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

Emergence of multiple Higgs modes due to spontaneous breakdown of a 2\mathbb{Z}_{2} symmetry in a superconductor

Shunji Tsuchiya [email protected] Department of Physics, Chuo University, 1-13-27 Kasuga, Bunkyo-ku, Tokyo 112-8551, Japan
Abstract

We study the Higgs mode in a Bardeen-Cooper-Schrieffer (BCS) superconductor. Motivated by the observation that U(1) symmetry of the BCS Hamiltonian is not essential for the Higgs mode, we study the Ising-like Hamiltonian in the pseudospin representation. We show that the Higgs mode emerges as the lowest excited state of the Ising-like Hamiltonian due to spontaneous breakdown of 2\mathbb{Z}_{2} symmetry under the time-reversal operation 𝒯\mathcal{T} in the pseudospin space. We further predict the existence of multiple Higgs modes that have quantized energy 2(n+1)Δ02(n+1)\Delta_{0} (0nNkF0\leq n\leq N_{k_{F}}), where Δ0\Delta_{0} is the superconducting gap, nn is an integer, and NkFN_{k_{F}} is the number of states on the Fermi surface.

Introduction.– Higgs modes and Nambu-Goldstone (NG) modes are ubiquitous collective excitations that arise in systems with spontaneous symmetry breaking anderson-58 ; nambu-60 ; goldstone-61 ; goldstone-62 ; higgs-64 . They are amplitude and phase modes of the order parameter and appear in a broad class of systems in condensed-matter and particle physics. Stimulated by the discovery of the Higgs boson in the Standard Model cms-12 , interests in Higgs modes in condensed-matter physics have grown recently pekker-15 ; shimano-20 . Higgs modes have been observed in various condensed-matter systems including superconductors sooryakumar-81 ; matsunaga-13 ; matsunaga-14 ; sherman-15 ; katsumi-18 ; krull-16 ; chu-20 , superfluids giannetta-80 ; zavjalov-16 ; nguyen-19 , quantum spin systems ruegg-08 ; jain-17 ; hong-17 ; souliou-17 , charge-density-wave (CDW) materials demsar-99 ; yusupov-10 , and ultracold atomic gases bissbort-11 ; endres-12 ; leonard-17 ; behrle-18 .

Although Higgs modes are believed to emerge with NG modes when continuous symmetries are spontaneously broken, fundamental questions on Higgs modes remain to be understood. Namely, in contrast with NG modes, whose existence is predicted by the Goldstone theorem goldstone-62 , Higgs modes do not necessarily appear in systems exhibiting spontaneous breakdown of continuous symmetries. For instance, whereas a Higgs mode appears in a Bardeen-Cooper-Schrieffer (BCS) superconductor littlewood-81 , it disappears in a Bose-Einstein condensate (BEC) varma-02 , although spontaneous breakdown of U(1) symmetry occurs in both systems and, furthermore, the former continuously evolves to the latter in the BCS-BEC crossover phenomenon leggett-80 ; nozieres-85 ; ohashi-02 ; regal-04 .

Particle-hole (p-h) symmetry is a crucial condition for the emergence of Higgs modes. In superconductors, for instance, Higgs modes appear when the fermion energy dispersion is p-h symmetric engelbrecht-97 ; varma-02 ; tsuchiya-13 ; cea-15 . In a bosonic superfluid in an optical lattice, the system exhibits the approximate p-h symmetry in the vicinity of the tip of the Mott lobe, where a Higgs mode appears sachdev-98 ; endres-12 ; nakayama-15 ; diliberto-18 ; nakayama-19 .

In the previous work tsuchiya-18 , we have shown that the BCS Hamiltonian for a superconductor with a p-h symmetric fermion energy dispersion has the non-trivial 2\mathbb{Z}_{2} symmetries under discrete operations. We refer to them as the charge-conjugation (𝒞\mathcal{C}), parity (𝒫\mathcal{P}), and time-reversal (𝒯\mathcal{T}) operations in analogy with the corresponding ones in the relativistic field theory lee-81 . We found that 𝒯\mathcal{T} and 𝒫\mathcal{P} are spontaneously broken, while 𝒞\mathcal{C} is unbroken in the BCS ground state. We further conjectured that the emergence of the Higgs mode may be a consequence of the spontaneous breakdown of 2\mathbb{Z}_{2} symmetry under 𝒯\mathcal{T}.

In this Letter, extending the previous work, we establish that the Higgs mode emerges in a superconductor due to spontaneous breakdown of 2\mathbb{Z}_{2} symmetry under 𝒯\mathcal{T}. Motivated by the observation that U(1) symmetry is not essential for the Higgs mode, we study the Ising-like Hamiltonian in the pseudospin representation I\mathcal{H}_{I} derived from the BCS Hamiltonian, which exhibits 2\mathbb{Z}_{2} symmetries under 𝒞\mathcal{C}, 𝒫\mathcal{P}, and 𝒯\mathcal{T}. We show that the Higgs mode appears as the lowest excited state of I\mathcal{H}_{I} due to spontaneous breakdown of 2\mathbb{Z}_{2} symmetry under 𝒯\mathcal{T} in the ground state. Furthermore, we predict the existence of the multiple Higgs modes that have quantized energy 2(n+1)Δ02(n+1)\Delta_{0} (0nNkF0\leq n\leq N_{k_{F}}) and include the original one (n=0n=0), where Δ0\Delta_{0} is the superconducting (SC) gap, nn is an integer, and NkFN_{k_{F}} is the number of states on the Fermi surface (FS). Our main results are schematically illustrated in Fig. 1 (a).

Refer to caption
Figure 1: (a) Schematic illustration of the effective double-well potential for I\mathcal{H}_{I}. |Ψ\ket{\Psi} (|Ψ¯\ket{\bar{\Psi}}) is the SC ground state for the gap function Δ0\Delta_{0} (Δ0-\Delta_{0}). The multiple Higgs modes with excitation energy 2(n+1)Δ02(n+1)\Delta_{0} (0nNkF0\leq n\leq N_{k_{F}}) emerge due to spontaneous breakdown of 2\mathbb{Z}_{2} symmetry under 𝒯\mathcal{T} in the ground state |Ψ\ket{\Psi}. (b) Pseudospin configuration for the SC ground state with broken 2\mathbb{Z}_{2} symmetry under 𝒯\mathcal{T}. The pseudospins for |Ψ\ket{\Psi} (|Ψ¯\ket{\bar{\Psi}}) turn towards positive (negative) xx-direction as 𝒌\bm{k} evolves from below to above the FS. We denote 𝑺𝒌=Ψ|𝑺𝒌|Ψ\langle\bm{S}_{\bm{k}}\rangle=\bra{\Psi}{\bm{S}}_{\bm{k}}\ket{\Psi} and ¯𝑺𝒌=Ψ¯|𝑺𝒌|Ψ¯\langle\bm{\bar{}}{\bm{S}}_{\bm{k}}\rangle=\bra{\bar{\Psi}}{\bm{S}}_{\bm{k}}\ket{\bar{\Psi}}.

Particle-hole symmetry and Higgs modes.–To illustrate the condition in which the Higgs mode is well-defined in a superconductor, let us first discuss the relation between p-h symmetry and the Higgs mode. For simplicity, we consider the BCS Hamiltonian in the pseudospin representation anderson-58 ; tsuchiya-18

BCS\displaystyle\mathcal{H}_{\rm BCS} =\displaystyle= K+XY,K=𝒌2ξ𝒌Sz𝒌,\displaystyle\mathcal{H}_{K}+\mathcal{H}_{XY}~{},\quad\mathcal{H}_{K}=\sum_{\bm{k}}2\xi_{\bm{k}}S_{z\bm{k}}~{}, (1)
XY\displaystyle\mathcal{H}_{XY} =\displaystyle= g𝒌,𝒌(Sx𝒌Sx𝒌+Sy𝒌Sy𝒌),\displaystyle-g\sum_{\bm{k},\bm{k}^{\prime}}(S_{x\bm{k}}S_{x\bm{k}^{\prime}}+S_{y\bm{k}}S_{y\bm{k}^{\prime}})~{}, (2)

where 𝑺𝒌=(Sx𝒌,Sy𝒌,Sz𝒌)\bm{S}_{\bm{k}}=(S_{x\bm{k}},S_{y\bm{k}},S_{z\bm{k}}) is the spin-1/2 pseudospin operator and Sμ=𝒌Sμ𝒌S_{\mu}=\sum_{\bm{k}}S_{\mu\bm{k}} is the total spin. We set =1\hbar=1 throughout the Letter. The fermion vacuum corresponds to the spin-down state (|0𝒌=|𝒌\ket{0}_{\bm{k}}=\ket{\downarrow}_{\bm{k}}) and the fully occupied state to the spin-up state (c𝒌c𝒌|0𝒌=|𝒌c_{\bm{k}\uparrow}^{\dagger}c_{-\bm{k}\downarrow}^{\dagger}\ket{0}_{\bm{k}}=\ket{\uparrow}_{\bm{k}}), where c𝒌sc_{\bm{k}s}^{\dagger} is the creation operator for a fermion with momentum 𝒌\bm{k} and spin ss(=,=\uparrow,\downarrow). K\mathcal{H}_{K} is the kinetic energy term and XY\mathcal{H}_{XY} is the interaction term. Here, ξ𝒌=ε𝒌μ\xi_{\bm{k}}=\varepsilon_{\bm{k}}-\mu is the energy dispersion of a fermion measured from the chemical potential μ\mu and g(>0)g(>0) is the coupling constant for the attractive interaction between fermions. We do not specify the form of ε𝒌\varepsilon_{\bm{k}} for generality of argument.

BCS\mathcal{H}_{\rm BCS} has rotational U(1) symmetry about the zz-axis in the pseudospin space that reflects U(1) gauge symmetry. It is spontaneously broken in the SC state anderson-58 ; nambu-60 . When ξ𝒌\xi_{\bm{k}} satisfies the condition for the p-h symmetry

ξ𝒌¯=ξ𝒌,-\xi_{\underline{\bm{k}}}=\xi_{\bm{k}}~{}, (3)

BCS\mathcal{H}_{\rm BCS} has additional 2\mathbb{Z}_{2} symmetries under 𝒞\mathcal{C} and 𝒯\mathcal{T} tsuchiya-18 . Here, 𝒌\bm{k} and 𝒌¯\underline{\bm{k}} are a pair of wave vectors that are located on the opposite side of the FS with the same distance from it (See Figs. 1 (a)-(c) in Ref. tsuchiya-18 ). 𝒞\mathcal{C}, 𝒫\mathcal{P}, and 𝒯\mathcal{T} are defined as

𝒞=𝒌σx𝒌,=𝒌f𝒌,𝒌¯,\displaystyle\mathcal{C}={\mathcal{F}}\prod_{\bm{k}}\sigma_{x\bm{k}}~{},\quad\mathcal{F}=\sideset{}{{}^{{}^{\prime}}}{\prod}_{\bm{k}}f_{\bm{k},\underline{\bm{k}}}~{}, (4)
𝒯=UT𝒦,UT=𝒌(iσy𝒌),\displaystyle\mathcal{T}=U_{T}\mathcal{K}~{},\quad U_{T}=\mathcal{F}\prod_{\bm{k}}(-i\sigma_{y\bm{k}})~{}, (5)
𝒫=𝒌σz𝒌,\displaystyle\mathcal{P}=\prod_{\bm{k}}\sigma_{z\bm{k}}~{}, (6)

where 𝒌ξ𝒌>0\prod_{\bm{k}}^{\prime}\equiv\prod_{\xi_{\bm{k}}>0}, σμ𝒌2Sμ𝒌\sigma_{\mu\bm{k}}\equiv 2S_{\mu\bm{k}}, f𝒌𝒌¯f_{\bm{k}\underline{\bm{k}}} is a swapping operator between the states of 𝒌\bm{k} and 𝒌¯\underline{\bm{k}}, and 𝒦\mathcal{K} is the complex conjugation operator. Note that 𝒫\mathcal{P} represents a π\pi rotation about the zz-axis that is an element of U(1). 𝒞𝒫𝒯\mathcal{CPT} and all other permutations of 𝒞\mathcal{C}, 𝒫\mathcal{P}, and 𝒯\mathcal{T} are exact symmetries. 2\mathbb{Z}_{2} symmetries under 𝒯\mathcal{T} and 𝒫\mathcal{P} are spontaneously broken with U(1) in the SC state, while that under 𝒞\mathcal{C} is unbroken tsuchiya-18 .

On the other hand, the Higgs mode appears as an amplitude mode only when ξ𝒌\xi_{\bm{k}} is p-h symmetric, i.e., it satisfies Eq. (3) varma-02 ; tsuchiya-18 . In the classical spin analysis, for a gap function in the ground state Δ0>0\Delta_{0}>0, amplitude and phase fluctuations of the gap function are proportional to those of the xx and yy components of the total spin as δΔδSx\delta\Delta\propto\delta S_{x} and δθδSy\delta\theta\propto\delta S_{y}, respectively. δΔ\delta\Delta and δθ\delta\theta are uncoupled and pure amplitude oscillations of the Higgs mode are allowed only when ξ𝒌\xi_{\bm{k}} is p-h symmetric, because the off-diagonal element of the dynamical spin susceptibility χxy(ω)\chi_{xy}(\omega) vanishes due to the opposite parity of SxS_{x} and SyS_{y} under 𝒞\mathcal{C} tsuchiya-18 ; supplement . When ξ𝒌\xi_{\bm{k}} is not p-h symmetric, finite χxy(ω)\chi_{xy}(\omega) couples amplitude and phase fluctuations.

Moreover, the Higgs mode is stable only when ξ𝒌\xi_{\bm{k}} is p-h symmetric tsuchiya-18 . It is prohibited to decay into single-particle excitations despite its degeneracy with the lower edge of the single-particle continuum at 2Δ02\Delta_{0}, because the Higgs mode has even parity and the single-particle excitations at 2Δ02\Delta_{0} have odd parity under 𝒞\mathcal{C}. When ξ𝒌\xi_{\bm{k}} is not p-h symmetric, however, the Higgs mode merges with the single-particle continuum and suffers from strong damping due to decay into single-particle excitations. Thus, the Higgs mode is well-defined only when 2\mathbb{Z}_{2} under 𝒯\mathcal{T} is spontaneously broken with U(1) in the SC state.

Minimal Hamiltonian for Higgs modes.– To argue the origin of the Higgs mode, we focus on a p-h symmetric system, in which the Higgs mode is well-defined, and assume Eq. (3) in the rest of the Letter.

