Next Article in Journal
Engineering and Materials: Editorial
Previous Article in Journal
Fractional Bernoulli and Euler Numbers and Related Fractional Polynomials—A Symmetry in Number Theory
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Generation of High-Energy Electron–Positron Pairs during the Breit–Wheeler Resonant Process in a Strong Field of an X-ray Electromagnetic Wave

by
Sergei P. Roshchupkin
*,
Vitalii D. Serov
and
Victor V. Dubov
Department of Theoretical Physics, Peter the Great St. Petersburg Polytechnic University, Polytechnicheskaya 29, 195251 Saint Petersburg, Russia
*
Author to whom correspondence should be addressed.
Symmetry 2023, 15(10), 1901; https://doi.org/10.3390/sym15101901
Submission received: 31 July 2023 / Revised: 6 September 2023 / Accepted: 9 October 2023 / Published: 10 October 2023
(This article belongs to the Section Physics)

Abstract

:
The Breit–Wheeler resonant process was theoretically studied in a strong X-ray electromagnetic wave field under conditions when the energy of one of the initial high-energy gamma quanta passes into the energy of a positron or electron. These conditions were realized when the energy of a high-energy gamma quantum significantly exceeded the characteristic Breit–Wheeler energy, which was determined using the parameters of the electromagnetic wave and the initial setup. Analytical formulas for the resonant differential cross-section were obtained. It is shown that the resonant differential cross-section significantly depends on the ratio between the energies of the initial gamma quanta and the characteristic Breit–Wheeler energy. With a decrease in the characteristic Breit–Wheeler energy, the resonant cross-section increases sharply and may exceed the corresponding non-resonant cross-section by several orders of magnitude. The results make it possible to obtain narrow beams of ultrarelativistic positrons (electrons) with energies of the order ∼ 10 2 GeV and could also be used to explain high-energy fluxes of positrons (electrons) near neutron stars, as well as to simulate QED processes in laser fusion.

1. Introduction

Obtaining narrow beams of ultrarelativistic positrons (electrons) is an important scientific problem. The Breit–Wheeler process (BWP) is well-known and involves the production of electron–positron pairs through the interaction of two gamma quanta [1]. In this process, the energy spectrum of the generated electrons and positrons is continuous (the energy of the initial gamma quanta is distributed between the electron and the positron). With the advent of high-intensity lasers [2,3,4,5,6,7,8,9,10,11], the study of quantum electrodynamics (QED) processes in external electromagnetic fields has been widely explored (see, for example, reviews [12,13,14,15,16,17] and articles [18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48]). Moreover, these processes can occur both resonantly and non-resonantly. The resonant behavior of high-order QED processes is associated with the entering of an intermediate particle into the mass shell. These resonances were first investigated by Oleinik [18,19]. As a result, higher-order processes by the fine-structure constant effectively reduce into several consecutive lower-order processes (see reviews [5,13,14]), as well as recent articles [26,27,28,29,42,43]). It is important to note that the probability of resonant processes can significantly exceed the corresponding probabilities of non-resonant processes.
Currently, there is a considerable body of work dedicated to studying BWP in an external electromagnetic field (see, for example, [30,31,32,33,34,35,36,37,38,39,40,41,42,43]). It is important to distinguish between the external field-stimulated BWP (a first-order process with the fine-structure constant) and the external field-assisted BWP (a second-order process with the fine-structure constant). Note that the production of narrow beams of ultrarelativistic electron–positron pairs in an external electromagnetic field has been studied in the articles in [27,28,29,41,43]. At the same time, in the articles in [27,28,29], the resonant process of photogeneration of ultrarelativistic electron–positron pairs in the field of the nucleus in a weak and strong electromagnetic field was considered. The article in [41] studied the Breit–Wheeler resonant process in a weak electromagnetic field (when the classical relativistically invariant parameter η (4) is much less than one).
In the recent article from the authors of [43], the resonant external field-assisted BWP in a strong field of a monochromatic circularly polarized wave, propagating along the z axis, was studied. However, in this paper, in contrast to previous articles devoted to the Breit–Wheeler resonant process in a strong electromagnetic field, we will study the case when the resonant energy of a positron (electron) is close to the energy of a high-energy initial gamma quantum. It is important to note that under the conditions of resonance and the absence of interference from different reaction channels, the original second-order process effectively reduces into two first-order processes: the external field-stimulated Breit–Wheeler process (EFSBWP), and the external field-stimulated Compton effect (EFSCE) [12] (see Figure 1). The resonant BWP for high-energy initial gamma quanta and ultrarelativistic final electron–positron pairs was considered
ħ ω 1 , 2 m c 2 , E ± m c 2 .
Here, ħ ω 1 , 2 and E ± are the energies of the initial gamma quanta and the final positron or electron. At the same time, all particles (the initial gamma quanta and the final electron–positron pair) move in a narrow cone away from the direction of wave propagation
θ j ± ( k j , p ± ) 1 , θ i ( k 1 , k 2 ) 1 ,
θ j ( k j , k ) 1 , j = 1 , 2 ; θ ± ( p ± , k ) 1 , θ 1 , 2 θ ± θ .
In expression (3), p ± are momenta of the positron or the electron. It was assumed that the classical relativistic-invariant parameter satisfies the following condition [12,27,28,29]:
η = e F ƛ m c 2 η m a x , η m a x = min E ± m c 2 .
Here, e and m are the charge and mass of the electron, F and ƛ = c / ω are the electric field strength and wavelength, and ω is the frequency of the wave.
It was shown that the resonant energy of positron and electron has a discrete spectrum. The outgoing angles of the electron and positron are uniquely related and determine their resonant energies. In addition, the resonant energies of the electron–positron pair for each reaction channel are determined by the quantum parameters ε 1 C ( r ) (at the first vertex for the EFSCE) and ε 2 B W ( r ) (at the second vertex for the EFSBWP):
ε 1 C ( r ) = r ε 1 C , ε 1 C = ħ ω 1 ħ ω C , ħ ω C = 1 4 ħ ω B W ,
ε 2 B W ( r ) = r ε 2 B W 1 , ε 2 B W = ħ ω 2 ħ ω B W ,
ħ ω B W = ( m c 2 ) 2 ( 1 + η 2 ) ħ ω sin 2 ( θ / 2 ) .
In expression (5), r = 1 , 2 , 3 represents the number of photons of the wave emitted at the first vertex, while ħ ω C is the characteristic energy of the EFSCE [27]. In Equation (6), r represents the number of photons absorbed at the second vertex, while ħ ω B W (7) is the characteristic energy of the EFSBWP [28,29]. Unlike the first vertex, the number of photons absorbed from the wave r at the second vertex significantly depends on the relationship between the energy of the second gamma quantum and the characteristic energy ħ ω B W . Thus, if the quantum parameter is ε 2 B W < 1 ( ω 2 < ω B W ) , then the number of absorbed photons of the wave at the second vertex starts from a certain minimum value: r r m i n = ε 2 B W 1 > 1 (see condition (6) for parameter ε 2 B W ( r ) ). However, if the quantum parameter is ε 2 B W 1 ( ω 2 ω B W ) , then the number of absorbed photons starts from one: r = 1 , 2 , 3 .
It should be emphasized that the case where the following conditions were imposed on the energies of the initial gamma quanta within framework (1) was thoroughly investigated:
ω 2 ω B W , ω 1 ω B W , ω 1 ω C ( ε 2 B W 1 , ε 1 B W 1 , ε 1 C 1 ) .
It should be noted that under these conditions, the exchange resonant reaction channels ( k 1 k 2 ) were suppressed and not considered.
It is important to highlight that from a physical point of view, the most interesting case is when the energy of the second gamma quantum significantly exceeds the characteristic energy of the EFSBWP, while the energy of the first gamma quantum is on the order of, or much smaller than, the characteristic energy of the EFSCE, which had not been studied previously. In this case, instead of relationships (8), the following conditions are obtained:
ω 2 ω B W , ω 1 ω B W , ω 1 ω C ( ε 2 B W 1 , ε 1 B W 1 , ε 1 C 1 ) ;
ω 2 ω B W , ω 1 ω C ( ε 2 B W 1 , ε 1 C 1 ) .
It is important to note that the characteristic Breit–Wheeler energy ħ ω B W (7), at a fixed parameter η value, is inversely proportional to the energy of the photon of the electromagnetic wave. As the photon energy of the wave increases from optical frequencies to X-ray frequencies, the characteristic Breit–Wheeler energy decreases, while the intensity of the wave increases. Let us estimate the characteristic energy of the EFSBWP ħ ω B W (7). In deriving the estimate, we will consider frequencies of the electromagnetic wave in the X-ray range, as well as the angle θ = π :
ħ ω B W = 17.4 G e V i f ω = 30 e V , I = 1.675 · 10 21 W c m 2 ( η 2 = 1 ) 522 M e V i f ω = 1 k e V , I = 1.861 · 10 24 W c m 2 ( η 2 = 1 ) 522 M e V i f ω = 10 k e V , I = 3.536 · 10 27 W c m 2 ( η 2 = 19 )
We note that in Equation (11), for frequencies of the X-ray wave ω = 30 eV and ω = 1 keV, the wave parameters were chosen such that the characteristic Breit–Wheeler energy decreases with increasing wave intensity, from ħ ω B W = 17.4 GeV to ħ ω B W = 522 MeV. At the same time, for X-ray frequencies ranging from ω = 1 keV to ω = 10 keV, the wave parameters were chosen such that the characteristic energy of the EFSBWP ħ ω B W = 522 MeV remains unchanged with increasing wave intensity.
In this article, we will study the resonant BWP in a strong X-ray field under the conditions of the initial gamma quanta energies specified by Equations (9) and (10). As will be shown later, under these conditions, the resonant energy of the positron (for Channel A) or electron (for Channel B) tends to be the energy of the high-energy second gamma quantum with the greatest probability.
Subsequently, the relativistic system of units is used: c = ħ = 1 .

