\pi T\omega}\int_{q,\Omega}\mathcal{L}^{R}_{q}(\Omega)\Biggl{\{}\Bigl{[}\mathcal{B}_{\Omega}-\mathcal{B}_{\Omega-\omega}\Bigr{]}\Bigl{[}\psi^{\prime}\left(\mathcal{X}_{q,\Omega}\right)-\psi^{\prime}\left(\mathcal{X}_{q,-\Omega+2\omega}\right)\Bigr{]}
+Ω[ψ(𝒳q,Ω)ψ(𝒳q,Ω+2ω)]}.\displaystyle+\mathcal{B}_{\Omega}\Bigl{[}\psi^{\prime}\left(\mathcal{X}_{q,\Omega}\right)-\psi^{\prime}\left(\mathcal{X}_{q,\Omega+2\omega}\right)\Bigr{]}\Biggr{\}}. (36)

Next, using Eq. (36), we split the regular Maki-Thompson contribution into three parts

δg~MT,reg(ω)=2DπTωq,ΩΩqR(Ω)ψ(𝒳q,Ω)+δgDOS(ω)+δgsc,1(ω),\displaystyle\delta\tilde{g}^{\rm MT,reg}(\omega)=-\frac{2D}{\pi T\omega}\int_{q,\Omega}\mathcal{B}_{\Omega}\mathcal{L}_{q}^{R}(\Omega)\psi^{\prime}(\mathcal{X}_{q,\Omega})+\delta{g}^{\rm DOS}(\omega)+\delta{g}^{\rm sc,1}(\omega), (37)

where

δgsc,1(ω)=D4πTωq,ΩqR(Ω){4ΩΦ2ω(Ω)+Ω[ψ(𝒳q,Ω)ψ(𝒳q,Ω+2ω)]\displaystyle\delta{g}^{\rm sc,1}(\omega)=\frac{D}{4\pi T\omega}\int_{q,\Omega}\mathcal{L}_{q}^{R}(\Omega)\Biggl{\{}4\mathcal{B}_{\Omega}\Phi_{-2\omega}(\Omega)+\mathcal{B}_{\Omega}\Bigl{[}\psi^{\prime}(\mathcal{X}_{q,\Omega})-\psi^{\prime}(\mathcal{X}_{q,\Omega+2\omega})\Bigr{]}
+[ΩΩω][ψ(𝒳q,Ω)ψ(𝒳q,2ωΩ)]}.\displaystyle+\Bigl{[}\mathcal{B}_{\Omega}-\mathcal{B}_{\Omega-\omega}\Bigr{]}\Bigl{[}\psi^{\prime}(\mathcal{X}_{q,\Omega})-\psi^{\prime}(\mathcal{X}_{q,2\omega-\Omega})\Bigr{]}\Biggr{\}}. (38)

3.2.4 DOS-type correction

The so-called DOS-type correction [1] is given by contributions (18d) - (18e). It is convenient to rewrite them as follows

δg~DOS(iωn)=4D2ωn2dqq2ωmq(iωm){ψ(𝒳q,i|ωm|)ψ(𝒳q,i|ωm+n|+iωn)+4πTωn[ψ(𝒳q,i|ωm+n|+iωn)\displaystyle\delta\tilde{g}^{\rm DOS}(i\omega_{n})=\frac{4D^{2}}{\omega_{n}^{2}d}\int_{q}q^{2}\sum_{\omega_{m}}\mathcal{L}_{q}(i\omega_{m})\Biggl{\{}\psi^{\prime}\left(\mathcal{X}_{q,i|\omega_{m}|}\right)-\psi^{\prime}\left(\mathcal{X}_{q,i|\omega_{m+n}|+i\omega_{n}}\right)+\frac{4\pi T}{\omega_{n}}\Bigl{[}\psi\left(\mathcal{X}_{q,i|\omega_{m+n}|+i\omega_{n}}\right)
ψ(𝒳q,i|ωm+n|)+ψ(𝒳q,i|ωm|+iωn)ψ(𝒳q,i|ωm|+2iωn)]}.\displaystyle-\psi\left(\mathcal{X}_{q,i|\omega_{m+n}|}\right)+\psi\left(\mathcal{X}_{q,i|\omega_{m}|+i\omega_{n}}\right)-\psi\left(\mathcal{X}_{q,i|\omega_{m}|+2i\omega_{n}}\right)\Bigr{]}\Biggr{\}}. (39)

The analytic continuation of Eq. (39) to the real frequency, iωnω+i0i\omega_{n}\to\omega+i0, yields

δg~DOS(ω)=iD2πTω2dq,Ωq2qR(Ω){(ΩΩω)[Φω(Ω)Φω(2ωΩ)]+2Ω[ψ(𝒳q,Ω)\displaystyle\delta\tilde{g}^{\rm DOS}(\omega)=\frac{iD^{2}}{\pi T\omega^{2}d}\int_{q,\Omega}q^{2}\mathcal{L}_{q}^{R}(\Omega)\Biggl{\{}\bigl{(}\mathcal{B}_{\Omega}-\mathcal{B}_{\Omega-\omega}\bigr{)}\Bigl{[}\Phi_{\omega}(\Omega)-\Phi_{\omega}(2\omega-\Omega)\Bigr{]}+2\mathcal{B}_{\Omega}\Bigl{[}\psi^{\prime}(\mathcal{X}_{q,\Omega})
ψ(𝒳q,Ω+2ω)]+Ω[Φω(2ω+Ω)Φω(Ω)]}.\displaystyle-\psi^{\prime}(\mathcal{X}_{q,\Omega+2\omega})\Bigr{]}+\mathcal{B}_{\Omega}\Bigl{[}\Phi_{\omega}(2\omega+\Omega)-\Phi_{\omega}(\Omega)\Bigr{]}\Biggr{\}}. (40)

It is useful to single out explicitly the part that diverges in the limit ω0\omega\to 0. Then we obtain

δg~DOS(ω)=D2dπ2T2ωq,Ωq2ΩqR(Ω)ψ′′(𝒳q,Ω)+δgsc,2(ω),\displaystyle\delta\tilde{g}^{\rm DOS}(\omega)=-\frac{D^{2}}{d\pi^{2}T^{2}\omega}\int_{q,\Omega}q^{2}\mathcal{B}_{\Omega}\mathcal{L}_{q}^{R}(\Omega)\psi^{\prime\prime}(\mathcal{X}_{q,\Omega})+\delta{g}^{\rm sc,2}(\omega), (41)

where

δgsc,2(ω)=iD2πTω2dq,Ωq2qR(Ω){(ΩΩω)[Φω(Ω)Φω(2ωΩ)]+2Ω[ψ(𝒳q,Ω)\displaystyle\delta{g}^{\rm sc,2}(\omega)=\frac{iD^{2}}{\pi T\omega^{2}d}\int_{q,\Omega}q^{2}\mathcal{L}_{q}^{R}(\Omega)\Biggl{\{}\bigl{(}\mathcal{B}_{\Omega}-\mathcal{B}_{\Omega-\omega}\bigr{)}\Bigl{[}\Phi_{\omega}(\Omega)-\Phi_{\omega}(2\omega-\Omega)\Bigr{]}+2\mathcal{B}_{\Omega}\Bigl{[}\psi^{\prime}(\mathcal{X}_{q,\Omega})
ψ(𝒳q,Ω+2ω)iω2πTψ′′(𝒳q,Ω)]+Ω[Φω(2ω+Ω)Φω(Ω)]}.\displaystyle-\psi^{\prime}(\mathcal{X}_{q,\Omega+2\omega})-\frac{i\omega}{2\pi T}\psi^{\prime\prime}(\mathcal{X}_{q,\Omega})\Bigr{]}+\mathcal{B}_{\Omega}\Bigl{[}\Phi_{\omega}(2\omega+\Omega)-\Phi_{\omega}(\Omega)\Bigr{]}\Biggr{\}}. (42)

3.2.5 Aslamazov-Larkin correction

The contribution (18f) is the correction due to Aslamazov–Larkin process [6]. It can be written as follows

δg~AL(iωn)\displaystyle\delta\tilde{g}^{\rm AL}(i\omega_{n}) =8πTdωn(D4πT)2qq2ωmq(iωm)q(iωm+n)Δq2(iωm,iωm+n,iωn),\displaystyle=-\frac{8\pi T}{d\omega_{n}}\left(\frac{D}{4\pi T}\right)^{2}\int_{q}q^{2}\sum_{\omega_{m}}\mathcal{L}_{q}(i\omega_{m})\mathcal{L}_{q}(i\omega_{m+n})\Delta_{q}^{2}(i\omega_{m},i\omega_{m+n},i\omega_{n}), (43)

where

Δq(iωm,iωk,iωn)=4πTωn[ψ(𝒳q,i|ωm|)+ψ(𝒳q,i|ωk|)ψ(𝒳q,i|ωm|+iωn)ψ(𝒳q,i|ωk|+iωn)].\displaystyle\Delta_{q}(i\omega_{m},i\omega_{k},i\omega_{n})=-\frac{4\pi T}{\omega_{n}}\Biggl{[}\psi\left(\mathcal{X}_{q,i|\omega_{m}|}\right)+\psi\left(\mathcal{X}_{q,i|\omega_{k}|}\right)-\psi\left(\mathcal{X}_{q,i|\omega_{m}|+i\omega_{n}}\right)-\psi\left(\mathcal{X}_{q,i|\omega_{k}|+i\omega_{n}}\right)\Biggr{]}. (44)

After the analytic continuation to the real frequency, iωnω+i0i\omega_{n}\to\omega+i0, we find

δg~AL(ω)=D28dπ2T2ωq,Ωq2qR(Ω){(Ω+Ωω)qR(Ωω)(ΔqRRR(Ωω,Ω,ω))2\displaystyle\delta\tilde{g}^{\rm AL}(\omega)=-\frac{D^{2}}{8d\pi^{2}T^{2}\omega}\int_{q,\Omega}q^{2}\mathcal{L}_{q}^{R}(\Omega)\Biggl{\{}\Bigl{(}\mathcal{B}_{\Omega}+\mathcal{B}_{\Omega-\omega}\Bigr{)}\mathcal{L}_{q}^{R}(\Omega-\omega)\bigl{(}\Delta^{RRR}_{q}(\Omega-\omega,\Omega,\omega)\bigr{)}^{2}
+(ΩωΩ)[qR(Ωω)(ΔqRRR(Ωω,Ω,ω))2qA(Ωω)(ΔqARR(Ωω,Ω,ω))2]}.\displaystyle+\Bigl{(}\mathcal{B}_{\Omega-\omega}-\mathcal{B}_{\Omega}\Bigr{)}\Bigl{[}\mathcal{L}_{q}^{R}(\Omega-\omega)\bigl{(}\Delta^{RRR}_{q}(\Omega-\omega,\Omega,\omega)\bigr{)}^{2}-\mathcal{L}_{q}^{A}(\Omega-\omega)\bigl{(}\Delta^{ARR}_{q}(\Omega-\omega,\Omega,\omega)\bigr{)}^{2}\Bigr{]}\Biggr{\}}. (45)

Here we introduced the function

ΔqRRR(Ω,Ω,ω)\displaystyle\Delta^{RRR}_{q}(\Omega,\Omega^{\prime},\omega) =4πTiω[ψ(𝒳q,Ω)ψ(𝒳q,Ω+ω)+ψ(𝒳q,Ω)ψ(𝒳q,Ω+ω)].\displaystyle=\frac{4\pi T}{i\omega}\Biggl{[}\psi\left(\mathcal{X}_{q,\Omega}\right)-\psi\left(\mathcal{X}_{q,\Omega+\omega}\right)+\psi\left(\mathcal{X}_{q,\Omega^{\prime}}\right)-\psi\left(\mathcal{X}_{q,\Omega^{\prime}+\omega}\right)\Biggr{]}. (46)

The function ΔqARR(Ω,Ω,ω)\Delta^{ARR}_{q}(\Omega,\Omega^{\prime},\omega) can be obtained from ΔqRRR(Ω,Ω,ω)\Delta^{RRR}_{q}(\Omega,\Omega^{\prime},\omega) according to the following prescription, ΔqARR(Ω,Ω,ω)=ΔqRRR(Ω,Ω,ω)\Delta^{ARR}_{q}(\Omega,\Omega^{\prime},\omega)=\Delta^{RRR}_{q}(-\Omega,\Omega^{\prime},\omega). We note that Eq. (45) coincides with the general result for the Aslamazov–Larkin contribution computed by the diagrammatic technique (see Eq. (7.105) in Ref. [1]). It is convenient to rewrite the correction (45) in the following way

δg~AL(ω)=D2dπ2T2ωq,Ωq2Ω[qR(Ω)ψ(Xq,Ω)]2+δgAL(ω)+δgsc,3(ω).\displaystyle\delta\tilde{g}^{\rm AL}(\omega)=-\frac{D^{2}}{d\pi^{2}T^{2}\omega}\int_{q,\Omega}q^{2}\mathcal{B}_{\Omega}\Bigl{[}\mathcal{L}_{q}^{R}(\Omega)\psi^{\prime}\left(X_{q,\Omega}\right)\Bigr{]}^{2}+\delta{g}^{\rm AL}(\omega)+\delta{g}^{\rm sc,3}(\omega). (47)

Here we single out the term which diverges in the limit of zero frequency, ω0\omega\to 0. Next, we introduce

δgAL(ω)=4D2dω3q,Ωq2(ΩωΩ)qR(Ω)[ψ(𝒳q,Ωω)ψ(𝒳q,Ω+ω)]ImqR(Ωω)\displaystyle\delta{g}^{\rm AL}(\omega)=-\frac{4D^{2}}{d\omega^{3}}\int_{q,\Omega}q^{2}\Bigl{(}\mathcal{B}_{\Omega-\omega}-\mathcal{B}_{\Omega}\Bigr{)}\mathcal{L}_{q}^{R}(\Omega)\Bigl{[}\psi\bigl{(}\mathcal{X}_{q,\Omega-\omega}\bigr{)}-\psi\bigl{(}\mathcal{X}_{q,\Omega+\omega}\bigr{)}\Bigr{]}\operatorname{Im}\mathcal{L}_{q}^{R}(\Omega-\omega)
×Im[ψ(𝒳q,Ωω)ψ(𝒳q,Ω+ω)]\displaystyle\times\operatorname{Im}\Bigl{[}\psi\bigl{(}\mathcal{X}_{q,\Omega-\omega}\bigr{)}-\psi\bigl{(}\mathcal{X}_{q,\Omega+\omega}\bigr{)}\Bigr{]} (48)

and

δgsc,3(ω)=D28dπ2T2ωq,Ωq2qR(Ω){(Ω+Ωω)qR(Ωω)(ΔqRRR(Ωω,Ω,ω))28Ω\displaystyle\delta{g}^{\rm sc,3}(\omega)=-\frac{D^{2}}{8d\pi^{2}T^{2}\omega}\int_{q,\Omega}q^{2}\mathcal{L}_{q}^{R}(\Omega)\Biggl{\{}\Bigl{(}\mathcal{B}_{\Omega}+\mathcal{B}_{\Omega-\omega}\Bigr{)}\mathcal{L}_{q}^{R}(\Omega-\omega)\bigl{(}\Delta^{RRR}_{q}(\Omega-\omega,\Omega,\omega)\bigr{)}^{2}-8\mathcal{B}_{\Omega}
×qR(Ω)ψ2(Xq,Ω)+(ΩωΩ)[qA(Ωω)[ΔqRRR(Ωω,Ω,ω)ReΔqRRR(Ωω,Ω,ω)\displaystyle\times\mathcal{L}_{q}^{R}(\Omega)\psi^{\prime 2}\left(X_{q,\Omega}\right)+\Bigl{(}\mathcal{B}_{\Omega-\omega}-\mathcal{B}_{\Omega}\Bigr{)}\Biggl{[}\mathcal{L}_{q}^{A}(\Omega-\omega)\Bigl{[}\Delta^{RRR}_{q}(\Omega-\omega,\Omega,\omega)\operatorname{Re}\Delta^{RRR}_{q}(\Omega-\omega,\Omega,\omega)
(ΔqARR(Ωω,Ω,ω))2]+iqR(Ωω)ΔqRRR(Ωω,Ω,ω)ImΔqRRR(Ωω,Ω,ω)]}.\displaystyle-\bigl{(}\Delta^{ARR}_{q}(\Omega-\omega,\Omega,\omega)\bigr{)}^{2}\Bigr{]}+i\mathcal{L}_{q}^{R}(\Omega-\omega)\Delta^{RRR}_{q}(\Omega-\omega,\Omega,\omega)\operatorname{Im}\Delta^{RRR}_{q}(\Omega-\omega,\Omega,\omega)\Biggr{]}\Biggr{\}}. (49)

