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

Chandrasekhar-Kendall-Woltjer-Taylor state in a resistive plasma

Ze-Yu Zhai [email protected] Interdisciplinary Center for Theoretical Study and Department of Modern Physics, University of Science and Technology of China, Hefei, Anhui 230026, China Peng Huanwu Center for Fundamental Theory, Hefei, Anhui 230026, China    Yang-Guang Yang [email protected] Interdisciplinary Center for Theoretical Study and Department of Modern Physics, University of Science and Technology of China, Hefei, Anhui 230026, China Peng Huanwu Center for Fundamental Theory, Hefei, Anhui 230026, China    Xiao-Liang Xia [email protected] Department of Physics and Center for Field Theory and Particle Physics, Fudan University, Shanghai 200433, China    Qun Wang [email protected] Interdisciplinary Center for Theoretical Study and Department of Modern Physics, University of Science and Technology of China, Hefei, Anhui 230026, China Peng Huanwu Center for Fundamental Theory, Hefei, Anhui 230026, China
Abstract

We give a criterion for the Chandrasekhar-Kendall-Woltjer-Taylor (CKWT) state in a resistive plasma. We find that the lowest momentum (longest wavelength) of the initial helicity amplitudes of magnetic fields are the key to the CKWT state which can be reached if one helicity is favored over the other. This indicates that the imbalance between two helicities at the lowest momentum or longest wavelength in the initial conditions is essential to the CKWT state. A few examples of initial conditions for helicity amplitudes are taken to support the above statement both analytically and numerically.

preprint: USTC-ICTS/PCFT-22-02

I Introduction

Many observations indicate that a magnetohydrodynamic (MHD) plasma or a fluid can evolve into a special static state (Robinson, 1969; Bodin and Newton, 1980; Ortolani and Schnack, 1993; Sarff et al., 1997; Yagi et al., 1999; Sarff et al., 2003; Ding et al., 2004; Prager et al., 2005; Lorenzini et al., 2009), in which a time-varying vector field is parallel to its curl,

𝑭×(×𝑭)=0.\boldsymbol{F}\times(\boldsymbol{\nabla}\times\boldsymbol{F})=0. (1)

This type of vector field is called the Beltrami field and was first studied by Beltrami (Beltrami, 1889). In contrast, when the vector field is orthogonal to its curl,

𝑭(×𝑭)=0,\boldsymbol{F}\cdot(\boldsymbol{\nabla}\times\boldsymbol{F})=0, (2)

the field is called the complex lamellar field.

In the MHD plasma, the Beltrami field is just a force-free (magnetic) field that was first discussed by Lust (Lüst and Schlüter, 1954) and Chandrasekhar (Chandrasekhar and Kendall, 1957) in the context of cosmology. Then Chandrasekhar and Woltjer (Chandrasekhar and Woltjer, 1958; Woltjer, 1958) gave the first analytical solution for such a force-free field. Such a state of the MHD plasma is later called Chandrasekhar-Kendall-Woltjer-Taylor (CKWT) state. The CKWT state satisfies the following equation for the magnetic field,

×𝑩=λ(𝒓)𝑩,\boldsymbol{\nabla}\times\boldsymbol{B}=\lambda(\boldsymbol{r})\boldsymbol{B}, (3)

where λ(𝒓)\lambda(\boldsymbol{r}) is a space-varying coefficient. If λ(𝒓)=λ\lambda(\boldsymbol{r})=\lambda is a space constant, we call the magnetic field is in a strong CKWT state. Otherwise, the magnetic field is in a general CKWT state.

Some natural questions arise: can the CKWT state be reached? Under what conditions can it be reached? Woltjer showed that the CKWT state has the minimum magnetic energy at a fixed magnetic helicity (Woltjer, 1958). As an invariant of plasma motion, magnetic helicity is associated with the topological properties of the magnetic field lines and measures them with the net twisting and braiding numbers (Wells and Norwood, 1969; Moffatt, 1978; Berger and Field, 1984; Arnold and Khesin, 1999). Later on, Taylor applied Woltjer’s idea to a plasma with small electrical resistance and found that Woltjer’s condition is valid (Taylor, 1974, 1986). To provide a natural way to minimize the magnetic energy while keeping the magnetic helicity fixed, Taylor speculated that the magnetic relaxation is caused by small-scale turbulence (Taylor, 1974, 1986; Schnack, 2009). However, both experimental and theoretical studies did not give conclusive evidence to support the hypothesis that the plasma relaxation should be dominated by short-wavelength properties (Robinson, 1969; Bodin and Newton, 1980; Caramana et al., 1983; Schnack et al., 1985; Strauss, 1985; Kusano and Sato, 1987; Ortolani and Schnack, 1993; Holmes et al., 1988; Ho and Craddock, 1991; Sarff et al., 1997; Yagi et al., 1999; Sarff et al., 2003; Diamond and Malkov, 2003; Ding et al., 2004; Prager et al., 2005; Marrelli et al., 2005; Lorenzini et al., 2009). The idea that the fluctuations seem to have a global long-wavelength structure is supported by extensive numerical simulations, which show that the relaxation is caused by the long-wavelength instability and nonlinear interaction.

To overcome the shortcoming of Taylor’s theory, the relaxation theory was developed using an infinite set of other approximate invariants by different authors (Bhattacharjee et al., 1980; Bhattacharjee and Dewar, 1982). Another study on how to reach the CKWT state in the resistive plasmas without Taylor’s conjecture is proposed in Ref. (Qin et al., 2012). Although the conditions in this work are not sufficient, the methods are useful and have been applied in subsequent studies. The authors of Ref. (Hirono et al., 2015) investigated the helicity evolution of an expanding chiral plasma in magnetic fields with the chiral magnetic effect (Vilenkin, 1980; Kharzeev et al., 2008; Fukushima et al., 2008) based on an expansion of the fields in the vector spherical harmonics (VSH) [for recent reviews of the chiral magnetic effect and related topics, see, e.g. Refs. (Kharzeev et al., 2013, 2016)]. The VSH method was later applied to study the CKWT state in a chiral plasma with the chiral magnetic effect by some of us (Xia et al., 2016), and it is found that the chiral magnetic effect plays the role of seed to the realization of the CKWT state.

A natural question arises: can the CKWT state be reached without the chiral magnetic effect? In this paper, we are going to answer this question by using the VSH method and a set of inequalities about magnetic fields and vector potentials. We will propose a criterion for the CKWT state, with which we find that the lowest momentum in the initial helicity amplitudes of magnetic fields is the key to the CKWT state.

The paper is organized as follows. In Section II, we will introduce the basic knowledge about the CKWT state. In Section III, we will give the criterion for the CKWT state through observables. In Section IV, we will introduce the VSH method to calculate the time evolution of these observables. In Section V, we will study under which initial conditions the CKWT state can be reached. We will summarize the main results of this paper in the final section.

II Basics of CKWT state

We start from Maxwell equations,

×𝑩\displaystyle\boldsymbol{\nabla}\times\boldsymbol{B} =\displaystyle= 𝑬t+𝒋,\displaystyle\frac{\partial\boldsymbol{E}}{\partial t}+\boldsymbol{j}, (4)
×𝑬\displaystyle\boldsymbol{\nabla}\times\boldsymbol{E} =\displaystyle= 𝑩t,\displaystyle-\frac{\partial\boldsymbol{B}}{\partial t}, (5)
𝑩\displaystyle\boldsymbol{\nabla}\cdot\boldsymbol{B} =\displaystyle= 0,\displaystyle 0, (6)
𝑬\displaystyle\boldsymbol{\nabla}\cdot\boldsymbol{E} =\displaystyle= 0,\displaystyle 0, (7)

where 𝑬\boldsymbol{E} and 𝑩\boldsymbol{B} are the electric and magnetic field respectively. The current 𝒋\boldsymbol{j} reads

𝒋=σ𝑬,\boldsymbol{j}=\sigma\boldsymbol{E}, (8)

where σ\sigma is the electric conductivity. After taking a curl of Eq. (4), we obtain an evolution equation for the magnetic field,

2t2𝑩+σt𝑩\displaystyle\frac{\partial^{2}}{\partial t^{2}}\boldsymbol{B}+\sigma\frac{\partial}{\partial t}\boldsymbol{B} =\displaystyle= 2𝑩.\displaystyle\nabla^{2}\boldsymbol{B}. (9)

In this paper we assume that σ\sigma is a constant. We also assume that terms of second-order time derivatives are much smaller than those of first-order one, which is valid for a slowly time-varying system. In this case, Eq. (9) is reduced to

t𝑩=η2𝑩,\frac{\partial}{\partial t}\boldsymbol{B}=\eta\nabla^{2}\boldsymbol{B}, (10)

where η=1/σ\eta=1/\sigma is the electrical resistance.

