1 Introduction

The discovery of gravitational waves by the interferometer detector LIGO in 2015 [1] opened up multi-messenger astronomy, where electromagnetic waves, gravitational waves, neutrinos, and cosmic rays are utilized to explore the universe. In the future, as the history of electromagnetic wave astronomy tells us, multi-frequency gravitational wave observations will be required to boost the multi-messenger astronomy.

It is useful to review the current status of gravitational wave observations [1]. Note that the lowest frequency we can measure is around \(10^{-18}\) Hz, below which the wavelength of gravitational waves exceeds the current Hubble horizon. Measuring the temperature anisotropy and the B-mode polarization of the cosmic microwave background [2, 3], we can probe gravitational waves with frequencies between \(10^{-18}\) and \(10^{-16}\) Hz. Astrometry of extragalactic radio sources is sensitive to gravitational waves with frequencies between \(10^{-16}\) and \(10^{-9}\) Hz [4, 5]. The pulsar timing arrays, like EPTA [6, 7], IPTA [8] and NANOGrav [9], observe gravitational waves in the frequency band from \(10^{-9}\) to \(10^{-7}\) Hz. Doppler tracking of a space craft, which uses a measurement method similar to the pulsar timing arrays, can search for gravitational waves in the frequency band from \(10^{-7}\) to \(10^{-3}\) Hz [10]. The space interferometers LISA [11] and DECIGO [12] can cover the range between \(10^{-3}\) and 10 Hz. The interferometer detectors LIGO [13], Virgo [14], and KAGRA [15] with km size arm lengths can search for gravitational waves with frequencies from 10 to 1 kHz. In this frequency band, resonant bar experiments [16] are complementary to the interferometers [17]. Furthermore, interferometers can be used to measure gravitational waves with the frequencies between 1 kHz and 100 MHz. Recently, a limit on gravitational waves at MHz was reported [18] and a 0.75 m arm length interferometer gave an upper limit on 100 MHz gravitational waves [19]. At 100 MHz, there is a waveguide experiment using an interaction between gravitational waves and electromagnetic fields [20]. The interaction of gravitational waves with electromagnetic fields is useful to explore high frequency gravitational waves and has been studied extensively [21, 22]. Indeed, the interaction is utilized to constrain very high frequency gravitational waves higher than \(10^{14}\) Hz [23]. Although gravitational waves in the GHz range are theoretically interesting [24], no detector for GHz gravitational waves has been constructed.

In order to explore the GHz range, it would be useful to consider condensed matter systems. In our previous work, we pointed out that magnons in a cavity can be utilized to detect GHz gravitational waves [24]. There, we gave observational constraints on GHz gravitational waves for the first time. In this paper, we present the method in detail. To treat the general coordinate invariance appropriately, we need to use Fermi normal coordinates, or more precisely detector coordinates. Furthermore, we study non-relativistic fermions to reveal the interaction between magnons and gravitational waves. As a result, we obtain a formalism for non-relativistic fermions in curved spacetime, including a gravitational wave background as a special case. Finally, as a demonstration, we will give upper limits on the spectral density of continuous gravitational waves (95% CL): \(7.5 \times 10^{-19} \ [\mathrm{Hz}^{-1/2}]\) at 14 GHz and \(8.7 \times 10^{-18} \ [\mathrm{Hz}^{-1/2}]\) at 8.2 GHz, respectively, by utilizing results of magnon experiments conducted recently [25, 26].

The organization of the paper is as follows. In Sect. 2, we study the Dirac equation in Fermi normal coordinates. In Sect. 3, we take the non-relativistic limit to obtain a Hamiltonian of the fermions. In Sect. 4, we explain how to use magnons for detecting high frequency gravitational waves. Furthermore, we give upper limits on continuous gravitational waves in the GHz range. The final section is devoted to the conclusion. In Appendices A and B, we review how to derive Fermi normal coordinates and proper detector coordinates, respectively. In particular in Appendix B, the reason why one can neglect gravity of the earth and use the Fermi normal coordinates as the proper detector frames approximately will be clarified. We also give a simple mathematical formula for calculations in Appendix C.

2 Dirac field in Fermi normal coordinates

   In order to study the effects of gravity on a fermion, we consider the Dirac equation in curved spacetime described by a metric \(g_{\mu \nu }\). It is given by

$$\begin{aligned} i \gamma ^{\hat{\alpha }} e^{\mu }_{\hat{\alpha }} \left( \partial _{\mu } - \Gamma _{\mu } - i e A_{\mu } \right) \psi = m \psi \ , \end{aligned}$$
(1)

where \(\gamma ^{\hat{\alpha }}\), e, m, \(A_{\mu }\) are the gamma matrices, the elementary charge, the mass of the fermion, and the vector potential of U(1) gauge theory, respectively. The tetrad \(e^{\mu }_{\hat{\alpha }}\) satisfies

$$\begin{aligned} e^{\hat{\alpha }}_{\mu } e^{\hat{\beta }}_{\nu } \eta _{\hat{\alpha }\hat{\beta }} = g_{\mu \nu } \ . \end{aligned}$$
(2)

Note that \(\eta _{\hat{\alpha }\hat{\beta }}\) is the Minkowski metric of a local inertial frame and a hat is used for the frame. The spin connection is defined by

$$\begin{aligned} \Gamma _{\mu } = \frac{i}{2} e^{\hat{\alpha }}_{\nu } \sigma _{\hat{\alpha }\hat{\beta }} \left( \partial _{\mu } e^{\nu \hat{\beta }} + \Gamma ^{\nu }_{\lambda \mu } e^{\lambda \hat{\beta }} \right) , \end{aligned}$$
(3)

where \(\sigma _{\hat{\alpha }\hat{\beta }} = \frac{i}{4} [ \gamma _{\hat{\alpha }}, \gamma _{\hat{\beta }} ] \) is a generator of the Lorentz group and \(\Gamma ^{\mu }_{\nu \lambda }\) is the Christoffel symbol.

Since there is the equivalence principle for gravity, the choice of coordinates is quite important. We should consider a proper reference frame, which coincides with the coordinates used in an experiment. Actually, the proper reference frame can be approximated by Fermi normal coordinates (see Appendix A) because the effects of the earth are negligible for our purposes, as discussed in Appendix B.

In Appendix A, we have derived an explicit expression of the metric in Fermi normal coordinates:

$$\begin{aligned} g_{00}= & {} - 1 - R_{0i0j} x^{i} x^{j} \ , \end{aligned}$$
(4)
$$\begin{aligned} g_{0i}= & {} -\frac{2}{3} R_{0jik} x^{j} x^{k} \ , \end{aligned}$$
(5)
$$\begin{aligned} g_{ij}= & {} \delta _{ij} - \frac{1}{3} R_{ikjl} x^{k} x^{l} \ , \end{aligned}$$
(6)

where the Riemann tensor is evaluated at \(\varvec{x} = 0\) and thus it only depends on time \(x^{0}\). Moreover, the inverse of the metric is approximately given by

$$\begin{aligned} g^{00}= & {} - 1 + R^{0i0j} x_{i} x_{j} \ , \end{aligned}$$
(7)
$$\begin{aligned} g^{0i}= & {} +\frac{2}{3} R^{0jik} x_{j} x_{k} \ , \end{aligned}$$
(8)
$$\begin{aligned} g^{ij}= & {} \delta _{ij} + \frac{1}{3} R^{ikjl} x_{k} x_{l} \ , \end{aligned}$$
(9)

where we neglected higher order terms with respect to the curvature. From the metric (4)–(9), one can obtain the Christoffel symbols:

$$\begin{aligned} \Gamma ^{0}_{00}= & {} 0 \ , \quad \Gamma ^{0}_{0i} = R_{0i0j} x^{j} \ , \quad \Gamma ^{0}_{ij} = \frac{1}{3} \left( R_{0ijk} + R_{0jik} \right) x^{k} \ , \nonumber \\ \Gamma ^{i}_{00}= & {} R_{0i0j} x^{j} \ , \quad \Gamma ^{i}_{0j} = R_{0kji} x^{k} \ , \quad \nonumber \\ \Gamma ^{i}_{jk}= & {} \frac{1}{3} \left( R_{kijl} + R_{jikl} \right) x^{l} \ . \end{aligned}$$
(10)