3.3 Final result

Naturally, one expects that the dc conductivity in the normal state of a disordered electron system is finite. We note that separate contributions to δgCC(ω)\delta g^{CC}(\omega) do not satisfy this requirement. In particular, there are terms in Eqs. (37), (41), and (48) which diverge as 1/(iω)1/(i\omega) in the limit ω0\omega\to 0. They can be summed up as follows:

2DπTωq,ΩΩqR(Ω){ψ(𝒳q,Ω)+Dq22dπTψ′′(𝒳q,Ω)+Dq22dπTqR(Ω)[ψ(Xq,Ω)]2}\displaystyle-\frac{2D}{\pi T\omega}\int_{q,\Omega}\mathcal{B}_{\Omega}\mathcal{L}_{q}^{R}(\Omega)\Bigl{\{}\psi^{\prime}(\mathcal{X}_{q,\Omega})+\frac{Dq^{2}}{2d\pi T}\psi^{\prime\prime}(\mathcal{X}_{q,\Omega})+\frac{Dq^{2}}{2d\pi T}\mathcal{L}_{q}^{R}(\Omega)\bigl{[}\psi^{\prime}\left(X_{q,\Omega}\right)\bigr{]}^{2}\Bigr{\}}
=4iωdImq,ΩΩqμqμlnqR(Ω).\displaystyle=\frac{4}{i\omega d}\operatorname{Im}\int_{q,\Omega}\mathcal{B}_{\Omega}\partial_{q_{\mu}}\partial_{q_{\mu}}\ln\mathcal{L}^{R}_{q}(\Omega). (50)

Thus the sum of all terms in δgCC(ω)\delta g^{CC}(\omega) which are proportional to 1/(iω)1/(i\omega) has the form of the total second derivative with respect to the momentum. This implies that the contribution (50) is determined by the ultraviolet and, consequently, cannot be accurately computed within NLσ\sigmaM that is the low-energy effective theory only. However, as one can check [22], the contribution from the ballistic scales has exactly the same form (of course with the ballistic fluctuation propagator) such that the 1/(iω)1/(i\omega) term (50) vanishes identically. This fact is intimately related with the gauge invariance (see Ref. [37, 38] for detailed discussion). Indeed the expression (50) can be written as the second derivative of the contribution to the thermodynamic potential from superconducting fluctuations with respect to a constant vector potential. Since the thermodynamic potential is independent of the constant vector potential in virtue of the gauge invariance, the expression (50) should be zero. We note that the result for quantum correction to the ac conductivity due to interaction in the Cooper channel reported in Ref. [23] diverges as 1/(iω)1/(i\omega) in the limit ω0\omega\to 0.

Gathering together the contributions (25), (37), (41), and (47) (disregarding the terms which sum up to zero as discussed above), we find the following final form of the correction to the ac conductivity due to the interaction in the Cooper channel:

δgCC(ω)=δgMT,an(ω)+δgDOS(ω)+δgAL(ω)+δgsc(ω).\delta g^{CC}(\omega)=\delta g^{\rm MT,an}(\omega)+\delta g^{\rm DOS}(\omega)+\delta g^{\rm AL}(\omega)+\delta g^{\rm sc}(\omega). (51)

Here we introduce δgsc(ω)=δgsc,1(ω)+δgsc,2(ω)+δgsc,3(ω)\delta g^{\rm sc}(\omega)=\delta g^{\rm sc,1}(\omega)+\delta g^{\rm sc,2}(\omega)+\delta g^{\rm sc,3}(\omega) that can be rewritten as the following lengthy expression:

δgsc(ω)=D4πdTωq,Ωqμ{qμqR(Ω)[4ΩΦ2ω(Ω)+Ω[ψ(𝒳q,Ω)ψ(𝒳q,Ω+2ω)]+[ΩΩω]\displaystyle\delta{g}^{\rm sc}(\omega)=\frac{D}{4\pi dT\omega}\int_{q,\Omega}\partial_{q_{\mu}}\Biggl{\{}q_{\mu}\mathcal{L}_{q}^{R}(\Omega)\Biggl{[}4\mathcal{B}_{\Omega}\Phi_{-2\omega}(\Omega)+\mathcal{B}_{\Omega}\Bigl{[}\psi^{\prime}(\mathcal{X}_{q,\Omega})-\psi^{\prime}(\mathcal{X}_{q,\Omega+2\omega})\Bigr{]}+\Bigl{[}\mathcal{B}_{\Omega}-\mathcal{B}_{\Omega-\omega}\Bigr{]}
×[ψ(𝒳q,Ω)ψ(𝒳q,2ωΩ)]]}+D28d(πT)2ωq,Ωq2qR(Ω)Ω[3ψ′′(𝒳q,Ω)+ψ′′(𝒳q,Ω+2ω)\displaystyle\times\Bigl{[}\psi^{\prime}(\mathcal{X}_{q,\Omega})-\psi^{\prime}(\mathcal{X}_{q,2\omega-\Omega})\Bigr{]}\Biggr{]}\Biggr{\}}+\frac{D^{2}}{8d(\pi T)^{2}\omega}\int_{q,\Omega}q^{2}\mathcal{L}_{q}^{R}(\Omega)\mathcal{B}_{\Omega}\Biggl{[}3\psi^{\prime\prime}(\mathcal{X}_{q,\Omega})+\psi^{\prime\prime}(\mathcal{X}_{q,\Omega+2\omega})
+2(4πTω)2[ψ(𝒳q,2ω+Ω)ψ(𝒳q,Ω+ω)ψ(𝒳q,Ω)+ψ(𝒳q,Ωω)]]+D28d(πT)2ωq,Ωq2qR(Ω)\displaystyle+2\left(\frac{4\pi T}{\omega}\right)^{2}\Bigl{[}\psi(\mathcal{X}_{q,2\omega+\Omega})-\psi(\mathcal{X}_{q,\Omega+\omega})-\psi(\mathcal{X}_{q,\Omega})+\psi(\mathcal{X}_{q,\Omega-\omega})\Bigl{]}\Biggl{]}+\frac{D^{2}}{8d(\pi T)^{2}\omega}\int_{q,\Omega}q^{2}\mathcal{L}_{q}^{R}(\Omega)
×[ΩωΩ][ψ′′(𝒳q,2ωΩ)ψ′′(𝒳q,Ω)8πTiω[Φω(Ω)Φω(2ωΩ)]]\displaystyle\times\Bigl{[}\mathcal{B}_{\Omega-\omega}-\mathcal{B}_{\Omega}\Bigr{]}\Biggl{[}\psi^{\prime\prime}(\mathcal{X}_{q,2\omega-\Omega})-\psi^{\prime\prime}(\mathcal{X}_{q,\Omega})-\frac{8\pi T}{i\omega}\Bigl{[}\Phi_{\omega}(\Omega)-\Phi_{\omega}(2\omega-\Omega)\Bigr{]}\Biggr{]}
D28d(πT)2ωq,Ωq2qR(Ω)Ω{qR(Ω)ψ(𝒳q,Ω)[4Φ2ω(Ω)ψ(𝒳q,Ω+2ω)7ψ(𝒳q,Ω)]\displaystyle-\frac{D^{2}}{8d(\pi T)^{2}\omega}\int_{q,\Omega}q^{2}\mathcal{L}_{q}^{R}(\Omega)\mathcal{B}_{\Omega}\Biggl{\{}\mathcal{L}_{q}^{R}(\Omega)\psi^{\prime}(\mathcal{X}_{q,\Omega})\Bigl{[}4\Phi_{-2\omega}(\Omega)-\psi^{\prime}(\mathcal{X}_{q,\Omega+2\omega})-7\psi^{\prime}(\mathcal{X}_{q,\Omega})\Bigr{]}
+2qR(Ωω)[ΔqRRR(Ωω,Ω,ω)]2}\displaystyle+2\mathcal{L}_{q}^{R}(\Omega-\omega)\bigl{[}\Delta^{RRR}_{q}(\Omega-\omega,\Omega,\omega)\bigr{]}^{2}\Biggr{\}}
D28d(πT)2ωq,Ωq2qR(Ω)[ΩΩω]{qR(Ω)ψ(𝒳q,Ω)[ψ(𝒳q,Ω)ψ(𝒳q,2ωΩ)]\displaystyle-\frac{D^{2}}{8d(\pi T)^{2}\omega}\int_{q,\Omega}q^{2}\mathcal{L}_{q}^{R}(\Omega)\Bigl{[}\mathcal{B}_{\Omega}-\mathcal{B}_{\Omega-\omega}\Bigr{]}\Biggl{\{}\mathcal{L}_{q}^{R}(\Omega)\psi^{\prime}(\mathcal{X}_{q,\Omega})\Bigl{[}\psi^{\prime}(\mathcal{X}_{q,\Omega})-\psi^{\prime}(\mathcal{X}_{q,2\omega-\Omega})\Bigr{]}
qR(Ωω)[(ΔqRRR(Ωω,Ω,ω))2+iΔqRRR(Ωω,Ω,ω)ImΔqRRR(Ωω,Ω,ω)]\displaystyle-\mathcal{L}_{q}^{R}(\Omega-\omega)\Bigl{[}\bigl{(}\Delta^{RRR}_{q}(\Omega-\omega,\Omega,\omega)\bigr{)}^{2}+i\Delta^{RRR}_{q}(\Omega-\omega,\Omega,\omega)\operatorname{Im}\Delta^{RRR}_{q}(\Omega-\omega,\Omega,\omega)\Bigr{]}
qA(Ωω)[ΔqRRR(Ωω,Ω,ω)ReΔqRRR(Ωω,Ω,ω)(ΔqARR(Ωω,Ω,ω))2]}.\displaystyle-\mathcal{L}_{q}^{A}(\Omega-\omega)\Bigl{[}\Delta^{RRR}_{q}(\Omega-\omega,\Omega,\omega)\operatorname{Re}\Delta^{RRR}_{q}(\Omega-\omega,\Omega,\omega)-\bigl{(}\Delta^{ARR}_{q}(\Omega-\omega,\Omega,\omega)\bigr{)}^{2}\Bigr{]}\Biggr{\}}. (52)

We note that the first term in the right hand side of Eq. (52) is the full derivative with respect to momentum and, thus, as discussed above should vanish being supplemented by the corresponding contribution from the ballistic scales. Therefore, we shall disregard the corresponding term below.

3.4 Corrections to the conductivity in the dc limit due to superconducting fluctuations

Although corrections to the static conductivity due to superconducting fluctuations were discussed many times in literature, it is instructive to check that our result (51) for an arbitrary frequency correctly reproduces the well-known corrections in the static limit. In particular, the static anomalous Maki–Thompson correction becomes

δgMT,an(ω=0)=4q,Ω𝒞qR(0)ΩΩImqA(Ω)Imψ(𝒳q,Ω).\delta g^{\rm MT,an}(\omega=0)=4\int_{q,\Omega}\mathcal{C}_{q}^{R}(0)\partial_{\Omega}\mathcal{B}_{\Omega}\operatorname{Im}\mathcal{L}_{q}^{A}(\Omega)\operatorname{Im}\psi(\mathcal{X}_{q,\Omega}). (53)

The DOS correction in the dc limit, ω0\omega\to 0, acquires the following form

δgDOS(ω=0)=D8π2T2Imq,ΩΩqR(Ω)ψ′′(𝒳q,Ω)D2πTq,ΩΩΩImqR(Ω)Imψ(𝒳q,Ω).\displaystyle\delta g^{\rm DOS}(\omega=0)=-\frac{D}{8\pi^{2}T^{2}}\operatorname{Im}\int_{q,\Omega}\mathcal{B}_{\Omega}\mathcal{L}^{R}_{q}(\Omega)\psi^{\prime\prime}\left(\mathcal{X}_{q,\Omega}\right)-\frac{D}{2\pi T}\int_{q,\Omega}\partial_{\Omega}\mathcal{B}_{\Omega}\operatorname{Im}\mathcal{L}^{R}_{q}(\Omega)\operatorname{Im}\psi^{\prime}(\mathcal{X}_{q,\Omega}). (54)

At ω0\omega\to 0 the Aslamazov–Larkin correction can be written as follows

δgAL(ω=0)=D2dπ2T2q,Ωq2ΩΩImqR(Ω)Im[qR(Ω)ψ(𝒳q,Ω)]Reψ(𝒳q,Ω).\displaystyle\delta{g}^{\rm AL}(\omega=0)=-\frac{D^{2}}{d\pi^{2}T^{2}}\int_{q,\Omega}q^{2}\partial_{\Omega}\mathcal{B}_{\Omega}\operatorname{Im}\mathcal{L}_{q}^{R}(\Omega)\operatorname{Im}\Bigl{[}\mathcal{L}_{q}^{R}(\Omega)\psi^{\prime}(\mathcal{X}_{q,\Omega})\Bigr{]}\operatorname{Re}\psi^{\prime}(\mathcal{X}_{q,\Omega}). (55)

Finally, the contribution δgsc\delta g^{\rm sc} in the dc limit becomes

δgsc(ω=0)\displaystyle\delta g^{\rm sc}(\omega=0) =D22d(2πT)3Imq,Ωq2ΩqR2(Ω)ψ(𝒳q,Ω)ψ′′(𝒳q,Ω)\displaystyle=-\frac{D^{2}}{2d(2\pi T)^{3}}\operatorname{Im}\int_{q,\Omega}q^{2}\mathcal{B}_{\Omega}\mathcal{L}^{R2}_{q}(\Omega)\psi^{\prime}\left(\mathcal{X}_{q,\Omega}\right)\psi^{\prime\prime}\left(\mathcal{X}_{q,\Omega}\right)
D2d(2πT)2q,Ωq2ΩΩIm[qR2(Ω)ψ(𝒳q,Ω)]Imψ(𝒳q,Ω).\displaystyle-\frac{D^{2}}{d(2\pi T)^{2}}\int_{q,\Omega}q^{2}\partial_{\Omega}\mathcal{B}_{\Omega}\operatorname{Im}\Bigl{[}\mathcal{L}^{R2}_{q}(\Omega)\psi^{\prime}\left(\mathcal{X}_{q,\Omega}\right)\Bigr{]}\operatorname{Im}\psi^{\prime}\left(\mathcal{X}_{q,\Omega}\right). (56)

We note that Eqs. (53), (54), (55), and (56) coincide with the zero magnetic field limit of corresponding fluctuation corrections found in Ref. [21] and with the fluctuation corrections in the diffusive regime computed in Ref. [22].