The authors of Ref. (Qin et al., 2012) studied the general conditions for the CKWT state in a MHD plasma. It is helpful to introduce the following inner products

W\displaystyle W =𝑩,𝑩=Ω𝑩2d3𝒙,\displaystyle=\left\langle\boldsymbol{B},\boldsymbol{B}\right\rangle=\int_{\Omega}\boldsymbol{B}^{2}d^{3}\boldsymbol{x},
Q\displaystyle Q =𝑨,𝑨=Ω𝑨2d3𝒙,\displaystyle=\left\langle\boldsymbol{A},\boldsymbol{A}\right\rangle=\int_{\Omega}\boldsymbol{A}^{2}d^{3}\boldsymbol{x},
H\displaystyle H =𝑨,𝑩=Ω𝑨𝑩d3𝒙,\displaystyle=\left\langle\boldsymbol{A},\boldsymbol{B}\right\rangle=\int_{\Omega}\boldsymbol{A}\cdot\boldsymbol{B}d^{3}\boldsymbol{x}, (11)

where WW is the magnetic energy, HH is the magnetic helicity, and Ω\Omega is the space volume. Using Eq. (10), we obtain

dQdt\displaystyle\frac{dQ}{dt} =2ηΩ𝑩2d3𝒙,\displaystyle=-2\eta\int_{\Omega}\boldsymbol{B}^{2}d^{3}\boldsymbol{x},
dWdt\displaystyle\frac{dW}{dt} =2ηΩ𝒋2d3𝒙,\displaystyle=-2\eta\int_{\Omega}\boldsymbol{j}^{2}d^{3}\boldsymbol{x},
dHdt\displaystyle\frac{dH}{dt} =2ηΩ𝒋𝑩d3𝒙.\displaystyle=-2\eta\int_{\Omega}\boldsymbol{j}\cdot\boldsymbol{B}d^{3}\boldsymbol{x}. (12)

After successively applying the Arithmetic Mean-Geometric Mean inequality and Cauchy-Schwarz inequality, one can prove (Qin et al., 2012)

ddt(WQH2)\displaystyle\frac{d}{dt}(WQ-H^{2}) \displaystyle\leq 4η[Ω𝑨𝑩d3𝒙Ω𝒋𝑩d3𝒙\displaystyle 4\eta\left[\int_{\Omega}\boldsymbol{A}\cdot\boldsymbol{B}d^{3}\boldsymbol{x}\int_{\Omega}\boldsymbol{j}\cdot\boldsymbol{B}d^{3}\boldsymbol{x}\right. (13)
Ω𝑨2d3𝒙Ω𝑩2d3𝒙Ω𝒋2d3𝒙Ω𝑩2d3𝒙]\displaystyle\left.-\sqrt{\int_{\Omega}\boldsymbol{A}^{2}d^{3}\boldsymbol{x}\int_{\Omega}\boldsymbol{B}^{2}d^{3}\boldsymbol{x}\int_{\Omega}\boldsymbol{j}^{2}d^{3}\boldsymbol{x}\int_{\Omega}\boldsymbol{B}^{2}d^{3}\boldsymbol{x}}\right]
\displaystyle\leq 0.\displaystyle 0.

The Cauchy-Schwartz inequality also gives the following inequality

WQ(Ω|𝑨||𝑩|d3𝒙)2(Ω𝑨𝑩d3𝒙)2=H2.WQ\geq\left(\int_{\Omega}|\boldsymbol{A}||\boldsymbol{B}|d^{3}\boldsymbol{x}\right)^{2}\geq\left(\int_{\Omega}\boldsymbol{A}\cdot\boldsymbol{B}d^{3}\boldsymbol{x}\right)^{2}=H^{2}. (14)

Inequalities (13) and (14) indicate that the quantity WQH2WQ-H^{2} is always positive and decreases with time until the condition 𝑩=λ𝑨\boldsymbol{B}=\lambda\boldsymbol{A} is reached, in which WQH2WQ-H^{2} is vanishing (Qin et al., 2012).

III Observables for CKWT State

As shown in Eqs. (13) and (14), QWH2QW-H^{2} is always positive and decreases with time unless 𝑩=λ𝑨\boldsymbol{B}=\lambda\boldsymbol{A} is reached. However, it is not sufficient to judge for the CKWT state only from a decreasing QWH2QW-H^{2}, since it can decrease as the magnitudes of 𝑨\boldsymbol{A} and 𝑩\boldsymbol{B} decrease while keeping a fixed angle between them (Chen and Fan, 2013). The sufficient condition for the CKWT state should be 𝑩\boldsymbol{B} and 𝑨\boldsymbol{A} are parallel. In this section, we propose to use the observable WQ/H21WQ/H^{2}-1 for the CKWT state provided H0H\neq 0 and it is non-negative with Cauchy-Schwartz inequality as we have shown in Section II. We will show in this section that the condition for the CKWT state should be

WQH2\displaystyle\frac{WQ}{H^{2}} 1=tan2(θ)t0,\displaystyle-1=\tan^{2}(\theta)\overset{t\rightarrow\infty}{\Longrightarrow}0, (15)

where θ\theta is an average angle between 𝑨\boldsymbol{A} and 𝑩\boldsymbol{B} defined through 𝑨,𝑩2=𝑨,𝑨𝑩,𝑩cos2(θ)\left\langle\boldsymbol{A},\boldsymbol{B}\right\rangle^{2}=\left\langle\boldsymbol{A},\boldsymbol{A}\right\rangle\left\langle\boldsymbol{B},\boldsymbol{B}\right\rangle\cos^{2}(\theta).

From QWH2=QWsin2(θ)QW-H^{2}=QW\sin^{2}(\theta), we see that the sufficient condition for the CKWT state is θ=0 or π\theta=0\text{ or }\pi. It is more convenient to introduce the quantity

tan2θQWH2H2=𝑨,𝑨𝑩,𝑩𝑨,𝑩2𝑨,𝑩2.\tan^{2}\theta\equiv\frac{QW-H^{2}}{H^{2}}=\frac{\left\langle\boldsymbol{A},\boldsymbol{A}\right\rangle\left\langle\boldsymbol{B},\boldsymbol{B}\right\rangle-\left\langle\boldsymbol{A},\boldsymbol{B}\right\rangle^{2}}{\left\langle\boldsymbol{A},\boldsymbol{B}\right\rangle^{2}}. (16)

Assuming that H0H\neq 0, the time rate of QWH2QW-H^{2} can be expressed as

ddt(QWH2)\displaystyle\frac{d}{dt}(QW-H^{2}) =\displaystyle= 2H2tan2θ(t)(dln|tanθ(t)|dt+dln|H|dt)0,\displaystyle 2H^{2}\tan^{2}\theta(t)\left(\frac{d\ln|\tan\theta(t)|}{dt}+\frac{d\ln|H|}{dt}\right)\leq 0, (17)

with two contributions: the angular one and helicity one. We can prove

dln|tanθ(t)|dt<0,\frac{d\ln|\tan\theta(t)|}{dt}<0, (18)

for tt\rightarrow\infty in order to approach the CKWT state, which means θ(t)<0\theta^{\prime}(t)<0 for θ[0,π/2)\theta\in[0,\pi/2) and θ(t)>0\theta^{\prime}(t)>0 for θ(π/2,π]\theta\in(\pi/2,\pi].