2. Resonant Energies of the Positron (Electron)

Oleinik resonances occur when an intermediate electron or positron in an external field enters the mass shell. As a result, for Channels A and B, we obtain (see Figure 1):
q ˜ 2 = m * 2 , q ˜ = k 2 p ˜ + + r k = p ˜ k 1 + r k ,
q ˜ + 2 = m * 2 , q ˜ + = k 2 p ˜ + r k = p ˜ + k 1 + r k .
In Expressions (12) and (13), k 1 , 2 = ( ω 1 , 2 , k 1 , 2 ) and k = ( ω , k ) are 4-momenta of the first and second gamma quanta and the external wave photon, p ˜ ± = ( E ˜ ± , p ˜ ± ) and q ˜ are 4-quasimomenta of the final positron (electron) and intermediate electron (positron), and m * is the effective mass of the electron in the field of a circularly polarized wave [12]:
p ˜ ± = p ± + η 2 m 2 2 ( k p ± ) k , q ˜ = q + η 2 m 2 2 ( k q ) k ,
p ˜ ± 2 = m * 2 , m * = m 1 + η 2 ,
p ˜ i , f 2 = m * 2 , m * = m 1 + η 2 .
In expression (14), p ± = ( E ± , p ± ) —4-momenta of the positron (electron).
In this article, within the framework of conditions (9) and (10), we will consider the energies of the second gamma quantum ω 2 10 3 GeV, as well as the range of X-ray frequencies 10 e V ω 10 5 e V . At the same time, we will consider the intensities of the electromagnetic wave to be significantly smaller than the critical intensities of the Schwinger [17,46] ( I I * 10 29 W/cm 2 ).
Considering Equations (1)–(4), (12) and (13), it is straightforward to obtain the resonant energies of positron E + ( r ) (for Channel A) or electron E ( r ) (for Channel B) at the second vertex, where the EFSBWP occurs (see Figure 1):
E ± ( r ) = ω 2 2 ( ε 2 B W ( r ) + δ 2 ± 2 ) ε 2 B W ( r ) + ε 2 B W ( r ) ( ε 2 B W ( r ) 1 ) δ 2 ± 2 ,
where the ultrarelativistic parameter
δ 2 ± 2 = ω 2 2 ( 2 m * ) 2 θ 2 ± 2 ,
associated with the outgoing angle of the positron (electron) relative to the momentum of the second gamma quantum, can vary in the interval:
0 δ 2 ± 2 δ 2 m a x 2 , δ 2 m a x 2 = ε 2 B W ( r ) ( ε 2 B W ( r ) 1 ) .
It should be noted that the quantity (17) determines the maximum energy of the positron (electron). There is also a solution with a minus sign in front of the square root, which determines the minimum value of the resonant energy. However, the minimum resonant energy is unlikely and not used in the text below. Under the conditions (9) and (10), the quantum parameter ε 2 B W ( r ) = r ε 2 B W 1 ( r 1 ) . As a result, the resonant energy of the positron or electron (17) will be close to the energy of the high-energy second gamma quantum:
E ± ( r ) ω 2 1 ( 1 + 4 δ 2 ± 2 ) 4 r ε 2 B W ω 2 ( δ 2 ± 2 r ε 2 B W ) .
Figure 2 illustrates the resonant energy of the positron (for Channel A) or electron (for Channel B) as a function of the corresponding ultrarelativistic parameter δ 2 ± 2 for the energies of the second gamma quantum ω 2 = 600 MeV and ω 2 = 100 GeV. This graph presents two solutions for the resonant energy, corresponding to the “plus” (solid curves) and “minus” (dashed curves) signs before the square root in Equation (17). As mentioned earlier, only the maximum energy of the positron (electron) will be used in future (solid curves in Figure 2). From Figure 2, it can be observed that there is a significant difference in the resonant energies of the positron (electron) for these second gamma quantum energies. Unlike the case with ω 2 = 600 MeV, the resonant energy of the positron (electron) for the case with ω 2 = 100 GeV tends to converge to the energy of the second gamma quantum at small values of the parameter δ 2 ± 2 (see solid curve 2 in Figure 2, as well as Formula (20)).
Considering Equations (1)–(4), (12) and (13), it is straightforward to obtain the resonant energies of electron E ( r ) (for Channel A) or positron E + ( r ) (for Channel B) at the first vertex, where the EFSCE occurs (see Figure 1):
E ( r ) = ω 1 2 ( ε 1 C ( r ) δ 1 2 ) ε 1 C ( r ) + ε 1 C ( r ) 2 + 4 ( ε 1 C ( r ) δ 1 2 ) .
Here, the ultrarelativistic parameter, associated with the outgoing angle of the electron (positron) relative to the momentum of the first gamma quantum, is defined:
δ 1 2 = ω 1 2 m * 2 θ 1 2 .
It should be noted that there are no constraints on the parameter ε 1 C ( r ) for the EFSCE. At the same time, δ 1 2 < ε 1 C ( r ) .
It is important to note that in the conditions of resonance (12) and (13) for each of the reaction channels, the resonant energies of the positron and electron are determined by different physical processes: the EFSBWP (17) and the EFSCE (21). At the same time, the energies of the electron–positron pair are related to each other by the general law of conservation of energy
E + + E ω 1 + ω 2 .
It should be noted that in Equation (23) we have neglected a small correction term | r r | ω / ω 2 1 . Taking into account relations (20) and (21), as well as the law of energy conservation (23) for Channels A and B, we obtain the following equation relating the outgoing angles of the positron and electron within the conditions (9) and (10):
δ 1 2 = ε 1 C r r ε 1 C 1 + 4 δ 2 ± 2 + r ε 1 C r + r 1 + 4 δ 2 ± 2 + r ε 1 C .
Here, the upper (lower) sign corresponds to Channel A (B). In relation (24), to the left of the equality sign is the ultrarelativistic parameter responsible for the outgoing angle of electron (positron) relative to the momentum of the first gamma quantum, and to the right of the equality sign is the function of the ultrarelativistic parameter δ 2 ± 2 , which corresponds to the outgoing angle of the positron (electron) relative to the momentum of the second gamma quantum. Under the given parameters ε 1 C and ( r , r ) , the relation (24) uniquely determines the outgoing angles of the electron and positron, and therefore their resonant energies. If we substitute expression (24) into relation (21), we find that, under conditions (9), the energy of the electron (for Channel A) or positron (for Channel B) is determined by the expression (also see relations (20), (23)):
E ω 1 + 1 + 4 δ 2 ± 2 r ω C .
Moreover, under conditions (10), the expression (25) simplifies:
E 1 + 4 δ 2 ± 2 r ω C .
Thus, the resonant energy of the electron–positron pair in conditions (9) is determined by Equations (20) and (25), and in conditions (10) by Equations (20) and (26). It is important to emphasize that in Conditions (9) and (10), the energy of the positron (for Channel A) or electron (for Channel B) is determined by the energy of the second high-energy gamma quantum, while the energy of the electron (for Channel A) or positron (for Channel B) in conditions (10) is primarily determined, not by the energy of the first gamma quantum, but by the characteristic energy of the EFSCE.

