\unnumbered
\equalcont

These authors contributed equally to this work.

\equalcont

These authors contributed equally to this work.

\equalcont

These authors contributed equally to this work.

[6]\fnmKin \surChung Fong

[1]\fnmPhilip \surKim

1]\orgdivDepartment of Physics, \orgnameHarvard University, \orgaddress\cityCambridge, \postcode02138, \stateMA, \countryUSA

2]\orgdivDepartment of Physics, \orgnameMassachusetts Institute of Technology, \orgaddress\cityCambridge, \postcode02138, \stateMA, \countryUSA

3]\orgdivResearch Center for Electronic and Optical Materials, \orgnameNational Institute for Materials Science, \orgaddress\street1-1 Namiki, \cityTsukuba, \postcode305-0044, \countryJapan

4]\orgdivResearch Center for Materials Nanoarchitectonics, \orgnameNational Institute for Materials Science, \orgaddress\street1-1 Namiki, \cityTsukuba, \postcode305-0044, \countryJapan

5]\orgdivDepartment of Physics, \orgnameUniversity of Connecticut, \orgaddress\stateMA, \postcode06269, \countryUSA

6]\orgdivRaytheon BBN Technologies, \orgaddress\stateMA, \countryUSA

Superfluid stiffness of twisted multilayer
graphene superconductors

\fnmAbhishek \surBanerjee    \fnmZeyu \surHao    \fnmMary \surKreidel    \fnmPatrick \surLedwith    \fnmIsabelle \surPhinney    \fnmJeong Min \surPark    \fnmAndrew \surZimmerman    \fnm Kenji \sur Watanabe    \fnmTakashi \surTaniguchi    \fnmRobert \surM Westervelt    \fnmPablo \surJarillo-Herrero    \fnmPavel A. \surVolkov    \fnmAshvin \surVishwanath    [email protected]    [email protected] [ [ [ [ [ [
Abstract

The robustness of the macroscopic quantum nature of a superconductor can be characterized by the superfluid stiffness, ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, a quantity that describes the energy required to vary the phase of the macroscopic quantum wave function. In unconventional superconductors, such as cuprates, the low-temperature behavior of ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT drastically differs from that of conventional superconductors due to quasiparticle excitations from gapless points (nodes) in momentum space. Intensive research on the recently discovered magic-angle twisted graphene family has revealed, in addition to superconducting states, strongly correlated electronic states associated with spontaneously broken symmetries, inviting the study of ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT to uncover the potentially unconventional nature of its superconductivity. Here we report the measurement of ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT in magic-angle twisted trilayer graphene (TTG), revealing unconventional nodal-gap superconductivity. Utilizing radio-frequency reflectometry techniques to measure the kinetic inductive response of superconducting TTG coupled to a microwave resonator, we find a linear temperature dependence of ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT at low temperatures and nonlinear Meissner effects in the current bias dependence, both indicating nodal structures in the superconducting order parameter. Furthermore, the do** dependence shows a linear correlation between the zero temperature ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and the superconducting transition temperature Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, reminiscent of Uemura’s relation in cuprates, suggesting phase-coherence-limited superconductivity. Our results provide strong evidence for nodal superconductivity in TTG and put strong constraints on the mechanisms of these graphene-based superconductors.

Superconductivity arises from the condensation of pairs of electrons, known as Cooper pairs, and is characterized by a complex order parameter, comprising an amplitude and a well-defined phase. Microscopically, the amplitude of the order parameter, ΔΔ\Deltaroman_Δ, is related to the strength of Cooper pairing, whereas the superfluid stiffness, ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, characterizes the phase-rigidity of the condensate order parameter. Measurements of ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT are highly motivated not only because phase rigidity is a defining property of the superconducting state, responsible for both zero resistance and the Meissner effect, but also because they help distinguish between conventional and unconventional superconducting orders [1, 2]. The dependence of ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT on temperature [1, 3] and superfluid velocity  [4, 5] carries distinctive signatures of the superconducting pairing symmetry. For instance, the linear temperature dependence of ρs(T)subscript𝜌𝑠𝑇\rho_{s}(T)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ) in cuprates is the hallmark of nodes in the superconducting order parameter, i.e., vanishing ΔΔ\Deltaroman_Δ along a specific direction in momentum space, resulting from the unconventional pairing in d-wave superconductivity in cuprates [6]. The magnitude and do** dependence of ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT also offer deep insights into the details of the superconducting gap and the underlying electron bands [1]. Notably, in a large class of unconventional superconductors, ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is orders of magnitude smaller than in conventional Bardeen-Cooper-Schrieffer (BCS)-type superconductors and follows unusual scaling laws [7] relating superconducting phase fluctuations with Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT [2, 8].

The recent discovery of magic-angle twisted multilayer graphene (MATMG)  [9, 10, 11, 12, 13] has provided an unprecedented opportunity to study the emergence of superconductivity due to strongly correlated electrons. However, the intense investigations thus far have relied primarily on electrical transport and scanning-tunneling-spectroscopy (STS) measurements, and the superconducting phase remains poorly understood. Electrical transport experiments varying displacement field, magnetic field, and carrier density have shown nematicity [14], Pauli limit violation [15], unusual behavior of the superconducting critical current [16], and correlation with symmetry breaking in the superconducting state [9, 10, 11], suggesting a paradigm beyond BCS physics. However, zero resistance of the superconducting state precludes any exploration of the internal structure of the superconducting dome using transport. On the other hand, STS studies showed V-shaped differential conductance spectra, consistent with nodal superconductivity [17, 18]. Yet this single-particle tunneling interpretation is complicated by a difficulty in distinguishing superconducting gaps from gaps due to non-superconducting orders such as flavor polarized correlated states. The need to better understand the superconducting order parameter in MATMG calls for measurements of thermodynamic quantities in the superconducting states such as ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. In particular, measuring ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT allows one to probe inside the superconducting dome, reveal the do** evolution of the superconducting state, and provide insights into the origin of superconductivity. However, this has been technically challenging due to the reduced sample dimensionality and size. Only indirect estimates of ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT based on critical-current density analyses are available so far [16].

In this work, we use radio-frequency (rf) resonant circuits [19, 20, 21, 22] to measure the superfluid response of micrometer-sized TTG superconductors. The microwave response of a superconductor at finite temperatures can be captured by a two-fluid model [23] — a parallel circuit of two channels: the superfluid channel composed of condensed Cooper pairs, and the normal fluid channel composed of quasi-particle excitations. The frequency (f𝑓fitalic_f) dependent complex rf conductivity in this model can be described by a dissipative real part, σ1=nne2τn/mesubscript𝜎1subscript𝑛𝑛superscript𝑒2subscript𝜏𝑛subscript𝑚𝑒\sigma_{1}=n_{n}e^{2}\tau_{n}/m_{e}italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_n start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_τ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT / italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT and dissipationless imaginary part, σ2=nse2/2πfmesubscript𝜎2subscript𝑛𝑠superscript𝑒22𝜋𝑓subscript𝑚𝑒\sigma_{2}=n_{s}e^{2}/2\pi fm_{e}italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 italic_π italic_f italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT, where e𝑒eitalic_e, mesubscript𝑚𝑒m_{e}italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT are electron charge and mass, respectively, nesubscript𝑛𝑒n_{e}italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT and nssubscript𝑛𝑠n_{s}italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT are the normal and superfluid electron densities and τnsubscript𝜏𝑛\tau_{n}italic_τ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is the scattering time for quasi-particle excitations. For the frequency regime f<fc=(ns/nn)/2πτn𝑓subscript𝑓𝑐subscript𝑛𝑠subscript𝑛𝑛2𝜋subscript𝜏𝑛f<f_{c}=(n_{s}/n_{n})/2\pi\tau_{n}italic_f < italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = ( italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_n start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) / 2 italic_π italic_τ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, σ1σ2much-less-thansubscript𝜎1subscript𝜎2\sigma_{1}\ll\sigma_{2}italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≪ italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, we expect the impedance of the superconducting sample to be dominated by a dissipationless inductive channel with the characteristic inductance LK=1/(2πfσ2)subscript𝐿𝐾12𝜋𝑓subscript𝜎2L_{K}=1/(2\pi f\sigma_{2})italic_L start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT = 1 / ( 2 italic_π italic_f italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ), often referred to as the kinetic inductance. LKsubscript𝐿𝐾L_{K}italic_L start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT is related to the superfluid stiffness by ρs=(w/l)(2/4e2LK)subscript𝜌𝑠𝑤𝑙superscriptPlanck-constant-over-2-pi24superscript𝑒2subscript𝐿𝐾\rho_{s}=(w/l)(\hbar^{2}/4e^{2}L_{K})italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = ( italic_w / italic_l ) ( roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 4 italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_L start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ), where w𝑤witalic_w and l𝑙litalic_l are the device width (perpendicular to the current flow) and length, respectively. Therefore, by incorporating the superconducting TTG sample into a microwave resonator with a resonance frequency that reflects LKsubscript𝐿𝐾L_{K}italic_L start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT, we can measure its superfluid stiffness.

Although techniques utilizing rf resonantors have been used to measure kinetic inductances [24, 25, 26], the large inherent contact resistance Rc110kΩsimilar-tosubscript𝑅𝑐110kΩR_{c}\sim 1-10~{}\mathrm{k\Omega}italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ∼ 1 - 10 roman_k roman_Ω and parasitic capacitive coupling CPsubscript𝐶𝑃C_{P}italic_C start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT in 2D material devices (see Fig. 1a) can completely damp out the resonance, thus forbidding the application of these techniques in measuring LKsubscript𝐿𝐾L_{K}italic_L start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT of 2D material superconductors. We solve this problem by using an impedance-matching network and adding a surface-mount inductor L0subscript𝐿0L_{0}italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Under our experimental conditions, CP110pFsimilar-tosubscript𝐶𝑃110pFC_{P}\sim 1-10~{}\mathrm{pF}italic_C start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ∼ 1 - 10 roman_pF and a choice of L0100200nHsimilar-tosubscript𝐿0100200nHL_{0}\sim 100-200~{}\mathrm{nH}italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∼ 100 - 200 roman_nH allows the combined impedance of the resonant circuit, ZDL0/(CPRC)similar-to-or-equalssubscript𝑍𝐷subscript𝐿0subscript𝐶𝑃subscript𝑅𝐶Z_{D}\simeq L_{0}/(C_{P}R_{C})italic_Z start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT ≃ italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / ( italic_C start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ), to match the characteristic rf matching impedance Z0=50Ωsubscript𝑍050ΩZ_{0}=50~{}\Omegaitalic_Z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 50 roman_Ω, resulting in lumped-element microwave resonant circuits with a resonance frequency fr0=1/2πfL0CP100300MHzsubscript𝑓𝑟012𝜋𝑓subscript𝐿0subscript𝐶𝑃similar-to100300MHzf_{r0}=1/\sqrt{2\pi fL_{0}C_{P}}\sim 100-300~{}\mathrm{MHz}italic_f start_POSTSUBSCRIPT italic_r 0 end_POSTSUBSCRIPT = 1 / square-root start_ARG 2 italic_π italic_f italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_C start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∼ 100 - 300 roman_MHz. Changes of kinetic inductance produce relative shifts of the microwave resonance frequency Δfr/fr0Δsubscript𝑓𝑟subscript𝑓𝑟0\Delta f_{r}/f_{r0}roman_Δ italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT / italic_f start_POSTSUBSCRIPT italic_r 0 end_POSTSUBSCRIPT, and they are related by a linear relationship ΔLK=2Rc2CP(Δfr/fr0)Δsubscript𝐿𝐾2superscriptsubscript𝑅𝑐2subscript𝐶𝑃Δsubscript𝑓𝑟subscript𝑓𝑟0\Delta L_{K}=2R_{c}^{2}C_{P}\left(\Delta f_{r}/f_{r0}\right)roman_Δ italic_L start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT = 2 italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_C start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ( roman_Δ italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT / italic_f start_POSTSUBSCRIPT italic_r 0 end_POSTSUBSCRIPT ) for expected values of LK10100similar-to-or-equalssubscript𝐿𝐾10100L_{K}\simeq 10-100italic_L start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ≃ 10 - 100s of nH 0.1Rc2CP1μabsent0.1superscriptsubscript𝑅𝑐2subscript𝐶𝑃similar-to-or-equals1μ\leq 0.1~{}R_{c}^{2}C_{P}\simeq 1~{}\mathrm{\upmu}≤ 0.1 italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_C start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ≃ 1 roman_μH. (see Methods). We contrast our approach with other experimental methods to measure superfluid stiffness that do not readily integrate with 2D material flakes [24, 25], have high operating frequencies [26], or require macroscopic-sized samples [27].

Our TTG device consists of three layers of graphene stacked with alternating twist angles of ±θ=1.55plus-or-minus𝜃superscript1.55\pm\theta=1.55^{\circ}± italic_θ = 1.55 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT [28], encapsulated within a layer of hBN and graphite on both sides, assembled on an insulating silicon substrate to minimize parasitic capacitance. The graphite layers serve as gate electrodes and can be independently controlled to tune the carrier density and displacement electric field D𝐷Ditalic_D. Fig. 1b shows an optical microscope image of the device, along with the measurement scheme, where we perform simultaneous DC electrical transport and microwave measurements (see Methods). Fig. 1c shows the 4-terminal DC resistance R𝑅Ritalic_R as a function of filling factor ν𝜈\nuitalic_ν, defined as the number of carriers per moiré unit cell. We observe superconductivity indicated by zero resistance in the range of filling factors 3<ν<23𝜈2-3<\nu<-2- 3 < italic_ν < - 2 for hole do** and 1.5<ν<31.5𝜈31.5<\nu<31.5 < italic_ν < 3 for electron do**, with a 2D νD𝜈𝐷\nu-Ditalic_ν - italic_D phase diagram (Fig. S1b) that is similar to our previous studies [10, 11]. Simultaneously, we measure the complex reflection coefficient of the resonator S21subscript𝑆21S_{21}italic_S start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT, defined as the ratio between the output (reflected) and input microwave voltage signal Vout/Vinsubscript𝑉𝑜𝑢𝑡subscript𝑉𝑖𝑛V_{out}/V_{in}italic_V start_POSTSUBSCRIPT italic_o italic_u italic_t end_POSTSUBSCRIPT / italic_V start_POSTSUBSCRIPT italic_i italic_n end_POSTSUBSCRIPT (see Methods). The amplitude |S21|subscript𝑆21\left|S_{21}\right|| italic_S start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT | as a function of ν𝜈\nuitalic_ν and frequency f𝑓fitalic_f, as shown in Fig. 1d, displays a dip around 350 MHz due to resonance through the whole range of ν𝜈\nuitalic_ν, and the resonance shifts to a lower frequency by Δfr1015similar-toΔsubscript𝑓𝑟1015\Delta f_{r}\sim 10-15roman_Δ italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ∼ 10 - 15 MHz as the sample transits from normal to superconducting states. The insets of Fig. 1c show two examples of the S21subscript𝑆21S_{21}italic_S start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT amplitude and phase response at the normal and superconducting states, respectively, with the filling indicated by the arrows of the corresponding color. We use a circle fitting method [29] to accurately extract the resonance frequency frsubscript𝑓𝑟f_{r}italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, from which we estimate LKsubscript𝐿𝐾L_{K}italic_L start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT and ρs=(w/l)(2/4e2LK)subscript𝜌𝑠𝑤𝑙superscriptPlanck-constant-over-2-pi24superscript𝑒2subscript𝐿𝐾\rho_{s}=(w/l)(\hbar^{2}/4e^{2}L_{K})italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = ( italic_w / italic_l ) ( roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 4 italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_L start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ). In this analysis, we employed w/l5similar-to-or-equals𝑤𝑙5w/l\simeq 5italic_w / italic_l ≃ 5 based on the device geometry (see Methods).

Focusing on the hole-side superconductor, we measure R𝑅Ritalic_R and frsubscript𝑓𝑟f_{r}italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT as a function of ν𝜈\nuitalic_ν and temperature T𝑇Titalic_T. R(ν,T)𝑅𝜈𝑇R(\nu,T)italic_R ( italic_ν , italic_T ) shows a superconducting dome manifesting as the zero resistance area in the dark-colored region in Fig. 2a, with a maximum Tc=1.2Ksubscript𝑇𝑐1.2KT_{c}=1.2~{}\mathrm{K}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 1.2 roman_K at the optimal do** νop=2.3subscript𝜈𝑜𝑝2.3\nu_{op}=-2.3italic_ν start_POSTSUBSCRIPT italic_o italic_p end_POSTSUBSCRIPT = - 2.3, where Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is defined as the onset temperature for non-zero values of the DC resistance. Correspondingly, fr(ν,T)subscript𝑓𝑟𝜈𝑇f_{r}(\nu,T)italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ( italic_ν , italic_T ) in Fig. 2b also displays a dome of lower resonance frequency in about the same area. Converting frsubscript𝑓𝑟f_{r}italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT to ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, in Fig. 2c we plot the temperature dependence of ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT in the units of both nH1superscriptnH1\mathrm{nH}^{-1}roman_nH start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and Kelvin (K) at several filling factors around νopsubscript𝜈𝑜𝑝\nu_{op}italic_ν start_POSTSUBSCRIPT italic_o italic_p end_POSTSUBSCRIPT within the superconducting dome. Across these filling factors, ρs(T)subscript𝜌𝑠𝑇\rho_{s}(T)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ) increases linearly with decreasing T𝑇Titalic_T, with no sign of saturation down to the lowest measurable temperature 30mKsimilar-toabsent30mK\sim 30~{}\mathrm{mK}∼ 30 roman_mK, strongly indicating that the superconducting gap is nodal across the entire superconducting dome. We emphasize that this linear-in-T𝑇Titalic_T behavior occurs below T0.20.5K0.4Tcsimilar-to𝑇0.20.5Ksimilar-to-or-equals0.4subscript𝑇𝑐T\sim 0.2-0.5~{}\mathrm{K}\simeq 0.4T_{c}italic_T ∼ 0.2 - 0.5 roman_K ≃ 0.4 italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, where the magnitude change of the order parameter is a subleading effect compared to the thermal activation of quasi-particles, and covers approximately an order of magnitude in temperature. At higher temperatures, ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT decreases in a reverse-S-shaped pattern and tends to flatten out, which coincides well with the onset of finite DC resistance (see Fig. S3).

The quasi-particle contribution to the linear-in-T𝑇Titalic_T suppression of ρs(T)ns(T)(ns(0)nqp(T))proportional-tosubscript𝜌𝑠𝑇subscript𝑛𝑠𝑇proportional-tosubscript𝑛𝑠0subscript𝑛𝑞𝑝𝑇\rho_{s}(T)\propto n_{s}(T)\propto(n_{s}(0)-n_{qp}(T))italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ) ∝ italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ) ∝ ( italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( 0 ) - italic_n start_POSTSUBSCRIPT italic_q italic_p end_POSTSUBSCRIPT ( italic_T ) ) implies a linear-in-T𝑇Titalic_T increase in the quasi-particle density, nqpsubscript𝑛𝑞𝑝n_{qp}italic_n start_POSTSUBSCRIPT italic_q italic_p end_POSTSUBSCRIPT. This is not expected in a fully gapped superconductor, where the quasi-particle population is exponentially suppressed for T0.3Tc𝑇0.3subscript𝑇𝑐T\leq 0.3~{}T_{c}italic_T ≤ 0.3 italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. Instead, it is consistent with a 2D nodal superconducting order, where the superconducting gap vanishes along certain directions — referred to as “nodes” — on the Fermi surface [1, 30], leading to a linear dispersion for the quasi-particle excitations. We illustrate this in Fig. 2c inset, where we plot the superconducting quasi-particle energy dispersion near two nodes that are symmetrically located at the Fermi momentum ±kFplus-or-minussubscript𝑘𝐹\pm k_{F}± italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT in the lattice momentum k𝑘kitalic_k-space. Two distinct velocities vΔsubscript𝑣Δv_{\Delta}italic_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT, corresponding to the slope of the superconducting gap, and vFsubscript𝑣𝐹v_{F}italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT, the Fermi velocity, describe the linear dispersion perpendicular and parallel to the axis connecting the two nodes, respectively. A non-zero T𝑇Titalic_T would immediately populate these cones with an equal number of quasi-particles and quasiholes, with their density linearly proportional to T𝑇Titalic_T.