To prove that the necessary condition (18) is achievable, we look at a simple case of the helicity time evolution. The long time behaviors of QQ and HH lead to dln|H|/dt0d\ln|H|/dt\leq 0 when tt\rightarrow\infty. Then the condition (18) can be rewritten as

1QWd(QW)dt<1H2dH2dt.\frac{1}{QW}\frac{d(QW)}{dt}<\frac{1}{H^{2}}\frac{dH^{2}}{dt}. (19)

We take a simple example of to illustrate the above condition. To obtain an upper bound of the left-hand side of the above inequality, we employ the Poincare inequality for the vector field 𝒇\boldsymbol{f} in following form (Poincaré, 1890)

Ω𝒇2d3𝒙qΩ2Ω(×𝒇)2d3𝒙,\int_{\Omega}\boldsymbol{f}^{2}d^{3}\boldsymbol{x}\leq q_{\Omega}^{2}\int_{\Omega}(\boldsymbol{\nabla\times f})^{2}d^{3}\boldsymbol{x}, (20)

where qΩq_{\Omega} is a Poincare constant associated with the space volume Ω\Omega. Then we obtain the upper bound as

1QWd(QW)dt\displaystyle\frac{1}{QW}\frac{d(QW)}{dt} =\displaystyle= 2η(Ω𝑩2d3𝒙Ω𝑨2d3𝒙+Ω𝒋2d3𝒙Ω𝑩2d3𝒙)\displaystyle-2\eta\left(\frac{\int_{\Omega}\boldsymbol{B}^{2}d^{3}\boldsymbol{x}}{\int_{\Omega}\boldsymbol{A}^{2}d^{3}\boldsymbol{x}}+\frac{\int_{\Omega}\boldsymbol{j}^{2}d^{3}\boldsymbol{x}}{\int_{\Omega}\boldsymbol{B}^{2}d^{3}\boldsymbol{x}}\right) (21)
\displaystyle\leq 2η(qΩ2+qΩ2)=4ηqΩ2.\displaystyle-2\eta(q_{\Omega}^{-2}+q_{\Omega}^{-2})=-4\eta q_{\Omega}^{-2}.

Since helicity is a topological quantity of plasma evolution, here we postulate a tighter inequality than (19)

4ηqΩ2<1H2dH2dt,-4\eta q_{\Omega}^{-2}<\frac{1}{H^{2}}\frac{dH^{2}}{dt}, (22)

which can lead to (19) and (18). So if the condition (22) is satisfied the angle between 𝑨\boldsymbol{A} and 𝑩\boldsymbol{B} decreases with time. Furthermore, if θ(t)\theta(t) decreases fast enough, the system will reach the CKWT state in a finite time. We still need to know the time limit of tan2θ(t)\tan^{2}\theta(t) in order to judge for the CKWT state, which we will study in the next section.

IV Methods

To study the criteria for CKWT states, we need to analyze the time evolution of WQ/H2WQ/H^{2}, it is convenient to expand WW, QQ and HH in (11) as well as their time rates in (12) on the VSH basis (Jackson, 1999). The VSH basis functions are the eigenfunctions of the curl operator in momentum space. They have been used to study the time evolution of the magnetic helicity and the CKWT state in chiral plasma (Hirono et al., 2015; Xia et al., 2016).

IV.1 VSH expansion

We now expand 𝑨\boldsymbol{A} and 𝑩\boldsymbol{B} in terms of the VSH basis functions 𝑾lms(𝒙,k)\boldsymbol{W}_{lm}^{s}(\boldsymbol{x},k) as

𝑩(t,𝒙)\displaystyle\boldsymbol{B}(t,\boldsymbol{x}) =\displaystyle= 1πl,m0𝑑kk2[αlm+(t,k)𝑾lm+(𝒙,k)+αlm(t,k)𝑾lm(𝒙,k)],\displaystyle\frac{1}{\pi}\sum_{l,m}\int_{0}^{\infty}dkk^{2}\left[\alpha_{lm}^{+}(t,k)\boldsymbol{W}_{lm}^{+}(\boldsymbol{x},k)+\alpha_{lm}^{-}(t,k)\boldsymbol{W}_{lm}^{-}(\boldsymbol{x},k)\right],
𝑨(t,𝒙)\displaystyle\boldsymbol{A}(t,\boldsymbol{x}) =\displaystyle= 1πl,m0𝑑kk[αlm+(t,k)𝑾lm+(𝒙,k)αlm(t,k)𝑾lm(𝒙,k)],\displaystyle\frac{1}{\pi}\sum_{l,m}\int_{0}^{\infty}dkk\left[\alpha_{lm}^{+}(t,k)\boldsymbol{W}_{lm}^{+}(\boldsymbol{x},k)-\alpha_{lm}^{-}(t,k)\boldsymbol{W}_{lm}^{-}(\boldsymbol{x},k)\right], (23)

where αlm±(t,k)\alpha_{lm}^{\pm}(t,k) denote the coefficients of the expansion, and 𝑾lms(𝒙,k)\boldsymbol{W}_{lm}^{s}(\boldsymbol{x},k) (with s=±s=\pm being the helicity) denote the complete set of eigenfunctions (vectors) of the curl operator and are divergence-free

×𝑾lms(𝒙,k)\displaystyle\boldsymbol{\nabla}\times\boldsymbol{W}_{lm}^{s}(\boldsymbol{x},k) =\displaystyle= sk𝑾lms(𝒙,k),\displaystyle sk\boldsymbol{W}_{lm}^{s}(\boldsymbol{x},k),
𝑾lms(𝒙,k)\displaystyle\boldsymbol{\nabla}\cdot\boldsymbol{W}_{lm}^{s}(\boldsymbol{x},k) =\displaystyle= 0.\displaystyle 0. (24)

In 𝑾lms(𝒙,k)\boldsymbol{W}_{lm}^{s}(\boldsymbol{x},k), l=0,1,l=0,1,\cdots denotes the orbital angular momentum quantum number, m=l,l+1,,lm=-l,-l+1,\cdots,l denotes the magnetic quantum number, and k|𝒌|k\equiv|\boldsymbol{k}| is the norm of the momentum. The orthogonormality relations read

d3𝒙𝑾l1m1s1(𝒙,k)𝑾l2m2s2(𝒙,k)\displaystyle\int d^{3}\boldsymbol{x}\boldsymbol{W}_{l_{1}m_{1}}^{s_{1}}(\boldsymbol{x},k)\cdot\boldsymbol{W}_{l_{2}m_{2}}^{s_{2}}(\boldsymbol{x},k^{\prime}) =\displaystyle= πk2δ(kk)δl1l2δm1m2δs1s2.\displaystyle\frac{\pi}{k^{2}}\delta\left(k-k^{\prime}\right)\delta_{l_{1}l_{2}}\delta_{m_{1}m_{2}}\delta_{s_{1}s_{2}}. (25)

To be specific, 𝑾lms(𝒙,k)\boldsymbol{W}_{lm}^{s}(\boldsymbol{x},k) can be put into the form

𝑾lms(𝒙,k)=𝑻lms(𝒙,k)+sk×𝑻lms(𝒙,k),\boldsymbol{W}_{lm}^{s}(\boldsymbol{x},k)=\boldsymbol{T}_{lm}^{s}(\boldsymbol{x},k)+\frac{s}{k}\boldsymbol{\nabla}\times\boldsymbol{T}_{lm}^{s}(\boldsymbol{x},k), (26)

where 𝑻lms(𝒙,k)\boldsymbol{T}_{lm}^{s}(\boldsymbol{x},k) are toroidal fields and can be expressed as a combination of spherical Bessel function jl(kr)j_{l}(kr) and spherical harmonic functions Ylm(θ,ϕ)Y_{lm}\left(\theta,\phi\right).

