This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Paper The following article is Open access

Ultrarelativistic electrons in counterpropagating laser beams

, , and

Published 11 June 2021 © 2021 The Author(s). Published by IOP Publishing Ltd on behalf of the Institute of Physics and Deutsche Physikalische Gesellschaft
, , Focus on Strong Field Quantum Electrodynamics with High Power Lasers and Particle Beams Citation Q Z Lv et al 2021 New J. Phys. 23 065005 DOI 10.1088/1367-2630/abfa60

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

1367-2630/23/6/065005

Abstract

The dynamics and radiation of ultrarelativistic electrons in strong counterpropagating laser beams are investigated. Assuming that the particle energy is the dominant scale in the problem, an approximate solution of classical equations of motion is derived and the characteristic features of the motion are examined. A specific regime is found with comparable strong field quantum parameters of the beams, when the electron trajectory exhibits ultrashort spike-like features, which bears great significance to the corresponding radiation properties. An analytical expression for the spectral distribution of spontaneous radiation is derived in the framework of the Baier–Katkov semiclassical approximation based on the classical trajectory. All the analytical results are further validated by exact numerical calculations. We consider a non-resonant regime of interaction, when the laser frequencies in the electron rest frame are far from each other, avoiding stimulated emission. Special attention is devoted to settings when the description of radiation via the local constant field approximation fails and to corresponding spectral features. Periodic and non-periodic regimes are considered, when lab frequencies of the laser waves are always commensurate. The sensitivity of spectra with respect to the electron beam spread, focusing and finite duration of the laser beams is explored.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Electromagnetic (EM) processes in strong laser fields are characterized by nonperturbative multiphoton dynamics. An efficient treatment of nonlinear processes in strong field quantum electrodynamics (QED) has been provided within the Furry picture [1], regarding the strong field as classical and employing the electron wave function in such fields for the calculation of amplitudes of QED processes. The Volkov wave function of an electron in a plane wave laser field [2] has been successfully and extensively employed to explore the nonlinear Compton effect, nonlinear Breit–Wheeler [35], and nonlinear Bethe–Heitler pair production processes [6]. The multiphoton processes in a plane wave field enter into play at large values of the classical strong field parameter ξ ≡ −ea/m ≫ 1, where a is the amplitude of the vector potential, $a\equiv \sqrt{{A}^{2}}$, while e and m the electron charge and mass, respectively. Relativistic units = c = 1 are used throughout the paper, unless specified otherwise. Present day laser facilities attain intensities of up to 5 × 1022 W cm−2 in optical wavelengths [7, 8], corresponding to ξ ∼ 100. For the next generation extreme laser infrastructures an order of magnitude increase of intensity is expected [9, 10], opening a bright avenue for investigation of extreme nonlinear strong field QED processes [1114] in laser–plasma or laser–electron beam interactions.

The desire to increase the effective laser field with a given laser beam energy gave rise to the concept of multi-beam configurations and to the notion of a dipole wave [1521]. The simplest case of a multi-beam configuration is the counterpropagating laser beam setup, which is an attractive setup to study QED effects [2231]. All of the above admit no exact analytical solutions for the wave function and are, therefore, not accessible to strong field QED calculations within the Furry picture. The common way of treating strong field QED processes in laser–plasma interaction is to approximate the emission by that in the presence of the local constant field when the field intensity is very high (ξ ≫ 1). The local constant field approximation (LCFA) is rigorously derived in the asymptotic limit ξ ≫ 1 for the plane wave case (more precise condition is (ξ/χ1/3)[ω/(ɛω)]1/3 ≫ 1 [32, 33], with typical emission frequencies ω/ɛχ/(χ + 1) and the electron energy ɛ). In this case the formation length of the process becomes smaller than the field wavelength, and the process probability depends solely on the quantum parameter $\chi =e\sqrt{-{\left({F}^{\mu \nu }{P}_{\nu }\right)}^{2}}/{m}^{3}$, where Fμν is the EM tensor and ${P}_{\mu }=\left(\varepsilon ,\mathbf{P}\right)$ the particle four-momentum. Due to its simplicity, this approximation allows for the inclusion of QED processes in kinetic Monte Carlo and particle-in cell (PIC) simulations involving fields of complex forms [3436]. However, recently deficiencies and failures of LCFA have been observed in low [32, 3739] and high energy limits [40]. LCFA violation in counterpropagating laser waves is demonstrated in [41] which is due to emergence of an additional small time scale in the electron dynamics.

Beyond LCFA treatment, one may apply the Wentzel–Kramers–Brillouin (WKB) approximation to describe the electron quantum (quasiclassical) dynamics [42, 43]. A similar high-energy approximation describing the electron dynamics in a focused laser field, when the electron longitudinal momentum dominates over transverse one, is developed in [44] and applied for description of corresponding nonlinear QED processes [4547]. As WKB approximation is closely connected with the classical description, a WKB wave function in closed analytical form can be derived in the cases when such solution is available for the electron classical trajectory. In the 60s' this observation motivated Baier and Katkov to develop the operator approach and with its help to express the amplitudes of strong field QED processes, such as radiation and pair production, as a function of the electron classical trajectory in the external field [4850].

We consider the setup of counterpropagating laser beams. Here one should distinguish resonant and non-resonant regimes of interaction. The resonance appears when the frequencies of the laser waves match in the average rest frame of the electron [51], which would lead to stimulated emission of laser photons [52, 53], to coherent electron scattering from the moving laser grating (Kapitza–Dirac effect [5457]). Rather than the widely explored topic of stimulated processes in the resonant regime, we discuss in this paper the non-resonant regime, relevant to the investigation of spontaneous radiation in this setup. The equation of motion is highly nonlinear and is known to exhibit chaotic dynamics when the corresponding field are strong [58, 59]. In the quantum domain, approximations to the wave function of a scalar particle experiencing this field have been discussed in [60, 61]. Radiation in this setup and its reaction to the electron dynamics have been investigated within LCFA via PIC-QED simulations [22, 2527, 29]. In particular, this configuration turned out to be favorable to QED cascades where the emitted γ-photons are energetic enough to produce electron–positron pairs, starting an avalanche-like dynamics. Moreover, it was shown that radiation reaction can essentially modify the trapping of particles in this field [24, 28].

In the present paper an electron interacting with counterpropagating laser beams in the non-resonant regime is considered, using laser fields of equal frequency in the laboratory frame and the ultrarelativistic electron moving initially along the propagation direction of the first laser beam. An approximate analytical solution to the classical equation of motion is derived, imposing a restriction on the laser parameters and electron initial momentum, in particular, demanding ξ1 ξ2γ2, for the lasers' field parameters ξ1, ξ2 and γ as the average Lorentz factor of the electron in the fields. Based on the approximated analytical trajectory, the radiation is calculated in the realm of the semiclassical Baier–Katkov formalism. We compare the obtained formula with a fully numerical calculation and discuss radiation features in different regimes. Furthermore, the influence of the pulses width and focusing, which cannot be accounted for analytically, are studied numerically.

The paper is organized as follows. In section 2 an approximate solution to the Lorentz equation in the counterpropagating beams is derived. The investigation of the photon emission is given in section 3. Radiation spectra in strong fields are discussed along with a numerical example. The validity of the analytical treatment and the deviations with respect to numerical calculations are analyzed. The impact of finite duration and focusing of the laser beam is investigated numerically. Conclusions are given in section 4.

2. The classical dynamics

The classical equation of motion for the particle in EM fields reads

Equation (1)

where τ is the proper time, γ = ɛ/m is the relativistic Lorentz-factor, $\varepsilon =\sqrt{{m}^{2}+{\mathbf{P}}^{2}}$ is the energy, P is the momentum, E, B are the electric and magnetic fields, correspondingly, and v = P/ɛ is the velocity. In the general case, equation (1) cannot be solved analytically because of its nonlinearity, as x(τ) depends on the momentum via x(τ) = ∫dτ P(τ)/m. In the following, we seek for an approximated solution in the presence of counterpropagating circularly polarized laser waves with the four-vector potential A = A1 + A2 where

Equation (2)

The four-wavevectors of the beams are k1 = (ω, 0, 0, ω), k2 = (ω, 0, 0, −ω) and ex = (0, 1, 0, 0), ey = (0, 0, 1, 0) are the unit vectors. The dimensionless functions g1(k1x)and g2(k2x) are slow wave envelopes. In this section they will be set to unity. We will refer to them when considering the influence of the turn-on process on the relation between the average momentum and its initial value in section 2.3. Here ab denotes the inner product of two four-vectors. The electric and magnetic fields are derived from the vector potential through $\mathbf{E}=-\frac{\partial \mathbf{A}}{\partial t}$ and B = ∇ × A:

Equation (3)

Equation (4)

Please note that here we have chosen the counterpropagating waves being co-rotating. However, the characterization of the electron dynamics is similar for counter-rotating waves in the considered regimes. The only difference is that the rotation caused by the ξ2-beam changes its directions but the physical properties of radiation remain the same.