The tetrad is constructed using Eq. (2):

$$\begin{aligned} e^{\hat{\alpha }}_{0}= & {} \delta ^{\hat{\alpha }}_{0} - \frac{1}{2} \delta ^{\hat{\alpha }}_{\alpha } R^{\alpha }_{\ k0l}x^{k} x^{l} \ , \end{aligned}$$
(11)
$$\begin{aligned} e^{\hat{\alpha }}_{i}= & {} \delta ^{\hat{\alpha }}_{i} - \frac{1}{6} \delta ^{\hat{\alpha }}_{\alpha } R^{\alpha }_{\ kil}x^{k} x^{l} \ . \end{aligned}$$
(12)
$$\begin{aligned} e^{0}_{\hat{\alpha }}= & {} \delta ^{0}_{\hat{\alpha }} + \frac{1}{2} \delta ^{0}_{\hat{\alpha }} R^{0}_{\ k0l} - \frac{1}{6} \delta ^{j}_{\hat{\alpha }} R_{jk0l} x^{k} x^{l} \ , \end{aligned}$$
(13)
$$\begin{aligned} e^{i}_{\hat{\alpha }}= & {} \delta ^{i}_{\hat{\alpha }} - \frac{1}{2} \delta ^{ 0}_{\hat{\alpha }} R^{0\ i}_{\ k\ l} x^{k} x^{l} + \frac{1}{6} \delta ^{j}_{\hat{\alpha }} R^{i}_{\ kjl} x^{k} x^{l} \ . \end{aligned}$$
(14)

Substituting Eqs. (10)–(14) into Eq. (3), we can evaluate the spin connection as

$$\begin{aligned} \Gamma _{0}= & {} \frac{1}{2} \gamma ^{{\hat{0}}} \gamma ^{{\hat{i}}} R_{0i0j} x^{j} + \frac{1}{4} \gamma ^{{\hat{i}}} \gamma ^{{\hat{j}}} R_{ij0k} x^{k} \ , \end{aligned}$$
(15)
$$\begin{aligned} \Gamma _{i}= & {} \frac{1}{4} \gamma ^{{\hat{0}}} \gamma ^{{\hat{j}}} R_{0jik} x^{k} + \frac{1}{8} \gamma ^{{\hat{j}}} \gamma ^{{\hat{k}}} R_{jkil} x^{l} \ . \end{aligned}$$
(16)

Here we have rewritten \(\delta _{\hat{\alpha }}^{\mu } \gamma ^{\hat{\alpha }}\) as \(\gamma ^{\hat{\mu }}\) and we will do so throughout.

On the other hand, the Dirac equation (1) can be rewritten as

$$\begin{aligned} i \gamma ^{0} \partial _{0} \psi= & {} \left[ i \gamma ^{0} \left( \Gamma _{0} + i e A_{0} \right) \nonumber \right. \\&\left. - i \gamma ^{j} \left( \partial _{j} - \Gamma _{j} - i e A_{j} \right) + m \right] \psi \nonumber \\= & {} \gamma ^{0} H \psi \ , \end{aligned}$$
(17)

where we defined a Hamiltonian H and the gamma matrices in curved spacetime \(\gamma ^{\mu } = e^{\mu }_{\hat{\alpha }} \gamma ^{\hat{\alpha }}\) satisfying the relation

$$\begin{aligned} \{ \gamma ^{\mu } , \gamma ^{\nu } \} = - 2 g^{\mu \nu } \ . \end{aligned}$$
(18)

Let us express the Hamiltonian in terms of the gamma matrices of the local inertial frame instead of those of curved spacetime. Because of \(\gamma ^{0}\gamma ^{0} = -g^{00}\), we obtain

$$\begin{aligned} H= & {} (g^{00})^{-1} \left[ i g^{00} \left( \Gamma _{0} + i e A_{0} \right) \nonumber \right. \\&\left. + i \gamma ^{0}\gamma ^{j} \left( \partial _{j} - \Gamma _{j} - i e A_{j} \right) - \gamma ^{0} m \right] \ . \end{aligned}$$
(19)

Using Eqs. (13) and (14), we calculate

$$\begin{aligned} \gamma ^{0} \gamma ^{j}= & {} \left( e^{0}_{\hat{\alpha }} \gamma ^{\hat{\alpha }} \right) \left( e^{j}_{\hat{\beta }} \gamma ^{\hat{\beta }} \right) \nonumber \\= & {} \left( e^{0}_{{\hat{0}}} \gamma ^{{\hat{0}}} + e^{0}_{{\hat{a}}} \gamma ^{{\hat{a}}} \right) \left( e^{j}_{{\hat{0}}} \gamma ^{{\hat{0}}} + e^{j}_{{\hat{b}}} \gamma ^{{\hat{b}}} \right) \nonumber \\= & {} \gamma ^{{\hat{0}}} \gamma ^{{\hat{j}}} - \frac{1}{2} \gamma ^{{\hat{0}}} \gamma ^{{\hat{0}}} R^{0\ j}_{\ k\ l} x^{k} x^{l}\nonumber \\&+ \frac{1}{6} \gamma ^{{\hat{0}}} \gamma ^{{\hat{b}}} R^{j}_{\ kbl} x^{k} x^{l} \nonumber \\&+ \frac{1}{2} \gamma ^{{\hat{0}}} \gamma ^{{\hat{j}}} R^{0}_{\ k0l} x^{k}x^{l} - \frac{1}{6} \gamma ^{{\hat{a}}} \gamma ^{{\hat{j}}} R_{ak0l} x^{k} x^{l} \ . \end{aligned}$$
(20)

Together with Eq. (7), we have

$$\begin{aligned} (g^{00})^{-1} \gamma ^{0} \gamma ^{j}\simeq & {} - \gamma ^{{\hat{0}}} \gamma ^{{\hat{j}}} - \frac{1}{2} R_{0kjl} x^{k} x^{l}\nonumber \\&- \frac{1}{6} \gamma ^{{\hat{0}}} \gamma ^{{\hat{a}}} R_{jkal} x^{k} x^{l} \nonumber \\&- \frac{1}{2} \gamma ^{{\hat{0}}} \gamma ^{{\hat{j}}} R_{0k0l} x^{k}x^{l} \nonumber \\&+ \frac{1}{6} \gamma ^{{\hat{a}}} \gamma ^{{\hat{j}}} R_{ak0l} x^{k} x^{l} \ . \end{aligned}$$
(21)

Similarly, one can obtain

$$\begin{aligned} (g^{00})^{-1} \gamma ^{0}\simeq & {} -\gamma ^{{\hat{0}}} - \frac{1}{2} \gamma ^{{\hat{0}}} R_{0k0l} x^{k}x^{l} \nonumber \\&+ \frac{1}{6} \gamma ^{{\hat{a}}} R_{ak0l} x^{k} x^{l} \ . \end{aligned}$$
(22)

Therefore, from Eqs. (19), (21) and (22), the Hamiltonian expressed in the local inertial coordinates becomes

$$\begin{aligned} H= & {} i \Gamma _{0} + i \gamma ^{{\hat{0}}} \gamma ^{{\hat{j}}} \Gamma _{j} - eA_{0} \nonumber \\&+ \bigg [ \gamma ^{{\hat{0}}} \gamma ^{{\hat{j}}} + \frac{1}{2} R_{0kjl} x^{k} x^{l} + \frac{1}{6} \gamma ^{{\hat{0}}} \gamma ^{{\hat{a}}} R_{jkal} x^{k} x^{l} \nonumber \\&+ \frac{1}{2} \gamma ^{{\hat{0}}} \gamma ^{{\hat{j}}} R_{0k0l} x^{k}x^{l} - \frac{1}{6} \gamma ^{{\hat{a}}} \gamma ^{{\hat{j}}} R_{ak0l} x^{k} x^{l} \bigg ]\nonumber \\&\left( -i \partial _{j} - eA_{j} \right) \nonumber \\&+ \left[ \gamma ^{{\hat{0}}} + \frac{1}{2} \gamma ^{{\hat{0}}} R_{0k0l} x^{k} x^{l} - \frac{1}{6} \gamma ^{{\hat{a}}} R_{ak0l} x^{k} x^{l} \right] m \ . \end{aligned}$$
(23)

