• Matter and Radiation at Extremes
  • Vol. 8, Issue 1, 014402 (2023)
A. S. Samsonova), E. N. Nerush, and I. Yu. Kostyukov
Author Affiliations
  • Institute of Applied Physics of the Russian Academy of Sciences, 46 Ulyanov St., Nizhny Novgorod 603950, Russia
  • show less
    DOI: 10.1063/5.0117504 Cite this Article
    A. S. Samsonov, E. N. Nerush, I. Yu. Kostyukov. High-order corrections to the radiation-free dynamics of an electron in the strongly radiation-dominated regime[J]. Matter and Radiation at Extremes, 2023, 8(1): 014402 Copy Citation Text show less

    Abstract

    A system of reduced equations is proposed for electron motion in the strongly radiation-dominated regime for an arbitrary electromagnetic field configuration. The approach developed here is used to analyze various scenarios of electron dynamics in this regime: motion in rotating electric and magnetic fields and longitudinal acceleration in a plane wave and in a plasma wakefield. The results obtained show that this approach is able to describe features of electron dynamics that are essential in certain scenarios, but cannot be captured in the framework of the original radiation-free approximation [Samsonov et al., Phys. Rev. A 98, 053858 (2018) and A. Gonoskov and M. Marklund, Phys. Plasmas 25, 093109 (2018)]. The results are verified by numerical integration of the nonreduced equations of motion with account taken of radiation reaction in both semiclassical and fully quantum cases.

    I. INTRODUCTION

    With upcoming laser facilities such as ELI,1 SULF,2 SEL,3 and XCELS,4 investigation of laser–matter interactions in the regime of extreme laser intensity will become feasible. Radiation reaction is expected to accompany such interactions, although its direct impact is usually quite hard to predict. It has been known for more than a century that a charged particle experiences a recoil force when radiating, but a consistent model describing this phenomenon in both theoretical and numerical studies has yet to be firmly established. Recently conducted experiments aimed at determining the correct model of radiation friction are still subject to a certain level of ambiguity5,6 and thus have not been able to solve this problem. The problem is becoming more and more acute with the growing number of studies discovering possible new effects caused by radiation reaction. It is clear that these effects vary greatly and include, among many others, alterations in particle acceleration mechanisms,7–16 highly efficient laser pulse absorption,17 relativistic transparency reduction,18,19 the inverse Faraday effect,20–22 particle polarization,23–31 initiation of quantum electrodynamic (QED) cascades.32–47 The signatures of these effects are expected to be most prominent in the so-called radiation-dominated regime, i.e., the regime in which radiative losses of charged particles are comparable to the energy gain in the electromagnetic (EM) field. Estimates show that the field amplitudes needed for realization of this regime can be achieved experimentally, either at future laser facilities such as ELI, SEL, and XCELS, or at future accelerators such as FACET-II.48

    Via QED, one can describe radiation reaction self-consistently and calculate the probability of radiation of a photon with a given energy. While a full QED description currently provides the most accurate description of radiation reaction, it is not usually applicable to practical problems involving complex light–matter interactions, since via QED one calculates scattering probabilities between some stationary (commonly Volkov) electron states. To describe a dynamic problem where these states evolve owing to the evolution of the EM fields, nonstationary Dirac equations have to be solved, which is usually either unfeasible or impractical. However, under some conditions, this is not necessary, since the problem can be significantly simplified. The first main parameter that defines such a condition is the dimensionless EM field amplitude a0:a0=eE0mcω,where m and e > 0 are respectively the electron mass and charge, and E0 and ω are respectively the characteristic strength and frequency of the EM field. In the regime a0 ≫ 1, the characteristic radiation formation length λf can in most cases be estimated as λ/a0λ, where λ = 2πc/ω, i.e., individual acts of radiation occur almost instantaneously compared with the scale of variation of the EM fields, and thus these fields can be assumed constant on a radiation formation length. In this locally constant field approximation (LCFA49–51), the total radiation probabilities and the shapes of emission spectra depend only on the QED parameter χ,χ=γES(E+v×B)2(vE)2,where γ and v are respectively the electron Lorentz factor and velocity normalized to c, E and B are respectively the electric and magnetic fields, ES = m2c3/eℏ is the critical Sauter–Schwinger field,50 and is Planck’s constant. These probabilities can be calculated analytically in either classical (χ ≪ 1) or quantum (χ ≫ 1) regimes:Wradαmc2γ×1.4χ,χ1,0.7χ2/3,χ1,where α = e2/ℏc is the fine structure constant. Note that a number of different approaches have been proposed for calculation of radiation probabilities when the LCFA is no longer valid.52–56 In LCFA, the characteristic distance traveled by an ultrarelativistic electron between two consecutive photon emissions, λW, can be estimated as c/Wrad, which in both classical and quantum cases is at least 1/α ≈ 137 times longer than the radiation formation length. Note, however, that the above-mentioned estimate for the radiation formation length actually depends on the frequency of the emitted radiation57 and can be inaccurate for χ ≳ 10. Furthermore, the ratio between the mean free path λW and the EM field wavelength can be estimated asλWλ1αa0×1,χ1,χ1/3,χ1.So, as χ ≲ 10 for most experiments that are likely to be performed in the near future, and for a0 ≫ 137, the hierarchy of the characteristic scales of the problem is as follows:λfλWλ,which means that the electron moves classically between short but frequent acts of photon emission.

    In that case, we can approximate the effect of radiation recoil as an additional continuous force acting on a particle, i.e., the equations of motion take the formdpdt=Ev×BFrrv,dγdt=vEFrrv2,where the electron momentum p is normalized to mc, the time t is normalized to 1/ω, and the electric and magnetic fields are normalized to mcω/e. In Eqs. (6) and (7), Frr is the total radiation power normalized to mc2ω and is given byFrr=αaS33π04u3+5u2+4u(1+u)4K2/32u3χdu,where aS = eES/mcωmc2/ℏω is the normalized Sauter–Schwinger field. In the limiting cases, this expression simplifies toFrrαaS×0.67χ2,χ1,0.37χ2/3,χ1.This approach to the description of the electron dynamics with account of radiation reaction is commonly referred to as semiclassical.58–62 In the quantum regime, radiated photons can carry away a significant portion of the electron energy, and as this radiation is stochastic, electrons contained in a small phase volume can significantly diverge in phase space after some time. Equations (6) and (7) essentially describe the zeroth moment, i.e., the trajectory of the center of mass of the electron distribution function, while effects caused by the probabilistic nature of radiation, such as straggling and quenching,62–65 lead to diffusion of the distribution function and thus cannot be captured using this approach. In that case, a more accurate description requires equations for higher moments of the distribution function. Such an approach was applied to calculate the mean particle energy and energy spread in different fields configurations in Refs. 61, 66, and 67. If, on the contrary, χ ≲ 1, then the recoil from a single-photon radiation is small, and the approximation of continuous recoil is sufficient to describe the electron dynamics.