2.1. Classical trajectory

For solving the equation of motion (1) the phases appearing in the fields arguments are expressed via the trajectory x(τ)

Equation (5)

where

Equation (6)

with $\delta {P}_{\mu }={P}_{\mu }\left(\tau \right)-{\bar{P}}_{\mu }$. The bar symbol designates time-averaged quantities. The key assumption lying in the basis of our derivation is

Equation (7)

as well as similar relations where in the right wing cos → sin and in the left wing sin → −cos. By employing this assumption, the four-momentum P of the particle can be derived. With the momentum, expressions for δϕ1, δϕ2 according to equation (6) are calculated under certain restrictions, which assure the validity of the assumption of equation (7).

Since the vector potential is independent on the transverse coordinates, the canonical momentum in these directions is conserved P(τ) = peA(τ). Without loss of generality, we choose the initial transverse momentum p to be on the x-axis. Then,

Equation (8)

Equation (9)

where −ea1,2 = 1,2. Applying the assumption of equation (7), the x, y components of the trajectory read

Equation (10)

Equation (11)

Now let us consider the oscillations on the z axis

Equation (12)

Employing equations (8), (9) and (4), one can find out that the terms scaling like ${\xi }_{1}^{2},{\xi }_{2}^{2}$ cancel. Therefore we have

Equation (13)

Accordingly, ${P}_{z}={\bar{P}}_{z}+\delta {P}_{z}$ with

Equation (14)

where $\overline{P}$ is the time-averaged momentum, whose relation to the initial momentum of the electron before interacting with the laser pulses will be discussed in section 2.3. Integrating over τ, one obtains the z-component of the trajectory

Equation (15)

Let us now calculate the energy and its oscillatory part: $\varepsilon =\sqrt{{m}^{2}+{P}_{x}^{2}+{P}_{y}^{2}+{\left({\bar{P}}_{z}+\delta {P}_{z}\right)}^{2}}$. With equations (8) and (9),

Equation (16)

Using δPz given by equation (14) and recalling that $\left({k}_{1}-{k}_{2}\right)\cdot \bar{P}=-2\omega {\bar{P}}_{z}$, one can find out that the terms proportional to cos(ϕ1ϕ2) cancel each other. The expression for the energy may be further simplified to

Equation (17)

with ${k}_{1}\cdot \bar{P}=\omega \bar{\varepsilon }\left(1-{\bar{v}}_{z}\right)$, ${k}_{2}\cdot \bar{P}=\omega \bar{\varepsilon }\left(1+{\bar{v}}_{z}\right)$, where the average velocity on the z axis is defined as ${\bar{v}}_{z}={\bar{P}}_{z}/\bar{\varepsilon }$. With a Taylor expansion the following expression is obtained

Equation (18)

where the average energy, effective mass, and the oscillatory part are defined as

Equation (19)

Notice that for vanishing transverse momentum px = 0 the energy is constant, in accordance with [51]. The expansion in equation (18) is justified if $\delta {P}_{z}\ll \bar{\varepsilon }$. Here we have taken into account that for an ultrarelativistic electron, the amplitude of δPz is always larger than δɛ according to equations (14) and (19) and thus δɛ/ɛ < δPz /ɛ ≪ 1. Taking into account the explicit form of δPz , equation (14), the validity condition $\delta {P}_{z}\ll \bar{\varepsilon }$ reads

Equation (20)

Equation (21)

Equation (22)

So far ϕ1, ϕ2 were not specified yet. With the help of δɛ, δPz we evaluate δϕ1, δϕ2 and thus obtain the phases ϕ1, ϕ2. Accordingly, the validity criterion for the basic assumption of this derivation, equation (7), is determined. Substituting equations (14) and (19) in (6) we have

Equation (23)

where Φ1, Φ2 are arbitrary constants and the coefficients are

Equation (24)

Equation (5) together with equation (23) form an implicit system for the solution of the phases. Without loss of generality, we assumed that ${\bar{v}}_{z}{ >}0$, i.e. the particle copropagates with the ξ1 beam, leading to asymmetry between the two beams. As a consequence, ${k}_{1}\cdot \bar{P}\ll {k}_{2}\cdot \bar{P}$, so that if the beams amplitudes are of the same order of magnitude, C2 is considerably larger than C1. In the following we assume that C1, C12 ≪ 1, yielding the following expressions

Equation (25)

In order to prove the consistency of this conclusion, one should accomplish two things. First, one has to show that the contributions of C1, C12 to the momentum are of second order, justifying the neglection. For this purpose, we consider a general function F with the following argument ϕ(τ) = ϕ0(τ) + ν sin f(τ), where ν is a small constant and ϕ0(τ), f(τ) are general functions. Taylor expanding with respect to ν yields

Equation (26)

In our case, ϕ0 designates the approximated phases ϕ1 or ϕ2 given in equation (25) and ν is either C1 or C12, ϕ stands for the full phases including the neglected terms proportional to C12, C1, and F(ϕ) either epsilon cos(ϕ) or epsilon sin ϕ, where epsilon stands for the amplitudes of the various momentum oscillations appearing in equations (8), (9) and (14). Since F' ∼ epsilon, the correction scales as O(epsilonν). One should notice that the amplitude of the momentum oscillations are assumed to be considerably smaller with respect to the particle energy, being the dominant energy scale. Hence, epsilon is a small parameter and the corrections corresponding to C1, C12 may be neglected, up to the second order.

Second, one should verify that the approximation of equation (7) indeed holds. Plugging the phases equation (25) into equation (7) we notice that all the three integrals take the form $\mathcal{I}\equiv \int \mathrm{cos}\left[\alpha \tau +\beta \right.$ $\left.\mathrm{sin}\left(\kappa \tau \right)\right]\mathrm{d}\tau $ with different choices of α, β, κ. In order to calculate this integral, we recall the identity

Equation (27)

where Js (β) is the Bessel function. Multiplying eiατ on both sides, one readily obtains the real and imaginary part, respectively, as

Equation (28)

The integral can thus be obtained as

Equation (29)

For a certain β we know that Js (β) vanishes if the index s is larger enough than β. Therefore, further simplification can be accomplished if

Equation (30)

The integral is thus approximated by

Equation (31)

where equation (28) has been considered. This result is in agreement with equation (7). Now let us find the conditions for which equation (30) is satisfied for all three cases. For the first integral, β vanishes and equation (30) is trivially fulfilled. For the second case, one has $\alpha =\left({k}_{2}\cdot \bar{P}\right)/{m}^{2},\kappa =\left({k}_{1}\cdot \bar{P}\right)/{m}^{2}$ and β = C2, so that equation (30) yields

Equation (32)

For the third integral β, κ are as in the second case but $\alpha =\left[\left({k}_{2}-{k}_{1}\right)\cdot \bar{P}\right]/{m}^{2}$, imposing the condition

Equation (33)

Thus, equation (7) was explicitly shown to be valid, given that equations (32) and (33) are satisfied. Combining C1, C12 ≪ 1 with equations (20)–(22), (32) and (33), yields the final validity criteria

Equation (34)

Let us conclude the derivation. The final expressions for the trajectory and momentum are equations (10), (11), and (15) and equations (8), (9), (14), (19), correspondingly. The phases ϕ1(τ) and ϕ2(τ) are given by equation (25). The validity criteria corresponding to this solution are equation (34). In the ultrarelativistic regime $1-{\bar{v}}_{z}\ll 1$ they are simplified to

Equation (35)

The above criteria can be fulfilled in a scenario where an ultrarelativistic electron moves along the laser propagation direction with a small deviating angle. Alternatively, one may write the instantaneous momentum in a covariant form as follows

Equation (36)

One may verify that in the case if one of the laser beams vanishes, our result equation (36) recovers the familiar plane wave solution [5].