3. Maximum Breit–Wheeler Resonant Cross-Section

In conditions (9) and (10), in Channels A and B the EFSBWP proceeds with the number of absorbed photons of the wave r 1 , and for exchange resonant diagrams A’ and B’ the number of absorbed photons is r r 1 m i n = ε 1 B W 1 1 . Therefore, the resonant Channels A’ and B’ will be suppressed, and it is sufficient to consider only two resonant Channels, A and B. It should also be noted that for Channel A, the resonant energy of the electron–positron pair is determined by the outgoing angle of the positron relative to the momentum of the second gamma quantum, while for Channel B—by the outgoing angle of the electron relative to the momentum of the second gamma quantum. Therefore, Channels A and B are distinguishable and do not interfere with each other.
In the article [43], a general relativistic expression for the resonant BWP modified by a strong electromagnetic field was obtained. The resonance infinity, which occurs in the field of a plane monochromatic wave, was eliminated using the Breit–Wigner procedure [27,29,48]. As a result, the maximum resonant differential cross-section (at the point of maximum Breit–Wigner distribution) for Channel A (with a “plus” sign) and Channel B (with a “minus” sign) has the following form:
R ± ( r r ) = d σ r r m a x d δ 2 ± 2 = r e 2 c η i ψ ± ( r r ) .
Here, r e = e 2 / m is the classical electron radius, and the function c η i is determined by the initial setup parameters:
c η i = 2 ( 4 π ) 3 α 2 K 2 ( ε 1 C ) m δ η i ω B W 2 0.745 · 10 8 m δ η i ω B W K ( ε 1 C ) 2 ,
where α is the fine-structure constant, and δ η i 2 is the ultrarelativistic parameter, which determines the angle between the momenta of the initial gamma quanta (see Equation (2));
δ η i 2 = ω 1 ω 2 m * 2 θ i 2 .
The function K ( ε 1 C ) is defined by the total probability (per unit of time) of the EFSCE on the intermediate electron (for Channel A) or positron (for Channel B):
K ( ε 1 C ) = s = 1 0 s ε 1 C d u ( 1 + u ) 2 K ( u , s ε 1 C ) .
Here, the function K ( u , s ε 1 C ) is determined by the expressions:
K ( u , s ε 1 C ) = 4 J s 2 ( y 1 ( s ) ) + η 2 2 + u 2 1 + u ( J s 1 2 + J s + 1 2 2 J s 2 ) ,
y 1 ( s ) = 2 s η 1 + η 2 u s ε 1 C 1 u s ε 1 C .
In Equation (27), functions Ψ ± ( r r ) determine the spectral-angular distribution of the generated electron–positron pair:
ψ ± ( r r ) = x ± ( r ) ε 2 B W 2 ( 1 x ± ( r ) ) K 1 ( r ) P 2 ± ( r ) .
In expression (33), the function P 2 ± ( r ) determines the probability of the EFSBWP, and the function K 1 ( r ) determines the probability of the EFSCE [12]:
P 2 ± ( r ) = J r 2 ( γ 2 ± ( r ) ) + η 2 ( 2 u 2 ± ( r ) 1 ) r 2 γ 2 ± ( r ) 2 1 J r 2 + J r 2 ,
K 1 ( r ) = 4 J r 2 ( γ 1 ( r ) ) + η 2 2 + u 1 ( r ) 2 1 + u 1 ( r ) ( J r 1 2 + J r + 1 2 2 J r 2 ) .
The arguments of the Bessel functions have the following form:
γ 2 ± ( r ) = 2 r η 1 + η 2 u 2 ± ( r ) v 2 ± ( r ) 1 u 2 ± ( r ) v 2 ± ( r ) ,
γ 1 ( r ) = 2 r η 1 + η 2 u 1 ( r ) v 1 ( r ) 1 u 1 ( r ) v 1 ( r ) .
Here, the relativistic-invariant parameters are equal to
u 1 ( r ) = ( k 1 k ) ( p k ) ω 1 E ( r ) ω 1 ω 1 + ω 2 E ± ( r ) ,
v 1 ( r ) = 2 r ( q k ) m * 2 ε 1 C ( r ) E ( r ) ω 1 1 ε 1 C ( r ) ω 2 E ± ( r ) ω 1 ,
u 2 ± ( r ) = ( k 2 k ) 2 4 ( p ± k ) ( q k ) ω 2 2 4 E ± ( r ) ω 2 E ± ( r ) , v 2 ± ( r ) = r ( k 2 k ) 2 m * 2 ε 2 B W ( r ) .
It should be noted that in the right-hand sides of Equations (38) and (39), the general law of conservation of energy (23) has been taken into account. Therefore, the relativistic-invariant parameters for the EFSCE u 1 ( r ) and v 1 ( r ) can be expressed in terms of the energy of the positron (electron) for the EFSBWP.
However, when considering the correlation of processes at the first and second vertices of the resonant Feynman diagrams (see Figure 1), it is necessary to require the fulfillment of the condition u 1 ( r ) v 1 ( r ) , to ensure that the argument of the Bessel functions γ 1 ( r ) (37) is a real quantity. This condition can be expressed as a requirement for the physically admissible number of emitted photons at the first vertex:
r r * , r * = ω 1 2 ε 1 C ω 1 + ω 2 E ± ( r ) m a x 1 ω 2 E ± ( r ) m a x 1 .
Here, E ± ( r ) m a x is the maximum value of the resonant energy of the positron (electron), which is obtained from expression (17) at δ 2 ± 2 = 0 . Thus, the conditions (24) and (41) uniquely determine the outgoing angle of the electron (positron).
From Equations (27), (28), and (33), it can be seen that the magnitude of the resonant differential cross-section is mainly determined by function c η i (28). With a fixed value of the initial ultrarelativistic parameter δ η i 2 (29), the function c η i is significantly dependent on the characteristic energy of the EFSBWP ω B W (7), as well as on function K ( ε 1 C ) (30), which determines the resonance width. As the characteristic energy of EFSBWP decreases, the resonant cross-section increases quite rapidly (see Figure 3, Figure 4 and Figure 5).
Within the scope of the case of the very high energies of the second gamma quantum (9) studied in this article, the resonant energy of the positron (electron) takes the form (20). Taking this into account, expressions (36)–(40) will take the following form:
γ 2 ± 4 r η 1 + η 2 δ 2 ± 1 + 4 δ 2 ± 2 ,
u 2 ± r ε 2 B W 1 + 4 δ 2 ± 2 1 ,
γ 1 2 r η 1 + η 2 ( r / r ) β ± 1 β ± ,
β ± = ( 1 + 4 δ 2 ± 2 ) 1 + 1 + 4 δ 2 ± 2 r ε 1 C ,
