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

Charge and spin density wave orders in field-biased Bernal bilayer graphene

Zhiyu Dong Department of Physics, Massachusetts Institute of Technology, Cambridge, MA 02139 Department of Physics and Institute for Quantum Information and Matter, California Institute of Technology, Pasadena, California 91125    Patrick A. Lee Department of Physics, Massachusetts Institute of Technology, Cambridge, MA 02139    Leonid Levitov Department of Physics, Massachusetts Institute of Technology, Cambridge, MA 02139
Abstract

This paper aims to clarify the nature of a surprising ordered phase recently reported in biased Bernal bilayer graphene that occurs at the phase boundary between the isospin-polarized and unpolarized phases. Strong nonlinearity of transport at abnormally small currents, with dI/dVdI/dV vs. II sharply rising and then falling back, is typical for a charge/spin-density-wave state (CDW or SDW) sliding transport. Here, however, it is observed at an isospin-order phase boundary, prompting a question about the CDW/SDW mechanism and its relation to the quantum critical point. We argue that the observed phase diagram cannot be understood within a standard weak-coupling picture. Rather, it points to a mechanism that relies on an effective interaction enhancement at a quantum critical point. We develop a detailed strong-coupling framework accounting for the soft collective modes that explain these observations.

Graphene-based materials provide new opportunities for studying exotic orders driven by electron-electron interaction. One system that received much attention is the twisted bilayer graphene[1], where the flattened electron bands lead to the dominant role of interaction[2], giving rise to a variety of correlated insulating and superconducting phases[3, 4, 5, 6, 7, 8, 9, 10, 11]. Another system that came to light recently is the nontwisted multilayer graphene[12, 13, 14, 15, 16]. When biased by a transverse electric field, a gap at charge neutrality opens up, a transformation that flattens the electron dispersion[17]. The flattened bands reduce the kinetic energy and boost the effects of interactions, leading to interesting strongly correlated phases[12, 13, 14, 15, 16]. Among the phases where the order type has been identified is a cascade of isospin-polarized phases and adjacent superconducting phases [14, 15, 16].

In addition, in some of the observed phases, the order type has not yet been identified. Interestingly, near one of the isospin phase transitions seen in Bernal bilayer graphene (BBG), the system exhibits a rise in resistivity below about 50 mK, together with nonlinear resistivity[14]. The differential resistivity dV/dIdV/dI starts at a constant value at a small current, and abruptly drops around a threshold current of 5nA\sim 5\rm{nA}, and recovers the high-temperature value at a current of 10nA\sim 10\rm{nA} (see Fig 5a). This has led the authors to propose an interpretation in terms of the charge or spin density wave (CDW/SDW) and its sliding motion due to depinning [18]. These data pose the question of the mechanism of this conjectured density-wave phase and why it is associated with the phase boundary of isospin (valley) polarization.

The answer to this question is nontrivial because the standard weak-coupling scenario for CDW/SDW cannot explain the measured phase diagram. Indeed, the CDW/SDW order predicted in this way would be insensitive to the proximity to the onset of isospin polarization: the TcT_{c} in a weak-coupling theory should not exhibit an abrupt enhancement at the isospin phase transition. This is discussed in detail in Sec.II.

Refer to caption
Figure 1: a) The Fermi surface in the absence of interaction. The red and blue contour corresponds to Fermi surfaces in valley KK and KK^{\prime} respectively. Q1Q_{1}, Q2Q_{2}, and Q3Q_{3} are the maxima of the radius of curvature on the Fermi surface where CDW/SDW gap tends to open. In this case, we obtain a triple-Q CDW/SDW instability. b) The case of concave Fermi surface, where the curvature has six minima. In general, this situation might occur but, for simplicity, we only focus on the triple-Q CDW/SDW in this paper. Dashed blue lines are the edges of the mini Brillouin zone where the gap opening distorts the Fermi surface. c) The Fermi surface in the presence of a triple-Q CDW/SDW. Gaps open at three points in each valley. d) The schematic finite-temperature phase diagram. The yellow line and the cyan dashed line are respectively the boundaries of the CDW phase predicted by strong-coupling theory and weak-coupling theory. The strong coupling theory predicts a significant enhancement of CDW critical temperature at the valley-polarizing criticality TcT0T_{c}\gg T_{0}. When temperature lies between TcT_{c} and T0T_{0}, we obtain a phase diagram resembling measured one [14].

To the contrary, a strong-coupling approach accounting for the e-e interaction enhanced at the quantum-critical point offers a natural framework to understand the experimental findings. The key assumption is that the observed isospin order phase transition is either continuous or weakly first-order. Starting with this assumption, we predict a strong instability in the CDW and SDW channels. Specifically, we demonstrate that the interaction enhanced by the soft critical modes results in an abrupt rise of TcT_{c} for the CDW/SDW order at the onset of isospin polarization. This prediction is in agreement with the observations (see Fig.(1) d)).

The predicted CDW/SDW phase has several unique features. First, it is a metallic state since the gaps only open in a segment of the Fermi surface. This is distinct from the CDW in 1D which gaps out the entire Fermi surface[19]. As a result, the Drude weight of the sliding mode in our system, which is proportional to the length of the gapped segment, is much smaller as compared to the 1D case [20]. Moreover, here CDW/SDW takes a triple-QQ form, with the density-wave order parameter

for CDW:ρ(r)j=1,2,3ρjcos(2Qjr+ϕj).\displaystyle\text{for CDW:}\quad\rho(r)\sim\sum_{j=1,2,3}\rho_{j}\cos(2Q_{j}r+\phi_{j}). (1)
for SDW:𝑺(r)j=1,2,3𝑺jcos(2Qjr+ϕj).\displaystyle\text{for SDW:}\quad{\boldsymbol{S}}(r)\sim\sum_{j=1,2,3}{\boldsymbol{S}}_{j}\cos(2Q_{j}r+\phi_{j}). (2)

The three QjQ_{j} vectors are given by three minima of curvature on the Fermi surface [see Fig.1 a)], which are related by the C3C_{3} rotation symmetry. In our theory, the three CDW amplitudes ρ1\rho_{1}, ρ2\rho_{2}, and ρ3\rho_{3} are equal, whereas the phases are decoupled at leading order in interaction. The same applies to SDW amplitudes and phases. The relation between phases ϕj\phi_{j} in the ground state depends on subleading interactions. The most likely ground-state configuration is when the three components respect the C3C_{3} rotation symmetry, i.e. ρ1=ρ2=ρ3\rho_{1}=\rho_{2}=\rho_{3} or 𝑺1=C3𝑺2=C31𝑺3{\boldsymbol{S}}_{1}=C_{3}{\boldsymbol{S}}_{2}=C_{3}^{-1}{\boldsymbol{S}}_{3}. The real-space patterns for these possible ground states are illustrated in Fig.2. In these states, the wavevectors QjQ_{j} form a triade around 2K2K, each QjQ_{j} deviating from 2K2K by a small correction of order kFk_{F}. The CDW and SDW order parameters correspond to atomic-scale patterns that triple the unit cell, forming a 3×3\sqrt{3}\times\sqrt{3} Kekulé-like charge-density-wave pattern and a similar 120120^{\circ} spin-density-wave pattern as shown in Fig.2. These orders are similar to the orders recently seen in twisted bilayer graphene[21]. Because of the kF\sim k_{F} corrections to the CDW/SDW wavevector lengths, these patterns are modulated with a much longer spatial periodicity set by the Fermi wavelength.

Refer to caption
Figure 2: The predicted CDW and SDW patterns in real space. Black dots represent carbon atoms of the top layer of Bernal bilayer graphene to which carriers are shifted upon application of a transverse field. Carriers predominantly populate one sublattice of the top layer, referred to as A, in which sites are not aligned with the sites at the bottom layer. Circles in a) and arrows in b) illustrate CDW and SDW order, respectively. Circles of different radii denote different local charge density. Different arrow orientations denote different local spin polarization. Sites of the other sublattice (B) support states with strong interlayer hybridization, which form a higher-energy band that does not play a role in electronic ordering Both CDW and SDW orders feature spatial modulation with periodicity set by the Fermi wavelength. Notably, they exhibit Kekulé-like atomic-scale patterns that triple the unit cell. Accordingly, in the SDW order, the spin density wave takes the form of a 120120^{\circ} order in which spins on three different sublattices are oriented at 120120^{\circ} with respect to each other.

Our analysis will proceed as follows. First, after introducing a model of Bernal bilayer graphene with Coulomb interactions, we consider the weak-coupling picture of the CDW/SDW order and discuss its limitations. Next, we develop a strong-coupling framework and show that the CDW/SDW susceptibility is significantly enhanced at the isospin phase transition. Finally, we argue that the CDW/SDW predicted in this framework can explain the dependence of differential resistivity dV/dIdV/dI vs. II reported in Ref.[13]. We discuss two possible explanations. One is the conventional scenario of CDW/SDW sliding. Another explanation relies on Landau-Zener tunneling processes occurring in the CDW/SDW phase. Such processes may significantly impact the VV-II dependence provided the CDW/SDW gap is small. We find that both scenario can reproduce the correct orders of magnitude of the measured quantities.

I model

Bernal bilayer graphene bands are conventionally described by quadratic Dirac bands with trigonal warping. However, at low carrier densities typical for isospin and CDW orders the system can be described by a one-band Hamiltonian with a short-range interaction: H=H0+HintH=H_{0}+H_{\rm int},

H0=ξ,sϵξ(𝒑)ψξ,s,𝒑ψξ,s,𝒑\displaystyle H_{0}=\sum_{\xi,s}\epsilon_{\xi}({\boldsymbol{p}})\psi_{\xi,s,{\boldsymbol{p}}}^{\dagger}\psi_{\xi,s,{\boldsymbol{p}}} (3)
Hint=U2ξξss𝒑𝒑𝒒ψξ,s,𝒑+𝒒ψξ,s,𝒑𝒒ψξ,s,𝒑ψξ,s,𝒑\displaystyle H_{\rm int}=\frac{U}{2}\sum_{\xi\xi^{\prime}ss^{\prime}{\boldsymbol{p}}{\boldsymbol{p}}^{\prime}{\boldsymbol{q}}}\psi_{\xi,s,{\boldsymbol{p}}+{\boldsymbol{q}}}^{\dagger}\psi_{\xi^{\prime},s^{\prime},{\boldsymbol{p}}^{\prime}-{\boldsymbol{q}}}^{\dagger}\psi_{\xi^{\prime},s^{\prime},{\boldsymbol{p}}^{\prime}}\psi_{\xi,s,{\boldsymbol{p}}} (4)

where ξ,ξ\xi,\xi^{\prime} label valleys KK and KK^{\prime}, s,s=,s,s^{\prime}=\uparrow,\downarrow label spin. Fermionic variables ψξ,s\psi_{\xi,s} represent the spin-ss electrons in the conduction band in valleys ξ=K,K\xi=K,\,K^{\prime}. In what follows we will refer to the indices ξ\xi and ss as isospin indices. The quantity ϵξ(𝒑)\epsilon_{\xi}({\boldsymbol{p}}) describes the dispersion in valleys ξ=K,K\xi=K,\,K^{\prime}, where momentum pp is measured from KK and KK^{\prime} points, respectively. Since here we are concerned with CDW order, we focus on the specific points on the Fermi surface where CDW nesting occurs. The dispersion at these points will be discussed shortly [see Eq.(6)].

The interaction term, Eq.(4), describes electron-electron repulsion potential projected onto the BBG conduction bands. Here, we ignore the intervalley scattering (KKK\rightarrow K^{\prime}, KKK^{\prime}\rightarrow K) on the grounds that such scattering transfers large momenta, which makes it a weak process compared to the intravalley scattering. Indeed, the Coulomb interaction in 2D scales inversely with the transferred momentum and is therefore small when the momentum transfer is large. We also ignore the interaction UU momentum dependence arising from projection onto the conduction band. This is a reasonable approximation because the electron Bloch function momentum dependence is weak in the realistic regime ϵFD\epsilon_{F}\ll D.