4 Corrections to the ac conductivity due to superconducting fluctuations

Now we discuss the dependence of corrections to the ac conductivity due to superconducting fluctuations. It is convenient to introduce the following dimensionless variables, ϵ=lnT/Tc\epsilon=\ln T/T_{c} and α=ω/(4πT)\alpha=\omega/(4\pi T).

4.1 Anomalous Maki–Thompson contribution

We start from the anomalous Maki–Thompson correction, Eq. (25). We note that the integral over momentum in Eq. (25) diverges in the infra-red. Therefore, we need to introduce a finite dephasing rate 1/τϕ1/\tau_{\phi} which cuts off the pole in the cooperon propagator. In what follows we shall use dimensionless variable γ=1/(4πTτϕ)\gamma=1/(4\pi T\tau_{\phi}).

The asymptotic behavior of δgMT,an(ω)\delta g^{\rm MT,an}(\omega) at large frequencies, α1\alpha\gg 1, and for an arbitrary distance from superconducting transition temperatue, ϵ\epsilon, is given as follows (see A)

δgMT,an(ω)=π28ln24π1ϵ+lnα.\displaystyle\delta g^{\rm MT,an}(\omega)=\frac{\pi^{2}-8\ln 2}{4\pi}\frac{1}{\epsilon+\ln\alpha}. (57)

The anomalous Maki–Thompson correction vanishes in the limit of large frequencies, ωT\omega\gg T. Away from the superconducting transition, ϵ1\epsilon\gg 1, and for small frequencies, α1\alpha\ll 1, the anomalous Maki–Thompson contribution becomes

δgMT,an(ω)=(π6ϵ22πiα3ϵ)ln1γiα.\displaystyle\delta g^{\rm MT,an}(\omega)=\left(\frac{\pi}{6\epsilon^{2}}-\frac{2\pi i\alpha}{3\epsilon}\right)\ln\frac{1}{\gamma-i\alpha}. (58)

We note that the first term in the r.h.s. of Eq. (58) dominates over the second one at α1/ϵ\alpha\ll 1/\epsilon. At small frequencies, α1\alpha\ll 1, and in the vicinity of the superconducting transition, ϵ1\epsilon\ll 1, the anomalous Maki-Thompson correction reads

δgMT,an(ω)=12π1ϵ¯γ+iαlnϵ¯γiα,\displaystyle\delta g^{\rm MT,an}(\omega)=\frac{1}{2\pi}\frac{1}{\bar{\epsilon}-\gamma+i\alpha}\ln\frac{\bar{\epsilon}}{\gamma-i\alpha}, (59)

where ϵ¯=2ϵ/π21/(4πTτGL)\bar{\epsilon}=2\epsilon/\pi^{2}\equiv 1/(4\pi T\tau_{GL}). We note that we omit subleading terms proportional to lnϵ\ln\epsilon in Eq. (59) (see Refs. [23, 22] for details).

Refer to captionRefer to caption

Figure 1: The dependence of the real (left panel) and imaginary (right panel) parts of the anomalous Maki-Thompson correction on the frequency at different temperatures. The ratio of the dephasing rate to the temperature is fixed to the value γ=0.01\gamma=0.01.

The overall behavior of δgMT,an(ω)\delta g^{\rm MT,an}(\omega) as a function of the dimensionless frequency α\alpha at different values of ϵ\epsilon is shown in Fig. 1. The real part of δgMT,an(ω)\delta g^{\rm MT,an}(\omega) has non-monotonous behavior for temperatures close to TcT_{c}, i.e. for ϵ1\epsilon\ll 1 (see the left panel in Fig. 1). For temperatures away from TcT_{c}, i.e. for ϵ1\epsilon\gg 1, ReδgMT,an(ω)\operatorname{Re}\delta g^{\rm MT,an}(\omega) is also non-monotonous function of ω\omega. In the case TTcT\gg T_{c}, provided 1/τϕT/ln(T/Tc)1/\tau_{\phi}\ll T/\ln(T/T_{c}), the real part of δgMT,an(ω)\delta g^{\rm MT,an}(\omega) has the minimum at ωT/ln(T/Tc)\omega\sim T/\ln(T/T_{c}). The dependence of the imaginary part of δgMT,an(ω)\delta g^{\rm MT,an}(\omega) on the dimensionless frequency α\alpha at different values of ϵ\epsilon is figured in the right panel of Fig. 1. Exactly at zero frequency the imaginary part vanishes, ImδgMT,an(ω=0)=0\operatorname{Im}\delta g^{\rm MT,an}(\omega=0)=0. The imaginary part of δgMT,an(ω)\delta g^{\rm MT,an}(\omega) demonstrates non-monotonous behavior with ω\omega. At ultra small frequencies, ω1/τϕ\omega\ll 1/\tau_{\phi}, the imaginary part of δgMT,an(ω)\delta g^{\rm MT,an}(\omega) increases linearly with ω\omega. For TTcT\gg T_{c}, ImδgMT,an(ω)\operatorname{Im}\delta g^{\rm MT,an}(\omega) has the maximum at the frequency of the order of (T/τϕ)/ln(T/Tc)\sqrt{(T/\tau_{\phi})/\ln(T/T_{c})}. For temperatures near TcT_{c}, i.e. for ϵ1\epsilon\ll 1, the imaginary part of δgMT,an(ω)\delta g^{\rm MT,an}(\omega) has the maximum at ω1/τϕτGL\omega\sim 1/\sqrt{\tau_{\phi}\tau_{GL}}.

4.2 DOS correction

Next we turn our attention to the DOS correction to the conductivity. We note that the integrals over momentum and frequency in Eq. (36) diverge at the ultraviolet. Therefore, we shall introduce a cut-off corresponding to the inverse elastic mean free time, 1/τ1/\tau. Then, we can single out the part of δgDOS(ω)\delta g^{\rm DOS}(\omega) that depends on the cut-off,

δgDOS(ω)=1πlnln14πTcτ+δgfDOS(ω),\delta g^{\rm DOS}(\omega)=-\frac{1}{\pi}\ln\ln\frac{1}{4\pi T_{c}\tau}+\delta g^{\rm DOS}_{f}(\omega), (60)

such that δgfDOS(ω)\delta g^{\rm DOS}_{f}(\omega) is finite in the ultraviolet. We mention that the first term in the right hand side of Eq. (60) corresponds to the one loop DOS correction in the renormalization group equations for the conductivity [39].

Refer to captionRefer to caption

Figure 2: The dependence of the real (left panel) and imaginary (right panel) parts of the DOS correction on the frequency at different temperatures.

At large frequencies, α1\alpha\gg 1, and for an arbitrary magnitude of ϵ\epsilon the asymptotic behavior of δgfDOS(ω)\delta g^{\rm DOS}_{f}(\omega) is given as (see B)

δgfDOS(ω)=1πln(ϵ+lnα)i21ϵ+lnα.\displaystyle\delta g^{\rm DOS}_{f}(\omega)=\frac{1}{\pi}\ln\Bigl{(}\epsilon+\ln\alpha\Bigr{)}-\frac{i}{2}\frac{1}{\epsilon+\ln\alpha}. (61)

In the case of high temperatures, ϵ1\epsilon\gg 1, but small frequencies, α1\alpha\ll 1, the DOS correction becomes

δgfDOS(ω)=1πlnϵ+ln2πϵ2πiα3ϵ.\displaystyle\delta g^{\rm DOS}_{f}(\omega)=\frac{1}{\pi}\ln\epsilon+\frac{\ln 2}{\pi\epsilon}-\frac{2\pi i\alpha}{3\epsilon}. (62)

Near the superconducting transition, ϵ1\epsilon\ll 1, and at small frequencies, α1\alpha\ll 1, one can derive the following expression

δgfDOS(ω)=(14ζ(3)π3+iπα)ln1ϵ.\displaystyle\delta g^{\rm DOS}_{f}(\omega)=-\left(\frac{14\zeta(3)}{\pi^{3}}+i\pi\alpha\right)\ln\frac{1}{\epsilon}. (63)

The overall dependence of the real and imaginary parts of δgfDOS(ω)\delta g^{\rm DOS}_{f}(\omega) on frequency is shown in Fig. 2. The real part of δgfDOS(ω)\delta g^{\rm DOS}_{f}(\omega) grows monotonically with increase of the frequency. The imaginary part has the minimum.

4.3 Aslamazov–Larkin contribution

Next we consider the Aslamazov–Larkin contribution to the conductivity. We note that this correction is finite both in the infrared and the ultraviolet. We start from the case of large frequencies, α1\alpha\gg 1, and arbitrary temperature above TcT_{c}. Then we find (see C)

δgAL(ω)=c3AL(ϵ+lnα)3,\displaystyle\delta g^{\rm AL}(\omega)=\frac{c^{\rm AL}_{3}}{(\epsilon+\ln\alpha)^{3}}, (64)

where numerical constant c3AL0.170.89ic^{\rm AL}_{3}\approx 0.17-0.89i. In the case of small frequencies, α1\alpha\ll 1, and temperatures away from the superconducting transition, ϵ1\epsilon\gg 1, we obtain

δgAL(ω)=c4ALic5ALαϵ3,\displaystyle\delta g^{\rm AL}(\omega)=\frac{c^{\rm AL}_{4}-ic^{\rm AL}_{5}\alpha}{\epsilon^{3}}, (65)

where magnitudes of the numerical constants are c4AL1.44c^{\rm AL}_{4}\approx 1.44 and c5AL9.23c^{\rm AL}_{5}\approx 9.23. For temperatures close to superconducting transitions, ϵ1\epsilon\ll 1, and for small frequencies, α1\alpha\ll 1, the Aslamazov–Larkin contribution becomes

δgAL(ω)=π8ϵW1(π2α2ϵ)iπ3α32ϵ2W2(π2α2ϵ).\displaystyle\delta g^{\rm AL}(\omega)=\frac{\pi}{8\epsilon}W_{1}\left(\frac{\pi^{2}\alpha}{2\epsilon}\right)-\frac{i\pi^{3}\alpha}{32\epsilon^{2}}W_{2}\left(\frac{\pi^{2}\alpha}{2\epsilon}\right). (66)

Here the functions W1,2(z)W_{1,2}(z) are defined as follows

W1(z)=4z[arctan(z/2)1zln(1+z2/4)],\displaystyle W_{1}(z)=\frac{4}{z}\Bigl{[}\arctan(z/2)-\frac{1}{z}\ln(1+z^{2}/4)\Bigr{]},
W2(z)=8z3[arctan(z)2arctan(z/2)+zarctan3z28+5z2].\displaystyle W_{2}(z)=\frac{8}{z^{3}}\Bigl{[}\arctan(z)-2\arctan(z/2)+z\arctan\frac{3z^{2}}{8+5z^{2}}\Bigr{]}. (67)

We note that the part of Eq. (66) proportional to the function W1W_{1} coincides with the result derived in Ref. [5] and with the contribution ReσAL(1,1)\operatorname{Re}\sigma^{(1,1)}_{AL} of Ref. [23]. We note that there is also subleading term proportional to lnϵ\ln\epsilon (see Eq. (41) of Ref. [23]).

Refer to captionRefer to caption

Figure 3: The dependence of the real (left panel) and imaginary (right panel) parts of the Aslamazov-Larkin correction on the frequency at different temperatures.

The overall dependence of the real and imaginary parts of δgAL(ω)\delta g^{\rm AL}(\omega) on frequency is shown in Fig. 3. The real part of δgAL(ω)\delta g^{\rm AL}(\omega) decreases monotonously with increase of α\alpha. The imaginary part of δgAL(ω)\delta g^{\rm AL}(\omega) has the minimum at some frequency for all temperatures above the superconducting transition. For TT close to TcT_{c} the maximum is at αϵ\alpha\sim\epsilon.

4.4 The correction δgsc(ω)\delta g^{\rm sc}(\omega)

Finally, we turn our attention to the contribution δgsc(ω)\delta g^{\rm sc}(\omega), cf. Eq. (52). Similar to the Aslamasov–Larkin contribution, the correction δgsc(ω)\delta g^{\rm sc}(\omega) has divergencies neither in the infrared nor in the ultraviolet. At first, we consider the case of large frequencies, α1\alpha\gg 1, and arbitrary temperatures above the superconducting transition. Then we find (see D)

δgsc(ω)=23π1+3ln2ϵ+lnα.\displaystyle\delta g^{\rm sc}(\omega)=\frac{2}{3\pi}\frac{1+3\ln 2}{\epsilon+\ln\alpha}. (68)

In the case of small frequencies, α1\alpha\ll 1, but high temperatures, ϵ1\epsilon\gg 1, we obtain

δgsc(ω)=12πϵ(114π2iα3).\displaystyle\delta g^{\rm sc}(\omega)=\frac{1}{2\pi\epsilon}\left(1-\frac{14\pi^{2}i\alpha}{3}\right). (69)

In the vicinity of the superconducting transition, ϵ1\epsilon\ll 1, and small frequencies, α1\alpha\ll 1, we find

δgsc(ω)=iπ3α24ϵ2W3(π2α2ϵ)28ζ(3)π3lnϵ,\displaystyle\delta g^{\rm sc}(\omega)=\frac{i\pi^{3}\alpha}{24\epsilon^{2}}W_{3}\left(\frac{\pi^{2}\alpha}{2\epsilon}\right)-\frac{28\zeta(3)}{\pi^{3}}\ln\epsilon, (70)

where

W3(z)=3z2[2arctanzz+ln(1+z2)2].W_{3}(z)=\frac{3}{z^{2}}\left[2\frac{\arctan z}{z}+\ln(1+z^{2})-2\right]. (71)

Refer to captionRefer to caption

Figure 4: The dependence of the real (left panel) and imaginary (right panel) parts of δgsc(ω)\delta g^{\rm sc}(\omega) on the frequency at different temperatures.

The overall dependence of the real and imaginary parts of δgsc(ω)\delta g^{\rm sc}(\omega) on frequency is shown in Fig. 4. Both the real and imaginary parts of δgsc(ω)\delta g^{\rm sc}(\omega) have non-monotonous behavior. For temperatures away from the superconducting transition, ϵ1\epsilon\gg 1, Reδgsc(ω)\operatorname{Re}\delta g^{\rm sc}(\omega) is positive and has the maximum at some frequency α1\alpha\sim 1. The imaginary part of δgsc(ω)\delta g^{\rm sc}(\omega) has the minimum at some frequency α\alpha of the order of unity. Near the superconducting transition, ϵ1\epsilon\ll 1, the real (imaginary) part of δgsc(ω)\delta g^{\rm sc}(\omega) is positive and has the minimum (maximum) at αϵ\alpha\sim\epsilon.

