Emergence of multiple Higgs modes due to spontaneous breakdown of a 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 symmetry under the time-reversal operation in the pseudospin space.
We further predict the existence of multiple Higgs modes that have quantized energy (), where is the superconducting gap, is an integer, and is the number of states on the Fermi surface.
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 symmetries under discrete operations. We refer to them as the charge-conjugation (), parity (), and time-reversal () operations in analogy with the corresponding ones in the relativistic field theory lee-81 .
We found that and are spontaneously broken, while 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 symmetry under .
In this Letter, extending the previous work, we establish that the Higgs mode emerges in a superconductor due to spontaneous breakdown of symmetry under .
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 derived from the BCS Hamiltonian, which exhibits symmetries under , , and .
We show that the Higgs mode appears as the lowest excited state of due to spontaneous breakdown of symmetry under in the ground state.
Furthermore, we predict the existence of the multiple Higgs modes that have quantized energy () and include the original one (), where is the superconducting (SC) gap, is an integer, and is the number of states on the Fermi surface (FS).
Our main results are schematically illustrated in Fig. 1 (a).
Figure 1: (a) Schematic illustration of the effective double-well potential for . () is the SC ground state for the gap function (). The multiple Higgs modes with excitation energy () emerge due to spontaneous breakdown of symmetry under in the ground state . (b) Pseudospin configuration for the SC ground state with broken symmetry under . The pseudospins for () turn towards positive (negative) -direction as evolves from below to above the FS. We denote and .
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
(1)
(2)
where is the spin-1/2 pseudospin operator and is the total spin. We set throughout the Letter.
The fermion vacuum corresponds to the spin-down state () and the fully occupied state to the spin-up state (), where is the creation operator for a fermion with momentum and spin (). is the kinetic energy term and is the interaction term. Here, is the energy dispersion of a fermion measured from the chemical potential and is the coupling constant for the attractive interaction between fermions. We do not specify the form of for generality of argument.
has rotational U(1) symmetry about the -axis in the pseudospin space that reflects U(1) gauge symmetry. It is spontaneously broken in the SC state anderson-58 ; nambu-60 .
When satisfies the condition for the p-h symmetry
(3)
has additional symmetries under and tsuchiya-18 .
Here, and 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 ). , , and are defined as
(4)
(5)
(6)
where , , is a swapping operator between the states of and , and is the complex conjugation operator. Note that represents a rotation about the -axis that is an element of U(1).
and all other permutations of , , and are exact symmetries.
symmetries under and are spontaneously broken with U(1) in the SC state, while that under is unbroken tsuchiya-18 .
On the other hand, the Higgs mode appears as an amplitude mode only when 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 , amplitude and phase fluctuations of the gap function are proportional to those of the and components of the total spin as and , respectively.
and are uncoupled and pure amplitude oscillations of the Higgs mode are allowed only when is p-h symmetric, because the off-diagonal element of the dynamical spin susceptibility vanishes due to the opposite parity of and under tsuchiya-18 ; supplement .
When is not p-h symmetric, finite couples amplitude and phase fluctuations.
Moreover, the Higgs mode is stable only when 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 , because the Higgs mode has even parity and the single-particle excitations at have odd parity under .
When 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 under 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 component of the interaction term may be irrelevant to the Higgs mode, because the Higgs mode involves out-of-phase oscillations of and , whereas induces in-phase oscillations of them anderson-58 ; tsuchiya-18 ; supplement .
Furthermore, the analysis based on the Holstein-Primakoff (H-P) theory suggests that is not necessary for constructing the creation operator of the Higgs mode tsuchiya-18 .
Motivated by these observations, we neglect and study the Ising-like Hamiltonian :
(7)
(8)
Note that is invariant under , , and , despite it loses U(1) symmetry. We derive the Higgs mode from to demonstrate that the emergence of the Higgs mode is not due to spontaneous breakdown of U(1) but to that of under .
Discrete symmetries of .–
In addition to symmetries under , , and , has symmetries under local charge-conjugation operations in momentum space and , where can be written as .