We further characterize the do** dependence of the low-temperature linear behavior of ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT by performing a linear fit, as indicated by the dashed lines in Fig. 2c. From this, we extract the zero-temperature limit of the superfluid stiffness ρs0=ρs(T=0)subscript𝜌𝑠0subscript𝜌𝑠𝑇0\rho_{s0}=\rho_{s}(T=0)italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT = italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T = 0 ) and the slope dρs/dT𝑑subscript𝜌𝑠𝑑𝑇d\rho_{s}/dTitalic_d italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_d italic_T, shown in Fig 2d and Fig 2e, respectively. In Fig. 2d, we compare ρs0subscript𝜌𝑠0\rho_{s0}italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT with Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT extracted from DC measurements. The two quantities roughly track each other in a bell-shaped curve centered around ν=2.3𝜈2.3\nu=-2.3italic_ν = - 2.3, suggesting that the same mechanism determines both ρs0(ν)subscript𝜌𝑠0𝜈\rho_{s0}(\nu)italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT ( italic_ν ) and Tc(ν)subscript𝑇𝑐𝜈T_{c}(\nu)italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_ν ). On the other hand, the low-temperature slope dρs/dT𝑑subscript𝜌𝑠𝑑𝑇d\rho_{s}/dTitalic_d italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_d italic_T is roughly constant 0.20.3similar-toabsent0.20.3\sim 0.2-0.3∼ 0.2 - 0.3 for 2.6<ν<2.42.6𝜈2.4-2.6<\nu<-2.4- 2.6 < italic_ν < - 2.4, and then shows a sharp change rising to nearly 0.9similar-toabsent0.9\sim 0.9∼ 0.9 around ν=2.3𝜈2.3\nu=-2.3italic_ν = - 2.3 before tapering off to 0similar-toabsent0\sim 0∼ 0 as we approach the right edge of the dome at ν=2𝜈2\nu=-2italic_ν = - 2.

Considering that TTG superconductivity is a strictly 2D phenomenon, the experimentally obtained ρs(T)subscript𝜌𝑠𝑇\rho_{s}(T)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ) allows us to estimate the Berezinskii–Kosterlitz–Thouless (BKT) transition temperature T0subscript𝑇0T_{0}italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT using the Nelson–Kosterlitz criterion: TBKT=πρs(TBKT)/2subscript𝑇𝐵𝐾𝑇𝜋subscript𝜌𝑠subscript𝑇𝐵𝐾𝑇2T_{BKT}=\pi\rho_{s}(T_{BKT})/2italic_T start_POSTSUBSCRIPT italic_B italic_K italic_T end_POSTSUBSCRIPT = italic_π italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T start_POSTSUBSCRIPT italic_B italic_K italic_T end_POSTSUBSCRIPT ) / 2  [31]. In the BKT theory, for T>TBKT𝑇subscript𝑇𝐵𝐾𝑇T>T_{BKT}italic_T > italic_T start_POSTSUBSCRIPT italic_B italic_K italic_T end_POSTSUBSCRIPT, the phase-coherence of a 2D superconductor is destroyed due to the proliferation of unbound vortices, restoring a finite resistance due to their dissipative motion. Experimentally, we can estimate the density-dependent BKT transition temperature from Fig. 3a by locating the temperature T0subscript𝑇0T_{0}italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT where ρs(T,ν)subscript𝜌𝑠𝑇𝜈\rho_{s}(T,\nu)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T , italic_ν ) (smoothed for clarity) intersects with the universal ρc=2T/πsubscript𝜌𝑐2𝑇𝜋\rho_{c}=2T/\piitalic_ρ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 2 italic_T / italic_π plane. As shown in the top panel of Fig. 3a, we find the T0(ν)subscript𝑇0𝜈T_{0}(\nu)italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_ν ) is smaller than Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT obtained from the DC transport, indicating that the TTG device is superconducting even at T>T0𝑇subscript𝑇0T>T_{0}italic_T > italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. This observation is rather surprising as the simple BKT model seems to break down. While this observation could be explained by an unusually large vortex-core energy suppressing the proliferation of vortices T>TBKT𝑇subscript𝑇𝐵𝐾𝑇T>T_{BKT}italic_T > italic_T start_POSTSUBSCRIPT italic_B italic_K italic_T end_POSTSUBSCRIPT [32, 33], we find an explanation associated with sample inhomogeneity more plausible, as discussed below.

It is known that superconductivity in MATMG samples suffers from twist angle disorder  [34], which can create inhomogeneous superconductivity [10], and percolation paths when a supercurrent flows in the sample. Given such a network of superconducting paths, as illusrated in Fig. 3b inset, the aspect ratio of the device w/𝑤w/\ellitalic_w / roman_ℓ for two-terminal DC and rf transport should be rescaled using the narrower effective sample width wsuperscript𝑤w^{*}italic_w start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT. Therefore, our estimated ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT from the measured kinetic response underestimates the true superfluid stiffness by a factor of α=w/w𝛼superscript𝑤𝑤\alpha=w^{*}/witalic_α = italic_w start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT / italic_w. We can estimate α𝛼\alphaitalic_α from our experiment by noting the remarkable relationships between experimentally measured ρs0subscript𝜌𝑠0\rho_{s0}italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT, T0subscript𝑇0T_{0}italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, and Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, as shown in Fig. 3b. In particular, we find Tc/T03.0subscript𝑇𝑐subscript𝑇03.0T_{c}/T_{0}\approx 3.0italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≈ 3.0 by taking the ratio between the slopes of the linear parts in Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT versus ρs0subscript𝜌𝑠0\rho_{s0}italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT and in T0subscript𝑇0T_{0}italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT versus ρs0subscript𝜌𝑠0\rho_{s0}italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT. Considering TBKTTcsubscript𝑇𝐵𝐾𝑇subscript𝑇𝑐T_{BKT}\approx T_{c}italic_T start_POSTSUBSCRIPT italic_B italic_K italic_T end_POSTSUBSCRIPT ≈ italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT in MATMG [10, 11] (i.e., zero resistivity for T<TBKT𝑇subscript𝑇𝐵𝐾𝑇T<T_{BKT}italic_T < italic_T start_POSTSUBSCRIPT italic_B italic_K italic_T end_POSTSUBSCRIPT), we estimate α=w/wTBKT/T0Tc/T03.0𝛼superscript𝑤𝑤subscript𝑇𝐵𝐾𝑇subscript𝑇0subscript𝑇𝑐subscript𝑇03.0\alpha=w^{*}/w\approx T_{BKT}/T_{0}\approx T_{c}/T_{0}\approx 3.0italic_α = italic_w start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT / italic_w ≈ italic_T start_POSTSUBSCRIPT italic_B italic_K italic_T end_POSTSUBSCRIPT / italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≈ italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≈ 3.0

Having obtained α𝛼\alphaitalic_α, we can now make quantitative comparison between ρs0subscript𝜌𝑠0\rho_{s0}italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT and Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. We find a linear correlation between ρs0subscript𝜌𝑠0\rho_{s0}italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT and Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT away from the optimal do** — in both underdoped (ν>νop𝜈subscript𝜈𝑜𝑝\nu>\nu_{op}italic_ν > italic_ν start_POSTSUBSCRIPT italic_o italic_p end_POSTSUBSCRIPT) and overdoped (ν<νop𝜈subscript𝜈𝑜𝑝\nu<\nu_{op}italic_ν < italic_ν start_POSTSUBSCRIPT italic_o italic_p end_POSTSUBSCRIPT) regimes. A linear dependence of Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT on ρs0subscript𝜌𝑠0\rho_{s0}italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT was first identified in underdoped cuprates  [7], and subsequently observed in a variety of unconventional superconductors including overdoped cuprates [35, 36, 27], suggesting a universal scaling behavior known as the Uemura’s relation. This phenomenological dependence suggests that long-range phase order, characterized by ρs0subscript𝜌𝑠0\rho_{s0}italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT, determines Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, in contrast to conventional BCS superconductors where the strength of Cooper pairing, characterized by the BCS gap ΔΔ\Deltaroman_Δ, determines Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. Following this interpretation and assuming parabolic band dispersions in TTG, we estimate Tc/EF=Tc/(α4πρs0)0.08subscript𝑇𝑐subscript𝐸𝐹subscript𝑇𝑐𝛼4𝜋subscript𝜌𝑠0similar-to-or-equals0.08T_{c}/E_{F}=T_{c}/(\alpha 4\pi\rho_{s0})\simeq 0.08italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / ( italic_α 4 italic_π italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT ) ≃ 0.08 based on the linear slope between Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and ρs0subscript𝜌𝑠0\rho_{s0}italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT. This result indicates strongly-coupled superconductivity, consistent with more indirect analyses in previous works [10, 11]. Note that for non-parabolic bands (see Supplementary Material), EFsubscript𝐸𝐹E_{F}italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT is larger and therefore the estimate will be somewhat lower. We have observed similar linear scaling relationships in other devices, in both electron and hole-doped sectors, even when effects of twist-angle disorder are evident (see Fig. S7). This suggests the generality of this effect in MATMG superconductivity.

We now focus on the density-dependent behaviors of ρs0subscript𝜌𝑠0\rho_{s0}italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT and dρs/dT𝑑subscript𝜌𝑠𝑑𝑇d\rho_{s}/dTitalic_d italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_d italic_T measured at the zero temperature limit. Recalling that ρs=2ns/(4me)subscript𝜌𝑠superscriptPlanck-constant-over-2-pi2subscript𝑛𝑠4subscript𝑚𝑒\rho_{s}=\hbar^{2}n_{s}/(4m_{e})italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / ( 4 italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ) for parabolic bands described by an effective mass mesubscript𝑚𝑒m_{e}italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT, we expect a linear relationship between ρs0subscript𝜌𝑠0\rho_{s0}italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT and νnsproportional-to𝜈subscript𝑛𝑠\nu\propto n_{s}italic_ν ∝ italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. The bell-shaped dependence observed in the experiment [Fig. 2d], particularly the drop of ρs0subscript𝜌𝑠0\rho_{s0}italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT with increasing hole do** for ν<2.4𝜈2.4\nu<-2.4italic_ν < - 2.4, suggests a departure from the parabolic-band approximation. Theoretical modeling of the TTG band structure for moiré fillings between 3<ν<23𝜈2-3<\nu<-2- 3 < italic_ν < - 2 suggests the presence of strongly renormalized band dispersion E(k)𝐸𝑘E(k)italic_E ( italic_k ) that are highly non-parabolic, as shown in Fig. 3c, obtained by Hartree-Fock calculation assuming flavor polarization of the TTG flat band. An interplay between strong Coulomb repulsion and quantum geometric effects, particularly the concentration of Berry curvature near k=0𝑘0k=0italic_k = 0, leads to a band structure that is highly dispersive close to ν=2𝜈2\nu=-2italic_ν = - 2 but becomes progressively flatter on approaching ν=3𝜈3\nu=-3italic_ν = - 3. For such non-parabolic bands, a more general formula is required for the superfluid stiffness: ρs(ν)=18𝒌nk𝒌2E=vF(ν)kF(ν)8πsubscript𝜌𝑠𝜈18subscript𝒌subscript𝑛𝑘superscriptsubscript𝒌2𝐸subscript𝑣𝐹𝜈subscript𝑘𝐹𝜈8𝜋\rho_{s}(\nu)=\frac{1}{8}\sum_{{{\boldsymbol{k}}}}n_{k}\nabla_{{\boldsymbol{k}% }}^{2}E=\frac{v_{F}(\nu)k_{F}(\nu)}{8\pi}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_ν ) = divide start_ARG 1 end_ARG start_ARG 8 end_ARG ∑ start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ∇ start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_E = divide start_ARG italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_ν ) italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_ν ) end_ARG start_ARG 8 italic_π end_ARG (for an isotropic E(𝒌)𝐸𝒌E({\boldsymbol{k}})italic_E ( bold_italic_k )), leading to a do** dependence as shown in the theoretical estimate in Fig. 3d, which is shown in yellow and plotted together with the smoothed experimental curve shown in red. Starting at ν=2𝜈2\nu=-2italic_ν = - 2, the initial steep rise of ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is enabled by a large positive band curvature 𝒌2Esuperscriptsubscript𝒌2𝐸\nabla_{{\boldsymbol{k}}}^{2}E∇ start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_E near k=0𝑘0k=0italic_k = 0. However, on further do**, 𝒌2Esuperscriptsubscript𝒌2𝐸\nabla_{{\boldsymbol{k}}}^{2}E∇ start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_E becomes negative, causing a reduction of the superfluid stiffness. Remarkably, the maximum of ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is reached well below the half-filling of the band due to its strong particle-hole asymmetry. However, the theoretical estimates are larger than the experimental values by a factor of ×\times×5 even after the geometrical factor correction by α𝛼\alphaitalic_α. This discrepancy indicates that only a fraction of the electronic spectral weight is condensed into the superfluid. We leave the understanding of this to future studies. It is worth noting that a similar reduction of superfluid stiffness with increasing do** is observed in overdoped cuprates and remains a subject of active investigation [27, 37].

With the geometrical correction α𝛼\alphaitalic_α, we can also now provide a quantitative analysis on the low temperature linear dependence of ρs(T)subscript𝜌𝑠𝑇\rho_{s}(T)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ). Within the physical picture of nodal superconducting gaps, the aforementioned gap and electron band velocities vΔsubscript𝑣Δv_{\Delta}italic_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT and vFsubscript𝑣𝐹v_{F}italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT are related to the linear slope of ρs(T)subscript𝜌𝑠𝑇\rho_{s}(T)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ) by vF/vΔ=(8π/Nlog(2))dρs/dT×αsubscript𝑣𝐹subscript𝑣Δ8𝜋𝑁2𝑑subscript𝜌𝑠𝑑𝑇𝛼v_{F}/v_{\Delta}=(8\pi/N\log(2))d\rho_{s}/dT\times\alphaitalic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT / italic_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT = ( 8 italic_π / italic_N roman_log ( 2 ) ) italic_d italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_d italic_T × italic_α, where N𝑁Nitalic_N is the number of nodes (see Methods for details). Tentatively assuming N=4𝑁4N=4italic_N = 4 (d-wave) and considering the scaling factor α=3𝛼3\alpha=3italic_α = 3, we obtain vF/vΔ520similar-to-or-equalssubscript𝑣𝐹subscript𝑣Δ520v_{F}/v_{\Delta}\simeq 5-20italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT / italic_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT ≃ 5 - 20 as a function of ν𝜈\nuitalic_ν, as shown in Fig. 3e. We also estimate vF/vΔsubscript𝑣𝐹subscript𝑣Δv_{F}/v_{\Delta}italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT / italic_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT from current-bias dependent measurements ρs(I)subscript𝜌𝑠𝐼\rho_{s}(I)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_I ) (discussed below), and obtain similar values [Fig. 3e], suggesting the robustness of our estimation of vF/vΔsubscript𝑣𝐹subscript𝑣Δv_{F}/v_{\Delta}italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT / italic_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT.

To solidify our interpretation of nodal superconductivity, we measure both R𝑅Ritalic_R and ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT in the presence of a finite DC supercurrent bias I𝐼Iitalic_I. In nodal superconductors, a current I𝐼Iitalic_I produces a Doppler shift of the quasi-particle spectrum due to finite superfluid velocity, resulting in a current-dependent quasi-particle population at the nodal points. This phenomenon is known as the non-linear Meissner effect and was initially proposed as a zero-temperature test for nodal superconductivity [4], and observed in later experiments in cuprates [38, 39, 40].

Figures 4a and  4b show R𝑅Ritalic_R as a function of I𝐼Iitalic_I and T𝑇Titalic_T, and a function of I𝐼Iitalic_I and ν𝜈\nuitalic_ν, respectively, with superconductivity indicated by the dark regions of zero resistance. Within the superconducting regime, we measure a rapid suppression of ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT with increasing I𝐼Iitalic_I [Fig. 4(c,d)], with the strongest signatures close to a range of filling factors around the optimal do** (ν=2.3𝜈2.3\nu=-2.3italic_ν = - 2.3, Ic0.12μsimilar-to-or-equalssubscript𝐼𝑐0.12μI_{c}\simeq 0.12~{}\mathrm{\upmu}italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≃ 0.12 roman_μA) [Fig. 4d]. The suppression of ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT follows a quadratic behavior, δρs=bI2𝛿subscript𝜌𝑠𝑏superscript𝐼2\delta\rho_{s}=-bI^{2}italic_δ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = - italic_b italic_I start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, close to I0𝐼0I\to 0italic_I → 0, which turns linear with δρs=cI𝛿subscript𝜌𝑠𝑐𝐼\delta\rho_{s}=-cIitalic_δ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = - italic_c italic_I after I𝐼Iitalic_I exceeds a cross-over scale Isuperscript𝐼I^{*}italic_I start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT. We further investigate the temperature dependence of this effect at ν=2.34𝜈2.34\nu=-2.34italic_ν = - 2.34, close to the optimal do**, and find that the quadratic behavior at low bias is strongly temperature dependent [Fig. 4c] with the quadratic coefficient b𝑏bitalic_b appearing to diverge as T0𝑇0T\to 0italic_T → 0 [Fig. 4e]. We also observe a non-trivial filling factor dependence of both the quadratic coefficient, b(ν)𝑏𝜈b(\nu)italic_b ( italic_ν ), and the linear coefficient, c(ν)𝑐𝜈c(\nu)italic_c ( italic_ν ), measured at T=30𝑇30T=30italic_T = 30 mK [Fig. 4(f,g)].

To understand these behaviors, we model the electrodynamic response of a 2D nodal superconductor in the presence of a DC supercurrent bias (see Methods), and obtain a quadratic dependence for I0𝐼0I\to 0italic_I → 0:

Δρs(I)=vF3vΔ1TN128(4eρs0w)2I2Δsubscript𝜌𝑠𝐼superscriptsubscript𝑣𝐹3subscript𝑣Δ1𝑇𝑁128superscript4𝑒subscript𝜌𝑠0superscript𝑤2superscript𝐼2\Delta\rho_{s}(I)=-\frac{v_{F}^{3}}{v_{\Delta}}\frac{1}{T}\frac{N}{128(4e\rho_% {s0}w^{*})^{2}}I^{2}roman_Δ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_I ) = - divide start_ARG italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT end_ARG divide start_ARG 1 end_ARG start_ARG italic_T end_ARG divide start_ARG italic_N end_ARG start_ARG 128 ( 4 italic_e italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT italic_w start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_I start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (1)

and a linear dependence for I>I𝐼superscript𝐼I>I^{*}italic_I > italic_I start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT,

Δρs(I)=vF2vΔN|cosθI,vn|16(4eρs0w)IΔsubscript𝜌𝑠𝐼superscriptsubscript𝑣𝐹2subscript𝑣Δ𝑁delimited-⟨⟩subscript𝜃𝐼subscript𝑣𝑛164𝑒subscript𝜌𝑠0superscript𝑤𝐼\Delta\rho_{s}(I)=-\frac{v_{F}^{2}}{v_{\Delta}}\frac{N\langle|\cos\theta_{I,v_% {n}}|\rangle}{16(4e\rho_{s0}w^{*})}Iroman_Δ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_I ) = - divide start_ARG italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT end_ARG divide start_ARG italic_N ⟨ | roman_cos italic_θ start_POSTSUBSCRIPT italic_I , italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_POSTSUBSCRIPT | ⟩ end_ARG start_ARG 16 ( 4 italic_e italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT italic_w start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) end_ARG italic_I (2)

with the crossover current I(T)=(2eρs0w/vF)Tsuperscript𝐼𝑇2𝑒subscript𝜌𝑠0superscript𝑤subscript𝑣𝐹𝑇I^{*}(T)=(2e\rho_{s0}w^{*}/v_{F})Titalic_I start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_T ) = ( 2 italic_e italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT italic_w start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT / italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) italic_T separating the quadratic and linear regimes of ρs(I)subscript𝜌𝑠𝐼\rho_{s}(I)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_I ). Here |cosθI,vn|delimited-⟨⟩subscript𝜃𝐼subscript𝑣𝑛\langle|\cos\theta_{I,v_{n}}|\rangle⟨ | roman_cos italic_θ start_POSTSUBSCRIPT italic_I , italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_POSTSUBSCRIPT | ⟩ represents the angle between the applied supercurrent direction and the Fermi velocity at the node, averaged over all gap nodes.

Our model captures all three experimental features: the quadratic suppression of stiffness at I0𝐼0I\to 0italic_I → 0 (Fig. 4c), the crossover to a linear dependence at a current threshold Isuperscript𝐼I^{*}italic_I start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT (Fig. 4d), and the divergence of the quadratic coefficient b𝑏bitalic_b for T0𝑇0T\to 0italic_T → 0 (Fig. 4e). Particularly, the divergence of b(T0)𝑏𝑇0b(T\to 0)italic_b ( italic_T → 0 ) is a strong signature of nodal superconductivity that cannot be easily mimicked by anisotropically gapped superconducting states, providing an additional signature of nodal pairing symmetry [41, 42, 40]. Similar divergences of b(T)𝑏𝑇b(T)italic_b ( italic_T ) are measured in a second device at optimal do**s of both the electron-like and hole-like superconducting sectors (see Fig. S8).