The one-band model, Eq.(3), is well justified in the low-carrier-density regime of interest. In this regime, Fermi energy ϵF\epsilon_{F} is much smaller than the BBG bandgap by 10 to 100 times [14, 13, 15, 16]. At low carrier density, the dispersion in the conduction band can be approximated as a sum of an isotropic and a trigonal warping term[22, jung2014accurate]:

ϵK,K(𝒑)=D2+(p22m)2±λp3cos3θ𝒑,\epsilon_{K,K^{\prime}}({\boldsymbol{p}})=\sqrt{D^{2}+\left(\frac{p^{2}}{2m}\right)^{2}}\pm\lambda p^{3}\cos 3\theta_{{\boldsymbol{p}}}, (5)

where 𝒑{\boldsymbol{p}} is the momentum measured from KK or KK^{\prime} valley center, θ𝒑\theta_{{\boldsymbol{p}}} is the azimuthal angle of 𝒑{\boldsymbol{p}}, and ±\pm corresponds to KK and KK^{\prime}, respectively. The first term gives an isotropic Fermi sea, the trigonal warping term lowers the symmetry down to a three-fold rotation symmetry C3C_{3}. This distortion generates several minima of the curvature on the Fermi surface [see Fig.1 c)] — acting as hotspots for nesting that drives CDW order. We call these points “hotspots”. As we will see, the hotspots are the regions where the CDW/SDW gap tends to open. In general, the shape of the Fermi surface can be either convex or partly concave [see Fig.1 b) and c)]. For the convex case [see Fig.1 b)], the Fermi surface hosts three hotspots Q1Q_{1}, Q2Q_{2}, Q3Q_{3} associated with each other through rotation symmetry. For the partly concave case [see Fig.1 c)], the Fermi surface hosts six hotspots centered at points where curvature vanishes. For simplicity, we will focus on the convex case Fig.1 b). A generalization of our analysis to the partly concave case is straightforward and will be discussed below.

In the analysis of CDW/SDW, we will need the dispersion expanded around the hotspots. To that end, we define pp_{*} as the radius of curvature at one of the QiQ_{i} points (i=1,2,3i=1,2,3), from now on for conciseness called QQ. Expanding the electron energy (measured from the Fermi level) around QQ gives

ϵK(QK+k)=vFk+k22m,m=pvF,\epsilon_{K}(Q-K+k)=v_{F}k_{\parallel}+\frac{k_{\perp}^{2}}{2m_{*}},\quad m_{*}=\frac{p_{*}}{v_{F}}, (6)

valid provided k|QK|k\ll|Q-K|. The dispersion in valley KK^{\prime} is of a similar form, related to that in Eq.(6) by time-reversal symmetry. The quantity vFv_{F} is Fermi velocity, kk_{\parallel} is the momentum parallel to the Fermi velocity and normal to the Fermi surface, kk_{\perp} is the momentum perpendicular to the Fermi velocity. For simplicity, we will take vFv_{F}to be constant everywhere on the Fermi surface.

II CDW instability at weak coupling

To motivate the strong-coupling framework developed in the next section, we start with a conventional weak-coupling analysis of the CDW/SDW instability. For simplicity, here we only focus on CDW. The analysis and results for SDW are similar and will be discussed later.

A CDW state with wavevector 2Q2Q and amplitude Δ\Delta is described by the mean-field Hamiltonian

Hk\displaystyle H_{k} =(ϵK(QK+k)ΔΔϵK(QK+k))\displaystyle=\left(\begin{matrix}\epsilon_{K}(Q-K+k)&-\Delta\\ -\Delta&\epsilon_{K^{\prime}}(-Q-K^{\prime}+k)\end{matrix}\right)
=k22mτ0Δτ1+vFkτ3,\displaystyle=\frac{k_{\perp}^{2}}{2m_{*}}\tau_{0}-\Delta\tau_{1}+v_{F}k_{\parallel}\tau_{3}, (7)

where τ1,2,3\tau_{1,2,3} are 2×22\times 2 Pauli matrices, τ0=12×2\tau_{0}=1_{2\times 2}, the CDW order parameter Δ\Delta opens a gap in small segments on the Fermi surface near each of the hotspots (see Fig.1 c)). The length of each segment is

kΔ=22mΔ.k_{\Delta}=2\sqrt{2m_{*}\Delta}. (8)

The Hamiltonian, Eq. (7), possesses a U(1)U(1) symmetry ΔΔeiϕ\Delta\rightarrow\Delta e^{i\phi}, resulting in a Goldstone mode corresponding to the sliding motion. As discussed below, this mode is crucial for understanding transport in the CDW phase.

Refer to caption
Figure 3: a) The self-consistency relation in weak-coupling theory. b) The interaction is written in matrix form.

The onset of CDW instability is predicted by the standard saddle-point self-consistency relation similar to the BCS self-consistency equation for superconductivity:

Δ=ϵ,kUΔ12tr[τ1G(iϵ,k)τ1G(iϵ,k)]\Delta=-\sum_{\epsilon,k}U\Delta\frac{1}{2}{\rm tr}\left[\tau_{1}G(i\epsilon,k)\tau_{1}G(i\epsilon,k)\right] (9)

where G(iϵ,k)G(i\epsilon,k) is a matrix electron Green’s function defined as

G(iϵ,k)=1/(iϵHk).G(i\epsilon,k)=1/\left(i\epsilon-H_{k}\right). (10)

This equation, in a linearized form, is illustrated diagrammatically in Fig.3. Linearizing Eq.(9), and using Eq.(6) and Eq.(7), we obtain the following linearized self-consistency relation

Δ=ϵ,kUΔz2vF2k2,\Delta=-\sum_{\epsilon,k}\frac{U\Delta}{z^{2}-v_{F}^{2}k_{\parallel}^{2}}, (11)

where z=iϵk2/2mz=i\epsilon-k_{\perp}^{2}/2m_{*}, summation over kk is carried out over the vicinity of the hotspots QiQ_{i} vectors.

To make a comparison with the isospin polarization instability more direct, we rewrite this equation in terms of the CDW polarization function Π2Q\Pi_{2Q}. this quantity is nothing but the charge polarization function evaluated at momentum 2Q2Q. Taking the latter in the free-electron approximation gives

1+UΠ2Q=0,Π2Q=12ϵ,ktr(τ1Gϵ,k(0)τ1Gϵ,k(0)).1+U\Pi_{2Q}=0,\quad\Pi_{2Q}=\frac{1}{2}\sum_{\epsilon,k}{\rm tr}\left(\tau_{1}G^{(0)}_{\epsilon,k}\tau_{1}G^{(0)}_{\epsilon,k}\right). (12)

Here Gϵ,k(0)G_{\epsilon,k}^{(0)} is a shorthand form of the free-electron Green’s function, Eq.(10), with Δ\Delta taken to be zero.

This equation is of the same form as the Stoner criterion for valley polarization

1+UΠv=0,Πv=12ϵ,ktr(τ3Gϵ,k(0)τ3Gϵ,k(0))1+U\Pi_{v}=0,\quad\Pi_{v}=\frac{1}{2}\sum_{\epsilon,k}{\rm tr}\left(\tau_{3}G^{(0)}_{\epsilon,k}\tau_{3}G^{(0)}_{\epsilon,k}\right) (13)

Since at weak coupling the interaction strength UU entering both instabilities is similar in magnitude, the competition between CDW instability and the valley polarization instability is decided by the relative strength of the polarization functions Π2Q\Pi_{2Q} and Πv\Pi_{v}. We compare these quantities below.

The local flatness of the Fermi surface near the hotspots facilitates nesting and enhances CDW instability. However, while the CDW polarization function Π2Q\Pi_{2Q} is enhanced, the valley polarization function Πv\Pi_{v}, which is proportional to the total density of states, is mostly insensitive to the curvature of the Fermi surface at the hotspots. As a result, at T=0T=0, one expects CDW instability to be stronger and occur before the valley polarization instability.

Yet, the measurement in BBG [14] does not find isospin-polarized phase being outmatched by CDW phase in most part of the phase diagram. This can be attributed to the critical temperature TcT_{c} of the CDW phase being lower than the base temperature of the measurements. Indeed, it is reasonable to expect that TcT_{c} values for a weak-coupling CDW are low, since the Fermi surface in BBG does not exhibit a perfect nesting — the CDW only opens a gap in a small segment on the Fermi surface near the hotspots. As a result, only a small fraction of carriers undergoes condensation, giving a relatively small condensation energy and, therefore, a lower critical temperature TcT_{c} as compared to the valley-polarized state in which a large fraction of electrons gains condensation energy.

However, careful examination of the observed phase diagram indicates that the situation is not all that simple. In particular, one question that the weak-coupling scenario does not address is why a CDW phase is seen at the onset of isospin order in the experiment[14]. In other words, why the TcT_{c} of the CDW state is enhanced at the isospin ordering phase transition. To illustrate the disagreement between the weak-coupling scenario and measurement, in Fig. 1 d), we show a schematic phase diagram, where the cyan dashed line represents the predicted CDW phase boundary in the weak-coupling theory. Here, T0T_{0} represents the expected TcT_{c} in weak-coupling theory which we assume to lie below the base temperature of the measurement. In comparison, the yellow curve illustrates the behavior of TcT_{c} suggested by the measurement[14], which lies below the base temperature everywhere except at the onset of isospin order where it is enhanced to a value TT_{*} in the observable range.

This disagreement between the weak-coupling theory and observation calls for a strong-coupling analysis. As we will see in the next section, a strong-coupling analysis predicts an abrupt enhancement of coupling near the onset of isospin order, which pushes the critical temperature of the CDW state there from the low value T0T_{0} to a higher value TT_{*}. This will resolve the issue pointed out above and explain the measured phase diagram.

III Strong-coupling framework

Here we introduce a strong-coupling framework which accounts for quantum fluctuations near the isospin phase transition. We assume this transition to be second order (i.e. continuous) or weakly first order, giving rise to soft modes that can assist the CDW/SDW instability. To understand the emergence of the CDW/SDW order at the onset of isospin polarization order, one must determine the order type which can be either a valley polarization (VP) or intervalley coherence (IVC) order, and in each case identify the soft modes originating from the order parameter fluctuations. As discussed below, in both cases soft modes arise in a natural way and have distinct impact on the CDW/SDW instability. Yet, we do not need to immediately commit to one of the two possibilities, since no matter which order wins, both VP and IVC order parameter fluctuations are likely to be soft near the phase transition. Furthermore, measurements[14] indicate that the phase transition occurs near the van Hove singularity, implying that the system is close to both VP and IVC instabilities. We will first focus on the soft modes associated with a nearby VP instability and then, in Sec.VI, discuss the effects due to a nearby IVC instability. As we will see, proximity to the VP instability enhances the CDW/SDW instability whereas the IVC instability suppresses the instability.

Near valley polarization transition, valley-polarizing modes are softened, giving rise to a significant renormalization of the interaction[23]. To see this, consider the renormalization of the bare intervalley Coulomb interaction. There are two types of renormalization — screening and vertex correction. Accounting for both of them, we find in Ref.23 and re-derive in Supplement[24] that the effective interaction VV between two electrons in two valleys (one is in valley KK, the other is in KK^{\prime}) takes the following form at the critical point:

V(iν,q)=U/Nκ|ν|2vFq+l02q2V(i\nu,q)=\frac{U/N}{\frac{\kappa|\nu|}{2v_{F}q}+l_{0}^{2}q^{2}} (14)

Here, qq is the momentum transfer, and the quantity κ\kappa is defined as κ=p+pp0κ+1\kappa=\frac{p_{*}+p_{*}^{\prime}}{p_{0}}\sim\kappa_{*}+1, where pp_{*}^{\prime} represents the radius of curvature at QQ^{\prime}, which is opposite to QQ on the Fermi surface in valley K. The quantity p0p_{0} is the average radius of curvature on the whole Fermi sea, κ\kappa_{*} is defined as κ=p/p0\kappa_{*}=p_{*}/p_{0}. We have assumed pp0p_{*}^{\prime}\sim p_{0}. The quantity l0l_{0} depends on the shape of the Fermi surface. For simplicity, we assume l0kFl_{0}\sim k_{F} throughout this paper. The integer NN represents the number of isospin species. Here and below, we focus on the unpolarized phase, so N=4N=4 (spin \uparrow/\downarrow, and valley KK/KK^{\prime}). This problem is closely related to a problem tackled by Altshuler, Ioffe, and Millis[25] who showed that in a Fermi sea coupled to gauge field fluctuations, the 2kF2k_{F} vertex is logarithmically enhanced. In Ref.[25], the transverse gauge propagator takes the same form as Eq. 14. Therefore, many of the results of Ref.[25] are directly applicable here, except that in our case the absence of a singular self-energy modifies the answer, as will be seen below.

