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

1/ff noise of a tiny tunnel magnetoresistance sensor originated from a wide distribution of bath correlation time

Hiroshi Imamura [email protected]    Hiroko Arai [email protected]    Rie Matsumoto    Toshiki Yamaji National Institute of Advanced Industrial Science and Technology (AIST), Research Center for Emerging Computing Technologies (RCECT), Tsukuba, Ibaraki 305-8568, Japan
Abstract

Tunnel magnetoresistance (TMR) sensor is a highly sensitive magnetic field sensor and is expected to be applied in various fields, such as magnetic recording, industrial sensing, and bio-medical sensing. To improve the detection capability of TMR sensors in low frequency regime it is necessary to suppress the 1/ff noise. We theoretically study 1/ff noise of a tiny TMR sensor using the macrospin model. Starting from the generalized Langevin equation, 1/ff noise power spectrum and the Hooge parameter are derived. The calculated Hooge parameter of a tiny TMR sensor is much smaller than that of a conventional TMR sensor with large junction area. The results provide a new perspective on magnetic 1/ff noise and will be useful for improvement of TMR sensors.

I INTRODUCTION

Tunnel magnetoresistance (TMR) sensor [1, 2, 3, 4, 5, 6, 7] is a highly sensitive magnetic field sensor where the magnetic field signal is converted to the change in resistance of a magnetic tunnel junction (MTJ) [8, 9, 10, 11, 12]. The most popular application of the TMR sensor is a reading head of hard disk drives. Because of its high sensitivity, small size, and low power consumption, the TMR sensors are expanding their applications into a variety of fields such as industrial sensing and bio-medical sensing. In the bio-medical applications such as magetocardiography and magnetoencephalography, TMR sensors detect the weak magnetic fields generated in the human heart and brain by electrophysiological activity of cardiac muscle and nerve cells [4, 5, 6]. The frequency range of the bio-magnetic signal is less than a few hundred Hz where the 1/ff noise is the dominant noise. Reduction of 1/ff noise is a key issue for bio-medical applications [13, 3].

1/ff noise is a ubiquitous low-frequency noise whose noise power is inversely proportional to the frequency, ff [14, 15, 16, 17]. A large number of theories have been developed to explain the mechanism of 1/ff noise as reviewed in Ref. [17]. An obvious way to obtain a 1/ff power spectrum is to superimpose a large number of Lorentzian power spectra produced by exponential relaxation processes [18, 19, 20, 14, 17]. The magnitude of the 1/ff noise in different devices and materials is characterized by the Hooge parameter [15].

The magnetic 1/ff noise derived from thermal fluctuation of magnetization in a TMR sensor has been studied by several groups [21, 22, 23, 24, 25, 2, 26, 27, 7, 28]. In most previous studies the TMR sensors exhibit clear hysteresis in the magnetic field dependence of resistance, and the 1/ff noise is observed within the hysteresis loop. The observed 1/ff noise has been attributed to thermally excited hopping of magnetic domain walls between pinning sites. It is natural to ask the question if the magnetic 1/ff noise appears in a tiny TMR sensor where the domain wall cannot be created. If 1/ff noise appears in a tiny TMR sensor, what is its power? To answer this question it is necessary to develop a theoretical model of magnetic 1/ff noise based on the macrospin model.

In this paper, we propose a theoretical model for the magnetic 1/ff noise of a tiny TMR sensor based on the macrospin model. Starting from the generalized Langevin equation, we derive an analytical expression of the voltage power spectrum in the low frequency regime. Assuming a wide distribution of bath correlation times, the derived voltage power spectrum is inversely proportional to the frequency, i.e. 1/ff noise. We also show that the Hooge parameter of a tiny TMR sensor is much smaller than that of a conventional TMR sensors with large junction area.

II THEORETICAL MODEL

The system we consider is the MTJ nano-pillar shown in Fig. 1 (a), which is the core element of a tiny TMR sensor. The nonmagnetic insulating layer is sandwiched by the ferromagnetic layers. The top ferromagnetic layer is the free layer (FL) of which magnetization is softly pinned by the orange peel coupling field, 𝑯p\bm{H}_{p}, directing in the zz direction and by the uniaxial anisotropy field, HkH_{k}, along the zz axis. The direction of the magnetization in the FL is denoted by 𝒎\bm{m}. To tune the sensitivity, the bias field, 𝑯b\bm{H}_{b}, is applied in the yy direction. The bottom ferromagnetic layer is the reference layer of which magnetization unit vector, 𝒑\bm{p}, is fixed to the negative zz direction [7]. The size of the TMR sensor is assumed to be so small that a domain wall cannot be created in the FL, i.e. about or less than 10 nm.

Refer to caption

Figure 1: (a) Schematic illustration of a magnetic tunnel junction. The insulating layer is sandwiched by the free layer (FL) and the reference layer (RL). The direction of the magnetization in the FL is denoted by the magnetization unit vector, 𝒎\bm{m}. The direction of the magnetization in the RL is denoted by the magnetization unit vector, 𝒑\bm{p}, and is fixed in the negative zz direction. The magnetization in the FL is pinned by the orange peel coupling field, 𝑯p\bm{H}_{p}, directing in the zz direction and by the uniaxial anisotropy field along the zz axis.

The bias field, 𝑯b\bm{H}_{b}, is applied in the yy direction. (b) Definition of the rotated coordinate system. The zz^{\prime} axis is aligned to the equilibrium direction of the magnetization in the FL, 𝒎eq\bm{m}_{\mathrm{eq}}, by rotating around the xx axis with the angle θeq\theta_{\mathrm{eq}}.

Assuming that the FL is a thin circular disk, the magnetic free energy density of the FL is given by

E\displaystyle E =μ0Ms𝒎(𝑯p+𝑯b)+12μ0Ms2mx2\displaystyle=-\mu_{0}M_{s}\bm{m}\cdot\left(\bm{H}_{p}+\bm{H}_{b}\right)+\frac{1}{2}\mu_{0}M_{s}^{2}m_{x}^{2}
12μ0MsHkmz2,\displaystyle\hskip 10.00002pt-\frac{1}{2}\mu_{0}M_{s}H_{k}m_{z}^{2}, (1)

where μ0\mu_{0} is the permeability of vacuum, and MsM_{s} is the saturation magnetization. The equilibrium direction, 𝒎eq=(0,sinθeq,cosθeq)\bm{m}_{\mathrm{eq}}=(0,\sin\theta_{\mathrm{eq}},\cos\theta_{\mathrm{eq}}), is obtained by minimizing EE.