Choosing N=4𝑁4N=4italic_N = 4 and |cosθI,vn|=0.7delimited-⟨⟩subscript𝜃𝐼subscript𝑣𝑛0.7\langle|\cos\theta_{I,v_{n}}|\rangle=0.7⟨ | roman_cos italic_θ start_POSTSUBSCRIPT italic_I , italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_POSTSUBSCRIPT | ⟩ = 0.7, we can estimate vFsubscript𝑣𝐹v_{F}italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT and vΔsubscript𝑣Δv_{\Delta}italic_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT using Eq. 1 and Eq. 2, and b(ν)𝑏𝜈b(\nu)italic_b ( italic_ν ) and c(ν)𝑐𝜈c(\nu)italic_c ( italic_ν ) extracted from experiments (see Methods for details). The estimated results are shown in Fig. 4h, where vFsubscript𝑣𝐹v_{F}italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT covers a range of (0.10.7)×1050.10.7superscript105(0.1-0.7)\times 10^{5}( 0.1 - 0.7 ) × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT m/s, and vΔsubscript𝑣Δv_{\Delta}italic_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT covers a range of (0.20.5)×1040.20.5superscript104(0.2-0.5)\times 10^{4}( 0.2 - 0.5 ) × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT m/s, roughly an order of magnitude smaller than vFsubscript𝑣𝐹v_{F}italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT. Remarkably, the density-dependent ratio vF/vΔsubscript𝑣𝐹subscript𝑣Δv_{F}/v_{\Delta}italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT / italic_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT estimated here yields similar magnitude (520similar-to-or-equalsabsent520\simeq 5-20≃ 5 - 20) and dependence on ν𝜈\nuitalic_ν to estimates in Fig. 3e, which was based on an independent analysis of ρs(T)subscript𝜌𝑠𝑇\rho_{s}(T)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ) measurements. An alternative way to estimate vFsubscript𝑣𝐹v_{F}italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT is to use experimental ρs0(ν)subscript𝜌𝑠0𝜈\rho_{s0}(\nu)italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT ( italic_ν ) and the aforemetioned expression ρs(ν)=18𝒌nk𝒌2E=vF(ν)kF(ν)8πsubscript𝜌𝑠𝜈18subscript𝒌subscript𝑛𝑘superscriptsubscript𝒌2𝐸subscript𝑣𝐹𝜈subscript𝑘𝐹𝜈8𝜋\rho_{s}(\nu)=\frac{1}{8}\sum_{{{\boldsymbol{k}}}}n_{k}\nabla_{{\boldsymbol{k}% }}^{2}E=\frac{v_{F}(\nu)k_{F}(\nu)}{8\pi}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_ν ) = divide start_ARG 1 end_ARG start_ARG 8 end_ARG ∑ start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ∇ start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_E = divide start_ARG italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_ν ) italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_ν ) end_ARG start_ARG 8 italic_π end_ARG. This approach produces a similar but slightly smaller vFsubscript𝑣𝐹v_{F}italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT (0.10.2×105similar-to-or-equalsabsent0.10.2superscript105\simeq 0.1-0.2\times 10^{5}≃ 0.1 - 0.2 × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT m/s), shown as the purple solid line in Fig. 4h.

To summarize, our superfluid stiffness measurements provide novel insights into the nature of superconductivity in MATMG. A confluence of signatures, including the linear-in-T𝑇Titalic_T suppression of superfluid stiffness, the observation of the nonlinear Meissner effect, and its temperature dependence provide strong evidence for a nodal pairing symmetry. The linear scaling relation between superfluid stiffness at the zero temperature limit and the superconducting transition temperature suggests an unusual scenario where the superconducting transition is controlled by phase fluctuations rather than Cooper-pair breaking.

1 Methods

1.1 Device fabrication

We prepare the twisted trilayer graphene (TTG) heterostructures using the standard dry-transfer method, where we use stamps consisting of polycarbonate (PC) polymer and polydimethylsiloxane (PDMS) to pick up each 2D material flake sequentially. The TTG heterostructures consist of hBN/graphite/hBN/TTG/hBN/graphite/hBN from top to bottom. Graphite flakes are used as top and bottom gates to control the density and displacement field in TTG. The three pieces of TTG come from the same monolayer graphene, which was pre-cut into three with a laser.

To make the TTG structure, we pick up each piece sequentially, with a rotation angle θ𝜃\thetaitalic_θ and θ𝜃-\theta- italic_θ applied to the stage before picking the second and third piece, respectively. After picking up all layers, we drop the stack onto an undoped silicon substrate, which creates a much less parasitic capacitance than a dope one in the rf measurements. We then etch the stack into a Hall bar geometry and deposit Cr/Au contacts through standard nanofabrication lithography. We intentionally make contacts wide, especially for those used for RF measurements, given that RF reflectometry is intrinsically a two-terminal measurement and benefits from wide uniform sample path. We also minimize the area that’s between the gold-TTG edge contact region and the double-gated TTG channel region so that the two-terminal signal is dominated by the channel region, in addition to the inevitable contact resistance.

1.2 Measurements

Measurements were performed in a Bluefors dilution refrigerator with a base electron temperature of T=20𝑇20T=20italic_T = 20 mK, achieved with extensive filtering, shielding, and thermal coupling of the sample within copper packaging. DC transport measurements are performed using the a.c. lock-in technique with an excitation current of 1-10 nA and a low excitation frequency similar-to\sim 17 Hz. All DC lines were filtered using a set of RC filters mounted on the 4 K plate and LC filters, to avoid dissipation, mounted at the mixing chamber. The device was embedded in the resonant circuit by way of bias tees and wire bonding the reflectometry line to one of the sample leads, while another lead provided the rf ground. Reflectometry measurements were performed using a vector network analyzer (VNA). The rf excitation signal generated from the VNA was first attenuated by 32 dB via cryostat attenuators installed at various stages of the fridge, and then by another 20 dB as it passed from the coupled-port to the in-port of the directional coupler installed below the mixing chamber. The direction coupler provides a means to remove any spurious resonances in the background of the measurement chain by subtracting the sample signal from measured gain line without the sample. This allows us to precisely quantify the sample resonance shift, and subsequently its superfluid stiffness, without contributions from the measurement chain. After reflection from the LCR circuit, the signal passes through a K&L filter (450 MHz cutoff), is amplified by 40 dB using a cryogenic amplifier (Weinreb CITLF3) mounted on the 4K plate, and amplified again by 40 dB using a room-temperature amplifier before being measured by the VNA (see Fig. S2). The rf power incident on the sample was varied from -112 dBm to -102 dBm. Measurements were performed with a bandwidth of 0.5-1 kHz, with each VNA trace averaged 5-10 times.

1.3 Analytical Model of resonant circuit

Fig. S3(a) shows a schematic of the effective microwave resonant circuit, where L0subscript𝐿0L_{0}italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the externally added inductor, CPsubscript𝐶𝑃C_{P}italic_C start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT represents the parasitic capacitance arising primarily from the circuit board, bond wires and the substrate, but may also be added externally in parallel, RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT is the contact resistance and LKsubscript𝐿𝐾L_{K}italic_L start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT is the inductance of the device in the superconducting state. This circuit can be simplified to an effective series RLC model, shown in Fig. S3(b) where, using the method of series-parallel transformations, we obtain Reff=L0/(RcCeff)subscript𝑅𝑒𝑓𝑓subscript𝐿0subscript𝑅𝑐subscript𝐶𝑒𝑓𝑓R_{eff}=L_{0}/(R_{c}C_{eff})italic_R start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT = italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / ( italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_C start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ) and Ceff=(CPL0/Rc2+LK/Rc2)subscript𝐶𝑒𝑓𝑓subscript𝐶𝑃subscript𝐿0superscriptsubscript𝑅𝑐2subscript𝐿𝐾superscriptsubscript𝑅𝑐2C_{eff}=(C_{P}-L_{0}/R_{c}^{2}+L_{K}/R_{c}^{2})italic_C start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT = ( italic_C start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT - italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_L start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ). Modeling the resonator as a series RLC circuit coupled to a Z0=50Ωsubscript𝑍050ΩZ_{0}=50~{}\Omegaitalic_Z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 50 roman_Ω transmission line, we can express the impedance of the resonator as:

Z=Reff+2πifL0+12πifCeffQCZ0(1QI+i2Δffr)𝑍subscript𝑅𝑒𝑓𝑓2𝜋𝑖𝑓subscript𝐿012𝜋𝑖𝑓subscript𝐶𝑒𝑓𝑓similar-to-or-equalssubscript𝑄𝐶subscript𝑍01subscript𝑄𝐼𝑖2Δ𝑓subscript𝑓𝑟Z=R_{eff}+2\pi ifL_{0}+\frac{1}{2\pi ifC_{eff}}\simeq Q_{C}Z_{0}\left(\frac{1}% {Q_{I}}+i\frac{2\Delta f}{f_{r}}\right)italic_Z = italic_R start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT + 2 italic_π italic_i italic_f italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + divide start_ARG 1 end_ARG start_ARG 2 italic_π italic_i italic_f italic_C start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT end_ARG ≃ italic_Q start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( divide start_ARG 1 end_ARG start_ARG italic_Q start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_ARG + italic_i divide start_ARG 2 roman_Δ italic_f end_ARG start_ARG italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_ARG ) (3)

where the right-hand side is valid near the resonance frequency, Δf=(ffr)frΔ𝑓𝑓subscript𝑓𝑟much-less-thansubscript𝑓𝑟\Delta f=(f-f_{r})\ll f_{r}roman_Δ italic_f = ( italic_f - italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) ≪ italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT. Here Qc=Zchar/Z0subscript𝑄𝑐subscript𝑍𝑐𝑎𝑟subscript𝑍0Q_{c}=Z_{char}/Z_{0}italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = italic_Z start_POSTSUBSCRIPT italic_c italic_h italic_a italic_r end_POSTSUBSCRIPT / italic_Z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the coupling quality factor, QI=Zchar/Reffsubscript𝑄𝐼subscript𝑍𝑐𝑎𝑟subscript𝑅𝑒𝑓𝑓Q_{I}=Z_{char}/R_{eff}italic_Q start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = italic_Z start_POSTSUBSCRIPT italic_c italic_h italic_a italic_r end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT is the internal quality factor, and the characteristic impedance Zchar=L0/Ceffsubscript𝑍𝑐𝑎𝑟subscript𝐿0subscript𝐶𝑒𝑓𝑓Z_{char}=\sqrt{L_{0}/C_{eff}}italic_Z start_POSTSUBSCRIPT italic_c italic_h italic_a italic_r end_POSTSUBSCRIPT = square-root start_ARG italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_C start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT end_ARG. We also define a loaded quality factor 1/QL=(1/QC+1/QI)1subscript𝑄𝐿1subscript𝑄𝐶1subscript𝑄𝐼1/Q_{L}=(1/Q_{C}+1/Q_{I})1 / italic_Q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT = ( 1 / italic_Q start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT + 1 / italic_Q start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT ).

The reflection coefficient of the resonator, measured by the reflectometry circuit, is given as:

Γ=ZZ0Z+Z0=12QL/QC1+2iQL(Δf/fr)Γ𝑍subscript𝑍0𝑍subscript𝑍012subscript𝑄𝐿subscript𝑄𝐶12𝑖subscript𝑄𝐿Δ𝑓subscript𝑓𝑟\Gamma=\frac{Z-Z_{0}}{Z+Z_{0}}=1-\frac{2Q_{L}/Q_{C}}{1+2iQ_{L}(\Delta f/f_{r})}roman_Γ = divide start_ARG italic_Z - italic_Z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_Z + italic_Z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG = 1 - divide start_ARG 2 italic_Q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT / italic_Q start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT end_ARG start_ARG 1 + 2 italic_i italic_Q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( roman_Δ italic_f / italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) end_ARG (4)

where the resonance frequency fr=1/L0Cefffr0(1+0.5LK/(Rc2CP))subscript𝑓𝑟1subscript𝐿0subscript𝐶𝑒𝑓𝑓similar-to-or-equalssubscript𝑓𝑟010.5subscript𝐿𝐾superscriptsubscript𝑅𝑐2subscript𝐶𝑃f_{r}=1/\sqrt{L_{0}C_{eff}}\simeq f_{r0}\left(1+0.5L_{K}/(R_{c}^{2}C_{P})\right)italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = 1 / square-root start_ARG italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_C start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT end_ARG ≃ italic_f start_POSTSUBSCRIPT italic_r 0 end_POSTSUBSCRIPT ( 1 + 0.5 italic_L start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT / ( italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_C start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ) ), and fr0=1L0CP1Rc2CP2subscript𝑓𝑟01subscript𝐿0subscript𝐶𝑃1superscriptsubscript𝑅𝑐2superscriptsubscript𝐶𝑃2f_{r0}=\sqrt{\frac{1}{L_{0}C_{P}}-\frac{1}{R_{c}^{2}C_{P}^{2}}}italic_f start_POSTSUBSCRIPT italic_r 0 end_POSTSUBSCRIPT = square-root start_ARG divide start_ARG 1 end_ARG start_ARG italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_C start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG - divide start_ARG 1 end_ARG start_ARG italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_C start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG. In Fig. S3(c), we numerically evaluate the behavior of the circuit with changing kinetic inductance, for L0=205subscript𝐿0205L_{0}=205italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 205 nH, CP=1.5subscript𝐶𝑃1.5C_{P}=1.5italic_C start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT = 1.5 pF, and Rc=3.5subscript𝑅𝑐3.5R_{c}=3.5italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 3.5 kΩΩ\Omegaroman_Ω. These values are very typical of a realistic experiment. We compare our analytical expression for frsubscript𝑓𝑟f_{r}italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT with the exact result and obtain a quantitative match for LK500subscript𝐿𝐾500L_{K}\leq 500italic_L start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ≤ 500 nH, the desired operation regime for our experiments, where the frequency shift is linearly related to the kinetic inductance, with Δfr/fr0=1/(2Rc2CP)ΔLKΔsubscript𝑓𝑟subscript𝑓𝑟012superscriptsubscript𝑅𝑐2subscript𝐶𝑃Δsubscript𝐿𝐾\Delta f_{r}/f_{r0}=1/(2R_{c}^{2}C_{P})\Delta L_{K}roman_Δ italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT / italic_f start_POSTSUBSCRIPT italic_r 0 end_POSTSUBSCRIPT = 1 / ( 2 italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_C start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ) roman_Δ italic_L start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT.

1.4 Circle fitting and analysis

Reflectometry measurement using a vector network analyzer provides the complex reflection coefficient S21subscript𝑆21S_{21}italic_S start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT as a function of frequency. Apart from the ideal resonator response described in Eq. 4, two additional contributions appear in standard reflectometry measurements [29]. First, we account for an environmental factor: aei(α2πifτ)𝑎superscript𝑒𝑖𝛼2𝜋𝑖𝑓𝜏ae^{i(\alpha-2\pi if\tau)}italic_a italic_e start_POSTSUPERSCRIPT italic_i ( italic_α - 2 italic_π italic_i italic_f italic_τ ) end_POSTSUPERSCRIPT, representing an additional amplitude a𝑎aitalic_a, a phase shift α𝛼\alphaitalic_α and a cable delay τ𝜏\tauitalic_τ determined by the speed of light and the finite length (similar-to\sim50 mm) of the coaxial cable connecting the device to the direction coupler mounted on the mixing chamber. Second, a phenomenological term ϕitalic-ϕ\phiitalic_ϕ is introduced to account for the finite asymmetry of the resonances. This asymmetry is usually small and appears in the complex plane represented by (S21)(S21)subscript𝑆21subscript𝑆21\Re({S_{21}})-\Im({S_{21}})roman_ℜ ( italic_S start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT ) - roman_ℑ ( italic_S start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT ) as a rotation of the resonance circle along the real axis by an amount ϕitalic-ϕ\phiitalic_ϕ [29].

S21=aei(α2πifτ)[12(QL/QC)eiϕ1+2iQL(f/fr1)]subscript𝑆21𝑎superscript𝑒𝑖𝛼2𝜋𝑖𝑓𝜏delimited-[]12subscript𝑄𝐿subscript𝑄𝐶superscript𝑒𝑖italic-ϕ12𝑖subscript𝑄𝐿𝑓subscript𝑓𝑟1S_{21}=ae^{i(\alpha-2\pi if\tau)}\left[1-\frac{2(Q_{L}/Q_{C})e^{i\phi}}{1+2iQ_% {L}(f/f_{r}-1)}\right]italic_S start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT = italic_a italic_e start_POSTSUPERSCRIPT italic_i ( italic_α - 2 italic_π italic_i italic_f italic_τ ) end_POSTSUPERSCRIPT [ 1 - divide start_ARG 2 ( italic_Q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT / italic_Q start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ) italic_e start_POSTSUPERSCRIPT italic_i italic_ϕ end_POSTSUPERSCRIPT end_ARG start_ARG 1 + 2 italic_i italic_Q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_f / italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT - 1 ) end_ARG ] (5)

We fit our S21subscript𝑆21S_{21}italic_S start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT data in the complex plane to Eq. 5 using the publicly available circle fitting package [29]. An example of the fitting in the complex plane is shown in Fig.  S4. The fit parameters frsubscript𝑓𝑟f_{r}italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, QLsubscript𝑄𝐿Q_{L}italic_Q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT, and QCsubscript𝑄𝐶Q_{C}italic_Q start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT are further analyzed using the analytical model described in Sec. 1.3. In Fig. S5 we show the analysis procedure for converting the resonance frequency into kinetic inductance and superfluid stiffness, as a function of temperature at a representative value of the filling factor ν=2.4𝜈2.4\nu=-2.4italic_ν = - 2.4. We can similarly perform a quality factor analysis using QLsubscript𝑄𝐿Q_{L}italic_Q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT and QCsubscript𝑄𝐶Q_{C}italic_Q start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT and estimate RF resistance (Rrfsubscript𝑅𝑟𝑓R_{rf}italic_R start_POSTSUBSCRIPT italic_r italic_f end_POSTSUBSCRIPT) of the sample, using the relation Rrf=L0Z0CP(QC/QL1)subscript𝑅𝑟𝑓subscript𝐿0subscript𝑍0subscript𝐶𝑃subscript𝑄𝐶subscript𝑄𝐿1R_{rf}=\frac{L_{0}}{Z_{0}C_{P}}(Q_{C}/Q_{L}-1)italic_R start_POSTSUBSCRIPT italic_r italic_f end_POSTSUBSCRIPT = divide start_ARG italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_Z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_C start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ( italic_Q start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT / italic_Q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT - 1 ). At perfect matching, QL=QC/2subscript𝑄𝐿subscript𝑄𝐶2Q_{L}=Q_{C}/2italic_Q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT = italic_Q start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT / 2 such that Rrf=L0Z0CPsubscript𝑅𝑟𝑓subscript𝐿0subscript𝑍0subscript𝐶𝑃R_{rf}=\frac{L_{0}}{Z_{0}C_{P}}italic_R start_POSTSUBSCRIPT italic_r italic_f end_POSTSUBSCRIPT = divide start_ARG italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_Z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_C start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG, the impedance matching criterion we used to design our circuits. In Fig. S6, we show this analysis as a function of T𝑇Titalic_T and ν𝜈\nuitalic_ν. Rrf(ν,T)subscript𝑅𝑟𝑓𝜈𝑇R_{rf}(\nu,T)italic_R start_POSTSUBSCRIPT italic_r italic_f end_POSTSUBSCRIPT ( italic_ν , italic_T ) [Fig. S6(e)] shows good qualitative agreement with both R2tsubscript𝑅2𝑡R_{2t}italic_R start_POSTSUBSCRIPT 2 italic_t end_POSTSUBSCRIPT and R4tsubscript𝑅4𝑡R_{4t}italic_R start_POSTSUBSCRIPT 4 italic_t end_POSTSUBSCRIPT [Figs. S6(c,d)]. In the superconducting state, when R4t=0subscript𝑅4𝑡0R_{4t}=0italic_R start_POSTSUBSCRIPT 4 italic_t end_POSTSUBSCRIPT = 0, we obtain a quantitative match between Rrfsubscript𝑅𝑟𝑓R_{rf}italic_R start_POSTSUBSCRIPT italic_r italic_f end_POSTSUBSCRIPT and R2tsubscript𝑅2𝑡R_{2t}italic_R start_POSTSUBSCRIPT 2 italic_t end_POSTSUBSCRIPT where RrfR2tRcsimilar-to-or-equalssubscript𝑅𝑟𝑓subscript𝑅2𝑡similar-to-or-equalssubscript𝑅𝑐R_{rf}\simeq R_{2t}\simeq R_{c}italic_R start_POSTSUBSCRIPT italic_r italic_f end_POSTSUBSCRIPT ≃ italic_R start_POSTSUBSCRIPT 2 italic_t end_POSTSUBSCRIPT ≃ italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT arises purely from the contact resistance Rc3.2similar-to-or-equalssubscript𝑅𝑐3.2R_{c}\simeq 3.2italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≃ 3.2 kΩΩ\Omegaroman_Ω.