    Another important consideration in studying the effect of radiation reaction is its dependence on the internal degree of freedom of the electron, i.e., spin. Strictly speaking, the quasiclassical limit of the Dirac equation leads to equations of motion where both the orbital motion of the electron and the evolution of its spin are coupled. In particular, one should add the Stern–Gerlach force68 to the equation for the electron momentum and describe the spin dynamics via the Thomas–Bargmann–Michel–Telegdi (T-BMT)69,70 equation. Note that although the latter is strictly valid only in homogeneous EM fields, it can still be used in heterogeneous fields if the Stern–Gerlach force can be neglected.71 The ratio between the Lorentz force and Stern–Gerlach force can be estimated to be of the order of ℏω/mc2, and thus for optical frequencies the latter can be neglected with a large margin of accuracy. In that case, spin dynamics is decoupled from electron orbital motion and can be calculated after the electron trajectory has been found. Radiation reaction can again couple spin dynamics and electron orbital motion, since radiation probabilities depend on the spin of the electron (and the polarization of the emitted photon). Note that the order-of-magnitude estimates made above where radiation probabilities are averaged over the initial and summed over the final polarization states of the electron remain valid. However, in certain scenarios, the assumption that electrons are generally not polarized may no longer hold, since the radiation probabilities of spin-up and spin-down electrons are different. Resolving electron polarization can lead to effects such as a significant increase in pair production during QED cascade development,72 production of polarized high-energy particles,28,73 and spatially inhomogeneous polarization.31 In this paper, such effects caused by spin dynamics are not covered.

    Previous studies have shown that the problem can be simplified even further. In particular, to some extent, radiation reaction can be accounted for implicitly, i.e., without specifying an expression for radiation reaction in the equations of motion.74,75 This is done by noticing that in constant homogeneous EM fields, electron motion is stable if the electron does not experience transverse acceleration. As the radiation probability depends on the parameter χ, which is essentially proportional to the transverse acceleration, the direction of such motion is called the radiation-free direction (RFD). As this direction corresponds to vanishing of the transverse acceleration, and radiation recoil is directed against the electron velocity, to find the RFD, one does not need to specify any expression for the radiation probability at all. The timescale τv on which the electron velocity approaches the RFD in constant fields is of the order of γmc/eE0. If the EM fields are varying with characteristic frequency ω, then the RFD defined by the local and instantaneous field configuration changes on the same timescale. In the absence of radiation reaction, one can estimate that γa0, and thus, by the time the electron velocity approaches the RFD, the latter itself changes, and so the geometric relation between the electron velocity and the RFD is arbitrary. This is not the case in the strongly radiation-dominated regime, however, when, by definition, γa0, and thus the EM field orients the electron velocity much faster than the field itself changes, and so the electron velocity quickly aligns to the RFD defined by the local and instantaneous electric and magnetic fields. Thus, to approximately determine the electron trajectory, one can assume that at each time instant, the electron velocity coincides with the RFD. While this approach allows one to describe the dynamics of the electron in the strongly radiation-dominated regime without specifying an expression for the radiation power, it is quite limited for a couple of reasons. First, the electron velocity converges to the RFD sufficiently fast only at extremely large intensities I ≳ 1025 W/cm2. Second, this approach does not allow one to find the electron energy and radiation losses as the RFD is approached, since the particle energy is assumed to be indefinitely large, albeit much smaller than the field amplitude at the same time. Despite its apparent drawbacks, this approach has recently been applied successfully to describe electron motion in an astrophysical environment.76 In this paper, we extend this radiation-free approach to overcome its inherent problems and to describe the dynamics of an electron in the strongly radiation-dominated regime more precisely.

    The remainder of the paper is organized as follows. In Sec. II, we reintroduce the concept of radiation-free dynamics and extend it by application of perturbation theory. In Sec. III, we consider several EM field configurations where the reduced equations of motion obtained can be explicitly solved. In Sec. IV, we discuss the domain of applicability of the proposed approach and draw conclusions.

    II. RADIATION-FREE APPROACH

    Let us start by introducing a radiation-free approach to the description of electron dynamics, loosely following the original papers.74,75 The equations governing electron dynamics in an EM field with account taken of radiation reaction can be written in terms of the electron velocity v and Lorentz factor γ:dγdt=vEFrrv2,dvdt=1γE+v×Bv(vE)+Frrvγ2.Since radiation reaction is sufficient to significantly alter dynamics only for ultrarelativistic particles (γ ≫ 1), the last term in Eq. (11) can be omitted. There exists a formal stationary solution v0 of these equations, corresponding to vanishing of the transverse force acting on the electron and in turn vanishing of radiation reaction. Because of that property, this solution is called the radiation-free direction (RFD):E+v0×Bv0(v0E)=0.Note, first, that there always exists a solution to this equation that can be calculated algebraically74 or geometrically75 and, second, that |v0| = 1, as can be shown by taking the scalar product of Eq. (12) with v0, which means that this solution is not entirely physical, i.e., an electron is unable to move in the EM fields without experiencing transverse acceleration. To understand how this solution relates to the actual solution of the equations of motion in the strongly radiation-dominated regime, let us consider the following. By definition, in the strongly radiation-dominated regime, the energy of an electron is significantly smaller than that of a hypothetical electron that is in the same EM fields but does not experience radiation reaction. One can roughly estimate that the characteristic energy of the latter electron is of the order of the dimensionless electric field amplitude E. Therefore, for a real electron in the radiation-dominated regime, we can assume that γE. Under this assumption, Eq. (11) then states that the EM fields orient the electron velocity much faster than they themselves change. So, on the timescale of velocity orientation, the EM fields can be assumed constant and homogeneous. In that case, the electron velocity tends asymptotically to the RFD. Neglecting the time taken for the electron velocity to approach the RFD, one can construct an asymptotic trajectory that in some sense serves as an attractor for real electron trajectories:drdt=v0(E(r,t),B(r,t)).As mentioned in Sec. I, although Eqs. (12) and (13) describe electron dynamics in the strongly radiation-dominated regime with radiation reaction being accounted for implicitly, there are two drawbacks to this approach. First, these asymptotic trajectories describe real particle trajectories well only at extremely large intensities I ≳ 1025 W/cm2. This is because, for most field configurations, the characteristic time at which the electron velocity approaches the RFD is underestimated by above reasoning. Second, this approach does not allow one to find the electron energy and radiation losses on approaching this asymptotic trajectory, since the particle energy is assumed to be indefinitely large (albeit much smaller than the field amplitude at the same time).