The voltage noise of the TMR sensor is induced by the resistance variation due to fluctuation of 𝒎\bm{m} around the equilibrium direction. To calculate the fluctuation of 𝒎\bm{m} we introduce the rotated coordinate system shown in Fig. 1(b), where the yy^{\prime} and zz^{\prime} axes are generated by rotating the yy and zz axes around the xx axis by the angle θeq\theta_{\mathrm{eq}}. The basis vectors of the xyzx-y^{\prime}-z^{\prime} coordinate system are defined as

(𝒆x𝒆y𝒆z)=(1000cosθeqsinθeq0sinθeqcosθeq)(𝒆x𝒆y𝒆z).\displaystyle\begin{pmatrix}\bm{e}_{x}\\ \bm{e}_{y^{\prime}}\\ \bm{e}_{z^{\prime}}\end{pmatrix}=\begin{pmatrix}1&0&0\\ 0&\cos\theta_{\mathrm{eq}}&-\sin\theta_{\mathrm{eq}}\\ 0&\sin\theta_{\mathrm{eq}}&\cos\theta_{\mathrm{eq}}\end{pmatrix}\begin{pmatrix}\bm{e}_{x}\\ \bm{e}_{y}\\ \bm{e}_{z}\end{pmatrix}. (2)

In the rotated coordinate system, the magnetization unit vector in the FL is represented as

𝒎\displaystyle\bm{m} =mx𝒆x+my𝒆y+mz𝒆z.\displaystyle=m_{x}\bm{e}_{x}+m_{y^{\prime}}\bm{e}_{y^{\prime}}+m_{z^{\prime}}\bm{e}_{z^{\prime}}. (3)

Since we are interested in the small fluctuation of 𝒎\bm{m} around 𝒎eq\bm{m}_{\mathrm{eq}}, we assume |mx|1|m_{x}|\ll 1, |my|1|m_{y^{\prime}}|\ll 1, and |mz|1|m_{z^{\prime}}|\simeq 1.

The resistance of the TMR sensor is given by [8, 9, 29]

R=R0+R¯1+P2𝒎𝒑,\displaystyle R=R_{0}+\frac{\bar{R}}{1+P^{2}\bm{m}\cdot\bm{p}}, (4)

where R0R_{0} is the resistance not caused by tunneling, R¯\bar{R} is the resistance due to tunneling at 𝒎𝒑=0\bm{m}\cdot\bm{p}=0, PP is the spin polarization of tunneling electrons. Substituting Eq. (3) into Eq. (4), the resistance is obtained as

R=R0+R¯1P2(cosθeqmzsinθeqmy).\displaystyle R=R_{0}+\frac{\bar{R}}{1-P^{2}\left(\cos\theta_{\mathrm{eq}}m_{z^{\prime}}-\sin\theta_{\mathrm{eq}}m_{y^{\prime}}\right)}. (5)

Up to the first order of mym_{y^{\prime}}, the resistance can be approximated as

R=R0+R¯1P2cosθeqR¯P2sinθeq(1P2cosθeq)2my.\displaystyle R=R_{0}+\frac{\bar{R}}{1-P^{2}\cos\theta_{\mathrm{eq}}}-\frac{\bar{R}P^{2}\sin\theta_{\mathrm{eq}}}{\left(1-P^{2}\cos\theta_{\mathrm{eq}}\right)^{2}}m_{y^{\prime}}. (6)

III Results

In this section, we show the results of our theoretical analysis on magnetic 1/ff noise of a tiny TMR sensor. We first show the relation between the voltage power spectrum and the power spectrum of mym_{y^{\prime}} in Sec. III.1. To calculate the power spectrum of mym_{y^{\prime}}, we solve the linearized equations of motion of 𝒎\bm{m} by using the Fourier transformation in Sec. III.2. Then we derive the Lorentzian power spectrum of mym_{y^{\prime}} in Sec. III.3. Assuming that bath correlation time, τc\tau_{c}, has a wide distribution, we derive the 1/ff power spectrum of voltage by superimposing the Lorentzian power spectra with different τc\tau_{c} in Sec. III.4. In Sec. III.5, we show that the Hooge parameter of a tiny TMR sensor is much smaller than the conventional TMR sensor with the same sensitivity by comparing with the experimental results of Ref. [7].

III.1 Power spectrum of voltage

In most experiments, the voltage noise of a TMR sensor is measured under a constant direct current, II. Assuming that the measured voltage, VV, is proportional to the resistance, RR, the power spectrum of voltag, SVV(f)S_{VV}(f), is proportional to the power spectrum of resistance, SRR(f)S_{RR}(f), as

SVV(f)=I2SRR(f).\displaystyle S_{VV}(f)=I^{2}S_{RR}(f). (7)

Introducing the angular frequency, ω=2πf\omega=2\pi f, the power spectrum of resistance is defined as

SRR(ω)=40R(t)R(0)cos(ωt)𝑑t,\displaystyle S_{RR}(\omega)=4\int_{0}^{\infty}\left\langle R(t)R(0)\right\rangle\cos(\omega t)dt, (8)

where \langle\ \rangle represents the statistical average. Substituting Eq. (6) into Eq. (8), SRR(ω)S_{RR}(\omega) is expressed as

SRR(ω)=[R¯P2sinθeq(1P2cosθeq)2]2Smymy(ω),\displaystyle S_{RR}(\omega)=\left[\frac{\bar{R}P^{2}\sin\theta_{\mathrm{eq}}}{\left(1-P^{2}\cos\theta_{\mathrm{eq}}\right)^{2}}\right]^{2}S_{m_{y^{\prime}}m_{y^{\prime}}}(\omega), (9)

where Smymy(ω)S_{m_{y^{\prime}}m_{y^{\prime}}}(\omega) is the power spectrum of mym_{y^{\prime}} defined as

Smymy(ω)=40my(t)my(0)cos(ωt)𝑑t.\displaystyle S_{m_{y^{\prime}}m_{y^{\prime}}}(\omega)=4\int_{0}^{\infty}\left\langle m_{y^{\prime}}(t)m_{y^{\prime}}(0)\right\rangle\cos(\omega t)dt. (10)

We define the Fourier transform of a function f(t)f(t) as f(ω)=f(t)exp(iωt)𝑑tf(\omega)=\int_{-\infty}^{\infty}f(t)\exp\left(-i\omega t\right)dt. Substituting the inverse Fourier transform of mym_{y^{\prime}} into Eq. (10) and performing some algebra, we obtain

Smymy(ω)\displaystyle S_{m_{y^{\prime}}m_{y^{\prime}}}(\omega) =12πmy(ω)my(ω)𝑑ω\displaystyle=\frac{1}{2\pi}\int_{-\infty}^{\infty}\langle m_{y^{\prime}}(\omega)m_{y^{\prime}}(\omega^{\prime})\rangle d\omega^{\prime}
+12πmy(ω)my(ω)𝑑ω.\displaystyle+\frac{1}{2\pi}\int_{-\infty}^{\infty}\langle m_{y^{\prime}}(-\omega)m_{y^{\prime}}(\omega^{\prime})\rangle d\omega^{\prime}. (11)

The Fourier transform of mym_{y^{\prime}} can be obtained by solving the equations of motion in the Fourier space.

