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

A theory of coupled dual dynamics of macroscopic phase coherence and microscopic electronic fluids: effect of dephasing on cuprate superconductivity

F. Yang [email protected]. Hefei National Laboratory for Physical Sciences at Microscale, Department of Physics, and CAS Key Laboratory of Strongly-Coupled Quantum Matter Physics, University of Science and Technology of China, Hefei, Anhui, 230026, China    M. W. Wu [email protected]. Hefei National Laboratory for Physical Sciences at Microscale, Department of Physics, and CAS Key Laboratory of Strongly-Coupled Quantum Matter Physics, University of Science and Technology of China, Hefei, Anhui, 230026, China
Abstract

By using the gauge-invariant kinetic equation approach [Yang and Wu, Phys. Rev. B 98, 094507 (2018); 100, 104513 (2019)], we construct the coupled dual dynamics of macroscopic phase coherence and microscopic electronic fluids in cuprate superconductors. We prove that the developed dual dynamics provides an efficient and simplified approach to formulate the dephasing process of macroscopic superconducting phase coherence, as well as its influence on microscopic electronic fluids (including gap, densities of superfluid and normal fluid, and in particular, the transport property to determine superconducting transition temperature TcT_{c}). We then present theoretical description of the preformed Cooper pairs in pseudogap state. The key origin of pseudogap state comes from the quantum effect of disorder, which excites the macroscopic inhomogeneous phase fluctuation through Josephson effect. Influenced by this phase fluctuation, there exist normal fluid and viscous superfluid below TcT_{c} in cuprate superconductors, in addition to conventional non-viscous superfluid. The normal fluid always emerges around nodal points even at zero temperature, whereas the viscous superfluid emerges due to the friction between superfluid and normal fluid. Particularly, the non-viscous superfluid gets suppressed when the phase fluctuation is enhanced by increasing temperature, until vanishes at TcT_{c}. Then, the system enters the pseudogap state, showing the nonzero resistivity as well as the finite gap from the viscous superfluid. By further increasing temperature to TosT^{\rm os}, the viscous superfluid and hence gap vanish. An experimental scheme to distinguish the densities of normal fluid as well as viscous and non-viscous superfluids is proposed. Finally, this theory is also applied to low-dimensional disordered ss-wave superconductors.

pacs:
74.72.Kf, 74.62.En, 74.72.-h, 74.40.+k, 74.78.-w

I Introduction

Within the framework of Bardeen, Cooper and Schrieffer (BCS) theory of conventional superconductivityBCS , the emergence of the superconducting order parameter by forming Cooper pairs occurs upon cooling at the critical temperature TcT_{c}, below which superconductivity (zero-resistance) phenomenon occurs. The high-TcT_{c} superconducting materials, like cuprates first discovered in 1986cuprate1 ; cuprate2 , are beyond the BCS mechanismcuprate3 ; cuprate4 ; cuprate5 . Specifically, it is established that the superconducting order parameter shows up as an energy gap in spectroscopic probes of quasiparticle energy spectrum. Whereas in cuprate superconductors, detected most directly by angle-resolved photo-emission spectroscopyARPES1 ; ARPES2 ; ARPES3 and scanning tunneling microscopeSTM1 ; STM2 ; STM3 ; STM4 , a finite normal-state gap (pseudogap) with dd-wave symmetry opens below a temperature TT^{*}, far above the superconducting transition temperature TcT_{c}.

The origin of the pseudogap and its relationship to superconductivity have attracted extensive experimental and theoretical interest. Transport properties like electrical resistivityresistivity1 ; resistivity2 , Hall conductivityHall1 ; Hall2 and Nernst coefficientNernst1 ; Nernst2 ; Nernst3 ; Nernst4 ; Nernst5 ; Nernst6 that are remarkably affected by the opening of the pseudogap have been used to detect TT^{*}. While among various theoretical models, significant superconducting phase fluctuation is suggested to be a possible attributionphase1 ; phase2 ; phase3 . Specifically, the generation of the superconducting order parameter breaks the continuous U(1)U(1) symmetry spontaneouslygi0 , and then, according to Goldstone theoremGm1 ; Gm2 , a collective gapless Bosonic excitation that describes the phase fluctuation of the order parameter emergesgi0 ; AK ; Ba0 ; Am0 . This excitation in conventional bulk superconductors is inactiveAK ; Ba0 ; Am0 , since its original low-energy spectrum is lifted to high-frequency plasma energy by Anderson-Higgs mechanismAHM , due to the coupling to longitudinal electromagnetic field and hence long-range Coulomb interactionAK ; Ba0 ; Am0 . Whereas in high-TcT_{c} superconductors, an active phase fluctuation becomes inevitable because of the low-dimensional layered structurecuprate4 ; cuprate5 . Then, in one view, the pseudogap state is a incoherent precursor of the superconducting statephase2 ; PCP1 ; PCP2 ; PCP3 ; PCP4 ; PCP5 , reflecting a state of preformed Cooper pairs without the phase coherence necessary to achieve superconductivity. The superconducting transition then occurs upon cooling below a lower temperature where the long-range phase coherence is established. TcT_{c} is therefore determined by the onset of the phase coherence rather than the formation of the Cooper pairsphase2 ; PCP1 ; PCP2 ; PCP3 ; PCP4 ; PCP5 .

The existence of the preformed Cooper pairing above TcT_{c} received a number of experimental supports by various approaches like diamagnetism probeTM1 ; TM2 ; TM3 ; TM4 , specific heatSH and paraconductivityPC1 measurements, Nernst effectNF1 , ultrafast pump-probe spectroscopiesPPS1 ; PPS2 and angle-resolved photo-emissionAF1 ; AF2 ; AF3 ; AF4 as well as detection of optical conductivity in infraredInfrad1 ; Infrad2 ; Infrad3 , microwaveMS1 ; MS2 ; MS3 and terahertzTHZ1 ; THZ2 ; THZ3 ; THZ4 frequency regimes. The most convincing evidence so far comes from a recent observation of the collective Higgs mode above TcT_{c}DHM1 ; DHM2 ; DHM3 , since this excitation describes the amplitude fluctuation of the order parameterAm0 and thus directly reflects the existence of the pairing. Nevertheless, most of these experimentsTM2 ; TM3 ; TM4 ; SH ; PC1 ; NF1 ; AF1 ; AF2 ; AF3 ; AF4 ; PPS1 ; PPS2 ; Infrad1 ; Infrad2 ; Infrad3 ; MS1 ; MS2 ; MS3 ; THZ1 ; THZ2 ; THZ3 ; THZ4 ; DHM3 realized that the regime of the significant superconducting phase fluctuation is in fact a relatively narrow one that tracks TcT_{c}, whereas the upper onset temperature of this regime TosT^{\rm os} lies well below the pseudogap temperature TT^{*} in underdoped regime and tends to coincide with TT^{*} in overdoped regimeTM4 ; SH ; PC1 ; NF1 ; AF4 ; MS1 ; MS2 ; MS3 ; THZ1 ; THZ2 ; THZ3 ; THZ4 ; DHM3 . Then, in an alternative view, the pseudogap state between TosT^{\rm os} and TT^{*} in underdoped regime, exhibiting various intertwined orders (nematicity, charge-density-wave and spin-density-wave orders), possibly represents another state of matter that competes/couples with superconductivitySTM1 ; AF4 ; NF1 . Whereas the precursor of the superconducting state with significant phase fluctuation (incoherent preformed Cooper pair) actually begins upon cooling at TosT^{\rm os}.

The phenomenological preformed Cooper-pair model with significant phase fluctuation is now widely accepted from experimental findings, but its microscopic theoretical description is not yet developed. Interestingly, from earlier thermodynamicNSB1 and recent opticalNSB2 measurements, substantial fraction of the uncondensed normal state, which exhibits TT-linear specific heat and Drude optical conductivity, persists down to temperatures far below TcT_{c}. This suggests that the phase fluctuation exists not only in the pseudogap state but also in the superconducting one. Therefore, to achieve superconductivity, the phase coherence must exceed a specific nonzero threshold. The theory of determining this threshold and in particular, TcT_{c}, requires coupled dual dynamics with different scales, i.e., in different Hilbert spaces: macroscopic phase-coherence dynamics and microscopic electronic fluid (including superfluid and normal fluidSNT ) dynamics. In the early-stage works, without distinguishing the amplitude and phase fluctuations, Ussishkin et al.GT1 ; GT2 applied the Gaussian approximationGT to calculate the contribution from the fluctuations of order parameter to Nernst signal. After that, within the path-integral method to study thermodynamics, Curty and BeckMT1 treated separately amplitude and phase fluctuations, and used the Monte Carlo procedure and Wolff algorithm for simulations, respectively. But in both approaches, the origin of the fluctuations is unclear. Recently, Li et al.disorder suggested that in cuprate superconductors, the disorder effect at low temperature plays the crucial role in determining the amplitude and phase fluctuations of the order parameter. By using tight-binding model with the random on-site potential (Anderson disorder)disorder , they numerically obtained granular superconducting islands to explain the substantial fraction of the normal state at T=0T=0, and then, suggested that the strong phase fluctuation can emerge in regions with small gap as a consequence. But this stationary-state calculation actually does not take account of the phase-coherence dynamics seriously, and is hard to extend for finite temperatures. Most importantly, the microscopic electronic fluid dynamics is decoupled and overlooked in all approaches above, inhibiting the deep insight into the key issue, separation between TcT_{c} and TosT^{\rm os}.

Actually, a microscopic gauge-invariant kinetic equation (GIKE) approach has been developed in conventional ss-wave superconductorsGIKE1 ; GIKE2 ; GIKE3 . This approach, as analytically demonstrated, not only involves both superfluid and normal-fluid dynamicsGIKE1 , but also is capable of formulating both phase and amplitude fluctuations of the order parameter (i.e., Nambu-Goldstone and Higgs modes)GIKE2 . The complete microscopic scattering is also constructed in GIKEGIKE1 ; GIKE3 . Very recently, this approach has also been extended into the dd-wave superconductors for studying the Higgs modesGIKE4 . It is therefore natural to further use this approach to study the phase-coherence dynamics and its influence on electronic fluid dynamics in cuprate superconductors, and then, elucidate the fundamental nature of the incoherent preformed Cooper pairs.

In this work, by using the GIKE approachGIKE1 ; GIKE2 ; GIKE3 ; GIKE4 , we construct the coupled dual dynamics of macroscopic phase coherence and microscopic electronic fluids (consisting of normal fluid and superfluid) in cuprate superconductors. Then, both the dephasing process of macroscopic superconducting phase coherence from long range to short range with the increase of temperature, and the influence of this dephasing on microscopic electronic fluids (including gap, densities of superfluid and normal fluid, and in particular, the transport property to determine superconducting transition temperature TcT_{c}) can be formulated.

Specifically, to develop the macroscopic phase-coherence dynamics, the equation of motion of the superconducting phase fluctuation, in which both disorder and long-range Coulomb interaction effects are considered, is derived analytically. We show that differing from the conventional bulk superconductors with inactive phase fluctuation due to Anderson-Higgs mechanismAHM ; AK ; Am0 ; Ba0 , the phase fluctuation in cuprate superconductors retains gapless energy spectrum after considering the long-range Coulomb interaction, thanks to the layered structurescuprate4 ; cuprate5 , and hence, is active. We also find that the superfluid density determines the superconducting phase stiffness in the phase-coherence dynamics, in consistency with the previous understanding in the literaturephase1 ; phase2 ; phase3 . The phase-coherence dynamics is therefore influenced by electronic fluids. The derived microscopic electronic-fluid dynamics includes two parts: the anomalous correlation to reflect pairing (i.e., distinguish superfluid and normal fluidSNT ) and determine gap and superfluid density; the microscopic scattering of the electronic fluids that is essential for studying the transport property and hence superconductivity. It is established that the spatial fluctuation of superconducting phase can generate a superconducting momentum 𝐩s{\bf p}_{s} gi0 ; Ba0 ; G1 , which drives the Doppler shift 𝐯𝐤𝐩s{\bf v}_{\bf k}\!\cdot\!{\bf p}_{s} in quasiparticle energy spectraFF4 ; FF5 ; FF6 ; GIKE1 , with 𝐯𝐤{\bf v}_{\bf k} being the group velocity. Then, both anomalous correlation and microscopic scattering in the electronic-fluid dynamics are affected by the superconducting phase fluctuation through this Doppler shift. Therefore, the macroscopic phase-coherence dynamics and microscopic electronic-fluid dynamics are mutually coupled.

It is noted that TosT^{\rm os} and TcT_{c} are determined by critical temperatures where the gap and resistivity vanish, respectively. Based on the developed dual dynamics, we present theoretical descriptions of the separation between TcT_{c} and TosT^{\rm os} as well as the emerged normal fluid in superconducting state (below TcT_{c}). The quantum effect of disorder, which excites a macroscopic inhomogeneous phase fluctuation through the Josephson effectJosephson , provides the key origin. This excited phase fluctuation drives the Doppler shift 𝐯𝐤𝐩s{\bf v}_{\bf k}\cdot{\bf p}_{s} mentioned above. Following the idea of Fulde-Ferrell-Larkin-Ovchinnikov (FFLO) state in conventional superconductors, the anomalous correlation vanishes in region where |𝐯𝐤𝐩s|>Δ𝐤|{\bf v}_{\bf k}\cdot{\bf p}_{s}|>\Delta_{\bf k} with Δ𝐤\Delta_{\bf k} being the superconducting gapFF1 ; FF2 ; FF7 ; FF8 ; FF9 ; GIKE1 . Particles then no longer participate in the pairing in this region and behave like the normal ones, leading to the emergence of normal fluidFF1 ; GIKE1 . Interestingly, with the phase fluctuation, we find that the condition of the unpairing region is always satisfied around nodal points in dd-wave superconductors. One therefore always finds a nonzero fraction of the normal fluid even at low temperature, in consistency with the experimentally observed substantial fraction of normal state at low temperatureNSB1 ; NSB2 .

Particles in regions with nonzero anomalous correlation contribute to the superconducting gap as superfluid. Moreover, we find that there exists the scattering between particles in pairing and unpairing regions in dd-wave superconductors, and this scattering behaves like the friction between superfluid and normal fluid. We prove analytically that due to this friction, part of superfluid becomes viscous with nonzero momentum-relaxation rate. Consequently, in addition to conventional non-viscous superfluid, there also exist normal fluid and viscous superfluid at small phase fluctuation in cuprate superconductors, similar to the three-fluid model proposed in our previous workGIKE1 in conventional superconductors which is caused by external electromagnetic field. A scheme to detect distinguish these three electronic fluids in cuprate superconductors is then proposed.

Particularly, as shown in Ref. GIKE1, , the increase of Doppler shift leads to the increases of normal fluid and viscous superfluid but the shrinkage of the non-viscous superfluid. When the non-viscous superfluid vanishes at large enough Doppler shift, one can find an exotic state with only normal fluid and viscous superfluid left, showing the nonzero resistivity as well as the finite gap from the viscous superfluid. Following the same idea, we demonstrate that by increasing the temperature in dd-wave cuprate superconductors, the suppressed superconducting gap and hence superfluid density weaken the phase stiffness, enhancing the phase fluctuation and hence Doppler shift. Once the phase fluctuation (i.e., temperature) exceeds the critical point, the non-viscous superfluid vanishes, leaving only normal fluid and viscous superfluid. The system then enters the pseudogap state with nonzero resistivity and finite gap due to the significant phase fluctuation. It is noted that in this circumstance, the viscous superfluid matches the description of the incoherent preformed Cooper pairs, as they both contribute to gap but experience the scattering. Whereas the existing normal fluid in our description implies the existence of normal particles in pseudogap state, which has been overlooked in previous preformed Cooper-pair model to describe pseudogap statephase2 ; PCP1 ; PCP2 ; PCP3 ; PCP4 ; PCP5 . With a further increase of temperature in pseudogap state, the viscous superfluid starts to shrink, until vanishes at TosT^{\rm os} where the gap is eventually destroyed.

To confirm the derivation from the GIKE approach, we also apply the standard path-integral approach to recover the equation of motion of the phase fluctuation as well as anomalous correlation, gap equation and superfluid density in the presence of the superconducting momentum, except the microscopic scattering of the electronic fluids which is hard to handle within the path-integral approach. A self-consistent numerical simulation by applying Anderson disorder is also addressed, to verify our theoretical description. Finally, we show that the developed dual dynamics can also be applied similarly to the low-dimensional disordered ss-wave superconductors.

II Coupled dual dynamics

In this section, from the rigorous analytic derivation within GIKE approach (refer to Sec. IV), we summarize the simplified results to present the coupled dual dynamics of macroscopic phase coherence and microscopic electronic fluids for a generalized order parameter:

Δ(x,x)=𝐤ei𝐤(𝐫𝐫)[Δ𝐤+δΔ𝐤(R)]eiδθ(R).\Delta(x,x^{\prime})=\sum_{\bf k}e^{i{\bf k}\cdot({\bf r-r^{\prime}})}[\Delta_{{\bf k}}+\delta\Delta_{\bf k}(R)]e^{i\delta\theta(R)}. (1)

Here, x=(x0,𝐫)x=(x_{0},{\bf r}) denotes the space-time vector; R=(x+x)/2=(t,𝐑)R={(x+x^{\prime})}/{2}=(t,{\bf R}) stands for the center-of-mass coordinate; Δ𝐤\Delta_{{\bf k}} denotes the equilibrium-state gap, independent of the center-of-mass coordinate due to the translational symmetry; δθ(R)\delta\theta(R) and δΔ𝐤(R)\delta\Delta_{\bf k}(R) represent the phase and amplitude fluctuations, respectively.

Specifically, the macroscopic phase-coherence dynamics involves the generation of the phase fluctuation δθ(R)\delta\theta(R), which along 𝐞ϕ{\bf e}_{\phi} direction is determined by

(psϕ)2=qq2[Uq𝐞ϕUq𝐞ϕ4CωN+2nB(ωN)+12ωN(1+2DVqD)].(p_{s}^{\phi})^{2}=\!\!\sum_{q}q^{2}\Big{[}\frac{U_{q{\bf e}_{\phi}}U_{-q{\bf e}_{\phi}}}{4C\omega_{N}}+\frac{2n_{B}(\omega_{N})\!+\!1}{2\omega_{N}}\Big{(}\frac{1\!+\!2DV_{q}}{D}\Big{)}\Big{]}. (2)

Here, psϕ𝐞ϕp_{s}^{\phi}{\bf e}_{\phi} denotes the generated superconducting momentum 𝐩s=𝐑δθ(R)/2{\bf p}_{s}={\bm{\nabla}}_{\bf R}\delta\theta(R)/2 by phase fluctuation along 𝐞ϕ{\bf e}_{\phi} direction; U𝐪U_{\bf q} represents the Fourier component of disorder-induced local electric potential; nB(x)n_{B}(x) stands for the Bose distribution; ωN=ωp2+nsq2/(2Dm)\omega_{N}=\sqrt{\omega_{p}^{2}+n_{s}q^{2}/(2Dm)} denotes the energy spectrum of the phase fluctuation, where ωp=q2Vqns/m\omega_{p}=\sqrt{q^{2}V_{q}n_{s}/m} represents the plasma frequency and nsn_{s} stands for the superfluid density, with VqV_{q} being the Coulomb potential; mm and DD denote the effective mass and density of states of carriers, respectively; 1/C1/C is the normalized factor in frequency-momentum space.

The microscopic electronic-fluid dynamics includes: the anomalous correlation F𝐤F_{\bf k}, used to characterize the pairing and determine gap Δ𝐤\Delta_{\bf k} equation and superfluid density nsn_{s}; the microscopic momentum-relaxation rate Γ𝐤\Gamma_{\bf k} in superfluid. The expressions of these quantities are written as

F𝐤\displaystyle F_{\bf k} =\displaystyle= f(E𝐤+)f(E𝐤)2E𝐤,\displaystyle\frac{f(E_{\bf k}^{+})-f(E_{\bf k}^{-})}{2E_{\bf k}}, (3)
Δ𝐤\displaystyle\Delta_{\bf k} =\displaystyle= 𝐤g𝐤𝐤Δ𝐤F𝐤,\displaystyle-{\sum_{\bf k^{\prime}}}^{\prime}g_{\bf kk^{\prime}}{\Delta_{\bf k^{\prime}}}F_{\bf k^{\prime}}, (4)
ns\displaystyle n_{s} =\displaystyle= kF2m𝐤Δ𝐤2E𝐤E𝐤F𝐤,\displaystyle\frac{k_{F}^{2}}{m}{\sum_{\bf k}}^{\prime}\frac{\Delta^{2}_{\bf k}}{E_{\bf k}}\partial_{E_{\bf k}}{F_{\bf k}}, (5)
Γ𝐤\displaystyle\Gamma_{\bf k}\! =\displaystyle= 𝐤η=±|M𝐤𝐤|2δ(E𝐤ηE𝐤η).\displaystyle\!-\!{\sum_{{\bf k^{\prime}}\eta=\pm}}^{\prime}|M_{\bf kk^{\prime}}|^{2}\delta(E^{\eta}_{\bf k}-E^{-\eta}_{\bf k^{\prime}}). (6)

Here, E𝐤±E^{\pm}_{\bf k} denotes the quasi-electron and quasi-hole energies; g𝐤𝐤g_{\bf kk^{\prime}} represents the pairing potential, in which we approximately taking k=k=kFk=k^{\prime}=k_{F} around the Fermi surface so that the gap Δ𝐤\Delta_{\bf k} only has angular dependence of the momentum; 𝐤\sum_{\bf k}^{\prime} here and hereafter stands for the summation restricted in the spherical shell (ξ𝐤ωD\xi_{\bf k}\leq\omega_{D}) with ωD\omega_{D} being the cutoff frequency, following the BCS theoryBCS ; G1 ; M𝐤𝐤M_{\bf kk^{\prime}} denotes the effective matrix element of the electron-impurity scattering. It is established that in the presence of the superconducting momentum 𝐩s{\bf p}_{s}, the quasiparticle energy is tilted as E𝐤±=(𝐯𝐤𝐩s)±E𝐤E_{\bf k}^{\pm}=({\bf v}_{\bf k}\cdot{\bf p}_{s})\pm{E_{\bf k}}, where E𝐤=ξ𝐤2+Δ𝐤2E_{\bf k}=\sqrt{\xi_{\bf k}^{2}+\Delta_{\bf k}^{2}} is the original Bogoliubov quasiparticle energy and 𝐯𝐤𝐩s{\bf v}_{\bf k}\cdot{\bf p}_{s} denotes the Doppler shiftFF1 ; FF4 ; FF5 ; FF6 ; GIKE1 , with the group velocity 𝐯𝐤=𝐤ξ𝐤{\bf v}_{\bf k}=\partial_{\bf k}\xi_{\bf k}. In the present work, we approximately take the parabolic spectrum, i.e., ξ𝐤=𝐤2/(2m)μ\xi_{\bf k}={\bf k}^{2}/(2m)-\mu with μ\mu being the chemical potential.

Particularly, it is noted that the generation of the phase fluctuation in Eq. (2) is coupled with electronic fluids through the superfluid density nsn_{s} in the energy spectrum ωN\omega_{N}, whereas the phase fluctuation is involved in the electronic-fluid dynamics in Eqs. (3)-(6) by Doppler shift. The macroscopic phase-coherence dynamics and microscopic electronic-fluid dynamics are therefore mutually coupled. For phase fluctuation along a certain direction, by self-consistently solving Eqs. (2) and (4) as well as (5), one can uniquely determine the superconducting gap Δ𝐤\Delta_{\bf k}, superfluid density nsn_{s} and superconducting phase fluctuation psp_{s}, and then, the microscopic momentum-relaxation rate Γ𝐤\Gamma_{\bf k} and various physical quantities are obtained. Nevertheless, in realistic situation, there exist phase fluctuations along all directions. Hence, the experimentally observed quantity is a statistical average of phase fluctuations in all directions. The expected value of the physical quantity XX therefore reads

X=12π𝑑ϕX(ϕ),\langle{X}\rangle=\frac{1}{2\pi}{{\int}d\phi}X(\phi), (7)

with X(ϕ)X(\phi) being the solved XX for phase fluctuation along 𝐞ϕ{\bf e}_{\phi} direction.

Equations (2)-(6) then provide an efficient and simplified way to understand the superconductivity properties in cuprate and disordered ss-wave superconductors with significant phase fluctuation.

III Application to dd-wave cuprate superconductors

In this section, we apply the developed dual dynamics into the dd-wave cuprate superconductors. Without losing generality, we choose dx2y2d_{x^{2}-y^{2}}-wave order parameter for analysis, i.e., Δ𝐤=Δcos(ζθ𝐤)\Delta_{{\bf k}}=\Delta\cos(\zeta\theta_{\bf k}) with ζ=2\zeta=2, and the pairing potential g𝐤𝐤gcos[ζ(θ𝐤θ𝐤)]g_{\bf kk^{\prime}}{\approx}g\cos[\zeta(\theta_{\bf k}-\theta_{\bf k^{\prime}})] as a consequence of the translational and time-reversal symmetriesSM1 .

III.1 Theoretical description of preformed Cooper pairs

In this part, by simply performing an analytic analysis on the coupled dual dynamics of macroscopic phase coherence and microscopic electronic fluids, we present the physical pictures of the preformed Cooper pair model in pseudogap state (i.e., separation between TcT_{c} and TosT^{\rm os}) and the emerged normal fluid in superconducting one in dd-wave cuprate superconductors.