1.5 Superfluid stiffness phenomenology and main equations

Here we derive the formulas for superfluid stiffness of a nodal superconductor, used in the main text. For the temperature and current dependence, we focus on the effects of quasi-particles near the nodes. We note that the effects of classical phase fluctuations can also produce a linear in T𝑇Titalic_T suppression of ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. However, these are expected to be suppressed below TQR/Lsimilar-tosubscript𝑇𝑄Planck-constant-over-2-pi𝑅𝐿T_{Q}\sim\hbar R/Litalic_T start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ∼ roman_ℏ italic_R / italic_L, which for our parameters (R1similar-to𝑅1R\sim 1italic_R ∼ 1 kΩΩ\Omegaroman_Ω, L100similar-to𝐿100L\sim 100italic_L ∼ 100 nH) yields about 500500500500 mK [43].

We begin with a mean-field model for superconducting pairing in the active hole bands of TBG. There are two doped hole bands on top of the ν=2𝜈2\nu=-2italic_ν = - 2 state. We assume these to be described by a single spin-degenerate band with singlet pairing, for simplicity and concreteness, though the formalism can be adapted to more general scenarios as well.

𝐤ξ(𝐤)(c𝐤,c𝐤,+c𝐤,c𝐤,)+Δ(𝐤)(c𝐤+𝐐/2,c𝐤+𝐐/2,+c𝐤+𝐐/2,c𝐤+𝐐/2,)=𝐤[c𝐤+𝐐/2,c𝐤+𝐐/2,][ξ(𝐤+𝐐/2)Δ(𝐤)Δ(𝐤)ξ(𝐤+𝐐/2)][c𝐤+𝐐/2,c𝐤+𝐐/2,]++𝐤ξ(𝐤+𝐐/2),subscript𝐤𝜉𝐤subscriptsuperscript𝑐𝐤subscript𝑐𝐤subscriptsuperscript𝑐𝐤subscript𝑐𝐤Δ𝐤subscriptsuperscript𝑐𝐤𝐐2subscriptsuperscript𝑐𝐤𝐐2subscript𝑐𝐤𝐐2subscript𝑐𝐤𝐐2subscript𝐤superscriptmatrixsubscript𝑐𝐤𝐐2superscriptsubscript𝑐𝐤𝐐2matrix𝜉𝐤𝐐2Δ𝐤Δ𝐤𝜉𝐤𝐐2matrixsubscript𝑐𝐤𝐐2superscriptsubscript𝑐𝐤𝐐2subscript𝐤𝜉𝐤𝐐2\begin{gathered}\sum_{\bf k}\xi({\bf k})(c^{\dagger}_{{\bf k},\uparrow}c_{{\bf k% },\uparrow}+c^{\dagger}_{{\bf k},\downarrow}c_{{\bf k},\downarrow})+\Delta({% \bf k})(c^{\dagger}_{{\bf k}+{\bf Q}/2,\uparrow}c^{\dagger}_{-{\bf k}+{\bf Q}/% 2,\downarrow}+c_{-{\bf k}+{\bf Q}/2,\downarrow}c_{{\bf k}+{\bf Q}/2,\uparrow})% =\\ \sum_{\bf k}\begin{bmatrix}c_{{\bf k}+{\bf Q}/2,\uparrow}\\ c_{-{\bf k}+{\bf Q}/2,\downarrow}^{\dagger}\end{bmatrix}^{\dagger}\begin{% bmatrix}\xi({\bf k}+{\bf Q}/2)&\Delta({\bf k})\\ \Delta({\bf k})&-\xi(-{\bf k}+{\bf Q}/2)\end{bmatrix}\begin{bmatrix}c_{{\bf k}% +{\bf Q}/2,\uparrow}\\ c_{-{\bf k}+{\bf Q}/2,\downarrow}^{\dagger}\end{bmatrix}+\\ +\sum_{\bf k}\xi(-{\bf k}+{\bf Q}/2),\end{gathered}start_ROW start_CELL ∑ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT italic_ξ ( bold_k ) ( italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k , ↑ end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT bold_k , ↑ end_POSTSUBSCRIPT + italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k , ↓ end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT bold_k , ↓ end_POSTSUBSCRIPT ) + roman_Δ ( bold_k ) ( italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k + bold_Q / 2 , ↑ end_POSTSUBSCRIPT italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - bold_k + bold_Q / 2 , ↓ end_POSTSUBSCRIPT + italic_c start_POSTSUBSCRIPT - bold_k + bold_Q / 2 , ↓ end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT bold_k + bold_Q / 2 , ↑ end_POSTSUBSCRIPT ) = end_CELL end_ROW start_ROW start_CELL ∑ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT [ start_ARG start_ROW start_CELL italic_c start_POSTSUBSCRIPT bold_k + bold_Q / 2 , ↑ end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL italic_c start_POSTSUBSCRIPT - bold_k + bold_Q / 2 , ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT end_CELL end_ROW end_ARG ] start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT [ start_ARG start_ROW start_CELL italic_ξ ( bold_k + bold_Q / 2 ) end_CELL start_CELL roman_Δ ( bold_k ) end_CELL end_ROW start_ROW start_CELL roman_Δ ( bold_k ) end_CELL start_CELL - italic_ξ ( - bold_k + bold_Q / 2 ) end_CELL end_ROW end_ARG ] [ start_ARG start_ROW start_CELL italic_c start_POSTSUBSCRIPT bold_k + bold_Q / 2 , ↑ end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL italic_c start_POSTSUBSCRIPT - bold_k + bold_Q / 2 , ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT end_CELL end_ROW end_ARG ] + end_CELL end_ROW start_ROW start_CELL + ∑ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT italic_ξ ( - bold_k + bold_Q / 2 ) , end_CELL end_ROW (6)

where ξ(𝐤)=ε(𝐤)μ𝜉𝐤𝜀𝐤𝜇\xi({\bf k})=\varepsilon({\bf k})-\muitalic_ξ ( bold_k ) = italic_ε ( bold_k ) - italic_μ, where ε(𝐤)𝜀𝐤\varepsilon({\bf k})italic_ε ( bold_k ) - is the single-particle dispersion. The momentum 𝐐=𝐪𝟎+𝐪𝐐subscript𝐪0𝐪\bf{Q}=\bf{q}_{0}+\bf{q}bold_Q = bold_q start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT + bold_q is the momentum of the Cooper pair and 𝐪𝟎subscript𝐪0\bf{q}_{0}bold_q start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT is the static current of the system.

While the last piece of (6) is a c-number it still contributes to the total energy of the system and needs to be included to avoid convergence issues in the final expressions. Note that, in general, Δ(𝒌)Δ𝒌\Delta({\boldsymbol{k}})roman_Δ ( bold_italic_k ) can be dependent on 𝐐𝐐\bf{Q}bold_Q as well, though we will ignore this dependence for simplicity. This assumption corresponds to neglecting the geometric contribution to the stiffness[44, 45, 46]; see e.g. Appendix A2 of Ref. [46] for a discussion in the projected-context. We neglect the geometric stiffness here because in generic models it depends on self-consistent energetics, and thus a microscopic model for the pairing, in addition to the active band wavefunctions [45]. In the case of the TBG bands we expect the geometric contribution to have the same qualitative dependence on do** as the kinetic contribution, as the dominant kinetic contribution is itself “geometric”: it arises from interactions and the 𝒌𝒌{\boldsymbol{k}}bold_italic_k dependence of single-particle wavefunctions. We will discuss this point further in Supplementary materials.

The superfluid stiffness can be obtained by expanding the grand potential in 𝐪𝐪{\bf q}bold_q: [ρs]αβ=1V2Ωqαqβ|𝐪=0subscriptdelimited-[]subscript𝜌𝑠𝛼𝛽evaluated-at1𝑉superscriptPlanck-constant-over-2-pi2Ωsubscript𝑞𝛼subscript𝑞𝛽𝐪0[\rho_{s}]_{\alpha\beta}=\left.\frac{1}{V\hbar^{2}}\frac{\partial\Omega}{% \partial q_{\alpha}\partial q_{\beta}}\right|_{{\bf q}=0}[ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ] start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_V roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG divide start_ARG ∂ roman_Ω end_ARG start_ARG ∂ italic_q start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ∂ italic_q start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT end_ARG | start_POSTSUBSCRIPT bold_q = 0 end_POSTSUBSCRIPT [44, 45] or calculating the current response. We note two convention differences from prior works: a different definition of Cooper pair momentum, as well as the computation of the superfluid stiffness instead of the superfluid weight Ds=4ρssubscript𝐷𝑠4subscript𝜌𝑠D_{s}=4\rho_{s}italic_D start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 4 italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. These differences correspond to equal and opposite factors of four in the previous expression.

From Eq. (6) we obtain:

ΩTωn,𝐤Trlog[iωnΔ(𝐤)τ^1{ξ(k^0)τ^3+αξ(k^0)qα/2+αβξ(k^0)τ^3qαqβ/8}]+𝐤ξ(𝐤+𝐪0/2)+αξ(𝐤+𝐪0/2)qα/2+αβξ(𝐤+𝐪0/2)qαqβ/8,k^0=𝐤+τ^3𝐪0/2.formulae-sequenceΩ𝑇subscriptsubscript𝜔𝑛𝐤Tr𝑖subscript𝜔𝑛Δ𝐤subscript^𝜏1𝜉subscript^𝑘0subscript^𝜏3subscript𝛼𝜉subscript^𝑘0subscript𝑞𝛼2subscript𝛼subscript𝛽𝜉subscript^𝑘0subscript^𝜏3subscript𝑞𝛼subscript𝑞𝛽8subscript𝐤𝜉𝐤subscript𝐪02subscript𝛼𝜉𝐤subscript𝐪02subscript𝑞𝛼2subscript𝛼subscript𝛽𝜉𝐤subscript𝐪02subscript𝑞𝛼subscript𝑞𝛽8subscript^𝑘0𝐤subscript^𝜏3subscript𝐪02\begin{gathered}\Omega\approx-T\sum_{\omega_{n},{\bf k}}{\rm Tr}\log\Big{[}i% \omega_{n}-\Delta({\bf k})\hat{\tau}_{1}\\ -\{\xi(\hat{k}_{0})\hat{\tau}_{3}+\partial_{\alpha}\xi(\hat{k}_{0})q_{\alpha}/% 2+\partial_{\alpha}\partial_{\beta}\xi(\hat{k}_{0})\hat{\tau}_{3}q_{\alpha}q_{% \beta}/8\}\Big{]}\\ +\sum_{\bf k}\xi({\bf k}+{\bf q}_{0}/2)+\partial_{\alpha}\xi({\bf k}+{\bf q}_{% 0}/2)q_{\alpha}/2+\partial_{\alpha}\partial_{\beta}\xi({\bf k}+{\bf q}_{0}/2)q% _{\alpha}q_{\beta}/8,\\ \hat{k}_{0}={\bf k}+\hat{\tau}_{3}{\bf q}_{0}/2.\end{gathered}start_ROW start_CELL roman_Ω ≈ - italic_T ∑ start_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , bold_k end_POSTSUBSCRIPT roman_Tr roman_log [ italic_i italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - roman_Δ ( bold_k ) over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL - { italic_ξ ( over^ start_ARG italic_k end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT + ∂ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ξ ( over^ start_ARG italic_k end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) italic_q start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT / 2 + ∂ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ∂ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT italic_ξ ( over^ start_ARG italic_k end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_q start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_q start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT / 8 } ] end_CELL end_ROW start_ROW start_CELL + ∑ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT italic_ξ ( bold_k + bold_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / 2 ) + ∂ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ξ ( bold_k + bold_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / 2 ) italic_q start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT / 2 + ∂ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ∂ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT italic_ξ ( bold_k + bold_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / 2 ) italic_q start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_q start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT / 8 , end_CELL end_ROW start_ROW start_CELL over^ start_ARG italic_k end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = bold_k + over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT bold_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / 2 . end_CELL end_ROW (7)

where τisubscript𝜏𝑖\tau_{i}italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are matrices in Gor’kov-Nambu space and we assume time reversal symmetry ξ(𝐤)=ξ(𝐤)𝜉𝐤𝜉𝐤\xi(-{\bf k})=\xi({\bf k})italic_ξ ( - bold_k ) = italic_ξ ( bold_k ).

The linear in 𝐪𝐪{\bf q}bold_q terms vanish after angle integration for 𝐪0=0subscript𝐪00{\bf q}_{0}=0bold_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0 but are nonzero otherwise, corresponding to a finite supercurrent. While ξ(𝐤)𝜉𝐤\xi({\bf k})italic_ξ ( bold_k ) vanishes at the Fermi surface, αξ(𝐤)subscript𝛼𝜉𝐤\partial_{\alpha}\xi({\bf k})∂ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ξ ( bold_k ) and αβξ(𝐤)subscript𝛼subscript𝛽𝜉𝐤\partial_{\alpha}\partial_{\beta}\xi({\bf k})∂ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ∂ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT italic_ξ ( bold_k ) in general do not. This allows us to take 𝐤±𝐪0/2𝐤plus-or-minus𝐤subscript𝐪02𝐤{\bf k}\pm{\bf q}_{0}/2\approx{\bf k}bold_k ± bold_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / 2 ≈ bold_k in those derivatives and take ξ(𝐤±𝐪0/2)ξ(𝐤)±𝐯𝐪0/2𝜉plus-or-minus𝐤subscript𝐪02plus-or-minus𝜉𝐤subscript𝐯𝐪02\xi({\bf k}\pm{\bf q}_{0}/2)\approx\xi({\bf k})\pm{\bf v}{\bf q}_{0}/2italic_ξ ( bold_k ± bold_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / 2 ) ≈ italic_ξ ( bold_k ) ± bold_vq start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / 2. Then we get:

=αβ[ρsdia]αβ+[ρspara]αβ,[ρsdia]αβ=14𝐤{1+TωnTr[G^(iωn,𝐤)τ^3]}αβξ(𝐤)==14𝐤,σc𝐤,σc𝐤,σαβξ(𝐤)=14𝐤(1ξ(𝐤)tanh[ξ2(𝐤)+|Δ(𝐤)|22T]ξ2(𝐤)+|Δ(𝐤)|2)αβξ(𝐤)[ρspara]αβ=14Tωn,𝐤Tr[G^(iωn,𝐤)αξG^(iωn,𝐤)βξ],\begin{gathered}{}_{\alpha\beta}=[\rho_{s}^{dia}]_{\alpha\beta}+[\rho_{s}^{% para}]_{\alpha\beta},\\ [\rho_{s}^{dia}]_{\alpha\beta}=\frac{1}{4}\sum_{{\bf k}}\left\{1+T\sum_{\omega% _{n}}{\rm Tr}[\hat{G}(i\omega_{n},{\bf k})\hat{\tau}_{3}]\right\}\partial_{% \alpha}\partial_{\beta}\xi({\bf k})=\\ =\frac{1}{4}\sum_{{\bf k},\sigma}\langle c^{\dagger}_{\bf k,\sigma}c_{\bf k,% \sigma}\rangle\partial_{\alpha}\partial_{\beta}\xi({\bf k})=\frac{1}{4}\sum_{% \bf k}\left(1-\frac{\xi({\bf k})\tanh\left[\frac{\sqrt{\xi^{2}({\bf k})+|% \Delta({\bf k})|^{2}}}{2T}\right]}{\sqrt{\xi^{2}({\bf k})+|\Delta({\bf k})|^{2% }}}\right)\partial_{\alpha}\partial_{\beta}\xi({\bf k})\\ [\rho_{s}^{para}]_{\alpha\beta}=\frac{1}{4}T\sum_{\omega_{n},{\bf k}}{\rm Tr}[% \hat{G}(i\omega_{n},{\bf k})\partial_{\alpha}\xi\hat{G}(i\omega_{n},{\bf k})% \partial_{\beta}\xi],\end{gathered}start_ROW start_CELL start_FLOATSUBSCRIPT italic_α italic_β end_FLOATSUBSCRIPT = [ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d italic_i italic_a end_POSTSUPERSCRIPT ] start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT + [ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_p italic_a italic_r italic_a end_POSTSUPERSCRIPT ] start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT , end_CELL end_ROW start_ROW start_CELL [ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d italic_i italic_a end_POSTSUPERSCRIPT ] start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 4 end_ARG ∑ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT { 1 + italic_T ∑ start_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_Tr [ over^ start_ARG italic_G end_ARG ( italic_i italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , bold_k ) over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ] } ∂ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ∂ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT italic_ξ ( bold_k ) = end_CELL end_ROW start_ROW start_CELL = divide start_ARG 1 end_ARG start_ARG 4 end_ARG ∑ start_POSTSUBSCRIPT bold_k , italic_σ end_POSTSUBSCRIPT ⟨ italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k , italic_σ end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT bold_k , italic_σ end_POSTSUBSCRIPT ⟩ ∂ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ∂ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT italic_ξ ( bold_k ) = divide start_ARG 1 end_ARG start_ARG 4 end_ARG ∑ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT ( 1 - divide start_ARG italic_ξ ( bold_k ) roman_tanh [ divide start_ARG square-root start_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_k ) + | roman_Δ ( bold_k ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG start_ARG 2 italic_T end_ARG ] end_ARG start_ARG square-root start_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_k ) + | roman_Δ ( bold_k ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG ) ∂ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ∂ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT italic_ξ ( bold_k ) end_CELL end_ROW start_ROW start_CELL [ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_p italic_a italic_r italic_a end_POSTSUPERSCRIPT ] start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 4 end_ARG italic_T ∑ start_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , bold_k end_POSTSUBSCRIPT roman_Tr [ over^ start_ARG italic_G end_ARG ( italic_i italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , bold_k ) ∂ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ξ over^ start_ARG italic_G end_ARG ( italic_i italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , bold_k ) ∂ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT italic_ξ ] , end_CELL end_ROW (8)

where

G^(iωn,𝐤)=iωn𝐯𝐪0/2+ξτ^3+Δ(𝐤)τ^1[ωn+i(𝐯𝐪0)/2]2+ξ2+Δ2(𝐤)^𝐺𝑖subscript𝜔𝑛𝐤𝑖subscript𝜔𝑛𝐯subscript𝐪02𝜉subscript^𝜏3Δ𝐤subscript^𝜏1superscriptdelimited-[]subscript𝜔𝑛𝑖𝐯subscript𝐪022superscript𝜉2superscriptΔ2𝐤\hat{G}(i\omega_{n},{\bf k})=-\frac{i\omega_{n}-{\bf v}\cdot{\bf q}_{0}/2+\xi% \hat{\tau}_{3}+\Delta({\bf k})\hat{\tau}_{1}}{[\omega_{n}+i({\bf v}\cdot{\bf q% }_{0})/2]^{2}+\xi^{2}+\Delta^{2}({\bf k})}over^ start_ARG italic_G end_ARG ( italic_i italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , bold_k ) = - divide start_ARG italic_i italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - bold_v ⋅ bold_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / 2 + italic_ξ over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT + roman_Δ ( bold_k ) over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG [ italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + italic_i ( bold_v ⋅ bold_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / 2 ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_k ) end_ARG (9)

For αβξ(𝐤)=constsubscript𝛼subscript𝛽𝜉𝐤𝑐𝑜𝑛𝑠𝑡\partial_{\alpha}\partial_{\beta}\xi({\bf k})=const∂ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ∂ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT italic_ξ ( bold_k ) = italic_c italic_o italic_n italic_s italic_t (such as for parabolic band), ρsdiasuperscriptsubscript𝜌𝑠𝑑𝑖𝑎\rho_{s}^{dia}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d italic_i italic_a end_POSTSUPERSCRIPT is simply proportional to the electron density, which is temperature-independent. Ignoring corrections to the chemical potential (assuming TEFmuch-less-than𝑇subscript𝐸𝐹T\ll E_{F}italic_T ≪ italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT), corrections to ρsdiasuperscriptsubscript𝜌𝑠𝑑𝑖𝑎\rho_{s}^{dia}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d italic_i italic_a end_POSTSUPERSCRIPT can arise from the vicinity of the nodes. However, the leading term will arise from the expansion αβξ(𝐤)a0+a1ξsubscript𝛼subscript𝛽𝜉𝐤subscript𝑎0subscript𝑎1𝜉\partial_{\alpha}\partial_{\beta}\xi({\bf k})\approx a_{0}+a_{1}\xi∂ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ∂ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT italic_ξ ( bold_k ) ≈ italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ξ, so that the integral over momentum will take the form 𝑑ξ𝑑Δexp[ξ2+Δ2/2T]a1ξ2/ξ2+Δ2differential-d𝜉differential-dΔsuperscript𝜉2superscriptΔ22𝑇subscript𝑎1superscript𝜉2superscript𝜉2superscriptΔ2\int d\xi d\Delta\exp[-\sqrt{\xi^{2}+\Delta^{2}}/2T]a_{1}\xi^{2}/\sqrt{\xi^{2}% +\Delta^{2}}∫ italic_d italic_ξ italic_d roman_Δ roman_exp [ - square-root start_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG / 2 italic_T ] italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / square-root start_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG and will scale as T3proportional-toabsentsuperscript𝑇3\propto T^{3}∝ italic_T start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT in temperature.

