Hostname: page-component-7bb8b95d7b-pwrkn Total loading time: 0 Render date: 2024-09-28T21:23:11.773Z Has data issue: false hasContentIssue false

Diffusive scattering of energetic electrons by intense whistler-mode waves in an inhomogeneous plasma

Published online by Cambridge University Press:  06 January 2023

Viktor A. Frantsuzov*
Affiliation:
Space Research Institute of the Russian Academy of Sciences (IKI), 84/32 Profsoyuznaya Str., Moscow 117997, Russia Faculty of Physics, National Research University Higher School of Economics, 21/4 Staraya Basmannaya Ulitsa, Moscow 105066, Russia
Anton V. Artemyev
Affiliation:
Space Research Institute of the Russian Academy of Sciences (IKI), 84/32 Profsoyuznaya Str., Moscow 117997, Russia Institute of Geophysics and Planetary Physics, UCLA, Los Angeles, CA 90095, USA
Xiao-Jia Zhang
Affiliation:
Institute of Geophysics and Planetary Physics, UCLA, Los Angeles, CA 90095, USA
Oliver Allanson
Affiliation:
Department of Mathematics, University of Exeter, Penryn/Cornwall Campus, Penryn, TR10 9FE, UK
Pavel I. Shustov
Affiliation:
Space Research Institute of the Russian Academy of Sciences (IKI), 84/32 Profsoyuznaya Str., Moscow 117997, Russia Faculty of Physics, National Research University Higher School of Economics, 21/4 Staraya Basmannaya Ulitsa, Moscow 105066, Russia
Anatoli A. Petrukovich
Affiliation:
Space Research Institute of the Russian Academy of Sciences (IKI), 84/32 Profsoyuznaya Str., Moscow 117997, Russia
*
Email address for correspondence: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

Electron resonant interactions with electromagnetic whistler-mode waves play an important role in electron flux dynamics in various space plasma systems: planetary radiation belts, bow shocks, solar wind and magnetic reconnection regions. Two key wave characteristics determining the regime of wave–particle interactions are the wave intensity and the wave coherency. The classical quasi-linear diffusion approach describes well electron diffusion by incoherent and low-amplitude waves, whereas the nonlinear resonant models describe electron phase bunching and trapping by highly coherent intense waves. This study is devoted to the investigation of the regime of electron resonant interactions with incoherent but intense waves. Although this regime is characterized by electron diffusion, we show that diffusion rates scale linearly with the wave amplitude, $D\propto B_w$, in contrast to the quasi-linear diffusion scaling $D_{QL}\propto B_w^2$. Using observed wave amplitude distributions, we demonstrate that the quasi-linear diffusion model significantly overestimates electron scattering by incoherent, but intense whistler-mode waves. We discuss the results obtained in the context of simulations of long-term electron flux dynamics in space plasma systems.

Type
Research Article
Copyright
Copyright © The Author(s), 2023. Published by Cambridge University Press

1. Introduction

Resonant electron interactions with whistler-mode waves are one of the main drivers of electron pitch-angle scattering and acceleration in various space plasma systems, e.g. solar flares (Bespalov, Zaitsev & Stepanov Reference Bespalov, Zaitsev and Stepanov1991; Filatov & Melnikov Reference Filatov and Melnikov2017; Melnikov & Filatov Reference Melnikov and Filatov2020), solar wind (Tong et al. Reference Tong, Vasko, Artemyev, Bale and Mozer2019; Cattell et al. Reference Cattell, Short, Breneman and Grul2020, Reference Cattell, Short, Breneman, Halekas, Whittesley, Larson, Kasper, Stevens, Case and Moncuquet2021; Mozer et al. Reference Mozer, Bonnell, Halekas, Rahmati, Schum and Vasko2021), shock waves (Hull et al. Reference Hull, Muschietti, Oka, Larson, Mozer, Chaston, Bonnell and Hospodarsky2012; Wilson et al. Reference Wilson, Koval, Szabo, Breneman, Cattell, Goetz, Kellogg, Kersten, Kasper and Maruca2013; Oka et al. Reference Oka, Wilson III, Phan, Hull, Amano, Hoshino, Argall, Le Contel, Agapitov and Gershman2017; Page et al. Reference Page, Vasko, Artemyev and Bale2021), planetary radiation belts (Li et al. Reference Li, Ma, Shen, Zhang, Mauk, Clark, Allegrini, Kurth, Hospodarsky and Hue2021; Menietti et al. Reference Menietti, Averkamp, Kurth, Imai, Faden, Hospodarsky, Santolik, Clark, Allegrini and Elliott2021; Thorne et al. Reference Thorne, Bortnik, Li and Ma2021) and magnetic reconnection regions (Le Contel et al. Reference Le Contel, Roux, Jacquey, Robert, Berthomier, Chust, Grison, Angelopoulos, Sibeck and Chaston2009; Breuillard et al. Reference Breuillard, Le Contel, Retino, Chasapis, Chust, Mirioni, Graham, Wilder, Cohen and Vaivads2016; Zhang et al. Reference Zhang, Angelopoulos, Artemyev and Liu2018a). The basic theoretical framework for description of such interactions is the quasi-linear model (Drummond & Pines Reference Drummond and Pines1962; Vedenov, Velikhov & Sagdeev Reference Vedenov, Velikhov and Sagdeev1962; Andronov & Trakhtengerts Reference Andronov and Trakhtengerts1964; Kennel & Engelmann Reference Kennel and Engelmann1966) that is based on the assumption of weak perturbation of particle dynamics by each single resonance. This assumption reduces the Vlasov equation to the Fokker–Planck diffusion equation (Drummond & Pines Reference Drummond and Pines1962; Vedenov et al. Reference Vedenov, Velikhov and Sagdeev1962) where the main characteristics of wave–particle resonant interactions are diffusion rates. The requirement of a weak perturbation of particle trajectories for a single resonance is equivalent to the requirement that each individual wave–particle resonant interaction should not last for a long time (so particle energy/pitch-angle change for a single resonance is sufficiently small), and there are several mechanisms responsible for particle escape from the resonance.

The original quasi-linear diffusion model assumes the broad spectrum of waves resonating with charged particles (Drummond & Pines Reference Drummond and Pines1962; Vedenov et al. Reference Vedenov, Velikhov and Sagdeev1962), when the resonance width in velocity space $\Delta v_R$ is equal to the difference of resonance $v_R$ velocity (determined for the cyclotron resonant conditions) and wave group velocity $v_g=\partial \omega /\partial k$ (where $\omega$ and $k$ are wave frequency and wavenumber). The estimate for the resonance width can be derived from the condition that a change of the resonant particle velocity, $\Delta v_R \sim |v_R-v_g|\Delta k/k$, will remove the particle from the cyclotron resonance (Karpman Reference Karpman1974). The small factor $\Delta k/k$ is determined by the wave spectrum width in wavenumber space, $\Delta k$. This mechanism determines the shortness of an individual resonance and justifies the applicability of the diffusion approximation for modelling the dynamics of the charged particle ensemble (Karpman Reference Karpman1974; Le Queau & Roux Reference Le Queau and Roux1987; Shapiro & Sagdeev Reference Shapiro and Sagdeev1997). This description works well for low-amplitude whistler-mode waves resonating with electrons in homogeneous systems (without spatial gradients of the background plasma and magnetic field), e.g. in the solar wind (see review by Verscharen et al. (Reference Verscharen, Chandran, Boella, Halekas, Innocenti, Jagarlamudi, Micera, Pierrard, Štverák and Vasko2022) and references therein).

The assumption of background magnetic field homogeneity, however, does not work for many space plasma systems. Resonant electron scattering by whistler-mode waves is often observed in magnetic field traps, regions with a spatially localized minimum of the magnetic field magnitude, where charged particles can be trapped and bouncing. Important examples of such traps are the radiation belt dipole field (Lyons & Williams Reference Lyons and Williams1984; Schulz & Lanzerotti Reference Schulz and Lanzerotti1974) and magnetic holes generated by compressional perturbations on a bow shock (Oka et al. Reference Oka, Otsuka, Matsukiyo, Wilson, Argall, Amano, Phan, Hoshino, Le Contel and Gershman2019; Hull et al. Reference Hull, Muschietti, Le Contel, Dorelli and Lindqvist2020; Yao et al. Reference Yao, Shi, Zong, Degeling, Guo, Li, Li, Tian, Zhang and Yao2021). Bouncing within magnetic traps, electrons periodically resonate with whistler-mode waves, and the resonance width for cyclotron resonance in an inhomogeneous field is determined from the condition that a change of the resonant particle velocity (due to the field spatial gradient) $\Delta v_R \sim |\partial v_R/\partial s|/k$ will remove the particle from the cyclotron resonance (Trakhtengerts & Rycroft Reference Trakhtengerts and Rycroft2008). If $\Delta v_R$ is finite, the quasi-linear diffusion model works even for monochromatic waves ($\Delta k \to 0$) resonating with electrons in magnetic traps (Karpman & Shklyar Reference Karpman and Shklyar1977; Albert Reference Albert2001, Reference Albert2010; Shklyar Reference Shklyar2021). Thus, the only condition required for the application of the quasi-linear diffusion model is that the mirror force due to the background magnetic field gradient should be stronger than the Lorentz force of the wave field (Karpman Reference Karpman1974).

The small wave intensity approximation, however, is often violated for whistler-mode waves observed in the highly unstable plasma of shock waves (Zhang et al. Reference Zhang, Matsumoto, Kojima and Omura1999; Artemyev et al. Reference Artemyev, Shi, Liu, Zhang, Vasko and Angelopoulos2022) and plasma injections (Zhang et al. Reference Zhang, Thorne, Artemyev, Mourenas, Angelopoulos, Bortnik, Kletzing, Kurth and Hospodarsky2018b, Reference Zhang, Mourenas, Artemyev, Angelopoulos, Bortnik, Thorne, Kurth, Kletzing and Hospodarsky2019). Such intense waves may resonate with electrons in a nonlinear regime, including effects of phase trapping and phase bunching (e.g. Nunn Reference Nunn1971, Reference Nunn1974; Karpman, Istomin & Shklyar Reference Karpman, Istomin and Shklyar1974; Inan, Bell & Helliwell Reference Inan, Bell and Helliwell1978). Although phase bunching is a strongly nonlinear effect (Albert Reference Albert1993; Bortnik, Thorne & Inan Reference Bortnik, Thorne and Inan2008), due to the smallness of the electron energy and pitch-angle changes in a single resonant phase bunching, it can be incorporated as a drift term into the Fokker–Planck equation (see discussion in Artemyev et al. (Reference Artemyev, Vasiliev, Mourenas, Agapitov and Krasnoselskikh2014), Gan et al. (Reference Gan, Li, Ma, Albert, Artemyev and Bortnik2020) and Allanson et al. (Reference Allanson, Watt, Allison and Ratcliffe2021)). Changes of electron energy and pitch angle due to phase trapping are comparable with the initial energies/pitch angles (Omura, Furuya & Summers Reference Omura, Furuya and Summers2007; Summers & Omura Reference Summers and Omura2007), and thus it is not clear how to include this effect into the Fokker–Planck equation. Several approaches with different integral operators describing the phase trapping contribution to the electron flux dynamics have been proposed (e.g. Omura et al. Reference Omura, Miyashita, Yoshikawa, Summers, Hikishima, Ebihara and Kubota2015; Artemyev et al. Reference Artemyev, Neishtadt, Vasiliev and Mourenas2018b; Vainchtein et al. Reference Vainchtein, Zhang, Artemyev, Mourenas, Angelopoulos and Thorne2018; Hsieh, Kubota & Omura Reference Hsieh, Kubota and Omura2020), but the evaluation of such operators is computationally expensive and significantly changes the Fokker–Planck equation. Thus, it is important and practically useful to propose an approach for incorporation of nonlinear effects without significantly altering models based on the Fokker–Planck equation.

The principal possibility for such an approach has been proposed in Solovev & Shkliar (Reference Solovev and Shkliar1986): namely, the total contribution of trapping and bunching may compensate each other. This idea has been reinvestigated in Mourenas et al. (Reference Mourenas, Zhang, Artemyev, Angelopoulos, Thorne, Bortnik, Neishtadt and Vasiliev2018), where effects of wave modulations were taken into account. Spacecraft observations (e.g. Oka et al. Reference Oka, Otsuka, Matsukiyo, Wilson, Argall, Amano, Phan, Hoshino, Le Contel and Gershman2019; Zhang et al. Reference Zhang, Mourenas, Artemyev, Angelopoulos, Bortnik, Thorne, Kurth, Kletzing and Hospodarsky2019, Reference Zhang, Mourenas, Artemyev, Angelopoulos, Kurth, Kletzing and Hospodarsky2020b; Foster, Erickson & Omura Reference Foster, Erickson and Omura2021; Artemyev et al. Reference Artemyev, Shi, Liu, Zhang, Vasko and Angelopoulos2022) and numerical simulations (e.g. Nunn & Omura Reference Nunn and Omura2012; Katoh & Omura Reference Katoh and Omura2016; Demekhov, Taubenschuss & Santolık Reference Demekhov, Taubenschuss and Santolık2017; Tao et al. Reference Tao, Zonca, Chen and Wu2020; Zhang et al. Reference Zhang, Demekhov, Katoh, Nunn, Tao, Mourenas, Omura, Artemyev and Angelopoulos2021) show that intense whistler-mode waves mostly propagate in a form of short, modulated wave packets. Typical wave packets include only a few wave periods (see figure 1), which can be an effect of sideband instability of wave generation (Nunn Reference Nunn1986) or overlapping of several waves with close wave frequencies (Zhang et al. Reference Zhang, Mourenas, Artemyev, Angelopoulos, Kurth, Kletzing and Hospodarsky2020b; Nunn et al. Reference Nunn, Zhang, Mourenas and Artemyev2021). Such modulation reduces the efficiency of phase trapping (Tao et al. Reference Tao, Bortnik, Thorne, Albert and Li2012b, Reference Tao, Bortnik, Albert, Thorne and Li2013), and can make the net effect of electron resonant interactions with waves more diffusive (Zhang et al. Reference Zhang, Agapitov, Artemyev, Mourenas, Angelopoulos, Kurth, Bonnell and Hospodarsky2020a; Allanson et al. Reference Allanson, Watt, Ratcliffe, Allison, Meredith, Bentley, Ross and Glauert2020, Reference Allanson, Watt, Allison and Ratcliffe2021; Gan et al. Reference Gan, Li, Ma, Albert, Artemyev and Bortnik2020; An, Wu & Tao Reference An, Wu and Tao2022; Mourenas et al. Reference Mourenas, Zhang, Nunn, Artemyev, Angelopoulos, Tsai and Wilkins2022). Thus, the derivation of diffusion rates is the main question for theoretical description of such regime of wave–particle interaction.