4.5 The asymptotic expressions for δgCC(ω)\delta g^{CC}(\omega)

Now we are ready to present the asymptotic expressions for the correction to the ac conductivity due to superconducting fluctuations, i.e. due to interaction in the Cooper channel. It is convenient to single out the term which depends on the ultraviolet cutoff 1/τ1/\tau,

δgCC(ω)=1πlnln[1/(4πTcτ)]+δgfCC(ω).\displaystyle\delta g^{CC}(\omega)=-\frac{1}{\pi}\ln\ln[1/(4\pi T_{c}\tau)]+\delta g^{CC}_{f}(\omega). (72)

The contribution δgfCC(ω)\delta g^{CC}_{f}(\omega) is finite in the ultraviolet. For large frequencies in comparison with the temperature, ωT\omega\gg T, we find from Eqs. (57), (61), (64), and (68),

δgfCC(ω)=1πlnln[ω/(4πTc)]+3π2+86πi12πln[ω/(4πTc)].\displaystyle\delta g^{CC}_{f}(\omega)=\frac{1}{\pi}\ln\ln[\omega/(4\pi T_{c})]+\frac{3\pi^{2}+8-6\pi i}{12\pi\ln[\omega/(4\pi T_{c})]}. (73)

As expected the real and imaginary parts of the conductivity correction is dominated by the DOS contribution, Eq. (61). At small frequencies, ωT\omega\ll T, but for temperatures away from the superconducting transition, TTcT\gg T_{c}, using Eqs. (58), (62), (65), and (69), we obtain

δgfCC(ω)=1πlnln(T/Tc)+2ln2+12πln(T/Tc)+16(iωTπln(T/Tc))ln[(τϕ1iω)/(4πT)]ln(T/Tc).\displaystyle\delta g^{CC}_{f}(\omega)=\frac{1}{\pi}\ln\ln(T/T_{c})+\frac{2\ln 2+1}{2\pi\ln(T/T_{c})}+\frac{1}{6}\left(\frac{i\omega}{T}-\frac{\pi}{\ln(T/T_{c})}\right)\frac{\ln[(\tau_{\phi}^{-1}-i\omega)/(4\pi T)]}{\ln(T/T_{c})}. (74)

The real part of the conductivity correction is dominated by the DOS contribution as in the static case. The imaginary part of the conductivity correction is dominated by the anomalous Maki–Thompson term. In the region close to the superconducting transition, TTcTcT-T_{c}\ll T_{c}, and for small frequencies, ωTc\omega\ll T_{c}, with the help of Eqs. (59), (63), (66), and (70), we find

δgfCC(ω)=2TτGL1τGL/τϕ+iωτGLln[τGL/τϕiωτGL]+TτGL[W1(ωτGL)iωτGL2W2(ωτGL)\displaystyle\delta g^{CC}_{f}(\omega)=-\frac{2T\tau_{GL}}{1-\tau_{GL}/\tau_{\phi}+i\omega\tau_{GL}}\ln\Bigl{[}\tau_{GL}/\tau_{\phi}-i\omega\tau_{GL}\Bigr{]}+T\tau_{GL}\Bigl{[}W_{1}(\omega\tau_{GL})-\frac{i\omega\tau_{GL}}{2}W_{2}(\omega\tau_{GL})
+iωτGL3W3(ωτGL)].\displaystyle+\frac{i\omega\tau_{GL}}{3}W_{3}(\omega\tau_{GL})\Bigr{]}. (75)

The dependence of real and imaginary parts of δgfCC(ω)\delta g^{CC}_{f}(\omega) on frequency for different temperatures is shown in Fig. 5. For all temperatures above TcT_{c} the real (imaginary) part of δgfCC(ω)\delta g^{CC}_{f}(\omega) has the minimum (maximum). At temperatures TTcT\gg T_{c}, the minimum of ReδgfCC(ω)\operatorname{Re}\delta g^{CC}_{f}(\omega) occurs at frequency ω1/ln(T/Tc)\omega\sim 1/\ln(T/T_{c}) whereas the maximum of ImδgfCC(ω)\operatorname{Im}\delta g^{CC}_{f}(\omega) is at ωT/[τϕln(T/Tc)]\omega\sim\sqrt{T/[\tau_{\phi}\ln(T/T_{c})]}. In the vicinity of the superconducting transition, TTcTcT-T_{c}\ll T_{c}, the real part of δgfCC(ω)\delta g^{CC}_{f}(\omega) has a shallow minimum at frequency of the order of TcT_{c}. The maximum of the imaginary part of δgfCC(ω)\delta g^{CC}_{f}(\omega) is at frequency ω1/τϕτGL\omega\sim 1/\sqrt{\tau_{\phi}\tau_{GL}}. The frequency dependence of the real and imaginary part of δgfCC(ω)\delta g^{CC}_{f}(\omega) shown in Fig. 5 is in qualitative agreement with the measured conductivity near superconducting transition in thin films (see, e.g. Refs. [11, 13, 40]).

Refer to captionRefer to caption

Figure 5: The dependence of the real (left panel) and imaginary (right panel) parts of δgfCC(ω)\delta g^{CC}_{f}(\omega) on the frequency at different temperatures. The ratio of the dephasing rate to the temperature is fixed to the value γ=0.01\gamma=0.01.

5 Conclusion

To summarize, we reported the general analytical expression for the quantum correction to the ac conductivity of a disordered electron system in the diffusive regime. In addition to the well established weak localization and Altshuler–Aronov corrections, we computed the contributions to the ac conductivity due to superconducting fluctuations above the transition temperature.

In the static case, ω=0\omega=0, the weak localization, Altshuler–Aronov, and DOS corrections can be resumed in the form of the one-loop terms in the renormalization group equation for the conductivity [39]. The fluctuation propagator (14) is also subjected to renormalization. In particular, the diffusion coefficient DD and dimensionless Cooper interaction γc\gamma_{c} become scale dependent. Therefore, the contribution δgfCC\delta g_{f}^{CC} should be computed with the properly renormalized fluctuation propagator. For δgfCC(ω=0)\delta g_{f}^{CC}(\omega=0) such calculation results in the substitution of lnT/Tc\ln T/T_{c} by 1/γc(LT)1/\gamma_{c}(L_{T}). Here LT=D/TL_{T}=\sqrt{D/T} stands for the length scale associated with the temperature (see Refs. [41, 42] for details).

Present work can be extended in several ways. Our analysis can be extended to the pairing ac conductivity in the presence of a static magnetic field [43, 21]. Also it would be tempting to study the effect of superconducting fluctuations on the physical observables in non-standard symmetry classes [44]. The ac Nernst effect measured recently in thin superconducting films [45] suggests an interesting problem for computation of ac thermoelectric and thermal responses.

6 Acknowledgements

The author is grateful to A. Levchenko, K. Tikhonov, A. Petković, M. Skvortsov, and N. Stepanov for useful discussions. The author is indebted to I. Gornyi and A. Mirlin for fruitful collaboration at the initial stage of this work. The research was partially supported by the Russian Foundation for Basic Research (grant No. 20-52-12013) – Deutsche Forschungsgemeinschaft (grant No. EV 30/14-1) cooperation and by the Alexander von Humboldt Foundation.

Appendix A Anomalous Maki–Thompson contribution

In this Appendix we present derivation for the asymptotic expression of the anomalous Maki–Thompson correction. Let us introduce the function

G(z)=ϵ+ψ(z+1/2)ψ(1/2),\displaystyle G(z)=\epsilon+\psi(z+1/2)-\psi(1/2), (76)

where z=x+iyz=x+iy. Then, we can rewrite Eq. (25) as follows

δgMT,an(ω)=sinh(2πα)2πα0dxxiα+γdysinh(2πy)1sinh(2π(y+α))G(z)G(z2iα)G(z).\displaystyle\delta g^{\rm MT,an}(\omega)=\frac{\sinh(2\pi\alpha)}{2\pi\alpha}\int\limits_{0}^{\infty}\frac{dx}{x-i\alpha+\gamma}\int\limits_{-\infty}^{\infty}\frac{dy}{\sinh(2\pi y)}\frac{1}{\sinh(2\pi(y+\alpha))}\frac{G(z)-G(z^{*}-2i\alpha)}{G(z)}. (77)

Here we introduced the following notations, x=Dq2/(4πT)x=Dq^{2}/(4\pi T) and y=Ω/(4πT)y=\Omega/(4\pi T).

In the case α1\alpha\gg 1 it is convenient to rescale the integration variables as xαxx\to\alpha x and yαyy\to\alpha y. Then, we find

δgMT,an(ω)sinh(2πα)2π0dxxidysinh(2παy)1sinh(2πα(y+1))ln[z/(z2i)]ϵ+lnα+lnzψ(1/2)\displaystyle\delta g^{\rm MT,an}(\omega)\approx\frac{\sinh(2\pi\alpha)}{2\pi}\int\limits_{0}^{\infty}\frac{dx}{x-i}\int\limits_{-\infty}^{\infty}\frac{dy}{\sinh(2\pi\alpha y)}\frac{1}{\sinh(2\pi\alpha(y+1))}\frac{\ln[z/(z^{*}-2i)]}{\epsilon+\ln\alpha+\ln z-\psi(1/2)}
10dyπ0dxxiln[z/(z2i)]ϵ+lnα+lnzψ(1/2)π28ln24π1ϵ+lnαcMT,an(ϵ+lnα)2,\displaystyle\approx-\int\limits_{-1}^{0}\frac{dy}{\pi}\int\limits_{0}^{\infty}\frac{dx}{x-i}\frac{\ln[{z}/({z^{*}-2i})]}{\epsilon+\ln\alpha+\ln z-\psi(1/2)}\approx\frac{\pi^{2}-8\ln 2}{4\pi}\frac{1}{\epsilon+\ln\alpha}-\frac{c^{\rm MT,an}}{(\epsilon+\ln\alpha)^{2}}, (78)

where cMT,an0.810.54ic^{\rm MT,an}\approx 0.81-0.54i. We note that the main contribution to the integral comes from the region xyα1x\sim y\sim\alpha\gg 1.

In the case of small frequencies, α1\alpha\ll 1, but away from the transition temperature, ϵ1\epsilon\gg 1, we can expand Eq. (77) in 1/ϵ1/\epsilon:

δgMT,an(ω)=αiπϵK1+1ϵ2K2.\displaystyle\delta g^{\rm MT,an}(\omega)=\frac{\alpha}{i\pi\epsilon}K_{1}+\frac{1}{\epsilon^{2}}K_{2}. (79)

Here the first integral in the r.h.s. can be computed as follows

K1=0dxxiα+γ0𝑑ycoth(2πy)Imψ′′(1/2+x+iy)=01dxxiα+γ0𝑑ycoth(2πy)π3sinh(πy)cosh4(πy)\displaystyle K_{1}=\int\limits_{0}^{\infty}\frac{dx}{x-i\alpha+\gamma}\int\limits_{0}^{\infty}dy\coth(2\pi y)\operatorname{Im}\psi^{\prime\prime}(1/2+x+iy)=\int\limits_{0}^{1}\frac{dx}{x-i\alpha+\gamma}\int\limits_{0}^{\infty}dy\coth(2\pi y)\frac{\pi^{3}\sinh(\pi y)}{\cosh^{4}(\pi y)}
+01dxx0𝑑ycoth(2πy)Im[ψ′′(1/2+x+iy)ψ′′(1/2+iy)]+1dxx0𝑑ycoth(2πy)\displaystyle+\int\limits_{0}^{1}\frac{dx}{x}\int\limits_{0}^{\infty}dy\coth(2\pi y)\operatorname{Im}\Bigl{[}\psi^{\prime\prime}(1/2+x+iy)-\psi^{\prime\prime}(1/2+iy)\Bigr{]}+\int\limits_{1}^{\infty}\frac{dx}{x}\int\limits_{0}^{\infty}dy\coth(2\pi y)
×Imψ′′(1/2+x+iy)=2π23(ln1γiαc1MT,an),\displaystyle\times\operatorname{Im}\psi^{\prime\prime}(1/2+x+iy)=\frac{2\pi^{2}}{3}\Bigl{(}\ln\frac{1}{\gamma-i\alpha}-c^{\rm MT,an}_{1}\Bigr{)}, (80)

where the numerical constant is equal c11.62c_{1}\approx 1.62. The second integral in the r.h.s. of Eq. (79) can be evaluated as

K2=40dxxiα+γ0dy[Imψ(1/2+x+iy)]2sinh2(2πy)=4{π2401dxxiα+γ0dytanh2(πy)sinh2(2πy)\displaystyle K_{2}=4\int\limits_{0}^{\infty}\frac{dx}{x-i\alpha+\gamma}\int\limits_{0}^{\infty}\frac{dy\bigl{[}\operatorname{Im}\psi(1/2+x+iy)\bigr{]}^{2}}{\sinh^{2}(2\pi y)}=4\Biggl{\{}\frac{\pi^{2}}{4}\int\limits_{0}^{1}\frac{dx}{x-i\alpha+\gamma}\int\limits_{0}^{\infty}\frac{dy\tanh^{2}(\pi y)}{\sinh^{2}(2\pi y)}
+1dxx0dy[Imψ(1/2+x+iy)]2sinh2(2πy)+01dxx0dysinh2(2πy)[[Imψ(1/2+x+iy)]2\displaystyle+\int\limits_{1}^{\infty}\frac{dx}{x}\int\limits_{0}^{\infty}\frac{dy\bigl{[}\operatorname{Im}\psi(1/2+x+iy)\bigr{]}^{2}}{\sinh^{2}(2\pi y)}+\int\limits_{0}^{1}\frac{dx}{x}\int\limits_{0}^{\infty}\frac{dy}{\sinh^{2}(2\pi y)}\Bigl{[}\bigl{[}\operatorname{Im}\psi(1/2+x+iy)\bigr{]}^{2}
[Imψ(1/2+iy)]2]}=π6(ln1γiαc2MT,an).\displaystyle-\bigl{[}\operatorname{Im}\psi(1/2+iy)\bigr{]}^{2}\Bigr{]}\Biggr{\}}=\frac{\pi}{6}\Bigl{(}\ln\frac{1}{\gamma-i\alpha}-c^{\rm MT,an}_{2}\Bigr{)}. (81)

Here the numerical constant is equal c2MT,an2.19c^{\rm MT,an}_{2}\approx 2.19. Finally,

δgMT,an(ω)=2πα3iϵ(ln1γiαc1MT,an)+π6ϵ2(ln1γiαc2MT,an).\displaystyle\delta g^{\rm MT,an}(\omega)=\frac{2\pi\alpha}{3i\epsilon}\Bigl{(}\ln\frac{1}{\gamma-i\alpha}-c^{\rm MT,an}_{1}\Bigr{)}+\frac{\pi}{6\epsilon^{2}}\Bigl{(}\ln\frac{1}{\gamma-i\alpha}-c^{\rm MT,an}_{2}\Bigr{)}. (82)