We thus focus on ρsparasuperscriptsubscript𝜌𝑠𝑝𝑎𝑟𝑎\rho_{s}^{para}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_p italic_a italic_r italic_a end_POSTSUPERSCRIPT and its temperature dependence. Evaluating the trace we get:

=αβT2ωn,𝐤vαvβ[ωn+i(𝐯𝐪0)/2]2+ξ2+Δ2(𝐤)([ωn+i(𝐯𝐪0)/2]2+ξ2+Δ2(𝐤))2=116T𝐤vαvβ(1cosh2ξ2+Δ2(𝐤)+(𝐯𝐪0)/22T+1cosh2ξ2+Δ2(𝐤)(𝐯𝐪0)/22T).\begin{gathered}{}_{\alpha\beta}=\frac{T}{2}\sum_{\omega_{n},{\bf k}}v_{\alpha% }v_{\beta}\frac{-[\omega_{n}+i({\bf v}\cdot{\bf q}_{0})/2]^{2}+\xi^{2}+\Delta^% {2}({\bf k})}{([\omega_{n}+i({\bf v}\cdot{\bf q}_{0})/2]^{2}+\xi^{2}+\Delta^{2% }({\bf k}))^{2}}\\ =-\frac{1}{16T}\sum_{\bf k}v_{\alpha}v_{\beta}\left(\frac{1}{\cosh^{2}\frac{% \sqrt{\xi^{2}+\Delta^{2}({\bf k})}+({\bf v}\cdot{\bf q}_{0})/2}{2T}}+\frac{1}{% \cosh^{2}\frac{\sqrt{\xi^{2}+\Delta^{2}({\bf k})}-({\bf v}\cdot{\bf q}_{0})/2}% {2T}}\right).\end{gathered}start_ROW start_CELL start_FLOATSUBSCRIPT italic_α italic_β end_FLOATSUBSCRIPT = divide start_ARG italic_T end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , bold_k end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT divide start_ARG - [ italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + italic_i ( bold_v ⋅ bold_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / 2 ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_k ) end_ARG start_ARG ( [ italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + italic_i ( bold_v ⋅ bold_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / 2 ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_k ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_CELL end_ROW start_ROW start_CELL = - divide start_ARG 1 end_ARG start_ARG 16 italic_T end_ARG ∑ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT ( divide start_ARG 1 end_ARG start_ARG roman_cosh start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG square-root start_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_k ) end_ARG + ( bold_v ⋅ bold_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / 2 end_ARG start_ARG 2 italic_T end_ARG end_ARG + divide start_ARG 1 end_ARG start_ARG roman_cosh start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG square-root start_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_k ) end_ARG - ( bold_v ⋅ bold_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / 2 end_ARG start_ARG 2 italic_T end_ARG end_ARG ) . end_CELL end_ROW (10)

Since the contribution is exponentially suppressed by the gap, we focus on the vicinity of the nodes, where Δ(𝐤)𝐯Δ(𝐤𝐤F)Δ𝐤subscript𝐯Δ𝐤subscript𝐤𝐹\Delta({\bf k})\approx{\bf v}_{\Delta}\cdot({\bf k}-{\bf k}_{F})roman_Δ ( bold_k ) ≈ bold_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT ⋅ ( bold_k - bold_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) and ξ𝐯(𝐤𝐤F)𝜉𝐯𝐤subscript𝐤𝐹\xi\approx{\bf v}\cdot({\bf k}-{\bf k}_{F})italic_ξ ≈ bold_v ⋅ ( bold_k - bold_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ).

δρs(T,I)/1 node/=[ρspara]αβnvnαvnβ/(vnvΔ)16Tsinθv,vΔdξdΔ(2π)2(1cosh2ξ2+Δ2+(𝐯n𝐪0)/22T+1cosh2ξ2+Δ2(𝐤)(𝐯n𝐪0)/22T)==nvnαvnβ/(vnvΔ)4πsinθv,vΔTlog[2cosh(𝐯n𝐪0)4T],\begin{gathered}\delta\rho_{s}(T,I)/\text{1 node}/=[\rho_{s}^{para}]_{\alpha% \beta}\approx\\ \approx\sum_{n}-\frac{v_{n}^{\alpha}v_{n}^{\beta}/(v_{n}v_{\Delta})}{16T\sin% \theta_{v,v_{\Delta}}}\int\frac{d\xi d\Delta}{(2\pi)^{2}}\left(\frac{1}{\cosh^% {2}\frac{\sqrt{\xi^{2}+\Delta^{2}}+({\bf v}_{n}\cdot{\bf q}_{0})/2}{2T}}+\frac% {1}{\cosh^{2}\frac{\sqrt{\xi^{2}+\Delta^{2}({\bf k})}-({\bf v}_{n}\cdot{\bf q}% _{0})/2}{2T}}\right)=\\ =-\sum_{n}\frac{v_{n}^{\alpha}v_{n}^{\beta}/(v_{n}v_{\Delta})}{4\pi\sin\theta_% {v,v_{\Delta}}}T\log\left[2\cosh\frac{({\bf v}_{n}\cdot{\bf q}_{0})}{4T}\right% ],\end{gathered}start_ROW start_CELL italic_δ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T , italic_I ) / 1 node / = [ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_p italic_a italic_r italic_a end_POSTSUPERSCRIPT ] start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT ≈ end_CELL end_ROW start_ROW start_CELL ≈ ∑ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - divide start_ARG italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT / ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT ) end_ARG start_ARG 16 italic_T roman_sin italic_θ start_POSTSUBSCRIPT italic_v , italic_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG ∫ divide start_ARG italic_d italic_ξ italic_d roman_Δ end_ARG start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( divide start_ARG 1 end_ARG start_ARG roman_cosh start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG square-root start_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + ( bold_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⋅ bold_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / 2 end_ARG start_ARG 2 italic_T end_ARG end_ARG + divide start_ARG 1 end_ARG start_ARG roman_cosh start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG square-root start_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_k ) end_ARG - ( bold_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⋅ bold_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / 2 end_ARG start_ARG 2 italic_T end_ARG end_ARG ) = end_CELL end_ROW start_ROW start_CELL = - ∑ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT divide start_ARG italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT / ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT ) end_ARG start_ARG 4 italic_π roman_sin italic_θ start_POSTSUBSCRIPT italic_v , italic_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG italic_T roman_log [ 2 roman_cosh divide start_ARG ( bold_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⋅ bold_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) end_ARG start_ARG 4 italic_T end_ARG ] , end_CELL end_ROW (11)

where 𝐪0𝐈proportional-tosubscript𝐪0𝐈{\bf q}_{0}\propto{\bf I}bold_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∝ bold_I (see Eq. S2). Thus, at zero current, one gets δρsTproportional-to𝛿subscript𝜌𝑠𝑇\delta\rho_{s}\propto-Titalic_δ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ∝ - italic_T, while at (𝐯𝐪0)Tmuch-greater-than𝐯subscript𝐪0𝑇({\bf v}\cdot{\bf q}_{0})\gg T( bold_v ⋅ bold_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ≫ italic_T one gets δρsIproportional-to𝛿subscript𝜌𝑠𝐼\delta\rho_{s}\propto-Iitalic_δ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ∝ - italic_I. Note that expression 11 can be modified by Fermi-liquid effects; these result in an additional multiplicative factor in the above formula [3]; in cuprates the value of this factor was found to be close to 1 [47, 48].

The results simplify considerably for: (a) symmetry-protected nodes, i.e. 𝐯𝐯Δperpendicular-to𝐯subscript𝐯Δ{\bf v}\perp{\bf v}_{\Delta}bold_v ⟂ bold_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT (b) non-nematic SC state, i.e. ραβδαβproportional-tosubscript𝜌𝛼𝛽subscript𝛿𝛼𝛽\rho_{\alpha\beta}\propto\delta_{\alpha\beta}italic_ρ start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT ∝ italic_δ start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT. From (b) it follows that n(vnα)2=Nv2/2subscript𝑛superscriptsuperscriptsubscript𝑣𝑛𝛼2𝑁superscript𝑣22\sum_{n}(v_{n}^{\alpha})^{2}=Nv^{2}/2∑ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = italic_N italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2, where N𝑁Nitalic_N is the number of nodes. For N=4𝑁4N=4italic_N = 4 (d-wave) we recover δρs(T,q0=0)=vlog22πvΔT𝛿subscript𝜌𝑠𝑇subscript𝑞00𝑣22𝜋subscript𝑣Δ𝑇\delta\rho_{s}(T,q_{0}=0)=-\frac{v\log 2}{2\pi v_{\Delta}}Titalic_δ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T , italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0 ) = - divide start_ARG italic_v roman_log 2 end_ARG start_ARG 2 italic_π italic_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT end_ARG italic_T[1].

Let us now estimate how the current-induced correction depends on temperature. We use q0=I2eρs0Wsubscript𝑞0𝐼2𝑒superscriptsubscript𝜌𝑠0𝑊q_{0}=\frac{I}{2e\rho_{s}^{0}W}italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = divide start_ARG italic_I end_ARG start_ARG 2 italic_e italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT italic_W end_ARG from Eq. S2 and expand (11) using the assumptions (a) and (b). We further compute the ρsTrραβ/2subscript𝜌𝑠Trsubscript𝜌𝛼𝛽2\rho_{s}\approx{\rm Tr}\rho_{\alpha\beta}/2italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ≈ roman_Tr italic_ρ start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT / 2 to average over current-induced anisotropies. The result is

δρs(Tvq0)=Nv8πvΔT(log2+v2I216(4eρs0W)2T2)𝛿subscript𝜌𝑠much-greater-than𝑇𝑣subscript𝑞0𝑁𝑣8𝜋subscript𝑣Δ𝑇2superscript𝑣2superscript𝐼216superscript4𝑒superscriptsubscript𝜌𝑠0𝑊2superscript𝑇2\begin{gathered}\delta\rho_{s}(T\gg vq_{0})=-\frac{Nv}{8\pi v_{\Delta}}T\left(% \log 2+\frac{v^{2}I^{2}}{16(4e\rho_{s}^{0}W)^{2}T^{2}}\right)\end{gathered}start_ROW start_CELL italic_δ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ≫ italic_v italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) = - divide start_ARG italic_N italic_v end_ARG start_ARG 8 italic_π italic_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT end_ARG italic_T ( roman_log 2 + divide start_ARG italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_I start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 16 ( 4 italic_e italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT italic_W ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) end_CELL end_ROW (12)

Now let’s express δρs𝛿subscript𝜌𝑠\delta\rho_{s}italic_δ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT at large current (compared to the effects of temperature):

δρs(vq0T)=Nv264vΔπIeρs0WcosθI,vn𝛿subscript𝜌𝑠much-greater-than𝑣subscript𝑞0𝑇𝑁superscript𝑣264subscript𝑣Δ𝜋𝐼𝑒superscriptsubscript𝜌𝑠0𝑊delimited-⟨⟩subscript𝜃𝐼subscript𝑣𝑛\delta\rho_{s}(vq_{0}\gg T)=\frac{Nv^{2}}{64v_{\Delta}\pi}\frac{I}{e\rho_{s}^{% 0}W}\langle\cos\theta_{I,v_{n}}\rangleitalic_δ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_v italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≫ italic_T ) = divide start_ARG italic_N italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 64 italic_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT italic_π end_ARG divide start_ARG italic_I end_ARG start_ARG italic_e italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT italic_W end_ARG ⟨ roman_cos italic_θ start_POSTSUBSCRIPT italic_I , italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_POSTSUBSCRIPT ⟩ (13)
\bmhead

Acknowledgements We thank Steve Kivelson, Erez Berg, Amir Yacoby, and Marie Wesson for helpful discussions. The major experimental work is supported by ARO MURI (W911NF-21-2-0147). P.K. and A.B. acknowledge support from the DOE (DE-SC0012260). AV, PL and PAV are supported by a Simons Investigator grant (AV) and by NSF-DMR 2220703. PAV acknowledges support by a Quantum-CT Quantum Regional Partnership Investments (QRPI) Award. K.W. and T.T. acknowledge support from the JSPS KAKENHI (Grant Numbers 21H05233 and 23H02052) and World Premier International Research Center Initiative (WPI), MEXT, Japan. Nanofabrication was performed at the Center for Nanoscale Systems at Harvard, supported in part by an NSF NNIN award ECS- 00335765.

Refer to caption
Figure 1: Experimental setup and device characterization: a, Schematic of the radio-frequency (rf) reflectometry setup. An LC matching network consisting of L0subscript𝐿0L_{0}italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and CPsubscript𝐶𝑃C_{P}italic_C start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT transforms the device impedance 1101101-101 - 10kΩΩ\Omegaroman_Ω to the 50 ΩΩ\Omegaroman_Ω characteristic impedance of the RF measurement circuit. b, Optical microscope image of the twisted-trilayer graphene device reported in the main text and a simple schematic of the measurement scheme. c, DC resistance R𝑅Ritalic_R of the device as a function of moiré filling factor ν𝜈\nuitalic_ν. The sample is superconducting (R=0𝑅0R=0italic_R = 0) in both the electron and hole-doped sectors. Insets show the amplitude and phase response of the complex reflection co-efficient S21subscript𝑆21S_{21}italic_S start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT showing shifts in resonance frequency between the normal and superconducting states. d, 2D map of the normalized amplitude |S21norm|subscriptsuperscript𝑆norm21|S^{\text{norm}}_{21}|| italic_S start_POSTSUPERSCRIPT norm end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT | shown as a function of frequency f𝑓fitalic_f and filling factor ν𝜈\nuitalic_ν. For clearer visualization, for each ν𝜈\nuitalic_ν, we define its |S21norm(f)|=1(max(|S21(f)|)|S21(f)|)/0.9subscriptsuperscript𝑆norm21𝑓1maxsubscript𝑆21𝑓subscript𝑆21𝑓0.9|S^{\text{norm}}_{21}(f)|=1-(\mathrm{max}(|S_{21}(f)|)-|S_{21}(f)|)/0.9| italic_S start_POSTSUPERSCRIPT norm end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT ( italic_f ) | = 1 - ( roman_max ( | italic_S start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT ( italic_f ) | ) - | italic_S start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT ( italic_f ) | ) / 0.9. The resonator undergoes large frequency shifts Δfr1015similar-to-or-equalsΔsubscript𝑓𝑟1015\Delta f_{r}\simeq 10-15roman_Δ italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ≃ 10 - 15 MHz when the sample changes from normal to superconducting states.
Refer to caption
Figure 2: Temperature and do** dependent superfluid stiffness: Two-dimensional map of the a, DC resistance R𝑅Ritalic_R and b, resonator frequency frsubscript𝑓𝑟f_{r}italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT measured as a function of moiré filling factor ν𝜈\nuitalic_ν and sample temperature T𝑇Titalic_T. c, Superfluid stiffness ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT as a function of T𝑇Titalic_T for a range of filling factors around ν=2.4𝜈2.4\nu=-2.4italic_ν = - 2.4. d, (Left-axis) Zero-temperature superfluid stiffness obtained by linear extrapolation of curves in c, as a function of ν𝜈\nuitalic_ν. (Right axis) Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT obtained from DC resistance measurements as a function of ν𝜈\nuitalic_ν. e, Low-temperature slope dρ/dT𝑑𝜌𝑑𝑇d\rho/dTitalic_d italic_ρ / italic_d italic_T obtained as a function of ν𝜈\nuitalic_ν.
Refer to caption
Figure 3: BKT transition, Uemura’s relation, and nodal pairing symmetry: a, Waterfall plot showing ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT as a function of T𝑇Titalic_T and ν𝜈\nuitalic_ν. To aid visualization, the curves have been interpolated and smoothed with 6-pt moving average along the T𝑇Titalic_T axis. The BKT transition temperature (T0subscript𝑇0T_{0}italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) is obtained from the intersection between the universal BKT plane represented as ρc=2T/πsubscript𝜌𝑐2𝑇𝜋\rho_{c}=2T/\piitalic_ρ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 2 italic_T / italic_π and the experimental stiffness curves. b, Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and T0subscript𝑇0T_{0}italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT plotted against ρs0subscript𝜌𝑠0\rho_{s0}italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT. (Inset) Schematic showing inhomogeneous superconducting paths that may lead to an underestimation of the superfluid stiffness. c, Hartree-Fock renormalized bands in TTG as a function of momentum k𝑘kitalic_k within the mini-Brillouin zone. The dashed lines mark the energy where ν=2,3𝜈23\nu=-2,-3italic_ν = - 2 , - 3 and the curvature k2E(k)=0subscriptsuperscript2𝑘𝐸𝑘0\partial^{2}_{k}E(k)=0∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_E ( italic_k ) = 0 occur. d, Comparing the theoretical estimation of ρs0(ν)subscript𝜌𝑠0𝜈\rho_{s0}(\nu)italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT ( italic_ν ) based on the Hartree-Forck renormalized band structure, with the experimentally obtained ρs0(ν)subscript𝜌𝑠0𝜈\rho_{s0}(\nu)italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT ( italic_ν ), which is also interpolated and smoothed for clarity. Both theory and experiment show a bell-shaped ν𝜈\nuitalic_ν dependence, with a lopsided maximum close to ν=2𝜈2\nu=-2italic_ν = - 2. Theory estimates are larger by ×\times×5 compared to the experimental value. e, vF/vΔsubscript𝑣𝐹subscript𝑣Δv_{F}/v_{\Delta}italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT / italic_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT as a function of ν𝜈\nuitalic_ν obtained from the temperature dependence ρs(T)subscript𝜌𝑠𝑇\rho_{s}(T)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ) and the current-bias dependence ρs(I)subscript𝜌𝑠𝐼\rho_{s}(I)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_I ) (described later). The two methods roughly agree with vF/vΔ220similar-to-or-equalssubscript𝑣𝐹subscript𝑣Δ220v_{F}/v_{\Delta}\simeq 2-20italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT / italic_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT ≃ 2 - 20.
Refer to caption
Figure 4: Nonlinear Meissner effect: a, Two-dimensional map of the DC resistance R𝑅Ritalic_R as a function of current bias I𝐼Iitalic_I and temperature T𝑇Titalic_T close to optimal do** ν=2.34𝜈2.34\nu=-2.34italic_ν = - 2.34 where critical current, Ic0.12similar-to-or-equalssubscript𝐼𝑐0.12I_{c}\simeq 0.12italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≃ 0.12 μμ\mathrm{\upmu}roman_μA is maximum. b, R𝑅Ritalic_R as a function of supercurrent bias I𝐼Iitalic_I and moiré filling factor ν𝜈\nuitalic_ν. c, ρs(I)subscript𝜌𝑠𝐼\rho_{s}(I)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_I ) measured at ν=2.34𝜈2.34\nu=-2.34italic_ν = - 2.34 at different values of sample temperature T𝑇Titalic_T. The curvature of ρs(I)subscript𝜌𝑠𝐼\rho_{s}(I)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_I ) at zero current decays with increasing temperature. d, Superfluid stiffness ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT as a function of I𝐼Iitalic_I for a range of filling factors ν=2.24𝜈2.24\nu=-2.24italic_ν = - 2.24 to ν=2.47𝜈2.47\nu=-2.47italic_ν = - 2.47 (T=30𝑇30T=30italic_T = 30 mK). For a range of filling factors around optimal do**, ρs(I)I2proportional-tosubscript𝜌𝑠𝐼superscript𝐼2\rho_{s}(I)\propto-I^{2}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_I ) ∝ - italic_I start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for I0𝐼0I\to 0italic_I → 0 with a crossover to linear dependence ρs(I)Iproportional-tosubscript𝜌𝑠𝐼𝐼\rho_{s}(I)\propto-Iitalic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_I ) ∝ - italic_I. The linear behavior persists for a large range of I<Ic0.12𝐼subscript𝐼𝑐similar-to-or-equals0.12I<I_{c}\simeq 0.12italic_I < italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≃ 0.12 μμ\mathrm{\upmu}roman_μA around ν=2.34𝜈2.34\nu=-2.34italic_ν = - 2.34. Curves are shifted for clarity. e, Temperature dependence of the curvature of ρs(I)subscript𝜌𝑠𝐼\rho_{s}(I)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_I ) at I=0𝐼0I=0italic_I = 0, b(T)𝑏𝑇b(T)italic_b ( italic_T ). Diverging behavior is observed as T0𝑇0T\to 0italic_T → 0, a strong signature of nodal superconductivity. Base temperature value of f, the quadratic curvature b𝑏bitalic_b, and g, the linear slope, c𝑐citalic_c, as a function of moiré filling factor. h, Fermi-velocity estimated from ρs(I)subscript𝜌𝑠𝐼\rho_{s}(I)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_I ) and ρs(T)subscript𝜌𝑠𝑇\rho_{s}(T)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ), indicating vF0.10.7×105similar-to-or-equalssubscript𝑣𝐹0.10.7superscript105v_{F}\simeq 0.1-0.7\times 10^{5}italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ≃ 0.1 - 0.7 × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT m/s and vΔ0.20.5×104similar-to-or-equalssubscript𝑣Δ0.20.5superscript104v_{\Delta}\simeq 0.2-0.5\times 10^{4}italic_v start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT ≃ 0.2 - 0.5 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT m/s. Solid line shows vFsubscript𝑣𝐹v_{F}italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT calculated using the interpolated and smoothed experimental ρs0(ν)subscript𝜌𝑠0𝜈\rho_{s0}(\nu)italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT ( italic_ν ). The schematics in the insets of (f) and (g) depict the two different regimes of the nonlinear Meissner effect. In (f), I0𝐼0I\to 0italic_I → 0, both temperature and current produce quasiparticle excitations. In (g), I>I𝐼superscript𝐼I>I^{*}italic_I > italic_I start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT, the current-induced effect dominates over temperature. In both cases, the current bias produces a finite population of quasi-particles in one node and quasi-holes in the opposite node.