We start with the influence on the microscopic electronic fluids from the macroscopic phase fluctuation. Specifically, we first focus on the anomalous correlation. At low temperature, for an excited superconducting fluctuation 𝐩s=psϕ𝐞ϕ{\bf p}_{s}=p_{s}^{\phi}{\bf e}_{\phi}, it is noted that considering the fact that E𝐤+E𝐤E_{\bf k}^{+}~{}{\geq}~{}E_{\bf k}^{-}, the anomalous correlation F𝐤F_{\bf k} in Eq. (3) vanishes in regions with |𝐯𝐤𝐩s|>E𝐤|{\bf v}_{\bf k}\cdot{\bf p}_{s}|>E_{\bf k}, where the quasielectron energy E𝐤+=(𝐯𝐤𝐩s)+E𝐤<0E_{\bf k}^{+}=({\bf v}_{\bf k}\cdot{\bf p}_{s})+{E_{\bf k}}<0 or quasihole energy E𝐤=(𝐯𝐤𝐩s)E𝐤>0E_{\bf k}^{-}=({\bf v}_{\bf k}\cdot{\bf p}_{s})-{E_{\bf k}}>0, whereas F𝐤F_{\bf k} in regions with |𝐯𝐤𝐩s|<E𝐤|{\bf v}_{\bf k}\cdot{\bf p}_{s}|<E_{\bf k} is always finite. It has been established in the literatureG1 ; FF1 ; GIKE1 that the nonzero anomalous correlation directly reflects the existence of the pairing as the characteristic quantity. Regions with nonzero and vanishing anomalous correlation are therefore referred to as the pairing and unpairing regionsFF1 ; FF2 ; FF7 ; FF8 ; FF9 ; GIKE1 , respectively. Particularly, particles in the unpairing region no longer participate in the pairing and behave like the normal ones, leading to the emergence of the normal fluidGIKE1 . Interestingly, it is noted that due to the anisotropy of dd-wave gap, the condition |𝐯𝐤𝐩s|>E𝐤|{\bf v}_{\bf k}\cdot{\bf p}_{s}|>E_{\bf k} of the unpairing region is always satisfied around the nodal points, irrelevant of the direction of the phase fluctuation. One therefore always finds a nonzero fraction of the normal fluid in cuprate superconductors even at low temperature, in consistency with the experimentally observed substantial fraction of the normal state far below TcT_{c}NSB1 ; NSB2 . This is very different from the ss-wave superconductors, where the emergence of the normal fluid requires 𝐩s>Δ/vF{\bf p}_{s}>\Delta/v_{F} by the unpairing-region condition |𝐯𝐤𝐩s|>E𝐤|{\bf v}_{\bf k}\cdot{\bf p}_{s}|>E_{\bf k}, leading to a threshold to realize normal fluid as revealed in our previous workGIKE1 .

Particles in the pairing regions with nonzero anomalous correlation contribute to the gap and superfluid density as superfluid. Moreover, according to the microscopic scattering of the momentum relaxation in superfluid [Eq. (6)], one can further divide the pairing region into two parts: viscous one, in which Γ𝐤0\Gamma_{\bf k}\neq 0; non-viscous one where Γ𝐤=0\Gamma_{\bf k}=0. Specifically, in Eq. (6), if 𝐤{\bf k} particle lies in the pairing region, one has E𝐤+>0E_{\bf k}^{+}>0 and E𝐤<0E_{\bf k}^{-}<0. Then, once the energy conservation is satisfied to give rise to nonzero momentum-relaxation rate Γ𝐤\Gamma_{\bf k} of 𝐤{\bf k} particle, one finds E𝐤>0E_{\bf k^{\prime}}^{-}>0 by δ(E𝐤+E𝐤)\delta(E_{\bf k}^{+}-E_{\bf k^{\prime}}^{-}) or E𝐤+<0E_{\bf k^{\prime}}^{+}<0 by δ(E𝐤E𝐤+)\delta(E_{\bf k}^{-}-E_{\bf k^{\prime}}^{+}), and hence, 𝐤{\bf k^{\prime}} particle lies in the unpairing region. This scattering between particles in pairing and unpairing regions, behaves like the friction between superfluid and normal fluid, leading to the viscous superfluid with finite momentum-relaxation rate Γ𝐤\Gamma_{\bf k}. But if the energy conservation can not be satisfied for any 𝐤{\bf k}^{\prime}, the 𝐤{\bf k} particle is free from the momentum-relaxation scattering, and one therefore gets the non-viscous superfluid with zero momentum-relaxation rate.

Refer to caption
Figure 1: (Color online) Schematic showing the division in momentum space at different phase fluctuations 𝐩s{\bf p}_{s}. In the figure, Ξ=ξ𝐤+2ωD\Xi=\xi_{\bf k}+2\omega_{D} and hence the original spherical shell is characterized by ωDΞ3ωD\omega_{D}\leq\Xi\leq 3\omega_{D}. The spherical shell is divided into three parts: unpairing region where anomalous correlation F𝐤=0F_{\bf k}=0, non-viscous pairing region with F𝐤0F_{\bf k}\neq 0 and momentum-relaxation rate Γ𝐤=0\Gamma_{\bf k}=0, viscous pairing region with both F𝐤F_{\bf k} and Γ𝐤\Gamma_{\bf k} being finite, represented by yellow, blue and orange regions, respectively. (a)-(d): 𝐩s{\bf p}_{s} along antinodal point (ϕ=0\phi=0); (e)-(h): 𝐩s{\bf p}_{s} along nodal point (ϕ=π/4\phi=\pi/4). Δ=0.5ωD\Delta=0.5\omega_{D} and T=0T=0.

Therefore, there exist three electronic fluids at small phase fluctuation in dd-wave cuprate superconductors: normal fluid, viscous and non-viscous superfluids, similar to the three-fluid model proposed in our previous workGIKE1 in conventional superconductors which is caused by external electromagnetic field. Consequently, in analogy to Eq. (5), we define the non-viscous and viscous superfluid densities as

nns=kF2m𝐤PnvΔ𝐤2E𝐤E𝐤F𝐤,n_{\rm ns}=\frac{k_{F}^{2}}{m}{\sum_{{\bf k}\in{P_{\rm nv}}}}^{\prime}\frac{\Delta^{2}_{\bf k}}{E_{\bf k}}\partial_{E_{\bf k}}F_{\bf k}, (8)

and

nvs=kF2m𝐤PvΔ𝐤2E𝐤E𝐤F𝐤,n_{\rm vs}=\frac{k_{F}^{2}}{m}{\sum_{{\bf k}\in{P_{\rm v}}}}^{\prime}\frac{\Delta^{2}_{\bf k}}{E_{\bf k}}\partial_{E_{\bf k}}F_{\bf k}, (9)

respectively. Whereas the normal-fluid density reads

nn=𝐤U1.n_{n}={\sum_{{\bf k}\in{U}}}^{\prime}1. (10)

Here, 𝐤Pnv{\bf k}\in{P_{\rm nv}}, 𝐤Pv{\bf k}\in{P_{\rm v}} and 𝐤U{\bf k}\in{U} denote the summations restricted in the non-viscous (F𝐤Pnv0F_{{\bf k}\in{P_{\rm nv}}}\neq 0 and Γ𝐤Pnv=0\Gamma_{{\bf k}\in{P_{\rm nv}}}=0) and viscous (F𝐤Pv0F_{{\bf k}\in{P_{\rm v}}}\neq 0 and Γ𝐤Pv0\Gamma_{{\bf k}\in{P_{\rm v}}}\neq 0) pairing regions as well as unpairing regions (F𝐤U=0F_{{\bf k}\in{U}}=0), respectively. Considering the short-circuit effect, if there exists the non-viscous pairing region in the spherical shell of the momentum space, i.e., non-viscous superfluid density, the system lies in the superconducting state, showing the zero-resistance (superconductivity) phenomenon.

It has been shown in Ref. GIKE1, that in conventional superconductors, the increase of Doppler shift leads to the increases of normal fluid and viscous superfluid but the shrinkage of the non-viscous superfluid. When the non-viscous superfluid vanishes at large enough Doppler shift, one can find an exotic state with only normal fluid and viscous superfluid left, showing the nonzero resistivity as well as the finite gap from the viscous superfluid. Following the same idea, in the spherical shell of the momentum space, a schematic illustration for the division of the unpairing, non-viscous and viscous pairing regions in dd-wave superconductors is plotted in Fig. (1). As seen from Figs. 1(a)-(d) [or Figs. 1(e)-(h)], the increase of ps{p}_{s} first gradually enlarges the unpairing region (yellow regions) and hence viscous pairing region (orange regions) in the spherical shell, leading to the shrinkage of the non-viscous pairing region (blue regions). Particularly, once the non-viscous pairing region vanishes when the phase fluctuation exceeds a critical point, as shown in Fig. 1(d) [Fig. 1(g)], only viscous pairing and unpairing regions are left. Then, the system enters the pseudogap state with nonzero resistivity and finite gap due to the significant phase fluctuation. Correspondingly, in this circumstance, the viscous superfluid matches the description of the incoherent preformed Cooper pairs, as they both contribute to gap but experience the scattering. Whereas the existing normal fluid in our description implies the existence of normal particles in pseudogap state, which has been overlooked in previous preformed Cooper-pair model to describe the pseudogap statephase2 ; PCP1 ; PCP2 ; PCP3 ; PCP4 ; PCP5 . By further increasing ps{p}_{s} in pseudogap state, the viscous pairing region starts to shrink, as shown in Figs. 1(g) and (h), until vanishes. The system with only unpairing region left eventually enters the normal state.

In addition, in comparison between Figs. 1(a)-(d) and Figs. 1(e)-(h) for phase fluctuations along antinodal and nodal points, respectively, one finds that the influence of the phase fluctuation is anisotropic with respect to its direction. This anisotropy in fact arises from the anisotropy of the dd-wave gap, similar to the observed anisotropic transport properties in the previous experimentRo . Specifically, at the same condition, in comparison to 𝐩s{\bf p}_{s} along the antinodal point, 𝐩s{\bf p}_{s} along the nodal point is easier to satisfy |𝐯𝐤𝐩s|>E𝐤|{\bf v}_{\bf k}\cdot{\bf p}_{s}|>E_{\bf k}, and hence, causes the larger unpairing region, making it earlier to enter the pseudogap and then normal states with the gradual increase of psp_{s}, as shown by comparison between Figs. 1(c) and (g). Nevertheless, the judgment to enter the pseudogap and normal states in realistic situation is given by the vanishing non-viscous and viscous superfluid densities after the statistical average of the phase fluctuations in all directions, respectively.

We next discuss the influence on the macroscopic phase fluctuation from the microscopic superfluid. For the source terms on the right-hand side of Eq. (2), the first term arises from the disorder-induced local electric potential, which excites the inhomogeneous phase fluctuation through Josephson effectJosephson . The second term comes from the zero-point energy and thermal excitation of the Bosonic phase fluctuation. Particularly, differing from the conventional bulk superconductors where the high-energy plasma frequency and hence ωNωp\omega_{N}\approx\omega_{p} causes the inactive phase fluctuation (Anderson-Higgs mechanism)AHM ; AK ; Am0 ; Ba0 , in consideration of the two-dimensional Vq=2πe2/(qε)V_{q}=2\pi{e^{2}}/({q\varepsilon}), ωN\omega_{N} retains gapless in cuprate superconductors thanks to the layered structurescuprate4 ; cuprate5 , leading to an active phase fluctuation.

For the convenience of analysis, by using the fact that ωN=ωp2+nsq2/(2Dm)=ns/mq2Vq+q2/(2D)\omega_{N}=\sqrt{\omega_{p}^{2}+n_{s}q^{2}/(2Dm)}=\sqrt{n_{s}/m}\sqrt{q^{2}V_{q}+q^{2}/(2D)}, we transform Eq. (2) into an equivalent form:

(psϕ)2ns/m=𝒯ϕ,(p_{s}^{\phi})^{2}\sqrt{n_{s}/m}={\cal{T}_{\phi}}, (11)

with 𝒯ϕ=q[q2Uq𝐞ϕUq𝐞ϕ/C+2q2/Dq+4q2n(ωN)/Dq4q2Vq+q2/(2D)]{\cal{T}}_{\phi}=\sum_{q}\big{[}\frac{q^{2}U_{q{\bf e}_{\phi}}U_{-q{\bf e}_{\phi}}/C+2q^{2}/D_{q}+4q^{2}n(\omega_{N})/D_{q}}{4\sqrt{q^{2}V_{q}+q^{2}/(2D)}}\big{]}. It is noted that the leading contribution of n(ωN)n(\omega_{N}) lies in the long-wave regime, leading to the marginal role of q2n(ωN)q^{2}n(\omega_{N}) in 𝒯ϕ{\cal{T}_{\phi}}. In this circumstance, 𝒯ϕ{\cal{T}_{\phi}} contributed by disorder and zero-point-energy effects is independent of the phase fluctuation and electronic fluids as well as the temperature, and hence, acts as a structure factor in the phase-coherence dynamics in Eq. (11). Consequently, from Eq. (11), the factor ns/mn_{s}/m (superfluid density over effective mass) plays a crucial role in determining the superconducting phase stiffness, i.e., an enhancement of ns/m{n_{s}/m} suppresses the phase fluctuation, as the early experiments in cuprate superconductors realizedphase1 ; phase2 ; phase3 .

In fact, the main thermal effect in the phase-coherence dynamics arises from the coupling to the electronic fluids. Specifically, with the increase of the temperature, the suppressed order parameter and hence the superfluid density [Eq. (5)] enlarge the phase fluctuation by weakening the phase stiffness. Once the temperature exceeds a critical point to generate the significant phase fluctuation, the system enters the pseudogap state with only viscous superfluid and normal fluid left. With further increase of temperature, the order parameter and hence superfluid density tend to vanish, and then, psp_{s} moves towards an infinitely large value, causing the unpairing region (i.e., normal fluid) left alone. Particularly, it is noted that the increased ps=𝐑δθp_{s}=\nabla_{\bf R}\delta\theta with temperature directly suggests a dephasing process of superconducting phase coherence from long range to short range.

III.2 Numerical Simulation

In this part, we perform a full numerical simulation to verify the theoretical analysis in the previous subsection. For the quantum effect of disorder on the macroscopic phase-coherence dynamics in Eq. (2), we apply the method of the Anderson disorder, which is introduced by generating random on-site potential at the structural sites, i.e., U(𝐑i)=γiWU({\bf R}_{i})=\gamma_{i}W, where γi\gamma_{i} denotes on-site random number with uniform probability in range (1,1)(-1,1) and WW represents the Anderson-disorder strength. Considering the crystal structure of cuprate superconductors, we take a finite square lattice system of large size 400×400400\times 400 with the periodic boundary condition. Then, in each random configuration, U𝐪=𝐑iei𝐪𝐑iU(𝐑i)U_{\bf q}=\sum_{{\bf R}_{i}}e^{i{\bf q}\cdot{\bf R}_{i}}U({\bf R}_{i}) in Eq. (2).

In each random configuration, we self-consistently solve the phase fluctuation [Eq. (2)] and gap equation [Eq. (4)] as well as superfluid density [Eq. (5)] for phase fluctuation along a certain direction. Then, with the solved 𝐩s{\bf p}_{s} and Δ\Delta, by determining the unpairing region as well as non-viscous and viscous pairing regions according to the anomalous correlation [Eq. (3)] and microscopic momentum-relaxation rate [Eq. (6)], we can calculate the densities of the non-viscous [Eq. (8)] and viscous [Eq. (9)] superfluids and normal-fluid [Eq. (10)]. After that, by varying the direction of the phase fluctuation, we calculate the various quantities for phase fluctuation along each direction, and then, take the statistical average of the phase fluctuations in all directions [Eq. (7)]. Finally, the gap Δ\langle\Delta\rangle as well as densities of the normal-fluid nn\langle{n_{n}}\rangle, viscous nvs\langle{n_{\rm vs}}\rangle and non-viscous nns\langle{n_{\rm ns}}\rangle superfluids are averaged over 400400 random configurations for convergence. Moreover, in the simulation, the gap at zero temperature in the absence of the phase fluctuation is taken to be Δ0=22\Delta_{0}=22~{}meV, which is close to the maximum value of the observed gap in YBa2Cu3O7-xYpa . C=2ωD/ξc2C=2\omega_{D}/\xi^{2}_{c} with ξc=vF/Δ0\xi_{c}=\hbar{v_{F}}/\Delta_{0} being the coherence lengthG1 . The Fermi energy EF=220E_{F}=220~{}meV and meff=1.9m0m_{\rm eff}=1.9m_{0}Ypa1 ; Ypa2 , with m0m_{0} representing the free-electron mass. The specific cutoff frequency ωD\omega_{D} requires the microscopic pairing mechanism, which still remains an open question in the literature. We chose ωD=Δ0\omega_{D}=\Delta_{0}, and then, due to the phase fluctuation, one has ΔωD\Delta\ll\omega_{D} in the simulation. With determined Δ0\Delta_{0} and ωD\omega_{D}, the pairing potential gg is determined by gap equation [Eq. (4)] at zero temperature in the absence of the phase fluctuation. Furthermore, our model focuses on the long-wave (i.e., low-frequency) regime (refer to Appendix B), and hence, we introduce a cutoff qcq_{c} in the summation of qq in Eq. (2) to approximately guarantee that the energy spectrum of the phase fluctuation does not enter the Bogoliubov quasiparticle continuum along antinodal points, i.e., ωN(qc)<2Δ0\omega_{N}(q_{c})<2\Delta_{0}.

Refer to caption
Figure 2: (Color online) The temperature and Anderson-disorder strength dependence of the expected value of the gap Δ\langle\Delta\rangle as well as densities of the non-viscous superfluid nns\langle{n_{\rm ns}}\rangle , viscous superfluid nvs\langle{n_{\rm vs}}\rangle and normal fluid nn\langle{n_{n}}\rangle after the statistical average of the phase fluctuations in all directions from the full and self-consistent numerical simulation. (a)-(d): phase diagram; (e)-(h): temperature dependence at different Anderson-disorder strengths; (i)-(l): Anderson-disorder strength dependence at different temperatures. In the figure, TBT_{B} denotes the critical temperature where gap vanishes in the absence of the phase fluctuation, i.e., BCS critical temperature of two-dimensional dd-wave superconductors; ns0n_{s0} represents the superfluid density at zero temperature in the absence of the phase fluctuation, and nn0n_{n0} denotes the normal-fluid density above TosT^{\rm os}; W0=0.5Δ0W_{0}=0.5\Delta_{0}.

III.2.1 Phase Diagram

The temperature and Anderson-disorder strength dependence of the expected values of the gap as well as densities of the normal-fluid, viscous and non-viscous superfluids after the statistical average of the phase fluctuations in all directions are plotted in Fig. 2.

We first discuss the gap. As seen from the phase diagram in Fig. 2(a), Δ\langle\Delta\rangle decreases with the increase of the Anderson-disorder strength WW or temperature TT, exhibiting a half-dome behavior. Specifically, with the increase of temperature, as shown in Fig. 2(e), the gap shows the BCS-like behavior at small WW (red solid curve) as it should be, but tends to become linear decrease when WW is enhanced (ochre dotted and purple chain curves). Particularly, the change into the linear decrease starts at high-TT regime, and moves towards low-TT one with the enhancement of WW. This temperature dependence from our numerical simulation qualitatively agrees with the previous experimental observationSTM4 where the temperature dependence of the gap shows a small platform in low-TT regime but linearly decreases in the remaining temperature regime. The deviation of the temperature dependence from the BCS-like behavior here arises from the phase fluctuation. By increasing temperature, the suppressed gap and hence superfluid density (especially in high-TT regime) enhances the phase fluctuation and hence Doppler shift, and then, the shrinkage of the pairing region as a result feeds back to suppress gap, speeding up the gap falling in comparison to BCS-like behavior. This effect by phase fluctuation can be more directly seen from the WW dependence in Fig. 2(i). By increasing WW from zero, Δ\langle\Delta\rangle first changes marginally, and then, exhibits an exponential decay at relatively large WW where phase fluctuation becomes important.

By comparing Figs. 2(a) and (b), one finds that the normalized non-viscous superfluid density nns/ns0\langle{n_{\rm ns}}\rangle/n_{s0} also exhibits a half-dome behavior in phase diagram, but is smaller than that of the normalized gap Δ/Δ0\langle\Delta\rangle/\Delta_{0}. By further comparing Figs. 2(e) and (f) as well as Figs. 2(i) and (j), with the increase of TT or WW, nns/ns0\langle{n_{\rm ns}}\rangle/{n_{s0}} exhibits a faster decrease than Δ/Δ0\langle\Delta\rangle/\Delta_{0}. This is because that in comparison to nns\langle{n_{\rm ns}}\rangle contributed by non-viscous pairing region alone, both viscous and non-viscous pairing regions contribute to the gap. Therefore, with the enhancement of psp_{s} by increasing TT or WW, the faster shrinkage of the non-viscous pairing region than the entire pairing region (as shown in Fig. 1) leads to the faster decrease of nns/ns0\langle{n_{\rm ns}}\rangle/n_{s0} than Δ/Δ0\langle\Delta\rangle/\Delta_{0}. We point out that the non-viscous superfluid density can be directly measured by detecting the 1/ω1/\omega-like divergent behavior in the imaginary part of the optical conductivity [i.e., σ2(ω)=e2nnsmω\sigma_{2}(\omega)=\frac{e^{2}{\langle}n_{\rm ns}{\rangle}}{m\omega}] at low-frequency regimeInfrad2 ; Infrad3 ; MS1 ; MS3 ; THZ3 ; THZ4 ; NSB2 .

The normal-fluid density in Figs. 2(d), (h) and (l) shows compensatory behavior in comparison to the non-viscous superfluid density in corresponding Figs. 2(b), (f) and (j), as it should be according to the analytic analysis in Sec. III.1. Particularly, the compensatory behavior between the superfluid and normal-fluid densities in the temperature dependence has been observed in the previous experimentNSB2 . Moreover, it is noted in Fig. 2(l) that at relatively large WW (i.e., phase fluctuation), the nonzero normal-fluid density at zero temperature (red solid curve) agrees with our analysis in Sec. II, in consistency with the previous experimental observationNSB1 ; NSB2 .

The viscous superfluid density nvs\langle{n_{\rm vs}}\rangle exhibits very differently, showing a flame-like behavior in phase diagram in Fig. 2(c). From Eq. (9), there exist two opposite effects on nvs\langle{n_{\rm vs}}\rangle by increasing phase fluctuation: (i) at small phase fluctuation, through the friction, the increase of the unpairing region enlarges the viscous pairing one and hence nvs\langle{n_{\rm vs}}\rangle; (ii) the suppressed gap directly reduces nvs\langle{n_{\rm vs}}\rangle. As shown in Fig. 2(k), with the increase of WW, effects (i) and (ii) dominate at small and large WW, leading to the increase and decrease of nvs{\langle}n_{\rm vs}\rangle, respectively. A peak is therefore observed, and the peak position moves towards small WW with the increase of temperature as effect (ii) is enhanced. As for the temperature dependence in Fig. 2(g), at small WW (red solid curve), with the increase of temperature, a platform due to the competition between effects (i) and (ii) is found in low-TT regime, whereas effect (ii) dominates around the critical temperature where the gap changes dramatically, leading to the decrease of nvs\langle{n_{\rm vs}}\rangle. At large W>0.5W>0.5, effect (ii) dominates, causing the decrease of nvs\langle{n_{\rm vs}}\rangle.

Refer to caption
Figure 3: (Color online) Critical temperatures TcT_{c} and TosT^{\rm os} as well as the difference between TosT^{\rm os} and TcT_{c} versus Anderson-disorder strength. In the figure, TBT_{B} denotes the critical temperature where gap vanishes in the absence of the phase fluctuation, i.e., BCS critical temperature of two-dimensional dd-wave superconductors. W0=0.5Δ0W_{0}=0.5\Delta_{0}.

Now we propose a scheme to detect the specific viscous superfluid density nvs\langle{n_{\rm vs}}\rangle. Following our previous work in conventional superconductorsGIKE1 , based on three-fluid model, in the linear optical response, the optical conductivity in low-frequency regime reads

σ(ω)=e2nnsimω+e2nvsm(iω+γvs)+e2nnm(iω+γn),\sigma(\omega)=\frac{e^{2}\langle{n_{\rm ns}}\rangle}{im\omega}+\frac{e^{2}\langle{n_{\rm vs}}\rangle}{m(i\omega+\gamma_{\rm vs})}+\frac{e^{2}\langle{n_{n}}\rangle}{m(i\omega+\gamma_{n})}, (12)

with γvs\gamma_{\rm vs} and γn\gamma_{n} being the relaxation rates of viscous superfluid and normal fluid, respectively. Here, the normal fluid (the third term) exhibits the well-known Drude-model behavior; the non-viscous superfluid (the first term) is free from the resistance; due to the friction between superfluid and normal fluid, the viscous superfluid (the second term) also shows the Drude-model behavior. Therefore, as mentioned above, nns\langle{n_{\rm ns}}\rangle can be directly measured by detecting the 1/ω1/\omega-like divergent behavior in the imaginary part of the optical conductivity at low-frequency regimeInfrad2 ; Infrad3 ; MS1 ; MS3 ; THZ3 ; THZ4 ; NSB2 . Whereas from Eq. (12), one has

0+𝑑ωσ1(ω)=2e2πm(nvs+nn),\int^{\infty}_{0+}d\omega\sigma_{1}(\omega)=\frac{2e^{2}}{{\pi}m}(\langle{n_{\rm vs}}\rangle+\langle{n_{n}}\rangle), (13)

with σ1(ω)\sigma_{1}(\omega) being the real part of the optical conductivity. Therefore, by detecting the area under σ1(ω)\sigma_{1}(\omega) curve in frequency dependenceNSB2 , nvs+nn\langle{n_{\rm vs}}\rangle+\langle{n_{n}}\rangle is obtained.