Finally, we consider the region ϵ1\epsilon\ll 1 and α1\alpha\ll 1. Then we can split the anomalous Maki-Thompson correction into four parts

δgMT,an(ω)I1+I2ln(γiα)+I3+I4.\displaystyle\delta g^{\rm MT,an}(\omega)\approx I_{1}+I_{2}\ln(\gamma-i\alpha)+I_{3}+I_{4}. (83)

The first contribution can be estimated as follows

I1=01dxxiα+γ11dysinh(2πy)sinh(2π(y+α))ψ(1/2+x+iy)ψ(1/2+xiy2iα)ϵ+ψ(1/2+x+iy)ψ(1/2)\displaystyle I_{1}=\int\limits_{0}^{1}\frac{dx}{x-i\alpha+\gamma}\int\limits_{-1}^{1}\frac{dy}{\sinh(2\pi y)\sinh(2\pi(y+\alpha))}\frac{\psi(1/2+x+iy)-\psi(1/2+x-iy-2i\alpha)}{\epsilon+\psi(1/2+x+iy)-\psi(1/2)}
i2π201dx(xiα+γ)dyy1ϵ¯+x+iy=12π1ϵ¯γ+iαlnϵ¯γiα,\displaystyle\approx\frac{i}{2\pi^{2}}\int\limits_{0}^{1}\frac{dx}{(x-i\alpha+\gamma)}\int\limits_{-\infty}^{\infty}\frac{dy}{y}\frac{1}{\bar{\epsilon}+x+iy}=-\frac{1}{2\pi}\frac{1}{\bar{\epsilon}-\gamma+i\alpha}\ln\frac{\bar{\epsilon}}{\gamma-i\alpha}, (84)

where ϵ¯=2ϵ/π2\bar{\epsilon}=2\epsilon/\pi^{2}. We note that there are also subleading terms proportional to lnϵ\ln\epsilon. The other three contributions can be approximated by their values at ϵ=α=0\epsilon=\alpha=0,

I2=41dysinh2(2πy)|Imψ(1/2+iy)ψ(1/2+iy)ψ(1/2)|21.7106,\displaystyle I_{2}=-4\int\limits_{1}^{\infty}\frac{dy}{\sinh^{2}(2\pi y)}\left|\frac{\operatorname{Im}\psi(1/2+iy)}{\psi(1/2+iy)-\psi(1/2)}\right|^{2}\approx-1.7\cdot 10^{-6}, (85)
I3=401dxx1dysinh2(2πy)[|Imψ(1/2+x+iy)ψ(1/2+x+iy)ψ(1/2)|2|Imψ(1/2+iy)ψ(1/2+iy)ψ(1/2)|2]\displaystyle I_{3}=4\int\limits_{0}^{1}\frac{dx}{x}\int\limits_{1}^{\infty}\frac{dy}{\sinh^{2}(2\pi y)}\Biggl{[}\left|\frac{\operatorname{Im}\psi(1/2+x+iy)}{\psi(1/2+x+iy)-\psi(1/2)}\right|^{2}-\left|\frac{\operatorname{Im}\psi(1/2+iy)}{\psi(1/2+iy)-\psi(1/2)}\right|^{2}\Biggr{]}
1.4106,\displaystyle\approx-1.4\cdot 10^{-6}, (86)

and

I4=41dxx0dysinh2(2πy)|Imψ(1/2+x+iy)ψ(1/2+x+iy)ψ(1/2)|20.0021.\displaystyle I_{4}=4\int\limits_{1}^{\infty}\frac{dx}{x}\int\limits_{0}^{\infty}\frac{dy}{\sinh^{2}(2\pi y)}\left|\frac{\operatorname{Im}\psi(1/2+x+iy)}{\psi(1/2+x+iy)-\psi(1/2)}\right|^{2}\approx 0.0021. (87)

Appendix B DOS correction

In this Appendix we present derivation for the asymptotic expression of the DOS correction. We start from splitting the expression (36) into two parts

δgDOS(ω)=δg1DOS(ω)+δg2DOS(ω),\delta g^{\rm DOS}(\omega)=\delta g^{\rm DOS}_{1}(\omega)+\delta g^{\rm DOS}_{2}(\omega), (88)

where

δg1DOS(ω)=0dx4παdy[coth(2π(yα)coth(2πy)]G(z)G(z2iα)G(z),\displaystyle\delta g^{\rm DOS}_{1}(\omega)=\int\limits_{0}^{\infty}\frac{dx}{4\pi\alpha}\int\limits_{-\infty}^{\infty}dy\bigl{[}\coth(2\pi(y-\alpha)-\coth(2\pi y)\bigr{]}\frac{G^{\prime}(z^{*})-G^{\prime}(z-2i\alpha)}{G(z^{*})},
δg2DOS(ω)=0Λdx4παΛΛ𝑑ycoth(2πy)G(z)G(z2iα)G(z).\displaystyle\delta g^{\rm DOS}_{2}(\omega)=-\int\limits_{0}^{\Lambda}\frac{dx}{4\pi\alpha}\int\limits_{-\Lambda}^{\Lambda}dy\coth(2\pi y)\frac{G^{\prime}(z^{*})-G^{\prime}(z^{*}-2i\alpha)}{G(z^{*})}. (89)

Here we introduced the dimensionless ultra-violet cut off Λ=1/(4πTτ)1\Lambda=1/(4\pi T\tau)\gg 1. Next we split δg2DOS(ω)\delta g^{\rm DOS}_{2}(\omega) into three terms

δg2DOS(ω)=δg2,1DOS(ω)+δg2,2DOS(ω)+δg2,3DOS(ω).\displaystyle\delta g^{\rm DOS}_{2}(\omega)=\delta g^{\rm DOS}_{2,1}(\omega)+\delta g^{\rm DOS}_{2,2}(\omega)+\delta g^{\rm DOS}_{2,3}(\omega). (90)

The first two term is organized in such a way that one can integrate over xx exactly,

δg2,1DOS(ω)=0Λdy4παlnG(iy2iα)G(iy)G(iy)G(iy2iα)02αdy2παlnG(iy).\displaystyle\delta g^{\rm DOS}_{2,1}(\omega)=-\int\limits_{0}^{\Lambda}\frac{dy}{4\pi\alpha}\ln\frac{G(-iy-2i\alpha)G(iy)}{G(-iy)G(iy-2i\alpha)}-\int\limits_{0}^{2\alpha}\frac{dy}{2\pi\alpha}\ln G(-iy). (91)

The other two contributions are given as

δg2,2DOS(ω)=02αdy2παlnG(iy)+0dx4πα0dy[1coth(2πy)][G(z)G(z2iα)G(z)\displaystyle\delta g^{\rm DOS}_{2,2}(\omega)=\int\limits_{0}^{2\alpha}\frac{dy}{2\pi\alpha}\ln G(-iy)+\int\limits_{0}^{\infty}\frac{dx}{4\pi\alpha}\int\limits_{0}^{\infty}dy\bigl{[}1-\coth(2\pi y)\bigr{]}\Bigl{[}\frac{G^{\prime}(z^{*})-G^{\prime}(z^{*}-2i\alpha)}{G(z^{*})}
G(z)G(z2iα)G(z)],\displaystyle-\frac{G^{\prime}(z)-G^{\prime}(z-2i\alpha)}{G(z)}\Bigr{]}, (92)

and

δg2,3DOS(ω)=0dx4πα0𝑑y[G(z2iα)G(z2iα)G(z2iα)G(z)G(z2iα)G(z2iα)+G(z2iα)G(z)].\displaystyle\delta g^{\rm DOS}_{2,3}(\omega)=-\int\limits_{0}^{\infty}\frac{dx}{4\pi\alpha}\int\limits_{0}^{\infty}dy\left[\frac{G^{\prime}(z^{*}-2i\alpha)}{G(z^{*}-2i\alpha)}-\frac{G^{\prime}(z^{*}-2i\alpha)}{G(z^{*})}-\frac{G^{\prime}(z-2i\alpha)}{G(z-2i\alpha)}+\frac{G^{\prime}(z-2i\alpha)}{G(z)}\right]. (93)

The integral over yy in the expression for δg2,1DOS(ω)\delta g^{\rm DOS}_{2,1}(\omega) can be performed exactly,

δg2,1DOS(ω)=ΛΛ+2αdy4παlnG(iy)Λ2αΛdy4παlnG(iy)=1πlnG(iΛ)=1πln(ϵ+lnΛ).\displaystyle\delta g^{\rm DOS}_{2,1}(\omega)=-\int\limits_{\Lambda}^{\Lambda+2\alpha}\frac{dy}{4\pi\alpha}\ln G(-iy)-\int\limits_{\Lambda-2\alpha}^{\Lambda}\frac{dy}{4\pi\alpha}\ln G(-iy)=-\frac{1}{\pi}\ln G(-i\Lambda)=-\frac{1}{\pi}\ln(\epsilon+\ln\Lambda). (94)

Then we obtain Eq. (60) in which δgfDOS(ω)=δg1DOS(ω)+δg2,2DOS(ω)+δg2,3DOS(ω)\delta g^{\rm DOS}_{f}(\omega)=\delta g^{\rm DOS}_{1}(\omega)+\delta g^{\rm DOS}_{2,2}(\omega)+\delta g^{\rm DOS}_{2,3}(\omega).

In the case of large frequencies, α1\alpha\gg 1, it is convenient to perform rescaling xαxx\to\alpha x and yαyy\to\alpha y. Then we obtain

δg1DOS(ω)=1ϵ+lnα0dx4π01𝑑y[lnyln(2y)]=ln2π1ϵ+lnα.\displaystyle\delta g^{\rm DOS}_{1}(\omega)=\frac{1}{\epsilon+\ln\alpha}\int\limits_{0}^{\infty}\frac{dx}{4\pi}\int\limits_{0}^{1}dy\bigl{[}\ln y-\ln(2-y)\bigr{]}=-\frac{\ln 2}{\pi}\frac{1}{\epsilon+\ln\alpha}. (95)

Neglecting the second integral in the right hand side of Eq. (92), we find in a similar way

δg2,2DOS(ω)=1πln(ϵ+lnα)i21ϵ+lnα.\delta g^{\rm DOS}_{2,2}(\omega)=\frac{1}{\pi}\ln(\epsilon+\ln\alpha)-\frac{i}{2}\frac{1}{\epsilon+\ln\alpha}. (96)

Next, we find

δg2,3DOS(ω)=0dx4π0dy[1z2iln[z/(z2i)](ϵ+lnα+lnz)(ϵ+lnα+ln(z2i))\displaystyle\delta g^{\rm DOS}_{2,3}(\omega)=-\int\limits_{0}^{\infty}\frac{dx}{4\pi}\int\limits_{0}^{\infty}dy\Biggl{[}\frac{1}{z^{*}-2i}\frac{\ln[{z^{*}}/({z^{*}-2i})]}{(\epsilon+\ln\alpha+\ln z^{*})(\epsilon+\ln\alpha+\ln(z^{*}-2i))}
1z2iln[z/(z2i)](ϵ+lnα+lnz)(ϵ+lnα+ln(z2i))]\displaystyle-\frac{1}{z-2i}\frac{\ln[{z}/({z-2i})]}{(\epsilon+\ln\alpha+\ln z)(\epsilon+\ln\alpha+\ln(z-2i))}\Biggr{]}
1π1dx1dyIm[1z(ϵ+lnα+lnz)]21π1drr1(ϵ+lnα+lnr)2=1π1ϵ+lnα\displaystyle\approx-\frac{1}{\pi}\int\limits_{\sim 1}^{\infty}dx\int\limits_{\sim 1}^{\infty}dy\operatorname{Im}\left[\frac{1}{z(\epsilon+\ln\alpha+\ln z)}\right]^{2}\approx\frac{1}{\pi}\int\limits_{\sim 1}^{\infty}\frac{dr}{r}\frac{1}{(\epsilon+\ln\alpha+\ln r)^{2}}=\frac{1}{\pi}\frac{1}{\epsilon+\ln\alpha} (97)

In the case α1\alpha\ll 1 and ϵ1\epsilon\gg 1, we expand the integrand in δg1DOS(ω)\delta g^{\rm DOS}_{1}(\omega) in series in 1/ϵ1/\epsilon and obtain

δg1DOS(ω)=i2ϵ2dysinh2(2πy)0𝑑xx[ψ(1/2+xiy)Imψ(1/2+xiy)]=π24ϵ2.\displaystyle\delta g^{\rm DOS}_{1}(\omega)=-\frac{i}{2\epsilon^{2}}\int\limits_{-\infty}^{\infty}\frac{dy}{\sinh^{2}(2\pi y)}\int\limits_{0}^{\infty}dx\,\partial_{x}\Bigl{[}\psi(1/2+x-iy)\operatorname{Im}\psi(1/2+x-iy)\Bigr{]}=-\frac{\pi}{24\epsilon^{2}}. (98)

In a similar way, we find

δg2,2DOS(ω)=1πlnϵ1πϵ0𝑑y[1coth(2πy)]Imψ(1/2+iy)πiα2ϵ+αiπϵ0𝑑y[1coth(2πy)]\displaystyle\delta g^{\rm DOS}_{2,2}(\omega)=\frac{1}{\pi}\ln\epsilon-\frac{1}{\pi\epsilon}\int\limits_{0}^{\infty}dy[1-\coth(2\pi y)]\operatorname{Im}\psi^{\prime}(1/2+iy)-\frac{\pi i\alpha}{2\epsilon}+\frac{\alpha i}{\pi\epsilon}\int\limits_{0}^{\infty}dy[1-\coth(2\pi y)]
×Imψ′′(1/2+iy)=1πlnϵ+ln21πϵ2πiα3ϵ,\displaystyle\times\operatorname{Im}\psi^{\prime\prime}(1/2+iy)=\frac{1}{\pi}\ln\epsilon+\frac{\ln 2-1}{\pi\epsilon}-\frac{2\pi i\alpha}{3\epsilon}, (99)

Next, we can write

δg2,3DOS(ω)1π0dx0dyIm[G(z)G(z)]21π1drr1(ϵ+lnr)2=1πϵ,\displaystyle\delta g^{\rm DOS}_{2,3}(\omega)\approx-\frac{1}{\pi}\int\limits_{0}^{\infty}dx\int\limits_{0}^{\infty}dy\operatorname{Im}\left[\frac{G^{\prime}(z)}{G(z)}\right]^{2}\approx\frac{1}{\pi}\int\limits_{\sim 1}^{\infty}\frac{dr}{r}\frac{1}{(\epsilon+\ln r)^{2}}=\frac{1}{\pi\epsilon}, (100)

Finally, we consider small frequencies, α1\alpha\ll 1, and temperatures close to the superconducting transition, ϵ1\epsilon\ll 1. At first, we split δg1DOS(ω)\delta g^{\rm DOS}_{1}(\omega) into three parts