The above derivation expresses the physical quantities of interest, namely the trajectory and the four-momentum, as a function of the proper time τ. However, for practical applications it is favorable to use the laboratory time as the independent variable. The two quantities are simply related through $\mathrm{d}t=\frac{\varepsilon }{m}\mathrm{d}\tau $. Performing the integration we obtain

Equation (37)

The latter along with x(τ) provides a parametric description of the particle coordinate as a function of the laboratory time. Alternatively, one may further approximate the phases. We start by writing equation (37) as

Equation (38)

Substituting (38) into the phase ϕ1 given in (25) one obtains

Equation (39)

where ${\omega }_{1}\equiv \left(1-{\bar{v}}_{z}\right)\omega $. This equation is implicit, since ϕ1 appears in both sides. Nevertheless, it proves useful as a starting point for approximation of the phases, as we immediately show. According to the validity condition equation (34), one notices that the coefficients of the sine functions in equation (39) are much smaller than 1. As a result, equation (26) may be employed here. The fact that equation (39) is implicit (ϕ1 appears in both sides) poses no difficulty, since the argument f(τ) in (26) is general and has no influence on the final result. Due to (26) and according to the same reasoning that led us to neglect C1, C12, they may be omitted, leading to

Equation (40)

where ${\omega }_{2}\equiv \left(1+{\bar{v}}_{z}\right)\omega $. Hence, one observes that ω1, ω2 are the characteristic oscillation frequencies associated with the ξ1, ξ2 beams, respectively. Notice that according to our convention the particle copropagates with the ξ1 beam, so that ${\bar{v}}_{z}$ is positive, and hence ω2 is considerably larger than ω1, which indicates the non-resonant regime of interaction.

2.2. Characteristics of the trajectories

With the obtained analytical expression for the electron momentum and coordinate, we study in this section the main characteristics of the motion. As the dynamics is strongly effected by both of the laser beams, we would expect to find some unusual features in the electron trajectory, where the acceleration is large and which may yield radiation emission deviating from the LCFA results based on the Baier–Katkov technique [49]. We inspect the electron velocity in all components for three different field parameters in figure 1, featuring various behaviors. The results shown in the figure are obtained within the analytical treatment presented above and proved by the fully numerical solutions of equation (1). In all cases the initial transverse momentum vanishes px = 0 and the energy is ɛ = 2.6m*, corresponding to ω2/ω1 $\approx $ 25. The plots present a time interval of 2π/ω1, so it consists of one cycle of the ξ1 beam and about 25 cycles of the ξ2 beam.

Figure 1.

Figure 1. The electron velocity for three different field configurations. In all cases, the electron is copropagating with the ξ1 laser beam and the velocity is shown for the time period during which the electron has travelled only for one cycle in the ξ1 laser beam. Simulation parameters are: (a) ξ1 = 50, ξ2 = 1; (b) ξ1 = 20, ξ2 = 20; (c) ξ1 = 1, ξ2 = 50. In all cases the electron has no transverse momentum and its energy in the field is ɛ = 2.6m*.

Standard image High-resolution image

In panel (a) the laser parameters are ξ1 = 50, ξ2 = 1. In the xy plane the particle performs a cyclic motion with a radius of 1/ɛ and a frequency ω1 and on top of it rapid oscillations with frequency ω2 and amplitude 2/ɛ. According to equation (14), the amplitude of the oscillation on the z axis scales as ∼m2 ξ1 ξ2/ɛ2 and is, therefore, considerably smaller as compared to those in the x, y axes.

Panel (b) depicts the case of ξ1 = ξ2 = 20. In the xy plane the oscillations amplitude are now identical, so that the particle moves in circles with frequency ω2 according to ξ2. An interesting point is that the origin of the circle also exhibits a cyclic motion due to ξ1 with a frequency of ω1. Both the fast ξ2 circle and the slow ξ1 circle have the same radius because of the identical oscillation amplitudes. In addition, one can observe that the tilting angle of the total velocity with respect to the z axis is gradually changing. The reason is that the oscillation frequency on the z axis is ω2ω1. As a result, the relative phase between vz and vx for example gradually increases during the time interval under consideration from 0 to 2π.

Panel (c) presents the dynamics for ξ1 = 1, ξ2 = 50. It is quite similar to the previous case, but now the radius of the slow ξ1 circle is negligible, such that the motion takes the form of a single circle with time dependent tilt.

With respect to radiation emission, the more irregular the trajectory is, the more interesting is the spectral shape. Hence, in the following we concentrate on the ξ1ξ2 case, like in panel (a) of figure 1, where the dynamics is much more complex. Figure 2 shows a two dimensional projection of the velocity on the xy plane for ξ1 = 50, ξ2 = 1 with three different particle energies ɛ.

Figure 2.

Figure 2. The transverse velocity in the xy plane for ξ1 = 50, ξ2 = 1, px = 0 with three electron energies: (a) ɛ = 130m; (b) ɛ = 182m; (c) ɛ = 250m. The initial transverse momentum in all panels is zero. The inset is just the zoom in of the velocity for a small time interval. The plots are for the time period during which the electron has travelled only for one cycle in the ξ1 laser beam.

Standard image High-resolution image

Panel (a) corresponds to ɛ = 130m. As mentioned in figure 1, one can see that the dynamics is a combination of a large circle due to ξ1 and rapid oscillations corresponding to ξ2, which have a smooth sine-shape, see in the inset. When the energy is increased, see panel (b) with ɛ = 182m, several interesting changes take place. First of all, the number of the ξ2 oscillations contained in one cycle of ξ1 increases since the ratio of the frequencies, ω2/ω1, is now about 51 instead of 25 in panel (a). Furthermore, the radius of the circle as well as the amplitude of the small oscillations becomes smaller. This is because the amplitude of the transverse velocity v1/ɛ decreases with the energy increase. More interestingly, a sharp spike-like feature emerges for each cycle of ξ2 oscillation. It should be emphasized that the time scale corresponding to these spike-like features is significantly shorter than both 1/ω1 and 1/ω2.

In order to shed light on this spike-like feature, we take advantage of the approximated phases equation (40) and derive from the y component of the trajectory (11) the corresponding acceleration

Equation (41)

where in the last expression we take into account that the quantum parameter is proportional to the acceleration $\chi ={\varepsilon }^{2}\vert \dot {v}\vert /{m}^{3}$ [49]. Please note that here $\dot {a}$ refers to the derivative of time t. Here χ1 = ξ1 ɛω1/m, and χ2 = ξ2 ɛω2/m are the quantum parameters induced, respectively, by beams 1 and 2. Let us take a close look at the time interval corresponding to 0 < ϕ1 < π/2. One can see that as long as χ2 < χ1, the acceleration does not change its sign. Namely, the velocity will monotonously decrease, as the case in figure 2(a). Increasing the energy results in higher values of the ratio ω2/ω1, and at a certain point χ2 exceeds χ1. When χ2 becomes large enough, the acceleration ${\dot {v}}_{y}$ will change its sign during the time interval. If χ2 is only slightly higher than χ1, the acceleration is positive for a very short time, leading to sharp spikes, as encountered in figure 2(b). In case χ2 is significantly larger than χ1, the acceleration is positive about half of the time, giving rise to the whirl appearing in figure 2(c), where the energy is further increased to ɛ = 250m. The impact of these phenomena on the radiation emission has been explored in reference [41].

Finally, let us examine the influence of the transverse momentum px . From the final expressions for the momentum and energy, one observes that this quantity has several contributions. First, it gives rise to the oscillations in the longitudinal momentum Pz and the energy ɛ, see equations (14) and (19), respectively. This means that the energy of the electron in the field is not constant anymore. Moreover, the non-zero transverse momentum also adds a slow sine-term (with frequency ω1) to the phase ϕ2, see equation (25). As a result, the rapid oscillations corresponding to ξ2 are periodically modulated. This phenomenon is demonstrated in figure 3, where the y-component of velocity is plotted as a function of time within half a cycle of ξ1. To verify our analytical results (black), the numerical solution is also shown in this figure as a blue line. The agreement between the numerical solution and the analytical one is excellent, as the two curves are on top of each other. In addition, as we expected, the frequency of the small oscillations increases with time up to t/T ≈ 7, with T = 2π/ω, and then gradually decreases again.

Figure 3.

Figure 3. The y component of the velocity as a function of time is shown for the time duration in which the electron travels in ξ1 laser beam for only half cycle. The electron has an average energy $\bar{\varepsilon }=182m$ copropagating with ξ1 and the transverse momentum is px = 2.5m. The field parameters are ξ1 = 50, ξ2 = 1. The time t is in units of T = 2π/ω.