IV.2 Solving Maxwell equation

Inserting Eq. (23) into Eq. (10), we obtain the evolution equation of the coefficients as

tαlm±(t,k)\displaystyle\frac{\partial}{\partial t}\alpha_{lm}^{\pm}(t,k) =\displaystyle= ηk2αlm±(t,k),\displaystyle-\eta k^{2}\alpha_{lm}^{\pm}(t,k), (27)

where η=1/σ\eta=1/\sigma. Once αlm±(t,k)\alpha_{lm}^{\pm}(t,k) are obtained by solving the above equation, the magnetic field 𝑩(t,𝒙)\boldsymbol{B}\left(t,\boldsymbol{x}\right) as a function of time is then known. The solutions of αlm±(t,k)\alpha_{lm}^{\pm}(t,k) are

αlm±(t,k)=eηk2tαlm±(0,k),\alpha_{lm}^{\pm}(t,k)=e^{-\eta k^{2}t}\alpha_{lm}^{\pm}(0,k), (28)

where αlm±(0,k)\alpha_{lm}^{\pm}(0,k) are the values at the initial time t=0t=0. We need to calculate inner products of two fields as in Eq. (11). It is convenient to introduce positive-definite functions g±(t,k)g_{\pm}(t,k) for the positive and negative helicity,

g±(t,k)=1πlm|αlm±(t,k)|2=e2ηk2tg±(0,k),g_{\pm}(t,k)=\frac{1}{\pi}\sum_{lm}\left|\alpha_{lm}^{\pm}(t,k)\right|^{2}=e^{-2\eta k^{2}t}g_{\pm}(0,k), (29)

where the initial values of g±(t,k)g_{\pm}(t,k) are g±(0,k)=(1/π)lm|αlm±(0,k)|2g_{\pm}(0,k)=(1/\pi)\sum_{lm}\left|\alpha_{lm}^{\pm}(0,k)\right|^{2}. In terms of g±(t,k)g_{\pm}(t,k), WW, QQ and HH in (11) can be put into the forms

W\displaystyle W =0𝑑kk2[g+(t,k)+g(t,k)],\displaystyle=\int_{0}^{\infty}dkk^{2}\left[g_{+}(t,k)+g_{-}(t,k)\right],
Q\displaystyle Q =0𝑑k[g+(t,k)+g(t,k)],\displaystyle=\int_{0}^{\infty}dk\left[g_{+}(t,k)+g_{-}(t,k)\right],
H\displaystyle H =0𝑑kk[g+(t,k)g(t,k)],\displaystyle=\int_{0}^{\infty}dkk\left[g_{+}(t,k)-g_{-}(t,k)\right], (30)

where we have used Eqs. (23-25). We see in the above equations that WW and QQ are invariant or symmetric under the interchange g+gg_{+}\leftrightarrow g_{-}, while HH is anti-symmetric under the the interchange g+gg_{+}\leftrightarrow g_{-}.

V Approach to CKWT State

In this section, we will investigate under what conditions the CKWT state is achieved.

V.1 Special initial conditions

We can explicitly express WQ/H2WQ/H^{2} in terms of g±(0,k)g_{\pm}(0,k) using Eq. (29),

WQH2=0𝑑k10𝑑k2e2ηt(k12+k22)k12[g+(0,k1)+g(0,k1)][g+(0,k2)+g(0,k2)]0𝑑k10𝑑k2e2ηt(k12+k22)k1k2[g+(0,k1)g(0,k1)][g+(0,k2)g(0,k2)],\frac{WQ}{H^{2}}=\frac{\int_{0}^{\infty}dk_{1}\int_{0}^{\infty}dk_{2}e^{-2\eta t(k_{1}^{2}+k_{2}^{2})}k_{1}^{2}\left[g_{+}(0,k_{1})+g_{-}(0,k_{1})\right]\left[g_{+}(0,k_{2})+g_{-}(0,k_{2})\right]}{\int_{0}^{\infty}dk_{1}\int_{0}^{\infty}dk_{2}e^{-2\eta t(k_{1}^{2}+k_{2}^{2})}k_{1}k_{2}\left[g_{+}(0,k_{1})-g_{-}(0,k_{1})\right]\left[g_{+}(0,k_{2})-g_{-}(0,k_{2})\right]}, (31)

with its time derivative given by

ddt(WQH2)=1H3(WQH+WQH2HWQ),\frac{d}{dt}\left(\frac{WQ}{H^{2}}\right)=\frac{1}{H^{3}}\left(W^{\prime}QH+WQ^{\prime}H-2H^{\prime}WQ\right), (32)

where we have used the notation XdX/dtX^{\prime}\equiv dX/dt with X=W,Q,HX=W,Q,H, and the numerator and denominator have the explicit forms

WQH+WQH2HWQ\displaystyle W^{\prime}QH+WQ^{\prime}H-2H^{\prime}WQ
=\displaystyle= 0𝑑k10𝑑k20𝑑k3e2ηt(k12+k22+k32)k12k3(k12+k222k32)\displaystyle\int_{0}^{\infty}dk_{1}\int_{0}^{\infty}dk_{2}\int_{0}^{\infty}dk_{3}e^{-2\eta t(k_{1}^{2}+k_{2}^{2}+k_{3}^{2})}k_{1}^{2}k_{3}\left(k_{1}^{2}+k_{2}^{2}-2k_{3}^{2}\right)
×[g+(0,k1)+g(0,k1)][g+(0,k2)+g(0,k2)][g+(0,k3)g(0,k3)],\displaystyle\times\left[g_{+}(0,k_{1})+g_{-}(0,k_{1})\right]\left[g_{+}(0,k_{2})+g_{-}(0,k_{2})\right]\left[g_{+}(0,k_{3})-g_{-}(0,k_{3})\right],
H3\displaystyle H^{3} =\displaystyle= 0𝑑k10𝑑k20𝑑k3e2ηt(k12+k22+k32)k1k2k3\displaystyle\int_{0}^{\infty}dk_{1}\int_{0}^{\infty}dk_{2}\int_{0}^{\infty}dk_{3}e^{-2\eta t(k_{1}^{2}+k_{2}^{2}+k_{3}^{2})}k_{1}k_{2}k_{3} (33)
×[g+(0,k1)g(0,k1)][g+(0,k2)g(0,k2)][g+(0,k3)g(0,k3)].\displaystyle\times\left[g_{+}(0,k_{1})-g_{-}(0,k_{1})\right]\left[g_{+}(0,k_{2})-g_{-}(0,k_{2})\right]\left[g_{+}(0,k_{3})-g_{-}(0,k_{3})\right].

In the following, we will look at the long-time behaviour of WQ/H2WQ/H^{2} at tt\rightarrow\infty under some initial conditions. In the following analysis and calculation, we use a typical length of the magnetic field LL to scale the physical quantities, and we substitute tt/Lt\rightarrow t/L, kkLk\rightarrow kL, ηη/L\eta\rightarrow\eta/L and g±g±/L2g_{\pm}\rightarrow g_{\pm}/L^{2}, so all re-scaled quantities are dimensionless.