In the classical spin analysis, one finds that the yy component of the interaction term Y=gSy2\mathcal{H}_{Y}=-gS_{y}^{2} may be irrelevant to the Higgs mode, because the Higgs mode involves out-of-phase oscillations of δSy𝒌\delta S_{y\bm{k}} and δSy𝒌¯\delta S_{y\underline{\bm{k}}}, whereas Y\mathcal{H}_{Y} induces in-phase oscillations of them anderson-58 ; tsuchiya-18 ; supplement . Furthermore, the analysis based on the Holstein-Primakoff (H-P) theory suggests that Y\mathcal{H}_{Y} is not necessary for constructing the creation operator of the Higgs mode tsuchiya-18 . Motivated by these observations, we neglect Y\mathcal{H}_{Y} and study the Ising-like Hamiltonian I\mathcal{H}_{I}:

I=K+X,\displaystyle\mathcal{H}_{I}=\mathcal{H}_{K}+\mathcal{H}_{X}~{}, (7)
X=g𝒌,𝒌Sx𝒌Sx𝒌=gSx2.\displaystyle\mathcal{H}_{X}=-g\sum_{\bm{k},\bm{k}^{\prime}}S_{x\bm{k}}S_{x\bm{k}^{\prime}}=-gS_{x}^{2}~{}. (8)

Note that I\mathcal{H}_{I} is invariant under 𝒞\mathcal{C}, 𝒫\mathcal{P}, and 𝒯\mathcal{T}, despite it loses U(1) symmetry. We derive the Higgs mode from I\mathcal{H}_{I} to demonstrate that the emergence of the Higgs mode is not due to spontaneous breakdown of U(1) but to that of 2\mathbb{Z}_{2} under 𝒯\mathcal{T}.

Discrete symmetries of I\mathcal{H}_{I}.– In addition to 2\mathbb{Z}_{2} symmetries under 𝒞\mathcal{C}, 𝒫\mathcal{P}, and 𝒯\mathcal{T}, I\mathcal{H}_{I} has 2\mathbb{Z}_{2} symmetries under local charge-conjugation operations in momentum space 𝒞𝒌𝒌¯=f𝒌,𝒌¯σx𝒌σx𝒌¯{\mathcal{C}}_{\bm{k}\underline{\bm{k}}}=f_{\bm{k},\underline{\bm{k}}}\sigma_{x\bm{k}}\sigma_{x\underline{\bm{k}}} and 𝒞𝒌F=σx𝒌F{\mathcal{C}}_{\bm{k}_{F}}=\sigma_{x\bm{k}_{F}}, where 𝒞\mathcal{C} can be written as 𝒞=𝒌𝒞𝒌𝒌¯𝒌F𝒞𝒌F\mathcal{C}=\prod_{\bm{k}}^{\prime}\mathcal{C}_{\bm{k}\underline{\bm{k}}}\otimes\prod_{\bm{k}_{F}}{\mathcal{C}}_{\bm{k}_{F}}. To prove the invariance under 𝒞𝒌𝒌¯{\mathcal{C}}_{\bm{k}\underline{\bm{k}}}, it is convenient to introduce a pseudospin operator

Sμ𝒌𝒌¯=Sμ𝒌(1)δμ,xSμ𝒌¯,(μ=x,y,z),\displaystyle S_{\mu\bm{k}\underline{\bm{k}}}=S_{\mu\bm{k}}-(-1)^{\delta_{\mu,x}}S_{\mu\underline{\bm{k}}}~{},~{}(\mu=x,y,z), (9)

where 𝒌\bm{k} is above the FS (ξ𝒌>0\xi_{\bm{k}}>0). It commutes with 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}.

The kinetic energy term can be written as

K=𝒌2ξ𝒌Sz𝒌𝒌¯,\displaystyle\mathcal{H}_{K}=\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}}2\xi_{\bm{k}}S_{z\bm{k}\underline{\bm{k}}}~{}, (10)

where 𝒌ξ𝒌>0\sum_{\bm{k}}^{\prime}\equiv\sum_{\xi_{\bm{k}}>0}. We obtain [K,𝒞𝒌𝒌¯]=[K,𝒞𝒌F]=0[\mathcal{H}_{K},{\mathcal{C}}_{\bm{k}\underline{\bm{k}}}]=[\mathcal{H}_{K},{\mathcal{C}}_{\bm{k}_{F}}]=0 from Eq. (10). Since Sx=𝒌Sx𝒌𝒌¯+𝒌FSx𝒌FS_{x}=\sum_{\bm{k}}^{\prime}S_{x\bm{k}\underline{\bm{k}}}+\sum_{\bm{k}_{F}}S_{x\bm{k}_{F}} commutes with either 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}} and 𝒞𝒌F\mathcal{C}_{\bm{k}_{F}}, we also obtain [X,𝒞𝒌𝒌¯]=[X,𝒞𝒌F]=0[\mathcal{H}_{X},{\mathcal{C}}_{\bm{k}\underline{\bm{k}}}]=[\mathcal{H}_{X},{\mathcal{C}}_{\bm{k}_{F}}]=0 and [I,𝒞𝒌𝒌¯]=[I,𝒞𝒌F]=0[\mathcal{H}_{I},{\mathcal{C}}_{\bm{k}\underline{\bm{k}}}]=[\mathcal{H}_{I},{\mathcal{C}}_{\bm{k}_{F}}]=0. Consequently, I\mathcal{H}_{I} has 2\mathbb{Z}_{2} symmetry under each of 𝒞𝒌𝒌¯{\mathcal{C}}_{\bm{k}\underline{\bm{k}}} and 𝒞𝒌F{\mathcal{C}}_{\bm{k}_{F}}. Y\mathcal{H}_{Y} and BCS\mathcal{H}_{\rm BCS}, in contrast, commute with neither 𝒞𝒌𝒌¯{\mathcal{C}}_{\bm{k}\underline{\bm{k}}} nor 𝒞𝒌F{\mathcal{C}}_{\bm{k}_{F}}.

Discrete symmetries in low-energy states of I\mathcal{H}_{I}.– We next study the symmetry of the ground state of I\mathcal{H}_{I}. The pseudospin configuration of it is shown in Fig. 1 (b): The pseudospins turn smoothly from up to down towards either positive or negative xx-direction as 𝒌\bm{k} evolves from below to above the FS, where the gap function ΔgSx\Delta\equiv g\langle S_{x}\rangle is positive and negative in the former and latter cases, respectively. One of them being chosen spontaneously, 2\mathbb{Z}_{2} under 𝒫\mathcal{P} is broken in the ground state. It leads to breakdown of 2\mathbb{Z}_{2} under 𝒯\mathcal{T} due to unbroken 2\mathbb{Z}_{2} under 𝒞\mathcal{C}, which is shown below, and the exact symmetry under 𝒞𝒫𝒯\mathcal{C}\mathcal{P}\mathcal{T}. The energy landscape of I\mathcal{H}_{I} is effectively described by the double-well potential, as schematically illustrated in Fig. 1 (a).

Let us show that 2\mathbb{Z}_{2} under each of 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}, 𝒞𝒌F\mathcal{C}_{\bm{k}_{F}}, and 𝒞\mathcal{C} is unbroken in the ground state. If we introduce a gap function Δ=Δ0>0\Delta=\Delta_{0}>0, I\mathcal{H}_{I} reduces to the mean-field (MF) Hamiltonian

MF=𝒌2E𝒌Sz𝒌=𝒌2E𝒌Sz𝒌𝒌¯2Δ0𝒌FSx𝒌F,\displaystyle\mathcal{H}_{\rm MF}=\sum_{\bm{k}}2E_{\bm{k}}S^{\prime}_{z\bm{k}}=\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}}2E_{\bm{k}}S^{\prime}_{z\bm{k}\underline{\bm{k}}}-2\Delta_{0}\sum_{\bm{k}_{F}}S_{x\bm{k}_{F}}, (11)

where E𝒌=ξ𝒌2+Δ02E_{\bm{k}}=\sqrt{\xi_{\bm{k}}^{2}+\Delta_{0}^{2}} is the dispersion of a single-particle excitation referred to as a bogolon. 𝑺𝒌\bm{S}^{\prime}_{\bm{k}} is the pseudospin operator for a pair of bogolons with |𝒌\ket{\uparrow^{\prime}}_{\bm{k}} and |𝒌\ket{\downarrow^{\prime}}_{\bm{k}} being the eigenstates of Sz𝒌S^{\prime}_{z\bm{k}} anderson-58 ; tsuchiya-18 ; supplement . Note that I\mathcal{H}_{I} and BCS\mathcal{H}_{\rm BCS} reduce to the same MF Hamiltonian for a real gap function. Here, we have introduced a spin operator

Sμ𝒌𝒌¯=Sμ𝒌(1)δz,μSμ𝒌¯,(μ=x,y,z),\displaystyle S_{\mu\bm{k}\underline{\bm{k}}}^{\prime}=S_{\mu\bm{k}}^{\prime}-(-1)^{\delta_{z,\mu}}S_{\mu\underline{\bm{k}}}^{\prime}~{},~{}(\mu=x,y,z), (12)

where 𝒌\bm{k} is above the FS (ξ𝒌>0\xi_{\bm{k}}>0). It commutes with 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}. Using Eq. (11), we obtain [MF,𝒞𝒌𝒌¯]=[MF,𝒞𝒌F]=0[\mathcal{H}_{\rm MF},\mathcal{C}_{\bm{k}\underline{\bm{k}}}]=[\mathcal{H}_{\rm MF},\mathcal{C}_{\bm{k}_{F}}]=0. Consequently, 2\mathbb{Z}_{2} symmetry under each of 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}} and 𝒞𝒌F\mathcal{C}_{\bm{k}_{F}} as well as under 𝒞\mathcal{C} is unbroken in the ground state of I\mathcal{H}_{I}.

The eigenstates of MF\mathcal{H}_{\rm MF} can be characterized by parity under 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}} and 𝒞𝒌F\mathcal{C}_{\bm{k}_{F}}. The ground state of Eq. (11) |Ψ=𝒌|1𝒌𝒌¯𝒌F|+𝒌F\ket{\Psi}=\prod_{\bm{k}}^{\prime}\ket{-1^{\prime}}_{\bm{k}\underline{\bm{k}}}\otimes\prod_{\bm{k}_{F}}\ket{+}_{\bm{k}_{F}} has even parity for all 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}} and 𝒞𝒌F\mathcal{C}_{\bm{k}_{F}}, where |±𝒌(|𝒌±|𝒌)/2\ket{\pm}_{\bm{k}}\equiv(\ket{\uparrow}_{\bm{k}}\pm\ket{\downarrow}_{\bm{k}})/\sqrt{2}. Here, the product states |s,t𝒌𝒌¯=|s𝒌|t𝒌¯\ket{s^{\prime},t^{\prime}}_{\bm{k}\underline{\bm{k}}}=\ket{s^{\prime}}_{\bm{k}}\otimes\ket{t^{\prime}}_{\underline{\bm{k}}} (s,t=,s,t=\uparrow,\downarrow) are decomposed into the even and odd parity states under 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}, where the former consist of the spin-1 triplet states |m𝒌𝒌¯\ket{m^{\prime}}_{\bm{k}\underline{\bm{k}}} (m=1,0,1m=1,0,-1) for Sz𝒌𝒌¯S^{\prime}_{z\bm{k}\underline{\bm{k}}} and the latter the spin-0 singlet state |0~𝒌𝒌¯\ket{\tilde{0}^{\prime}}_{\bm{k}\underline{\bm{k}}} supplement . The ground state |Ψ¯=𝒯|Ψ\ket{\bar{\Psi}}=\mathcal{T}\ket{\Psi} for Δ=Δ0\Delta=-\Delta_{0} has even and odd parity under 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}} and 𝒞𝒌F\mathcal{C}_{\bm{k}_{F}}, respectively.

Parity of single-particle states plays a key role for the stability of the Higgs mode discussed below. |e𝒌𝒌¯l=|l𝒌𝒌¯𝒌𝒌|1𝒌𝒌¯𝒌F|+𝒌F\ket{e_{\bm{k}\underline{\bm{k}}}^{l}}=\ket{l^{\prime}}_{\bm{k}\underline{\bm{k}}}\otimes\prod_{\bm{k}^{\prime}\neq\bm{k}}^{\prime}\ket{-1^{\prime}}_{\bm{k}^{\prime}\underline{\bm{k}}^{\prime}}\otimes\prod_{\bm{k}_{F}}\ket{+}_{\bm{k}_{F}} (l=0,0~l=0,\tilde{0}) are degenerate single-particle states with excitation energy 2E𝒌2E_{\bm{k}} that form the two-particle continuum. They have even (l=0l=0) and odd (l=0~l=\tilde{0}) parity under 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}. They both have even parity under 𝒞𝒌F\mathcal{C}_{\bm{k}_{F}}. |e𝒌F=|𝒌F𝒌|1𝒌𝒌¯𝒌F𝒌F|+𝒌F\ket{e_{\bm{k}_{F}}}=\ket{-}_{\bm{k}_{F}}\otimes\prod_{\bm{k}}^{\prime}\ket{-1^{\prime}}_{\bm{k}\underline{\bm{k}}}\otimes\prod_{\bm{k}_{F}^{\prime}\neq\bm{k}_{F}}\ket{+}_{\bm{k}_{F}^{\prime}} forms the lower edge of the two-particle continuum at 2Δ02\Delta_{0} that is degenerate with the Higgs mode. It has odd parity under 𝒞𝒌F\mathcal{C}_{\bm{k}_{F}} and 𝒞\mathcal{C}.

Emergence of Higgs mode due to breakdown of 2\mathbb{Z}_{2} under 𝒯\mathcal{T}.– To derive the Higgs mode, we apply the H-P theory to I\mathcal{H}_{I} holstein-40 ; supplement . The second-order term in spin fluctuation can be diagonalized as I(2)=ωHβHβH\mathcal{H}_{I}^{(2)}=\omega_{\rm H}\beta_{\rm H}^{\dagger}\beta_{\rm H}. The collective mode has excitation energy ωH=2Δ0\omega_{\rm H}=2\Delta_{0} and its creation operator is given as

βH=𝒌ξ𝒌E𝒌(S𝒌𝒌¯+2|Δ0|2E𝒌+S𝒌𝒌¯2|Δ0|+2E𝒌),\displaystyle\beta_{\rm H}^{\dagger}=\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}}\frac{\xi_{\bm{k}}}{E_{\bm{k}}}\left(\frac{S_{\bm{k}\underline{\bm{k}}}^{{}^{\prime}+}}{2|\Delta_{0}|-2E_{\bm{k}}}+\frac{S_{\bm{k}\underline{\bm{k}}}^{{}^{\prime}-}}{2|\Delta_{0}|+2E_{\bm{k}}}\right)~{}, (13)

where S𝒌𝒌¯±=Sx𝒌𝒌¯±iSy𝒌𝒌¯S_{\bm{k}\underline{\bm{k}}}^{{}^{\prime}\pm}=S^{\prime}_{x\bm{k}\underline{\bm{k}}}\pm iS^{\prime}_{y\bm{k}\underline{\bm{k}}}. We find that βH\beta_{\rm H}^{\dagger} coincides exactly with the creation operator of the Higgs mode derived from BCS\mathcal{H}_{\rm BCS} (See Eq. (27) in Ref. tsuchiya-18 ). Thus, βH\beta_{\rm H}^{\dagger} represents the Higgs mode. The NG mode does not appear, because I\mathcal{H}_{I} does not have U(1) symmetry.

The emergence of the Higgs mode from I\mathcal{H}_{I} clearly shows that U(1) symmetry breaking is not essential for it. Now, the origin of the Higgs mode can be attributed to the spontaneous breakdown of 𝒯\mathcal{T} or 𝒫\mathcal{P}. However, the Higgs mode disappears, for example, in the BEC regime of the BCS-BEC crossover, despite 𝒫\mathcal{P} is broken with U(1) in the SC phase engelbrecht-97 . Therefore, the emergence of the Higgs mode is considered due to breakdown of 2{\mathbb{Z}}_{2} symmetry under 𝒯\mathcal{T}. This conclusion, together with the fact that the Higgs mode arises from BCS\mathcal{H}_{\rm BCS}, indicates that the SC phase transition for BCS\mathcal{H}_{\rm BCS} with a p-h symmetric ξ𝒌\xi_{\bm{k}} is associated with breakdown of not U(1), but 2{\mathbb{Z}}_{2} under 𝒯\mathcal{T}. Namely, BCS\mathcal{H}_{\rm BCS} effectively reduces to I\mathcal{H}_{I} for the SC phase transition when ξ𝒌\xi_{\bm{k}} is p-h symmetric.