Standard image High-resolution image

2.3. Drift momentum

From the discussion above, we can see that the drift momentum of the particle in the laser fields, especially the average energy in the field, is an essential parameter for our approximation. However, the drift momentum depends on the asymptotic momentum of the particle before entering in the laser fields as well as the way of switching on the laser pulses. In this section, we will derive the relation explicitly. The relation between ${\bar{P}}_{\mu }$ and the asymptotic momentum of the particle pμ is governed by the ponderomotive force [62, 63], arising from the turn on process of the laser fields:

Equation (42)

Substituting equation (2) and keeping the envelope functions g1, g2, we have

Equation (43)

Suppose that the copropagating laser pulse with the amplitude ξ1 is turned on first. During this process the second integral for ${\bar{P}}_{z}$ is vanishing, and the particle momentum reads

Equation (44)

Here ϕ1 = (kp)τ/m, because in the absence of the counterpropagating pulse k1P is exactly conserved. It is worthwhile to mention that this result is similar to the one corresponding to the plane wave case. Now the second pulse is turned on. Its contribution to the momentum is given by

Equation (45)

Recalling the approximation derived above equation (26), and assuming that the pulse is turned on adiabatically, namely g2'/g2 → 0, the oscillatory part of the phase may be omitted, yielding ${g}_{2}\left({\phi }_{2}\right)\approx {g}_{2}$ $\left(\frac{{k}_{2}\cdot \bar{P}}{m}\tau \right)$. Since the first pulse effect comes into play through the neglected oscillatory term in ϕ2, it does not influence the integration. We further assume that ${k}_{2}\cdot \bar{P}$ remains constant during the turn on of the second pulse. Then, the integral in (45) is straightforwardly carried out, yielding for ${\bar{P}}_{z}$ and $\bar{\varepsilon }$:

Equation (46)

where equation (44), and $\bar{\varepsilon }=\sqrt{{m}_{{\ast}}^{2}+{\bar{P}}_{z}^{2}}$ were employed. Hence

Equation (47)

Examining the final momentum (47), one may observe that our assumption ${k}_{2}\cdot \bar{P}={k}_{2}\cdot {\bar{P}}^{\left(1\right)}$ was justified. We underline that

Equation (48)

Namely, neither k1P nor k2P are conserved. One may observe that k1P is modified during the rise of the counterpropagating pulse and vice versa. In case the counterpropagating beam is turned on first, an analogous derivation leads to

Equation (49)

where ${\bar{P}}_{\mu }^{\left(2\right)}={p}_{\mu }+\frac{{m}^{2}{\xi }_{1}^{2}}{2\left({k}_{2}\cdot p\right)}{k}_{2\mu }$.

The relation between the drift momentum and the asymptotic initial momentum has been also investigated by numerically solving the Lorentz equation (1) and comparing with the analytical results. Table 1 presents the average four-momentum of the electron after both laser beams are turned-on, corresponding to different initial momenta and intensities of the lasers. Since the order by which the lasers are turned-on affects the final state, the table contains both options. For the sake of simplicity, we assume the initial p = 0 for all situations. From the expression for the final momentum equations (47) and (49), one can see that two factors determine which of the beams will be dominant. The obvious one is the corresponding field intensity. The surprising one is the relative direction between the propagation direction of the particle and the beam under consideration. It stems from the denominator $k\cdot \bar{P}$, namely, counterpropagating beams have lower influence than copropagating beams.

Table 1. The average four-momentum of an electron after both laser beams are turned-on. We consider three different cases for different initial momentum pz and field parameters. In all the cases, the initial transverse momentum is chosen to be zero such that ${\bar{P}}_{x}={\bar{P}}_{y}=0$. The fifth column, named order, indicates the order by which the two laser beams are turned-on; (a) the ξ1 beam is turned-on first. (b) The ξ2 beam is turned-on first. The four-momentum is given in units of the electron rest mass m. The superscript N designates the numerical calculation and A the analytical one.

Case ξ1 ξ2 pz Order ${\left(\bar{\varepsilon },{\bar{P}}_{z}\right)}^{\mathrm{A}}$ ${\left(\bar{\varepsilon },{\bar{P}}_{z}\right)}^{\mathrm{N}}$
110100a(51.495, 49.505)(51.502, 49.450)
b(51.495, −49.505)(51.502, −49.450)
232020a(200.637, 199.613)(200.637, 199.612)
b(25.471, 15.452)(25.835, 15.321)
3302−1a(187.816, 185.391)(187.815, 185.390)
b(43.522, 31.451)(29.531, −0.396)

In the first case the two beams have identical intensity and the particle is initially at rest, so the only thing that breaks the symmetry is the turn-on order. It demonstrates that a given beam will have a stronger influence if it is the first to be turned on. The reason is that after the turned-on, the particle will copropagate with the first beam and thus this beam will have a large influence in the final results. This also reflects in the direction of the average momentum as in this case the particle always copropagates with the first beam at the end, see in table 1.

For the second case appearing in the table, one may naively assume the ξ2 beams should be dominant, since ξ2/ξ1 ∼ 7 and the contribution to the final momentum of each beam scales like ∼ξ2. However, due to the fact that ξ1 is copropagating, its effect is actually of the same order of magnitude as of the ξ2 beam. This can be seen by the fact that the order of the turn-on causes an order of magnitude difference between the final energies. Namely, when ξ1 is turned-on first (order (a)), the particle is first accelerated to ultrarelativistic energy and then slightly decelerated when ξ2 is turned-on. The final energy is about ɛ ≈ 200m, which is much larger than the final energy ɛ ≈ 25m of the second scenario (order (b)), where the particle is first decelerated and then accelerated.

In the third case a new situation is encountered. The particle flips its direction of motion during the turn-on of the second beam if the ξ2 beam is turn-on first, order (b) in the table. One may see that it initially propagates to the left and only after the second pulse rises it flips direction and propagates to the right. Both from analytical and from experimental perspectives, such a scenario should be avoided. From an experimental point of view, as it will lead to collisions of electrons in the beam with those following them. From analytical perspective, since a direction flip implies that the particle average velocity should vanish at a certain point in the middle of the turn-on process, violating the validity conditions. Indeed, the analytical expression in this case fails to reproduce the numerical result.

To complete the discussion, we specify several considerations which were taken into account when choosing the above parameters. First, we made sure that the validity criteria derived above are met. Second, the final propagation direction is always copropagating with the first turn-on beam, in agreement with the convention introduced in the previous subsection. Third, both laser amplitudes were chosen to be higher than 1. Since the contribution of each beam scales as ∝ξ2, the influence of a beam with nonrelativistic intensity on the final momentum can be neglected.

2.4. Systematic errors analysis of the trajectory

In the following, the accuracy of the analytical solution derived in the previous sections is systematically put to a test. For the sake of this purpose, we define the relative deviation of the analytical prediction (subscript a) of a quantity X with respect to the numerically calculated value (subscript n) as follows

Equation (50)

with X being either the transverse velocity ${v}_{\perp }\equiv \sqrt{{v}_{x}^{2}+{v}_{y}^{2}}$, the longitudinal one vz or the energy ɛ. The integration time is taken to infinity, i.e. T. These deviations are explored as a function of ${{\epsilon}}_{1}\equiv 2{p}_{x}m{\xi }_{1}/{m}_{{\ast}}^{2}$ and ${{\epsilon}}_{2}\equiv {m}^{2}{\xi }_{1}{\xi }_{2}/{\bar{\varepsilon }}^{2}$, being the small parameters of the derivation for ultrarelativistic particles, see equation (35). In the following we restrict ourselves to ξ1 > ξ2, and hence the third small parameter is by definition smaller than epsilon2, namely ${p}_{x}m{\xi }_{2}/\left(2{\bar{\varepsilon }}^{2}\right){< }{{\epsilon}}_{2}$. For quantitative comparison of the analytical and numerical quantities, we specify the arbitrary constants Φ1, Φ2 in the phases ϕ1, ϕ2, which is accomplished in two ways, see figure 4. The first way (blue solid line) is to write down Φ1 = k1x0, Φ2 = k2x0 where x0 denotes the temporal and spatial location of the particle at the moment when the turn-on process is over, associated with the numerical calculation. The second way (blue dashed line) is to advance the analytical solution in time from t → − using the relation (47). It should be mentioned that this approach is not fully analytical since ${k}_{1}\cdot \bar{P}$ is not constant during the turn-on of the second pulse, but rather depends on ξ1, which in its turn depends on ${k}_{1}\cdot \bar{P}$ through the phase ϕ1. Nevertheless, assuming that the turn-on process is adiabatic, ξ1 barely changes from one time step to its subsequent one, so that one can use the value of the field amplitude from the previous time step.