III.2 Equations of motion and the Fourier transforms of mxm_{x} and mym_{y^{\prime}}

The equations of motion of 𝒎\bm{m} is given by the following generalized Langevin equation [30, 31, 32],

𝒎˙(t)\displaystyle\dot{\bm{m}}(t) =γ𝒎(t)×(𝑯eff+𝒓)\displaystyle=-\gamma\bm{m}(t)\times\left(\bm{H}_{\mathrm{eff}}+\bm{r}\right)
+α𝒎×tν(tt)𝒎˙(t)𝑑t,\displaystyle+\alpha\bm{m}\times\int_{-\infty}^{t}\nu(t-t^{\prime})\dot{\bm{m}}(t^{\prime})dt^{\prime}, (12)

where 𝒎˙(t)\dot{\bm{m}}(t) is the time derivative of 𝒎(t)\bm{m}(t), γ\gamma is the gyromagnetic ratio, and α\alpha is the Gilbert damping constant. The effective magnetic field acting on 𝒎\bm{m} is given by

𝑯eff=Msmx𝒆x+Hb𝒆y+(Hp+Hkmz)𝒆z.\displaystyle\bm{H}_{\mathrm{eff}}=-M_{s}m_{x}\bm{e}_{x}+H_{b}\bm{e}_{y}+\left(H_{p}+H_{k}m_{z}\right)\bm{e}_{z}. (13)

The memory function is defined as

ν(tt)=1τcexp(|tt|τc),\displaystyle\nu(t-t^{\prime})=\frac{1}{\tau_{c}}\exp\left(-\frac{\left|t-t^{\prime}\right|}{\tau_{c}}\right), (14)

where τc\tau_{c} is the bath correlation time. The thermal agitation field, 𝒓\bm{r}, is a random field satisfying rj=0\langle r_{j}\rangle=0 and

rjrk=μ2δj,kν(tt),\displaystyle\langle r_{j}r_{k}\rangle=\frac{\mu}{2}\delta_{j,k}\nu(t-t^{\prime}), (15)

where subscripts jj and kk denotes xx, yy, zz, yy^{\prime}, or zz^{\prime}. The constant μ\mu is defined as

μ=2αkBTγμ0MsΩ,\displaystyle\mu=\frac{2\alpha k_{B}T}{\gamma\mu_{0}M_{s}\Omega}, (16)

where kBk_{B} is the Boltzmann constant, TT is temperature, and Ω\Omega is the volume of the FL. From Eqs. (15) and (16) we see that the magnitude of the thermal agitation field is of the order of α\sqrt{\alpha} because μ\mu is of the order of α\alpha. The stochastic LLG equation with the Markovian damping derived by Brown [33] is reproduced in the limit of τc0\tau_{c}\to 0 because limτc0ν(tt)=2δ(tt)\lim_{\tau_{c}\to 0}\nu(t-t^{\prime})=2\delta(t-t^{\prime}), where δ(tt)\delta(t-t^{\prime}) is Dirac’s delta function. It should be noted that 1/ff noise cannot be derived from the LLG equation with the Markovian damping because many physical processes with different time scale is required to generate 1/ff noise.

Since the FL of a typical TMR sensor is made of a ferromagnetic material with α1\alpha\ll 1, we focus on terms up to the first order of α\alpha in the equations of motion. We also assume that mxm_{x}, mym_{y^{\prime}}, rxr_{x}, ryr_{y^{\prime}}, and rzr_{z^{\prime}} are small enough to linearize the equations of motion in terms of these small variables. Equation (III.2) can be approximated as

m˙x(t)=ω0my(t)+γry(t)\displaystyle\dot{m}_{x}(t)=-\omega_{0}m_{y^{\prime}}(t)+\gamma r_{y^{\prime}}(t)
αtν(tt)m˙y(t)𝑑t\displaystyle\hskip 40.00006pt-\alpha\int_{-\infty}^{t}\nu(t-t^{\prime})\dot{m}_{y^{\prime}}(t^{\prime})dt^{\prime} (17)
m˙y(t)=ω1mx(t)γrx(t)\displaystyle\dot{m}_{y^{\prime}}(t)=\omega_{1}m_{x}(t)-\gamma r_{x}(t)
+αtν(tt)m˙x(t)𝑑t\displaystyle\hskip 40.00006pt+\alpha\int_{-\infty}^{t}\nu(t-t^{\prime})\dot{m}_{x}(t^{\prime})dt^{\prime} (18)
m˙z(t)=0,\displaystyle\dot{m}_{z^{\prime}}(t)=0, (19)

where

ω0\displaystyle\omega_{0} =γ(Hbsinθeq+Hpcosθeq+Hkcos2θeq),\displaystyle=\gamma\left(H_{b}\sin\theta_{\rm eq}+H_{p}\cos\theta_{\rm eq}+H_{k}\cos 2\theta_{\rm eq}\right), (20)
ω1\displaystyle\omega_{1} =γ(Ms+Hbsinθeq+Hpcosθeq+Hkcos2θeq).\displaystyle=\gamma\left(M_{s}\!+\!H_{b}\sin\theta_{\rm eq}\!+\!H_{p}\cos\theta_{\rm eq}\!+\!H_{k}\cos^{2}\theta_{\rm eq}\right). (21)