The Higgs mode is uncoupled with the single-particle states |e𝒌F\ket{e_{\bm{k}_{F}}} due to their opposite parity under 𝒞𝒌F\mathcal{C}_{\bm{k}_{F}} despite their degeneracy at 2Δ02\Delta_{0}. The Higgs mode has even parity for all 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}} and 𝒞𝒌F\mathcal{C}_{\bm{k}_{F}} because [𝒞𝒌𝒌¯,βH]=[𝒞𝒌F,βH]=0[\mathcal{C}_{\bm{k}\underline{\bm{k}}},\beta_{\rm H}^{\dagger}]=[\mathcal{C}_{\bm{k}_{F}},\beta_{\rm H}^{\dagger}]=0. A single Higgs mode βH|Ψ\beta_{\rm H}^{\dagger}\ket{\Psi} in fact consists of the single-particle states |e𝒌𝒌¯0\ket{e_{\bm{k}\underline{\bm{k}}}^{0}} supplement . Thus, the Higgs mode is a stable excitation of I\mathcal{H}_{I}.

Strong-coupling perturbation theory.

Refer to caption
Figure 2: Low-lying energy eigenstates of X\mathcal{H}_{X} and their energy shifts by K\mathcal{H}_{K}. We denote |m𝒌,n𝒌,|mx𝒌𝒌¯|nx𝒌𝒌¯𝒌′′𝒌,𝒌|1x𝒌′′𝒌¯′′𝒌F|+𝒌F\ket{m_{\bm{k}},n_{\bm{k}^{\prime}},\cdots}\equiv\ket{m_{x}}_{\bm{k}\underline{\bm{k}}}\otimes\ket{n_{x}}_{\bm{k}^{\prime}\underline{\bm{k}^{\prime}}}\otimes\cdots\otimes\prod^{\prime}_{\bm{k}^{\prime\prime}\neq\bm{k},\bm{k}^{\prime}}\ket{1_{x}}_{\bm{k}^{\prime\prime}\underline{\bm{k}}^{\prime\prime}}\otimes\prod_{\bm{k}_{F}}\ket{+}_{\bm{k}_{F}} (m,n=0,1m,n=0,-1). The eigenstates of X\mathcal{H}_{X} can be decomposed into the blocks represented by squares, where each block is characterized by the parity under 𝒞𝒌F\mathcal{C}_{\bm{k}_{F}}. The multiple Higgs modes |Φ\ket{\Phi} and σz𝒌F|Φ\sigma_{z\bm{k}_{F}}\ket{\Phi} are the bound states formed at the lower edges of the continuum in each block shown as the gray regions.

Having shown the emergence of the Higgs mode due to breakdown of 2\mathbb{Z}_{2} in the weak-coupling theory, we demonstrate it as well in a strong-coupling approach, where we treat K\mathcal{H}_{K} as a perturbation to the unperturbed Hamiltonian X\mathcal{H}_{X}. It is valid when NgNg is much larger than the band width, i.e., Ngmax|ξ𝒌|Ng\gg{\rm max}|\xi_{\bm{k}}|, where NN is the total number of wave vectors. Given that the Higgs mode has even parity for all 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}, we treat 𝑺𝒌𝒌¯\bm{S}_{\bm{k}\underline{\bm{k}}} as a spin-1 operator to take into account only even parity states under 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}.

Each eigenstate of MF\mathcal{H}_{\rm MF} has a corresponding eigenstate of X\mathcal{H}_{X}, where the former transforms to the latter in the strong-coupling limit. 2\mathbb{Z}_{2} under 𝒯\mathcal{T} is broken in the unperturbed ground states |G=𝒌|1x𝒌𝒌¯𝒌F|+𝒌F\ket{G}=\prod^{\prime}_{\bm{k}}\ket{1_{x}}_{\bm{k}\underline{\bm{k}}}\otimes\prod_{\bm{k}_{F}}\ket{+}_{\bm{k}_{F}} and |G¯=𝒯|G\ket{\bar{G}}=\mathcal{T}\ket{G}, where |mx𝒌𝒌¯\ket{m_{x}}_{\bm{k}\underline{\bm{k}}} (m=1,0,1m=1,0,-1) denote the spin-1 triplet states for Sx𝒌𝒌¯S_{x\bm{k}\underline{\bm{k}}} supplement . They have gap functions Δ=Ng/2Δ~0\Delta=Ng/2\equiv\tilde{\Delta}_{0} and Δ=Δ~0\Delta=-\tilde{\Delta}_{0}, respectively. We consider perturbation to |G\ket{G} and the low-lying states above it shown in Fig. 2.

The energy eigenstates of X\mathcal{H}_{X} can be decomposed into the blocks characterized by the parity under 𝒞𝒌F\mathcal{C}_{\bm{k}_{F}}, which are represented as the squares in Fig. 2. K\mathcal{H}_{K} has no matrix elements between states in different blocks. If we focus on a block that has odd parity under 𝒞𝒌F\mathcal{C}_{\bm{k}_{F}} for a set of Fermi wavevectors {𝒌F}\{\bm{k}_{F}\}, each unperturbed state in this block can be obtained by operating 𝒵({𝒌F})𝒌F{𝒌F}σz𝒌F\mathcal{Z}(\{\bm{k}_{F}\})\equiv\prod_{\bm{k}_{F}^{\prime}\in\{\bm{k}_{F}\}}\sigma_{z\bm{k}_{F}^{\prime}} on a corresponding state in the block of |G\ket{G}. Flipping a pseudospin on the FS from |+𝒌F\ket{+}_{\bm{k}_{F}} to |𝒌F\ket{-}_{\bm{k}_{F}} by σz𝒌F\sigma_{z\bm{k}_{F}} yields a higher excited state as

X(σz𝒌F|ϕ)=(gSx(ϕ)2+ΔEϕ)(σz𝒌F|ϕ),\displaystyle\mathcal{H}_{X}(\sigma_{z\bm{k}_{F}}\ket{\phi})=(-gS_{x}(\phi)^{2}+\Delta E_{\phi})(\sigma_{z\bm{k}_{F}}\ket{\phi})~{}, (14)

where 𝒞𝒌F|ϕ=|ϕ\mathcal{C}_{\bm{k}_{F}}\ket{\phi}=\ket{\phi}, Sx|ϕ=Sx(ϕ)|ϕS_{x}\ket{\phi}=S_{x}(\phi)\ket{\phi}, and ΔEϕ=g(2Sx(ϕ)1)\Delta E_{\phi}=g(2S_{x}(\phi)-1). The energy spectrum of X\mathcal{H}_{X} in Fig. 2 is obtained by using Eq. (14).

The degeneracy among the first-excited states {|0𝒌}\{\ket{0_{\bm{k}}}\} is lifted by diagonalizing the second-order effective Hamiltonian effK(E(0𝒌)0X)1K\mathcal{H}_{\rm eff}\equiv\mathcal{H}_{K}(E^{0}_{(0_{\bm{k}})}-\mathcal{H}_{X})^{-1}\mathcal{H}_{K}, where E(0𝒌)0(N1)gE^{0}_{(0_{\bm{k}})}\equiv(N-1)g is the unperturbed energy of |0𝒌\ket{0_{\bm{k}}}. The effective Hamiltonian can be written as

eff=𝒌J𝒌|0𝒌0𝒌|𝒌𝒌J𝒌,𝒌|0𝒌0𝒌|,\displaystyle\mathcal{H}_{\rm eff}=\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}}J_{\bm{k}}\ket{0_{\bm{k}}}\bra{0_{\bm{k}}}-\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}\neq\bm{k}^{\prime}}J_{\bm{k},\bm{k}^{\prime}}\ket{0_{\bm{k}}}\bra{0_{\bm{k}^{\prime}}}~{}, (15)

where

J𝒌=2ξ𝒌2(N1)g,J𝒌,𝒌=4ξ𝒌ξ𝒌(N1)(N3)g.\displaystyle J_{\bm{k}}=\frac{2\xi_{\bm{k}}^{2}}{(N-1)g},\quad J_{\bm{k},\bm{k}^{\prime}}=\frac{4\xi_{\bm{k}}\xi_{\bm{k}^{\prime}}}{(N-1)(N-3)g}. (16)

Note that energy in Eq. (15) is measured from E(0𝒌)0ΔE𝒌FE^{0}_{(0_{\bm{k}})}-\Delta E_{\bm{k}_{F}}, where ΔE𝒌F=2𝒌ξ𝒌2(N3)g\Delta E_{\bm{k}_{F}}=\frac{2\sum_{\bm{k}^{\prime}}^{\prime}\xi_{\bm{k}^{\prime}}^{2}}{(N-3)g} is the second-order energy shift of σz𝒌F|G\sigma_{z\bm{k}_{F}}\ket{G} that corresponds to |e𝒌F\ket{e_{\bm{k}_{F}}}. The first term in Eq. (15) represents the energy continuum above the threshold E(0𝒌)0ΔE𝒌FE^{0}_{(0_{\bm{k}})}-\Delta E_{\bm{k}_{F}}, as shown in Fig. 2.

We consider the following state:

|Φ=𝒌1ξ𝒌|0𝒌.\ket{\Phi}=\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}}\frac{1}{\xi_{\bm{k}}}\ket{0_{\bm{k}}}. (17)

Using Eq. (15), we obtain

eff|Φ=𝒌C𝒌|0𝒌,C𝒌=J𝒌ξ𝒌+𝒌𝒌(J𝒌𝒌ξ𝒌).\displaystyle\mathcal{H}_{\rm eff}\ket{\Phi}=\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}}C_{\bm{k}}\ket{0_{\bm{k}}},~{}C_{\bm{k}}=\frac{J_{\bm{k}}}{\xi_{\bm{k}}}+\sum_{\bm{k}^{\prime}\neq\bm{k}}\left(-\frac{J_{\bm{k}\bm{k}^{\prime}}}{\xi_{\bm{k}^{\prime}}}\right). (18)

In the limit NNkFN\gg N_{k_{F}}, using 𝒌𝒌1N/2\sum_{\bm{k}^{\prime}\neq\bm{k}}^{\prime}1\simeq N/2, Eq. (18) reduces to eff|Φ=0\mathcal{H}_{\rm eff}|\Phi\rangle=0. Remarkably, |Φ|\Phi\rangle represents a bound state of {|0𝒌}\{\ket{0_{\bm{k}}}\} that is formed at the lower edge of the energy continuum at 2Δ~02\tilde{\Delta}_{0}. On the other hand, we find βH|Ψ|Φ\beta_{\rm H}^{\dagger}|\Psi\rangle\to|\Phi\rangle in the strong-coupling limit using βHi𝒌(Sy𝒌𝒌¯iSz𝒌𝒌¯)/ξ𝒌\beta_{\rm H}^{\dagger}\to-i\sum_{\bm{k}}^{\prime}(S_{y\bm{k}\underline{\bm{k}}}-iS_{z\bm{k}\underline{\bm{k}}})/\xi_{\bm{k}}. Thus, the Higgs mode |Φ\ket{\Phi} arises in the strong-coupling theory. |Φ\ket{\Phi} is indeed stable, because it is uncoupled with σz𝒌F|G\sigma_{z\bm{k}_{F}}\ket{G}. In the weak-coupling regime, since |0𝒌\ket{0_{\bm{k}}} corresponds to |e𝒌𝒌¯0\ket{e^{0}_{\bm{k}\underline{\bm{k}}}}, βH|Ψ\beta_{\rm H}^{\dagger}\ket{\Psi} is considered a bound state of {|e𝒌𝒌¯0}\{\ket{e^{0}_{\bm{k}\underline{\bm{k}}}}\} formed at 2Δ02\Delta_{0}.

Multiple Higgs modes.– We expect that a bound state analogous to the Higgs mode may be formed among the first excited states in each block. The effective Hamiltonian eff(n)\mathcal{H}^{(n)}_{\rm eff} for the degenerate first excited states {𝒵({𝒌F})|0𝒌}\{\mathcal{Z}(\{\bm{k}_{F}\})\ket{0_{\bm{k}}}\} in the block of 𝒵({𝒌F})|G\mathcal{Z}(\{\bm{k}_{F}\})\ket{G} is given by

eff(n)=𝒌J𝒌(n)𝒵|0𝒌0𝒌|𝒵𝒌𝒌J𝒌,𝒌(n)𝒵|0𝒌0𝒌|𝒵,\displaystyle\mathcal{H}_{\rm eff}^{(n)}=\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}}J^{(n)}_{\bm{k}}\mathcal{Z}\ket{0_{\bm{k}}}\bra{0_{\bm{k}}}\mathcal{Z}-\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}\neq\bm{k}^{\prime}}J^{(n)}_{\bm{k},\bm{k}^{\prime}}\mathcal{Z}\ket{0_{\bm{k}}}\bra{0_{\bm{k}^{\prime}}}\mathcal{Z}~{}, (19)

where

J𝒌(n)=2ξ𝒌2(N(2n+1))g,\displaystyle J^{(n)}_{\bm{k}}=\frac{2\xi_{\bm{k}}^{2}}{(N-(2n+1))g}~{}, (20)
J𝒌,𝒌(n)=4ξ𝒌ξ𝒌(N(2n+1))(N(2n+3))g.\displaystyle J^{(n)}_{\bm{k},\bm{k}^{\prime}}=\frac{4\xi_{\bm{k}}\xi_{\bm{k}^{\prime}}}{(N-(2n+1))(N-(2n+3))g}~{}. (21)

Here, nn is the number of wave vectors in {𝒌F}\{\bm{k}_{F}\}. In the limit NNkFN\gg N_{k_{F}}, we find that eff(n)\mathcal{H}^{(n)}_{\rm eff} has a Higgs-like bound state

|Φ(n)=𝒵({𝒌F})|Φ.\displaystyle\ket{\Phi^{(n)}}=\mathcal{Z}(\{\bm{k}_{F}\})\ket{\Phi}~{}. (22)

Thus, combining with the original Higgs mode |Φ(0)=|Φ\ket{\Phi^{(0)}}=\ket{\Phi}, the multiple Higgs modes |Φ(n)\ket{\Phi^{(n)}} (0nNkF0\leq n\leq N_{k_{F}}) have quantized energy 2(n+1)Δ~02(n+1)\tilde{\Delta}_{0}. They have odd parity under 𝒞𝒌F\mathcal{C}_{\bm{k}_{F}} for nn Fermi wavevectors in {𝒌F}\{\bm{k}_{F}\}. They should appear in the weak-coupling regime. They correspond to the states 𝒵({𝒌F})βH|Ψ\mathcal{Z}(\{\bm{k}_{F}\})\beta_{\rm H}^{\dagger}\ket{\Psi} with quantized energy 2(n+1)Δ02(n+1)\Delta_{0} in the weak-coupling regime.