Figure 4.

Figure 4. The analytical (blue) and numerical (red) longitudinal velocity for (a) epsilon1 = 0.05 and (b) epsilon1 = 0.1. In both cases epsilon2 = 0. The dashed line corresponds to advancing the analytical solution in time through the turn-on process and the solid line to determining Φ1, Φ2 according to the particle location when the turn on is over, see the main text. The time is in the unit of T = 2π/ω.

Standard image High-resolution image

Before our broader parameter survey, it is worthwhile to take a close look at a specific case in order to gain intuition regarding the nature of deviation. Figure 4 presents the longitudinal velocity for ξ1 = 50, ɛ = 200m, px = 0 for a different ξ2 value: ξ2 = 20 and ξ2 = 40 corresponding to epsilon1 = 0.05 (upper plot) and epsilon1 = 0.1 (lower plot), respectively. One can see that the main deviation stems from an inaccuracy in the phases ϕ1, ϕ2, rather than in the amplitudes. Moreover, the comparison between the dashed and solid curves in figure 4 demonstrates that the two approaches to determine the phases Φ1, Φ2 give similar results and the first method was employed in the following calculations.

Figure 5 depicts the relative deviations Δ, Δz , Δepsilon defined in equation (50) as a function of epsilon1, epsilon2. We fixed the parameters ξ1 = 100, ɛ = 200m and varied ξ2, px in the ranges 0 < ξ2 < 60, 0 < px < 15m, respectively. The initial momentum on the z axis was tuned in order to keep the energy ɛ constant. One may notice that for vanishing ξ2, the analytical calculation is accurate regardless of the value of px . This occurs due to the fact that vanishing of ξ2 corresponds to the plane wave limit, where the analytical solution equation (36) is exact without any restriction.

Figure 5.

Figure 5. The relative deviation between analytical and numerical calculations of the following quantities: (a) the transverse velocity; (b) the longitudinal velocity; (c) the energy. The x, y axes are the small parameters stemming from the analytical derivation as ${{\epsilon}}_{1}\equiv 2{p}_{x}m{\xi }_{1}/{m}_{{\ast}}^{2}$ and ${{\epsilon}}_{2}\equiv {m}^{2}{\xi }_{1}{\xi }_{2}/{\bar{\varepsilon }}^{2}$. The average energy is fixed to be ɛ = 200m and the amplitude of the first beam is ξ1 = 100. The initial momentum and the ξ2 beam amplitude are in the range 0 < px < 15m and 0 < ξ2 < 60, respectively.

Standard image High-resolution image

Furthermore, we can see from all three panels in figure 5 that the influence of the small parameter epsilon2 is considerably stronger as compared to that of epsilon1. This is because epsilon2 affects the phase while epsilon1 stems from the Taylor expansion of the energy (equation (18)) and the discrepancy mainly originates from the dephasing in time, as shown in figure 4. The amplitude of the oscillation in v, vz and ɛ, on the other hand, can be predicted quite well by the analytical expression even for nonnegligible epsilon1, epsilon2, which provides us a way to crudely estimate the relative error. For example, the average and oscillatory parts of v may be roughly estimated, respectively, as ∼1/ɛ and ∼2/ɛ, and therefore the relative error is approximately Δξ2/ξ1. The relative error for vz is $\sim {m}^{2}{\xi }_{1}{\xi }_{2}/\left({\bar{v}}_{z}{\varepsilon }^{2}\right)$, which is smaller than Δ. For the energy, according to (19), the oscillations are closely related to epsilon1, i.e. δɛ/ɛepsilon1/2. If we plug in the simulation parameters, these estimations qualitatively explain that for the same small parameter epsilon1 and epsilon2, v has the largest deviation. For figure 5(c), we can see that the analytical results predict a very good approximation when epsilon2 = 0 no matter how large epsilon1 is. The reason is that when px is zero the energy is constant (see equation (19)). Consequently, the phase plays no role and the approximation is quite good for the entire range of epsilon1 values presented in the figure.

3. Radiation

Using the classical trajectory developed in section 2.1, the radiation is calculated according to the Baier–Katkov method [49]. For the sake of simplicity, we start with a spinless particle. Analogous derivation for the spinor case is given later. The Baier–Katkov expression for the emitted intensity I reads

Equation (51)

where α is the fine structure constant, T0 is the interaction time, ɛ' = ɛω', and

Equation (52)

where vμ = dxμ /dt, ${k}_{\mu }^{\prime }$ is the emitted photon four-momentum characterized by its energy ω' and the emission direction n as

Equation (53)

with $\mathbf{n}=\left(\mathrm{cos}\enspace \varphi \enspace \mathrm{sin}\enspace \theta ,\mathrm{sin}\enspace \varphi \enspace \mathrm{sin}\enspace \theta ,\mathrm{cos}\enspace \theta \right)$. In the realm of this theory, the oscillation δɛ are assumed to be small as compared to ɛ, which holds in our case as shown in the previous section. Accordingly, the factor appearing in the phase may be approximated as $\frac{\varepsilon }{{\varepsilon }^{\prime }}\approx \frac{\bar{\varepsilon }}{{\bar{\varepsilon }}^{\prime }}\left[1+\frac{{\left(\delta \varepsilon \right)}^{2}}{\bar{\varepsilon }{\bar{\varepsilon }}^{\prime }}\right]$. In the following derivation the second order correction is neglected. Moreover, for simplicity reasons, the average energy $\bar{\varepsilon }$ is replaced from now on by ɛ. Furthermore, since the trajectory is given in terms of the proper time τ, we change the integration variable in equation (52), leading to

Equation (54)

The actual calculation takes two steps. First, the Lorentz equation for a single particle is solved and the trajectory is obtained. Second, the time integration in the photon emission amplitude is calculated using the time-dependent momentum and coordinate of the particle. Both steps can be proceeded either analytically when an analytical trajectory is available or numerically for general laser field and electron beam parameters. In order to make sure that the integral in equation (52) in our numerical calculation is converged, we have to solve the classical trajectory numerically with sufficiently small Δt between two time steps, especially with large emitted photon energy as Δt should be much smaller than 2π/ω'. For a high electron energy (ɛ ranging from several hundreds of MeV to GeV) and a strong laser pulse (the classical parameter ξ being around several hundreds) the analytical model is indispensable to calculate the whole spectrum, and numerically we confirm the result at discrete points in the spectrum. Below we will give an analytical derivation of the spectrum using the approximated trajectory of the electron given above. By substituting the trajectory and the emitted wavevector (53) into expressions (54), the phase can be obtained like

Equation (55)

where $u\equiv \frac{{\omega }^{\prime }}{\varepsilon -{\omega }^{\prime }}$, and the following quantities were introduced

Equation (56)

The linear term coefficient in (55) reads

Equation (57)

The phase ψ may be simplified by introducing

Equation (58)

Therefore, the phase takes the form

Equation (59)

Notice that in the particular case of px = 0, the second term in the expressions for ${z}_{1}^{x},{z}_{2}^{x}$ vanishes, leading to

Equation (60)

as well as φ1 = φ2 = φ. Let us calculate the y component of $\mathcal{T}$ in detail. Employing (9) and (52) we obtain

Equation (61)

In order to analytically solve this integral, the identity [5]

Equation (62)

is invoked. The functions B0, B1, B2 are related to the Bessel function and its first derivative ${J}_{s}\left(z\right),{J}_{s}^{\prime }\left(z\right)$ through

Equation (63)

As a result, the integral in (61) is solved, yielding

Equation (64)

where 1 ≡ (s1, z1, φ), 2 ≡ (s2, z2, φ) and 3 ≡ (s3, z3, 0) respectively, and the δ function argument is given by

Equation (65)

One may notice that different combinations of the indices s1, s2, s3 may yield the same δ function argument. As a result, when squaring $\mathcal{T}$, interference terms will arise. This interference depends on the quantity ω2/ω1. If this ratio is a rational number, the motion is periodic with the frequency 2π/(1) with n being the decimal part of the rational number. Otherwise, the motion is non-periodic. In the following we discuss each of the cases separately.