To prove the invariance under , it is convenient to introduce a pseudospin operator
(9)
where is above the FS (). It commutes with .
The kinetic energy term can be written as
(10)
where . We obtain from Eq. (10). Since commutes with either and , we also obtain and .
Consequently, has symmetry under each of and .
and , in contrast, commute with neither nor .
Discrete symmetries in low-energy states of .–
We next study the symmetry of the ground state of .
The pseudospin configuration of it is shown in Fig. 1 (b): The pseudospins turn smoothly from up to down towards either positive or negative -direction as evolves from below to above the FS, where the gap function is positive and negative in the former and latter cases, respectively.
One of them being chosen spontaneously, under is broken in the ground state.
It leads to breakdown of under due to unbroken under , which is shown below, and the exact symmetry under .
The energy landscape of is effectively described by the double-well potential, as schematically illustrated in Fig. 1 (a).
Let us show that under each of , , and is unbroken in the ground state.
If we introduce a gap function , reduces to the mean-field (MF) Hamiltonian
(11)
where is the dispersion of a single-particle excitation referred to as a bogolon.
is the pseudospin operator for a pair of bogolons with and being the eigenstates of anderson-58 ; tsuchiya-18 ; supplement .
Note that and reduce to the same MF Hamiltonian for a real gap function.
Here, we have introduced a spin operator
(12)
where is above the FS ().
It commutes with . Using Eq. (11), we obtain . Consequently, symmetry under each of and as well as under is unbroken in the ground state of .
The eigenstates of can be characterized by parity under and .
The ground state of Eq. (11) has even parity for all and , where . Here, the product states
() are decomposed into the even and odd parity states under , where the former consist of the spin-1 triplet states () for and the latter the spin-0 singlet state supplement .
The ground state for has even and odd parity under and , respectively.
Parity of single-particle states plays a key role for the stability of the Higgs mode discussed below. () are degenerate single-particle states with excitation energy that form the two-particle continuum. They have even () and odd () parity under . They both have even parity under .
forms the lower edge of the two-particle continuum at that is degenerate with the Higgs mode. It has odd parity under and .
Emergence of Higgs mode due to breakdown of under .–
To derive the Higgs mode, we apply the H-P theory to holstein-40 ; supplement .
The second-order term in spin fluctuation can be diagonalized as . The collective mode has excitation energy and its creation operator is given as
(13)
where .
We find that coincides exactly with the creation operator of the Higgs mode derived from (See Eq. (27) in Ref. tsuchiya-18 ). Thus, represents the Higgs mode. The NG mode does not appear, because does not have U(1) symmetry.
The emergence of the Higgs mode from 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 or .
However, the Higgs mode disappears, for example, in the BEC regime of the BCS-BEC crossover, despite is broken with U(1) in the SC phase engelbrecht-97 .
Therefore, the emergence of the Higgs mode is considered due to breakdown of symmetry under .
This conclusion, together with the fact that the Higgs mode arises from , indicates that the SC phase transition for with a p-h symmetric is associated with breakdown of not U(1), but under . Namely, effectively reduces to for the SC phase transition when is p-h symmetric.
The Higgs mode is uncoupled with the single-particle states due to their opposite parity under despite their degeneracy at .
The Higgs mode has even parity for all and because . A single Higgs mode in fact consists of the single-particle states supplement .
Thus, the Higgs mode is a stable excitation of .
Strong-coupling perturbation theory.–
Figure 2: Low-lying energy eigenstates of and their energy shifts by . We denote (). The eigenstates of can be decomposed into the blocks represented by squares, where each block is characterized by the parity under . The multiple Higgs modes and 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 in the weak-coupling theory, we demonstrate it as well in a strong-coupling approach, where we treat as a perturbation to the unperturbed Hamiltonian . It is valid when is much larger than the band width, i.e., , where is the total number of wave vectors.
Given that the Higgs mode has even parity for all , we treat as a spin-1 operator to take into account only even parity states under .
