Linear spectroscopy of collective modes and the gap structure
in two-dimensional superconductors

B. A. Levitan Department of Condensed Matter Physics, Weizmann Institute of Science, Rehovot 7610001, Israel    Y. Oreg Department of Condensed Matter Physics, Weizmann Institute of Science, Rehovot 7610001, Israel    E. Berg Department of Condensed Matter Physics, Weizmann Institute of Science, Rehovot 7610001, Israel    M. S. Rudner Department of Physics, University of Washington, Seattle, WA 98195-1560, USA    I. Iorsh Department of Condensed Matter Physics, Weizmann Institute of Science, Rehovot 7610001, Israel Faculty of Physics, ITMO University, St. Petersburg 197101, Russia Department of Physics, Engineering Physics and Astronomy, Queen’s University, Kingston, Ontario K7L 3N6, Canada
(July 3, 2024)
Abstract

We consider optical response in multi-band, multi-layer two-dimensional superconductors. Within a simple model, we show that linear response to AC gating can detect collective modes of the condensate, such as Leggett and clap** modes. We show how trigonal war** of the superconducting order parameter can help facilitate detection of clap** modes. Taking rhombohedral trilayer graphene as an example, we consider several possible pairing mechanisms and show that all-electronic mechanisms may produce in-gap clap** modes. These modes, if present, should be detectable in the absorption of microwaves applied via the gate electrodes, which are necessary to enable superconductivity in this and many other settings; their detection would constitute strong evidence for unconventional pairing. Lastly, we show that absorption at frequencies above the superconducting gap 2|Δ|2Δ2|\Delta|2 | roman_Δ | also contains a wealth of information about the gap structure. Our results suggest that linear spectroscopy can be a powerful tool for the characterization of unconventional two-dimensional superconductors.

Superconductivity, Leggett mode, clap** mode, rhombohedral trilayer graphene, collective mode spectroscopy.

In superconductors, a minimum of two collective modes arise: the Anderson-Bogoliubov-Goldstone (ABG) mode [1, 2, 3, 4], corresponding to phase fluctuations of the complex order parameter, and the Higgs mode [5, 6], corresponding to amplitude fluctuations. When the superconducting order parameter has more structure, its enlarged configuration space can produce a richer diversity of collective modes [7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32]. Measurements of the collective mode spectrum (e.g., by electromagnetic absorption) can therefore provide a valuable tool for diagnosing the nature of the order parameter, and even the underlying pairing mechanism [11, 12, 13, 33, 34, 35, 36, 37, 38, 30, 39, 40]. In this work, we show how to excite collective modes in gated two-dimensional superconductors using the gate electrodes themselves, providing a natural way to probe the nature of superconductivity in a wide range of emerging new materials.

As a prototypical example, consider the Leggett mode a in multi-band superconductor [8]. One can think of this mode as a momentum-space analogue of the Josephson effect, consisting of relative number/phase fluctuations between the condensates formed from different bands. Assuming that intra-band interactions are attractive, when the Josephson coupling (i.e., inter-band pair swap** interaction) is not too strong, the Leggett mode can arise inside the superconducting gap, and it can therefore be underdamped. However, even when this mode is a well-defined excitation, crystal symmetries often render it invisible to linear electromagnetic response at 𝐪0𝐪0\mathbf{q}\rightarrow 0bold_q → 0 [41, 42]. As we show below, in a gated 2D superconductor, static gating can break Mzsubscript𝑀𝑧M_{z}italic_M start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT mirror symmetry, permitting linear access to the Leggett mode at 𝐪0𝐪0\mathbf{q}\rightarrow 0bold_q → 0. The same gate electrodes can then be used to apply an AC probe field, thereby exciting the mode. This approach, which is uniquely well-suited to gated superconductors, represents an alternative to Raman [43, 34, 44, 21, 35, 28, 36, 45, 37, 38, 46, 47, 48, 39] and tunneling [33, 49] spectroscopy.

Refer to caption
Figure 1: (a) A two-layer superconductor under a large constant electric displacement field D𝐷Ditalic_D, driven by a perturbing AC field δD(t)𝛿𝐷𝑡\delta D(t)italic_δ italic_D ( italic_t ). (b) Single-electron dispersion in the toy model (black), showing the effect of a static perturbing field (red). (c) Schematic electron-electron scattering between Fermi sheets, indicating the relative phase (δθ𝛿𝜃\delta\thetaitalic_δ italic_θ) and number (δn𝛿𝑛\delta nitalic_δ italic_n) fluctuations of the Leggett mode. (d) Absorption spectrum given ν1=ν2subscript𝜈1subscript𝜈2\nu_{1}=\nu_{2}italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, g11(s)=g22(s)superscriptsubscript𝑔11𝑠superscriptsubscript𝑔22𝑠g_{11}^{(s)}=g_{22}^{(s)}italic_g start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT = italic_g start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT, and Δ1=Δ2=ΔsubscriptΔ1subscriptΔ2Δ\Delta_{1}=\Delta_{2}=\Deltaroman_Δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = roman_Δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = roman_Δ, for different values of interband coupling g12(s)subscriptsuperscript𝑔𝑠12g^{(s)}_{12}italic_g start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT and g(p)=0superscript𝑔𝑝0g^{(p)}=0italic_g start_POSTSUPERSCRIPT ( italic_p ) end_POSTSUPERSCRIPT = 0. (e, f) The clap** modes in a chiral p𝑝pitalic_p-wave superconductor, using the toy model with gαα(s)=0>g(p)superscriptsubscript𝑔𝛼superscript𝛼𝑠0superscript𝑔𝑝g_{\alpha\alpha^{\prime}}^{(s)}=0>g^{(p)}italic_g start_POSTSUBSCRIPT italic_α italic_α start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT = 0 > italic_g start_POSTSUPERSCRIPT ( italic_p ) end_POSTSUPERSCRIPT and |V|μmuch-greater-than𝑉𝜇|V|\gg\mu| italic_V | ≫ italic_μ. (e) Assuming a px+ipysubscript𝑝𝑥𝑖subscript𝑝𝑦p_{x}+ip_{y}italic_p start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT + italic_i italic_p start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ground state, the clap** modes at 𝐪0𝐪0\mathbf{q}\rightarrow 0bold_q → 0 correspond to fluctuations of the form Δ𝐤(t)=Δ0eiφ𝐤+(a(t)+ib(t))eiφ𝐤subscriptΔ𝐤𝑡subscriptΔ0superscript𝑒𝑖subscript𝜑𝐤𝑎𝑡𝑖𝑏𝑡superscript𝑒𝑖subscript𝜑𝐤\Delta_{\mathbf{k}}(t)=\Delta_{0}e^{i\varphi_{\mathbf{k}}}+\left(a(t)+ib(t)% \right)e^{-i\varphi_{\mathbf{k}}}roman_Δ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT ( italic_t ) = roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_φ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT end_POSTSUPERSCRIPT + ( italic_a ( italic_t ) + italic_i italic_b ( italic_t ) ) italic_e start_POSTSUPERSCRIPT - italic_i italic_φ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT end_POSTSUPERSCRIPT. Solid lines show a polar plot of the equilibrium order parameter, dashed lines show the distortion associated with each mode. (f) Real (dashed) and imaginary (solid) parts of the clap** mode contribution to the electronic compressibility for small but nonzero 𝐪𝐪\mathbf{q}bold_q in the circularly-symmetric model, Eq. (14); see text. We give ΩΩ\Omegaroman_Ω a small imaginary part by hand to give the pole a finite width.

Besides the Leggett mode, another family of exotic collective modes arise in superconductors with spontaneously broken time-reversal symmetry. Known as “generalized clap** modes” [9, 10, 11, 13, 12, 14, 15, 16, 17, 30], these amount to fluctuations in the uncondensed time-reversed partner of the condensed ground-state channel. For example, in a px+ipysubscript𝑝𝑥𝑖subscript𝑝𝑦p_{x}+ip_{y}italic_p start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT + italic_i italic_p start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT chiral superconductor, clap** modes arise as fluctuations in the pxipysubscript𝑝𝑥𝑖subscript𝑝𝑦p_{x}-ip_{y}italic_p start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT - italic_i italic_p start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT channel, as shown in Fig. 1e. Linear coupling of light to clap** modes at 𝐪0𝐪0\mathbf{q}\rightarrow 0bold_q → 0 generically requires breaking of rotational symmetry due to the non-vanishing angular momentum of these modes. Direct observation of clap** modes in linear electromagnetic response is therefore typically not possible at 𝐪0𝐪0\mathbf{q}\rightarrow 0bold_q → 0 in 2D materials with Cnsubscript𝐶𝑛C_{n}italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT rotational symmetry, where n>2𝑛2n>2italic_n > 2. Working at finite 𝐪𝐪\mathbf{q}bold_q, we focus on the experimentally-relevant case where (discrete) rotational symmetry is preserved, and show that the wavevector required for an appreciable signal corresponds to length scales comparable to device size in the case of superconducting rhombohedral trilayer graphene (RTG), which may host a chiral p𝑝pitalic_p-wave order parameter [50]. Our results therefore point towards a particularly convenient method for characterizing pairing in gated 2D superconductors.