V.1.1 Delta functions

First of all, we consider an ideal case of initial functions g±(0,k)g_{\pm}\left(0,k\right) with two different discrete momentum values aa and bb

g+(0,k)\displaystyle g_{+}(0,k) =\displaystyle= aδ(ka)+b2δ(kb),\displaystyle a\delta(k-a)+\frac{b}{2}\delta(k-b),
g(0,k)\displaystyle g_{-}(0,k) =\displaystyle= b2δ(kb).\displaystyle\frac{b}{2}\delta(k-b). (34)

We see that the positive helicity part has two momentum values while the negative helicity part has only one value. It is easy to obtain

WQH21=1a4[(a3b+ab3)e2(a2b2)ηt+b4e4(a2b2)ηt],\frac{WQ}{H^{2}}-1=\frac{1}{a^{4}}\left[\left(a^{3}b+ab^{3}\right)e^{2\left(a^{2}-b^{2}\right)\eta t}+b^{4}e^{4\left(a^{2}-b^{2}\right)\eta t}\right], (35)

with the tt\rightarrow\infty limit

limtWQH21{0a<ba>b.\lim_{t\rightarrow\infty}\frac{WQ}{H^{2}}-1\rightarrow\begin{cases}0&a<b\\ \infty&a>b\end{cases}. (36)

We can see that only when a<ba<b the CKWT state can be achieved at tt\rightarrow\infty. In this case, helicity is dominated by the low momentum mode. On the other hand, if the high momentum mode is dominant, the CKWT state cannot be reached. Nevertheless, since the delta function is not mathematically well-defined and should be replaced by more physical initial conditions, this simple case still provides a clue to more general conditions.

V.1.2 Two-band functions

As a more general case than delta-functions, we consider Heaviside step functions for g±(0,k)g_{\pm}(0,k) with two bands (the lower momentum band and higher momentum band),

g+(0,k)\displaystyle g_{+}(0,k) ={d1+,forkd1k<kd2d2+,forkd2k<kd3,\displaystyle=\begin{cases}d_{1}^{+},&\mathrm{for}\;\;k_{d1}\leq k<k_{d2}\\ d_{2}^{+},&\mathrm{for}\;\;k_{d2}\leq k<k_{d3}\end{cases}, (37)
g(0,k)\displaystyle g_{-}(0,k) ={d1,forkd1k<kd2d2,forkd2k<kd3,\displaystyle=\begin{cases}d_{1}^{-},&\mathrm{for}\;\;k_{d1}\leq k<k_{d2}\\ d_{2}^{-},&\mathrm{for}\;\;k_{d2}\leq k<k_{d3}\end{cases}, (38)

where k0k\geq 0, kd3>kd2>kd10k_{d3}>k_{d2}>k_{d1}\geq 0, d1++d1>0d_{1}^{+}+d_{1}^{-}>0 and di±0d_{i}^{\pm}\geq 0 for i=1,2i=1,2. We can verify the following limit when tt is sent to infinity,

limtWQH21{(d1++d1)2(d1+d1)21,forkd1>0π2(d1++d1)2(d1+d1)21,forkd1=0.\lim_{t\rightarrow\infty}\frac{WQ}{H^{2}}-1\rightarrow\begin{cases}\frac{\left(d_{1}^{+}+d_{1}^{-}\right)^{2}}{\left(d_{1}^{+}-d_{1}^{-}\right)^{2}}-1,&\mathrm{for}\;\;k_{d1}>0\\ \frac{\pi}{2}\frac{\left(d_{1}^{+}+d_{1}^{-}\right)^{2}}{\left(d_{1}^{+}-d_{1}^{-}\right)^{2}}-1,&\mathrm{for}\;\;k_{d1}=0\end{cases}. (39)

We see that such a limit is determined by the amplitudes of the lower momentum bands. The conditions for the CKWT state would be

Δ(d1++d1)2(d1+d1)2={1,forkd1>02/π,forkd1=0.\Delta\equiv\frac{\left(d_{1}^{+}+d_{1}^{-}\right)^{2}}{\left(d_{1}^{+}-d_{1}^{-}\right)^{2}}=\left\{\begin{array}[]{ll}1,&\mathrm{for}\;\;k_{d1}>0\\ 2/\pi,&\mathrm{for}\;\;k_{d1}=0\end{array}\right.. (40)

Because d1±0d_{1}^{\pm}\geq 0 and d1++d1>0d_{1}^{+}+d_{1}^{-}>0, Δ\Delta must not be less than 1 or Δ1\Delta\geq 1, so Δ\Delta cannot be 2/π2/\pi for the case kd1=0k_{d1}=0, which means that the CKWT state cannot be reached for kd1=0k_{d1}=0. For kd1>0k_{d1}>0, the CKWT state can be reached if and only if either of d1+d_{1}^{+} or d1d_{1}^{-} is vanishing. This observation can be verified by the numerical results in Fig. 1 for different sets of values of d1±d_{1}^{\pm} and d2±d_{2}^{\pm}.

Figure 1: Numerical results for tan2(θ)\tan^{2}(\theta) with two-band initial conditions for different set of amplitude values. The parameters are: η=0.1\eta=0.1, kd1=0or 1k_{d1}=0\;\mathrm{or}\;1, kd2=2k_{d2}=2, kd3=4k_{d3}=4. A: d1+=1d_{1}^{+}=1, d2+=1/4d_{2}^{+}=1/4, d1=0d_{1}^{-}=0, d2=1/4d_{2}^{-}=1/4, B: d1+=1/4d_{1}^{+}=1/4, d2+=1d_{2}^{+}=1, d1=0d_{1}^{-}=0, d2=1/4d_{2}^{-}=1/4, C: d1+=1/4d_{1}^{+}=1/4, d2+=1d_{2}^{+}=1, d1=1/4d_{1}^{-}=1/4, d2=0d_{2}^{-}=0, D: d1+=1d_{1}^{+}=1, d2+=1/4d_{2}^{+}=1/4, d1=1/4d_{1}^{-}=1/4, d2=0d_{2}^{-}=0, E: d1+=1d_{1}^{+}=1, d2+=1/4d_{2}^{+}=1/4, d1=1/3d_{1}^{-}=1/3, d2=1/3d_{2}^{-}=1/3. (a) kd1=1k_{d1}=1. In case A and B, tan2(θ)\tan^{2}(\theta) tends to zero as tt\rightarrow\infty indicating that the CKWT state can be reached, while tan2(θ)\tan^{2}(\theta) tends to infinity, 16/916/9 and 33 as tt\rightarrow\infty in case C, D and E, respectively, which indicates that the CKWT state cannot be reached. (b) kd1=0k_{d1}=0. The CKWT state cannot be reached in all cases.
Refer to caption

V.1.3 Multi-band functions

We now generalize two-steps functions to multi-steps functions,

g+(0,k)\displaystyle g_{+}\left(0,k\right) ={d1+,forkd1k<kd2d2+,forkd2k<kd3dn+,forkd(n)k<kd(n+1),\displaystyle=\begin{cases}d_{1}^{+},&\mathrm{for}\;\;k_{d1}\leq k<k_{d2}\\ d_{2}^{+},&\mathrm{for}\;\;k_{d2}\leq k<k_{d3}\\ ...&...\\ d_{n}^{+},&\mathrm{for}\;\;k_{d(n)}\leq k<k_{d(n+1)}\end{cases}, (41)
g(0,k)={d1,forkd1k<kd2d2,forkd2k<kd3dn,forkd(n)k<kd(n+1),g_{-}\left(0,k\right)=\begin{cases}d_{1}^{-},&\mathrm{for}\;\;k_{d1}\leq k<k_{d2}\\ d_{2}^{-},&\mathrm{for}\;\;k_{d2}\leq k<k_{d3}\\ ...&...\\ d_{n}^{-},&\mathrm{for}\;\;k_{d(n)}\leq k<k_{d(n+1)}\end{cases}, (42)

where k0k\geq 0, kd(n+1)>kd(n)>>kd2>kd10k_{d(n+1)}>k_{d(n)}>...>k_{d2}>k_{d1}\geq 0, d1++d1>0d_{1}^{+}+d_{1}^{-}>0 and di±0d_{i}^{\pm}\geq 0 for i=1,2,,ni=1,2,...,n. The result is similar to the case of two-bands functions,

limtWQH21{(d1++d1)2(d1+d1)21,forkd1>0π2(d1++d1)2(d1+d1)21,forkd1=0.\lim_{t\rightarrow\infty}\frac{WQ}{H^{2}}-1\rightarrow\begin{cases}\frac{\left(d_{1}^{+}+d_{1}^{-}\right)^{2}}{\left(d_{1}^{+}-d_{1}^{-}\right)^{2}}-1,&\mathrm{for}\;\;k_{d1}>0\\ \frac{\pi}{2}\frac{\left(d_{1}^{+}+d_{1}^{-}\right)^{2}}{\left(d_{1}^{+}-d_{1}^{-}\right)^{2}}-1,&\mathrm{for}\;\;k_{d1}=0\end{cases}. (43)

We see that the limit for WQ/H2WQ/H^{2} is also determined by the amplitudes of the lowest bands. Similar to the analysis in the previous subsection that the CKWT state can only be reached for kd1>0k_{d1}>0 under the condition that either of d1+d_{1}^{+} or d1d_{1}^{-} is vanishing. This observation can be verified by the numerical results in Fig. 2 for different sets of values of di±d_{i}^{\pm} for i=1,,4i=1,\cdots,4.