Furthermore, in the magnetic response, based on three-fluid model, the excited current is written asGIKE1

𝐣=e2𝐀nnsme2𝐀nvsm(1ξcl)+e2𝐀nnmξcl,{\bf j}=-\frac{e^{2}{\bf A}\langle{n_{\rm ns}}\rangle}{m}-\frac{e^{2}{\bf A}\langle{n_{\rm vs}}\rangle}{m}\Big{(}1-\frac{\xi_{c}}{l}\Big{)}+\frac{e^{2}{\bf A}\langle{n_{\rm n}}\rangle}{m}\frac{\xi_{c}}{l}, (14)

with ξc\xi_{c} and ll being the coherence length and mean-free path, respectively. Here, the first and second terms come from the excited supercurrent in non-viscous and viscous superfluids by the Meissner effectG1 , respectively. It is noted that the viscous superfluid experiences the resistance due to the friction between superfluid and normal fluid. The magnetic flux can not drive the normal-fluid current directly, but through the friction drag with the superfluid current, a normal-fluid current (the third term, proportional to 1/l1/l) is excited. From Eq. (14), the magnetic penetration depth λ\lambda is determined by

1λ2=e2(nns+nvs)me2(nvs+nn)mξcl.\frac{1}{\lambda^{2}}=\frac{e^{2}(\langle{n_{\rm ns}}\rangle+\langle{n_{\rm vs}}\rangle)}{m}-\frac{e^{2}(\langle{n_{\rm vs}}\rangle+\langle{n_{\rm n}}\rangle)}{m}\frac{\xi_{c}}{l}. (15)

Consequently, from above equation, with the established ξc\xi_{c} and ll as well as the obtained densities nvs+nn\langle{n_{\rm vs}}\rangle+\langle{n_{n}}\rangle and nns\langle{n_{\rm ns}}\rangle from the optical detection as mentioned above, one can determine the viscous superfluid density nvs\langle{n_{\rm vs}}\rangle and also the normal-fluid density nn\langle{n_{n}}\rangle by measuring the penetration depth λ\lambda.

III.2.2 Separation between TcT_{c} and TosT^{\rm os}

We next discuss the separation between TcT_{c} and TosT^{\rm os}. In the full numerical simulation, TcT_{c} and TosT^{\rm os} are chosen at the critical temperatures where the normalized non-viscous superfluid density nns/ns0\langle{n_{\rm ns}}\rangle/{n_{s0}} and normalized gap Δ/Δ0\langle\Delta\rangle/\Delta_{0} vanish, respectively, and are plotted in Fig. 3. As mentioned above, with the increase of temperature, the decrease of nns/ns0\langle{n_{\rm ns}}\rangle/{n_{s0}} is faster than Δ/Δ0\langle\Delta\rangle/\Delta_{0}, due to the additional contribution in gap from the viscous pairing region. TcT_{c} is therefore smaller than TosT^{\rm os}, as shown in Fig. 3. Whereas when Tos>T>TcT^{\rm os}>T>T_{c}, since the non-viscous superfluid vanishes, the system with only normal fluid and viscous superfluid left enters the pseudogap state, showing both nonzero resistivity and finite gap.

To enlarge the separation between TcT_{c} and TosT^{\rm os} (i.e., to lower TcT_{c}), one needs to enhance WW in order to generate significant phase fluctuation at low temperature, so that the system enters the pseudogap state at lower temperature. As seen from Fig. 3, the separation between TcT_{c} and TosT^{\rm os} at small WW is marginal. By increasing WW, both TosT^{\rm os} (ochre solid curve with triangles) and TcT_{c} (blue solid curve with squares) decrease due to the suppressed gap [Fig. 2(i)] and non-viscous superfluid density [Fig. 2(j)], respectively. Whereas the separation between TcT_{c} and TosT^{\rm os} (red solid curve with circles) is enlarged until Tc=0T_{c}=0, thanks to the enhanced phase fluctuation at low temperature. Particularly, with the large enough phase fluctuation at 1.9<W/W0<2.31.9<W/W_{0}<2.3, one can find the emergence of the pseudogap state even at zero temperature.

In our simulation with the superconducting-material parameters of YBa2Cu3O7-x (Δ0=22\Delta_{0}=22~{}meV and hence TB=110T_{B}=110~{}K)Ypa , TosT^{\rm os} tracks TcT_{c} closely in Fig. 3 (3535 K above TcT_{c} at most when Tc>16T_{c}>16~{}K and 5050~{}K above TcT_{c} at most when Tc=0T_{c}=0), and lies well below the experimental pseudogap temperature TT^{*} in underdoped regimeNF1 , in consistency with the current experimental findings and understandingTM2 ; TM3 ; TM4 ; SH ; PC1 ; NF1 ; AF1 ; AF2 ; AF3 ; AF4 ; PPS1 ; PPS2 ; Infrad1 ; Infrad2 ; Infrad3 ; MS1 ; MS2 ; MS3 ; THZ1 ; THZ2 ; THZ3 ; THZ4 ; DHM3 mentioned in the introduction. This close track in fact arises from the fast shrinkage of the remaining viscous pairing region in the pseudogap state with the increase of temperature and hence phase fluctuation, as shown in Figs. 1(g) and (h), leading to a strong suppression on the gap. In other words, once in the pseudogap state, because of the significant phase fluctuation and hence remarkable Doppler shift, the gap can not survive far above TcT_{c} in our simplified model.

Refer to caption
Figure 4: (Color online) Dependence of the gap and densities of the normal fluid and superfluid on the phase-fluctuation direction after the average over 400400 random configurations. W/W0=0.72W/W_{0}=0.72 and T/TB=0.18T/T_{B}=0.18. In the figure, ns0n_{s0} represents the superfluid density at zero temperature in the absence of the phase fluctuation, and nn0n_{n0} denotes the normal-fluid density above TosT^{\rm os}. TBT_{B} denotes the critical temperature where gap vanishes in the absence of the phase fluctuation, i.e., BCS critical temperature of the two-dimensional dd-wave superconductors. W0=0.5Δ0W_{0}=0.5\Delta_{0}.

III.2.3 Anisotropic influence of phase fluctuation

It is noted that in our numerical calculation, after the statistical average of the phase fluctuations in all directions, there is no current excitation, i.e., 𝐣=0\langle{\bf j}\rangle=0, since 𝐩s=0\langle{\bf p}_{s}\rangle=0 as a consequence of the inversion symmetry in our system. Whereas the nonzero gap Δ\langle{\Delta}\rangle and densities of the normal-fluid nn\langle{n_{n}}\rangle, viscous nvs\langle{n_{\rm vs}}\rangle and non-viscous nns\langle{n_{\rm ns}}\rangle superfluids are due to the fact that the anomalous correlation [Eq. (3)] and hence the gap equation [Eq. (4)] and superfluid density [Eq. (5)], as well as the microscopic momentum-relaxation rate Γ𝐤\Gamma_{\bf k} [Eq. (6)] are even functions with respect to 𝐩s{\bf p}_{s}, whereas ps20\langle{p_{s}^{2}}\rangle\neq 0 by Eq. (2).

In Fig. 4, after the average over 400400 random configurations, we plot the dependence of the gap and densities of the normal fluid and superfluid on the phase-fluctuation direction. As seen from the figure, all three quantities exhibit C4C_{4}-symmetric dependence on the phase-fluctuation direction. The superfluid density shows similar dependence to the gap, but exhibits compensatory behavior in comparison with the normal-fluid density, similar to the results in Sec. III.2.1. The C4C_{4}-symmetric dependence with respect to the phase-fluctuation direction arises from the dd-wave gap, since the phase fluctuation along the nodal point is easier to generate the unpairing region (normal fluid) and reduce the pairing region (i.e., suppress the gap and hence superfluid density) than that along the antinodal point, as analyzed in Sec. II.

For experimental probes that are related to the intrinsic characters of system, the superconducting momentum 𝐩s{\bf p}_{s} only arises from the phase fluctuation, and then, the C4C_{4}-symmetric anisotropies in Fig. 4 does not manifest directly, since there exist phase fluctuations along all directions and the observed quantity is a statistical average of phase fluctuations in all directions. However, one can apply the external stationary magnetic vector potential 𝐀{\bf A} or inject the background supercurrent 𝐈{\bf I} to generate the extrinsic superconducting momentum 𝐩sext=e𝐀{\bf p}_{s}^{\rm ext}=-e{\bf A} or 𝐩sext=𝐈m/(ens){\bf p}_{s}^{\rm ext}={\bf I}m/(en_{s}). Consequently, the total superconducting momentum, including the intrinsic part from phase fluctuation and extrinsic one, is enhanced along the direction of 𝐩sext{\bf p}_{s}^{\rm ext}. In this circumstance, by varying the direction of the vector potential or injected supercurrent, all C4C_{4}-symmetric anisotropies in Fig. 4, i.e, the anisotropic influence of 𝐩s{\bf p}_{s} mentioned in Sec. II can be observed.

IV Derivation of dual dynamics

In this section, we present the rigorous derivation of the coupled dual dynamics from the GIKE approach. The recovery from the path-integral approach is also addressed to confirm our derivation.

IV.1 Hamiltonian

We begin with a generalized Bogoliubov-de Gennes (BdG) HamiltonianG1 ; ER :

H0=𝑑𝐫𝑑𝐫ψ(x)[ξ𝐩^τ3δ(xx)+Δ^(x,x)]ψ(x).H_{0}=\int{d{\bf r}}{d{\bf r^{\prime}}}\psi^{\dagger}(x)[\xi_{\hat{\bf p}}\tau_{3}\delta(x-x^{\prime})+{\hat{\Delta}}(x,x^{\prime})]\psi(x^{\prime}). (16)

Here, ψ(x)=[ψ(x),ψ(x)]\psi(x)=[\psi_{\uparrow}(x),\psi^{\dagger}_{\downarrow}(x)] represents the field operator in Nambu space; the momentum operator 𝐩^=i{\hat{\bf p}}=-i\hbar{\bm{\nabla}}; Δ^(x,x)=Δ(x,x)τ++Δ(x,x)τ{\hat{\Delta}}(x,x^{\prime})=\Delta(x,x^{\prime})\tau_{+}+\Delta^{*}(x,x^{\prime})\tau_{-}; τi\tau_{i} are the Pauli matrices in Nambu space.

It is established that the phase fluctuation in Eq. (16) can be effectively removed by a unitary transformation:

ψ(x)eiτ3δθ(R)/2ψ(x),\psi(x){\rightarrow}e^{i\tau_{3}\delta\theta(R)/2}\psi(x), (17)

and then, one hasgi0 ; AK

H0\displaystyle H_{0} =\displaystyle= d𝐫d𝐫ψ(x){(ξ𝐩^+𝐩sτ3+tδθ/2)τ3δ(xx)\displaystyle\int{d{\bf r}}{d{\bf r^{\prime}}}\psi^{\dagger}(x)\Big{\{}(\xi_{{\hat{\bf p}}+{\bf p}_{s}\tau_{3}}+\partial_{t}\delta\theta/2)\tau_{3}\delta(x-x^{\prime}) (18)
+𝐤ei𝐤(𝐫𝐫)[Δ𝐤+δΔ𝐤(R)]τ1}ψ(x).\displaystyle\mbox{}+\sum_{\bf k}e^{i{\bf k}\cdot({\bf r-r^{\prime}})}[\Delta_{{\bf k}}+\delta\Delta_{\bf k}(R)]\tau_{1}\Big{\}}\psi(x^{\prime}).

with the superconducting momentum 𝐩s=𝐑δθ(R)/2{\bf p}_{s}={\bm{\nabla}}_{\bf R}\delta\theta(R)/2. Then, it is clearly seen that as a consequence of the spontaneous breaking of the U(1)U(1) symmetry by order parametergi0 ; AK , the superconducting phase fluctuation provides an effective electromagnetic potential eAμeff=(tδθ/2,𝐩s)eA^{\rm eff}_{\mu}=(\partial_{t}\delta\theta/2,-{\bf p}_{s}), in consistency with the gauge structure in superconductors first revealed by Nambugi0 :

eAμ\displaystyle eA_{\mu} \displaystyle\rightarrow eAμμχ(R),\displaystyle eA_{\mu}-\partial_{\mu}\chi(R), (19)
δθ(R)\displaystyle\delta\theta(R) \displaystyle\rightarrow δθ(R)+2χ(R).\displaystyle\delta\theta(R)+2\chi(R). (20)

Here, the standard electromagnetic potential eAμ=(eϕ,e𝐀)eA_{\mu}=(e\phi,e{\bf A}) and μ=(t,𝐑)\partial_{\mu}=(\partial_{t},-{\bm{\nabla}}_{\bf R}).

The electron-electron Coulomb interaction HeeH_{\rm ee} and electron-impurity interaction HeiH_{\rm ei} are written asG1

Hee=12𝑑𝐫𝑑𝐫V(xx)[ψ(x)τ3ψ(x)][ψ(x)τ3ψ(x)],\displaystyle H_{\rm ee}=\frac{1}{2}{\int}d{\bf r}d{\bf r^{\prime}}V(x-x^{\prime})[\psi^{\dagger}({x})\tau_{3}\psi({x})][\psi^{\dagger}({x}^{\prime})\tau_{3}\psi({x}^{\prime})],
(21)
Hei=𝑑𝐫ψ(x)U(x)τ3ψ(x).\displaystyle H_{\rm ei}=\int{d{\bf r}}\psi^{\dagger}(x)U(x)\tau_{3}\psi(x). (22)

Here, V(xx)V(x-x^{\prime}) and U(x)U(x) denote the Coulomb and impurity potentials in space-time coordinate. We have kept the time dependence of U(x)U(x) for general situation.

IV.2 GIKE approach

To derive the macroscopic phase-coherence dynamics and microscopic electronic superfluid dynamics, we first use the GIKE approachGIKE1 ; GIKE2 ; GIKE3 ; GIKE4 . In this microscopic approach, the response of the system is described by the density of matrix ρ𝐤\rho_{\bf k} in Nambu space. Considering the fluctuations, the density of matrix reads

ρ𝐤=ρ𝐤(0)+δρ𝐤(R).\rho_{\bf k}=\rho_{\bf k}^{(0)}+\delta\rho_{\bf k}(R). (23)

Here, δρ𝐤\delta\rho_{\bf k} stands for the non-equilibrium response from the equilibrium part ρ𝐤(0)\rho_{\bf k}^{(0)}, which is written asGIKE1

ρ𝐤(0)=F𝐤(E𝐤τ0+ξ𝐤τ3+Δ𝐤τ1).\rho_{\bf k}^{(0)}=F_{\bf k}(E_{\bf k}\tau_{0}+{\xi_{\bf k}}\tau_{3}+{\Delta_{\bf k}}\tau_{1}). (24)

To determine the density of matrix, one needs to solve δρ𝐤\delta\rho_{\bf k} from the GIKEGIKE1 ; GIKE2 ; GIKE3 ; GIKE4 :

tρ𝐤+i[(ξk+μeff+ps22m)τ3+Δ𝐤τ1,ρ𝐤][i𝐑28mτ3,ρ𝐤]\displaystyle\partial_{t}\rho_{\bf k}\!+\!i\left[\left(\xi_{k}\!+\!\mu_{\rm eff}\!+\!\frac{p^{2}_{s}}{2m}\right)\tau_{3}\!+\!\Delta_{\bf k}\tau_{1},\rho_{\bf k}\right]\!-\!\Big{[}\frac{i\nabla^{2}_{\bf R}}{8m}\tau_{3},\rho_{\bf k}\Big{]}
+12{e𝐄τ3(𝐑+2i𝐩sτ3)Δ𝐤τ1,𝐤ρ𝐤}\displaystyle\mbox{}\!+\!\frac{1}{2}\left\{e{\bf E}\tau_{3}\!-\!({\bm{\nabla}}_{\bf R}\!+\!2i{\bf p}_{s}\tau_{3})\Delta_{\bf k}\tau_{1},\partial_{\bf k}\rho_{\bf k}\right\}\!
i8[(𝐑+2i𝐩sτ3)(𝐑+2i𝐩sτ3)Δ𝐤τ1,𝐤𝐤ρ𝐤]\displaystyle\mbox{}\!-\!\frac{i}{8}\left[({\bm{\nabla}}_{\bf R}\!+\!2i{\bf p}_{s}\tau_{3})({\bm{\nabla}}_{\bf R}\!+\!2i{\bf p}_{s}\tau_{3})\Delta_{\bf k}\tau_{1},\partial_{\bf k}\partial_{\bf k}\rho_{\bf k}\right]
+{𝐤2mτ3,𝐑ρ𝐤}+[𝐑𝐩s4mτ3,τ3ρ𝐤]=tρ𝐤|scat,\displaystyle\mbox{}\!+\!\Big{\{}\frac{\bf k}{2m}\tau_{3},{\bm{\nabla}}_{\bf R}\rho_{\bf k}\Big{\}}\!+\!\Big{[}\frac{{\nabla_{\bf R}}\circ{\bf p}_{s}}{4m}\tau_{3},\tau_{3}\rho_{\bf k}\Big{]}\!=\!\partial_{t}\rho_{\bf k}\Big{|}_{\rm scat}, (25)

where we have effectively removed the phase fluctuation from the order parameter through the unitary transformation in Eq. (17). Here, [A,B]=ABBA[A,B]=AB-BA and {A,B}=AB+BA\{A,B\}=AB+BA represent the commutator and anti-commutator, respectively; 𝐑𝐩s=2𝐩s𝐑+𝐑𝐩s{\nabla_{\bf R}}\circ{\bf p}_{s}=2{\bf p}_{s}\cdot{\bm{\nabla}}_{\bf R}+{\bm{\nabla}}_{\bf R}\!\cdot\!{\bf p}_{s}; the effective field μeff(R)=tδθ(R)/2+μH(R)+U(R)\mu_{\rm eff}(R)={\partial_{t}\delta\theta(R)}/{2}+\mu_{H}(R)+U(R) with the Hartree field written as

μH(R)=RV(RR)n(R).\mu_{H}(R)=\sum_{R^{\prime}}V(R-R^{\prime})n({R^{\prime}}). (26)

The induced electric field e𝐄=𝐑[μH(R)+U(R)]e{\bf E}=-{\bm{\nabla}}_{\bf R}[\mu_{H}(R)+U(R)].

The density nn and current 𝐣{\bf j} are given byGIKE1 ; GIKE2 ; GIKE3 ; GIKE4

n=𝐤(1+2ρ𝐤3),\displaystyle n=\sum_{\bf k}(1+2\rho_{{\bf k}3}), (27)
𝐣=𝐤(e𝐤mρ𝐤0),\displaystyle{\bf j}=\sum_{\bf k}\big{(}\frac{e{\bf k}}{m}\rho_{{\bf k}0}\big{)}, (28)

respectively. The equation of the order parameter reads

(Δ𝐤+δΔ𝐤Δ𝐤+δΔ𝐤)=𝐤g𝐤𝐤(ρ𝐤+ρ𝐤),\left(\begin{array}[]{cc}&\Delta_{\bf k}\!+\!\delta\Delta_{\bf k}\\ \Delta_{\bf k}\!+\!\delta\Delta_{\bf k}&\end{array}\right)=-{\sum_{\bf k^{\prime}}}^{\prime}g_{\bf kk^{\prime}}\left(\begin{array}[]{cc}&\rho_{{\bf k}^{\prime}+}\\ \rho_{{\bf k}^{\prime}-}&\end{array}\right), (29)

whose components are given by

𝐤g𝐤𝐤ρ𝐤1\displaystyle{\sum_{\bf k^{\prime}}}^{\prime}g_{\bf kk^{\prime}}\rho_{{\bf k^{\prime}}1} =\displaystyle= Δ𝐤,\displaystyle-\Delta_{\bf k}, (30)
𝐤g𝐤𝐤ρ𝐤2\displaystyle{\sum_{\bf k^{\prime}}}^{\prime}g_{\bf kk^{\prime}}\rho_{{\bf k^{\prime}}2} =\displaystyle= 0.\displaystyle 0. (31)

Here, ρ𝐤i\rho_{{\bf k}i} stands for the τi\tau_{i} component of ρ𝐤=i=04ρ𝐤iτi\rho_{{\bf k}}=\sum_{i=0}^{4}\rho_{{\bf k}i}\tau_{i}; It is noted that Eq. (30) gives the gap equation, whereas Eq. (31) determines the phase fluctuation as revealed in our previous work GIKE2 .

The impurity scattering tρ𝐤|scat\partial_{t}\rho_{\bf k}\big{|}_{\rm scat} is derived based on the generalized Kadanoff-Baym ansatz with the random-phase and Markovian approximationsGQ2 ; GQ3 . The specific impurity-scattering term readsGIKE1

tρ𝐤|scat\displaystyle\partial_{t}\rho_{\bf k}|_{\rm scat} =\displaystyle\!=\! niπ𝐤η=±|U𝐤𝐤s|2δ(EkEk)(τ3Γ𝐤ητ3Γ𝐤ηρ𝐤\displaystyle\!-n_{i}\pi\sum_{{\bf k^{\prime}}\eta=\pm}|U^{s}_{\bf k\!-\!k^{\prime}}|^{2}\delta(E_{k^{\prime}}\!-\!E_{k})(\tau_{3}\Gamma^{\eta}_{\bf k^{\prime}}\tau_{3}\Gamma^{\eta}_{\bf k}\rho_{\bf k} (32)
τ3ρ𝐤Γ𝐤ητ3Γ𝐤η+h.c.).\displaystyle\mbox{}\!-\!\tau_{3}\rho_{\bf k^{\prime}}\Gamma^{\eta}_{\bf k^{\prime}}\tau_{3}\Gamma^{\eta}_{\bf k}\!+\!h.c.).

Here, nin_{i} and U𝐤𝐤sU^{s}_{\bf kk^{\prime}} stand for the impurity density and matrix element of the electron-impurity scattering, respectively; the projection operators Γ𝐤±=𝒰𝐤(1±τ3)𝒰𝐤/2\Gamma^{\pm}_{\bf k}=\mathscr{U}_{\bf k}^{\dagger}(1\pm\tau_{3})\mathscr{U}_{\bf k}/{2} with 𝒰𝐤=u𝐤τ0iv𝐤τ2\mathscr{U}_{\bf k}=u_{\bf k}\tau_{0}-iv_{\bf k}\tau_{2} being the unitary transformation matrix from the particle space to the quasiparticle one. u𝐤=1/2+ξ𝐤/(2E𝐤)u_{\bf k}=\sqrt{1/2+\xi_{\bf k}/(2E_{\bf k})} and v𝐤=1/2ξ𝐤/(2E𝐤)v_{\bf k}=\sqrt{1/2-\xi_{\bf k}/(2E_{\bf k})}.

To solve GIKE, by expanding δρ𝐤=δρ𝐤(1)+δρ𝐤(2)\delta\rho_{\bf k}=\delta\rho_{\bf k}^{(1)}+\delta\rho_{\bf k}^{(2)} with δρ𝐤(1)\delta\rho_{\bf k}^{(1)} and δρ𝐤(2)\delta\rho_{\bf k}^{(2)} standing for the linear and second-order terms of the nonequilibrium response, the GIKE becomes a chain of equations, whose first order only involves δρ𝐤(1)\delta\rho^{(1)}_{\bf k} and ρ𝐤(0)\rho^{(0)}_{\bf k} and second order involves δρ𝐤(2)\delta\rho^{(2)}_{\bf k}, δρ𝐤(1)\delta\rho^{(1)}_{\bf k} and ρ𝐤(0)\rho^{(0)}_{\bf k}. Then, one can solve δρ𝐤(1)\delta\rho_{\bf k}^{(1)} and δρ𝐤(2)\delta\rho_{\bf k}^{(2)} in sequence, whose specific lengthy expressions are presented in Appendix A.

IV.2.1 Gap equation and superfluid density

We first derive the gap equation and supercurrent. Substituting the solved ρ𝐤1=ρ𝐤1(0)+δρ𝐤1\rho_{{\bf k}1}=\rho_{{\bf k}1}^{(0)}+\delta\rho_{{\bf k}1} into Eq. (30), with ρ𝐤1(0)=Δ𝐤F𝐤\rho_{{\bf k}1}^{(0)}=\Delta_{\bf k}F_{\bf k} from Eq. (24), one can directly obtain the gap equation in Eq. (4), and find a vanishing amplitude fluctuation δΔ𝐤(R)=0\delta\Delta_{\bf k}(R)=0 as δρ𝐤1\delta\rho_{{\bf k}1} makes no contribution (refer to Appendix B). This is because of the particle-hole symmetry in our derivation. With this symmetry, the amplitude and phase fluctuations represent mutually orthogonal excitationsAm0 , and hence, are decoupledsymmetry . Moreover, with the particle-hole symmetry, the disorder-induced local potential can not excite the amplitude fluctuation, since the charge-amplitude correlation vanishes according to recent symmetry analysissymmetry .

For the excited phase fluctuation 𝐩s{\bf p}_{s}, the solved ρ𝐤0\rho_{{\bf k}0} in clean case reads (refer to Appendix B)

ρ𝐤0=(𝐯𝐤𝐩s)Δ𝐤2E𝐤E𝐤F𝐤.\rho_{{\bf k}0}=({\bf v}_{\bf k}\cdot{\bf p}_{s})\frac{\Delta^{2}_{\bf k}}{E_{\bf k}}\partial_{E_{\bf k}}F_{\bf k}. (33)

Then, from Eq. (28), the generated supercurrent is given by

𝐣=e𝐩𝐬mkF2m𝐤Δ𝐤2E𝐤E𝐤F𝐤=ensm𝐩s,{\bf j}=\frac{e{\bf p_{s}}}{m}\frac{k_{F}^{2}}{m}{\sum_{\bf k}}^{\prime}\frac{\Delta^{2}_{\bf k}}{E_{\bf k}}\partial_{E_{\bf k}}{F_{\bf k}}=\frac{{e}n_{s}}{m}{\bf p}_{s}, (34)

from which one obtains the superfluid density in Eq. (5).