Due to the frequency dependence of V(iν,q)V(i\nu,q), the CDW/SDW vertex function Γ(ϵ)\Gamma(\epsilon) acquires a singular frequency dependence. To see this, we consider the first-order vertex correction shown in Fig.4 a). In this diagram, the frequencies carried by external legs are ϵ\sim\epsilon, whereas the internal frequencies are considerably higher. This first-order vertex correction, evaluated in the Supplement [24], is given by

δΓ(ϵ)=UmπNκln(ϵFϵ)Γ0.\delta\Gamma(\epsilon)=\frac{Um_{*}}{\pi N\kappa}\ln\left(\frac{\epsilon_{F}}{\epsilon}\right)\Gamma_{0}. (15)

where the high-energy cutoff ϵF\epsilon_{F} is the Fermi energy, Γ0\Gamma_{0} is the bare CDW/SDW vertex at the high energy scale ϵF\epsilon_{F}. The log divergence suggests treating higher-order corrections to the vertex function by a renormalization group (RG) approach. Namely, the RG flow of the renormalized vertex ΓR\Gamma^{R} in the regime of ϵϵF\epsilon\ll\epsilon_{F} is of the form

dΓR(ϵ)dlnϵ=ηΓR(ϵ),η=UmπNκ.\frac{d\Gamma^{R}(\epsilon)}{d\ln{\epsilon}}=-\eta\Gamma^{R}(\epsilon),\quad\eta=\frac{Um_{*}}{\pi N\kappa}. (16)

This gives rise to an infrared power-law divergence:

ΓR(ϵ)=Λ(ϵ)Γ0,Λ(ϵ)=(ϵFϵ)η.\Gamma^{R}(\epsilon)=\Lambda(\epsilon)\Gamma_{0},\quad\Lambda(\epsilon)=\left(\frac{\epsilon_{F}}{\epsilon}\right)^{\eta}. (17)

This divergence and its implications for CDW order was first discussed in Ref.[25].

The power-law divergence of the vertex function gives rise to a singular CDW/SDW susceptibility. The renormalized CDW susceptibility Π2QR\Pi_{2Q}^{R} is obtained by absorbing the vertex correction into the bare CDW susceptibility (see Fig.4 b)). As the energy on ladder in Fig.4 b) flows from high-energy to low-energy and then back to high-energy, Π2QR\Pi_{2Q}^{R} is enhanced by a factor of Λ2(ϵ)\Lambda^{2}(\epsilon),

Π2QR=ϵ,kΛ2(ϵ)(iϵk22m)2vF2k2,\Pi^{R}_{2Q}=\sum_{\epsilon,k}\frac{\Lambda^{2}(\epsilon)}{\left(i\epsilon-\frac{k_{\perp}^{2}}{2m_{*}}\right)^{2}-v_{F}^{2}k_{\parallel}^{2}}, (18)

which is formally similar to the renormalized CDW susceptibility considered in the problem of fermions coupled to a gauge field [see Ref.[25], where the renormalized susceptibility is denoted Π(ω,q)\Pi(\omega,q)]. A difference between our results and those in Ref.[25] is that the self-energy that plays a major role in Ref.[25] does not exist in our problem. This is because the singular interaction here is an intervalley coupling which does not contribute to self-energy. We will discuss this in detail shortly.

We stress that the changes in the interaction due to screening effects and vertex corrections are negligible at a frequency scale of ϵF\epsilon_{F}. Therefore, the interaction vertex at ϵF\epsilon_{F} is not the collective-mode-mediated interaction Eq.(14), but a bare interaction denoted as WcW_{c} and WsW_{s} for CDW and SDW channels respectively. These couplings are represented by dashed line in Fig.4c), whose strength we will discuss in Sec.IV. Accounting for the ladder diagram in Fig.4c), we obtain a linearized self-consistency equation in a strong-coupling framework:

1+Wc/sΠ2QR=01+W_{c/s}\Pi_{2Q}^{R}=0 (19)

This is essentially the same as in Eq.(11) up to extra vertex correction factors in Π2QR\Pi_{2Q}^{R}. By power counting, we find the left-hand side is proportional to T1/22ηT^{1/2-2\eta}. Therefore, we conclude that either CDW or SDW instability occurs at a significantly enhanced TcT_{c} when dimensionless coupling η>1/4\eta>1/4. The resulting CDW/SDW phase boundary is sketched in Fig.1 d) [yellow curve].

The condition η>1/4\eta>1/4 is achievable at the isospin phase transition in BBG. To see this, we plug the Stoner threshold Up02πvF=1\frac{Up_{0}}{2\pi v_{F}}=1 into the expression of η\eta, and find η=2κNκ\eta=\frac{2\kappa_{*}}{N\kappa}. Using N=4N=4, κκ+1\kappa\sim\kappa_{*}+1 (see text after Eq.(14)) and κ1\kappa_{*}\gg 1, we obtain η12\eta\sim\frac{1}{2}. We note parenthetically that the trigonal warping term in the electron Hamiltonian, which breaks the rotation symmetry into a three-fold rotational symmetry, is crucial for CDW. If there was no trigonal warping, then we would find p=pp_{*}^{\prime}=p_{*}, so κ=2κ\kappa=2\kappa_{*} and η=14\eta=\frac{1}{4}, which gives a marginal (logarithmic) divergence.

Refer to caption
Figure 4: A diagrammatic derivation of the self-consistency equation in strong-coupling theory. a) The renormalization by soft-mode mediated interaction VV gives a power-law correction to the CDW vertex Γ(ϵ)\Gamma(\epsilon) for ΔϵϵF\Delta\ll\epsilon\ll\epsilon_{F} see Eq.(17). b) the vertex correction can be absorbed into polarization ladder Π2QR\Pi^{R}_{2Q}. See Eq.(18). c) The TcT_{c} of CDW/SDW can be obtained using the renormalized CDW/SDW polarization function ΠR\Pi^{R}. See Eq.(19).

We emphasize that the self-energy correction is ignored in our analysis. This is justified because the self-energy is governed by the intravalley interaction, which does not take the same singular form as the intervalley interaction Eq.(14). Although the same process that generates intervalley interaction still contributes to the intravalley interaction, it is no longer the only leading process. There are those exchange processes that are allowed by the intravalley scattering, which cancel the divergent contribution. This behavior is different from that discussed in Ref.25 where the self-energy correction due to gauge fluctuations is singular and leads to a reduction of the power-law singularity of the response function. The value of their exponent is 2/32η2/3-2\eta as opposed to our 1/22η1/2-2\eta which requires a larger η\eta value greater than 1/3 for a divergent 2kF2k_{F} response.

IV Competition between CDW and SDW orders

The CDW and SDW instabilities are nondegenerate due to the difference in the bare interaction in these two channels WcW_{c} and WsW_{s}. Indeed, both intervalley Coulomb-induced intervalley scattering UinterU_{\rm{inter}} and phonon-induced intervalley scattering and UphU_{\rm ph} differentiate WcW_{c} and WsW_{s}. These two interactions can be written as follows:

H=12(Uinter+Uph)ψKψKψKψKH^{\prime}=\frac{1}{2}\left(U_{\rm{inter}}+U_{\rm{ph}}\right)\psi^{\dagger}_{K}\psi^{\dagger}_{K^{\prime}}\psi_{K}\psi_{K^{\prime}} (20)

where the valley-exchanging electron-electron interaction UinterU_{\rm{inter}} and the phonon mediated interaction[26] UphU_{\rm ph} are given by

Uinter\displaystyle U_{\rm{inter}} =e22πKϵ0100 meV nm2,ϵ0=5,\displaystyle=\frac{e^{2}}{2\pi K\epsilon_{0}}\sim 100\text{ meV nm}^{2},\quad\epsilon_{0}=5, (21)
Uph\displaystyle U_{\rm ph} 100 meV nm2\displaystyle\sim-100\text{ meV nm}^{2} (22)

Here, UphU_{ph} has a negative sign because phonon mediates an attraction. The interaction vertex for CDW instability is now corrected by these two interactions as follows:

Wc=UUinterUph.W_{\rm{c}}=U-U_{\rm{inter}}-U_{\rm{ph}}. (23)

Here, the extra minus signs are due to enclosing an extra fermionic loop. In comparison, the interaction vertex for SDW is not changed since these two interactions preserve spins

Ws=U.W_{s}=U. (24)

As a result, the competition between CDW and SDW depends on the competition between Coulomb-induced and phonon-induced intervalley scattering. If the former is stronger, then SDW wins, otherwise, the CDW wins.

V Observables

Experiment[14] reports on a non-ohmic behavior of conductivity observed near the onset of the isospin-ordered phase, which is reproduced in Fig.5 panels a and b. As discussed below, there are several ways this behavior can emerge in the CDW phase proposed above.

Before we begin, it is important to clarify that, unlike CDW in 1D, the CDW phase here is in a metallic state because CDW only gaps out a small segment on the Fermi surface, with the remaining part of the Fermi surface being ungapped. Therefore, conductivity always has an ohmic component, which is not of interest to us. In the analysis below, we will study the contribution from the collective phase mode due to the gapped segment on the Fermi surface, which gives a nonlinear dV/dIdV/dI (see Fig.5).

On general grounds, CDW phase is expected to feature sliding dynamics. As is well known, this dynamics will give rise to a delta-function contribution to conductivity σ(ω)=Aδ(ω)\sigma(\omega)=A\delta(\omega)[20]. In one dimension, the current response mediated by CDW sliding is described by the Drude weight A=ne2/mA=ne^{2}/m which is independent of Δ\Delta. For our problem, estimates yield AA that depends on Δ\Delta and becomes small when Δ\Delta decreases. This unconventional behavior arises due to the smallness of the region in kk-space where nesting occurs. Namely, we obtain

A=Ce2mΔA=Ce^{2}\sqrt{m_{*}\Delta} (25)

where the prefactor CC is an order-one dimensionless numerical factor. The Δ\sqrt{\Delta} dependence arises from an estimate of the size of the region at the Fermi surface in which nesting occurs. This spectral weight AA given in Eq.(25) is in general much smaller than the standard result[20] for CDW in one dimension. Details for the derivation of Eq.(25) are given in AppendixVII.

The CDW sliding conductivity can readily explain the measured nonlinear dV/dIdV/dI. Namely, at a low in-plane DC electric field, the CDW is pinned by defects, so the sliding contribution vanishes. At a threshold current Ic1I_{c1}, the sliding motion is activated, and the dV/dIdV/dI drops abruptly. When the current is too high, the CDW gap can be suppressed. This can be seen by considering a momentum-2K2K interband particle-hole pair excitation energy. Such a pair has energy 2Δ2\Delta in the absence of current. When a current is applied, the standard Doppler shift argument shows that the energy reduces to Δ=ΔKvd\Delta=\Delta-Kv_{d}, where vdv_{d} is the drift velocity. Therefore, when the applied current reaches a threshold of Ic2=neWΔKI_{c2}=\frac{neW\Delta}{K} (where WW represents the width of the sample), the CDW breaks down, and the conductivity dV/dIdV/dI restores the ohmic value of the high-temperature state. In summary, this mechanism of sliding CDW predicts a dV/dIdV/dI that rises abruptly at Ic1I_{c1} and goes to background value at Ic2I_{c2}. This behavior qualitatively agrees with measured dV/dIdV/dI.