Furthermore, substituting Eqs. (15) and (16) into the above Hamiltonian and rearranging terms, we have

$$\begin{aligned} H= & {} \frac{i}{2} \gamma ^{{\hat{0}}} \gamma ^{{\hat{i}}} R_{0i0j} x^{j} + \frac{i}{4} \gamma ^{{\hat{i}}} \gamma ^{{\hat{j}}} R_{0ikj} x^{k}\nonumber \\&+ \frac{i}{8} \gamma ^{{\hat{0}}} \gamma ^{{\hat{i}}} \gamma ^{{\hat{j}}} \gamma ^{{\hat{k}}} R_{jkil} x^{l} - eA_{0} \nonumber \\&+ \bigg [ \gamma ^{{\hat{0}}} \gamma ^{{\hat{i}}} \left( \delta ^{j}_{i} + \theta ^{j}_{i} \right) \nonumber \\&+ \frac{1}{2} R_{0kjl} x^{k} x^{l} - \frac{1}{6} \gamma ^{{\hat{i}}} \gamma ^{{\hat{j}}} R_{ik0l} x^{k} x^{l} \bigg ] \left( -i \partial _{j} - eA_{j} \right) \nonumber \\&+ \gamma ^{{\hat{0}}} \left[ 1 + \frac{1}{2} R_{0k0l} x^{k} x^{l} - \frac{1}{6} \gamma ^{{\hat{0}}} \gamma ^{{\hat{i}}} R_{ik0l} x^{k} x^{l} \right] m \ , \end{aligned}$$
(24)

where we defined

$$\begin{aligned} \theta ^{j}_{i} = \frac{1}{2} \delta ^{j}_{i} R_{0k0l} x^{k}x^{l} + \frac{1}{6} R_{jkil} x^{k} x^{l} \ . \end{aligned}$$
(25)

The result is consistent with the earlier work [27] where the Hamiltonian (24) was obtained to examine the effects of gravity on an atom.

The Hamiltonian we have obtained is the \(4 \times 4\) matrix including both the particle and the anti-particle. What we will consider is the non-relativistic fermion. To take the non-relativistic limit of the Hamiltonian of the fermion, we have to separate the particle and the anti-particle while expanding the Hamiltonian in powers of 1/m. We will explicitly see how to perform this in the next section.

3 Non-relativistic limit of Dirac equation

In the previous section, we derived the Hamiltonian of a Dirac field in general curved spacetime with Fermi normal coordinates. Assuming that a fermion has a velocity well below the speed of light, which is the situation we will consider in the Sect. 4, we take the non-relativistic limit of the Hamiltonian. The procedure in flat spacetime is known as the Foldy–Wouthuysen transformation [28, 29]. We generalize it to the case of curved spacetime.

We first separate the Hamiltonian (24) into the even part, the odd part and the terms multiplied by m as

$$\begin{aligned} H= & {} \frac{i}{2} \alpha ^{i} R_{0i0j} x^{j} - \frac{i}{8} \alpha ^{i} \alpha ^{j} \alpha ^{k} R_{jkil} x^{l}\nonumber \\&+ \alpha ^{i} \left( \delta ^{j}_{i} + \theta ^{j}_{i} \right) \Pi _{j} \nonumber \\&- eA_{0} - \frac{i}{4} \alpha ^{i} \alpha ^{j} R_{0ikj} x^{k} \nonumber \\&+ \bigg [ \frac{1}{2} R_{0kjl} x^{k} x^{l} + \frac{1}{6} \alpha ^{i} \alpha ^{j} R_{ik0l} x^{k} x^{l} \bigg ] \Pi _{j} \nonumber \\&+ \left[ \beta \left( 1 + \frac{1}{2} R_{0k0l} x^{k} x^{l} \right) - \frac{1}{6} \beta \alpha ^{i} R_{ik0l} x^{k} x^{l} \right] m \nonumber \\= & {} {\mathcal {O}} + {\mathcal {E}} + \left[ \beta \left( 1 + \frac{1}{2} R_{0k0l} x^{k} x^{l} \right) \nonumber \right. \\&\left. - \frac{1}{6} \beta \alpha ^{i} R_{ik0l} x^{k} x^{l} \right] m \ , \end{aligned}$$
(26)

where we have defined \(\beta = \gamma ^{{\hat{0}}}\), \(\alpha ^{i} = \gamma ^{{\hat{0}}} \gamma ^{{\hat{i}}}\) and \(\Pi _{j} = -i \partial _{j} - eA_{j}\) for brevity. The even part, \({\mathcal {E}}\), means that the matrix has only block diagonal elements and the odd part, \({\mathcal {O}}\), means that the matrix has only block off-diagonal elements. Any product of two even (odd) matrices is even and a product of even (odd) and odd (even) matrices becomes odd. To take the non-relativistic limit of the Hamiltonian, we have to diagonalize the Hamiltonian (26) and expand the upper block diagonal part in powers of 1/m. More precisely, 1/m expansion is recognized as an expansion with respect to two parameters, \((m x)^{-1}\) and v/c. Here, x represents a typical length scale of the system which can be specified by Fermi normal coordinates, i.e., \(x \sim \sqrt{x^{i} x_{i}}\), v is the velocity of the fermion and c denotes the speed of light. Assuming \(1/m x \ll 1 \) and \(v/c \ll 1\), which hold in the situation of the Sect. 4, we will perform the 1/m expansion. It is known that this can be done in flat spacetime by repeating unitary transformations order by order in powers of 1/m [28, 29]. Let us generalize the method to the case of curved spacetime in Fermi normal coordinates.

We consider a unitary transformation,

$$\begin{aligned} \psi ' = e^{iS} \psi \ , \end{aligned}$$
(27)

where S is a time-dependent Hermitian 4 \(\times \) 4 matrix. Observing that

$$\begin{aligned} i \frac{\partial \psi '}{\partial t}= & {} i \frac{\partial }{\partial t} \left( e^{iS} \psi \right) \nonumber \\= & {} e^{iS} \left( i \frac{\partial \psi }{\partial t} \right) + i \left( \frac{\partial }{\partial t} e^{iS} \right) \psi \nonumber \\= & {} \left[ e^{iS} H e^{-iS} + i \left( \frac{\partial }{\partial t} e^{iS} \right) e^{-iS} \right] \psi ' \ , \end{aligned}$$
(28)

we find that the Hamiltonian after the unitary transformation is given by

$$\begin{aligned} H' = e^{iS} H e^{-iS} + i \left( \frac{\partial }{\partial t} e^{iS} \right) e^{-iS} \ . \end{aligned}$$
(29)

We now assume that S is proportional to powers of 1/m and expand the transformed Hamiltonian (29) in powers of S up to the order of 1/m. Using Eqs. (C4) and (C7) in Eq. (29), we obtain

$$\begin{aligned} H'= & {} H + i \big [ S,H \big ] - \frac{1}{2} \big [ S,\big [S,H \big ] \big ]\nonumber \\&- \frac{i}{6} \big [S,\big [S, \big [S , H \big ] \big ] \big ] + \cdots \nonumber \\&- {\dot{S}} - \frac{i}{2} \big [ S,{\dot{S}} \big ] + \cdots \ . \end{aligned}$$
(30)

