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

Topological charge pumping in excitonic insulators

Zhiyuan Sun Department of Physics, Columbia University, 538 West 120th Street, New York, New York 10027    Andrew J. Millis Department of Physics, Columbia University, 538 West 120th Street, New York, New York 10027 Center for Computational Quantum Physics, Flatiron Institute, 162 5th Avenue, New York, NY 10010
Abstract

We show that in excitonic insulators with ss-wave electron-hole pairing, an applied electric field (either pulsed or static) can induce a pp-wave component to the order parameter, and further drive it to rotate in the s+ips+ip plane, realizing a Thouless charge pump. In one dimension, each cycle of rotation pumps exactly two electrons across the sample. Higher dimensional systems can be viewed as a stack of one dimensional chains in momentum space in which each chain crossing the fermi surface contributes a channel of charge pumping. Physics beyond the adiabatic limit, including in particular dissipative effects is discussed.

Controlling many-body systems, and in particular using appropriately applied external fields to ‘steer’ order parameters of symmetry broken phases, has emerged as a central theme in current physics Kirilyuk et al. (2010); Mentink et al. (2015); Byrnes et al. (2014); Zhang and Averitt (2014); Basov et al. (2017); Golež et al. (2016); Claassen et al. (2019); Sun and Millis (2020a). The excitonic insulator (EI) is state of matter first proposed in the 1960s Mott (1961); Kozlov and Maksimov (1965); Jérome et al. (1967); Portengen et al. (1996) with an order parameter defined as a condensate of bound electron hole pairs that activates a hybridization between two otherwise (in the simplest case) decoupled bands and opens a gap in the electronic spectrum. Several candidate materials including electron-hole bilayers Fogler et al. (2014); Li et al. (2017); Du et al. (2017), Ta2NiSe5 Lu et al. (2017); Werdehausen et al. (2018); Wakisaka et al. (2009); Kaneko et al. (2013); Sugimoto et al. (2018); Mazza et al. (2020), 1T1T-TiSe2 Kogar et al. (2017); Cercellier et al. (2007); Kaneko et al. (2018); Chen et al. (2018) and monolayer WTe2 Jia et al. (2020) are objects of current intensive study; recent work Du et al. (2017); Hu et al. (2018); Wang et al. (2019); Hu et al. (2020); Varsano et al. (2020); Perfetto and Stefanucci (2020); Hou et al. (2019) has pointed out their possible topological aspects. While the early theories of EI considered a one component order parameter, typically of inversion symmetric ss-wave type, realistic interactions also allow for electron-hole pairing in sub-dominant channels including pp-wave (inversion-odd) ones. In equilibrium, the ss-wave ground state is favored, with the potential for pp-wave order revealed by its fluctuations accompanied by dipole moment oscillations: the ‘Bardasis-Schrieffer’ collective mode Sun and Millis (2020b).

In this paper we show that applied electric fields can steer the order parameter to rotate in the space of s and p symmetry components, as shown in Fig. 1(a), leading to a realization of the ‘Thouless charge pump’ Thouless (1983); Rice and Mele (1982); King-Smith and Vanderbilt (1993); Zhang et al. (2020), providing quantized charge transport across an insulating sample.

The minimal model of an EI involves two electron bands shown in Fig. 1(b): a valence band with energy ξv,k\xi_{v,k} that disperses downwards from a high symmetry point (taken to have zero momentum) and a conduction band (ξc,k\xi_{c,k}) that disperses upwards. For simplicity we assume that their energies are equal and opposite (ξc=ξv=ξ\xi_{c}=-\xi_{v}=\xi). Relaxing this assumption does not change our results in an essential way. Defining the overlap G=2ξv,0G=2\xi_{v,0}, we distinguish the ‘BCS’ case G>0G>0 where the two bands cross at a fermi wavevector kFk_{F} with fermi velocity vFv_{F} as shown by the dashed lines, leading to electron and hole pockets, and the ‘BEC’ case where G<0G<0 and the bands do not cross. Excitonic order corresponds to the spontaneous formation of a hybridization between the two bands due to the electron-electron interaction VV, leading to an order parameter Δ(k)=kVkkψc,kψv,k+c.c.\Delta(k)=\sum_{k^{\prime}}V_{k-k^{\prime}}\left<\psi^{\dagger}_{c,k^{\prime}}\psi_{v,k^{\prime}}\right>+c.c. where ψc/v,k\psi_{c/v,k} is the electron annihilation operator at momentum kk of the conduction/valence band and VqV_{q} is the Fourier transform of the density-density interaction potential V(r)V(r). The ss-wave order parameter Δs(k)\Delta_{s}(k) is invariant under crystal symmetry operations while pp-wave order parameters are odd under inversion: Δp(k)=Δp(k)\Delta_{p}(k)=-\Delta_{p}(-k), and often transform as a multi-dimensional representation of the crystal symmetry group. For simplicity we neglect the kk-dependence of Δs\Delta_{s}, and define Δp(k)=Δpfk\Delta_{p}(k)=\Delta_{p}f_{k} where the pairing function fkf_{k} carries the momentum dependence and satisfies max(|fkF|)=1\text{max}(|f_{k_{F}}|)=1. We focus on the pxp_{x} pairing channel which is induced by the x-direction electric fields we consider here. While the qualitative conclusions hold generically for all spatial dimensions, we will indicate the dimensionality if a specific d-dimensional system is discussed where the momentum kk means a d-dimensional vector.

Refer to caption
Figure 1: (a) The s+ips+ip plane for the excitonic order parameter, with electric field-driven evolution shown as dashed line. (b) The quasiparticle dispersion in a one dimensional EI (solid lines) along with bands in metallic phase (dashed lines). (c) The band dispersion of a two dimensional EI with an s+ips+ip order parameter and ΔsΔp\Delta_{s}\ll\Delta_{p}.

Writing the partition function ZZ as a path integral over fermion fields ψ=(ψc,ψv)\psi=(\psi_{c},\,\psi_{v}), performing a Hubbard-Stratonovich transformation of the interaction term in the excitonic pairing channel and subsuming the intraband interaction into ξ\xi one obtains the action (see SI Sec. I)

S=dτdr{\displaystyle S=\int d\tau dr\Bigg{\{} ψ(τ+Hm)ψ+1gs|Δs|2+1gp|Δp|2}\displaystyle\psi^{\dagger}\left(\partial_{\tau}+H_{m}\right)\psi+\frac{1}{g_{s}}|\Delta_{s}|^{2}+\frac{1}{g_{p}}|\Delta_{p}|^{2}\Bigg{\}} (1)

as an integral over space-time (r,τ)(r,\tau) and the partition function is Z=D[ψ¯,ψ]D[Δ¯,Δ]eSZ=\int D[\bar{\psi},\psi]D[\bar{\Delta},\Delta]e^{-S}. For physically reasonable interactions such as the screened Coulomb interaction, the ss-wave pairing interaction gsg_{s} is typically the strongest while gpg_{p} is the leading subdominant one. We may write the mean field Hamiltonian as 𝑑rψHmψ=kψkHmkψk\int dr\psi^{\dagger}H_{m}\psi=\sum_{k}\psi_{k}^{\dagger}H_{m}^{k}\psi_{k} with

Hmk[Δs,Δp]=ξkσ3+Δsσ1+Δpfkσ2H_{m}^{k}[\Delta_{s},\Delta_{p}]=\xi_{k}\sigma_{3}+\Delta_{s}\sigma_{1}+\Delta_{p}f_{k}\sigma_{2} (2)

where σi\sigma_{i} are the Pauli matrices acting in the c/v band space. The vector potential AA enters Eq. (2) through the minimal coupling kkAk\rightarrow k-A required by local gauge invariance (electric field is E=tAE=-\partial_{t}A) and we set electron charge ee, speed of light and the Planck constant \hbar to be one. Interband dipolar couplings could also occur Golež et al. (2016); Murakami et al. (2020) but do not affect our results. Since the global phase is not important, we choose the ss-wave order parameter to be real. As we will show, the system develops an electrical polarization as a pp-wave component π/2\pi/2 out of phase with the equilibrium Δs\Delta_{s} is introduced. Due to an emergent ‘particle hole’ symmetry in the BCS weak coupling case defined as |Δs|,|Δp|G|\Delta_{s}|,|\Delta_{p}|\ll G which we focus on, applied electric fields create Δp\Delta_{p} primarily in this channel (see Sec. VI A of SI for a rigorous proof), so we write pp-wave pairing in the σ2\sigma_{2} channel Sun and Millis (2020b). The quasiparticle spectrum is Ek=±ξk2+Δs2+Δp2fk2E_{k}=\pm\sqrt{\xi_{k}^{2}+\Delta_{s}^{2}+\Delta_{p}^{2}f_{k}^{2}}. As shown by Fig. 1(c), the spectrum will have gapless points (nodes) at (kx,ky)=(0,±kF)(k_{x},k_{y})=(0,\pm k_{F}) in the pure pp-wave state (Δs=0\Delta_{s}=0) in two dimensions (2D).

Charge pump—Spatially uniform changes in Δs,p\Delta_{s,p} produce uniform currents J=kkHmkJ=\langle\sum_{k}\partial_{k}H_{m}^{k}\rangle (see SI Sec. II), whose time integral from the initial (Δs,Δp)=(Δ,0)(\Delta_{s},\Delta_{p})=(\Delta,0) to the final point then gives the pumped charge PP (difference of polarization between the final state and the initial state). In a one dimensional (1D) system, PP has a geometrical meaning King-Smith and Vanderbilt (1993); Resta (1994) in the limit of slow order parameter dynamics. It is the flux of the Berry curvature 2-form BB through the 2D surface SS spanned by the occupied 1D crystal momentum kk and the time varying trajectory of Δs,p\Delta_{s,p}, or alternatively by the line integral of the Berry connection 𝒜μ=iψ|μ|ψ\mathcal{A}_{\mu}=i\langle\psi|\partial_{\mu}|\psi\rangle around its boundary:

P=12πS𝑑SB=12π𝑑l𝒜\displaystyle P=\frac{1}{2\pi}\int_{S}dS\cdot B=\frac{1}{2\pi}\oint dl\cdot\mathcal{A}\, (3)

where μ=(k,Δs,Δp)\mu=(k,\Delta_{s},\Delta_{p}) (see Fig. 2).

The Berry curvature BB from the valence band of Eq. (2) is sourced by monopoles at the points ξk=Δs=Δp=0\xi_{k}=\Delta_{s}=\Delta_{p}=0, i.e., the points (k,Δs,Δp)=(±kF,0,0)(k,\Delta_{s},\Delta_{p})=(\pm k_{F},0,0) each of which has monopole charge 11. If the order parameter evolution completes a full cycle on the s+ips+ip plane, SS becomes the surface of the 2-torus shown in Fig. 2(a) and the net charge pumped is the total flux from the enclosed monopoles which is an integer N=2N=2, the Chern number of the process. This quantized change in the polarization is known as the Thouless pump Thouless (1983), a topological phenomenon immune to disorder. Note that the monopoles exist only for the ‘BCS’ (G>0G>0, band inversion) case where the excitons strongly overlap such that charge can jump between them. In the ‘BEC’ case ξk0\xi_{k}\neq 0 for all kk and there are no monopoles enclosed in SS (see SI Sec. II C).

Refer to caption
Figure 2: (a) The surface SS in the (k,Δs,Δp)(k,\Delta_{s},\Delta_{p}) space used to calculate the flux of the Berry curvature for a 1D EI for which the order parameter evolution completes a full cycle in the s+ips+ip plane. The left and right ends of the cylinder are identified so that SS is a 2-torus. Blue dots are Berry curvature monopoles and red dashed lines are ‘Dirac strings’ with direction shown by black arrows. (b) The surface of the torus shown in (a) parametrized by kk and θ\theta with k=±πk=\pm\pi and θ=0, 2π\theta=0,\,2\pi identified. The blue contours yield the charge pumped during a full cycle. The red rectangles are used to compute the flux for a partial cycle in the BCS limit.

To compute the polarization for the case the order parameter does not complete a full cycle, we use the line integral approach. For notational simplicity, we suppress the subscripts ‘kk’ without causing ambiguities. An explicit expression for the valence band wave function from (2) at (k,Δs,Δp)(k,\Delta_{s},\Delta_{p}) is

|ψ=(v,u)=12E(Eξ)(ξE,Δ)\displaystyle|\psi\rangle=(-v^{\ast},\,u^{\ast})=\frac{1}{\sqrt{2E(E-\xi)}}\left(\xi-E,\,\Delta^{\ast}\right) (4)

where Δ=Δs+iΔpfk|Δ|eiϕ\Delta=\Delta_{s}+i\Delta_{p}f_{k}\equiv|\Delta|e^{i\phi} and |u|2(|v2|)=12(1±ξE)|u|^{2}(|v^{2}|)=\frac{1}{2}\left(1\pm\frac{\xi}{E}\right). The Berry connection 𝒜μ=|u|2μϕ\mathcal{A}_{\mu}=|u|^{2}\partial_{\mu}\phi has singularities associated with the Dirac strings, the intersections of which with SS (marked by crosses in Fig. 2(b)) must be correctly treated in the evaluation of the line integral. Noting that |u|20|u|^{2}\rightarrow 0 when ξ|Δ|\xi\ll-|\Delta| and |u|21|u|^{2}\rightarrow 1 when ξ|Δ|\xi\gg|\Delta|, we see that in the weak coupling BCS limit the contour can be collapsed to the red rectangles in Fig. 2(b). Parameterizing SS using kk and the angle θ\theta defined by Δs+iΔp=Reiθ\Delta_{s}+i\Delta_{p}=Re^{i\theta} in Fig. 1(a), one observes that the polarization of an state on the s+ips+ip plane depends only on the angle θ\theta. Specifically, we found

P=θ/π\displaystyle P=\theta/\pi (5)

for a 1D EI (see SI Sec. II). This may be understood by noting that the low energy physics around ±kF\pm k_{F} is of two massive Dirac models, each of which realizes a Goldstone-Wilczek Goldstone and Wilczek (1981) mechanism of charge pumping.

Higher dimensional systems can be viewed as 1D chains along xx direction stacked in momentum space. For a 2D circular fermi surface one finds

P(θ)={kF2πtanθ2(0<θ<π/2)kF2π(2cotθ2)(π/2<θ<π)kFπ+P(θπ)(π<θ<2π).\displaystyle P(\theta)=\left\{\begin{array}[]{lc}\frac{k_{F}}{2\pi}\tan\frac{\theta}{2}&\,(0<\theta<\pi/2)\\ \frac{k_{F}}{2\pi}\left(2-\cot\frac{\theta}{2}\right)&(\pi/2<\theta<\pi)\\ \frac{k_{F}}{\pi}+P(\theta-\pi)&(\pi<\theta<2\pi)\end{array}\right.\,. (9)

A full cycle pumps exactly two electrons along each 1D momentum chain that crosses the fermi surface, giving

P1D=2,P2D=2kFπ,P3D=kF22π\displaystyle P_{1D}=2\,,\quad P_{2D}=\frac{2k_{F}}{\pi}\,,\quad P_{3D}=\frac{k_{F}^{2}}{2\pi} (10)

for 1D, 2D and three dimensional (3D) isotropic systems respectively.

Refer to caption
Figure 3: Left is the evolution of the edge state energies as a function of θ\theta. Right is the spatial profile of one of the two edge states (labeled by red line on the left) neglecting its quick oscillating detail. Being filled means the state is occupied. The blue edge state is not shown but is the mirror image of the red one.

Although the charge pump is a bulk property carried by all valence band electrons, it is also revealed by the evolution of edge states Yang et al. (2020) as Δs\Delta_{s} and Δp\Delta_{p} are varied, as shown in Fig. 3 for a 1D wire connected with reservoirs. In the BCS limit, with open boundary conditions ψ(0)=ψ(L)=0\psi(0)=\psi(L)=0, there are two edge states

ψ±=1C±(1,±1)sin(kFx)exΔp/vF,E=±Δs\displaystyle\psi_{\pm}=\frac{1}{C_{\pm}}\left(1,\pm 1\right)\sin(k_{F}x)e^{\mp x\Delta_{p}/v_{F}}\,,\quad E=\pm\Delta_{s} (11)

where C±C_{\pm} is a normalization constant. We suppose Δs+iΔp=Reiθ\Delta_{s}+i\Delta_{p}=Re^{i\theta} and follow the evolution of ψ+\psi_{+} as θ\theta is varied (see Fig.3). At θ=0\theta=0 the state is delocalized and unoccupied with energy RR. As θ\theta is increased the state becomes localized near x=0x=0 and decreases in energy. When θ\theta passes through π/2\pi/2, the state becomes maximally localized and becomes occupied by an electron from the left reservoir since its energy crosses the chemical potential. As θ\theta further increases the state becomes delocalized and then localized at the right edge, delivering its electron to the right reservoir when θ\theta crosses 3π/23\pi/2. Considering the ψ\psi_{-} state during the same cycle, two electrons in total are pumped. In higher dimensions, each 1D kxk_{x} chain crossing the fermi surface has a similar edge state evolution (see SI Sec. III).

Refer to caption
Figure 4: Electric field pulse induced dynamics of a 2D isotropic EI. (a) The polarization as a function of time during the dynamics with PdisP_{\text{dis}} being small. Inset is the free energy landscape F(Δs,Δp)F(\Delta_{s},\Delta_{p}) plotted on the s+ips+ip plane. Lower energy appears bluer. The black dashed order parameter trajectory is caused by a pulse E(t)=Emaxtanh((tt0)/w)E(t)=E_{\text{max}}\tanh^{\prime}\left((t-t_{0})/w\right) with maximum electric field Emax=0.39E0E_{\text{max}}=0.39E_{0}. (b) The pumped charge by a single pulse as a function of EmaxE_{\text{max}}. The units are P2D=2kF/πP_{2D}=2k_{F}/\pi and E0=Δ2/vFE_{0}=\Delta^{2}/v_{F}. Top inset is a schematic of a train of well separated pulses which can induce a ‘steady’ current. Bottom inset is a schematic of the device with EI shown in blue and the contacts in gold. The parameters are gsν=0.3g_{s}\nu=0.3, gpν=0.58g_{p}\nu=0.58, Δ=2Λe1/(gsν)\Delta=2\Lambda e^{-1/(g_{s}\nu)}, γ=0.07Δ\gamma=0.07\Delta and w=2π/(2Δ)w=2\pi/(2\Delta).