Figure 1. Examples of typical wave packets of whistler-mode waves captured by THEMIS spacecraft (Angelopoulos Reference Angelopoulos2008) in the Earth bow shock (a), foreshock transient (b), outer radiation belt (c) and plasma injection region (d). These events are picked up from statistics published in Artemyev et al. (Reference Artemyev, Shi, Liu, Zhang, Vasko and Angelopoulos2022), Shi et al. (Reference Shi, Liu, Angelopoulos and Zhang2020), Zhang et al. (Reference Zhang, Thorne, Artemyev, Mourenas, Angelopoulos, Bortnik, Kletzing, Kurth and Hospodarsky2018b) and Zhang et al. (Reference Zhang, Angelopoulos, Artemyev and Liu2018a).

In this paper, we propose an approach for the evaluation of diffusion rates including nonlinear effects for intense, but strongly modulated waves. First, in § 2 we describe the concept of the diffusion coefficient model. Then, in § 3 we provide the basic model equations for the diffusion rate and evaluate diffusion rates for arbitrary wave intensity. Finally, in § 4 we show the contribution of nonlinear effects to diffusion rates averaged over wave intensity distributions and discuss the obtained results.

2. Basic concept

To propose the approach for the evaluation of such diffusion rates, let us illustrate the wave modulation effect on nonlinear wave–particle interactions. We consider electrons bouncing in a magnetic trap modelled by a curvature-free dipole field (Bell Reference Bell1984) and their resonant interaction with a monochromatic and intense whistler-mode wave. To evaluate a set of test particle trajectories resonating once with whistler-mode waves, we use the approximation of a monochromatic field-aligned wave. The wave field distribution along the magnetic field lines and the concept of description of wave packets are taken from Mourenas et al. (Reference Mourenas, Zhang, Nunn, Artemyev, Angelopoulos, Tsai and Wilkins2022). We start with the Hamiltonian of a relativistic electron (rest mass is $m_e$ and charge is $-e$) bouncing in a magnetic trap and interacting with a field-aligned whistler-mode wave:

(2.1)\begin{equation} H = \sqrt{\left({\boldsymbol p} + \frac{e}{c} {\boldsymbol A}\right)^2 c^2 + m^2 c^4}, \end{equation}

where ${\boldsymbol p}$ is a canonical momentum and ${\boldsymbol A}$ is a vector potential that can be derived from ${\boldsymbol B} = \boldsymbol {\nabla }\times {\boldsymbol A}$ with ${\boldsymbol B} = {\boldsymbol B}_0 + {\boldsymbol B}_w$. Here ${\boldsymbol B}_0$ is the background magnetic field of Earth's dipole and ${\boldsymbol B}_w$ describes the wave field. As the electron gyroradius is significantly smaller than the dipole magnetic field line curvature, $\sim LR_E$ where $R_E$ is Earth's radius and $L$-shell, we consider a curvature-free magnetic field with a pair $(z, p_z)$ of Cartesian coordinates and momentum playing a role of field-aligned coordinate and momentum $(s,p_{\parallel })$ (see Bell Reference Bell1984). The equatorial magnetic field is determined by a dipole model, $B_0(0)\propto L^{-3}$. To define a $B_0(z)$ function, we introduce a geomagnetic latitude $\lambda$:

(2.2a,b)\begin{equation} \frac{{\rm d}z}{{\rm d}\lambda} = \sqrt{1+3 \sin^2\lambda} \cos\lambda,\quad \frac{B_0(\lambda)}{B_0(0)} = \frac{\sqrt{1+3 \sin^2\lambda}}{\cos^6\lambda}. \end{equation}

The smallness of electron gyroradius allows us to write ${\boldsymbol A}_0=x\tilde {B}_0(z) {\boldsymbol e}_y$ (${\boldsymbol B}_0 = \boldsymbol {\nabla }\times {\boldsymbol A}_0$), where ${\boldsymbol e}_{r_i}$ is the unit vector along the $r_i$ axis, $r_i = (x, y, z)$, and $x$ is the cross-field coordinate. As the magnetic field ${\boldsymbol B}_0$ is mainly oriented along the $z$ axis, we use the approximation $\tilde {B}_0(z) \approx B_0(z)$. To derive the equation for the wave vector potential ${\boldsymbol A}_w$, we introduce the wave phase $\phi$ and define the magnetic vector ${\boldsymbol B}_w$ as $B_w {\boldsymbol e}_x \cos \phi - B_w {\boldsymbol e}_y \sin \phi$. Note that $B_w$ may be a function of $z$ (or/and $\phi$), but for the following derivations we assume that $|{\rm d}B_w/{\rm d}z| \ll |(\partial \phi /\partial z)B_w|$. The wave frequency $\omega$ and the wave vector ${\boldsymbol k}$ are defined by equations $\omega = -\partial \phi /\partial t$ and ${\boldsymbol k} = \boldsymbol {\nabla } \phi \approx (\partial \phi /\partial z) {\boldsymbol e}_z$ for field-aligned waves. We consider $\omega ={\rm const.}$ and determine ${\boldsymbol k}$ from a cold plasma dispersion: $k c / \varOmega _{pe} = ( \varOmega _{ce}/\omega - 1)^{-{1}/{2}}$, where $\varOmega _{pe} = \sqrt {4 {\rm \pi}n_e e^2 / m_e}$ is the plasma frequency and $\varOmega _{ce} = e B_0 / m_e c$ is the cyclotron frequency (Stix Reference Stix1962). Therefore, ${\boldsymbol A}_w\approx B_w {\boldsymbol e}_x \cos \phi / k - B_w {\boldsymbol e}_y \sin \phi / k$ and Hamiltonian (2.1) can be presented as

(2.3)\begin{align} H = m_e c^2 \sqrt{1 + \left( \frac{p_z}{m_e c} \right)^2 + \left( \frac{p_x}{m_e c} + \frac{\varOmega_{ce}}{kc} \frac{B_w}{B_0} \cos\phi \right)^2 + \left( \frac{p_y}{m_e c} + \frac{x\varOmega_{ce}}{c} - \frac{\varOmega_{ce}}{kc} \frac{B_w}{B_0} \sin\phi \right)^2}. \end{align}

This Hamiltonian does not depend on $y$, and thus canonical momentum $p_y$ is a constant: $\dot {p}_y = -\partial H/\partial y = 0$. Without loss of generality, we can set $p_y = 0$.

In the absence of a wave, this Hamiltonian describes fast $(x,p_x)$ oscillations and slow $(z,p_z)$ oscillations. Thus, we can introduce an adiabatic invariant $I_x=(2{\rm \pi} )^{-1}\oint \,{{{\rm d} x} p_x}$ as an area surrounded by a closed trajectory in the $(x,p_x)$ plane (Landau & Lifshitz Reference Landau and Lifshitz1988):

(2.4)\begin{equation} I_x = \frac{1}{2 {\rm \pi}} \oint\,{{{\rm d} x} p_x} = \frac{m_e c^2}{2 \varOmega_{ce}} \left( \frac{H^2}{m_e^2 c^4} - 1 - \left( \frac{p_z}{m_e c} \right)^2 \right) \end{equation}

with

(2.5a,b)\begin{equation} H = m_e c^2 \sqrt{1 + \left( \frac{p_z}{m_e c} \right)^2 + \frac{2 I_x \varOmega_{ce}}{m_e c^2}},\quad \frac{2 I_x \varOmega_{ce}}{m_e c^2} = \left(\frac{p_x}{m_e c}\right)^2 + \left(\frac{x\varOmega_{ce}}{c}\right)^2. \end{equation}

We consider the canonical transformation $(x, p_x) \to (\psi, I_x)$ given by a generating function $F_2(x, I_x)=(2{\rm \pi} )^{-1}\int \,{{{\rm d} x} p_x}$ (Landau & Lifshitz Reference Landau and Lifshitz1988):

(2.6)\begin{gather} F_2(x, I_x) ={\pm} \frac{ m_e }{2 {\rm \pi}}\int \,{{\rm d} x} \sqrt{\frac{2 I_x \varOmega_{ce}}{m_e} - x^2 \varOmega_{ce}^2} ={\pm} I_x \left[ \sqrt{\frac{m_e \varOmega_{ce}}{2 I_x}} x \sqrt{1 - \frac{m_e \varOmega_{ce}}{2 I_x} x^2} \right.\nonumber\\ + \left. {\rm arcsin}\left( \sqrt{\frac{m_e \varOmega_{ce}}{2 I_x}} x \right) \right]. \end{gather}

The corresponding variable transformations are

(2.7a,b)\begin{equation} p_x ={\pm} m_e \sqrt{\frac{2 I_x \varOmega_{ce}}{m_e} - x^2 \varOmega_{ce}^2},\quad \psi ={\pm} {\rm arcsin}\left( \sqrt{\frac{m_e \varOmega_{ce}}{2 I_x}} x \right). \end{equation}

Thus, equations for $x$ and $p_x$ are

(2.8a,b)\begin{equation} x = \frac{c}{\varOmega_{ce}} \sqrt{\frac{2 I_x \varOmega_{ce}}{m_e c^2}} \sin\psi,\quad p_x = m_e c \sqrt{\frac{2 I_x \varOmega_{ce}}{m_e c^2}} \cos\psi. \end{equation}

A new equation for $\varGamma = H / m_e c^2$ in terms of variables $(z, p_z)$ and $(\psi, I_x)$ can be written as

(2.9)\begin{equation} \varGamma = \sqrt{1 + \left( \frac{p_z}{m_e c} \right)^2 + \frac{2 I_x \varOmega_{ce}}{m_e c^2} + 2 \sqrt{\frac{2 I_x \varOmega_{ce}}{m_e c^2}} \frac{\varOmega_{ce}}{k c} \frac{B_w}{B_0} \cos(\phi + \psi) + \left( \frac{\varOmega_{ce}}{k c} \frac{B_w}{B_0} \right)^2}. \end{equation}

The Hamiltonian $H=m_ec^2\varGamma$ describes the dynamics of two pairs of conjugated variables, $(z, p_z)$ and $(\psi, I_x)$. The system of Hamiltonian equations can be solved with respect to these variables and time $t$. The general approach consists of expanding $H$ over $B_w/B_0$, keeping only the linear $\sim B_w/B_0$ term:

(2.10a,b)\begin{align} H = m_ec^2\gamma + m_e c^2\sqrt{\frac{2 I_x \varOmega_{ce}}{m_e c^2}} \frac{\varOmega_{ce}}{\gamma k c} \frac{B_w}{B_0} \cos(\phi + \psi),\quad \gamma=\sqrt{1 + \left( \frac{p_z}{m_e c} \right)^2 + \frac{2 I_x \varOmega_{ce}}{m_e c^2}}. \end{align}

We introduce the wave modulation through the $B_w$ dependence on the wave phase $\phi$ ($B_w(\phi )$ periodicity mimics effect if multiple wave packets):

(2.11)\begin{equation} B_w = B_{m} \frac{1 - e^{{-}h \sin^2(\phi / 2l)}}{1 - e^{{-}h}}, \end{equation}

where $B_m$ is the peak wave amplitude, $l$ defines the number of wave oscillations (periods) within one wave packet and $h$ controls the intensity of modulations. The effective wave amplitude $B_{{\rm eff}}$ can be determined by averaging $B_w$ over the period of modulations:

(2.12)\begin{equation} B_{{\rm eff}} = \sqrt{\left\langle B_w^2 \right\rangle}_{\phi\in[0,2{\rm \pi} l]}= B_{m} \frac{\sqrt{1 - 2 {\rm I}_0\left( h/2 \right) e^{ {-}h/2 } + {\rm I}_0\left( h \right) e^{ {-}h }}}{1 - e^{{-}h}}, \end{equation}

where ${\rm I}_n(z)$ is the modified Bessel function of the first kind. The average intensity of the plane wave with $B_w = B_{{\rm eff}} = {\rm const.}$ equals the intensity of the modulated wave with $B_w$ given by (2.11). Thus, to describe the electron diffusion theoretically, the approximation $B_w \approx B_{{\rm eff}} = {\rm const.}$ can be used. However, this assumption neglects nonlinear effects and, therefore, to determine whether the theory is applicable, we perform numerical simulations of modulated waves. Parameter $h$ in (2.11) determines the depth of the modulation: at $h \to \infty$ and $l\to \infty$ the wave packet reduces to a plane wave. Parameter $l$ determines the wave packet size with typical values $l\in [10,30]$ (see Zhang et al. Reference Zhang, Mourenas, Artemyev, Angelopoulos, Bortnik, Thorne, Kurth, Kletzing and Hospodarsky2019, Reference Zhang, Demekhov, Katoh, Nunn, Tao, Mourenas, Omura, Artemyev and Angelopoulos2021). For numerical simulations we use $h=1$ and $l=20$. Note that these parameters well satisfy the condition $|{\rm d}B_w/{\rm d}z| \ll |(\partial \phi /\partial z)B_w|$, because ${\rm d}B_w/{\rm d}z=(\partial \phi /\partial z) ({\rm d}B_w/{\rm d}\phi )$ and $({\rm d}B_w/{\rm d}\phi )B_w^{-1}\sim h/l \ll 1$.