First, let us eliminate the off-diagonal part of the Hamiltonian (26) at the order of m by a unitary transformation. Then we will drop the higher order terms with respect to the Riemann tensor, which only depends on time, and derivatives of the Riemann tensor with respect to the time by assuming that they are small enough.Footnote 1

To cancel the last term in the square bracket of (26), we take

$$\begin{aligned} S = - \frac{i}{2m}\beta \left( - \frac{1}{6} \beta \alpha ^{i} R_{ik0l} x^{k} x^{l} m \right) \ . \end{aligned}$$
(31)

We then obtain

$$\begin{aligned} i \big [ S,H \big ]\simeq & {} \frac{1}{6} \beta \alpha ^{i} R_{ik0l} x^{k}x^{l} m - \frac{1}{12} \big [ \alpha ^{i} , \alpha ^{j} \big ] R_{ik0l} x^{k} x^{l} \Pi _{j} \nonumber \\&+ \frac{i}{6} \alpha ^{i}\alpha ^{j} R_{0ikj} x^{k} + \frac{i}{12} \alpha ^{i}\alpha ^{j} R_{0jik} x^{k} \ . \end{aligned}$$
(32)

Therefore, from Eqs. (30) and (32), we have the transformed Hamiltonian as

$$\begin{aligned} H'\simeq & {} H + i \big [ S,H \big ] \nonumber \\\simeq & {} \frac{i}{2} \alpha ^{i} R_{0i0j} x^{j} - \frac{i}{8} \alpha ^{i} \alpha ^{j} \alpha ^{k} R_{jkil} x^{l} + \alpha ^{i} \left( \delta ^{j}_{i} + \theta ^{j}_{i} \right) \Pi _{j} \nonumber \\&- eA_{0} - \frac{i}{6} R_{0iki} x^{k} + \frac{2}{3} R_{0kil} x^{k} x^{l} \Pi _{i} \nonumber \\&+ \beta \left( 1 + \frac{1}{2} R_{0k0l} x^{k} x^{l} \right) m \nonumber \\= & {} {\mathcal {O}} + {\mathcal {E}}' + \beta \left( 1 + \frac{1}{2} R_{0k0l} x^{k} x^{l} \right) m \ , \end{aligned}$$
(33)

where we have used the relation \(\big \{ \alpha ^{i} ,\alpha ^{j} \big \} = 2 \delta ^{ij}\). One can see that only even terms remain at the order of m, as expected.

Next, we focus on the order of \(m^{0}\) and eliminate the odd terms by a unitary transformation. In order to do so, we choose the Hermitian operator to be

$$\begin{aligned} S'= & {} -\frac{i}{2m} \beta \left( {\mathcal {O}} -\frac{1}{2} \alpha ^{j} R_{0k0l} x^{k} x^{l} \Pi _{j} \nonumber \right. \\&\left. + \frac{i}{2} \alpha ^{j} R_{0k0j} x^{k} \right) \ . \end{aligned}$$
(34)

It is straightfoward to obtain

$$\begin{aligned} i \big [ S',H' \big ]\simeq & {} - {\mathcal {O}} + \frac{1}{m} \beta {\mathcal {O}}^{2} + \frac{1}{2m} \beta \big [ {\mathcal {O}}, {\mathcal {E}}' \big ] \nonumber \\&- \frac{1}{2m} \beta \alpha ^{i} \alpha ^{j} R_{0k0l} x^{k} x^{l} \Pi _{i} \Pi _{j} + \frac{i}{m} \beta R_{0k0i} x^{k} \Pi _{i} \nonumber \\&+ \frac{i}{4m} \beta \big [ \alpha ^{i} ,\alpha ^{j} \big ] R_{0k0i} x^{k} \Pi _{j} + \frac{1}{4m} \beta R_{0i0i} \nonumber \\&- \frac{i}{4m} \beta \alpha ^{j} R_{0k0l} x^{k} x^{l} \left( \partial _{j} e A_{0} \right) \ , \end{aligned}$$
(35)

Furthermore, up to the order of 1/m, we can deduce

$$\begin{aligned} - \frac{1}{2} \big [ S',\big [S',H' \big ] \big ]\simeq & {} - \frac{1}{2} \big [ S', i {\mathcal {O}} \big ] \nonumber \\\simeq & {} - \frac{1}{2m} \beta {\mathcal {O}}^{2} \nonumber \\&+ \frac{1}{4m} \beta \alpha ^{i} \alpha ^{j} R_{0k0l} x^{k} x^{l} \Pi _{i} \Pi _{j} \nonumber \\&- \frac{i}{2m} \beta R_{0k0i} x^{k} \Pi _{i} \nonumber \\&- \frac{i}{8m} \beta \big [ \alpha ^{i}, \alpha ^{j} \big ] R_{0k0i} x^{k} \Pi _{j} \nonumber \\&- \frac{1}{8m} \beta R_{0i0i} \ , \end{aligned}$$
(36)

and

$$\begin{aligned} - {\dot{S}}' \simeq \frac{i}{2m} \beta \dot{{\mathcal {O}}} + \frac{i}{4m} \beta \alpha ^{j} R_{0k0l} x^{k} x^{l} e {\dot{A}}_{j} \ . \end{aligned}$$
(37)

Therefore, the Hamiltonian after the unitary transformation is given by

$$\begin{aligned} H''\simeq & {} H' + i \big [ S',H' \big ] - \frac{1}{2} \big [ S',\big [S',H' \big ] \big ] - {\dot{S}}' \nonumber \\\simeq & {} - \frac{i}{4m} \beta \alpha ^{j} R_{0k0l} x^{k} x^{l} e E_{j} + \frac{1}{2m} \beta \left( \big [ {\mathcal {O}}, {\mathcal {E}}' \big ] + i \dot{{\mathcal {O}}} \right) \nonumber \\&+ {\mathcal {E}}' + \frac{1}{2m} \beta {\mathcal {O}}^{2} - \frac{1}{4m} \beta \alpha ^{i} \alpha ^{j} R_{0k0l} x^{k} x^{l} \Pi _{i} \Pi _{j} \nonumber \\&+ \frac{i}{2m} \beta R_{0k0i} x^{k} \Pi _{i} + \frac{i}{8m} \beta \big [ \alpha ^{i} ,\alpha ^{j} \big ] R_{0k0i} x^{k} \Pi _{j}\nonumber \\&+ \frac{1}{8m} \beta R_{0i0i} \nonumber \\&+ \beta \left( 1 + \frac{1}{2} R_{0k0l} x^{k} x^{l} \right) m \nonumber \\= & {} {\mathcal {O}}' + {\mathcal {E}}'' + \beta \left( 1 + \frac{1}{2} R_{0k0l} x^{k} x^{l} \right) m \ , \end{aligned}$$
(38)

where \(E_{j} \equiv \partial _{j} A_{0} - {\dot{A}}_{j}\) is an electric field. We see that \({\mathcal {O}}'\) has only terms of order of 1/m, so that odd terms at the order of \(m^{0}\) have been eliminated.