Dynamics—The coupled dynamics of electrons and the order parameters in the presence of an applied electric field is described by the action Eq. (1). To understand the qualitative dynamics, we use a low energy effective Ginzburg-Landau Lagrangian

L(Δs,Δp;E)=FK+Ldrive\displaystyle L(\Delta_{s},\Delta_{p};E)=F-K+L_{\text{drive}} (12)

for fields Δs,Δp\Delta_{s},\,\Delta_{p} obtained by integrating out the Fermions (see SI Sec. IV). The dynamics is given by the standard Euler-Lagrange equation ddtδLδΔ˙i=δLδΔi\frac{d}{dt}\frac{\delta L}{\delta\dot{\Delta}_{i}}=\frac{\delta L}{\delta\Delta_{i}} and is that of a point particle moving in the landscape defined by FF, with kinetic energy KK and driven by an electric field through LdriveL_{\text{drive}}. We find

Ldrive=P(θ)Es(Δs,Δp)E2+O(E3)L_{\text{drive}}=-P(\theta)E-s(\Delta_{s},\Delta_{p})E^{2}+O(E^{3}) (13)

where PP is the adiabatic polarization in Eqs. (5) or (9), s=limω0σ(ω)/(2iω)s=\lim_{\omega\rightarrow 0}\sigma(\omega)/(2i\omega) and σ(ω)\sigma(\omega) is the optical conductivity from virtual interband excitations (see SI Sec. IV). It is natural that electric field couples linearly to the polarization and therefore provides a ‘force’ EθPE\partial_{\theta}P to rotate the order parameter in the Δs,Δp\Delta_{s},\Delta_{p} plane.

F(Δs,Δp)F(\Delta_{s},\Delta_{p}) gives the potential landscape in which the dynamics takes place; it has the anisotropic ‘Mexican hat’ form shown in Fig. 4(a). For (quasi) 1D systems in the weak coupling BCS limit Altland and Simons (2010); Sun et al. (2020):

F=ν(Δs2+Δp2)ln2ΛΔs2+Δp2+1gsΔs2+1gpΔp2\displaystyle F=-\nu\left(\Delta_{s}^{2}+\Delta_{p}^{2}\right)\ln\frac{2\Lambda}{\sqrt{\Delta_{s}^{2}+\Delta_{p}^{2}}}+\frac{1}{g_{s}}\Delta_{s}^{2}+\frac{1}{g_{p}}\Delta_{p}^{2} (14)

where ν\nu is density of states in the metallic phase and ΛΔs2+Δp2\Lambda\gg\sqrt{\Delta_{s}^{2}+\Delta_{p}^{2}} is a high energy cutoff Kozlov and Maksimov (1965). The first term becomes νdθk2π(Δs2+Δp2cos2θk)ln(2Λ/Δs2+Δp2cos2θk)-\nu\int\frac{d\theta_{k}}{2\pi}\left(\Delta_{s}^{2}+\Delta_{p}^{2}\cos^{2}\theta_{k}\right)\ln\left(2\Lambda/\sqrt{\Delta_{s}^{2}+\Delta_{p}^{2}\cos^{2}\theta_{k}}\right) for a 2D isotropic Fermi surface and dθk2πsinθkdθkdϕk4π\frac{d\theta_{k}}{2\pi}\rightarrow\frac{\sin\theta_{k}d\theta_{k}d\phi_{k}}{4\pi} for 3D where θk\theta_{k} and ϕk\phi_{k} are angular variables on the Fermi surface. The landscape has a local maximum at R=0R=0 surrounded by a trough at R(θ)R(\theta) of lower values of FF. The ground state minima are at (±Δ,0)(\pm\Delta,0) and the pure p-wave phases at (0,±Δp0)(0,\pm\Delta_{p0}) are saddle points with energy higher by Fb=ν(Δ2cΔp02)/2F_{b}=\nu(\Delta^{2}-c\Delta_{p0}^{2})/2 where cc is a constant depending on the space dimension.

We may estimate the minimal electric field required to drive the system from the minimum through the pp-wave saddle point by equating the potential energy barrier FbF_{b} to the work EP(θ=π/2)+𝒪(E2)EP(\theta=\pi/2)+\mathcal{O}(E^{2}) done by the electric field, obtaining

EcκE0,E0=Δ/ξ0=Δ2/vF\displaystyle E_{c}\approx\kappa E_{0},\quad E_{0}=\Delta/\xi^{0}=\Delta^{2}/v_{F} (15)

where ξ0=vF/Δ\xi^{0}=v_{F}/\Delta is the coherence length (exciton size), κ=1π(1Δp02/Δ2)\kappa=\frac{1}{\pi}(1-\Delta_{p0}^{2}/\Delta^{2}) in 1D and κ=1214Δp02Δ2\kappa=\frac{1}{2}-\frac{1}{4}\frac{\Delta_{p0}^{2}}{\Delta^{2}} in 2D, and E0E_{0} is at the order of the dielectric breakdown field. For vF=106m/sv_{F}=10^{6}\,\mathrm{m/s}, Δ=10meV\Delta=10\,\mathrm{meV} and Δp0Δ\Delta_{p0}\ll\Delta, such as the case of electron hole bilayers, the threshold field is Ec103V/cmE_{c}\sim 10^{3}\,\mathrm{V/cm} which can be easily achieved by modern optical technique. For a 100meV100\,\mathrm{meV} gap such as that in Ta2NiSe5 Lu et al. (2017); Werdehausen et al. (2018) (assuming it is in the BCS regime), the threshold field is about 105V/cm10^{5}\,\mathrm{V/cm}. At such large field, O(E2)O(E^{2}) terms in the Lagrangian will be important, which pushes the order parameter closer to zero but does not destroy the qualitative dynamics in the transient regime (See SI Sec. IV D).

The dynamical term KK has a relatively simple form if the gap never closes on the Fermi surface and the order parameter variation timescale is long compared to the inverse of the gap. For example for (quasi) 1D

Kν(R˙2/R2+3θ˙2)/12K\approx\nu\left(\dot{R}^{2}/R^{2}+3\dot{\theta}^{2}\right)/12 (16)

to lowest order in time derivatives. For higher dimensions with closed Fermi surfaces, there are O(1)O(1) changes to the coefficients and, crucially, dissipation and time non-locality arises from quasiparticle excitations near the nodes of the p-wave gap when Δs\Delta_{s} passes zero. This dissipation also brings a correction to the pumped charge: PP2D+PdisP\rightarrow P_{2D}+P_{\text{dis}}. To estimate PdisP_{\text{dis}}, we observe that as the order parameter passes this gapless regime with a velocity Δ˙s\dot{\Delta}_{s}, the probability for exciting a particle-hole pair at kk is given by the Landau-Zener formula Wittig (2005): Pk=e2πδk2/|t2Δs|P_{k}=e^{-2\pi\delta_{k}^{2}/|\partial_{t}2\Delta_{s}|} where δk=ξk2+Δp2fk2\delta_{k}=\sqrt{\xi_{k}^{2}+\Delta_{p}^{2}f_{k}^{2}} is half of its minimal excitation energy during the dynamics. In 2D, summing over momenta, one obtains the number of excited quasi particles N=kF2π2vF|Δ˙sΔp|N=\frac{k_{F}}{2\pi^{2}v_{F}}|\frac{\dot{\Delta}_{s}}{\Delta_{p}}| and the non-adiabatic correction to the pumped charge

Pdis=P2D18π2|Δ˙s|Δp2\displaystyle P_{\text{dis}}=-P_{2D}\frac{1}{8\pi^{2}}\frac{|\dot{\Delta}_{s}|}{\Delta_{p}^{2}}\, (17)

valid if |Δ˙s||Δp|\sqrt{|\dot{\Delta}_{s}|}\ll|\Delta_{p}| (see SI Sec. VI C). Therefore, if the sub-dominant pp-wave coupling constant is too small such that Δp0Λe1gpν<|Δ˙s|/(8π2)\Delta_{p0}\sim\Lambda e^{-\frac{1}{g_{p}\nu}}<\sqrt{|\dot{\Delta}_{s}|/(8\pi^{2})}, this dissipative correction will dominate over the adiabatic charge.

Numerics and Experiment—We numerically solved the mean field dynamics implied by Eq. (1) for a BCS weak coupling EI in 2D, driven by a train of widely separated electric field pulses (Fig. 4(b)). Mean field dynamics Barankov et al. (2004) means that each momentum state evolves in the time dependent mean field (Δs,ΔpfkA,ξkA)(\Delta_{s},\,\Delta_{p}f_{k-A},\,\xi_{k-A}) with Δs,p\Delta_{s,p} determined self consistently by the gap equation, neglecting any spatial fluctuations. We include a weak phenomenological damping γ\gamma to represent energy loss caused by, e.g., a phonon bath (see SI Sec. VI). Each pulse drives the order parameter along the trajectory shown as the black dashed line in Fig. 4(a), advancing it by θ=π\theta=\pi to stabilize the system in the other ss-wave ground state. The total duration of the evolution from one minimum to the next is Ts20π/ΔT_{s}\approx 20\pi/\Delta and the amount of charge pumped is WP2D/2WP_{2D}/2 where WW is the width of the sample, as shown by Fig. 4(a). In a train of pulses with inter pulse separation T0TsT_{0}\gg T_{s}, each pulse advances the order parameter from one minimum to the next and allows it to stabilize before next pulse arrives, leading to a time-averaged current I0=eWkF/(πT0)I_{0}=eWk_{F}/(\pi T_{0}). For a 10μm10\,\mathrm{\mu m} wide sample with normal state carrier density of 1012cm210^{12}\,\mathrm{cm^{-2}} and inter pulse time T0=1nsT_{0}=1\,\mathrm{ns}, the current is I0=255nAI_{0}=255\,\mathrm{nA} considering spin degeneracy.

A minimum field strength Ec\sim E_{c} (15) is required: as the maximum electric field EmaxE_{\text{max}} of the pulse is increased beyond the threshold, the charge pumping (DC current) will onset sharply, as shown in Fig. 4(b). As EmaxE_{\text{max}} further increases, each pulse induces a rotation of more cycles which pumps more charge, giving rise to the step structure. Deviations from perfect quantization arise from fast order parameter dynamics caused by the short duration pulse. A precisely engineered long duration pulse can substantially reduce these deviations; see SI Sec. V.

A static electric field in the DC transport regime could also drive such an order parameter rotation and charge pumping. However, unlike the case of well separated pulses, there is no time break to dump the generated heat into the environment which might destroy the system.

Discussion—We have shown theoretically that a Thouless charge pump may be realized as a collective many-body effect arising from order parameter steering in BCS type excitonic insulators. Similar dynamics and charge pumping can occur in general when the ground state order parameter and the sub dominant one have different parities under inversion. Its observation would provide both a verification of order parameter steering and a probe of the excitonic insulating state, in particular, distinguishing BCS and BEC states. It is interesting to study the dynamics in the vicinity of the BCS-BEC crossover and effects beyond mean field.

Acknowledgements.
We acknowledge support from the Energy Frontier Research Center on Programmable Quantum Materials funded by the US Department of Energy (DOE), Office of Science, Basic Energy Sciences (BES), under award No. DE-SC0019443. We thank W. Yang, D. Golez and T. Kaneko for helpful discussions.

References

Supplemental Material for ‘Topological charge pumping in excitonic insulators’

I The Hamiltonian

We base our discussion on the two-band spinless Fermion Hamiltonian that is a minimal model for excitonic insulators:

H=\displaystyle H= 𝑑r[ψ(σ3ξpA+ϕ0)ψ]+𝑑r𝑑rV(rr)ρ(r)ρ(r)\displaystyle\int{dr}\left[\psi^{\dagger}\left(\sigma_{3}\xi_{p-A}+\phi_{0}\right)\psi\right]+\int{drdr^{\prime}}V(r-r^{\prime})\rho(r)\rho(r^{\prime}) (S1)

where ρ(r)=ψ(r)ψ(r)\rho(r)=\psi^{\dagger}(r)\psi(r) is the local density at position rr, ψ=(ψc,ψv)\psi^{\dagger}=(\psi_{c}^{\dagger},\,\psi_{v}^{\dagger}) is the two component electron creation operator with c/v labeling the conduction/valence band, ξp=εpμ\xi_{p}=\varepsilon_{p}-\mu is the kinetic energy shifted by the chemical potential, p=ip=-i\hbar\nabla, σi\sigma_{i} are the Pauli matrices in c-v space, (ϕ0,A)Aμ(\phi_{0},\,A)\equiv A_{\mu} is the electromagnetic (EM) potential and we have set the electron charge, the speed of light and Planck constant \hbar to be 11. In the non interacting case, the overlap of the bands gives rise to an electron and a hole pocket, each with the Fermi momentum kFk_{F}, Fermi velocity vFv_{F}, Fermi level density of states ν\nu and carrier density n/2n/2. We choose the gauge ϕ0=0\phi_{0}=0 in this paper.

The interaction is assumed to be density-density type, e.g., V(r)=1/|r|V(r)=1/|r| for the Coulomb interaction. We denote its Fourier transform at momentum qq as VqV_{q}. The repulsive interaction is effectively attractive between electrons and holes and can induce pairing in several angular momentum channels, in formal analogy to Cooper pairing in superconductors. We write the model as a fermionic path integral so the partition function is Z=D[ψ¯,ψ]eψτψH[ψ]Z=\int D[\bar{\psi},\psi]e^{\psi^{\dagger}\partial_{\tau}\psi-H[\psi]} and decouple the interaction in the electron hole pairing channel: Z=D[ψ¯,ψ]D[Δ¯,Δ]eS[ψ,Δ]Z=\int D[\bar{\psi},\psi]D[\bar{\Delta},\Delta]e^{-S[\psi,\Delta]}. Note that ‘DD’ means functional field integral. The Hubbard-Stratonovich fields Δ\Delta then represent the order parameters. The interband interaction term in Eq. (S1) can be decomposed into symmetry channels labeled by ll as V^inter=lglb^lb^l\hat{V}_{\text{inter}}=\sum_{l}g_{l}\hat{b}_{l}^{\dagger}\hat{b}_{l} where b^l=kfklψc,kψv,k\hat{b}_{l}=\sum_{k}f^{l}_{k}\psi^{\dagger}_{c,k}\psi_{v,k} and fklf^{l}_{k} are the representation functions (pairing functions) of the point symmetry group. Therefore, we can resolve Δ\Delta into pairing functions as Δk=lΔlfkl\Delta_{k}=\sum_{l}\Delta_{l}f^{l}_{k}. We assume for notational simplicity that the excitonic effects occur near a high symmetry point so that lattice effects are unimportant and that the interaction effects may be restricted to the Fermi surface. In this case flf^{l} become the usual d-dimensional rotational harmonics and the interaction VqV_{q} is parameterized by the momentum transfer qq connecting two points on the Fermi surface. We focus on ss pairing (fksf^{s}_{k} has the full point symmetry of the lattice; we take fks=1f^{s}_{k}=1) and pxp_{x} pairing fkp=kx/kFf^{p}_{k}=k_{x}/k_{F}. We denote the latter as fkf_{k} for notational simplicity.

The Hubbard Stratonovich transformed action is

S[ψ,Δs,Δp,A]=dτdr{\displaystyle S[\psi,\Delta_{s},\Delta_{p},A]=\int d\tau dr\Bigg{\{} ψ(τ+Hm)ψ+1gs|Δs|2+1gp|Δp|2}.\displaystyle\psi^{\dagger}\left(\partial_{\tau}+H_{m}\right)\psi+\frac{1}{g_{s}}|\Delta_{s}|^{2}+\frac{1}{g_{p}}|\Delta_{p}|^{2}\Bigg{\}}\,. (S2)

Since the overall phase of the Hubbard Stratonovich field is not relevant for our considerations, we choose Δs\Delta_{s} to be real. As we will see in Sec. VI.1, due to a ‘particle hole’ symmetry in the BCS weak coupling case, the main effect of an electric field is to induce a pp-wave field that is π/2\pi/2 out of phase with Δs\Delta_{s}. We restrict our attention to this case. The mean field Hamiltonian is thus Hmk=ξkσ3+Δsσ1+Δpfkσ2H_{m}^{k}=\xi_{k}\sigma_{3}+\Delta_{s}\sigma_{1}+\Delta_{p}f_{k}\sigma_{2} with real Δs,Δp\Delta_{s},\,\Delta_{p} and the EM vector potential enters as kkAk\rightarrow k-A. Note that minimal coupling substitution is also applied to the pp-wave decoupling term: ΔpfkA\Delta_{p}f_{k-A} although this term comes from the electron-electron interaction that contains no EM field. We discuss this choice here in terms of local gauge invariance.

In the full functional integral, the general gauge invariant form of the decoupling term is eir1r2𝑑lA(l)Δ(r1,r2)ψv(r1)ψc(r2)e^{-i\int_{r_{1}}^{r_{2}}dlA(l)}\Delta(r_{1},r_{2})\psi^{\dagger}_{v}(r_{1})\psi_{c}(r_{2}), which preserves its form under the usual local gauge transformation Ug:ψ(r)ψ(r)eiθ(r),AμAμ+μθ(r)U_{g}:\psi(r)\rightarrow\psi(r)e^{i\theta(r)},\,A_{\mu}\rightarrow A_{\mu}+\partial_{\mu}\theta(r). We write Δ(r1,r2)=|Δ(r1,r2)|eiφ(r1,r2)\Delta(r_{1},r_{2})=|\Delta(r_{1},r_{2})|e^{i\varphi(r_{1},r_{2})}; both the amplitude |Δ(r1,r2)||\Delta(r_{1},r_{2})| and the phase φ(r1,r2)\varphi(r_{1},r_{2}) are dynamical variables. The dependence on ‘center of mass’ coordinate r=(r1+r2)/2r=(r_{1}+r_{2})/2 gives the spatial variation of the order parameter while the dependence on r1r2r_{1}-r_{2} gives the internal structure of the electron hole pair (the momentum dependence of the pairing function). Writing the phase degrees of freedom as φ(r1,r2)=φ0(r)+(r1r2)α(r)\varphi(r_{1},r_{2})=\varphi_{0}(r)+(r_{1}-r_{2})\alpha(r) in the slow varying limit, the φ0\varphi_{0} is the usual order parameter phase and the combination α+A\alpha+A enters structure of pairing wave function as fkAαf_{k-A-\alpha}. Although α=0\alpha=0 in the initial ground state, one should in principle track the dynamics of α\alpha. In the weak coupling BCS limit, we may neglect the dynamics of α\alpha since its appearance is equivalent to a Δs\Delta_{s} in the σ2\sigma_{2} channel which vanishes as we will prove in Sec. VI. Even in the general case when the appearance of α\alpha must be considered at intermediate stages of the dynamics, the amount of charge pumping during a full cycle is still the quantized value given that the system finally returns to its initial state, as shown in general by Thouless Thouless (1983).

If the two bands are formed by atomic orbitals having different parities, e.g., p and d orbitals, an interband dipolar moment D0D_{0} adds the extra term D0Eψσ1ψD_{0}E\psi^{\dagger}\sigma_{1}\psi to the Hamiltonian. This term also contributes to the EM response due to change of inter orbital hybridization. However, for a full cycle of order parameter dynamics, the amount of pumped charge won’t be affected if the initial and final states are the same such that they have the same inter orbital hybridization. The dynamics itself won’t be qualitatively affected if the interband dipole D0D_{0} is not large compared to the dipole formed between ss and pp-symmetry electron and hole bound states that produce the order parameters. This is true in the BCS case since the former is proportional to the size of atomic orbitals while the latter is the size ξ0\xi^{0} of the extended electron hole bound state.