u 1 1 + 1 + 4 δ 2 ± 2 r ε 1 C 1 .
Due to this, the functions ψ ± ( r r ) (33) in the expression for the resonant differential cross-section (27) is simplified and takes the following form:
ψ ± ( r r ) = 8 η 2 r 2 K 1 ( r ) ( 1 + 4 δ 2 ± 2 + r ε 1 C ) ( 1 + 4 δ 2 ± 2 ) r 2 γ 2 ± 2 1 J r 2 ( γ 2 ± ) + J r 2 .
Here, the argument of the Bessel functions γ 2 ± takes the form (42). It should be noted that the obtained resonant cross-section (27), (28), (47) depends on the frequency of the high-energy second gamma quantum (9) only through the ultrarelativistic parameters (29) and (18).
Let us consider the case when the energies of the initial gamma quanta satisfy the conditions (10). Then, the parameter u 1 (46) takes the following form:
u 1 r ε 1 C 1 + 4 δ 2 ± 2 1 .
As a result, expressions (35) and (44) are simplified:
K 1 ( r ) = 4 J r 2 ( γ 1 ) + 2 η 2 ( J r 1 2 + J r + 1 2 2 J r 2 ) .
γ 1 2 r η 1 + η 2 r ε 1 C 1 + 4 δ 2 ± 2 .
Thus, the resonant cross-section for the initial gamma quanta energies (10) is given by Equations (27), (28) and (47) in which the functions K 1 ( r ) , γ 1 and γ 2 ± are determined by relationships (49), (50), and (42), respectively.
Figure 3, Figure 4 and Figure 5 show the distributions of the resonant cross-section in the strong X-ray field with varying numbers of absorbed and emitted photons as a function of the square of the positron (for Channel A) or electron (for Channel B) outgoing angle relative to the momentum of the second gamma quantum under the conditions (8), (9), and (10) for various characteristic energies of the EFSBWP. The initial ultrarelativistic parameter δ η i 2 = 10 2 . Table 1, Table 2 and Table 3 represent the maximum values of the resonant cross-sections R ± ( r , r ) * in the corresponding peaks of the distributions as functions of δ 2 ± 2 . From the data in the figures and corresponding tables, it is evident how the resonant cross-sections change when the characteristic Breit–Wheeler energy varies from ω B W = 17.4 GeV to ω B W = 522 MeV. In this case, the magnitude of ω B W changes due to both the variation of the wave frequency and the parameter η (see Equation (11)). Thus, if the wave intensity is confined to the interval I = 1.675 · 10 21 ÷ 1.861 · 10 24 ( 3.536 · 10 27 ) Wcm 2 , the maximum resonant cross-sections (in units r e 2 ) for r = r = 1 change as follows: under the conditions of the initial gamma quanta energies (8) in the interval R ± ( 1 , 1 ) * 1.13 · 10 2 ÷ 4.36 · 10 4 ( 1.41 · 10 4 ) ; under the conditions of the initial gamma quanta energies (9) in the interval R ± ( 1 , 1 ) * 1.68 · 10 2 ÷ 6.65 · 10 4 ( 2.00 · 10 4 ) ; under the conditions of the initial gamma quanta energies (10) in the interval R ± ( 1 , 1 ) * 1.43 · 10 3 ÷ 1.27 · 10 6 ( 3.39 · 10 5 ) . Therefore, when the characteristic energy of the EFSBWP is reduced by a factor of 33, the maximum resonant cross-section increases by a factor of 3.86 · 10 2 ( 1.25 · 10 2 ) under the conditions of the initial gamma quanta energies (8), by a factor of 3.96 · 10 2 ( 1.19 · 10 2 ) under the conditions of the initial gamma quanta energies (9), and by a factor of 0.89 · 10 3 ( 2.37 · 10 2 ) under the conditions of the initial gamma quanta energies (10).
Consequently, the maximum value of the resonant cross-section occurs under the conditions of the initial gamma quanta energies (10) ( ω 1 ω C , ω 2 ω B W ) and, depending on the value of the characteristic Breit–Wheeler energy in the interval ω B W 10 ÷ 1 GeV, can reach a magnitude of R ± ( 1 , 1 ) * 10 3 ÷ 10 6 r e 2 . At the same time, for Channel A, narrow beams of high-energy positrons ( E + ω 2 ) are obtained, while for Channel B, narrow beams of high-energy electrons ( E ω 2 ) are obtained (see relation (20)).
However, if the characteristic Breit–Wheeler energy remains unchanged with changes in the frequency and intensity of the wave, then the resonant cross-section changes less significantly (compare the values of the resonant cross-section without parentheses and within parentheses in the above text).
It should be emphasized that in strong electromagnetic fields, various QED processes can take place, both stimulated and modified by an external field (see, for example, [5,12,13,14,15,16,17]). At the same time, the resonant QED processes modified by an external field, such as the Breit–Wheeler process, photogeneration of electron–positron pairs on the nucleus, spontaneous bremshtralung radiation during electron scattering on nuclei, the Compton effect, the generation of electron–positron pairs when an electron beam collides with an electromagnetic wave, and others, will have the greatest probability (see, for example, [13,14]). These resonant processes, depending on specific conditions, can occur in parallel with different or comparable probabilities.
We would like to note that the experimental implementation of the Breit–Wheeler resonant process in strong X-ray fields is currently very problematic. This is due to the need to have a powerful source of gamma radiation with energies, as well as strong X-ray fields with intensities. However, this process can take place near neutron stars, where appropriate resonant conditions can be created. It should also be noted that the Breit–Wheeler resonant process, as one of the resonant QED processes, may influence the implementation of laser fusion, the course of which is accompanied by various QED processes in the electromagnetic wave field.