    To deal with these problems, we develop a perturbation theory, assuming that the electron velocity deviates from the RFD but that this deviation is small, i.e.,v=1δ22v0+v1,where v1 ⊥ v0, and δ can be found from the condition that |v|2 = 1 − γ−2, from which we getδ2v12+γ2.Substituting this into Eqs. (10) and (11), utilizing Eq. (12) and keeping only terms of first order in v1 (see the expansion up to terms of second order in Appendix A), we obtain a set of general equations governing electron dynamics in the strongly radiation-dominated regime:dv1dt=F1γdv0dtv0v1dv0dt,dγdt=v0E1v12212γ2v1EFrr(χ),F1=(v0E)v1+(v0B)[v0×v1]+O(δ2),χ=γ|F1|aS.Note that although χ is proportional to a small term v1, it can be arbitrarily large owing to the factor γ, and therefore the term Frr should be kept at all expansion orders, as a consequence of which the equations of motion remain nonlinear. This does not contradict the fact that the last term in Eq. (11) is smaller than the first three, which are equal to F1, and thus it can be omitted. Indeed, |Frrv/γ2| ≲ αaSχ2/γ2 = α|F1|2/aSαEv1|F1|/aS ⋘ |F1|, since EaS, α ≪ 1, v1 ≲ 1. The full time derivatives should be considered as derivatives of the vector field v0 along the real electron trajectory r(t), i.e.,dv0dt=v0t+(v)v0.Let us consider the equation for the magnitude of the vector v1:12dv12dt=v1dv0dt+v12(v0E)γ.In a constant EM fields (dv0/dt = 0), one can estimate the characteristic timescale on which the electron velocity approaches the RFD:τv=γ|v0E|γa0.However, for varying EM fields, the sign of the first term in Eq. (21) can be arbitrary and its magnitude can be as large as v1, and so the condition γa0 alone is not enough to justify the description of the electron dynamics by Eq. (13) in an arbitrary field configuration. Instead, one should use the set of Eqs. (16) and (17), where the variation of the RFD is taken into account. Moreover, these equations allow one to find the electron energy and radiative losses.

    The procedure to obtain the reduced equations of motion can be explained in few simple steps. First, it is shown that there exists a preferred RFD that the electron velocity approaches in constant EM fields. By decomposing the electron velocity in a new basis in which one axis coincides with the RFD, the equations of motion can be split. Motion along the RFD is essentially described via the particle energy, while the equations for the transverse velocity can be expanded in series, which clearly converge, since the magnitude of the velocity is strictly smaller than unity. Although the final set of equations remain nonlinear and cannot be solved explicitly in an arbitrary field configuration, the examples considered below show that this approach can be superior to solving the nonreduced Newton equations. It is worth noting that recently, in Ref. 77, a similar decomposition of the velocity vector was used to explore equilibrium solutions of Eqs. (16) and (17).

    III. EXAMPLE PROBLEMS

    Below, we consider several examples of field configurations in which Eqs. (16) and (17) can be solved explicitly.

    A. Generalized Zeldovich problem

    The equations of electron motion with a radiation reaction force can be integrated analytically for a rotating uniform electric field, as was first demonstrated by Zeldovich.78 Recently Zeldovich’s solution has been extended to a configuration with rotating electric and magnetic fields that are parallel to each other.37 Let us analyze the latter configuration within our approach. We assume that the electric and magnetic fields are uniform and parallel, and rotate with velocity Ω. The RFD in this configuration is antiparallel to the electric field: v0 = −E/E = −e. Let us consider a stationary solution in which the deviation vector v1 rotates synchronously with the electric and magnetic fields. In this case, all the time derivatives can be replaced by the cross product Ω×. Equation (16) then becomesΩ×v1=Eγv1+Bγe×v1+Ω×e+v1e.Keeping in mind that v1 ⊥ v0, we can express v1 as follows:v1=vΩ×e+vxΩ.Equation (23) can then be split into a set of linear equations, the solution of which is easily found:vx=γBE2+B2,v=γEE2+B2.This stationary solution corresponds to constant radiative losses and constant energy. Consequently, the final relation between the electron energy and the EM fields can be obtained from Eq. (17):E=Frr(χ)=Frrγ2aS.This result coincides exactly with the result obtained in Ref. 37 and, in the special case B = 0, with Zeldovich’s original solution.78 A comparison of the solution obtained here with numerical solutions of the nonreduced equations of motion (10) and (11) for both semiclassical and quantum approaches to radiation reaction is presented in Fig. 1. The magnitude of the EM fields has been chosen in such a way that the average value of the χ parameter of an electron is around 5. This has been done deliberately to show where the assumptions made in our approach start to break down. In particular, our analytical solution and the quasiclassical numerical solutions assume that the χ parameter does not deviate significantly between particles, and thus it is assumed that ⟨Frr(χ)⟩ ≈ Frr(⟨χ⟩), where the angular brackets denote averaging over the particle distribution function. However, as can be seen from Fig. 1, the parameters of individual electrons with the same initial conditions spread quite significantly, owing to the stochastic nature of radiation in the quantum regime, as mentioned in Sec. I. Because of this and the nonlinearity of Frr, in fact ⟨Frr(χ)⟩ < Frr(⟨χ⟩), which explains the discrepancy in Fig. 1 between the values averaged over quantum solutions and those obtained analytically or via numerical solution in the quasiclassical approximation.

    Electron dynamics in an electric field with dimensionless amplitude eE/mcΩ = 2500 and a parallel magnetic field with dimensionless amplitude eB/mcΩ = 2000 rotating with angular frequency Ω, corresponding to the wavelength λ = 1 µm: (a) component of electron velocity transverse to the electric field; (b) component of electron velocity along the angular velocity vector Ω. Orange and cyan lines correspond to numerical solution of the nonreduced equations of motion (10) and (11) with radiation reaction taken into account via semiclassical and quantum approaches, respectively. Blue lines correspond to the average value of 100 “quantum” solutions. Black dashed lines correspond to the analytical solution (25) and (26).

    Figure 1.Electron dynamics in an electric field with dimensionless amplitude eE/mcΩ = 2500 and a parallel magnetic field with dimensionless amplitude eB/mcΩ = 2000 rotating with angular frequency Ω, corresponding to the wavelength λ = 1 µm: (a) component of electron velocity transverse to the electric field; (b) component of electron velocity along the angular velocity vector Ω. Orange and cyan lines correspond to numerical solution of the nonreduced equations of motion (10) and (11) with radiation reaction taken into account via semiclassical and quantum approaches, respectively. Blue lines correspond to the average value of 100 “quantum” solutions. Black dashed lines correspond to the analytical solution (25) and (26).

    B. Monochromatic linearly polarized plane wave

    Some interesting results can be obtained if our approach is applied to electron motion in a plane wave. In this configuration, the RFD coincides with the direction of the Poynting vectorv0=E×BE2,where both E and B are functions of the phase φ = xt. For simplicity, let us assume that initially the deviation vector v1 is parallel to the electric field, in which case it will remain so at any time instant. It is also more convenient to write down equations in terms of momentum p1 = γv1 and phase φ. Since χ oscillates with constant amplitude in a plane wave without account of radiation reaction and radiation leads only to decrease χ, eventually any electron will reach the classical regime when χ ≪ 1, and so we will only consider electron dynamics in that regime. In this case, Eqs. (16) and (17) take the formdγdφ=2pγE1+p22ArrE2(1+p2),dpdφ=E2ArrE2pγ(1+p2),where Arr = α/6ES. Note that in a plane wave configuration, F1 as defined in Eq. (18) contains only terms of second and higher orders of smallness in v1, the derivation of which can be found in Appendix A. Let us start solving Eqs. (29) and (30) without account of radiation reaction for a linearly polarized monochromatic plane wave, i.e., assuming Arr = 0 and E = a0 cos φ. In this case, it is easy to obtain the following solution:ppw=a0sinφ,γpw=γ0(1+ppw2),where it has been assumed that initially the electron momentum has only a component along the direction of plane wave propagation. Note that the exact solution of the electron equations of motion in a linearly polarized plane wave (see, e.g., Ref. 79) coincides exactly with the solution given by Eqs. (31) and (32) in the limit γ0 ≫ 1. Thus, our method, which is essentially a series expansion, can be applied even for certain problems where radiation reaction is not taken into account.