In order to elucidate the basic ideas at work, we begin with a toy model which captures both the Leggett and clap** modes. The Hamiltonian is H=H0+Hint𝐻subscript𝐻0subscript𝐻intH=H_{0}+H_{\text{int}}italic_H = italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT int end_POSTSUBSCRIPT, with

H0=𝐤σ(f𝐤tσf𝐤bσ)(ξ𝐤0+VJJξ𝐤0V)(f𝐤tσf𝐤bσ)=𝐤σ(c𝐤1σc𝐤2σ)(ξ𝐤100ξ𝐤2)(c𝐤1σc𝐤2σ)subscript𝐻0subscript𝐤𝜎matrixsubscriptsuperscript𝑓𝐤𝑡𝜎subscriptsuperscript𝑓𝐤𝑏𝜎matrixsubscript𝜉𝐤0𝑉𝐽𝐽subscript𝜉𝐤0𝑉matrixsubscript𝑓𝐤𝑡𝜎subscript𝑓𝐤𝑏𝜎subscript𝐤𝜎matrixsubscriptsuperscript𝑐𝐤1𝜎subscriptsuperscript𝑐𝐤2𝜎matrixsubscript𝜉𝐤100subscript𝜉𝐤2matrixsubscript𝑐𝐤1𝜎subscript𝑐𝐤2𝜎H_{0}=\sum_{\mathbf{k}\sigma}\begin{pmatrix}f^{{\dagger}}_{\mathbf{k}t\sigma}&% f^{{\dagger}}_{\mathbf{k}b\sigma}\end{pmatrix}\begin{pmatrix}\xi_{\mathbf{k}0}% +V&J\\ J&\xi_{\mathbf{k}0}-V\end{pmatrix}\begin{pmatrix}f_{\mathbf{k}t\sigma}\\ f_{\mathbf{k}b\sigma}\end{pmatrix}\\ =\sum_{\mathbf{k}\sigma}\begin{pmatrix}c^{{\dagger}}_{\mathbf{k}1\sigma}&c^{{% \dagger}}_{\mathbf{k}2\sigma}\end{pmatrix}\begin{pmatrix}\xi_{\mathbf{k}1}&0\\ 0&\xi_{\mathbf{k}2}\end{pmatrix}\begin{pmatrix}c_{\mathbf{k}1\sigma}\\ c_{\mathbf{k}2\sigma}\end{pmatrix}start_ROW start_CELL italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT bold_k italic_σ end_POSTSUBSCRIPT ( start_ARG start_ROW start_CELL italic_f start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k italic_t italic_σ end_POSTSUBSCRIPT end_CELL start_CELL italic_f start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k italic_b italic_σ end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ) ( start_ARG start_ROW start_CELL italic_ξ start_POSTSUBSCRIPT bold_k 0 end_POSTSUBSCRIPT + italic_V end_CELL start_CELL italic_J end_CELL end_ROW start_ROW start_CELL italic_J end_CELL start_CELL italic_ξ start_POSTSUBSCRIPT bold_k 0 end_POSTSUBSCRIPT - italic_V end_CELL end_ROW end_ARG ) ( start_ARG start_ROW start_CELL italic_f start_POSTSUBSCRIPT bold_k italic_t italic_σ end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL italic_f start_POSTSUBSCRIPT bold_k italic_b italic_σ end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ) end_CELL end_ROW start_ROW start_CELL = ∑ start_POSTSUBSCRIPT bold_k italic_σ end_POSTSUBSCRIPT ( start_ARG start_ROW start_CELL italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k 1 italic_σ end_POSTSUBSCRIPT end_CELL start_CELL italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k 2 italic_σ end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ) ( start_ARG start_ROW start_CELL italic_ξ start_POSTSUBSCRIPT bold_k 1 end_POSTSUBSCRIPT end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL italic_ξ start_POSTSUBSCRIPT bold_k 2 end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ) ( start_ARG start_ROW start_CELL italic_c start_POSTSUBSCRIPT bold_k 1 italic_σ end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL italic_c start_POSTSUBSCRIPT bold_k 2 italic_σ end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ) end_CELL end_ROW (1)

describing spin-1/2 fermions (σ{,}𝜎\sigma\in\{\uparrow,\downarrow\}italic_σ ∈ { ↑ , ↓ }) in two layers (l{t,b}𝑙𝑡𝑏l\in\{t,b\}italic_l ∈ { italic_t , italic_b }), each with bare dispersion ξ𝐤0=ϵ𝐤0μsubscript𝜉𝐤0subscriptitalic-ϵ𝐤0𝜇\xi_{\mathbf{k}0}=\epsilon_{\mathbf{k}0}-\muitalic_ξ start_POSTSUBSCRIPT bold_k 0 end_POSTSUBSCRIPT = italic_ϵ start_POSTSUBSCRIPT bold_k 0 end_POSTSUBSCRIPT - italic_μ relative to the Fermi level. We assume isotropic ξ𝐤0subscript𝜉𝐤0\xi_{\mathbf{k}0}italic_ξ start_POSTSUBSCRIPT bold_k 0 end_POSTSUBSCRIPT for simplicity. Single-particle tunneling (J𝐽Jitalic_J) couples the layers, and the electric potential drops by 2V2𝑉2V2 italic_V from top to bottom (we set the electron charge to 1111). Including interlayer tunneling, the band energies are ξ𝐤,α{1,2}=ξ𝐤0±J2+V2subscript𝜉𝐤𝛼12plus-or-minussubscript𝜉𝐤0superscript𝐽2superscript𝑉2\xi_{\mathbf{k},\alpha\in\{1,2\}}=\xi_{\mathbf{k}0}\pm\sqrt{J^{2}+V^{2}}italic_ξ start_POSTSUBSCRIPT bold_k , italic_α ∈ { 1 , 2 } end_POSTSUBSCRIPT = italic_ξ start_POSTSUBSCRIPT bold_k 0 end_POSTSUBSCRIPT ± square-root start_ARG italic_J start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_V start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG. Notice that starting from V=0𝑉0V=0italic_V = 0, perturbing with a small V𝑉Vitalic_V produces energy shifts at second order in V𝑉Vitalic_V. This reflects a more general principle: barring accidental degeneracies, any system which respects the mirror symmetry Mzsubscript𝑀𝑧M_{z}italic_M start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT (which here exchanges the two layers) must have a spectrum which is even in the Mzsubscript𝑀𝑧M_{z}italic_M start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT-breaking out-of-plane displacement field, preventing linear access to the Leggett mode. In Fig. 1a, we assume a strong field (|V||J|much-greater-than𝑉𝐽|V|\gg|J|| italic_V | ≫ | italic_J |), strongly breaking Mzsubscript𝑀𝑧M_{z}italic_M start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, and rendering the band splitting essentially linear in |V|𝑉|V|| italic_V |. This assumption is not essential to our calculations; any V0𝑉0V\neq 0italic_V ≠ 0 will enable linear response to a subsequent perturbation δV𝛿𝑉\delta Vitalic_δ italic_V.

We model interactions using

Hint=1L2𝐤𝐤𝐪αα𝒱𝐤𝐤ααc𝐤αc𝐤+𝐪,αc𝐤+𝐪,αc𝐤α,subscript𝐻int1superscript𝐿2subscript𝐤𝐤𝐪𝛼superscript𝛼superscriptsubscript𝒱𝐤𝐤𝛼superscript𝛼subscriptsuperscript𝑐𝐤𝛼absentsubscriptsuperscript𝑐𝐤𝐪𝛼absentsubscript𝑐𝐤𝐪superscript𝛼absentsubscript𝑐𝐤superscript𝛼absentH_{\text{int}}=\frac{1}{L^{2}}\sum_{\begin{subarray}{c}\mathbf{kk^{\prime}q}\\ \alpha\alpha^{\prime}\end{subarray}}\mathcal{V}_{\mathbf{kk^{\prime}}}^{\alpha% \alpha^{\prime}}c^{{\dagger}}_{\mathbf{k}\alpha\uparrow}c^{{\dagger}}_{-% \mathbf{k}+\mathbf{q},\alpha\downarrow}c_{-\mathbf{k^{\prime}}+\mathbf{q},% \alpha^{\prime}\downarrow}c_{\mathbf{k^{\prime}}\alpha^{\prime}\uparrow},start_ROW start_CELL italic_H start_POSTSUBSCRIPT int end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ∑ start_POSTSUBSCRIPT start_ARG start_ROW start_CELL bold_kk start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT bold_q end_CELL end_ROW start_ROW start_CELL italic_α italic_α start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_CELL end_ROW end_ARG end_POSTSUBSCRIPT caligraphic_V start_POSTSUBSCRIPT bold_kk start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α italic_α start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k italic_α ↑ end_POSTSUBSCRIPT italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - bold_k + bold_q , italic_α ↓ end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT - start_ID bold_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ID + bold_q , italic_α start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ↓ end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT start_ID bold_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ID italic_α start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ↑ end_POSTSUBSCRIPT , end_CELL end_ROW (2)

focusing on small 𝐪𝐪\mathbf{q}bold_q (i.e. the Cooper channel). L𝐿Litalic_L is the linear dimension of the sample. For concreteness, we take

𝒱𝐤𝐤αα=gαα(s)+2g(p)cos(φ𝐤φ𝐤)δαα,superscriptsubscript𝒱𝐤𝐤𝛼superscript𝛼subscriptsuperscript𝑔𝑠𝛼superscript𝛼2superscript𝑔𝑝subscript𝜑𝐤subscript𝜑𝐤subscript𝛿𝛼superscript𝛼\mathcal{V}_{\mathbf{kk^{\prime}}}^{\alpha\alpha^{\prime}}=g^{(s)}_{\alpha% \alpha^{\prime}}+2g^{(p)}\cos(\varphi_{\mathbf{k}}-\varphi_{\mathbf{k^{\prime}% }})\delta_{\alpha\alpha^{\prime}},caligraphic_V start_POSTSUBSCRIPT bold_kk start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α italic_α start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT = italic_g start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_α start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT + 2 italic_g start_POSTSUPERSCRIPT ( italic_p ) end_POSTSUPERSCRIPT roman_cos ( start_ARG italic_φ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT - italic_φ start_POSTSUBSCRIPT bold_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT end_ARG ) italic_δ start_POSTSUBSCRIPT italic_α italic_α start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT , (3)

where gαα(s)<0subscriptsuperscript𝑔𝑠𝛼𝛼0g^{(s)}_{\alpha\alpha}<0italic_g start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_α end_POSTSUBSCRIPT < 0 binds s𝑠sitalic_s-wave pairs in band α𝛼\alphaitalic_α, g12(s)=(g21(s))subscriptsuperscript𝑔𝑠12superscriptsubscriptsuperscript𝑔𝑠21g^{(s)}_{12}=\left(g^{(s)}_{21}\right)^{*}italic_g start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT = ( italic_g start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT Josephson couples the two bands, and g(p)<0superscript𝑔𝑝0g^{(p)}<0italic_g start_POSTSUPERSCRIPT ( italic_p ) end_POSTSUPERSCRIPT < 0 binds p𝑝pitalic_p-wave pairs. Here φ𝐤subscript𝜑𝐤\varphi_{\mathbf{k}}italic_φ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT is the angle of 𝐤𝐤\mathbf{k}bold_k relative to the kxsubscript𝑘𝑥k_{x}italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT axis, i.e., 𝐤=|𝐤|(cosφ𝐤,sinφ𝐤)𝐤𝐤subscript𝜑𝐤subscript𝜑𝐤\mathbf{k}=|\mathbf{k}|(\cos\varphi_{\mathbf{k}},\sin\varphi_{\mathbf{k}})bold_k = | bold_k | ( roman_cos italic_φ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT , roman_sin italic_φ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT ). Note that with the spin indices as written in Eq. (2), when considering p𝑝pitalic_p-wave pairing, we implicitly consider the mz=0subscript𝑚𝑧0m_{z}=0italic_m start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0 triplet component only; the extension to the more general case is straightforward.