For a more quantitative comparison, we extract realistic parameters from experiments. From the measured TcT_{c}, we estimate the CDW gap as Δ=0.01meV\Delta=0.01\rm{meV}. Use realistic parameters W=1μmW=1\rm{\mu m}, n=0.5×1012cm2n=0.5\times 10^{12}\rm{cm}^{-2}, we find Ic210nAI_{c2}\sim 10\rm{nA}, which agrees with the measurement. Meanwhile, we can also calculate the sliding current, which should be compared with the measured value of Ic2Ic1I_{c2}-I_{c1}. We find the ratio between sliding current and normal current is given by ΔϵF0.1\sqrt{\frac{\Delta}{\epsilon_{F}}}\sim 0.1, predicting a sliding current of 1nA\sim 1\rm{nA} when normal current is 10nA10\rm{nA}. The sliding current extracted from the experiment is Ic2Ic15nAI_{c2}-I_{c1}\approx 5\rm{nA} at the same total current, which is a bit larger but still comparable to the expected value.

However, withing the framework of CDW order, there a mechanism that can explain the nonlinear I(V)I(V) without invoking CDW sliding dynamics. This alternative mechanism relies the Landau-Zener (LZ) tunneling of electrons through the CDW gap. Namely, since the CDW gap is quite small, the electrons can tunnel across the CDW gap when the applied electric field becomes sufficiently strong. The LZ tunneling probability for an electron is[27, 28]

PLZ=exp(2πΔ2eEvF)P_{\rm LZ}=\exp\left(-\frac{2\pi\Delta^{2}}{eEv_{F}\hbar}\right) (26)

Therefore, we expect activation of the carriers on the gapped segment of the band for fields above the critical value

Ec=Δ2evFΔ2kFeϵF.E_{c}=\frac{\Delta^{2}}{ev_{F}\hbar}\sim\frac{\Delta^{2}k_{F}}{e\epsilon_{F}}. (27)

The activated carrier density increases abruptly when EE approaches EcE_{c}, leading to an increase in the conductivity I/VI/V. Conductivity quickly saturates when PLZ1P_{\rm LZ}\rightarrow 1. This behavior predicts an abrupt step in IVI-V, or equivalently, a narrow dip in the differential resistivity dV/dIdV/dI, resembling the observed dependence (see Fig.5).

Refer to caption
Figure 5: Nonlinear II-VV dependence extracted from experiment[14]: a) dV/dIdV/dI measured at varying carrier density; b) individual line cuts at the three densities indicated by arrows in panel a). c) Schematic dependence II vs. VV inferred from b), illustrating an abrupt drop in dV/dIdV/dI in a finite interval of currents.This behavior agrees with the predictions from two possible scenarios within the CDW framework (see text). In the sliding current scenario, Ic1I_{c1} arises from the activation of sliding due to depinning, Ic2I_{c2} arises from the CDW breakdown. In the Landau-Zener scenario, the differential resistivity is expected to drop and recover in a narrow regime of current [see Eq.(26)]. The only quantity that needs to be compared with the experiment is the electric field EcE_{c} at which the abrupt change of resistivity occurs.

To assess applicability of this scenario, we estimate the critical field EcE_{c} and compare to the measured value. Starting with Eq.(26), we use the measured TcT_{c} value as a crude estimate for the CDW gap: Δ0.01meV\Delta\sim 0.01\rm{meV}. The band dispersion and Fermi surface calculated numerically[14] give values ϵF1meV\epsilon_{F}\sim 1\rm{meV} and kF106cm1k_{F}\sim 10^{6}\rm{cm}^{-1}. Plugging these into Eq.(27) gives Ec=10mV/mmE_{c}=10\rm{mV/mm}, whereas the measured value of EcE_{c} extracted from the dV/dIdV/dI measurement[14] is 3mV/mm3\rm{mV/mm}. The calculated and measured EcE_{c} are of the same order of magnitude. Their difference can be attributed to the uncertainty in our estimate of Δ\Delta.

Furthermore, beyond the nonlinear effects in transport, the CDW/SDW order can be probed by measuring narrow-band noise under an applied current[29, 30, 31, 32]. The narrow-band noise frequency ν\nu is given by

ν=vd/λ,\nu=v_{d}/\lambda, (28)

where vdv_{d} is the drift velocity and λ\lambda is the CDW wavelength. Within the CDW sliding scenario, the current above which the resistivity starts to drop can be interpreted as sliding current. The value of this current I5nAI\approx 5\rm{nA} allows us to extract the value of vdv_{d} through the relation vd=I/neWv_{d}=I/neW, where WW is the width of the current-carrying region and nn is the carrier density. In Ref.[14] W=1μmW=1\rm{\mu m}, n=0.5×1012cm2n=0.5\times 10^{12}\rm{cm}^{-2}. These values yield vd=6m/sv_{d}=6\rm{m}/\rm{s}.

However, identifying the right value of the modulation wavelength λ\lambda is somewhat subtle. The physical meaning of λ\lambda is the spatial periodicity seen by carriers. Therefore, nominally, its value is nothing but the CDW/SDW wavelength λ=π/K5Å\lambda=\pi/K\sim 5\rm{\AA}, which is on the order of carbon atom spacing. Plugging this into Eq.(28) yields a high narrow-band noise frequency of ν15GHz\nu\sim 15\rm{GHz}. Yet, in addition to this frequency, the CDW sliding can also generate a lower frequency. This is because in the presence of a IVC order (long-range or short-range) the carriers can also sense the long-wavelength modulation of periodicity which is comparable to the Fermi wavelength λλF30nm\lambda\sim\lambda_{F}\sim 30\rm{nm} (see Fig.1). This would give a lower frequency on the order of ν250MHz\nu\sim 250\,\rm{MHz} which is comfortably low for detection.

We note parenthetically that a short-range IVC order is sufficient for generating this narrow-band noise frequency so long as the correlation length of the IVC order exceeds the periodicity of 30nm30\rm{nm}. Such a correlation length is achievable since the system is near a van Hove singularity (vHS). The large density of states at vHS leads to a proximity to the IVC instability as discussed in Sec.III, resulting in a large IVC correlation length.

VI CDW instability suppression in proximity to IVC order

The analysis in Secs.III and IV is based on the assumption that the region of CDW/SDW onset is close to a VP instability. However, as discussed in Sec.III, a competing intervalley coherence (IVC) instability near the region of interest may exist. Does a similar strong-coupling CDW/SDW scenario work near an IVC instability? It can be seen that the answer to this question is negative. As a matter of fact, the CDW/SDW instability is suppressed near IVC instability because the soft-mode interaction arising from this instability[33, 34] has an opposite sign compared to that of the VP instability. Indeed, each IVC soft-mode interaction line when plugged into the CDW vertex correction introduces an extra fermionic loop compared to the VP soft-mode interaction (Eq.14), giving an extra minus sign. As a result, the IVC soft mode tends to suppress the CDW/SDW instability.

It is instructive to compare this conclusion to the scenario of a proximal superconducting (SC) phase developed in Ref.[34]. As a reminder, this SC phase is observed at almost the same carrier density and transverse field as the resistive phase, except that the SC phase occurs above a finite in-plane magnetic field BB_{\parallel} whereas the resistive phase in question occurs at B=0B_{\parallel}=0[14]. In Ref.[34] it was shown that the measured phase diagram points to a scenario where this superconductivity is mediated by soft modes near an IVC instability. However, here we interpret the resistive phase as a CDW/SDW phase that arises near the onset of VP. At a first glance it might seem that these two approaches contradict each other, as they resort to different types of isospin instabilities. However, the two scenarios are not mutually exclusive once we recognize that the phase transition through which SC and CDW/SDW order occurs is close to the van Hove singularity[14]. In this situation, the divergent density of states predicts that both IVC and VP instabilities are strong, however which of them actually wins is unimportant.

Furthermore, the competition between SC and CDW/SDW and its dependence on an in-plane magnetic field BB_{\parallel} can also be understood within a scenario of coexisting and competing IVC and VP instabilities. Namely, it is shown in Ref.[35] that BB_{\parallel} tends to suppress the low-frequency singularity of VP-mode interaction, which is the interaction used by CDW/SDW. This suppression arises due to the screening effect between two spin species that are nondegenerate under BB_{\parallel}. In comparison, we do not expect the low-frequency singularity of the IVC-soft-mode interaction to be suppressed since this interaction is not related to screening. This distinction between the two soft-mode interactions predicts a suppression of CDW/SDW upon applying BB_{\parallel}. This prediction agrees with the behavior reported in Ref.[14].

VII Conductivity due to sliding mode

In order to compare with the measurement, we calculate the sliding mode contribution to conductivity δσ\delta\sigma. Although the exact value of it depends on relaxation time, the ratio between δσ\delta\sigma and the normal ohmic conductivity σ0\sigma_{0} is nothing else but the ratio between the Drude weights of these two contributions, which is independent of the relaxation time. In this section, we will obtain this ratio, which is used in main text when comparing with measurement.

As a reminder, the conductivity of CDW/SDW consists of two parts: the single-particle contribution and the collective mode contribution. For CDW/SDW in 2D, the CDW/SDW gap only gaps out a segment on the Fermi surface, therefore the single particle contribution consists of two parts: one arises from the gapped segment, which gives two wings outside the CDW/SDW gap Δ\Delta, the other comes from the gapless segments, which gives a Drude peak. Similar to the one dimensional case, the collective mode contribution to 2D CDW/SDW also gives a correction to the Drude weight. As we will see below, the quantity that can be directly compared with the measurement is the collective mode contribution to Drude weight. Therefore, we will focus on this contribution below.

To calculate conductivity, we consider the AC susceptibility

χ(iω)=δjδA,\chi(i\omega)=\frac{\delta j}{\delta A}, (29)

where AA is the electromagnetic vector potential. The quantity χ(iω)\chi(i\omega) is related, up to an iωi\omega factor, to AC conductivity σ(iω)\sigma(i\omega). Naturally, here and below we focus on the longitudinal susceptibility in which the vectors 𝒋{\boldsymbol{j}} and 𝑨{\boldsymbol{A}} are parallel. The collective mode contribution to χ\chi is diagrammatically expressed in Fig. 6 a), where each line represents the electron Green’s function

G(iϵ,k)=zΔτ1+vFkτ3z2E2,\displaystyle G(i\epsilon,k)=\frac{z-\Delta\tau_{1}+v_{F}k_{\parallel}\tau_{3}}{z^{2}-E^{2}}, (30)
z=iϵk22m,E=vF2k2+Δ2.\displaystyle z=i\epsilon-\frac{k_{\perp}^{2}}{2m_{*}},\quad E=\sqrt{v_{F}^{2}k_{\parallel}^{2}+\Delta^{2}}. (31)

Then the collective mode contribution to χ\chi defined in Eq.(29) can be written as:

δχ(iω)\displaystyle\delta\chi(i\omega) =vF2j,j=0,1,2,3ϵtr(Gτ3G+τj(ϵFϵ)η)\displaystyle=-v_{F}^{2}\sum_{j,j^{\prime}=0,1,2,3}\sum_{\epsilon}{\rm tr}\left(G_{-}\tau_{3}G_{+}\tau_{j}\left(\frac{\epsilon_{F}}{\epsilon}\right)^{\eta}\right)
×(χjj(iω))ϵtr(G+τ3Gτj(ϵFϵ)η)\displaystyle\times(-\chi_{jj^{\prime}}(i\omega))\sum_{\epsilon^{\prime}}{\rm tr}\left(G_{+}\tau_{3}G_{-}\tau_{j^{\prime}}\left(\frac{\epsilon_{F}}{\epsilon^{\prime}}\right)^{\eta}\right) (32)