These multiple Higgs modes may be observable in superconductors with CDW-order such as NbSe2 using Raman spectroscopy sooryakumar-81 . Their signature may appear as multiple resonance peaks at frequencies 2(n+1)Δ02(n+1)\Delta_{0} in Raman spectrum. It may be also possible to observe them using recently developed terahertz spectroscopy techniques matsunaga-13 .

Conclusions and conjectures.– We have shown that the Higgs mode emerges due to spontaneous breakdown of 2\mathbb{Z}_{2} under 𝒯\mathcal{T} in a superconductor. We further predicted the existence of the multiple Higgs modes that have quantized energy 2(n+1)Δ02(n+1)\Delta_{0} (0nNkF0\leq n\leq N_{k_{F}}).

The analysis in the present work has broad applicability in condensed-matter and particle physics. Multiple Higgs modes may appear in systems for which the BCS theory is applicable, such as fermionic superfluids, atomic nuclei, and quark matters. It would be reasonable to conjecture that Higgs modes in other condensed-matter systems, such as a bosonic superfluid in an optical lattice and quantum spin systems, also arise due to spontaneous breakdown of 2\mathbb{Z}_{2} symmetries. Extensions of the present work to these systems are left for the future.

Acknowledgements.
The author wishes to thank I. Danshita, D. Yamamoto, and R. Yoshii for fruitful discussions. This work is supported by the Japan Society for the Promotion of Science Grant-in-Aid for Scientific Research (KAKENHI Grant No. 19K03691).

References

  • (1) P. W. Anderson, Phys. Rev. 112, 1900 (1958).
  • (2) Y. Nambu, Phys. Rev. 117, 648 (1960).
  • (3) J. Goldstone, Nuovo Cimento 19, 154 (1961).
  • (4) J. Goldstone, A. Salam, and S. Weinberg, Phys. Rev. 127, 965 (1962).
  • (5) P. W. Higgs, Phys. Rev. Lett. 13, 508 (1964).
  • (6) The CMS Collaboration, Science 338, 1569 (2012);The ATLAS Collaboration, Science 338, 1576 (2012).
  • (7) D. Pekker and C. M. Varma, Annu. Rev. Condens. Matter Phys. 6, 269 (2015).
  • (8) R. Shimano and N. Tsuji, Annu. Rev. Condens. Matter Phys. 11, 103 (2020).
  • (9) R. Sooryakumar and M. V. Klein, Phys. Rev. Lett. 45, 660 (1980); Phys. Rev. B 23, 3213 (1981).
  • (10) R. Matsunaga, Y. I. Hamada, K. Makise, Y. Uzawa, H. Terai, Z. Wang, and R. Shimano, Phys. Rev. Lett 111, 057002 (2013).
  • (11) R. Matsunaga, N. Tsuji, H. Fujita, A. Sugioka, K. Makise, Y. Uzawa, H. Terai, Z. Wang, H. Aoki, and R. Shimano, Science 345, 1145 (2014).
  • (12) D. Sherman, U. S. Pracht, B. Gorshunov, S. Poran, J. Jesudasan, M. Chand, P. Raychaudhuri, M. Swanson, N. Trivedi, A. Auerbach, M. Scheffler, A. Frydman, and M. Dressel, Nat. Phys. 11, 188 (2015).
  • (13) H. Krull, N. Bittner, G. S. Uhrig, D. Manske, and A. P. Schnyder, Nat. Commun. 7, 11921 (2016).
  • (14) K. Katsumi, N. Tsuji, Y. I. Hamada, R. Matsunaga, J. Schneeloch, R. D. Zhong, G. D. Gu, H. Aoki, Y. Gallais, and R. Shimano, Phys. Rev. Lett. 120, 117001 (2018).
  • (15) H. Chu, M.-J. Kim, K. Katsumi, S. Kovalev, R. D. Dawson, L. Schwarz, N. Yoshikawa, G. Kim, D. Putzky, Z. Z. Li, H. Raffy, S. Germanskiy, J.-C. Deinert, N. Awari, I. Ilyakov, B. Green, M. Chen, M. Bawatna, G. Cristiani, G. Logvenov, Y. Gallais, A. V. Boris, B. Keimer, A. P. Schnyder, D. Manske, M. Gensch, Z. Wang, R. Shimano and S. Kaiser, Nat. Commun. 11, 1793 (2020).
  • (16) R. W. Giannetta, A. Ahonen, E. Polturak, J. Saunders, E. K. Zeise, R. C. Richardson, and D. M. Lee, Phys. Rev. Lett. 45, 262 (1980).
  • (17) V. V. Zavjalov, S. Autti, V. B. Eltsov, P. Heikkinen, and G. E. Volovik, Nat. Comm. 7, 10294 (2016).
  • (18) M. D. Nguyen, A. M. Zimmerman, and W. P. Halperin, Phys. Rev. B 99, 054510 (2019).
  • (19) Ch. Rüegg, B. Normand, M. Matsumoto, A. Furrer, D. F. McMorrow, K. W. Kramer, H. U. Gudel, S. N. Gvasaliya, H. Mutka, and M. Boehm, Phys. Rev. Lett. 100, 205701 (2008).
  • (20) A. Jain, M. Krautloher, J. Porras, G. H. Ryu, D. P. Chen, D. L. Abernathy, J. T. Park, A. Ivanov, J. Chaloupka, G. Khaliullin, B. Keimer, and B. J. Kim, Nat. Phys. 13, 633 (2017).
  • (21) T. Hong, M. Matsumoto, Y. Qiu, W. Chen, T. R. Gentile, S. Watson, F. F. Awwadi, M. M. Turnbull, S. E. Dissanayake, H. Agrawal, R. Toft-Petersen, B. Klemke, K. Coester, K. P. Schmidt, and D. A. Tennant, Nat. Phys. 13, 638 (2017).
  • (22) S.-M. Souliou, J. Chaloupka, G. Khaliullin, G. Ryu, A. Jain, B. J. Kim, M. Le Tacon, and B. Keimer, Phys. Rev. Lett. 119, 067201 (2017)
  • (23) J. Demsar, K. Biljaković, and D. Mihailovic, Phys. Rev. Lett. 83, 800 (1999)
  • (24) R. Yusupov, T. Mertelj, V. V. Kabanov, S. Brazovskii, P. Kusar, J.-H. Chu, I. R. Fisher, and D. Mihailovic, Nat. Phys, 6, 681 (2010).
  • (25) U. Bissbort, S. Götze, Y. Li, J. Heinze, J. S. Krauser, M. Weinberg, C. Becker, K. Sengstock, and W. Hofstetter, Phys. Rev. Lett. 106, 205303 (2011).
  • (26) M. Endres, T. Fukuhara, D. Pekker, M. Cheneau, P.Schauß, C. Gross, E. Demler, S. Kuhrm, and I. Bloch, Nature 487, 454 (2012).
  • (27) J. Lëonard, A. Morales, P. Zupancic, T. Donner, and T. Esslinger, Science 358 1415 (2017).
  • (28) A. Behrle, T. Harrison, J. Kombe, K. Gao, M. Link, J.-S. Bernier, C. Kollath, and M. Köhl, Nat. Phys. 14, 781 (2018).
  • (29) P. B. Littlewood and C. M. Varma, Phys. Rev. Lett. 47, 811 (1981); Phys. Rev. B 26, 4883 (1982).
  • (30) C. M. Varma, J. Low Temp. Phys. 126, 901 (2002).
  • (31) A. Leggett, in Modern Trends in the Theory of Condensed Matter, edited by A. Pekalski and R. Przystawa (Springer-Verlag, Berlin, 1980).
  • (32) P. Noziéres and S. Schmitt-Rink, J. Low Temp. Phys. 59, 195 (1985).
  • (33) Y. Ohashi and A. Griffin, Phys. Rev. Lett. 89, 130402 (2002).
  • (34) C. A. Regal, M. Greiner, and D. S. Jin, Phys. Rev. Lett. 92, 040403 (2004).
  • (35) J. R. Engelbrecht, M. Randeria, and C. A. R. Sá de Melo Phys. Rev. B 55, 15153 (1997).
  • (36) S. Tsuchiya, R. Ganesh, and T. Nikuni, Phys. Rev. B 88, 014527 (2013).
  • (37) T. Cea, C. Castellani, G. Seibold, and L. Benfatto, Phys. Rev. Lett. 115, 157002 (2015).
  • (38) S. Sachdev, “Quantum Phase Transitions”, (Cambridge University Press, 1998).
  • (39) T. Nakayama, I. Danshita, T. Nikuni, and S. Tsuchiya, Phys. Rev. A 92, 043610 (2015).
  • (40) T. Nakayama and S. Tsuchiya, Phys. Rev. A 100, 063612 (2019).
  • (41) M. Di Liberto, A. Recati, N. Trivedi, I. Carusotto, and C. Menotti, Phys. Rev. Lett. 120, 073201 (2018).
  • (42) S. Tsuchiya, D. Yamamoto, R. Yoshii, and M. Nitta, Phys. Rev. B 98, 094503 (2018).
  • (43) T. D. Lett, Particle Physics and Introduction to Field Theory (Harwood Academic, Reading, UK, 1981).
  • (44) See Supplemental Materials for details.
  • (45) T. Holstein and H. Primakoff, Phys. Rev. 58, 1098 (1940).

Appendix A Supplemental Materials

A.1 Pseudospin Operators

In this Section, we summarize the algebra for the pseudospin operators. We introduce 𝑺𝒌𝒌¯\bm{S}_{\bm{k}\underline{\bm{k}}} in Eq. (9) in the main text. We also introduce another pseudospin operator 𝑺~𝒌𝒌¯\tilde{\bm{S}}_{\bm{k}\underline{\bm{k}}} as

S~μ𝒌𝒌¯=Sμ𝒌+(1)δμ,xSμ𝒌¯,(μ=x,y,z),\displaystyle\tilde{S}_{\mu\bm{k}\underline{\bm{k}}}=S_{\mu\bm{k}}+(-1)^{\delta_{\mu,x}}S_{\mu\underline{\bm{k}}}~{},~{}(\mu=x,y,z), (23)

where 𝒌\bm{k} is above the Fermi surface (FS) (ξ𝒌>0\xi_{\bm{k}}>0). 𝑺𝒌\bm{S}_{\bm{k}} is transformed by 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}} and 𝒞𝒌F\mathcal{C}_{\bm{k}_{F}} as

𝒞𝒌𝒌¯Sμ𝒌𝒞𝒌𝒌¯={(1)δμ,x+1Sμ𝒌¯,(𝒌=𝒌,𝒌¯),Sμ𝒌,otherwise,\displaystyle\mathcal{C}_{\bm{k}\underline{\bm{k}}}S_{\mu\bm{k}^{\prime}}\mathcal{C}_{\bm{k}\underline{\bm{k}}}=\left\{\begin{array}[]{ll}(-1)^{\delta_{\mu,x}+1}S_{\mu\underline{\bm{k}}^{\prime}}~{},&(\bm{k}^{\prime}=\bm{k},\underline{\bm{k}})~{},\\ S_{\mu{\bm{k}}^{\prime}}~{},&{\rm otherwise},\end{array}\right. (26)
𝒞𝒌FSμ𝒌𝒞𝒌F={(1)δμ,x+1Sμ𝒌F,(𝒌=𝒌F),Sμ𝒌,otherwise.\displaystyle\mathcal{C}_{\bm{k}_{F}}S_{\mu\bm{k}}\mathcal{C}_{\bm{k}_{F}}=\left\{\begin{array}[]{ll}(-1)^{\delta_{\mu,x}+1}S_{\mu\bm{k}_{F}}~{},&(\bm{k}=\bm{k}_{F})~{},\\ S_{\mu\bm{k}}~{},&{\rm otherwise}.\end{array}\right. (29)

Using Eqs. (26) and (29), one can derive the commutation relations

[𝒞𝒌𝒌¯,Sμ𝒌𝒌¯]={𝒞𝒌𝒌¯,S~μ𝒌𝒌¯}=0.\displaystyle[\mathcal{C}_{\bm{k}\underline{\bm{k}}},S_{\mu\bm{k}\underline{\bm{k}}}]=\{\mathcal{C}_{\bm{k}\underline{\bm{k}}},\tilde{S}_{\mu\bm{k}\underline{\bm{k}}}\}=0~{}. (30)

𝑺𝒌𝒌¯{\bm{S}}_{\bm{k}\underline{\bm{k}}} and 𝑺~𝒌𝒌¯\tilde{\bm{S}}_{\bm{k}\underline{\bm{k}}} also satisfy

[Sμ𝒌𝒌¯,Sν𝒌𝒌¯]=[S~μ𝒌𝒌¯,S~ν𝒌𝒌¯]=iρεμνρSρ𝒌𝒌¯,\displaystyle[S_{\mu\bm{k}\underline{\bm{k}}},S_{\nu\bm{k}\underline{\bm{k}}}]=[\tilde{S}_{\mu\bm{k}\underline{\bm{k}}},\tilde{S}_{\nu\bm{k}\underline{\bm{k}}}]=i\sum_{\rho}\varepsilon_{\mu\nu\rho}S_{\rho\bm{k}\underline{\bm{k}}}~{}, (31)
[Sμ𝒌𝒌¯,S~ν𝒌𝒌¯]=iρεμνρS~ρ𝒌𝒌¯,\displaystyle[S_{\mu\bm{k}\underline{\bm{k}}},\tilde{S}_{\nu\bm{k}\underline{\bm{k}}}]=i\sum_{\rho}\varepsilon_{\mu\nu\rho}\tilde{S}_{\rho\bm{k}\underline{\bm{k}}}~{}, (32)
Sμ𝒌𝒌¯S~μ𝒌𝒌¯=S~μ𝒌𝒌¯Sμ𝒌𝒌¯=0,Sμ𝒌𝒌¯2+S~μ𝒌𝒌¯2=1.\displaystyle S_{\mu\bm{k}\underline{\bm{k}}}\tilde{S}_{\mu\bm{k}\underline{\bm{k}}}=\tilde{S}_{\mu\bm{k}\underline{\bm{k}}}S_{\mu\bm{k}\underline{\bm{k}}}=0~{},\quad S_{\mu\bm{k}\underline{\bm{k}}}^{2}+\tilde{S}_{\mu\bm{k}\underline{\bm{k}}}^{2}=1~{}. (33)

The spin-1 triplet states for Sz𝒌𝒌¯S_{z\bm{k}\underline{\bm{k}}} can be written as

|1𝒌𝒌¯=|𝒌𝒌¯,\displaystyle|1\rangle_{\bm{k}\underline{\bm{k}}}=|\uparrow\downarrow\rangle_{\bm{k}\underline{\bm{k}}}, (34)
|0𝒌𝒌¯=12(|𝒌𝒌¯+|𝒌𝒌¯),\displaystyle|0\rangle_{\bm{k}\underline{\bm{k}}}=\frac{1}{\sqrt{2}}(|\uparrow\uparrow\rangle_{\bm{k}\underline{\bm{k}}}+|\downarrow\downarrow\rangle_{\bm{k}\underline{\bm{k}}}), (35)
|1𝒌𝒌¯=|𝒌𝒌¯,\displaystyle|-1\rangle_{\bm{k}\underline{\bm{k}}}=|\downarrow\uparrow\rangle_{\bm{k}\underline{\bm{k}}}~{}, (36)

Here, we denote |st𝒌𝒌¯=|s𝒌|t𝒌¯\ket{s~{}t}_{\bm{k}\underline{\bm{k}}}=\ket{s}_{\bm{k}}\otimes\ket{t}_{\underline{\bm{k}}} (s,t=,s,t=\uparrow,\downarrow). They satisfy Sz𝒌𝒌¯|m𝒌𝒌¯=m|m𝒌𝒌¯S_{z\bm{k}\underline{\bm{k}}}|m\rangle_{\bm{k}\underline{\bm{k}}}=m|m\rangle_{\bm{k}\underline{\bm{k}}} (m=1,0,1m=1,0,-1) and have even parity under 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}: 𝒞𝒌𝒌¯|m𝒌𝒌¯=|m𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}|m\rangle_{\bm{k}\underline{\bm{k}}}=|m\rangle_{\bm{k}\underline{\bm{k}}}.