Figure 2: Numerical results for tan2(θ)\tan^{2}(\theta) with multi-band initial conditions for different sets of amplitude values. The parameters are: η=0.1\eta=0.1, kd1=0or 1k_{d1}=0\;\mathrm{or}\;1, kd2=2k_{d2}=2, kd3=5k_{d3}=5, kd4=10k_{d4}=10, kd5=16k_{d5}=16. Case A: {d1+,d2+,d3+,d4+}={1,1,1,0}\{d_{1}^{+},\text{$d_{2}^{+}$},d_{3}^{+},d_{4}^{+}\}=\{1,1,1,0\}, {d1,d2,d3,d4}={0,1/4,0,1}\{d_{1}^{-},\text{$d_{2}^{-}$},d_{3}^{-},d_{4}^{-}\}=\{0,1/4,0,1\}; Case B: {d1+,d2+,d3+,d4+}={1/4,2,1,0}\{d_{1}^{+},\text{$d_{2}^{+}$},d_{3}^{+},d_{4}^{+}\}=\{1/4,2,1,0\}, {d1,d2,d3,d4}={0,3,0,1/4}\{d_{1}^{-},\text{$d_{2}^{-}$},d_{3}^{-},d_{4}^{-}\}=\{0,3,0,1/4\}; Case C: {d1+,d2+,d3+,d4+}={1/4,1,0,1}\{d_{1}^{+},\text{$d_{2}^{+}$},d_{3}^{+},d_{4}^{+}\}=\{1/4,1,0,1\}, {d1,d2,d3,d4}={1/4,4,0,1}\{d_{1}^{-},\text{$d_{2}^{-}$},d_{3}^{-},d_{4}^{-}\}=\{1/4,4,0,1\}; Case D: {d1+,d2+,d3+,d4+}={1,2,1,0}\{d_{1}^{+},\text{$d_{2}^{+}$},d_{3}^{+},d_{4}^{+}\}=\{1,2,1,0\}, {d1,d2,d3,d4}={1/4,3,1,1}\{d_{1}^{-},\text{$d_{2}^{-}$},d_{3}^{-},d_{4}^{-}\}=\{1/4,3,1,1\}; Case E: {d1+,d2+,d3+,d4+}={1,1/4,1,0}\{d_{1}^{+},\text{$d_{2}^{+}$},d_{3}^{+},d_{4}^{+}\}=\{1,1/4,1,0\}, {d1,d2,d3,d4}={1/3,1/4,0,1}\{d_{1}^{-},\text{$d_{2}^{-}$},d_{3}^{-},d_{4}^{-}\}=\{1/3,1/4,0,1\}. (a) kd1=1k_{d1}=1. In case A and B, the CKWT state can be reached as tt\rightarrow\infty. In other cases the CKWT state cannot be reached. (b) kd1=0k_{d1}=0. The CKWT state cannot be reached for all cases.
Refer to caption

V.2 Analytic functions as initial conditions

Based on the previous discussion about step functions as initial conditions, it is natural to generalize it to the limit of infinitely small intervals, i.e. analytic functions. As discussed in the previous subsection, the starting point of the integral can make a difference in the limit of WQ/H2WQ/H^{2}. In this subsection, we will do the same thing by distinguishing two cases: a>0a>0 and a=0a=0 for the starting point of the integral range [a,)[a,\infty).

V.2.1 Integration range [a,)[a,\infty) with a>0a>0

The physical quantities in our consideration are all in the integrated form

X=a𝑑kkne2ηtk2f(k),X=\int_{a}^{\infty}dkk^{n}e^{-2\eta tk^{2}}f(k), (44)

where the lower bound aa of the integration range is a positive number, and f(k)f(k) is an analytic function, meaning that the Taylor expansion is valid at any value k0k_{0} in the range [a,)[a,\infty),

f(k)=n=01n!(kk0)nf(n)(k0).f(k)=\sum_{n=0}^{\infty}\frac{1}{n!}\left(k-k_{0}\right)^{n}f^{(n)}(k_{0}). (45)

For WW, QQ and HH which we are considering in this paper, f(k)f(k) can be either ϕ(k)\phi(k) or φ(k)\varphi(k),

ϕ(k)\displaystyle\phi(k) =\displaystyle= g+(0,k)+g(0,k),\displaystyle g_{+}(0,k)+g_{-}(0,k),
φ(k)\displaystyle\varphi(k) =\displaystyle= g+(0,k)g(0,k).\displaystyle g_{+}(0,k)-g_{-}(0,k). (46)

So WW, QQ and HH can be put into the forms,

W\displaystyle W =\displaystyle= a𝑑kk2e2ηtk2[n=01n!ϕ(n)(a)(ka)n],\displaystyle\int_{a}^{\infty}dkk^{2}e^{-2\eta tk^{2}}\left[\sum_{n=0}^{\infty}\frac{1}{n!}\phi^{(n)}(a)\left(k-a\right)^{n}\right],
Q\displaystyle Q =\displaystyle= a𝑑ke2ηtk2[n=01n!ϕ(n)(a)(ka)n],\displaystyle\int_{a}^{\infty}dke^{-2\eta tk^{2}}\left[\sum_{n=0}^{\infty}\frac{1}{n!}\phi^{(n)}(a)\left(k-a\right)^{n}\right],
H\displaystyle H =\displaystyle= a𝑑kke2ηtk2[n=01n!φ(n)(a)(ka)n].\displaystyle\int_{a}^{\infty}dkke^{-2\eta tk^{2}}\left[\sum_{n=0}^{\infty}\frac{1}{n!}\varphi^{(n)}\left(a\right)\left(k-a\right)^{n}\right]. (47)

Then we obtain the long time limit

limtWQH21[g+(i)(0,a)+g(i)(0,a)g+(i)(0,a)g(i)(0,a)]21,\lim_{t\rightarrow\infty}\frac{WQ}{H^{2}}-1\rightarrow\left[\frac{g_{+}^{(i)}(0,a)+g_{-}^{(i)}(0,a)}{g_{+}^{(i)}(0,a)-g_{-}^{(i)}(0,a)}\right]^{2}-1, (48)

where one can explicitly define the derivative index ii (i0i\geq 0) with the following two cases: (a) i=min(i+,i)i=\min(i_{+},i_{-}) if g+(0,k)g_{+}(0,k) and g(0,k)g_{-}(0,k) are all non-vanishing, where isi_{s} is the index denote for a lowest order isi_{s}-th derivative that makes gs(is)(0,a)g_{s}^{(i_{s})}(0,a) non-vanishing with s=±s=\pm, respectively; (b) If one of g+(0,k)g_{+}(0,k) and g(0,k)g_{-}(0,k) is vanishing, for example, g+(0,k)=0g_{+}(0,k)=0, then for a lowest order ii-th derivative g(i)(0,a)0g_{-}^{(i)}(0,a)\neq 0. So is the case g(0,k)=0g_{-}(0,k)=0.

We see that the CKWT can be reached at tt\rightarrow\infty if and only if either g+(i)(0,a)g_{+}^{(i)}(0,a) or g(i)(0,a)g_{-}^{(i)}(0,a) is vanishing. The proof of the result (48) is given in Appendix A.