We consider particles having the same initial energy $E_0$ and pitch angle $\alpha _0$, but random wave phase $\phi$ and gyrophase $\psi$. Figure 2 shows typical examples of electron resonance interactions with whistler-mode waves obtained by a numerical integration of Hamiltonian equations: diffusive electron scattering by low-amplitude wave (figure 2a), nonlinear resonant interactions with intense coherent wave (figure 2b) and nonlinear resonant interactions with well-modulated intense wave (figure 2c). The resonant interaction occurs once per simulation interval (half of the bounce period), i.e. there is only one point along the unperturbed particle trajectory where the resonant condition $\dot {\phi } + \dot {\psi } = 0$ is satisfied. The diffusive scattering is characterized by a symmetric (relative to zero) distribution of energy changes, and thus this process should be described by a diffusion rate $\sim \langle (\Delta E)^2\rangle$. The nonlinear resonances with an intense coherent wave are characterized by a small population changing the energy significantly ($\Delta E>0$, the phase trapping effect) and a large population with a small energy change, but almost identical for all particles ($\Delta E<0$, the phase bunching effect). Thus, nonlinear resonances with a coherent wave should be described separately for trapped and bunched particle populations, and due to the large energy change of the trapped population it is not thought to be possible to include this into the Fokker–Planck equation (see Hsieh & Omura Reference Hsieh and Omura2017a,Reference Hsieh and Omurab; Artemyev et al. Reference Artemyev, Neishtadt, Vasiliev, Zhang, Mourenas and Vainchtein2021b; Zhang et al. Reference Zhang, Artemyev, Angelopoulos, Tsai, Wilkins, Kasahara, Mourenas, Yokota, Keika and Hori2022). The nonlinear resonances with modulated waves are characterized by: (i) an increase of probability for the $\Delta E>0$ changes; (ii) but also with a decrease in size of $|\Delta E|$ itself; and (iii) further a randomization of energy change for this population. Thus, for modulated waves, the resonant wave–particle interaction is closer to diffusive scattering (Tao et al. Reference Tao, Bortnik, Albert, Thorne and Li2013; Zhang et al. Reference Zhang, Agapitov, Artemyev, Mourenas, Angelopoulos, Kurth, Bonnell and Hospodarsky2020a; An et al. Reference An, Wu and Tao2022; Mourenas et al. Reference Mourenas, Zhang, Nunn, Artemyev, Angelopoulos, Tsai and Wilkins2022).

Figure 2. A set of trajectories obtained by the numerical integration of Hamiltonian equations for a system (2.10a,b). Results are shown for (a) low-amplitude wave with diffusive scattering, (b) coherent high-amplitude wave with trapping and bunching and (c) modulated high-amplitude wave with diffusive-like scattering. System parameters are: electron energy $E_0=100$ keV, equatorial pitch angle $\alpha _0 = 40^\circ$, number of particles $N = 100$, wave amplitude $B_w = 5$ pT (a), $B_w = 500$ pT (b,c). System parameters correspond to $L\text {-shell} = 6$, whistler-mode waves with frequency equal to $0.35$ of the electron cyclotron frequency at the equator, and the constant plasma frequency equal to $10$ times the electron cyclotron frequency at the equator.

In Figure 3 we show how wave intensity and wave modulation control the efficiency of the nonlinear interactions. Figure 3(ac) shows distributions of energy changes $\Delta E$ depending on the normalized wave amplitude $\varepsilon =B_w/B_{0}(0)$. Below $\varepsilon \sim 10^{-4}$$10^{-3}$, the resonant interaction is diffusive with a symmetric $\Delta E$ distribution and $\langle {(\Delta E)^2}\rangle ^{1/2}$ linearly growing with $\varepsilon$. This dependence $\langle {(\Delta E)^2}\rangle ^{1/2}\propto \varepsilon$ demonstrates the applicability of the unperturbed trajectory approximation, a core assumption of the quasi-linear diffusion theory (Tao et al. Reference Tao, Bortnik, Albert, Liu and Thorne2011, Reference Tao, Bortnik, Albert and Thorne2012a; Allanson et al. Reference Allanson, Watt, Ratcliffe, Allison, Meredith, Bentley, Ross and Glauert2020). After the wave amplitude reaches a certain threshold (depending on the electron energy, pitch angle and system characteristics; see Omura et al. Reference Omura, Matsumoto, Nunn and Rycroft1991; Shklyar & Matsumoto Reference Shklyar and Matsumoto2009), the resonant interaction becomes nonlinear with a clear formation of a population of trapped electrons (a separate group of large positive $\Delta E$ in figure 3ac) and a highly asymmetric $\Delta E$ distribution (most of the electrons experience phase bunching and form a large maximum at $\Delta E<0$). For such nonlinear resonant interactions the $\Delta E$ distribution with phase-trapped and phase-bunched populations cannot be characterized by a diffusion $\langle {(\Delta E)^2}\rangle$ only, and thus it is not known how to include this regime of resonant interactions into the Fokker–Planck equation.

Figure 3. Distributions of the energy change $\Delta E$ for different $B_{{\rm eff}}$ with $h=1$ and $l=20$ for coherent (ac) and modulated (df) waves with $\Delta E \propto B_w$ fitting (black dashed lines) and with $\langle {(\Delta E)^2}\rangle ^{1/2}$ profiles (black dots): (a,d) $E_0 = 100$ keV, $\alpha _0 = 40^{\circ }$; (b,e) $E_0 = 100$ keV, $\alpha _0 = 60^{\circ }$; (cf) $E_0 = 300$ keV, $\alpha _0 = 60^{\circ }$. For each $B_w$ we use $10^4$ trajectories to evaluate the $\Delta E$ distribution, each particle resonates with the wave only once and $\Delta E$ is the energy change for a single resonance, initial particle phases and wave phases are random and thus for a modulated case the actual wave amplitude is different for different particles having the same energy/pitch angle. For each $B_{{\rm eff}}$ the $\Delta E$ distribution is normalized to one. System parameters correspond to $L\text {-shell} = 6$, whistler-mode waves with frequency equal to $0.35$ of the electron cyclotron frequency at the equator, and the constant plasma frequency equal to $10$ times the electron cyclotron frequency at the equator.

Figure 3(df) shows $\Delta E$ distributions as a function of wave amplitude for strongly modulated waves. For small wave intensity, $\varepsilon <10^{-4}$, there is the same diffusive regime of wave–particle resonant interactions as for non-modulated waves: a symmetric $\Delta E$ distribution with $\langle {(\Delta E)^2}\rangle ^{1/2}\propto \varepsilon$. With amplitude increasing, the regime of wave–particle resonant interaction changes. However, the wave modulation does not allow strong trapping acceleration (there is no population with large positive $\Delta E$), but increases the number of trapped particles (the populations of particles with $\Delta E>0$ and with $\Delta E<0$ are comparable even for $\varepsilon >10^{-3}$). Thus, for intense modulated waves, the $\Delta E$ distribution remains almost symmetric and can be characterized by $\langle {(\Delta E)^2}\rangle$. Despite such a symmetric $\Delta E$ distribution, the resonant interaction for large $\varepsilon$ is nonlinear, and the wave field alters electron dynamics in the resonance. This results in inapplicability of the unperturbed trajectory approximation, and thus $\langle {(\Delta E)^2}\rangle ^{1/2}\propto \varepsilon ^A$ with $A<1$. Therefore, for intense modulated waves we deal with diffusion of resonant electrons, but it is not a quasi-linear diffusion. In this study, we aim to derive the diffusion coefficient $\sim \langle {(\Delta E)^2}\rangle$ as a function of $\varepsilon$ for a wide $\varepsilon$ range. Figure 3(df) shows that $\langle {(\Delta E)^2}\rangle$ for small $\varepsilon$ should be similar to that for the quasi-linear diffusion (see also Albert Reference Albert2010), whereas for large $\varepsilon$ the dispersion $\langle {(\Delta E)^2}\rangle$ should be about $\langle {\Delta E}\rangle ^2$ of the phase-bunched particle population. We have checked this assumption with an analytical approach, which we further introduce.

3. The diffusion coefficient model for an arbitrary wave intensity

The Hamiltonian $H=m_ec^2\gamma$ with $\gamma$ given by (2.9) determines the electron dynamics. For this Hamiltonian system, we derive an analytical approximation for the energy change $\Delta E = m_e c^2 \Delta \gamma$ due to a single resonant interaction. Such a change depends on the initial electron energy and pitch angle, and on the initial phase $\phi +\psi$. Thus, we finally aim to find a mean value $V_\gamma =\langle \Delta \gamma \rangle$ and variance $D_{\gamma \gamma }=\langle (\Delta \gamma )^2 \rangle$ with the averaging over initial phases.

To derive the $\Delta \gamma$ equation for $H=m_ec^2\gamma$, we follow the approach from Neishtadt & Vasiliev (Reference Neishtadt and Vasiliev2006) and Artemyev et al. (Reference Artemyev, Neishtadt, Vainchtein, Vasiliev, Vasko and Zelenyi2018a) for the perturbation theory application to a resonant system containing a phase and a small wave amplitude, $B_w / B_0 \ll 1$ (more precisely, $\varepsilon = B_w / B_0(0) \ll 1$). For analytical consideration we assume that $B_w$ is a constant or a function of the magnetic latitude with spatial gradient much weaker than wave phase gradients $\partial \phi /\partial s=k$.

The Hamiltonian from (2.10a,b) is time-dependent as $\phi = \phi (t)$. To find the invariant equation, we define a generation function of the second kind $F_2(s, \psi, P_z, I)$ for $(s, p_z, \psi, I_x) \to (\tilde {z}, P_z, \zeta, I)$:

(3.1)\begin{equation} F_2(s, \psi, P_z, I) = (\phi + \psi)I + P_z z \quad \begin{cases} p_z = \dfrac{\partial F_2}{\partial z} = P_z + k I ,\\ I_x = \dfrac{\partial F_2}{\partial \psi} = I ,\\ \tilde{z} = \dfrac{\partial F_2}{\partial P_z} = z ,\\ \zeta = \dfrac{\partial F_2}{\partial I} = \phi + \psi ,\\ {\mathcal{H}} - H = \dfrac{\partial F_2}{\partial t} ={-}\omega I. \end{cases} \end{equation}

As a result, a modified Hamiltonian ${\mathcal {H}}$ is

(3.2)\begin{equation} \left. \begin{aligned} {\mathcal{H}} & = m_e c^2 \gamma - \omega I + m_e c^2 \sqrt{\frac{2 I \varOmega_{ce}}{m_e c^2}} \frac{\varOmega_{ce}}{\gamma k c} \frac{B_w}{B_0} \cos\zeta, \\ \gamma & = \sqrt{1 + \left( \frac{P_z + k I}{m_e c} \right)^2 + \frac{2 I \varOmega_{ce}}{m_e c^2}}. \end{aligned} \right\} \end{equation}

Note that $\partial _t {\mathcal {H}} = 0$. Thus, in the zeroth order of $\varepsilon$ the first invariant of this system can be written as

(3.3)\begin{equation} m_e c^2 \gamma - \omega I = {\rm const.} \end{equation}

Without a wave perturbation, $I$ is invariant ($I = \textrm {const.}$) by its definition, and $\gamma$ has a constant value, considering (3.3). Thus, all derivatives of those variables are first-order terms of $\varepsilon$ or higher. Resonant wave–particle interactions change the particle energy $E = m_e c^2 \gamma$: $\gamma _0 \to \gamma _0 + \Delta \gamma$, where $\gamma _0$ is the initial Lorentz factor (at $t = 0$). The resonance equation can be written as $\dot {\zeta } = \partial {\mathcal {H}}/\partial I = 0$. This equation characterizes the dynamics of the system near the resonance point and, considering (3.2), can be written as (subscript $R$ means that the function is evaluated at resonance)

(3.4a,b)\begin{equation} m_e c^2 \left. \frac{\partial \gamma}{\partial I} \right\vert_R - \omega = 0, \quad \gamma_R= \frac{k_R c}{\omega} \frac{P_z + k_R I_R}{m_e c} + \frac{\varOmega_{ce,R}}{\omega}. \end{equation}

The Hamiltonian ${\mathcal {H}}$ has to be expanded near the resonance point $\dot {\zeta } = 0$ with values $\gamma _R = \gamma _0 + O(\varepsilon )$, $I_R = I_0 + O(\varepsilon )$:

(3.5)\begin{equation} \left. \begin{aligned} {\mathcal{H}} & \approx {\mathcal{H}}_R + \frac{m_e c^2}{2} \left.\frac{\partial^2 \gamma}{\partial I^2}\right\vert_{R} (I - I_R)^2 + m_e c^2 \sqrt{\frac{2 I_R \varOmega_{ce,R}}{m_e c^2}} \frac{\varOmega_{ce,R}}{\gamma_R k_R c} \frac{B_w}{B_{0,R}} \cos\zeta, \\ {\mathcal{H}}_R & = m_e c^2 \gamma_R - \omega I_R, \end{aligned} \right\} \end{equation}

where

(3.6)\begin{equation} \left.\frac{\partial^2 \gamma}{\partial I^2}\right\vert_{R} = \frac{k_R^2 c^2 - \omega^2}{m_e^2 c^4 \gamma_R}. \end{equation}

Thus, the final form of ${\mathcal {H}}$ is

(3.7)\begin{equation} \left. \begin{aligned} {\mathcal{H}} & = {\mathcal{H}}_R + \frac{(I - I_R)^2}{2 g} + m_e c^2\sqrt{\frac{2 I_R \varOmega_{ce,R}}{m_e c^2}} \frac{\varOmega_{ce,R}}{\gamma_R k_R} \frac{B_w}{B_{0,R}} \cos\zeta, \\ g^{{-}1} & \equiv m_e c^2 \left.\frac{\partial^2 \gamma}{\partial I^2}\right\vert_{R} =\frac{k_R^2 c^2 - \omega^2}{m_e c^2 \gamma_R}. \end{aligned} \right\} \end{equation}

The second term of ${\mathcal {H}}$ from (3.7) is an analogue of the kinetic energy of the particle motion near the resonance, with $I - I_R$ being the canonical momentum and $g$ playing the role of mass. Thus, we introduce a generating function of the third kind $F_3(\tilde {z}, \tilde {\zeta }, P_z, I)$ for $(z, P_z, \zeta, I) \to (\tilde {z}, \tilde {P}_z, \tilde {\zeta }, P_{\zeta })$:

(3.8)\begin{equation} F_3(\tilde{z}, \tilde{\zeta}, P_z, I) ={-}(I - I_R) \tilde{\zeta} - P_z \tilde{z} \quad \begin{cases} z ={-}\dfrac{\partial F_3}{\partial P_z} = \tilde{z} - \dfrac{\partial I_R}{\partial P_z} \tilde{\zeta},\\ \zeta ={-}\dfrac{\partial F_3}{\partial I} = \tilde{\zeta} ,\\ \tilde{P}_z ={-}\dfrac{\partial F_3}{\partial \tilde{z}} = P_z - \dfrac{\partial I_R}{\partial \tilde{z}} \tilde{\zeta} ,\\ P_{\zeta} ={-}\dfrac{\partial F_3}{\partial \tilde{\zeta}} = I - I_R . \end{cases} \end{equation}

As $I$ is an invariant in the unperturbed system, $\partial _{P_z} I_R \sim \varepsilon$ and $\partial _{\tilde {z}} I_R \sim \varepsilon$. This means that the first term of the Hamiltonian ${\mathcal {H}}$ can be expanded with respect to $\varepsilon$:

(3.9)\begin{align} {\mathcal{H}}_R & = {\mathcal{H}}_R\left(z, P_z\right) = {\mathcal{H}}_R\left(\tilde{z} - \frac{\partial I_R}{\partial P_z} \zeta, \tilde{P}_z + \frac{\partial I_R}{\partial \tilde{z}} \zeta\right) \nonumber\\ & \approx {\mathcal{H}}_R\left(\tilde{z}, \tilde{P}_z\right) - \frac{\partial {\mathcal{H}}_R}{\partial \tilde{z}} \frac{\partial I_R}{\partial \tilde{P}_z} \zeta + \frac{\partial {\mathcal{H}}_R}{\partial \tilde{P}_z} \frac{\partial I_R}{\partial \tilde{z}} \zeta \nonumber\\ & = {\mathcal{H}}_R\left(\tilde{z}, \tilde{P}_z\right) + \left\{{\mathcal{H}}_R, I_R \right\}_{\tilde{z}, \tilde{P}_z} \zeta . \end{align}

Substituting the expanded form of ${\mathcal {H}}_R$ into (3.7), we obtain two separate parts of the Hamiltonian ${\mathcal {H}}={\mathcal {H}}_R+{\mathcal {H}}_\zeta$, ${\mathcal {H}}_R = m_e c^2 \gamma _R - \omega I_R$ with canonical variables $(\tilde {z}, \tilde {P}_z, \zeta, P_{\zeta })$:

(3.10)\begin{equation} {\mathcal{H}}_{\zeta} = \frac{P_{\zeta}^2}{2g} + \left\{{\mathcal{H}}_R, I_R \right\}_{\tilde{z}, \tilde{P}_z} \zeta + m_e c^2 \sqrt{\frac{2 I_R \varOmega_{ce,R}}{m_e c^2}} \frac{\varOmega_{ce,R}}{\gamma_R k_R c} \frac{B_w}{B_{0,R}} \cos\zeta. \end{equation}

Hamiltonian ${\mathcal {H}}_R$ describes $(\tilde {z}, \tilde {P}_z)\approx (z, P_z)$ dynamics in the resonance, and this Hamiltonian does not depend on the fast phase $\zeta$. Hamiltonian ${\mathcal {H}}_{\zeta }$ is a $\zeta$-dependent pendulum Hamiltonian that describes fast phase and conjugated momentum dynamics around the resonance. Coefficients of Hamiltonian ${\mathcal {H}}_\zeta$ depend on $(\tilde {z}, \tilde {P}_z)$ and slowly change along the resonant trajectory.

In the Hamiltonian ${\mathcal {H}}_\zeta$, an effective potential energy, ${\mathcal {H}}_\zeta -P_\zeta ^2/2g$, contains two terms: the first term $\sim \{{\mathcal {H}}_R,I_R\}_{\tilde {z}, \tilde {P}_z}$ describes the impact of the background magnetic field gradient and the second term $\sim B_w$ describes the effect of the wave's field. The important system parameter is the ratio of magnitudes of these two terms:

(3.11)\begin{equation} a = m_e c^2 \sqrt{\frac{2 I_R \varOmega_{ce,R}}{m_e c^2}} \frac{\varOmega_{ce,R}}{\gamma_R k_R c} \frac{B_w}{B_{0,R}} \left\{{\mathcal{H}}_R, I_R \right\}_{\tilde{z}, \tilde{P}_z}^{{-}1}. \end{equation}

Taking into account ${\mathcal {H}}_R=m_ec^2\gamma _R-\omega I_R$, we can rewrite the Poisson bracket:

(3.12)\begin{equation} \left\{{\mathcal{H}}_R, I_R \right\}_{\tilde{z}, \tilde{P}_z} = m_e c^2 \left\{ \gamma_R, I_R \right\}_{\tilde{z},\tilde{P}_z} \approx m_e c^2 \left\{ \gamma_R, I_R \right\}_{z, P_z}, \end{equation}

where $\gamma _R$ and $I_R$ are given by (3.2) and (3.4a,b). Combining these equations, we obtain (subscript $R$ is omitted in equations below)

(3.13a,b)\begin{equation} \gamma = \frac{k c}{\omega} \frac{P_z + k I}{m_e c} + \frac{\varOmega_{ce}}{\omega},\quad I = \frac{m_e c^2}{2 \varOmega_{ce}} \left[ \gamma^2 - \left(\frac{\omega}{k c} \right)^2 \left( \gamma - \frac{\varOmega_{ce}}{\omega} \right)^2 - 1 \right]. \end{equation}

Therefore, the Poisson bracket $\{\cdots \} = \{\cdots \}_{s,P_{\parallel }}$ can be rewritten as

(3.14)\begin{align} \left\{ \gamma, I \right\} & = I \varOmega_{ce} \left\{ \gamma, \frac{1}{\varOmega_{ce}} \right\} - \frac{m_e c^2}{2 \varOmega_{ce}} \left\{ \gamma, \left(\frac{\omega}{k c} \right)^2 \left( \gamma - \frac{\varOmega_{ce}}{\omega} \right)^2 \right\} ={-}I \frac{\partial \gamma}{\partial P_z} \partial_z \left[ \ln \varOmega_{ce} \right] \nonumber\\ & \quad - \frac{m_e c^2}{\varOmega_{ce}} \frac{\omega}{k c} \left( \gamma - \frac{\varOmega_{ce}}{\omega} \right) \left\{ \gamma, \frac{\omega}{k c} \left( \gamma - \frac{\varOmega_{ce}}{\omega} \right) \right\} ={-}I \frac{\partial \gamma}{\partial P_z} \partial_z \left[ \ln \varOmega_{ce} \right] \nonumber\\ & \quad - \frac{m_e c^2}{\varOmega_{ce}} \frac{\omega}{k c} \left( \gamma - \frac{\varOmega_{ce}}{\omega} \right) \left[ \left( \gamma - \frac{\varOmega_{ce}}{\omega} \right) \frac{\partial \gamma}{\partial P_z} \partial_z \left[ \frac{\omega}{k c} \right] - \frac{\omega}{k c} \frac{\varOmega_{ce}}{\omega} \frac{\partial \gamma}{\partial P_z} \partial_z \left[ \ln \varOmega_{ce} \right] \right] \nonumber\\ & = \frac{m_e c^2}{\omega} \frac{\partial \gamma}{\partial P_z} \left[ \left(\left(\frac{\omega}{kc}\right)^2 \left( \gamma - \frac{\varOmega_{ce}}{\omega} \right) - \frac{\omega I}{m_e c^2} \right) \partial_z \left[ \ln \varOmega_{ce} \right] - \frac{\omega}{\varOmega_{ce}} \left(\frac{\omega}{kc}\right)^2 \right. \nonumber\\ & \quad \times \left. \left( \gamma - \frac{\varOmega_{ce}}{\omega} \right)^2 \partial_z \left[ \ln \frac{\omega}{k c} \right] \right]. \end{align}

To determine $\partial _z [ \ln ({\omega }/{k c}) ]$ we use the cold plasma approximation, $k c / \omega = (\varOmega _{pe} / \omega ) ( \varOmega _{ce}/\omega - 1)^{-{1}/{2}}$, with constant plasma frequency $\varOmega _{pe}$. Therefore, for partial derivatives in $\{ \gamma, I \}$ we can write

(3.15)\begin{equation} \left. \begin{aligned} \frac{\partial \gamma}{\partial P_z} & = \frac{1}{m_e c} \frac{k c}{\omega} \left( 1 + k \frac{\partial I}{\partial P_z} \right),\quad \frac{\partial I}{\partial P_z} = \frac{m_e c^2}{\varOmega_{ce}} \frac{\partial \gamma}{\partial P_z} \left[ \gamma - \left(\frac{\omega}{k c} \right)^2 \left( \gamma - \frac{\varOmega_{ce}}{\omega} \right) \right],\\ \frac{\partial \gamma}{\partial P_z} & ={-}\frac{1}{\gamma m_e c} \frac{\varOmega_{ce}}{\omega} \frac{k c}{\omega} \left[ \left(\frac{k c}{\omega}\right)^2 - 1 \right]^{{-}1},\quad \frac{\partial }{\partial s} \left[ \ln \frac{\omega}{k c} \right] = \frac{1}{2} \frac{\partial_z \left[ \ln \varOmega_{ce} \right]}{1 - \omega/\varOmega_{ce}} \end{aligned} \right\} \end{equation}

and

(3.16)\begin{gather} \left\{ \gamma, I \right\} = \frac{\partial_z \left[ \ln \varOmega_{ce} \right]}{\gamma k} \left[\frac{\left( \gamma - \varOmega_{ce}/\omega \right)^2}{2 \left( 1 - \omega/\varOmega_{ce} \right)} - \varOmega_{ce}/\omega \left(\gamma - \frac{\varOmega_{ce}}{\omega} - \frac{\omega I}{m_e c^2} \left(\frac{k c}{\omega}\right)^2 \right) \right] \nonumber\\ \times \left[ \left(\frac{k c}{\omega}\right)^2 - 1 \right]^{{-}1}. \end{gather}

Having $\{ \gamma, I \}$, we can determine $a$ from (3.11) at the resonance. As we show, this is the main parameter controlling the energy change due to the resonant wave–particle interaction.

To write an equation for the resonant energy change, $\Delta \gamma$, we use the invariant from (3.3):

(3.17)\begin{align} \Delta \gamma & = \frac{\omega}{m_e c^2} \Delta I = \frac{2 \omega}{m_e c^2} \int_{-\infty}^{t_R} \,{\rm d}t \ \dot{I} ={-}\frac{2 \omega}{m_e c^2} \int_{-\infty}^{t_R} \,{\rm d}t \frac{\partial {\mathcal{H}}}{\partial \zeta} \nonumber\\ & \approx 2 \omega \sqrt{\frac{2 I_R \varOmega_{ce,R}}{m_e c^2}} \frac{\varOmega_{ce,R}}{\gamma_R k_R c} \frac{B_w}{B_{0,R}} \int_{-\infty}^{\zeta_R} \,{\rm d}\zeta \frac{\sin\zeta}{\dot{\zeta}}, \end{align}

where $\varOmega _{ce,R}=eB_{0,R}/m_ec$ and $B_{0,R}$ is defined at the resonant $z=z(t_R)$ for given initial energy and pitch angle. For $\dot \zeta$ we use the Hamiltonian equation $\dot \zeta =\partial {\mathcal {H}}_\zeta /\partial P_\zeta$. The resonance is defined by $\dot {\zeta } = P_{\zeta } / g = 0$ with the solution $\zeta _R$ according to $\zeta _R + a\cos \zeta _R=\xi$ with the resonant energy $\xi = {\mathcal {H}}_\zeta / \{ {\mathcal {H}}_R, I_R \}$. Thus, (3.17) can be written as

(3.18)\begin{equation} \left. \begin{aligned} \Delta \gamma & = 2 \omega \sqrt{\frac{g I_R \varOmega_{ce,R}}{m_e c^2}} \frac{\varOmega_{ce,R}}{\gamma_R k_R c} \frac{B_w}{B_{0,R}} \left\{ {\mathcal{H}}_R, I_R \right\}_{\tilde{z}, \tilde{P}_z}^{{-}1/2} \int_{-\infty}^{\zeta_R} \,{\rm d}\zeta \frac{\sin\zeta}{\sqrt{\xi - \zeta - a \cos\zeta}} \\ & = \left( \frac{2 I_R \varOmega_{ce,R}}{m_e c^2} \right)^{\frac{1}{4}} \sqrt{\frac{2 \varOmega_{ce,R} / k_R c}{k_R^2 c^2 / \omega^2 - 1} \frac{B_w}{B_{0,R}}} f(a,\xi), \\ f(a,\xi) & = \int_{-\infty}^{\zeta_R} \,{\rm d}\zeta \frac{\sqrt{a} \sin\zeta}{\sqrt{\xi - \zeta - a \cos\zeta}} . \end{aligned} \right\} \end{equation}

Figure 4(a) shows the $f(a,\xi )$ function. This function is periodic with period $2{\rm \pi}$ for $\xi$ (this can be shown analytically; see Appendix A).

Figure 4. (a) Profiles of $f(a,\xi )$ for three $a$ values. (b) Profiles of $\langle f(a,\xi )\rangle _{\xi }$ and $\langle (f(a,\xi ))^2\rangle _{\xi }$.

The value of $\xi$ varies with the initial conditions, i.e. with wave phase $\phi$, gyrophase $\psi$ and electron position on the trajectory in $(s,P_\parallel )$ plane far from the resonance. Assuming a uniform distribution of these parameters, we numerically integrate an ensemble of trajectories as determined by (2.10a,b), and so determine the probability function for $\xi$. Figure 5 shows such distributions of $\xi$ for three sets of the system parameters. These distributions are very close to uniform distributions with $\xi \in [0,2{\rm \pi} ]$, which suggests that averaging over initial parameters $\phi$, $\psi$ and $s$ can be substituted with averaging over $\xi$ with constant weights (see also discussion in Itin, Neishtadt & Vasiliev (Reference Itin, Neishtadt and Vasiliev2000) and Albert et al. (Reference Albert, Artemyev, Li, Gan and Ma2022)):

(3.19)\begin{equation} \iiint_\varPi \,{\rm d}\phi \, {\rm d}\psi \, {\rm d}\lambda \to \int_{\xi_0}^{\xi_0 + 2 {\rm \pi}} \,{\rm d}\xi, \quad \begin{cases} \xi_0 - \zeta_0 - a \cos\zeta_0 = 0 ,\\ 1 = a \sin \zeta_0, \end{cases} \end{equation}

where $\varPi$ is the parametric range ($3D$ uniform distribution) and $\xi _0 = \xi _0(a)$ determines the case when the integral diverges near the resonance point (see Appendix A). Thus, the equation for the variance $D_{\gamma \gamma }$ can be written as

(3.20)\begin{equation} D_{\gamma \gamma}= \langle (\Delta \gamma)^2 \rangle_{\xi} = \frac{2 \varOmega_{ce,R} / k_R c}{k_R^2 c^2 / \omega^2 - 1} \frac{B_w}{B_{0,R}} \sqrt{ \frac{2 I_R \varOmega_{ce,R}}{m_e c^2} } \langle f^2(a,\xi) \rangle_{\xi}, \end{equation}

where $D_{\gamma \gamma }$ can be considered as a diffusion rate for a unit time interval between two resonant interactions (the actual diffusion rate is the ratio of $\langle D_{\gamma \gamma } \rangle$ and a fraction of electron bounce period).