We base our analysis [51] on the imaginary-time path integral. To identify the collective modes, we perform a Hubbard-Stratonovich decoupling in the Cooper channel, and expand the effective action to quadratic order in fluctuations ϕitalic-ϕ\phiitalic_ϕ around the mean-field saddle point [52, 53, 54, 25, 28, 30]. Analytically continuing the Gaussian effective action to real frequency (iΩmΩ+i0+𝑖subscriptΩ𝑚Ω𝑖superscript0i\Omega_{m}\rightarrow\Omega+i0^{+}italic_i roman_Ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT → roman_Ω + italic_i 0 start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT), the collective modes appear as poles of the bosonic propagator for pair fluctuations. To evaluate the linear response to applied electric fields, we also compute the couplings between the collective modes and such fields.

First we focus on the Leggett mode, by setting g(p)=0superscript𝑔𝑝0g^{(p)}=0italic_g start_POSTSUPERSCRIPT ( italic_p ) end_POSTSUPERSCRIPT = 0 and gαα(s)0subscriptsuperscript𝑔𝑠𝛼superscript𝛼0g^{(s)}_{\alpha\alpha^{\prime}}\neq 0italic_g start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_α start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ≠ 0. The Leggett mode requires μ>J2+V2𝜇superscript𝐽2superscript𝑉2\mu>\sqrt{J^{2}+V^{2}}italic_μ > square-root start_ARG italic_J start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_V start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG such that both bands are occupied. Leggett’s original paper analyzed essentially this model—at 𝐪=0𝐪0\mathbf{q}=0bold_q = 0 and with g(p)=0superscript𝑔𝑝0g^{(p)}=0italic_g start_POSTSUPERSCRIPT ( italic_p ) end_POSTSUPERSCRIPT = 0—and found its relative-phase mode [8]. We assume gαα<0subscript𝑔𝛼𝛼0g_{\alpha\alpha}<0italic_g start_POSTSUBSCRIPT italic_α italic_α end_POSTSUBSCRIPT < 0. Taking |Δ1||Δ2|subscriptΔ1subscriptΔ2|\Delta_{1}|\leq|\Delta_{2}|| roman_Δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | ≤ | roman_Δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT | without further loss of generality, we find [51] that an in-gap Leggett mode requires 0|g12(s)|/(g11(s)g22(s)|g12(s)|2)<ν2arcsin(|Δ1/Δ2|)/1(Δ1/Δ2)20subscriptsuperscript𝑔𝑠12subscriptsuperscript𝑔𝑠11subscriptsuperscript𝑔𝑠22superscriptsubscriptsuperscript𝑔𝑠122subscript𝜈2arcsinesubscriptΔ1subscriptΔ21superscriptsubscriptΔ1subscriptΔ220\leq|g^{(s)}_{12}|/(g^{(s)}_{11}g^{(s)}_{22}-|g^{(s)}_{12}|^{2})<\nu_{2}% \arcsin(|\Delta_{1}/\Delta_{2}|)/\sqrt{1-(\Delta_{1}/\Delta_{2})^{2}}0 ≤ | italic_g start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT | / ( italic_g start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT italic_g start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT - | italic_g start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) < italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_arcsin ( start_ARG | roman_Δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / roman_Δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT | end_ARG ) / square-root start_ARG 1 - ( roman_Δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / roman_Δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG, where ναsubscript𝜈𝛼\nu_{\alpha}italic_ν start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT is the density of states per unit area per spin on the Fermi surface of band α𝛼\alphaitalic_α in the normal state. (When |Δ1|=|Δ2|subscriptΔ1subscriptΔ2|\Delta_{1}|=|\Delta_{2}|| roman_Δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | = | roman_Δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT |, the second inequality is irrelevant). Fermi-liquid interactions can modify this condition [45]; see Sec. II of the Supplemental Material [51] for details.

We now show how an AC modulation of the displacement field can probe the Leggett mode. The drive field corresponds to a small time-dependent perturbation around the static background, VVDC+δV(t)𝑉subscript𝑉DC𝛿𝑉𝑡V\rightarrow V_{\rm DC}+\delta V(t)italic_V → italic_V start_POSTSUBSCRIPT roman_DC end_POSTSUBSCRIPT + italic_δ italic_V ( italic_t ). The physical idea is that in the presence of the static VDC0subscript𝑉DC0V_{\text{DC}}\neq 0italic_V start_POSTSUBSCRIPT DC end_POSTSUBSCRIPT ≠ 0 (which breaks Mzsubscript𝑀𝑧M_{z}italic_M start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT), the perturbation δV𝛿𝑉\delta Vitalic_δ italic_V produces linear energy shifts in opposite directions for the two bands, as illustrated in Fig. 1a, directly driving a phase difference between the two condensates and hence exciting the Leggett mode. Kee** only terms up to linear order in δV𝛿𝑉\delta Vitalic_δ italic_V and working in the band basis, driving corresponds to a Hamiltonian term HδV=𝐤σδV(t)(VDC/J2+VDC2)(c𝐤1σc𝐤1σc𝐤2σc𝐤2σ)subscript𝐻𝛿𝑉subscript𝐤𝜎𝛿𝑉𝑡subscript𝑉DCsuperscript𝐽2superscriptsubscript𝑉DC2subscriptsuperscript𝑐𝐤1𝜎subscript𝑐𝐤1𝜎subscriptsuperscript𝑐𝐤2𝜎subscript𝑐𝐤2𝜎H_{\delta V}=\sum_{\mathbf{k}\sigma}\delta V(t)(V_{\rm DC}/\sqrt{J^{2}+V_{\rm DC% }^{2}})(c^{{\dagger}}_{\mathbf{k}1\sigma}c_{\mathbf{k}1\sigma}-c^{{\dagger}}_{% \mathbf{k}2\sigma}c_{\mathbf{k}2\sigma})italic_H start_POSTSUBSCRIPT italic_δ italic_V end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT bold_k italic_σ end_POSTSUBSCRIPT italic_δ italic_V ( italic_t ) ( italic_V start_POSTSUBSCRIPT roman_DC end_POSTSUBSCRIPT / square-root start_ARG italic_J start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_V start_POSTSUBSCRIPT roman_DC end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) ( italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k 1 italic_σ end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT bold_k 1 italic_σ end_POSTSUBSCRIPT - italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k 2 italic_σ end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT bold_k 2 italic_σ end_POSTSUBSCRIPT ). Integrating out the fermions in the presence of the drive yields a quadratic term δV2similar-toabsent𝛿superscript𝑉2\sim\delta V^{2}∼ italic_δ italic_V start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and source terms ϕδVsimilar-toabsentitalic-ϕ𝛿𝑉\sim\phi\,\delta V∼ italic_ϕ italic_δ italic_V in the effective action, on top of the Gaussian action for pair fluctuations ϕ𝒟1ϕsimilar-toabsentsuperscriptitalic-ϕsuperscript𝒟1italic-ϕ\sim-\phi^{{\dagger}}\mathcal{D}^{-1}\phi∼ - italic_ϕ start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT caligraphic_D start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_ϕ. Integrating out the fluctuations ϕitalic-ϕ\phiitalic_ϕ then yields the full quadratic free energy functional [δV\mathcal{F}[\delta Vcaligraphic_F [ italic_δ italic_V], describing a capacitance (dependent on frequency ΩΩ\Omegaroman_Ω):

1L2[δV(Ω)]=12𝒞(Ω)δV(Ω)δV(Ω).1superscript𝐿2delimited-[]𝛿𝑉Ω12𝒞Ω𝛿𝑉Ω𝛿𝑉Ω\frac{1}{L^{2}}\mathcal{F}[\delta V(\Omega)]=\frac{1}{2}\mathcal{C}(\Omega)% \delta V(-\Omega)\delta V(\Omega).divide start_ARG 1 end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG caligraphic_F [ italic_δ italic_V ( roman_Ω ) ] = divide start_ARG 1 end_ARG start_ARG 2 end_ARG caligraphic_C ( roman_Ω ) italic_δ italic_V ( - roman_Ω ) italic_δ italic_V ( roman_Ω ) . (4)

We provide an explicit calculation of the capacitance per unit area 𝒞(Ω)𝒞Ω\mathcal{C}(\Omega)caligraphic_C ( roman_Ω ) in the Supplemental Material [51]. In the symmetric case where Δ1=Δ2=Δ>0subscriptΔ1subscriptΔ2Δ0\Delta_{1}=\Delta_{2}=\Delta>0roman_Δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = roman_Δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = roman_Δ > 0 and ν1=ν2=νsubscript𝜈1subscript𝜈2𝜈\nu_{1}=\nu_{2}=\nuitalic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_ν, the collective-mode contribution is