    When Arr ≠ 0, it can be shown from Eqs. (29) and (30) that there is an asymmetry in the particle motion in the accelerating and decelerating phases. This leads to a nonzero energy gain in a single period. Under several additional assumptions, we can obtain the following solution (see the detailed step-by-step solution in Appendix B):γγ0(1+a02sin2φ)1+Arra02γ0φ,va0sinφγ.Rewriting the solution in terms of laboratory time t yields the following resultχ,v1,γ1(Arrt)1/3.Note that the average behavior of the dependences in Eq. (35) asymptotically coincides with that extracted from the exact solution derived in Ref. 80.

    In obtaining the above solution, we have assumed that at the initial moment, the transverse momentum of the particle is equal to zero and the longitudinal momentum is sufficiently large and positive. To obtain a solution with arbitrary initial conditions, one can perform a Lorentz boost to an auxiliary reference frame where the above assumptions are satisfied and then transform the solution obtained there back to an initial reference frame. The solution shown in Fig. 2 was obtained in such a way, with the auxiliary reference frame taken to be moving with a velocity corresponding to a Lorentz factor 1000 along the negative x axis in the laboratory reference frame, and so the initial longitudinal electron momentum in the auxiliary reference frame was approximately 5mc.

    Dynamics of an electron with initial momentum px = −100 mc in a plane wave with amplitude a0 = 500 and wavelength λ = 1 µm propagating along the x axis: (a) energy of electron normalized to its initial value; (b) maximum value of QED parameter χ: a0(γ − px)/ES. Red lines correspond to the classical solution without radiation reaction. Orange and blue lines correspond to numerical solution of the nonreduced equations of motion (10) and (11) with radiation reaction taken into account via semiclassical and quantum approaches, respectively. Black dashed lines correspond to the analytical solution (33) and (34). Note that for visual clarity, the black dashed line only depicts the amplitude of the oscillations of γ in (a).

    Figure 2.Dynamics of an electron with initial momentum px = −100 mc in a plane wave with amplitude a0 = 500 and wavelength λ = 1 µm propagating along the x axis: (a) energy of electron normalized to its initial value; (b) maximum value of QED parameter χ: a0(γpx)/ES. Red lines correspond to the classical solution without radiation reaction. Orange and blue lines correspond to numerical solution of the nonreduced equations of motion (10) and (11) with radiation reaction taken into account via semiclassical and quantum approaches, respectively. Black dashed lines correspond to the analytical solution (33) and (34). Note that for visual clarity, the black dashed line only depicts the amplitude of the oscillations of γ in (a).

    The obtained solution is not only nonperiodic, but also features quite unexpected behavior: instead of slowing down the electron, radiation reaction actually allows it to gain infinite energy (in infinite time, obviously). Although this behavior has been reported before78,80–84 and has been confirmed by numerical solution of the nonreduced equations of motion (10) and (11) (see Fig. 2), it does not appear to be widely acknowledged. Simple reasoning can explain this seemingly controversial phenomenon. For this, it is more convenient to resort to a quantum description of radiation reaction. In a relativistically strong plane wave (E ≫ 1), the formation length of the radiation can be estimated as λ/Eλ, which can be interpreted as indicating that an electron moves classically between short acts of photon emission. Without radiation reaction, the light-front momentum γpx is a constant of motion, where px is the electron momentum along the direction of plane wave propagation [see the red line in Fig. 2(b)]. The radiation probability depends on the QED parameter χ, which in the plane wave configuration is given byχ=E(φ)ES(γpx).As the radiation formation length is much smaller than the wavelength, the EM fields can be assumed constant during a single act of photon emission, and it follows from energy and momentum conservation that the parameter χ of the electron strictly decreases after this act of emission [see the distinct jumps corresponding to emission of individual photons in the blue line in Fig. 2(b)]. We can therefore conclude that owing to radiation reaction, the light-front momentum γpx tends asymptotically to zero, which can be satisfied only when px (and correspondingly γ) grows indefinitely.

    C. Plasma accelerator

    Finally, let us consider a toy model of a plasma accelerator and derive conditions for a known stable solution in a radiation-dominated regime9,16 using our approach. For this, we assume that the EM fields are a uniform accelerating field z0Eacc and a linear focusing field yEfoc in which the electron undergoes betatron oscillations. To find a solution that corresponds to the averaged radiation losses being constant over many betatron periods, we will assume that any function of the QED parameter χ is a strictly periodic function of time, and thus the average of any function of χ is also constant, i.e.,dχ2dt=0,where χ2 is used for convenience. In the field configuration under consideration, from Eq. (19), we getχ=γEv1aS.Hereinafter, we will assume that γ ≈ ⟨γ⟩, i.e., oscillations of the particle energy are much smaller than the energy itself, and so we can extract γ from the angular brackets. Also, as the particle is accelerated, γ grows with time, while χ remains constant on average. This means that the oscillation amplitude of v1 decreases, and therefore at later times we can safely assume that the electric field experienced by the electron is mostly accelerating, so EEacc = const. Both of these assumptions are reliably confirmed by numerical simulations and remain valid in the final solution that we obtain below. Utilizing these assumptions and expanding Eq. (37) yields the following:Frrv12+γv1dv0dt=0.Differentiating this expression again, we obtain the following result:Frrγ(3Frr2Eacc)v12+γv1d2v0dt2dv0dt2=0.To deal with the last two terms in this expression, let us write the equation for the electron trajectory:drdt=v0+v1,wherev0=z0Eacc+yEfocEz0yEfocEaccz0κy.If we assume that the betatron oscillations are harmonic, i.e.,y=y0cosωt,then from the y component of Eq. (41) we getv1,y=y0(κcosωtωsinωt).To calculate the average of the last two terms in Eq. (40), we note that dv0/dt = −κdy/dt:v1d2v0dt2dv0dt2=y02ω2κκcos2ωtsin2ωtωsinωtcosωt=0.So, finally, in a model accelerator, we getFrr=23Eacc.Thus, on average, the particle is accelerated only at one-third of the classical rate of acceleration.9,16Figure 3 shows that the solution obtained here coincides quite well with the numerical solution of the nonreduced equations of motion (10) and (11). It should be noted that we have not used a specific expression for the power of radiative losses, although a more rigorous derivation of the relation (46) shows that the result actually depends on the scaling law of the radiation power with the parameter χ: for example, according to Ref. 16, in the fully quantum regime when χ ≫ 1 and Frrχ2/3, the relation (46) should in fact be slightly different, specifically,Frr=1219Eacc,although this is only 5% different from (46). However, reasoning using our approach cannot exactly reproduce this minor difference, owing to the approximations used throughout, in particular the neglect of terms proportional to 1/γ2 when calculating χ.