Figure 5. Probability distributions of $\xi$ for $10^5$ trajectories and (a) $E_0 = 100$ keV, $\alpha _0 = 40^{\circ }$, $B_w = 50$ pT, (b) $E_0 = 100$ keV, $\alpha _0 = 40^{\circ }$, $B_w = 500$ pT, (c) $E_0 = 300$ keV, $\alpha _0 = 60^{\circ }$, $B_w = 500$ pT.

The function $f(a,\xi )$ determines the difference of the diffusion coefficient $\sim D_{\gamma \gamma }$ and the quasi-linear model. For the case of $a \ll 1$, $D_{\gamma \gamma }$ asymptotically tends to the quasi-linear equation $D_{\gamma \gamma } = \langle (\Delta \gamma )^2 \rangle \sim (B_w / B_0)^2$ (Albert Reference Albert2010). To verify that, we have to expand $f(a,\xi )$ in a Taylor series:

(3.21)\begin{equation} \left. \begin{aligned} \lim_{a\to0} \frac{f(a,\xi)}{\sqrt{a}} & = \int_{-\infty}^{\xi} \,{\rm d}\zeta \frac{\sin\zeta}{\sqrt{\xi - \zeta}} = \sqrt{\rm \pi} \sin(\xi - {\rm \pi}/ 4),\\ \lim_{a\to0} \langle f(a,\xi) \rangle_{\xi} & = 0, \quad \lim_{a\to0} \frac{\langle f^2(a,\xi) \rangle_{\xi}}{a} = \frac{\rm \pi}{2}, \end{aligned} \right\} \end{equation}

and thus $D_{\gamma \gamma } \propto (B_w/B_0)\langle f^2(a,\xi ) \rangle _{\xi } \propto (B_w/B_0)^2$, because $a\propto B_w/B_0$ (see (3.11)).

Figure 4(b) shows that there are two $a$ ranges with different wave–particle resonant effects: for $a<1$ the mean value of $\langle f(a,\xi ) \rangle _{\xi }$ equals zero and there is only a particle diffusion $\propto a \langle f^2(a,\xi ) \rangle _{\xi }$, whereas for $a>1$ there is a non-zero negative $\langle f(a,\xi ) \rangle _{\xi }$. The diffusion $\propto a \langle f^2(a,\xi ) \rangle _{\xi }$ has a local maximum at $a\approx 1.0392$, a local minimum at $a\approx 1.5923$ and then increases with $a$ as $\propto a$. The mean value $\langle f(a,\xi ) \rangle _{\xi }$ has an asymptote $4 \sqrt {2} / {\rm \pi}$ for $a\gg 1$. There is an important property of $f(a,\xi )$: $\langle f^2(a,\xi ) \rangle _{\xi }-\langle f(a,\xi ) \rangle _{\xi }^2 \to 0$ for $a\to \infty$ (see Appendix A for details). This property defines the behaviour of the diffusion rate $\propto a\langle f^2(a,\xi ) \rangle _{\xi }$ for $a\gg 1$.

As shown in figure 3, the strong wave modulation should result in a symmetric distribution of $\Delta \gamma$ with a zero mean value and with the dispersion $\langle (\Delta \gamma )^2 \rangle$ about $\langle (\Delta \gamma )^2 \rangle _{\xi }$ where $\xi$ averaging is performed for population of phase-bunched particles (i.e. particles with $\Delta \gamma <0$). Therefore, for such strongly modulated waves, we consider $\langle (\Delta \gamma )^2 \rangle _{\xi }$ as a diffusion rate for both $a<1$ and $a>1$ parametric ranges. Figure 6 shows $\langle (\Delta \gamma )^2 \rangle _{\xi }$ distributions in energy and pitch-angle space for three typical wave intensities. Electrons of lower energy/higher pitch angle resonate with waves closer to the equatorial plane, where $a$ is large because $\{ \gamma _R, I_R \}_{s, P_{\parallel }} \propto \partial \varOmega _{ce}/\partial s$ and tends to zero around the equator.

Figure 6. Two-dimensional energy/pitch-angle maps of $(E_0,\alpha _{0})$ for (a) $B_w = 100$ pT, (b) $B_w = 500$ pT and (c) $B_w = 1000$ pT. Black curve shows $a=1$.

4. Discussion and conclusions

In this study, we derive the diffusion rate for electrons resonantly interacting with intense whistler-mode waves. Although such intense waves may resonate with electrons nonlinearly, the efficiency of this interaction would be significantly reduced by wave modulation (Zhang et al. Reference Zhang, Agapitov, Artemyev, Mourenas, Angelopoulos, Kurth, Bonnell and Hospodarsky2020a; Tao et al. Reference Tao, Bortnik, Thorne, Albert and Li2012b; Allanson et al. Reference Allanson, Watt, Allison and Ratcliffe2021; An et al. Reference An, Wu and Tao2022; Gan et al. Reference Gan, Li, Ma, Albert, Artemyev and Bortnik2020, Reference Gan, Li, Ma, Artemyev and Albert2022). Indeed, most of the intense whistler-mode waves that are observed in space plasma systems (like Earth's radiation belts, bow shock, foreshock transients and plasma injections) are presented in the form of short well-modulated wave packets (see examples in figure 1 and references in the figure caption).

Using test particle simulations, we show that such modulation will reduce the difference between energy changes of phase trapping and phase bunching electrons, and make the energy change distribution more symmetric (see figure 3). In the limit of total symmetrization of energy change distribution (extremely modulated wave packets), the main (and the only one) characteristic of electron resonant scattering will be the diffusion rate describing the dispersion of this distribution. Such a diffusion rate can be evaluated analytically: the dispersion of the energy changes $\langle (\Delta \gamma )^2\rangle _\xi$ tends to $\langle \Delta \gamma \rangle _\xi ^2$ for large wave amplitudes. That is to say, $\langle (\Delta \gamma )^2\rangle _\xi \to \langle \Delta \gamma \rangle _\xi ^2 \propto B_w$ for $a\gg 1$ (i.e. when wave amplitude is sufficiently large).

Using this theoretical result, we can calculate the ratio of such a nonlinear diffusion rate and the quasi-linear rate, $\langle (\Delta \gamma )^2\rangle _\xi ^{QL} \sim B_w^2$. We can extrapolate to the large-wave-amplitude limit using a renormalization:

(4.1)\begin{equation} \langle(\Delta\gamma)^2\rangle_\xi^{QL}(B_w) =\left(B_w/B_{w,\min}\right)^2\langle(\Delta\gamma)^2\rangle_\xi^{QL}(B_{w,\min}), \end{equation}

where $B_{w,\min }$ corresponds to the $a\ll 1$ limit. The ratio $\langle (\Delta \gamma )^2\rangle _\xi /\langle (\Delta \gamma )^2\rangle _\xi ^{QL}$ should show how quasi-linear diffusion models overestimate the diffusion rates for intense waves, because such models scale $\langle (\Delta \gamma )^2\rangle _\xi ^{QL}$ with wave intensity $B_w^2$. Both the numerator and the denominator of this ratio should be weighted with the actual distribution of observed wave intensities, ${\mathcal {F}}(B_w/B_0)$, and we use two distributions of whistler-mode wave packets collected in the inner magnetosphere.

Figure 7(a) shows two examples of ${\mathcal {F}}(B_w/B_0)$: the main difference between these distributions is in the definition of wave packets used in the two statistics (see details in Zhang et al. (Reference Zhang, Mourenas, Artemyev, Angelopoulos, Bortnik, Thorne, Kurth, Kletzing and Hospodarsky2019, Reference Zhang, Mourenas, Artemyev, Angelopoulos, Kurth, Kletzing and Hospodarsky2020b)). Using these distributions, we plot the ratio $\langle \langle (\Delta \gamma )^2\rangle _{\xi }/\langle (\Delta \gamma )^2\rangle _{\xi }^{QL} \rangle _{\varepsilon }$ as a function of energy and pitch angle in figure 7(b,c). The region with $\langle \langle (\Delta \gamma )^2\rangle _{\xi }/\langle (\Delta \gamma )^2\rangle _{\xi }^{QL} \rangle _{\varepsilon } \approx 1$ corresponds to the dominant contribution of waves with insufficiently large wave amplitude, where $a<1$ for most part of $\varepsilon$, and the diffusion rate is $\langle (\Delta \gamma )^2\rangle _{\xi } \propto B_w^2$. The region with $\langle \langle (\Delta \gamma )^2\rangle _{\xi }/\langle (\Delta \gamma )^2\rangle _{\xi }^{QL} \rangle _{\varepsilon } < 1$ corresponds to the dominant contribution of high-intensity waves, where $a>1$ for a significant fraction of the ${\mathcal {F}}(B_w/B_0)$ distribution, and the diffusion rate $\langle (\Delta \gamma )^2\rangle _{\xi } \propto B_w$. Note that electrons of smaller energy/larger pitch angle resonate with waves closer to the equator, where $a<1$ for the larger part of the ${\mathcal {F}}(B_w/B_0)$ distribution.

Figure 7. (a) Probability distribution functions ${\mathcal {F}}(B_w/B_0)$ of whistler-mode wave intensities from Zhang et al. (Reference Zhang, Mourenas, Artemyev, Angelopoulos, Bortnik, Thorne, Kurth, Kletzing and Hospodarsky2019) (blue) and Zhang et al. (Reference Zhang, Mourenas, Artemyev, Angelopoulos, Kurth, Kletzing and Hospodarsky2020b) (red). (b,c) Two-dimensional distributions of $\langle \langle (\Delta \gamma )^2\rangle _{\xi }/\langle (\Delta \gamma )^2\rangle _{\xi }^{QL} \rangle _{\varepsilon }$ in $(E_0,\alpha _{0})$ space for the two ${\mathcal {F}}(B_w/B_0)$ distributions from (a).

Figure 7(b,c) demonstrates clearly that the quasi-linear diffusion model significantly overestimates the real diffusion of electrons of smaller energy/larger pitch angle. Note that a similar effect of diffusion rate reduction relative to the quasi-linear theory predictions has been obtained for broadband waves (see Tao et al. Reference Tao, Bortnik, Albert, Liu and Thorne2011, Reference Tao, Bortnik, Albert and Thorne2012a). This overestimation will be stronger for active geomagnetic conditions with higher wave intensity (Meredith et al. Reference Meredith, Horne, Thorne and Anderson2003, Reference Meredith, Horne, Sicard-Piet, Boscher, Yearby, Li and Thorne2012; Agapitov et al. Reference Agapitov, Artemyev, Krasnoselskikh, Khotyaintsev, Mourenas, Breuillard, Balikhin and Rolland2013, Reference Agapitov, Mourenas, Artemyev, Mozer, Hospodarsky, Bonnell and Krasnoselskikh2018), because $\langle (\Delta \gamma )^2\rangle _{\xi }/\langle (\Delta \gamma )^2\rangle _{\xi }^{QL} \propto 1/B_{w}$. Figure 7(b,c) is plotted under the assumption that there is no net contribution of nonlinear resonant phase bunching and phase trapping effects, and that all wave–particle interactions can be solely described by diffusion because the wave field is dominated by well-modulated short wave packets. The opposite assumption consists of a dominant role of phase trapping and phase bunching effects of electron resonant interactions with coherent long wave packets. For long-term electron dynamics, such effects also can be fitted by diffusion, but the diffusion rate will be much larger than the quasi-linear one (Artemyev et al. Reference Artemyev, Neishtadt, Vasiliev and Mourenas2021a).

The schematic of figure 8 generalizes both regimes of wave–particle resonant interactions for large-amplitude waves: $\langle (\Delta \gamma )^2\rangle _{\xi } \propto B_{w}$ for well-modulated short wave packets and $\langle (\Delta \gamma )^2\rangle _{\xi } \propto B_{w}^{1/2}$ for highly coherent long wave packets. For Earth's radiation belts, the quasi-linear simulations of long-term electron flux dynamics generally describe observed electron fluxes with a reasonable tuning of the averaged wave intensity (e.g. Thorne et al. Reference Thorne, Li, Ni, Ma, Bortnik, Chen, Baker, Spence, Reeves and Henderson2013; Li et al. Reference Li, Thorne, Ma, Ni, Bortnik, Baker, Spence, Reeves, Kanekal and Green2014; Drozdov et al. Reference Drozdov, Shprits, Orlova, Kellerman, Subbotin, Baker, Spence and Reeves2015; Ma et al. Reference Ma, Li, Bortnik, Thorne, Chu, Ozeke, Reeves, Kletzing, Kurth and Hospodarsky2018; Allison et al. Reference Allison, Shprits, Zhelavskaya, Wang and Smirnov2021). This will not be possible if either of the two limiting cases shown in figure 8 would work. Therefore, we conclude that wave–particle resonant interactions include both diffusion by short wave packets and nonlinear phase trapping/bunching by rare long wave packets, and a fine balance of these two regimes results in electron diffusion that can be mimicked by the quasi-linear diffusion models. However, for each specific event, such mimicking would require a proper tuning of wave intensity. This underlines the importance of investigations of nonlinear resonant interactions for accurate inclusion of the net effects of phase trapping and phase bunching into wave–particle interaction models.

Figure 8. A schematic view of diffusion rate scaling with wave amplitude normalized to the amplitude threshold for nonlinear resonant interactions, $B_w^{*}$. Quasi-linear diffusion rate $D_{QL}\propto B_w^2$ works only for $B_w/B_w^{*}<1$. For $B_w/B_w^{*}>1$ and incoherent waves should work $D\propto B_w$, whereas for coherent waves should work $D_{NL}\propto \sqrt {B_w}$ (see Artemyev et al. Reference Artemyev, Neishtadt, Vasiliev and Mourenas2021a).

5. Summary

We have derived the diffusion rate for relativistic electron scattering by intense whistler-mode waves of arbitrary amplitude. This diffusion rate repeats the $D\propto B_w^2$ scaling of the quasi-linear diffusion rate $D_{QL}$ for small amplitudes, and tends to $D\propto B_w$ scaling for amplitudes exceeding the threshold of nonlinear wave–particle interactions. Therefore, under the assumption of the absence of main nonlinear resonant effects (phase trapping and phase bunching) due to low wave coherence, the quasi-linear diffusion model will overestimate the diffusion rate for large amplitudes, because $D/D_{QL}\propto B_w^{-1}$. This result demonstrates that the extrapolation of quasi-linear diffusion models should not work for high wave amplitudes, whereas the approximation of the total destruction of nonlinear resonant effects (phase trapping and phase bunching) due to low wave coherence/strong wave modulation will underestimate the rates of electron flux dynamics. Thus, for accurate inclusion of statistics of high wave amplitudes into electron flux models, we must account for the contribution of nonlinear resonant effects.

Acknowledgements

Editor Dmitri Uzdensky thanks the referees for their advice in evaluating this article.