Here and below the electron Green’s function G±=G(iϵ±)=G(iϵ±iω2)G_{\pm}=G(i\epsilon_{\pm})=G(i\epsilon\pm i\frac{\omega}{2}). The first minus sign arises from the fermionic loop, the trace notation is defined so that it includes integrals over momentum tr()=ktr(){\rm tr}(...)=\sum_{k}{\rm tr}(...), and the collective-mode propagators χjj\chi_{jj^{\prime}}’s are diagrammatically defined as in Fig.6 b), and can be written as

χjj(iω)\displaystyle\chi_{jj^{\prime}}(i\omega) =W2𝑑τeiωτTρj(τ)ρj(0),\displaystyle=-W^{2}\int d\tau e^{i\omega\tau}\langle T\rho_{j}(\tau)\rho_{j^{\prime}}(0)\rangle, (33)
ρj\displaystyle\rho_{j} =ΨτjΨ,j=0,1,2,3.\displaystyle=\Psi^{\dagger}\tau_{j}\Psi,\quad j=0,1,2,3. (34)

we find only χ22\chi_{22} gives a contribution since χ31,χ30=0\chi_{31},\chi_{30}=0. In the

δχ(iω)\displaystyle\delta\chi(i\omega) =vF2χ32(iω)χ22(iω)χ32(iω),\displaystyle=-v_{F}^{2}\chi_{32}(i\omega)\chi_{22}(i\omega)\chi_{32}(-i\omega), (35)
χ32(iω)\displaystyle\chi_{32}(i\omega) =ϵtr(Gτ3G+τ2(ϵFϵ)2η)\displaystyle=\sum_{\epsilon}{\rm tr}\left(G_{-}\tau_{3}G_{+}\tau_{2}\left(\frac{\epsilon_{F}}{\epsilon}\right)^{2\eta}\right) (36)
Refer to caption
Figure 6: a) The collective mode’s contribution to current-current susceptibility b) the collective mode propagator. See Eq.(33)

This expression of δχ(iω)\delta\chi(i\omega) is overall similar to the usual response function in CDW[20], except that both current vertices in it are renormalized by the interaction at critical point. Therefore, this expression contains extra the power-law renormalization factors. Here, the quantity χ22(iω)\chi_{22}(i\omega) is nothing but the sliding mode propagator, which is diagrammatically described in Fig.6 and is expressed as follows:

χ22(iω)=W1+WΠ2R(iω),\chi_{22}(i\omega)=\frac{W}{1+W\Pi_{2}^{R}(i\omega)}, (37)

where WW is the bare interaction defined in Eq.(23) and Eq.(24) for CDW and SDW respectively. The quantity Π2R(iω)\Pi_{2}^{R}(i\omega) is the intervalley phase susceptibility renormalized by the critical fluctuations, which is defined as follows:

Π2R(iω)=12k,ϵtr[Gτ2G+τ2(ϵFϵ)2η]\Pi_{2}^{R}(i\omega)=\frac{1}{2}\sum_{k,\epsilon}{\rm tr}\left[G_{-}\tau_{2}G_{+}\tau_{2}\left(\frac{\epsilon_{F}}{\epsilon}\right)^{2\eta}\right] (38)

As a reminder G±G_{\pm} is defined as G±=G(iϵ±iω2)G_{\pm}=G(i\epsilon\pm i\frac{\omega}{2}).

On general grounds, we expect that the sliding-mode propagator χ22(iω)\chi_{22}(i\omega) has a pole at ω=0\omega=0 as the CDW sliding mode is a Goldstone mode. In the ladder response framework used to analyze χ22(iω)\chi_{22}(i\omega) the condition for having a pole at ω=0\omega=0 is that Π2R\Pi_{2}^{R} satisfies

1+WΠ2R(0)=0.1+W\Pi_{2}^{R}(0)=0. (39)

To confirm that this relation holds, we calculate Π2R(0)\Pi_{2}^{R}(0) explicitly:

Π2R(0)\displaystyle\Pi_{2}^{R}(0) =12ϵ,ktr(Gτ2Gτ2(ϵFϵ)2η)\displaystyle={\color[rgb]{0,0,0}\frac{1}{2}}\sum_{\epsilon,k}{\rm tr}(G\tau_{2}G\tau_{2}\left(\frac{\epsilon_{F}}{\epsilon}\right)^{2\eta})
=ϵ,k(iϵk22m)2vF2k2Δ2[(iϵk22m)2vF2k2Δ2]2(ϵFϵ)2η\displaystyle=\sum_{\epsilon,k}\frac{\left(i\epsilon-\frac{k_{\perp}^{2}}{2m_{*}}\right)^{2}-v_{F}^{2}k_{\parallel}^{2}-\Delta^{2}}{\left[\left(i\epsilon-\frac{k_{\perp}^{2}}{2m_{*}}\right)^{2}-v_{F}^{2}k_{\parallel}^{2}-\Delta^{2}\right]^{2}}\left(\frac{\epsilon_{F}}{\epsilon}\right)^{2\eta}
=ϵ,k(ϵFϵ)2η(iϵk22m)2vF2k2Δ2\displaystyle=\sum_{\epsilon,k}\frac{\left(\frac{\epsilon_{F}}{\epsilon}\right)^{2\eta}}{\left(i\epsilon-\frac{k_{\perp}^{2}}{2m_{*}}\right)^{2}-v_{F}^{2}k_{\parallel}^{2}-\Delta^{2}} (40)

Importantly, the quantity Δ\Delta is itself determined by the CDW/SDW self-consistency equation, which takes the form:

ϵ,kW(ϵFϵ)2η(iϵk22m)2vF2k2Δ2=1\sum_{\epsilon,k}-\frac{W\left(\frac{\epsilon_{F}}{\epsilon}\right)^{2\eta}}{\left(i\epsilon-\frac{k_{\perp}^{2}}{2m_{*}}\right)^{2}-v_{F}^{2}k_{\parallel}^{2}-\Delta^{2}}=1 (41)

Comparing Eq.(40) and Eq.(41) we confirm the Goldstone mode condition given in Eq.(39).

To summarize this part of the analysis, Eq.(39) is not a coincidence because it follows from a physical argument: The sliding mode, whose propagator is χ22(iω)\chi_{22}(i\omega), is the Goldstone mode arising from the system’s U(1)U(1) symmetry ΔΔeiϕ\Delta\rightarrow\Delta e^{i\phi} (see Hamiltonian Eq.(7)). Therefore, its frequency vanishes, indicating that the pole of χ22(iω)\chi_{22}(i\omega) at ω=0\omega=0. In other words, the denominator of χ22(iω)\chi_{22}(i\omega), which equals to 1+WΠ22(0)1+W\Pi_{22}(0), must vanish at ω=0\omega=0.

Given these observations, the expression for χ22(iω)\chi_{22}(i\omega) in Eq.(37) can be put in the form that makes the Goldstone mode pole more apparent. Using the relation Eq.(39), we can write χ22(iω)\chi_{22}(i\omega) as

χ22(iω)=1Π2R(iω)Π2R(0).\chi_{22}(i\omega)=\frac{1}{\Pi_{2}^{R}(i\omega)-\Pi_{2}^{R}(0)}. (42)

This expression does not include the interaction strength WW, which enables us to evaluate χ22(iω)\chi_{22}(i\omega) directly from ΠR(iω)\Pi^{R}(i\omega). Both the ω=0\omega=0 pole and the residue at this pole can be obtained by analyzing the behavior of ΠR(iω)\Pi^{R}(i\omega) at small ω\omega.

We calculate the quantities χ32(iω)\chi_{32}(i\omega) and χ22(iω)\chi_{22}(i\omega), which is presented in detail in next section. Plugging the results into Eq.(35), we find that the collective mode contribution to dynamical current-current susceptibility is

δχ(iω)=CvFmΔ12\delta\chi(i\omega)=-Cv_{F}\sqrt{m_{*}}\Delta^{\frac{1}{2}} (43)

where CO(1)C\sim O(1) is an order-one factor. This indicates an extra Drude weight, which gives an extra contribution to zero-frequency conductivity:

δσ(ω)=e2vFmΔ12iω\delta\sigma(\omega)=-\frac{e^{2}v_{F}\sqrt{m_{*}}\Delta^{\frac{1}{2}}}{i\omega} (44)

In next section, we will detail the derivation of Eq.(43).

VIII The collective mode contribution to the AC susceptibility

In last section, we have outlined the steps to obtain the collective mode’s contribution to AC susceptibility δχ(iω)\delta\chi(i\omega) Eq.(44). In this section, we detail the derivation of this result explicitly. From Eq.(35) we see that to calculate δχ(iω)\delta\chi(i\omega) we need to evaluate χ32(iω)\chi_{32}(i\omega) and χ22(iω)\chi_{22}(i\omega). Below, we evaluate these two quantities.

First, the quantity χ32(iω)\chi_{32}(i\omega), which is represented by a bubble at each end of the diagram in Fig.6 a), is given by

χ32(iω)\displaystyle\chi_{32}(i\omega) =dϵ2πk12tr(G(iϵ)τ3G(iϵ+)τ2)(ϵFϵ)η\displaystyle=\int\frac{d\epsilon}{2\pi}\sum_{k}\frac{1}{2}{\rm tr}(G(i\epsilon_{-})\tau_{3}G(i\epsilon_{+})\tau_{2})\left(\frac{\epsilon_{F}}{\epsilon}\right)^{\eta}
=ϵFϵFdϵ2πkiz(Δ)iz+(Δ)(z+2E2)(z2E2)(ϵFϵ)η\displaystyle=\int_{-\epsilon_{F}}^{\epsilon_{F}}\frac{d\epsilon}{2\pi}\sum_{k}\frac{iz_{-}(-\Delta)-iz_{+}(-\Delta)}{\left(z_{+}^{2}-E^{2}\right)\left(z_{-}^{2}-E^{2}\right)}\left(\frac{\epsilon_{F}}{\epsilon}\right)^{\eta}
=ϵFϵFdϵ2πkωΔ(ϵFϵ)η(z+2E2)(z2E2)\displaystyle=\int_{-\epsilon_{F}}^{\epsilon_{F}}\frac{d\epsilon}{2\pi}\sum_{k}\frac{-\omega\Delta\left(\frac{\epsilon_{F}}{\epsilon}\right)^{\eta}}{\left(z_{+}^{2}-E^{2}\right)\left(z_{-}^{2}-E^{2}\right)}

In the first line, ϵ±=ϵ±ω2\epsilon_{\pm}=\epsilon\pm\frac{\omega}{2}, in the second line we defined z±=iϵ±k22mz_{\pm}=i\epsilon_{\pm}-\frac{k_{\perp}^{2}}{2m_{*}}. As a reminder here EE is the quasiparticle energy defined through E=vF2k2+Δ2E=\sqrt{v_{F}^{2}k_{\parallel}^{2}+\Delta^{2}}. To proceed, we keep only the terms leading-order in ω\omega, which gives:

χ32(iω)=ωΔϵFϵFdϵdkdk(2π)3(ϵFϵ)η(vFkz)2(vFk+z)2,\displaystyle\chi_{32}(i\omega)=-\omega\Delta\int\limits_{-\epsilon_{F}}^{\epsilon_{F}}\frac{d\epsilon dk_{\perp}dk_{\parallel}}{(2\pi)^{3}}\frac{\left(\frac{\epsilon_{F}}{\epsilon}\right)^{\eta}}{\left(v_{F}k_{\parallel}-z\right)^{2}\left(v_{F}k_{\parallel}+z\right)^{2}}, (45)

where we suppressed the higher order terms O(ω3)O(\omega^{3}). As a reminder zz is defined in the same way as in Eq.(30), namely, z=iϵk22mz=i\epsilon-\frac{k_{\perp}^{2}}{2m_{*}}.