Each eigenstate of has a corresponding eigenstate of , where the former transforms to the latter in the strong-coupling limit.
under is broken in the unperturbed ground states and , where () denote the spin-1 triplet states for supplement . They have gap functions and , respectively.
We consider perturbation to and the low-lying states above it shown in Fig. 2.
The energy eigenstates of can be decomposed into the blocks characterized by the parity under , which are represented as the squares in Fig. 2. has no matrix elements between states in different blocks.
If we focus on a block that has odd parity under for a set of Fermi wavevectors , each unperturbed state in this block can be obtained by operating on a corresponding state in the block of .
Flipping a pseudospin on the FS from to by yields a higher excited state as
(14)
where , , and .
The energy spectrum of in Fig. 2 is obtained by using Eq. (14).
The degeneracy among the first-excited states is lifted by diagonalizing the second-order effective Hamiltonian , where is the unperturbed energy of .
The effective Hamiltonian can be written as
(15)
where
(16)
Note that energy in Eq. (15) is measured from , where is the second-order energy shift of that corresponds to .
The first term in Eq. (15) represents the energy continuum above the threshold , as shown in Fig. 2.
In the limit , using , Eq. (18) reduces to . Remarkably, represents a bound state of that is formed at the lower edge of the energy continuum at .
On the other hand, we find in the strong-coupling limit using .
Thus, the Higgs mode arises in the strong-coupling theory.
is indeed stable, because it is uncoupled with .
In the weak-coupling regime, since corresponds to , is considered a bound state of formed at .
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 for the degenerate first excited states in the block of is given by
(19)
where
(20)
(21)
Here, is the number of wave vectors in .
In the limit , we find that has a Higgs-like bound state
(22)
Thus, combining with the original Higgs mode , the multiple Higgs modes () have quantized energy . They have odd parity under for Fermi wavevectors in . They should appear in the weak-coupling regime. They correspond to the states with quantized energy 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 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 under in a superconductor. We further predicted the existence of the multiple Higgs modes that have quantized energy ().
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 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).
(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 in Eq. (9) in the main text. We also introduce another pseudospin operator as
(23)
where is above the Fermi surface (FS) ().
is transformed by and as
(26)
(29)
Using Eqs. (26) and (29), one can derive the commutation relations
(30)
and also satisfy
(31)
(32)
(33)
The spin-1 triplet states for can be written as
(34)
(35)
(36)
Here, we denote ().
They satisfy () and have even parity under : .
The spin-1 triplet states for can be written as
(37)
(38)
(39)
where . They satisfy () and have even parity under : .
The spin-0 singlet state can be written as
(40)
(41)
It satisfies () and has odd parity under : . We note that the pseudospins for and are entangled in , , and .
are the ladder operators for the spin-1 triplet states :
(42)
(43)
(46)
transform the spin-1 triplet states into the spin-0 singlet state and vice versa as
(47)
(48)
(49)
(50)
In general, operating on a certain state changes its parity under , while operating does not as
(51)
(52)
where () denotes an even (odd) parity state under . 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 can be written as , where denotes the wave function for the pseudospins on the FS, which are zero modes for .
Using Eqs. (31) and (32), we obtain
(53)
(54)
The above equations show that and are the raising and lowering operators for the eigenstates of . Thus, and are the degenerate excited states with excitation energy involving a pair of fermions, where they have even and odd parity under , respectively.
We derive the fermion occupation number for . Using Eq. (52), we obtain
(55)
Specifically, from the -component of Eq. (55) we obtain the quantization of fermion occupation number
(56)
where is the fermion number operator.
We also obtain
(57)
where () is an even (odd) parity state under . The upper (lower) sign on the right-hand-side of Eq. (57) is for ().
From the -component of Eq. (57), we obtain
(58)
Therefore, the states and that appear in the eigenstates of are half-filled.
One can also derive some rigorous relations for the dynamical spin susceptibility
(59)
where is the Heisenberg representation and denotes the thermal average. We focus on and evaluate
(60)
where is the inverse temperature and denotes an eigenstate of : . Since operating changes the parity of the state under , we obtain and therefore and . On the other hand, one can also derive using .