IV.2.2 Scattering of momentum relaxation

For the scattering part, we focus on the momentum (current) relaxation. Then, according to Eq. (28), one only needs to keep the τ0\tau_{0} components of ρ𝐤\rho_{\bf k} and tρ𝐤|scat\partial_{t}\rho_{\bf k}\big{|}_{\rm scat}, and hence, Eq. (32) becomes

tρ𝐤|scatmr\displaystyle\!\!\!\!\partial_{t}\rho_{\bf k}\Big{|}^{\rm mr}_{\rm scat}\!\! =\displaystyle= 2niπ𝐤ηη|U𝐤𝐤s|2Tr(τ3Γ𝐤ητ3Γ𝐤η)(ρ𝐤0ρ𝐤0)δ(E𝐤ηE𝐤η)\displaystyle\!\!-2n_{i}\pi\sum_{{\bf k^{\prime}}\eta\eta^{\prime}}|U^{s}_{\bf kk^{\prime}}|^{2}{\rm Tr}(\tau_{3}\Gamma^{\eta}_{\bf k}\tau_{3}\Gamma^{\eta^{\prime}}_{\bf k^{\prime}})(\rho_{{\bf k}0}-\rho_{{\bf k}^{\prime}0})\delta(E^{\eta}_{\bf k}-E^{\eta^{\prime}}_{\bf k^{\prime}}) (35)
=\displaystyle= 2niπ𝐤|U𝐤𝐤s|2(ρ𝐤0ρ𝐤0)[e𝐤𝐤δ(E𝐤+E𝐤+)+e𝐤𝐤δ(E𝐤E𝐤)+e𝐤𝐤+δ(E𝐤+E𝐤)+e𝐤𝐤+δ(E𝐤E𝐤+)],\displaystyle\!\!-2n_{i}\pi\sum_{{\bf k^{\prime}}}|U^{s}_{\bf kk^{\prime}}|^{2}(\rho_{{\bf k}0}\!-\!\rho_{{\bf k}^{\prime}0})[{e^{-}_{\bf kk^{\prime}}}\delta(E^{+}_{\bf k}\!-\!E^{+}_{\bf k^{\prime}})\!+\!{e^{-}_{\bf kk^{\prime}}}\delta(E^{-}_{\bf k}\!-\!E^{-}_{\bf k^{\prime}})\!+\!{e^{+}_{\bf kk^{\prime}}}\delta(E^{+}_{\bf k}\!-\!E^{-}_{\bf k^{\prime}})\!+\!{e^{+}_{\bf kk^{\prime}}}\delta(E^{-}_{\bf k}\!-\!E^{+}_{\bf k^{\prime}})],~{}~{}~{}

where e𝐤𝐤±=12(1±Δ𝐤Δ𝐤ξ𝐤ξ𝐤E𝐤E𝐤)e_{\bf kk^{\prime}}^{\pm}=\frac{1}{2}(1\pm\frac{\Delta_{\bf k}\Delta_{\bf k^{\prime}}-\xi_{\bf k}\xi_{\bf k^{\prime}}}{E_{\bf k}E_{\bf k^{\prime}}}). It is noted that the term e𝐤𝐤(ρ𝐤0ρ𝐤0)[δ(E𝐤+E𝐤+)+δ(E𝐤E𝐤)]{e^{-}_{\bf kk^{\prime}}}(\rho_{{\bf k}0}\!-\!\rho_{{\bf k}^{\prime}0})[\delta(E^{+}_{\bf k}\!-\!E^{+}_{\bf k^{\prime}})\!+\!\delta(E^{-}_{\bf k}\!-\!E^{-}_{\bf k^{\prime}})] vanishes around the Fermi surface. Consequently, one has

tρ𝐤|scatmr=𝐤η=±|M𝐤𝐤|2(ρ𝐤0ρ𝐤0)δ(E𝐤ηE𝐤η),\partial_{t}\rho_{\bf k}\Big{|}^{\rm mr}_{\rm scat}\!=\!-\!{\sum_{{\bf k^{\prime}}\eta=\pm}}^{\prime}|M_{\bf kk^{\prime}}|^{2}(\rho_{{\bf k}0}-\rho_{{\bf k}^{\prime}0})\delta(E^{\eta}_{\bf k}-E^{-\eta}_{\bf k^{\prime}}), (36)

with |M𝐤𝐤|2=2niπ|U𝐤𝐤s|2e𝐤𝐤+|M_{\bf kk^{\prime}}|^{2}=2n_{i}\pi|U^{s}_{\bf kk^{\prime}}|^{2}e^{+}_{\bf kk^{\prime}}.

We next focus on the momentum-relaxation rate of the 𝐤{\bf k} particle in superfluid. In Eq. (36), if 𝐤{\bf k} particle lies in the pairing region, one has 𝐤{\bf k}^{\prime} particle lying in the unpairing region, as analyzed in Sec. II. Then, considering the situation with the drive effect from superconducting momentum 𝐩s{\bf p}_{s}, from Eq. (33), one has vanishing ρ𝐤0\rho_{{\bf k}^{\prime}0} and finite ρ𝐤0\rho_{{\bf k}0}. Consequently, Eq. (36) becomes

tρ𝐤|scatmr=Γ𝐤ρ𝐤0,\partial_{t}\rho_{\bf k}\Big{|}^{\rm mr}_{\rm scat}\!=\!-\Gamma_{\bf k}\rho_{{\bf k}0}, (37)

with the momentum-relaxation rate of superfluid given by Eq. (6). For weak external probe that is related to the intrinsic character of system, one can use Γ𝐤\Gamma_{\bf k} to discuss the superconductivity phenomenon, with 𝐩s{\bf p}_{s} in Γ𝐤\Gamma_{\bf k} arising from the phase fluctuation.

IV.2.3 Phase-coherence dynamics

We next construct the macroscopic phase-coherence dynamics by deriving the equation of motion of the phase fluctuation. We neglect the mutual interaction between phase fluctuations by only keeping the linear order of phase fluctuation in its equation of motion. Then, in center-of-mass frequency and momentum space [(t,𝐑)(ω,𝐪)][(t,{\bf R})\rightarrow(\omega,{\bf q})], substituting the solved ρ𝐤2\rho_{{\bf k}2} into Eq. (31), the equation of motion of the phase fluctuation is written as (refer to Appendix B)

2iω(iωδθ𝐪/2+μH)D+i𝐪𝐩sns/m=2iωDU𝐪.2i\omega({i\omega\delta\theta_{\bf q}}/{2}+\mu_{H})D+i{\bf q}\cdot{\bf p}_{s}n_{s}/m=-2i\omega{DU_{\bf q}}. (38)

As seen from the left-hand side of above equation, without the Hartree field μH\mu_{H}, one immediately finds a gapless energy spectrum of the phase fluctuation with the linear dispersion, in consistency with the Goldstone theoremGm1 ; Gm2 by the spontaneous breaking of continuous U(1)U(1) symmetry in superconductorsgi0 .

Particularly, substituting the solution of ρ𝐤3\rho_{{\bf k}3} into Eq. (27), the density fluctuation reads (refer to Appendix B)

δn=2Dμeff.\delta{n}=-2D\mu_{\rm eff}. (39)

Then, it is noted that from the expressions of the density [Eq. (39)] and current [Eq. (34)], the equation of motion of the phase fluctuation in Eq. (38) is a direct consequence of the charge conservation tδn+𝐑𝐣=0\partial_{t}\delta{n}+{\bm{\nabla}_{\bf R}}\cdot{\bf j}=0. The charge conservation in the gauge-invariant kinetic theory is naturalGIKE4 , as it has been proved long time ago by Nambu through the generalized Ward identity that the gauge invariance in the superconducting states is equivalent to the charge conservation gi0 .

From Eq. (39), the Hartree field [Eq. (26)] is therefore written as μH=2VqDμeff\mu_{H}=-2V_{q}D\mu_{\rm eff}. Then, Eq. (38) becomes

Dq(ωp2+nsq22Dmω2)δθ𝐪2=iωU𝐪Dq,D_{q}\Big{(}\omega_{p}^{2}+\frac{n_{s}q^{2}}{2Dm}-\omega^{2}\Big{)}\frac{\delta\theta_{\bf q}}{2}=-i\omega{U_{\bf q}}D_{q}, (40)

with Dq=D/(1+2DVq)D_{q}=D/(1+2DV_{q}). As seen from the left-hand side of above equation, when the long-range Coulomb interaction is included, the original energy spectrum of the phase fluctuation is altered as ωN=ωp2+nsq2/(2Dm)\omega_{N}=\sqrt{\omega_{p}^{2}+n_{s}q^{2}/(2Dm)}. The right-hand side of above equation represents the source term from impurity potential. It is noted that additional source terms emerge on the right-hand side of Eq. (40) if other quantum disorder/fluctuation effects that generate electric potential or couple to phase fluctuation are considered, so UqU_{q} here can represent an effective electric potential by related quantum disorder/fluctuation effects. The source term from the disorder-induced local electric potential excites the macroscopic inhomogeneous phase fluctuation through the Josephson effect, since without the long-range Coulomb interaction and kinetic term nsq2/(2Dm){n_{s}q^{2}}/{(2Dm)}, Eq. (40) reduces to tδθ/2=U(R)\partial_{t}\delta\theta/2=-U(R) in space-time coordinate, same as the Josephson effectJosephson .

From the equation of motion in Eq. (40), in principle, one can directly solve the generated phase fluctuation δθ(R)\delta\theta(R) from disorder effect, and then, calculate the induced superconducting momentum 𝐩s(R)=𝐑δθ(R)/2{\bf p}_{s}(R)=\nabla_{\bf R}\delta\theta(R)/2 and hence Doppler shift, in order to further formulate the influence of the phase fluctuation on electronic fluids at each 𝐑{\bf R}. Nevertheless, for the experimental detections that usually are long-wave measurement, one only needs to consider the long-wave component of the Doppler shift effect, which leads to a homogeneous influence on the electronic fluids. In this circumstance, we apply a simplified way by using the equation of motion in Eq. (40) to construct the action of the phase fluctuation:

S=𝑑t𝐪Dq(|tθ2|2ωN2|θ𝐪2|2+U𝐪tθ𝐪2+U𝐪tθ𝐪2).S\!=\!\int{dt}\!\sum_{\bf q}D_{q}\Big{(}\Big{|}\frac{\partial_{t}\theta}{2}\Big{|}^{2}\!-\!\omega^{2}_{N}\Big{|}\frac{\theta_{\bf q}}{2}\Big{|}^{2}\!+\!U^{*}_{\bf q}\frac{\partial_{t}\theta_{\bf q}}{2}\!+\!U_{\bf q}\frac{\partial_{t}\theta_{\bf q}^{*}}{2}\Big{)}. (41)

Then, on one hand, one can directly use above action to derive the expected value of the generated superconducting momentum from disorder effect. On the other hand, by mapping the action in Eq. (41) into the imaginary-time one 𝒮\mathcal{S}, the expected value of the generated superconducting momentum from the thermal excitation can also be determined.

Consequently, considering the anisotropy in dd-wave superconductors, the generated superconducting momentum by phase fluctuation along 𝐞ϕ{\bf e}_{\phi} direction reads

(psϕ)2\displaystyle(p_{s}^{\phi})^{2}\!\! =\displaystyle\!=\! qq2[|δθq𝐞ϕ(t+0+)2δθq𝐞ϕ(t)2eiS|\displaystyle\!\!\sum_{q}q^{2}\Big{[}\Big{\langle}\Big{|}\frac{\delta\theta^{*}_{q{\bf e}_{\phi}}(t+0^{+})}{2}\frac{\delta\theta_{q{\bf e}_{\phi}}(t)}{2}e^{iS}\Big{|}\Big{\rangle} (42)
+|δθq𝐞ϕ(τ)2δθq𝐞ϕ(τ)2e𝒮|]\displaystyle\mbox{}+\Big{\langle}\Big{|}\frac{\delta\theta^{*}_{q{\bf e}_{\phi}}(\tau)}{2}\frac{\delta\theta_{q{\bf e}_{\phi}}(\tau)}{2}e^{-\mathcal{S}}\Big{|}\Big{\rangle}\Big{]}
=\displaystyle\!=\! qq2[1Z0DθDθδθq𝐞ϕ(t+0+)2δθq𝐞ϕ(t)2eiS\displaystyle\!\!\sum_{q}q^{2}\Big{[}\frac{1}{Z_{0}}{\int}D\theta{D\theta^{*}}\frac{\delta\theta^{*}_{q{\bf e}_{\phi}}(t+0^{+})}{2}\frac{\delta\theta_{q{\bf e}_{\phi}}(t)}{2}e^{iS}
+1𝒵0DθDθδθq𝐞ϕ(τ)2δθq𝐞ϕ(τ)2e𝒮],\displaystyle\mbox{}+\frac{1}{\mathcal{Z}_{0}}{\int}D\theta{D\theta^{*}}\frac{\delta\theta^{*}_{q{\bf e}_{\phi}}(\tau)}{2}\frac{\delta\theta_{q{\bf e}_{\phi}}(\tau)}{2}e^{-\mathcal{S}}\Big{]},

with Z0Z_{0} and 𝒵0\mathcal{Z}_{0} standing for the corresponding partition functions. Following the standard generating functional methodFT to handle above equation (refer to Appendix C), one has

(psϕ)2=iCqωq2ω2Uq𝐞ϕUq𝐞ϕ(ω2ωN2+i0+)21βqωnq2Dq1(iωn)2ωN2,(p_{s}^{\phi})^{2}\!=\!\frac{i}{C}\sum_{q\omega}\frac{q^{2}\omega^{2}U_{q{\bf e}_{\phi}}U_{-q{\bf e}_{\phi}}}{(\omega^{2}\!-\!\omega_{N}^{2}+i0^{+})^{2}}\!-\!\frac{1}{\beta}\sum_{q\omega_{n}}\frac{q^{2}}{D_{q}}\frac{1}{(i\omega_{n})^{2}\!-\!\omega_{N}^{2}}, (43)

where ωn=2nπT\omega_{n}=2n\pi{T} denotes the Matsubara frequency and β=1/(kBT)\beta=1/(k_{B}T) with kBk_{B} being the Boltzmann constant. Then, after the frequency and Matsubara-frequency summations, Eq. (2) can be directly obtained.

IV.3 Path integral approach

In this part, in order to confirm our derivation from the GIKE approach, we use the path integral approachpi1 ; pi4 to derive the macroscopic phase-coherence dynamics and microscopic electronic-fluid dynamics. We start with the generalized action of superconductors:

S[ϕ,ϕ]=s=,𝑑xψs(x)[itξ𝐩^U(x)]ψs(x)\displaystyle S[\phi,\phi^{*}]=\sum_{s=\uparrow,\downarrow}\int{dx}\psi^{*}_{s}(x)[i\partial_{t}-\xi_{\hat{\bf p}}-U(x)]\psi_{s}(x)
12ss𝑑x𝑑xV(xx)ψs(x)ψs(x)ψs(x)ψs(x)\displaystyle\mbox{}-\frac{1}{2}\sum_{ss^{\prime}}\int{dxdx^{\prime}}V(x-x^{\prime})\psi_{s}^{*}(x)\psi_{s^{\prime}}^{*}(x^{\prime})\psi_{s^{\prime}}(x^{\prime})\psi_{s}(x)
+𝑑x𝑑xg(xx)ψ(x)ψ(x)ψ(x)ψ(x).\displaystyle\mbox{}+\int{dxdx^{\prime}}g(x-x^{\prime})\psi^{*}_{\uparrow}(x)\psi^{*}_{\downarrow}(x^{\prime})\psi_{\downarrow}(x^{\prime})\psi_{\uparrow}(x). (44)

All symbols used here are same as the previous ones. Applying the Hubbard-Stratonovich transformation, the above action becomes

S[ψ,ψ]\displaystyle S[\psi,\psi^{*}]\!\! =\displaystyle= s=,𝑑xψs(x)[itξ𝐩^U(x)μH(x)]ψs(x)\displaystyle\!\!\!\!\sum_{s=\uparrow,\downarrow}\int{dx}\psi^{*}_{s}(x)[i\partial_{t}\!-\!\xi_{\hat{\bf p}}\!-\!U(x)\!-\!\mu_{H}(x)]\psi_{s}(x) (45)
𝑑x𝑑xψ(x)Δ^(x,x)ψ(x)\displaystyle\mbox{}-\int{dxdx^{\prime}}\psi^{*}(x){\hat{\Delta}}(x,x^{\prime})\psi(x^{\prime})
d4R|Δ|2g+12ω,𝐪|μH(q)|2Vq.\displaystyle\mbox{}-\!\int{d^{4}R}\frac{|\Delta|^{2}}{g}+\frac{1}{2}\sum_{\omega,{\bf q}}\frac{|\mu_{H}(q)|^{2}}{V_{q}}.

in which we have substituted the pairing potential g𝐤𝐤=gcos(ζθ𝐤+α)cos(ζθ𝐤+α)g_{\bf kk^{\prime}}=g\cos(\zeta\theta_{\bf k}+\alpha)\cos(\zeta\theta_{\bf k^{\prime}}+\alpha) to give rise to Δ𝐤=Δcos(ζθ𝐤+α)\Delta_{{\bf k}}=\Delta\cos(\zeta\theta_{\bf k}+\alpha), with ζ\zeta being the orbital angular momentum of the pairing symmetry. By further using the unitary transformation in Eq. (17) to effectively remove the phase fluctuation from the order parameter, one has

S[ψ,ψ]=𝑑x𝑑xψ(x)[G01(x,x)Σ(R)τ3]ψ(x)\displaystyle S[\psi,\psi^{*}]\!=\!\int{dxdx^{\prime}}\psi^{*}(x)[G^{-1}_{0}(x,x^{\prime})\!-\!\Sigma(R)\tau_{3}]\psi(x^{\prime})
Vfd4RΣ(R)d4R|Δ|2g+12ω,𝐪|μH(q)|2Vq,\displaystyle\mbox{}-\!V_{f}\!\int{d^{4}R}\Sigma(R)\!-\!\int{d^{4}R}\frac{|\Delta|^{2}}{g}\!+\!\frac{1}{2}\sum_{\omega,{\bf q}}\frac{|\mu_{H}(q)|^{2}}{V_{q}}, (46)

where Vf=𝐤1V_{f}=\sum_{\bf k}1 arises from the anti-commutation of the Fermi field; the Green function is written as

G1(x,x)=[it𝐩s𝐩^/mξ𝐩^τ3]δ(xx)|Δ(x,x)|τ1,G^{-1}(x,x^{\prime})=[i\partial_{t}-{\bf p}_{s}\cdot{\hat{\bf p}}/m-\xi_{\hat{\bf p}}\tau_{3}]\delta(x-x^{\prime})-|\Delta(x,x^{\prime})|\tau_{1}, (47)

and the self-energy reads

Σ(R)=μeff(R)+ps22m.\Sigma(R)=\mu_{\rm eff}(R)+\frac{p_{s}^{2}}{2m}. (48)

It is noted that in the previous workspi1 ; pi4 , the Doppler-shift term 𝐩s𝐩^/m{\bf p}_{s}\cdot{\hat{\bf p}}/m is placed into the self-energy Σ(R)\Sigma(R) and then treated as small quantity to take perturbation expansion. This approximation holds only if |𝐩s𝐯𝐤|<Δ𝐤|{\bf p}_{s}\cdot{{\bf v}_{\bf k}}|<\Delta_{\bf k}, while this condition is usually satisfied in conventional ss-wave superconductors with inactive phase fluctuation or weak external electromagnetic field. In the present work, considering the anisotropy and strong phase fluctuation in dd-wave cuprate superconductors, we sublate this approximation by taking the Doppler-shift term into the Green function (i.e., quasiparticle energy spectra).

Then, after the standard integration over the Fermi field, one has

S\displaystyle S =\displaystyle= T¯rln[G01Στ3]\displaystyle{\rm{\bar{T}r}}\ln{[G_{0}^{-1}-\Sigma\tau_{3}]} (49)
Vfd4RΣ(R)d4R|Δ|2g+12ω,𝐪|μH(q)|2Vq\displaystyle\mbox{}-\!V_{f}\!\int{d^{4}R}\Sigma(R)-\!\int{d^{4}R}\frac{|\Delta|^{2}}{g}+\frac{1}{2}\sum_{\omega,{\bf q}}\frac{|\mu_{H}(q)|^{2}}{V_{q}}
=\displaystyle= T¯rlnG011nn=1T¯r[(G0Στ3)n]\displaystyle{\rm{\bar{T}r}}\ln{G_{0}^{-1}}-\frac{1}{n}\sum_{n=1}^{\infty}{\rm{\bar{T}r}}[(G_{0}\Sigma\tau_{3})^{n}]
Vfd4RΣ(R)d4R|Δ|2g+12ω,𝐪|μH(q)|2Vq\displaystyle\mbox{}-\!V_{f}\!\int{d^{4}R}\Sigma(R)-\!\int{d^{4}R}\frac{|\Delta|^{2}}{g}+\frac{1}{2}\sum_{\omega,{\bf q}}\frac{|\mu_{H}(q)|^{2}}{V_{q}}
\displaystyle\approx d4Rpn,𝐤ln[(ipnE𝐤+)(ipnE𝐤)]\displaystyle\int{d^{4}{R}}{\sum_{p_{n},{\bf k}}}\ln[(ip_{n}-E_{\bf k}^{+})(ip_{n}-E_{\bf k}^{-})]
Q1d4RΣ(R)+Q2d4R[Σ(R)]2\displaystyle\mbox{}-\!Q_{1}\int{d^{4}R}\Sigma(R)\!+\!Q_{2}\int{d^{4}R}[\Sigma(R)]^{2}
d4R|Δ|2g+12ω,𝐪|μH(q)|2Vq,\displaystyle\mbox{}-\!\int{d^{4}R}\frac{|\Delta|^{2}}{g}+\frac{1}{2}\sum_{\omega,{\bf q}}\frac{|\mu_{H}(q)|^{2}}{V_{q}},

in which we have neglected the mutual interaction between phase fluctuations by only keeping the lowest two orders (i.e., n=1n=1 and n=2n=2). Here, pn=(2n+1)πTp_{n}=(2n+1)\pi{T} denotes the Matsubara frequency and the coefficients read (refer to Appendix D)

Q1\displaystyle Q_{1}\!\!\! =\displaystyle= Vf+pTr[G0(p)τ3]=kF2m𝐤ξ𝐤(ξ𝐤F𝐤),\displaystyle\!\!\!V_{f}\!+\!\sum_{p}{\rm Tr}[G_{0}(p)\tau_{3}]=\!-\frac{k_{F}^{2}}{m}{\sum_{\bf k}}^{\prime}\partial_{\xi_{\bf k}}\big{(}\xi_{\bf k}F_{\bf k}\big{)},~{}~{}~{}~{} (50)
Q2\displaystyle Q_{2}\!\!\! =\displaystyle= 12pTr[G0(p)τ3G0(p)τ3]=𝐤ξ𝐤(ξ𝐤F𝐤).\displaystyle\!\!\!-\frac{1}{2}\sum_{p}{\rm Tr}[G_{0}(p)\tau_{3}G_{0}(p)\tau_{3}]=\!-\!{\sum_{\bf k}}^{\prime}\partial_{\xi_{\bf k}}\big{(}\xi_{\bf k}F_{\bf k}\big{)}.~{}~{}~{}~{} (51)

Here, p=(ipn,𝐤)p=(ip_{n},{\bf k}) and the Green function G(p)G(p) from Eq. (47) is given by

G0(p)=ipnτ0𝐩s𝐯𝐤τ0+ξ𝐤τ3+Δ𝐤τ1(ipnE𝐤+)(ipnE𝐤).G_{0}(p)=\frac{ip_{n}\tau_{0}-{\bf p}_{s}\cdot{\bf v_{k}}\tau_{0}+\xi_{\bf k}\tau_{3}+\Delta_{\bf k}\tau_{1}}{(ip_{n}-E_{\bf k}^{+})(ip_{n}-E_{\bf k}^{-})}. (52)

As μH(q=0)=Vq=0δnq=0=0\mu_{H}(q=0)=V_{q=0}\delta{n_{q=0}}=0, it can be easily demonstrated that d4RΣ(R)=d4R[U(R)+ps2/(2m)]\int{d^{4}R}\Sigma(R)=\int{d^{4}R}[U(R)+p^{2}_{s}/(2m)]. Then, one obtains the effective action of superconductors:

Seff\displaystyle S_{\rm eff} =\displaystyle= d4R{pn,𝐤ln[(ipnE𝐤+)(ipnE𝐤)]|Δ|2g\displaystyle\int{d^{4}{R}}\Big{\{}{\sum_{p_{n},{\bf k}}}\ln[(ip_{n}-E_{\bf k}^{+})(ip_{n}-E_{\bf k}^{-})]\!-\!\frac{|\Delta|^{2}}{g} (53)
+Q2(μeff2kF2ps22m2)}+12ω,𝐪|μH(q)|2Vq.\displaystyle\mbox{}\!+\!Q_{2}\Big{(}\mu_{\rm eff}^{2}\!-\!\frac{k_{F}^{2}p_{s}^{2}}{2m^{2}}\Big{)}\Big{\}}\!+\!\frac{1}{2}\sum_{\omega,{\bf q}}\frac{|\mu_{H}(q)|^{2}}{V_{q}}.

We next handle the Hartree field. Through the integration over the Hartree field in Eq. (53), one gets

S¯eff=d4Rpn,𝐤{ln[(ipnE𝐤+)(ipnE𝐤)]|Δ|2g}\displaystyle{\bar{S}}_{\rm eff}=\int{d^{4}{R}}{\sum_{p_{n},{\bf k}}}\Big{\{}\ln[(ip_{n}-E_{\bf k}^{+})(ip_{n}-E_{\bf k}^{-})]\!-\!\frac{|\Delta|^{2}}{g}\Big{\}}
+𝑑t𝐪[Q21+2Q2Vq(tδθ𝐪2+U𝐪)2Q2kF2ps22m2].\displaystyle\mbox{}+\!\!\int{dt}\sum_{\bf q}\Big{[}\frac{Q_{2}}{1\!+\!2Q_{2}V_{q}}\Big{(}\frac{\partial_{t}\delta\theta_{\bf q}}{2}\!+\!U_{\bf q}\Big{)}^{2}\!-\!\frac{Q_{2}k_{F}^{2}p_{s}^{2}}{2m^{2}}\Big{]}.~{}~{}~{}~{} (54)

From the action above, we prove in the following that the previous gap equation [Eq. (4)], superfluid density [Eq. (5)] as well as the equation of motion of the phase fluctuation [Eq. (40)] derived from the GIKE approach can all be recovered.

Specifically, through the variation δS¯eff=0\delta{{\bar{S}}_{\rm eff}}=0 with respect to the gap, one has

Δ\displaystyle\Delta =\displaystyle= g𝐤Δcos2(ζθ𝐤+α)pn1(ipnE𝐤+)(ipnE𝐤)\displaystyle-g{\sum_{\bf k}}^{\prime}\Delta\cos^{2}(\zeta\theta_{\bf k}+\alpha)\sum_{p_{n}}\frac{1}{(ip_{n}-E_{\bf k}^{+})(ip_{n}-E_{\bf k}^{-})} (55)
=\displaystyle= g𝐤Δ𝐤cos(ζθ𝐤+α)F𝐤,\displaystyle-g{\sum}^{\prime}_{\bf k}\Delta_{\bf k}\cos(\zeta\theta_{\bf k}+\alpha)F_{\bf k},

which with g𝐤𝐤=gcos(ζθ𝐤+α)cos(ζθ𝐤+α)g_{\bf kk^{\prime}}=g\cos(\zeta\theta_{\bf k}+\alpha)\cos(\zeta\theta_{\bf k^{\prime}}+\alpha), is exactly same as the previous gap equation in Eq. (4). The supercurrent is given by

𝐣\displaystyle{\bf j} =\displaystyle= e𝐩sS¯eff=eQ2kF2𝐩𝐬m2+pn𝐤2e𝐤(ipn𝐤𝐩s/m)m(ipnE𝐤+)(ipnE𝐤)\displaystyle-e{\partial_{{\bf p}_{s}}{{\bar{S}}_{\rm eff}}}\!=\!\frac{eQ_{2}k_{F}^{2}{\bf p_{s}}}{m^{2}}\!+\!{\sum_{p_{n}{\bf k}}}\frac{2e{\bf k}(ip_{n}\!-\!{\bf k}\!\cdot\!{\bf p}_{s}/m)}{m(ip_{n}\!-\!E_{\bf k}^{+})(ip_{n}\!-\!E_{\bf k}^{-})} (56)
=\displaystyle= ekF2𝐩𝐬m2𝐤{ξ𝐤(ξ𝐤F𝐤)E𝐤[f(E𝐤+)f(E𝐤)2]}\displaystyle-\frac{ek_{F}^{2}{\bf p_{s}}}{m^{2}}{\sum_{\bf k}}^{\prime}\Big{\{}\partial_{\xi_{\bf k}}(\xi_{\bf k}F_{\bf k})-\partial_{E_{\bf k}}\Big{[}\frac{f(E_{\bf k}^{+})-f(E_{\bf k}^{-})}{2}\Big{]}\Big{\}}
=\displaystyle= e𝐩𝐬mkF2m𝐤Δ𝐤2E𝐤E𝐤F𝐤=ens𝐩𝐬m.\displaystyle\frac{e{\bf p_{s}}}{m}\frac{k_{F}^{2}}{m}{\sum_{\bf k}}^{\prime}\frac{\Delta_{\bf k}^{2}}{E_{\bf k}}\partial_{E_{\bf k}}F_{\bf k}=\frac{en_{s}{\bf p_{s}}}{m}.

Then, the previous supercurrent in Eq. (34) and hence superfluid density in Eq. (5) are recovered.

Through the variation δS¯eff=0\delta{{\bar{S}}_{\rm eff}}=0 with respect to the phase fluctuation, one obtains

0\displaystyle 0 =\displaystyle= t[S¯eff(tδθ𝐪/2)]i𝐪(S¯eff𝐩s)\displaystyle\partial_{t}\Big{[}\frac{\partial{\bar{S}}_{\rm eff}}{\partial(\partial_{t}\delta\theta_{\bf q}/2)}\Big{]}-i{\bf q}\cdot\Big{(}\frac{\partial{\bar{S}}_{\rm eff}}{\partial{\bf p}_{s}}\Big{)} (57)
=\displaystyle= 2Q21+2Q2Vqt(tδθ𝐪2+U𝐪)i𝐪𝐩snsm\displaystyle\frac{2Q_{2}}{1+2Q_{2}V_{q}}\partial_{t}\Big{(}\frac{\partial_{t}\delta\theta_{\bf q}}{2}+U_{\bf q}\Big{)}-{i{\bf q}\cdot{\bf p}_{s}}\frac{n_{s}}{m}
=\displaystyle= Dq(q2ns2Dm+q2Vqnsm+t2)δθ𝐪2+DqtU𝐪,\displaystyle D_{q}\Big{(}\frac{q^{2}n_{s}}{2Dm}+\frac{q^{2}V_{q}n_{s}}{m}+\partial^{2}_{t}\Big{)}\frac{\delta\theta_{\bf q}}{2}+D_{q}\partial_{t}U_{\bf q},

in which we have used Q2=𝐤ξ𝐤(ξ𝐤F𝐤)DQ_{2}=-\sum_{\bf k}^{\prime}\partial_{\xi_{\bf k}}(\xi_{\bf k}F_{\bf k})\approx{D}. Then, the previous equation of motion of the phase fluctuation in Eq. (40) is recovered.

Consequently, the previous gap equation [Eq. (4)], superfluid density [Eq. (5)] as well as the equation of motion of the phase fluctuation [Eq. (40)] derived from the GIKE approach can all be recovered by the path-integral approach. Whereas the microscopic scattering of electronic fluids (i.e., microscopic momentum-relaxation rate in superfluid) is hard to handle within the path-integral approach. The conventional Kubo diagrammatic formalism is also difficult to track the microscopic scattering, as the inevitable calculation of the vertex correction becomes hard to deal with in superconductorsG1 , especially considering the anisotropy and significant phase fluctuation (i.e., Doppler shift) in dd-wave case. Thus, the GIKE approach in fact provides an efficient way to deal with the microscopic scattering in superconductors for investigating the superconductivity (zero-resistance) phenomenon.

V Application to disordered ss-wave superconductors

Finally, considering the recent experimental progress of realizing the atomically thin superconductorsatsc01 ; atsc02 ; atsc03 ; atsc04 ; atsc1 and disordered superconducting filmsDSTF1 ; DSTF2 ; DSTF3 ; DSTF4 with ss-wave gap, we briefly investigate the two-dimensional disordered ss-wave superconductors. In this circumstance, the phase fluctuation retains gapless energy spectrum even after considering the long-range Coulomb interaction, and hence, is active. We therefore apply the developed dual dynamics with gap Δ𝐤=Δ\Delta_{\bf k}=\Delta and pairing potential g𝐤𝐤=gg_{\bf kk^{\prime}}=g. By using the similar numerical simulation as the dd-wave case in Sec. III.2, the obtained results in disordered ss-wave superconductors are plotted in Fig. 5.

Refer to caption
Figure 5: (Color online) (a) Anderson-disorder strength and (b) temperature dependence of gap Δ\langle\Delta\rangle and densities of the non-viscous superfluid nns\langle{n_{\rm ns}}\rangle, viscous superfluid nvs\langle{n_{\rm vs}}\rangle and normal fluid nn\langle{n_{n}}\rangle as well as (c) phase diagram of Δ\langle\Delta\rangle from the full and self-consistent numerical simulation in disordered ss-wave superconductors. The results of nvs\langle{n_{\rm vs}}\rangle in figures (a) and (b) are multiplied by 2525 for the visibility. Note that the small deviation of Δ/Δ0\Delta/\Delta_{0} from one at W=0W=0 in figure (a) originates from the finite temperature. (d) Schematic showing the division in the momentum space at 𝐩s=0.55ωD/vF𝐞ϕ=π/4{\bf p}_{s}=0.55\omega_{D}/v_{F}{\bf e}_{\phi=\pi/4} and Δ=0.5ωD\Delta=0.5\omega_{D} in ss-wave superconductors. In the figure, TBT_{B} denotes the critical temperature where gap vanishes in the absence of the phase fluctuation, i.e., BCS critical temperature of the two-dimensional ss-wave superconductors; ns0n_{s0} represents the superfluid density at zero temperature in the absence of the phase fluctuation, and nn0n_{n0} denotes the normal-fluid density above TosT^{\rm os}; W0=0.5Δ0W_{0}=0.5\Delta_{0}.

Actually, based on the three-fluid model, by applying the analysis as the dd-wave case in Sec. III.1, one also expects a similar behavior between the non-viscous superfluid density nns\langle{n_{\rm ns}}\rangle and gap Δ\langle\Delta\rangle and a compensatory behavior between nns\langle{n_{\rm ns}}\rangle and normal-fluid density nn\langle{n_{n}}\rangle in temperature (Anderson-disorder strength) dependence of disordered ss-wave superconductors, as confirmed by numerical results in Fig. 5(b) [Fig. 5(a)].

The difference from the dd-wave case in disordered ss-wave superconductors includes three aspects. Firstly, as mentioned in Sec. III.1, differing from the dd-wave case where nonzero normal fluid always emerges around nodal points, the emergence of the normal fluid in ss-wave superconductors requires ps>Δ/vF{p}_{s}>\Delta/v_{F}, leading to a threshold to realize normal fluidGIKE1 . Therefore, at small phase fluctuation with ps<Δ/vFp_{s}<\Delta/v_{F}, no normal fluid and hence viscous superfluid are generated in ss-wave superconductors, while the gap and non-viscous superfluid density are free from the influence of the phase fluctuation (i.e., increase of WW), as shown in Fig. 5(a) at W<0.24W<0.24.

Secondly, with the increase of phase fluctuation (by enhancing WW) at ps>Δ/vFp_{s}>\Delta/v_{F}, the gap Δ\langle\Delta\rangle in disordered ss-wave superconductors [red solid curve in Fig. 5(a)] shows a much faster decrease than that in dd-wave case [Fig. 2(i)]. This implies that the ss-wave gap with the higher rotational symmetry is much more fragile against the phase fluctuation. At large phase fluctuation, as shown in Fig. 5(b), the temperature dependence of Δ\langle\Delta\rangle (red solid curve) deviates from the BCS-like behavior and exhibits a faster decrease with temperature, and a nonzero fraction of normal fluid emerges even at zero temperature. These two behaviors provide a scheme for experimental detection to justify the existence of phase fluctuation in disordered ss-wave superconductors.

Furthermore, as seen from Figs. 5(a) and (b), in disordered ss-wave superconductors, the viscous superfluid density nvs\langle{n_{\rm vs}}\rangle is marginal, and hence, the normalized non-viscous superfluid density nns/ns0\langle{n_{\rm ns}}\rangle/n_{s0} (blue dashed curve) nearly coincides with the normalized gap Δ/Δ0\langle\Delta\rangle/\Delta_{0} (red solid curve), in sharp contrast to the results in dd-wave case. This arises from the isotropy of the ss-wave gap. Specifically, in disordered ss-wave superconductors, as shown by the schematic illustration in Fig. 5(d), at phase fluctuation ps>Δ/vFp_{s}>\Delta/v_{F}, even the normal fluid (yellow regions) is excited, the viscous pairing (orange) region that can experience the friction with normal fluid [i.e., satisfying the energy conservation of momentum-relaxation rate in Eq. (6)] lies around the unpairing one and is very small, in sharp contrast to the large fraction in anisotropic dd-wave case at the same condition [Figs. 1(b) and (f)]. Moreover, the anomalous correlation F𝐤F_{\bf k} adjacent to the unpairing region is very small. Therefore, due to the small area and weak F𝐤F_{\bf k}, the viscous pairing region in disordered ss-wave superconductors makes a neglectable contribution in the gap equation [Eq. (4)] and leads to a marginal nvsn_{\rm vs} [Eq. (9)]. With the marginal nvs\langle{n_{\rm vs}}\rangle, the separation between TcT_{c} and TosT^{\rm os} is remarkably small (0.01TB0.01T_{B} at most) in our simplified model.

VI Summary

In summary, by using the GIKE approachGIKE1 ; GIKE2 ; GIKE3 ; GIKE4 , we construct the coupled dual dynamics of macroscopic phase coherence and microscopic electronic fluids (consisting of normal fluid and superfluid) in cuprate superconductors. The macroscopic phase-coherence dynamics is developed by deriving the equation of motion of the phase fluctuation, in which both disorder and long-range Coulomb interaction effects are considered. We show that the phase fluctuation in cuprate superconductors retains gapless energy spectrum thanks to the layered structurescuprate4 ; cuprate5 , and hence, is active, differing from the conventional bulk superconductors with inactive phase fluctuation due to Anderson-Higgs mechanismAHM ; AK ; Am0 ; Ba0 . Moreover, the superfluid density determines the phase stiffness in the phase-coherence dynamics, in consistency with the previous understanding in the literaturephase1 ; phase2 ; phase3 . The microscopic electronic-fluid dynamics includes two parts: the anomalous correlation to determine gap and superfluid density; the microscopic scattering of the electronic fluids that is essential for studying the transport property and hence superconductivity. It is shown that both anomalous correlation and microscopic scattering in the electronic-fluid dynamics are affected by the phase fluctuation as it drives the Doppler shift in quasiparticle energy spectraFF4 ; FF5 ; FF6 ; GIKE1 by generating a superconducting momentumgi0 ; Ba0 ; G1 .

Based on the developed dual dynamics, we present theoretical descriptions of the separation between TcT_{c} and TosT^{\rm os} as well as the emerged normal fluid below TcT_{c} in cuprate superconductors. We find that the key origin of both phenomena comes from the quantum effect of disorder, which excites the macroscopic inhomogeneous phase fluctuation through Josephson effect.Josephson With this excited phase fluctuation, we prove that the unpairing region with vanishing anomalous correlationFF1 ; FF2 ; FF7 ; FF8 ; FF9 ; GIKE1 , i.e., normal fluid, always emerges around nodal points in dd-wave cuprate superconductors even at zero temperature, in consistency with the experimentally observed substantial fraction of normal state at low temperatureNSB1 ; NSB2 . Whereas according to the microscopic scattering, the pairing region, i.e., superfluid, is divided into two parts: viscous and non-viscous ones. Particles in the viscous pairing region experience a finite momentum relaxation, due to the scattering with the ones in unpairing region, behaving like the friction between the superfluid and normal fluid. The non-viscous superfluid is free from the momentum relaxation scattering. Therefore, in addition to conventional non-viscous superfluid, there also exist normal fluid and viscous superfluid at small phase fluctuation in cuprate superconductors, similar to the three-fluid model proposed in our previous workGIKE1 in conventional superconductors which is caused by external electromagnetic field. An experimental scheme to distinguish the densities of these three electronic fluids is proposed.

We further demonstrate that by increasing the temperature in cuprate superconductors, the suppressed gap and hence superfluid density weaken the phase stiffness, enhancing the phase fluctuation. Once the phase fluctuation (i.e., temperature) exceeds the critical point, the non-viscous superfluid vanishes, leaving only normal fluid and viscous superfluid. The system then enters the pseudogap state with nonzero resistivity and finite gap due to the significant phase fluctuation. It is noted that in this circumstance, the viscous superfluid matches the description of the incoherent preformed Cooper pairs, as they both contribute to gap but experience the scattering. Whereas the existing normal fluid in our description implies the existence of normal particles in pseudogap state, which has been overlooked in previous preformed Cooper-pair model to describe pseudogap statephase2 ; PCP1 ; PCP2 ; PCP3 ; PCP4 ; PCP5 . The viscous superfluid starts to shrink by further increasing temperature in pseudogap state, until vanishes at TosT^{\rm os} where gap is eventually destroyed.

To confirm the derivation from the GIKE approach, we also apply the path-integral approach to recover the equation of motion of the phase fluctuation as well as anomalous correlation, gap equation and superfluid density in the presence of the superconducting momentum. Holding the pairing potential fixed, a self-consistent numerical simulation by applying Anderson disorder is also addressed, to verify our theoretical description. Then, both the separation between TcT_{c} and TosT^{\rm os} as well as the emerged normal fluid below TcT_{c} are confirmed. Particularly, TosT^{\rm os} tracks TcT_{c} closely in our numerical results, and lies well below the experimental pseudogap temperature TT^{*}, in consistency with the current experimental findings and understanding in the literatureTM2 ; TM3 ; TM4 ; SH ; PC1 ; NF1 ; AF1 ; AF2 ; AF3 ; AF4 ; PPS1 ; PPS2 ; Infrad1 ; Infrad2 ; Infrad3 ; MS1 ; MS2 ; MS3 ; THZ1 ; THZ2 ; THZ3 ; THZ4 ; DHM3 . This is because that once in the pseudogap state, due to the significant phase fluctuation and hence Doppler shift, the gap can not survive far above TcT_{c}.

Consequently, when the pairing potential is determined, the coupled dual dynamics in the present work provides an efficient and simplified approach to formulate the dephasing process of macroscopic superconducting phase coherence with the increase of temperature, as well as the influence of this dephasing on microscopic electronic fluids (including gap, densities of superfluid and normal fluid, and in particular, the transport property to determine superconducting transition temperature TcT_{c}). This theory distinguishingly takes the Anderson-disorder approach and in particular, impurity scattering treatment to discuss the disorder effects on macroscopic phase coherence and microscopic electronic fluids, respectively. The Anderson-disorder approach that calculates the converged quantities by averaging over numerous random impurity configuration is applied in order to characterize the quantum disorder/fluctuation effects from the impurity as well as possible couplings to charge-density wavecuprate5 ; CDW0 ; CDW1 ; CDW2 ; CDW3 , spin-density wavecuprate5 ; SDW1 ; SDW2 ; SDW3 ; SDW4 , electronic nematicitycuprate5 ; Nernst6 ; NO1 ; NO2 and/or theoretically proposed spinon-vortices excitationVP1 ; VP2 ; VP3 . Whereas the microscopic scattering, as the essential transport property for studying the superconductivity (zero-resistance) phenomenon, is still absent in the literature, since it is hard to handle in path-integral approach and conventional Kubo diagrammatic formalism. But the GIKE approach straightly tackles this crucial problem. Further determining the pairing potential and Anderson-disorder strength at different dopings requires the microscopic pairing mechanism, which still remains an open question in the literature and is beyond the scope of present work.

Considering the recent experimental progress of realizing the atomically thin superconductorsatsc01 ; atsc02 ; atsc03 ; atsc04 ; atsc1 and disordered superconducting filmsDSTF1 ; DSTF2 ; DSTF3 ; DSTF4 with the ss-wave gap, we show that the developed dual dynamics can also be applied similarly to the low-dimensional disordered ss-wave superconductors, which exhibits an active phase fluctuation due to the gapless energy spectrum. Differing from the dd-wave case where nonzero normal fluid always emerges around nodal points, there exists a threshold for phase fluctuation to induce normal fluid in isotropic ss-wave superconductorsGIKE1 . Consequently, at small phase fluctuation below threshold, no normal fluid and hence viscous superfluid are generated in ss-wave superconductors, while the gap and non-viscous superfluid density are free from the influence of the phase fluctuation. But for large phase fluctuation above threshold, we find that the ss-wave gap with higher rotational symmetry is much more fragile against the phase fluctuation. In this situation, the temperature dependence of gap deviates from the BCS-like behavior and exhibits a faster decrease with temperature, and a nonzero fraction of normal fluid emerges even at zero temperature. Nevertheless, due to the isotropy of ss-wave gap, the viscous superfluid density is marginal in our simplified model, leading to a remarkably small separation between TcT_{c} and TosT^{\rm os}.

VII Discussion

Finally, we would like to discuss recent theoretical and experimental progresses in cuprate superconductors.

Role of bandstructure.—The present work approximately takes the parabolic spectrum. Whereas recently, from the experimental angle-resolved photo-emission measurement, He et al. reported an important role of the flat energy dispersion near antinodal point in determining TcT_{c} (i.e., phase coherence) in cuprate superconductorsFB . Here, based on our coupled dual dynamics, we briefly discuss this effect. Specifically, around Fermi surface, with the flat (i.e., v𝐤0v_{\bf k}\approx 0) and steep (i.e., large v𝐤v_{\bf k}) energy dispersions near antinodal and nodal points, respectively, a large unpairing region (|𝐯𝐤𝐩s|>E𝐤|{\bf v}_{\bf k}\cdot{\bf p}_{s}|>E_{\bf k}) around nodal points can be achieved easily in the presence of significant phase fluctuation. Whereas the region around antinodal point is the pairing one (|𝐯𝐤𝐩s|<E𝐤|{\bf v}_{\bf k}\cdot{\bf p}_{s}|<E_{\bf k}), but can become viscous through the friction with the large unpairing region. Particularly, since the Doppler shift 𝐯𝐤𝐩s0{\bf v}_{\bf k}\cdot{\bf p}_{s}\approx 0 nearby, this viscous pairing region can survive against very large phase fluctuation. Consequently, a viscous superfluid that contributes to gap can exist far above TcT_{c}, enlarging the separation between TcT_{c} and TosT^{\rm os}. Moreover, following the similar analysis, one can also expect a larger separation between TcT_{c} and TosT^{\rm os} in low-dimensional disordered ss-wave superconductors with complex Fermi surface. The numerical simulation of this scheme in cuprate or disordered ss-wave superconductors requires a specific calculation of the bandstructure, which is beyond the scope of the present work.

Disorder treatment.—Recently, Lee-Hone et al.DS1 ; DS2 calculated the self-energy from the impurity and then self-consistently formulated the renormalizations of the gap and superfluid density in dd-wave cuprate superconductors. A finite disorder influence is obtained in their results, differing from the vanishing one in conventional ss-wave superconductors (Anderson theorem)AT1 ; AT2 ; AT3 . It is noted that this disorder treatment which takes the random phase approximation focuses on the diffusive-motion influence on the equilibrium gap Δ𝐤\Delta_{\bf k}. The disorder effect on the phase coherence or amplitude fluctuation as well as the microscopic scattering of the nonequilibrium response are beyond their equilibrium calculation that carried out with the translational symmetry.

Amplitude fluctuation.—Another issue concerns the amplitude fluctuation of the order parameter. As mentioned in Sec. IV.2.1, in the leading order of the disorder effect where the particle-hole symmetry is present, one finds a vanishing amplitude fluctuation. But the phase fluctuation is generated by Josephson effect and then influences the electronic fluids by inducing a superconducting momentum, as the present work addressed. Whereas in higher-order contribution with the broken particle-hole symmetry, the disorder-induced local potential can directly excite the amplitude fluctuation, as the charge-amplitude correlation becomes nonzerosymmetry .

Recently, by using the tight-binding model in real space and applying the Anderson disorder, Li et al. performed a self-consistent numerical calculation to solve the wave function and hence gapdisorder . A drastic amplitude fluctuation emerges in their results by disorder effect, leading to granular superconducting islands where the gap is destroyed in strong-disorder regions. They attributed this formation to the pairing-breaking effect on anisotropic gap from disorder, and then, suggested that the strong phase fluctuation emerges in regions with small gap (i.e., superfluid stiffness) as a consequence. This approach clearly handles the high-order disorder effect (i.e., generation of amplitude fluctuation) well, but the leading contribution via Josephson effect [i.e., tδθ/2=U(R)\partial_{t}\delta\theta/2=-U(R)] to excite the phase fluctuation is not involved in this stationary calculation. Particularly, in this situation with strong disorder, the phase fluctuation is likely to destroy the global gap before the drastic amplitude fluctuation emerges. Even the global gap survives, the influence of the phase fluctuation on such state with formation of granular superconducting islands can not be overlooked.

Acknowledgements.
The authors acknowledge financial support from the National Natural Science Foundation of China under Grants No. 11334014 and No. 61411136001.

Appendix A Solution of GIKE

We present the solution of GIKE in this part. The first order of GIKE [Eq. (25)] in clean limit reads

tδρ𝐤(1)+i[ξ𝐤τ3+Δ𝐤τ1,δρ𝐤(1)]+i[μeffτ3+δΔ𝐤(1)τ1,ρ𝐤(0)]i8m[τ3,𝐑2δρ𝐤(1)]+12{𝐯𝐤τ3,𝐑δρ𝐤(1)}+12{𝐯𝐤τ3,𝐑ρ𝐤(0)}\displaystyle\partial_{t}\delta\rho^{(1)}_{\bf k}\!+\!i[\xi_{\bf k}\!\tau_{3}\!+\!\Delta_{{\bf k}}\tau_{1},\delta\rho^{(1)}_{\bf k}]\!+\!i[\mu_{\rm eff}\tau_{3}\!+\!\delta\Delta^{(1)}_{\bf k}\tau_{1},\rho^{(0)}_{\bf k}]\!-\!\frac{i}{8m}[\tau_{3},\nabla_{\bf R}^{2}\delta\rho_{\bf k}^{(1)}]\!+\!\frac{1}{2}\{{\bf v_{k}}\tau_{3},{\bm{\nabla}}_{\bf R}\delta\rho^{(1)}_{\bf k}\}\!+\!\frac{1}{2}\{{\bf v_{k}}\tau_{3},{\bm{\nabla}}_{\bf R}\rho^{(0)}_{\bf k}\}
+12{e𝐄τ32i𝐩sτ3Δ𝐤τ1,𝐤ρ𝐤(0)}i8[𝐑𝐑δΔ𝐤(1)τ1+2i𝐑𝐩sτ3Δ𝐤τ1,𝐤𝐤ρ𝐤(0)]14m[𝐑𝐩sτ3,τ3ρ𝐤(0)]=0.\displaystyle\mbox{}\!+\!\frac{1}{2}\{e{\bf E}\tau_{3}\!-\!2i{\bf p}_{s}\tau_{3}\Delta_{\bf k}\tau_{1},\partial_{\bf k}\rho^{(0)}_{\bf k}\}\!-\!\frac{i}{8}[{\bm{\nabla}}_{\bf R}{\bm{\nabla}}_{\bf R}\delta\Delta^{(1)}_{\bf k}\tau_{1}\!+\!2i{\bm{\nabla}}_{\bf R}{\bf p}_{s}\tau_{3}\Delta_{{\bf k}}\tau_{1},\partial_{\bf k}\partial_{\bf k}\rho^{(0)}_{\bf k}]\!-\!\frac{1}{4m}[\nabla_{\bf R}\!\cdot{\bf p}_{s}\tau_{3},\tau_{3}\rho_{\bf k}^{(0)}]=0. (58)

In center-of-mass frequency and momentum space [R=(t,𝐑)q=(ω,𝐪)][R=(t,{\bf R})\rightarrow{q=(\omega,{\bf q})}], by keeping the lowest three orders of 𝐪{\bf q}, the components of above equation are written as

iωδρ𝐤0(1)=i𝐯𝐤𝐪δρ𝐤3(1)e𝐄𝐤ρ𝐤3(0),\displaystyle i\omega\delta\rho^{(1)}_{{\bf k}0}={i{\bf v_{k}}\!\cdot\!{\bf q}}\delta\rho^{(1)}_{{\bf k}3}-e{\bf E}\!\cdot\!\partial_{\bf k}\rho^{(0)}_{{\bf k}3}, (59)
iωδρ𝐤3(1)=2Δ𝐤δρ𝐤2(1)+i𝐯𝐤𝐪δρ𝐤0(1)+i𝐪𝐩sΔ𝐤:𝐤𝐤ρ𝐤1(0)/2+(𝐯𝐤i𝐪)(𝐯𝐤𝐩s)L𝐤,\displaystyle i\omega\delta\rho^{(1)}_{{\bf k}3}=2\Delta_{{\bf k}}\delta\rho^{(1)}_{{\bf k}2}+{i{\bf v_{k}}\!\cdot\!{\bf q}}\delta\rho^{(1)}_{{\bf k}0}+{i}{\bf q}{\bf p}_{s}\Delta_{{\bf k}}:\partial_{\bf k}\partial_{\bf k}\rho^{(0)}_{{\bf k}1}/2+({\bf v_{k}}\cdot{i{\bf q}})({\bf v_{k}}\cdot{\bf p}_{s})L_{\bf k}, (60)
iωδρ𝐤1(1)=2ξ𝐤δρ𝐤2(1)i𝐪𝐩sΔ𝐤:𝐤𝐤ρ𝐤3(0)/2i𝐪𝐩sρ𝐤1(0)/(2m),\displaystyle i\omega\delta\rho^{(1)}_{{\bf k}1}=-2\xi_{\bf k}\delta\rho_{{\bf k}2}^{(1)}-{i}{\bf q}{\bf p}_{s}\Delta_{{\bf k}}:\partial_{\bf k}\partial_{\bf k}\rho^{(0)}_{{\bf k}3}/2-{i{\bf q}\!\cdot\!{\bf p}_{s}}\rho^{(0)}_{{\bf k}1}/{(2m)}, (61)
iωδρ𝐤2(1)=2Δ𝐤δρ𝐤3(1)+2ξ𝐤δρ𝐤1(1)2ρ𝐤3(0)δΔ𝐤(1)+2ρ𝐤1(0)μeffδΔ𝐤(1)(𝐪𝐤)2ρ𝐤3(0)/4,\displaystyle i\omega\delta\rho^{(1)}_{{\bf k}2}=-2\Delta_{{\bf k}}\delta\rho_{{\bf k}3}^{(1)}+2\xi_{\bf k}\delta\rho_{{\bf k}1}^{(1)}-2\rho^{(0)}_{{\bf k}3}\delta\Delta^{(1)}_{\bf k}+2\rho^{(0)}_{{\bf k}1}\mu_{\rm eff}-\delta\Delta_{\bf k}^{(1)}({\bf q}\cdot\partial_{\bf k})^{2}\rho_{{\bf k}3}^{(0)}/4, (62)

where L𝐤=𝐯𝐤𝐩sρ𝐤0(0)L_{\bf k}=\partial_{{\bf v_{k}}\cdot{\bf p}_{s}}\rho^{(0)}_{{\bf k}0}. Substituting Eqs. (60) and (61) into Eq. (62), one has

(iω)2δρ𝐤2+2Δ𝐤[2Δ𝐤δρ𝐤2(1)+i𝐯𝐤𝐪δρ𝐤0(1)+i𝐪𝐩sΔ𝐤:𝐤𝐤ρ𝐤1(0)/2+(𝐯𝐤i𝐪)(𝐯𝐤𝐩s)L𝐤]+2ξ𝐤[2ξ𝐤δρ𝐤2(1)\displaystyle(i\omega)^{2}\delta\rho_{{\bf k}2}+2\Delta_{\bf k}[2\Delta_{{\bf k}}\delta\rho^{(1)}_{{\bf k}2}+{i{\bf v_{k}}\!\cdot\!{\bf q}}\delta\rho^{(1)}_{{\bf k}0}+{i}{\bf q}{\bf p}_{s}\Delta_{{\bf k}}:\partial_{\bf k}\partial_{\bf k}\rho^{(0)}_{{\bf k}1}/2+({\bf v_{k}}\cdot{i{\bf q}})({\bf v_{k}}\cdot{\bf p}_{s})L_{\bf k}]+2\xi_{\bf k}[2\xi_{\bf k}\delta\rho_{{\bf k}2}^{(1)}
+i𝐪𝐩sΔ𝐤:𝐤𝐤ρ𝐤3(0)/2+i𝐪𝐩sρ𝐤1(0)/(2m)]=2iωρ(0)𝐤1μeff2iωρ(0)𝐤3δΔ(1)𝐤iωδΔ𝐤(1)(𝐪𝐤)2ρ𝐤3(0)/4.\displaystyle\mbox{}+{i}{\bf q}{\bf p}_{s}\Delta_{{\bf k}}:\partial_{\bf k}\partial_{\bf k}\rho^{(0)}_{{\bf k}3}/2+{i{\bf q}\!\cdot\!{\bf p}_{s}}\rho^{(0)}_{{\bf k}1}/{(2m)}]=2i\omega\rho^{(0)}_{{\bf k}1}\mu_{\rm eff}-2i\omega\rho^{(0)}_{{\bf k}3}\delta\Delta^{(1)}_{\bf k}-i\omega\delta\Delta_{\bf k}^{(1)}({\bf q}\cdot\partial_{\bf k})^{2}\rho_{{\bf k}3}^{(0)}/4. (63)

Further substituting Eq. (59) into above equation to replace δρ𝐤0(1)\delta\rho_{{\bf k}0}^{(1)}, with Eq. (62), one immediately obtains

δρ𝐤2(1)\displaystyle\delta\rho_{{\bf k}2}^{(1)} =\displaystyle= 14E𝐤2ω2{2iωρ𝐤1(0)μeff2iωρ𝐤3(0)δΔ𝐤(1)iω4δΔ𝐤(1)(𝐪𝐤)2ρ𝐤3(0)iΔ𝐤ξ𝐤𝐪𝐩s:𝐤𝐤ρ𝐤3(0)iΔ𝐤2𝐪𝐩s:𝐤𝐤ρ𝐤1(0)\displaystyle\frac{1}{4E^{2}_{\bf k}\!-\!\omega^{2}}\Big{\{}2i\omega\rho^{(0)}_{{\bf k}1}\mu_{\rm eff}\!-\!2i\omega\rho^{(0)}_{{\bf k}3}\delta\Delta^{(1)}_{\bf k}\!-\!\frac{i\omega}{4}\delta\Delta_{\bf k}^{(1)}({\bf q}\cdot\partial_{\bf k})^{2}\rho_{{\bf k}3}^{(0)}\!-\!i\Delta_{\bf k}\xi_{\bf k}{\bf q}{\bf p}_{s}:{\partial_{\bf k}}{\partial_{\bf k}}\rho^{(0)}_{{\bf k}3}\!-\!i\Delta_{\bf k}^{2}{\bf q}{\bf p}_{s}:{\partial_{\bf k}}{\partial_{\bf k}}\rho^{(0)}_{{\bf k}1} (64)
i𝐪𝐩smξ𝐤ρ𝐤1(0)2i(𝐪𝐯𝐤)(𝐯𝐤𝐩s)Δ𝐤L𝐤+2Δ𝐤iω(i𝐪𝐯𝐤)(e𝐄𝐤)ρ𝐤3(0)(𝐯𝐤𝐪ω)2[(ω24ξ𝐤2)δρ𝐤2(1)\displaystyle\mbox{}-\!\frac{i{\bf q}\cdot{\bf p}_{s}}{m}\xi_{\bf k}\rho^{(0)}_{{\bf k}1}-2i({\bf q}\!\cdot\!{\bf v_{k}})({\bf v}_{\bf k}\!\cdot\!{\bf p}_{s})\Delta_{\bf k}L_{\bf k}\!+\!\frac{2\Delta_{\bf k}}{i\omega}(i{\bf q}\!\cdot\!{\bf v_{k}})(e{\bf E}\!\cdot\!{\partial_{\bf k}})\rho^{(0)}_{{\bf k}3}\!-\!\Big{(}\frac{{\bf v_{k}}\cdot{\bf q}}{\omega}\Big{)}^{2}[(\omega^{2}\!-\!4\xi_{\bf k}^{2})\delta\rho_{{\bf k}2}^{(1)}
+2iωρ𝐤1(0)μeff2iωρ𝐤3(0)δΔ𝐤(1)]}.\displaystyle\mbox{}+\!2i\omega\rho^{(0)}_{{\bf k}1}\mu_{\rm eff}\!-\!2i\omega\rho^{(0)}_{{\bf k}3}\delta\Delta^{(1)}_{\bf k}]\}.

Then, after the first-order iteration, one obtains the solution of δρ𝐤2(1)\delta\rho_{{\bf k}2}^{(1)}:

δρ𝐤2(1)\displaystyle\delta\rho_{{\bf k}2}^{(1)} =\displaystyle= 14E𝐤2ω2{2iωρ𝐤1(0)μeff2iωρ𝐤3(0)δΔ𝐤(1)iω4δΔ𝐤(1)(𝐪𝐤)2ρ𝐤3(0)iΔ𝐤ξ𝐤𝐪𝐩s:𝐤𝐤ρ𝐤3(0)iΔ𝐤2𝐪𝐩s:𝐤𝐤ρ𝐤1(0)\displaystyle\frac{1}{4E^{2}_{\bf k}\!-\!\omega^{2}}\Big{\{}2i\omega\rho^{(0)}_{{\bf k}1}\mu_{\rm eff}\!-\!2i\omega\rho^{(0)}_{{\bf k}3}\delta\Delta^{(1)}_{\bf k}\!-\!\frac{i\omega}{4}\delta\Delta_{\bf k}^{(1)}({\bf q}\cdot\partial_{\bf k})^{2}\rho_{{\bf k}3}^{(0)}\!-\!i\Delta_{\bf k}\xi_{\bf k}{\bf q}{\bf p}_{s}:{\partial_{\bf k}}{\partial_{\bf k}}\rho^{(0)}_{{\bf k}3}\!-\!i\Delta_{\bf k}^{2}{\bf q}{\bf p}_{s}:{\partial_{\bf k}}{\partial_{\bf k}}\rho^{(0)}_{{\bf k}1} (65)
i𝐪𝐩smξ𝐤ρ𝐤1(0)2i(𝐪𝐯𝐤)(𝐯𝐤𝐩s)Δ𝐤L𝐤+2Δ𝐤iω(i𝐪𝐯𝐤)(e𝐄𝐤)ρ𝐤3(0)(𝐯𝐤𝐪ω)24Δ𝐤2/(4E𝐤2ω2)\displaystyle\mbox{}-\!\frac{i{\bf q}\cdot{\bf p}_{s}}{m}\xi_{\bf k}\rho^{(0)}_{{\bf k}1}\!-\!2i({\bf q}\!\cdot\!{\bf v_{k}})({\bf v}_{\bf k}\!\cdot\!{\bf p}_{s})\Delta_{\bf k}L_{\bf k}\!+\!\frac{2\Delta_{\bf k}}{i\omega}(i{\bf q}\cdot{\bf v_{k}})(e{\bf E}\cdot{\partial_{\bf k}})\rho^{(0)}_{{\bf k}3}\!-\!\Big{(}\frac{{\bf v_{k}}\cdot{\bf q}}{\omega}\Big{)}^{2}4\Delta_{\bf k}^{2}/(4E^{2}_{\bf k}\!-\!\omega^{2})
×[2iωρ𝐤1(0)μeff2iωρ𝐤3(0)δΔ𝐤(1)]},\displaystyle\mbox{}\times[2i\omega\rho^{(0)}_{{\bf k}1}\mu_{\rm eff}\!-\!2i\omega\rho^{(0)}_{{\bf k}3}\delta\Delta^{(1)}_{\bf k}]\Big{\}},

and hence, the solution of δρ𝐤1(1)\delta\rho_{{\bf k}1}^{(1)} from Eq. (61) is given by

δρ𝐤1(1)\displaystyle\delta\rho_{{\bf k}1}^{(1)} =\displaystyle= 2ξ𝐤/(iω)4E𝐤2ω2{2iωρ𝐤1(0)μeff2iωρ𝐤3(0)δΔ𝐤(1)iω4δΔ𝐤(1)(𝐪𝐤)2ρ𝐤3(0)iΔ𝐤ξ𝐤𝐪𝐩s:𝐤𝐤ρ𝐤3(0)iΔ𝐤2𝐪𝐩s:𝐤𝐤ρ𝐤1(0)\displaystyle\frac{-2\xi_{\bf k}/(i\omega)}{4E^{2}_{\bf k}\!-\!\omega^{2}}\Big{\{}2i\omega\rho^{(0)}_{{\bf k}1}\mu_{\rm eff}\!-\!2i\omega\rho^{(0)}_{{\bf k}3}\delta\Delta^{(1)}_{\bf k}\!-\!\frac{i\omega}{4}\delta\Delta_{\bf k}^{(1)}({\bf q}\cdot\partial_{\bf k})^{2}\rho_{{\bf k}3}^{(0)}\!-\!i\Delta_{\bf k}\xi_{\bf k}{\bf q}{\bf p}_{s}:{\partial_{\bf k}}{\partial_{\bf k}}\rho^{(0)}_{{\bf k}3}\!-\!i\Delta_{\bf k}^{2}{\bf q}{\bf p}_{s}:{\partial_{\bf k}}{\partial_{\bf k}}\rho^{(0)}_{{\bf k}1} (66)
i𝐪𝐩smξ𝐤ρ𝐤1(0)2i(𝐪𝐯𝐤)(𝐯𝐤𝐩s)Δ𝐤L𝐤+2Δ𝐤iω(i𝐪𝐯𝐤)(e𝐄𝐤)ρ𝐤3(0)(𝐯𝐤𝐪ω)24Δ𝐤2/(4E𝐤2ω2)\displaystyle\mbox{}-\!\frac{i{\bf q}\cdot{\bf p}_{s}}{m}\xi_{\bf k}\rho^{(0)}_{{\bf k}1}\!-\!2i({\bf q}\!\cdot\!{\bf v_{k}})({\bf v}_{\bf k}\!\cdot\!{\bf p}_{s})\Delta_{\bf k}L_{\bf k}\!+\!\frac{2\Delta_{\bf k}}{i\omega}(i{\bf q}\cdot{\bf v_{k}})(e{\bf E}\cdot{\partial_{\bf k}})\rho^{(0)}_{{\bf k}3}\!-\!\Big{(}\frac{{\bf v_{k}}\cdot{\bf q}}{\omega}\Big{)}^{2}4\Delta_{\bf k}^{2}/(4E^{2}_{\bf k}\!-\!\omega^{2})
×[2iωρ𝐤1(0)μeff2iωρ𝐤3(0)δΔ𝐤(1)]}i𝐪𝐩s2iωΔ𝐤:𝐤𝐤ρ(0)𝐤3i𝐪𝐩s2imωρ(0)𝐤1.\displaystyle\mbox{}\times[2i\omega\rho^{(0)}_{{\bf k}1}\mu_{\rm eff}\!-\!2i\omega\rho^{(0)}_{{\bf k}3}\delta\Delta^{(1)}_{\bf k}]\Big{\}}-\frac{{i}{\bf q}{\bf p}_{s}}{2i\omega}\Delta_{{\bf k}}:\partial_{\bf k}\partial_{\bf k}\rho^{(0)}_{{\bf k}3}-\frac{{i{\bf q}\!\cdot\!{\bf p}_{s}}}{2im\omega}\rho^{(0)}_{{\bf k}1}.

Moreover, for density-related δρ𝐤3(1)\delta\rho_{{\bf k}3}^{(1)} [Eq. (27)] and current-related δρ𝐤0(1)\delta\rho_{{\bf k}0}^{(1)} [Eq. (28)], substituting the solved δρ𝐤2(1)\delta\rho_{{\bf k}2}^{(1)} and δρ𝐤1(1)\delta\rho_{{\bf k}1}^{(1)} into Eq. (62) and keeping the lowest order of 𝐪{\bf q}, one obtains the solution of δρ𝐤3(1)\delta\rho_{{\bf k}3}^{(1)}:

δρ𝐤3(1)=[(ω24ξ𝐤2)/(4E𝐤2ω2)+1](2ρ𝐤1(0)μeff2ρ𝐤3(0)δΔ𝐤(1))/(2Δ𝐤)=4Δ𝐤/(4E𝐤2ω2)(ρ𝐤1(0)μeffρ𝐤3(0)δΔ𝐤(1)),\delta\rho_{{\bf k}3}^{(1)}=[(\omega^{2}-4\xi^{2}_{\bf k})/(4E_{\bf k}^{2}-\omega^{2})+1](2\rho^{(0)}_{{\bf k}1}\mu_{\rm eff}\!-\!2\rho^{(0)}_{{\bf k}3}\delta\Delta^{(1)}_{\bf k})/({2\Delta_{\bf k}})=4\Delta_{\bf k}/(4E_{\bf k}^{2}-\omega^{2})(\rho^{(0)}_{{\bf k}1}\mu_{\rm eff}\!-\!\rho^{(0)}_{{\bf k}3}\delta\Delta^{(1)}_{\bf k}), (67)

and then, the solution of δρ𝐤0(1)\delta\rho_{{\bf k}0}^{(1)} from Eq. (59) reads

iωδρ𝐤0(1)=i𝐯𝐤𝐪4Δ𝐤/(4E𝐤2ω2)(ρ𝐤1(0)μeffρ𝐤3(0)δΔ𝐤(1))e𝐄𝐤ρ𝐤3(0).i\omega\delta\rho^{(1)}_{{\bf k}0}={i{\bf v_{k}}\!\cdot\!{\bf q}}4\Delta_{\bf k}/(4E_{\bf k}^{2}-\omega^{2})(\rho^{(0)}_{{\bf k}1}\mu_{\rm eff}\!-\!\rho^{(0)}_{{\bf k}3}\delta\Delta^{(1)}_{\bf k})-e{\bf E}\!\cdot\!\partial_{\bf k}\rho^{(0)}_{{\bf k}3}. (68)

The second-order GIKE [Eq. (25)] in center-of-mass frequency and momentum space is written as

2iωδρ𝐤(2)+i[ξ𝐤τ3+Δ𝐤τ1,δρ𝐤(2)]+i[ps22mτ3+δΔ𝐤(2)τ1,ρ𝐤(0)]+12{e𝐄τ32i𝐩sτ3Δ𝐤τ1}+i2[𝐩s𝐩sΔ𝐤τ1,𝐤𝐤ρ𝐤(0)]=0,\displaystyle 2i\omega\delta\rho_{\bf k}^{(2)}+i[\xi_{\bf k}\tau_{3}+\Delta_{\bf k}\tau_{1},\delta\rho_{\bf k}^{(2)}]+i\Big{[}\frac{p_{s}^{2}}{2m}\tau_{3}+\delta\Delta_{\bf k}^{(2)}\tau_{1},\rho_{\bf k}^{(0)}\Big{]}+\frac{1}{2}\{e{\bf E}\tau_{3}-2i{\bf p}_{s}\tau_{3}\Delta_{\bf k}\tau_{1}\}+\frac{i}{2}[{\bf p}_{s}{\bf p}_{s}\Delta_{\bf k}\tau_{1},\partial_{\bf k}\partial_{\bf k}\rho_{\bf k}^{(0)}]=0,

in which we only keep the lowest order of 𝐪{\bf q} to consider a homogeneous influence/excitation from 𝐩s{\bf p}_{s}. The components of above equation are written as

2iωδρ𝐤0(2)=e𝐄𝐤δρ𝐤3(1),\displaystyle 2i\omega\delta\rho^{(2)}_{{\bf k}0}=-e{\bf E}\cdot\partial_{\bf k}\delta\rho^{(1)}_{{\bf k}3}, (70)
2iωδρ𝐤3(2)=2Δ𝐤δρ𝐤2(2)e𝐄𝐤δρ𝐤0(1),\displaystyle 2i\omega\delta\rho^{(2)}_{{\bf k}3}=2\Delta_{{\bf k}}\delta\rho^{(2)}_{{\bf k}2}-e{\bf E}\cdot{\partial_{\bf k}}\delta\rho^{(1)}_{{\bf k}0}, (71)
2iωδρ𝐤1(2)=2ξ𝐤δρ𝐤2(2),\displaystyle 2i\omega\delta\rho^{(2)}_{{\bf k}1}=-2\xi_{\bf k}\delta\rho_{{\bf k}2}^{(2)}, (72)
2iωδρ𝐤2(2)=2Δ𝐤δρ𝐤3(2)+2ξ𝐤δρ𝐤1(2)2ρ𝐤3(0)δΔ𝐤(2)+2ρ𝐤1(0)ps2/(2m)2Δ𝐤𝐩s𝐤δρ𝐤0(1)Δ𝐤(𝐩s𝐤)2ρ𝐤3(0).\displaystyle 2i\omega\delta\rho^{(2)}_{{\bf k}2}=-2\Delta_{{\bf k}}\delta\rho_{{\bf k}3}^{(2)}+2\xi_{\bf k}\delta\rho_{{\bf k}1}^{(2)}-2\rho^{(0)}_{{\bf k}3}\delta\Delta^{(2)}_{\bf k}+2\rho^{(0)}_{{\bf k}1}p_{s}^{2}/(2m)-2\Delta_{\bf k}{\bf p}_{s}\cdot{\partial_{\bf k}}\delta\rho^{(1)}_{{\bf k}0}-\Delta_{\bf k}({\bf p}_{s}\cdot\partial_{\bf k})^{2}\rho_{{\bf k}3}^{(0)}. (73)

Carrying Eqs. (71)-(72) and Eq. (59) into Eq. (73), one has

δρ𝐤2(2)=14E𝐤24ω2{2iω[2ρ𝐤1(0)μeff(2)2ρ𝐤3(0)δΔ𝐤(2)]+2iωΔ𝐤[(e𝐄iω𝐩s)𝐤]2ρ𝐤3(0)},\delta\rho^{(2)}_{{\bf k}2}=\frac{1}{4E_{\bf k}^{2}-4\omega^{2}}\Big{\{}2i\omega[2\rho^{(0)}_{{\bf k}1}\mu^{(2)}_{\rm eff}-2\rho^{(0)}_{{\bf k}3}\delta\Delta^{(2)}_{\bf k}]+2i\omega\Delta_{\bf k}\Big{[}\Big{(}\frac{e{\bf E}}{i\omega}-{\bf p}_{s}\Big{)}\cdot{\partial_{\bf k}}\Big{]}^{2}\rho_{{\bf k}3}^{(0)}\Big{\}}, (74)

and then, from Eq. (72), one obtains

δρ𝐤1(2)=ξ𝐤/(iω)4E𝐤24ω2{2iω[2ρ𝐤1(0)μeff(2)2ρ𝐤3(0)δΔ𝐤(2)]+2iωΔ𝐤[(e𝐄iω𝐩s)𝐤]2ρ𝐤3(0)}.\delta\rho^{(2)}_{{\bf k}1}=-\frac{\xi_{\bf k}/(i\omega)}{4E_{\bf k}^{2}-4\omega^{2}}\Big{\{}2i\omega[2\rho^{(0)}_{{\bf k}1}\mu^{(2)}_{\rm eff}-2\rho^{(0)}_{{\bf k}3}\delta\Delta^{(2)}_{\bf k}]+2i\omega\Delta_{\bf k}\Big{[}\Big{(}\frac{e{\bf E}}{i\omega}-{\bf p}_{s}\Big{)}\cdot{\partial_{\bf k}}\Big{]}^{2}\rho_{{\bf k}3}^{(0)}\Big{\}}. (75)

Appendix B Phase and gap equations as well as superfluid density

In this part, we present the derivation of the phase and gap equations as well as superfluid density. We first solve the Hartree field. Substituting ρ𝐤3(0)\rho_{{\bf k}3}^{(0)} [Eq. (24)] and the solved δρ𝐤3(1)\delta\rho_{{\bf k}3}^{(1)} [Eq. (67)] and Eq. (71) into Eq. (27), at low-frequency regime, one has

δn=2𝐤(ξ𝐤E𝐤μeff(1)ξ𝐤ρ𝐤3q0+Δ𝐤E𝐤2ρ𝐤1(0)μeff(1))+2iω𝐤Δ𝐤δρ𝐤2(2)/(iω),\delta{n}=2\sum_{\bf k}\Big{(}\frac{\xi_{\bf k}}{E_{\bf k}}\mu^{(1)}_{\rm eff}\partial_{\xi_{\bf k}}\rho_{{\bf k}3}^{q0}+\frac{\Delta_{\bf k}}{E_{\bf k}^{2}}\rho_{{\bf k}1}^{(0)}\mu^{(1)}_{\rm eff}\Big{)}+\frac{2}{i\omega}{\sum_{\bf k}}^{\prime}\Delta_{\bf k}\delta\rho_{{\bf k}2}^{(2)}/(i\omega), (76)

where ρ𝐤3q0=[f(E𝐤+)f(E𝐤)]/2\rho_{{\bf k}3}^{q0}=[{f(E^{+}_{\bf k})-f(E^{-}_{\bf k})}]/{2} represents the τ3\tau_{3} component of the equilibrium density of matrix in quasiparticle space. On the right-hand side of above equation, substituting Eq. (30) to replace Δ𝐤\Delta_{\bf k} in the last term, with Eq. (31), one has

δn\displaystyle\delta{n} =\displaystyle= 2𝐤[ξ𝐤E𝐤μeff(1)ξ𝐤ρ𝐤3q0+Δ𝐤E𝐤2ρ𝐤1(0)μeff(1)]2iω𝐤𝐤g𝐤𝐤ρ𝐤1δρ𝐤2(2)\displaystyle 2\sum_{\bf k}\Big{[}\frac{\xi_{\bf k}}{E_{\bf k}}\mu^{(1)}_{\rm eff}\partial_{\xi_{\bf k}}\rho_{{\bf k}3}^{q0}+\frac{\Delta_{\bf k}}{E_{\bf k}^{2}}\rho_{{\bf k}1}^{(0)}\mu^{(1)}_{\rm eff}\Big{]}-\frac{2}{i\omega}{\sum_{{\bf k}^{\prime}{\bf k}}}^{\prime}g_{\bf kk^{\prime}}\rho_{{\bf k^{\prime}}1}\delta\rho_{{\bf k}2}^{(2)} (77)
=\displaystyle= 2𝐤[ξ𝐤E𝐤μeff(1)ξ𝐤ρ𝐤3q0+Δ𝐤E𝐤2ρ𝐤1(0)μeff(1)]2iω𝐤ρ𝐤1[𝐤g𝐤𝐤δρ𝐤2(2)]\displaystyle 2\sum_{\bf k}\Big{[}\frac{\xi_{\bf k}}{E_{\bf k}}\mu^{(1)}_{\rm eff}\partial_{\xi_{\bf k}}\rho_{{\bf k}3}^{q0}+\frac{\Delta_{\bf k}}{E_{\bf k}^{2}}\rho_{{\bf k}1}^{(0)}\mu^{(1)}_{\rm eff}\Big{]}-\frac{2}{i\omega}{\sum_{{\bf k}^{\prime}}}^{\prime}\rho_{{\bf k^{\prime}}1}\Big{[}{\sum_{\bf k}}^{\prime}g_{\bf k^{\prime}k}\delta\rho_{{\bf k}2}^{(2)}\Big{]}
=\displaystyle= 2𝐤ξ𝐤(ξ𝐤F𝐤)μeff(1)=2Dμeff(1).\displaystyle 2\sum_{\bf k}\partial_{\xi_{\bf k}}(\xi_{\bf k}F_{\bf k})\mu^{(1)}_{\rm eff}=-2D\mu^{(1)}_{\rm eff}.

Consequently, Eq. (39) and Hartree field μH=2VqDμeff=2DVq(iωδθ/2+U)/(1+2DVq)\mu_{H}=-2V_{q}D\mu_{\rm eff}=-2DV_{q}(i\omega\delta\theta/2+U)/(1+2DV_{q}) are obtained. Then, the electric field reads

e𝐄=i𝐪(μH+U)=i𝐪(U2DVqiωθ/2)1+2DVqiω𝐩s,e{\bf E}=i{\bf q}(\mu_{H}+U)=\frac{i{\bf q}(U-2DV_{q}i\omega\theta/2)}{1+2DV_{q}}\approx{i\omega}{\bf p}_{s}, (78)

in which we have taken the long-wave approximation (DVq1DV_{q}\gg 1).

Substituting the solved δρ𝐤2(1)\delta\rho_{{\bf k}2}^{(1)} [Eq. (65)] and electric field [Eq. (78)] into Eq. (31), at low-frequency regime, one has

𝐤g𝐤𝐤{2iωμeffρ𝐤1(0)4E𝐤2+2(𝐯𝐤i𝐪)(𝐯𝐤𝐩s)4E𝐤2[Δ𝐤3E𝐤3ρ𝐤3q0Δ𝐤L𝐤+2Δ𝐤(ξ𝐤E𝐤)2E𝐤ρ𝐤3q012(Δ𝐤2ξ𝐤2ρ𝐤1(0)+Δ𝐤ξ𝐤ξ𝐤2ρ𝐤3(0))]}=0,{\sum_{\bf k}}^{\prime}g_{\bf k^{\prime}k}\Big{\{}2i\omega\mu_{\rm eff}\frac{\rho_{{\bf k}1}^{(0)}}{4E_{\bf k}^{2}}\!+\!\frac{2({\bf v_{k}}\!\cdot\!{i{\bf q}})({\bf v_{k}}\!\cdot\!{\bf p}_{s})}{4E_{\bf k}^{2}}\Big{[}\frac{\Delta_{\bf k}^{3}}{E_{\bf k}^{3}}\rho^{q0}_{{\bf k}3}\!-\!\Delta_{\bf k}L_{\bf k}\!+\!2\Delta_{\bf k}\Big{(}\frac{\xi_{\bf k}}{E_{\bf k}}\Big{)}^{2}\partial_{E_{\bf k}}\rho^{q0}_{{\bf k}3}\!-\!\frac{1}{2}(\Delta_{\bf k}^{2}\partial^{2}_{\xi_{\bf k}}\rho_{{\bf k}1}^{(0)}\!+\!\Delta_{\bf k}\xi_{\bf k}\partial^{2}_{\xi_{\bf k}}\rho_{{\bf k}3}^{(0)})\Big{]}\Big{\}}=0, (79)

where we have taken care of the particle-hole symmetry to remove terms with the odd order of ξ𝐤\xi_{\bf k} in the summation of 𝐤{\bf k}. Taking a generalized pairing potential g𝐤𝐤=gcos(ζθ𝐤+α)cos(ζθ𝐤+α)g_{\bf kk^{\prime}}=g\cos(\zeta\theta_{\bf k}+\alpha)\cos(\zeta\theta_{\bf k^{\prime}}+\alpha) which gives rise to Δ𝐤=Δcos(ζθ𝐤+α)\Delta_{{\bf k}}=\Delta\cos(\zeta\theta_{\bf k}+\alpha), the above equation becomes

2iωμeffZ1+i𝐪𝐩sZ2/m=0,2i\omega\mu_{\rm eff}Z_{1}+{i{\bf q}\cdot{\bf p}_{s}}Z_{2}/m=0, (80)

with Z1=𝐤Δ𝐤2E𝐤3ρ𝐤3q0𝐤ξ𝐤(ξ𝐤F𝐤)DZ_{1}={\sum_{\bf k}}^{\prime}\frac{\Delta_{\bf k}^{2}}{E_{\bf k}^{3}}\rho_{{\bf k}3}^{q0}\approx{\sum_{\bf k}}^{\prime}\partial_{\xi_{\bf k}}(\xi_{\bf k}F_{\bf k})\approx{-D} around Fermi surface and

Z2=kF2m𝐤Δ𝐤2E𝐤2[Δ𝐤2E𝐤3ρ𝐤3q0L𝐤+ξ𝐤2E𝐤2E𝐤ρ𝐤3q012(Δ𝐤ξ𝐤2ρ𝐤1(0)+ξ𝐤ξ𝐤2ρ𝐤3(0))]=kF2m𝐤Δ𝐤2E𝐤2(Δ𝐤2E𝐤3+ξ𝐤2E𝐤3E𝐤)ρ𝐤3q0=ns.Z_{2}=\!\frac{k_{F}^{2}}{m}{\sum_{\bf k}}^{\prime}\frac{\Delta_{\bf k}^{2}}{E_{\bf k}^{2}}\Big{[}\frac{\Delta_{\bf k}^{2}}{E_{\bf k}^{3}}\rho_{{\bf k}3}^{q0}\!-\!L_{\bf k}\!+\!\frac{\xi_{\bf k}^{2}}{E_{\bf k}^{2}}\partial_{E_{\bf k}}\rho_{{\bf k}3}^{q0}\!-\!\frac{1}{2}(\Delta_{\bf k}\partial^{2}_{\xi_{\bf k}}\rho_{{\bf k}1}^{(0)}\!+\!\xi_{\bf k}\partial^{2}_{\xi_{\bf k}}\rho_{{\bf k}3}^{(0)})\Big{]}\!=\!\frac{k_{F}^{2}}{m}{\sum_{\bf k}}^{\prime}\frac{\Delta_{\bf k}^{2}}{E_{\bf k}^{2}}\Big{(}\frac{\Delta_{\bf k}^{2}}{E_{\bf k}^{3}}\!+\!\frac{\xi_{\bf k}^{2}}{E_{\bf k}^{3}}\!-\!\partial_{E_{\bf k}}\Big{)}\rho_{{\bf k}3}^{q0}\!=\!-n_{s}. (81)

Then, Eq. (38) is obtained.

Substituting the solved δρ𝐤1(1)\delta\rho_{{\bf k}1}^{(1)} [Eq. (66)] into Eq. (30) and taking care of the particle-hole symmetry to remove terms with the odd order of ξ𝐤\xi_{\bf k} in the summation of 𝐤{\bf k}, at low-frequency regime, one has

δΔ𝐤(1)+𝐤g𝐤𝐤δΔ𝐤(1)ξ𝐤2E𝐤2[2ρ𝐤3(0)+(𝐪𝐤2)2ρ𝐤3(0)2(𝐯𝐤𝐪ω)2Δ𝐤2ρ𝐤3(0)E𝐤2]=i𝐪𝐩s2imω𝐤g𝐤𝐤{ξ𝐤E𝐤2[Δ𝐤ξ𝐤ξ𝐤ρ𝐤3(0)\displaystyle\delta\Delta_{\bf k^{\prime}}^{(1)}+\sum_{\bf k}g_{\bf k^{\prime}k}\delta\Delta_{\bf k}^{(1)}\frac{\xi_{\bf k}}{2E^{2}_{\bf k}}\Big{[}2\rho^{(0)}_{{\bf k}3}\!+\!\Big{(}\frac{{\bf q}\cdot\partial_{\bf k}}{2}\Big{)}^{2}\rho_{{\bf k}3}^{(0)}\!-\!2\Big{(}\frac{{\bf v_{k}}\cdot{\bf q}}{\omega}\Big{)}^{2}\frac{\Delta_{\bf k}^{2}\rho^{(0)}_{{\bf k}3}}{E^{2}_{\bf k}}\Big{]}=\frac{i{\bf q}\cdot{\bf p}_{s}}{2im\omega}\sum_{\bf k}g_{\bf k^{\prime}k}\Big{\{}\frac{-\xi_{\bf k}}{E^{2}_{\bf k}}\Big{[}\Delta_{\bf k}\xi_{\bf k}\partial_{\xi_{\bf k}}\rho^{(0)}_{{\bf k}3}
+Δ𝐤2ξ𝐤ρ𝐤1(0)+ξ𝐤ρ𝐤1(0)]+Δ𝐤ξ𝐤ρ𝐤3(0)+ρ𝐤1(0)}.\displaystyle\mbox{}\!+\!\Delta_{\bf k}^{2}\partial_{\xi_{\bf k}}\rho^{(0)}_{{\bf k}1}\!+\!\xi_{\bf k}\rho^{(0)}_{{\bf k}1}\Big{]}+\Delta_{{\bf k}}\partial_{\xi_{\bf k}}\rho^{(0)}_{{\bf k}3}+\rho^{(0)}_{{\bf k}1}\Big{\}}. (82)

Then, consider a homogeneous influence/excitation from 𝐩s{\bf p}_{s}, one finds a vanishing δΔ𝐤(1)\delta\Delta_{\bf k}^{(1)}. Moreover, with Eq. (78), substituting the solved δρ𝐤1(2)\delta\rho_{{\bf k}1}^{(2)} [Eq. (75)] into Eq. (30) and taking care of the particle-hole symmetry to remove terms with the odd order of ξ𝐤\xi_{\bf k} in the summation of 𝐤{\bf k}, one immediately finds δΔ𝐤(2)=0\delta\Delta_{\bf k}^{(2)}=0.

Furthermore, with δΔ𝐤(1)=0\delta\Delta_{\bf k}^{(1)}=0 and the solved Hartree field as well as Eq. (78), Eq. (68) at low-frequency regime becomes:

δρ𝐤0(1)=𝐯𝐤𝐩s𝐤ρ𝐤3(0),\delta\rho^{(1)}_{{\bf k}0}=-{{\bf v_{k}}\!\cdot\!{\bf p}_{s}}\partial_{\bf k}\rho^{(0)}_{{\bf k}3}, (83)

in which we have taken the long-wave approximation (DVq1DV_{q}\gg 1). Consequently, the current-related ρ𝐤0\rho_{{\bf k}0} [Eq. (28)] reads

ρ𝐤0=ρ𝐤0(0)+δρ𝐤0(1)=(𝐯𝐤𝐩s)(E𝐤ρ𝐤3q0𝐤ρ𝐤3(0))=(𝐯𝐤𝐩s)Δ𝐤2E𝐤2(E𝐤1E𝐤)ρ𝐤3q0=(𝐯𝐤𝐩s)Δ𝐤2E𝐤E𝐤F𝐤.\rho_{{\bf k}0}=\rho^{(0)}_{{\bf k}0}+\delta\rho^{(1)}_{{\bf k}0}=({{\bf v_{k}}\!\cdot\!{\bf p}_{s}})(\partial_{E_{\bf k}}\rho_{{\bf k}3}^{q0}-\partial_{\bf k}\rho^{(0)}_{{\bf k}3})=({{\bf v_{k}}\!\cdot\!{\bf p}_{s}})\frac{\Delta^{2}_{\bf k}}{E_{\bf k}^{2}}\Big{(}\partial_{E_{\bf k}}-\frac{1}{E_{\bf k}}\Big{)}\rho_{{\bf k}3}^{q0}=({{\bf v_{k}}\!\cdot\!{\bf p}_{s}})\frac{\Delta^{2}_{\bf k}}{E_{\bf k}}\partial_{E_{\bf k}}F_{\bf k}. (84)

Then, Eq. (33) is obtained.

Appendix C Derivation of Eq. (43)

In this part, we present the derivation of Eq. (43). With the action in Eq. (41), we first calculate the quantum fluctuation part in Eq. (42):

(psϕ)2|Q\displaystyle(p_{s}^{\phi})^{2}|_{\rm Q}\!\! =\displaystyle\!=\! qq2[|δθq𝐞ϕ(t+0+)2δθq𝐞ϕ(t)2eiS|=qq21Z0DθDθδθq𝐞ϕ(t+0+)2δθq𝐞ϕ(t)2eiS\displaystyle\!\!\sum_{q}q^{2}\Big{[}\Big{\langle}\Big{|}\frac{\delta\theta^{*}_{q{\bf e}_{\phi}}(t+0^{+})}{2}\frac{\delta\theta_{q{\bf e}_{\phi}}(t)}{2}e^{iS}\Big{|}\Big{\rangle}\!=\!\!\!\sum_{q}q^{2}\frac{1}{Z_{0}}{\int}D\theta{D\theta^{*}}\frac{\delta\theta^{*}_{q{\bf e}_{\phi}}(t+0^{+})}{2}\frac{\delta\theta_{q{\bf e}_{\phi}}(t)}{2}e^{iS} (85)
=\displaystyle= qi2q21Z0DθDθδJq𝐞ϕ(t+0+)δJq𝐞ϕ(t)ei[S+𝑑t𝐪(J𝐪θ𝐪/2+J𝐪θ𝐪/2)]|J=J=0\displaystyle\!\!\!\sum_{q}i^{2}q^{2}\frac{1}{Z_{0}}{\int}D\theta{D\theta^{*}}\delta_{J_{q^{*}{\bf e}_{\phi}}(t+0^{+})}\delta_{J_{q{\bf e}_{\phi}}(t)}e^{i[S+\int{dt^{\prime}}\sum_{\bf q^{\prime}}(J_{\bf q^{\prime}}\theta_{\bf q^{\prime}}/2+J_{\bf q^{\prime}}^{*}\theta^{*}_{\bf q^{\prime}}/2)]}\Big{|}_{J=J^{*}=0}
=\displaystyle= qi2q2δJq𝐞ϕ(t+0+)δJq𝐞ϕ(t)exp{i𝑑t𝐪(J𝐪+DqU𝐪t)D(t,𝐪)(J𝐪+DqU𝐪t)}|J=J=0\displaystyle\!\!\!\sum_{q}i^{2}q^{2}\delta_{J_{q^{*}{\bf e}_{\phi}}(t+0^{+})}\delta_{J_{q{\bf e}_{\phi}}(t)}\exp\Big{\{}i\int{dt^{\prime}}\sum_{\bf q^{\prime}}(J_{\bf q^{\prime}}+D_{q^{\prime}}U_{\bf q^{\prime}}^{*}\partial_{t^{\prime}})D(t^{\prime},{\bf q}^{\prime})(J^{*}_{\bf q^{\prime}}+D_{q^{\prime}}U_{\bf q^{\prime}}\partial_{t^{\prime}})\Big{\}}\Big{|}_{J=J^{*}=0}
=\displaystyle= qq2Trω[D2(t,q𝐞ϕ)|DqUq𝐞ϕt|2]=iCqωq2ω2Uq𝐞ϕUq𝐞ϕ(ω2ωN2+i0+)2.\displaystyle\!\!\!\sum_{q}q^{2}{\rm Tr}_{\omega}[D^{2}(t,q{\bf e}_{\phi})|D_{q}U_{q{\bf e}_{\phi}}\partial_{t}|^{2}]=\frac{i}{C}\sum_{q\omega}\frac{q^{2}\omega^{2}U_{q{\bf e}_{\phi}}U_{-q{\bf e}_{\phi}}}{(\omega^{2}\!-\!\omega_{N}^{2}+i0^{+})^{2}}.

Here, the Green function D(t,𝐪)=Dq1(t2+ωN2)1D(t,{\bf q})=D_{q}^{-1}(\partial_{t}^{2}+\omega_{N}^{2})^{-1}; J𝐪J_{\bf q} denotes the generating functional and δJ𝐪\delta_{J_{\bf q}} represents the functional derivativeFT .

We next derive the thermal fluctuation part in Eq. (42). By mapping the action in Eq. (41) into the imaginary-time one 𝒮\mathcal{S}:

𝒮=0β𝑑τ𝐪Dq[θ𝐪(τ2ωN2)θ𝐪/4iU𝐪τθ𝐪/2+iU𝐪τθ𝐪/2].\mathcal{S}=\int^{\beta}_{0}d\tau\sum_{\bf q}D_{q}\Big{[}{\theta_{\bf q}^{*}}(\partial_{\tau}^{2}-\omega_{N}^{2}){\theta_{\bf q}}/4-iU_{\bf q}^{*}\partial_{\tau}\theta_{\bf q}/2+iU_{\bf q}\partial_{\tau}\theta_{\bf q}^{*}/2\Big{]}. (86)

Since the Bosonic field θ(τ=β)=θ(τ=0)\theta(\tau=\beta)=\theta(\tau=0)G1 , the Josephson coupling terms (the last two terms) vanish in the imaginary-time space, i.e., the Josephson effect from electric potential, as a quantum effect, can not excite the imaginary-time fluctuation of the superconducting phase. The thermal fluctuation part in Eq. (42) then reads

(psϕ)2|T\displaystyle(p_{s}^{\phi})^{2}|_{\rm T}\!\! =\displaystyle\!=\! qq2[|δθq𝐞ϕ(τ)2δθq𝐞ϕ(τ)2e𝒮|]=qq2[1𝒵0DθDθδθq𝐞ϕ(τ)2δθq𝐞ϕ(τ)2e𝒮]\displaystyle\!\!\sum_{q}q^{2}\Big{[}\Big{\langle}\Big{|}\frac{\delta\theta^{*}_{q{\bf e}_{\phi}}(\tau)}{2}\frac{\delta\theta_{q{\bf e}_{\phi}}(\tau)}{2}e^{-\mathcal{S}}\Big{|}\Big{\rangle}\Big{]}\!=\!\!\!\sum_{q}q^{2}\Big{[}\frac{1}{\mathcal{Z}_{0}}{\int}D\theta{D\theta^{*}}\frac{\delta\theta^{*}_{q{\bf e}_{\phi}}(\tau)}{2}\frac{\delta\theta_{q{\bf e}_{\phi}}(\tau)}{2}e^{-\mathcal{S}}\Big{]} (87)
=\displaystyle= qq21𝒵0DθDθδJq𝐞ϕδJq𝐞ϕe[𝒮+𝑑τ𝐪(J𝐪θ𝐪/2+J𝐪θ𝐪/2)]|J=J=0\displaystyle\!\!\!\sum_{q}q^{2}\frac{1}{\mathcal{Z}_{0}}{\int}D\theta{D\theta^{*}}\delta_{J_{q^{*}{\bf e}_{\phi}}}\delta_{J_{q{\bf e}_{\phi}}}e^{-[\mathcal{S}+\int{d\tau}\sum_{\bf q^{\prime}}(J_{\bf q^{\prime}}\theta_{\bf q^{\prime}}/2+J_{\bf q^{\prime}}^{*}\theta^{*}_{\bf q^{\prime}}/2)]}\Big{|}_{J=J^{*}=0}
=\displaystyle= qq2δJq𝐞ϕδJq𝐞ϕexp{𝑑t𝐪J𝐪D(τ,𝐪)J𝐪}|J=J=0\displaystyle\!\!\!\sum_{q}q^{2}\delta_{J_{q^{*}{\bf e}_{\phi}}}\delta_{J_{q{\bf e}_{\phi}}}\exp\Big{\{}-\int{dt^{\prime}}\sum_{\bf q^{\prime}}J_{\bf q^{\prime}}D(\tau,{\bf q}^{\prime})J^{*}_{\bf q^{\prime}}\Big{\}}\Big{|}_{J=J^{*}=0}
=\displaystyle= qq2Trω[D(τ,q𝐞ϕ)]=1βqωnq2Dq1(iωn)2ωN2.\displaystyle\!\!-\!\sum_{q}q^{2}{\rm Tr}_{\omega}[D(\tau,q{\bf e}_{\phi})]=\!-\!\frac{1}{\beta}\sum_{q\omega_{n}}\frac{q^{2}}{D_{q}}\frac{1}{(i\omega_{n})^{2}\!-\!\omega_{N}^{2}}.

Consequently, with Eqs. (85) and (87), Eq. (43) is derived.

Appendix D Derivation of Eqs. (50) and (51)

In this part, we present the derivation of Eqs. (50) and (51). Following the standard Matsubara-frequency summationG1 , from Eq. (52) one has

Q1\displaystyle Q_{1} =\displaystyle= Vf+pTr[G0(p)τ3]=Vf+𝐤1βipn2ξ𝐤(ipnE𝐤+)(ipnE𝐤)=Vf+𝐤2ξ𝐤[f(E𝐤+)(E𝐤)]2E𝐤\displaystyle V_{f}+\sum_{p}{\rm Tr}[G_{0}(p)\tau_{3}]=V_{f}+\sum_{\bf k}\frac{1}{\beta}\sum_{ip_{n}}\frac{2\xi_{\bf k}}{(ip_{n}-E_{\bf k}^{+})(ip_{n}-E_{\bf k}^{-})}=V_{f}+\sum_{\bf k}\frac{2\xi_{\bf k}[f(E_{\bf k}^{+})-(E_{\bf k}^{-})]}{2E_{\bf k}} (88)
=\displaystyle= 𝐤(1+2ξ𝐤F𝐤)=kF2m𝐤ξ𝐤(ξ𝐤F𝐤),\displaystyle\sum_{\bf k}(1+2\xi_{\bf k}F_{\bf k})=-\frac{k_{F}^{2}}{m}{\sum_{\bf k}}^{\prime}\partial_{\xi_{\bf k}}\big{(}\xi_{\bf k}F_{\bf k}\big{)},

and

Q2=12pTr[G0(p)τ3G0(p)τ3]=𝐤1βipn(ipn𝐩s𝐯𝐤)2+ξ𝐤2Δ𝐤2(ipnE𝐤+)2(ipnE𝐤)2=𝐤η=±[E𝐤2+ξ𝐤2Δ𝐤24E𝐤2E𝐤ηf(E𝐤η)\displaystyle Q_{2}=-\frac{1}{2}\sum_{p}{\rm Tr}[G_{0}(p)\tau_{3}G_{0}(p)\tau_{3}]=-\sum_{\bf k}\frac{1}{\beta}\sum_{ip_{n}}\frac{(ip_{n}-{\bf p}_{s}\cdot{\bf v_{k}})^{2}+\xi_{\bf k}^{2}-\Delta_{\bf k}^{2}}{(ip_{n}-E_{\bf k}^{+})^{2}(ip_{n}-E_{\bf k}^{-})^{2}}=-\sum_{\bf k}\sum_{\eta=\pm}\Big{[}\frac{E_{\bf k}^{2}+\xi_{\bf k}^{2}-\Delta_{\bf k}^{2}}{4E_{\bf k}^{2}}\partial_{E_{\bf k}^{\eta}}f(E_{\bf k}^{\eta})
+η(2E𝐤4E𝐤2E𝐤2+ξ𝐤2Δ𝐤24E𝐤3)f(E𝐤η)]=𝐤{Δ𝐤2E𝐤2F𝐤+ξ𝐤2E𝐤2E𝐤[f(E𝐤+)f(E𝐤)2]}=𝐤ξ𝐤(ξ𝐤F𝐤).\displaystyle\mbox{}+\eta\Big{(}\frac{2E_{\bf k}}{4E_{\bf k}^{2}}-\frac{E_{\bf k}^{2}+\xi_{\bf k}^{2}-\Delta_{\bf k}^{2}}{4E_{\bf k}^{3}}\Big{)}f(E_{\bf k}^{\eta})\Big{]}=-{\sum_{\bf k}}\Big{\{}\frac{\Delta_{\bf k}^{2}}{E_{\bf k}^{2}}F_{\bf k}+\frac{\xi_{\bf k}^{2}}{E_{\bf k}^{2}}\partial_{E_{\bf k}}\Big{[}\frac{f(E_{\bf k}^{+})-f(E_{\bf k}^{-})}{2}\Big{]}\Big{\}}=-{\sum_{\bf k}}^{\prime}\partial_{\xi_{\bf k}}\big{(}\xi_{\bf k}F_{\bf k}\big{)}. (89)

Then, Eqs. (50) and (51) are obtained.