Following Ref. [32], we approximate the non-Markovian damping term in Eqs. (III.2) and (III.2) up to the first order of α\alpha. Successive application of the integration by parts gives the following linearized equations of motion up to the order of α\alpha,

m˙x(t)=γ^1ω0my(t)+γry(t)α~ω1mx(t)\displaystyle\dot{m}_{x}(t)=-\hat{\gamma}_{1}\omega_{0}m_{y^{\prime}}(t)+\gamma r_{y^{\prime}}(t)-\tilde{\alpha}\omega_{1}m_{x}(t) (22)
m˙y(t)=γ^0ω1mx(t)γrx(t)α~ω0my(t),\displaystyle\dot{m}_{y^{\prime}}(t)=\hat{\gamma}_{0}\omega_{1}m_{x}(t)-\gamma r_{x}(t)-\tilde{\alpha}\omega_{0}m_{y^{\prime}}(t), (23)

where

γ^0=(1+αξ01+ξ0ξ1)\displaystyle\hat{\gamma}_{0}=\left(1+\frac{\alpha\xi_{0}}{1+\xi_{0}\xi_{1}}\right) (24)
γ^1=(1+αξ11+ξ0ξ1)\displaystyle\hat{\gamma}_{1}=\left(1+\frac{\alpha\xi_{1}}{1+\xi_{0}\xi_{1}}\right) (25)
α~=α1+ξ0ξ1\displaystyle\tilde{\alpha}=\frac{\alpha}{1+\xi_{0}\xi_{1}} (26)
ξ0=τcω0\displaystyle\xi_{0}=\tau_{c}\omega_{0}
ξ1=τcω1.\displaystyle\xi_{1}=\tau_{c}\omega_{1}. (27)

Details of the derivation of the above equations will be provided in Appendix A. In the Fourier space, the equations of motion are expressed as

iωmx(ω)=γ^1ω0my(ω)+γry(ω)α~ω1mx(ω),\displaystyle i\omega m_{x}(\omega)=-\hat{\gamma}_{1}\omega_{0}m_{y^{\prime}}(\omega)+\gamma r_{y^{\prime}}(\omega)-\tilde{\alpha}\omega_{1}m_{x}(\omega), (28)
iωmy(ω)=γ^0ω1mx(ω)γrx(ω)α~ω0my(ω).\displaystyle i\omega m_{y^{\prime}}(\omega)=\hat{\gamma}_{0}\omega_{1}m_{x}(\omega)-\gamma r_{x}(\omega)-\tilde{\alpha}\omega_{0}m_{y^{\prime}}(\omega). (29)

The solutions are obtained as

mx(ω)=γ^1ω0γrx(ω)+(α~ω0+iω)γry(ω)A(ω)\displaystyle m_{x}(\omega)=\frac{\hat{\gamma}_{1}\omega_{0}\gamma r_{x}(\omega)+\left(\tilde{\alpha}\omega_{0}+i\omega\right)\gamma r_{y^{\prime}}(\omega)}{A(\omega)} (30)
my(ω)=γ^0ω1γry(ω)(α~ω1+iω)γrx(ω)A(ω),\displaystyle m_{y^{\prime}}(\omega)=\frac{\hat{\gamma}_{0}\omega_{1}\gamma r_{y^{\prime}}(\omega)-\left(\tilde{\alpha}\omega_{1}+i\omega\right)\gamma r_{x}(\omega)}{A(\omega)}, (31)

where

A(ω)=(γ^0γ^1+α~2)ω0ω1ω2+iα~(ω0+ω1)ω.\displaystyle A(\omega)=\left(\hat{\gamma}_{0}\hat{\gamma}_{1}+\tilde{\alpha}^{2}\right)\omega_{0}\omega_{1}-\omega^{2}+i\tilde{\alpha}(\omega_{0}+\omega_{1})\omega. (32)

III.3 Power spectrum of mym_{y^{\prime}}

From Eq. (31), the correlation of my(ω)m_{y^{\prime}}(\omega) and my(ω)m_{y^{\prime}}(\omega^{\prime}) is expressed as

my(ω)my(ω)=(γ^0ω1)2A(ω)A(ω)γ2ry(ω)ry(ω)\displaystyle\langle m_{y^{\prime}}(\omega)m_{y^{\prime}}(\omega^{\prime})\rangle=\frac{\left(\hat{\gamma}_{0}\omega_{1}\right)^{2}}{A(\omega)A(\omega^{\prime})}\gamma^{2}\left\langle r_{y^{\prime}}(\omega)r_{y^{\prime}}(\omega^{\prime})\right\rangle
+(α~ω1+iω)(α~ω1+iω)A(ω)A(ω)γ2rx(ω)rx(ω),\displaystyle+\frac{\left(\tilde{\alpha}\omega_{1}+i\omega\right)\left(\tilde{\alpha}\omega_{1}+i\omega^{\prime}\right)}{A(\omega)A(\omega^{\prime})}\gamma^{2}\left\langle r_{x}(\omega)r_{x}(\omega^{\prime})\right\rangle, (33)

where we use the fact that rxr_{x} and ryr_{y^{\prime}} do not correlate with each other. The correlation my(ω)my(ω)\langle m_{y^{\prime}}(-\omega)m_{y^{\prime}}(\omega^{\prime})\rangle is obtained by replacing ω\omega with ω-\omega in Eq. (III.3)

Following Ref. [34], the correlation of thermal agitation fields in the Fourier space is obtained as

rj(ω)rk(ω)=2πμδj,k11+iτcωδ(ω+ω).\displaystyle\langle r_{j}(\omega)r_{k}(\omega^{\prime})\rangle=2\pi\mu\delta_{j,k}\frac{1}{1+i\tau_{c}\omega}\delta(\omega+\omega^{\prime}). (34)

The correlation rj(ω)rk(ω)\langle r_{j}(-\omega)r_{k}(\omega^{\prime})\rangle is obtained by replacing ω\omega with ω-\omega in Eq. (34).

Substituting Eqs. (III.3) and (34) into Eq. (III.1), the power spectrum of mym_{y^{\prime}} is expressed as

Smymy(ω)\displaystyle S_{m_{y^{\prime}}m_{y^{\prime}}}(\omega) =(γ^02+α~2)ω12+ω2B(ω)2γ2μ1+(τcω)2,\displaystyle=\frac{\left(\hat{\gamma}_{0}^{2}+\tilde{\alpha}^{2}\right)\omega_{1}^{2}+\omega^{2}}{B(\omega)}\frac{2\gamma^{2}\mu}{1+(\tau_{c}\omega)^{2}}, (35)

where

B(ω)\displaystyle B(\omega) =[(γ^0γ^1+α~2)ω0ω1ω2]2\displaystyle=\left[\left(\hat{\gamma}_{0}\hat{\gamma}_{1}+\tilde{\alpha}^{2}\right)\omega_{0}\omega_{1}-\omega^{2}\right]^{2}
+[α~(ω0+ω1)ω]2.\displaystyle\hskip 10.00002pt+\left[\tilde{\alpha}(\omega_{0}+\omega_{1})\omega\right]^{2}. (36)

In the low frequency regime satisfying ωω0\omega\ll\omega_{0} and ωω1\omega\ll\omega_{1}, Eq. (35) can be approximated by the Lorentzian function as

Smymy(ω)\displaystyle S_{m_{y^{\prime}}m_{y^{\prime}}}(\omega) =2γ2μ(γ^1ω0)211+(τcω)2.\displaystyle=\frac{2\gamma^{2}\mu}{\left(\hat{\gamma}_{1}\omega_{0}\right)^{2}}\frac{1}{1+(\tau_{c}\omega)^{2}}. (37)

Since ω0\omega_{0} and ω1\omega_{1} are of the order of 0.1 GHz \sim 10 GHz for conventional TMR sensors [7], the low frequency condition is clearly satisfied for the frequency range of the bio-magnetic signal, i.e. less than a few hundred Hz.