The spin-1 triplet states for Sx𝒌𝒌¯S_{x\bm{k}\underline{\bm{k}}} can be written as

|1x𝒌𝒌¯\displaystyle\ket{1_{x}}_{\bm{k}\underline{\bm{k}}} =\displaystyle= |++𝒌𝒌¯,\displaystyle\ket{++}_{\bm{k}\underline{\bm{k}}}~{}, (37)
|0x𝒌𝒌¯\displaystyle\ket{0_{x}}_{\bm{k}\underline{\bm{k}}} =\displaystyle= 12(|+𝒌𝒌¯|+𝒌𝒌¯),\displaystyle\frac{1}{\sqrt{2}}(\ket{-+}_{\bm{k}\underline{\bm{k}}}-\ket{+-}_{\bm{k}\underline{\bm{k}}})~{}, (38)
|1x𝒌𝒌¯\displaystyle\ket{-1_{x}}_{\bm{k}\underline{\bm{k}}} =\displaystyle= |𝒌𝒌¯,\displaystyle-\ket{--}_{\bm{k}\underline{\bm{k}}}~{}, (39)

where |±𝒌=(|𝒌±|𝒌)/2\ket{\pm}_{\bm{k}}=(\ket{\uparrow}_{\bm{k}}\pm\ket{\downarrow}_{\bm{k}})/\sqrt{2}. They satisfy Sx𝒌𝒌¯|mx𝒌𝒌¯=m|mx𝒌𝒌¯S_{x\bm{k}\underline{\bm{k}}}|m_{x}\rangle_{\bm{k}\underline{\bm{k}}}=m|m_{x}\rangle_{\bm{k}\underline{\bm{k}}} (m=1,0,1m=1,0,-1) and have even parity under 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}: 𝒞𝒌𝒌¯|mx𝒌𝒌¯=|mx𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}|m_{x}\rangle_{\bm{k}\underline{\bm{k}}}=|m_{x}\rangle_{\bm{k}\underline{\bm{k}}}.

The spin-0 singlet state |0~𝒌𝒌¯|\tilde{0}\rangle_{\bm{k}\underline{\bm{k}}} can be written as

|0~𝒌𝒌¯\displaystyle|\tilde{0}\rangle_{\bm{k}\underline{\bm{k}}} =\displaystyle= 12(|𝒌𝒌¯|𝒌𝒌¯)\displaystyle\frac{1}{\sqrt{2}}(\ket{\uparrow\uparrow}_{\bm{k}\underline{\bm{k}}}-\ket{\downarrow\downarrow}_{\bm{k}\underline{\bm{k}}}) (40)
=\displaystyle= 12(|+𝒌𝒌¯+|+𝒌𝒌¯).\displaystyle\frac{1}{\sqrt{2}}(\ket{+-}_{\bm{k}\underline{\bm{k}}}+\ket{-+}_{\bm{k}\underline{\bm{k}}}). (41)

It satisfies Sμ𝒌𝒌¯|0~𝒌𝒌¯=0S_{\mu\bm{k}\underline{\bm{k}}}|\tilde{0}\rangle_{\bm{k}\underline{\bm{k}}}=0 (μ=x,y,z\mu=x,y,z) and has odd parity under 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}: 𝒞𝒌𝒌¯|0~𝒌𝒌¯=|0~𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}|\tilde{0}\rangle_{\bm{k}\underline{\bm{k}}}=-|\tilde{0}\rangle_{\bm{k}\underline{\bm{k}}}. We note that the pseudospins for 𝒌\bm{k} and 𝒌¯\underline{\bm{k}} are entangled in |0𝒌𝒌¯|0\rangle_{\bm{k}\underline{\bm{k}}}, |0x𝒌𝒌¯|0_{x}\rangle_{\bm{k}\underline{\bm{k}}}, and |0~𝒌𝒌¯|\tilde{0}\rangle_{\bm{k}\underline{\bm{k}}}.

S𝒌𝒌¯±=Sx𝒌𝒌¯±iSy𝒌𝒌¯S^{\pm}_{\bm{k}\underline{\bm{k}}}=S_{x\bm{k}\underline{\bm{k}}}\pm iS_{y\bm{k}\underline{\bm{k}}} are the ladder operators for the spin-1 triplet states {|m𝒌𝒌¯}\{\ket{m}_{\bm{k}\underline{\bm{k}}}\}:

S𝒌𝒌¯|1𝒌𝒌¯=S𝒌𝒌¯+|1𝒌𝒌¯=2|0𝒌𝒌¯,\displaystyle S^{-}_{\bm{k}\underline{\bm{k}}}\ket{1}_{\bm{k}\underline{\bm{k}}}=S^{+}_{\bm{k}\underline{\bm{k}}}\ket{-1}_{\bm{k}\underline{\bm{k}}}=\sqrt{2}\ket{0}_{\bm{k}\underline{\bm{k}}}~{}, (42)
S𝒌𝒌¯+|1𝒌𝒌¯=S𝒌𝒌¯|1𝒌𝒌¯=0,\displaystyle S^{+}_{\bm{k}\underline{\bm{k}}}\ket{1}_{\bm{k}\underline{\bm{k}}}=S^{-}_{\bm{k}\underline{\bm{k}}}\ket{-1}_{\bm{k}\underline{\bm{k}}}=0~{}, (43)
S𝒌𝒌¯±|0𝒌𝒌¯={2|1𝒌𝒌¯,2|1𝒌𝒌¯.\displaystyle S^{\pm}_{\bm{k}\underline{\bm{k}}}\ket{0}_{\bm{k}\underline{\bm{k}}}=\left\{\begin{array}[]{l}\sqrt{2}\ket{1}_{\bm{k}\underline{\bm{k}}}~{},\\ \sqrt{2}\ket{-1}_{\bm{k}\underline{\bm{k}}}~{}.\end{array}\right. (46)

S~𝒌𝒌¯±=S~x𝒌𝒌¯±iS~y𝒌𝒌¯\tilde{S}^{\pm}_{\bm{k}\underline{\bm{k}}}=\tilde{S}_{x\bm{k}\underline{\bm{k}}}\pm i\tilde{S}_{y\bm{k}\underline{\bm{k}}} transform the spin-1 triplet states into the spin-0 singlet state |0~𝒌𝒌¯\ket{\tilde{0}}_{\bm{k}\underline{\bm{k}}} and vice versa as

S~𝒌𝒌¯|1𝒌𝒌¯=S~𝒌𝒌¯+|1𝒌𝒌¯=2|0~𝒌𝒌¯,\displaystyle\tilde{S}^{-}_{\bm{k}\underline{\bm{k}}}\ket{1}_{\bm{k}\underline{\bm{k}}}=-\tilde{S}^{+}_{\bm{k}\underline{\bm{k}}}\ket{-1}_{\bm{k}\underline{\bm{k}}}=-\sqrt{2}\ket{\tilde{0}}_{\bm{k}\underline{\bm{k}}}~{}, (47)
S~𝒌𝒌¯+|1𝒌𝒌¯=S~𝒌𝒌¯|1𝒌𝒌¯=S~𝒌𝒌¯±|0𝒌𝒌¯=0,\displaystyle\tilde{S}^{+}_{\bm{k}\underline{\bm{k}}}\ket{1}_{\bm{k}\underline{\bm{k}}}=\tilde{S}^{-}_{\bm{k}\underline{\bm{k}}}\ket{-1}_{\bm{k}\underline{\bm{k}}}=\tilde{S}^{\pm}_{\bm{k}\underline{\bm{k}}}\ket{0}_{\bm{k}\underline{\bm{k}}}=0~{}, (48)
S~𝒌𝒌¯+|0~𝒌𝒌¯=2|1𝒌𝒌¯,\displaystyle\tilde{S}^{+}_{\bm{k}\underline{\bm{k}}}\ket{\tilde{0}}_{\bm{k}\underline{\bm{k}}}=-\sqrt{2}\ket{1}_{\bm{k}\underline{\bm{k}}}, (49)
S~𝒌𝒌¯|0~𝒌𝒌¯=2|1𝒌𝒌¯.\displaystyle\tilde{S}^{-}_{\bm{k}\underline{\bm{k}}}\ket{\tilde{0}}_{\bm{k}\underline{\bm{k}}}=\sqrt{2}\ket{-1}_{\bm{k}\underline{\bm{k}}}~{}. (50)

In general, operating S~μ𝒌𝒌¯\tilde{S}_{\mu\bm{k}\underline{\bm{k}}} on a certain state changes its parity under 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}, while operating Sμ𝒌𝒌¯S_{\mu\bm{k}\underline{\bm{k}}} does not as

𝒞𝒌𝒌¯Sμ𝒌𝒌¯|ψe(o)𝒌𝒌¯\displaystyle\mathcal{C}_{\bm{k}\underline{\bm{k}}}S_{\mu\bm{k}\underline{\bm{k}}}\ket{\psi_{\rm e(o)}}_{\bm{k}\underline{\bm{k}}} =\displaystyle= ±Sμ𝒌𝒌¯|ψe(o)𝒌𝒌¯,\displaystyle\pm S_{\mu\bm{k}\underline{\bm{k}}}\ket{\psi_{\rm e(\rm o)}}_{\bm{k}\underline{\bm{k}}},~{} (51)
𝒞𝒌𝒌¯S~μ𝒌𝒌¯|ψe(o)𝒌𝒌¯\displaystyle\mathcal{C}_{\bm{k}\underline{\bm{k}}}\tilde{S}_{\mu\bm{k}\underline{\bm{k}}}\ket{\psi_{\rm e(\rm o)}}_{\bm{k}\underline{\bm{k}}} =\displaystyle= S~μ𝒌𝒌¯|ψe(o)𝒌𝒌¯,\displaystyle\mp\tilde{S}_{\mu\bm{k}\underline{\bm{k}}}\ket{\psi_{\rm e(\rm o)}}_{\bm{k}\underline{\bm{k}}}~{}, (52)

where |ψe𝒌𝒌¯\ket{\psi_{\rm e}}_{\bm{k}\underline{\bm{k}}} (|ψo𝒌𝒌¯\ket{\psi_{\rm o}}_{\bm{k}\underline{\bm{k}}}) denotes an even (odd) parity state under 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}. The upper (lower) signs on the right-hand-sides of Eqs. (51) and (52) are for the even (odd) parity state.

A.2 Applications of Pseudospin Operators

In this Section, for illustration of applications of the pseudospin operators introduced in the previous Section, we describe the normal state in terms of them and also derive some rigorous relations for occupation numbers of fermions and dynamical spin susceptibilities.

We first consider the normal state. Given the kinetic energy term in Eq. (10), the ground state of K\mathcal{H}_{K} can be written as |Ψ0=𝒌|1𝒌𝒌¯|ΦF\ket{\Psi_{0}}=\prod_{\bm{k}}^{\prime}\ket{-1}_{\bm{k}\underline{\bm{k}}}\otimes\ket{\Phi_{F}}, where |ΦF\ket{\Phi_{F}} denotes the wave function for the pseudospins on the FS, which are zero modes for K\mathcal{H}_{K}. Using Eqs. (31) and (32), we obtain

[K,S𝒌𝒌¯±]\displaystyle[\mathcal{H}_{K},S_{\bm{k}\underline{\bm{k}}}^{\pm}] =\displaystyle= ±2ξ𝒌S𝒌𝒌¯±,\displaystyle\pm 2\xi_{\bm{k}}S_{\bm{k}\underline{\bm{k}}}^{\pm}~{}, (53)
[K,S~𝒌𝒌¯±]\displaystyle[\mathcal{H}_{K},\tilde{S}_{\bm{k}\underline{\bm{k}}}^{\pm}] =\displaystyle= ±2ξ𝒌S~𝒌𝒌¯±.\displaystyle\pm 2\xi_{\bm{k}}\tilde{S}_{\bm{k}\underline{\bm{k}}}^{\pm}~{}. (54)

The above equations show that S𝒌𝒌¯±S_{\bm{k}\underline{\bm{k}}}^{\pm} and S~𝒌𝒌¯±\tilde{S}_{\bm{k}\underline{\bm{k}}}^{\pm} are the raising and lowering operators for the eigenstates of K\mathcal{H}_{K}. Thus, S𝒌𝒌¯+|Ψ0/2S_{\bm{k}\underline{\bm{k}}}^{+}\ket{\Psi_{0}}/\sqrt{2} and S~𝒌𝒌¯+|Ψ0/2\tilde{S}_{\bm{k}\underline{\bm{k}}}^{+}\ket{\Psi_{0}}/\sqrt{2} are the degenerate excited states with excitation energy 2ξ𝒌2\xi_{\bm{k}} involving a pair of fermions, where they have even and odd parity under 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}, respectively.

We derive the fermion occupation number for |ψe(o)𝒌𝒌¯\ket{\psi_{\rm e(o)}}_{\bm{k}\underline{\bm{k}}}. Using Eq. (52), we obtain

ψe(o)|𝒌𝒌¯𝑺~𝒌𝒌¯|ψe(o)𝒌𝒌¯=0.{}_{\bm{k}\underline{\bm{k}}}\bra{\psi_{\rm e(o)}}\tilde{\bm{S}}_{\bm{k}\underline{\bm{k}}}\ket{\psi_{\rm e(o)}}_{\bm{k}\underline{\bm{k}}}=0~{}. (55)

Specifically, from the zz-component of Eq. (55) we obtain the quantization of fermion occupation number

ψe(o)|𝒌𝒌¯(n𝒌+n𝒌+n𝒌¯+n𝒌¯)|ψe(o)𝒌𝒌¯=2,{}_{\bm{k}\underline{\bm{k}}}\bra{\psi_{\rm e(o)}}(n_{\bm{k}\uparrow}+n_{-\bm{k}\downarrow}+n_{\underline{\bm{k}}\uparrow}+n_{-\underline{\bm{k}}\downarrow})\ket{\psi_{\rm e(o)}}_{\bm{k}\underline{\bm{k}}}=2~{}, (56)

where n𝒌σ=c𝒌σc𝒌σn_{\bm{k}\sigma}=c_{\bm{k}\sigma}^{\dagger}c_{\bm{k}\sigma} is the fermion number operator. We also obtain

±|𝒌F𝑺𝒌F|±𝒌F=(±1/2,0,0),{}_{\bm{k}_{F}}\bra{\pm}\bm{S}_{\bm{k}_{F}}\ket{\pm}_{\bm{k}_{F}}=(\pm 1/2,0,0)~{}, (57)

where |+𝒌F\ket{+}_{\bm{k}_{F}} (|𝒌F\ket{-}_{\bm{k}_{F}}) is an even (odd) parity state under 𝒞𝒌F\mathcal{C}_{\bm{k}_{F}}. The upper (lower) sign on the right-hand-side of Eq. (57) is for |+𝒌F\ket{+}_{\bm{k}_{F}} (|𝒌F\ket{-}_{\bm{k}_{F}}). From the zz-component of Eq. (57), we obtain

±|𝒌F(n𝒌F+n𝒌F)|±𝒌F=1.{}_{\bm{k}_{F}}\bra{\pm}(n_{\bm{k}_{F}\uparrow}+n_{-\bm{k}_{F}\downarrow})\ket{\pm}_{\bm{k}_{F}}=1~{}. (58)

Therefore, the states |ψe(o)𝒌𝒌¯\ket{\psi_{\rm e(o)}}_{\bm{k}\underline{\bm{k}}} and |±𝒌F\ket{\pm}_{\bm{k}_{F}} that appear in the eigenstates of I\mathcal{H}_{I} are half-filled.