    Electron dynamics in a model accelerator with Eacc = 30 TV/m and Efoc growing linearly from 0 to 30 TV/m at displacement 0.1 µm: (a) average rate of acceleration; (b) average value of QED parameter χ. Time is normalized to the initial value of the inverse betatron frequency ωb/γ0 of the electron. Blue lines correspond to the solution of the nonreduced equations of motion (10) and (11) without account of radiation reaction Orange lines correspond to the solution with radiation reaction taken into account via a semiclassical approach. The black dashed line corresponds to the analytical solution (46).

    Figure 3.Electron dynamics in a model accelerator with Eacc = 30 TV/m and Efoc growing linearly from 0 to 30 TV/m at displacement 0.1 µm: (a) average rate of acceleration; (b) average value of QED parameter χ. Time is normalized to the initial value of the inverse betatron frequency ωb/γ0 of the electron. Blue lines correspond to the solution of the nonreduced equations of motion (10) and (11) without account of radiation reaction Orange lines correspond to the solution with radiation reaction taken into account via a semiclassical approach. The black dashed line corresponds to the analytical solution (46).

    IV. DISCUSSION AND CONCLUSION

    In this paper, we have developed an approach to tackle the problem of single-electron dynamics in arbitrary EM fields in the strongly radiation-dominated regime. We have shown that the electron velocity approaches a certain direction, moving along which an electron does not experience transverse acceleration and thus does not radiate. If we assume that the electron velocity deviates slightly from this radiation-free direction (RFD), then the equations of motion can be simplified. In certain EM field configurations, this simplification is enough to allow an analytical solution of the electron equations of motion to be obtained. Remarkably, in a plane wave example, the solution obtained is valid even if radiation reaction is not taken into account. This shows that the domain of applicability of our method is wider than was initially expected. This can be partially attributed to the fact that our approach is based on an expansion of the equations of motion in terms of the electron velocity. Since the magnitude of this velocity vector is smaller than unity, series expansions in terms of velocity should converge. The rates at which these series converge depend on how close the zeroth order is to the real value. We have shown that in the strongly radiation-dominated regime, the RFD can be chosen as a zeroth-order approximation of the direction of the electron velocity. However, a plane wave example shows that the same expansion can be valid even without account of radiation reaction in certain field configurations.

    It should be noted that the approach developed here is valid when the continuous radiation recoil approximation is justified. The validity of this approximation is mostly determined by the value of the QED parameter χ. In particular, in a sufficiently quantum regime, when χ ≫ 1, electron dynamics can become stochastic, and thus the electron distribution function can evolve in a complex way. In that case, equations for higher moments of the distribution function can provide a more accurate description, but this lies outside the scope of this paper.

    In conclusion, we have proposed a general approach for theoretical investigation of single-particle dynamics in the strongly radiation-dominated regime. Most importantly, the method developed here allows one to obtain qualitatively new results compared with the radiation-free approach originally developed in Refs. 74 and 75. We have demonstrated the applicability of our method in different EM field configurations. In particular, we have reproduced the generalized Zeldovich solution in rotating parallel electric and magnetic fields,37,78 the damping of the average rate of electron acceleration in a model plasma accelerator in the radiation-dominated regime,9,16 and a peculiar feature of the electron motion in a strong plane wave, namely, unlimited longitudinal acceleration.80–84 Utilizing this approach to explore plasma behavior in a radiation-dominated regime is planned for future work.

    ACKNOWLEDGMENTS

    Acknowledgment. This work has been supported by the Ministry of Science and Higher Education of the Russian Federation (Agreement No. 075-15-2020-906, Center of Excellence “Center of Photonics”). We are grateful to Anton Golovanov for fruitful discussions.

    APPENDIX A: DERIVING REDUCED EQUATIONS OF MOTION UP TO TERMS OF THE SECOND ORDER OF SMALLNESS

    To obtain the reduced equations of motion, we consider the following representation of the electron velocity:v=v01v12212γ2+v1.Let us substitute this into Eq. (10) and expand it, keeping only terms up to second order (v12) and larger:1v12212γ2dv0dtv02dv12dt+dγ2dt+dv1dt=1γv122+12γ2v0×B+v1×B+v12+1γ2v0(v0E)v0(v1E)v1(v0E)v1(v1E).To eliminate the term dv12/dt from this equation, we take the scalar product of the equation with v1, keeping in mind that v1v0 = 0 and neglecting higher-order terms:12dv12dt=v1dv0dt+v12(v0E)γ.We expand the term dγ−2/dt:dγ2dt=2γ3dγdt2v0Eγ3.Substituting Eqs. (A3) and (A4) back into Eq. (A2), we getdv1dt=F1γ1v12212γ2dv0dtv0v1dv0dt,F1=v122+12γ2v0×B+v1×Bv0(v0E)γ2v0(v1E)v1(v0E)v1(v1E).Let us separately examine the vector v1 × B:v1×B=(v0B)[v0×v1]+v0(v0[v1×B]).We now take the scalar product of Eq. (12) with v1:v1E+v1[v0×B]=0.Performing a cyclic permutation of the scalar triple product, we getv0[v1×B]=v1E.Substituting this relation into Eq. (A7), we obtainv1×B=(v0B)[v0×v1]+v0(v1E).Finally, substituting this into Eq. (A6), we getF1=(v0B)[v0×v1](v0E)v1v122+12γ2v0×B+v0(v0E)γ2(v1E)v1.