Finally, we will eliminate the odd term \({\mathcal {O}}'\) and then the Hamiltonian will consist of only even terms up to the order of 1/m, which we want to get. To this end, we now choose the Hermitian operator of a unitary transformation as

$$\begin{aligned} S'' = -\frac{i}{2m} \beta \left( {\mathcal {O}}' - \frac{i}{4m} \beta \alpha ^{i} e E_{i} R_{0k0l} x^{k} x^{l} \right) \ . \end{aligned}$$
(39)

Then, up to the order of 1/m, we have

$$\begin{aligned} i \big [ S'' , H'' \big ] \simeq - {\mathcal {O}}' \ . \end{aligned}$$
(40)

Therefore, we have the transformed Hamiltonian as

$$\begin{aligned} H'''\simeq & {} H'' + i \big [ S'' , H'' \big ] \nonumber \\\simeq & {} {\mathcal {E}}'' + \beta \left( 1 + \frac{1}{2} R_{0k0l} x^{k} x^{l} \right) m \ , \end{aligned}$$
(41)

where \({\mathcal {E}}''\) is given by

$$\begin{aligned} {\mathcal {E}}''= & {} - eA_{0} - \frac{i}{6} R_{0iki} x^{k} + \frac{2}{3} R_{0kil} x^{k} x^{l} \Pi _{i} + \frac{1}{2m} \beta {\mathcal {O}}^{2}\nonumber \\&- \frac{1}{4m} \beta \alpha ^{i} \alpha ^{j} R_{0k0l} x^{k} x^{l} \Pi _{i} \Pi _{j} \nonumber \\&+ \frac{i}{2m} \beta R_{0k0i} x^{k} \Pi _{i} + \frac{i}{8m} \beta \big [ \alpha ^{i} ,\alpha ^{j} \big ] R_{0k0i} x^{k} \Pi _{j} \nonumber \\&+ \frac{1}{8m} \beta R_{0i0i} \ . \end{aligned}$$
(42)

Moreover, the fourth term in the first line of Eq. (42) can be evaluated as

$$\begin{aligned} \frac{1}{2m} \beta {\mathcal {O}}^{2}\simeq & {} \frac{i}{8m}\beta \big [ \alpha ^{i} ,\alpha ^{j} \big ] \epsilon _{ilm} e B^{m} \left( \delta _{lj} + 2 \theta _{lj} \right) \nonumber \\&-\frac{i}{4m}\beta \big [ \alpha ^{i} ,\alpha ^{j} \big ] \left( \frac{1}{4} R_{lmji} + \delta ^{l}_{j}R_{0i0m} \right) x^{m} \Pi _{l} \nonumber \\&+ \frac{1}{2m}\beta \left( \delta _{ij} + 2 \theta _{ij} \right) \Pi _{i} \Pi _{j} \nonumber \\&+ \frac{i}{12 m}\beta R_{kikj} x^{j} \Pi _{i} + \frac{1}{4m} \beta R_{0i0i} \nonumber \\&+ \frac{1}{16m} \beta \alpha ^{i} \alpha ^{j} \alpha ^{k} \alpha ^{l} R_{ijkl} \nonumber \\&- \frac{i}{16m} \beta \big \{ \alpha ^{i} , \alpha ^{j} \alpha ^{k} \alpha ^{l} \big \} R_{kljm} x^{m} \Pi _{i} \ , \end{aligned}$$
(43)

where \(B^{i} \equiv \frac{1}{2} \epsilon ^{ijk} (\partial _{j}A_{k} - \partial _{k}A_{j})\) is a magnetic field. Using Eqs. (42), (43) and the relation \(\big [ \alpha ^{i} ,\alpha ^{j} \big ] = 2 i \epsilon _{ijk} \sigma ^{k}\) in the transformed Hamiltonian (41), we finally arrive at the Hamiltonian for a non-relativistic fermion up to the order of 1/m as

$$\begin{aligned} H'''= & {} \left( 1 + \frac{1}{2} R_{0k0l} x^{k} x^{l} \right) m - eA_{0} - \frac{i}{6} R_{0iki} x^{k}\nonumber \\&+ \frac{2}{3} R_{0kil} x^{k} x^{l} \Pi _{i} \nonumber \\&+ \frac{1}{2m} \left[ \delta _{ij} \left( 1 + \frac{1}{2} R_{0k0l} x^{k}x^{l} \right) \nonumber \right. \\&\left. + \frac{1}{3} R_{jkil} x^{k} x^{l} \right] \Pi _{i} \Pi _{j} \nonumber \\&- \frac{e}{2m} \sigma ^{i} B^{j} \left[ \delta _{ij} \left( 1 + \frac{1}{2} R_{0k0l} x^{k} x^{l} \nonumber \right. \right. \\&\left. \left. + \frac{1}{6} R_{mkml} x^{k} x^{l} \right) - \frac{1}{6} R_{ikjl} x^{k} x^{l} \right] \nonumber \\&+ \frac{1}{8m} \epsilon _{ijk} \sigma ^{k} \left( R_{ijlm} + 2 \delta _{jm} R_{0i0l} \right) x^{l} \Pi _{m} \nonumber \\&+ \frac{1}{8 m} \left( 3 R_{0i0i} - R_{ijij} \right) \nonumber \\&+ \frac{i}{2 m} \left( R_{0i0j} - \frac{1}{3} R_{kikj} \right) x^{i} \Pi _{j} \ . \end{aligned}$$
(44)

The first term is the rest mass and its correction from the gravity at a point \(x^{i}\). The third term represents gravitational redshift, namely energy shift due to gravity. The first term in the last line gives the same effect at the order of 1/m. We find that the fourth and the fifth terms describe gravitational effects on the motion of a particle. However, we notice that the former contains the time derivative of the curvature, which has been assumed to be small. The second term in the last line is also small. The third line represents interactions between gravity and a spin in the presence of an external magnetic field. This is what causes the spin resonance and/or the excitation of magnons as we will see in the next section. The fourth line is a spin–orbit coupling mediated by gravity.

In vacuum, the Riemann tensor coincides with the Weyl tensor. Then it may be useful to rewrite the Riemann tensor of the Hamiltonian (44) in terms of the electric \(E_{ij}\) and magnetic \(H_{ij}\) components of the Weyl tensor \(C_{\mu \nu }^{\ \ \rho \sigma }\), defined by

$$\begin{aligned} R_{\mu \nu }^{\ \ \rho \sigma }= & {} C_{\mu \nu }^{\ \ \rho \sigma } = 4 \left( \gamma ^{[\rho }_{[\mu } - \delta ^{|0|}_{[\mu } \delta ^{[\rho }_{|0|} \right) E^{\sigma ]}_{\nu ]} \nonumber \\&+ 2 \epsilon _{\mu \nu \alpha 0} \delta ^{[\rho }_{0} H^{\sigma ]\alpha } + 2 \epsilon ^{\rho \sigma \alpha 0} \delta ^{0}_{[\mu } H_{\nu ]\alpha } \ , \end{aligned}$$
(45)

where \(\gamma _{\mu \nu }\) is the induced three dimensional metric, i.e., \(\gamma _{00} = \gamma _{0i} = 0, \gamma _{ij} \simeq \delta _{ij}\). Substituting the above relation into the Hamiltonian (44), we obtain

$$\begin{aligned} H'''= & {} \left( 1 + \frac{1}{2} E_{kl} x^{k} x^{l} \right) m - eA_{0} - \frac{i}{6} H_{il}\epsilon _{kil} x^{k}\nonumber \\&+ \frac{2}{3} H_{km} \epsilon _{ilm} x^{k} x^{l} \Pi _{i} \nonumber \\&+ \frac{1}{2m} \left[ \delta _{ij} \left( 1 + \frac{1}{2} E_{kl} x^{k} x^{l} \right) \nonumber \right. \\&\left. + \frac{4}{3} \delta _{[j|[i} E_{l]|k]} x^{k} x^{l} \right] \Pi _{i}\Pi _{j} \nonumber \\&- \frac{e}{2m} \sigma ^{i} B^{j} \left[ \delta _{ij} \left( 1 + \frac{1}{2} E_{kl} x^{k} x^{l}\nonumber \right. \right. \\&\left. \left. + \frac{2}{3} \delta _{[m|[m} E_{l]|k]} x^{k} x^{l} \right) - \frac{2}{3} \delta _{[i|[j} E_{l]|k]} x^{k} x^{l} \right] \nonumber \\&+ \frac{1}{4m} \epsilon _{ijk} \sigma ^{k} \left( 2 \delta _{[i|[l} E_{m]|j]} + \delta _{jm} E_{il} \right) x^{l} \Pi _{m} \nonumber \\&+ \frac{1}{2m} \left( \frac{3}{4} E_{ii} - \delta _{[i|[i} E_{j]|j]} \right) \nonumber \\&+ \frac{i}{2m} \left( E_{ij} - \frac{4}{3} \delta _{[k|[k} E_{j]|i]} \right) x^{i} \Pi _{j} \ . \end{aligned}$$
(46)