3.1. The non-periodic case

In the non-periodic case, when the ratio ω2/ω1 is not an integer, it is convenient to define sLs1 + s3, sRs2s3. Hence, one may write

Equation (66)

with ${{\Omega}}_{{s}_{\mathrm{L}},{s}_{\mathrm{R}}}\equiv {\psi }_{np}-\frac{\varepsilon }{m}\left({s}_{\mathrm{L}}{\omega }_{1}+{s}_{\mathrm{R}}{\omega }_{2}\right)$. The matrix element takes the form

Equation (67)

An analogous procedure may be applied for the other components as well, yielding

Equation (68)

where

Equation (69)

Equation (70)

Equation (71)

As the squaring $\mathcal{T}$ does not mix terms associated with different sL, sR indices, the interference takes place only between terms included within $\mathcal{M}\left({s}_{\mathrm{L}},{s}_{\mathrm{R}}\right)$. Finally, the emitted intensity may be obtained by integrating (51) over the polar angle.

Equation (72)

where the identity ${\delta }^{2}\left({{\Omega}}_{{s}_{\mathrm{L}},{s}_{\mathrm{R}}}\right)=\frac{{\tau }_{0}}{2\pi }\delta \left({{\Omega}}_{{s}_{\mathrm{L}},{s}_{\mathrm{R}}}\right)$ is used. The proper interaction time is given by τ0 = (m/ɛ)T0. The condition imposed by the δ function, ${{\Omega}}_{{s}_{\mathrm{L}},{s}_{\mathrm{R}}}=0$, determines the relation between cos θ and ω', φ

Equation (73)

Squaring and solving this equation one obtains two possible angles

Equation (74)

where the following quantities were introduced

Equation (75)

Notice that when squaring (73) a redundant solution may be added, which solves the equation

Equation (76)

rather than the original one. Thus, the solutions given in (74) are physical only when a positive results appear after substituting it into the right wing of (73). In quantitative terms, this condition reads

Equation (77)

A solution that does not meet this criterion is therefore excluded. Employing the δ function to perform the integration leads to

Equation (78)

The reciprocal of the derivative of the δ function, required for the integration, reads

Equation (79)

Plugging (79) into (78) the final result follows

Equation (80)

For spinor particle the initial emission expression (51) is modified as follows

Equation (81)

Therefore, the final results for scalars (80) is multiplied by $\left(\frac{{\varepsilon }^{\prime 2}+{\varepsilon }^{2}}{2\varepsilon {\varepsilon }^{\prime }}\right)$ and a second term is added

Equation (82)

3.2. The periodic case

Now, we consider the case of periodic motion when ω2 = 1, with an integer n. As a result, the kinematic relation which follows from the δ function is modified. Using the relation between sL, sR and s1, s2, s3 one obtains sL ω1 + sR ω2s* ω1, with the definition s*s1 + ns2s3(n − 1). Accordingly, ρ in (75) is replaced by $\rho =\frac{{\varepsilon }^{\prime }}{\varepsilon {\omega }^{\prime }}{s}_{{\ast}}{\omega }_{1}$. As a consequence of the periodicity, the summation over s2 takes place inside the matrix element, similarly to s3. Correspondingly, we have

Equation (83)

Equation (84)

Equation (85)

Equation (86)

Compared to the non-periodic case, the interference between different harmonics in the spectrum is much more complicated in the periodic case as there is a double summation inside the squaring of the matrix elements. The final result, analogous to (82) of the non-periodic case, is given by

Equation (87)

It is worth to point out that the periodic case is most likely to be observed in a short laser pulse, when the condition ω2 = 1 can be fulfilled within the broad bandwidth of the laser pulse. We discuss this issue below.

3.3. Vanishing initial transverse momentum

In this subsection several quantities are explicitly evaluated for the particular case of vanishing initial transverse momentum, px = 0. It allows us to simplify the expressions and thus to obtain order of magnitude estimations which will prove useful later on. Substituting px = 0 to (74), the emitted photon angle reduces to

Equation (88)

Since ρ ≪ 1, the corresponding sine function is approximately given by $\mathrm{sin}\enspace \theta \approx \sqrt{1-\frac{1}{{\bar{v}}_{z}^{2}}+\frac{2\rho }{{\bar{v}}_{z}^{2}}}$. Substituting this expression to the Bessel arguments definitions (60) one obtains

Equation (89)

where we have defined ${u}_{s}\equiv \frac{2\varepsilon \left({s}_{\mathrm{L}}{\omega }_{1}+{s}_{\mathrm{R}}{\omega }_{2}\right)}{{m}_{{\ast}}^{2}}$ and the relation ${\bar{v}}_{z}^{2}=1-\frac{{m}_{{\ast}}^{2}}{{\varepsilon }^{2}}$ was employed. The maximal value of z1, z2 corresponds to u = us /2, namely

Equation (90)

3.4. Spectra in the strong field regime: ξ1 ≫ 1

In what follows we consider in detail the case where the copropagating beam is of relativistic intensity. It should be stressed that the spectrum may not be approximated by LCFA even though ξ1 ≫ 1. The physical conditions and the nature of this specific LCFA violation is discussed in [41].

In the strong field regime the argument of the Bessel function in equation (58) can be the order of 108 or even larger with the increasing of the laser field strength. This means the sum over the harmonics in the emission spectrum covers an extremely large region. In order to make the calculation feasible, we have employed an optimised scheme for the calculation, based on the logic proposed by Ritus and Sov [5].

It is well known that an ultrarelativistic particle emits mainly within a cone of angle ∼1/γ along its propagation direction. Hence, the emission angle θ may be approximated by the angle of the particle's momentum between P with respect to the z axis. Examining the classical momentum P, one observes that this angle lies in the range sin θd < sin θ < sin θu and its time-averaged value is sin θc , where

Equation (91)

In the case considered here, namely ξ ≫ 1, ξ2ξ1, and due to px 1 (see equation (20)), this range is very narrow and the angle may be crudely estimated according to the average value θc . Accordingly, one may show that the second term in the brackets appearing in the expression for ${z}_{1}^{x},{z}_{2}^{x}$ is negligible. As a result, the px = 0 expressions (60) provides an order of magnitude estimation for z1, z2, z3. Plugging in sin θc m*/ɛ, cos θc ≈ 1 one obtains

Equation (92)

Notice that ${z}_{1}^{c},{z}_{2}^{c}$ coincide with the maximal value possible for these quantities, see equation (90). Furthermore, one may observe that since ${\bar{v}}_{z}\approx 1$ and m*1 we have ${z}_{3}^{c}\approx {z}_{2}^{c}$.

In the following we take advantage of these relations in order to accelerate the harmonics summation (sL, sR, s3) appearing in the final emission formula (82) as well as derive simplified validity conditions. We follow the logic presented by Ritus and Sov [5] for emission in a circularly polarized laser. Since uχ, these arguments may be much larger than 1. As a result, the number of harmonics contributing to the emission may be enormous, and an efficient way to carry out the summation is required. First, we replace the summation by integration. Second, since Bessel function of high order is maximal for zs and strongly suppressed for either zs or zs, the integration is centered around

Equation (93)

In order to estimate the integration range, we define ${z}_{1}^{d},{z}_{2}^{d},{z}_{3}^{d}$ and ${z}_{1}^{u},{z}_{2}^{u},{z}_{3}^{u}$ analogously to ${z}_{1}^{c},{z}_{2}^{c},{z}_{3}^{c}$ appearing in equation (92) with θ = θd and θ = θu respectively. Accordingly, the upper and lower limits of the integration are respectively

Equation (94)

In mathematical terms, our improved summation scheme may be formulated as

Equation (95)

The replacement of the summation by integral in the calculation is appropriate only when su sd is large enough, which strongly depends on the chosen parameters.

In the following we present typical spectra in the strong field regime (ξ1 = 20, ξ2 = 0.3, px = 0), and use it to discuss the differences between the periodic and non-periodic cases derived above. Please note that the radiation reaction is neglected in the applied parameter regime, as the energy emitted during one laser cycle is very small compared with the electron energy.