References

  • \bibcommenthead
  • Lee and Wen [1997] Lee, P.A., Wen, X.-G.: Unusual superconducting state of underdoped cuprates. Physical Review Letters 78(21), 4111 (1997)
  • Emery and Kivelson [1995] Emery, V., Kivelson, S.: Importance of phase fluctuations in superconductors with small superfluid density. Nature 374(6521), 434–437 (1995)
  • Millis et al. [1998] Millis, A.J., Girvin, S.M., Ioffe, L.B., Larkin, A.I.: Anomalous charge dynamics in the superconducting state of underdoped cuprates. Journal of Physics and Chemistry of Solids 59(10), 1742–1744 (1998) https://doi.org/10.1016/S0022-3697(98)00094-8
  • Yip and Sauls [1992] Yip, S., Sauls, J.: Nonlinear Meissner effect in CuO superconductors. Physical Review Letters 69(15), 2264 (1992)
  • Xu et al. [1995] Xu, D., Yip, S., Sauls, J.: Nonlinear Meissner effect in unconventional superconductors. Physical Review B 51(22), 16233 (1995)
  • Hardy et al. [1993] Hardy, W., Bonn, D., Morgan, D., Liang, R., Zhang, K.: Precision measurements of the temperature dependence of λ𝜆\lambdaitalic_λ in YBCO: Strong evidence for nodes in the gap function. Physical Review Letters 70(25), 3999 (1993)
  • Uemura et al. [1989] Uemura, Y., Luke, G., Sternlieb, B., Brewer, J., Carolan, J., Hardy, W., Kadono, R., Kempton, J., Kiefl, R., Kreitzman, S., et al.: Universal correlations between tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and ns/msubscript𝑛𝑠superscript𝑚n_{s}/m^{*}italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_m start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT (carrier density over effective mass) in high-Tc cuprate superconductors. Physical Review Letters 62(19), 2317 (1989)
  • Keimer et al. [2015] Keimer, B., Kivelson, S.A., Norman, M.R., Uchida, S., Zaanen, J.: From quantum matter to high-temperature superconductivity in copper oxides. Nature 518(7538), 179–186 (2015)
  • Cao et al. [2018] Cao, Y., Fatemi, V., Fang, S., Watanabe, K., Taniguchi, T., Kaxiras, E., Jarillo-Herrero, P.: Unconventional superconductivity in magic-angle graphene superlattices. Nature 556(7699), 43–50 (2018)
  • Hao et al. [2021] Hao, Z., Zimmerman, A., Ledwith, P., Khalaf, E., Najafabadi, D.H., Watanabe, K., Taniguchi, T., Vishwanath, A., Kim, P.: Electric field–tunable superconductivity in alternating-twist magic-angle trilayer graphene. Science 371(6534), 1133–1138 (2021)
  • Park et al. [2021] Park, J.M., Cao, Y., Watanabe, K., Taniguchi, T., Jarillo-Herrero, P.: Tunable strongly coupled superconductivity in magic-angle twisted trilayer graphene. Nature 590(7845), 249–255 (2021)
  • Park et al. [2022] Park, J.M., Cao, Y., Xia, L.-Q., Sun, S., Watanabe, K., Taniguchi, T., Jarillo-Herrero, P.: Robust superconductivity in magic-angle multilayer graphene family. Nature Materials 21(8), 877–883 (2022)
  • Zhang et al. [2022] Zhang, Y., Polski, R., Lewandowski, C., Thomson, A., Peng, Y., Choi, Y., Kim, H., Watanabe, K., Taniguchi, T., Alicea, J., et al.: Promotion of superconductivity in magic-angle graphene multilayers. Science 377(6614), 1538–1543 (2022)
  • Cao et al. [2021a] Cao, Y., Rodan-Legrain, D., Park, J.M., Yuan, N.F., Watanabe, K., Taniguchi, T., Fernandes, R.M., Fu, L., Jarillo-Herrero, P.: Nematicity and competing orders in superconducting magic-angle graphene. Science 372(6539), 264–271 (2021)
  • Cao et al. [2021b] Cao, Y., Park, J.M., Watanabe, K., Taniguchi, T., Jarillo-Herrero, P.: Pauli-limit violation and re-entrant superconductivity in moiré graphene. Nature 595(7868), 526–531 (2021)
  • Tian et al. [2023] Tian, H., Gao, X., Zhang, Y., Che, S., Xu, T., Cheung, P., Watanabe, K., Taniguchi, T., Randeria, M., Zhang, F., et al.: Evidence for Dirac flat band superconductivity enabled by quantum geometry. Nature 614(7948), 440–444 (2023)
  • Oh et al. [2021] Oh, M., Nuckolls, K.P., Wong, D., Lee, R.L., Liu, X., Watanabe, K., Taniguchi, T., Yazdani, A.: Evidence for unconventional superconductivity in twisted bilayer graphene. Nature 600(7888), 240–245 (2021)
  • Kim et al. [2022] Kim, H., Choi, Y., Lewandowski, C., Thomson, A., Zhang, Y., Polski, R., Watanabe, K., Taniguchi, T., Alicea, J., Nadj-Perge, S.: Evidence for unconventional superconductivity in twisted trilayer graphene. Nature 606(7914), 494–500 (2022)
  • Schoelkopf et al. [1998] Schoelkopf, R., Wahlgren, P., Kozhevnikov, A., Delsing, P., Prober, D.: The radio-frequency single-electron transistor (RF-SET): A fast and ultrasensitive electrometer. Science 280(5367), 1238–1242 (1998)
  • Reilly et al. [2007] Reilly, D., Marcus, C., Hanson, M., Gossard, A.: Fast single-charge sensing with a rf quantum point contact. Applied Physics Letters 91(16) (2007)
  • Vigneau et al. [2023] Vigneau, F., Fedele, F., Chatterjee, A., Reilly, D., Kuemmeth, F., Gonzalez-Zalba, M.F., Laird, E., Ares, N.: Probing quantum devices with radio-frequency reflectometry. Applied Physics Reviews 10(2) (2023)
  • Crossno et al. [2016] Crossno, J., Shi, J.K., Wang, K., Liu, X., Harzheim, A., Lucas, A., Sachdev, S., Kim, P., Taniguchi, T., Watanabe, K., et al.: Observation of the Dirac fluid and the breakdown of the Wiedemann-Franz law in graphene. Science 351(6277), 1058–1061 (2016)
  • Tinkham [2004] Tinkham, M.: Introduction to Superconductivity. Courier Corporation, ??? (2004)
  • Annunziata et al. [2010] Annunziata, A.J., Santavicca, D.F., Frunzio, L., Catelani, G., Rooks, M.J., Frydman, A., Prober, D.E.: Tunable superconducting nanoinductors. Nanotechnology 21(44), 445202 (2010)
  • Singh et al. [2018] Singh, G., Jouan, A., Benfatto, L., Couëdo, F., Kumar, P., Dogra, A., Budhani, R., Caprara, S., Grilli, M., Lesne, E., et al.: Competition between electron pairing and phase coherence in superconducting interfaces. Nature Communications 9(1), 407 (2018)
  • Phan et al. [2022] Phan, D., Senior, J., Ghazaryan, A., Hatefipour, M., Strickland, W.M., Shabani, J., Serbyn, M., Higginbotham, A.P.: Detecting induced p±ipplus-or-minus𝑝𝑖𝑝p\pm{}ipitalic_p ± italic_i italic_p pairing at the Al-InAs interface with a quantum microwave circuit. Phys. Rev. Lett. 128, 107701 (2022)
  • Božović et al. [2016] Božović, I., He, X., Wu, J., Bollinger, A.: Dependence of the critical temperature in overdoped copper oxides on superfluid density. Nature 536(7616), 309–311 (2016)
  • Khalaf et al. [2019] Khalaf, E., Kruchkov, A.J., Tarnopolsky, G., Vishwanath, A.: Magic angle hierarchy in twisted graphene multilayers. Phys. Rev. B 100, 085109 (2019) https://doi.org/10.1103/PhysRevB.100.085109
  • Probst et al. [2015] Probst, S., Song, F., Bushev, P.A., Ustinov, A.V., Weides, M.: Efficient and robust analysis of complex scattering data under noise in microwave resonators. Review of Scientific Instruments 86(2) (2015)
  • Millis et al. [1998] Millis, A., Girvin, S., Ioffe, L., Larkin, A.: Anomalous charge dynamics in the superconducting state of underdoped cuprates. Journal of Physics and Chemistry of Solids 59(10-12), 1742–1744 (1998)
  • Nelson and Kosterlitz [1977] Nelson, D.R., Kosterlitz, J.: Universal jump in the superfluid density of two-dimensional superfluids. Physical Review Letters 39(19), 1201 (1977)
  • Benfatto et al. [2007] Benfatto, L., Castellani, C., Giamarchi, T.: Kosterlitz-Thouless behavior in layered superconductors: The role of the vortex core energy. Phys. Rev. Lett. 98, 117008 (2007)
  • Hetel et al. [2007] Hetel, I., Lemberger, T.R., Randeria, M.: Quantum critical behaviour in the superfluid density of strongly underdoped ultrathin copper oxide films. Nature Physics 3(10), 700–702 (2007)
  • Turkel et al. [2022] Turkel, S., Swann, J., Zhu, Z., Christos, M., Watanabe, K., Taniguchi, T., Sachdev, S., Scheurer, M.S., Kaxiras, E., Dean, C.R., et al.: Orderly disorder in magic-angle twisted trilayer graphene. Science 376(6589), 193–199 (2022)
  • Homes et al. [2004] Homes, C., Dordevic, S., Strongin, M., Bonn, D., Liang, R., Hardy, W., Komiya, S., Ando, Y., Yu, G., Kaneko, N., et al.: A universal scaling relation in high-temperature superconductors. Nature 430(6999), 539–541 (2004)
  • Dordevic et al. [2013] Dordevic, S., Basov, D., Homes, C.: Do organic and other exotic superconductors fail universal scaling relations? Scientific Reports 3(1), 1713 (2013)
  • Mahmood et al. [2019] Mahmood, F., He, X., Bozovic, I., Armitage, N.P.: Locating the missing superconducting electrons in the overdoped cuprates La2xSrxCuO4subscriptLa2𝑥subscriptSr𝑥subscriptCuO4{\mathrm{La}}_{2-x}{\mathrm{Sr}}_{x}{\mathrm{CuO}}_{4}roman_La start_POSTSUBSCRIPT 2 - italic_x end_POSTSUBSCRIPT roman_Sr start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT roman_CuO start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT. Phys. Rev. Lett. 122, 027003 (2019)
  • Bidinosti et al. [1999] Bidinosti, C.P., Hardy, W.N., Bonn, D.A., Liang, R.: Magnetic field dependence of λ𝜆\lambdaitalic_λ in YBa2Cu3O6.95subscriptYBa2subscriptCu3subscript𝑂6.95{\mathrm{YBa}}_{2}{\mathrm{Cu}}_{3}{O}_{6.95}roman_YBa start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_Cu start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_O start_POSTSUBSCRIPT 6.95 end_POSTSUBSCRIPT: Results as a function of temperature and field orientation. Phys. Rev. Lett. 83, 3277–3280 (1999) https://doi.org/10.1103/PhysRevLett.83.3277
  • Oates et al. [2004] Oates, D.E., Park, S.-H., Koren, G.: Observation of the Nonlinear Meissner effect in YBCO thin films: Evidence for a d𝑑ditalic_d-wave order parameter in the bulk of the cuprate superconductors. Phys. Rev. Lett. 93, 197001 (2004) https://doi.org/10.1103/PhysRevLett.93.197001
  • Wilcox et al. [2022] Wilcox, J., Grant, M., Malone, L., Putzke, C., Kaczorowski, D., Wolf, T., Hardy, F., Meingast, C., Analytis, J., Chu, J.-H., et al.: Observation of the non-linear Meissner effect. Nature communications 13(1), 1201 (2022)
  • Dahm and Scalapino [1996] Dahm, T., Scalapino, D.: Theory of microwave intermodulation in a high-Tc superconducting microstrip resonator. Applied physics letters 69(27), 4248–4250 (1996)
  • Bae et al. [2019] Bae, S., Tan, Y., Zhuravel, A.P., Zhang, L., Zeng, S., Liu, Y., Lograsso, T.A., Ariando, A., Venkatesan, T., Anlage, S.M.: Dielectric resonator method for determining gap symmetry of superconductors through anisotropic nonlinear Meissner effect. Review of Scientific Instruments 90(4) (2019)
  • Turneaure et al. [2000] Turneaure, S.J., Lemberger, T.R., Graybeal, J.M.: Effect of thermal phase fluctuations on the superfluid density of two-dimensional superconducting films. Phys. Rev. Lett. 84, 987–990 (2000) https://doi.org/10.1103/PhysRevLett.84.987
  • Peotta and Törmä [2015] Peotta, S., Törmä, P.: Superfluidity in topologically nontrivial flat bands. Nature Communications 6(1), 8944 (2015)
  • Peotta et al. [2023] Peotta, S., Huhtinen, K.-E., Törmä, P.: Quantum geometry in superfluidity and superconductivity (2023)
  • Verma et al. [2023] Verma, N., Guerci, D., Queiroz, R.: Geometric Stiffness in Interlayer Exciton Condensates (2023)
  • Ioffe and Millis [2002] Ioffe, L., Millis, A.: d-wave superconductivity in doped Mott insulators. Journal of Physics and Chemistry of Solids 63(12), 2259–2268 (2002)
  • Lee et al. [2006] Lee, P.A., Nagaosa, N., Wen, X.-G.: Do** a Mott insulator: Physics of high-temperature superconductivity. Rev. Mod. Phys. 78, 17–85 (2006) https://doi.org/10.1103/RevModPhys.78.17
  • Lee and Wen [1997] Lee, P.A., Wen, X.-G.: Unusual superconducting state of underdoped cuprates. Phys. Rev. Lett. 78, 4111–4114 (1997) https://doi.org/10.1103/PhysRevLett.78.4111
  • Bultinck et al. [2020] Bultinck, N., Khalaf, E., Liu, S., Chatterjee, S., Vishwanath, A., Zaletel, M.P.: Ground state and hidden symmetry of magic-angle graphene at even integer filling. Physical Review X 10(3), 031034 (2020) https://doi.org/10.1103/PhysRevX.10.031034
  • Vafek and Kang [2020] Vafek, O., Kang, J.: Renormalization Group Study of Hidden Symmetry in Twisted Bilayer Graphene with Coulomb Interactions. Physical Review Letters 125(25), 257602 (2020) https://doi.org/10.1103/PhysRevLett.125.257602
  • Lian et al. [2020] Lian, B., Song, Z.-D., Regnault, N., Efetov, D.K., Yazdani, A., Bernevig, B.A.: TBG IV: Exact insulator ground states and phase diagram of twisted bilayer graphene. arXiv preprint arXiv:2009.13530 (2020) arxiv:2009.13530
  • Ledwith et al. [2021] Ledwith, P.J., Khalaf, E., Vishwanath, A.: Strong coupling theory of magic-angle graphene: A pedagogical introduction. Annals of Physics 435, 168646 (2021) https://doi.org/10.1016/j.aop.2021.168646
  • Guinea and Walet [2018] Guinea, F., Walet, N.R.: Electrostatic effects, band distortions, and superconductivity in twisted graphene bilayers. Proceedings of the National Academy of Sciences 115(52), 13174–13179 (2018) https://doi.org/10.1073/pnas.1810947115
  • Carr et al. [2019] Carr, S., Fang, S., Po, H.C., Vishwanath, A., Kaxiras, E.: Derivation of Wannier orbitals and minimal-basis tight-binding Hamiltonians for twisted bilayer graphene: First-principles approach. Physical Review Research 1(3), 033072 (2019) https://doi.org/10.1103/PhysRevResearch.1.033072
  • Rademaker et al. [2019] Rademaker, L., Abanin, D.A., Mellado, P.: Charge smoothening and band flattening due to Hartree corrections in twisted bilayer graphene. Physical Review B 100(20), 205114 (2019) https://doi.org/%****␣Superfluid_stiffness_arxiv.bbl␣Line␣950␣****10.1103/PhysRevB.100.205114
  • Cea and Guinea [2020] Cea, T., Guinea, F.: Band structure and insulating states driven by Coulomb interaction in twisted bilayer graphene. Physical Review B 102(4), 045107 (2020) https://doi.org/10.1103/PhysRevB.102.045107
  • Goodwin et al. [2020] Goodwin, Z.A.H., Vitale, V., Liang, X., Mostofi, A.A., Lischner, J.: Hartree theory calculations of quasiparticle properties in twisted bilayer graphene. Electronic Structure 2(3), 034001 (2020) https://doi.org/10.1088/2516-1075/ab9f94
  • Kang et al. [2021] Kang, J., Bernevig, B.A., Vafek, O.: Cascades between Light and Heavy Fermions in the Normal State of Magic-Angle Twisted Bilayer Graphene. Physical Review Letters 127(26), 266402 (2021) https://doi.org/10.1103/PhysRevLett.127.266402
  • Song and Bernevig [2022] Song, Z.-D., Bernevig, B.A.: Magic-Angle Twisted Bilayer Graphene as a Topological Heavy Fermion Problem. Physical Review Letters 129(4), 047601 (2022) https://doi.org/10.1103/PhysRevLett.129.047601
  • Pierce et al. [2021] Pierce, A.T., Xie, Y., Park, J.M., Khalaf, E., Lee, S.H., Cao, Y., Parker, D.E., Forrester, P.R., Chen, S., Watanabe, K., Taniguchi, T., Vishwanath, A., Jarillo-Herrero, P., Yacoby, A.: Unconventional sequence of correlated Chern insulators in magic-angle twisted bilayer graphene. Nature Physics 17(11), 1210–1215 (2021) https://doi.org/10.1038/s41567-021-01347-4
  • Parker et al. [2021] Parker, D., Ledwith, P., Khalaf, E., Soejima, T., Hauschild, J., Xie, Y., Pierce, A., Zaletel, M.P., Yacoby, A., Vishwanath, A.: Field-tuned and zero-field fractional Chern insulators in magic angle graphene. arXiv:2112.13837 [cond-mat] (2021) arxiv:2112.13837 [cond-mat]
  • Kwan et al. [2021] Kwan, Y.H., Wagner, G., Soejima, T., Zaletel, M.P., Simon, S.H., Parameswaran, S.A., Bultinck, N.: Kekul\’e Spiral Order at All Nonzero Integer Fillings in Twisted Bilayer Graphene. Physical Review X 11(4), 041063 (2021) https://doi.org/10.1103/PhysRevX.11.041063
  • Wagner et al. [2022] Wagner, G., Kwan, Y.H., Bultinck, N., Simon, S.H., Parameswaran, S.A.: Global Phase Diagram of the Normal State of Twisted Bilayer Graphene. Physical Review Letters 128(15), 156401 (2022) https://doi.org/10.1103/PhysRevLett.128.156401
  • Wang et al. [2023] Wang, T., Parker, D.E., Soejima, T., Hauschild, J., Anand, S., Bultinck, N., Zaletel, M.P.: Ground-state order in magic-angle graphene at filling $\ensuremath{\nu}=\ensuremath{-}3$: A full-scale density matrix renormalization group study. Physical Review B 108(23), 235128 (2023) https://doi.org/10.1103/PhysRevB.108.235128
  • Nuckolls et al. [2023] Nuckolls, K.P., Lee, R.L., Oh, M., Wong, D., Soejima, T., Hong, J.P., Călugăru, D., Herzog-Arbeitman, J., Bernevig, B.A., Watanabe, K., Taniguchi, T., Regnault, N., Zaletel, M.P., Yazdani, A.: Quantum textures of the many-body wavefunctions in magic-angle graphene. Nature 620(7974), 525–532 (2023) https://doi.org/10.1038/s41586-023-06226-x
  • Kim et al. [2023] Kim, H., Choi, Y., Lantagne-Hurtubise, É., Lewandowski, C., Thomson, A., Kong, L., Zhou, H., Baum, E., Zhang, Y., Holleis, L., Watanabe, K., Taniguchi, T., Young, A.F., Alicea, J., Nadj-Perge, S.: Imaging inter-valley coherent order in magic-angle twisted trilayer graphene. Nature 623(7989), 942–948 (2023) https://doi.org/10.1038/s41586-023-06663-8

2 Superfluid stiffness of twisted multilayer graphene superconductors: Supplementary Material

2.1 Device DC transport characterization

We characterize the TTG device presented in the main text through 4-terminal DC transport measurements, which show results consistent with our previous study on magic-angle TTG. Fig. S1a shows the 4-terminal resistance R𝑅Ritalic_R as a function of ν𝜈\nuitalic_ν and magnetic field B𝐵Bitalic_B at zero displacement field and a temperature of 2 K. Characteristic of magic-angle TTG, there appear two sets of Landau fans. One set has fans emerging from integer fillings of the TTG flat bands. The sequences in these fans generally have larger slopes in Bν𝐵𝜈B-\nuitalic_B - italic_ν plane, in contrast to the other set that appears at low magnetic fields as ”arc-like” quantum oscillations features. These features emerge from the dispersive Dirac cone sector of the TTG electron bands, in addition to the flat band sector.