Although Eq. (44) is applicable to a general curved spacetime, let us focus on gravitational waves as gravitational effects from now on. The Riemann tensor for a perturbed metric \(g_{\mu \nu } = \eta _{\mu \nu } + h_{\mu \nu }\) at the linear order is given by

$$\begin{aligned} R^{\alpha }{}_{\mu \beta \nu } = \frac{1}{2}(h^{\alpha }_{\ \nu ,\mu \beta }-h_{\mu \nu \ \beta }^{\ \ ,\alpha } -h^{\alpha }_{\ \beta ,\mu \nu }+h_{\mu \beta \ ,\nu }^{\ \ ,\alpha }) \ , \end{aligned}$$
(47)

where \(\eta _{\mu \nu }\) stands for a flat spacetime metric and \(h_{\mu \nu }\) represents a deviation from the flat spacetime. Because the Riemann tensor (47) is invariant under gauge transformations, we can use any coordinate to evaluate the Riemann tensor included in Eq. (44). We then take the transverse traceless gauge, i.e., \(h_{0\mu } = h_{ii} = h_{ij,j} = 0\). As a result, one can obtain

$$\begin{aligned} R_{0i0j}= & {} - \frac{1}{2} \ddot{h}_{ij} \ , \nonumber \\ R_{0ijk}= & {} \frac{1}{2} \left( {\dot{h}}_{ij,k} - {\dot{h}}_{ik,j} \right) \ , \nonumber \\ R_{ijkl}= & {} \frac{1}{2} \left( h_{il,jk} + h_{jk,il} - h_{jl,ik} - h_{ik,jl} \right) \ . \end{aligned}$$
(48)

Note that they are evaluated at the origin, \(x^{i} = 0\), so that they do not depend on spatial coordinates. Substituting (48) into (44), we finally obtain

$$\begin{aligned} H'''= & {} \left( 1 - \frac{1}{4} \ddot{h}_{ij} x^{i} x^{j} \right) m - eA_{0}\nonumber \\&+ \frac{1}{3} \left( h_{ki,l} - h_{kl,i} \right) x^{k} x^{l} \Pi _{i} \nonumber \\&+ \frac{1}{2m} \left[ \delta _{ij} \left( 1 - \frac{1}{4} \ddot{h}_{kl} x^{k} x^{l} \right) \nonumber \right. \\&\left. + \frac{1}{6} \left( h_{jl,ki} + h_{ki,jl} - h_{kl,ij} - h_{ij,kl} \right) x^k x^l \right] \Pi _{i} \Pi _{j} \nonumber \\&- \frac{e}{2m} \sigma ^{i} B^{j} \left[ \delta _{ij} \left( 1 - \frac{1}{3} \ddot{h}_{kl} x^{k} x^{l} \right) \nonumber \right. \\&\left. - \frac{1}{12} \left( h_{il,kj} + h_{kj,il} - h_{kl,ij} - h_{ij,kl} \right) x^{k} x^{l} \right] \nonumber \\&+ \frac{1}{8m} \epsilon _{ijk} \sigma ^{k} \left( h_{im,jl} - h_{il,jm} - \delta _{jm} \ddot{h}_{il} \right) x^{l} \Pi _{m} \nonumber \\&- \frac{i}{6m} \ddot{h}_{ij} x^{i} \Pi _{j} \ , \end{aligned}$$
(49)

where we have used the equation of motion for gravitational waves, i.e., \(\Box h_{ij} = 0\).

In the next section, we will see that gravitational waves excite magnons, which are collective excitation of spins through the interaction in the third line in Eq. (49).

4 Magnon gravitational wave detectors

In Sect. 3, we revealed gravitational effects on a non-relativistic Dirac fermion in Fermi normal coordinates. As you can see in Eq. (49), if one consider a freely falling point particle and set Fermi normal coordinates, the particle does not feel perturbative gravity \(h_{ij}\) at the origin because of the equivalence principle. However, gravitational effects are canceled, of course, only at one point and thus an object with finite dimension feels gravitation. In the case of magnons, we prepare, for example, a ferromagnetic sample in an external magnetic field and then the sample feels gravity since it has finite size. Thus, magnons can be excited by gravitational waves. To examine the effect of gravitational waves on magnons, it is appropriate to set a Fermi normal coordinate with the origin placed at the center of the ferromagnetic sample. Then we can use the discussion of Sect. 3.

We consider a ferromagnetic sample in an external magnetic field. Such a system is described by the Heisenberg model [31]:

$$\begin{aligned} H_{\mathrm{spin}} = - 2\mu _{B} B_{z} \sum _{i} {\hat{S}}^{z}_{(i)} - \sum _{i,j} J_{ij} \hat{\varvec{S}}_{(i)} \cdot \hat{\varvec{S}}_{(j)} \ , \end{aligned}$$
(50)

where the Bohr magneton \(\mu _{B} = e/2m_{e}\) is defined by the elementary charge e and the mass of electrons \(m_{e}\). We applied an external magnetic field along the z-direction, \(B_{z}\), without loss of generality because of isotropy. Here, i specifies each site of spins. The first term is the conventional Pauli term, which turns the spin direction to be along the external magnetic field. The second term represents the exchange interactions between spins with the strength \(J_{ij}\).

Next, we take into account the effect of gravitational waves on the system. From Eq. (49), the interaction Hamiltonian between gravitational waves and a spin in the ferromagnetic sample is

$$\begin{aligned} H_{\mathrm{GW}} = - \mu _{B} B_{a} {\hat{S}}^{b}_{(i)} Q_{ab} \ , \end{aligned}$$
(51)

where we have defined

$$\begin{aligned} Q_{ij}= & {} - \frac{2}{3} \delta _{ij} \ddot{h}_{kl}|_{\varvec{x}=0} \, x^{k} x^{l} \nonumber \\&- \frac{1}{6} \left( h_{il,kj} + h_{kj,il} - h_{kl,ij} - h_{ij,kl} \right) |_{\varvec{x}=0} \, x^{k} x^{l} \ . \end{aligned}$$
(52)

It represents the effect of gravitational waves on a spin located at \(x^{i}\) in Fermi normal coordinates. Indeed, at the origin, \(x^{i}=0\), we see that \(Q_{ij}=0\). From Eqs. (50) and (51), the total Hamiltonian of the system is

$$\begin{aligned} H_{\mathrm{tot}}= & {} H_{\mathrm{spin}} + H_{\mathrm{GW}} \nonumber \\= & {} - \mu _{B} \left( 2 \delta _{za} + Q_{za} \right) B_{z} \sum _{i} {\hat{S}}^{a}_{(i)} - \sum _{i,j} J_{ij} \hat{\varvec{S}}_{(i)} \cdot \hat{\varvec{S}}_{(j)} \ . \nonumber \\ \end{aligned}$$
(53)