Next, we carry out integral over kk_{\parallel} and kk_{\perp} step by step as follows:

χ32(iω)=ωΔvFϵFϵFdϵ2πdk2π(2)i(2z)3(ϵFϵ)η\displaystyle\chi_{32}(i\omega)=-\frac{\omega\Delta}{v_{F}}\int_{-\epsilon_{F}}^{\epsilon_{F}}\frac{d\epsilon}{2\pi}\frac{dk_{\perp}}{2\pi}\frac{(-2)i}{\left(2z\right)^{3}}\left(\frac{\epsilon_{F}}{\epsilon}\right)^{\eta} (46)
=iωΔ4vFϵFϵFdϵ2π0dϵ2π2mϵϵ33ϵ2ϵ(ϵ2+ϵ2)3(ϵFϵ)η.\displaystyle=\frac{i\omega\Delta}{4v_{F}}\int_{-\epsilon_{F}}^{\epsilon_{F}}\frac{d\epsilon}{2\pi}\int_{0}^{\infty}\frac{d\epsilon_{\perp}}{2\pi}\frac{\sqrt{2m_{*}}}{\sqrt{\epsilon_{\perp}}}\frac{\epsilon_{\perp}^{3}-3\epsilon^{2}\epsilon_{\perp}}{-\left(\epsilon^{2}_{\perp}+\epsilon^{2}\right)^{3}}\left(\frac{\epsilon_{F}}{\epsilon}\right)^{\eta}.

Here, we defined ϵ=k22m\epsilon_{\perp}=\frac{k_{\perp}^{2}}{2m_{*}}. To carry out integration, we nondimensionalize ϵ=xϵ\epsilon_{\perp}=x\epsilon and integrate over xx as follows:

iωΔ2m4vFϵFϵFdϵ2π0dx2πx523x12(x2+1)3|ϵ|52|ϵ||ϵ|6(ϵFϵ)η\displaystyle-\frac{i\omega\Delta\sqrt{2m_{*}}}{4v_{F}}\int_{-\epsilon_{F}}^{\epsilon_{F}}\frac{d\epsilon}{2\pi}\int_{0}^{\infty}\frac{dx}{2\pi}\frac{x^{\frac{5}{2}}-3x^{\frac{1}{2}}}{\left(x^{2}+1\right)^{3}}\frac{|\epsilon|^{\frac{5}{2}}|\epsilon|}{|\epsilon|^{6}}\left(\frac{\epsilon_{F}}{\epsilon}\right)^{\eta}
=3162iωΔ2m4vFϵFϵFdϵ2π|ϵ|52(ϵFϵ)η.\displaystyle=\frac{3}{16\sqrt{2}}\frac{i\omega\Delta\sqrt{2m_{*}}}{4v_{F}}\int_{-\epsilon_{F}}^{\epsilon_{F}}\frac{d\epsilon}{2\pi}|\epsilon|^{-\frac{5}{2}}\left(\frac{\epsilon_{F}}{\epsilon}\right)^{\eta}. (47)

To arrive at the expression in the last line [Eq.(47)] we have used the identity

0dx2πx523x12(x2+1)3=364215642=3162.\int_{0}^{\infty}\frac{dx}{2\pi}\frac{x^{\frac{5}{2}}-3x^{\frac{1}{2}}}{\left(x^{2}+1\right)^{3}}=\frac{3}{64\sqrt{2}}-\frac{15}{64\sqrt{2}}=-\frac{3}{16\sqrt{2}}. (48)

Note that the factor of (ϵFϵ)η\left(\frac{\epsilon_{F}}{\epsilon}\right)^{\eta} needs to be cut off at an energy scale of ϵΔ\epsilon\sim\Delta. Therefore, we carry out the integral over ϵ\epsilon exactly while approximately cutting off the integral in Eq.(47) at a lower-bound of ϵΔ\epsilon\sim\Delta:

χ32(iω)\displaystyle\chi_{32}(i\omega) =iω3Δm64πvF2ΔϵF(ϵF|ϵ|)η|ϵ|52𝑑ϵ\displaystyle=i\omega\frac{3\Delta\sqrt{m_{*}}}{64\pi v_{F}}2\int_{\Delta}^{\epsilon_{F}}\left(\frac{\epsilon_{F}}{|\epsilon|}\right)^{\eta}|\epsilon|^{-\frac{5}{2}}d\epsilon
=iω3Δm32vF1η+3/2(ϵFΔ)ηΔ32.\displaystyle=i\omega\frac{3\Delta\sqrt{m_{*}}}{32v_{F}}\frac{1}{\eta+3/2}\left(\frac{\epsilon_{F}}{\Delta}\right)^{\eta}\Delta^{-\frac{3}{2}}. (49)

Next, we evaluate the quantity χ22(iω)\chi_{22}(i\omega), which is the sliding mode propagator diagrammatically represented by the double line in the middle of Fig.6 c). As argued above [see Eq.(42) and accompanying discussion], this quantity is the inverse of the difference

Π(iω)2RΠ2R(ω0).\Pi{{}_{2}^{R}}(i\omega)-\Pi_{2}^{R}(\omega\to 0). (50)

We therefore proceed to evaluate the quantity Π22(iω)\Pi_{22}(i\omega), which is given by

Π(iω)2R=12d2k(2π)2ϵFϵFdϵ2πtr(G(iϵ+)τ2G(iϵ)τ2))(ϵFϵ)2η\displaystyle\Pi{{}_{2}^{R}}(i\omega)={\color[rgb]{0,0,0}\frac{1}{2}}\int\frac{d^{2}k}{(2\pi)^{2}}\int\limits^{\epsilon_{F}}_{-\epsilon_{F}}\frac{d\epsilon}{2\pi}{\rm tr}(G(i\epsilon_{+})\tau_{2}G(i\epsilon_{-})\tau_{2}))\left(\frac{\epsilon_{F}}{\epsilon}\right)^{2\eta}
=12d2k(2π)2ϵFϵFdϵ2π[1iϵ+k22mE1iϵk22m+E\displaystyle={\color[rgb]{0,0,0}\frac{1}{2}}\int\frac{d^{2}k}{(2\pi)^{2}}\int\limits^{\epsilon_{F}}_{-\epsilon_{F}}\frac{d\epsilon}{2\pi}\left[\frac{1}{i\epsilon_{+}-\frac{k_{\perp}^{2}}{2m_{*}}-E}\frac{1}{i\epsilon_{-}-\frac{k_{\perp}^{2}}{2m_{*}}+E}\right. (51)
+(EE)](ϵFϵ)2η.\displaystyle\left.+\left(E\rightarrow-E\right)\right]\left(\frac{\epsilon_{F}}{\epsilon}\right)^{2\eta}.

Here ϵ±\epsilon_{\pm} is defined as ϵ±=ϵ±ω2\epsilon_{\pm}=\epsilon\pm\frac{\omega}{2}. Next, we carry out the integral over ϵ\epsilon. This cannot be done analytically, but we observe that the contribution to the integral over ϵ\epsilon predominantly comes from energy ϵΔ\epsilon{\color[rgb]{0,0,0}\sim\Delta}. This is because of the following two reasons, a) the power-law factor (ϵF/ϵ)η(\epsilon_{F}/\epsilon)^{\eta} becomes largest at small energy, b) however, this power-law behavior is valid only for ϵΔ\epsilon\gtrsim\Delta and therefore should be cutoff at an energy ofϵΔ\epsilon\sim\Delta. Consequently, as an approximation, we can replace (ϵFϵ)η\left(\frac{\epsilon_{F}}{\epsilon}\right)^{\eta} with (ϵFΔ)η\left(\frac{\epsilon_{F}}{\Delta}\right)^{\eta}. With this, we can proceed analytically as follows:

Π(iω)2R=12(ϵFΔ)2ηd2k(2π)21iω+2E[f(k22m+E)\displaystyle\Pi{{}_{2}^{R}}(i\omega)={\color[rgb]{0,0,0}\frac{1}{2}}\left(\frac{\epsilon_{F}}{\Delta}\right)^{2\eta}\int\frac{d^{2}k}{(2\pi)^{2}}\frac{1}{-i\omega+2E}\left[f\left(\frac{k_{\perp}^{2}}{2m_{*}}+E\right)\right.
f(k22mE)]+(EE)\displaystyle\left.-f\left(\frac{k_{\perp}^{2}}{2m_{*}}-E\right)\right]+\left(E\rightarrow-E\right)
=(ϵFΔ)2ηd2k(2π)2[1iω+2E1iω2E]\displaystyle=\left(\frac{\epsilon_{F}}{\Delta}\right)^{2\eta}\int\frac{d^{2}k}{(2\pi)^{2}}\left[\frac{1}{-i\omega+2E}-\frac{1}{-i\omega-2E}\right]
×[f(k22m+E)f(k22mE)]\displaystyle\times\left[f\left(\frac{k_{\perp}^{2}}{2m_{*}}+E\right)-f\left(\frac{k_{\perp}^{2}}{2m_{*}}-E\right)\right] (52)

As EE is independent of kk_{\perp} (see Eq.(31)), it is convenient to first integrate over kk_{\perp}. This integration can be carried out exactly because the last factor simply takes value of 1-1 for |k2/2m|<E|k_{\perp}^{2}/2m|<E, and vanishes for |k2/2m|>E|k_{\perp}^{2}/2m|>E. Consequently, the integral over kk_{\perp} yields a factor of 22mE2\sqrt{2m_{*}E}, which physically represents the length of the segment on the Fermi surface that is gapped out by CDW

Π(iω)2R=(ϵFΔ)2ηdk2π2Eω24E222mE2π\displaystyle\Pi{{}_{2}^{R}}(i\omega)=-\left(\frac{\epsilon_{F}}{\Delta}\right)^{2\eta}\int\frac{dk_{\parallel}}{2\pi}\frac{-{\color[rgb]{0,0,0}2}E}{-\frac{\omega^{2}}{4}-E^{2}}\frac{2\sqrt{2m_{*}E}}{2\pi} (53)

Next, to integrate over kk_{\parallel}, we define ϵ=vFk\epsilon_{\parallel}=v_{F}k_{\parallel}, and rewrite the expression as an integral over ϵ\epsilon_{\parallel}

Π(iω)2R=2mπ2vF(ϵFΔ)2ηdϵ(ϵ2+Δ2)34ω24+ϵ2+Δ2\displaystyle\Pi{{}_{2}^{R}}(i\omega)=-\frac{\sqrt{2m_{*}}}{\pi^{2}v_{F}}\left(\frac{\epsilon_{F}}{\Delta}\right)^{2\eta}\int_{{\color[rgb]{0,0,0}-\infty}}^{{\color[rgb]{0,0,0}\infty}}\frac{d\epsilon_{\parallel}\left(\epsilon_{\parallel}^{2}+\Delta^{2}\right)^{\frac{3}{4}}}{\frac{\omega^{2}}{4}+\epsilon_{\parallel}^{2}+\Delta^{2}} (54)

Taking the difference Π(iω)2RΠ(0)2R\Pi{{}_{2}^{R}}(i\omega)-\Pi{{}_{2}^{R}}(0) and keeping only the leading-order terms in ω\omega, we find

Π(iω)2RΠ(0)2R\displaystyle\Pi{{}_{2}^{R}}(i\omega)-\Pi{{}_{2}^{R}}(0) =ω242mπ2vF(ϵFΔ)2ηdϵ(ϵ2+Δ2)54\displaystyle=\frac{\omega^{2}}{4}\frac{\sqrt{2m_{*}}}{\pi^{2}v_{F}}\left(\frac{\epsilon_{F}}{\Delta}\right)^{2\eta}\int\limits_{{\color[rgb]{0,0,0}-\infty}}^{{\color[rgb]{0,0,0}\infty}}\frac{d\epsilon_{\parallel}}{\left(\epsilon_{\parallel}^{2}+\Delta^{2}\right)^{\frac{5}{4}}}
ζω222mπ2vF(ϵFΔ)2ηΔ32\displaystyle\sim\zeta\frac{\omega^{2}}{2}\frac{\sqrt{2m_{*}}}{\pi^{2}v_{F}}\left(\frac{\epsilon_{F}}{\Delta}\right)^{2\eta}\Delta^{-\frac{3}{2}} (55)

where the dimensionless constant ζ\zeta is defined as ζ=0𝑑x(x2+1)5/4\zeta=\int^{\infty}_{0}dx(x^{2}+1)^{-5/4}, which arises from the integral over ϵ\epsilon_{\parallel}. As ζ\zeta is an order-1 prefactor, we will simply take it to be 11 in the followings steps. Plugging this into Eq.(42) yields