4. Conclusions

We considered the resonant BWP modified by a strong X-ray field for high-energy initial gamma quanta. The following results were obtained:
  • It is shown that the resonant cross-section significantly depends on the magnitude of the characteristic Breit–Wheeler energy ω B W (7) and the characteristic energy of the Compton effect ω C (5). The ratios of the initial energies of gamma quanta with these characteristic energies significantly affect the magnitude of the resonant cross-section.
  • Under the conditions when the energy of the second gamma quantum significantly exceeds the characteristic Breit–Wheeler energy ( ω 2 ω B W ) , the resonant energy of the positron (for Channel A) or electron (for Channel B) tends toward the energy of the high-energy second gamma quantum (20) ( E ± ω 2 ) ;
  • The magnitude of the resonant cross-section significantly depends on the characteristic Breit–Wheeler energy, as well as the width of the resonance ( R ± ( r r ) ω B W 2 K 2 ( ε 1 C ) ) (see Equations (27) and (28)). Consequently, by decreasing the characteristic Breit–Wheeler energy by one order of magnitude, the resonant cross-section increases by two orders of magnitude. Moreover, the highest resonant differential cross-section is achieved when the energy of the second gamma quantum significantly exceeds the characteristic Breit–Wheeler energy ( ω 2 ω B W ) , while the energy of the first gamma quantum is significantly lower than the characteristic Compton effect energy ( ω 1 ω C ) . In this case, when reducing the characteristic Breit–Wheeler energy within the range ω B W 10 ÷ 1 GeV, the maximum resonant cross-section can reach values of R ± ( 1 , 1 ) * 10 3 ÷ 10 6 r e 2 .
The obtained results could be used to explain the narrow fluxes of high-energy ultrarelativistic positrons (electrons) in the vicinity of neutron stars, such as accretion-powered X-ray binaries [49,50], rotation-powered X/Gamma pulsars [51,52] and magnetic-field-powered magnetars [53,54], as well as to simulate physical processes in laser thermonuclear fusion [55].

Author Contributions

Conceptualization, S.P.R.; methodology, S.P.R. and V.V.D.; software, V.D.S.; validation, S.P.R., V.V.D., and V.D.S.; formal analysis, S.P.R. and V.D.S.; investigation, S.P.R., V.V.D., and V.D.S.; resources, V.V.D.; data curation, S.P.R. and V.V.D.; writing—original-draft preparation, V.D.S.; writing—review and editing, S.P.R. and V.D.S.; visualization, V.D.S.; supervision, S.P.R. and V.V.D.; project administration, V.V.D.; funding acquisition, V.V.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Higher Education of the Russian Federation under the strategic academic leadership program “Priority 2030” (Agreement 075-15-2023-380 dated 20 February 2023).