These two distinct quantum oscillation features with large and small slopes, respectively, are because the flat and Dirac cone sectors are filled simultaneously and a larger proportion of carriers goes into the flat bands until they are fully filled. We can calculate the moiré unit cell area Amsubscript𝐴𝑚A_{m}italic_A start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT based on the flat band fan sequences, which satisfy BAm/ϕ0=Cν+s𝐵subscript𝐴𝑚subscriptitalic-ϕ0𝐶𝜈𝑠BA_{m}/\phi_{0}=C\nu+sitalic_B italic_A start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT / italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_C italic_ν + italic_s, where ϕ0=e/hsubscriptitalic-ϕ0𝑒\phi_{0}=e/hitalic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_e / italic_h is the magnetic flux quantum, C𝐶Citalic_C is the Chern number (and the slope), and s𝑠sitalic_s is the intercept on the ν𝜈\nuitalic_ν axis from which the fans emerge. Then we can obtain the twist angle θ=1.55𝜃1.55\theta=1.55italic_θ = 1.55° from the relation Am=3a22θsubscript𝐴𝑚3superscript𝑎22𝜃A_{m}=\frac{\sqrt{3}a^{2}}{2\theta}italic_A start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT = divide start_ARG square-root start_ARG 3 end_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_θ end_ARG, where a=0.246𝑎0.246a=0.246italic_a = 0.246 nm is the graphene lattice constant. We have measured the fan diagrams for multiple pairs of contacts in different areas of the sample and they all show similar features and twist angles, indicating the TTG sample is highly uniform. Fig. S1b shows R𝑅Ritalic_R as a function of ν𝜈\nuitalic_ν and displacement field D𝐷Ditalic_D at a dilution fridge base temperature of 30 mK, with a resistive feature at the charge neutrality and superconducting states within 3<ν<23𝜈2-3<\nu<-2- 3 < italic_ν < - 2 on the hole-doped side and 1.5<ν<31.5𝜈31.5<\nu<31.5 < italic_ν < 3 on the electron-doped side.

2.2 Superfluid Stiffness, superfluid Density and kinetic Inductance

Here we write out the general relations between superfluid stiffness ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, superfluid weight Dssubscript𝐷𝑠D_{s}italic_D start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and kinetic inductance. In particular, the change in free energy of a superconductor under current is related to Cooper pair momentum as [48, 44]:

ΔF=SDsQ28=SρsQ22LkinI22,Δ𝐹𝑆subscript𝐷𝑠superscript𝑄28𝑆subscript𝜌𝑠superscript𝑄22subscript𝐿𝑘𝑖𝑛superscript𝐼22\Delta F=SD_{s}\frac{Q^{2}}{8}=S\rho_{s}\frac{Q^{2}}{2}\equiv\frac{L_{kin}I^{2% }}{2},roman_Δ italic_F = italic_S italic_D start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT divide start_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 8 end_ARG = italic_S italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT divide start_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG ≡ divide start_ARG italic_L start_POSTSUBSCRIPT italic_k italic_i italic_n end_POSTSUBSCRIPT italic_I start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG , (S1)

where S𝑆Sitalic_S is the area of the sample and Q𝑄Qitalic_Q is the Cooper pair momentum [note that definition of superfluid weight [47] is 4 times smaller than of superfluid density ρs=Ds/4subscript𝜌𝑠subscript𝐷𝑠4\rho_{s}=D_{s}/4italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = italic_D start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / 4]. We will assume a rectangular sample with S=lW𝑆𝑙𝑊S=lWitalic_S = italic_l italic_W, l𝑙litalic_l being the length (along the current direction) and W𝑊Witalic_W - width (perpendicular to the current).

Current can be obtained from gauge coupling QQ2ecA𝑄𝑄2𝑒𝑐𝐴Q\to Q-\frac{2e}{c}\vec{A}italic_Q → italic_Q - divide start_ARG 2 italic_e end_ARG start_ARG italic_c end_ARG over→ start_ARG italic_A end_ARG and using j2D=cΔFAsubscript𝑗2𝐷𝑐Δ𝐹𝐴j_{2D}=-c\frac{\partial\Delta F}{\partial A}italic_j start_POSTSUBSCRIPT 2 italic_D end_POSTSUBSCRIPT = - italic_c divide start_ARG ∂ roman_Δ italic_F end_ARG start_ARG ∂ italic_A end_ARG [49]:

I=j2DW=2eρsQ,𝐼subscript𝑗2𝐷𝑊2𝑒subscript𝜌𝑠𝑄I=j_{2D}W=2e\rho_{s}Q,italic_I = italic_j start_POSTSUBSCRIPT 2 italic_D end_POSTSUBSCRIPT italic_W = 2 italic_e italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_Q , (S2)

which leads to the relation:

Lkin=l4e2ρsW.subscript𝐿𝑘𝑖𝑛𝑙4superscript𝑒2subscript𝜌𝑠𝑊L_{kin}=\frac{l}{4e^{2}\rho_{s}W}.italic_L start_POSTSUBSCRIPT italic_k italic_i italic_n end_POSTSUBSCRIPT = divide start_ARG italic_l end_ARG start_ARG 4 italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_W end_ARG . (S3)

For the case of a Gallilean invariant superconductor where mCP=2mesubscript𝑚𝐶𝑃2subscript𝑚𝑒m_{CP}=2m_{e}italic_m start_POSTSUBSCRIPT italic_C italic_P end_POSTSUBSCRIPT = 2 italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT one gets Ds=ne/mesubscript𝐷𝑠subscript𝑛𝑒subscript𝑚𝑒D_{s}=n_{e}/m_{e}italic_D start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT / italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT, ρs=ne/(4me)subscript𝜌𝑠subscript𝑛𝑒4subscript𝑚𝑒\rho_{s}=n_{e}/(4m_{e})italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT / ( 4 italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ) and Lkin=lmenee2nWsubscript𝐿𝑘𝑖𝑛𝑙subscript𝑚𝑒subscript𝑛𝑒superscript𝑒2𝑛𝑊L_{kin}=\frac{lm_{e}}{n_{e}e^{2}nW}italic_L start_POSTSUBSCRIPT italic_k italic_i italic_n end_POSTSUBSCRIPT = divide start_ARG italic_l italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_ARG start_ARG italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_n italic_W end_ARG.

2.3 Dispersion on top of correlated insulator

In this section we describe how we calculate the hole dispersion on top of the ν=2𝜈2\nu=-2italic_ν = - 2 correlated insulator. While we will make several assumptions in order to arrive at a simple, easy to compute, dispersion, we will later comment on its essential features and why we expect them to be generic amongst most ν=2𝜈2\nu=-2italic_ν = - 2 states.

We will take the “strong coupling” limit[50, 51, 52, 53], under which the interaction strength is much larger than the bare single-particle dispersion but much smaller than the band gap to the remote bands. This is a good approximation when the bands are flat: close to the magic angle and for not-too-large sample strain. The resulting dispersion we obtain will be entirely generated from interactions and the 𝒌𝒌{\boldsymbol{k}}bold_italic_k-dependence of the single particle wavefunctions (“quantum geometry”). It is convenient to work in the sublattice or Chern basis in this limit, which consists of four bands with C=+1𝐶1C=+1italic_C = + 1 and four bands with C=1𝐶1C=-1italic_C = - 1, each set of which can be labeled by spin and valley. The Chern basis is convenient for Coulomb interactions because the form factors, which determine the projected density operators

Λ𝒒αβ(𝒌)=u𝒌αu𝒌+𝒒βρ𝒒=𝒌c𝒌Λ𝒒(𝒌)c𝒌+𝒒,formulae-sequencesubscriptΛ𝒒𝛼𝛽𝒌delimited-⟨⟩subscript𝑢𝒌𝛼subscript𝑢𝒌𝒒𝛽subscript𝜌𝒒subscript𝒌subscriptsuperscript𝑐𝒌subscriptΛ𝒒𝒌subscript𝑐𝒌𝒒\Lambda_{{\boldsymbol{q}}\alpha\beta}({\boldsymbol{k}})=\langle u_{{% \boldsymbol{k}}\alpha}u_{{\boldsymbol{k}}+{\boldsymbol{q}}\beta}\rangle\quad% \rho_{\boldsymbol{q}}=\sum_{\boldsymbol{k}}c^{\dagger}_{{\boldsymbol{k}}}% \Lambda_{{\boldsymbol{q}}}({\boldsymbol{k}})c_{{\boldsymbol{k}}+{\boldsymbol{q% }}},roman_Λ start_POSTSUBSCRIPT bold_italic_q italic_α italic_β end_POSTSUBSCRIPT ( bold_italic_k ) = ⟨ italic_u start_POSTSUBSCRIPT bold_italic_k italic_α end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT bold_italic_k + bold_italic_q italic_β end_POSTSUBSCRIPT ⟩ italic_ρ start_POSTSUBSCRIPT bold_italic_q end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT roman_Λ start_POSTSUBSCRIPT bold_italic_q end_POSTSUBSCRIPT ( bold_italic_k ) italic_c start_POSTSUBSCRIPT bold_italic_k + bold_italic_q end_POSTSUBSCRIPT , (S4)

are approximately diagonal,

Λ𝒒(𝒌)=(λ𝒒(𝒌)I4×400λ𝒒(𝒌)¯).subscriptΛ𝒒𝒌matrixsubscript𝜆𝒒𝒌subscript𝐼4400¯subscript𝜆𝒒𝒌\Lambda_{{\boldsymbol{q}}}({\boldsymbol{k}})=\begin{pmatrix}\lambda_{{% \boldsymbol{q}}}({\boldsymbol{k}})I_{4\times 4}&0\\ 0&\overline{\lambda_{{\boldsymbol{q}}}({\boldsymbol{k}})}\end{pmatrix}.roman_Λ start_POSTSUBSCRIPT bold_italic_q end_POSTSUBSCRIPT ( bold_italic_k ) = ( start_ARG start_ROW start_CELL italic_λ start_POSTSUBSCRIPT bold_italic_q end_POSTSUBSCRIPT ( bold_italic_k ) italic_I start_POSTSUBSCRIPT 4 × 4 end_POSTSUBSCRIPT end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL over¯ start_ARG italic_λ start_POSTSUBSCRIPT bold_italic_q end_POSTSUBSCRIPT ( bold_italic_k ) end_ARG end_CELL end_ROW end_ARG ) . (S5)

The Hamiltonian is

=12A𝒒V𝒒δρ𝒒δρ𝒒,δρ𝒒=ρ𝒒ρ𝒒CNP,formulae-sequence12𝐴subscript𝒒subscript𝑉𝒒𝛿subscript𝜌𝒒𝛿subscript𝜌𝒒𝛿subscript𝜌𝒒subscript𝜌𝒒superscriptsubscript𝜌𝒒CNP\mathcal{H}=\frac{1}{2A}\sum_{\boldsymbol{q}}V_{\boldsymbol{q}}\delta\rho_{{% \boldsymbol{q}}}\delta\rho_{-{\boldsymbol{q}}},\qquad\delta\rho_{\boldsymbol{q% }}=\rho_{\boldsymbol{q}}-\rho_{\boldsymbol{q}}^{\rm CNP},caligraphic_H = divide start_ARG 1 end_ARG start_ARG 2 italic_A end_ARG ∑ start_POSTSUBSCRIPT bold_italic_q end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT bold_italic_q end_POSTSUBSCRIPT italic_δ italic_ρ start_POSTSUBSCRIPT bold_italic_q end_POSTSUBSCRIPT italic_δ italic_ρ start_POSTSUBSCRIPT - bold_italic_q end_POSTSUBSCRIPT , italic_δ italic_ρ start_POSTSUBSCRIPT bold_italic_q end_POSTSUBSCRIPT = italic_ρ start_POSTSUBSCRIPT bold_italic_q end_POSTSUBSCRIPT - italic_ρ start_POSTSUBSCRIPT bold_italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_CNP end_POSTSUPERSCRIPT , (S6)

where ρ𝒒CNP=12𝑮δ𝒒,𝑮trΛ𝒒(𝒌)superscriptsubscript𝜌𝒒CNP12subscript𝑮subscript𝛿𝒒𝑮trsubscriptΛ𝒒𝒌\rho_{\boldsymbol{q}}^{\rm CNP}=\frac{1}{2}\sum_{{\boldsymbol{G}}}\delta_{{% \boldsymbol{q}},{\boldsymbol{G}}}\mathrm{tr}\,\Lambda_{\boldsymbol{q}}({% \boldsymbol{k}})italic_ρ start_POSTSUBSCRIPT bold_italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_CNP end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT bold_italic_G end_POSTSUBSCRIPT italic_δ start_POSTSUBSCRIPT bold_italic_q , bold_italic_G end_POSTSUBSCRIPT roman_tr roman_Λ start_POSTSUBSCRIPT bold_italic_q end_POSTSUBSCRIPT ( bold_italic_k ) is half the total density of the flat bands and A𝐴Aitalic_A is the sample area. There is a U(4)×U(4)𝑈4𝑈4U(4)\times U(4)italic_U ( 4 ) × italic_U ( 4 ) symmetry of \mathcal{H}caligraphic_H consisting of rotating the bands in each Chern sector into each other, which is broken by off-diagonal terms in the form factors, single particle dispersion, strain, and other corrections such as those from K𝐾Kitalic_K-point phonons.

Exact eigenstates of \mathcal{H}caligraphic_H are generalized quantum Hall ferromagnets, consisting of filling entirely some number of the eight Chern bands. This large manifold of states is split by the aforementioned symmetry breaking perturbations, and the resulting ground state is determined by their competition. In this supplement we will be agnostic about the precise ν=2𝜈2\nu=-2italic_ν = - 2 ground state. Thus we will simply use the symmetric hole dispersion, which is the same for all strong coupling ground states in the U(4)×U(4)𝑈4𝑈4U(4)\times U(4)italic_U ( 4 ) × italic_U ( 4 ) symmetric limit and can be computed exactly. The dispersion hνsubscript𝜈h_{\nu}italic_h start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT for holes at filling ν𝜈\nuitalic_ν is then

hν(𝒌)subscript𝜈𝒌\displaystyle h_{\nu}({\boldsymbol{k}})italic_h start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( bold_italic_k ) =hF(𝒌)+νhH(𝒌),absentsubscript𝐹𝒌𝜈subscript𝐻𝒌\displaystyle=-h_{F}({\boldsymbol{k}})+\nu h_{H}({\boldsymbol{k}}),= - italic_h start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( bold_italic_k ) + italic_ν italic_h start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT ( bold_italic_k ) , (S7)
hF(𝒌)subscript𝐹𝒌\displaystyle h_{F}({\boldsymbol{k}})italic_h start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( bold_italic_k ) =12A𝒒V𝒒|λ𝒒(𝒌)|2,absent12𝐴subscript𝒒subscript𝑉𝒒superscriptsubscript𝜆𝒒𝒌2\displaystyle=\frac{1}{2A}\sum_{{\boldsymbol{q}}}V_{\boldsymbol{q}}|\lambda_{% \boldsymbol{q}}({\boldsymbol{k}})|^{2},= divide start_ARG 1 end_ARG start_ARG 2 italic_A end_ARG ∑ start_POSTSUBSCRIPT bold_italic_q end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT bold_italic_q end_POSTSUBSCRIPT | italic_λ start_POSTSUBSCRIPT bold_italic_q end_POSTSUBSCRIPT ( bold_italic_k ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ,
hH(𝒌)subscript𝐻𝒌\displaystyle h_{H}({\boldsymbol{k}})italic_h start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT ( bold_italic_k ) =1A𝑮V𝑮Λ𝑮(𝒌)𝒌λ𝑮(𝒌),absent1𝐴subscript𝑮subscript𝑉𝑮subscriptΛ𝑮𝒌subscriptsuperscript𝒌subscript𝜆𝑮superscript𝒌\displaystyle=\frac{1}{A}\sum_{{\boldsymbol{G}}}V_{\boldsymbol{G}}\Lambda_{% \boldsymbol{G}}({\boldsymbol{k}})\sum_{{\boldsymbol{k}}^{\prime}}\lambda_{-{% \boldsymbol{G}}}({\boldsymbol{k}}^{\prime}),= divide start_ARG 1 end_ARG start_ARG italic_A end_ARG ∑ start_POSTSUBSCRIPT bold_italic_G end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT bold_italic_G end_POSTSUBSCRIPT roman_Λ start_POSTSUBSCRIPT bold_italic_G end_POSTSUBSCRIPT ( bold_italic_k ) ∑ start_POSTSUBSCRIPT bold_italic_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT - bold_italic_G end_POSTSUBSCRIPT ( bold_italic_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ,

where hHsubscript𝐻h_{H}italic_h start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT and hFsubscript𝐹h_{F}italic_h start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT are the Hartree and Fock contributions respectively.

Both hHsubscript𝐻h_{H}italic_h start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT and hFsubscript𝐹h_{F}italic_h start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT are strongly peaked in a small region near 𝒌=Γ𝒌Γ{\boldsymbol{k}}=\Gammabold_italic_k = roman_Γ, and quite flat away from ΓΓ\Gammaroman_Γ. We expect this to be a generic feature of most band dispersions of translationally symmetric correlated insulators. Indeed, away from ΓΓ\Gammaroman_Γ, the wavefunctions are largely 𝒌𝒌{\boldsymbol{k}}bold_italic_k independent and are well described by linear combinations of Wannier states localized at the AA sites of the moiré lattice. The region near the Gamma point is distinct, since ψΓ(𝒓=AA)subscript𝜓Γ𝒓𝐴𝐴\psi_{\Gamma}({\boldsymbol{r}}=AA)italic_ψ start_POSTSUBSCRIPT roman_Γ end_POSTSUBSCRIPT ( bold_italic_r = italic_A italic_A ) vanishes, and has strong 𝒌𝒌{\boldsymbol{k}}bold_italic_k dependence [54, 55, 56, 57, 58, 59, 60, 61, 62]. Thus, any translationally-symmetric mean-field potential projected to the flat bands, Δ(𝒌)=u𝒌|Δ^|u𝒌Δ𝒌quantum-operator-productsubscript𝑢𝒌^Δsubscript𝑢𝒌\Delta({\boldsymbol{k}})=\langle u_{{\boldsymbol{k}}}|\hat{\Delta}|u_{% \boldsymbol{k}}\rangleroman_Δ ( bold_italic_k ) = ⟨ italic_u start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT | over^ start_ARG roman_Δ end_ARG | italic_u start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT ⟩, will in general be flat away from ΓΓ\Gammaroman_Γ and have a bump or dip near ΓΓ\Gammaroman_Γ. The effect is especially intuitive for the Hartree dispersion, where ΓΓ\Gammaroman_Γ-point electrons are cheaper to dope than the rest of the Brillouin Zone because their wavefunctions vanish at the AA charge density peak, thereby evading a large electrostatic penalty. We pause to comment that the intervalley kekulé spiral state is a translation-breaking candidate state stabilized by strain that may have a distinct dispersion[63, 64, 65, 66, 67]. As we alluded to in the previous section, we expect that the geometric contribution to the superfluid stiffness will have a similar do** dependence as the interaction-induced kinetic part. Indeed, both the dispersion of the band and the quantum geometry are generated by the strong 𝒌𝒌{\boldsymbol{k}}bold_italic_k-dependence of the wavefunctions near the ΓΓ\Gammaroman_Γ point band edge.