𝒞ϕ(Ω)=4VDC2VDC2+J2(νγsecγ)2νγtanγ2g~1,subscript𝒞italic-ϕΩ4superscriptsubscript𝑉DC2superscriptsubscript𝑉DC2superscript𝐽2superscript𝜈𝛾𝛾2𝜈𝛾𝛾2superscript~𝑔1\mathcal{C}_{\phi}(\Omega)=\frac{4V_{\text{DC}}^{2}}{V_{\text{DC}}^{2}+J^{2}}% \frac{(\nu\gamma\sec\gamma)^{2}}{\nu\gamma\tan\gamma-2\tilde{g}^{-1}},caligraphic_C start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT ( roman_Ω ) = divide start_ARG 4 italic_V start_POSTSUBSCRIPT DC end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_V start_POSTSUBSCRIPT DC end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_J start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG divide start_ARG ( italic_ν italic_γ roman_sec italic_γ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ν italic_γ roman_tan italic_γ - 2 over~ start_ARG italic_g end_ARG start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT end_ARG , (5)

where

g~1=|g12(s)|/(g11(s)g22(s)|g12(s)|2)superscript~𝑔1superscriptsubscript𝑔12𝑠superscriptsubscript𝑔11𝑠superscriptsubscript𝑔22𝑠superscriptsuperscriptsubscript𝑔12𝑠2\tilde{g}^{-1}=\big{|}g_{12}^{(s)}\big{|}/\Big{(}g_{11}^{(s)}g_{22}^{(s)}-\big% {|}g_{12}^{(s)}\big{|}^{2}\Big{)}over~ start_ARG italic_g end_ARG start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = | italic_g start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT | / ( italic_g start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT italic_g start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT - | italic_g start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) (6)

and

sinγ=(Ω+i0+)/(2|Δ|).𝛾Ω𝑖superscript02Δ\sin\gamma=(\Omega+i0^{+})/(2|\Delta|).roman_sin italic_γ = ( roman_Ω + italic_i 0 start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ) / ( 2 | roman_Δ | ) . (7)

The Leggett mode frequency ΩLsubscriptΩL\Omega_{\text{L}}roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT solves νγLtanγL=2g~1𝜈subscript𝛾Lsubscript𝛾L2superscript~𝑔1\nu\gamma_{\text{L}}\tan\gamma_{\text{L}}=2\tilde{g}^{-1}italic_ν italic_γ start_POSTSUBSCRIPT L end_POSTSUBSCRIPT roman_tan italic_γ start_POSTSUBSCRIPT L end_POSTSUBSCRIPT = 2 over~ start_ARG italic_g end_ARG start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, with sinγL=ΩL/(2Δ)subscript𝛾LsubscriptΩL2Δ\sin\gamma_{\text{L}}=\Omega_{\text{L}}/(2\Delta)roman_sin italic_γ start_POSTSUBSCRIPT L end_POSTSUBSCRIPT = roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT / ( 2 roman_Δ ). Near the positive-frequency pole, we can approximate

𝒞ϕ(Ω)4VDC2VDC2+J2νγL2sinγL+γLsecγL2ΔΩΩL+i0+.subscript𝒞italic-ϕΩ4superscriptsubscript𝑉DC2superscriptsubscript𝑉DC2superscript𝐽2𝜈superscriptsubscript𝛾L2subscript𝛾Lsubscript𝛾Lsubscript𝛾L2ΔΩsubscriptΩL𝑖superscript0\mathcal{C}_{\phi}(\Omega)\approx\frac{4V_{\text{DC}}^{2}}{V_{\text{DC}}^{2}+J% ^{2}}\frac{\nu\gamma_{\text{L}}^{2}}{\sin\gamma_{\text{L}}+\gamma_{\text{L}}% \sec\gamma_{\text{L}}}\frac{2\Delta}{\Omega-\Omega_{\text{L}}+i0^{+}}.caligraphic_C start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT ( roman_Ω ) ≈ divide start_ARG 4 italic_V start_POSTSUBSCRIPT DC end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_V start_POSTSUBSCRIPT DC end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_J start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG divide start_ARG italic_ν italic_γ start_POSTSUBSCRIPT L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG roman_sin italic_γ start_POSTSUBSCRIPT L end_POSTSUBSCRIPT + italic_γ start_POSTSUBSCRIPT L end_POSTSUBSCRIPT roman_sec italic_γ start_POSTSUBSCRIPT L end_POSTSUBSCRIPT end_ARG divide start_ARG 2 roman_Δ end_ARG start_ARG roman_Ω - roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT + italic_i 0 start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT end_ARG . (8)

In the general (asymmetric) case, the key results are that: i) The amplitude modes and ABG mode (at 𝐪0𝐪0\mathbf{q}\rightarrow 0bold_q → 0) do not contribute to 𝒞(Ω)𝒞Ω\mathcal{C}(\Omega)caligraphic_C ( roman_Ω ); and ii) The Leggett mode does contribute an in-gap pole to 𝒞(Ω)𝒞Ω\mathcal{C}(\Omega)caligraphic_C ( roman_Ω ), provided that the background field VDC0subscript𝑉DC0V_{\rm DC}\neq 0italic_V start_POSTSUBSCRIPT roman_DC end_POSTSUBSCRIPT ≠ 0 (and that the mode is in-gap). We make no attempt to explicitly compute the width of the pole, which is controlled by higher-order processes, and is model-dependent. However, we expect the resonance to be narrow, since no quasiparticle excitations are available to facilitate decay through low-order processes.

Fig. 1d shows the total absorption (Im𝒞𝒞-\imaginary\mathcal{C}- start_OPERATOR roman_Im end_OPERATOR caligraphic_C, including quasiparticle and collective-mode contributions) for the symmetric toy model. In this case, when the interactions become uniform (g12(s)g11(s)=g22(s)subscriptsuperscript𝑔𝑠12subscriptsuperscript𝑔𝑠11subscriptsuperscript𝑔𝑠22g^{(s)}_{12}\rightarrow g^{(s)}_{11}=g^{(s)}_{22}italic_g start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT → italic_g start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT = italic_g start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT), the absorption peak corresponding to the Leggett mode approaches the gap edge. For illustrative purposes, in this figure we add a small imaginary part to ΩΩ\Omegaroman_Ω, tuned by hand to give the pole an approximately equal width in each of the curves.

We now turn our attention to the clap** modes, setting gαα(s)=0>g(p)subscriptsuperscript𝑔𝑠𝛼superscript𝛼0superscript𝑔𝑝g^{(s)}_{\alpha\alpha^{\prime}}=0>g^{(p)}italic_g start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_α start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT = 0 > italic_g start_POSTSUPERSCRIPT ( italic_p ) end_POSTSUPERSCRIPT and |V|μmuch-greater-than𝑉𝜇|V|\gg\mu| italic_V | ≫ italic_μ. We project into the occupied (lower) band, and hence drop the band index α𝛼\alphaitalic_α. The gap equation has two degenerate solutions, corresponding to px±ipyplus-or-minussubscript𝑝𝑥𝑖subscript𝑝𝑦p_{x}\pm ip_{y}italic_p start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ± italic_i italic_p start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT pairing. We assume the px+ipysubscript𝑝𝑥𝑖subscript𝑝𝑦p_{x}+ip_{y}italic_p start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT + italic_i italic_p start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT channel condenses, i.e., Δ𝐤=Δ0eiφ𝐤subscriptΔ𝐤subscriptΔ0superscript𝑒𝑖subscript𝜑𝐤\Delta_{\mathbf{k}}=\Delta_{0}e^{i\varphi_{\mathbf{k}}}roman_Δ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT = roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_φ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT end_POSTSUPERSCRIPT with Δ0>0subscriptΔ00\Delta_{0}>0roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT > 0. As detailed in the Supplemental Material [51], fluctuations in the uncondensed pxipysubscript𝑝𝑥𝑖subscript𝑝𝑦p_{x}-ip_{y}italic_p start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT - italic_i italic_p start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT channel give rise to two real bosons a𝑎aitalic_a and b𝑏bitalic_b; these are the clap** modes, which in this symmetric minimal model are degenerate at Ω=2Δ0Ω2subscriptΔ0\Omega=\sqrt{2}\Delta_{0}roman_Ω = square-root start_ARG 2 end_ARG roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT.

To show how an AC electric field can probe the clap** modes, as in the Leggett case, we perturb around the static background field, VVDC+δV(t)𝑉subscript𝑉DC𝛿𝑉𝑡V\rightarrow V_{\text{DC}}+\delta V(t)italic_V → italic_V start_POSTSUBSCRIPT DC end_POSTSUBSCRIPT + italic_δ italic_V ( italic_t ). Since we projected out the upper band, the only effect of the modulation is to shift the energy of the lower band; when the displacement field-induced interlayer potential is very large compared with the interlayer hop** matrix element, the electrons only live on one layer, so they only feel the perturbing field through its local potential. Therefore, we calculate the clap** mode contribution to the electronic compressibility Π00superscriptΠ00\Pi^{00}roman_Π start_POSTSUPERSCRIPT 00 end_POSTSUPERSCRIPT. Unlike the Leggett mode, the clap** modes carry angular momentum 2222 relative to the ground state; this means that with circular symmetry, their couplings to the scalar potential must vanish at 𝐪=0𝐪0\mathbf{q}=0bold_q = 0. Hence, we calculate their compressibility contributions to leading nonvanishing order in |𝐪|𝐪|\mathbf{q}|| bold_q |.