I.1 The pairing interactions

For isotropic systems in dd dimensions, the pairing interaction in channel ll is

gl=12|fl|2𝑑Ω1𝑑Ω2fk(Ω1)lVk(Ω1)k(Ω2)fk(Ω2)l\displaystyle g_{l}=\frac{1}{2|f^{l}|^{2}}\int d\Omega_{1}d\Omega_{2}f^{l\ast}_{k(\Omega_{1})}V_{k(\Omega_{1})-k(\Omega_{2})}f^{l}_{k(\Omega_{2})} (S3)

where Ω\Omega is the angular variable on the d1d-1 dimensional Fermi surface, k(Ω)k(\Omega) is the electron momentum at angle Ω\Omega and |fl|2=𝑑Ω|fk(Ω)l|2|f^{l}|^{2}=\int d\Omega|f^{l}_{k(\Omega)}|^{2} is the normalization factor. Since the symmetry group is O(d)O(d), the dd-dimensional rotations and inversions, the pairing functions flf^{l} are dd-dimensional spherical harmonics. Below we discuss 1D and 2D separately.

One dimension—There are only two pairing functions: the inversion symmetric one fkFs=1,fkFs=1f^{s}_{k_{F}}=1,\,f^{s}_{-k_{F}}=1 and the antisymmetric one fkFp=1,fkFp=1f^{p}_{k_{F}}=1,\,f^{p}_{-k_{F}}=-1. Thus Eq. (S3) leads to gs=Vq=0+Vq=2kFg_{s}=V_{q=0}+V_{q=2k_{F}} and gp=Vq=0Vq=2kFg_{p}=V_{q=0}-V_{q=2k_{F}}. Since the Coulumb interaction in 1D is Vqln(aq)V_{q}\sim\ln(aq) where aa is a short distance cutoff, Vq=0V_{q=0} diverges Logarithmically which results in gs=gpg_{s}=g_{p}. If the 1D system is put parallel to a metallic gate at a distance hh, the screening kills the divergence of Vq=0V_{q=0} and renders gp/gs<1g_{p}/g_{s}<1. As hh decreases, the screening get stronger and gp/gsg_{p}/g_{s} decreases continuously until reaching zero at h1/kFh\ll 1/k_{F}. Therefore, gp/gsg_{p}/g_{s} is experimentally tunable by hh.

Two dimensions—Examples are electron hole bilayers and other 2D excitonic insulators such as monolayer WTe2 Jia et al. (2020). The natural pairing functions are fkl=eilθkf_{k}^{l}=e^{il\theta_{k}} which plugged into Eq. (S3) renders gl=12π𝑑θkcos(lθk)V2kFsin(θk/2)g_{l}=\frac{1}{2\pi}\int d\theta_{k}\cos(l\theta_{k})V_{2k_{F}\sin(\theta_{k}/2)}. Note that for l=0l=0, the 1/2π1/2\pi factor should be changed to 1/4π1/4\pi. Since eilθke^{il\theta_{k}} and eilθke^{-il\theta_{k}} have the same glg_{l}, we will use the real pairing functions fkl=cos(lθk),sin(lθk)f_{k}^{l}=\cos(l\theta_{k}),\sin(l\theta_{k}) for convenience. For Thomas-Fermi screened interaction Vq=2πϵ(q+qTF)V_{q}=\frac{2\pi}{\epsilon(q+q_{\text{TF}})} in 2D where qTF/(2kF)=e2/(ϵvF)αq_{\text{TF}}/(2k_{F})=e^{2}/(\epsilon\hbar v_{F})\equiv\alpha is the ‘fine structure constant’ in this system and ϵ\epsilon is the dielectric constant of the environment, the ss-wave pairing strength is

νgs\displaystyle\nu g_{s} =ν14π𝑑θk2π2kF|sin(θk/2)|+qTF=α1α21πTanh1(1α2)\displaystyle=\nu\frac{1}{4\pi}\int d\theta_{k}\frac{2\pi}{2k_{F}|\sin(\theta_{k}/2)|+q_{\text{TF}}}=\frac{\alpha}{\sqrt{1-\alpha^{2}}}\frac{1}{\pi}\mathrm{Tanh}^{-1}\left(\sqrt{1-\alpha^{2}}\right) (S4)

and the pp-wave one is

νgp\displaystyle\nu g_{p} =ν12π𝑑θk2πcosθk2kF|sin(θk/2)|+qTF\displaystyle=\nu\frac{1}{2\pi}\int d\theta_{k}\frac{2\pi\cos\theta_{k}}{2k_{F}|\sin(\theta_{k}/2)|+q_{\text{TF}}}
=α[4π+2α+4π(12α2)1α2(Tanh1(1α2)Tanh1(1α21+α))].\displaystyle=\alpha\Bigg{[}-\frac{4}{\pi}+2\alpha+\frac{4}{\pi}\frac{\left(1-2\alpha^{2}\right)}{\sqrt{1-\alpha^{2}}}\Bigg{(}\mathrm{Tanh}^{-1}\left(\sqrt{1-\alpha^{2}}\right)-\mathrm{Tanh}^{-1}\left(\frac{\sqrt{1-\alpha^{2}}}{1+\alpha}\right)\Bigg{)}\Bigg{]}\,. (S5)

These equations were previously given Sun and Millis (2020b) and are reproduced here for convenience.

The pairing interactions are shown in Fig. S1 for the screened Coulomb interaction in 2D (reproduced from the supplemental material of Ref. Sun and Millis (2020b)). To obtain a substantial gp/(2gs)g_{p}/(2g_{s}), one needs the high density case where the fermi velocity is large so that the Thomas fermi wave vector is smaller than the fermi momentum: qTF/(2kF)=α=e2/(ϵvF)1q_{\text{TF}}/(2k_{F})=\alpha=e^{2}/(\epsilon\hbar v_{F})\ll 1. Stronger dielectric screening of the environment can further reduce α\alpha and increase gp/(2gs)g_{p}/(2g_{s}). Moreover, a non-negligible interlayer distance aa changes the bare electron-hole Coulomb attraction into V(r)=1/r2+a2V(r)=1/\sqrt{r^{2}+a^{2}}, making it more nonlocal and thus can lead to a larger gp/(2gs)g_{p}/(2g_{s}). Other types of interactions such as nearest neighbor Hubbard interaction (although originating from Coulomb) could give a very strong gpg_{p}, given that the band overlapping is suitable.

Refer to caption
Figure S1: Pairing interaction plots reproduced from Ref. Sun and Millis (2020b). (a) The ss,pp,dd-wave components of the screened Coulomb interaction in 2D and the ‘fine structure constant’ α=e2/(ϵvF)=qTF/(2kF)\alpha=e^{2}/(\epsilon\hbar v_{F})=q_{\text{TF}}/(2k_{F}) as functions of electron density n/2=m2vF2/(4π2)n/2=m^{2}v_{F}^{2}/(4\pi\hbar^{2}) computed from Eqs. (S4) and (S5) using m=0.05mem=0.05m_{e} and ϵ=10\epsilon=10 where mm is the electron mass in the model and mem_{e} is the vacuum electron mass. (b) The ratio gp/(2gs)g_{p}/(2g_{s}) as a function of α=qTF/(2kF)\alpha=q_{\text{TF}}/(2k_{F}). For α1\alpha\ll 1, i.e., in the high density case, gp/(2gs)g_{p}/(2g_{s}) becomes considerable and approaches one in the high density limit. Spin degeneracy is neglected.

II Computation of the polarization

In this section, we explicitly derive the charge pumping in an 1D excitonic insulator by computing the polarization PP (charge pumped) as a time integral of the current JJ induced by adiabatic changes to the order parameter over a time interval from 0 to tt and comparing the result to the formula in terms of Berry curvature, consistent with previous results King-Smith and Vanderbilt (1993).

II.1 Polarization

For convenience we reproduce the mean field Hamiltonian HmkH_{m}^{k} and current operator JJ here:

Hmk=ξkσ3+Δsσ1+Δpfkσ2,j^k=ψkjkψk=ψk(kHmk)ψk=ψk(vkσ3+Δpkfkσ2)ψk,J=kj^k\displaystyle H_{m}^{k}=\xi_{k}\sigma_{3}+\Delta_{s}\sigma_{1}+\Delta_{p}f_{k}\sigma_{2}\,,\quad\hat{j}_{k}=\psi_{k}^{\dagger}j_{k}\psi_{k}=\psi_{k}^{\dagger}\left(\partial_{k}H_{m}^{k}\right)\psi_{k}=\psi_{k}^{\dagger}\left(v_{k}\sigma_{3}+\Delta_{p}\partial_{k}f_{k}\sigma_{2}\right)\psi_{k}\,,\quad J=\sum_{k}\hat{j}_{k} (S6)

where vk=kξkv_{k}=\partial_{k}\xi_{k} is the velocity and the energy eigenvalues are Ek=±ξk2+Δs2+Δp2fk2E_{k}=\pm\sqrt{\xi_{k}^{2}+\Delta_{s}^{2}+\Delta_{p}^{2}f^{2}_{k}}. The current δJ\delta J in response to a change in order parameter δΔk=δΔs(t)σ1+δΔp(t)fkσ2\delta\Delta_{k}=\delta\Delta_{s}(t)\sigma_{1}+\delta\Delta_{p}(t)f_{k}\sigma_{2} is

δJ(Ωn)=TωnkTr[jk(iωn+iΩnHmk)δΔk(iωHmk)((ωn+Ωn)2+Ek2)(ωn2+Ek2)]\delta J(\Omega_{n})=-T\sum_{\omega_{n}}\sum_{k}\mathrm{Tr}\left[\frac{j_{k}\left(i\omega_{n}+i\Omega_{n}-H_{m}^{k}\right)\delta\Delta_{k}\left(i\omega-H_{m}^{k}\right)}{\left(\left(\omega_{n}+\Omega_{n}\right)^{2}+E_{k}^{2}\right)\left(\omega_{n}^{2}+E_{k}^{2}\right)}\right] (S7)

in frequency representation where ωn=(2n+1)πT\omega_{n}=(2n+1)\pi T, Ωn=2nπT\Omega_{n}=2n\pi T are the Fermion and Boson Matsubara frequencies with nn\in\mathbb{Z} (not to be confused with the carrier density), TT is the temperature and we have set the Boltzmann constant to be 11. Carrying out the trace over band indices, performing the frequency summation at T=0T=0 and analytically continuing iΩni\Omega_{n} to ω\omega, one obtains

δJ(ω)=iω2(kvkfkξkkfkEk(ω24+Ek2)ΔpδΔs(ω)kvkfkEk(ω24+Ek2)ΔsδΔp(ω))\delta J(\omega)=\frac{i\omega}{2}\left(\sum_{k}\frac{v_{k}f_{k}-\xi_{k}\partial_{k}f_{k}}{E_{k}\left(-\frac{\omega^{2}}{4}+E_{k}^{2}\right)}\Delta_{p}\delta\Delta_{s}(\omega)-\sum_{k}\frac{v_{k}f_{k}}{{E_{k}\left(-\frac{\omega^{2}}{4}+E_{k}^{2}\right)}}\Delta_{s}\delta\Delta_{p}(\omega)\right) (S8)

Note that if the argument is ω\omega in Δs,p(ω)\Delta_{s,p}(\omega), the latter means Δs,p(t)\Delta_{s,p}(t) Fourier transformed into frequency representation. In the adiabatic limit we may neglect the ω2\omega^{2} in the denominators; then transforming to the time domain we obtain

J(t)=12(kvkfkξkkfkEk3ΔpΔstkvkfkEk3ΔsΔpt)J(t)=-\frac{1}{2}\left(\sum_{k}\frac{v_{k}f_{k}-\xi_{k}\partial_{k}f_{k}}{E_{k}^{3}}\Delta_{p}\frac{\partial\Delta_{s}}{\partial t}-\sum_{k}\frac{v_{k}f_{k}}{{E_{k}^{3}}}\Delta_{s}\frac{\partial\Delta_{p}}{\partial t}\right)\, (S9)

where ‘JJ’ now means the expectation value of the current operator. Integrating in time gives the change in polarization:

P=(Δp(vkfkξkkfk)2Ek3dΔsdk2π+Δsvkfk2Ek3dΔpdk2π).P=\int\left(-\frac{\Delta_{p}\left(v_{k}f_{k}-\xi_{k}\partial_{k}f_{k}\right)}{2E_{k}^{3}}\frac{d\Delta_{s}dk}{2\pi}+\frac{\Delta_{s}v_{k}f_{k}}{2E_{k}^{3}}\frac{d\Delta_{p}dk}{2\pi}\right)\,. (S10)

II.2 Berry Connection and Berry Curvature

The Berry connection 𝒜μ\mathcal{A}_{\mu} is given in terms of the change in wave function under infinitesimal variation of the parameters μ=(k,Δs,Δp)\mu=(k,\Delta_{s},\Delta_{p}) as 𝒜μ=iψ|μ|ψ\mathcal{A}_{\mu}=i\langle\psi|\partial_{\mu}|\psi\rangle. We suppress the momentum subscripts whenever possible without causing ambiguity. Defining Δ=Δs+iΔpfk|Δ|eiϕ\Delta=\Delta_{s}+i\Delta_{p}f_{k}\equiv|\Delta|e^{i\phi} we may write the valence band wave function as

|ψ=(v,u)=12E(Eξ)(ξE,Δ)\displaystyle|\psi\rangle=(-v^{\ast},\,u^{\ast})=\frac{1}{\sqrt{2E(E-\xi)}}\left(\xi-E,\,\Delta^{\ast}\right) (S11)

implying 𝒜μ=|u|2μϕ\mathcal{A}_{\mu}=|u|^{2}\partial_{\mu}\phi where |u|2=12(1+ξE)|u|^{2}=\frac{1}{2}\left(1+\frac{\xi}{E}\right). Explicitly,

(𝒜k,𝒜Δs,𝒜Δp)=|u|2Δs2+fk2Δp2(ΔsΔpkfk,fkΔp,fkΔs).\displaystyle\left(\mathcal{A}_{k},\mathcal{A}_{\Delta_{s}},\mathcal{A}_{\Delta_{p}}\right)=\frac{|u|^{2}}{\Delta_{s}^{2}+f_{k}^{2}\Delta_{p}^{2}}\left(\Delta_{s}\Delta_{p}\partial_{k}f_{k},\,-f_{k}\Delta_{p},\,f_{k}\Delta_{s}\right)\,. (S12)

Note that 𝒜\mathcal{A} has singularities (“Dirac strings") along the line Δs=Δp=0\Delta_{s}=\Delta_{p}=0 and also along the lines Δs=fk=0\Delta_{s}=f_{k}=0 unless |u|2|u|^{2} vanishes on parts of these lines. These are shown as dashed lines in Fig. S2(a) for the BEC case (G<0G<0). The Berry curvature B=dAB=dA is then

(Bk,BΔs,BΔp)=12Ek3(ξkfk,Δsvkfk,Δp(vkfkξkkfk)).\displaystyle\left(B_{k},B_{\Delta_{s}},B_{\Delta_{p}}\right)=-\frac{1}{2E_{k}^{3}}\left(\xi_{k}f_{k},\,\Delta_{s}v_{k}f_{k},\,\Delta_{p}(v_{k}f_{k}-\xi_{k}\partial_{k}f_{k})\right)\,. (S13)

Considering now the flux of BB through a surface element of an oriented 2D manifold in (k,Δs,Δp)(k,\Delta_{s},\Delta_{p}) space defined by a function s(Δs,Δp)=constants(\Delta_{s},\Delta_{p})=\text{constant} and choosing the orientation to be pointing ‘inside’ the cylinder in Fig. S2(a) we see by comparison to Eq. (S10) that the flux through the surface is just the polarization PP. This conclusion is independent of the choice of coordinate.

II.3 BCS-BEC crossover

If the numbers of electrons and holes are separately conserved, the total number n=nelectron+nhole=σ3+n0n=\langle n_{electron}+n_{hole}\rangle=-\langle\sigma_{3}\rangle+n_{0} is also conserved where n0n_{0} is the particle number of a completely occupied band. nn is the analogy to the total charge in a superconductor, and gives the constraint that shifts GG from positive to negative as interaction becomes stronger such that the system crossovers from a BCS to a BEC type condensate. This is the situation in electron hole bilayers with no interlayer tunneling. Moreover, nn can also be approximately fixed by gate voltage. For natural crystals, nn is not fixed since there are always interband conversion mechanisms breaking this U(1)U(1) symmetry. Hartree terms due to Coulomb repulsion between a/b orbitals will shift down GG and induce such a crossover in this case.

In the BEC case (G<0G<0, no band inversion), there are no monopoles and the Dirac string structure looks like that in Fig. S2, rending zero pumped charge. Intuitively, the excitons in the BEC state are tightly bound electron hole pairs that don’t overlap with other, and can be viewed as charge neutral point particles. Thus no charge transport can occur.

Therefore, there is a topological transition at G=0G=0 during the BCS-BEC crossover, and the charge pumping PP can be viewed as an ‘order parameter’ that separate these two regimes, as shown in Fig. S2(c). However, we have focused on the dynamics in the BCS limit in this paper while it is interesting to investigate similar dynamics in the crossover regime.

Refer to caption
Figure S2: (a) The (k,θ)(k,\theta) surface SS embedded in the (k,Δs,Δp)(k,\Delta_{s},\Delta_{p}) space. Shown is the BEC case where the effective GG is negative in the mean field equation (band gap G-G is positive). In the Berry connection convention Eq. (S12), there are ‘Dirac strings’ as shown by the red dashed lines, whose intersection with SS lead to the singular vortices in (b). The black arrows indicate the directions of the Dirac strings. (b) The torus parametrized with (k,θ)(k,\theta). The integral of the Berry curvature over the torus is converted into the loop integral on its boundary and around the singular vortices. The vortices at k=0k=0 contribute opposite values to those at k=π/ak=\pi/a, rendering the net result to be P=0P=0. (c) Schematic of the pumped charge in BCS and BEC regimes of excitonic insulators.

II.4 Pumped charge for arbitrary rotation angle

In this section we provide the details leading to Eqs. (5) and (6) of the main text. Parameterizing SS using kk and the angle θarg(Δs,Δp)\theta\equiv\arg(\Delta_{s},\Delta_{p}) defined in Fig. 1(a) of the main text, the Berry connection and curvature can be projected onto the (k,θ)(k,\theta) space. In other words, The wave function can now be viewed as a function of (k,θ)(k,\theta) and Δ(k,θ)=Δs+iΔpfk=|Δ(k,θ)|eiϕ(k,θ)\Delta(k,\theta)=\Delta_{s}+i\Delta_{p}f_{k}=|\Delta(k,\theta)|e^{i\phi(k,\theta)} is the pairing field at (k,θ)(k,\theta). Note that fkf_{k} reverses sign for positive and negative kk, and that |u|2=12(1+ξE)|u|^{2}=\frac{1}{2}\left(1+\frac{\xi}{E}\right) is nearly zero deep inside the fermi sea (ξ|Δ|\xi\ll-|\Delta|) and |u|21|u|^{2}\rightarrow 1 outside when ξ|Δ|\xi\gg|\Delta|.