Funding

V.A.F., P.I.S. and A.A.P. acknowledge support from the Russian Science Foundation through grant no. 19-12-00313 covering the theoretical part of this work. A.V.A. and X.-J.Z. acknowledge support from NASA HGI 80NSSC22K0522 covering spacecraft data analysis. O.A. gratefully acknowledges financial support from the University of Exeter, and also from the United Kingdom Research and Innovation (UKRI) Natural Environment Research Council (NERC) Independent Research Fellowship NE/V013963/1.

Declaration of interests

The authors report no conflict of interest.

Appendix A.

This appendix is devoted to the investigation of the properties of $f(a,\xi )$:

(A1)\begin{equation} f(a,\xi) = \int_{-\infty}^{\zeta_R} \,{\rm d}\zeta \frac{\sqrt{a} \sin\zeta}{\sqrt{\xi - \zeta - a \cos\zeta}}. \end{equation}

Considering that $\xi$ is a uniformly distributed random variable (see figure 5), we determine the mean value $\langle f(a,\xi ) \rangle _{\xi }$ and the mean square value $\langle f^2(a,\xi ) \rangle _{\xi }$.

First, let us discuss the convergence of $f(a,\xi )$ integral. At $\zeta \to -\infty$, the function $\sin \zeta / \sqrt {\xi -\zeta -a \cos \zeta }$ tends to $\sin \zeta / \sqrt {|\zeta |}$, corresponding to the Fresnel integral, and therefore the integral converges. At $\zeta \to \zeta _R$, the convergence depends on the behaviour of $\xi -\zeta -a \cos \zeta$. We introduce $\zeta = \zeta _R - \delta \zeta$ and consider $\delta \zeta$ to be sufficiently small:

(A2)\begin{align} \xi - \zeta - a \cos\zeta & = \xi - \zeta_R + \delta\zeta - a \cos(\zeta_R - \delta\zeta) \nonumber\\ & \approx \xi - \zeta_R + \delta\zeta - a \cos\zeta_R - a \delta\zeta \sin\zeta_R \nonumber\\ & = (1 - a \sin\zeta_R) \delta\zeta . \end{align}

If $1 - a \sin \zeta _R \neq 0$, the integral converges near $\zeta _R$ as $\delta \zeta ^{-1/2}$. The equation $1 - a \sin \zeta _R = 0$ determines cases when $f(a,\xi )$ has an infinite value. Thus, these points have to be excluded from the averaging procedure. By definition, $\zeta _R$ has to be the only solution of $\xi -\zeta -a\cos \zeta =0$ on the interval of integration. Therefore, we expect $\zeta _R$ to satisfy the equation $\zeta _R=\arcsin ({1}/{a}) + 2{\rm \pi} n$, where $n\in \mathbb {Z}$. Values of $\xi$, corresponding to the diverging cases (noted as $\xi _0$), can be simply determined from the following system of equations:

(A3)\begin{equation} \left. \begin{aligned} \xi_0 & = \zeta_0 + a \cos\zeta_0, \\ 1 & = a \sin\zeta_0. \end{aligned} \right\} \end{equation}

In the case of $a < 1$, there is no solution and the interval of integration for $\xi$ can be performed, i.e. $\xi \in ({\rm \pi} /2, 5 {\rm \pi}/2)$ (we take $\xi _0(a<1) \equiv \lim _{a \to 1^+} \xi _0(a) = {\rm \pi}/2$). For $a\geqslant 1$, this system has infinitely many solutions, separated by $2{\rm \pi}$. The function $f(a,\xi )$ is periodic with period $2{\rm \pi}$ for $\xi$:

(A4)\begin{align} f(a,\xi + 2 {\rm \pi}) & = \int_{-\infty}^{\zeta_R} \,{\rm d}\zeta \frac{\sqrt{a} \sin\zeta}{\sqrt{\xi + 2 {\rm \pi}- \zeta - a \cos\zeta}} = \int_{-\infty}^{\zeta_R - 2 {\rm \pi}} \,{\rm d}\tilde{\zeta} \frac{\sqrt{a} \sin\tilde{\zeta}}{\sqrt{\xi - \tilde{\zeta} - a \cos\tilde{\zeta}}} \nonumber\\ & = \int_{-\infty}^{\tilde{\zeta}_R} \,{\rm d}\tilde{\zeta} \frac{\sqrt{a} \sin\tilde{\zeta}}{\sqrt{\xi - \tilde{\zeta} - a \cos\tilde{\zeta}}} = f(a,\xi). \end{align}

Thus, without loss of generality, we consider only one value of $\xi _0$ and the corresponding phase $\zeta _0$. As a result, the mean value $\langle f(a,\xi ) \rangle _{\xi }$ and the mean square value $\langle f^2(a,\xi ) \rangle _{\xi }$ can be defined as

(A5)\begin{gather} \langle f(a,\xi) \rangle_{\xi} = \frac{1}{2 {\rm \pi}}\int_{\xi_0}^{\xi_0 + 2{\rm \pi}} \,{\rm d}\xi \int_{-\infty}^{\zeta_R} \,{\rm d}\zeta \frac{\sqrt{a} \sin\zeta}{\sqrt{\xi - \zeta - a \cos\zeta}}, \end{gather}
(A6)\begin{gather}\langle f^2(a,\xi) \rangle_{\xi} = \frac{1}{2 {\rm \pi}}\int_{\xi_0}^{\xi_0 + 2{\rm \pi}} \,{\rm d}\xi \left[ \int_{-\infty}^{\zeta_R} \,{\rm d}\zeta \frac{\sqrt{a} \sin\zeta}{\sqrt{\xi - \zeta - a \cos\zeta}} \right]^2. \end{gather}

To proceed, it is necessary to determine $\zeta _R$ as a function of $\xi$ on the interval $(\xi _0, \xi _0 + 2{\rm \pi} )$:

(A7)\begin{equation} \left. \begin{aligned} & \xi = \zeta_R + a \cos\zeta_R \Rightarrow 1 = (1 - a \sin\zeta_R) \partial_{\xi} \zeta_R \Rightarrow \partial_{\xi} \zeta_R = \frac{1}{1 - a \cos\zeta_R},\\ & \lim_{\xi\to\xi_0^+} \zeta_R(a,\xi) = \zeta_m, \quad \lim_{\xi\to(\xi_0 + 2{\rm \pi})^-} \zeta_R(a,\xi) = \zeta_0 + 2{\rm \pi}. \end{aligned} \right\} \end{equation}

Thus, $\zeta _R$ is bounded on the interval $(\zeta _m, \zeta _0 + 2{\rm \pi} )$ and is monotonic on it, considering the equation for $\partial _{\xi } \zeta _R$. This implies that the integral interval can be expanded (Neishtadt Reference Neishtadt1999; Artemyev et al. Reference Artemyev, Neishtadt, Vainchtein, Vasiliev, Vasko and Zelenyi2018a; Albert et al. Reference Albert, Artemyev, Li, Gan and Ma2022):

(A8)\begin{align} & \langle f(a,\xi) \rangle_{\xi} = \frac{1}{2 {\rm \pi}}\int_{\xi_0}^{\xi_0 + 2{\rm \pi}} \,{\rm d}\xi \int_{-\infty}^{\zeta_0 + 2{\rm \pi}} \,{\rm d}\zeta \frac{\sqrt{a} \sin\zeta}{\sqrt{\xi - \zeta - a \cos\zeta}} \theta(\xi - \zeta - a \cos\zeta) \nonumber\\ & \quad = \frac{1}{2 {\rm \pi}}\int_{\xi_0}^{\xi_0 + 2{\rm \pi}} \,{\rm d}\xi \left[ \int_{-\infty}^{\zeta_m} \,{\rm d}\zeta \frac{\sqrt{a} \sin\zeta}{\sqrt{\xi - \zeta - a \cos\zeta}} \theta(\xi - \zeta - a \cos\zeta) \right.\nonumber\\ & \qquad \left. + \int_{\zeta_m}^{\zeta_0+2{\rm \pi}} \,{\rm d}\zeta \frac{\sqrt{a} \sin\zeta}{\sqrt{\xi - \zeta - a \cos\zeta}} \theta(\xi - \zeta - a \cos\zeta)\right] \nonumber\\ & \quad = \frac{1}{2 {\rm \pi}} \int_{-\infty}^{\zeta_m} \,{\rm d}\zeta \int_{\xi_0}^{\xi_0 + 2{\rm \pi}} \,{\rm d}\xi \frac{\sqrt{a} \sin\zeta}{\sqrt{\xi - \zeta - a \cos\zeta}} + \frac{1}{2 {\rm \pi}} \int_{\zeta_m}^{\zeta_0 + 2{\rm \pi}} \,{\rm d}\zeta \int_{\zeta + a\cos\zeta}^{\xi_0 + 2{\rm \pi}} \,{\rm d}\xi \frac{\sqrt{a} \sin\zeta}{\sqrt{\xi - \zeta - a \cos\zeta}} \nonumber\\ & \quad = \frac{\sqrt{a}}{\rm \pi} \left[ -\int_{\zeta_m - 2{\rm \pi}}^{\zeta_m} \,{\rm d}\zeta \sin\zeta \sqrt{\xi_0 - \zeta - a \cos\zeta} + \int_{\zeta_m}^{\zeta_0 + 2{\rm \pi}} \,{\rm d}\zeta \sin\zeta \sqrt{\xi_0 + 2{\rm \pi} - \zeta - a \cos\zeta}\right] \nonumber\\ & \quad ={-}\frac{\sqrt{a}}{\rm \pi} \int_{\zeta_0}^{\zeta_m} \,{\rm d}\zeta \sin\zeta \sqrt{\xi_0 - \zeta - a \cos\zeta} ={-}\frac{1}{{\rm \pi} \sqrt{a}} \int_{\zeta_0}^{\zeta_m} \,{\rm d}\zeta \sqrt{\xi_0 - \zeta - a \cos\zeta} \nonumber\\ & \quad ={-}\frac{1}{{\rm \pi} \sqrt{a}} \int_{\zeta_0}^{\zeta_m} \,{\rm d}\zeta \sqrt{\zeta_0 - \zeta + a(\cos\zeta_0 - \cos\zeta)}. \end{align}

The integral from (A8) does not have any singularities on the interval of integration and, thus, can be easily computed. To neglect the dependency on $\zeta _m$, an integral in a complex plane can be introduced:

(A9)\begin{equation} \langle f(a,\xi) \rangle_{\xi} ={-}\frac{1}{{\rm \pi} \sqrt{a}} \mathrm{Re} \left[\int_{\zeta_0}^{\zeta_0 + 2{\rm \pi}} \,{\rm d}\zeta \sqrt{\zeta_0 - \zeta + a(\cos\zeta_0 - \cos\zeta)} \right]. \end{equation}

Additionally, (A8) determines $\langle f(a,\xi ) \rangle _{\xi }$ for $a < 1$: $\langle f(a,\xi ) \rangle _{\xi } = 0$ as $\zeta _m = \zeta _0$. For $a\to \infty$, we get

(A10)\begin{equation} \left. \begin{aligned} & \zeta_0 = \arcsin\left(\frac{1}{a}\right) \Rightarrow \lim_{a \to \infty} \zeta_0 = 0, \quad \lim_{a \to \infty} \zeta_m = 2 {\rm \pi}, \\ & \lim_{a \to \infty} \langle f(a,\xi) \rangle_{\xi} ={-}\frac{1}{\rm \pi} \int_{0}^{2{\rm \pi}} \,{\rm d}\zeta \sqrt{1 - \cos\zeta} ={-}\frac{4\sqrt{2}}{\rm \pi}. \end{aligned} \right\} \end{equation}

Figure 9 shows that the asymptote $a \gg 1$ is the same for $\langle f^2(a,\xi ) \rangle _{\xi }$ and $\langle f(a,\xi ) \rangle _{\xi }^2$ functions and, considering (A10), $\lim _{a \to \infty } \langle f^2(a,\xi ) \rangle _{\xi } = {32}/{{\rm \pi} ^2}$.

Figure 9. Comparison between $\langle f^2(a,\xi ) \rangle _{\xi }$ and $\langle f(a,\xi ) \rangle _{\xi }^2$ as functions of $a$.