Our calculation [51] mirrors that of Ref. [30], generalized to nonzero momentum. After Hubbard-Stratonovich decoupling, the fermions encounter the total (fluctuating) pairing field

Δ𝐤,q=Δq(+)eiφ𝐤+Δq()eiφ𝐤.subscriptΔ𝐤𝑞subscriptsuperscriptΔ𝑞superscript𝑒𝑖subscript𝜑𝐤subscriptsuperscriptΔ𝑞superscript𝑒𝑖subscript𝜑𝐤\Delta_{\mathbf{k},q}=\Delta^{(+)}_{q}e^{i\varphi_{\mathbf{k}}}+\Delta^{(-)}_{% q}e^{-i\varphi_{\mathbf{k}}}.roman_Δ start_POSTSUBSCRIPT bold_k , italic_q end_POSTSUBSCRIPT = roman_Δ start_POSTSUPERSCRIPT ( + ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_φ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT end_POSTSUPERSCRIPT + roman_Δ start_POSTSUPERSCRIPT ( - ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_φ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT end_POSTSUPERSCRIPT . (9)

Here q=(iΩm,𝐪)𝑞𝑖subscriptΩ𝑚𝐪q=(i\Omega_{m},\mathbf{q})italic_q = ( italic_i roman_Ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT , bold_q ) contains both the bosonic Matsubara frequency Ωm=2πmTsubscriptΩ𝑚2𝜋𝑚𝑇\Omega_{m}=2\pi mTroman_Ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT = 2 italic_π italic_m italic_T and momentum 𝐪𝐪\mathbf{q}bold_q. In coordinate space, writing x=(τ,𝐫)𝑥𝜏𝐫x=(\tau,\mathbf{r})italic_x = ( italic_τ , bold_r ), the Hubbard-Stratonovich bosons corresponding to the two pairing channels are

Δ(+)(x)=eiθ(x)(Δ0+h(x))superscriptΔ𝑥superscript𝑒𝑖𝜃𝑥subscriptΔ0𝑥\Delta^{(+)}(x)=e^{i\theta(x)}(\Delta_{0}+h(x))roman_Δ start_POSTSUPERSCRIPT ( + ) end_POSTSUPERSCRIPT ( italic_x ) = italic_e start_POSTSUPERSCRIPT italic_i italic_θ ( italic_x ) end_POSTSUPERSCRIPT ( roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_h ( italic_x ) ) (10)

and

Δ()(x)=eiθ(x)(a(x)+ib(x)).superscriptΔ𝑥superscript𝑒𝑖𝜃𝑥𝑎𝑥𝑖𝑏𝑥\Delta^{(-)}(x)=e^{i\theta(x)}(a(x)+ib(x)).roman_Δ start_POSTSUPERSCRIPT ( - ) end_POSTSUPERSCRIPT ( italic_x ) = italic_e start_POSTSUPERSCRIPT italic_i italic_θ ( italic_x ) end_POSTSUPERSCRIPT ( italic_a ( italic_x ) + italic_i italic_b ( italic_x ) ) . (11)

Here θ𝜃\thetaitalic_θ is the global phase (ABG) mode, hhitalic_h the Higgs mode, and a𝑎aitalic_a and b𝑏bitalic_b the clap** modes. After integrating out the fermions, and neglecting the (massive) Higgs mode for brevity, the long-wavelength action at quadratic order is

S=q(𝒟a,q1aqaq𝒟b,q1bqbq+Πqμν𝒜qμ𝒜qν+Πqμa𝒜qμaq+Πqμb𝒜qμbq),𝑆subscript𝑞subscriptsuperscript𝒟1𝑎𝑞subscript𝑎𝑞subscript𝑎𝑞subscriptsuperscript𝒟1𝑏𝑞subscript𝑏𝑞subscript𝑏𝑞subscriptsuperscriptΠ𝜇𝜈𝑞subscriptsuperscript𝒜𝜇𝑞subscriptsuperscript𝒜𝜈𝑞subscriptsuperscriptΠ𝜇𝑎𝑞subscriptsuperscript𝒜𝜇𝑞subscript𝑎𝑞subscriptsuperscriptΠ𝜇𝑏𝑞subscriptsuperscript𝒜𝜇𝑞subscript𝑏𝑞S=\sum_{q}\Bigg{(}-\mathcal{D}^{-1}_{a,q}a_{-q}a_{q}-\mathcal{D}^{-1}_{b,q}b_{% -q}b_{q}\\ +\Pi^{\mu\nu}_{q}\mathcal{A}^{\mu}_{-q}\mathcal{A}^{\nu}_{q}+\Pi^{\mu a}_{q}% \mathcal{A}^{\mu}_{-q}a_{q}+\Pi^{\mu b}_{q}\mathcal{A}^{\mu}_{-q}b_{q}\Bigg{)},start_ROW start_CELL italic_S = ∑ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( - caligraphic_D start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_a , italic_q end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT - italic_q end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT - caligraphic_D start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b , italic_q end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT - italic_q end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL + roman_Π start_POSTSUPERSCRIPT italic_μ italic_ν end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT caligraphic_A start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - italic_q end_POSTSUBSCRIPT caligraphic_A start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT + roman_Π start_POSTSUPERSCRIPT italic_μ italic_a end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT caligraphic_A start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - italic_q end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT + roman_Π start_POSTSUPERSCRIPT italic_μ italic_b end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT caligraphic_A start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - italic_q end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ) , end_CELL end_ROW (12)

where 𝒜=(𝒜0,𝓐)=(A0+τθ,𝐀θ)𝒜superscript𝒜0𝓐superscript𝐴0subscript𝜏𝜃𝐀𝜃\mathcal{A}=(\mathcal{A}^{0},\bm{\mathcal{A}})=(A^{0}+\partial_{\tau}\theta,% \mathbf{A}-\gradient\theta)caligraphic_A = ( caligraphic_A start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT , bold_caligraphic_A ) = ( italic_A start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT + ∂ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT italic_θ , bold_A - start_OPERATOR ∇ end_OPERATOR italic_θ ) is the gauge-invariant combination of the electromagnetic gauge potential A𝐴Aitalic_A and the spacetime gradient of the global phase θ𝜃\thetaitalic_θ, and summation over μ,ν{0,x,y}𝜇𝜈0𝑥𝑦\mu,\nu\in\{0,x,y\}italic_μ , italic_ν ∈ { 0 , italic_x , italic_y } is implied. Since we are only interested in the impact of the clap** modes on the compressibility, we drop all terms involving the spatial component 𝓐𝓐\bm{\mathcal{A}}bold_caligraphic_A for simplicity. Physically, we expect that this approximation should not drastically change the result, since the ABG mode associated with θ𝜃\gradient\thetastart_OPERATOR ∇ end_OPERATOR italic_θ is either at a much lower energy (in a neutral superfluid) or much higher energy (after the Anderson-Higgs mechanism in a superconductor) than the clap** mode, for small but nonzero 𝐪𝐪\mathbf{q}bold_q. We provide explicit expressions for the propagators 𝒟Xqsubscript𝒟𝑋𝑞\mathcal{D}_{Xq}caligraphic_D start_POSTSUBSCRIPT italic_X italic_q end_POSTSUBSCRIPT and couplings Πq0XsubscriptsuperscriptΠ0𝑋𝑞\Pi^{0X}_{q}roman_Π start_POSTSUPERSCRIPT 0 italic_X end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT (X{a,b}𝑋𝑎𝑏X\in\{a,b\}italic_X ∈ { italic_a , italic_b }) in the Supplemental Material [51]. Note that with circular symmetry, Πq0XsubscriptsuperscriptΠ0𝑋𝑞\Pi^{0X}_{q}roman_Π start_POSTSUPERSCRIPT 0 italic_X end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT vanishes at 𝐪=0𝐪0\mathbf{q}=0bold_q = 0 and first shows up at O(|𝐪|2)𝑂superscript𝐪2O(|\mathbf{q}|^{2})italic_O ( | bold_q | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ). The product Πq0X𝒜q0XqsubscriptsuperscriptΠ0𝑋𝑞subscriptsuperscript𝒜0𝑞subscript𝑋𝑞\Pi^{0X}_{q}\mathcal{A}^{0}_{-q}X_{q}roman_Π start_POSTSUPERSCRIPT 0 italic_X end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT caligraphic_A start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - italic_q end_POSTSUBSCRIPT italic_X start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT must be scalar, so two factors of momentum are required to balance the angular momentum 2 of the clap** modes.

Refer to caption
Figure 2: (a) Electron dispersion of RTG under a static perpendicular electric field corresponding to a potential drop of 20202020 meV (black) or 30303030 meV (orange) between the top and bottom layers. Here, kxsubscript𝑘𝑥k_{x}italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT is measured relative to the K𝐾Kitalic_K point. (b) Fermi surfaces of RTG in each of its two valleys, for a potential drop of 20202020 meV and a Fermi energy of 29.329.3-29.3- 29.3 meV. The line cut through the dispersion relation (panel (a)) is along the dotted line schematically indicated by the dashed line in panel (b). (c-d) Absorption spectrum and gap structure for (c) phonon-mediated s𝑠sitalic_s-wave pairing, and (d) intervalley coherence fluctuation-mediated chiral p𝑝pitalic_p-wave pairing. Note the in-gap clap** mode in the p𝑝pitalic_p-wave case. The gap curves are colored according to the Fermi surface component on which they reside, matching panel (b).