The spin system (53) can be rewritten by using the Holstein–Primakoff transformation [32]:

$$\begin{aligned} {\left\{ \begin{array}{ll} {\hat{S}}_{(i)}^{z} = \frac{1}{2} - {\hat{C}}_{i}^{\dagger } {\hat{C}}_{i} \ , &{} \\ {\hat{S}}_{(i)}^{+} = \sqrt{1- {\hat{C}}_{i}^{\dagger } {\hat{C}}_{i}} \ {\hat{C}}_{i} \ , &{} \\ {\hat{S}}_{(i)}^{-} = {\hat{C}}_{i}^{\dagger } \sqrt{1- {\hat{C}}_{i}^{\dagger } {\hat{C}}_{i}} \ , &{} \end{array}\right. } \end{aligned}$$
(54)

where bosonic operators \({\hat{C}}_{i}\) and \({\hat{C}}^{\dagger }_{i}\) satisfy commutation relations \([{\hat{C}}_{i}, {\hat{C}}^{\dagger }_{j}]=\delta _{ij}\) and \({\hat{S}}_{(j)}^{\pm } = {\hat{S}}_{(j)}^{x} \pm i {\hat{S}}_{(j)}^{y}\) are the ladder operators. It is easy to check that the SU(2) algebra, \([{\hat{S}}^{i} , {\hat{S}}^{j}] = i \epsilon _{ijk} {\hat{S}}^{k}\) (\(i,j,k = x, y, z\)), is satisfied even after the transformation (54). We note that \({\hat{C}}_{i}^{\dagger } {\hat{C}}_{i}\) represents the particle numbers of the boson created by the creation operator \({\hat{C}}_{i}^{\dagger }\). The bosonic operators describe spin waves with dispersion relations determined by \(B_{z}\) and \(J_{ij}\). Furthermore, provided that contributions from the surface of the sample are negligible, one can expand the bosonic operators by plane waves as

$$\begin{aligned} {\hat{C}}_{i} = \sum _{\varvec{k}} \frac{e^{-i\varvec{k} \cdot \varvec{r}_{i}}}{\sqrt{N}} {\hat{c}}_{k} \ , \end{aligned}$$
(55)

where \(\varvec{r}_{i}\) is the position vector of the i spin. The excitation of the spin waves created by \({\hat{c}}_{k}^{\dagger }\) is called a magnon.

We now rewrite the spin system (53) by magnons with the Holstein–Primakoff transformation (54) and then we only focus on the homogeneous mode of magnons, i.e., \(k = 0\) mode. Then the second term in the total Hamiltonian (53) is irrelevant because it does not contribute to the homogeneous mode. Furthermore, because \(Q_{zz}\) does not contribute to the resonance of the spins, namely excitation of magnons, we will drop it. Thus we have

$$\begin{aligned} H_{\mathrm{tot}}= & {} \mu _{B} B_{z} \sum _{i} \left[ 2 {\hat{C}}_{i}^{\dagger } {\hat{C}}_{i} + \frac{{\hat{C}}_{i} + {\hat{C}}_{i}^{\dagger }}{2} Q_{zx} \nonumber \right. \\&\left. + \frac{{\hat{C}}_{i} - {\hat{C}}_{i}^{\dagger }}{2i} Q_{zy} \right] \ . \end{aligned}$$
(56)

Now let us consider a planar gravitational wave propagating in the zx plane, namely, the wave number vector of the gravitational wave \(\varvec{k}\) has a direction \({\hat{k}} = ( \sin \theta , 0 , \cos \theta )\). Moreover, we postulate that the wavelength of the gravitational wave is much longer than the dimension of the sample and it is necessary for the validity of the Fermi normal coordinates. This situation is actually satisfied in the case of usual cavity experiments for magnons. We can expand the gravitational wave \(h_{ij}\) in terms of linear polarization tensors satisfying \(e^{(\sigma )}_{ij} e^{(\sigma ')}_{ij} = \delta _{\sigma \sigma '}\) as

$$\begin{aligned} h_{ij}(\varvec{x}, t) = h^{(+)}(\varvec{x}, t) e^{(+)}_{ij} + h^{(\times )}(\varvec{x}, t) e^{(\times )}_{ij} \ . \end{aligned}$$
(57)

More explicitly, we took the representation

figure a

where \(\omega _{h}\) is an angular frequency of the gravitational wave and \(\alpha \) represents a difference of the phases of polarizations. Note that the polarization tensors can be explicitly constructed as

$$\begin{aligned} e _{ij}^{(+)}= & {} \frac{1}{\sqrt{2}}\left( \begin{array}{ccc} \cos \theta ^{2} &{} 0 &{} -\cos \theta \sin \theta \\ 0 &{} -1 &{} 0 \\ -\cos \theta \sin \theta &{} 0 &{} \sin \theta ^{2} \end{array} \right) , \end{aligned}$$
(60)
$$\begin{aligned} e _{ij}^{(\times )}= & {} \frac{1}{\sqrt{2}}\left( \begin{array}{ccc} 0 &{} \cos \theta &{} 0 \\ \cos \theta &{} 0 &{} -\sin \theta \\ 0 &{} -\sin \theta &{} 0 \end{array} \right) . \end{aligned}$$
(61)

In Eqs. (60) and (61), we defined the \(+\) mode as a deformation in the y-direction.

Then substituting Eqs. (57)–(61) into the total Hamiltonian (56), moving on to the Fourier space and using the rotating wave approximation, one can deduce

$$\begin{aligned} H_{\mathrm{tot}} \simeq 2\mu _{B} B_{z} {\hat{c}}^{\dagger } {\hat{c}} + g_{eff} \left( {\hat{c}}^{\dagger } e^{-i\omega _{h}t} + {\hat{c}} e^{i\omega _{h}t} \right) , \end{aligned}$$
(62)

where \({\hat{c}} = {\hat{c}}_{k=0}\) and

$$\begin{aligned} g_{eff}= & {} \frac{\sqrt{2}\pi ^{2}}{60} \left( \frac{l}{\lambda } \right) ^{2} \mu _{B} B_{z} \sin \theta \sqrt{N} \left[ \cos ^{2} \theta \, (h^{(+)})^{2} \nonumber \right. \\&\left. + (h^{(\times )})^{2} + 2 \cos \theta \sin \alpha \, h^{(+)}h^{(\times )} \right] ^{1/2} \ , \end{aligned}$$
(63)

is an effective coupling constant between the gravitational waves and the magnons. The parameters l and \(\lambda = 2\pi / \omega _{h}\) are the radius of the (spherical) ferromagnetic sample and the wavelength of the gravitational wave. We note that the sum over the spin sites i was evaluated as

$$\begin{aligned} \sum _{i} xx= & {} \sum _{i} yy = \sum _{i} zz \simeq \frac{1}{L^{3}} \iiint ^{l}_{0} r^{2} \nonumber \\&\sin \zeta \left( r \cos \zeta \right) ^{2} \mathrm{{d}}r \mathrm{{d}}\zeta \mathrm{{d}}\phi = \frac{4\pi }{15} \frac{l^{5}}{L^{3}} \ , \end{aligned}$$
(64)

where L is a lattice constant, which is related to the number of spins as \(N = \left( \frac{4\pi }{3}l^{3} \right) / L^{3}\).

From Eq. (63), we see that the effective coupling constant has gotten a huge factor \(\sqrt{N}\). Moreover, in order to obtain a coordinate-independent expression of \(g_{eff}\), it is useful to use the Stokes parameters:

$$\begin{aligned} g_\mathrm{{eff}}= & {} \frac{\sqrt{2}\pi ^{2}}{60} \left( \frac{l}{\lambda } \right) ^{2} \mu _{B} B_{z} \sin \theta \sqrt{N} \nonumber \\&\times \left[ \frac{1+\cos ^{2}\theta }{2} \, I - \frac{\sin ^{2}\theta }{2} \, Q + \cos \theta \, V \right] ^{1/2} \ , \end{aligned}$$
(65)

where the Stokes parameters are defined by

$$\begin{aligned} {\left\{ \begin{array}{ll} I = (h^{(+)})^{2} + (h^{(\times )})^{2} \ , &{} \\ Q = (h^{(+)})^{2} - (h^{(\times )})^{2} \ , &{} \\ U = 2 \cos \alpha \, h^{(+)} h^{(\times )} \ , &{} \\ V = 2 \sin \alpha \, h^{(+)} h^{(\times )} \ . &{} \end{array}\right. } \end{aligned}$$
(66)

They satisfy \(I^{2} = U^{2} + Q^{2} +V^{2}\). We see that the effective coupling constant depends on the polarizations. Note that the stokes parameters Q and U transform as

$$\begin{aligned} \left( {\begin{array}{c}Q'\\ U'\end{array}}\right) = \begin{pmatrix} \cos 4\psi &{} \sin 4\psi \\ -\sin 4\psi &{} \cos 4\psi \end{pmatrix} \left( {\begin{array}{c}Q\\ U\end{array}}\right) \end{aligned}$$
(67)

where \(\psi \) is the rotation angle around \(\varvec{k}\).