Non-periodic case versus periodic case with px = 0. In figures 6(a) and (c), we consider the non-periodic case. The energy was chosen to be ɛ = 4m*, so that the ratio ω1/ω2 is approximately 60.1. Since the radiation of an ultrarelativistic electron emitted to a certain direction originates from the vicinity of the location where the particle velocity points to the detector, it implies that the emission should depend on φ. However, due to the non-periodicity, we can see from figure 6(a) that the emission at a given φ takes place with different χ values in different ω1 cycles. As a result of this χ-averaging, the difference between emissions at various φ disappears with long enough pulse of ξ1, see the spectra in figure 6(c) for three different φ. The spectrum was evaluated with the aid of the non-periodic formula (82) together with (95).

Figure 6.

Figure 6. The time-dependent quantum parameter χ(t) as a function of φ for both non-periodic case (a) and periodic case (b). The solid black curves are for the one cycle of the laser pulse and the dashed blue curves are for another cycle that is ten period away. The emission spectra for non-periodic case (c) and periodic case (d). Here φ1 is corresponding to χ(t) near maximum, φ2 is for χ(t) in the middle and φ3 is for χ(t) near minimum. The other parameters are ξ1 = 20 and ξ2 = 0.3 with px = 0. The non-periodic case is for ɛ = 4m* with ω2/ω1 ≈ 60.1, while the periodic case is for ω2/ω1 = 60 (ɛ = 4.02m*).

Standard image High-resolution image

In the periodic case, figures 6(b) and (d), the electron energy is tuned ɛ = 4.02m* to fulfil the integer ratio ω2/ω1 = 60. As opposed to the non-periodic case, here a particular value of χ parameter corresponds to the emission at a given φ at any period of the trajectory (see figure 6(b)), and therefore the emission depends on φ. In figure 6(d) the black, blue and red curves are calculated with different values of φ, respectively. One may see that these three curves significantly differ from each other.

Now we examine numerically the spectrum obtained for the non-periodic case, but with finite number of cycles in the laser pulse (as compared to the infinite pulse assumed by the analytical derivation). For the numerical calculations, we have evaluated equation (51) numerically, employing the numerical trajectory for the electron, as for realistic laser pulses the trajectory is not available analytically. In figure 7 full (hollow) circles designate ten cycles with φ corresponding to, respectively, the minimum (maximum) χ and full (hollow) squares are for five cycles with the same φ. First of all, the non-periodic spectrum, which represents averaging over φ, lies indeed in the middle between those curves, as expected. Secondly, one may see that the spectra for the finite laser pulse are far from the infinite pulse calculation. Moreover, the shorter the pulse is, the closer the results are to the periodic case. The reason is that the averaging out of the azimuthal dependence, as explained above, requires many cycles of interaction. The criterion which determines when one may employ the periodic formula is that the χ-averaging is not significant, namely

Equation (96)

where N is the number of cycles in the laser pulse, n = ω2/ω1 and n* is the closest integer number to n. For the parameters considered above this quantity reads 0.5 and 1, respectively. Consequently, the periodic expression provides a good estimation to the final result for short laser pulse, provided that the condition (96) is fulfilled.

Figure 7.

Figure 7. The emission spectra for the same parameters as in figure 6. The blue curve corresponds to the non-periodic case and the black solid and dashed lines correspond to the periodic case for specific values of φ corresponding to the minimum χmin and the maximum χmax, respectively. The circles (squares) designate the numerical calculation of a finite pulse of ten (five) cycles. The filling (open) markers are related to φ for χmin (χmax).

Standard image High-resolution image

It is worth to point out that there is a certain regime where the emission for χmin is larger than for χmax in figure 6(d). This is because in the region of χmin along the electron's trajectory, the emission of the electron is not uniquely determined by the quantum parameter χ as commonly believed. This is because of the violation of the LCFA. In the formation length around χmin, χ(t) changes rapidly and increases up to the order of χmax and thus the emission is also similar near χmax or even larger, see more discussions in reference [41].

Non-periodic case versus periodic case with px ≠ 0. Previously, we have discussed the emission of an electron in counterpropagating waves with vanishing transverse momentum. However, in a realistic experimental setup, the electrons in a beam always have non vanishing transverse momentum because of the angle spreading of the beam. In order to study the influence of the transverse momentum on the radiation process, we have in this section calculated the emission spectrum of an electron with px ≠ 0 for both non-periodic and periodic cases.

In figure 8, the spectra for px being 0.25% of the total energy have been investigated. Both of the spectra are not symmetric with respect to the azimuthal angle φ as the x-direction is favorable. For the non-periodic case, even the gradual shift of χ regarding the azimuthal angle still happens for px ≠ 0, the spectrum is nevertheless φ dependent because the transverse momentum breaks the symmetry. Furthermore, the spectrum for the periodic case with nonzero px has fringes with respect to φ. This means that the quantum parameter χ still has the similar dependence on the azimuthal angle like in figure 6(b).

Figure 8.

Figure 8. The emission spectra for the non-periodic case (a) and periodic case (b). Here we have px = 0.25m, being about 0.25% of the total energy, for both cases. The non-periodic case is for ω2/ω1 ≈ 60.2, while the periodic case is for ω2/ω1 = 60. The other parameters are the same as in figure 6.

Standard image High-resolution image

3.5. Validity condition

In the following we derive the validity conditions for the emission formula obtained in the previous section. For this purpose, we recall that the next order correction to the trajectory employed in this paper reads

Equation (97)

Let us substitute these modifications into the expression (55) for the phase ψ and examine the additional terms. Next, we take advantage of the identity

Equation (98)

Please note that κ here is just a parameter and not related to equation (79). Consequently, sin, cos functions in (55) are replaced according to

Equation (99)

Equation (100)

with Λ denoting $\left(\alpha ,{\beta }_{1},{\kappa }_{1},{\beta }_{2},{\kappa }_{2},\varphi \right)$. Using (98) one obtains

Equation (101)

Equation (102)

In the previous section the trigonometric identity ${z}_{1}^{x}\enspace \mathrm{cos}\enspace {\phi }_{1}+{z}_{1}^{y}\enspace \mathrm{sin}\enspace {\phi }_{1}={z}_{1}\enspace \mathrm{sin}\left({\phi }_{1}-{\varphi }_{1}\right)$ was employed, where z1, φ1 are given by (58) respectively. Analogously, in this case we have

Equation (103)

where

Equation (104)

As a result, the modified phase may be written as

Equation (105)

where

Equation (106)

Let us estimate the neglected contribution to the phase, namely the difference between (59) and (105). For the sake of simplicity, we split the corrections to three contributions, Δψ1, Δψ2, Δψ3 associated with z1, z2, z3, respectively.

Equation (107)

In explicit terms, the corrections take the form ${\Delta}{\psi }_{1}=-{z}_{1}\left[{\mathcal{I}}_{2}\left({{\Lambda}}_{1}\right)-\mathrm{sin}\left({\phi }_{1}-{\varphi }_{1}\right)\right]$, and for Δψ2, Δψ3 we have z1z2, z3. Since C1, C12 ≪ 1, we consider only first order corrections, namely s1 = 0, s2 = ±1 and s1 = ±1, s2 = 0. Therefore, one readily obtains

Equation (108)

where J1(β) = −J−1(β) ≈ β/2 was used. Using equations (104) and (108) yields

Equation (109)

where ω1/ω2 ≪ 1 was employed. Analogously, for the other contribution one finds

Equation (110)

The following coefficients were defined

In order to formulate the general validity condition, we notice that the phase (59) contains a linear term with low (ω1) and high (ω2, ω2ω1) frequencies. Therefore, we require that the coefficients of the high frequency corrections, will be smaller as compared to z2, z3. Similarly, the coefficients of the low frequency should be lower than z1, and the one corresponding to the linear term smaller than ψnp . Hence the general validity condition may be cast in the form

Equation (111)

We call attention to the fact that these conditions depend on the emitted photon properties ω, θ, φ. As a result, for given interaction parameters (laser amplitudes, particle energy), part of the spectrum may be described by our analytical expression whereas a different part may exhibit deviations. Hence, one should verify that (111) holds for the entire spectral range of interest. In the strong field case, however, the situation is much simplified and simple criteria are derived, which hold for the entire spectrum.