The flux can be converted to the line integral of the Berry connection over the edge of SS in Fig. 2(b) of the main text. The singularities (marked by crosses in Fig. 2(b)) are from the intersections with the Dirac strings, and must be correctly treated in the evaluation of the line integral, although they do not contain fluxes in BB. For a full cycle, the edge is the blue contour together with the two small circles surrounding the two vortices. The former gives zero net contribution due to periodicity in kk and θ\theta while the latter contributes N=2N=2, recovering the number of monopoles.

The polarization at arbitrary angle θ\theta can be evaluated analytically in the BCS limit in which |Δs|,|Δp|G|\Delta_{s}|,|\Delta_{p}|\ll G where aa is the lattice constant. The Berry curvature is concentrated in the region |ξk||Δ||\xi_{k}|\lesssim|\Delta| so we may perform the integral in Eq. (3) of the main text only on the red rectangles centered on k=±kFk=\pm k_{F} in Fig. 2(b) of the main text, chosen such that u0u\approx 0 on the vertical edges inside the fermi momentum ±kF\pm k_{F} and thus 𝒜θ=0\mathcal{A}_{\theta}=0 from Eq. (S12), while |u|1|u|\approx 1 on the vertical edges outside the fermi momentum and 𝒜θ=±1\mathcal{A}_{\theta}=\pm 1. The top and bottom edges all contribute zero since kfk\partial_{k}f_{k} and thus 𝒜k\mathcal{A}_{k} are negligible around k=±kFk=\pm k_{F}. Therefore, the contour integral gives

P=θ/π\displaystyle P=\theta/\pi (S14)

for an 1D excitonic insulator. This result may also be understood by noting that the low energy physics around ±kF\pm k_{F} is of two massive Dirac models, each of which realizes a Goldstone-Wilczek Goldstone and Wilczek (1981) mechanism of charge pumping.

In a 2D system one has two momenta, which we choose to be parallel (kxk_{x}) and vertical (ky)(k_{y}) to the direction defined by the antisymmetry of fkf_{k}. The net charge pumped is then an integral over kyk_{y} of the previously obtained formula. The only change is that now ξkxξkx,ky\xi_{k_{x}}\rightarrow\xi_{k_{x},k_{y}} and it may be that for some values of kyk_{y} the sign of ξ\xi does not change as a function of kxk_{x}, meaning that for these kyk_{y} the monopoles lie outside the torus of integration so no charge pumping occurs. In the weak coupling limit the issue may be discussed in terms of the Fermi surface of the metallic (Δ=0\Delta=0) phase. If the fermi surface (fermi crossings for each kyk_{y} as kxk_{x} is varied) is open the density of transferred charge is 2/ay2/a_{y} during a full cycle where aya_{y} is the lattice constant in yy direction; if the fermi surface is closed, then only the range of kyk_{y} where crossings occur gives rise to a charge pumping; thus the net density of pumped charge during a full cycle is 2kFy/π2k_{Fy}/\pi where kFyk_{Fy} is the maximum extent of the fermi surface in the yy direction.

In an isotropic 2D system, for an incomplete cycle with arbitrary θ\theta, note that each 1D momentum chain crossing the fermi surface at (±kF2ky2,ky)=kF(±cosθk,sinθk)\left(\pm\sqrt{k_{F}^{2}-k_{y}^{2}},k_{y}\right)=k_{F}\left(\pm\cos\theta_{k},\sin\theta_{k}\right) contributes a charge pumping channel described by Eq. (S14), with effective rotation angle φky=tan1(cosθktanθ)\varphi_{k_{y}}=\tan^{-1}\left(\cos\theta_{k}\tan\theta\right). Summing over all the chains, one obtains

P=12πkFkF𝑑kyφkyπ=kF2π211𝑑ttan1(1t2tanθ)=kF2πtanθ2\displaystyle P=\frac{1}{2\pi}\int_{-k_{F}}^{k_{F}}dk_{y}\frac{\varphi_{k_{y}}}{\pi}=\frac{k_{F}}{2\pi^{2}}\int_{-1}^{1}dt\tan^{-1}\left(\sqrt{1-t^{2}}\text{tan}\theta\right)=\frac{k_{F}}{2\pi}\tan\frac{\theta}{2} (S15)

for 0<θ<π/20<\theta<\pi/2. Extending the above integral to higher θ\theta, one obtains Eq. (6) of the main text.

II.5 Current response in time domain

In this section, we try to expand Eq. (S7) to higher orders in frequency and show that this won’t give corrections to the adiabatic result. We focus on the nodes at k=(0,±kF)k=(0,\pm k_{F}) in 2D when the system is close to pure pxp_{x}-wave order. In 3D, the nodes become a nodal line and the result stays the same up to some O(1)O(1) constants. Close to (0,Δp)(0,\Delta_{p}), the trajectory of motion is nearly along the Δs\Delta_{s} direction due to the saddle point structure on the free energy landscape in Fig. 4(a) of the main text. Thus it is enough to consider the current response to Δ˙stΔs\dot{\Delta}_{s}\equiv\partial_{t}\Delta_{s} in Eq. (S8): J(ω)=χjx,Δs(ω)Δs(ω)J(\omega)=\chi_{j_{x},\Delta_{s}}(\omega)\Delta_{s}(\omega). We assume the order parameter passes the point (0,Δp)(0,\Delta_{p}) with nearly constant velocity Δ˙s\dot{\Delta}_{s}. In the BCS limit, the second term vanishes due to the ξ\xi factor and what remains is

χjx,Δs(ω,0)=kvkfkΔpEk2iωω24Ek2=C0(Δp,Δs)(iω)+C1(Δp,Δs)(iω)2+O(ω3)\displaystyle\chi_{j_{x},\Delta_{s}}(\omega,0)=\sum_{k}v_{k}f_{k}\frac{\Delta_{p}}{E_{k}}\frac{-2i\omega}{\omega^{2}-4E_{k}^{2}}=C_{0}(\Delta_{p},\Delta_{s})(-i\omega)+C_{1}(\Delta_{p},\Delta_{s})(-i\omega)^{2}+O(\omega^{3}) (S16)

where

C0=12Δpkvkfk1Ek3=Δpν𝑑θk12πvFcos2θk1Δp2cos2θk=νvF1Δp=12πkFΔp\displaystyle C_{0}=-\frac{1}{2}\Delta_{p}\sum_{k}v_{k}f_{k}\frac{1}{E_{k}^{3}}=-\Delta_{p}\nu\int d\theta_{k}\frac{1}{2\pi}v_{F}\cos^{2}\theta_{k}\frac{1}{\Delta_{p}^{2}\cos^{2}\theta_{k}}=-\nu v_{F}\frac{1}{\Delta_{p}}=-\frac{1}{2\pi}\frac{k_{F}}{\Delta_{p}} (S17)

is the adiabatic current leading to Eq. (S15) and C1C_{1} is a dissipative term that arises from quasiparicles excitations. At exactly (0,Δp)(0,\Delta_{p}), it is

C1\displaystyle C_{1} =1iωIm[kvkfkΔpEk22Ekω24Ek2]=πωΔpkvkfkEk2(δ(ω2Ek)δ(ω+2Ek))18kFΔp2.\displaystyle=\frac{1}{-i\omega}\mathrm{Im}\left[\sum_{k}\frac{v_{k}f_{k}\Delta_{p}}{E_{k}^{2}}\frac{2E_{k}}{\omega^{2}-4E_{k}^{2}}\right]=\frac{\pi}{\omega}\Delta_{p}\sum_{k}\frac{v_{k}f_{k}}{E_{k}^{2}}\left(\delta(\omega-2E_{k})-\delta(\omega+2E_{k})\right)\approx\frac{1}{8}\frac{k_{F}}{\Delta_{p}^{2}}. (S18)

Another source for dissipative current is the quasiparticle contributed optical conductivity from the node:

σxx\displaystyle\sigma_{xx} =iωχΔpkxfkσ2,Δpkxfkσ2=iωΔp2kF2χσ2σ2=iωΔp2kF2k4Ekω24Ek2ξk2Ek2=18ΔpkFvF+iIm[σxx].\displaystyle=\frac{i}{\omega}\chi_{\Delta_{p}\partial_{k_{x}}f_{k}\sigma_{2},\,\Delta_{p}\partial_{k_{x}}f_{k}\sigma_{2}}=\frac{i}{\omega}\frac{\Delta_{p}^{2}}{k_{F}^{2}}\chi_{\sigma_{2}\sigma_{2}}=\frac{i}{\omega}\frac{\Delta_{p}^{2}}{k_{F}^{2}}\sum_{k}\frac{4E_{k}}{\omega^{2}-4E_{k}^{2}}\frac{\xi_{k}^{2}}{E^{2}_{k}}=\frac{1}{8}\frac{\Delta_{p}}{k_{F}v_{F}}+i\text{Im}[\sigma_{xx}]\,. (S19)

Its real part is suppressed by the small number Δp/εF=Δp/(kFvF)\Delta_{p}/\varepsilon_{F}=\Delta_{p}/(k_{F}v_{F}) and can thus be neglected.

It appears from Eq. (S18) that there is a correction to the pumped charge as δP=𝑑tC1t2Δs\delta P=\int dtC_{1}\partial_{t}^{2}\Delta_{s}. However, if one includes higher order terms in frequency, the current response from Eq. (S16) can be written in time domain:

J(t)=t𝑑tχ(t,t)tΔs\displaystyle J(t)=\int_{-\infty}^{t}dt^{\prime}\chi(t,t^{\prime})\partial_{t^{\prime}}\Delta_{s} (S20)

where

χ(t+t,t)\displaystyle\chi(t+t^{\prime},t^{\prime}) =kvx,kfkΔpEk2sin(2Ekt)=ν12π𝑑ξ𝑑θkΔpvFcos2θksin(2Et)ξ2+Δp2cos2θk+Δs2\displaystyle=\sum_{k}v_{x,k}f_{k}\frac{\Delta_{p}}{E_{k}^{2}}\sin(2E_{k}t)=\nu\frac{1}{2\pi}\int d\xi d\theta_{k}\Delta_{p}v_{F}\cos^{2}\theta_{k}\frac{\sin(2Et)}{\xi^{2}+\Delta_{p}^{2}\cos^{2}\theta_{k}+\Delta_{s}^{2}}
=node contribution+high energy state contribution\displaystyle=\text{node contribution}+\text{high energy state contribution}
ν22πΔp2vFΔpΔp𝑑ξ𝑑xx2sin(2Et)ξ2+x2+Δs2+ν2π2πΔpvFΔp𝑑ξsin(2Et)ξ2+Δ2\displaystyle\approx\nu\frac{2}{2\pi}\Delta_{p}^{-2}v_{F}\int_{-\Delta_{p}}^{\Delta_{p}}d\xi dxx^{2}\frac{\sin(2Et)}{\xi^{2}+x^{2}+\Delta_{s}^{2}}+\nu\frac{2\pi}{2\pi}\Delta_{p}v_{F}\int_{\Delta_{p}}^{\infty}d\xi\frac{\sin(2Et)}{\xi^{2}+\Delta^{2}}
νΔp2vF0Δp𝑑uu3sin(2Et)u2+Δs2+high energy state contribution\displaystyle\approx\nu\Delta_{p}^{-2}v_{F}\int^{\Delta_{p}}_{0}duu^{3}\frac{\sin(2Et)}{u^{2}+\Delta_{s}^{2}}+\text{high energy state contribution}
νΔp2vFΔsΔp𝑑E(EΔs2/E)sin(2Et)\displaystyle\approx\nu\Delta_{p}^{-2}v_{F}\int^{\Delta_{p}}_{\Delta_{s}}dE(E-\Delta_{s}^{2}/E)\sin(2Et)
=14νvFΔp2(t+4Δs2𝑑t)1t(sin(2Δpt)sin(2Δst))\displaystyle=-\frac{1}{4}\nu v_{F}\Delta_{p}^{-2}\left(\partial_{t}+4\Delta_{s}^{2}\int dt\right)\frac{1}{t}\left(\sin(2\Delta_{p}t)-\sin(2\Delta_{s}t)\right) (S21)

is the response kernel which is time dependent due to the fact that Δs\Delta_{s},Δp\Delta_{p} changes with time. Note that E=ξk2+Δs2+Δp2fk2E=\sqrt{\xi_{k}^{2}+\Delta_{s}^{2}+\Delta_{p}^{2}f_{k}^{2}} in the above integrals although we neglected its subscripts for notational simplicity. If one uses the values of Δs,Δp\Delta_{s},\,\Delta_{p} at tt^{\prime} and evaluate the polarization at tt\rightarrow\infty by 𝑑tJ(t)\int dtJ(t), one recovers exactly the adiabatic current and the topological charge pumping. Therefore, the non-adiabatic correction is beyond the scope of Eq. (S21), but lies in the fact that the state at tt^{\prime} is not the ground state of the instantaneous mean field Hamiltonian assumed here. We will show that this physics can be addressed in terms of exact dynamics of pseudo-spins.

III Edge states

In this section we analyse the behavior of edge states. For simplicity we foucs on the weak coupling BCS limit of Eq. (S2) with open boundary condition. Linearizing the Hamiltonian near the two fermi points ±kF\pm k_{F}, we find that an edge state wave function may be written

ψ(x)=ϕ1eikFx+k0x+ϕ2eikFx+k0x\psi(x)=\phi_{1}e^{-ik_{F}x+k_{0}x}+\phi_{2}e^{ik_{F}x+k_{0}x} (S22)

with energy E2=Δ2vF2k02E^{2}=\Delta^{2}-v_{F}^{2}k_{0}^{2} where Δ2=Δs2+Δp2fkF2=Δs2+Δp2\Delta^{2}=\Delta_{s}^{2}+\Delta_{p}^{2}f_{k_{F}}^{2}=\Delta_{s}^{2}+\Delta_{p}^{2} and we have made use of our convention fkF=1f_{k_{F}}=1. The spinor part of the wave function is

ϕ1=(Δs+iΔp,ivFk0+E),ϕ2=(ΔsiΔp,ivFk0+E).\phi_{1}=(\Delta_{s}+i\Delta_{p},\,-iv_{F}k_{0}+E)\,,\quad\phi_{2}=(\Delta_{s}-i\Delta_{p},\,iv_{F}k_{0}+E)\,. (S23)

To satisfy the open boundary condition ψ(0)=0\psi(0)=0, one requires ϕ1+ϕ2=0\phi_{1}+\phi_{2}=0 which yields

Δs+iΔpΔsiΔp=EivFk0E+ivFk0.\frac{\Delta_{s}+i\Delta_{p}}{\Delta_{s}-i\Delta_{p}}=\frac{E-iv_{F}k_{0}}{E+iv_{F}k_{0}}\,. (S24)

This and the relation E2=Δ2vF2k02E^{2}=\Delta^{2}-v_{F}^{2}k_{0}^{2} is satisfied by two solutions: (k0,E+)=(Δp/vF,Δs)(k_{0},\,E_{+})=(-\Delta_{p}/v_{F},\,\Delta_{s}) and (k0,E)=(Δp/vF,Δs)(k_{0},\,E_{-})=(\Delta_{p}/v_{F},\,-\Delta_{s}). The corresponding wave functions are

ψ±=1C±(1,±1)sin(kFx)exΔp/vF.\displaystyle\psi_{\pm}=\frac{1}{C_{\pm}}\left(1,\pm 1\right)\sin(k_{F}x)e^{\mp x\Delta_{p}/v_{F}}\,. (S25)

Note that the subscript ±\pm tracks each wave function smoothly as θ\theta varies, but does not specify either the energy or the side where the state is localized at. They are determined by the signs of their energies and the exponential factors.

The relation between the two edge states follows from symmetries. One may define two unitary operations, the ‘phase rotated inversion’ P^1:(ψa(x),ψb(x))(ψa(x),ψb(x))\hat{P}_{1}:(\psi_{a}(x),\psi_{b}(x))\rightarrow(\psi_{a}(-x),-\psi_{b}(-x)) and the P^2:(ψa(x),ψb(x))(ψb(x),ψa(x))\hat{P}_{2}:(\psi_{a}(x),\psi_{b}(x))\rightarrow(\psi_{b}(-x),-\psi_{a}(-x)). Both operators inter converts the two edge states. P^1\hat{P}_{1} is a symmetry of the mean field Hamiltonian HH if the system is in a pure pp-wave state while P^2\hat{P}_{2} always anti commutes with HH. Therefore, ψ±\psi_{\pm} have opposite energies and will be at zero energy in a pure pp-wave state.

Note that in open 1D wires connecting two reservoirs in Fig. 3 of the main text, although the edge states seem to be responsible for the charge pumping, the actual carries are all electrons in the valence band moved by continuous deformation of their wave functions, which is a bulk property. Indeed, in macroscopically long wires, the expansion and shrinking of edges states happen only in a tiny vicinity of θ=0,π\theta=0,\pi, while the charge pumping is a continuous process as θ\theta varies. For example in the θ=0+\theta=0_{+} state, although there is an occupied edge state localized on the right, the other electrons in the valence band form a density distribution that has a ‘hole’ on the right, such that the total polarization is still nearly zero. In the θπ/2\theta\rightarrow\pi/2 state, the background density distribution has a ‘half’ hole on each edge. Together with the occupied edge state on the right, it look like there is a half charge on the right edge and a half hole on the left, so that the polarization P=1/2P=1/2.

IV The Ginzburg-Landau action

In this section we present the derivation of the semiclassical action used in the main text. The Ginzburg-Landau action for order parameter fields and the EM field is obtained by integrating out the fermions (eS[Δs,Δp,A]D[ψ¯,ψ]eS[ψ,Δs,Δp,A]e^{-S[\Delta_{s},\Delta_{p},A]}\equiv\int D[\bar{\psi},\psi]e^{-S[\psi,\Delta_{s},\Delta_{p},A]}) in Eq. (S2), resulting in

S[Δs,Δp,A]=Trln[τ+ξkσ3+Δsσ1+Δpfkσ2]+𝑑r𝑑τ(Δs2gs+Δp2gp)𝑑r𝑑τL(Δs,Δp,A).S[\Delta_{s},\Delta_{p},A]=\mathrm{Tr\,ln}\left[\partial_{\tau}+\xi_{k}\sigma_{3}+\Delta_{s}\sigma_{1}+\Delta_{p}f_{k}\sigma_{2}\right]+\int drd\tau\left(\frac{\Delta_{s}^{2}}{g_{s}}+\frac{\Delta_{p}^{2}}{g_{p}}\right)\equiv\int drd\tau L(\Delta_{s},\Delta_{p},A)\,. (S26)

The Trln\mathrm{Tr\,ln} means trace of logarithm of the infinite dimensional matrix where kk should be interpreted as the spatial derivative i-i\nabla acting on the fermion fields, i.e., the matrix is just the kernel τ+Hm\partial_{\tau}+H_{m} for all Fermion fields at all (r,t)(r,t) in Eq. (S2) (See Sec. 6.4 of Ref. Altland and Simons (2010)). Performed in Fourier basis, it involves a summation over momenta kk, the fermion Matsubara frequencies ωn=(2n+1)πT\omega_{n}=(2n+1)\pi T (nn\in\mathbb{Z}, TT is the temperature and we have set the Boltzmann constant to be 11) and a trace of logarithm of the 2×22\times 2 matrices. Note that we have removed the absolute value symbol from Eq. (S2) since Δs,Δp\Delta_{s},\Delta_{p} are real numbers. We interpret the action as the Lagrangian for the order parameter fields moving in the presence of electric field E=tAE=-\partial_{t}A. Changing the argument AA to EE, we write the Lagrangian

L(Δs,Δp;E)=FK+Ldrive\displaystyle L(\Delta_{s},\Delta_{p};E)=F-K+L_{\text{drive}}\, (S27)

as the sum of three terms: the static free energy landscape FF, the ‘Kinetic energy’ KK and drive terms. We consider each in turn.

IV.1 Static free energy landscape

By integrating out the fermions for static Hubbard-Stratonovich fields in Eq. (S26) we obtain the free energy density (see Ref. Sun et al. (2020) or Sec. 6.4, page 276 of Ref. Altland and Simons (2010)):

F\displaystyle F =Tdkd(2π)d(ωnTrln[iωn+ξkσ3+Δsσ1+Δpfkσ2]+ξk)+Δs2gs+Δp2gp\displaystyle=T\int\frac{dk^{d}}{(2\pi)^{d}}\left(\sum_{\omega_{n}}\mathrm{Trln}\left[i\omega_{n}+\xi_{k}\sigma_{3}+\Delta_{s}\sigma_{1}+\Delta_{p}f_{k}\sigma_{2}\right]+\xi_{k}\right)+\frac{\Delta_{s}^{2}}{g_{s}}+\frac{\Delta_{p}^{2}}{g_{p}}
=Tdkd(2π)d(ωnln(iωn2Ek2)+ξk)+Δs2gs+Δp2gp\displaystyle=T\int\frac{dk^{d}}{(2\pi)^{d}}\left(\sum_{\omega_{n}}\ln\left(i\omega_{n}^{2}-E_{k}^{2}\right)+\xi_{k}\right)+\frac{\Delta_{s}^{2}}{g_{s}}+\frac{\Delta_{p}^{2}}{g_{p}}
T0dkd(2π)d(Ekξk)+Δs2gs+Δp2gp\displaystyle\xrightarrow{T\rightarrow 0}-\int\frac{dk^{d}}{(2\pi)^{d}}\left(E_{k}-\xi_{k}\right)+\frac{\Delta_{s}^{2}}{g_{s}}+\frac{\Delta_{p}^{2}}{g_{p}} (S28)

where Trln\mathrm{Trln} now means the trace of logarithm of the 2×22\times 2 matrix at momentum kk. Note that the momentum integral has an ultraviolet (UV) cutoff Λ\Lambda which is at the order of fermi energy but also affected by the Thomas-Fermi screening length Kozlov and Maksimov (1965). Converting the integral variable to an energy ξ\xi, we find for (quasi) 1D systems in the BCS limit:

F\displaystyle F =2ν0Λ𝑑ξ(ξ2+Δs2+Δp2ξ)+1gsΔs2+1gpΔp2Δs2+Δp2Λν(Δs2+Δp2)ln2ΛΔs2+Δp2+1gsΔs2+1gpΔp2.\displaystyle=-2\nu\int_{0}^{\Lambda}d\xi\left(\sqrt{\xi^{2}+\Delta_{s}^{2}+\Delta_{p}^{2}}-\xi\right)+\frac{1}{g_{s}}\Delta_{s}^{2}+\frac{1}{g_{p}}\Delta_{p}^{2}\xrightarrow{\sqrt{\Delta_{s}^{2}+\Delta_{p}^{2}}\ll\Lambda}-\nu\left(\Delta_{s}^{2}+\Delta_{p}^{2}\right)\ln\frac{2\Lambda}{\sqrt{\Delta_{s}^{2}+\Delta_{p}^{2}}}+\frac{1}{g_{s}}\Delta_{s}^{2}+\frac{1}{g_{p}}\Delta_{p}^{2}\,. (S29)

For a 2D isotropic Fermi surface, the first term is replaced by

νdθk2π(Δs2+Δp2cos2θk)ln2ΛΔs2+Δp2cos2θk\displaystyle-\nu\int\frac{d\theta_{k}}{2\pi}\left(\Delta_{s}^{2}+\Delta_{p}^{2}\cos^{2}\theta_{k}\right)\ln\frac{2\Lambda}{\sqrt{\Delta_{s}^{2}+\Delta_{p}^{2}\cos^{2}\theta_{k}}} (S30)

where θk\theta_{k} runs from 0 to 2π2\pi. In 3D, one just needs to make the replacement dθk2πsinθkdθkdϕk4π\frac{d\theta_{k}}{2\pi}\rightarrow\frac{\sin\theta_{k}d\theta_{k}d\phi_{k}}{4\pi} where θk\theta_{k} is the polar angle ranging from 0 to π\pi and ϕk\phi_{k} is the azimuthal angle ranging from 0 to 2π2\pi. In 2D, as long as gp<2gsg_{p}<2g_{s} Sun and Millis (2020b), the ss-wave phase at Δ=2Λe1gsν12\Delta=2\Lambda e^{-\frac{1}{g_{s}\nu}-\frac{1}{2}} is the ground state with energy νΔ2/2-\nu\Delta^{2}/2 while the pp-wave phase at Δp0=4Λe2gpν1\Delta_{p0}=4\Lambda e^{-\frac{2}{g_{p}\nu}-1} is a saddle point that has energy νΔp02/4-\nu\Delta_{p0}^{2}/4. For gp>2gsg_{p}>2g_{s}, the ground state minimum shifts to an s+ips+ip one. Note that in d dimensions, the density of states is ν=Ωd(2π)dkFd1/vF\nu=\frac{\Omega_{d}}{(2\pi)^{d}}k_{F}^{d-1}/v_{F} where Ωd\Omega_{d} is the surface area of the dd dimensional sphere with radius 11. For example, in 2D, Ωd=2π\Omega_{d}=2\pi and ν=12πkF/vF\nu=\frac{1}{2\pi}k_{F}/v_{F}.

IV.2 Kinetic energy

The action for order parameter fluctuations is obtained by expanding Eq. (S26) as

S=F(Δs,Δp)+S2(δΔs,δΔp)S=F(\Delta_{s},\Delta_{p})+S_{2}(\delta\Delta_{s},\delta\Delta_{p})\, (S31)

around the mean field configuration to second order in δΔs\delta\Delta_{s}, δΔp\delta\Delta_{p}, temporarily neglecting the EM field. We assume spatially uniform, time-dependent order fluctuations Δ^k(iΩn)=(Δs+δΔs(iΩn))σ1+(Δp+δΔp(iΩn))fkσ2=Δsσ1+Δpfkσ2+δΔ^k(iΩn)\hat{\Delta}_{k}(i\Omega_{n})=\left(\Delta_{s}+\delta\Delta_{s}(i\Omega_{n})\right)\sigma_{1}+\left(\Delta_{p}+\delta\Delta_{p}(i\Omega_{n})\right)f_{k}\sigma_{2}=\Delta_{s}\sigma_{1}+\Delta_{p}f_{k}\sigma_{2}+\delta\hat{\Delta}_{k}(i\Omega_{n}) and find the fluctuation term

S2=12TωnkTr[δΔ^k(iωn+iΩnHmk)δΔ^k(iωHmk)((ωn+Ωn)2+Ek2)(ωn2+Ek2)]+(δΔs)2gs+(δΔp)2gp.S_{2}=-\frac{1}{2}T\sum_{\omega_{n}}\sum_{k}\mathrm{Tr}\left[\frac{\delta\hat{\Delta}_{k}\left(i\omega_{n}+i\Omega_{n}-H^{k}_{m}\right)\delta\hat{\Delta}_{k}\left(i\omega-H^{k}_{m}\right)}{\left(\left(\omega_{n}+\Omega_{n}\right)^{2}+E_{k}^{2}\right)\left(\omega_{n}^{2}+E_{k}^{2}\right)}\right]+\frac{\left(\delta\Delta_{s}\right)^{2}}{g_{s}}+\frac{\left(\delta\Delta_{p}\right)^{2}}{g_{p}}\,. (S32)

For convenience we will include the fkf_{k} in the definition of Δp\Delta_{p} and δΔp\delta\Delta_{p} in this subsection. Evaluating the frequency integral at T=0T=0, taking the trace explicitly, rearranging and keeping only the terms with Ω\Omega dependence gives

S2(Ω)S2(Ω=0)=kΩ24((δΔs)2+(δΔp)2)+(δΔsΔs+δΔpΔp)22Ek(Ek2+Ω24).S_{2}(\Omega)-S_{2}(\Omega=0)=-\sum_{k}\frac{\frac{\Omega^{2}}{4}\left(\left(\delta\Delta_{s}\right)^{2}+\left(\delta\Delta_{p}\right)^{2}\right)+\left(\delta\Delta_{s}\Delta_{s}+\delta\Delta_{p}\Delta_{p}\right)^{2}}{2E_{k}\left(E_{k}^{2}+\frac{\Omega^{2}}{4}\right)}\,. (S33)

Writing k=ν𝑑ξ𝑑Ωk\sum_{k}=\nu\int d\xi d\Omega_{k} with Ωk\Omega_{k} the angular coordinates on the contours of constant energy, one obtains

S2(Ω)S2(Ω=0)=ν𝑑Ωk𝑑ξΩ24((δΔs)2+(δΔp)2)+(δΔsΔs+δΔpΔp)22ξ2+Δ2(ξ2+Δ2+Ω24)S_{2}(\Omega)-S_{2}(\Omega=0)=\nu\int d\Omega_{k}\int d\xi\frac{\frac{\Omega^{2}}{4}\left(\left(\delta\Delta_{s}\right)^{2}+\left(\delta\Delta_{p}\right)^{2}\right)+\left(\delta\Delta_{s}\Delta_{s}+\delta\Delta_{p}\Delta_{p}\right)^{2}}{2\sqrt{\xi^{2}+\Delta^{2}}\left(\xi^{2}+\Delta^{2}+\frac{\Omega^{2}}{4}\right)} (S34)

Defining ξ=Δtanx\xi=\Delta\tan x we find for the energy integral

12𝑑xcosx1+Ω24Δ2cos2x=01d(sinx)11+Ω24Δ2Ω24Δ2sin2x=122Δ|Ω|1+Ω24Δ2ln1+Ω24Δ2+|Ω|2Δ1+Ω24Δ2|Ω|2Δ.\frac{1}{2}\int dx\frac{\cos x}{1+\frac{\Omega^{2}}{4\Delta^{2}}\cos^{2}x}=\int_{0}^{1}d(\sin x)\frac{1}{1+\frac{\Omega^{2}}{4\Delta^{2}}-\frac{\Omega^{2}}{4\Delta^{2}}\sin^{2}x}=\frac{1}{2}\frac{\frac{2\Delta}{|\Omega|}}{\sqrt{1+\frac{\Omega^{2}}{4\Delta^{2}}}}\ln\frac{\sqrt{1+\frac{\Omega^{2}}{4\Delta^{2}}}+\frac{|\Omega|}{2\Delta}}{\sqrt{1+\frac{\Omega^{2}}{4\Delta^{2}}}-\frac{|\Omega|}{2\Delta}}\,. (S35)

In the adiabatic limit (lowest order in Ω\Omega expansion), the kinetic energy is thus

K=ν𝑑Ωk112Δ4(3Δ2((tΔs)2+(tΔp)2)2(ΔstΔs+ΔptΔp)2)K=\nu\int d\Omega_{k}\frac{1}{12\Delta^{4}}\left(3\Delta^{2}\left(\left(\partial_{t}\Delta_{s}\right)^{2}+\left(\partial_{t}\Delta_{p}\right)^{2}\right)-2\left(\Delta_{s}\partial_{t}\Delta_{s}+\Delta_{p}\partial_{t}\Delta_{p}\right)^{2}\right)\, (S36)

where Δ2=Δs2+Δp2\Delta^{2}=\Delta_{s}^{2}+\Delta_{p}^{2}. In 1D, writing Δs+iΔp=Reiθ\Delta_{s}+i\Delta_{p}=Re^{i\theta}, the kinetic term becomes

K=ν12R2((tR)2+3R2(tθ)2).K=\frac{\nu}{12R^{2}}\left((\partial_{t}R)^{2}+3R^{2}(\partial_{t}\theta)^{2}\right)\,. (S37)

If Δs\Delta_{s} is very small and the system has a closed Fermi surface in d=2d=2 or d=3d=3 then the adiabatic expansion breaks down in the regions where the gap vanishes. In this case the operator KK becomes nonlocal in time, and the physics is most efficiently treated directly from the action Eq. (S2).

IV.2.1 Dissipative terms

For convenience, we first introduce the general form of correlation functions at zero temperature:

χσi,σj(ω,q)=12k1ω2(Ek+Ek)2{\displaystyle\chi_{\sigma_{i},\,\sigma_{j}}(\omega,q)=\frac{1}{2}\sum_{k}\frac{1}{\omega^{2}-(E_{k}+E_{k^{\prime}})^{2}}\Bigg{\{} (Ek+Ek)Tr[σiσjHmkσiHmkσjEkEk]+ωTr[σiHmkσjEkHmkσiσjEk]}\displaystyle(E_{k}+E_{k^{\prime}})\mathrm{Tr}\left[\sigma_{i}\sigma_{j}-\frac{H_{m}^{k}\sigma_{i}H_{m}^{k^{\prime}}\sigma_{j}}{E_{k}E_{k^{\prime}}}\right]+\omega\mathrm{Tr}\left[\frac{\sigma_{i}H_{m}^{k^{\prime}}\sigma_{j}}{E_{k^{\prime}}}-\frac{H_{m}^{k}\sigma_{i}\sigma_{j}}{E_{k}}\right]\Bigg{\}} (S38)

where Hmk=ξkσ3+Δsσ1+Δpfkσ2H_{m}^{k}=\xi_{k}\sigma_{3}+\Delta_{s}\sigma_{1}+\Delta_{p}f_{k}\sigma_{2}. We re-derive the kinetic terms by expanding the order parameter correlation functions in frequency:

S2=ω(Δs(ω)Δp(ω))(1gs+χΔs,Δs(ω)χΔs,Δp(ω)χΔp,Δs(ω)1gp+χΔp,Δp(ω))(Δs(ω)Δp(ω))\displaystyle S_{2}=\sum_{\omega}\left(\begin{array}[]{cc}\Delta_{s}(-\omega)&\Delta_{p}(-\omega)\end{array}\right)\left(\begin{array}[]{cc}\frac{1}{g_{s}}+\chi_{\Delta_{s},\Delta_{s}}(\omega)&\chi_{\Delta_{s},\Delta_{p}}(\omega)\\ \chi_{\Delta_{p},\Delta_{s}}(\omega)&\frac{1}{g_{p}}+\chi_{\Delta_{p},\Delta_{p}}(\omega)\end{array}\right)\left(\begin{array}[]{c}\Delta_{s}(\omega)\\ \Delta_{p}(\omega)\end{array}\right)\, (S44)

where the subscripts of the correlation functions correspond to their channels in HmkH_{m}^{k}. For example, χΔs,Δpχσ1,fkσ2\chi_{\Delta_{s},\Delta_{p}}\equiv\chi_{\sigma_{1},f_{k}\sigma_{2}} which is Eq. (S38) with σi,σj\sigma_{i},\,\sigma_{j} replaced by σ1,fkσ2\sigma_{1},f_{k}\sigma_{2} everywhere. The Δs2\Delta_{s}^{2} term in S2S_{2} is

χΔs,Δs(ω,0)\displaystyle\chi_{\Delta_{s},\Delta_{s}}(\omega,0) =4kξk2+Δp2fk2(ω24Ek2)Ek=k1EkdΩkΩd(ω24Δs2)F0(ΔΩk,ω)=χΔs,Δs(0,0)ω2ν{12Δ2Δs23Δ4d=1Δs2+2Δp26|Δs|Δ3d=2+O(ω4)\displaystyle=4\sum_{k}\frac{\xi_{k}^{2}+\Delta_{p}^{2}f^{2}_{k}}{(\omega^{2}-4E_{k}^{2})E_{k}}=-\sum_{k}\frac{1}{E_{k}}-\int\frac{d\Omega_{k}}{\Omega_{d}}(\omega^{2}-4\Delta_{s}^{2})F_{0}(\Delta_{\Omega_{k}},\omega)=\chi_{\Delta_{s},\Delta_{s}}(0,0)-\omega^{2}\nu\left\{\begin{array}[]{lc}\frac{1}{2\Delta^{2}}-\frac{\Delta_{s}^{2}}{3\Delta^{4}}&\,d=1\\ \frac{\Delta_{s}^{2}+2\Delta_{p}^{2}}{6|\Delta_{s}|\Delta^{3}}&\,d=2\end{array}\right.+O(\omega^{4}) (S47)

where ΔΩk2=Δs2+Δp2fk(Ωk)2\Delta_{\Omega_{k}}^{2}=\Delta_{s}^{2}+\Delta_{p}^{2}f^{2}_{k({\Omega_{k}})}, Δ2=Δs2+Δp2\Delta^{2}=\Delta_{s}^{2}+\Delta_{p}^{2}, F0(Δ,ω)=ν4Δ22Δωsin1(ω2Δ)1(ω2Δ)2F_{0}(\Delta,\omega)=\frac{\nu}{4\Delta^{2}}\frac{2\Delta}{\omega}\frac{\mathrm{sin}^{-1}\left(\frac{\omega}{2\Delta}\right)}{\sqrt{1-\left(\frac{\omega}{2\Delta}\right)^{2}}} and Ωk\Omega_{k} is the angular variable in dd-dimension. The Δp2\Delta_{p}^{2} term is

χΔp,Δp(ω,0)\displaystyle\chi_{\Delta_{p},\Delta_{p}}(\omega,0) =4kfk2(ξk2+Δs2)Ek(ω24Ek2)=kfk2EkdΩkΩD(ω24Δp2fk(Ωk)2)F0(ΔΩk,ω)\displaystyle=4\sum_{k}\frac{f^{2}_{k}\left({\xi_{k}}^{2}+\Delta_{s}^{2}\right)}{{E_{k}}(\omega^{2}-4{E_{k}}^{2})}=-\sum_{k}\frac{f^{2}_{k}}{E_{k}}-\int\frac{d\Omega_{k}}{\Omega_{D}}(\omega^{2}-4\Delta_{p}^{2}f^{2}_{k({\Omega_{k}})})F_{0}(\Delta_{\Omega_{k}},\omega)
=χΔp,Δp(0,0)ω2ν{12Δ2Δp23Δ4d=116Δp2[1Δs3Δ3]d=2+O(ω4).\displaystyle=\chi_{\Delta_{p},\Delta_{p}}(0,0)-\omega^{2}\nu\left\{\begin{array}[]{lc}\frac{1}{2\Delta^{2}}-\frac{\Delta_{p}^{2}}{3\Delta^{4}}&\,d=1\\ \frac{1}{6\Delta_{p}^{2}}\left[1-\frac{\Delta_{s}^{3}}{\Delta^{3}}\right]&\,d=2\end{array}+O(\omega^{4})\right.\,. (S50)

The ΔpΔs\Delta_{p}\Delta_{s} term is

χΔs,Δp(ω,0)\displaystyle\chi_{\Delta_{s},\Delta_{p}}(\omega,0) =4kfk2ΔsΔpEk(ω24Ek2)=4ΔsΔpdΩkΩDfk(Ωk)2F0(ΔΩk,ω)\displaystyle=4\sum_{k}f^{2}_{k}\frac{-\Delta_{s}\Delta_{p}}{{E_{k}}(\omega^{2}-4{E_{k}}^{2})}=4\Delta_{s}\Delta_{p}\int\frac{d{\Omega_{k}}}{\Omega_{D}}f^{2}_{k({\Omega_{k}})}F_{0}(\Delta_{\Omega_{k}},\omega)
=χΔs,Δp(0,0)+ω2ν{ΔsΔp3Δ4d=1Δs3Δp3[1Δs(2Δs2+3Δp2)2Δ3]d=2+O(ω4).\displaystyle=\chi_{\Delta_{s},\Delta_{p}}(0,0)+\omega^{2}\nu\left\{\begin{array}[]{lc}\frac{\Delta_{s}\Delta_{p}}{3\Delta^{4}}&\,d=1\\ \frac{\Delta_{s}}{3\Delta_{p}^{3}}\left[1-\frac{\Delta_{s}(2\Delta_{s}^{2}+3\Delta_{p}^{2})}{2\Delta^{3}}\right]&\,d=2\end{array}\right.+O(\omega^{4})\,. (S53)

The above expansions in ω\omega fails as ωΔs\omega\sim\Delta_{s}, the minimal gap around the fermi surface, especially when Δs=0\Delta_{s}=0 such that there are nodes at k=(0,±kF)k=(0,\pm k_{F}) in 2D. We next evaluate the kernels in the pure pp-wave case Δs=0\Delta_{s}=0 to gain a rough idea of the crossover of dynamical behavior. The dissipative part of Δs\Delta_{s} kernel is

Im[χΔs,Δs(ω,0)]\displaystyle\mathrm{Im}\left[\chi_{\Delta_{s},\Delta_{s}}(\omega,0)\right] =Im[4kEk(ω+iη)24Ek2]=πk(δ(ω2Ek)δ(ω+2Ek))ωΔp,d=212νωΔp\displaystyle=\mathrm{Im}\left[4\sum_{k}\frac{{E_{k}}}{(\omega+i\eta)^{2}-4{E_{k}}^{2}}\right]=-\pi\sum_{k}\left(\delta(\omega-2{E_{k}})-\delta(\omega+2{E_{k}})\right)\xrightarrow{\omega\ll\Delta_{p},\,d=2}-\frac{1}{2}\nu\frac{\omega}{\Delta_{p}} (S54)

where η\eta is an infinitesimal positive number and we have made use of the quasi-particle density of states due to the nodes: g(E)=12πkFE/(vFΔp)g(E)=\frac{1}{2\pi}k_{F}E/(v_{F}\Delta_{p}). The linear in frequency dissipation continues with a cutoff of about Δp\Delta_{p} beyond which it scales as a constant. Kramers-Kronig relation implies that

χΔs,Δs(ω,0)12ν(iωΔp+ω2Δp2).\displaystyle\chi_{\Delta_{s},\Delta_{s}}(\omega,0)\approx-\frac{1}{2}\nu\left(i\frac{\omega}{\Delta_{p}}+\frac{\omega^{2}}{\Delta_{p}^{2}}\right)\,. (S55)

The dissipative part of Δp\Delta_{p} kernel is

Im[χΔp,Δp(ω,0)]\displaystyle\mathrm{Im}[\chi_{\Delta_{p},\Delta_{p}}(\omega,0)] =Im[4kfk2ξk2Ek2Ek(ω+iη)24Ek2]=πkfk2ξk2Ek2(δ(ω2Ek)δ(ω+2Ek))ωΔp,d=2π27νω3Δp3\displaystyle=\mathrm{Im}\left[4\sum_{k}\frac{f^{2}_{k}{\xi_{k}}^{2}}{{E_{k}}^{2}}\frac{{E_{k}}}{(\omega+i\eta)^{2}-4{E_{k}}^{2}}\right]=-\pi\sum_{k}\frac{f^{2}_{k}{\xi_{k}}^{2}}{{E_{k}}^{2}}\left(\delta(\omega-2{E_{k}})-\delta(\omega+2{E_{k}})\right)\xrightarrow{\omega\ll\Delta_{p},\,d=2}-\frac{\pi}{2^{7}}\nu\frac{\omega^{3}}{\Delta_{p}^{3}}\, (S56)

and the cubic behavior has the cutoff Δp\Delta_{p}. This together with the Δs=0\Delta_{s}=0 limit of Eq. (S50) gives

χΔp,Δp(ω,0)π27ν(iω3Δp3+ω26Δp2).\displaystyle\chi_{\Delta_{p},\Delta_{p}}(\omega,0)\approx-\frac{\pi}{2^{7}}\nu\left(i\frac{\omega^{3}}{\Delta_{p}^{3}}+\frac{\omega^{2}}{6\Delta_{p}^{2}}\right)\,. (S57)

IV.2.2 In time domain

With the adiabatic approximation so at time t0+tt_{0}+t we can write Δ(t0+t)=Δ(t0)+δΔ(t0+t)\Delta(t_{0}+t)=\Delta(t_{0})+\delta\Delta(t_{0}+t), the action from Eq. (S31) reads

S=𝑑tF[Δ]+12𝑑t𝑑tδΔtMR(tt)δΔtS=\int dtF\left[\Delta\right]+\frac{1}{2}\int dtdt^{\prime}\frac{\partial\delta\Delta}{\partial t}\ M^{R}(t-t^{\prime})\frac{\partial\delta\Delta}{\partial t^{\prime}} (S58)

where MRM^{R} is the retarded kernel in time domain as a 2×22\times 2 matrix and Δ(Δs,Δp)\Delta\equiv(\Delta_{s},\Delta_{p}) in this subsection. The instantaneous (force) term in the Euler-Lagrange equations comes from the equal time correlator (potential) and the dynamics comes from expanding in derivatives, in other words

δFδΔ=tt𝑑tMR(tt)tδΔ(t)\frac{\delta F}{\delta\Delta}=\partial_{t}\int^{t}dt^{\prime}M^{R}(t-t^{\prime})\partial_{t^{\prime}}\delta\Delta(t^{\prime}) (S59)

Noting that MR(0)=0M^{R}(0)=0 we have

δFδΔ=t𝑑ttMR(tt)tδΔ(t)\frac{\delta F}{\delta\Delta}=\int^{t}dt^{\prime}\partial_{t}M^{R}(t-t^{\prime})\partial_{t^{\prime}}\delta\Delta(t^{\prime}) (S60)

The adiabatic approximation is reasonable if the change in Δ\Delta over a time corresponding to the range of MM is small (tΔ/|Δ|1\partial_{t}\Delta/|\Delta|\ll 1) so that we can evaluate MM at fixed Δ\Delta. If we have a fully gapped configuration (open Fermi surface or Δs\Delta_{s} not small), MM decays on times larger than |Δ|1=1/Δs2+Δp2|\Delta|^{-1}=1/\sqrt{\Delta_{s}^{2}+\Delta_{p}^{2}} so we can shift the derivative to the tt^{\prime} and integrate by parts to get

δVδΔ=t𝑑tMR(tt)t2δΔ(t)Mt2Δ\frac{\delta V}{\delta\Delta}=\int^{t}dt^{\prime}M^{R}(t-t^{\prime})\partial^{2}_{t^{\prime}}\delta\Delta(t^{\prime})\rightarrow M\partial_{t}^{2}\Delta (S61)

with M=t𝑑tMR(tt)M=\int^{t}dt^{\prime}M^{R}(t-t^{\prime}). However, for closed Fermi surfaces, the vanishing of Δpfk\Delta_{p}f_{k} at some Fermi surface points means that when Δs\Delta_{s} is small MM has a part that decays slowly, actually on the time-scale of 1/Δs1/\Delta_{s} and a more careful analysis is needed. In the isotropic 2D case, we have

S2=12𝑑t1𝑑t2(tδΔs(t1)tδΔp(t1))𝐌R(t1t2)(tδΔs(t2)tδΔp(t2))S_{2}=\frac{1}{2}\int dt_{1}dt_{2}\left(\begin{array}[]{cc}\partial_{t}\delta\Delta_{s}(t_{1})&\partial_{t}\delta\Delta_{p}(t_{1})\end{array}\right)\mathbf{M}_{R}(t_{1}-t_{2})\left(\begin{array}[]{c}\partial_{t}\delta\Delta_{s}(t_{2})\\ \partial_{t}\delta\Delta_{p}(t_{2})\end{array}\right) (S62)

and the (retarded) correlator is given by

𝐌R(t)=Θ(t)ksin2Ekt4Ek4(ξk2+Δp2ΔsΔpfkΔsΔpfk(ξk2+Δs2)fk2)\mathbf{M}_{R}(t)={\Theta}(t)\sum_{k}\frac{\sin 2E_{k}t}{4E_{k}^{4}}\left(\begin{array}[]{cc}\xi_{k}^{2}+\Delta_{p}^{2}&-\Delta_{s}\Delta_{p}f_{k}\\ -\Delta_{s}\Delta_{p}f_{k}&(\xi_{k}^{2}+\Delta_{s}^{2})f^{2}_{k}\end{array}\right)\, (S63)

where Θ\Theta is the Heaviside step function. Performing the integral over momentum, one obtains the low energy kernel

tMR11(t)\displaystyle\partial_{t}M_{R}^{11}(t) \displaystyle\approx Θ(t)ν2ΔpΔsΔp2𝑑v(1Δs2v2)cos2vt+high energy contribution\displaystyle\Theta(t)\frac{\nu}{2\Delta_{p}}\int_{\Delta_{s}}^{\Delta_{p}}2dv\left(1-\frac{\Delta_{s}^{2}}{v^{2}}\right)\cos 2vt+\text{high energy contribution}
=\displaystyle= Θ(t)ν2Δp[sin2Δptsin2Δstt+Δs[2Δst(π2Si[2Δst])cos2Δst]]+ν6(Δs2+Δp2)tδ(t)\displaystyle\Theta(t)\frac{\nu}{2\Delta_{p}}\left[\frac{\sin 2\Delta_{p}t-\sin 2\Delta_{s}t}{t}+\Delta_{s}\left[2\Delta_{s}t\left(\frac{\pi}{2}-Si[2\Delta_{s}t]\right)-\cos 2\Delta_{s}t\right]\right]+\frac{\nu}{6(\Delta_{s}^{2}+\Delta_{p}^{2})}\partial_{t}\delta(t)
\displaystyle\approx Θ(t)ν2Δpsin2Δptsin2Δstt+ν6(Δs2+Δp2)tδ(t),\displaystyle\Theta(t)\frac{\nu}{2\Delta_{p}}\frac{\sin 2\Delta_{p}t-\sin 2\Delta_{s}t}{t}+\frac{\nu}{6(\Delta_{s}^{2}+\Delta_{p}^{2})}\partial_{t}\delta(t)\,,
tMR22(t)\displaystyle\partial_{t}M_{R}^{22}(t) \displaystyle\approx ν6(Δs2+Δp2)tδ(t)\displaystyle\frac{\nu}{6(\Delta_{s}^{2}+\Delta_{p}^{2})}\partial_{t}\delta(t)

The off diagonal terms don’t affect the qualitative dynamics which we neglect. At small Δs\Delta_{s} we can neglect the second term of tMR11\partial_{t}M_{R}^{11}. Therefore, in 2D, a smooth crossover between non dissipative and dissipative behaviors during the swiping across θ=π/2\theta=\pi/2 can be described by the retarded Kinetic kernel

Sdis\displaystyle S_{\text{dis}} =12𝑑t𝑑tΔ˙s(t)MR(tt)Δ˙s(t),MR(t)ν2|Δp|0t𝑑tsin2Δptsin2Δstt.\displaystyle=\frac{1}{2}\int dtdt^{\prime}\dot{\Delta}_{s}(t)\ M_{R}(t-t^{\prime})\dot{\Delta}_{s}(t^{\prime})\,,\quad M_{R}(t)\approx\frac{\nu}{2|\Delta_{p}|}\int_{0}^{t}dt^{\prime}\frac{\sin 2\Delta_{p}t^{\prime}-\sin 2\Delta_{s}t^{\prime}}{t^{\prime}}\,. (S65)

Eq. (IV.2.2) implies the equation of motion

δFδΔi=ν6(Δs2+Δp2)t2Δi+νδi,s2Δpt𝑑tsin[2Δp(tt)]sin[2Δs(tt)]ttΔi(t)\frac{\delta F}{\delta\Delta_{i}}=\frac{\nu}{6(\Delta_{s}^{2}+\Delta_{p}^{2})}\partial_{t}^{2}\Delta_{i}+\frac{\nu\delta{i,s}}{2\Delta_{p}}\int^{t}_{-\infty}dt^{\prime}\frac{\sin\left[2\Delta_{p}(t-t^{\prime})\right]-\sin\left[2\Delta_{s}(t-t^{\prime})\right]}{t-t^{\prime}}\Delta_{i}(t^{\prime}) (S66)

which describes the dissipationless-dissipative crossover behavior when Δs\Delta_{s} crosses zero during the dynamics.

IV.3 The drive term

In the drive term Ldrive=P(θ)Es(Δs,Δp)E2+O(E3)L_{\text{drive}}=-P(\theta)E-s(\Delta_{s},\Delta_{p})E^{2}+O(E^{3}), the linear coupling of electric field to the polarization is obvious. We derive the second term in this section. The kernel of the O(A2)O(A^{2}) term is Sun and Millis (2020b)

Kij(ω)=(nm+χJi,Jj(ω))δij\displaystyle K_{ij}(\omega)=\left(\frac{n}{m}+\chi_{J_{i},J_{j}}(\omega)\right)\delta_{ij} (S67)

where JJ is the current operator in Eq. (S6) and mm is the electron mass in our model. Since the second term in the current in Eq. (S6) is suppressed by the factor Δp/εF\Delta_{p}/\varepsilon_{F} in the BCS limit, its contribution can be neglected. In 1D, the current correlation function is thus

χJ,J(ω)=χσ3vk,σ3vk(ω)=4vF2Δ2F0(Δ,ω)=vF2ν(1+23(ω2Δ)2+O((ω2Δ)4))\displaystyle\chi_{J,J}(\omega)=\chi_{\sigma_{3}v_{k},\sigma_{3}v_{k}}(\omega)=-4v_{F}^{2}\Delta^{2}F_{0}(\Delta,\omega)=-v_{F}^{2}\nu\left(1+\frac{2}{3}\left(\frac{\omega}{2\Delta}\right)^{2}+O\left(\left(\frac{\omega}{2\Delta}\right)^{4}\right)\right) (S68)

where Δ2=Δs2+Δp2\Delta^{2}=\Delta_{s}^{2}+\Delta_{p}^{2} and F0(ω)=k1Ek(4Ek2ω2)=ν4Δ22Δωsin1(ω2Δ)1(ω2Δ)2=ν4Δ2(1+23(ω2Δ)2+O((ω2Δ)4))F_{0}(\omega)=\sum_{k}\frac{1}{E_{k}(4E_{k}^{2}-\omega^{2})}=\frac{\nu}{4\Delta^{2}}\frac{2\Delta}{\omega}\frac{\mathrm{sin}^{-1}\left(\frac{\omega}{2\Delta}\right)}{\sqrt{1-\left(\frac{\omega}{2\Delta}\right)^{2}}}=\frac{\nu}{4\Delta^{2}}\left(1+\frac{2}{3}\left(\frac{\omega}{2\Delta}\right)^{2}+O\left(\left(\frac{\omega}{2\Delta}\right)^{4}\right)\right)\,. The constant term cancels the diamagnetic contribution n/mn/m and what remains in the kernel is the O(ω2)O(\omega^{2}) term that corresponds to the static polarizability from ‘scattering states’ of the electron hole pair. In 2D, the current correlator up to O(ω2)O(\omega^{2}) is

χJi,Jj(ω)=δij1dvF2νdθk2π(1+2cos2θk3ω24(Δs2+Δp2cos2θk))=δij1dvF2ν(1+16ω2Δs2+Δp2+|Δs|Δs2+Δp2).\displaystyle\chi_{J_{i},J_{j}}(\omega)=-\delta_{ij}\frac{1}{d}v_{F}^{2}\nu\int\frac{d\theta_{k}}{2\pi}\left(1+\frac{2\cos^{2}\theta_{k}}{3}\frac{\omega^{2}}{4(\Delta_{s}^{2}+\Delta_{p}^{2}\cos^{2}\theta_{k})}\right)=-\delta_{ij}\frac{1}{d}v_{F}^{2}\nu\left(1+\frac{1}{6}\frac{\omega^{2}}{\Delta_{s}^{2}+\Delta_{p}^{2}+|\Delta_{s}|\sqrt{\Delta_{s}^{2}+\Delta_{p}^{2}}}\right)\,. (S69)

Therefore, the O(E2)O(E^{2}) term in the action reads

L2=1ω2KijEiEj=16νΔ2(EE0)2{Δ2Δs2+Δp2d=1Δ2Δs2+Δp2+|Δs|Δs2+Δp2d=2\displaystyle L_{2}=\frac{1}{\omega^{2}}K_{ij}E_{i}E_{j}=-\frac{1}{6}\nu\Delta^{2}\left(\frac{E}{E_{0}}\right)^{2}\left\{\begin{array}[]{lc}\frac{\Delta^{2}}{\Delta_{s}^{2}+\Delta_{p}^{2}}&\,d=1\\ \frac{\Delta^{2}}{\Delta_{s}^{2}+\Delta_{p}^{2}+|\Delta_{s}|\sqrt{\Delta_{s}^{2}+\Delta_{p}^{2}}}&d=2\end{array}\right.\,\, (S72)

where the coefficient can be interpreted as s=limω0σ(ω)/(2iω)s=\lim_{\omega\rightarrow 0}\sigma(\omega)/(2i\omega). The higher order terms is EE are in higher powers of (EE0)2Δ2Δs2+Δp2\left(\frac{E}{E_{0}}\right)^{2}\frac{\Delta^{2}}{\Delta_{s}^{2}+\Delta_{p}^{2}} where E0E_{0} is defined in the main text.

V The adiabatic transport scheme

V.1 Description

If the sin pulse is wide enough in time, it is possible to make the dynamics perfectly adiabatic since the system simply follows the instantaneous minimum on the free energy landscape. As the field increases, the minimum shifts away from (Δ,0)(\Delta,0) counter clockwisely while the maximum at (0,Δp0)(0,\Delta_{p0}) shifts clockwisely. The maximum field needed is simply that making the instantaneous minimum and maximum coincide. In 1D, this field can be computed analytically:

Em(gp)=2t1x2e1/2t+t2+1/4\displaystyle E_{m}(g_{p})=2t\sqrt{1-x^{2}}e^{-1/2-t+\sqrt{t^{2}+1/4}} (S73)

where x=(1/t+1/t2+4)/2x=(-1/t+\sqrt{1/t^{2}+4})/2 and t=1/(νgp)1/(νgs)t=1/(\nu g_{p})-1/(\nu g_{s}). After reaching the maximum value (a little higher than that), the field starts to decrease, shifting back the two extrema. The order parameter is moved to the immediate left of the maximum, which gradually shifts back to (0,Δp0)(0,\Delta_{p0}) as the field decreases to zero. The second half of the sin pulse would therefore transport the order parameter to the minimum at (Δ,0)(-\Delta,0), completing a half cycle. However, if the decreasing field phase of the pulse is too slow, unstable fluctuations of order parameter tend to grow exponentiallySun and Millis (2020a) and get comparable to its mean field value within the ‘spinodal time’ 1|Δ|ln1G0\frac{1}{|\Delta|}\ln\frac{1}{G_{0}} where G0|Δ|εF1G_{0}\sim\frac{|\Delta|}{\varepsilon_{F}}\ll 1 is the Ginzburg parameter of the landau theory. Therefore, the time scale of the pulse has to be smaller than the spinodal time.

If gpg_{p} is too small, the requires maximum field is so large that the O(E2)O(E^{2}) term L2=16νE2E02Δ4Δs2+Δp2L_{2}=-\frac{1}{6}\nu\frac{E^{2}}{E_{0}^{2}}\frac{\Delta^{4}}{\Delta_{s}^{2}+\Delta_{p}^{2}} would pull the order parameter to the origin and destroy the above adiabatic trajectory. This imposes an lower bound for the pp-wave pairing strength gpc=gs/(1+3/8νgs)g_{pc}=g_{s}/(1+\sqrt{3/8}\nu g_{s}). For the adiabatic transport scheme to work, gpg_{p} has to be larger than gpcg_{pc}. These conclusions apply qualitatively to higher dimensions.

If the adiabatic scheme is realized, experimental measurement of the threshold electric field gives the estimation of gpg_{p} through Eq. (S73). In the fast scheme described in the main text, if the full frequency spectrum of the current can be measured, it is possible to reconstruct the angular dynamics through, e.g., Eq. (6) of the main text.

V.2 Derivation

Refer to caption
Figure S3: (a) The curves r1r_{1} and r2r_{2} on the (r,θ)(r,\theta) plane at various values of electric field EE^{\prime}, neglecting O(E2)O(E^{2}) terms in the free energy. Solid curves are r1r_{1} and dashed curves are r2r_{2}. The intersections between the solid black curve and the dashed curves are the saddle points. The parameters are gs=0.2g_{s}=0.2 and gp=0.18g_{p}=0.18. (b) The maximum field EmE_{m} as a function of gpg_{p} for gs=0.2g_{s}=0.2. (c) Same as (a) but with the O(E2)O(E^{2}) terms taken into account. The r2r_{2} curves are not affected by the O(E2)O(E^{2}) terms while r1r_{1} curves are deformed. Each r1r_{1} curve can be separated into two branches: the left branch has r2f<0\partial_{r}^{2}f<0 (maxima) while the right branch has r2f>0\partial_{r}^{2}f>0 (minima).

In 1D, incorporating the effect of a static electric field up to O(E2)O(E^{2}), the free energy is

f(Δs,Δp)=ν(r2ln2Λr+1νgsr2+(1νgp1νgs)r2sin2θ12Δ02Eθ16E2Δ04r2)\displaystyle f(\Delta_{s},\Delta_{p})=\nu\left(-r^{2}\ln\frac{2\Lambda}{r}+\frac{1}{\nu g_{s}}r^{2}+\left(\frac{1}{\nu g_{p}}-\frac{1}{\nu g_{s}}\right)r^{2}\sin^{2}\theta-\frac{1}{2}\Delta_{0}^{2}E^{\prime}\theta-\frac{1}{6}E^{\prime 2}\frac{\Delta_{0}^{4}}{r^{2}}\right) (S74)

where the ‘polar’ coordinate is defined as (Δs,Δp)=r(cosθ,sinθ)(\Delta_{s},\Delta_{p})=r(\cos\theta,\sin\theta), the dimensionless electric field is E=E/E0E^{\prime}=E/E_{0}, E0=Δ02/vFE_{0}=\Delta_{0}^{2}/v_{F} and Δ0=Δ\Delta_{0}=\Delta. Note that rr is RR in the main text. We look for saddle points on the free energy landscape within the domain θ[0,π/2]\theta\in[0,\pi/2]. There are two curves defined by rf=0\partial_{r}f=0 and θf=0\partial_{\theta}f=0 respectively, whose solutions read

r1(θ)=Λe1gsνtsin2θ12,r2(θ)=12tΔ02Esin2θ\displaystyle r_{1}(\theta)=\Lambda e^{-\frac{1}{g_{s}\nu}-t\sin^{2}\theta-\frac{1}{2}},\quad r_{2}(\theta)=\sqrt{\frac{1}{2t}\frac{\Delta_{0}^{2}E^{\prime}}{\sin 2\theta}} (S75)

where t=1νgp1νgst=\frac{1}{\nu g_{p}}-\frac{1}{\nu g_{s}}, as shown in Fig. S3(a). Note that we temporarily neglected the O(E2)O(E^{2}) terms in the free energy. The intersections of the two curves are the saddle points. At zero field, the two saddle point are just the two minima at (θ,r)=(0,Δ0),(π/2,Δp0)(\theta,r)=(0,\Delta_{0}),\,(\pi/2,\Delta_{p0}). For weak field, the two saddle points shift towards each other in angular direction. As the field further increases to the critical value EmE_{m}, the two saddle point meet which means the two lines are tangent to each other: r1=r2,θr1=θr2r_{1}=r_{2},\,\partial_{\theta}r_{1}=\partial_{\theta}r_{2} is satisfied at the intersection. This condition gives the angle at intersection as cos(2θm)=12(1t+1t2+4)\cos(2\theta_{m})=\frac{1}{2}\left(-\frac{1}{t}+\sqrt{\frac{1}{t^{2}}+4}\right) and critical field

Em=2tsin(2θm)e1/2t+t2+1/4.\displaystyle E_{m}^{\prime}=2t\sin(2\theta_{m})e^{-1/2-t+\sqrt{t^{2}+1/4}}\,. (S76)

It increases from zero as gpg_{p} decreases from gsg_{s}, and diverges as 1/gp1/\sqrt{g_{p}} as gp0g_{p}\rightarrow 0, as shown in Fig. S3(b).

The O(E2)O(E^{2}) term in Eq. (S74) lowers the energy dramatically close to r=0r=0, and therefore tends to pull the system to the zero order state. As a result, the free energy has a maximum in the rr direction, followed by the minimum as rr increases. Thus the r1r_{1} curve has two branches: the left one has r2f<0\partial_{r}^{2}f<0 (maxima) while the right one has r2f>0\partial_{r}^{2}f>0 (minima), as shown by the solid curves in Fig. S3(c) for weak fields. For strong enough field EE, it can happen that the two branches meet each other at certain θ(E)\theta(E) such that there will be no saddle points along rr if θ>θ(E)\theta>\theta(E), as shown by the solid curves in Fig. S3(c) for stronger fields. The summits of those curves satisfy (r,r2)f=(0,0)(\partial_{r},\partial_{r}^{2})f=(0,0) which yields

E=32r4Δ04,r=2Λe(1νgs+tsin2θ)34.\displaystyle E^{\prime}=\frac{3}{2}\frac{r^{4}}{\Delta_{0}^{4}},\quad r=2\Lambda e^{-\left(\frac{1}{\nu g_{s}}+t\sin^{2}\theta\right)-\frac{3}{4}}\,. (S77)

Making the summit at θ=π/2\theta=\pi/2, the pure pp-wave order line, one obtains the minimal field Ec=3/2e2t1/2E^{\prime}_{c}=\sqrt{3/2}e^{-2t-1/2} for the r1r_{1} curves to be closed, i.e., for the minima in rr direction to disappear at certain angles.

As the field increases, the intersections A,BA,B between the r2r_{2} curve and the right branch of r1r_{1} curve moves towards each other. If they successfully meet each other at certain field EmE_{m}, the order parameter is handed by BB to AA and the subsequent decreasing field phase pushes AA back to the pp-wave order, i.e., adiabatic transport works. However, if gpg_{p} is too weak, it can happen that AA annihilates with another intersection on the left branch of r1r_{1}. In this situation, the order parameter will be transported to zero order instead of to the pp-wave state. The critical pp-wave pairing strength can be estimated roughly in this way: the summit of r1r_{1} collides with the left most point of r2r_{2} as field increases. This condition leads to the equality r2=2/3Δ02E=12tΔ02Er^{2}=\sqrt{2/3}\Delta_{0}^{2}E^{\prime}=\frac{1}{2t}\Delta_{0}^{2}E^{\prime} which renders gpc=gs/(1+3/8νgs)g_{pc}=g_{s}/(1+\sqrt{3/8}\nu g_{s}).

VI Exact mean field dynamics

VI.1 Pseudo spin representation

The degree of freedom at each momentum kk in Eq. (2) of the main text can be mapped to an Anderson pseudo spin 𝐬k\mathbf{s}_{k}. In the second quantized language, each momentum kk labels two single particle states from the two bands whose annihilation operators are ψc,k\psi_{c,k} and ψv,k\psi_{v,k}, giving a 4-dimensional Hilbert space. However, the mean field dynamics here implies that the total occupation number nk=ψc,kψc,k+ψv,kψv,k=ψkσ0ψkn_{k}=\psi^{\dagger}_{c,k}\psi_{c,k}+\psi^{\dagger}_{v,k}\psi_{v,k}=\psi^{\dagger}_{k}\sigma_{0}\psi_{k} at kk is always one (nkn_{k} commutes with the Hamiltonian) where ψk=(ψc,k,ψv,k)\psi^{\dagger}_{k}=(\psi^{\dagger}_{c,k},\,\psi^{\dagger}_{v,k}) and σ0\sigma_{0} is the identity matrix. Therefore, it is enough to consider a 2-dimensional subspace of single occupancy which can be mapped to a pseudo spin-1/21/2 defined as 𝐬^kψk𝝈ψk\hat{\mathbf{s}}_{k}\equiv\psi^{\dagger}_{k}\bm{\sigma}\psi_{k} where 𝝈\bm{\sigma} are the three Pauli matrices.

The general mean field Hamiltonian at momentum kk can be written as

ψkHmkψk=ψk(ξkΔsiΔpfkΔs+iΔpfkξk)ψk=𝐛k𝐬^k\displaystyle\psi^{\dagger}_{k}H_{m}^{k}\psi_{k}=\psi^{\dagger}_{k}\begin{pmatrix}\xi_{k}&\Delta_{s}-i\Delta_{p}f_{k}\\ \Delta_{s}^{\ast}+i\Delta_{p}^{\ast}f_{k}&-\xi_{k}\end{pmatrix}\psi_{k}=\mathbf{b}_{k}\cdot\hat{\mathbf{s}}_{k} (S78)

and the dynamics implied by Eq. (1) of the main text is mapped to the procession of the Anderson pseudo spins 𝐬k=𝐬^k\mathbf{s}_{k}=\langle\hat{\mathbf{s}}_{k}\rangle in the time dependent self consistent mean field 𝐛k\mathbf{b}_{k} Barankov et al. (2004):

𝐬˙k\displaystyle\dot{\mathbf{s}}_{k} =(𝐛kγ𝐛k×𝐬k)×𝐬k,\displaystyle=(\mathbf{b}_{k}-\gamma\mathbf{b}_{k}\times\mathbf{s}_{k})\times\mathbf{s}_{k}\,,
𝐛k\displaystyle\mathbf{b}_{k} =(Re[Δs]+Im[Δp]fk,Im[Δs]+Re[Δp]fk,ξk)=(gs2ks1k+gp2fkkfks1k,gs2ks2k+gp2fkkfks2k,ξk)\displaystyle=\left(\mathrm{Re}[\Delta_{s}]+\mathrm{Im}[\Delta_{p}]f_{k},\,-\mathrm{Im}[\Delta_{s}]+\mathrm{Re}[\Delta_{p}]f_{k},\,\xi_{k}\right)=\left(\frac{g_{s}}{2}\sum_{k^{\prime}}s_{1k^{\prime}}+\frac{g_{p}}{2}f_{k}\sum_{k^{\prime}}f_{k^{\prime}}s_{1k^{\prime}},\,\,\,\,\frac{g_{s}}{2}\sum_{k^{\prime}}s_{2k^{\prime}}+\frac{g_{p}}{2}f_{k}\sum_{k^{\prime}}f_{k^{\prime}}s_{2k^{\prime}},\,\,\,\,\xi_{k}\right) (S79)

where the last equality is known as the gap equation. The EM vector potential A(t)A(t) enters by kkA(t)k\rightarrow k-A(t) and we use a phenomenological damping γ\gamma to account for the effect of energy loss due to, e.g., the phonon bath. Starting with the ground state with a real Δs\Delta_{s}, we simulated the dynamics induced by a pulse and the current ji=k𝐬kki𝐛kj_{i}=\sum_{k}\mathbf{s}_{k}\cdot\partial_{k_{i}}\mathbf{b}_{k} is integrated over time to obtain the pumped charge. Some numerical solutions to Eq. (S79) are shown in Figs. S5 and S6.

Note that Eqs. (S78) and (S79) are general which allows for appearance of imaginary components of Δs\Delta_{s} and Δp\Delta_{p}. However, as a result of an emergent ‘particle-hole’ symmetry in the BCS weak coupling case that maps (Δs,Δp)(\Delta_{s},\Delta_{p}) to (Δs,Δp)(\Delta_{s}^{\ast},\Delta_{p}^{\ast}), the order parameter dynamics is restricted to the s+ips+ip plane where Δs(t),Δp(t)\Delta_{s}(t),\Delta_{p}(t) are always real numbers which we prove in the next subsection. Away from BCS weak coupling, the order parameter trajectory departs from the s+ips+ip plane in a continuous way: e.g., Δp\Delta_{p} can temporarily develop an imaginary component that couples to σ1\sigma_{1} and Δs\Delta_{s} can have an imaginary component that couples to σ2\sigma_{2}. However, since the initial and final states are the same ones, and the charge pumping is still quantized for 1D systems because it is a topological effect.

VI.1.1 Proof of why the dynamics is confined within the s+ips+ip plane in the BCS weak coupling case

Refer to caption
Figure S4: (a) Illustration of the momentum 𝐤\mathbf{k} and its ‘particle hole’ image 𝐤\mathbf{k}^{\prime} on the diagram of quasi particle energy dispersion as a function of momentum. (b) Solid arrows are the pseudo spins 𝐬𝐤\mathbf{s_{k}} and 𝐬𝐤\mathbf{s_{k^{\prime}}} in the 1-2-3 space. Dashed arrows are their projections on the 2-3 plane.

The physical argument is that in the BCS weak coupling case, there will be no force pushing the order parameter away from the s+ips+ip plane during the dynamics. Following we provide a more mathematical proof. Define the ‘particle-hole’ operation

P^h:(ψc,𝐤,ψv,𝐤)(ψv,𝐤,ψc,𝐤),Δ𝐤Δ𝐤,𝐤=𝐤2kF𝐤^\displaystyle\hat{P}_{h}:\quad\left(\psi_{c,\mathbf{k}}\,,\,\psi_{v,\mathbf{k}}\right)\rightarrow\left(\psi_{v,\mathbf{k}^{\prime}}\,,\,\psi_{c,\mathbf{k}^{\prime}}\right)\,,\quad\Delta_{\mathbf{k}}\rightarrow\Delta^{\ast}_{\mathbf{k}^{\prime}}\,,\quad\mathbf{k}^{\prime}=\mathbf{k}-2k_{F}\hat{\mathbf{k}} (S80)

where kFk_{F} is the magnitude of Fermi momentum, 𝐤^\hat{\mathbf{k}} is the unit vector along the direction of 𝐤\mathbf{k} and we have used bold fonts for the momentum to emphasize its vector nature (although every kk in the main text and supplemental material already means a vector). The operation P^h\hat{P}_{h} changes 𝐤\mathbf{k} to its ‘particle-hole’ image 𝐤\mathbf{k}^{\prime}, as illustrated in Fig. S4. In the BCS weak coupling case (|Δs|,|Δp|G|\Delta_{s}|,|\Delta_{p}|\ll G), what is relevant are the low energy states close to the Fermi surface (band crossing point) where P^h\hat{P}_{h} becomes a symmetry of the action in Eq. (1) of the main text (note that ξ𝐤=ξ𝐤\xi_{\mathbf{k}}=-\xi_{\mathbf{k}^{\prime}}). Considering that f𝐤=f𝐤f_{\mathbf{k}^{\prime}}=-f_{\mathbf{k}} in the BCS weak coupling case, (Δs,Δp)(\Delta_{s},\Delta_{p}) is changed to its complex conjugate (Δs,Δp)(\Delta_{s}^{\ast},\Delta_{p}^{\ast}) under operation of P^h\hat{P}_{h} according to Eq. (S80) and Eq. (S78). Therefore, given a solution (Δs(t),Δp(t))(\Delta_{s}(t),\Delta_{p}(t)) to the order parameter dynamics, its ‘image’ (Δs(t),Δp(t))(\Delta_{s}^{\ast}(t),\Delta_{p}^{\ast}(t)) must also be a solution. Given the initial condition of (Δs(0),Δp(0))(\Delta_{s}(0),\Delta_{p}(0))=(|Δ|,0)(|\Delta|,0), since the solution is unique, the order parameter trajectory must lie within the plane of Δs,Δp\Delta_{s},\Delta_{p}\in\mathbb{R}, i.e., the s+ips+ip plane studied in the main text.

An alternative way to prove this is using the pseudo-spin language. According to Eq. (S79), to prove that (Δs(t),Δp(t))(\Delta_{s}(t),\Delta_{p}(t)) is always real, we just need to prove 𝐬𝐤\mathbf{s}_{\mathbf{k}} and 𝐬𝐤\mathbf{s}_{\mathbf{k}^{\prime}} are always related to each other by the ‘mirror’ operation M^\hat{M} with respect to ‘232-3’ plane: (s1,s2,s3)𝐤=(s1,s2,s3)𝐤(s_{1},s_{2},s_{3})_{\mathbf{k}}=(s_{1},-s_{2},-s_{3})_{\mathbf{k}^{\prime}}, as shown in Fig. S4(b). Note that the pseudo spins are defined as axial vectors such that the mirror operation M^=s^1\hat{M}=\hat{s}_{1} transforms the spins as (s1,s2,s3)(s1,s2,s3)(s_{1},s_{2},s_{3})\rightarrow(s_{1},-s_{2},-s_{3}). For the initial ground state, this is obviously true and the ‘magnetic field’ 𝐛𝐤/𝐤\mathbf{b}_{\mathbf{k}/\mathbf{k}^{\prime}} on the two pseudo-spins are also mirror images of each other under MM, so are (𝐛×𝐬)𝐤/𝐤(\mathbf{b}\times\mathbf{s})_{\mathbf{k}/\mathbf{k}^{\prime}}. Therefore, this ‘mirror’ relation between the pseudo spins at 𝐤/𝐤\mathbf{k}/\mathbf{k}^{\prime} is sustainable during the dynamics according to Eq. (S79), and guarantees that the order parameter stays on the s+ips+ip plane.

Refer to caption
Refer to caption
Refer to caption
Figure S5: Order parameter dynamics of an 1D excitonic insulator subject to a pump pulse described by the vector potential A(t)=Emaxw(tanh(tt0w)+1)A(t)=-E_{\text{max}}w\left(\tanh(\frac{t-t_{0}}{w})+1\right). Left panel is the trajectory on the free energy landscape plotted on the s+ips+ip plane for Emax=0.22E0E_{\text{max}}=0.22E_{0}. Middle panel is the polarization as a function of time. Right panel is the pumped charge as a function of EmaxE_{\text{max}}. The parameters are w=1/(2Δ)w=1/(2\Delta), gsν=0.3g_{s}\nu=0.3, gpν=0.28g_{p}\nu=0.28, Δ=2Λe1/(gsν)=0.071Λ\Delta=2\Lambda e^{-1/(g_{s}\nu)}=0.071\Lambda, γ=0.07Δ\gamma=0.07\Delta, Emax=0.22E0E_{\text{max}}=0.22E_{0}. The grid in time direction is 10410^{4}.
Refer to caption
Refer to caption
Refer to caption
Figure S6: Order parameter dynamics of a 2D excitonic insulator subject to a pump pulse described by the vector potential A(t)=Emaxw(tanh(tt0w)+1)A(t)=-E_{\text{max}}w\left(\tanh(\frac{t-t_{0}}{w})+1\right). Left panel is the trajectory on the free energy landscape plotted on the s+ips+ip plane for Emax=0.686E0E_{\text{max}}=0.686E_{0}. Middle panel is the polarization as a function of time. Right panel is the pumped charge as a function of EmaxE_{\text{max}}. The parameters are w=1/(2Δ)w=1/(2\Delta), gsν=0.3g_{s}\nu=0.3, gpν=0.5g_{p}\nu=0.5, Δ=2Λe1/(gsν)=0.071Λ\Delta=2\Lambda e^{-1/(g_{s}\nu)}=0.071\Lambda, γ=0.07Δ\gamma=0.07\Delta. The grid in time direction is 10410^{4}. Here the pp-wave pairing interaction gpg_{p} is weaker than Fig. 4 of the main text, so that stronger field is needed to induce the dynamics and the nonadiabatic correction to polarization is larger.

VI.2 ‘Super-current’ in 1D systems

The solution is trivial in the degenerate case gs=gpg_{s}=g_{p} in the BCS limit where the effect of the pairing function fkf_{k} is captured by f±kF=±1f_{\pm k_{F}}=\pm 1 on the right/left fermi point. We start from a ground state (Δs,Δp)=(Δ,0)(\Delta_{s},\Delta_{p})=(\Delta,0) where all spins are pointing in 131-3 plane: 𝐬k=(Δ,0,ξk)/Ek\mathbf{s}_{k}=(\Delta,0,\xi_{k})/E_{k}. Consider the electric field pulse at t=0t=0 applied through A=A0Θ(t)A=A_{0}\Theta(t). The leading driving term due to electric field is 𝐛k=(0,0,vkA)\mathbf{b}_{k}=(0,0,v_{k}A) where vk=±vFv_{k}=\pm v_{F} around the right/left fermi point. The diamagnetic term A2\sim A^{2} is subleading in driving the spinor dynamics but contributes a diamagnetic current we will discuss in the end. After the kick, the spinors start to rotate around the ‘33’ axis with angular frequency ω=2vFA0\omega=2v_{F}A_{0}. The mean field rotates at the same speed: (Δs,Δp)=Δ(cos(ωt),sin(ωt))(\Delta_{s},\Delta_{p})=\Delta(\cos(\omega t),\sin(\omega t)) such that 𝐛k(0,0,vkA)\mathbf{b}_{k}-(0,0,v_{k}A) is always parallel to each spinor, not affecting the spin rotation. Thus the solution is that each spin synchronize and keeps rotating around ‘33’ with angular frequency vFA0v_{F}A_{0}. Now we evaluate the current J=JP+JDJ=J_{P}+J_{D}. The paramagnetic current JP=kvkσ3J_{P}=\sum_{k}\langle v_{k}\sigma_{3}\rangle vanishes in this state. The diamagnetic current is jD=2πvFA0=2ω/(2π)j_{D}=\frac{2}{\pi}v_{F}A_{0}=2\omega/(2\pi). Therefore, the system behaves like a ‘superconductor’ with the superfluid density nn.

VI.3 Dynamics of the node in 2D: Landau-Zener formula

The node contribution to the polarization is captured by a Dirac Hamiltonian with time dependent gap:

Hk(t)=ΔpkFvFvFkxσ2+vFkyσ3+Δs(t)σ1\displaystyle H_{k}(t)=\frac{\Delta_{p}}{k_{F}v_{F}}v_{F}k_{x}\sigma_{2}+v_{F}k_{y}\sigma_{3}+\Delta_{s}(t)\sigma_{1} (S81)

which is an approximation to Eq. (2) of the main text around k0=(0,kF)k_{0}=(0,k_{F}), and is valid for kx,kykFk_{x},k_{y}\ll k_{F}. Taking into account the higher order term kx22mσ3\frac{k_{x}^{2}}{2m}\sigma_{3} in the Hamiltonian, the current in xx direction is Jx=vFkFkxσ3+ΔpkFσ2J_{x}=\frac{v_{F}}{k_{F}}k_{x}\sigma_{3}+\frac{\Delta_{p}}{k_{F}}\sigma_{2}. In the BCS limit we are concerned here, during the dynamics, the spinor at (kx,ky)(k_{x},k_{y}) is always the mirror image (under M^\hat{M} defined in Sec. VI) of that at (kx,ky)(-k_{x},-k_{y}) with respect to the 232-3 plane. Therefore, the σ2\sigma_{2} contributions to the current will always sum to zero, and it is enough to consider Jx=vFkFkxσ3J_{x}=\frac{v_{F}}{k_{F}}k_{x}\sigma_{3}. Define the energy variables kx=ΔpkFkxk_{x}^{\prime}=\frac{\Delta_{p}}{k_{F}}k_{x}, ky=vFkyk_{y}^{\prime}=v_{F}k_{y}, the Hamiltonian becomes

Hk(t)=kxσ2+kyσ3+Δs(t)σ1\displaystyle H_{k}(t)=k_{x}^{\prime}\sigma_{2}+k_{y}^{\prime}\sigma_{3}+\Delta_{s}(t)\sigma_{1} (S82)

and the current reads Jx=vFΔpkxσ3J_{x}=\frac{v_{F}}{\Delta_{p}}k_{x}^{\prime}\sigma_{3}. We now use Eq. (S82) to study the spinor dynamics and the current generated.

As the order parameter passes the (0,Δp)(0,\Delta_{p}) point with nearly constant velocity, the nodal gap Δs\Delta_{s} changes sign. For the spinor at certain kk, as Δs\Delta_{s} swipes from the positive value Δ\Delta at time t0-t_{0} to negative value Δ-\Delta at time t0t_{0}, the energy splitting starts from δk2+Δ2\sqrt{\delta_{k}^{2}+\Delta^{2}}, passes through the minimal splitting |δk|=|k||\delta_{k}|=|k^{\prime}|, and ends up with δk2+Δ2\sqrt{\delta_{k}^{2}+\Delta^{2}}. If the initial state is the low energy state, the probability of finally tunneling into the high energy state is given by the Landau-Zener formula Wittig (2005):

Pk=e2πδk2|t2Δs|\displaystyle P_{k}=e^{-2\pi\frac{\delta_{k}^{2}}{|\partial_{t}2\Delta_{s}|}} (S83)

which is exact if Δδk\Delta\gg\delta_{k}. Therefore, the tunneling probability is unity at the node and decays to zero away from the node within a range of |tΔs|\sim\sqrt{|\partial_{t}\Delta_{s}|}. Considering there are two nodes, the total number of quasiparticles excited is thus

N=2kPk=24π2kFvFΔp𝑑kx𝑑kyeπk2|tΔs|=24π2kFvFΔpπ|tΔs|π=12π2kFvF|tΔs|Δp=kF22π21kFvF|tΔs|Δp.\displaystyle N=2\sum_{k}P_{k}=\frac{2}{4\pi^{2}}\frac{k_{F}}{v_{F}\Delta_{p}}\int dk_{x}^{\prime}dk_{y}^{\prime}e^{-\pi\frac{k^{\prime 2}}{|\partial_{t}\Delta_{s}|}}=\frac{2}{4\pi^{2}}\frac{k_{F}}{v_{F}\Delta_{p}}\pi\frac{|\partial_{t}\Delta_{s}|}{\pi}=\frac{1}{2\pi^{2}}\frac{k_{F}}{v_{F}}\frac{|\partial_{t}\Delta_{s}|}{\Delta_{p}}=\frac{k_{F}^{2}}{2\pi^{2}}\frac{1}{k_{F}v_{F}}\frac{|\partial_{t}\Delta_{s}|}{\Delta_{p}}\,. (S84)

Since we have assumed Dirac dispersion in the integral, Eq. (S84) is accurate if |tΔs|Δp\sqrt{|\partial_{t}\Delta_{s}|}\ll\Delta_{p}.

VI.3.1 The pumped charge around the node

We now compute the pumped charge, which reads

P=2k𝑑tJx(k)t=24π2kFΔp2𝑑kx𝑑ky𝑑tkxσ3k,t=P0+Pdis.\displaystyle P=2\sum_{k}\int dt\langle J_{x}(k)\rangle_{t}=\frac{2}{4\pi^{2}}\frac{k_{F}}{\Delta_{p}^{2}}\int dk_{x}^{\prime}dk_{y}^{\prime}dtk_{x}^{\prime}\langle\sigma_{3}\rangle_{k,t}=P_{0}+P_{dis}\,. (S85)

The integral is completely determined by the dynamics governed by Eq. (S82), the evalution of which requires more detailed analysis of the time evolution of each spinor. Before that, we can guess the result simply from dimensional analysis. The nonadiabatic correction PdisP_{dis} comes from spinors with δkΔ\delta_{k}\ll\Delta, and the contribution arises during the anti crossing time regime when Δs(t)\Delta_{s}(t) is not much larger than δk\delta_{k}. Therefore, neither the momentum cutoff nor the maximum value of Δs\Delta_{s} should enter the result. The only remaining energy scale in Eq. (S82) is provide by tΔs\partial_{t}\Delta_{s} which has the unit of energy squared. Since the integral in PdisP_{dis} has the unit of energy squared, one obtains Pdis=κkF2π2|tΔs|Δp2P_{dis}=\kappa\frac{k_{F}}{2\pi^{2}}\frac{|\partial_{t}\Delta_{s}|}{\Delta_{p}^{2}} where κ\kappa is a universal O(1)O(1) constant.

Now we compute PdisP_{dis} exactly. It is more convenient to perform a permutation of the Pauli matrices: (σ2,σ3,σ1)(σ1,σ2,σ3)(\sigma_{2},\sigma_{3},\sigma_{1})\rightarrow(\sigma_{1},\sigma_{2},\sigma_{3}) such that the node Hamiltonian reads

Hk(t)=kxσ1+kyσ2+Δs(t)σ3\displaystyle H_{k}(t)=k_{x}^{\prime}\sigma_{1}+k_{y}^{\prime}\sigma_{2}+\Delta_{s}(t)\sigma_{3} (S86)

and the current becomes Jx=vFΔpkxσ2J_{x}=\frac{v_{F}}{\Delta_{p}}k_{x}^{\prime}\sigma_{2}. The dynamics of the pseudo spin at kk^{\prime} is a Landau-Zener problem Wittig (2005). At time t0-t_{0}, we have Δs=Δk\Delta_{s}=\Delta\gg k^{\prime} and the spin is in the ground state: ψ=(0,1)T\psi=(0,1)^{T}. The time evolution can be written as ψ=(A(t)eiϕ(t),B(t)eiϕ(t))T\psi=(A(t)e^{-i\phi(t)},\,B(t)e^{i\phi(t)})^{T} where ϕ(t)=𝑑tΔs(t)\phi(t)=\int dt\Delta_{s}(t) should not be confused with that from Sec. II. The Schrodinger equation for the amplitudes reads

tA=i(kxiky)Beiϕ,tB=i(kx+iky)Aeiϕ\displaystyle\partial_{t}A=-i(k_{x}^{\prime}-ik_{y}^{\prime})Be^{i\phi}\,,\quad\partial_{t}B=-i(k_{x}^{\prime}+ik_{y}^{\prime})Ae^{-i\phi}\, (S87)

which leads to

t2Ai2Δs(t)tA+k2A=0,t2B+i2Δs(t)tB+k2B=0.\displaystyle\partial_{t}^{2}A-i2\Delta_{s}(t)\partial_{t}A+k^{\prime 2}A=0\,,\quad\partial_{t}^{2}B+i2\Delta_{s}(t)\partial_{t}B+k^{\prime 2}B=0\,. (S88)

The current involves the expectation value of σ2\sigma_{2}:

σ2=i(BAeiθc.c.)=(1kxikyBtB+c.c.)\displaystyle\langle\sigma_{2}\rangle=i\left(B^{\ast}Ae^{i\theta}-c.c.\right)=-\left(\frac{1}{k_{x}^{\prime}-ik_{y}^{\prime}}B\partial_{t}B^{\ast}+c.c.\right)\, (S89)

whose time integral gives the charge:

𝑑tσ2=Re[1kxiky](|B(t0)|2|B(t0)|2)iIm[1kxiky]𝑑t|B(t)|2tln(BB).\displaystyle\int dt\langle\sigma_{2}\rangle=-\text{Re}\left[\frac{1}{k_{x}^{\prime}-ik_{y}^{\prime}}\right]\left(|B(t_{0})|^{2}-|B(-t_{0})|^{2}\right)-i\text{Im}\left[\frac{1}{k_{x}^{\prime}-ik_{y}^{\prime}}\right]\int dt|B(t)|^{2}\partial_{t}\ln\left(\frac{B^{\ast}}{B}\right)\,. (S90)

It can be seen from Eq. (S88) that the time dependent wave function is the same between the spins at (kx,ky)(k_{x}^{\prime},k_{y}^{\prime}) and (kx,ky)(-k_{x}^{\prime},k_{y}^{\prime}). Since the current is jx=vFΔpkxσ2j_{x}=\frac{v_{F}}{\Delta_{p}}k_{x}^{\prime}\sigma_{2}, the second term in Eq. (S90) will be canceled out by the two spins. The first term just needs the initial and final state information:

𝑑tσ2=Re[1kxiky](|B(t0)|2|B(t0)|2)=Re[1kxiky](1Pk).\displaystyle\int dt\langle\sigma_{2}\rangle=-\text{Re}\left[\frac{1}{k_{x}^{\prime}-ik_{y}^{\prime}}\right]\left(|B(t_{0})|^{2}-|B(-t_{0})|^{2}\right)=\text{Re}\left[\frac{1}{k_{x}^{\prime}-ik_{y}^{\prime}}\right]\left(1-P_{k}\right)\,. (S91)

which is provide by the Landau-Zener formula. Summing over all the spins, the pumped charge reads

P\displaystyle P =24π2kFΔp2𝑑kx𝑑kykxRe[1kxiky](1Pk)=P0+Pdis\displaystyle=\frac{2}{4\pi^{2}}\frac{k_{F}}{\Delta_{p}^{2}}\int dk_{x}^{\prime}dk_{y}^{\prime}k_{x}^{\prime}\text{Re}\left[\frac{1}{k_{x}-ik_{y}}\right](1-P_{k})=P_{0}+P_{dis} (S92)

where the nonadiabatic correction is identified as

Pdis\displaystyle P_{dis} =24π2kFΔp2𝑑kx𝑑kykxRe[1kxiky]Pk\displaystyle=-\frac{2}{4\pi^{2}}\frac{k_{F}}{\Delta_{p}^{2}}\int dk_{x}^{\prime}dk_{y}^{\prime}k_{x}^{\prime}\text{Re}\left[\frac{1}{k_{x}-ik_{y}}\right]P_{k}
=24π2kFΔp2𝑑kx𝑑kykx2k2eπk2|tΔs|=24π2kFΔp2𝑑k𝑑θkcos2θeπk2|tΔs|=kF8π3|tΔs|Δp2.\displaystyle=-\frac{2}{4\pi^{2}}\frac{k_{F}}{\Delta_{p}^{2}}\int dk_{x}^{\prime}dk_{y}^{\prime}\frac{k_{x}^{2}}{k^{2}}e^{-\pi\frac{k^{\prime 2}}{|\partial_{t}\Delta_{s}|}}=-\frac{2}{4\pi^{2}}\frac{k_{F}}{\Delta_{p}^{2}}\int dk^{\prime}d\theta k^{\prime}\cos^{2}\theta e^{-\pi\frac{k^{\prime 2}}{|\partial_{t}\Delta_{s}|}}=-\frac{k_{F}}{8\pi^{3}}\frac{|\partial_{t}\Delta_{s}|}{\Delta_{p}^{2}}\,. (S93)

Due to the negative relative sign of the non-adiabatic correction to the adiabatic one, we conclude that

Pdis=kF8π3|tΔs|Δp2=P018π2|tΔs|Δp2.\displaystyle P_{dis}=-\frac{k_{F}}{8\pi^{3}}\frac{|\partial_{t}\Delta_{s}|}{\Delta_{p}^{2}}=-P_{0}\frac{1}{8\pi^{2}}\frac{|\partial_{t}\Delta_{s}|}{\Delta_{p}^{2}}\,. (S94)

Therefore, Eq. (S94) gives the non adiabatic correction during each half cycle of order parameter rotation, which is valid if |tΔs|Δp\sqrt{|\partial_{t}\Delta_{s}|}\ll\Delta_{p}. This formula is nonperturbative in the swiping speed in the sense that, it can not be obtained by integrating over instantaneous linear or nonlinear current response functions perturbatively over the time evolution.