χ22=2π2vF(ΔϵF)2ηΔ322m1ω2\chi_{22}=\frac{{\color[rgb]{0,0,0}2}\pi^{2}v_{F}\left(\frac{\Delta}{\epsilon_{F}}\right)^{2\eta}\Delta^{\frac{3}{2}}}{\sqrt{2m_{*}}}\frac{1}{\omega^{2}} (56)

Finally, using Eq.(35), we find that the collective mode contribution to dynamical susceptibility is given by

δχ(iω)\displaystyle\delta\chi(i\omega) =vF2χ32(iω)χ22(iω)χ32(iω)\displaystyle=-v_{F}^{2}\chi_{32}(i\omega)\chi_{22}(i\omega)\chi_{32}(-i\omega) (57)
=2π2vFΔ322m(3Δm32Δ32)2\displaystyle=-\frac{{\color[rgb]{0,0,0}2}\pi^{2}v_{F}\Delta^{\frac{3}{2}}}{\sqrt{2m_{*}}}\left(\frac{3\Delta\sqrt{m_{*}}}{32}\Delta^{-\frac{3}{2}}\right)^{2}
=9π2vFmΔ125122\displaystyle=-\frac{9\pi^{2}v_{F}\sqrt{m_{*}}\Delta^{\frac{1}{2}}}{512\sqrt{2}}

As shown in Eq.(44), this quantity represents the contribution to the spectral weight of the AC conductivity arising from CDW sliding mode. As discussed above, this spectral weight is independent of total carrier density since the CDW gap is weak, ΔϵF\Delta\ll\epsilon_{F}. Instead, the spectral weight is determined by the length of the segment on the Fermi surface that is gapped out by CDW order, which makes it proportional to Δ12\Delta^{\frac{1}{2}}.

IX Summary

Motivated by recent experimental observations [14, 36], we considered a density-wave phase (charge or spin) near the isospin phase transition in BBG driven by the critical valley fluctuations. Strong interactions between critical fluctuations and band electrons define a unique strong-coupling CDW phase in which the interactions are characterized by momentum transfer much smaller than kFk_{F}. We use a renormalization group approach to show that the CDW and SDW susceptibilities have a power-law divergence, indicating that TcT_{c} is enhanced compared to CDW/SDW emerging away from criticality. The predicted phase diagram is consistent with the transport measurement[14], where a peculiar resistive state is observed directly at an onset of isospin order.

We use this strong-coupling framework to analyze nonlionear transport. In particular, we explain the anomalous non-monotonic differential conductivity behavior reported in Ref.14 and argue that it can be understood in the framework of the conventional CDW/SDW phase mode sliding mechanism. We estimate the sliding conductivity in the strong-coupling framework and arrive at a result which is in reasonable agreement with measurements. We also consider an alternative explanation of the nonlinear transport based on the interband Landau-Zener tunneling across the minigaps induced by the CDW/SDW perturbation, finding that this scenario can also reproduce the observed nonlinear behavior. Together, these results provide a strong indication that the resistive state seen in the experiment is indeed related to CDW/SDW order mediated by quantum-critical interactions.

References

  • Andrei and MacDonald [2020] E. Y. Andrei and A. H. MacDonald, Graphene bilayers with a twist, Nature materials 19, 1265 (2020).
  • Bistritzer and MacDonald [2011] R. Bistritzer and A. H. MacDonald, Moiré bands in twisted double-layer graphene, Proceedings of the National Academy of Sciences 108, 12233 (2011)https://www.pnas.org/content/108/30/12233.full.pdf .
  • Cao et al. [2018a] Y. Cao, V. Fatemi, A. Demir, S. Fang, S. L. Tomarken, J. Y. Luo, J. D. Sanchez-Yamagishi, K. Watanabe, T. Taniguchi, E. Kaxiras, and et al., Correlated insulator behaviour at half-filling in magic-angle graphene superlattices, Nature 556, 80–84 (2018a).
  • Cao et al. [2021] Y. Cao, D. Rodan-Legrain, J. M. Park, N. F. Q. Yuan, K. Watanabe, T. Taniguchi, R. M. Fernandes, L. Fu, and P. Jarillo-Herrero, Nematicity and competing orders in superconducting magic-angle graphene, Science 372, 264 (2021).
  • Jaoui et al. [2022] A. Jaoui, I. Das, G. Di Battista, J. Díez-Mérida, X. Lu, K. Watanabe, T. Taniguchi, H. Ishizuka, L. Levitov, and D. K. Efetov, Quantum critical behaviour in magic-angle twisted bilayer graphene, Nature Physics , 1 (2022).
  • Cao et al. [2018b] Y. Cao, V. Fatemi, S. Fang, K. Watanabe, T. Taniguchi, E. Kaxiras, and P. Jarillo-Herrero, Unconventional superconductivity in magic-angle graphene superlattices, Nature 556, 43–50 (2018b).
  • Lu et al. [2019] X. Lu, P. Stepanov, W. Yang, M. Xie, M. A. Aamir, I. Das, C. Urgell, K. Watanabe, T. Taniguchi, G. Zhang, et al., Superconductors, orbital magnets and correlated states in magic-angle bilayer graphene, Nature 574, 653 (2019).
  • Stepanov et al. [2020] P. Stepanov, I. Das, X. Lu, A. Fahimniya, K. Watanabe, T. Taniguchi, F. H. Koppens, J. Lischner, L. Levitov, and D. K. Efetov, Untying the insulating and superconducting orders in magic-angle graphene, Nature 583, 375 (2020).
  • Saito et al. [2020] Y. Saito, J. Ge, K. Watanabe, T. Taniguchi, and A. F. Young, Independent superconductors and correlated insulators in twisted bilayer graphene, Nature Physics 16, 926 (2020).
  • Oh et al. [2021] M. Oh, K. P. Nuckolls, D. Wong, R. L. Lee, X. Liu, K. Watanabe, T. Taniguchi, and A. Yazdani, Evidence for unconventional superconductivity in twisted bilayer graphene, Nature 600, 240 (2021).
  • Liu et al. [2021] X. Liu, Z. Wang, K. Watanabe, T. Taniguchi, O. Vafek, and J. Li, Tuning electron correlation in magic-angle twisted bilayer graphene using coulomb screening, Science 371, 1261 (2021).
  • Zhou et al. [2021a] H. Zhou, T. Xie, A. Ghazaryan, T. Holder, J. R. Ehrets, E. M. Spanton, T. Taniguchi, K. Watanabe, E. Berg, M. Serbyn, et al., Half-and quarter-metals in rhombohedral trilayer graphene, Nature 598, 429 (2021a).
  • Zhou et al. [2021b] H. Zhou, T. Xie, T. Taniguchi, K. Watanabe, and A. F. Young, Superconductivity in rhombohedral trilayer graphene, Nature 598, 434 (2021b).
  • Zhou et al. [2022] H. Zhou, L. Holleis, Y. Saito, L. Cohen, W. Huynh, C. L. Patterson, F. Yang, T. Taniguchi, K. Watanabe, and A. F. Young, Isospin magnetism and spin-polarized superconductivity in bernal bilayer graphene, Science 375, 774 (2022)https://www.science.org/doi/pdf/10.1126/science.abm8386 .
  • de la Barrera et al. [2021] S. C. de la Barrera, S. Aronson, Z. Zheng, K. Watanabe, T. Taniguchi, Q. Ma, P. Jarillo-Herrero, and R. Ashoori, Cascade of isospin phase transitions in bernal bilayer graphene at zero magnetic field, arXiv preprint arXiv:2110.13907  (2021).
  • Seiler et al. [2021] A. M. Seiler, F. R. Geisenhof, F. Winterer, K. Watanabe, T. Taniguchi, T. Xu, F. Zhang, and R. T. Weitz, Quantum cascade of new correlated phases in trigonally warped bilayer graphene, arXiv preprint arXiv:2111.06413  (2021).
  • McCann and Koshino [2013a] E. McCann and M. Koshino, The electronic properties of bilayer graphene, Reports on Progress in physics 76, 056503 (2013a).
  • Fleming and Grimes [1979] R. M. Fleming and C. C. Grimes, Sliding-mode conductivity in nbse3{\mathrm{se}}_{3}: Observation of a threshold electric field and conduction noise, Phys. Rev. Lett. 42, 1423 (1979).
  • Peierls and Peierls [1955] R. Peierls and R. E. Peierls, Quantum theory of solids (Oxford University Press, 1955).
  • Lee et al. [1974] P. Lee, T. Rice, and P. Anderson, Conductivity from charge or spin density waves, Solid State Communications 14, 703 (1974).
  • Călugăru et al. [2022] D. Călugăru, N. Regnault, M. Oh, K. P. Nuckolls, D. Wong, R. L. Lee, A. Yazdani, O. Vafek, and B. A. Bernevig, Spectroscopy of twisted bilayer graphene correlated insulators, Phys. Rev. Lett. 129, 117602 (2022).
  • McCann and Koshino [2013b] E. McCann and M. Koshino, The electronic properties of bilayer graphene, Reports on Progress in Physics 76, 056503 (2013b).
  • Dong et al. [2022] Z. Dong, A. V. Chubukov, and L. Levitov, Spin-triplet superconductivity at the onset of isospin order in biased bilayer graphene, arXiv preprint arXiv:2205.13353  (2022).
  • [24] Supplemental material [url will be inserted by publisher].
  • Altshuler et al. [1994] B. L. Altshuler, L. B. Ioffe, and A. J. Millis, Low-energy properties of fermions with singular interactions, Phys. Rev. B 50, 14048 (1994).
  • Chou et al. [2022] Y.-Z. Chou, F. Wu, J. D. Sau, and S. D. Sarma, Acoustic-phonon-mediated superconductivity in moir\\backslash’eless graphene multilayers, arXiv preprint arXiv:2204.09811  (2022).
  • Landau [1932] L. D. Landau, Zur theorie der energieubertragung ii, Z. Sowjetunion 2, 46 (1932).
  • Zener and Fowler [1932] C. Zener and R. H. Fowler, Non-adiabatic crossing of energy levels, Proceedings of the Royal Society of London. Series A, Containing Papers of a Mathematical and Physical Character 137, 696 (1932).
  • Grüner et al. [1981] G. Grüner, A. Zawadowski, and P. Chaikin, Nonlinear conductivity and noise due to charge-density-wave depinning in nb se 3, Physical Review Letters 46, 511 (1981).
  • Grü ner et al. [1981] G. Grü ner, A. Zettl, W. Clark, and A. Thompson, Observation of narrow-band charge-density-wave noise in tas3tas_{3}, Physical Review B 23, 6813 (1981).
  • Bardeen et al. [1982] J. Bardeen, E. Ben-Jacob, A. Zettl, and G. Grüner, Current oscillations and stability of charge-density-wave motion in nbse3{\mathrm{nbse}}_{3}Phys. Rev. Lett. 49, 493 (1982).
  • Nomura et al. [1989] K. Nomura, T. Shimizu, K. Ichimura, T. Sambongi, M. Tokumoto, H. Anzai, and N. Kinoshita, Narrow band noise in sdw sliding, Solid State Communications 72, 1123 (1989).
  • Chatterjee et al. [2022] S. Chatterjee, T. Wang, E. Berg, and M. P. Zaletel, Inter-valley coherent order and isospin fluctuation mediated superconductivity in rhombohedral trilayer graphene, Nature communications 13, 6013 (2022).
  • Dong et al. [2023] Z. Dong, P. A. Lee, and L. S. Levitov, Signatures of cooper pair dynamics and quantum-critical superconductivity in tunable carrier bands, arXiv preprint arXiv:2304.09812  (2023).
  • Dong and Levitov [2021] Z. Dong and L. Levitov, Superconductivity in the vicinity of an isospin-polarized state in a cubic dirac band, arXiv preprint arXiv:2109.01133  (2021).
  • Seiler et al. [2023] A. M. Seiler, M. Statz, I. Weimer, N. Jacobsen, K. Watanabe, T. Taniguchi, Z. Dong, L. S. Levitov, and R. T. Weitz, Interaction-driven (quasi-) insulating ground states of gapped electron-doped bilayer graphene, arXiv preprint arXiv:2308.00827  (2023).
  • Coleman [2015] P. Coleman, Introduction to Many-Body Physics (Cambridge University Press, 2015).