δg1DOS(ω)=sinh(2πα)4πα01𝑑x11dy[G(z)G(z2iα)]sinh(2π(yα))sinh(2πy)G(z)+21𝑑x01dysinh2(2πy)\displaystyle\delta g^{\rm DOS}_{1}(\omega)=\frac{\sinh(2\pi\alpha)}{4\pi\alpha}\int\limits_{0}^{1}dx\int\limits_{-1}^{1}\frac{dy\,[G^{\prime}(z^{*})-G^{\prime}(z-2i\alpha)]}{\sinh(2\pi(y-\alpha))\sinh(2\pi y)G(z^{*})}+2\int\limits_{1}^{\infty}dx\int\limits_{0}^{1}\frac{dy}{\sinh^{2}(2\pi y)}
×Imψ(1/2+z)Imψ(1/2+z)|ψ(1/2+z)ψ(1/2)|2+20𝑑x1dysinh2(2πy)Imψ(1/2+z)Imψ(1/2+z)|ψ(1/2+z)ψ(1/2)|2.\displaystyle\times\frac{\operatorname{Im}\psi^{\prime}(1/2+z)\operatorname{Im}\psi(1/2+z)}{\bigl{|}\psi(1/2+z)-\psi(1/2)\bigr{|}^{2}}+2\int\limits_{0}^{\infty}dx\int\limits_{1}^{\infty}\frac{dy}{\sinh^{2}(2\pi y)}\frac{\operatorname{Im}\psi^{\prime}(1/2+z)\operatorname{Im}\psi(1/2+z)}{\bigl{|}\psi(1/2+z)-\psi(1/2)\bigr{|}^{2}}. (101)

Here we neglected α\alpha and ϵ\epsilon whenever it is possible. Next, we omit the terms independent of α\alpha and ϵ\epsilon and expand the integrand in the first line of Eq. (101) to the lowest order in xx, yy, and α\alpha. Then we find with the logarithmic accuracy,

δg1DOS(ω)=14π201𝑑x11dyyiψ′′(1/2)+αψ′′′(1/2)ϵ+ψ(1/2)(xiy)=(7ζ(3)π3+iπα2)ln1ϵ.\displaystyle\delta g^{\rm DOS}_{1}(\omega)=\frac{1}{4\pi^{2}}\int\limits_{0}^{1}dx\int\limits_{-1}^{1}\frac{dy}{y}\frac{i\psi^{\prime\prime}(1/2)+\alpha\psi^{\prime\prime\prime}(1/2)}{\epsilon+\psi^{\prime}(1/2)(x-iy)}=-\left(\frac{7\zeta(3)}{\pi^{3}}+\frac{i\pi\alpha}{2}\right)\ln\frac{1}{\epsilon}. (102)

Next we find

δg2,2DOS(ω)=02αdy2παln[ϵiψ(1/2)y]+01dx4πα0dy[1coth(2πy)][G(z)G(z2iα)G(z)\displaystyle\delta g^{\rm DOS}_{2,2}(\omega)=\int\limits_{0}^{2\alpha}\frac{dy}{2\pi\alpha}\ln[\epsilon-i\psi^{\prime}(1/2)y]+\int\limits_{0}^{1}\frac{dx}{4\pi\alpha}\int\limits_{0}^{\infty}dy\bigl{[}1-\coth(2\pi y)\bigr{]}\Bigl{[}\frac{G^{\prime}(z^{*})-G^{\prime}(z^{*}-2i\alpha)}{G(z^{*})}
G(z)G(z2iα)G(z)]+1dxπ0dy[1coth(2πy)]ImG′′(z)G(z)\displaystyle-\frac{G^{\prime}(z)-G^{\prime}(z-2i\alpha)}{G(z)}\Bigr{]}+\int\limits_{1}^{\infty}\frac{dx}{\pi}\int\limits_{0}^{\infty}dy\bigl{[}1-\coth(2\pi y)\bigr{]}\operatorname{Im}\frac{G^{\prime\prime}(z)}{G(z)}
1πlnϵ+1π(1+iϵπ2α)ln(1iπ2αϵ)+ψ′′(1/2)iπ4α2πψ(1/2)01𝑑x0dy(ϵ¯+x)2+y2\displaystyle\approx\frac{1}{\pi}\ln\epsilon+\frac{1}{\pi}\left(1+\frac{i\epsilon}{\pi^{2}\alpha}\right)\ln\left(1-\frac{i\pi^{2}\alpha}{\epsilon}\right)+\frac{\psi^{\prime\prime}(1/2)-i\pi^{4}\alpha}{2\pi\psi^{\prime}(1/2)}\int\limits_{0}^{1}dx\int\limits_{0}^{\infty}\frac{dy}{(\bar{\epsilon}+x)^{2}+y^{2}} (103)

Hence, we find with the logarithmic accuracy

δg2,2DOS(ω)=(7ζ(3)π3+1π+iπα2)ln1ϵ+1π(1+iϵπ2α)ln(1iπ2αϵ).\displaystyle\delta g^{\rm DOS}_{2,2}(\omega)=-\left(\frac{7\zeta(3)}{\pi^{3}}+\frac{1}{\pi}+\frac{i\pi\alpha}{2}\right)\ln\frac{1}{\epsilon}+\frac{1}{\pi}\left(1+\frac{i\epsilon}{\pi^{2}\alpha}\right)\ln\left(1-\frac{i\pi^{2}\alpha}{\epsilon}\right). (104)

Also, we obtain with logarithmic accuracy

δg2,3DOS(ω)=1π0dx0dyIm[G(z)G(z)]2=1π01dy0dxIm[1ϵ¯+x+iy]2\displaystyle\delta g^{\rm DOS}_{2,3}(\omega)=-\frac{1}{\pi}\int\limits_{0}^{\infty}dx\int\limits_{0}^{\infty}dy\operatorname{Im}\left[\frac{G^{\prime}(z)}{G(z)}\right]^{2}=-\frac{1}{\pi}\int\limits_{0}^{1}dy\int\limits_{0}^{\infty}dx\operatorname{Im}\left[\frac{1}{\bar{\epsilon}+x+iy}\right]^{2}
1π0dx1dyIm[G(z)G(z)]2=1π01dyyϵ¯2+y21π0dx1dyIm[G(z)G(z)]2=1πln1ϵ.\displaystyle-\frac{1}{\pi}\int\limits_{0}^{\infty}dx\int\limits_{1}^{\infty}dy\operatorname{Im}\left[\frac{G^{\prime}(z)}{G(z)}\right]^{2}=\frac{1}{\pi}\int\limits_{0}^{1}dy\frac{y}{\bar{\epsilon}^{2}+y^{2}}-\frac{1}{\pi}\int\limits_{0}^{\infty}dx\int\limits_{1}^{\infty}dy\operatorname{Im}\left[\frac{G^{\prime}(z)}{G(z)}\right]^{2}=\frac{1}{\pi}\ln\frac{1}{\epsilon}. (105)

Appendix C Aslamazov–Larkin contribution

In this Appendix we present derivation for the asymptotic expression of the Aslamazov–Larkin contribution. This correction can be written as

δgAL(ω)=sinh(2πα)2πα30𝑑xxdysinh(2πy)sinh(2π(yα))ImG(ziα)|G(ziα)|2G(z+iα)G(ziα)G(z)\displaystyle\delta g^{\rm AL}(\omega)=-\frac{\sinh(2\pi\alpha)}{2\pi\alpha^{3}}\int\limits_{0}^{\infty}dx\ x\int\limits_{-\infty}^{\infty}\frac{dy}{\sinh(2\pi y)\sinh\bigl{(}2\pi(y-\alpha)\bigr{)}}\frac{\operatorname{Im}G(z-i\alpha)}{|G(z-i\alpha)|^{2}}\frac{G(z^{*}+i\alpha)-G(z^{*}-i\alpha)}{G(z^{*})}
×Im[G(z+iα)G(ziα)].\displaystyle\times\operatorname{Im}\Bigr{[}G(z^{*}+i\alpha)-G(z^{*}-i\alpha)\Bigr{]}. (106)

In the case of large frequencies, α1\alpha\gg 1, it is convenient to perform rescaling xαxx\to\alpha x and yαyy\to\alpha y. Then we obtain

δgAL(ω)=1π1(ϵ+lnα)30dxx01dyarctan(y1x)lnxiy+ixiyi[arctan(1yx)\displaystyle\delta g^{\rm AL}(\omega)=\frac{1}{\pi}\frac{1}{(\epsilon+\ln\alpha)^{3}}\int\limits_{0}^{\infty}dx\ x\int\limits_{0}^{1}dy\arctan\left(\frac{y-1}{x}\right)\ln\frac{x-iy+i}{x-iy-i}\,\Biggl{[}\arctan\left(\frac{1-y}{x}\right)
+arctan(1+yx)]c3AL(ϵ+lnα)3,\displaystyle+\arctan\left(\frac{1+y}{x}\right)\Biggr{]}\approx\frac{c^{\rm AL}_{3}}{(\epsilon+\ln\alpha)^{3}}, (107)

where the constant c3AL0.170.89ic^{\rm AL}_{3}\approx 0.17-0.89i.

In the case of small frequencies, α1\alpha\ll 1, and high temperatures, ϵ1\epsilon\gg 1, we can approximate the function G(z)G(z) in denominators of the integrand in Eq. (106) by ϵ\epsilon,

δgAL(ω)4iϵ30𝑑xx𝑑yxf(x,y)Imf(x,yα)Reψ(1/2+xiy),\displaystyle\delta g^{\rm AL}(\omega)\approx\frac{4i}{\epsilon^{3}}\int\limits_{0}^{\infty}dx\ x\int\limits_{-\infty}^{\infty}dy\ \partial_{x}f(x,y)\operatorname{Im}f(x,y-\alpha)\operatorname{Re}\psi^{\prime}(1/2+x-iy), (108)

where f(x,y)=ψ(1/2+xiy)/sinh(2πy)f(x,y)=\psi(1/2+x-iy)/\sinh(2\pi y). Expanding in α\alpha in the right hand side of Eq. (108), we obtain

δgAL(ω)c4ALc5ALiαϵ3,\displaystyle\delta g^{\rm AL}(\omega)\approx\frac{c^{\rm AL}_{4}-c^{\rm AL}_{5}i\alpha}{\epsilon^{3}}, (109)

where c4AL1.44c^{\rm AL}_{4}\approx 1.44 and c5AL9.23c^{\rm AL}_{5}\approx 9.23.

In the vicinity of the superconducting transition, ϵ1\epsilon\ll 1, and for small frequencies, α1\alpha\ll 1, we can expand the integrand in Eq. (106) in yy and xx,

δgAL(ω)iπ20𝑑xdyyx[(ϵ¯+x)2+(yα)2][ϵ¯+xiy]=π8ϵW1(π2α2ϵ)iπ3α32ϵ2W2(π2α2ϵ),\displaystyle\delta g^{\rm AL}(\omega)\approx-\frac{i}{\pi^{2}}\int\limits_{0}^{\infty}dx\int\limits_{-\infty}^{\infty}\frac{dy}{y}\frac{x}{[(\bar{\epsilon}+x)^{2}+(y-\alpha)^{2}][\bar{\epsilon}+x-iy]}=\frac{\pi}{8\epsilon}W_{1}\left(\frac{\pi^{2}\alpha}{2\epsilon}\right)-\frac{i\pi^{3}\alpha}{32\epsilon^{2}}W_{2}\left(\frac{\pi^{2}\alpha}{2\epsilon}\right), (110)

where the functions W1(X)W_{1}(X) and W2(X)W_{2}(X) are defined in Eq. (67). We note that there are also subleading terms proportional to lnϵ\ln\epsilon.

Appendix D The correction δgsc(ω)\delta g^{\rm sc}(\omega)

In this Appendix we present derivation for the asymptotic expression of the correction δgsc(ω)\delta g^{\rm sc}(\omega). It is convenient to split the expression (52) into four parts, δgsc(ω)=δgIsc(ω)+δgIIsc(ω)+δgIIIsc(ω)+δgIVsc(ω)\delta g^{\rm sc}(\omega)=\delta g^{\rm sc}_{I}(\omega)+\delta g^{\rm sc}_{II}(\omega)+\delta g^{\rm sc}_{III}(\omega)+\delta g^{\rm sc}_{IV}(\omega), and discuss each of them separately.

D.1 δgIsc(ω)\delta g^{\rm sc}_{I}(\omega)

The first contribution δgIsc(ω)\delta g^{\rm sc}_{I}(\omega) can be expressed in terms of the dimensionless parameters in the following way

δgIsc(ω)=14πα0dxxdyG(z)coth(2πy){3G′′(z)+G′′(z2iα)+2α2[G(z2iα)G(ziα)\displaystyle\delta g^{\rm sc}_{I}(\omega)=\frac{1}{4\pi\alpha}\int\limits_{0}^{\infty}dx\ x\int\limits_{-\infty}^{\infty}\frac{dy}{G(z)}\coth(2\pi y)\Biggl{\{}3G^{\prime\prime}(z)+G^{\prime\prime}(z-2i\alpha)+\frac{2}{\alpha^{2}}\Bigl{[}G(z-2i\alpha)-G(z-i\alpha)
G(z)+G(z+iα)]}.\displaystyle-G(z)+G(z+i\alpha)\Bigr{]}\Biggr{\}}. (111)

In the case of large frequencies, α1\alpha\gg 1, we perform rescaling xαxx\to\alpha x and yαyy\to\alpha y. Then we find

δgIsc(ω)14π1ϵ+lnα0𝑑xx𝑑ysgny[3z21(z2i)2+2ln(z2i)(z+i)(zi)z]\displaystyle\delta g^{\rm sc}_{I}(\omega)\approx\frac{1}{4\pi}\frac{1}{\epsilon+\ln\alpha}\int\limits_{0}^{\infty}dx\ x\int\limits_{-\infty}^{\infty}dy\operatorname{sgn}y\Biggl{[}-\frac{3}{z^{2}}-\frac{1}{(z-2i)^{2}}+2\ln\frac{(z-2i)(z+i)}{(z-i)z}\Biggr{]}
=16π52ln2+iπϵ+lnα.\displaystyle=\frac{1}{6\pi}\frac{5-2\ln 2+i\pi}{\epsilon+\ln\alpha}. (112)

At low frequencies, α1\alpha\ll 1, and at high temperatures, ϵ1\epsilon\gg 1, we expand Eq. (111) in α\alpha and obtain

δgIsc(ω)α6πϵ0𝑑xx𝑑ycoth(2πy)G′′′′(z)+5iα224πϵ0𝑑xx𝑑ycoth(2πy)G′′′′′(z)\displaystyle\delta g^{\rm sc}_{I}(\omega)\approx-\frac{\alpha}{6\pi\epsilon}\int\limits_{0}^{\infty}dx\ x\int\limits_{-\infty}^{\infty}dy\coth(2\pi y)\ G^{\prime\prime\prime\prime}(z)+\frac{5i\alpha^{2}}{24\pi\epsilon}\int\limits_{0}^{\infty}dx\ x\int\limits_{-\infty}^{\infty}dy\coth(2\pi y)\ G^{\prime\prime\prime\prime\prime}(z)
=2πiα9ϵ+c6α2ϵ,\displaystyle=-\frac{2\pi i\alpha}{9\epsilon}+\frac{c_{6}\alpha^{2}}{\epsilon}, (113)

where c63.83c_{6}\approx 3.83.