References

  • (1) J. Bardeen, L. N. Cooper, and J. R. Schrieffer, Phys. Rev. 106, 162 (1957).
  • (2) J. G. Bednorz and K. A. Müller, Z. Phys. B 64, 189 (1986).
  • (3) M. K. Wu, J. R. Ashburn, C. J. Torng, P. H. Hor, R. L. Meng, L. Gao, Z. J. Huang, Y. Q. Wang, and C. W. Chu, Phys. Rev. Lett. 58, 908 (1987).
  • (4) T. Timusk and B. Statt, Rep. Prog. Phys. 62, 61 (1999).
  • (5) A. Damascelli, Z. Hussain, and Z. X. Shen, Rev. Mod. Phys. 75, 473 (2003).
  • (6) B. Keimer, S. A. Kivelson, M. R. Norman, S. Uchida, and J. Zaanen, Nature 518, 179 (2015).
  • (7) H. Ding, T. Yokoya, J. C. Campuzano, T. Takahashi, M. Randeria, M. R. Norman, T. Mochiku, K. Kadowaki, and J. Giapintzakis, Nature (London) 382, 51 (1996).
  • (8) A. G. Loeser, Z. X. Shen, D. S. Dessau, D. S. Marshall, C. H. Park, P. Fournier, and A. Kapitulnik, Science 273, 325 (1996).
  • (9) M. R. Norman, H. Ding, M. Randeria, J. C. Campuzano, T. Yokoya, T. Takeuchi, T. Takahashi, T. Mochiku, K. Kadowaki, P. Guptasarma, and D. G. Hinks, Nature (London) 392, 157 (1998).
  • (10) T. Valla, A. V. Fedorov, J. Lee, J. C. Davis, and G. D. Gu, Science 314, 1914 (2006).
  • (11) K. K. Gomes, A. N. Pasupathy, A. Pushp, S. Ono, Y. Ando, and A. Yazdani, Nature 447, 569 (2007).
  • (12) J. W. Alldredge, J. Lee, K. McElroy, M. Wang, K. Fujita, Y. Kohsaka, C. Taylor, H. Eisaki, S. Uchida, P. J. Hirschfeld, and J. C. Davis, Nat. Phys. 4, 319 (2008).
  • (13) A. N. Pasupathy, A. Pushp, K. K. Gomes, C. V. Parker, J. S. Wen, Z. Xu, G. Gu, S. Ono, Y. Ando, and A. Yazdani, Science 320, 196 (2008).
  • (14) R. Daou, N. D. Leyraud, D. LeBoeuf, S. Y. Li, F. Laliberté, O. C. Choiniére, Y. J. Jo, L. Balicas, J.-Q. Yan, J.-S. Zhou, J. B. Goodenough, and L. Taillefer, Nat. Phys. 5, 31 (2009).
  • (15) C. Collignon, S. Badoux, S. A. A. Afshar, B. Michon, F. Laliberté, O. C. Choiniére, J.-S. Zhou, S. Licciardello, S. Wiedmann, N. D. Leyraud, and L. Taillefer, Phys. Rev. B 95, 224517 (2017).
  • (16) S. Badoux, W. Tabis, F. Laliberté, G. Grissonnanche, B. Vignolle, D. Vignolles, J. Bérd, D. A. Bonn, W. N. Hardy, R. Liang, N. D. Leyraud, L. Taillefer, and C. Proust, Nature 531, 210 (2016).
  • (17) N. D. Leyraud, O. C. Choiniére, S. Badoux, A. Ataei, C. Collignon, A. Gourgout, S. D. Beauséjour, F. F. Tafti, F. Laliberté, M. E. Boulanger, M. Matusiak, D. Graf, M. Kim, J.-S. Zhou, N. Momono, T. Kurosawa, H. Takagi, and L. Taillefer, Nat. Commun. 8, 2044 (2017).
  • (18) Z. A. Xu, N. P. Ong, Y. Wang, T. Kakeshita, and S. Uchida, Nature (London) 406, 486 (2000).
  • (19) Y. Wang, Z. A. Xu, T. Kakeshita, S. Uchida, S. Ono, Y. Ando, and N. P. Ong, Phys. Rev. B 64, 224519 (2001)
  • (20) Y. Wang, N. P. Ong, Z. A. Xu, T. Kakeshita, S. Uchida, D. A. Bonn, R. Liang, and W. N. Hardy, Phys. Rev. Lett. 88, 257003 (2002).
  • (21) Y. Wang, S. Ono, Y. Onose, G. Gu, Y. Ando, Y. Tokura, S. Uchida, and N. P. Ong, Science 299, 86 (2003).
  • (22) Y. Wang, L. Li, and N. P. Ong, Phys. Rev. B 73, 024510 (2006).
  • (23) R. Daou, J. Chang, D. LeBoeuf, O. C. Choiniére, F. Laliberté, N. D. Leyraud, B. J. Ramshaw, R. Liang, D. A. Bonn, W. N. Hardy, and L. Taillefer, Nature (London) 463, 519 (2010).
  • (24) Y. J. Uemura, G. M. Luke, B. J. Sternlieb, J. H. Brewer, J. F. Carolan, W. N. Hardy, R. Kadono, J. R. Kempton, R. F. Kiefl, S. R. Kreitzman, P. Mulhern, T. M. Riseman, D. Ll. Williams, B. X. Yang, S. Uchida, H. Takagi, J. Gopalakrishnan, A. W. Sleight, M. A. Subramanian, C. L. Chien, M. Z. Cieplak, G. Xiao, V. Y. Lee, B. W. Statt, C. E. Stronach, W. J. Kossler, and X. H. Yu, Phys. Rev. Lett. 62, 2317 (1989).
  • (25) Y. J. Uemura, L. P. Le, G. M. Luke, B. J. Sternlieb, W. D. Wu, J. H. Brewer, T. M. Riseman, C. L. Seaman, M. B. Maple, M. Ishikawa, D. G. Hinks, J. D. Jorgensen, G. Saito, and H. Yamochi, Phys. Rev. Lett. 66, 2665 (1992).
  • (26) V. J. Emery and S. A. Kivelson, Nature 374, 434 (1995).
  • (27) Y. Nambu, Phys. Rev. 117, 648 (1960).
  • (28) J. Goldstone, Nuovo Cimento 19, 154 (1961).
  • (29) J. Goldstone, A. Salam, and S. Weinberg, Phys. Rev. 127, 965 (1962).
  • (30) V. Ambegaokar and L. P. Kadanoff, Nuovo Cimento 22, 914 (1961).
  • (31) J. R. Schrieffer, Theory of Superconductivity (W. A. Benjamin, New York, 1964).
  • (32) P. B. Littlewood and C. M. Varma, Phys. Rev. Lett. 47, 811 (1981); Phys. Rev. B 26, 4883 (1982).
  • (33) P. W. Anderson, Phys. Rev. 130, 439 (1963).
  • (34) M. Randeria, N. Trivedi, A. Moreo, and R. T. Scalettar, Phys. Rev. Lett. 69, 2001 (1992).
  • (35) V. B. Geshkenbein, L. B. Iofe, and A. I. Larkin, Phys. Rev. B 55, 3173 (1997).
  • (36) H. B. Yang, J. D. Rameau, P. D. Johnson, T. Valla, A. Tsvelik, and G. D. Gu, Nature (London) 456, 77 (2008).
  • (37) Y. I. Seo, W. J. Choi, S. I. Kimura, Y. Bang, and Y. S. Kwon, Phys. Rev. B 95, 094510 (2017).
  • (38) Y. I. Seo, W. J. Choi, S. I. Kimura, and Y. S. Kwon, Sci. Rep. 9, 3987 (2019).
  • (39) L. Li, Y. Wang, M. J. Naughton, S. Ono, Y. Ando, and N. P. Ong, Europhys. Lett. 72, 451 (2005).
  • (40) Y. Wang, L. Li, M. J. Naughton, G. D. Gu, S. Uchida, and N. P. Ong, Phys. Rev. Lett. 95, 247002 (2005).
  • (41) L. Li, Y. Wang, S. Komiya, S. Ono, Y. Ando, G. D. Gu, and N. P. Ong, Phys. Rev. B 81, 054510 (2010).
  • (42) G. Yu, D.-D. Xia, D. Pelc, R.-H. He, N.-H. Kaneko, T. Sasagawa, Y. Li, X. Zhao, N. Barišić, A. Shekhter, and M. Greven, Phys. Rev. B 99, 214502 (2019).
  • (43) J. L. Tallon, J. G. Storey, and J. W. Loram, Phys. Rev. B 83, 092502 (2011).
  • (44) Y. Ando, S. Komiya, K. Segawa, S. Ono, and Y. Kurita, Phys. Rev. Lett. 93, 267001 (2004).
  • (45) O. C.-Choiniére, R. Daou, F. Laliberté, C. Collignon, S. Badoux, D. LeBoeuf, J. Chang, B. J. Ramshaw, D. A. Bonn, W. N. Hardy, R. Liang, J.-Q. Yan, J.-G. Cheng, J.-S. Zhou, J. B. Goodenough, S. Pyon, T. Takayama, H. Takagi, N. D. Leyraud, and L. Taillefer, Phys. Rev. B 97, 064502 (2018).
  • (46) W. Zhang, C. L. Smallwood, C. Jozwiak, T. L. Miller, Y. Yoshida, H. Eisaki, D.-H. Lee, and A. Lanzara, Phys. Rev. B 88, 245132 (2013).
  • (47) L. Perfetti, B. Sciolla, G. Biroli, C. J. Van der Beek, C. Piovera, M. Wolf, and T. Kampfrath, Phys. Rev. Lett. 114, 067003 (2015).
  • (48) T. Kondo, Y. Hamaya, A. D. Palczewski, T. Takeuchi, J. S. Wen, Z. J. Xu, G. Gu, J. Schmalian, and A. Kaminski, Nat. Phys. 7, 21 (2011).
  • (49) T. Kondo, A. D. Palczewski, Y. Hamaya, T. Takeuchi, J. S. Wen, Z. J. Xu, G. D. Gu, and A. Kaminski, Phys. Rev. Lett. 111, 157003 (2013).
  • (50) T. Kondo, W. Malaeb, Y. Ishida, T. Sasagawa, H. Sakamoto, T. Takeuchi, T. Tohyama, and S. Shin, Nat. Commun. 6, 7699 (2015).
  • (51) S. D. Chen, M. Hashimoto, Y. He, D. Song, K.-J. Xu, J.-F. He, T. P. Devereaux, H. Eisaki, D.-H. Lu, J. Zaanen, and Z. X. Shen, Science 366, 1099 (2019).
  • (52) A. Dubroka, M. Rössle, K. W. Kim, V. K. Malik, D. Munzar, D. N. Basov, A. A. Schafgans, S. J. Moon, C. T. Lin, D. Haug, V. Hinkov, B. Keimer, T. Wolf, J. G. Storey, J. L. Tallon, and C. Bernhard, Phys. Rev. Lett. 106, 047006 (2011).
  • (53) E. Uykur, K. Tanaka, T. Masui, S. Miyasaka, and S. Tajima, Phys. Rev. Lett. 112, 127003 (2014).
  • (54) K. Lee, K. Kamiya, M. Nakajima, S. Miyasaka, and S. Tajima, J. Phys. Soc. Jpn. 86, 023701 (2017).
  • (55) H. Kitano, T. Ohashi, A. Maeda, and I. Tsukada, Phys. Rev. B 73, 092504 (2006).
  • (56) M. S. Grbić, N. Barišić, A. Dulčić, I. Kupčić, Y. Li, X. Zhao, G. Yu, M. Dressel, M. Greven, and M. Požek, Phys. Rev. B 80, 094511 (2009).
  • (57) M. S. Grbić, M. Požek, D. Paar, V. Hinkov, M. Raichle, D. Haug, B. Keimer, N. Barišić, and A. Dulčić, Phys. Rev. B 83, 144508 (2011).
  • (58) J. Corson, R. Mallozzi, J. Orenstein, J. N. Eckstein, and I. Božović, Nature (London) 398, 221 (1999).
  • (59) J. Orenstein, J. Corson, S. Oh, and J. N. Eckstein, Ann. Phys. 15, 596 (2006).
  • (60) L. S. Bilbro, R. V. Aguilar, G. Logvenov, O. Pelleg, I. Božović, and N. P. Armitage, Nat. Phys. 7, 298 (2011).
  • (61) D. Nakamura, Y. Imai, A. Maeda, and I. Tsukada, J. Phys. Soc. Jpn. 81, 044709 (2012).
  • (62) K. Katsumi, N. Tsuji, Y. I. Hamada, R. Matsunaga, J. Schneeloch, R. D. Zhong, G. D. Gu, H. Aoki, Y. Gallais, and R. Shimano, Phys. Rev. Lett. 120, 117001 (2018).
  • (63) H. Chu, M. J. Kim, K. Katsumi, S. Kovalev, R. D. Dawson, L. Schwarz, N. Yoshikawa, G. Kim, D. Putzky, Z. Z. Li, H. Raffy, S. Germanskiy, J. C. Deinert, N. Awari, I. Ilyakov, B. Green, M. Chen, M. Bawatna, G. Christiani, G. Logvenov, Y. Gallais, A. V. Boris, B. Keimer, A. Schnyder, D. Manske, M. Gensch, Z. Wang, R. Shimano, and S. Kaiser, Nat. Commun. 11, 1793 (2020).
  • (64) K. Katsumi, Z. Z. Li, H. Raffy, Y. Gallais, and R. Shimano, Phys. Rev. B 102, 054510 (2020).
  • (65) J. M. Wade, J. W. Loram, K. A. Mirza, J. R. Cooper, and J. L. Tallon, J. Supercond. 7, 261 (1994).
  • (66) F. Mahmood, X. He, I. Božović, and N. P. Armitage, Phys. Rev. Lett. 122, 027003 (2019).
  • (67) In the original phenomenological two-fluid model which was first proposed by Tisza and LondonTisza in superconductors and then developed by LandauLandau in Bosonic liquid helium, the superfluid and normal fluid in superconductors describe the excitations in superconducting and normal statesBa0 , respectively. From our recent study in microscopic viewGIKE1 , we followed the idea of Fulde-Ferrell-Larkin-Ovchinnikov stateFF1 ; FF2 and then divided the BCS spherical shell in momentum space into the pairing and unpairing regions with finite and vanishing anomalous correlation, respectively. It was further shown that the particles in unpairing region behave like normal ones, whereas the dynamics in the pairing region consists of the contributions from Meissner density and Bogoliubov quasiparticlesGIKE1 . We therefore considered the excitations in the pairing and unpairing regions as superfluid and normal fluid, and then, recovered the optical conductivity of the two-fluid model at low temperatureGIKE1 .
  • (68) I. Ussishkin, S. L. Sondhi, and D. A. Huse, Phys. Rev. Lett. 89, 287001 (2002).
  • (69) M. N. Serbyn, M. A. Skvortsov, A. A. Varlamov, and V. Galitski, Phys. Rev. Lett. 102, 067001 (2009).
  • (70) A. I. Larkin and A. A. Varlamov, Theory of Fluctuations in Superconductors, (Oxford University Press, New York, 2002).
  • (71) P. Curty and H. Beck, Phys. Rev. Lett. 91, 257002 (2003).
  • (72) Z. X. Li, S. A. Kivelson, and D. H. Lee, npj Quantum Mater. 6, 36 (2021).
  • (73) F. Yang and M. W. Wu, Phys. Rev. B 98, 094507 (2018).
  • (74) F. Yang and M. W. Wu, Phys. Rev. B 100, 104513 (2019).
  • (75) F. Yang and M. W. Wu, Phys. Rev. B 102, 144508 (2020).
  • (76) F. Yang and M. W. Wu, Phys. Rev. B 102, 014511 (2020).
  • (77) A. A. Abrikosov, L. P. Gor’kov, and I. E. Dzyaloshinski, Methods of Quantum Field Theory in Statistical Physics (Prentice Hall, Englewood Cliffs, 1963).
  • (78) I. Khavkine, H. Y. Kee, and K. Maki, Phys. Rev. B 70, 184521 (2004).
  • (79) G. Tkachov and V. I. Fal’ko, Phys. Rev. B 69, 092503 (2004).
  • (80) F. Rohlfing, G. Tkachov, F. Otto, K. Richter, D. Weiss, G. Borghs, and C. Strunk, Phys. Rev. B 80, 220507(R) (2009).
  • (81) B. D. Josephson, Rev. Mod. Phys. 46, 251 (1974).
  • (82) P. Fulde and R. A. Ferrell, Phys. Rev. 135, 550 (1964).
  • (83) A. I. Larkin and Y. N. Ovchinnikov, Zh. Eksp. Teor. Fiz. 47, 1136 (1964) [Sov. Phys. JETP 20, 762 (1965)].
  • (84) T. Yu and M. W. Wu, Phys. Rev. B 94, 205305 (2016).
  • (85) F. Yang and M. W. Wu, Phys. Rev. B 95, 075304 (2017).
  • (86) F. Yang and M. W. Wu, J. Low Temp. Phys. 192, 241 (2018).
  • (87) The translational symmetry leads to g𝐤𝐤=g𝐤𝐤g_{\bf kk^{\prime}}=g_{\bf k-k^{\prime}}, whereas g𝐤𝐤=g𝐤𝐤g_{\bf k-k^{\prime}}=g_{\bf k^{\prime}-k} by time-reversal symmetry.
  • (88) J. Wu, A. T. Bollinger, X. He, and I. Božović, Nature 547, 432 (2017).
  • (89) Y. Dagan, R. Krupke, and G. Deutscher, Phys. Rev. B 62, 146 (2000).
  • (90) S. E. Sebastian, N. Harrison, M. M. Altarawneh, C. H. Mielke, R. Liang, D. A. Bonn, W. N. Hardy, and G. G. Lonzarich, Proc. Natl. Acad. Sci. USA 107, 6175 (2010).
  • (91) B. J. Ramshaw, S. E. Sebastian, R. D. McDonald, J. Day, B. S. Tan, Z. Zhu, J. B. Betts, R. Liang, D. A. Bonn, W. N. Hardy, and N. Harrison, Science 348, 317 (2015).
  • (92) The prefactor 1/21/2 in Nambu-space BdG Hamiltonian [Eq. (1)] in Refs. GIKE1, and GIKE2, should be removed.
  • (93) H. Haug and A. P. Jauho, Quantum Kinetics in Transport and Optics of Semiconductors (Springer, Berlin, 1996).
  • (94) M. W. Wu, J. H. Jiang, and M. Q. Weng, Phys. Rep. 493, 61 (2010).
  • (95) S. Tsuchiya, D. Yamamoto, R. Yoshii, and M. Nitta, Phys. Rev. B 98, 094503 (2018).
  • (96) M. E. Peskin and D. V. Schroeder, An Introduction to Quantum Field Theory (Addison-Wesley, New York, 1995).
  • (97) I. J. R. Aitchison, P. Ao, D. J. Thouless, and X. M. Zhu, Phys. Rev. B 51, 6531 (1995).
  • (98) I. J. R. Aitchison, G. Metikas, and D. J. Lee, Phys. Rev. B 62, 6638 (2000).
  • (99) S. Y. Qin, J. Kim, Q. Niu, and C. K. Shih, Science 324, 1314 (2009).
  • (100) T. Zhang, P. Cheng, W. J. Li, Y. J. Sun, G. Wang, X. G. Zhu, K. He, L. L. Wang, X. C. Ma, X. Chen, Y. Y. Wang, Y. Liu, H. Q. Lin, J. F. Jia, and Q. K. Xue, Nat. Phys. 6, 104 (2010).
  • (101) M. H. Liao, Y. Y. Zang, Z. Y. Guan, H. W. Li, Y. Gong, K. J. Zhu, X. P. Hu, D. Zhang, Y. Xu, Y. Y. Wang, K. He, X. C. Ma, S. C. Zhang, and Q. K. Xue, Nat. Phys. 14, 344 (2018).
  • (102) H. Nam, H. Chen, P. W. Adams, S. Y. Guan, T. M. Chuang, C. S. Chang, A. H. MacDonald, and C. K. Shih, Nat. Commun. 9, 5431 (2018).
  • (103) M. Liu, H. Nam, J. Kim, G. A. Fiete, and C. K. Shih, arXiv:2008.12733.
  • (104) M. Mondal, A. Kamlapure, M. Chand, G. Saraswat, S. Kumar, J. Jesudasan, L. Benfatto, V. Tripathi, and P. Raychaudhuri, Phys. Rev. Lett. 106, 047001 (2011).
  • (105) M. Chand, G. Saraswat, A. Kamlapure, M. Mondal, S. Kumar, J. Jesudasan, V. Bagwe, L. Benfatto, V. Tripathi, and P. Raychaudhuri, Phys. Rev. B 85, 014508 (2012).
  • (106) M. Mondal, A. Kamlapure, S. C. Ganguli, J. Jesudasan, V. Bagwe, L. Benfatto, and P. Raychaudhuri, Sci. Rep. 3, 1357 (2013).
  • (107) T. Dubouchet, B. Sacépé, J. Seidemann, D. Shahar, M. Sanquer, and C. Chapelier, Nat. Phys. 15, 233 (2019).
  • (108) J. Chang, E. Blackburn, A. T. Holmes, N. B. Christensen, J. Larsen, J. Mesot, R. X. Liang, D. A. Bonn, W. N. Hardy, A. Watenphul, M. v. Zimmermann, E. M. Forgan, and S. M. Hayden, Nat. Phys. 8, 871 (2012).
  • (109) R. Comin, A. Frano, M. M. Yee, Y. Yoshida, H. Eisaki, E. Schierle, E. Weschke, R. Sutarto, F. He, A. Soumyanarayanan, Y. He, M. L. Tacon, I. S. Elfimov, J. E. Hoffman, G. A. Sawatzky, B. Keimer, and A. Damascelli, Sci. Rep. 343, 390 (2014).
  • (110) X. M. Chen, C. Mazzoli, Y. Cao, V. Thampy, A. M. Barbour, W. Hu, M. Lu, T. A. Assefa, H. Miao, G. Fabbris, G. D. Gu, J. M. Tranquada, M. P. M. Dean, S. B. Wilkins, and I. K. Robinson, Nat. Commun. 10, 1435 (2019).
  • (111) H. Miao, G. Fabbris, R. J. Koch, D. G. Mazzone, C. S. Nelson, R. A. Esteves, G. D. Gu, Y. Li, T. Yilimaz, K. Kaznatcheev, E. Vescovo, M. Oda, T. Kurosawa, N. Momono, T. Assefa, I. K. Robinson, E. S. Bozin, J. M. Tranquada, P. D. Johnson, and M. P. M. Dean, npj Quantum Mater. 6, 31 (2021).
  • (112) J. M. Tranquada, H. Woo, T. G. Perring, H. Goka, G. D. Gu, G. Xu, M. Fujita, and K. Yamada, Nature 429, 534 (2004).
  • (113) S. M. Hayden, H. A. Mook, P. C. Dai, T. G. Perring, and F. Doǧan, Nature 429, 531 (2004).
  • (114) B. Khaykovich, S. Wakimoto, R. J. Birgeneau, M. A. Kastner, Y. S. Lee, P. Smeibidl, P. Vorderwisch, and K. Yamada, Phys. Rev. B 71, 220508 (2005).
  • (115) C. V. Parker, P. Aynajian, E. H. D. S. Neto, A. Pushp, S. Ono, J. S. Wen, Z. J. Xu, G. Gu, and A. Yazdani, Nature 468, 677 (2010).
  • (116) M. J. Lawler, K. Fujita, J. Lee, A. R. Schmidt, Y. Kohsaka, C. K. Kim, H. Eisaki, S. Uchida, J. C. Davis, J. P. Sethna, and E.-A. Kim, Nature 466, 347 (2010).
  • (117) E. H. D. S. Neto, P. Aynajian, R. E. Baumbach, E. D. Bauer, J. Mydosh, S. Ono, and A. Yazdani, Phys. Rev. B 87, 161117 (2013).
  • (118) Z. Y. Weng and V. N. Muthukumar, Phys. Rev. B 66, 094509 (2002).
  • (119) J.-H. Zhang, S. Li, Y. Ma, Y. G. Zhong, H. Ding, and Z. Y. Weng, Phys. Rev. Res. 2, 023398 (2020).
  • (120) J. Y. Zhao, S. A. Chen, H. K. Zhang, Z. Y. Weng, arXiv:2106.14898.
  • (121) Y. He, S. D. Chen, Z. X. Li, D. Zhao, D. Song, Y. Yoshida, H. Eisaki, T. Wu, X. H. Chen, D. H. Lu, C. Meingast, R. J. Birgeneau, T. P. Devereaux, M. Hashimoto, D. H. Lee, and Z. X. Shen, arXiv:2009.10932.
  • (122) N. R. Lee-Hone, J. S. Dodge, and D. M. Broun, Phys. Rev. B 96, 024501 (2017).
  • (123) N. R. Lee-Hone, H. U. Özdemir, V. Mishra, D. M. Broun, and P. J. Hirschfeld, Phys. Rev. Res. 2, 013228 (2020).
  • (124) P. W. Anderson, J. Phys. Chem. Solids 11, 26 (1959).
  • (125) H. Suhl and B. T. Matthias, Phys. Rev. 114, 977 (1959).
  • (126) S. Skalski, O. B. Matibet, and P. R. Weiss, Phys. Rev. 136, A1500 (1964).
  • (127) L. Tisza, C. R. Acad. Sci. 207, 1035 (1938); 207, 1186 (1938); F. London, Phys. Rev. 54, 947 (1938).
  • (128) L. D. Landau, Zh. Eksp. Teor. Fiz. 11, 592 (1941); J. Phys. USSR 5, 71 (1941).