We obtain the clap** mode contribution to the compressibility δΠq00𝛿subscriptsuperscriptΠ00𝑞\delta\Pi^{00}_{q}italic_δ roman_Π start_POSTSUPERSCRIPT 00 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT by integrating out those modes, yielding

δΠq00=14Πq0a𝒟a,qΠq0a+14Πq0b𝒟b,qΠq0b.𝛿subscriptsuperscriptΠ00𝑞14subscriptsuperscriptΠ0𝑎𝑞subscript𝒟𝑎𝑞subscriptsuperscriptΠ0𝑎𝑞14subscriptsuperscriptΠ0𝑏𝑞subscript𝒟𝑏𝑞subscriptsuperscriptΠ0𝑏𝑞\delta\Pi^{00}_{q}=\frac{1}{4}\Pi^{0a}_{-q}\mathcal{D}_{a,q}\Pi^{0a}_{q}+\frac% {1}{4}\Pi^{0b}_{-q}\mathcal{D}_{b,q}\Pi^{0b}_{q}.italic_δ roman_Π start_POSTSUPERSCRIPT 00 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 4 end_ARG roman_Π start_POSTSUPERSCRIPT 0 italic_a end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - italic_q end_POSTSUBSCRIPT caligraphic_D start_POSTSUBSCRIPT italic_a , italic_q end_POSTSUBSCRIPT roman_Π start_POSTSUPERSCRIPT 0 italic_a end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT + divide start_ARG 1 end_ARG start_ARG 4 end_ARG roman_Π start_POSTSUPERSCRIPT 0 italic_b end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - italic_q end_POSTSUBSCRIPT caligraphic_D start_POSTSUBSCRIPT italic_b , italic_q end_POSTSUBSCRIPT roman_Π start_POSTSUPERSCRIPT 0 italic_b end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT . (13)

Analytically continuing to real time (iΩmΩ+i0+𝑖subscriptΩ𝑚Ω𝑖superscript0i\Omega_{m}\rightarrow\Omega+i0^{+}italic_i roman_Ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT → roman_Ω + italic_i 0 start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT), we find to leading order in 𝐪𝐪\mathbf{q}bold_q and for large μ𝜇\muitalic_μ:

δΠq00=2ν(π4ξBCS|𝐪|)41γ2(cotγtanγ)2γcos2γsin4γ.𝛿subscriptsuperscriptΠ00𝑞2𝜈superscript𝜋4subscript𝜉BCS𝐪41superscript𝛾2superscript𝛾𝛾2𝛾superscript2𝛾4𝛾\delta\Pi^{00}_{q}=2\nu\left(\frac{\pi}{4}\xi_{\text{BCS}}|\mathbf{q}|\right)^% {4}\frac{1-\gamma^{2}(\cot\gamma-\tan\gamma)^{2}}{\gamma\cos^{2}\gamma\sin 4% \gamma}.italic_δ roman_Π start_POSTSUPERSCRIPT 00 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT = 2 italic_ν ( divide start_ARG italic_π end_ARG start_ARG 4 end_ARG italic_ξ start_POSTSUBSCRIPT BCS end_POSTSUBSCRIPT | bold_q | ) start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT divide start_ARG 1 - italic_γ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( roman_cot italic_γ - roman_tan italic_γ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_γ roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_γ roman_sin 4 italic_γ end_ARG . (14)

Here ξBCS=vF/(πΔ0)subscript𝜉BCSsubscript𝑣F𝜋subscriptΔ0\xi_{\text{BCS}}=v_{\text{F}}/(\pi\Delta_{0})italic_ξ start_POSTSUBSCRIPT BCS end_POSTSUBSCRIPT = italic_v start_POSTSUBSCRIPT F end_POSTSUBSCRIPT / ( italic_π roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) is the BCS coherence length, and vF=2μ/msubscript𝑣F2𝜇𝑚v_{\text{F}}=\sqrt{2\mu/m}italic_v start_POSTSUBSCRIPT F end_POSTSUBSCRIPT = square-root start_ARG 2 italic_μ / italic_m end_ARG is the Fermi velocity. This result neglects the dispersion of the clap** modes. Note the pole at γ=π/4𝛾𝜋4\gamma=\pi/4italic_γ = italic_π / 4, corresponding to Ω=±2Δ0Ωplus-or-minus2subscriptΔ0\Omega=\pm\sqrt{2}\Delta_{0}roman_Ω = ± square-root start_ARG 2 end_ARG roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT: the clap** modes are visible in the compressibility, and hence in the capacitance. In the vicinity of the positive-frequency pole,

δΠq00ν(π4)3(ξBCS|𝐪|)42Δ0Ω2Δ0+i0+.𝛿subscriptsuperscriptΠ00𝑞𝜈superscript𝜋43superscriptsubscript𝜉BCS𝐪42subscriptΔ0Ω2subscriptΔ0𝑖superscript0\delta\Pi^{00}_{q}\approx\nu\left(\frac{\pi}{4}\right)^{3}\left(\xi_{\text{BCS% }}|\mathbf{q}|\right)^{4}\frac{-\sqrt{2}\Delta_{0}}{\Omega-\sqrt{2}\Delta_{0}+% i0^{+}}.italic_δ roman_Π start_POSTSUPERSCRIPT 00 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ≈ italic_ν ( divide start_ARG italic_π end_ARG start_ARG 4 end_ARG ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ( italic_ξ start_POSTSUBSCRIPT BCS end_POSTSUBSCRIPT | bold_q | ) start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT divide start_ARG - square-root start_ARG 2 end_ARG roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG roman_Ω - square-root start_ARG 2 end_ARG roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_i 0 start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT end_ARG . (15)

As explained above, the factor of |𝐪|4superscript𝐪4|\mathbf{q}|^{4}| bold_q | start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT in Eq. (14) is a consequence of circular symmetry. To obtain a signal at lower order in |𝐪|𝐪|\mathbf{q}|| bold_q |, something must break that symmetry in order to relax the angular momentum constraint. If the symmetry is broken completely, for example by a cos(ζ)px+isin(ζ)py𝜁subscript𝑝𝑥𝑖𝜁subscript𝑝𝑦\cos(\zeta)p_{x}+i\sin(\zeta)p_{y}roman_cos ( start_ARG italic_ζ end_ARG ) italic_p start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT + italic_i roman_sin ( start_ARG italic_ζ end_ARG ) italic_p start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT order parameter with ζ𝜁\zetaitalic_ζ away from high-symmetry values (as in Ref. [30]), then a signal can even arise at 𝐪0𝐪0\mathbf{q}\rightarrow 0bold_q → 0. A less drastic scenario is that rotation is explicitly broken down to a discrete subgroup dictated by the lattice structure. With C3vsubscript𝐶3𝑣C_{3v}italic_C start_POSTSUBSCRIPT 3 italic_v end_POSTSUBSCRIPT symmetry (possessed by RTG, for example), the couplings Πq0XsubscriptsuperscriptΠ0𝑋𝑞\Pi^{0X}_{q}roman_Π start_POSTSUPERSCRIPT 0 italic_X end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT need only contain a single factor of 𝐪𝐪\mathbf{q}bold_q, since 2=1(mod3)2annotated1pmod32=-1\pmod{3}2 = - 1 start_MODIFIER ( roman_mod start_ARG 3 end_ARG ) end_MODIFIER. We find [51] that trigonal war** indeed yields a nonvanishing result at O(|𝐪|2)𝑂superscript𝐪2O(|\mathbf{q}|^{2})italic_O ( | bold_q | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ), of similar qualitative form to Eq. (14), but suppressed by the small factor 1/(ξBCSkF)2(Δ0/μ)2similar-to1superscriptsubscript𝜉BCSsubscript𝑘F2superscriptsubscriptΔ0𝜇21/(\xi_{\text{BCS}}k_{\text{F}})^{2}\sim(\Delta_{0}/\mu)^{2}1 / ( italic_ξ start_POSTSUBSCRIPT BCS end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∼ ( roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_μ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.

To go beyond our simple toy model, we consider superconducting RTG as an illustrative case study. We model the single-particle bandstructure of RTG using a 6×6666\times 66 × 6 tight-binding Hamiltonian in its continuum limit, written in the basis of two atoms per unit cell in each of the three layers [55, 56, 51]. We limit ourselves to the range of parameters (static perpendicular displacement field and do** level) where superconductivity was observed [50], focusing on the SC2 phase. In this regime, prior to the onset of superconductivity, RTG contains one annular Fermi sea in each valley (two in total), whose Fermi surfaces show substantial trigonal war** (see Fig. 2b). The specific pairing mechanism responsible for superconductivity in RTG is yet to be determined, so we consider conventional (s𝑠sitalic_s wave, mediated by phonons) and unconventional (chiral p𝑝pitalic_p wave, mediated by fluctuations of either charge density, as in the Kohn-Luttinger mechanism [57, 58, 59, 60], or intervalley coherence [61, 62]) scenarios [51]. We construct a general formalism for calculating superconducting collective modes and their response functions in the Supplemental Material [51].

Figures 2c,d show our numerical results for the absorption spectrum of superconducting RTG in the SC2 phase, accounting for the Anderson-Higgs mechanism, and working to leading order in |Δ|/μΔ𝜇|\Delta|/\mu| roman_Δ | / italic_μ [51] (hence neglecting the weak O(|𝐪|2)𝑂superscript𝐪2O(|\mathbf{q}|^{2})italic_O ( | bold_q | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) contribution discussed above, which is permitted by the substantial trigonal war** present in RTG). As in the toy model discussion, the absorption corresponds to the imaginary part of the capacitance. Fig. 2c shows the results for the s𝑠sitalic_s-wave case. Fig. 2d shows the results for the IVC-mediated p𝑝pitalic_p-wave case; the Kohn-Luttinger mechanism gives qualitatively similar results. Most importantly, the all-electronic mechanisms both show an in-gap clap** mode. While the clap** mode is visible only away from 𝐪=0𝐪0\mathbf{q}=0bold_q = 0, since superconducting RTG devices are comparable in size to the superconducting coherence length ξBCSsubscript𝜉BCS\xi_{\text{BCS}}italic_ξ start_POSTSUBSCRIPT BCS end_POSTSUBSCRIPT, we expect finite size effects to render the clap** mode visible even without any special patterning of the gate electrodes.

Figs. 2c,d also show that even without the collective modes, the absorption spectrum contains rich information about the gap structure. Extrema of the superconducting gap reveal themselves as sharp features in the absorption.

To conclude, we have shown that simple linear spectroscopy via gate electrodes can be an invaluable tool for studying the pairing mechanisms in low-dimensional superconductors. Using a simple but powerful toy model, we showed how various exotic collective modes should be detectable in AC capacitance measurements. By considering the example of superconducting rhombohedral trilayer graphene (RTG) under a perpendicular displacement field, we further showed how the observation of an in-gap clap** mode could yield compelling evidence for unconventional, all-electronic pairing mechanisms. We also showed how linear spectroscopy above the quasiparticle excitation gap reveals information on the gap structure, in the form of sharp features at frequencies corresponding to gap extrema over the Fermi surface.