In our computations of the Hartree Fock dispersion we use gate screened Coulomb interactions V𝒒=12ϵϵ0qtanhqdsubscript𝑉𝒒12italic-ϵsubscriptitalic-ϵ0𝑞𝑞𝑑V_{\boldsymbol{q}}=\frac{1}{2\epsilon\epsilon_{0}q}\tanh qditalic_V start_POSTSUBSCRIPT bold_italic_q end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 italic_ϵ italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_q end_ARG roman_tanh italic_q italic_d with gate distance d=20𝑑20d=20italic_d = 20nm, a relative permittivity ϵ=20italic-ϵ20\epsilon=20italic_ϵ = 20 (coming from screening of hBN, remote band electrons, and the decoupled Dirac cone sector). Our BM model wavefunctions correspond to a twist angle of 1.5superscript1.51.5^{\circ}1.5 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT and AA-AB tunneling ratio of w0/w1=0.5subscript𝑤0subscript𝑤10.5w_{0}/w_{1}=0.5italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_w start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0.5.

We see that the resulting band structure is highly non-parabolic. To visualize the non-parabolicity, we plot 12vFkF12subscript𝑣𝐹subscript𝑘𝐹\frac{1}{2}v_{F}k_{F}divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT and EFsubscript𝐸𝐹E_{F}italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT, equal for a parabolic band, versus the filling ν𝜈\nuitalic_ν. See Fig.S11. While the Fermi energy monotonically increases, 12vFkF12subscript𝑣𝐹subscript𝑘𝐹\frac{1}{2}v_{F}k_{F}divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT peaks at a small filling factor and subsequently decreases. Note that the kinetic contribution to the superfluid stiffness is 18πvFkF18𝜋subscript𝑣𝐹subscript𝑘𝐹\frac{1}{8\pi}v_{F}k_{F}divide start_ARG 1 end_ARG start_ARG 8 italic_π end_ARG italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT.

2.4 Superfluid stiffness in other devices

We measured a total of 4 devices comprising three twisted trilayer graphene devices (TTG1, TTG3, TTG-MIT) and one twisted quadrilayer graphene device (T4G). Within these devices, six independent superconducting domes, including both electron (e) and hole (h) doped sectors were investigated using simultaneous DC transport and rf-reflectometry. In Fig. S7, we show the zero-temperature superfluid stiffness ρs(0)subscript𝜌𝑠0\rho_{s}(0)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( 0 ) as a function of TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT for all measured superconducting domes. A roughly linear trend is observed, even for devices where twist-disorder appears to be strong (see discussion below).

In Figs. S8 and S9 we show data from all measured superconducting domes where the top panel represents resistance measurements as a function of ν𝜈\nuitalic_ν and T𝑇Titalic_T, the middle panel is the frequency shift of the resonator and the bottom panel is the temperature derivative of stiffness (dρs/dT𝑑subscript𝜌𝑠𝑑𝑇d\rho_{s}/dTitalic_d italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_d italic_T) or equivalently the frequency shift (dfr/dT𝑑subscript𝑓𝑟𝑑𝑇df_{r}/dTitalic_d italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT / italic_d italic_T). The derivative plots allow us to identify regions within the superconducting dome where the stiffness varies strongly as a function of temperature, with red regions indicating zero slope. All measured superconducting domes show large values of dρs/dT𝑑subscript𝜌𝑠𝑑𝑇d\rho_{s}/dTitalic_d italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_d italic_T at low temperatures, consistent with nodal superconductivity. However, in some devices, for certain values of ν𝜈\nuitalic_ν, we measure dρs/dT0similar-to-or-equals𝑑subscript𝜌𝑠𝑑𝑇0d\rho_{s}/dT\simeq 0italic_d italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_d italic_T ≃ 0 representing a saturated ρs(T)subscript𝜌𝑠𝑇\rho_{s}(T)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ) at low temperature (bottom panel of Figs. S7(b) and S7(c)). On inspection, we find that these regions coincide with fillings where the two-terminal resistance also shows signatures of large twist-angle disorder, identified as a cascade of resistance transitions as a function of temperature and do** (top panel of Figs. S7(b) and S7(c)). For example, in T4G-e, we observe saturated ρ(T)𝜌𝑇\rho(T)italic_ρ ( italic_T ) in the same region of (ν,T)(23,0.020.4K)similar-to-or-equals𝜈𝑇230.020.4𝐾(\nu,T)\simeq(2-3,0.02-0.4K)( italic_ν , italic_T ) ≃ ( 2 - 3 , 0.02 - 0.4 italic_K ) where R2tsubscript𝑅2𝑡R_{2t}italic_R start_POSTSUBSCRIPT 2 italic_t end_POSTSUBSCRIPT shows signatures of strong twist-disorder induced inhomogeneity in DC transport. Similar behavior is seen in TTG-MIT where we observe T-linear suppression of ρs(T)subscript𝜌𝑠𝑇\rho_{s}(T)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ) in the gate range VG=1719subscript𝑉𝐺1719V_{G}=17-19italic_V start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT = 17 - 19 V where Tc0.7similar-to-or-equalssubscript𝑇𝑐0.7T_{c}\simeq 0.7italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≃ 0.7 K is highest, but a low-temperature saturation of ρs(T)subscript𝜌𝑠𝑇\rho_{s}(T)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ) is observed in the gate range VG=1317subscript𝑉𝐺1317V_{G}=13-17italic_V start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT = 13 - 17 V, where Tc0.3similar-to-or-equalssubscript𝑇𝑐0.3T_{c}\simeq 0.3italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≃ 0.3 K is lower, and also cascades of resistance transitions are observed in R2tsubscript𝑅2𝑡R_{2t}italic_R start_POSTSUBSCRIPT 2 italic_t end_POSTSUBSCRIPT. Our data suggests that strong twist disorder tends to produce a low-temperature saturation of ρs(T)subscript𝜌𝑠𝑇\rho_{s}(T)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ) whereas regions of clean superconductivity show linear-T behavior consistent with nodal superconductivity.

In Fig. S10 we show the nonlinear Meissner effect observed at optimal do** in TTG1 in both the electron and hole doped sectors. Red dashed lines show fits to a quadratic dependence, δρs=bIb2𝛿subscript𝜌𝑠𝑏superscriptsubscript𝐼𝑏2\delta\rho_{s}=-bI_{b}^{2}italic_δ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = - italic_b italic_I start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The quadratic coefficient b𝑏bitalic_b as a function of T𝑇Titalic_T shows diverging behavior as T0𝑇0T\to 0italic_T → 0, indicating nodal superconductivity. Nonlinear Meissner effect measurements are particularly sensitive to the magnitude of the critical current and require relatively large values of Icsubscript𝐼𝑐absentI_{c}\geqitalic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≥100 nA. Moreover, twist-disorder-induced inhomogeneity causes cascades of superconducting transitions as a function of I𝐼Iitalic_I that are not conducive to a reliable investigation of ρs(I)subscript𝜌𝑠𝐼\rho_{s}(I)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_I ). Only the cleanest devices, TTG1 and TTG3, showed the nonlinear Meissner effect, whereas in the more twist-disordered devices we were not able to perform reliable ρs(I)subscript𝜌𝑠𝐼\rho_{s}(I)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_I ) measurements.

Refer to caption
Figure S1: Device DC transport characterization. a, TTG 4-terminal R𝑅Ritalic_R as a function of ν𝜈\nuitalic_ν and magnetic field B𝐵Bitalic_B at zero displacement field and a temperature of 2 K. The Landau fans show two set of structures: one set with large slopes and the other that appear at low B𝐵Bitalic_B with shallow slopes. b, TTG 4-terminal R𝑅Ritalic_R as a function of ν𝜈\nuitalic_ν and displacement field D𝐷Ditalic_D at zero B𝐵Bitalic_B and a temperature of 30 mK.
Refer to caption
Figure S2: Measurement setup Circuit model of measurement setup for superfluid stiffness measurements. The moiré graphene sample is indicated by dashed pink lines while the circuit board–containing sample, LC matching network, and bias tees to allow impedance matching for microwave measurement–is indicated by dashed purple lines. DC and RF filtering is used to reduce noise throughout the measurement chain. The RF signal is sent through an attenuated input line before entering a directional coupler. Sample contact resistance of 3.2similar-to-or-equalsabsent3.2\simeq 3.2≃ 3.2 kΩΩ\Omegaroman_Ω is given by Rcsubscript𝑅𝑐R_{c}italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. Graphite top and bottom gates are represented by ”gate”. The final signal is amplified both at 4 K and room temperature before being measured by a vector network analyzer (VNA).
Refer to caption
Figure S3: Analytical circuit model: (a) Model of the resonant circuit with an externally added inductor L0subscript𝐿0L_{0}italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, capacitance CPsubscript𝐶𝑃C_{P}italic_C start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT, contact resistance RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT and the kinetic inductance LKsubscript𝐿𝐾L_{K}italic_L start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT (b) Series RLC circuit equivalent to the circuit in (a). (c) Numerical simulation of the circuit in (a) for realistic parameters and changing kinetic inductance. (d) Resonance frequency frsubscript𝑓𝑟f_{r}italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT comparing the linear analytic formula presented above and the exact result.
Refer to caption
Figure S4: Circle fitting: (a) Circle fitting of S21subscript𝑆21S_{21}italic_S start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT in the complex plane. (b) Fitting of the reflection amplitude |S21(f)|subscript𝑆21𝑓|S_{21}(f)|| italic_S start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT ( italic_f ) |. (c) Fitting of the phase expressed in degrees argS21(f)subscript𝑆21𝑓\arg{S_{21}(f)}roman_arg italic_S start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT ( italic_f ).
Refer to caption
Figure S5: Frequency shift analysis: (a) Resonance frequency frsubscript𝑓𝑟f_{r}italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT as a function of T𝑇Titalic_T measured at ν=2.4𝜈2.4\nu=-2.4italic_ν = - 2.4 compared with simultaneously measured 4-terminal DC resistance R4tsubscript𝑅4𝑡R_{4t}italic_R start_POSTSUBSCRIPT 4 italic_t end_POSTSUBSCRIPT. The strong downward shift of frsubscript𝑓𝑟f_{r}italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT at T0.9similar-to-or-equals𝑇0.9T\simeq 0.9italic_T ≃ 0.9 K coincides with the onset of zero four-terminal resistance. (b) LKsubscript𝐿𝐾L_{K}italic_L start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT extracted from frsubscript𝑓𝑟f_{r}italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT shows large kinetic inductances 120350similar-to-or-equalsabsent120350\simeq 120-350≃ 120 - 350 nH. (c) Superfluid stiffness extracted from LKsubscript𝐿𝐾L_{K}italic_L start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT, accounting for the geometric aspect ratio of the device ρs=(w/l)1/LKsubscript𝜌𝑠𝑤𝑙1subscript𝐿𝐾\rho_{s}=(w/l)1/L_{K}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = ( italic_w / italic_l ) 1 / italic_L start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT, where w/l5similar-to-or-equals𝑤𝑙5w/l\simeq 5italic_w / italic_l ≃ 5.
Refer to caption
Figure S6: Quality factor analysis: (a) Coupling quality factor QCsubscript𝑄𝐶Q_{C}italic_Q start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT and (b) loaded quality factor QLsubscript𝑄𝐿Q_{L}italic_Q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT obtained from circle fitting. (c) Simulatanously measured two-terminal resistance R2tsubscript𝑅2𝑡R_{2t}italic_R start_POSTSUBSCRIPT 2 italic_t end_POSTSUBSCRIPT and (d) four-termianl resistance R4tsubscript𝑅4𝑡R_{4t}italic_R start_POSTSUBSCRIPT 4 italic_t end_POSTSUBSCRIPT. (e) Extracted rf resistance Rrfsubscript𝑅𝑟𝑓R_{rf}italic_R start_POSTSUBSCRIPT italic_r italic_f end_POSTSUBSCRIPT showing a qualitative agreement with (c) and (d) in the νT𝜈𝑇\nu-Titalic_ν - italic_T plane. (f) Quantitative agreement is achieved in the superconducting state where RrfR2tRcsimilar-to-or-equalssubscript𝑅𝑟𝑓subscript𝑅2𝑡similar-to-or-equalssubscript𝑅𝑐R_{rf}\simeq R_{2t}\simeq R_{c}italic_R start_POSTSUBSCRIPT italic_r italic_f end_POSTSUBSCRIPT ≃ italic_R start_POSTSUBSCRIPT 2 italic_t end_POSTSUBSCRIPT ≃ italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, where Rcsubscript𝑅𝑐R_{c}italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is the contact resistance.
Refer to caption
Figure S7: Uemura’s law: Scaling of ρs0subscript𝜌𝑠0\rho_{s0}italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT versus Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT (obtained from resistance measurements) for both electron and hole side superconductors in all measured devices. A roughly linear trend is generally observed, even for devices with twist-angle disorder.
Refer to caption
Figure S8: Superconducting domes in other devices: Superconducting domes measured as a function of temperature T𝑇Titalic_T and carrier density for (a) TTG1-e (b)TTG-MIT-e and (c) T4G-e. Top panel: Two terminal resistance R2tsubscript𝑅2𝑡R_{2t}italic_R start_POSTSUBSCRIPT 2 italic_t end_POSTSUBSCRIPT measured with DC transport. Middle panel: Resonance frequency frsubscript𝑓𝑟f_{r}italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT measured simultaneously with rf-reflectometry. Bottom panel: Temperature derivative of the superfluid stiffness, dρs/dT𝑑subscript𝜌𝑠𝑑𝑇d\rho_{s}/dTitalic_d italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_d italic_T (for TTG1-e), or equivalently the resonance frequency, dfr/dT𝑑subscript𝑓𝑟𝑑𝑇df_{r}/dTitalic_d italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT / italic_d italic_T (for TTG-MIT-e and T4G-e). dρs/dT=0𝑑subscript𝜌𝑠𝑑𝑇0d\rho_{s}/dT=0italic_d italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_d italic_T = 0 (dfr/dT=0𝑑subscript𝑓𝑟𝑑𝑇0df_{r}/dT=0italic_d italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT / italic_d italic_T = 0) is indicated with red on the colorscale. The smaller frequency shift in the latter devices makes dfr/dT𝑑subscript𝑓𝑟𝑑𝑇df_{r}/dTitalic_d italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT / italic_d italic_T more reliable than dρ/dT𝑑𝜌𝑑𝑇d\rho/dTitalic_d italic_ρ / italic_d italic_T. We observe non-zero derivatives within most of the νT𝜈𝑇\nu-Titalic_ν - italic_T phase diagram, especially at optimal do** (max Tc) for every superconducting dome. At optimal do**, all superconducting domes show linear-T low-temperature stiffness variation, with a roughly constant slope. On the other hand, away from optimal do**, several parts of the νT𝜈𝑇\nu-Titalic_ν - italic_T phase diagram, especially in TTG-MIT-e and T4G-e show strong resistivity variations within the superconducting dome, identified from the two-terminal resistance measurements (top panel). We attribute these resistivity variations to twist angle disorder, where multiple superconducting grains with different transition temperatures produce smaller overlap** superconducting domes. Corresponding twist-disordered parts of the phase-diagram usually show saturated frequency dependence as a function of temperature, indicated in the bottom panel as regions with zero slope (red on the colorscale).
Refer to caption
Figure S9: Superconducting domes in other devices: Superconducting domes measured as a function of temperature T𝑇Titalic_T and carrier density for (a) TTG1-h (b)TTG3-e and (c) TTG3-h (discussed in the main text). Top panel: Device resistance (R2tsubscript𝑅2𝑡R_{2t}italic_R start_POSTSUBSCRIPT 2 italic_t end_POSTSUBSCRIPT or R4tsubscript𝑅4𝑡R_{4t}italic_R start_POSTSUBSCRIPT 4 italic_t end_POSTSUBSCRIPT) measured with DC transport. Middle panel: Resonance frequency frsubscript𝑓𝑟f_{r}italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT measured simultaneously with rf-reflectometry. Bottom panel: Temperature derivative of the superfluid stiffness, dρs/dT𝑑subscript𝜌𝑠𝑑𝑇d\rho_{s}/dTitalic_d italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_d italic_T, with dρs/dT=0𝑑subscript𝜌𝑠𝑑𝑇0d\rho_{s}/dT=0italic_d italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_d italic_T = 0 indicated with red on the colorscale. We observe non-zero derivatives within most of the νT𝜈𝑇\nu-Titalic_ν - italic_T phase diagram, especially at optimal do** (max Tc) for every superconducting dome.
Refer to caption
Figure S10: Nonlinear Meissner effect at optimal do** in TTG-1: (a-c) Nonlinear Meissner effect measurement at optimal hole do** in TTG-1, ν=2.45𝜈2.45\nu=-2.45italic_ν = - 2.45 (a) ρs(T)subscript𝜌𝑠𝑇\rho_{s}(T)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ) measured at different values of supercurrent bias Ibsubscript𝐼𝑏I_{b}italic_I start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT. (b) ρs(I)subscript𝜌𝑠𝐼\rho_{s}(I)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_I ) measured at different values of temperature T𝑇Titalic_T. Red dashed lines show fits to a quadratic dependence, δρs=bIb2𝛿subscript𝜌𝑠𝑏superscriptsubscript𝐼𝑏2\delta\rho_{s}=-bI_{b}^{2}italic_δ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = - italic_b italic_I start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (c) Quadratic coefficient b𝑏bitalic_b as a function of T𝑇Titalic_T, show diverging behavior as T0𝑇0T\to 0italic_T → 0. (d-f) Nonlinear Meissner effect measurement at optimal electron do** in TTG-1, ν=+2.6𝜈2.6\nu=+2.6italic_ν = + 2.6 (a) ρs(T)subscript𝜌𝑠𝑇\rho_{s}(T)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ) measured at different values of supercurrent bias Ibsubscript𝐼𝑏I_{b}italic_I start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT. (b) ρs(I)subscript𝜌𝑠𝐼\rho_{s}(I)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_I ) measured at different values of temperature T𝑇Titalic_T. Red dashed lines show fits to a quadratic dependence, δρs=bIb2𝛿subscript𝜌𝑠𝑏superscriptsubscript𝐼𝑏2\delta\rho_{s}=-bI_{b}^{2}italic_δ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = - italic_b italic_I start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (c) Quadratic coefficient b𝑏bitalic_b as a function of T𝑇Titalic_T, show diverging behavior as T0𝑇0T\to 0italic_T → 0.
Refer to caption
Figure S11: Non-parabolicity of Hartree Fock band structure. While EF=12vFkFsubscript𝐸𝐹12Planck-constant-over-2-pisubscript𝑣𝐹subscript𝑘𝐹E_{F}=\frac{1}{2}\hbar v_{F}k_{F}italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_ℏ italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT for a parabolic band, the Hartree Fock dispersion of the TTG flat bands is much more concentrated at the ΓΓ\Gammaroman_Γ point, and flattens out elsewhere. Thus, vFsubscript𝑣𝐹v_{F}italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT begins decreasing at rather-small do**. For a non-parabolic band, EF12vFkF=4πρs0diasubscript𝐸𝐹12Planck-constant-over-2-pisubscript𝑣𝐹subscript𝑘𝐹4𝜋superscriptsubscript𝜌𝑠0𝑑𝑖𝑎E_{F}\neq\frac{1}{2}\hbar v_{F}k_{F}=4\pi\rho_{s0}^{dia}italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ≠ divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_ℏ italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = 4 italic_π italic_ρ start_POSTSUBSCRIPT italic_s 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d italic_i italic_a end_POSTSUPERSCRIPT.