References

REFERENCES

Agapitov, O.V., Artemyev, A., Krasnoselskikh, V., Khotyaintsev, Y.V., Mourenas, D., Breuillard, H., Balikhin, M. & Rolland, G. 2013 Statistics of whistler mode waves in the outer radiation belt: cluster STAFF-SA measurements. J. Geophys. Res. 118, 34073420.CrossRefGoogle Scholar
Agapitov, O.V., Mourenas, D., Artemyev, A.V., Mozer, F.S., Hospodarsky, G., Bonnell, J. & Krasnoselskikh, V. 2018 Synthetic empirical chorus wave model from combined van allen probes and cluster statistics. J. Geophys. Res. 123 (1), 297314.CrossRefGoogle Scholar
Albert, J.M. 1993 Cyclotron resonance in an inhomogeneous magnetic field. Phys. Fluids B 5, 27442750.CrossRefGoogle Scholar
Albert, J.M. 2001 Comparison of pitch angle diffusion by turbulent and monochromatic whistler waves. J. Geophys. Res. 106, 84778482.CrossRefGoogle Scholar
Albert, J.M. 2010 Diffusion by one wave and by many waves. J. Geophys. Res. 115, A00F05.Google Scholar
Albert, J.M., Artemyev, A., Li, W., Gan, L. & Ma, Q. 2022 Analytical results for phase bunching in the pendulum model of wave–particle interactions. Front. Astron. Space Sci. 9, 971358.CrossRefGoogle Scholar
Allanson, O., Watt, C.E.J., Allison, H.J. & Ratcliffe, H. 2021 Electron diffusion and advection during nonlinear interactions with whistler mode waves. J. Geophys. Res. 126 (5), e28793.CrossRefGoogle Scholar
Allanson, O., Watt, C.E.J., Ratcliffe, H., Allison, H.J., Meredith, N.P., Bentley, S.N., Ross, J.P.J. & Glauert, S.A. 2020 Particle-in-cell experiments examine electron diffusion by whistler-mode waves: 2. Quasi-linear and nonlinear dynamics. J. Geophys. Res. 125 (7), e27949.CrossRefGoogle Scholar
Allison, H.J., Shprits, Y.Y., Zhelavskaya, I.S., Wang, D. & Smirnov, A.G. 2021 Gyroresonant wave–particle interactions with chorus waves during extreme depletions of plasma density in the Van Allen radiation belts. Sci. Adv. 7 (5), eabc0380.CrossRefGoogle ScholarPubMed
Andronov, A.A. & Trakhtengerts, V.Y. 1964 Kinetic instability of the Earth's outer radiation belt. Geomagn. Aeron. 4, 233242.Google Scholar
Angelopoulos, V. 2008 The THEMIS mission. Space Sci. Rev. 141, 534.CrossRefGoogle Scholar
An, Z., Wu, Y. & Tao, X. 2022 Electron dynamics in a chorus wave field generated from particle-in-cell simulations. Geophys. Res. Lett. 49 (3), e97778.CrossRefGoogle Scholar
Artemyev, A.V., Neishtadt, A.I., Vainchtein, D.L., Vasiliev, A.A., Vasko, I.Y. & Zelenyi, L.M. 2018 a Trapping (capture) into resonance and scattering on resonance: summary of results for space plasma systems. Commun. Nonlinear Sci. Numer. Simul. 65, 111160.CrossRefGoogle Scholar
Artemyev, A.V., Neishtadt, A.I., Vasiliev, A.A. & Mourenas, D. 2018 b Long-term evolution of electron distribution function due to nonlinear resonant interaction with whistler mode waves. J. Plasma Phys. 84, 905840206.CrossRefGoogle Scholar
Artemyev, A.V., Neishtadt, A.I., Vasiliev, A.A. & Mourenas, D. 2021 a Transitional regime of electron resonant interaction with whistler-mode waves in inhomogeneous space plasma. Phys. Rev. E 104 (5), 055203.CrossRefGoogle ScholarPubMed
Artemyev, A.V., Neishtadt, A.I., Vasiliev, A.A., Zhang, X.-J., Mourenas, D. & Vainchtein, D. 2021 b Long-term dynamics driven by resonant wave–particle interactions: from Hamiltonian resonance theory to phase space mapping. J. Plasma Phys. 87 (2), 835870201.CrossRefGoogle Scholar
Artemyev, A.V., Shi, X., Liu, T.Z., Zhang, X.J., Vasko, I. & Angelopoulos, V. 2022 Electron resonant interaction with whistler waves around foreshock transients and the bow shock behind the terminator. J. Geophys. Res. 127 (2), e29820.CrossRefGoogle Scholar
Artemyev, A.V., Vasiliev, A.A., Mourenas, D., Agapitov, O.V. & Krasnoselskikh, V.V. 2014 Electron scattering and nonlinear trapping by oblique whistler waves: the critical wave intensity for nonlinear effects. Phys. Plasmas 21 (10), 102903.CrossRefGoogle Scholar
Bell, T.F. 1984 The nonlinear gyroresonance interaction between energetic electrons and coherent VLF waves propagating at an arbitrary angle with respect to the Earth's magnetic field. J. Geophys. Res. 89, 905918.CrossRefGoogle Scholar
Bespalov, P.A., Zaitsev, V.V. & Stepanov, A.V. 1991 Consequences of strong pitch-angle diffusion of particles in solar flares. Astrophys. J. 374, 369.CrossRefGoogle Scholar
Bortnik, J., Thorne, R.M. & Inan, U.S. 2008 Nonlinear interaction of energetic electrons with large amplitude chorus. Geophys. Res. Lett. 35, 21102.CrossRefGoogle Scholar
Breuillard, H., Le Contel, O., Retino, A., Chasapis, A., Chust, T., Mirioni, L., Graham, D.B., Wilder, F.D., Cohen, I., Vaivads, A., et al. 2016 Multispacecraft analysis of dipolarization fronts and associated whistler wave emissions using MMS data. Geophys. Res. Lett. 43, 72797286.CrossRefGoogle Scholar
Cattell, C.A., Short, B., Breneman, A.W. & Grul, P. 2020 Narrowband large amplitude whistler-mode waves in the solar wind and their association with electrons: STEREO waveform capture observations. Astrophys. J. 897 (2), 126.CrossRefGoogle Scholar
Cattell, C., Short, B., Breneman, A., Halekas, J., Whittesley, P., Larson, D., Kasper, J.C., Stevens, M., Case, T., Moncuquet, M., et al. 2021 Narrowband oblique whistler-mode waves: comparing properties observed by Parker Solar Probe at $<0.3$ AU and STEREO at 1 AU. Astron. Astrophys. 650, A8.CrossRefGoogle Scholar
Demekhov, A.G., Taubenschuss, U. & Santolık, O. 2017 Simulation of VLF chorus emissions in the magnetosphere and comparison with THEMIS spacecraft data. J. Geophys. Res. 122, 166184.CrossRefGoogle Scholar
Drozdov, A.Y., Shprits, Y.Y., Orlova, K.G., Kellerman, A.C., Subbotin, D.A., Baker, D.N., Spence, H.E. & Reeves, G.D. 2015 Energetic, relativistic, and ultrarelativistic electrons: comparison of long-term VERB code simulations with Van Allen Probes measurements. J. Geophys. Res. 120, 35743587.CrossRefGoogle Scholar
Drummond, W.E. & Pines, D. 1962 Nonlinear stability of plasma oscillations. Nucl. Fusion Suppl. 3, 10491058.Google Scholar
Filatov, L.V. & Melnikov, V.F. 2017 Influence of whistler turbulence on fast electron distribution and their microwave emissions in a flare loop. Geomagn. Aeron. 57 (8), 10011008.CrossRefGoogle Scholar
Foster, J.C., Erickson, P.J. & Omura, Y. 2021 Subpacket structure in strong VLF chorus rising tones: characteristics and consequences for relativistic electron acceleration. Earth Planet. Space 73 (1), 140.CrossRefGoogle ScholarPubMed
Gan, L., Li, W., Ma, Q., Albert, J.M., Artemyev, A.V. & Bortnik, J. 2020 Nonlinear interactions between radiation belt electrons and chorus waves: dependence on wave amplitude modulation. Geophys. Res. Lett. 47 (4), e85987.CrossRefGoogle Scholar
Gan, L., Li, W., Ma, Q., Artemyev, A.V. & Albert, J.M. 2022 Dependence of nonlinear effects on whistler-mode wave bandwidth and amplitude: a perspective from diffusion coefficients. J. Geophys. Res. 127 (5), e30063.CrossRefGoogle Scholar
Hsieh, Y.-K., Kubota, Y. & Omura, Y. 2020 Nonlinear evolution of radiation belt electron fluxes interacting with oblique whistler mode chorus emissions. J. Geophys. Res 125, e2019JA027465.CrossRefGoogle Scholar
Hsieh, Y.-K. & Omura, Y. 2017 a Nonlinear dynamics of electrons interacting with oblique whistler mode chorus in the magnetosphere. J. Geophys. Res. 122, 675694.CrossRefGoogle Scholar
Hsieh, Y.-K. & Omura, Y. 2017 b Study of wave–particle interactions for whistler mode waves at oblique angles by utilizing the gyroaveraging method. Radio Sci. 52 (10), 12681281.CrossRefGoogle Scholar
Hull, A.J., Muschietti, L., Le Contel, O., Dorelli, J.C. & Lindqvist, P.A. 2020 MMS observations of intense whistler waves within Earth's supercritical bow shock: source mechanism and impact on shock structure and plasma transport. J. Geophys. Res. 125 (7), e27290.CrossRefGoogle Scholar
Hull, A.J., Muschietti, L., Oka, M., Larson, D.E., Mozer, F.S., Chaston, C.C., Bonnell, J.W. & Hospodarsky, G.B. 2012 Multiscale whistler waves within Earth's perpendicular bow shock. J. Geophys. Res. 117, 12104.Google Scholar
Inan, U.S., Bell, T.F. & Helliwell, R.A. 1978 Nonlinear pitch angle scattering of energetic electrons by coherent VLF waves in the magnetosphere. J. Geophys. Res. 83 (A7), 32353254.CrossRefGoogle Scholar
Itin, A.P., Neishtadt, A.I. & Vasiliev, A.A. 2000 Captures into resonance and scattering on resonance in dynamics of a charged relativistic particle in magnetic field and electrostatic wave. Physica D 141, 281296.CrossRefGoogle Scholar
Karpman, V.I. 1974 Nonlinear effects in the elf waves propagating along the magnetic field in the magnetosphere. Space Sci. Rev. 16, 361388.CrossRefGoogle Scholar
Karpman, V.I., Istomin, J.N. & Shklyar, D.R. 1974 Nonlinear theory of a quasi-monochromatic whistler mode packet in inhomogeneous plasma. Plasma Phys. 16, 685703.CrossRefGoogle Scholar
Karpman, V.I. & Shklyar, D.R. 1977 Particle precipitation caused by a single whistler-mode wave injected into the magnetosphere. Planet. Space Sci. 25, 395403.CrossRefGoogle Scholar
Katoh, Y. & Omura, Y. 2016 Electron hybrid code simulation of whistler-mode chorus generation with real parameters in the Earth's inner magnetosphere. Earth Planet. Space 68 (1), 192.CrossRefGoogle Scholar
Kennel, C.F. & Engelmann, F. 1966 Velocity space diffusion from weak plasma turbulence in a magnetic field. Phys. Fluids 9, 23772388.CrossRefGoogle Scholar
Landau, L.D. & Lifshitz, E.M. 1988 Vol. 1: Mechanics. Pergamon.Google Scholar
Le Contel, O., Roux, A., Jacquey, C., Robert, P., Berthomier, M., Chust, T., Grison, B., Angelopoulos, V., Sibeck, D., Chaston, C.C., et al. 2009 Quasi-parallel whistler mode waves observed by THEMIS during near-earth dipolarizations. Ann. Geophys. 27, 22592275.CrossRefGoogle Scholar
Le Queau, D. & Roux, A. 1987 Quasi-monochromatic wave–particle interactions in magnetospheric plasmas. Sol. Phys. 111, 5980.CrossRefGoogle Scholar
Li, W., Ma, Q., Shen, X.C., Zhang, X.J., Mauk, B.H., Clark, G., Allegrini, F., Kurth, W.S., Hospodarsky, G.B., Hue, V., et al. 2021 Quantification of diffuse auroral electron precipitation driven by whistler mode waves at jupiter. Geophys. Res. Lett. 48 (19), e95457.CrossRefGoogle Scholar
Li, W., Thorne, R.M., Ma, Q., Ni, B., Bortnik, J., Baker, D.N., Spence, H.E., Reeves, G.D., Kanekal, S.G., Green, J.C., et al. 2014 Radiation belt electron acceleration by chorus waves during the 17 March 2013 storm. J. Geophys. Res. 119, 46814693.CrossRefGoogle Scholar
Lyons, L.R. & Williams, D.J. 1984 Quantitative Aspects of Magnetospheric Physics. D. Reidel Publishing Company.CrossRefGoogle Scholar
Ma, Q., Li, W., Bortnik, J., Thorne, R.M., Chu, X., Ozeke, L.G., Reeves, G.D., Kletzing, C.A., Kurth, W.S., Hospodarsky, G.B., et al. 2018 Quantitative evaluation of radial diffusion and local acceleration processes during gem challenge events. J. Geophys. Res. 123 (3), 19381952.CrossRefGoogle Scholar
Melnikov, V.F. & Filatov, L.V. 2020 Conditions for whistler generation by nonthermal electrons in flare loops. Geomagn. Aeron. 60 (8), 11261131.CrossRefGoogle Scholar
Menietti, J.D., Averkamp, T.F., Kurth, W.S., Imai, M., Faden, J.B., Hospodarsky, G.B., Santolik, O., Clark, G., Allegrini, F., Elliott, S.S., et al. 2021 Analysis of whistler-mode and Z-mode emission in the juno primary mission. J. Geophys. Res. 126 (11), e29885.CrossRefGoogle Scholar
Meredith, N.P., Horne, R.B., Sicard-Piet, A., Boscher, D., Yearby, K.H., Li, W. & Thorne, R.M. 2012 Global model of lower band and upper band chorus from multiple satellite observations. J. Geophys. Res. 117, 10225.Google Scholar
Meredith, N.P., Horne, R.B., Thorne, R.M. & Anderson, R.R. 2003 Favored regions for chorus-driven electron acceleration to relativistic energies in the Earth's outer radiation belt. Geophys. Res. Lett. 30 (16), 160000.CrossRefGoogle Scholar
Mourenas, D., Zhang, X.-J., Artemyev, A.V., Angelopoulos, V., Thorne, R.M., Bortnik, J., Neishtadt, A.I. & Vasiliev, A.A. 2018 Electron nonlinear resonant interaction with short and intense parallel chorus wave packets. J. Geophys. Res. 123, 49794999.CrossRefGoogle Scholar
Mourenas, D., Zhang, X.J., Nunn, D., Artemyev, A.V., Angelopoulos, V., Tsai, E. & Wilkins, C. 2022 Short chorus wave packets: generation within chorus elements, statistics, and consequences on energetic electron precipitation. J. Geophys. Res. 127 (5), e30310.CrossRefGoogle ScholarPubMed
Mozer, F.S., Bonnell, J.W., Halekas, J.S., Rahmati, A., Schum, G. & Vasko, I.V. 2021 Whistlers in the solar vicinity that are spiky in time and frequency. Astrophys. J. 908 (1), 26.CrossRefGoogle Scholar
Neishtadt, A.I. 1999 On adiabatic invariance in two-frequency systems. In Hamiltonian Systems with Three or More Degrees of Freedom (ed. C. Simo), NATO ASI Series C, vol. 533, pp. 193–213. Kluwer Academic Publishers.CrossRefGoogle Scholar
Neishtadt, A.I. & Vasiliev, A.A. 2006 Destruction of adiabatic invariance at resonances in slow fast Hamiltonian systems. Nucl. Instrum. Meth. Phys. Res. A 561, 158165.CrossRefGoogle Scholar
Nunn, D. 1971 Wave–particle interactions in electrostatic waves in an inhomogeneous medium. J. Plasma Phys. 6, 291.CrossRefGoogle Scholar
Nunn, D. 1974 A self-consistent theory of triggered VLF emissions. Planet. Space Sci. 22, 349378.CrossRefGoogle Scholar
Nunn, D. 1986 A nonlinear theory of sideband stability in ducted whistler mode waves. Planet. Space Sci. 34, 429451.CrossRefGoogle Scholar
Nunn, D. & Omura, Y. 2012 A computational and theoretical analysis of falling frequency VLF emissions. J. Geophys. Res. 117, 8228.Google Scholar
Nunn, D., Zhang, X.J., Mourenas, D. & Artemyev, A.V. 2021 Generation of realistic short chorus wave packets. Geophys. Res. Lett. 48 (7), e92178.CrossRefGoogle Scholar
Oka, M., Otsuka, F., Matsukiyo, S., Wilson, L.B.I., Argall, M.R., Amano, T., Phan, T.D., Hoshino, M., Le Contel, O., Gershman, D.J., et al. 2019 Electron scattering by low-frequency whistler waves at Earth's bow shock. Astrophys. J. 886 (1), 53.CrossRefGoogle Scholar
Oka, M., Wilson III, L.B., Phan, T.D., Hull, A.J., Amano, T., Hoshino, M., Argall, M.R., Le Contel, O., Agapitov, O., Gershman, D.J., et al. 2017 Electron scattering by high-frequency whistler waves at Earth's bow shock. Astrophys. J. Lett. 842, L11.CrossRefGoogle Scholar
Omura, Y., Furuya, N. & Summers, D. 2007 Relativistic turning acceleration of resonant electrons by coherent whistler mode waves in a dipole magnetic field. J. Geophys. Res. 112, 6236.Google Scholar
Omura, Y., Matsumoto, H., Nunn, D. & Rycroft, M.J. 1991 A review of observational, theoretical and numerical studies of VLF triggered emissions. J. Atmos. Terr. Phys. 53, 351368.CrossRefGoogle Scholar
Omura, Y., Miyashita, Y., Yoshikawa, M., Summers, D., Hikishima, M., Ebihara, Y. & Kubota, Y. 2015 Formation process of relativistic electron flux through interaction with chorus emissions in the Earth's inner magnetosphere. J. Geophys. Res. 120, 95459562.CrossRefGoogle Scholar
Page, B., Vasko, I.Y., Artemyev, A.V. & Bale, S.D. 2021 Generation of high-frequency whistler waves in the Earth's quasi-perpendicular bow shock. Astrophys. J. Lett. 919 (2), L17.CrossRefGoogle Scholar
Schulz, M. & Lanzerotti, L.J. 1974 Particle Diffusion in the Radiation Belts. Springer.CrossRefGoogle Scholar
Shapiro, V.D. & Sagdeev, R.Z. 1997 Nonlinear wave–particle interaction and conditions for the applicability of quasilinear theory. Phys. Rep. 283, 4971.CrossRefGoogle Scholar
Shi, X., Liu, T.Z., Angelopoulos, V. & Zhang, X.-J. 2020 Whistler mode waves in the compressional boundary of foreshock transients. J. Geophys. Res. 125 (8), e27758.CrossRefGoogle Scholar
Shklyar, D.R. 2021 A theory of interaction between relativistic electrons and magnetospherically reflected whistlers. J. Geophys. Res. 126 (2), e28799.CrossRefGoogle Scholar
Shklyar, D.R. & Matsumoto, H. 2009 Oblique whistler-mode waves in the inhomogeneous magnetospheric plasma: resonant interactions with energetic charged particles. Surv. Geophys. 30, 55104.CrossRefGoogle Scholar
Solovev, V.V. & Shkliar, D.R. 1986 Particle heating by a low-amplitude wave in an inhomogeneous magnetoplasma. Sov. Phys. JETP 63, 272277.Google Scholar
Stix, T.H. 1962 The Theory of Plasma Waves. McGraw-Hill.Google Scholar
Summers, D. & Omura, Y. 2007 Ultra-relativistic acceleration of electrons in planetary magnetospheres. Geophys. Res. Lett. 34, 24205.CrossRefGoogle Scholar
Tao, X., Bortnik, J., Albert, J.M., Liu, K. & Thorne, R.M. 2011 Comparison of quasilinear diffusion coefficients for parallel propagating whistler mode waves with test particle simulations. Geophys. Res. Lett. 38, 6105.CrossRefGoogle Scholar
Tao, X., Bortnik, J., Albert, J.M. & Thorne, R.M. 2012 a Comparison of bounce-averaged quasi-linear diffusion coefficients for parallel propagating whistler mode waves with test particle simulations. J. Geophys. Res. 117, 10205.Google Scholar
Tao, X., Bortnik, J., Albert, J.M., Thorne, R.M. & Li, W. 2013 The importance of amplitude modulation in nonlinear interactions between electrons and large amplitude whistler waves. J. Atmos. Sol.-Terr. Phys. 99, 6772.CrossRefGoogle Scholar
Tao, X., Bortnik, J., Thorne, R.M., Albert, J.M. & Li, W. 2012 b Effects of amplitude modulation on nonlinear interactions between electrons and chorus waves. Geophys. Res. Lett. 39, 6102.CrossRefGoogle Scholar
Tao, X., Zonca, F., Chen, L. & Wu, Y. 2020 Theoretical and numerical studies of chorus waves: a review. Sci. China Earth Sci. 63 (1), 7892.CrossRefGoogle Scholar
Thorne, R.M., Bortnik, J., Li, W. & Ma, Q. 2021 Wave–Particle Interactions in the Earth's Magnetosphere, chap. 6, pp. 93108. American Geophysical Union (AGU).Google Scholar
Thorne, R.M., Li, W., Ni, B., Ma, Q., Bortnik, J., Chen, L., Baker, D.N., Spence, H.E., Reeves, G.D., Henderson, M.G., et al. 2013 Rapid local acceleration of relativistic radiation-belt electrons by magnetospheric chorus. Nature 504, 411414.CrossRefGoogle ScholarPubMed
Tong, Y., Vasko, I.Y., Artemyev, A.V., Bale, S.D. & Mozer, F.S. 2019 Statistical Study of Whistler Waves in the Solar Wind at 1 au. Astrophys. J. 878 (1), 41.CrossRefGoogle Scholar
Trakhtengerts, V.Y. & Rycroft, M.J. 2008 Whistler and Alfvén Mode Cyclotron Masers in Space. Cambridge University Press.CrossRefGoogle Scholar
Vainchtein, D., Zhang, X.J., Artemyev, A.V., Mourenas, D., Angelopoulos, V. & Thorne, R.M. 2018 Evolution of electron distribution driven by nonlinear resonances with intense field-aligned chorus waves. J. Geophys. Res. 123 (10), 81498169.CrossRefGoogle Scholar
Vedenov, A.A., Velikhov, E. & Sagdeev, R. 1962 Quasilinear theory of plasma oscillations. Nucl. Fusion Suppl. 2, 465475.Google Scholar
Verscharen, D., Chandran, B.D.G., Boella, E., Halekas, J., Innocenti, M.E., Jagarlamudi, V.K., Micera, A., Pierrard, V., Štverák, Š., Vasko, I.Y., et al. 2022 Electron-driven instabilities in the solar wind. Front. Astron. Space Sci. 9, 951628.CrossRefGoogle Scholar
Wilson, L.B., Koval, A., Szabo, A., Breneman, A., Cattell, C.A., Goetz, K., Kellogg, P.J., Kersten, K., Kasper, J.C., Maruca, B.A., et al. 2013 Electromagnetic waves and electron anisotropies downstream of supercritical interplanetary shocks. J. Geophys. Res. 118, 516.CrossRefGoogle Scholar
Yao, S.T., Shi, Q.Q., Zong, Q.G., Degeling, A.W., Guo, R.L., Li, L., Li, J.X., Tian, A.M., Zhang, H., Yao, Z.H., et al. 2021 Low-frequency whistler waves modulate electrons and generate higher-frequency whistler waves in the solar wind. Astrophys. J. 923 (2), 216.CrossRefGoogle Scholar
Zhang, X.-J., Artemyev, A., Angelopoulos, V., Tsai, E., Wilkins, C., Kasahara, S., Mourenas, D., Yokota, S., Keika, K., Hori, T., et al. 2022 Superfast precipitation of energetic electrons in the radiation belts of the Earth. Nat. Commun. 13, 1611.CrossRefGoogle ScholarPubMed
Zhang, X., Angelopoulos, V., Artemyev, A.V. & Liu, J. 2018 a Whistler and electron firehose instability control of electron distributions in and around dipolarizing flux bundles. Geophys. Res. Lett. 45, 93809389.CrossRefGoogle Scholar
Zhang, X.J., Agapitov, O., Artemyev, A.V., Mourenas, D., Angelopoulos, V., Kurth, W.S., Bonnell, J.W. & Hospodarsky, G.B. 2020 a Phase decoherence within intense chorus wave packets constrains the efficiency of nonlinear resonant electron acceleration. Geophys. Res. Lett. 47 (20), e89807.CrossRefGoogle Scholar
Zhang, X.J., Demekhov, A.G., Katoh, Y., Nunn, D., Tao, X., Mourenas, D., Omura, Y., Artemyev, A.V. & Angelopoulos, V. 2021 Fine structure of chorus wave packets: comparison between observations and wave generation models. J. Geophys. Res. 126 (8), e29330.Google Scholar
Zhang, X.J., Mourenas, D., Artemyev, A.V., Angelopoulos, V., Bortnik, J., Thorne, R.M., Kurth, W.S., Kletzing, C.A. & Hospodarsky, G.B. 2019 Nonlinear electron interaction with intense chorus waves: statistics of occurrence rates. Geophys. Res. Lett. 46 (13), 71827190.CrossRefGoogle Scholar
Zhang, X.J., Mourenas, D., Artemyev, A.V., Angelopoulos, V., Kurth, W.S., Kletzing, C.A. & Hospodarsky, G.B. 2020 b Rapid frequency variations within intense chorus wave packets. Geophys. Res. Lett. 47 (15), e88853.Google Scholar
Zhang, X.J., Thorne, R., Artemyev, A., Mourenas, D., Angelopoulos, V., Bortnik, J., Kletzing, C.A., Kurth, W.S. & Hospodarsky, G.B. 2018 b Properties of intense field-aligned lower-band chorus waves: implications for nonlinear wave–particle interactions. J. Geophys. Res. 123 (7), 53795393.CrossRefGoogle Scholar
Zhang, Y., Matsumoto, H., Kojima, H. & Omura, Y. 1999 Extremely intense whistler mode waves near the bow shock: geotail observations. J. Geophys. Res. 104, 449462.CrossRefGoogle Scholar
Figure 0