    APPENDIX B: APPROXIMATE ELECTRON MOTION IN A PLANE MONOCHROMATIC WAVE

    Let us consider the following plane wave configuration: E = E(φ)y0, B = E(φ)z0, v0 = E × B/E2, φ = tx. For the sake of simplicity, we take the initial electron velocity to lie in the xy plane, in which case the y component of the electron velocity will always remain zero. The reduced equations of motion (16) and (17) in the considered configuration are as follows:dvdt=E2γv21γ2,dγdt=vEFrr(χ),dφdt=12v2+1γ2,where v = v1. Changing the integration variable from t to φ, we getdvdφ=E2γv21γ22v2+1/γ2dγdφ=[vE+Frr(χ)]2v2+1/γ2.The expression for the QED parameter χ is as follows:χ=γEES12v2+1γ2.As χ decreases owing to radiation, any electron will eventually reach the classical regime when χ ≪ 1. In that case,Frr(χ)=23αESχ2=Arrγ2E2v2+1γ22.Equations (B4) and (B5) can then be rewritten asdvdφ=Eγ(vγ)21(vγ)2+1,dγdφ=2(vγ)γE(vγ)2+12ArrE2[(vγ)2+1].Introducing the momentum p = , we obtain the final set of equations:dγdφ=2pγE1+p22ArrE2(1+p2),dpdφ=E2ArrE2pγ(1+p2).Let us start by solving Eqs. (B10) and (B11) without account of radiation reaction for a linearly polarized monochromatic plane wave, i.e., assuming Arr = 0 and E = a0 cos φ:dγpwdφ=2ppwγpwa0cosφ1+ppw2,dppwdφ=a0cosφ.Equation (B11) has the solutionppw=a0sinφ.Substituting this into Eq. (B10), we getdγpwdφ=γpw2a02sinφcosφ1+a02sin2φ,γpw=γ0(1+a02sin2φ)=γ0(1+ppw2).To find the solution when A ≠ 0, let us assumep=ppw+u,γ=γpwΓ,where uppw. Equations (B10) and (B11) transform todΓdφ=2Arra02cos2φγ02uΓa0cosφ1a02cos2φ(1+a02cos2φ)2,dudφ=2Arra03cos2φsinφγ0Γ.Assuming that Γ changes only slightly during a single period, it can be factored outside the integration, i.e.,u2Arra03γ0Γ0φcos2φsinφdφ=2Arra033γ0Γ(1cos3φ).Substituting this into Eq. (B19) givesdΓdφ=2Arra02cos2φγ04Arra043γ0cosφ(1cos3φ)(1a02sin2φ)(1+a02sin2φ)2.Since we assume that Γ changes on a timescale much longer than a single wave period, we can replace the right-hand side of this equation by its average value, i.e.,dΓdφ12π02π2Arra02cos2φγ04Arra043γ0cosφ(1cos3φ)(1a02sin2φ)(1+a02sin2φ)2dφ=Arrγ0a02+441+a02a01Arra02γ0.Thus,Γ1+Arra02γ0φ.The final solution is thenγγ0(1+a02sin2φ)1+Arra02γ0φ,va0sinφγ0(1+a02sin2φ)1+Arra02γ0φ.To express the solution in terms of the laboratory time instead of the wave phase, we solve the following equation:dφdt=12v2+1γ2=12γ02(1+a02sin2φ)1+Arra02γ0φ2,t=2γ02(1+a02sin2φ)1+Arra02γ0φ2dφ.Averaging over a wave period, we get the approximate relationta06Arr2φ3.

    References

    [1]

    [2] Y.Xu, X.Liu, W.Li, S.Li, Y.Liu, Z.Gan, C.Wang, X.Wang, L.Yu, L.Yuet?al.. The Shanghai Superintense Ultrafast Laser Facility (SULF) project. Progress in Ultrafast Intense Laser Science XVI, 199-217(2021).

    [3] R.Li, Y.Leng, Y.Peng, B.Shao, Y.Li, J.Qian, P.Wang. Broad-bandwidth high-temporal-contrast carrier-envelope-phase-stabilized laser seed for 100 PW lasers. Opt. Lett., 45, 2215-2218(2020).

    [4]

    [5] E.Gerstmayr, S.Kuschel, M.Marklund, J. C.Wood, C. D.Murphy, M.Zepf, A.Ilderton, K.Poder, G. M.Samarin, D. R.Symes, P.McKenna, J.Warwick, Z.Najmudin, S. P. D.Mangles, C.Harvey, J. M.Cole, T. G.Blackburn, K. T.Behm, A. G. R.Thomas, G.Sarri, A. S.Joglekar, C. P.Ridgers, C. D.Baird, M. J.Duff, K.Krushelnick. Experimental evidence of radiation reaction in the collision of a high-intensity laser pulse with a laser-wakefield accelerated electron beam. Phys. Rev. X, 8, 011020(2018).

    [6] A. G. R.Thomas, M.Tamburini, S.Kuschel, C. H.Keitel, M.Duff, S.Bohlen, E.Gerstmayr, C. P.Ridgers, C. D.Baird, M.Zepf, K.Poder, D. J.Corvan, S. P. D.Mangles, G. M.Samarin, J. M.Cole, K.Krushelnick, P.McKenna, K.Behm, G.Sarri, Z.Najmudin, A.Di Piazza, D. R.Symes, J.Warwick, C. D.Murphy. Experimental signatures of the quantum nature of radiation reaction in the field of an ultraintense laser. Phys. Rev. X, 8, 031004(2018).

    [7] A.Macchi, M.Tamburini, A.Di Piazza, F.Pegoraro, C. H.Keitel. Radiation reaction effects on radiation pressure acceleration. New J. Phys., 12, 123005(2010).

    [8] F.Pegoraro, M.Tamburini, A.Macchi, T. V.Liseykina. Radiation-pressure-dominant acceleration: Polarization and radiation reaction effects and energy increase in three-dimensional simulations. Phys. Rev. E, 85, 016407(2012).

    [9] E. N.Nerush, I. Y.Kostyukov, A. G.Litvak. Radiative damping in plasma-based accelerators. Phys. Rev. Spec. Top.--Accel. Beams, 15, 111001(2012).

    [10] V. T.Tikhonchuk, R.Capdessus, E.d’Humières. Modeling of radiation losses in ultrahigh power laser-matter interaction. Phys. Rev. E, 86, 036401(2012).

    [11] R.Capdessus, P.McKenna. Influence of radiation reaction force on ultraintense laser-driven ion acceleration. Phys. Rev. E, 91, 053105(2015).

    [12] I. Y.Kostyukov, E. N.Nerush. Laser-driven hole boring and gamma-ray emission in high-density plasmas. Plasma Phys. Controlled Fusion, 57, 035007(2015).

    [13] S.Weber, E. G.Gelfer, A. M.Fedotov. Theory and simulations of radiation friction induced enhancement of laser-driven longitudinal fields. Plasma Phys. Controlled Fusion, 60, 064005(2018).

    [14] A.Fedotov, E.Gelfer, N.Elkina. Unexpected impact of radiation friction: Enhancing production of longitudinal plasma waves. Sci. Rep., 8, 6478(2018).

    [15] A. M.Fedotov, S.Weber, E. G.Gelfer. Radiation induced acceleration of ions in a laser irradiated transparent foil. New J. Phys., 23, 095002(2021).

    [16] I. Y.Kostyukov, A. A.Golovanov, E. N.Nerush. Radiation reaction-dominated regime of wakefield acceleration. New J. Phys., 24, 033011(2022).

    [17] J. L.Martins, T.Grismayer, R. A.Fonseca, L. O.Silva, M.Vranic. Laser absorption via quantum electrodynamics cascades in counter propagating laser pulses. Phys. Plasmas, 23, 056706(2016).

    [18] A. G. R.Thomas, C. P.Ridgers, P.Zhang. The effect of nonlinear quantum electrodynamics on relativistic transparency and laser absorption in ultra-relativistic plasmas. New J. Phys., 17, 043051(2015).

    [19] A.Samsonov, I. Y.Kostyukov, M.Serebryakov, E.Nerush. Opacity of relativistically underdense plasmas for extremely intense laser pulses(2022).

    [20] T. V.Liseykina, S. V.Popruzhenko, A.Macchi. Inverse Faraday effect driven by radiation friction. New J. Phys., 18, 072001(2016).