Near the superconducting transition, ϵ1\epsilon\ll 1, and for small frequencies, α1\alpha\ll 1, the correction δgIsc(ω)\delta g^{\rm sc}_{I}(\omega) does not diverge in the limit ϵ0\epsilon\to 0.

D.2 δgIIsc(ω)\delta g^{\rm sc}_{II}(\omega)

The contribution δgIIsc(ω)\delta g^{\rm sc}_{II}(\omega) reads

δgIIsc(ω)=14πα0dxxdyG(z)[coth(2π(yα))coth(2πy)]{G′′(ziα)G′′(z)\displaystyle\delta g^{\rm sc}_{II}(\omega)=\frac{1}{4\pi\alpha}\int\limits_{0}^{\infty}dx\ x\int\limits_{-\infty}^{\infty}\frac{dy}{G(z)}\Bigl{[}\coth\bigl{(}2\pi(y-\alpha)\bigr{)}-\coth(2\pi y)\Bigr{]}\Biggl{\{}G^{\prime\prime}(z-i\alpha)-G^{\prime\prime}(z^{*})
+2iα[G(z)+G(z)G(z+iα)iαG(z2iα)G(z2iα)G(ziα)iα]}.\displaystyle+\frac{2i}{\alpha}\Bigl{[}G^{\prime}(z^{*})+\frac{G(z^{*})-G(z^{*}+i\alpha)}{i\alpha}-G^{\prime}(z-2i\alpha)-\frac{G(z-2i\alpha)-G(z-i\alpha)}{i\alpha}\Bigr{]}\Biggr{\}}. (114)

In the case of large frequencies, α1\alpha\gg 1, it is convenient to rescale integration variables xαxx\to\alpha x and yαyy\to\alpha y. Hence, we obtain

δgIIsc(ω)12π1ϵ+lnα0𝑑xx01𝑑y{1(zi)2+1z2+2iz2iz2i+2lnz(zi)(z+i)(z2i)}\displaystyle\delta g^{\rm sc}_{II}(\omega)\approx-\frac{1}{2\pi}\frac{1}{\epsilon+\ln\alpha}\int\limits_{0}^{\infty}dx\ x\int\limits_{0}^{1}dy\Biggl{\{}-\frac{1}{(z-i)^{2}}+\frac{1}{z^{*2}}+\frac{2i}{z^{*}}-\frac{2i}{z-2i}+2\ln\frac{z^{*}(z-i)}{(z^{*}+i)(z-2i)}\Biggr{\}}
=16π14ln24iπϵ+lnα.\displaystyle=\frac{1}{6\pi}\frac{14\ln 2-4-i\pi}{\epsilon+\ln\alpha}. (115)

In the case of small frequencies, α1\alpha\ll 1, but well above the superconductivity transition temperature, ϵ1\epsilon\gg 1, we expand δgIIsc(ω)\delta g^{\rm sc}_{II}(\omega) in α\alpha. Then we find

δgIIsc(ω)sinh(2πα)4παϵdysinh(2π(yα))sinh(2πy){iα3[ψ(1/2iy)+2ψ(1/2+iy)]\displaystyle\delta g^{\rm sc}_{II}(\omega)\approx\frac{\sinh(2\pi\alpha)}{4\pi\alpha\epsilon}\int\limits_{-\infty}^{\infty}\frac{dy}{\sinh\bigl{(}2\pi(y-\alpha)\bigr{)}\sinh(2\pi y)}\Biggl{\{}\frac{i\alpha}{3}\Bigl{[}\psi^{\prime}\left({1}/{2}-iy\right)+2\psi^{\prime}\left({1}/{2}+iy\right)\Bigr{]}
α212[ψ′′(1/2iy)11ψ′′(1/2+iy)]}=iπα3ϵ+c6α2ϵ.\displaystyle-\frac{\alpha^{2}}{12}\Bigl{[}\psi^{\prime\prime}\left({1}/{2}-iy\right)-11\psi^{\prime\prime}\left({1}/{2}+iy\right)\Bigr{]}\Biggr{\}}=-\frac{i\pi\alpha}{3\epsilon}+\frac{c_{6}\alpha^{2}}{\epsilon}. (116)

The correction δgIIsc(ω)\delta g^{\rm sc}_{II}(\omega) becomes a constant in the limit α1\alpha\ll 1 and ϵ1\epsilon\ll 1.

D.3 δgIIIsc(ω)\delta g^{\rm sc}_{III}(\omega) and δgIVsc(ω)\delta g^{\rm sc}_{IV}(\omega)

The contributions δgIIIsc(ω)\delta g^{\rm sc}_{III}(\omega) and δgIVsc(ω)\delta g^{\rm sc}_{IV}(\omega) are given as

δgIIIsc(ω)=14πα0dxxdyG(z)coth(2πy){G(z)G(z)[3G(z)+G(z2iα)+2G(z)G(z2iα)iα]\displaystyle\delta g^{\rm sc}_{III}(\omega)=-\frac{1}{4\pi\alpha}\int\limits_{0}^{\infty}dx\ x\int\limits_{-\infty}^{\infty}\frac{dy}{G(z)}\coth(2\pi y)\Biggl{\{}\frac{G^{\prime}(z)}{G(z)}\Biggl{[}3G^{\prime}(z)+G^{\prime}(z-2i\alpha)+2\frac{G(z)-G(z-2i\alpha)}{i\alpha}\Biggr{]}
+2[G(z+iα)G(ziα)]2α2G(z+iα)}\displaystyle+2\frac{\bigl{[}G(z+i\alpha)-G(z-i\alpha)\bigr{]}^{2}}{\alpha^{2}G(z+i\alpha)}\Biggr{\}} (117)

and

δgIVsc(ω)=14πα0dxxdyG(z)[coth(2π(yα))coth(2πy)]{G(z)G(z)[G(z)G(z2iα)]\displaystyle\delta g^{\rm sc}_{IV}(\omega)=\frac{1}{4\pi\alpha}\int\limits_{0}^{\infty}dx\ x\int\limits_{-\infty}^{\infty}\frac{dy}{G(z^{*})}\Bigl{[}\coth\bigl{(}2\pi(y-\alpha)\bigr{)}-\coth(2\pi y)\Bigr{]}\Biggl{\{}\frac{G^{\prime}(z^{*})}{G(z^{*})}\Bigl{[}G^{\prime}(z^{*})-G^{\prime}(z-2i\alpha)\Bigr{]}
+G(z+iα)G(ziα)α2G(z+iα)[G(z+iα)G(ziα)+ReG(z+iα)ReG(ziα)]\displaystyle+\frac{G(z^{*}+i\alpha)-G(z^{*}-i\alpha)}{\alpha^{2}G(z^{*}+i\alpha)}\Bigl{[}G(z^{*}+i\alpha)-G(z^{*}-i\alpha)+\operatorname{Re}G(z^{*}+i\alpha)-\operatorname{Re}G(z^{*}-i\alpha)\Bigr{]}
G(z+iα)G(ziα)iα2G(ziα)Im[G(z+iα)G(ziα)]\displaystyle-\frac{G(z^{*}+i\alpha)-G(z^{*}-i\alpha)}{i\alpha^{2}G(z-i\alpha)}\operatorname{Im}\Bigl{[}G(z^{*}+i\alpha)-G(z^{*}-i\alpha)\Bigr{]}
[G(ziα)G(z2iα)+G(z)G(ziα)]2α2G(ziα)}.\displaystyle-\frac{\bigl{[}G(z-i\alpha)-G(z-2i\alpha)+G(z^{*})-G(z^{*}-i\alpha)\bigr{]}^{2}}{\alpha^{2}G(z-i\alpha)}\Biggr{\}}. (118)

It is convenient to split the contribution δgIVsc(ω)\delta g^{\rm sc}_{IV}(\omega) into two parts. The first part is given as

δgIV,1sc(ω)=sinh(2πα)4πα0𝑑xx𝑑yG(z)G2(z)G(z)G(z2iα)sinh(2π(yα))sinh(2πy).\displaystyle\delta g^{\rm sc}_{IV,1}(\omega)=\frac{\sinh(2\pi\alpha)}{4\pi\alpha}\int\limits_{0}^{\infty}dx\ x\int\limits_{-\infty}^{\infty}dy\frac{G^{\prime}(z^{*})}{G^{2}(z^{*})}\frac{G^{\prime}(z^{*})-G^{\prime}(z-2i\alpha)}{\sinh\bigl{(}2\pi(y-\alpha)\bigr{)}\sinh(2\pi y)}. (119)

The second part of δgIVsc(ω)\delta g^{\rm sc}_{IV}(\omega) can be combined with the term δgIIIsc(ω)\delta g^{\rm sc}_{III}(\omega). Then we find

δgIIIsc(ω)+δgIV,2sc(ω)=14πα0dxxdycoth(2πy){G(z)G2(z)[3G(z)+G(z2iα)\displaystyle\delta g^{\rm sc}_{III}(\omega)+\delta g^{\rm sc}_{IV,2}(\omega)=-\frac{1}{4\pi\alpha}\int\limits_{0}^{\infty}dx\ x\int\limits_{-\infty}^{\infty}dy\coth(2\pi y)\Biggl{\{}\frac{G^{\prime}(z)}{G^{2}(z)}\Biggl{[}3G^{\prime}(z)+G^{\prime}(z-2i\alpha)
+2G(z)G(z2iα)iα]+2[G(z+iα)G(ziα)]2α2G(z)G(z+iα)+G(z+iα)G(ziα)α2G(z)G(z+iα)\displaystyle+2\frac{G(z)-G(z-2i\alpha)}{i\alpha}\Biggr{]}+2\frac{\bigl{[}G(z+i\alpha)-G(z-i\alpha)\bigr{]}^{2}}{\alpha^{2}G(z)G(z+i\alpha)}+\frac{G(z^{*}+i\alpha)-G(z^{*}-i\alpha)}{\alpha^{2}G(z^{*})G(z^{*}+i\alpha)}
×[G(z+iα)G(ziα)+ReG(z+iα)ReG(ziα)]\displaystyle\times\Bigl{[}G(z^{*}+i\alpha)-G(z^{*}-i\alpha)+\operatorname{Re}G(z^{*}+i\alpha)-\operatorname{Re}G(z^{*}-i\alpha)\Bigr{]}
G(z+iα)G(ziα)iα2G(z)G(ziα)Im[G(z+iα)G(ziα)]\displaystyle-\frac{G(z^{*}+i\alpha)-G(z^{*}-i\alpha)}{i\alpha^{2}G(z^{*})G(z-i\alpha)}\operatorname{Im}\Bigl{[}G(z^{*}+i\alpha)-G(z^{*}-i\alpha)\Bigr{]}
[G(ziα)G(z2iα)+G(z)G(ziα)]2α2G(z)G(ziα)G(z)G(z2iα)α2G(ziα)G(z)\displaystyle-\frac{\bigl{[}G(z-i\alpha)-G(z-2i\alpha)+G(z^{*})-G(z^{*}-i\alpha)\bigr{]}^{2}}{\alpha^{2}G(z^{*})G(z-i\alpha)}-\frac{G(z^{*})-G(z^{*}-2i\alpha)}{\alpha^{2}G(z^{*}-i\alpha)G(z^{*})}
×[G(z)G(z2iα)+ReG(z)ReG(z2iα)]\displaystyle\times\Bigl{[}G(z^{*})-G(z^{*}-2i\alpha)+\operatorname{Re}G(z^{*})-\operatorname{Re}G(z^{*}-2i\alpha)\Bigr{]}
+G(z)G(z2iα)iα2G(ziα)G(z)Im[G(z)G(z2iα)]\displaystyle+\frac{G(z^{*})-G(z^{*}-2i\alpha)}{i\alpha^{2}G(z^{*}-i\alpha)G(z)}\operatorname{Im}\Bigl{[}G(z^{*})-G(z^{*}-2i\alpha)\Bigr{]}
+[G(z)G(ziα)+G(ziα)G(z2iα)]2α2G(ziα)G(z)}.\displaystyle+\frac{\bigl{[}G(z)-G(z-i\alpha)+G(z^{*}-i\alpha)-G(z^{*}-2i\alpha)\bigr{]}^{2}}{\alpha^{2}G(z^{*}-i\alpha)G(z)}\Biggr{\}}. (120)

At first, we consider the regime of large frequencies, α1\alpha\gg 1. It is convenient to make the rescaling xαxx\to\alpha x and yαyy\to\alpha y. Then we obtain

δgIV,1sc(ω)iπ(ϵ+lnα)20𝑑xx01𝑑y1z2(z2i)=18π9ln34ln22(ϵ+lnα)2.\displaystyle\delta g^{\rm sc}_{IV,1}(\omega)\approx\frac{i}{\pi(\epsilon+\ln\alpha)^{2}}\int\limits_{0}^{\infty}dx\ x\int\limits_{0}^{1}dy\frac{1}{z^{*2}(z^{*}-2i)}=-\frac{1}{8\pi}\frac{9\ln 3-4\ln 2-2}{(\epsilon+\ln\alpha)^{2}}. (121)

Also, we find

δgIIIsc(ω)+δgIV,2sc(ω)14π0dxxdysgny(ϵ+lnα+ln|z|)2{1z[3z+1z2i2ilnzz2i]\displaystyle\delta g^{\rm sc}_{III}(\omega)+\delta g^{\rm sc}_{IV,2}(\omega)\approx-\frac{1}{4\pi}\int\limits_{0}^{\infty}dx\ x\int\limits_{-\infty}^{\infty}dy\frac{\operatorname{sgn}y}{(\epsilon+\ln\alpha+\ln|z|)^{2}}\Biggl{\{}\frac{1}{z}\Bigl{[}\frac{3}{z}+\frac{1}{z-2i}-2i\ln\frac{z}{z-2i}\Bigr{]}
+2ln2z+izi+2ln2z+iziln2(zi)z(zi)(z2i)2ln2zz2i+ln2z(zi)(z2i)(zi)]\displaystyle+2\ln^{2}\frac{z+i}{z-i}+2\ln^{2}\frac{z^{*}+i}{z^{*}-i}-\ln^{2}\frac{(z-i)z^{*}}{(z^{*}-i)(z-2i)}-2\ln^{2}\frac{z^{*}}{z^{*}-2i}+\ln^{2}\frac{z(z^{*}-i)}{(z^{*}-2i)(z-i)}\Biggr{]}
12π1drr1(ϵ+lnα+lnr)2=12π1ϵ+lnα.\displaystyle\approx\frac{1}{2\pi}\int\limits_{\sim 1}^{\infty}\frac{dr}{r}\frac{1}{(\epsilon+\ln\alpha+\ln r)^{2}}=\frac{1}{2\pi}\frac{1}{\epsilon+\ln\alpha}. (122)

Next, we consider the case of small frequencies, α1\alpha\ll 1, and high temperatures, ϵ1\epsilon\gg 1. Then, expanding in α\alpha, we find

δgIV,1sc(ω)2ϵ20𝑑xx𝑑y[ImG(z)]2sinh2(2πy)c7ϵ2,\displaystyle\delta g^{\rm sc}_{IV,1}(\omega)\approx-\frac{2}{\epsilon^{2}}\int\limits_{0}^{\infty}dx\ x\int\limits_{-\infty}^{\infty}dy\frac{[\operatorname{Im}G^{\prime}(z)]^{2}}{\sinh^{2}(2\pi y)}\approx-\frac{c_{7}}{\epsilon^{2}}, (123)

where c70.047c_{7}\approx 0.047. For the contribution δgIIIsc(ω)+δgIV,2sc(ω)\delta g^{\rm sc}_{III}(\omega)+\delta g^{\rm sc}_{IV,2}(\omega) the integrals over xx and yy are dominated by their large values of the order of expϵ\exp\epsilon. Therefore, after expansion in α\alpha, we obtain

δgIIIsc(ω)+δgIV,2sc(ω)1π0𝑑xx0𝑑ycoth(2πy)(ϵ+ln|z|)2ImG(z)[3G′′(z)+2G′′(z)]\displaystyle\delta g^{\rm sc}_{III}(\omega)+\delta g^{\rm sc}_{IV,2}(\omega)\approx\frac{1}{\pi}\int\limits_{0}^{\infty}dx\ x\int\limits_{0}^{\infty}dy\frac{\coth(2\pi y)}{(\epsilon+\ln|z|)^{2}}\operatorname{Im}G^{\prime}(z)\Bigl{[}3G^{\prime\prime}(z)+2G^{\prime\prime}(z^{*})\Bigr{]}
12π1drr1(ϵ+lnr)2=12πϵ.\displaystyle\approx\frac{1}{2\pi}\int\limits_{\sim 1}^{\infty}\frac{dr}{r}\frac{1}{(\epsilon+\ln r)^{2}}=\frac{1}{2\pi\epsilon}. (124)

We note that the terms of the next order in α\alpha has additional smallness in 1/ϵ1/\epsilon.