One can also derive some rigorous relations for the dynamical spin susceptibility

χμ𝒌,ν(ω)i0[Sν,Sμ𝒌(t)]eiωt𝑑t,\chi_{\mu\bm{k},\nu}(\omega)\equiv-i\int_{0}^{\infty}\langle[S_{\nu},S_{\mu\bm{k}}(t)]\rangle e^{-i\omega t}dt~{}, (59)

where Sμ𝒌(t)=eiItSμ𝒌eiItS_{\mu\bm{k}}(t)=e^{-i\mathcal{H}_{I}t}S_{\mu\bm{k}}e^{i\mathcal{H}_{I}t} is the Heisenberg representation and \langle\cdots\rangle denotes the thermal average. We focus on χy𝒌,x(ω)\chi_{y\bm{k},x}(\omega) and evaluate

[Sx,Sy𝒌(t)]neβEnn|[Sx,Sy𝒌(t)]|n,\langle[S_{x},S_{y\bm{k}}(t)]\rangle\propto\sum_{n}e^{-\beta E_{n}}\bra{n}[S_{x},S_{y\bm{k}}(t)]\ket{n}~{}, (60)

where β\beta is the inverse temperature and |n\ket{n} denotes an eigenstate of I\mathcal{H}_{I}: I|n=En|n\mathcal{H}_{I}\ket{n}=E_{n}\ket{n}. Since operating S~y𝒌𝒌¯\tilde{S}_{y\bm{k}\underline{\bm{k}}} changes the parity of the state under 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}, we obtain n|SxS~y𝒌𝒌¯(t)|n=n|S~y𝒌𝒌¯(t)Sx|n=0\bra{n}S_{x}\tilde{S}_{y\bm{k}\underline{\bm{k}}}(t)\ket{n}=\bra{n}\tilde{S}_{y\bm{k}\underline{\bm{k}}}(t)S_{x}\ket{n}=0 and therefore n|SxSy𝒌(t)|n=n|SxSy𝒌¯(t)|n\bra{n}S_{x}S_{y\bm{k}}(t)\ket{n}=-\bra{n}S_{x}S_{y\underline{\bm{k}}}(t)\ket{n} and n|Sy𝒌(t)Sx|n=n|Sy𝒌¯(t)Sx|n\bra{n}S_{y\bm{k}}(t)S_{x}\ket{n}=-\bra{n}S_{y\underline{\bm{k}}}(t)S_{x}\ket{n}. On the other hand, one can also derive n|SxSy𝒌F(t)|n=0\bra{n}S_{x}S_{y\bm{k}_{F}}(t)\ket{n}=0 using 𝒞𝒌FSy𝒌F(t)𝒞𝒌F=Sy𝒌F(t)\mathcal{C}_{\bm{k}_{F}}S_{y\bm{k}_{F}}(t)\mathcal{C}_{\bm{k}_{F}}=-S_{y\bm{k}_{F}}(t). We thus obtain

χy𝒌,x(ω)=χy𝒌¯,x(ω),\displaystyle\chi_{y\bm{k},x}(\omega)=-\chi_{y\underline{\bm{k}},x}(\omega)~{}, (61)
χy𝒌F,x(ω)=0.\displaystyle\chi_{y\bm{k}_{F},x}(\omega)=0~{}. (62)

Equations (61) and (62) lead to the out-of-phase oscillations of spin fluctuations for the Higgs mode δSy𝒌¯=δSy𝒌\delta S_{y\underline{\bm{k}}}=-\delta S_{y\bm{k}} in the classical spin analysis [42]. Analogously, one can show

χx𝒌,y(ω)=χx𝒌¯,y(ω),\displaystyle\chi_{x\bm{k},y}(\omega)=-\chi_{x\underline{\bm{k}},y}(\omega)~{}, (63)
χx𝒌F,y(ω)=0,\displaystyle\chi_{x\bm{k}_{F},y}(\omega)=0~{}, (64)

using n|S~y𝒌𝒌¯Sx𝒌𝒌¯(t)|n=n|Sx𝒌𝒌¯(t)S~y𝒌𝒌¯|n=0\bra{n}\tilde{S}_{y\bm{k}\underline{\bm{k}}}S_{x\bm{k}\underline{\bm{k}}}(t)\ket{n}=\bra{n}S_{x\bm{k}\underline{\bm{k}}}(t)\tilde{S}_{y\bm{k}\underline{\bm{k}}}\ket{n}=0, n|Sy𝒌FSx𝒌𝒌¯(t)|n=n|Sx𝒌𝒌¯(t)Sy𝒌F|n=0\bra{n}S_{y\bm{k}_{F}}S_{x\bm{k}\underline{\bm{k}}}(t)\ket{n}=\bra{n}S_{x\bm{k}\underline{\bm{k}}}(t)S_{y\bm{k}_{F}}\ket{n}=0, and n|SySx𝒌𝒌¯(t)|n=n|Sx𝒌𝒌¯(t)Sy|n=0\bra{n}S_{y}S_{x\bm{k}\underline{\bm{k}}}(t)\ket{n}=\bra{n}S_{x\bm{k}\underline{\bm{k}}}(t)S_{y}\ket{n}=0. We thus find that the dynamical spin susceptibilities for total spin vanish as χxy(ω)=𝒌χx𝒌,y(ω)=0\chi_{xy}(\omega)=\sum_{\bm{k}}\chi_{x\bm{k},y}(\omega)=0 and χyx(ω)=𝒌χy𝒌,x(ω)=0\chi_{yx}(\omega)=\sum_{\bm{k}}\chi_{y\bm{k},x}(\omega)=0 [42]. The expressions for the MF theory

χy𝒌,x(ω)χx𝒌,y(ω)ωξ𝒌E𝒌(4E𝒌2ω2),\chi_{y\bm{k},x}(\omega)\propto\chi_{x\bm{k},y}(\omega)\propto\frac{\omega\xi_{\bm{k}}}{E_{\bm{k}}(4E_{\bm{k}}^{2}-\omega^{2})}~{}, (65)

indeed satisfy Eqs. (61)\sim(64) [42].

A.3 Pseudospin Operators for Bogolons

In this Section, we summarize the algebra for the pseudospin operators for bogolons. 𝑺𝒌\bm{S}^{\prime}_{\bm{k}} is introduced as [42]

(Sz𝒌Sx𝒌Sy𝒌)=(cosφ𝒌sinφ𝒌0sinφ𝒌cosφ𝒌0001)(Sz𝒌Sx𝒌Sy𝒌),\displaystyle\left(\begin{array}[]{ccc}S_{z\bm{k}}^{\prime}\\ S_{x\bm{k}}^{\prime}\\ S_{y\bm{k}}^{\prime}\end{array}\right)=\left(\begin{array}[]{ccc}-\cos\varphi_{\bm{k}}&-\sin\varphi_{\bm{k}}&0\\ \sin\varphi_{\bm{k}}&-\cos\varphi_{\bm{k}}&0\\ 0&0&1\end{array}\right)\left(\begin{array}[]{cc}S_{z\bm{k}}\\ S_{x\bm{k}}\\ S_{y\bm{k}}\end{array}\right), (75)

where cosφ𝒌=ξ𝒌/E𝒌\cos\varphi_{\bm{k}}=-\xi_{\bm{k}}/E_{\bm{k}} and sinφ𝒌=Δ0/E𝒌\sin\varphi_{\bm{k}}=\Delta_{0}/E_{\bm{k}}. 𝑺𝒌\bm{S}_{\bm{k}} is rotated about the angle πφ𝒌\pi-\varphi_{\bm{k}} in the xzxz-plane in Eq. (75). Using Eq. (75), the MF Hamiltonian can be written in the form of Eq. (11). The eigenstates of Sz𝒌S_{z\bm{k}}^{\prime} are given as

|𝒌\displaystyle|\uparrow^{\prime}\rangle_{\bm{k}} =\displaystyle= u𝒌|𝒌v𝒌|𝒌,\displaystyle u_{\bm{k}}|\uparrow\rangle_{\bm{k}}-v_{\bm{k}}|\downarrow\rangle_{\bm{k}}~{}, (76)
|𝒌\displaystyle|\!\downarrow^{\prime}\rangle_{\bm{k}} =\displaystyle= u𝒌|𝒌+v𝒌|𝒌,\displaystyle u_{\bm{k}}|\downarrow\rangle_{\bm{k}}+v_{\bm{k}}|\uparrow\rangle_{\bm{k}}~{}, (77)

where |𝒌\ket{\downarrow^{\prime}}_{\bm{k}} represents the vacuum of bogolons and |𝒌\ket{\uparrow^{\prime}}_{\bm{k}} the excited state with a pair of bogolons. All the pseudospins 𝑺𝒌{\bm{S}}_{\bm{k}}^{\prime} are aligned downward in the zz direction in the superconducting (SC) ground state. The BCS wave function can be thus written as

|Ψ=𝒌|𝒌.|\Psi\rangle=\prod_{\bm{k}}\ket{\downarrow^{\prime}}_{\bm{k}}. (78)

Note that (Sx𝒌F,Sy𝒌F,Sz𝒌F)=(Sz𝒌F,Sy𝒌F,Sx𝒌F)(S^{\prime}_{x\bm{k}_{F}},S^{\prime}_{y\bm{k}_{F}},S^{\prime}_{z\bm{k}_{F}})=(S_{z\bm{k}_{F}},S_{y\bm{k}_{F}},-S_{x\bm{k}_{F}}), |𝒌F=|𝒌F\ket{\uparrow^{\prime}}_{\bm{k}_{F}}=\ket{-}_{\bm{k}_{F}}, and |𝒌F=|+𝒌F\ket{\downarrow^{\prime}}_{\bm{k}_{F}}=\ket{+}_{\bm{k}_{F}}.

We introduce another pseudospin operator 𝑺~𝒌𝒌¯\tilde{\bm{S}}^{\prime}_{\bm{k}\underline{\bm{k}}} as

S~μ𝒌𝒌¯=Sμ𝒌+(1)δμ,zSμ𝒌¯,(μ=x,y,z),\displaystyle\tilde{S}^{\prime}_{\mu\bm{k}\underline{\bm{k}}}=S^{\prime}_{\mu\bm{k}}+(-1)^{\delta_{\mu,z}}S^{\prime}_{\mu\underline{\bm{k}}}~{},~{}(\mu=x,y,z), (79)

where 𝒌\bm{k} is above the FS (ξ𝒌>0\xi_{\bm{k}}>0). 𝑺𝒌\bm{S}^{\prime}_{\bm{k}} is transformed by 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}} and 𝒞𝒌F\mathcal{C}_{\bm{k}_{F}} as

𝒞𝒌𝒌¯Sμ𝒌𝒞𝒌𝒌¯={(1)δμ,z+1Sμ𝒌¯,(𝒌=𝒌,𝒌¯),Sμ𝒌,otherwise,\displaystyle\mathcal{C}_{\bm{k}\underline{\bm{k}}}S^{\prime}_{\mu\bm{k}^{\prime}}\mathcal{C}_{\bm{k}\underline{\bm{k}}}=\left\{\begin{array}[]{ll}(-1)^{\delta_{\mu,z}+1}S^{\prime}_{\mu\underline{\bm{k}}^{\prime}}~{},&(\bm{k}^{\prime}=\bm{k},\underline{\bm{k}})~{},\\ S^{\prime}_{\mu{\bm{k}}^{\prime}}~{},&{\rm otherwise},\end{array}\right. (82)
𝒞𝒌FSμ𝒌𝒞𝒌F={(1)δμ,z+1Sμ𝒌F,(𝒌=𝒌F),Sμ𝒌,otherwise.\displaystyle\mathcal{C}_{\bm{k}_{F}}S^{\prime}_{\mu\bm{k}}\mathcal{C}_{\bm{k}_{F}}=\left\{\begin{array}[]{ll}(-1)^{\delta_{\mu,z}+1}S^{\prime}_{\mu\bm{k}_{F}}~{},&(\bm{k}=\bm{k}_{F})~{},\\ S_{\mu\bm{k}}~{},&{\rm otherwise}.\end{array}\right. (85)

Using Eqs. (82) and (85), one can derive the commutation relations

[𝒞𝒌𝒌¯,Sμ𝒌𝒌¯]={𝒞𝒌𝒌¯,S~μ𝒌𝒌¯}=0.\displaystyle[\mathcal{C}_{\bm{k}\underline{\bm{k}}},S^{\prime}_{\mu\bm{k}\underline{\bm{k}}}]=\{\mathcal{C}_{\bm{k}\underline{\bm{k}}},{\tilde{S}}^{\prime}_{\mu\bm{k}\underline{\bm{k}}}\}=0~{}. (86)

𝑺𝒌𝒌¯\bm{S}^{\prime}_{\bm{k}\underline{\bm{k}}} and 𝑺~𝒌𝒌¯\tilde{\bm{S}}^{\prime}_{\bm{k}\underline{\bm{k}}} obey the commutation relations

[Sμ𝒌𝒌¯,Sν𝒌𝒌¯]=[S~μ𝒌𝒌¯,S~ν𝒌𝒌¯]=iρεμνρSρ𝒌𝒌¯,\displaystyle[S^{\prime}_{\mu\bm{k}\underline{\bm{k}}},S^{\prime}_{\nu\bm{k}\underline{\bm{k}}}]=[\tilde{S}^{\prime}_{\mu\bm{k}\underline{\bm{k}}},\tilde{S}^{\prime}_{\nu\bm{k}\underline{\bm{k}}}]=i\sum_{\rho}\varepsilon_{\mu\nu\rho}S^{\prime}_{\rho\bm{k}\underline{\bm{k}}}~{}, (87)
[Sμ𝒌𝒌¯,S~ν𝒌𝒌¯]=iρεμνρS~ρ𝒌𝒌¯.\displaystyle[S^{\prime}_{\mu\bm{k}\underline{\bm{k}}},\tilde{S}^{\prime}_{\nu\bm{k}\underline{\bm{k}}}]=i\sum_{\rho}\varepsilon_{\mu\nu\rho}\tilde{S}^{\prime}_{\rho\bm{k}\underline{\bm{k}}}~{}. (88)

The spin-1 triplet states for Sz𝒌𝒌¯S^{\prime}_{z\bm{k}\underline{\bm{k}}} can be written as

|1𝒌𝒌¯=|𝒌𝒌¯,\displaystyle\ket{1^{\prime}}_{\bm{k}\underline{\bm{k}}}=\ket{\uparrow^{\prime}\uparrow^{\prime}}_{\bm{k}\underline{\bm{k}}}, (89)
|0𝒌𝒌¯=12(|𝒌𝒌¯|𝒌𝒌¯),\displaystyle\ket{0^{\prime}}_{\bm{k}\underline{\bm{k}}}=\frac{1}{\sqrt{2}}(\ket{\uparrow^{\prime}\downarrow^{\prime}}_{\bm{k}\underline{\bm{k}}}-\ket{\downarrow^{\prime}\uparrow^{\prime}}_{\bm{k}\underline{\bm{k}}}), (90)
|1𝒌𝒌¯=|𝒌𝒌¯.\displaystyle\ket{-1^{\prime}}_{\bm{k}\underline{\bm{k}}}=\ket{\downarrow^{\prime}\downarrow^{\prime}}_{\bm{k}\underline{\bm{k}}}~{}. (91)

They satisfy Sz𝒌𝒌¯|m𝒌𝒌¯=m|m𝒌𝒌¯S^{\prime}_{z\bm{k}\underline{\bm{k}}}\ket{m^{\prime}}_{\bm{k}\underline{\bm{k}}}=m\ket{m^{\prime}}_{\bm{k}\underline{\bm{k}}} (m=1,0,1m=1,0,-1). They have even parity under 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}: 𝒞𝒌𝒌¯|m𝒌𝒌¯=|m𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}\ket{m^{\prime}}_{\bm{k}\underline{\bm{k}}}=\ket{m^{\prime}}_{\bm{k}\underline{\bm{k}}}.