The physics we have described should be accessible in experiments. The superconducting phases of RTG [50] require dual gating to be realized, so RTG is an ideal candidate material for the gate-based detection scheme we have proposed. These phases have transition temperatures between 40similar-toabsent40\sim 40∼ 40 and 100similar-toabsent100\sim 100∼ 100 mK, so one expects the gap and possible in-gap clap** modes to reside at the scale of a few GHz to 10s of GHz, easily within the range of microwave function generators. Further, we expect that similar phenomenology may occur in other quasi-two-dimensional superconductors which are sensitive to a displacement field, such as Bernal bilayer [63, 64] and twisted trilayer [65, 66, 67, 68] graphene, as well as twisted bilayer WSe2 [69, 70]. Future work should study the collective mode spectroscopy of these systems.

Acknowledgements.
We thank Peter Hirschfeld, Saurabh Maiti, and Surajit Sarkar for their helpful comments on this manuscript. B. A. L. was hosted at the Institute for Theoretical Physics at the University of Cologne during a considerable portion of the preparation of this manuscript, and gratefully acknowledges their generous support, as well as the financial support of the Zuckerman STEM Leadership Program. This work was supported by NSF-BSF award DMR-2310312, by the European Union’s Horizon 2020 research and innovation programme (grant agreements LEGOTOP No. 788715 and HQMAT No. 817799), and the DFG CRC SFB/TRR183. M. R. gratefully acknowledges support from the Brown Investigator Award, a program of the Brown Science Foundation, the University of Washington College of Arts and Sciences, and the Kenneth K. Young Memorial Professorship.