V.2.2 Integration range [0,)[0,\infty)

In this section we consider the integration range [0,)[0,\infty), in which g+(0,k)g_{+}(0,k) and g(0,k)g_{-}(0,k) are analytic functions and can be expanded in a Taylor expansion. In the following we use the shorthand notation g±(k)g_{\pm}(k) for g±(0,k)g_{\pm}(0,k).

At k=0k=0, the Taylor expansion of ϕ(k)\phi(k) and φ(k)\varphi(k) in Eq. (46) reads

ϕ(k)\displaystyle\phi(k) =\displaystyle= ϕ(0)+ϕ(0)k+12ϕ′′(0)k2+,\displaystyle\phi(0)+\phi^{\prime}(0)k+\frac{1}{2}\phi^{\prime\prime}(0)k^{2}+...,
φ(k)\displaystyle\varphi(k) =\displaystyle= φ(0)+φ(0)k+12φ′′(0)k2+.\displaystyle\varphi(0)+\varphi^{\prime}(0)k+\frac{1}{2}\varphi^{\prime\prime}(0)k^{2}+.... (49)

By switching the order of the summation and the integration, one can calculate the integration of every term,

1n!0e2ηtk2ϕ(n)(0)kp+n𝑑k\displaystyle\frac{1}{n!}\int_{0}^{\infty}e^{-2\eta tk^{2}}\phi^{(n)}(0)k^{p+n}dk t(1+p+n)/2,\displaystyle\propto t^{-(1+p+n)/2},
1n!0e2ηtk2φ(n)(0)k1+n𝑑k\displaystyle\frac{1}{n!}\int_{0}^{\infty}e^{-2\eta tk^{2}}\varphi^{(n)}(0)k^{1+n}dk t(2+n)/2,\displaystyle\propto t^{-(2+n)/2}, (50)

where p=0,2p=0,2 for Q,WQ,W respectively. One can get rid of the term t(1+p+n)/2t^{-(1+p+n)/2} for n>in>i when tt goes to infinity, where ii is the the derivative index denote for a lowest order ii-th (i0i\geq 0) derivative that makes ϕ(i)(0)\phi^{(i)}(0) non-vanishing.

Similarly, we obtains the final result

limtWQH21[g+(i)(0)+g(i)(0)g+(i)(0)g(i)(0)]2Γ(3+i2)Γ(1+i2)Γ(2+i2)Γ(2+i2)1.\lim_{t\rightarrow\infty}\frac{WQ}{H^{2}}-1\rightarrow\left[\frac{g_{+}^{(i)}(0)+g_{-}^{(i)}(0)}{g_{+}^{(i)}(0)-g_{-}^{(i)}(0)}\right]^{2}\frac{\Gamma\left(\frac{3+i}{2}\right)\Gamma\left(\frac{1+i}{2}\right)}{\Gamma\left(\frac{2+i}{2}\right)\Gamma\left(\frac{2+i}{2}\right)}-1. (51)

One can prove that the first factor in the right-hand-side of Eq. (51) is always larger than or equal to 1, and it is 1 if and only if either g+(i)(0)=0g_{+}^{(i)}(0)=0 or g(i)(0)=0g_{-}^{(i)}(0)=0. The second factor can be easily proved to be larger than 1. As a consequence, WQ/H2WQ/H^{2} is always larger than 1 and will not reach 11 as tt\rightarrow\infty, so the CKWT state cannot be reached in this case.

We see that the result for a>0a>0 cannot be simply extended to that for a=0a=0 by taking the limit a0a\rightarrow 0. The analytical result can be verified numerically as presented in Fig. 3.

Figure 3: Numerical results for tan2(θ)\tan^{2}(\theta) for different continuous functions as initial conditions. The parameters are: η=1\eta=1, g1+(0,k)=cos(kd)+1g_{1+}(0,k)=\cos(k-d)+1, g2+(0,k)=e(kd)g_{2+}(0,k)=e^{-(k-d)}, g3+(0,k)=f(k)ekg_{3+}(0,k)=f(k)e^{-k}, g1(0,k)=(kd)e(kd)g_{1-}(0,k)=(k-d)e^{-(k-d)}, g2(0,k)=|sin(kd)|g_{2-}(0,k)=|\sin(k-d)|, g3(0,k)=(k3/10)ek2g_{3-}(0,k)=(k^{3}/10)e^{-k^{2}}. (a) Case A: g±(0,k)=g1±(0,k)g_{\pm}(0,k)=g_{1\pm}(0,k), d=0.1d=0.1; Case B: g±(0,k)=g2±(0,k)g_{\pm}(0,k)=g_{2\pm}(0,k), d=0.1d=0.1; Case C: g±(0,k)=g1±(0,k)g_{\pm}(0,k)=g_{1\pm}(0,k), d=0d=0; Case D: g±(0,k)=g2±(0,k)g_{\pm}(0,k)=g_{2\pm}(0,k), d=0d=0. The integral range of them are all [d,)[d,\infty). In case A and B, the CKWT state can be reached as tt\rightarrow\infty. In other cases the CKWT state cannot be reached. (b) Case E: g±(0,k)=g3±(0,k)g_{\pm}(0,k)=g_{3\pm}(0,k), f(k)=1f(k)=1; Case F: g±(0,k)=g3±(0,k)g_{\pm}(0,k)=g_{3\pm}(0,k), f(k)=kf(k)=k; Case G: g±(0,k)=g3±(0,k)g_{\pm}(0,k)=g_{3\pm}(0,k), f(k)=k2/2f(k)=k^{2}/2. The integral range of them are all [0,)[0,\infty). The CKWT state cannot be reached for all cases of (b). The black dashed lines in figures show the analytical results for each curve with l1(0)=0l1(0)=0, l2(0)=Γ(3/2)Γ(1/2)/(Γ(1)Γ(1))10.571l2(0)=\Gamma(3/2)\Gamma(1/2)/(\Gamma(1)\Gamma(1))-1\approx 0.571, l2(1)=Γ(2)Γ(1)/(Γ(3/2)Γ(3/2))10.273l2(1)=\Gamma(2)\Gamma(1)/(\Gamma(3/2)\Gamma(3/2))-1\approx 0.273 and l2(2)=Γ(5/2)Γ(3/2)/(Γ(2)Γ(2))10.178l2(2)=\Gamma(5/2)\Gamma(3/2)/(\Gamma(2)\Gamma(2))-1\approx 0.178.
Refer to caption

V.3 Special non-analytic function

For non-analytic functions as initial conditions, it is difficult to reach a similar conclusion as in previous sections. We can only take an example and carry out our numerical calculations. We consider the following function as the initial condition

g+(0,k)={e1/k2,k>00,k=0.g_{+}(0,k)=\begin{cases}e^{-1/k^{2}},&k>0\\ 0,&k=0\end{cases}. (52)

What is special for this function is that it has infinite order of derivatives at k=0k=0 which are vanishing. Thus g+(0,k)g_{+}(0,k) is non-analytic at k=0k=0 and cannot be expanded into a Taylor series because zero is the essential singularity in the complex domain. For convenience, we assume g(0,k)=0g_{-}(0,k)=0 and the integration range is [0,)[0,\infty), then we can calculate WQ/H2WQ/H^{2} directly and find the long time limit with the second kind modified Bessel function Kν(z)K_{\nu}(z)

limtWQH21=limte42ηtπ(1+22ηt)16ηt[K1(22ηt)]21\displaystyle\lim_{t\rightarrow\infty}\frac{WQ}{H^{2}}-1=\lim_{t\rightarrow\infty}\frac{e^{-4\sqrt{2\eta t}}\pi\left(1+2\sqrt{2\eta t}\right)}{16\eta t\left[K_{1}(2\sqrt{2\eta t})\right]^{2}}-1 \displaystyle\rightarrow 0.\displaystyle 0. (53)

We see that under this condition the CKWT state can be reached.