Let us consider explicitly the strong field regime (ξ1 ≫ 1). As explained in section 3.4, in this regime the emission is restricted to a limited angle range, for which the Bessel coefficients may be approximated by ${z}_{1}^{c},{z}_{2}^{c},{z}_{3}^{c}$. Substituting these expressions to the requirement (111) and employing the trajectory validity conditions in section 2.4 as well as the approximation $\frac{{\omega }_{2}}{{\omega }_{1}}=\frac{1+{\bar{v}}_{z}}{1-{\bar{v}}_{z}}\approx \frac{4{\varepsilon }^{2}}{{m}_{{\ast}}^{2}}$, we find that η4, η5, η7, η8 obey (111) by definition. Employing equation (92) as well as the expressions for C1, C12, the validity condition is simplified to

Equation (112)

The last three conditions are automatically fulfilled according to the validity conditions for the trajectory. Hence, only a single additional condition, corresponding to η1 in equation (112), is required to validate the applied formalism:

Equation (113)

As demonstrated above, the analytical approximation depends on several criteria being fulfilled. In the following, we examine in detail the strong field case, where the number of quantities required to be low is relatively small, allowing for a tractable study of the error. The main quantities, which stem from the trajectory approximation, are epsilon1, epsilon2 given in section 2.4. In the following we investigate systematically and quantitatively the relation between these parameters and the corresponding error. For the sake of this purpose, a new quantity is introduced

Equation (114)

measuring the relative difference Δ (in percent) between the analytical (Ia) and numerical (In) results. It is evaluated in three points ui on the spectrum. Figure 9 shows the deviation Δ as a function of epsilon1, assuming that epsilon2 vanishes. One can see that error grows monotonically, and that epsilon1 = 0.1 yields a deviation of 15%. In order to examine the influence of epsilon2, the calculation was generalized to two dimensions, and the results are presented in table 2. One may see that the impact of epsilon2 is significantly smaller as compared to epsilon1. The results presented in this subsection provides quantitative information which may be valuable when applying our expressions in practice.

Figure 9.

Figure 9. The relative deviation in percentage between the analytical and the numerical spectrum as a function of the small parameter epsilon1. See text for details.

Standard image High-resolution image

Table 2. The relative difference between the analytical and numerical results as a function of the parameters epsilon1 and epsilon2.

  epsilon2
epsilon1 0.030.060.090.120.15
0.033.0110.2513.8518.3525.46
0.063.0510.6014.1919.1026.19
0.093.1011.0014.7219.8326.99
0.123.1911.4015.5720.5627.81
0.153.3211.8016.7921.4528.90

3.6. Realistic pulse effects

The analytical derivation presented above assumes that the laser fields are monochromatic plain waves. This approximation is appropriate for long pulses (dozens of cycles) which are focused on large spots (radius of dozens of wavelengths). However, realistic pulses tend to be short and tightly focused, in order to maximize the obtained intensity for a given pulse energy. Therefore, for practical reasons it is highly important to thoroughly examine the dependence of the emission on pulse duration and focal size. In particular, we wish to establish qualitatively which spectral features are affected by shortening/focusing the laser pulse, what is the amplitude of the deviation and to find the conditions for which the spectrum recovers the analytical result.

In order to specify the spatial and temporal shape for the realistic laser pulse, the following quantities are introduced

Equation (115)

The vector potential corresponding to this pulse reads [64]

Equation (116)

where A0 denotes the amplitude. Since we consider a circular polarization, the y component is given by Ay = iAx . The EM fields can thus be derived from the above vector potential by

Equation (117)

In the calculation below, we choose a sin2-function with σ0 denoting the pulse length for g(kx) as the temporal envelope. Figure 10 depicts the angle integrated emission of a particle interacting with pulses with normalized amplitudes ξ1 = 12.5, ξ2 = 0.1, ɛ = 80m, respectively. The solid line stands for the analytical expression. Numerical calculations corresponding to variety of pulse durations and focal radii were carried out as well. For the sake of comparison, we wanted to keep the energy of the particle in the main part of the pulse identical for all compared cases. For this purpose, the initial electron energy was zero and its initial location z0 was tuned, namely the distance to the beginning of the ξ1 and ξ2 beams. From an experimental point of view, it may be realized by placing atoms which are ionized by the laser field.

Figure 10.

Figure 10. The emitted spectra for various pulse duration and spot sizes. w0 is the spot radius and σ0 the pulse duration. Here w0 and σ0 are presented in the unit of laser wave length λ and period T, respectively. The solid line stands for the analytical expression (w0, σ0).

Standard image High-resolution image

As expected, the analytical formula coincides fairly well with the numerical calculation for a long pulse with large focus w0 = 50λ, σ0 = 20T, with λ and T as the laser wave length and laser period, respectively. Let us examine the influence of the temporal width first. Decreasing the duration to σ = 10T does not change much the spectrum. However, for ultrashort pulses (full circles) with σ = 5T, the emission significantly increases. This may be explained by the fact that the rapid rise of the pulse is accompanied by stronger acceleration and enhanced χ value. Moreover, the shorter the pulse is, the larger the edge effect will be in the emission spectrum. This edge effect will induce deviations of the spectrum from the LCFA predictions and enhance the emission, especially in the high energy domain [41].

As for the spatial focusing, one observes an opposite trend. Namely, a small spot results in a significant decrease in the emitted spectrum, as well as in a deformation of its spectral shape. We suggest that this outcome stems from the fact that tightly focused beams rapidly expel the particle from the focus due to the transverse ponderomotive force. Furthermore, one may notice that even moderate focusing, w0 = 20λ, results in a considerable deviation from the one dimensional case. Thus, figure 10 shows that finite duration yields significant deviation from the analytical expression only for ultrashort pulses, whereas the focal radius has greater influence and should be fairly large in order to recover the theoretical result.

4. Summary and conclusion

We have investigated the radiation properties of a relativistic electron in counterpropagating laser waves within the semiclassical formalism introduced by Baier and Katkov. This formalism is valid when the electron dynamics in the background classical fields is quasiclassical. It treats a photon emission quantum mechanically, fully taking into account the quantum recoil of the emitted photon. As the formalism employs the electron classical trajectory in the given fields, we firstly investigate in detail the electron classical dynamics in the counterpropagating laser beam setup. The classical momentum and trajectory are analytically derived assuming that the particle energy is the dominant scale and that the angle between the particle propagation direction and the beams axis is small (see the exact conditions in equation (34)). The trajectory characteristics as a function of the laser parameters and the particle energy are discussed. In particular, we show that in the case when the quantum parameters induced by each of the beams are comparable, χ1χ2, a peculiar spike-like feature arises. Since its typical time scale is significantly shorter as compared to 1/ω1, 1/ω2, it will bear great significance to the corresponding radiation properties. Moreover, a detailed comparison with the full numerical solution was carried out resulting in a good agreement and validating our analytical solution in the given conditions. The dependence of the small deviations with respect to the exact solution on the parameters has been systematically investigated. We have observed an interesting relationship of the cycle-averaged momentum in the field to the asymptotic one. We show that the final average momentum depends on the order by which the laser beams are turned on.

Further, employing the approximated analytical trajectory, the radiation has been calculated in the Baier–Katkov semiclassical framework. The Baier–Katkov integrals were analytically solved yielding closed formulas in terms of sums over Bessel functions. Different regimes, periodic and non-periodic, are explored.

We concentrated on the strong field regime, which was found to be of particular interest for anomalous LCFA violation [41]. An optimised calculation method based on a physical reasoning is suggested, which enables quick summation over the numerous Bessel harmonics appearing in the analytical formula. The result is employed to compare in detail the periodic and the non-periodic regimes. We have observed that as opposed to the non-periodic case, where non-uniformity in the azimuthal direction finally averages to zero, in the periodic case considerable dependence on the azimuthal angle appears. We found that in a rather short laser pulse the emission in the non-periodic case becomes similar to the periodic one. Furthermore, we analyze numerically the effect introduced by a finite duration and spot size of the beams, which are not included in the analytical derivation. We demonstrate that the ultrashort pulse results in enhanced emission while tightly focused beam reduce the emitted energy and give physical explanations.

Finally, elaborated analytical analysis of the validity condition is presented. In the general case, it depends on the energy and angle of the emitted photon. In the strong field case, it reduces to a simple restriction on the ratio between the energy and the laser amplitude. The error in the spectrum is evaluated numerically and systematically explored as a function of the small quantities lying in the foundation of the theoretical approximation.

Acknowledgments

QZL and ER contributed equally to the work, to numerical and analytical calculations, respectively. ER acknowledges partial support from the Alexander von Humboldt Foundation.

Data availability statement

The data that support the findings of this study are available upon reasonable request from the authors.

Please wait… references are loading.