The spin-0 singlet state |0~𝒌𝒌¯\ket{\tilde{0}^{\prime}}_{\bm{k}\underline{\bm{k}}} is given as

|0~𝒌𝒌¯=12(|𝒌𝒌¯+|𝒌𝒌¯).\displaystyle\ket{{\tilde{0}}^{\prime}}_{\bm{k}\underline{\bm{k}}}=\frac{1}{\sqrt{2}}(\ket{\uparrow^{\prime}\downarrow^{\prime}}_{\bm{k}\underline{\bm{k}}}+\ket{\downarrow^{\prime}\uparrow^{\prime}}_{\bm{k}\underline{\bm{k}}})~{}. (92)

It satisfies Sμ𝒌𝒌¯|0~𝒌𝒌¯=0S^{\prime}_{\mu\bm{k}\underline{\bm{k}}}\ket{\tilde{0}^{\prime}}_{\bm{k}\underline{\bm{k}}}=0 (μ=x,y,z\mu=x,y,z). It has odd parity under 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}: 𝒞𝒌𝒌¯|0~𝒌𝒌¯=|0~𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}\ket{\tilde{0}^{\prime}}_{\bm{k}\underline{\bm{k}}}=-\ket{\tilde{0}^{\prime}}_{\bm{k}\underline{\bm{k}}}.

S𝒌𝒌¯±=Sx𝒌𝒌¯±iSy𝒌𝒌¯S^{\prime\pm}_{\bm{k}\underline{\bm{k}}}=S^{\prime}_{x\bm{k}\underline{\bm{k}}}\pm iS^{\prime}_{y\bm{k}\underline{\bm{k}}} are the ladder operators for the spin-1 triplet states {|m𝒌𝒌¯}\{\ket{m^{\prime}}_{\bm{k}\underline{\bm{k}}}\}:

S𝒌𝒌¯|1𝒌𝒌¯=S𝒌𝒌¯+|1𝒌𝒌¯=2|0𝒌𝒌¯,\displaystyle S^{\prime-}_{\bm{k}\underline{\bm{k}}}\ket{1^{\prime}}_{\bm{k}\underline{\bm{k}}}=S^{\prime+}_{\bm{k}\underline{\bm{k}}}\ket{-1^{\prime}}_{\bm{k}\underline{\bm{k}}}=\sqrt{2}\ket{0^{\prime}}_{\bm{k}\underline{\bm{k}}}~{}, (93)
S𝒌𝒌¯+|1𝒌𝒌¯=S𝒌𝒌¯|1𝒌𝒌¯=0,\displaystyle S^{\prime+}_{\bm{k}\underline{\bm{k}}}\ket{1^{\prime}}_{\bm{k}\underline{\bm{k}}}=S^{\prime-}_{\bm{k}\underline{\bm{k}}}\ket{-1^{\prime}}_{\bm{k}\underline{\bm{k}}}=0~{}, (94)
S𝒌𝒌¯±|0𝒌𝒌¯={2|1𝒌𝒌¯,2|1𝒌𝒌¯.\displaystyle S^{\prime\pm}_{\bm{k}\underline{\bm{k}}}\ket{0^{\prime}}_{\bm{k}\underline{\bm{k}}}=\left\{\begin{array}[]{l}\sqrt{2}\ket{1^{\prime}}_{\bm{k}\underline{\bm{k}}}~{},\\ \sqrt{2}\ket{-1^{\prime}}_{\bm{k}\underline{\bm{k}}}~{}.\end{array}\right. (97)

S~𝒌𝒌¯±=S~x𝒌𝒌¯±iS~y𝒌𝒌¯\tilde{S}^{\prime\pm}_{\bm{k}\underline{\bm{k}}}=\tilde{S}^{\prime}_{x\bm{k}\underline{\bm{k}}}\pm i\tilde{S}^{\prime}_{y\bm{k}\underline{\bm{k}}} transform |m𝒌𝒌¯\ket{m^{\prime}}_{\bm{k}\underline{\bm{k}}} into |0~𝒌𝒌¯\ket{\tilde{0}^{\prime}}_{\bm{k}\underline{\bm{k}}} and vice versa:

S~𝒌𝒌¯|1𝒌𝒌¯=S~𝒌𝒌¯+|1𝒌𝒌¯=2|0~𝒌𝒌¯,\displaystyle\tilde{S}^{\prime-}_{\bm{k}\underline{\bm{k}}}\ket{1^{\prime}}_{\bm{k}\underline{\bm{k}}}=-\tilde{S}^{\prime+}_{\bm{k}\underline{\bm{k}}}\ket{-1^{\prime}}_{\bm{k}\underline{\bm{k}}}=-\sqrt{2}\ket{\tilde{0}^{\prime}}_{\bm{k}\underline{\bm{k}}}~{}, (98)
S~𝒌𝒌¯+|1𝒌𝒌¯=S~𝒌𝒌¯|1𝒌𝒌¯=S~𝒌𝒌¯±|0𝒌𝒌¯=0,\displaystyle\tilde{S}^{\prime+}_{\bm{k}\underline{\bm{k}}}\ket{1^{\prime}}_{\bm{k}\underline{\bm{k}}}=\tilde{S}^{\prime-}_{\bm{k}\underline{\bm{k}}}\ket{-1^{\prime}}_{\bm{k}\underline{\bm{k}}}=\tilde{S}^{\prime\pm}_{\bm{k}\underline{\bm{k}}}\ket{0^{\prime}}_{\bm{k}\underline{\bm{k}}}=0~{}, (99)
S~𝒌𝒌¯+|0~𝒌𝒌¯=2|1𝒌𝒌¯,\displaystyle\tilde{S}^{\prime+}_{\bm{k}\underline{\bm{k}}}\ket{\tilde{0}^{\prime}}_{\bm{k}\underline{\bm{k}}}=-\sqrt{2}\ket{1^{\prime}}_{\bm{k}\underline{\bm{k}}}, (100)
S~𝒌𝒌¯|0~𝒌𝒌¯=2|1𝒌𝒌¯.\displaystyle\tilde{S}^{\prime-}_{\bm{k}\underline{\bm{k}}}\ket{\tilde{0}^{\prime}}_{\bm{k}\underline{\bm{k}}}=\sqrt{2}\ket{-1^{\prime}}_{\bm{k}\underline{\bm{k}}}~{}. (101)

Given the MF Hamiltonian in Eq. (11), using Eqs. (87) and (88), one obtains

[MF,S𝒌𝒌¯±]=±2E𝒌S𝒌𝒌¯±,\displaystyle\left[\mathcal{H}_{\rm MF},S^{\prime\pm}_{\bm{k}\underline{\bm{k}}}\right]=\pm 2E_{\bm{k}}S^{\prime\pm}_{\bm{k}\underline{\bm{k}}}~{}, (102)
[MF,S~𝒌𝒌¯±]=±2E𝒌S~𝒌𝒌¯±,\displaystyle\left[\mathcal{H}_{\rm MF},{\tilde{S}}^{\prime\pm}_{\bm{k}\underline{\bm{k}}}\right]=\pm 2E_{\bm{k}}\tilde{S}^{\prime\pm}_{\bm{k}\underline{\bm{k}}}~{}, (103)
[MF,S𝒌F±]=±2Δ0S𝒌F±.\displaystyle\left[\mathcal{H}_{\rm MF},S^{\prime\pm}_{\bm{k}_{F}}\right]=\pm 2\Delta_{0}S^{\prime\pm}_{\bm{k}_{F}}~{}. (104)

S𝒌𝒌¯±S^{\prime\pm}_{\bm{k}\underline{\bm{k}}}, S~𝒌𝒌¯±{\tilde{S}}^{\prime\pm}_{\bm{k}\underline{\bm{k}}}, and S𝒌F±S^{\prime\pm}_{\bm{k}_{F}} are the raising and lowering operators for the eigenstates of MF\mathcal{H}_{\rm MF}. |e𝒌𝒌¯0=S𝒌𝒌¯+|Ψ/2\ket{e_{\bm{k}\underline{\bm{k}}}^{0}}=S_{\bm{k}\underline{\bm{k}}}^{\prime+}\ket{\Psi}/\sqrt{2} and |e𝒌𝒌¯0~=S~𝒌𝒌¯+|Ψ/2\ket{e_{\bm{k}\underline{\bm{k}}}^{\tilde{0}}}={\tilde{S}}_{\bm{k}\underline{\bm{k}}}^{\prime+}\ket{\Psi}/\sqrt{2} are the degenerate excited states with excitation energy 2E𝒌2E_{\bm{k}} that involves a single pair of bogolons. They have even and odd parity under 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}, respectively. |e𝒌F=S𝒌F+|Ψ\ket{e_{\bm{k}_{F}}}={S^{\prime}}^{+}_{\bm{k}_{F}}\ket{\Psi} is an excited state with excitation energy 2Δ02\Delta_{0} that has odd parity under 𝒞𝒌F\mathcal{C}_{\bm{k}_{F}}.

Appendix B Holstein-Primakoff Theory applied to I\mathcal{H}_{I}

In this Section, we apply the Holstein-Primakoff theory to I\mathcal{H}_{I} in order to derive the Higgs mode. Substituting Eq. (75) into I\mathcal{H}_{I}, it can be written as

I=𝒌2ξ𝒌(cosφ𝒌Sz𝒌𝒌¯+sinφ𝒌Sx𝒌𝒌¯)\displaystyle\mathcal{H}_{I}=\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}}2\xi_{\bm{k}}(-\cos\varphi_{\bm{k}}S^{\prime}_{z\bm{k}\underline{\bm{k}}}+\sin\varphi_{\bm{k}}S^{\prime}_{x\bm{k}\underline{\bm{k}}})
g{𝒌,𝒌(cosφ𝒌cosφ𝒌Sx𝒌𝒌¯Sx𝒌𝒌¯+sinφ𝒌cosφ𝒌\displaystyle-g\left\{\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k},\bm{k}^{\prime}}(\cos\varphi_{\bm{k}}\cos\varphi_{\bm{k}^{\prime}}S^{\prime}_{x\bm{k}\underline{\bm{k}}}S^{\prime}_{x\bm{k}^{\prime}\underline{\bm{k}}^{\prime}}+\sin\varphi_{\bm{k}}\cos\varphi_{\bm{k}^{\prime}}\right.
×{Sz𝒌𝒌¯,Sx𝒌𝒌¯}+sinφ𝒌sinφ𝒌Sz𝒌𝒌¯Sz𝒌𝒌¯)\displaystyle\times\{S^{\prime}_{z\bm{k}\underline{\bm{k}}},S^{\prime}_{x\bm{k}^{\prime}\underline{\bm{k}}^{\prime}}\}+\sin\varphi_{\bm{k}}\sin\varphi_{\bm{k}^{\prime}}S^{\prime}_{z\bm{k}\underline{\bm{k}}}S^{\prime}_{z\bm{k}^{\prime}\underline{\bm{k}}^{\prime}})
+2𝒌𝒌F(sinφ𝒌Sz𝒌𝒌¯cosφ𝒌Sx𝒌𝒌¯)Sx𝒌F\displaystyle+2\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}}\sum_{\bm{k}_{F}}(-\sin\varphi_{\bm{k}}S^{\prime}_{z\bm{k}\underline{\bm{k}}}-\cos\varphi_{\bm{k}}S^{\prime}_{x\bm{k}\underline{\bm{k}}})S_{x\bm{k}_{F}}
+𝒌F,𝒌FSx𝒌FSx𝒌F}.\displaystyle\left.+\sum_{\bm{k}_{F},\bm{k}_{F}^{\prime}}S_{x\bm{k}_{F}}S_{x\bm{k}_{F}^{\prime}}\right\}. (105)

Since the eigenstates of I\mathcal{H}_{I} can be characterized by the parity under 𝒞𝒌F\mathcal{C}_{\bm{k}_{F}}, we can treat Sx𝒌FS_{x\bm{k}_{F}} as a cc-number Sx𝒌F=±1/2S_{x\bm{k}_{F}}=\pm 1/2. We quantize spin fluctuations around the ground state |Ψ\ket{\Psi} by the Holstein-Primakoff transformation [45]:

S𝒌𝒌¯+=2Sγ𝒌1γ𝒌γ𝒌2S,S𝒌𝒌¯=(S𝒌𝒌¯+),\displaystyle S^{\prime+}_{\bm{k}\underline{\bm{k}}}=\sqrt{2S}\gamma_{\bm{k}}^{\dagger}\sqrt{1-\frac{\gamma_{\bm{k}}^{\dagger}\gamma_{\bm{k}}}{2S}}~{},\quad S^{\prime-}_{\bm{k}\underline{\bm{k}}}=(S^{\prime+}_{\bm{k}\underline{\bm{k}}})^{\dagger}~{}, (106)
Sz𝒌𝒌¯=(Sγ𝒌γ𝒌),\displaystyle S^{\prime}_{z\bm{k}\underline{\bm{k}}}=-\left(S-\gamma_{\bm{k}}^{\dagger}\gamma_{\bm{k}}\right)~{}, (107)

where γ𝒌\gamma_{\bm{k}}^{\dagger} and γ𝒌\gamma_{\bm{k}} denote, respectively, the creation and annihilation operators of a boson that represents spin fluctuations. They satisfy the usual commutation relations [γ𝒌,γ𝒌]=δ𝒌,𝒌[\gamma_{\bm{k}},\gamma_{\bm{k}^{\prime}}^{\dagger}]=\delta_{\bm{k},\bm{k}^{\prime}} and [γ𝒌,γ𝒌]=[γ𝒌,γ𝒌]=0[\gamma_{\bm{k}},\gamma_{\bm{k}^{\prime}}]=[\gamma_{\bm{k}}^{\dagger},\gamma_{\bm{k}^{\prime}}^{\dagger}]=0. In view of the fact that the creation operator of the Higgs mode derived from BCS\mathcal{H}_{\rm BCS} in Ref. [42] has even parity under 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}, we restrict the Hilbert space spanned by even-parity states under 𝒞𝒌𝒌¯\mathcal{C}_{\bm{k}\underline{\bm{k}}}, in which 𝑺𝒌𝒌¯\bm{S}^{\prime}_{\bm{k}\underline{\bm{k}}} represents a spin-1 operator. We thus set S=1S=1. If spin fluctuation is small (γ𝒌γ𝒌1\gamma_{\bm{k}}^{\dagger}\gamma_{\bm{k}}\ll 1), S𝒌𝒌¯±S_{\bm{k}\underline{\bm{k}}}^{\prime\pm} can be approximated as S𝒌𝒌¯+2γ𝒌S_{\bm{k}\underline{\bm{k}}}^{\prime+}\simeq\sqrt{2}\gamma_{\bm{k}}^{\dagger} and S𝒌𝒌¯2γ𝒌S_{\bm{k}\underline{\bm{k}}}^{\prime-}\simeq\sqrt{2}\gamma_{\bm{k}}.