Appendix A The intervalley interaction mediated by valley-polarization soft mode

In this section, we describe the origin of singular intervalley interaction used above in Eq.(14) to analyze CDW and SDW instabilities. This analysis follows closely the analysis carried out in Ref.[23].

The intervalley interaction is dominated by the intravalley scattering processes, where an electron in valley KK is scattered to valley KK, and the other electron in valley KK^{\prime} is scattered to valley KK^{\prime}. Here we can safely neglect the valley-exchange interaction that transfers a large momentum KK since the Coulomb interaction is proportional to the inverse of momentum transfer. At the isospin-polarizing phase transition, such intervalley scattering is renormalized by two effects: the vertex corrections and screening. The renormalized interaction is diagrammatically expressed in Fig.7. Here, the wavy lines represent bare Coulomb interaction UU.

Refer to caption
Figure 7: [Figure adapted from Ref.[23]] Diagrams describing the effective interaction between electrons in valleys KK and KK^{\prime} mediated by quantum-critical modes, Eq.(62). Here, the wavy lines represent bare Coulomb interaction UU. These processes give an enhancement to forward scattering divergent near the valley-polarization instability (see discussion in Ref.[23]). (b)The diagrammatic representation of the irreducible part of the charge susceptibility (the shaded ellipse), summed over s=,s=\uparrow,\downarrow.

Below we calculate the renormalized intervalley interaction. First, the effect of each vertex correction can be described using a factor γ\gamma:

γ(iν,q)=11+UΠ0(iν,q)\gamma(i\nu,q)=\frac{1}{1+U\Pi_{0}(i\nu,q)} (58)

Here, Π0\Pi_{0} is the bare polarization function of one spin species, Π0(iν,q)=ϵ,kG(iϵ+iν,k+q)G(iϵ,k)\Pi_{0}(i\nu,q)=\sum_{\epsilon,k}G(i\epsilon+i\nu,k+q)G(i\epsilon,k). In our setting, spin up and spin down are degenerate. Standard calculation[37] gives

Π0(iν,q)=(ν0(1l02q2)|ν|qp1+p24πvF2)\Pi_{0}(i\nu,q)=-\left(\nu_{0}(1-l_{0}^{2}q^{2})-\frac{|\nu|}{q}\frac{p_{1}+p_{2}}{4\pi v_{F}^{2}}\right) (59)

where ν0\nu_{0} is the density of state at the Fermi level, p1p_{1} and p2p_{2} are the radii of curvature on the Fermi surface where qq is perpendicular to the fermi velocity. The quantity l0l_{0} is a length scale depending on band dispersion.

Importantly, the momentum qq relevant for analysis is restricted to a certain direction. Namely, the momentum qq, which is the momentum transfer of the interaction used to generate CDW, is parallel to the tangential of the Fermi surface at QQ(see main text), where the curvature of radius is maximized. Therefore, for such a qq, one of p1p_{1} and p2p_{2} is pp_{*}, which is the radius of curvature at QQ point, whereas the other is pp_{*}^{\prime} which is the radius of curvature at the point opposite to QQ point on the Fermi surface. This definition of pp_{*} and pp_{*}^{\prime} is the same as the definition in the main text. Therefore, the polarization function at the relevant qq can be written as

Π0(iν,q)=(ν0(1q2l02)|ν|qp+p4πvF2)\Pi_{0}(i\nu,q)=-\left(\nu_{0}(1-q^{2}l_{0}^{2})-\frac{|\nu|}{q}\frac{p_{*}+p_{*}^{\prime}}{4\pi v_{F}^{2}}\right) (60)

Using the Stoner condition ν0U=1\nu_{0}U=1, and plugging in the density of state in our model ν0=m/2π\nu_{0}=m/2\pi, we find

γ(iν,q)=1κ+κ2|ν|vFq+q2l02\gamma(i\nu,q)=\frac{1}{\frac{\kappa_{*}+\kappa^{\prime}}{2}\frac{|\nu|}{v_{F}q}+q^{2}l_{0}^{2}} (61)

Further accounting for the screening effect, we arrive at the following form of effective coupling

V(iν,q)=Uγ(iν,q)21NUγ(iν,q)Π0(iν,q)Uγ(iν,q)/N=U/Nκ+κ2|ν|vFq+q2l02V(i\nu,q)=\frac{U\gamma(i\nu,q)^{2}}{1-NU\gamma(i\nu,q)\Pi_{0}(i\nu,q)}\sim U\gamma(i\nu,q)/N=\frac{U/N}{\frac{\kappa_{*}+\kappa^{\prime}}{2}\frac{|\nu|}{v_{F}q}+q^{2}l_{0}^{2}} (62)

where N=4N=4 represents the number of isospin species in the unpolarized phase. Eq.(62) is the form we used in the main text Eq.(14). Here we have used the stoner condition UΠ0(0,0)=1-U\Pi_{0}(0,0)=1 and that each polarization bubble is renormalized by one vertex correction γ(iν,q)\gamma(i\nu,q), which is divergent at small momentum and low frequency.

Appendix B Renormalization of the CDW/SDW vertex correction

In this section, we derive the first-order correction for CDW/SDW vertex function, which is a result used in main text Eq.(15). This derivation is similar to the analysis of the gauge-fluctuation problem carried out in Ref.[25]. We start from the following vertex correction which is read off from the diagram in main text Fig.4 a):

δΓ(ϵ)=ν,𝒌V(iνiϵ,k)Γ0(ν)(iνk22m)2vF2k2\delta\Gamma(\epsilon)=-\sum_{\nu,{\boldsymbol{k}}}\frac{V(i\nu-i\epsilon,k)\Gamma_{0}(\nu)}{\left(i\nu-\frac{k_{\perp}^{2}}{2m_{*}}\right)^{2}-v_{F}^{2}k_{\parallel}^{2}} (63)

To proceed analytically, the following approximation can be employed. First, we focus on the frequency regime νϵ\nu\gg\epsilon which , as we will see, is the origin of log-divergence. In this regime, V(iνiϵ)V(i\nu-i\epsilon) can be substituted with V(iν)V(i\nu). Second, the contribution to the right-hand side decays much faster along kk_{\parallel} direction than along kk_{\perp} direction. As a result, processes with kkk_{\perp}\gg k_{\parallel} dominates the right-hand side. Therefore, we can replace |k||k| in the expression of V(iν,k)V(i\nu,k) with kk_{\perp}. After that, Eq.(63) becomes

δΓ(ϵ)=4UvFNκϵdν2πk|k|(iνk22m)2vF2k21|ν|+α|k|3Γ0,\displaystyle\delta\Gamma(\epsilon)=-\frac{4Uv_{F}}{N\kappa}\int_{\epsilon}\frac{d\nu}{2\pi}\sum_{k_{\perp}}\frac{|k_{\perp}|}{\left(i\nu-\frac{k_{\perp}^{2}}{2m_{*}}\right)^{2}-v_{F}^{2}k_{\parallel}^{2}}\frac{1}{|\nu|+\alpha|k_{\perp}|^{3}}\Gamma_{0}, (64)

The competition of two denominators in the integral define a new energy scale. Namely, the integrand is suppressed by the first factor at a momentum scale of kk1=ϵ1/3α1/3k_{\perp}\gtrsim k_{1}=\epsilon^{1/3}\alpha^{-1/3} whereas the second factor starts to suppress the integrand at kk2=(2mϵ)1/2k_{\perp}\gtrsim k_{2}=(2m\epsilon)^{1/2}. The comparison of these two momentum scales k1k_{1} and k2k_{2} sets a characteristic frequency

ϵ=α2(2m)3.\epsilon_{*}=\alpha^{-2}(2m)^{-3}. (65)

which is set by the details in band dispersion. From now on we focus on the case where ϵ>ϵF\epsilon_{*}>\epsilon_{F} because it simplifies the discussion as we always have ϵ<ϵ\epsilon<\epsilon_{*} in this case. In this regime, k1>k2k_{1}>k_{2}, so the k3k_{\perp}^{3} term in the denominator can be safely neglected, yielding

δΓ(ϵ)=4UvFNκϵϵFdν2πS(ν)|ν|Γ0,S(ν)=k|k|(iνk22m)2vF2k2\delta\Gamma(\epsilon)=-\frac{4Uv_{F}}{N\kappa}\int^{\epsilon_{F}}_{\epsilon}\frac{d\nu}{2\pi}\frac{S(\nu)}{|\nu|}\Gamma_{0},\quad S(\nu)=\sum_{k_{\perp}}\frac{|k_{\perp}|}{\left(i\nu-\frac{k_{\perp}^{2}}{2m_{*}}\right)^{2}-v_{F}^{2}k_{\parallel}^{2}} (66)

We consider exclusively the case where Γ(ϵ)\Gamma(\epsilon) is even in ϵ\epsilon since the odd-frequency channel is weak due to an extra zero at ϵ=0\epsilon=0. Therefore, we define a quantity SS^{\prime} to describe the symmetrized part of SS:

S(ν)=S(ν)+S(ν)2S^{\prime}(\nu)=\frac{S(\nu)+S(-\nu)}{2} (67)

Below, we evaluate SS^{\prime} by working out the summation over kk step by step. First, we integrate over kk_{\parallel} to obtain

S(ν)=Re[isgn(ν)2vFk|k|iνk22m]S^{\prime}(\nu)={\rm Re}\left[-\frac{i{\rm sgn}(\nu)}{2v_{F}}\sum_{k_{\perp}}\frac{|k_{\perp}|}{i\nu-\frac{k_{\perp}^{2}}{2m_{*}}}\right] (68)

Next, we integrate over kk_{\perp}, finding

S(ν)\displaystyle S^{\prime}(\nu) =Re[isgn(ν)vF20kdk2π(iνk22m)]\displaystyle={\rm Re}\left[-\frac{i{\rm sgn}(\nu)}{v_{F}}2\int_{0}^{\infty}\frac{k_{\perp}dk_{\perp}}{2\pi\left(i\nu-\frac{k_{\perp}^{2}}{2m_{*}}\right)}\right]
=isgn(ν)2m2πvF04imνkdk(2mν)2+k4\displaystyle=-\frac{i{\rm sgn}(\nu)2m_{*}}{2\pi v_{F}}\int_{0}^{\infty}\frac{-4im_{*}\nu k_{\perp}dk_{\perp}}{\left(2m_{*}\nu\right)^{2}+k_{\perp}^{4}}
=2m2πvFπ2=m2vF\displaystyle=-\frac{2m_{*}}{2\pi v_{F}}\frac{\pi}{2}=-\frac{m_{*}}{2v_{F}} (69)

Plugging back into Eq.(66) we find

δΓ(ϵ)=2UmNκϵϵFdν2πνΓ0=UmπNκln(ϵFϵ)Γ0.\delta\Gamma(\epsilon)=\frac{2Um_{*}}{N\kappa}\int^{\epsilon_{F}}_{\epsilon}\frac{d\nu}{2\pi\nu}\Gamma_{0}=\frac{Um_{*}}{\pi N\kappa}\ln\left(\frac{\epsilon_{F}}{\epsilon}\right)\Gamma_{0}. (70)

This is the result used in main text Eq.(15). Interestingly, we arrive at the same logarithmically divergent vertex correction as in the problem of Fermi sea coupled by gauge-field fluctuation studied in Ref.[25], despite the absence of singular self-energy in our analysis. The presence or absence of self-energy does not affect the logarithmic divergence because the self-energy ultimate goes into S(ν)S(\nu), in which the integral over kk_{\perp} always yields a constant as shown in Eq.(69), so that the self-energy drops out.