References

  • Anderson [1958a] P. W. Anderson, Random-phase approximation in the theory of superconductivity, Phys. Rev. 112, 1900 (1958a).
  • Anderson [1958b] P. W. Anderson, New method in the theory of superconductivity, Phys. Rev. 110, 985 (1958b).
  • Bogoljubov et al. [1958] N. Bogoljubov, V. V. Tolmachov, and D. Širkov, A new method in the theory of superconductivity, Fortschr. Phys. 6, 605 (1958).
  • Rickayzen [1959] G. Rickayzen, Collective excitations in the theory of superconductivity, Phys. Rev. 115, 795 (1959).
  • Littlewood and Varma [1982] P. B. Littlewood and C. M. Varma, Amplitude collective modes in superconductors and their coupling to charge-density waves, Phys. Rev. B 26, 4883 (1982).
  • Shimano and Tsuji [2020] R. Shimano and N. Tsuji, Higgs mode in superconductors, Annu. Rev. Condens. Matter Phys. 11, 103 (2020).
  • Bardasis and Schrieffer [1961] A. Bardasis and J. R. Schrieffer, Excitons and plasmons in superconductors, Phys. Rev. 121, 1050 (1961).
  • Leggett [1966] A. Leggett, Number-phase fluctuations in two-band superconductors, Prog. Theor. Phys. 36, 901 (1966).
  • Wölfle [1976] P. Wölfle, Order-parameter collective modes in 3HeA𝐴-{A}- italic_APhys. Rev. Lett. 37, 1279 (1976).
  • Hirashima and Namaizawa [1988] D. S. Hirashima and H. Namaizawa, Collective modes and a theory of response of d-wave superconductors, J. Low Temp. Phys. 73, 137 (1988).
  • Hirschfeld et al. [1989] P. J. Hirschfeld, P. Wölfle, J. A. Sauls, D. Einzel, and W. O. Putikka, Electromagnetic absorption in anisotropic superconductors, Phys. Rev. B 40, 6695 (1989).
  • Hirschfeld et al. [1992] P. J. Hirschfeld, W. O. Putikka, and P. Wölfle, Electromagnetic power absorption by collective modes in unconventional superconductors, Phys. Rev. Lett. 69, 1447 (1992).
  • Yip and Sauls [1992] S. Yip and J. Sauls, Circular dichroism and birefringence in unconventional superconductors, J. Low Temp. Phys. 86, 257 (1992).
  • Kee et al. [2000] H.-Y. Kee, Y. B. Kim, and K. Maki, Collective modes and sound propagation in a p-wave superconductor: Sr2RuO4Phys. Rev. B 62, 5877 (2000).
  • Higashitani and Nagai [2000a] S. Higashitani and K. Nagai, Order parameter collective modes in Sr2RuO4Physica B: Condensed Matter 284, 539 (2000a).
  • Higashitani and Nagai [2000b] S. Higashitani and K. Nagai, Electromagnetic response of a kx±ikyplus-or-minussubscript𝑘𝑥subscriptik𝑦{k}_{x}\pm{}{\mathrm{ik}}_{y}italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ± roman_ik start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT superconductor: Effect of order-parameter collective modes, Phys. Rev. B 62, 3042 (2000b).
  • Balatsky et al. [2000] A. V. Balatsky, P. Kumar, and J. R. Schrieffer, Collective mode in a superconductor with mixed-symmetry order parameter components, Phys. Rev. Lett. 84, 4445 (2000).
  • Roy and Kallin [2008] R. Roy and C. Kallin, Collective modes and electromagnetic response of a chiral superconductor, Phys. Rev. B 77, 174513 (2008).
  • Ota et al. [2011] Y. Ota, M. Machida, T. Koyama, and H. Aoki, Collective modes in multiband superfluids and superconductors: Multiple dynamical classes, Phys. Rev. B 83, 060507(R) (2011).
  • Carlström et al. [2011] J. Carlström, J. Garaud, and E. Babaev, Length scales, collective modes, and type-1.5 regimes in three-band superconductors, Phys. Rev. B 84, 134518 (2011).
  • Lin and Hu [2012] S.-Z. Lin and X. Hu, Massless Leggett mode in three-band superconductors with time-reversal-symmetry breaking, Phys. Rev. Lett. 108, 177005 (2012).
  • Stanev [2012] V. Stanev, Model of collective modes in three-band superconductors with repulsive interband interactions, Phys. Rev. B 85, 174520 (2012).
  • Puetter and Kee [2012] C. M. Puetter and H.-Y. Kee, Identifying spin-triplet pairing in spin-orbit coupled multi-band superconductors, Europhys. Lett. 98, 27010 (2012).
  • Kobayashi et al. [2013] K. Kobayashi, M. Machida, Y. Ota, and F. Nori, Massless collective excitations in frustrated multiband superconductors, Phys. Rev. B 88, 224516 (2013).
  • Marciani et al. [2013] M. Marciani, L. Fanfarillo, C. Castellani, and L. Benfatto, Leggett modes in iron-based superconductors as a probe of time-reversal symmetry breaking, Phys. Rev. B 88, 214508 (2013).
  • Lin [2014] S.-Z. Lin, Ground state, collective mode, phase soliton and vortex in multiband superconductors, J. Phys. Condens. Matter 26, 493202 (2014).
  • Maiti and Hirschfeld [2015] S. Maiti and P. J. Hirschfeld, Collective modes in superconductors with competing s𝑠sitalic_s- and d𝑑ditalic_d-wave interactions, Phys. Rev. B 92, 094506 (2015).
  • Cea and Benfatto [2016] T. Cea and L. Benfatto, Signature of the Leggett mode in the A1gsubscript𝐴1𝑔{A}_{1g}italic_A start_POSTSUBSCRIPT 1 italic_g end_POSTSUBSCRIPT Raman response: From MgB2 to iron-based superconductors, Phys. Rev. B 94, 064512 (2016).
  • Sun et al. [2020] Z. Sun, M. M. Fogler, D. N. Basov, and A. J. Millis, Collective modes and terahertz near-field response of superconductors, Phys. Rev. Res. 2, 023413 (2020).
  • Poniatowski et al. [2022] N. R. Poniatowski, J. B. Curtis, A. Yacoby, and P. Narang, Spectroscopic signatures of time-reversal symmetry breaking superconductivity, Commun. Phys. 5, 44 (2022).
  • Gabriele et al. [2022] F. Gabriele, C. Castellani, and L. Benfatto, Generalized plasma waves in layered superconductors: A unified approach, Phys. Rev. Res. 4, 023112 (2022).
  • Sellati et al. [2023] N. Sellati, F. Gabriele, C. Castellani, and L. Benfatto, Generalized Josephson plasmons in bilayer superconductors, Phys. Rev. B 108, 014503 (2023).
  • Ponomarev et al. [2004] Y. G. Ponomarev, S. A. Kuzmichev, M. G. Mikheev, M. V. Sudakova, S. N. Tchesnokov, N. Z. Timergaleev, A. V. Yarigin, E. G. Maksimov, S. I. Krasnosvobodtsev, A. V. Varlashkin, M. A. Hein, G. Müller, H. Piel, L. G. Sevastyanova, O. V. Kravchenko, K. P. Burdina, and B. M. Bulychev, Evidence for a two-band behavior of MgB2 from point-contact and tunneling spectroscopy, Solid State Commun. 129, 85 (2004).
  • Chubukov et al. [2009] A. V. Chubukov, I. Eremin, and M. M. Korshunov, Theory of Raman response of a superconductor with extended s𝑠sitalic_s-wave symmetry: Application to the iron pnictides, Phys. Rev. B 79, 220501(R) (2009).
  • Böhm et al. [2014] T. Böhm, A. F. Kemper, B. Moritz, F. Kretzschmar, B. Muschler, H.-M. Eiter, R. Hackl, T. P. Devereaux, D. J. Scalapino, and H.-H. Wen, Balancing act: Evidence for a strong subdominant d𝑑ditalic_d-wave pairing channel in Ba0.6K0.4Fe2As2Phys. Rev. X 4, 041046 (2014).
  • Maiti et al. [2016] S. Maiti, T. A. Maier, T. Böhm, R. Hackl, and P. J. Hirschfeld, Probing the pairing interaction and multiple Bardasis-Schrieffer modes using Raman spectroscopy, Phys. Rev. Lett. 117, 257001 (2016).
  • Böhm et al. [2018] T. Böhm, F. Kretzschmar, A. Baum, M. Rehm, D. Jost, R. Hosseinian Ahangharnejhad, R. Thomale, C. Platt, T. A. Maier, W. Hanke, B. Moritz, T. P. Devereaux, D. J. Scalapino, S. Maiti, P. J. Hirschfeld, P. Adelmann, T. Wolf, H.-H. Wen, and R. Hackl, Microscopic origin of cooper pairing in the iron-based superconductor Ba1-xKxFe2As2npj Quantum Mater. 3, 48 (2018).
  • Huang et al. [2018] W. Huang, M. Sigrist, and Z.-Y. Weng, Identifying the dominant pairing interaction in high-Tcsubscript𝑇𝑐{T}_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT FeSe superconductors through Leggett modes, Phys. Rev. B 97, 144507 (2018).
  • Sarkar and Maiti [2024] S. Sarkar and S. Maiti, Electronic Raman response of a superconductor across a time reversal symmetry breaking phase transition, Phys. Rev. B 109, 094515 (2024).
  • Sellati et al. [2024] N. Sellati, J. Fiore, C. Castellani, and L. Benfatto, Optical absorption in tilted geometries as an indirect measurement of longitudinal plasma waves in layered cuprates, Nanomater. 1410.3390/nano14121021 (2024).
  • Kamatani et al. [2022] T. Kamatani, S. Kitamura, N. Tsuji, R. Shimano, and T. Morimoto, Optical response of the Leggett mode in multiband superconductors in the linear response regime, Phys. Rev. B 105, 094520 (2022).
  • Nagashima et al. [2024] R. Nagashima, S. Tian, R. Haenel, N. Tsuji, and D. Manske, Classification of Lifshitz invariant in multiband superconductors: An application to Leggett modes in the linear response regime in Kagome lattice models, Phys. Rev. Res. 6, 013120 (2024).
  • Blumberg et al. [2007] G. Blumberg, A. Mialitsin, B. S. Dennis, M. V. Klein, N. D. Zhigadlo, and J. Karpinski, Observation of Leggett’s collective mode in a multiband MgB2 superconductor, Phys. Rev. Lett. 99, 227002 (2007).
  • Klein [2010] M. V. Klein, Theory of Raman scattering from Leggett’s collective mode in a multiband superconductor: Application to MgB2Phys. Rev. B 82, 014507 (2010).
  • Maiti et al. [2017] S. Maiti, A. V. Chubukov, and P. J. Hirschfeld, Conservation laws, vertex corrections, and screening in Raman spectroscopy, Phys. Rev. B 96, 014503 (2017).
  • Giorgianni et al. [2019] F. Giorgianni, T. Cea, C. Vicario, C. P. Hauri, W. K. Withanage, X. Xi, and L. Benfatto, Leggett mode controlled by light pulses, Nat. Phys. 15, 341 (2019).
  • Zhao et al. [2020] S. Z. Zhao, H.-Y. Song, L. L. Hu, T. Xie, C. Liu, H. Q. Luo, C.-Y. Jiang, X. Zhang, X. C. Nie, J.-Q. Meng, Y.-X. Duan, S.-B. Liu, H.-Y. Xie, and H. Y. Liu, Observation of soft Leggett mode in superconducting CaKFe4As4Phys. Rev. B 102, 144519 (2020).
  • Benek-Lins and Maiti [2024] I. Benek-Lins and S. Maiti, Many-body physics-induced selection rules: Application to Raman spectroscopy, Phys. Rev. B 109, 104505 (2024).
  • Lee and Steiner [2023] P. A. Lee and J. F. Steiner, Detection of collective modes in unconventional superconductors using tunneling spectroscopy, Phys. Rev. B 108, 174503 (2023).
  • Zhou et al. [2021a] H. Zhou, T. Xie, T. Taniguchi, K. Watanabe, and A. F. Young, Superconductivity in rhombohedral trilayer graphene, Nature 598, 434 (2021a).
  • [51] Please find the details of our analysis and our RTG model in the Supplementary Material at URL_will_be_inserted_by_publisher.
  • De Palo et al. [1999] S. De Palo, C. Castellani, C. Di Castro, and B. K. Chakraverty, Effective action for superconductors and BCS-Bose crossover, Phys. Rev. B 60, 564 (1999).
  • Paramekanti et al. [2000] A. Paramekanti, M. Randeria, T. V. Ramakrishnan, and S. S. Mandal, Effective actions and phase fluctuations in d𝑑ditalic_d-wave superconductors, Phys. Rev. B 62, 6786 (2000).
  • Sharapov et al. [2002] S. Sharapov, V. Gusynin, and H. Beck, Effective action approach to the Leggett’s mode in two-band superconductors, Eur. Phys. J. B 30, 45 (2002).
  • Zhang et al. [2010] F. Zhang, B. Sahu, H. Min, and A. H. MacDonald, Band structure of ABC𝐴𝐵𝐶{A}{B}{C}italic_A italic_B italic_C-stacked graphene trilayers, Phys. Rev. B 82, 035409 (2010).
  • Zhou et al. [2021b] H. Zhou, T. Xie, A. Ghazaryan, T. Holder, J. R. Ehrets, E. M. Spanton, T. Taniguchi, K. Watanabe, E. Berg, M. Serbyn, and A. F. Young, Half-and quarter-metals in rhombohedral trilayer graphene, Nature 598, 429 (2021b).
  • Ghazaryan et al. [2021] A. Ghazaryan, T. Holder, M. Serbyn, and E. Berg, Unconventional superconductivity in systems with annular fermi surfaces: Application to rhombohedral trilayer graphene, Phys. Rev. Lett. 127, 247001 (2021).
  • Cea et al. [2022] T. Cea, P. A. Pantaleón, V. T. Phong, and F. Guinea, Superconductivity from repulsive interactions in rhombohedral trilayer graphene: A Kohn-Luttinger-like mechanism, Phys. Rev. B 105, 075432 (2022).
  • Qin et al. [2023] W. Qin, C. Huang, T. Wolf, N. Wei, I. Blinov, and A. H. MacDonald, Functional renormalization group study of superconductivity in rhombohedral trilayer graphene, Phys. Rev. Lett. 130, 146001 (2023).
  • Ghazaryan et al. [2023] A. Ghazaryan, T. Holder, E. Berg, and M. Serbyn, Multilayer graphenes as a platform for interaction-driven physics and topological superconductivity, Phys. Rev. B 107, 104502 (2023).
  • Chatterjee et al. [2022] S. Chatterjee, T. Wang, E. Berg, and M. P. Zaletel, Inter-valley coherent order and isospin fluctuation mediated superconductivity in rhombohedral trilayer graphene, Nat. Commun. 13, 6013 (2022).
  • Dong et al. [2023] Z. Dong, P. A. Lee, and L. S. Levitov, Signatures of cooper pair dynamics and quantum-critical superconductivity in tunable carrier bands, Proc. Natl. Acad. Sci. U.S.A. 120, e2305943120 (2023).
  • Zhou et al. [2022] H. Zhou, L. Holleis, Y. Saito, L. Cohen, W. Huynh, C. L. Patterson, F. Yang, T. Taniguchi, K. Watanabe, and A. F. Young, Isospin magnetism and spin-polarized superconductivity in Bernal bilayer graphene, Science 375, 774 (2022).
  • Zhang et al. [2023] Y. Zhang, R. Polski, A. Thomson, É. Lantagne-Hurtubise, C. Lewandowski, H. Zhou, K. Watanabe, T. Taniguchi, J. Alicea, and S. Nadj-Perge, Enhanced superconductivity in spin–orbit proximitized bilayer graphene, Nature 613, 268 (2023).
  • Chen et al. [2019] G. Chen, A. L. Sharpe, P. Gallagher, I. T. Rosen, E. J. Fox, L. Jiang, B. Lyu, H. Li, K. Watanabe, T. Taniguchi, J. Jung, Z. Shi, D. Goldhaber-Gordon, Y. Zhang, and F. Wang, Signatures of tunable superconductivity in a trilayer graphene moiré superlattice, Nature 572, 215 (2019).
  • Park et al. [2021] J. M. Park, Y. Cao, K. Watanabe, T. Taniguchi, and P. Jarillo-Herrero, Tunable strongly coupled superconductivity in magic-angle twisted trilayer graphene, Nature 590, 249 (2021).
  • Hao et al. [2021] Z. Hao, A. Zimmerman, P. Ledwith, E. Khalaf, D. H. Najafabadi, K. Watanabe, T. Taniguchi, A. Vishwanath, and P. Kim, Electric field–tunable superconductivity in alternating-twist magic-angle trilayer graphene, Science 371, 1133 (2021).
  • Kim et al. [2022] H. Kim, Y. Choi, C. Lewandowski, A. Thomson, Y. Zhang, R. Polski, K. Watanabe, T. Taniguchi, J. Alicea, and S. Nadj-Perge, Evidence for unconventional superconductivity in twisted trilayer graphene, Nature 606, 494 (2022).
  • Guo et al. [2024] Y. Guo, J. Pack, J. Swann, L. Holtzman, M. Cothrine, K. Watanabe, T. Taniguchi, D. Mandrus, K. Barmak, J. Hone, A. J. Millis, A. N. Pasupathy, and C. R. Dean, Superconductivity in twisted bilayer WSe2arXiv preprint arXiv:2406.03418  (2024).
  • Xia et al. [2024] Y. Xia, Z. Han, K. Watanabe, T. Taniguchi, J. Shan, and K. F. Mak, Unconventional superconductivity in twisted bilayer WSe2, arXiv preprint arXiv:2405.14784  (2024).