VI Conclusion

We have studied how the Chandrasekhar-Kendall-Woltjer-Taylor (CKWT) state can be reached in the time evolution of a resistive plasma. We propose a criterion for the CKWT state as the destination of the time evolution, limttan2(θ)0\lim_{t\rightarrow\infty}\tan^{2}(\theta)\rightarrow 0, where θ\theta is the average angle between the magnetic field and the vector potential. We find that the initial conditions for the helicity amplitudes of the magnetic field and the vector potential are essential to the CKWT state. Our analysis is based on an expansion in the vector spherical harmonics for magnetic fields and vector potentials.

The asymptotic form of tan2(θ)\tan^{2}(\theta) is dominated by the lowest momentum kmink_{\mathrm{min}} of the initial helicity amplitudes g±(0,k)g_{\pm}(0,k) as functions of the scalar momentum kk. For those initial helicity amplitudes that can be expanded into a Taylor series, the CKWT state cannot be reached if kmin=0k_{\mathrm{min}}=0, while it can be reached for kmin>0k_{\mathrm{min}}>0 if and only if either g+(i)(0,kmin)=0g_{+}^{(i)}(0,k_{\mathrm{min}})=0 or g(i)(0,kmin)=0g_{-}^{(i)}(0,k_{\mathrm{min}})=0 with the lowest ii-th non-zero derivative (i0i\geq 0) explained in Section V. In other words, the CKWT state can be reached if one helicity is favored over the other at the lowest momentum in the initial helicity amplitudes of the magnetic field. This indicates that the imbalance between two helicities at the lowest momentum (longest wavelength) in the initial helicity amplitudes is the key factor for the CKWT state.

Acknowledgments. Z.Y.Z., Y.G.Y. and Q.W. are supported in part by National Natural Science Foundation of China (NSFC) under Grant Nos. 12135011, 11890713 (a subgrant of 11890710) and 12047502, and by the Strategic Priority Research Program of Chinese Academy of Sciences under Grant No. XDB34030102.

References

Appendix A Proof of Eq. (48)

Using the asymptotic series of the incomplete gamma function Γ(s,z)\Gamma(s,z) when zz\rightarrow\infty we obtain

akpe2k2ηt𝑑k\displaystyle\int_{a}^{\infty}k^{p}e^{-2k^{2}\eta t}dk =\displaystyle= 12(2ηt)(p+1)/2Γ(1+p2,2a2ηt)\displaystyle\frac{1}{2}(2\eta t)^{-(p+1)/2}\Gamma\left(\frac{1+p}{2},2a^{2}\eta t\right) (54)
\displaystyle\sim e2a2ηtj=1ap(2j1)(4ηt)jh=1j1[p(2h1)],\displaystyle e^{-2a^{2}\eta t}\sum_{j=1}^{\infty}\frac{a^{p-(2j-1)}}{(4\eta t)^{j}}\prod_{h=1}^{j-1}\left[p-\left(2h-1\right)\right],

where the index jj is called the integral approximation order. The integration of the ii-th order term of Taylor expansion of f(k)f(k) is evaluated as

Xi\displaystyle X_{i} =akn[1i!(ka)if(i)(a)]e2k2ηt𝑑k\displaystyle=\int_{a}^{\infty}k^{n}\left[\frac{1}{i!}(k-a)^{i}f^{(i)}(a)\right]e^{-2k^{2}\eta t}dk
=1i!f(i)(a)m=0iCim(akn+me2k2ηt𝑑k)aim(1)im,\displaystyle=\frac{1}{i!}f^{(i)}(a)\sum_{m=0}^{i}C_{i}^{m}\left(\int_{a}^{\infty}k^{n+m}e^{-2k^{2}\eta t}dk\right)a^{i-m}(-1)^{i-m}, (55)

where n=0,1,2n=0,1,2 for Q,H,WQ,H,W respectively.

Now we consider the jj-th order term in XiX_{i} with the expansion of akn+me2k2ηt𝑑k\int_{a}^{\infty}k^{n+m}e^{-2k^{2}\eta t}dk following Eq. (54),

Xij\displaystyle X_{ij} =\displaystyle= 1i!f(i)(a)m=0iCime2a2ηtan+m(2j1)(4ηt)j\displaystyle\frac{1}{i!}f^{(i)}(a)\sum_{m=0}^{i}C_{i}^{m}e^{-2a^{2}\eta t}\frac{a^{n+m-(2j-1)}}{(4\eta t)^{j}} (56)
×h=1j1[n+m(2h1)]aim(1)im\displaystyle\times\prod_{h=1}^{j-1}\left[n+m-(2h-1)\right]a^{i-m}(-1)^{i-m}
=\displaystyle= 1i!f(i)(a)e2a2ηtan+i(2j1)(4ηt)j\displaystyle\frac{1}{i!}f^{(i)}(a)e^{-2a^{2}\eta t}\frac{a^{n+i-(2j-1)}}{(4\eta t)^{j}}
×m=0iCim(1)im(h=1j1[n+m(2h1)])\displaystyle\times\sum_{m=0}^{i}C_{i}^{m}(-1)^{i-m}\left(\prod_{h=1}^{j-1}\left[n+m-(2h-1)\right]\right)
=\displaystyle= 1i!f(i)(a)e2a2ηtan+i(2j1)(4ηt)jm=0j1i!S(m,i)qm(j,n).\displaystyle\frac{1}{i!}f^{(i)}(a)e^{-2a^{2}\eta t}\frac{a^{n+i-(2j-1)}}{(4\eta t)^{j}}\sum_{m=0}^{j-1}i!S(m,i)q_{m}(j,n).

Here we have used the second kind Stirling number S(m,i)S(m,i), with the function qq being defined as

h=1j1[n+m(2h1)]=i=0j1qi(j,n)mi.\prod_{h=1}^{j-1}\left[n+m-(2h-1)\right]=\sum_{i^{\prime}=0}^{j-1}q_{i^{\prime}}(j,n)m^{i^{\prime}}. (57)

Since S(m,i)=0S(m,i)=0 for m<im<i, the lowest order term of 1/t1/t among XijX_{ij} must require j=i+1j=i+1 with the lowest ii-th non-zero derivative (i0i\geq 0), which is

f(i)(a)e2a2ηtani1(4ηt)i+1.f^{(i)}(a)e^{-2a^{2}\eta t}\frac{a^{n-i-1}}{(4\eta t)^{i+1}}. (58)

And then as tt goes to infinity, the leading term of WQ/H2WQ/H^{2} becomes

WQH2\displaystyle\frac{WQ}{H^{2}} ϕ(i)(a)e2a2ηta1i(4ηt)i+1ϕ(i)(a)e2a2ηta1i(4ηt)i+1φ(i)(a)e2a2ηtai(4ηt)i+1φ(i)(a)e2a2ηtai(4ηt)i+1\displaystyle\sim\frac{\phi^{(i)}(a)e^{-2a^{2}\eta t}\frac{a^{1-i}}{(4\eta t)^{i+1}}\phi^{(i)}(a)e^{-2a^{2}\eta t}\frac{a^{-1-i}}{(4\eta t)^{i+1}}}{\varphi^{(i)}(a)e^{-2a^{2}\eta t}\frac{a^{-i}}{(4\eta t)^{i+1}}\varphi^{(i)}(a)e^{-2a^{2}\eta t}\frac{a^{-i}}{(4\eta t)^{i+1}}}
=[ϕ(i)(a)φ(i)(a)]2,\displaystyle=\left[\frac{\phi^{(i)}(a)}{\varphi^{(i)}(a)}\right]^{2}, (59)

where ϕ(i)(a)\phi^{(i)}(a) and φ(i)(a)\varphi^{(i)}(a) are the same ii-th order derivatives of ϕ(k)\phi(k) and φ(k)\varphi(k) defined in Eq. (46) at k=ak=a.