Data Availability Statement

The data presented in this study are openly available.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Breit, G.; Wheeler, J.A. Collision of two light quanta. Phys. Rev. 1934, 46, 1087–1091. [Google Scholar] [CrossRef]
  2. Bula, C.; McDonald, K.T.; Prebys, E.J.; Bamber, C.; Boege, S.; Kotseroglou, T.; Melissinos, A.C.; Meyerhofer, D.D.; Ragg, W.; Burke, D.L.; et al. Observation of Nonlinear Effects in Compton Scattering. Phys. Rev. Lett. 1996, 76, 3116–3119. [Google Scholar] [CrossRef] [PubMed]
  3. Mourou, G.A.; Tajima, T.; Bulanov, S.V. Optics in the relativistic regime. Rev. Mod. Phys. 2006, 78, 309–371. [Google Scholar] [CrossRef]
  4. Bagnoud, V.; Aurand, B.; Blazevic, A.; Borneis, S.; Bruske, C.; Ecker, B.; Eisenbarth, U.; Fils, J.; Frank, A.; Gaul, E.; et al. Commissioning and early experiments of the PHELIX facility. Appl. Phys. B 2010, 100, 137–150. [Google Scholar] [CrossRef]
  5. Di Piazza, A.; Müller, C.; Hatsagortsyan, K.Z.; Keitel, C.H. Extremely high-intensity laser interactions with fundamental quantum systems. Rev. Mod. Phys. 2012, 84, 1117–1228. [Google Scholar] [CrossRef]
  6. Burke, D.L.; Field, R.C.; Horton-Smith, G.; Spencer, J.E.; Walz, D.; Berridge, S.C.; Bugg, W.M.; Shmakov, K.; Weidemann, A.W.; Bula, C.; et al. Positron Production in Multiphoton Light-by-Light Scattering. Phys. Rev. Lett. 1997, 79, 1626–1629. [Google Scholar] [CrossRef]
  7. Bamber, C.; Boege, S.J.; Koffas, T.; Kotseroglou, T.; Melissinos, A.C.; Meyerhofer, D.D.; Reis, D.A.; Ragg, W.; Bula, C.; McDonald, K.T.; et al. Studies of nonlinear QED in collisions of 46.6 GeV electrons with intense laser pulses. Phys. Rev. D 1999, 60, 092004. [Google Scholar] [CrossRef]
  8. Kanya, R.; Morimoto, Y.; Yamanouchi, K. Observation of Laser-Assisted Electron-Atom Scattering in Femtosecond Intense Laser Fields. Phys. Rev. Lett. 2010, 105, 123202. [Google Scholar] [CrossRef]
  9. Hartin, A. Strong field QED in lepton colliders and electron/laser interactions. Int. J. Mod. Phys. A 2018, 33, 1830011. [Google Scholar] [CrossRef]
  10. Magnusson, J.; Gonoskov, A.; Marklund, M.; Esirkepov, T.Z.; Koga, J.K.; Kondo, K.; Kando, M.; Bulanov, S.V.; Korn, G.; Bulanov, S.S. Laser-Particle Collider for Multi-GeV Photon Production. Phys. Rev. Lett. 2019, 122, 254801. [Google Scholar] [CrossRef]
  11. Berestetskii, V.B.; Lifshitz, E.M.; Pitaevskii, L.P. Quantum Electrodynamics; Butterworth-Heinemann: Oxford, UK, 1982; pp. 1–652. [Google Scholar]
  12. Ritus, V.I.; Nikishov, A.I. Quantum electrodynamics phenomena in the intense field. In Trudy FIAN; Nauka: Moscow, Russia, 1979; Volume 111, pp. 1–276. [Google Scholar]
  13. Roshchupkin, S.P. Resonant effects in collisions of relativistic electrons in the field of a light wave. Laser Phys. 1996, 6, 837–858. [Google Scholar]
  14. Roshchupkin, S.P.; Tsybul’nik, V.A.; Chmirev, A.N. The Probability of Multiphoton Processes in Quantum-Electrodynamic Phenomena in a Strong Light Field. Laser Phys. 2000, 10, 1256–1272. [Google Scholar]
  15. Mironov, A.A.; Meuren, S.; Fedotov, A.M. Resummation of QED radiative corrections in a strong constant crossed field. Phys. Rev. D 2020, 102, 053005. [Google Scholar] [CrossRef]
  16. Gonoskov, A.; Blackburn, T.G.; Marklund, M.; Bulanov, S.S. Charged particle motion and radiation in strong electromagnetic fields. Rev. Mod. Phys. 2022, 94, 045001. [Google Scholar] [CrossRef]
  17. Fedotov, A.; Ilderton, A.; Karbstein, F.; King, B.; Seipt, D.; Taya, H.; Torgrimsson, G. Advances in QED with intense background fields. Phys. Rep. 2023, 1010, 1–138. [Google Scholar]
  18. Oleinik, V.P. Resonance effects in the field of an intense laser beam. Sov. Phys. JETP 1967, 25, 697–708. [Google Scholar]
  19. Oleinik, V.P. Resonance effects in the field of an intense laser ray ii. Sov. Phys. JETP 1968, 26, 1132–1138. [Google Scholar]
  20. Florescu, A.; Florescu, V. Laser-modified electron bremsstrahlung in a Coulomb field. Phys. Rev. A 2000, 61, 033406. [Google Scholar] [CrossRef]
  21. Flegel, A.V.; Frolov, M.V.; Manakov, N.L.; Starace, A.F.; Zheltukhin, A.N. Analytic description of elastic electron-atom scattering in an elliptically polarized laser field. Phys. Rev. A 2013, 87, 013404. [Google Scholar] [CrossRef]
  22. Zheltukhin, A.N.; Flegel, A.V.; Frolov, M.V.; Manakov, N.L.; Starace, A.F. Resonant electron-atom bremsstrahlung in an intense laser field. Phys. Rev. A 2014, 89, 023407. [Google Scholar] [CrossRef]
  23. Zheltukhin, A.N.; Flegel, A.V.; Frolov, M.V.; Manakov, N.L.; Starace, A.F. Rescattering effects in laser-assisted electron–atom bremsstrahlung. J. Phys. B 2015, 48, 075202. [Google Scholar] [CrossRef]
  24. Li, A.; Wang, J.; Ren, N.; Wang, W.; Zhu, W.; Li, X.; Hoehn, R.; Kais, S. The interference effect of laser-assisted bremsstrahlung emission in Coulomb fields of two nuclei. J. Appl. Phys. 2013, 114, 124904. [Google Scholar] [CrossRef]
  25. Heinzl, T.; Ilderton, A. Exact Classical and Quantum Dynamics in Background Electromagnetic Fields. Phys. Rev. Lett. 2017, 118, 113202. [Google Scholar] [CrossRef] [PubMed]
  26. Krachkov, P.A.; Di Piazza, A.; Milstein, A.I. High-energy bremsstrahlung on atoms in a laser field. Phys. Lett. B 2019, 797, 134814. [Google Scholar] [CrossRef]
  27. Roshchupkin, S.P.; Starodub, S.S. The effect of generation of narrow ultrarelativistic beams of positrons (electrons) in the process of resonant photoproduction of pairs on nuclei in a strong electromagnetic field. Laser Phys. Lett. 2022, 19, 115301. [Google Scholar] [CrossRef]
  28. Roshchupkin, S.P.; Larin, N.R.; Dubov, V.V. Resonant effect of the ultrarelativistic electron–positron pair production by gamma quanta in the field of a nucleus and a pulsed light wave. Laser Phys. 2021, 31, 045301. [Google Scholar] [CrossRef]
  29. Roshchupkin, S.P.; Larin, N.R.; Dubov, V.V. Resonant photoproduction of ultrarelativistic electron-positron pairs on a nucleus in moderate and strong monochromatic light fields. Phys. Rev. D 2021, 104, 116011. [Google Scholar] [CrossRef]
  30. Zhao, Q.; Wu, Y.X.; Ababekri, M.; Li, Z.P.; Tang, L.; Li, J.X. Angle-dependent pair production in the polarized two-photon Breit-Wheeler process. Phys. Rev. D 2023, 107, 096013. [Google Scholar] [CrossRef]
  31. He, Y.; Yeh, I.L.; Blackburn, T.G.; Arefiev, A. A single-laser scheme for observation of linear Breit–Wheeler electron–positron pair creation. New J. Phys. 2021, 23, 115005. [Google Scholar] [CrossRef]