    [21] S. V.Popruzhenko, T. V.Liseykina, A.Macchi. Quantum effects on radiation friction driven magnetic field generation. Eur. Phys. J. Plus, 136, 170(2021).

    [22] E. N.Nerush, I. Y.Kostyukov, A. S.Samsonov. Effect of electron–positron plasma production on the generation of a magnetic field in laser-plasma interactions. Quantum Electron., 51, 861-865(2021).

    [23] A. G. R.Thomas, D.Del Sorbo, D.Seipt, T. G.Blackburn, C. P.Ridgers, C. D.Murphy, J. G.Kirk. Spin polarization of electrons by ultraintense lasers. Phys. Rev. A, 96, 043407(2017).

    [24] C. P.Ridgers, D.Seipt, D.Del Sorbo, A. G. R.Thomas. Electron spin polarization in realistic trajectories around the magnetic node of two counter-propagating, circularly polarized, ultra-intense lasers. Plasma Phys. Controlled Fusion, 60, 064003(2018).

    [25] C. H.Keitel, Y.-Y.Chen, P.-L.He, K. Z.Hatsagortsyan, R.Shaisultanov. Polarized positron beams via intense two-color laser pulses. Phys. Rev. Lett., 123, 174801(2019).

    [26] A. G. R.Thomas, C. P.Ridgers, D.Seipt, D.Del Sorbo. Ultrafast polarization of an electron beam in an intense bichromatic laser field. Phys. Rev. A, 100, 061402(2019).

    [27] M.Büscher, L.Ji, J.Thomas, W.Wang, N.Wang, A.Hützen, Z.Guo, R.Li, Y.Wu, Q.Yu, B.Feng, A.Pukhov, C.Qin, T. P.Rakitzis, X.Geng, L.Zhang, B.Shen, X.Yan. Polarized electron-beam acceleration driven by vortex laser pulses. New J. Phys., 21, 073052(2019).

    [28] J.-X.Li, C. H.Keitel, F.Wan, K. Z.Hatsagortsyan, R.Shaisultanov, Y.-F.Li. Ultrarelativistic electron-beam polarization in single-shot interaction with an ultraintense laser pulse. Phys. Rev. Lett., 122, 154801(2019).

    [29] W.-M.Wang, H.-S.Hu, Y.-Y.Chen, Y.-F.Li. Production of highly polarized positron beams via helicity transfer from polarized electrons in a strong laser field. Phys. Rev. Lett., 125, 044802(2020).

    [30] C. H.Keitel, Y.-F.Li, K. Z.Hatsagortsyan, J.-X.Li, F.Wan, R.Shaisultanov. Ultrarelativistic polarized positron jets via collision of electron and ultraintense laser beams. Phys. Lett. B, 800, 135120(2020).

    [31] C. H.Keitel, Z.Gong, K. Z.Hatsagortsyan. Retrieving transient magnetic fields of ultrarelativistic laser plasma via ejected electron polarization. Phys. Rev. Lett., 127, 165002(2021).

    [32] E. N.Nerush, I. Yu.Kostyukov. Radiation emission by extreme relativistic electrons and pair production by hard photons in a strong plasma wakefield. Phys. Rev. E, 75, 057401(2007).

    [33] A. R.Bell, J. G.Kirk. Possibility of prolific pair production with high-power lasers. Phys. Rev. Lett., 101, 200403(2008).

    [34] N. V.Elkina, I. Y.Kostyukov, N. B.Narozhny, E. N.Nerush, A. M.Fedotov, H.Ruhl. Laser field absorption in self-generated electron-positron pair plasma. Phys. Rev. Lett., 106, 035001(2011).

    [35] K.Bennett, J. G.Kirk, C. P.Ridgers, R.Duclous, A. P. L.Robinson, T. D.Arber, A. R.Bell, C. S.Brady. Dense electron-positron plasmas and ultraintense γ rays from laser-irradiated solids. Phys. Rev. Lett., 108, 165006(2012).

    [36] A. M.Fedotov, N. B.Narozhny. Quantum-electrodynamic cascades in intense laser fields. Phys.-Usp., 58, 95(2015).

    [37] I. Y.Kostyukov, E. N.Nerush. Production and dynamics of positrons in ultrahigh intensity laser-foil interactions. Phys. Plasmas, 23, 093119(2016).

    [38] T.Grismayer, R. A.Fonseca, M.Vranic, L. O.Silva, J. L.Martins. Seeded QED cascades in counterpropagating laser pulses. Phys. Rev. E, 95, 023210(2017).

    [39] M.Vranic, G.Korn, O.Klimo, S.Weber, M.Jirka. QED cascade with 10 PW-class lasers. Sci. Rep., 7, 15302(2017).

    [40] W.Luo, Z.-M.Sheng, F.-Y.Li, C. P.Ridgers, D.Del Sorbo, T.Yuan, J.-Y.Yu, M.Chen, W.-Y.Liu. QED cascade saturation in extreme high fields. Sci. Rep., 8, 8400(2018).

    [41] J.Zhang, W. Y.Liu, X. H.Yuan, T.Yuan, Z. M.Sheng, W.Luo, S. M.Weng, J. Y.Yu, M.Chen. Spatiotemporal distributions of pair production and cascade in solid targets irradiated by ultra-relativistic lasers with different polarizations. Plasma Phys. Controlled Fusion, 60, 065003(2018).

    [42] Z.-M.Sheng, D.Del Sorbo, K.Small, P.McKenna, A. P. L.Robinson, C.Slade-Lowther, M. J.Duff, R.Capdessus, W.Luo, D. R.Blackmanet?al.. Efficient ion acceleration and dense electron–positron plasma creation in ultra-high intensity laser-solid interactions. New J. Phys., 20, 033014(2018).

    [43] F.-Q.Shao, J.-X.Liu, K.Liu, T.-P.Yu, W.-Q.Wang, L.-X.Hu, Y.Yin, Z.-Y.Ge, Y.Lu. Enhanced copious electron–positron pair production via electron injection from a mass-limited foil. Plasma Phys. Controlled Fusion, 60, 125008(2018).

    [44] Z.-M.Sheng, W.Luo, W.-Y.Liu, M.Chen, S.-D.Wu, T.Yuan, F.-Y.Li, J.-Y.Yu, Y.-Y.Ma. Enhanced electron-positron pair production by two obliquely incident lasers interacting with a solid target. Plasma Phys. Controlled Fusion, 60, 095006(2018).

    [45] A. A.Gonoskov, A. V.Bashinov, S. I.Bastrakov, E. S.Efimenko, A. M.Sergeev, A. V.Kim, A. A.Muraviev, I. B.Meyerov. Laser-driven plasma pinching in ee+ cascade. Phys. Rev. E, 99, 031201(2019).

    [46] A. S.Samsonov, E. N.Nerush, I. Y.Kostyukov. Laser-driven vacuum breakdown waves. Sci. Rep., 9, 11133(2019).

    [47] I. Y.Kostyukov, A. S.Samsonov, E. N.Nerush. Hydrodynamical model of QED cascade expansion in an extremely strong laser pulse. Matter Radiat. Extremes, 6, 034401(2021).