We thus obtain
(61)
(62)
Equations (61) and (62) lead to the out-of-phase oscillations of spin fluctuations for the Higgs mode in the classical spin analysis [42].
Analogously, one can show
(63)
(64)
using , , and . We thus find that the dynamical spin susceptibilities for total spin vanish as and [42].
The expressions for the MF theory
In this Section, we summarize the algebra for the pseudospin operators for bogolons. is introduced as [42]
(75)
where and .
is rotated about the angle in the -plane in Eq. (75).
Using Eq. (75), the MF Hamiltonian can be written in the form of Eq. (11).
The eigenstates of are given as
(76)
(77)
where represents the vacuum of bogolons and the excited state with a pair of bogolons. All the pseudospins are aligned downward in the direction in the superconducting (SC) ground state. The BCS wave function can be thus written as
(78)
Note that , , and .
We introduce another pseudospin operator as
(79)
where is above the FS ().
is transformed by and as
(82)
(85)
Using Eqs. (82) and (85), one can derive the commutation relations
(86)
and obey the commutation relations
(87)
(88)
The spin-1 triplet states for can be written as
(89)
(90)
(91)
They satisfy (). They have even parity under : .
The spin-0 singlet state is given as
(92)
It satisfies (). It has odd parity under : .
are the ladder operators for the spin-1 triplet states :
(93)
(94)
(97)
transform into and vice versa:
(98)
(99)
(100)
(101)
Given the MF Hamiltonian in Eq. (11), using Eqs. (87) and (88), one obtains
(102)
(103)
(104)
, , and are the raising and lowering operators for the eigenstates of . and are the degenerate excited states with excitation energy that involves a single pair of bogolons. They have even and odd parity under , respectively. is an excited state with excitation energy that has odd parity under .
Appendix B Holstein-Primakoff Theory applied to
In this Section, we apply the Holstein-Primakoff theory to in order to derive the Higgs mode.
Substituting Eq. (75) into , it can be written as
(105)
Since the eigenstates of can be characterized by the parity under , we can treat as a -number .
We quantize spin fluctuations around the ground state by the Holstein-Primakoff transformation [45]:
(106)
(107)
where and denote, respectively, the creation and annihilation operators of a boson that represents spin fluctuations. They satisfy the usual commutation relations and .
In view of the fact that the creation operator of the Higgs mode derived from in Ref. [42] has even parity under , we restrict the Hilbert space spanned by even-parity states under , in which represents a spin-1 operator. We thus set .
If spin fluctuation is small (), can be approximated as and .
We expand Eq. (105) in terms of and . The zeroth and first-order terms read
(108)
(109)
Here, we have used the approximation
(110)
Since should vanish, we obtain and . Equation (110) thus reduces to the gap equation
(111)
The second-order term reads
(112)
We diagonalize by the Bogoliubov transformation
(113)
(114)
where labels the normal modes. The bosonic operator satisfies the commutation relations
(115)
(116)
One can derive the inverse transformation from Eqs. (113) and (114) as
(117)
(118)
Assuming that the second order term is diagonalized as , we obtain
For Eqs. (119) and (120) to be consistent, and should satisfy
(121)
(122)
where the coefficients and are given by
(123)
(124)
If , Eqs (121) and (122) can be formally solved as
(125)
(126)
We omit below.
Substituting Eqs. (125) and (126) into Eqs. (123) and (124), we obtain
(127)
The above equation has a solution , for which it reduces to the MF gap equation (111).
We thus obtain
(128)
where is the normalization constant. is determined by the normalization condition (115) as
(129)
The creation operator for the collective excitation with is thus given by
(130)
coincides with the creation operator of the Higgs mode derived from (See Eq. (C31) in Ref. [42]) using . Here, is a creation operator of a boson that describes fluctuations of the spin-1/2 operator (See Eqs. (C2) and (C3) in Ref. [42]).
In the limit of small fluctuations (), using and , the creation operator for the Higgs mode can be written as
(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 as