We expand Eq. (105) in terms of γ𝒌\gamma_{\bm{k}} and γ𝒌\gamma_{\bm{k}}^{\dagger}. The zeroth and first-order terms read

I(0)\displaystyle\mathcal{H}_{I}^{(0)} =\displaystyle= 2𝒌ξ𝒌2E𝒌Δ02g,\displaystyle-2\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}}\frac{\xi_{\bm{k}}^{2}}{E_{\bm{k}}}-\frac{\Delta_{0}^{2}}{g}~{}, (108)
I(1)\displaystyle\mathcal{H}_{I}^{(1)} =\displaystyle= 2𝒌(ξ𝒌sinφ𝒌+Δ0cosφ𝒌)(γ𝒌+γ𝒌).\displaystyle\sqrt{2}\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}}(\xi_{\bm{k}}\sin\varphi_{\bm{k}}+\Delta_{0}\cos\varphi_{\bm{k}})(\gamma_{\bm{k}}+\gamma_{\bm{k}}^{\dagger})~{}. (109)

Here, we have used the approximation

Δ0=g(𝒌sinφ𝒌+𝒌FSx𝒌F)g𝒌sinφ𝒌.\Delta_{0}=g(\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}}\sin\varphi_{\bm{k}}+\sum_{\bm{k}_{F}}S_{x\bm{k}_{F}})\simeq g\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}}\sin\varphi_{\bm{k}}~{}. (110)

Since I(1)\mathcal{H}_{I}^{(1)} should vanish, we obtain sinφ𝒌=Δ0/E𝒌\sin\varphi_{\bm{k}}=\Delta_{0}/E_{\bm{k}} and cosφ𝒌=ξ𝒌/E𝒌\cos\varphi_{\bm{k}}=-\xi_{\bm{k}}/E_{\bm{k}}. Equation (110) thus reduces to the gap equation

1=g𝒌1E𝒌.1=g\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}}\frac{1}{E_{\bm{k}}}~{}. (111)

The second-order term reads

I(2)\displaystyle\mathcal{H}_{I}^{(2)} =\displaystyle= 2𝒌E𝒌γ𝒌γ𝒌g2𝒌,𝒌cosφ𝒌cosφ𝒌\displaystyle 2\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}}E_{\bm{k}}\gamma_{\bm{k}}^{\dagger}\gamma_{\bm{k}}-\frac{g}{2}\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k},\bm{k}^{\prime}}\cos\varphi_{\bm{k}}\cos\varphi_{\bm{k}^{\prime}} (112)
×(γ𝒌γ𝒌+γ𝒌γ𝒌+γ𝒌γ𝒌+γ𝒌γ𝒌).\displaystyle\times(\gamma_{\bm{k}}\gamma_{\bm{k}^{\prime}}+\gamma_{\bm{k}}\gamma_{\bm{k}^{\prime}}^{\dagger}+\gamma_{\bm{k}}^{\dagger}\gamma_{\bm{k}^{\prime}}+\gamma_{\bm{k}}^{\dagger}\gamma_{\bm{k}^{\prime}}^{\dagger})~{}.

We diagonalize I(2)\mathcal{H}_{I}^{(2)} by the Bogoliubov transformation

βλ=𝒌(Uλ𝒌α𝒌+Vλ𝒌α𝒌),\displaystyle\beta_{\lambda}=\sum_{\bm{k}}(U_{\lambda\bm{k}}^{*}\alpha_{\bm{k}}+V_{\lambda\bm{k}}^{*}\alpha_{\bm{k}}^{\dagger})~{}, (113)
βλ=𝒌(Uλ𝒌α𝒌+Vλ𝒌α𝒌),\displaystyle\beta_{\lambda}^{\dagger}=\sum_{\bm{k}}(U_{\lambda\bm{k}}\alpha_{\bm{k}}^{\dagger}+V_{\lambda\bm{k}}\alpha_{\bm{k}})~{}, (114)

where λ\lambda labels the normal modes. The bosonic operator βλ\beta_{\lambda} satisfies the commutation relations

[βλ,βλ]=𝒌(Uλ𝒌Uλ𝒌Vλ𝒌Vλ𝒌)=δλ,λ,\displaystyle[\beta_{\lambda},\beta_{\lambda^{\prime}}^{\dagger}]=\sum_{\bm{k}}(U_{\lambda\bm{k}}^{*}U_{\lambda^{\prime}\bm{k}}-V_{\lambda\bm{k}}^{*}V_{\lambda^{\prime}\bm{k}})=\delta_{\lambda,\lambda^{\prime}}~{}, (115)
[βλ,βλ]=𝒌(Uλ𝒌Vλ𝒌+Vλ𝒌Uλ𝒌)=0.\displaystyle[\beta_{\lambda}^{\dagger},\beta_{\lambda^{\prime}}^{\dagger}]=\sum_{\bm{k}}(-U_{\lambda\bm{k}}V_{\lambda^{\prime}\bm{k}}+V_{\lambda\bm{k}}U_{\lambda^{\prime}\bm{k}})=0~{}. (116)

One can derive the inverse transformation from Eqs. (113) and (114) as

γ𝒌=λ(Uλ𝒌βλVλ𝒌βλ),\displaystyle\gamma_{\bm{k}}=\sum_{\lambda}(U_{\lambda\bm{k}}\beta_{\lambda}-V_{\lambda\bm{k}}^{*}\beta_{\lambda}^{\dagger})~{}, (117)
γ𝒌=λ(Uλ𝒌βλVλ𝒌βλ).\displaystyle\gamma_{\bm{k}}^{\dagger}=\sum_{\lambda}(U_{\lambda\bm{k}}^{*}\beta_{\lambda}^{\dagger}-V_{\lambda\bm{k}}\beta_{\lambda})~{}. (118)

Assuming that the second order term is diagonalized as I(2)=λωλβλβλ+const.\mathcal{H}_{I}^{(2)}=\sum_{\lambda}\omega_{\lambda}\beta_{\lambda}^{\dagger}\beta_{\lambda}+{\rm const.}, we obtain

[γ𝒌,I(2)]=λωλ(Uλ𝒌βλ+Vλ𝒌βλ).\displaystyle[\gamma_{\bm{k}},\mathcal{H}_{I}^{(2)}]=\sum_{\lambda}\omega_{\lambda}(U_{\lambda\bm{k}}\beta_{\lambda}+V_{\lambda\bm{k}}^{*}\beta_{\lambda}^{\dagger})~{}. (119)

On the other hand, using Eq. (112), we obtain

[γ𝒌,I(2)]\displaystyle[\gamma_{\bm{k}},\mathcal{H}_{I}^{(2)}]
=λ{(2E𝒌Uλ𝒌g𝒌cosφ𝒌cosφ𝒌(Uλ𝒌Vλ𝒌))βλ\displaystyle=\sum_{\lambda}\left\{\left(2E_{\bm{k}}U_{\lambda\bm{k}}-g\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}^{\prime}}\cos\varphi_{\bm{k}}\cos\varphi_{\bm{k}^{\prime}}(U_{\lambda\bm{k}^{\prime}}-V_{\lambda\bm{k}^{\prime}})\right)\beta_{\lambda}\right.
+(2E𝒌Vλ𝒌g𝒌cosφ𝒌cosφ𝒌(Uλ𝒌Vλ𝒌))}βλ.\displaystyle\left.+\left(-2E_{\bm{k}}V^{*}_{\lambda\bm{k}}-g\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}^{\prime}}\cos\varphi_{\bm{k}}\cos\varphi_{\bm{k}^{\prime}}(U^{*}_{\lambda\bm{k}^{\prime}}-V^{*}_{\lambda\bm{k}^{\prime}})\right)\right\}\beta^{\dagger}_{\lambda}~{}. (120)

For Eqs. (119) and (120) to be consistent, Xλ𝒌X_{\lambda\bm{k}} and Yλ𝒌Y_{\lambda\bm{k}} should satisfy

2E𝒌Uλ𝒌g(eλfλ)cosφ𝒌=ωλUλ𝒌,\displaystyle 2E_{\bm{k}}U_{\lambda\bm{k}}-g(e_{\lambda}-f_{\lambda})\cos\varphi_{\bm{k}}=\omega_{\lambda}U_{\lambda\bm{k}}~{}, (121)
2E𝒌Vλ𝒌g(eλfλ)cosφ𝒌=ωλVλ𝒌,\displaystyle-2E_{\bm{k}}V_{\lambda\bm{k}}-g(e_{\lambda}-f_{\lambda})\cos\varphi_{\bm{k}}=\omega_{\lambda}V_{\lambda\bm{k}}~{}, (122)

where the coefficients eλe_{\lambda} and fλf_{\lambda} are given by

eλ=𝒌cosφ𝒌Uλ𝒌,\displaystyle e_{\lambda}=\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}}\cos\varphi_{\bm{k}}U_{\lambda\bm{k}}~{}, (123)
fλ=𝒌cosφ𝒌Vλ𝒌.\displaystyle f_{\lambda}=\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}}\cos\varphi_{\bm{k}}V_{\lambda\bm{k}}~{}. (124)

If eλfλ0e_{\lambda}-f_{\lambda}\neq 0, Eqs (121) and (122) can be formally solved as

Uλ𝒌=g(eλfλ)cosφ𝒌2E𝒌ωλ,\displaystyle U_{\lambda\bm{k}}=g\frac{(e_{\lambda}-f_{\lambda})\cos\varphi_{\bm{k}}}{2E_{\bm{k}}-\omega_{\lambda}}~{}, (125)
Vλ𝒌=g(eλfλ)cosφ𝒌2E𝒌+ωλ.\displaystyle V_{\lambda\bm{k}}=-g\frac{(e_{\lambda}-f_{\lambda})\cos\varphi_{\bm{k}}}{2E_{\bm{k}}+\omega_{\lambda}}~{}. (126)

We omit λ\lambda below.

Substituting Eqs. (125) and (126) into Eqs. (123) and (124), we obtain

14g𝒌ξ𝒌2E𝒌2E𝒌4E𝒌2ω2=0.\displaystyle 1-4g\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}}\frac{\xi_{\bm{k}}^{2}}{E_{\bm{k}}^{2}}\frac{E_{\bm{k}}}{4E_{\bm{k}}^{2}-\omega^{2}}=0~{}. (127)

The above equation has a solution ω=2Δ0\omega=2\Delta_{0}, for which it reduces to the MF gap equation (111). We thus obtain

U𝒌\displaystyle U_{\bm{k}} =\displaystyle= Bcosφ𝒌2Δ02E𝒌,V𝒌=Bcosφ𝒌2Δ0+2E𝒌,\displaystyle\frac{{B}\cos\varphi_{\bm{k}}}{2\Delta_{0}-2E_{\bm{k}}},\quad V_{\bm{k}}=\frac{{B}\cos\varphi_{\bm{k}}}{2\Delta_{0}+2E_{\bm{k}}}~{}, (128)

where BB is the normalization constant. BB is determined by the normalization condition (115) as

B=1𝒌Δ0E𝒌ξ𝒌2.\displaystyle B=\frac{1}{\sqrt{\sum_{\bm{k}}^{\prime}\frac{\Delta_{0}}{E_{\bm{k}}\xi_{\bm{k}}^{2}}}}~{}. (129)

The creation operator for the collective excitation with ω=2Δ0\omega=2\Delta_{0} is thus given by

βH=B𝒌ξ𝒌E𝒌(γ𝒌2Δ02E𝒌+γ𝒌2Δ0+2E𝒌).\displaystyle\beta_{\rm H}^{\dagger}=B\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}}\frac{\xi_{\bm{k}}}{E_{\bm{k}}}\left(\frac{\gamma_{\bm{k}}^{\dagger}}{2\Delta_{0}-2E_{\bm{k}}}+\frac{\gamma_{\bm{k}}}{2\Delta_{0}+2E_{\bm{k}}}\right)~{}. (130)

βH\beta_{\rm H}^{\dagger} coincides with the creation operator of the Higgs mode derived from BCS\mathcal{H}_{\rm BCS} (See Eq. (C31) in Ref. [42]) using α𝒌α𝒌¯=2γ𝒌\alpha_{\bm{k}}^{\dagger}-\alpha_{\underline{\bm{k}}}^{\dagger}=\sqrt{2}\gamma_{\bm{k}}^{\dagger}. Here, α𝒌\alpha_{\bm{k}}^{\dagger} is a creation operator of a boson that describes fluctuations of the spin-1/2 operator 𝑺𝒌{\bm{S}}^{\prime}_{\bm{k}} (See Eqs. (C2) and (C3) in Ref. [42]).

In the limit of small fluctuations (γ𝒌γ𝒌1\gamma_{\bm{k}}^{\dagger}\gamma_{\bm{k}}\ll 1), using γ𝒌S𝒌𝒌¯+/2\gamma_{\bm{k}}^{\dagger}\simeq S^{\prime+}_{\bm{k}\underline{\bm{k}}}/\sqrt{2} and γ𝒌S𝒌𝒌¯/2\gamma_{\bm{k}}\simeq S^{\prime-}_{\bm{k}\underline{\bm{k}}}/\sqrt{2}, the creation operator for the Higgs mode can be written as

βH=B2𝒌ξ𝒌E𝒌(S𝒌𝒌¯+2Δ02E𝒌+S𝒌𝒌¯2Δ0+2E𝒌).\displaystyle\beta_{\rm H}^{\dagger}=\frac{B}{\sqrt{2}}\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}}\frac{\xi_{\bm{k}}}{E_{\bm{k}}}\left(\frac{S^{\prime+}_{\bm{k}\underline{\bm{k}}}}{2\Delta_{0}-2E_{\bm{k}}}+\frac{S^{\prime-}_{\bm{k}\underline{\bm{k}}}}{2\Delta_{0}+2E_{\bm{k}}}\right)~{}. (131)

We omit the normalization constant in Eq. (13) in the main text. The excited state with a single Higgs mode can be written in terms of |e𝒌𝒌¯0\ket{e^{0}_{\bm{k}\underline{\bm{k}}}} as

βH|Ψ=B𝒌ξ𝒌E𝒌12Δ02E𝒌|e𝒌𝒌¯0.\beta_{\rm H}^{\dagger}\ket{\Psi}=B\sideset{}{{}^{{}^{\prime}}}{\sum}_{\bm{k}}\frac{\xi_{\bm{k}}}{E_{\bm{k}}}\frac{1}{2\Delta_{0}-2E_{\bm{k}}}\ket{e^{0}_{\bm{k}\underline{\bm{k}}}}~{}. (132)