    [48] (2016).

    [49] V.Ritus, A.Nikishov. Quantum processes in the field of a plane electromagnetic wave and in a constant field I. Sov. Phys. JETP, 19, 529-541(1964).

    [50] E. M.Lifshitz, L. P.Pitaevskii, V. B.Berestetskii. Quantum Electrodynamics(1982).

    [51] V.Ritus. Quantum effects of the interaction of elementary particles with an intense electromagnetic field. J. Sov. Laser Res., 6, 497(1985).

    [52] H.Nitta, M. K.Khokonov. Standard radiation spectrum of relativistic electrons: Beyond the synchrotron approximation. Phys. Rev. Lett., 89, 094801(2002).

    [53] B.King, A.Ilderton, D.Seipt. Extended locally constant field approximation for nonlinear Compton scattering. Phys. Rev. A, 99, 042121(2019).

    [54] A.MacLeod, B.King, T.Heinzl. Locally monochromatic approximation to QED in intense laser fields. Phys. Rev. A, 102, 063110(2020).

    [55] E. G.Gelfer, A. A.Mironov, A. M.Fedotov, S.Weber. Nonlinear Compton scattering in time-dependent electric fields beyond the locally constant crossed field approximation. Phys. Rev. D, 106, 056013(2022).

    [56] T.Podszus, A.Di Piazza. High-energy behavior of strong-field QED in an intense plane wave. Phys. Rev. D, 99, 076004(2019).

    [57] E. N.Nerush, I. I.Artemenko, I. Yu.Kostyukov. Quasiclassical approach to synergic synchrotron-cherenkov radiation in polarized vacuum. New J. Phys, 22, 093072(2020).

    [58] I.Arka, A. R.Bell, J. G.Kirk. Pair production in counter-propagating laser beams. Plasma Phys. Controlled Fusion, 51, 085008(2009).

    [59] W.Leemans, E.Esarey, S.Bulanov, C.Schroeder. Electromagnetic cascade in high-energy electron, positron, and photon interactions with intense laser pulses. Phys. Rev. A, 87, 062110(2013).

    [60] J. K.Koga, K.Kondo, S. V.Bulanov, S. S.Bulanov, G.Korn, N. N.Rosanov, T. Z.Esirkepov, M.Kando. Attractors and chaos of electron dynamics in electromagnetic standing wave. Phys. Lett. A, 379, 2044(2015).

    [61] M.Grech, C.Riconda, R.Duclous, F.Niel, F.Amiranoff. From quantum to classical modeling of radiation reaction: A focus on stochasticity effects. Phys. Rev. E, 97, 043209(2018).

    [62] A.Gonoskov, T.Blackburn, M.Marklund, S.Bulanov. Charged particle motion and radiation in strong electromagnetic fields. Rev. Mod. Phys, 94, 045001(2022).

    [63] C. S.Shen, D.White. Energy straggling and radiation reaction for magnetic bremsstrahlung. Phys. Rev. Lett., 28, 455(1972).

    [64] R.Duclous, A. R.Bell, J. G.Kirk. Monte Carlo calculations of pair production in high-intensity laser–plasma interactions. Plasma Phys. Controlled Fusion, 53, 015009(2010).

    [65] A.Ilderton, A.Gonoskov, M.Marklund, C. N.Harvey. Quantum quenching of radiation losses in short laser pulses. Phys. Rev. Lett., 118, 105004(2017).

    [66] A.Di Piazza, N.Neitz. Stochasticity effects in quantum radiation reaction. Phys. Rev. Lett., 111, 054802(2013).

    [67] C. D.Murphy, T. G.Blackburn, C. P.Ridgers, A. G. R.Thomas, P.McKenna, C.Slade-Lowther, L. E.Bradley, S. P. D.Mangles, C. D.Baird, M.Marklund, D.Del Sorbo. Signatures of quantum effects on radiation reaction in laser–electron-beam collisions. J. Plasma Phys., 83, 715830502(2017).

    [68] W.Gerlach, O.Stern. Der experimentelle Nachweis der Richtungsquantelung im Magnetfeld. Z. Phys., 9, 349-352(1922).

    [69] L. H.Thomas. The motion of the spinning electron. Nature, 117, 514(1926).

    [70] V.Bargmann, L.Michel, V. L.Telegdi. Precession of the polarization of particles moving in a homogeneous electromagnetic field. Phys. Rev. Lett., 2, 435(1959).

    [71] S. R.Mane, Yu. M.Shatunov, K.Yokoya. Spin-polarized charged particle beams in high-energy accelerators. Rep. Prog. Phys, 68, 1997(2005).

    [72] D.Seipt, D.Del Sorbo, A. G. R.Thomas, C. P.Ridgers. Polarized QED cascades. New J. Phys., 23, 053025(2021).

    [73] C. H.Keitel, M.Wen, M.Tamburini. Polarized laser-wakefield-accelerated kiloampere electron beams. Phys. Rev. Lett., 122, 214801(2019).

    [74] E. N.Nerush, I. Y.Kostyukov, A. S.Samsonov. Asymptotic electron motion in the strongly-radiation-dominated regime. Phys. Rev. A, 98, 053858(2018).

    [75] A.Gonoskov, M.Marklund. Radiation-dominated particle and plasma dynamics. Phys. Plasmas, 25, 093109(2018).

    [76] P.Jér?me. Particle acceleration and radiation reaction in a strongly magnetised rotating dipole. Astron. Astrophys, 666, A5(2022).

    [77] S. E.Gralla, V.Paschalidis, Y.Cai. Dynamics of ultrarelativistic charged particles with strong radiation reaction. I. Aristotelian equilibrium state(2022).

    [78] Y. B.Zel’dovich. Interaction of free electrons with electromagnetic radiation. Sov. Phys. Usp, 18, 79(1975).

    [79] L. D.Landau. The Classical Theory of Fields(2013).

    [80] A. D.Piazza. Exact solution of the Landau-Lifshitz equation in a plane wave. Lett. Math. Phys., 83, 305-313(2008).

    [81] J. P.Ostriker, J. E.Gunn. On the motion and radiation of charged particles in strong electromagnetic waves. I. Motion in plane and spherical waves. Astrophys. J., 165, 523(1971).

    [82] M.Grewing, H.Heintzmann, E.Schrüfer. Acceleration of charged particles and radiation reaction in strong plane and spherical waves. II. Z. Phys. A: Hadrons Nucl., 260, 375-384(1973).

    [83] K.Thielheim. Particle acceleration in extremely strong electromagnetic wave fields, 276-278(1993).

    [84] R.Ekman, T.Heinzl, A.Ilderton. Exact solutions in radiation reaction and the radiation-free direction. New J. Phys., 23, 055001(2021).

    A. S. Samsonov, E. N. Nerush, I. Yu. Kostyukov. High-order corrections to the radiation-free dynamics of an electron in the strongly radiation-dominated regime[J]. Matter and Radiation at Extremes, 2023, 8(1): 014402
    Download Citation