III.4 Superimposition of Lorentzian power spectra

Bath correlation time, τc\tau_{c}, is the decay time of the correlation of thermal agitation field as shown in Eq. (15). Thermal agitation field is produced by many kinds of sources or baths such as dipolar coupling with magnons in the reference layer and spin orbit coupling with phonons. Since τc\tau_{c} depends on the relaxation mechanism of the bath, different relaxation modes in different baths have their own τc\tau_{c}. Instead of discussing τc\tau_{c} for some specific types of baths, we just assume a distribution of τc\tau_{c} and analyze the effect of the distribution of τc\tau_{c} on the low frequency power spectrum of voltage. Assuming a wide distribution of τc\tau_{c}, we derive an analytical expression of the power spectrum of the magnetic 1/ff noise.

We assume that τc\tau_{c} is uniformly distributed in the range of τc,minτcτc,max\tau_{c,\mathrm{min}}\leq\tau_{c}\leq\tau_{c,\mathrm{max}} and has the probability distribution defined as ρ(τc)=1/(τc,maxτc,min)\rho(\tau_{c})=1/(\tau_{c,\mathrm{max}}-\tau_{c,\mathrm{min}}). The superimposition of Smymy(ω)S_{m_{y^{\prime}}m_{y^{\prime}}}(\omega) for all τc\tau_{c} is given by

Smymy(ω)\displaystyle S_{m_{y^{\prime}}m_{y^{\prime}}}(\omega) =2γ2μω0201γ^12ρ(τc)1+(τcω)2𝑑τc.\displaystyle=\frac{2\gamma^{2}\mu}{\omega_{0}^{2}}\int_{0}^{\infty}\frac{1}{\hat{\gamma}_{1}^{2}}\frac{\rho(\tau_{c})}{1+(\tau_{c}\omega)^{2}}d\tau_{c}. (38)

As a function of τc\tau_{c}, γ^1\hat{\gamma}_{1} is almost unity except around the peak at τc=1/ω0ω1\tau_{c}=1/\sqrt{\omega_{0}\omega_{1}}, which is of the order of ns. Since the peak value of γ^1\hat{\gamma}_{1} is as small as 1+(α/2)ω1/ω01+(\alpha/2)\sqrt{\omega_{1}/\omega_{0}}, and ω\omega is assumed to be much smaller than ω0\omega_{0} and ω1\omega_{1}, Eq. (38) can be approximated as

Smymy(ω)\displaystyle S_{m_{y^{\prime}}m_{y^{\prime}}}(\omega) =2γ2μω021τc,diffτc,minτc,max11+(τcω)2𝑑τc\displaystyle=\frac{2\gamma^{2}\mu}{\omega_{0}^{2}}\frac{1}{\tau_{c,\mathrm{diff}}}\int_{\tau_{c,\mathrm{min}}}^{\tau_{c,\mathrm{max}}}\frac{1}{1+(\tau_{c}\omega)^{2}}d\tau_{c}
=2γ2μω021τc,diff[arctan(ωτc,max)ω\displaystyle=\frac{2\gamma^{2}\mu}{\omega_{0}^{2}}\frac{1}{\tau_{c,\mathrm{diff}}}\left[\frac{\arctan\left(\omega\tau_{c,\mathrm{max}}\right)}{\omega}\right.
arctan(ωτc,min)ω],\displaystyle\hskip 10.00002pt\left.-\frac{\arctan\left(\omega\tau_{c,\mathrm{min}}\right)}{\omega}\right], (39)

where τc,diff=τc,maxτc,min\tau_{c,\mathrm{diff}}=\tau_{c,\mathrm{max}}-\tau_{c,\mathrm{min}}. Since arctan(x)/x\arctan(x)/x is a monotonically decreasing function of xx for x>0x>0 and limx0arctan(x)/x=1\lim_{x\to 0}\arctan(x)/x=1, Smymy(ω)S_{m_{y^{\prime}}m_{y^{\prime}}}(\omega) is a monotonically decreasing function of ω\omega and take a maximum value of 2γ2μ/ω022\gamma^{2}\mu/\omega_{0}^{2} in the limit of ω0\omega\to 0.

Refer to caption

Figure 2: Power spectrum of mym_{y^{\prime}}, Smymy(ω)S_{m_{y^{\prime}}m_{y^{\prime}}}(\omega) given by Eq. (40) normalized by 2γ2μ/ω022\gamma^{2}\mu/\omega_{0}^{2}. The black solid, red dotted, and blue dashed curves represent the results for τc,max\tau_{c,\mathrm{max}}=10, 1, and 0.1 s, respectively. The green circles indicate the values at ω\omega = 1/τc,max1/\tau_{c,\mathrm{max}}.

When ωτc,min1\omega\tau_{c,\mathrm{min}}\ll 1, the second term in the square bracket of Eq. (III.4) can be neglected and Smymy(ω)S_{m_{y^{\prime}}m_{y^{\prime}}}(\omega) is approximated as

Smymy(ω)=2γ2μω02arctan(ωτc,max)ωτc,max.\displaystyle S_{m_{y^{\prime}}m_{y^{\prime}}}(\omega)=\frac{2\gamma^{2}\mu}{\omega_{0}^{2}}\frac{\arctan\left(\omega\tau_{c,\mathrm{max}}\right)}{\omega\tau_{c,\mathrm{max}}}. (40)

Figure 2 shows Smymy(ω)S_{m_{y^{\prime}}m_{y^{\prime}}}(\omega) given by Eq. (40) normalized by 2γ2μ/ω022\gamma^{2}\mu/\omega_{0}^{2} for τc,max\tau_{c,\mathrm{max}}=10 s (black solid), 1 s (red dotted), and 0.1 s (blue dashed). The values at ω\omega = 1/τc,max1/\tau_{c,\mathrm{max}} are indicated by the green circles. All curves are almost flat for ω1/τc,max\omega\ll 1/\tau_{c,\mathrm{max}} and inversely proportional to ω\omega for ω1/τc,max\omega\gg 1/\tau_{c,\mathrm{max}}.