The second term in Eq. (62) shows that planar gravitational waves induce the resonant spin precessions and/or the excitation of magnons if the angular frequency of the gravitational waves is near the Lamor frequency, \(2\mu _{B} B_{z}\). It is worth noting that the situation is similar to the resonant bar experiments [16] where planar gravitational waves excite phonons in a bar detector.

Let us show the ability of magnon gravitational detectors by giving constraints on high frequency gravitational waves. Recently, measurements of resonance fluorescence of magnons induced by the axion dark matter was conducted and upper bounds on an axion–electron coupling constant have been obtained [25, 26]. Such an axion–magnon resonance [33] has a similar mechanism to our graviton–magnon resonance. Therefore, we can utilize these experimental results to give the upper bounds on the amplitude of GHz gravitational waves [24].

The interaction hamiltonian which describe the axion–magnon resonance is given by

$$\begin{aligned} {\mathcal {H}}_{a} = {\tilde{g}}_\mathrm{{eff}} \left( {\hat{c}}^{\dagger } e^{-im_{a}t} + {\hat{c}} e^{im_{a}t} \right) \ , \end{aligned}$$
(68)

where \({\tilde{g}}_\mathrm{{eff}}\) is an effective coupling constant between an axion and a magnon. Notice that the axion oscillates with a frequency determined by the axion mass \(m_{a}\). One can see that this form is the same as the interaction term in Eq. (62). Through the hamiltonian (68), \({\tilde{g}}_\mathrm{{eff}}\) is related to an axion–electron coupling constant in [25, 26]. Then the axion–electron coupling constant can be converted to \({\tilde{g}}_\mathrm{{eff}}\) by using parameters, such as the energy density of the axion dark matter, which are explicitly given in [25, 26]. Therefore constraints on \({\tilde{g}}_{eff}\) (95% CL) can be read from the constraints on the axion–electron coupling constant given in [25, 26], respectively, as follows:

$$\begin{aligned} {\tilde{g}}_\mathrm{{eff}} < {\left\{ \begin{array}{ll} 3.5 \times 10^{-12} \ \mathrm{eV} \ , &{} \\ 3.1 \times 10^{-11} \ \mathrm{eV} \ . &{} \end{array}\right. } \end{aligned}$$
(69)

It is easy to convert the above constraints to those on the amplitude of gravitational waves appearing in the effective coupling constant (65). Indeed, we can read off the external magnetic field \(B_{z}\) and the number of electrons N as \((B_{z},N) = (0.5 \,\mathrm{T}, \ 5.6\times 10^{19} )\) from [25] and \((B_{z},N) = (0.3 \,\mathrm{T}, \ 9.2\times 10^{19} )\) from [26], respectively. The external magnetic field \(B_{z}\) determines the frequency of gravitational waves we can detect. Therefore, using Eqs. (65), (69) and the above parameters, one can put upper limits on gravitational waves at frequencies determined by \(B_{z}\). Since [25] and [26] focused on the direction of Cygnus and set the external magnetic fields to be perpendicular to it, we probe continuous gravitational waves coming from Cygnus with \(\theta = \frac{\pi }{2}\) (more precisely, \(\sin \theta =0.9\) in [26]). We also assume no linear and circular polarizations, i.e., \(Q'=U'=V=0\). Consequently, experimental data [25] and [26] enable us to constrain the characteristic amplitude of gravitational waves defined by \(h_{c} = h^{(+)} = h^{(\times )}\) as

$$\begin{aligned} h_{c} \sim {\left\{ \begin{array}{ll} 1.3 \times 10^{-13} \quad \mathrm{at} \ 14 \ \mathrm{GHz} \ , &{} \\ 1.1 \times 10^{-12} \quad \mathrm{at} \ 8.2 \ \mathrm{GHz} \ , &{} \end{array}\right. } \end{aligned}$$
(70)

at 95% CL, respectively. In terms of the spectral density defined by \(S_{h} = h_{c}^{2}/2f\) and the energy density parameter defined by \(\Omega _{GW} = 2\pi ^{2} f^{2} h_{c}^{2} / 3 H_{0}^{2}\) (\(H_{0}\) is the Hubble parameter), the upper limits at 95% CL are

$$\begin{aligned} \sqrt{S_{h}} \sim {\left\{ \begin{array}{ll} 7.5 \times 10^{-19} \ [\mathrm{Hz}^{-1/2}] \quad \mathrm{at} \ 14 \ \mathrm{GHz} \ , &{} \\ 8.7 \times 10^{-18} \ [\mathrm{Hz}^{-1/2}] \quad \mathrm{at} \ 8.2 \ \mathrm{GHz} \ , &{} \end{array}\right. } \end{aligned}$$
(71)

and

$$\begin{aligned} h_{0}^{2} \Omega _{GW} \sim {\left\{ \begin{array}{ll} 2.1 \times 10^{29} \quad \mathrm{at} \ 14 \ \mathrm{GHz} \ , &{} \\ 5.5 \times 10^{30} \quad \mathrm{at} \ 8.2 \ \mathrm{GHz} \ . &{} \end{array}\right. } \end{aligned}$$
(72)

We depict the limits on the spectral density with several other gravitational wave experiments in Fig. 1.

Fig. 1
figure 1

Several experimental sensitivities and constraints on high frequency gravitational waves are depicted. The blue color represents an upper limit on stochastic gravitational waves by waveguide experiment using an interaction between electromagnetic fields and gravitational waves [20]. The green one is the upper limit on stochastic gravitational waves, obtained by the 0.75 m interferometer [19]. Our new constraints on continuous gravitational waves are plotted with a red color, which also represent the sensitivity of the magnon gravitational wave detector for stochastic gravitational waves

5 Conclusion

In order to detect high frequency gravitational waves, we developed a new detection method. Using Fermi normal coordinates and taking the non-relativistic limit, we obtained the Hamiltonian for non-relativistic fermions in Fermi normal coordinates for general curved spacetime. This Hamiltonian is applicable for any curved spacetime background as long as one can treat a curvature perturbatively. Therefore, our formalism is useful to consider gravitational effects on non-relativistic fermions, which is usual in condensed matter systems.

In Sect. 4, we focused on the interaction between a spin of a fermion and gravitational waves, expressed by the third line in Eq. (49). It turned out that gravitational waves can excite magnons. Moreover, we explicitly demonstrated how to use magnons for detecting high frequency gravitational waves and gave upper limits on the spectral density of continuous gravitational waves (95 % CL): \(7.5 \times 10^{-19} \ [\mathrm{Hz}^{-1/2}]\) at 14 GHz and \(8.7 \times 10^{-18} \ [\mathrm{Hz}^{-1/2}]\) at 8.2 GHz, respectively, by utilizing results of magnon experiments. Interestingly, there are several theoretical models predicting high frequency gravitational waves which are within the scope of our method [1].

The graviton–magnon resonance is also useful for probing stochastic gravitational waves with almost the same sensitivity illustrated in Fig. 1. Although the current sensitivity is still not sufficient for putting a meaningful constraint on stochastic gravitational waves, it is important to pursue the high frequency stochastic gravitational wave search for future gravitational wave physics. Moreover, we can probe a burst of gravitational waves of any wave form if the duration time is smaller than the relaxation time of a system. The situation is the same as for resonant bar detectors [34, 35]. For instance, in the measurements [25, 26], the relaxation time is about 0.1 \(\upmu \)s which is determined by the line width of the ferromagnetic sample and the cavity. If the duration of the burst of gravitational waves is smaller than 0.1 \(\upmu \)s, we can detect it. Furthermore, improving the line width of the sample and the cavity not only leads to detecting a burst of gravitational waves but also to increasing the sensitivity. As another way to improve the sensitivity of the magnon gravitational wave detector, quantum nondemolition measurement may be promising [36,37,38].