Figure 1. Examples of typical wave packets of whistler-mode waves captured by THEMIS spacecraft (Angelopoulos 2008) in the Earth bow shock (a), foreshock transient (b), outer radiation belt (c) and plasma injection region (d). These events are picked up from statistics published in Artemyev et al. (2022), Shi et al. (2020), Zhang et al. (2018b) and Zhang et al. (2018a).

Figure 1

Figure 2. A set of trajectories obtained by the numerical integration of Hamiltonian equations for a system (2.10a,b). Results are shown for (a) low-amplitude wave with diffusive scattering, (b) coherent high-amplitude wave with trapping and bunching and (c) modulated high-amplitude wave with diffusive-like scattering. System parameters are: electron energy $E_0=100$ keV, equatorial pitch angle $\alpha _0 = 40^\circ$, number of particles $N = 100$, wave amplitude $B_w = 5$ pT (a), $B_w = 500$ pT (b,c). System parameters correspond to $L\text {-shell} = 6$, whistler-mode waves with frequency equal to $0.35$ of the electron cyclotron frequency at the equator, and the constant plasma frequency equal to $10$ times the electron cyclotron frequency at the equator.

Figure 2

Figure 3. Distributions of the energy change $\Delta E$ for different $B_{{\rm eff}}$ with $h=1$ and $l=20$ for coherent (ac) and modulated (df) waves with $\Delta E \propto B_w$ fitting (black dashed lines) and with $\langle {(\Delta E)^2}\rangle ^{1/2}$ profiles (black dots): (a,d) $E_0 = 100$ keV, $\alpha _0 = 40^{\circ }$; (b,e) $E_0 = 100$ keV, $\alpha _0 = 60^{\circ }$; (cf) $E_0 = 300$ keV, $\alpha _0 = 60^{\circ }$. For each $B_w$ we use $10^4$ trajectories to evaluate the $\Delta E$ distribution, each particle resonates with the wave only once and $\Delta E$ is the energy change for a single resonance, initial particle phases and wave phases are random and thus for a modulated case the actual wave amplitude is different for different particles having the same energy/pitch angle. For each $B_{{\rm eff}}$ the $\Delta E$ distribution is normalized to one. System parameters correspond to $L\text {-shell} = 6$, whistler-mode waves with frequency equal to $0.35$ of the electron cyclotron frequency at the equator, and the constant plasma frequency equal to $10$ times the electron cyclotron frequency at the equator.

Figure 3

Figure 4. (a) Profiles of $f(a,\xi )$ for three $a$ values. (b) Profiles of $\langle f(a,\xi )\rangle _{\xi }$ and $\langle (f(a,\xi ))^2\rangle _{\xi }$.

Figure 4

Figure 5. Probability distributions of $\xi$ for $10^5$ trajectories and (a) $E_0 = 100$ keV, $\alpha _0 = 40^{\circ }$, $B_w = 50$ pT, (b) $E_0 = 100$ keV, $\alpha _0 = 40^{\circ }$, $B_w = 500$ pT, (c) $E_0 = 300$ keV, $\alpha _0 = 60^{\circ }$, $B_w = 500$ pT.

Figure 5

Figure 6. Two-dimensional energy/pitch-angle maps of $(E_0,\alpha _{0})$ for (a) $B_w = 100$ pT, (b) $B_w = 500$ pT and (c) $B_w = 1000$ pT. Black curve shows $a=1$.

Figure 6

Figure 7. (a) Probability distribution functions ${\mathcal {F}}(B_w/B_0)$ of whistler-mode wave intensities from Zhang et al. (2019) (blue) and Zhang et al. (2020b) (red). (b,c) Two-dimensional distributions of $\langle \langle (\Delta \gamma )^2\rangle _{\xi }/\langle (\Delta \gamma )^2\rangle _{\xi }^{QL} \rangle _{\varepsilon }$ in $(E_0,\alpha _{0})$ space for the two ${\mathcal {F}}(B_w/B_0)$ distributions from (a).

Figure 7

Figure 8. A schematic view of diffusion rate scaling with wave amplitude normalized to the amplitude threshold for nonlinear resonant interactions, $B_w^{*}$. Quasi-linear diffusion rate $D_{QL}\propto B_w^2$ works only for $B_w/B_w^{*}<1$. For $B_w/B_w^{*}>1$ and incoherent waves should work $D\propto B_w$, whereas for coherent waves should work $D_{NL}\propto \sqrt {B_w}$ (see Artemyev et al.2021a).

Figure 8

Figure 9. Comparison between $\langle f^2(a,\xi ) \rangle _{\xi }$ and $\langle f(a,\xi ) \rangle _{\xi }^2$ as functions of $a$.