Finally, we consider the vicinity of the superconducting transition, ϵ1\epsilon\ll 1, and small frequencies, α1\alpha\ll 1. Then expanding in xx and yy we find

δgIV,1sc(ω)ψ′′(1/2)π2ψ(1/2)01𝑑x0𝑑yx(ϵ¯+x)[(ϵ¯+x)2+y2]2=7ζ(3)π3lnϵ.\displaystyle\delta g^{\rm sc}_{IV,1}(\omega)\approx\frac{\psi^{\prime\prime}(1/2)}{\pi^{2}\psi^{\prime}(1/2)}\int\limits_{0}^{\sim 1}dx\int\limits_{0}^{\infty}dy\frac{x(\bar{\epsilon}+x)}{[(\bar{\epsilon}+x)^{2}+y^{2}]^{2}}=\frac{7\zeta(3)}{\pi^{3}}\ln\epsilon. (125)

Here we neglected the dependence on α\alpha since it does not lead to terms divergent for ϵ0\epsilon\to 0. In order to analyse the term δgIIIsc(ω)+δgIV,2sc(ω)\delta g^{\rm sc}_{III}(\omega)+\delta g^{\rm sc}_{IV,2}(\omega), at first, we perform expansion of enumerators in α\alpha in the right hand side of Eq. (120),

δgIIIsc(ω)+δgIV,2sc(ω)18π2α0dxxdyy{8G2(z)G2(z)4G(z)ReG(z)G(z)G(z+iα)8G2(z)G(z)G(ziα)\displaystyle\delta g^{\rm sc}_{III}(\omega)+\delta g^{\rm sc}_{IV,2}(\omega)\approx-\frac{1}{8\pi^{2}\alpha}\int\limits_{0}^{\infty}dx\ x\int\limits_{-\infty}^{\infty}\frac{dy}{y}\Biggl{\{}\frac{8G^{\prime 2}(z)}{G^{2}(z)}-\frac{4G^{\prime}(z)\operatorname{Re}G^{\prime}(z)}{G(z)G(z+i\alpha)}-\frac{8G^{\prime 2}(z)}{G(z)G(z-i\alpha)}
+4G(z)ReG(z)G(z)G(ziα)+2iαG(z)G′′(z)G2(z)+iα2G(z)G′′(z)+iG′′(z)ImG(z)+G(z)ReG′′(z)G(z)G(ziα)\displaystyle+\frac{4G^{\prime}(z)\operatorname{Re}G^{\prime}(z)}{G(z)G(z-i\alpha)}+\frac{2i\alpha G^{\prime}(z)G^{\prime\prime}(z)}{G^{2}(z)}+i\alpha\frac{2G^{\prime}(z)G^{\prime\prime}(z)+iG^{\prime\prime}(z)\operatorname{Im}G^{\prime}(z)+G^{\prime}(z)\operatorname{Re}G^{\prime\prime}(z)}{G(z)G(z-i\alpha)}
iα4[3G′′(z)+G′′(z)]ReG(z)G′′(z)ReG(z)iG(z)ImG′′(z)G(z)G(ziα)}.\displaystyle-i\alpha\frac{4[3G^{\prime\prime}(z)+G^{\prime\prime}(z^{*})]\operatorname{Re}G^{\prime}(z)-G^{\prime\prime}(z)\operatorname{Re}G^{\prime}(z)-iG^{\prime}(z)\operatorname{Im}G^{\prime\prime}(z)}{G(z^{*})G(z-i\alpha)}\Biggr{\}}. (126)

Expanding the function GG in powers of its argument, we obtain

δgIIIsc(ω)+δgIV,2sc(ω)απ20𝑑xxdyy1(ϵ¯+z)2(ϵ¯+z+iα)(ϵ¯+ziα)\displaystyle\delta g^{\rm sc}_{III}(\omega)+\delta g^{\rm sc}_{IV,2}(\omega)\approx-\frac{\alpha}{\pi^{2}}\int\limits_{0}^{\infty}dx\ x\int\limits_{-\infty}^{\infty}\frac{dy}{y}\frac{1}{(\bar{\epsilon}+z)^{2}(\bar{\epsilon}+z+i\alpha)(\bar{\epsilon}+z-i\alpha)}
5iψ′′(1/2)8π2ψ(1/2)0𝑑xxdyy1(ϵ¯+z)2=iα6πϵ¯2W3(αϵ¯)35ζ(3)π3lnϵ,\displaystyle-\frac{5i\psi^{\prime\prime}(1/2)}{8\pi^{2}\psi^{\prime}(1/2)}\int\limits_{0}^{\infty}dx\ x\int\limits_{-\infty}^{\infty}\frac{dy}{y}\frac{1}{(\bar{\epsilon}+z)^{2}}=\frac{i\alpha}{6\pi\bar{\epsilon}^{2}}W_{3}\left(\frac{\alpha}{\bar{\epsilon}}\right)-\frac{35\zeta(3)}{\pi^{3}}\ln\epsilon, (127)

where the function W3(z)W_{3}(z) is given by Eq.(71). Here we neglected terms of the order of α/ϵ¯\alpha/\bar{\epsilon}.

References

  • [1] A. I. Larkin and A. A. Varlamov, Theory of fluctuations in superconductors, OUP Oxford (2009).
  • [2] A. A. Varlamov, A. Galda, A. Glatz, Fluctuation spectroscopy: From Rayleigh–Jeans waves to Abrikosov vortex clusters, Rev. Mod. Phys. 90, 015009 (2018).
  • [3] K. Maki, The critical fluctuation of the order parameter in type-II superconductor, Progr. Theor. Phys. 39, 897 (1968).
  • [4] R.S. Thompson, Microwave, flux flow, and fluctuation resistance of dirty type-II superconductors, Phys. Rev. B 1, 327 (1970).
  • [5] H. Schmidt, The onset of superconductivity in the time-dependent Ginzburg–Landau Theory, Z. Phys. 216, 336 (1968).
  • [6] L.G. Aslamazov and A.I. Larkin, Effect of fluctuations on the properties of a superconductor above the critical temperature, Sov. Phys. Solid State 10, 875 (1968).
  • [7] R. V. D’Aiello and S. J. Freedman, Microwave conductivity of granular aluminum films in the superconducting transition region, Phys. Rev. Lett. 22, 515 (1969).
  • [8] S. L. (A.) Lehoczky and C. V. Briscoe, Fluctuation effects in the ac conductivity of thin lead films above the superconducting transition temperature, Phys. Rev. Lett. 23, 695 (1969).
  • [9] S. L. Lehoczky and C. V. Briscoe, Fluctuation effects in the ac conductivity of thin superconducting lead films at microwave frequencies, Phys. Rev. B 4, 3938 (1971).
  • [10] T. Ohashi, H. Kitano, A. Maeda, H. Akaike, and A. Fujimaki, Dynamic fluctuations in the superconductivity of NbN films from microwave conductivity measurements, Phys. Rev. B 73, 174522 (2006).
  • [11] W. Liu, M. Kim, G. Sambandamurthy, and N. P. Armitage, Dynamical study of phase fluctuations and their critical slowing down in amorphous superconducting films, Phys. Rev. B 84, 024511 (2011).
  • [12] U. S. Pracht, M. Scheffler, M. Dressel, D. F. Kalok, C. Strunk, and T. I. Baturina, Direct observation of the superconducting gap in a thin film of titanium nitride using terahertz spectroscopy, Phys. Rev. B 86, 184503 (2012).
  • [13] M. Mondal, A. Kamlapure, S. C. Ganguli, J. Jesudasan, V. Bagwe, L. Benfatto and P. Raychaudhuri, Enhancement of the finite-frequency superfluid response in the pseudogap regime of strongly disordered superconducting films, Sci. Rep. 3, 1357 (2013).
  • [14] D. Sherman, B. Gorshunov, S. Poran, N. Trivedi, E. Farber, M. Dressel, and A. Frydman, Effect of Coulomb interactions on the disorder-driven superconductor-insulator transition, Phys. Rev. B 89, 035149 (2014).
  • [15] J. Simmendinger, U. S. Pracht, L. Daschke, T. Proslier, J. A. Klug, M. Dressel, and M. Scheffler, Superconducting energy scales and anomalous dissipative conductivity in thin films of molybdenum nitride, Phys. Rev. B 94, 064506 (2016).
  • [16] B. L. Altshuler, A. A. Varlamov, and M. Yu. Reizer, Interelectron effects and the conductivity of disordered two-dimensional electron systems, Sov. Phys. JETP 57, 1329 (1983).
  • [17] L. P. Gor’kov, A. I. Larkin, and D.E. Khmelnitskii, Particle conductivity in a two-dimensional random potential, JETP Lett. 30, 248 (1979).
  • [18] B. L. Altshuler and A. G. Aronov, Contribution to the theory of disordered metals in strongly doped semiconductors, Sov. Phys. JETP 50, 968 (1979).
  • [19] P. A. Lee and T. V. Ramakrishnan, Disordered electronic systems, Rev. Mod. Phys. 57, 287 (1985).
  • [20] B. L. Altshuler and A. G. Aronov, Electron-electron interactions in disordered conductors, eds. A. J. Efros and M. Pollack, Elsevier Science Publishers, North-Holland (1985).
  • [21] K. S. Tikhonov, G. Schwiete, A. M. Finkelstein, Fluctuation conductivity in disordered superconducting films, Phys. Rev. B 85, 174527 (2012).
  • [22] N. A. Stepanov, M. A. Skvortsov, Superconducting fluctuations at arbitrary disorder strength, Phys. Rev. B 97, 144517 (2018).
  • [23] A. Petković and V. M. Vinokur, Pairing fluctuation ac conductivity of disordered thin films, J. Phys.: Condens. Matter, 25, 355701 (2013).
  • [24] A. Kamenev and A. Levchenko, Keldysh technique and non-linear σ\sigma-model: basic principles and applications, Adv. Phys. 58, 197 (2009).
  • [25] A. M. Finkel’stein, Electron Liquid in Disordered Conductors, vol. 14 of Soviet Scientific Reviews, ed. by I.M.  Khalatnikov, Harwood Academic Publishers, London, (1990).
  • [26] D. Belitz, T. R. Kirkpatrick, The Anderson-Mott transition, Rev. Mod. Phys. 66, 261 (1994).
  • [27] I. S. Burmistrov, Finkel’stein nonlinear sigma model: interplay of disorder and interaction in 2D electron systems, ZhETF 156, 724 (2019).
  • [28] I. S. Burmistrov, I. V. Gornyi, A. D. Mirlin, Superconductor-insulator transitions: Phase diagram and magnetoresistance, Phys. Rev. B 92, 014506 (2015).
  • [29] A. M. Finkelstein, On the frequency and temperature dependence of the conductivity near a metal-insulator transition, JETP Lett. 37, 517 (1983).
  • [30] A. M. Finkelstein, Influence of coulomb interaction on the properties of disordered metals, Sov. Phys. JETP 53, 97 (1983).
  • [31] A. M. Finkelstein, Metal-insulator transition in a disordered system, Sov. Phys. JETP 59, 212 (1984).
  • [32] C. Castellani, C. Di Castro, P. A. Lee, and M. Ma, Interaction-driven metal-insulator transitions in disordered fermion systems, Phys. Rev. B 30, 527 (1984).
  • [33] L. G. Aslamazov and A. A. Varlamov, Fluctuation conductivity in intercalated superconductors, J. Low Temp. Phys. 38, 223 (1979).
  • [34] A. I. Larkin, Reluctance of two-dimensional systems, JETP Lett. 31, 219 (1980) .
  • [35] E. Abrahams, M. Redi, J. W. F. Woo, Effect of fluctuations n electronic properties above the superconducting transition, Phys. Rev. B 1, 208 (1970).
  • [36] C. Di Castro, R. Raimondi, C. Castellani, and A. A. Varlamov, Superconductive fluctuations in the density of states and tunneling resistance in high-Tc superconductors, Phys. Rev. B 42, 10211 (1990).
  • [37] B. M. Anderson, R. Boyack, C.-Te Wu, and K. Levin, Going beyond the BCS level in the superfluid path integral: A consistent treatment of electrodynamics and thermodynamics, Phys. Rev. B 93, 180504(R) (2016).
  • [38] R. Boyack, Restoring gauge invariance in conventional fluctuation corrections to a superconductor, Phys. Rev. B 98, 184504 (2018).
  • [39] A.M. Finkelstein, Weak-localization and Coulomb interactions in disordered systems, Z. Phys. B 56, 189 (1984).
  • [40] F. Nabeshima, K. Nagasawa, A. Maeda, and Y. Imai, Superconducting fluctuations in FeSe0.5Te0.5 thin films probed via microwave spectroscopy, Phys. Rev. B 97, 024504 (2018).
  • [41] I. S. Burmistrov, I. V. Gornyi, A. D. Mirlin, Superconductor-insulator transitions: Phase diagram and magnetoresistance, Phys. Rev. B 92, 014506 (2015).
  • [42] I. S. Burmistrov, I. V. Gornyi, A. D. Mirlin, Local density of states and its mesoscopic fluctuations near the transition to a superconducting state in disordered systems, Phys. Rev. B 93, 205432 (2016).
  • [43] V. M. Galitski and A. I. Larkin, , Phys. Rev. B 63, 174506 (2001).
  • [44] L. Dell’Anna, Beta-functions of non-linear models for disordered and interacting electron systems, Ann. der Phys. 529, 1600317 (2017).
  • [45] Y. Wu, A. Frydman, and A. Roy, AC Measurement of the Nernst effect of thin films at low temperatures, arXiv:2003.07946 .