Assuming a wide distribution of τc\tau_{c} satisfying ωτc,min1\omega\tau_{c,\mathrm{min}}\ll 1 and ωτc,max1\omega\tau_{c,\mathrm{max}}\gg 1, we have arctan(ωτc,min)=0\arctan\left(\omega\tau_{c,\mathrm{min}}\right)=0 and arctan(ωτc,max)=π/2\arctan\left(\omega\tau_{c,\mathrm{max}}\right)=\pi/2. Then the power spectrum can be approximated as

Smymy(ω)\displaystyle S_{m_{y^{\prime}}m_{y^{\prime}}}(\omega) =2γ2μω021τc,maxπ2ω,\displaystyle=\frac{2\gamma^{2}\mu}{\omega_{0}^{2}}\frac{1}{\tau_{c,\mathrm{max}}}\frac{\pi}{2\omega}, (41)

which is inversely proportional to the angular frequency, ω\omega (=2πf2\pi f). From Eqs. (7), (9) and (41) the voltage power spectrum is given by

SVV(f)\displaystyle S_{VV}(f) =[IR¯P2sinθeq(1P2cosθeq)2]2γ2μ2ω021τc,max1f.\displaystyle=\left[\frac{I\bar{R}P^{2}\sin\theta_{\mathrm{eq}}}{\left(1-P^{2}\cos\theta_{\mathrm{eq}}\right)^{2}}\right]^{2}\frac{\gamma^{2}\mu}{2\omega_{0}^{2}}\frac{1}{\tau_{c,\mathrm{max}}}\frac{1}{f}. (42)

This is the main result of this paper. The obvious difference from other models of low frequency magnetic noise [35, 21, 36, 37, 23, 38] is that Eq. (42) has the term 1/τc,max1/\tau_{c,\mathrm{max}} as information of the distribution of the bath correlation time. It should be noted that the 1/f noise of a tiny TMR sensor we derived is response to the thermal agitation fields that exhibit 1/f power spectrum as the superimposition of the Lorentzian power spectrum.

III.5 Comparison with a conventional TMR sensor with large junction area

We compare the derived 1/ff noise of the macrospin model with the experimental results of a conventional TMR sensor with large junction area reported in Ref. [7]. The Hooge parameter, αH\alpha_{H}, is a convenient measure to compare the 1/f noise between different MTJs, which is defined as

SVV(f)=SVVwh+αHVb2A1f1,\displaystyle S_{VV}(f)=S_{VV}^{\rm wh}+\alpha_{H}V_{b}^{2}A^{-1}f^{-1}, (43)

where SVVwhS_{VV}^{\rm wh} is the power spectral density of the white noise, VbV_{b} is the bias voltage, and AA is the area of the MTJ. The typical value of the Hooge parameter of conventional TMR sensors is about 106101110^{-6}\sim 10^{-11} μ\mum2 [26, 27, 7, 28]. From Eq. (6), the bias voltage is given by

Vb=I(R0+R¯1P2cosθeq).\displaystyle V_{b}=I\left(R_{0}+\frac{\bar{R}}{1-P^{2}\cos\theta_{\mathrm{eq}}}\right). (44)

From Eqs. (16), (42), (43), and (44), the Hooge parameter of a tiny TMR sensor is obtained as

αH\displaystyle\alpha_{H} ={R¯P2sinθeq(1P2cosθeq)[R¯+R0(1P2cosθeq)]}2\displaystyle=\left\{\frac{\bar{R}P^{2}\sin\theta_{\rm eq}}{(1-P^{2}\cos\theta_{\rm eq})\left[\bar{R}+R_{0}(1-P^{2}\cos\theta_{\rm eq})\right]}\right\}^{2}
×αγkBTμ0Msd1ω021τc,max,\displaystyle\times\frac{\alpha\,\gamma\,k_{\rm B}T}{\mu_{0}\,M_{s}\,d}\frac{1}{\omega_{0}^{2}}\frac{1}{\tau_{c,\mathrm{max}}}, (45)

where dd is the thickness of the FL.

To compare Eq. (III.5) with the experimental results of a conventional TMR sensor, we determine the junction parameters by fitting the bias field dependence of the resistance shown in Figs. 2(a) of Ref. [7]. The parameters are determined as MsM_{s}=0.93 MA/m, μ0Hk\mu_{0}H_{k}=1.0 mT, μ0Hp\mu_{0}H_{p}= 2.15 mT, R0R_{0}=10.7 Ω\Omega, R¯\bar{R}=10.3 Ω\Omega, P2P^{2}=0.74. Figure 3(a) shows the bias field, μ0Hb\mu_{0}H_{b}, dependence of the sensitivity defined as

Sensitivity=1RmaxdRdμ0Hb,\displaystyle{\rm Sensitivity}=\frac{1}{R_{\rm max}}\frac{dR}{d\,\mu_{0}H_{b}}, (46)

where RmaxR_{\rm max} is the maximum value of the resistance. The experimental results indicated by the yellow curves are well reproduced by the theoretical results represented by the black curves.

Refer to caption

Figure 3: (a) Sensitivity as a function of the bias field, μ0Hb\mu_{0}H_{b}. The top panel shows the results for the signal field along the magnetization hard axis, i.e the yy axis. The bottom panel shows the results for the signal field along the magnetization easy axis, i.e zz axis. In both panels, the yellow and the black curves represent the experimental and theoretical results, respectively. (b) Hooge parameter, αH\alpha_{H}, as a function of the bias field, μ0Hb\mu_{0}H_{b}. The yellow circles and the black curve represent the experimental and theoretical results, respectively. Note that the theoretical results are multiplied by 10610^{6}. In all panels, the experimental results are the same as those shown in Fig. 3(b) in Ref. [7].

Figure 3(b) shows the bias field dependence of the Hooge parameter, αH\alpha_{H}. The experimental results are indicated by the yellow circles. The black solid curve represents the theoretical results multiplied by 10610^{6}. For calculation of the Hooge parameter, the following parameters are assumed: α\alpha = 0.05, dd = 2 nm, τc,max\tau_{c,\mathrm{max}} = 1 s, and TT = 300 K. The field dependences of the Hooge parameter and the sensitivity look very similar because the Hooge parameter is proportional to the square of the sensitivity along the yy^{\prime} axis as indicated by Eqs. (6), (42), and (46).