  32. Ivanov, D.Y.; Kotkin, G.L.; Serbo, V.G. Complete description of polarization effects in e+ e-pair productionby a photon in the field of a strong laser wave. Eur. Phys. J. C 2005, 40, 27–40. [Google Scholar] [CrossRef]
  33. Krajewska, K.; Kaminski, J.Z. Breit-Wheeler process in intense short laser pulses. Phys. Rev. A 2012, 86, 052104. [Google Scholar] [CrossRef]
  34. Bragin, S.; Di Piazza, A. Electron-positron annihilation into two photons in an intense plane-wave field. Phys. Rev. D 2021, 102, 116012. [Google Scholar] [CrossRef]
  35. Titov, A.I.; Takabe, H.; Kämpfer, B. Nonlinear Breit-Wheeler process in short laser double pulses. Phys. Rev. D 2018, 98, 036022. [Google Scholar] [CrossRef]
  36. Titov, A.I.; Kämpfer, B. Non-linear Breit–Wheeler process with linearly polarized beams. Eur. Phys. J. D 2020, 74, 218. [Google Scholar] [CrossRef]
  37. Tang, S. Fully polarized nonlinear Breit-Wheeler pair production in pulsed plane waves. Phys. Rev. D 2022, 105, 056018. [Google Scholar] [CrossRef]
  38. Blackburn, T.G.; King, B. Higher fidelity simulations of nonlinear Breit–Wheeler pair creation in intense laser pulses. Eur. Phys. J. C 2022, 82, 44. [Google Scholar] [CrossRef]
  39. Seipt, D.; King, B. Spin-and polarization-dependent locally-constant-field-approximation rates for nonlinear Compton and Breit-Wheeler processes. Phys. Rev. A 2020, 102, 052805. [Google Scholar] [CrossRef]
  40. Pustyntsev, A.A.; Dubov, V.V.; Roshchupkin, S.P. Resonant Breit-Wheeler process in an external electromagnetic field. Mod. Phys. Lett. A 2020, 35, 2040027. [Google Scholar] [CrossRef]
  41. Serov, V.D.; Roshchupkin, S.P.; Dubov, V.V. Resonant Effect for Breit–Wheeler Process in the Field of an X-ray Pulsar. Universe 2020, 6, 190. [Google Scholar] [CrossRef]
  42. Serov, V.D.; Roshchupkin, S.P.; Dubov, V.V. Resonant Breit–Wheeler process in a strong electromagnetic field. TMF 2023, 216, 577–589. [Google Scholar] [CrossRef]
  43. Roshchupkin, S.P.; Serov, V.D.; Dubov, V.V. Generation of narrow beams of ultrarelativistic positrons (electrons) in the Breit-Wheeler resonant process, modified by the field of a strong electromagnetic wave. Photonics 2023, 10, 949. [Google Scholar] [CrossRef]
  44. Volkov, D. On a class of solutions of the Dirac equation. Z. Phys. 1935, 94, 250–260. [Google Scholar]
  45. Wang, H.; Zhong, M.; Gan, L.F. Orthonormality of Volkov Solutions and the Sufficient Condition. Commun. Theor. Phys. 2019, 71, 1179–1186. [Google Scholar] [CrossRef]
  46. Schwinger, J. On Gauge Invariance and Vacuum Polarization. Phys. Rev. 1951, 82, 664–679. [Google Scholar] [CrossRef]
  47. Brown, L.S.; Kibble, T.W.B. Interaction of Intense Laser Beams with Electrons. Phys. Rev. 1964, 133, A705–A719. [Google Scholar] [CrossRef]
  48. Breit, G.; Wigner, E. Capture of Slow Neutrons. Phys. Rev. 1936, 49, 519–531. [Google Scholar] [CrossRef]
  49. Deng, Z.-L.; Gao, Z.-F.; Li, X.-D.; Shao, Y. On the Formation of PSR J1640+2224: A Neutron Star Born Massive? Astrophys. J. 2020, 892, 4. [Google Scholar] [CrossRef]
  50. Deng, Z.-L.; Gao, Z.-F.; Li, X.-D.; Shao, Y. Evolution of LMXBs under Different Magnetic Braking Prescriptions. Astrophys. J. 2021, 909, 174. [Google Scholar] [CrossRef]
  51. Gao, Z.-F.; Wang, N.; Shan, H.; Li, X.-D.; Wang, W. The Dipole Magnetic Field and Spin-down Evolutions of the High Braking Index Pulsar PSR J1640–4631. Astrophys. J. 2017, 849, 19. [Google Scholar] [CrossRef]
  52. Wang, H.; Gao, Z.-F.; Jia, H.-Y.; Wang, N.; Li, X.-D. Estimation of Electrical Conductivity and Magnetization Parameter of Neutron Star Crusts and Applied to the High-Braking-Index Pulsar PSR J1640-4631. Universe 2020, 6, 63. [Google Scholar] [CrossRef]
  53. Gao, Z.-F.; Li, X.-D.; Wang, N.; Yuan, J.P.; Wang, P.; Peng, Q.H.; Du, Y.J. Constraining the braking indices of magnetars. Mon. Notices Royal Astron. Soc. 2016, 456, 55–65. [Google Scholar] [CrossRef]
  54. Yan, F.-Z.; Gao, Z.-F.; Yang, W.-S.; Dong, A.-J. Explaining high braking indices of magnetars SGR 0501+4516 and 1E 2259+586 using the double magnetic-dipole model. Astron. Nachrichten 2021, 342, 249–254. [Google Scholar] [CrossRef]
  55. Gus’kov, S.Y. Laser fusion and high energy density physics. Kvantovaya Elektron. 2022, 52, 1070–1078. [Google Scholar] [CrossRef]
Figure 1. Feynman diagrams of the resonant external field–assisted Breit–Wheeler process, Channels A and B; for Channels A’ and B’ substitution k 1 k 2 is needed.
Figure 1. Feynman diagrams of the resonant external field–assisted Breit–Wheeler process, Channels A and B; for Channels A’ and B’ substitution k 1 k 2 is needed.
Symmetry 15 01901 g001
Figure 2. The dependence of the resonant energy of the positron (electron) (in units of ω 2 ) on the ultrarelativistic parameter δ 2 ± 2 for a single absorbed photon of the wave with significantly different energies of the second gamma quantum.
Figure 2. The dependence of the resonant energy of the positron (electron) (in units of ω 2 ) on the ultrarelativistic parameter δ 2 ± 2 for a single absorbed photon of the wave with significantly different energies of the second gamma quantum.
Symmetry 15 01901 g002
Figure 3. The dependence of the maximum resonant differential cross-section (27), (28), (33) (in units of r e 2 ) on the square of the positron outgoing angle (Channel A) or the square of the electron outgoing angle (Channel B) for various intensities of the X-ray wave and the characteristic energy of the EFSBWP, as well as the numbers of absorbed (r) and emitted ( r ) photons in conditions (8) for the energies of the initial gamma quanta. The corresponding energies: (a)— ω B W = 17.4 G e V ,   ω 2 = 35 G e V ,   ω 1 = 1 G e V ; (b,c)— ω B W = 522 M e V ,   ω 2 = 1 G e V ,   ω 1 = 50 M e V .
Figure 3. The dependence of the maximum resonant differential cross-section (27), (28), (33) (in units of r e 2 ) on the square of the positron outgoing angle (Channel A) or the square of the electron outgoing angle (Channel B) for various intensities of the X-ray wave and the characteristic energy of the EFSBWP, as well as the numbers of absorbed (r) and emitted ( r ) photons in conditions (8) for the energies of the initial gamma quanta. The corresponding energies: (a)— ω B W = 17.4 G e V ,   ω 2 = 35 G e V ,   ω 1 = 1 G e V ; (b,c)— ω B W = 522 M e V ,   ω 2 = 1 G e V ,   ω 1 = 50 M e V .