The results show that the Hooge parameter of a tiny TMR sensor is much smaller than that of the conventional TMR sensor with the same sensitivity by a factor of 10610^{-6}. If a number of tiny MTJs are connected in parallel to reproduce the same resistance as a conventional TMR sensor, the power spectrum of the magnetic 1/ff noise can be reduced by a factor of 10610^{-6} without reducing sensitivity.

IV Summary

In summary, we propose a theoretical model for magnetic 1/ff noise of a tiny TMR sensor originated from distribution of bath correlation time. Starting from the generalized Langevin equation, we derive an analytical expression of the low frequency power spectrum of voltage. Assuming a wide distribution of the bath correlation times, the derived voltage power spectrum is inversely proportional to the frequency. We also show that the Hooge parameter of a tiny TMR sensor is much smaller than that of a conventional TMR sensor with large junction area. The power spectrum of the 1/ff noise can be substantially reduced without reducing sensitivity by connecting tiny TMR sensors in parallel. The results provides a new perspective on magnetic 1/ff noise and will be useful for reduction of 1/ff noise of TMR sensors. The presented theoretical framework is applicable not only to the magnetic 1/ff noise of a tiny TMR sensor, but also to the low frequency fluctuation of any tiny magnetic devices where the macrospin model is appropriate.

Acknowledgements.
The authors thank T. Nakatani for providing experimental data and for valuable discussions. This work was partly supported by JSPS KAKENHI Grant No. JP23K04575.

Appendix A Derivation of Eqs. (22) and (23)

In this section, we provide the details of the derivation of Eqs. (22) and (23). Following Ref. [32], different approaches are used depending on the value of τc\tau_{c}: in the short τc\tau_{c} regime and in the long τc\tau_{c} regime. Introducing ξ0=τcω0\xi_{0}=\tau_{c}\omega_{0} and ξ1=τcω1\xi_{1}=\tau_{c}\omega_{1}, the short τc\tau_{c} regime is defined as ξ0ξ1<1\xi_{0}\xi_{1}<1, and the long τc\tau_{c} regime is defined as ξ0ξ1>1\xi_{0}\xi_{1}>1. For both τc\tau_{c} regimes we can derive the same equations of motion as Eqs. (22) and (23).

A.0.1 Short τc\tau_{c} regime: ξ0ξ1<1\xi_{0}\xi_{1}<1

We approximate the non-Markovian damping term in Eq. (III.2) up to the first order of α\alpha. Successive application of the integration by parts using ν(tt)=τc[dν(tt)/dt]\nu(t-t^{\prime})=\tau_{c}\left[d\nu(t-t^{\prime})/dt^{\prime}\right] the integral part of the non-Markovian damping term in Eq. (III.2) is expressed as

tν(tt)m˙y(t)𝑑t=n=1(τc)n1dndtnmy(t).\displaystyle\int_{-\infty}^{t}\nu(t-t^{\prime})\dot{m}_{y^{\prime}}(t^{\prime})dt^{\prime}=\sum_{n=1}^{\infty}(-\tau_{c})^{n-1}\frac{d^{n}}{dt^{n}}m_{y^{\prime}}(t). (47)

Since the integral part of the non-Markovian damping is multiplied by α\alpha we approximate the time derivative of my(t)m_{y^{\prime}}(t) in the 0th order of α\alpha as

d2ndt2nmy(t)=(1)nω0nω1nmy(t),\displaystyle\frac{d^{2n}}{dt^{2n}}m_{y^{\prime}}(t)=(-1)^{n}\omega_{0}^{n}\omega_{1}^{n}m_{y^{\prime}}(t), (48)

and

d2n+1dt2n+1my(t)=(1)nω0nω1n+1mx(t).\displaystyle\frac{d^{2n+1}}{dt^{2n+1}}m_{y^{\prime}}(t)=(-1)^{n}\omega_{0}^{n}\omega_{1}^{n+1}m_{x}(t). (49)

Substituting Eqs. (48) and (49) into Eq. (47), the integral part of the non-Markovian damping term in Eq. (III.2) is expressed as

tν(tt)m˙y(t)dt=[ξ1ω0my(t)\displaystyle\int_{-\infty}^{t}\nu(t-t^{\prime})\dot{m}_{y^{\prime}}(t^{\prime})dt^{\prime}=\Bigl{[}\xi_{1}\omega_{0}m_{y^{\prime}}(t)
+ω1mx(t)]n=1(ξ0ξ1)n1.\displaystyle+\omega_{1}m_{x}(t)\Bigr{]}\sum_{n=1}^{\infty}(-\xi_{0}\xi_{1})^{n-1}. (50)

The summation in Eq. (A.0.1) converges under the condition of ξ0ξ1<1\xi_{0}\xi_{1}<1 as

n=1(ξ0ξ1)n1=11+ξ0ξ1.\displaystyle\sum_{n=1}^{\infty}(-\xi_{0}\xi_{1})^{n-1}=\frac{1}{1+\xi_{0}\xi_{1}}. (51)

Then Eq. (A.0.1) becomes

tν(tt)m˙y(t)𝑑t=ξ1ω0my(t)+ω1mx(t)1+ξ0ξ1.\displaystyle\int_{-\infty}^{t}\nu(t-t^{\prime})\dot{m}_{y^{\prime}}(t^{\prime})dt^{\prime}=\frac{\xi_{1}\omega_{0}m_{y^{\prime}}(t)+\omega_{1}m_{x}(t)}{1+\xi_{0}\xi_{1}}. (52)

Substituting Eq. (52) into Eq. (III.2) and performing some algebra, we obtain the following linearized equation of motion for mx(t)m_{x}(t) up to the first order of α\alpha:

m˙x(t)\displaystyle\dot{m}_{x}(t) =(1+αξ11+ξ0ξ1)ω0my(t)\displaystyle=-\left(1+\frac{\alpha\xi_{1}}{1+\xi_{0}\xi_{1}}\right)\omega_{0}m_{y^{\prime}}(t)
+γry(t)α1+ξ0ξ1ω1mx(t).\displaystyle+\gamma r_{y^{\prime}}(t)-\frac{\alpha}{1+\xi_{0}\xi_{1}}\omega_{1}m_{x}(t). (53)

Similarly the following linearized equation of motion for my(t)m_{y^{\prime}}(t) up to the first order of α\alpha is obtained as

m˙y(t)\displaystyle\dot{m}_{y^{\prime}}(t) =(1+αξ01+ξ0ξ1)ω1mx(t)\displaystyle=\left(1+\frac{\alpha\xi_{0}}{1+\xi_{0}\xi_{1}}\right)\omega_{1}m_{x}(t)
γrx(t)α1+ξ0ξ1ω0my(t).\displaystyle-\gamma r_{x}(t)-\frac{\alpha}{1+\xi_{0}\xi_{1}}\omega_{0}m_{y^{\prime}}(t). (54)

Using the symbols defined by Eqs. (24), (25), and (26), one can easily confirm that Eqs. (A.0.1) and (A.0.1) are the same as Eqs. (22) and (23), respectively.

A.0.2 Long τc\tau_{c} regime: ξ0ξ1>1\xi_{0}\xi_{1}>1

In the long bath τc\tau_{c} regime satisfying ξ0ξ1>1\xi_{0}\xi_{1}>1, we expand Eq. (III.2) in power series of 1/(ξ0ξ1)1/(\xi_{0}\xi_{1}). Using the integration by parts with dν(tt)/dt=ν(tt)/τcd\nu(t-t^{\prime})/dt^{\prime}=\nu(t-t^{\prime})/\tau_{c} the integral part of the non-Markovian damping in Eq. (III.2) can be written as

tν(tt)m˙y(t)𝑑t=1τctm˙y(t)𝑑t\displaystyle\int_{-\infty}^{t}\nu(t-t^{\prime})\dot{m}_{y^{\prime}}(t^{\prime})dt^{\prime}=\frac{1}{\tau_{c}}\int_{-\infty}^{t}\dot{m}_{y^{\prime}}(t^{\prime})dt^{\prime}
1τctν(tt)[tm˙y(t′′)𝑑t′′]𝑑t.\displaystyle-\frac{1}{\tau_{c}}\int_{-\infty}^{t}\nu(t-t^{\prime})\left[\int_{-\infty}^{t^{\prime}}\dot{m}_{y^{\prime}}(t^{\prime\prime})dt^{\prime\prime}\right]dt^{\prime}. (55)

Successive application of the integration by parts gives

tν(tt)m˙y(t)𝑑t=n=1(1τc)nJn,\displaystyle\int_{-\infty}^{t}\nu(t-t^{\prime})\dot{m}_{y^{\prime}}(t^{\prime})dt^{\prime}=-\sum_{n=1}^{\infty}\left(-\frac{1}{\tau_{c}}\right)^{n}J_{n}, (56)

where JnJ_{n} is the nnth order multiple integral defined as

Jn=tt1tn1m˙y(tn)𝑑tn𝑑t2𝑑t1.\displaystyle J_{n}=\int_{-\infty}^{t}\int_{-\infty}^{t_{1}}\cdots\int_{-\infty}^{t_{n-1}}\dot{m}_{y^{\prime}}(t_{n})dt_{n}\cdots dt_{2}dt_{1}. (57)

From Eq. (48), on the other hand, m˙y(t)\dot{m}_{y^{\prime}}(t) is expressed as

m˙y(t)=1(1)nω0nω1n(d2n+1dt2n+1my(t)).\displaystyle\dot{m}_{y^{\prime}}(t)=\frac{1}{(-1)^{n}\omega_{0}^{n}\omega_{1}^{n}}\left(\frac{d^{2n+1}}{dt^{2n+1}}m_{y^{\prime}}(t)\right). (58)

Substituting Eq. (58) into Eq. (57) the multiple integrals are calculated as

J2n\displaystyle J_{2n} =1(1)nω0nω1nm˙y(t),\displaystyle=\frac{1}{(-1)^{n}\omega_{0}^{n}\omega_{1}^{n}}\dot{m}_{y^{\prime}}(t), (59)

and

J2n1\displaystyle J_{2n-1} =1(1)nω0nω1nm¨y(t).\displaystyle=\frac{1}{(-1)^{n}\omega_{0}^{n}\omega_{1}^{n}}\ddot{m}_{y^{\prime}}(t). (60)

Substituting the time derivative of my(t)m_{y^{\prime}}(t) in the 0th order of α\alpha into Eqs. (59) and (60), we obtain

J2n=ω1mx(t)(1)nω0nω1n,\displaystyle J_{2n}=\frac{\omega_{1}m_{x}(t)}{(-1)^{n}\omega_{0}^{n}\omega_{1}^{n}}, (61)

and

J2n1=my(t)(1)n1ω0n1ω1n1.\displaystyle J_{2n-1}=\frac{m_{y^{\prime}}(t)}{(-1)^{n-1}\omega_{0}^{n-1}\omega_{1}^{n-1}}. (62)

Substituting Eqs. (61) and (62) into Eq. (56) and performing some algebra, the integral part of the non-Markovian damping in Eq. (III.2) can be expressed as

tν(tt)m˙y(t)𝑑t\displaystyle\int_{-\infty}^{t}\nu(t-t^{\prime})\dot{m}_{y^{\prime}}(t^{\prime})dt^{\prime}
=n=1[(1τc)2n1J2n1+(1τc)2nJ2n]\displaystyle=-\sum_{n=1}^{\infty}\left[\left(-\frac{1}{\tau_{c}}\right)^{2n-1}J_{2n-1}+\left(-\frac{1}{\tau_{c}}\right)^{2n}J_{2n}\right]
=[n=1(1ξ0ξ1)n][ξ1ω0my(t)+ω1mx(t)]\displaystyle=-\left[\sum_{n=1}^{\infty}\left(-\frac{1}{\xi_{0}\xi_{1}}\right)^{n}\right]\Bigl{[}\xi_{1}\omega_{0}m_{y^{\prime}}(t)+\omega_{1}m_{x}(t)\Bigr{]}
=11+ξ0ξ1[ξ1ω0my(t)+ω1mx(t)].\displaystyle=\frac{1}{1+\xi_{0}\xi_{1}}\left[\xi_{1}\omega_{0}m_{y^{\prime}}(t)+\omega_{1}m_{x}(t)\right]. (63)

In the last equality, we use the following relation:

n=1(1ξ0ξ1)n=11+ξ0ξ1,\displaystyle\sum_{n=1}^{\infty}\left(-\frac{1}{\xi_{0}\xi_{1}}\right)^{n}=-\frac{1}{1+\xi_{0}\xi_{1}}, (64)

which holds under the condition that ξ0ξ1>1\xi_{0}\xi_{1}>1. Equation (A.0.2) is the same as Eq. (52). Substituting Eq. (A.0.2) into Eq. (III.2) and preforming some algebra, we obtain the same linearized equation of motion of mx(t)m_{x}(t) as Eq. (22). Equation (23) can also be obtained by similar calculations.

References