Symmetry 15 01901 g003
Figure 4. The dependence of the maximum resonant differential cross-section (27), (28), (47) (in units of r e 2 ) on the square of the positron outgoing angle (Channel A) or the square of the electron outgoing angle (Channel B) for various intensities of the X-ray wave and the characteristic energy of EFSBWP, as well as the numbers of absorbed (r) and emitted ( r ) photons in conditions (9) for the energies of the initial gamma quanta. The energies of the initial gamma quanta: (a)— ω 2 ω B W = 17.4 G e V , ω 1 = 1 G e V ; (b,c)— ω 2 ω B W = 522 M e V , ω 1 = 50 M e V .
Figure 4. The dependence of the maximum resonant differential cross-section (27), (28), (47) (in units of r e 2 ) on the square of the positron outgoing angle (Channel A) or the square of the electron outgoing angle (Channel B) for various intensities of the X-ray wave and the characteristic energy of EFSBWP, as well as the numbers of absorbed (r) and emitted ( r ) photons in conditions (9) for the energies of the initial gamma quanta. The energies of the initial gamma quanta: (a)— ω 2 ω B W = 17.4 G e V , ω 1 = 1 G e V ; (b,c)— ω 2 ω B W = 522 M e V , ω 1 = 50 M e V .
Symmetry 15 01901 g004
Figure 5. The dependence of the maximum resonant differential cross-section (27), (28), (47) (in units of r e 2 ) on the square of the positron outgoing angle (Channel A) or the square of the electron outgoing angle (Channel B) for various intensities of the X-ray wave and the characteristic energy of the EFSBWP, as well as the numbers of absorbed (r) and emitted ( r ) photons in conditions (10) for the energies of the initial gamma quanta. The energies of the initial gamma quanta: (a)— ω 2 ω B W = 17.4 G e V , ω 1 = 0.35 G e V ; (b,c)— ω 2 ω B W = 522 M e V , ω 1 = 12 M e V .
Figure 5. The dependence of the maximum resonant differential cross-section (27), (28), (47) (in units of r e 2 ) on the square of the positron outgoing angle (Channel A) or the square of the electron outgoing angle (Channel B) for various intensities of the X-ray wave and the characteristic energy of the EFSBWP, as well as the numbers of absorbed (r) and emitted ( r ) photons in conditions (10) for the energies of the initial gamma quanta. The energies of the initial gamma quanta: (a)— ω 2 ω B W = 17.4 G e V , ω 1 = 0.35 G e V ; (b,c)— ω 2 ω B W = 522 M e V , ω 1 = 12 M e V .
Symmetry 15 01901 g005
Table 1. The values of the ultrarelativistic parameter δ 2 ± 2 ( * ) that correspond to maximum values R ± ( r , r ) ( * ) of the spectral-angular distribution for the resonant differential cross-sections (27) under the conditions (8) (see Figure 3 and expression (11)).
Table 1. The values of the ultrarelativistic parameter δ 2 ± 2 ( * ) that correspond to maximum values R ± ( r , r ) ( * ) of the spectral-angular distribution for the resonant differential cross-sections (27) under the conditions (8) (see Figure 3 and expression (11)).
(r,r’) δ 2 ± 2 ( * ) R ± ( r , r ) ( * )
I = 1.675 · 10 21 W c m 2 ,
ω 1 = 1 G e V ,
ω 2 = 35 G e V
(1,1)0 1.13 · 10 2
(2,1)0.095 3.90 · 10 1
(1,2)0 1.69 · 10 1
(2,2)0.060 1.53 · 10 1
I = 1.861 · 10 24 W c m 2 ,
ω 1 = 50 M e V ,
ω 2 = 1 G e V
(1,1)0 4.36 · 10 4
(2,1)0.047 1.66 · 10 4
(1,2)0 9.85 · 10 3
(2,2)0.069 6.83 · 10 3
I = 3.536 · 10 27 W c m 2 ,
ω 1 = 50 M e V ,
ω 2 = 1 G e V
(1,1)0 1.41 · 10 4
(2,1)0.025 6.41 · 10 3
(1,2)0 6.13 · 10 3
(2,2)0.048 3.49 · 10 3
Table 2. The values of the ultrarelativistic parameter δ 2 ± 2 ( * ) that correspond to maximum values R ± ( r , r ) ( * ) of the spectral-angular distribution for the resonant differential cross-sections (27) under the conditions (9) (see Figure 4 and expression (11)).
Table 2. The values of the ultrarelativistic parameter δ 2 ± 2 ( * ) that correspond to maximum values R ± ( r , r ) ( * ) of the spectral-angular distribution for the resonant differential cross-sections (27) under the conditions (9) (see Figure 4 and expression (11)).
(r,r’) δ 2 ± 2 ( * ) R ± ( r , r ) ( * )
I = 1.675 · 10 21 W c m 2 ,
ω 1 = 1 G e V
(1,1)0 1.68 · 10 2
(2,1)0.070 4.66 · 10 1
(1,2)0 3.40 · 10 1
(2,2)0.055 1.99 · 10 1
I = 1.861 · 10 24 W c m 2 ,
ω 1 = 50 M e V
(1,1)0 6.65 · 10 4
(2,1)0.033 2.59 · 10 4
(1,2)0 2.02 · 10 4
(2,2)0.063 8.84 · 10 3
I = 3.536 · 10 27 W c m 2 ,
ω 1 = 50 M e V
(1,1)0 2.00 · 10 4
(2,1)0.020 1.15 · 10 4
(1,2)0 1.18 · 10 4
(2,2)0.043 4.19 · 10 3
Table 3. The values of the ultrarelativistic parameter δ 2 ± 2 ( * ) that correspond to the maximum values R ± ( r , r ) ( * ) of the spectral-angular distribution for the resonant differential cross-sections (27) under the conditions (10) (see Figure 5 and expression (11)).
Table 3. The values of the ultrarelativistic parameter δ 2 ± 2 ( * ) that correspond to the maximum values R ± ( r , r ) ( * ) of the spectral-angular distribution for the resonant differential cross-sections (27) under the conditions (10) (see Figure 5 and expression (11)).
(r,r’) δ 2 ± 2 ( * ) R ± ( r , r ) ( * )
I = 1.675 · 10 21 W c m 2 ,
ω 1 = 0.35 G e V
(1,1)0 1.43 · 10 3
(2,1)0.070 5.01 · 10 2
(1,2)0 1.10 · 10 2
(2,2)0.043 1.05 · 10 2
I = 1.861 · 10 24 W c m 2 ,
ω 1 = 12 M e V
(1,1)0 1.27 · 10 6
(2,1)0.073 2.33 · 10 5
(1,2)0 1.12 · 10 5
(2,2)0.044 9.88 · 10 4
I = 3.536 · 10 27 W c m 2 ,
ω 1 = 12 M e V
(1,1)0 3.39 · 10 5
(2,1)0.073 1.08 · 10 5
(1,2)0 6.11 · 10 4
(2,2)0.044 5.92 · 10 4
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Roshchupkin, S.P.; Serov, V.D.; Dubov, V.V. The Generation of High-Energy Electron–Positron Pairs during the Breit–Wheeler Resonant Process in a Strong Field of an X-ray Electromagnetic Wave. Symmetry 2023, 15, 1901. https://doi.org/10.3390/sym15101901

AMA Style

Roshchupkin SP, Serov VD, Dubov VV. The Generation of High-Energy Electron–Positron Pairs during the Breit–Wheeler Resonant Process in a Strong Field of an X-ray Electromagnetic Wave. Symmetry. 2023; 15(10):1901. https://doi.org/10.3390/sym15101901

Chicago/Turabian Style

Roshchupkin, Sergei P., Vitalii D. Serov, and Victor V. Dubov. 2023. "The Generation of High-Energy Electron–Positron Pairs during the Breit–Wheeler Resonant Process in a Strong Field of an X-ray Electromagnetic Wave" Symmetry 15, no. 10: 1901. https://doi.org/10.3390/sym15101901

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop