Next Article in Journal
Dynamics of Neutron Transfer in the Reaction 3He + 9Be
Previous Article in Journal
The Two-Thirds Power Law Derived from a Higher-Derivative Action
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Accelerating Charge: Add-Ons to Rest Mass and Field Energy

by
Elizabeth P. Tito
1 and
Vadim I. Pavlov
1,2,*
1
Scientific Advisory Group, Pasadena, CA 91125, USA
2
Laboratoire de Mécanique des Fluides de Lille—Kampé de Fériet (LMFL—KF)—UMR 9014 ONERA, CNRS (Office National d’Études et de Recherches Aérospatiales, Centre National de la Recherche Scientifique), University of Lille, F-59000 Lille, France
*
Author to whom correspondence should be addressed.
Physics 2024, 6(4), 1264-1280; https://doi.org/10.3390/physics6040078
Submission received: 15 July 2024 / Revised: 4 November 2024 / Accepted: 6 November 2024 / Published: 19 November 2024
(This article belongs to the Section Classical Physics)

Abstract

:
We present—in the framework of classical theory—a self-consistent derivation scheme which produces equations for the calculation of add-ons to the full field energy and to the effective mass of a charge moving with acceleration, which may be practically used for analyses in various scenarios. The charge is treated as a quasi-point-like charge; this helps to resolve the complications of the “infinite” electromagnetic energy, which are avoided by the procedure of slightly “spreading” the charge. As a result, the concept of the size of the particle takes a straightforward physical interpretation. Indeed, it is within the charge spread, at scales smaller than Compton’s length, where the quantum-field-mechanics approach to be applied. Beyond this region, no “infinite” tails of quantities accumulate. The seeming divergences of the integrals at the upper limits are not physical if one takes into account that the charge moves with acceleration only for a finite duration of time; every real physical process has its beginning and its end. The key focus of this paper is on the methodological aspects of the calculations.

1. Introduction

A charged particle radiates when it moves (in a vacuum) with acceleration. When this acceleration is caused by a collision of charged particles, the radiation is called bremsstrahlung. When this acceleration is caused by an external magnetic field, the radiation is called the magnetic bremsstrahlung or, more commonly, the synchrotron radiation. Nowadays, the term “electromagnetic bremsstrahlung” is often used in the much wider context, regardless of the specific mechanism that causes the acceleration of the charged particle.
Quite a number of studies has been devoted to this effect; for some notable papers and reviews, see, e.g., [1,2,3,4,5,6,7,8,9,10,11,12,13]. Bremsstrahlung is used as a source of X-radiation production [14,15,16,17]. It appears in high-power laser–plasma interactions [18,19], is used for diagnostic purposes [20,21,22,23,24,25,26,27,28,29], and determines the dynamics of fast (relativistic) runaway electrons [30,31]. Bremsstrahlung is frequently observed and has been extensively studied in astrophysics (see, e.g., [4,32,33,34,35,36,37]). Bremsstrahlung plays a role in the physics of cosmic microwave background distortions [38,39], can be an essential electromagnetic radiation process for plasma around compact objects [40,41], and is responsible for high-energy gamma-radiation production in galaxy clusters [42,43,44] and stellar flares [29,45,46,47]. Within the realm of accelerating charges, electrons have received high attention. Accelerated electrons produce radio emission when they move in a magnetic field, especially near the magnetic poles of pulsars. These electrons are a source of synchrotron radiation when they are affected by external magnetic field (see, e.g., [48]). Accelerated electrons in high-energy environments, such as near the magnetic poles of pulsars or within the magnetic fields of magnetars, produce Roentgen radiation and are accelerated to extremely high energies (see, e.g., [49,50,51]). Bremsstrahlung is one of the primary mechanisms of radiative energy loss from plasma. The effect is mainly characterized by its differential cross section (DCS), whose relativistic expansion was derived in the seminal publications by Hans Bethe and Walter Heitler [52,53]. The inverse process, the so-called bremsstrahlung absorption, is one of the principal mechanisms of laser energy transfer in experiments of inertial confinement fusion (ICF) [54,55,56].
Given the unconditional importance of the issues related to this topic, in this paper, we address a significant methodological issue. In the process, the charged particle interacts with the field which it generates, and this interaction leads to the deceleration of the particle. Naturally, there should be no self-acceleration. Generally, one can immediately write the equations for the energy balance as follows:
d E p d t = P , d W f d t = P d Σ · S ,
where E p is the full/kinetic energy of the charged particle (in the relativistic case, E p = m c 2 γ , where m is the effective rest mass of the charged particle, c denotes the speed of light, and γ is the Lorentz factor); W f = ( 8 π ) 1 ( E 2 + H 2 ) ¯ is the energy of the field surrounding the charge, with E and H the strengths of electric and magnetic fields, respectively, and the bar shows that integration is taken over the volume occupied by the field; the positively defined quantity P is the power of the force of radiation friction (in the first equation of the system (1), it leads to the particle’s slowdown, whereas in the second equation, it leads to the increase in the magnitude of the electromagnetic field); and the surface integral on the right hand side (r.h.s.) of the second equation is an “outflow” (described by Poynting vector S = ( c / 4 π ) [ E , H ] , where the square brackets denote the vector cross product) of the field energy through a surrounding surface Σ located at a large distance from the charge.
The sense of expressions (1) might seem trivial; however, as a result of the interaction of the charge with its accompanying field and with the radiated field, the charge’s rest “effective mass” changes ( m m + δ m ) and the field’s energy density also changes ( W f W f + δ W f ). The task of obtaining the explicit expressions for δ m and δ W f is not trivial.
In this paper, we analytically give—in the framework of classical electrodynamics and relativity—the methodology and the derivations for δ m and δ W f in the spirit of the founding paper by Paul Dirac [57] with details of the derivations; see [58,59,60,61,62,63,64,65,66,67] for a more in-depth examination.
This paper is organized as follows. Section 2 formulates the model. Section 3 shows how the energy–momentum tensors for the particle/charge and the field are constructed. Section 4 discusses the effect of the radiation friction force. Section 5 analyzes the structure of the field in the vicinity of the charge. Section 6 concludes with final remarks. Appendix A gives the details of the calculations.

2. Model

We consider a flat spacetime characterized by the Galilean metrics with the positive time-space signature ( + ) (in contrast to the often-used “negative” signature ( + + + ) ; see, e.g., [61]). Each world-point is described by a vector pointing to a position/state with contra-variant components q i = ( c t , q α ) , t is the time measured by a remote observer. The Latin letter indices take the values 0 (time), 1 , 2 , and 3 (space), while the Greek letter indices take the values 1 , 2 , and 3. With this definition of coordinates, s (see just below) and q i have the length dimension, L. The spacetime metric is fixed by the interval d s : d s 2 = g i k d q i d q k , where g i k is the metric tensor. For a flat spacetime, g i k = ( 1 , δ μ ν ) , where δ μ ν is the Kronecker delta. In the latter case, d s = c d t 1 β 2 c γ 1 d t . The contra-variant components of 4-velocity are u 0 = γ , u α = γ β α , where β α = v α / c , and the covariant ones are u 0 = u 0 , u α = u α . Similarly to the definition of 4-speed u i = d q i / d s = ( γ , γ β ) (certainly, u i u i = g i k u i u k = u 0 u 0 u α u α = γ 2 γ 2 β 2 = 1 ), the quantity w i = d u i / d s = d 2 q i / d s 2 can be called the contra-variant components of 4-acceleration (we follow here definitions of Ref. [59] (pp. 41 and 318). The dimensions of the quantities are [ u i ] = 1 , [ w i ] = L 1 . The components of acceleration are
w i = ( w 0 , w ) = γ 4 c 2 β · a , γ 2 c 2 a + γ 4 c 2 β ( β · a ) .
Here, a = c ( d β / d t ) is the common three-dimensional acceleration with the dimension of L 1 T 2 , where T denotes the time dimension.

3. Energy–Momentum Tensors for Particle and Field

For the electromagnetic field without particles, the Lagrangian density is the scalar:
Λ = 1 16 π F i k F i k .
Let us remind that, symbolically, F i j = ( E , H ) , F i j = ( E , H ) , i.e., F 0 α = E α , F 0 α = E α , etc. Expressed in terms of E and H , the convolution F i k F i k = 2 ( E 2 H 2 ) . The sign of Equation (3) is explained in Ref. [59] (p. 69). The electromagnetic tensor F k l is expressed via potentials A l as
F k l = k A l l A k .
Hereafter, the operator i / q i where q i denotes the contravariant 4-coordinate.
Let us stress that the energy–momentum tensor is determined according to the standard scheme: from action to Lagrangian to extremum of action to evolution equation to derivative of the Lagrangian (with respect to the spacetime coordinates) to consideration of the evolution equation to transfer of the terms to the law of conservation for the introduced tensor in the 4-divergence form.
The generalized coordinates for the field are the components of the 4-potential of the field, A k . Thus, the mixed energy–momentum tensor T i k [ f ] is introduced as
T i k [ f ] = ( i A l ) Λ ( k A l ) Λ δ i k .
The symbolic argument f here indicates that the marked quantity refers to the electromagnetic field. After calculations—and keeping in mind that the variation of Λ (which gives T i k [ f ] ) is defined, with accuracy, up to the term that is the divergence of something—we find the following expressions for the energy–momentum tensor of the electromagnetic field:
T i k [ f ] = 1 4 π ( F i l F l k + 1 4 g i k F l m F l m )
and for the mixed tensor:
T i k [ f ] = 1 4 π ( F i l F l k + 1 4 δ i k F l m F l m ) .
The components of the tensor in terms of the electric and magnetic field intensities are the quantity T 00 [ f ] , which coincides with the energy density ( 1 / 8 π ) ( E 2 + H 2 ) , and the components T 0 α [ f ] , which are the spatial components of the Poynting vector divided by c, i.e., S α / c . Next, we differentiate Equation (6) and take into account the Maxwell equations:
k F k l = 4 π c j l , i F l m = l F m i m F i l ,
where j k is the current, and then permute the indices, cancel certain terms out, and obtain the result—the divergence of the mixed tensor:
k T i k [ f ] = 1 c F i k j k .
The sense of the term on the r.h.s. of Equation (9) is straightforward: it is the “source” of the electromagnetic field—the field density changes if there is a “collision” of a moving charge ( j k 0 ) with an electromagnetic field ( F i k 0 ). Equation (9) shows that indeed no separate conservation of energy exists for only the electromagnetic field. The particles that create this field must also be included in the consideration.
The total energy–momentum for the closed system of particle plus field to be conserved: k ( T i k [ p ] + T i k [ f ] ) = 0 . Then, for the particle, one must certainly have
k T i k [ p ] = + 1 c F i k j k ,
with the plus sign in front of the source-term on the r.h.s. of Equation (10).
Now, let us turn to the question of the structure of the energy–momentum tensor for the particle, T i k [ p ] . First, one must move from the model of a point-like particle whose mass distribution is described by the Dirac delta-function as μ = m δ ( r r p ) , where r p is the radius vector for the particle, to a model with an arbitrary density distribution μ , which is normalized by the condition d V μ = m .
However, expressing the energy–momentum tensor for the mass distribution in the form T i k [ p ] = μ c 2 u i u k is a mistake. The correct expression is given by the formula
T i k [ p ] = μ c d x i d s d x k d t = μ c 2 u i u k ( 1 c d s d t ) .
The rationale for this is given, e.g., in Ref. [59], p. 88, which states: “The ‘four-momentum density’ of the particles is given by μ c u i . We know that this density is the component T 0 α / c of the energy–momentum tensor, i.e., T 0 α = μ c 2 u α ( α = 1 , 2 , 3 ) . But the mass density is the time component of the four-vector ( μ / c ) ( d q i / d t ) (in analogy to the charge density; see §28). Therefore the energy–momentum tensor of the system of non-interacting particles is T i k = μ c u i u k ( d s / d t ) (33.5). As expected, this tensor is symmetric”.
This result can also be obtained by direct calculation. We take into consideration the condition u k u k = 1 , which holds along the particle’s real world-line.
For a free non-point-like particle, the integral of action is written as
S [ p ] = m c d s S [ p ] = d t d V μ c u k u k d s d t ,
where V is the volume bounded by the surface Σ .
As soon as the integral S [ p ] is the relativistic-invariant quantity, and d t d V and u k u k are also invariants in Equation (12), then the invariant is μ * = μ ( d s / c d t ) and not μ . Thus, for the continuous distribution of mass, the Lagrangian density (per 4-volume unit) is
Λ [ p ] = μ * c 2 u k u k .
Including the constraint of being on the real world-line, the quantity Λ has to be written as
Λ [ p ] = μ * c 2 u k u k + λ 1 ( u l u l 1 ) + λ 2 .
The second term in Equation (14) is zero on the real world-line but its derivative is not:
Λ [ p ] u k = μ * c 2 u k + λ 1 u k .
The energy–momentum tensor is
u i Λ [ p ] u k Λ δ i k = u i μ * c 2 u k + λ 1 u i u k + δ i k μ * c 2 u l u l δ i k λ 1 ( u l u l 1 ) δ i k λ 2 .
For indeterminate coefficients λ 1 and λ 2 , we require that
u i μ * c 2 u k + λ 1 u i u k + δ i k μ * c 2 u l u l δ i k λ 1 ( u l u l 1 ) δ i k λ 2 = u i μ * c 2 u k ,
i.e.,
u i μ * c 2 u k u i μ * c 2 u k + λ 1 u i u k + δ i k μ * c 2 u l u l δ i k λ 1 ( u l u l 1 ) δ i k λ 2 = 0 .
From Equation (18), one finds that the constant coefficients λ 1 , 2 may be fixed as
λ 1 = 2 μ * c 2 , λ 2 = μ * c 2 .
Thus, as soon as u l u l = 1 , the mixed tensor for the particle, as formulated in Ref. [59], is
T i k [ p ] = μ * c 2 u i u k = μ c u i d x k d t .
Now, let us calculate the derivative of the tensor T i k [ p ] :
k T i k [ p ] = c u i k μ d x k d t + μ c d x k d t k u i .
The first term in this expression is zero because of the conservation of mass k ( μ d x k / d t ) = 0, and therefore,
k T i k [ p ] = μ c d x k d t u i x k = μ c d u i d t = d s d t μ c d u i d s .

4. Effect of Radiation Friction Force

For the next step, we use the 4-dimensional equation of motion of the charge in the electromagnetic field, written in the form that includes the radiation friction force:
m c d u i d s = e c F i k u k + m c f i .
Surely, the radiation friction force f i has to satisfy the condition u i f i = 0 because of u i u i = 1 and F i k = F k i . By transitioning to the continuous distributions of charge and mass and using the equality μ / m = ρ / e , where e is the particle charge and ρ is the charge density, one obtains
μ c d u i d s = ρ c F i k u k + μ c f i μ c d u i d t = 1 c F i k ρ u k d s d t + μ c f i d s d t ,
or, introducing the current j k ,
μ c d u i d t = 1 c F i k j k + μ c f i d s d t .
Thus, for the particle in the field, one finds the equation:
k T i k [ p ] = μ c d u i d t = 1 c F i k j k + μ c f i d s d t .
In order for the conservation law to be satisfied for the particle–field system,
k ( T i k [ f ] + T i k [ p ] ) = 0 ,
it is necessary to add the compensating term Q i = μ c f i ( d s / d t ) to the r.h.s. of Equation (9). Then, Equation (9) becomes
k T i k [ f ] = 1 c F i k j k μ c f i d s d t .
Let us parameterize the effect of the radiation friction in the leading approximation by the following expression (see, e.g., [57,58,59]):
f i = 2 e 2 3 m c 2 ( δ k i u i u k ) d 2 u k d s 2 or f i = 2 e 2 3 m c 2 ( g i k u i u k ) d 2 u k d s 2 ,
which certainly satisfies the mandatory requirement u i f i = 0 . The quantity 2 e 2 / 3 m c 2 here is nothing but the so-called “classical radius of the electron”, r e (with the accuracy within a numerical coefficient of the order of 1). Variants of Equation (29), in which additional nonlocal terms are added, are also presented in the literature, but discussion of this issue is beyond the scope of this paper. Equation (29) can be rewritten in an alternative form, taking into account the feature that u k u k = 1 and, consequently, for the derivatives u k ( d u k / d s ) = 0 and u k ( d 2 u k / d s 2 ) = ( d u i / d s ) ( d u i / d s ) .
Thus, for the particle, Equation (23) written for the covariant component of 4-velocity becomes
μ c d u i d s = ρ c F i k u k + μ c r e ( d 2 u i d s 2 + u i d u k d s d u k d s ) .
The analogous equation for the field is
k T i k [ f ] = ρ c F i l u l d s d t μ c r e ( d 2 u i d s 2 + u i d u k d s d u k d s ) d s d t .
For the time component of Equation (31) with u 0 = c d t / d s (which is equal to γ ), one explicitly obtains:
1 c t T 0 0 [ f ] + α T 0 α [ f ] = ρ c F 0 α u α d s d t μ c r e ( d w 0 d s + u 0 w k w k ) d s d t ,
where T 0 0 [ f ] = ( 8 π ) 1 ( E 2 + H 2 ) W f is the spatial density of the electromagnetic field energy, T 0 α [ f ] = c 1 S α . Here, F 0 α = E α , F 0 α = E α are the spatial components of the anti-symmetrical electromagnetic tensor; and u 0 = γ , u α = γ β α . One has
1 c t W f + μ c r e d w 0 d t = ρ c E α γ β α d s d t + μ c r e w k w k u 0 d s d t 1 c · S ,
i.e.,
1 c t W f + μ c r e d w 0 d t = ρ c c E α β α + μ c 2 r e ( w k w k ) 1 c · S ,
i.e.,
t W f + μ c 2 r e d w 0 d t = ρ c ( β · E ) + μ c 3 r e ( w k w k ) · S .
After integration over a certain volume V, one obtains:
d d t ( W f ¯ + m c 2 r e w 0 ) = d Σ · S + m c 3 r e ( w k w k ) e c β · E ¯ , where W f ¯ = d V E 2 + H 2 8 π .
The first term in the r.h.s. of Equation (36) has a straightforward interpretation, namely it gives the electromagnetic energy flow “outflowing” from the volume V through the remote surface Σ . The contribution to this term is made only by the most slowly decreasing part of the field created by the accelerating charge.
The second term in the r.h.s. of Equation (36) represents the power of the radiation friction force: it increases the total energy of the electromagnetic field surrounding the moving charge. It is straightforward to verify using the definition of w i in Equation (2) that
P = m c 3 r e ( w i w i ) = m c 3 r e ( g i k w k w i ) m c 3 r e ( w 0 w 0 + w μ w μ ) = m c 1 r e ( 1 β 2 ) a 2 γ 8 ( 1 β 2 sin 2 ψ ) = 2 3 e 2 c 3 γ 6 ( a 2 [ β , a ] 2 ) ,
where ψ is the angle between the directions of the spatial velocity, v = c β , and the spatial acceleration a of the particle.
Let us illustrate the elementary calculation so that to follow the details (here, in this calculation only, v β , in order to not be confused with the indices). The first term in the r.h.s. of Equation (37) gives γ 8 v 2 a 2 cos 2 ψ due to the scalar product feature. The next term is γ 8 ( a β γ 2 + v β v a cos ψ ) ( a β γ 2 + v β v a cos ψ ) , i.e., γ 8 ( a 2 γ 4 + 2 γ 2 a 2 v 2 cos 2 ψ + a 2 v 4 cos 2 ψ ) = γ 8 a 2 ( 1 2 v 2 + v 4 + 2 v 2 cos 2 ψ 2 v 4 cos 2 ψ + v 4 cos 2 ψ ) = γ 8 a 2 ( 1 2 v 2 + 2 v 2 cos 2 ψ + v 4 sin 2 ψ ) = γ 8 a 2 ( 1 2 v 2 sin 2 ψ + v 4 sin 2 ψ ) . Next, it is transformed into γ 8 a 2 ( v 2 cos 2 ψ + 1 2 v 2 + 2 v 2 cos 2 ψ + v 4 sin 2 ψ ) = γ 8 a 2 ( 1 2 v 2 + v 2 cos 2 ψ + v 4 sin 2 ψ ) = γ 8 a 2 ( 1 v 2 v 2 sin 2 ψ + v 4 sin 2 ψ ) = γ 8 a 2 ( 1 v 2 ) ( 1 v 2 sin 2 ψ ) . Finally, one finds the resulting expression:
2 3 e 2 c w i w i = 2 3 e 2 c 3 γ 6 ( a 2 a 2 β 2 sin 2 ψ ) .
The expression (38) can be rewritten as
P = 2 3 e 2 c 3 ( a 2 [ a , β ] 2 ) ( 1 β 2 ) 3 .
It is necessary to remember the features of the cross product in Cartesian and curvilinear coordinates.
When the charged particle moves with much slower than the speed of light ( β 1 ) in an inertial frame of reference, Equation (37) gives
d E = 2 3 e 2 c 3 a 2 d t .
Equation (40) is the known Larmor’s formula [68] for bremsstrachlung. Here, certainly, a = a ( t ) , t = t R / c and d t = d t .
In the r.h.s. of Equation (36), the third term, E = e c β · E ¯ = E β + E a + E f , represents the work performed by the charge in a unit of time against the field E in order to generate an additional electromagnetic field. Section 5 discusses in detail how to evaluate the work E and clarifies on each term.

5. Structure of Field in Vicinity of Charge

In Equation (36), the field E = E β + E a + E f is measured at the location of the charge and is composed of three terms. The last term, E f , is the free (external) field satisfying the homogenous-wave equation. The first two terms, E β and E a , result from the Lienard–Wichert formulation.
Indeed, when a point-like charge e moves in vacuum along a certain trajectory, r ( t ) , the electromagnetic fields, E and H , are determined by the known expressions that follow from the Lienard–Wiechert potentials [59]. More precisely, since the field potential A satisfies an inhomogeneous-wave equation in which the source is the charge, then the potential is found using the Green function method [69]. For a point-like source, by rearranging the order of integration, the mentioned Lienard–Wiechert potentials are found, and then the fields E and H are calculated. The resulting expressions are known (see, e.g., [59]):
E LW = E β + E a = e γ 2 R 2 n β ( 1 n · β ) 3 + e c 2 R [ n , [ ( n β ) , a ] ] ( 1 n · β ) 3 , H = [ n , E ] .
Here, R = | R | is the distance between the observation point described by r ( t ) , and the charge location described by r 0 ( t ) . The vectors in the expressions (41) are, therefore, R = r 0 ( t ) r ( t ) , n = R / R . The field E β R 2 describes the field accompanying the charge, and the field E a a / R describes the field radiated by the charge. The fields E and H in the left hand side (l.h.s.) of the expressions (41) are taken here at the point of observation r ( t ) at the instant t. The quantities in the r.h.s. of expressions (41) are taken at the instant t , which satisfies the equation t = t + c 1 | r 0 ( t ) r ( t ) | . Vector R = r 0 ( t ) r ( t ) is drawn from the point where the charge is located, to the observation point. Also, the derivative v ( t ) v = d r ( t ) / d t is the velocity of the charge, and the acceleration is a = d v / d t . Let us also write v = c β .
Hereafter, we focus on the terms E β and E a satisfying the inhomogeneous-wave equation, and not on E f , which satisfies the homogeneous-wave equation.
In the expression E = e c β · E ¯ = E β + E a + E f , the first term E β is zero. Indeed,
E β 0 π d θ sin θ β · n β 2 ( 1 β · n ) 3 = 0 ,
where the relations for the components β = ( 0 , 0 , β ) and n = ( cos ϕ sin θ , sin ϕ sin θ , cos θ ) have been used.
The term E a is not zero. Here, we are speaking about the interaction of the charge with the generated radiation field: just below in this Section, the charge density is parameterized by a “spread-out” three-dimensional delta-function in order to emphasize the point that the charge is localized in a relatively small space-region of the size ϵ . Then, in the third term in the r.h.s. of Equation (36), E is the field to be integrated over the domain (quasi-point-like) where this charge is located. To evaluate this term, we parameterize the delta-function as follows: δ ϵ ( r ) = ( 4 π ) 1 3 ϵ 2 / ( r 2 + ϵ 2 ) 5 / 2 . This function satisfies all the necessary requirements: it is symmetrical with respect to the angular variables, has a sharp maximum at r = 0 , quickly falls to zero when r , and the volume integral of δ ϵ ( r ) (at d V = d r r 2 sin θ d θ d ϕ ) results in 1. Note that r 2 r r 2 r ( ( 4 π ) 1 ( r 2 + ϵ 2 ) = δ ϵ ( r ) . Therefore, in the denominator of the expression for the field E a , one should put R 2 R 2 + ϵ 2 ; that is, the nonphysical singularity at r = 0 is avoided.
Generally, it can be assumed that the parameter ϵ / m c , with the reduced Planck constant. Indeed, classical electrodynamics is limited to scales larger than the Compton length. At smaller scales, quantum effects come into play.
Calculation of the integral E a results into the following:
E a = e c β · E a ¯ = e c 0 d R R 2 ( 1 4 π 3 ϵ 2 ( R 2 + ϵ 2 ) 5 / 2 d Ω ( e c 2 ) 1 ( R 2 + ϵ 2 ) 1 / 2 β · [ n , [ ( n β ) , a ] ] ( 1 n · β ) 3 ,
i.e.,
E a = 3 e 2 4 π c ϵ 0 d y y 2 ( y 2 + 1 ) 3 d Ω β · [ n , [ ( n β ) , a ] ] ( 1 n · β ) 3 = 3 4 π π 16 e 2 c ϵ d Ω β · [ n , [ ( n β ) , a ] ] ( 1 n · β ) 3 .
From the decomposition [ n , [ ( n β ) , a ] ] = ( n β ) ( n · a ) a ( n · ( n β ) ) = n ( n · a ) β ( n · a ) a ( n · n ) + a ( β · n ) , we rewrite the numerator in the angular integral of Equation (44) as K ( θ , ϕ ) = β · [ n , [ ( n β ) , a ] ] = ( β · n ) ( n · a ) ( β · β ) ( n · a ) ( β · a ) ( n · n ) + ( β · a ) ( β · n ) = ( β · a ) β 2 ( n · a ) + ( β · n ) ( n · a ) + ( β · a ) ( β · n ) . Then, we express all vectors in the components β = ( 0 , 0 , β ) , a = ( a sin ψ , 0 , a cos ψ ) , and n = ( cos ϕ sin θ , sin ϕ sin θ , cos θ ) , and calculate the integrals. Then, one finds that
E a = 3 π 16 e 2 c ϵ 1 β 2 ( 1 β 2 ) 1 1 β 2 β ln 1 + β 1 β ( β · a ) .
Taking into account that the time derivative of the modulus of the vector β ( t ) is written as d β / d t = d β · β / d t = ( β ) 1 β · d β / d t , Equation (45) can be rewritten as
E a = 3 π 16 e 2 ϵ 1 Arctanh β β , β d β d t = d d t 3 π 16 e 2 ϵ 1 Arctanh β β .
This expression describes the work of the charged particle against the (radiated) electromagnetic field generated by the particle. Insert this expression into the equation of particle motion:
m c d u i d s = e c F i l u l ¯ + m c r e d 2 u i d s 2 + u i d u k d s d u k d s .
For the time component of Equation (47) with u 0 = c d t / d s (which is equal to γ ), one explicitly obtains
d d s m c γ = e c F 0 α u α ¯ + μ c r e d w 0 d s + γ w k w k .
After some straightforward transformations, Equation (48) is reduced to
d d t m c 2 1 β 2 2 e 2 3 c 2 β · a ( 1 β 2 ) 2 + 3 π 16 e 2 ϵ Arctanh β β 1 = P .
The expression (49) gives the link between the change of full/kinetic energy E p E p * (the expression in the square brackets, which includes the rest energy) of the charged particle and the power P of the radiation friction force acting onto the particle.

6. Results and Discussion

This analysis employed the framework of classical electrodynamics and relativity.
Equation (49) can be formally written as ( d / d t ) ( m * c 2 / 1 β 2 ) = P , where the notation m * for the “effective mass” ( m * = m + δ m ) is introduced:
m * = m + 3 π 16 e 2 c 2 ϵ 1 β 2 Arctanh β β 1 2 e 2 3 c 4 β · a ( 1 β 2 ) 3 / 2 .
The expression within the square brackets in Equation (49)—denoted as E p * —just signifies that the link between the full/kinetic energy of a moving charged ( e 2 0 ) particle and its velocity β , becomes more complicated than for a moving neutral ( e 2 = 0 ) particle when the last two terms in E p * disappear. Equation (50), rewritten in the form
m * m = 1 + 9 π 32 r e ϵ 1 β 2 Arctanh β β 1 r e c τ β 2 ( 1 β 2 ) 3 / 2 β β · a τ β c = 1 + S 1 ( β ) S 2 ( β ) β β · a τ β c ,
allows us to estimate the field’s contribution to the change: E p 0 = m c 2 E p = m c 2 γ E p * . Here, τ is the characteristic time interval during which the particle velocity changes significantly. The second (always positive) term in the r.h.s. of Equation (51) shows that the “mass”—which is an inertial property—“increases” even more when the particle is moving ( β 0 or γ 1 ) because of the presence of the field ( e 2 0 ). If the space scale ϵ is assumed to be of the order of the Compton length λ C = / ( m c ) , then the coefficient r e / ϵ in the second term of Equation (51) is of the order of e 2 / ( c ) 1 / 137.04 which is the fine-structure constant ( 7 · 10 3 ). Then, for all values of β , the mass add-on from the second term in Equation (51) is small ( δ m < 0.003 ).
A similar reasoning may be applied to the third term in Equation (51), where the scalar product β · a may be either positive or negative.
Let us estimate the order of the magnitude of the acceleration as a c Δ β / τ c β / τ . Here, τ is the characteristic time of the meaningful change in the particle velocity. Then, the coefficient in front of the function β 2 / ( 1 β 2 ) 3 / 2 , resulting from the third term in the r.h.s. of Equation (51), namely r e / c τ , can be written as r e / c τ 10 23 / τ . Here, numerical values of r e 2.8 × 10 13 and c 3 · 10 10 in the CGS (cm–g–s) system are used. On the other hand, just as there exists the characteristic space-scale ϵ / ( m c ) , so does the characteristic time-scale τ ϵ ϵ / c = ( / ( m c ) ) / c = / m c 2 10 21 exist, where the numerical values m c 2 0.5 × 10 6 e v and 6.6 × 10 16 e v · s are used. Thus, τ must be larger than 10 21 s in order for the classical electrodynamics to be applicable. Therefore, coefficient r e / c τ has to be less than 10 2 . Then—in order to remain within the bounds of applicability of the classical theory—the acceleration a (which is of the order of the ratio of the velocity change to the time-interval over which this change occurs) must be less than the characteristic value c β / τ ϵ . Thus, for the moderate values of velocity ( 0 < β < 0.9 ), the third term in Equation (51) is of the order of the fine-structure constant, i.e., the term is small relative to 1.
Thus, we may state that in most common cases, the add-on δ m to the inertial rest mass m is small and the approximation m * m may be used. The add-on can be significant for the very fast processes (with very small τ 10 21 ) occurring within the ultra-relativistic domain ( β 1 ). But that is already stepping into the topic of nuclear collision.
The part of the parameter m * that is linearly dependent on the acceleration—i.e., the third term in Equation (51)—can be interpreted as the so-called “added mass”. It appears only when the charge is moving with acceleration. To make quite a rough analogy, a similar phenomenon occurs in classical hydrodynamics (see, e.g., [70,71]) when a material body moving with acceleration appears to “pull” into its motion some part of the fluid surrounding the body. More accurately, the “added mass” is a quantity which has the dimension of a mass and which is “added” to the mass of the body (which moves unevenly in a liquid medium) in order to take into account the influence of the medium on this body. For example, if an absolutely rigid body of mass m moves rectilinearly in an ideal fluid under the influence of force f, then the resistance of the medium is proportional to the body’s acceleration a, i.e., m a = f m a a or ( m + m a ) a = f , where the coefficient of proportionality m a is called the “added mass”. Consequently, a body with mass m moves in a liquid in the same way as it would have moved in the emptiness and had the mass m + m a . The value of the “added mass” depends on the shape of the body, its orientation relative to the direction of movement, the density of the medium, etc. Thus, one has to pay particular attention when applying such an interpretation to electrodynamic systems.
We return to Equation (36) and, after substituting Equation (46) into Equation (36), one obtains
d W d t = P r + P ,
where
W = W f ¯ ( m * m ) c 2 γ W f ¯ + m c 2 r e w 0 + 3 π 16 e 2 ϵ 1 Arctanh β β , P r = d Σ · S and P = m c 3 r e ( w k w k ) , w i = ( w 0 , w ) = γ 4 c 2 β · a , γ 2 c 2 a + γ 4 c 2 β ( β · a ) .
Equation (52) has a straightforward physical sense. It explains how the energy is balanced between three processes: radiation escaping through the remote spherical surface Σ , excitation of the electromagnetic field by the particle decelerating by the action of the radiative friction force, and the change in the field’s volume density. On the other hand, Equation (52) shows that the three quantities—the change in the field energy per unit time ( d W / d t ), the total energy flux of the outflowing field ( P r ), and the power of the “radiative friction force” (P)—are connected by only one relationship. In a general case, they are far from being equal to each other in terms of the absolute magnitude. Only if a stationary motion with d W / d t = 0 is considered, then the equality P = P r takes place. This circumstance was noted in a methodological note [72].
To renormalize the energy of the electromagnetic field accompanying the moving charge, one has to replace W f ¯ with W in Equation (53).
The field energy W, as it follows from Equation (53), is composed of three components: the first component does not depend on the charge’s acceleration a, the second component is linearly proportional to a, and the third component is proportional to a 2 . To calculate the first component, the integration over the volume is carried out over the entire space; the singularity at the location of the charge is avoided via the use of a small spatial parameter ϵ . As for the components with a and a 2 , one has to take into account that any physical process is the process with initial conditions and with a finite duration. In the case under study, the initial time-moment is the moment when the charge begins to move with acceleration. It is from this moment that the radiation field emerges and the terms, linear and quadratic with respect to the acceleration, appear in the expression for W. Furthermore, all field quantities calculated from the Lienard–Wiechert potentials are taken at time t , which is calculated from the equation R ( t ) = c ( t t ) . This means that the size of the space-region, where the field generated by the accelerating charge exists, is finite, so the volume integral is taken not from zero to infinity, but from zero to a finite value calculated from the equation R ( t ) = c ( t t ) . In such a case, there is no divergence of the volume integral at the upper limit.
If one associates the field energy with some effective mass (of the electromagnetic field) using the relation W = m f c 2 , then Equations (51) and (53) actually just denotes a renormalization of the field mass due to the appearance of the add-on dependent on both the velocity of the charge and its acceleration.
Note that the first and the second terms in the r.h.s. of Equation (52), cannot be equal to each other because, for the term P, the force is applied to the charge along its trajectory, while for the term P r , the energy flux of the field is integrated over the surface with a large characteristic radius. It is certain that, in full accord with the spirit of the field theory, the energy flux through the surface is determined by the field in the vicinity of this surface and not by the field along the charge’s trajectory, which lays inside the volume bounded by the surface.
To conclude, this paper aimed to present—in the framework of classical theory—a self-consistent derivation scheme which produces equations that may be practically used for analyses of various scenarios. The charge is treated as a quasi-point-like charge; this helps to resolve the complications of the “infinite” electromagnetic energy, which are avoided by the procedure of slightly “spreading” the charge. As a result, the physical interpretation of the finite size of a particle is presented in a new light. Indeed, it is within the charge, at scales smaller than Compton’s length of an order of / m c , where the quantum-mechanics approach to be applied. Beyond this region, no “infinite” tails of quantities accumulate. The seeming divergences of the integrals at the upper limits are not physical if one takes into account that the charge moves with acceleration only for a finite duration of time. In every real physical scenario, there is always a starting and an end moment.

Author Contributions

Conceptualization and writing, E.P.T. and V.I.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

No new data were created or analyzed in this study.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A. The Green Function: Interpretation and Application

When a neutral particle is point-like, its spatial mass density distribution is μ = m δ ( 3 ) ( x x p ( t ) ) ; its integral of action is S p = m c d s (see [59], §27).
When one considers a particle with an arbitrary mass density distribution μ , its integral of action can be written as
S p = c d Ω μ 1 β α β α = c d Ω μ γ 1 .
Here, d Ω = d ( c t ) d V , μ is normalized by the condition m = d V μ . The interval in the 4-dimensional spacetime d s = d q i d q i = c d t β i β i = c d t 1 β α β α . We introduce β α = c 1 ( d q α / d t ) , β i = ( 1 , β ) , β i = ( 1 , β ) . The particle 4-velocity is u i = γ β i . The quantity γ = ( 1 β 2 ) 1 / 2 is the Lorentz factor.
For the next step, the action of the charged particle plus an electromagnetic field is
S p S = S p + S int + S f = c d Ω μ 1 β α β α 1 c d Ω ρ A i β i 1 16 π c d Ω F i k F i k ,
where F i k = i A k k A i is the electromagnetic field strength tensor. The particle charge density ρ is normalized by the condition e = d V ρ , similar to the condition for the mass m = d V μ .
From Equation (A2), the Lagrangian 4-volume density is
Λ = c μ 1 β α β α 1 c ρ A 0 1 c ρ A ν β ν 1 16 π c F i k F i k .
The first two terms in Equation (A2) depend on the 4-spacetime location of the particle q i , on its velocity β i = c 1 ( d q i / d t ) , and on the field via the 4-potential A i . The third term depends only on the field. When varying the action S along coordinates δ q k , with δ β i = c 1 δ ( d q i / d t ) = c 1 ( d ( δ q i ) / d t ) , it is necessary to take into account that the last term does not contribute to the variation, but the second term does because when the position of the particle changes virtually, the field value at this new point differs from the prior one. We also take into account that according to the procedure for calculating variations and variational derivatives, the quantities β i and q k are considered as independent variables.
We remind that, generally speaking, the Lagrangian density is not uniquely determined; for instance, S S + λ d Ω ρ + const .
Next, the functional derivative of the action
δ S δ q i = 1 c d d t Λ β i + Λ q i
turns out to be equal to
δ S δ q i = 1 c d d t c μ β i 1 β α β α 1 c ρ A i 1 c i ( ρ A k β k ) = d d t μ β i 1 β α β α + 1 c β k k ( ρ A i ) 1 c β k i ( ρ A k ) 1 c ρ A k i β k .
The derivative (A5) has to be zero. Note that the last term in expression (A5) has to be taken as zero because the quantities β i and q k during the derivation of the equation are considered as independent quantities.
Thus, the equation for the charged particle motion, with d t = γ d s / c , becomes
d d s ( μ c u i ) = 1 c u k ( i ( ρ A k ) k ( ρ A i ) ) .
The l.h.s. of (A6) can be rewritten as μ c d u i / d s + c u i d μ / d s . However, the second term here is equal to zero for the charged particle, which moves along the trajectory x = x p ( t ) . Indeed, in this case, one has d μ / d s = γ β k k μ ( x x p ( t ) ) . Here, one just has to take into account that the action of the operator β k k on any function, including the delta-function whose argument is θ = q ν q 0 ν ( t ) , yields zero. Indeed, from the definition of the dimensionless velocity of the particle, β , it follows that β k k ( q μ q 0 μ ( t ) ) = β k k q μ β k k q 0 μ ( t ) = β k δ k μ 0 q 0 μ ( t ) = 0 . Thus, the equation of motion for a non-point--like particle is
μ c d u i d s = 1 c u k ( i ( ρ A k ) k ( ρ A i ) ) ,
which must be solved together with the electromagnetic field equations for potentials A i and their derivatives.
As for the point-like particle, i.e., when ( μ , ρ ) = ( m , e ) δ ( 3 ) ( x x p ( t ) ) , the situation here is quite nuanced and require higher attention. Let us note here that the use of the delta-function, or its spread-out analogue, is nothing more than a case-appropriate approximation of an object, made for studying processes in the spacetime region where classical electrodynamics still applies, i.e., with the scales exceeding the Compton length. For smaller scales, quantum-field physics must be used.
Let us rewrite Equation (A7) so that the terms dependent on ( μ , ρ ) and on their derivatives are arranged at opposing sides:
μ c d u i d s ρ 1 c u k ( i A k k A i ) = 1 c ( A k u k i ρ u k A i k ρ ) .
Now, on the r.h.s. of Equation (A8), only the terms proportional to the potential are present. One can then conclude that the r.h.s. must equal zero because the potentials—and not their derivatives, which determine the “force characteristics”—must not appear directly in the equations of motion for “measurable” dynamic quantities, such as the equation of motion for a particle.
The feature that the potential A k is equal to zero at the location of the point-like charge can be understood from the following considerations.
First, the electromagnetic field equations are written for the electromagnetic stress–energy tensor as k F i k = ( 4 π / c ) j i (see also Equation (8)). Second, after imposing the Lorentz condition k A k = 0 on the potentials A k , one obtains
k k A i = 4 π c j i .
Here, k k = c 2 t 2 Δ ; the symbol Δ = α α is the Laplace operator; the current components, j k , are j 0 = c ρ , j = ρ c β ; and the tensor F i k = i A k k A i . Raising and lowering the indices are carried out according to the standard procedure using the metric tensor g i k = ( 1 , δ α β ) . Then,
A i = 4 π d Ω G ( t t , x x ) j i ( t , x ) + A 0 i ,
where d Ω = c d t d x d y d z ; A 0 is the solution of the homogenous-wave equation; and the Green function is introduced:
k k G ( t t , x x ) = δ ( 1 ) ( c ( t t ) ) δ ( 3 ) ( x x ) .
Using the Fourier transform, the full Green function may be found quite elementarily:
G ( t t , x x ) = G r + G a = c 4 π | x x | δ ( 1 ) ( c ( t t ) | x x | ) + c 4 π | x x | δ ( 1 ) ( c ( t t ) + | x x | ) .
Using the properties of the delta-function, the full Green function of the wave equation in the spacetime with Galilean metrics can be rewritten as
G ( t t , x x ) = c 2 π δ ( 1 ) ( c 2 ( t t ) 2 | x x | 2 ) .
From Equation (A13), it follows that the Green function is not zero when its argument is zero. The argument of the first delta-function in Equation (A12) can become zero when t < t , i.e., at the previous moments of time; the argument of the second term in the same equation can be zero only when t > t . The first term on the r.h.s. of the expression (A12) represents the so-called retarded Green function, the second term represents the advanced Green function. Because for all physical processes, the principle of causality must hold, one should deal only with the first term; the field at the observation point at instant t is determined by the superposition of impacts from the states of the system/particle collected only along its world-line at the preceding, t < t , moments of time. The unnecessary branch of the Green function is cut off by choosing such an integration contour along the time axis so that the singularity point is bypassed (i.e., by the deformation of the contour in the vicinity of the singularity in the complex time domain). To perform this, a Heaviside cutoff function H ( ξ ) is introduced with the following properties: H ( ξ ) = 1 when ξ > 0 ; H ( ξ ) = 0 when ξ < 0 ; and notably, H ( 0 ) = 0 . Then, the Green function in Equation (A10) is written as
G r = c 4 π δ ( 1 ) ( c 2 ( t t ) 2 | x x | 2 ) H ( t t ) .
Now, for the point-like charge, we substitute Equation (A14) into Equation (A10) (without any free term A 0 i ) and integrate it: first over the volume, and then over time. Hence, one obtains the Lienard–Wichert potentials. One finds that A k = ( e u k / | R i u i | ) H ( t t * ) . Here, u k is the 4-velocity of the charge at t = t * , R k = ( c ( t t * ) , r r 0 ( t * ) ) and parameter t * satisfies equation R k R k = 0 . Certainly, R k 0 at the point where the charge is located, but at this point, the Heavyside function H ( 0 ) = 0 is strictly equal to zero. Therefore, potential A i is zero at the point where the point-like charge is located. Thus, the r.h.s. of Equation (A8) is zero for the point-like charge.
The above then points that, for the point-like charged particle, Equation (A8) degenerates into the classical expression:
m c d d s u i = e c u k ( i A k k A i ) = e c F i k u k .

References

  1. Sommerfeld, A. Über die Beugung und Bremsung der Elektronen. Ann. Phys. 1931, 403, 257–330. [Google Scholar] [CrossRef]
  2. Bethe, H.A.; Maximon, L.C. Theory of bremsstrahlung and pair production. I. Differential cross section. Phys. Rev. 1954, 93, 768–784. [Google Scholar] [CrossRef]
  3. Davies, H.; Bethe, H.A.; Maximon, L.C. Theory of bremsstrahlung and pair production. II. Integral cross section for pair production. Phys. Rev. 1954, 93, 788–795. [Google Scholar] [CrossRef]
  4. Karzas, W.J.; Latter, R. Electron radiative transitions in a Coulomb field. Astrophys. J. Suppl. 1962, 6, 167–212. [Google Scholar] [CrossRef]
  5. Haug, E. Bremsstrahlung and pair production in the field of free electrons. Z. Naturforsch. A 1975, 30, 1099–1113. [Google Scholar] [CrossRef]
  6. Haug, E. Electron-electron bremsstrahlung in a hot plasma. Z. Naturforsch. A 1975, 30, 1546–1552. [Google Scholar] [CrossRef]
  7. Koch, H.W.; Motz, J.W. Bremsstrahlung cross-section formulas and related data. Rev. Mod. Phys. 1959, 31, 920–955. [Google Scholar] [CrossRef]
  8. Akhiezer, A.I.; Berestetskii, V.B. Quantum Electrodynamics; Intersience Publishers/John Wiley & Sons, Inc.: New York, NY, USA, 1965; Available online: https://archive.org/details/akhiezer-berestetskii-quantum-electrodynamics/ (accessed on 5 November 2024).
  9. Blumenthal, G.R.; Gould, R.J. Bremsstrahlung, synchrotron radiation, and Compton scattering of high-energy electrons traversing dilute gases. Rev. Mod. Phys. 1970, 42, 237–271. [Google Scholar] [CrossRef]
  10. Bekefi, G. Radiation Processes in Plasmas; John Wiley & Sons, Inc.: New York, NY, USA, 1966; Available online: https://archive.org/details/radiationprocess0000beke/ (accessed on 5 November 2024).
  11. Jauch, J.M.; Rohrlich, F. The Theory of Photons and Electrons. The Relativistic Quantum Field Theory of Charged Particles with Spin One-Half; Springer: Berlin/Hedelberg, Germany, 1976. [Google Scholar] [CrossRef]
  12. Haug, E.; Nakel, W. The Elementary Process of Bremsstrahlung; World Scientific: Singapore, 2004. [Google Scholar] [CrossRef]
  13. Munirov, V.R.; Fisch, N.J. Suppression of bremsstrahlung losses from relativistic plasma with energy cutoff. Phys. Rev. E 2023, 107, 065205. [Google Scholar] [CrossRef]
  14. Jarrott, L.C.; Kemp, A.J.; Divol, L.; Mariscal, D.; Westover, B.; McGuffey, C.; Beg, F.N.; Suggit, M.; Chen, C.; Hey, D.; et al. Kα and bremsstrahlung X-ray radiation backlighter sources from short pulse laser driven silver targets as a function of laser pre-pulse energy. Phys. Plasmas 2014, 21, 031211. [Google Scholar] [CrossRef]
  15. Cheng, S.; Hill, F.A.; Heubel, E.V.; Velásquez-García, L.F. Low-bremsstrahlung X-ray source using a low-voltage high-current-density nanostructured field emission cathode and a transmission anode for markerless soft tissue imaging. J. Microelectromech. Syst. 2015, 24, 373–383. [Google Scholar] [CrossRef]
  16. Huntington, C.M.; McNaney, J.M.; Gumbrell, E.; Krygier, A.; Wehrenberg, C.; Park, H.-S. Bremsstrahlung X-ray generation for high optical depth radiography applications on the National Ignition Facility. Rev. Sci. Instrum. 2018, 89, 10G121. [Google Scholar] [CrossRef] [PubMed]
  17. Underwood, C.I.D.; Baird, C.D.; Murphy, C.D.; Armstrong, C.D.; Thornton, C.; Finlay, O.J.; Streeter, M.J.V.; Selwood, M.P.; Brierley, N.; Cipiccia, S.; et al. Development of control mechanisms for a laser wakefield accelerator-driven bremsstrahlung x-ray source for advanced radiographic imaging. Plasma Phys. Control. Fusion 2020, 62, 124002. [Google Scholar] [CrossRef]
  18. Singh, S.; Armstrong, C.D.; Kang, N.; Ren, L.; Liu, H.; Hua, N.; Rusby, D.R.; Klimo, O.; Versaci, R.; Zhang, Y.; et al. Bremsstrahlung emission and plasma characterization driven by moderately relativistic laser–plasma interactions. Plasma Phys. Control. Fusion 2021, 63, 035004. [Google Scholar] [CrossRef]
  19. Vyskočil, J.; Klimo, O.; Weber, S. Simulations of bremsstrahlung emission in ultra-intense laser interactions with foil targets. Plasma Phys. Control. Fusion 2018, 60, 054013. [Google Scholar] [CrossRef]
  20. Stevens, J.; von Goeler, S.; Bernabei, S.; Bitter, M.; Chu, T.; Efthimion, P.; Fisch, N.; Hooke, W.; Hosea, J.; Jobes, F.; et al. Modelling of the electron distribution based on bremsstrahlung emission during lower-hybrid current drive on PLT. Nucl. Fusion 1985, 25, 1529–1541. [Google Scholar] [CrossRef]
  21. Voss, K.E.; Fisch, N.J. Inverse problem for bremsstrahlung radiation. Phys. Fluids B Plasma Phys. 1992, 4, 762–763. [Google Scholar] [CrossRef]
  22. Peysson, Y.; Arslanbekov, R. Measurement of the non-thermal bremsstrahlung emission between 30 and 200 keV with a high time-space resolution on the tokamak TORE SUPRA. Nucl. Instrum. Meth. Phys. Res. A Accelerat. Spectrom. Detect. Assoc. Equip. 1996, 380, 423–426. [Google Scholar] [CrossRef]
  23. Chen, Z.Y.; Wan, B.N.; Lin, S.Y.; Shi, Y.J.; Hu, L.Q.; Ma, T.P. Measurement of the non-thermal bremsstrahlung emission between 20 and 7000 keV in the HT-7 Tokamak. Nucl. Instrum. Meth. Phys. Res. A Accelerat. Spectrom. Detect. Assoc. Equip. 2006, 560, 558–563. [Google Scholar] [CrossRef]
  24. Chen, C.D.; King, J.A.; Key, M.H.; Akli, K.U.; Beg, F.N.; Chen, H.; Freeman, R.R.; Link, A.; Mackinnon, A.J.; MacPhee, A.G.; et al. A Bremsstrahlung spectrometer using k-edge and differential filters with image plate dosimeters. Rev. Sci. Instrum. 2006, 79, 10E305. [Google Scholar] [CrossRef]
  25. Meadowcroft, A.L.; Edwards, R.D. High-energy bremsstrahlung diagnostics to characterize hot-electron production in short-pulse laser-plasma experiments. IEEE Trans. Plasma Sci. 2012, 40, 1992–2001. [Google Scholar] [CrossRef]
  26. Swanson, C.; Jandovitz, P.; Cohen, S.A. Using Poisson-regularized inversion of Bremsstrahlung emission to extract full electron energy distribution functions from x-ray pulse-height detector data. AIP Adv. 2018, 8, 025222. [Google Scholar] [CrossRef]
  27. Kagan, G.; Landen, O.L.; Svyatskiy, D.; Sio, H.; Kabadi, N.V.; Simpson, R.A.; Gatu Johnson, M.; Frenje, J.A.; Petrasso, R.D.; Shah, R.C.; et al. Inference of the electron temperature in inertial confinement fusion implosions from the hard X-ray spectral continuum. Contrib. Plasma Phys. 2019, 59, 181–188. [Google Scholar] [CrossRef]
  28. Kumar, J.; Pandya, S.P.; Sharma, P. Conceptual design of multichannel fast electron bremsstrahlung detection system to study fast electron dynamics during lower hybrid current drive in ADITYA-U tokamak. J. Instrum. 2023, 18, P03040. [Google Scholar] [CrossRef]
  29. Brown, J.C. The deduction of energy spectra of non-thermal electrons in flares from the observed dynamic spectra of hard X-ray bursts. Sol. Phys. 1971, 18, 489–502. [Google Scholar] [CrossRef]
  30. Bakhtiari, M.; Kramer, G.J.; Takechi, M.; Tamai, H.; Miura, Y.; Kusama, Y.; Kamada, Y. Role of bremsstrahlung radiation in limiting the energy of runaway electrons in tokamaks. Phys. Rev. Lett. 2005, 94, 215003. [Google Scholar] [CrossRef]
  31. Embréus, O.; Stahl, A.; Fülöp, T. Effect of bremsstrahlung radiation emission on fast electrons in plasmas. New J. Phys. 2016, 18, 093023. [Google Scholar] [CrossRef]
  32. Kellogg, E.; Baldwin, J.R.; Koch, D. Studies of cluster X-ray sources. Energy spectra for the Perseus, Virgo, and Coma clusters. Astrophys. J. 1975, 199, 299–306. [Google Scholar] [CrossRef]
  33. Svensson, R. Electron-positron pair equilibria in relativistic plasmas. Astrophys. J. 1982, 258, 335–348. [Google Scholar] [CrossRef]
  34. Chluba, J.; Ravenni, A.; Bolliet, B. Improved calculations of electron–ion bremsstrahlung Gaunt factors for astrophysical applications. Mon. Not. R. Astron. Soc. 2020, 34921, 177–194. [Google Scholar] [CrossRef]
  35. Pradler, J.; Semmelrock, L. Exact Theory of nonrelativistic quadrupole bremsstrahlung. Astrophys. J. 2021, 909, 134. [Google Scholar] [CrossRef]
  36. Pradler, J.; Semmelrock, L. Accurate Gaunt Factors for nonrelativistic quadrupole bremsstrahlung. Astrophys. J. 2021, 916, 105. [Google Scholar] [CrossRef]
  37. Pradler, J.; Semmelrock, L. Nonrelativistic electron–ion bremsstrahlung: An approximate formula for all parameters. Astrophys. J. 2021, 922, 57. [Google Scholar] [CrossRef]
  38. Sunyaev, R.A.; Zeldovich, Y.B. The interaction of matter and radiation in the hot model of the Universe, II. Astrophys. Space Sci. 1970, 7, 20–30. [Google Scholar] [CrossRef]
  39. Chluba, J.; Sunyaev, R.A. The evolution of CMB spectral distortions in the early Universe. Mon. Not. R. Astron. Soc. 2012, 419, 1294–1314. [Google Scholar] [CrossRef]
  40. Narayan, R.; Yi, I. Advection-dominated accretion: Underfed black holes and neutron stars. Astrophys. J. 1995, 452, 710–735. [Google Scholar] [CrossRef]
  41. Yarza, R.; Wong, G.N.; Ryan, B.R.; Gammie, C.F. Bremsstrahlung in GRMHD models of accreting black holes. Astrophys. J. 2020, 898, 50. [Google Scholar] [CrossRef]
  42. Sarazin, C.L. X-ray emission from clusters of galaxies. Rev. Mod. Phys. 1986, 58, 1–115. [Google Scholar] [CrossRef]
  43. Sarazin, C.L.; Kempner, J.C. Nonthermal bremsstrahlung and hard X-Ray emission from clusters of galaxies. Astrophys. J. 2000, 533, 73–83. [Google Scholar] [CrossRef]
  44. Pearce, F.R.; Thomas, P.A.; Couchman, H.M.P.; Edge, A.C. The effect of radiative cooling on the X-ray properties of galaxy clusters. Mon. Not. R. Astron. Soc. 2000, 317, 1029–1040. [Google Scholar] [CrossRef]
  45. Haug, E. Contribution of electron-electron bremsstrahlung to solar hard X-radiation during flares. Solar Phys. 1975, 45, 453–458. [Google Scholar] [CrossRef]
  46. Holman, G.D.; Sui, L.; Schwartz, R.A.; Emslie, A.G. Electron bremsstrahlung hard X-Ray spectra, electron distributions, and energetics in the 2002 July 23 solar flare. Astrophys. J. 2003, 595, L97–L101. [Google Scholar] [CrossRef]
  47. Kontar, E.P.; Emslie, A.G.; Massone, A.M.; Piana, M.; Brown, J.C.; Prato, M. Electron-electron bremsstrahlung emission and the inference of electron flux spectra in solar flares. Astrophys. J. 2007, 670, 857–861. [Google Scholar] [CrossRef]
  48. Wen, Z.G.; Yuan, J.P.; Wang, N.; Li, D.; Chen, J.L.; Wang, P.; Wu, Q.D.; Yan, W.M.; Yuen, R.; Wang, Z. A Single-pulse study of the subpulse drifter PSR J1631+1252 discovered at FAST. Astrophys. J. 2022, 929, 71. [Google Scholar] [CrossRef]
  49. Deng, Z.-L.; Li, X.-D.; Gao, Z.-F.; Shao, Y. Evolution of LMXBs under different magnetic braking prescriptions. Astrophys. J. 2021, 909, 174. [Google Scholar] [CrossRef]
  50. 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]
  51. 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. Not. R. Astron. Soc. 2016, 456, 55–65. [Google Scholar] [CrossRef]
  52. Bethe, H.; Heitler, W. On the stopping of fast particles and on the creation of positive electrons. Proc. R. Soc. Lond. A Math. Phys. Engin. Sci. 1934, 146, 83–112. [Google Scholar] [CrossRef]
  53. Heitler, W. The Quantum Theory of Radiation; Oxford University Press: Oxford, UK, 1947; Available online: https://archive.org/details/in.ernet.dli.2015.37198/ (accessed on 5 November 2024).
  54. Keefe, D. Inertial confinement fusion. Annu. Rev. Nucl. Part. Sci. 1982, 32, 391–441. [Google Scholar] [CrossRef]
  55. Craxton, R.S.; Anderson, K.S.; Boehly, T.R.; Goncharov, V.N.; Harding, D.R.; Knauer, J.P.; McCrory, R.L.; McKenty, P.W.; Meyerhofer, D.D.; Myatt, J.F.; et al. Direct-drive inertial confinement fusion: A review. Phys. Plasmas 2015, 22, 110501. [Google Scholar] [CrossRef]
  56. Firouzi Farrashbandi, N.; Eslami-Kalantari, M. Inverse bremsstrahlung absorption in laser-fusion plasma. J. Theor. Appl. Phys. 2020, 14, 261–264. [Google Scholar] [CrossRef]
  57. Dirac, P.A.M. Classical theory of radiating electrons. Proc. R. Soc. Lond. A Math. Phys. Engin. Sci. 1938, 167, 148–169. [Google Scholar] [CrossRef]
  58. Jackson, J.D. Classical Electrodynamics; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 1999; Available online: https://archive.org/details/john-david-jackson-classical-electrodynamics-wiley-1999/ (accessed on 5 November 2024).
  59. Landau, L.D.; Lifshitz, E.M. The Classical Theory of Fields; Pergamon Press plc/Elsevier Ltd: Oxford, UK, 2013; Available online: https://www.sciencedirect.com/book/9780080250724/the-classical-theory-of-fields (accessed on 5 November 2024).
  60. Landau, L.D.; Lifshitz, E.M. Mechanics; Elsevier Butterworth-Heinemann/Elsevier Ltd.: Oxford, UK, 1982. [Google Scholar] [CrossRef]
  61. Misner, C.W.; Thorne, K.S.; Wheeler, J.A. Gravitation; W.H. Freeman and Company: San Francisco, CA, USA, 1973; Available online: https://physicsgg.me/wp-content/uploads/2023/05/misner_thorne_wheeler_gravitation_freema.pdf (accessed on 5 November 2024).
  62. Feynman, R.P.; Leighton, R.B.; Sands, M. The Feynman Lectures on Physics; The California Institute of Technology: Pasadena, CA, USA, 2013; Volume 2, Ch. 28; Available online: https://www.feynmanlectures.caltech.edu/II_28.html (accessed on 5 November 2024).
  63. Fermi, E. Über einen Widerspruch zwischen der elektrodynamischen und der relativistischen Theorie der elektromagnetischen Masse. Phys. Z. 1922, 23, 340–344. Available online: https://archive.org/details/physikalische-zeitschrift-vol-23/page/340/mode/2up (accessed on 5 November 2024).
  64. Fermi, E. Concerning a contradiction between the electrodynamic and relativistic theory of electromagnetic mass. Wikisource. 15 September 2023. Available online: https://en.wikisource.org/wiki/Translation:Electrodynamic_and_Relativistic_Theory_of_Electromagnetic_Mass (accessed on 5 November 2024).
  65. Schwinger, J. Electromagnetic mass revisited. Found. Phys. 1983, 13, 373–383. [Google Scholar] [CrossRef]
  66. Rohrlich, F. The dynamics of a charged sphere and the electron. Am. J. Phys. 1997, 65, 1051–1056. [Google Scholar] [CrossRef]
  67. Gralla, S.E.; Harte, A.I.; Wald, R.M. Rigorous derivation of electromagnetic self-force. Phys. Rev. D 2009, 80, 024031. [Google Scholar] [CrossRef]
  68. Larmor, J.J. On the theory of the magnetic influence on spectra; and on the radiation from moving ions. Philos. Mag. 1897, 44, 503–512. [Google Scholar] [CrossRef]
  69. Morse, P.M.; Feshbach, H. Methods of Theoretical Physics. Part I; McGraw-Hill Book Company, Inc.: New York, NY, USA, 1953; Available online: https://archive.org/details/methodsoftheoret0000phil_w6e3 (accessed on 5 November 2024).
  70. Lamb, H. Hydrodynamics; Cambridge University Press: New York, NY, USA, 1997; Available online: https://archive.org/details/hydrodynamics00lamb/ (accessed on 5 November 2024).
  71. Landau, L.D.; Lifshitz, E.M. Fluid Mechanics; Pergamon Press plc/Elsevier Ltd.: Oxford, UK, 1987. [Google Scholar] [CrossRef]
  72. Ginzburg, V.L. Radiation and radiation friction force in uniformly accelerated motion of a charge. Sov. Phys. Uspekhi 1970, 12, 565–574. [Google Scholar] [CrossRef]
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

Tito, E.P.; Pavlov, V.I. Accelerating Charge: Add-Ons to Rest Mass and Field Energy. Physics 2024, 6, 1264-1280. https://doi.org/10.3390/physics6040078

AMA Style

Tito EP, Pavlov VI. Accelerating Charge: Add-Ons to Rest Mass and Field Energy. Physics. 2024; 6(4):1264-1280. https://doi.org/10.3390/physics6040078

Chicago/Turabian Style

Tito, Elizabeth P., and Vadim I. Pavlov. 2024. "Accelerating Charge: Add-Ons to Rest Mass and Field Energy" Physics 6, no. 4: 1264-1280. https://doi.org/10.3390/physics6040078

APA Style

Tito, E. P., & Pavlov, V. I. (2024). Accelerating Charge: Add-Ons to Rest Mass and Field Energy. Physics, 6(4), 1264-1280. https://doi.org/10.3390/physics6040078

Article Metrics

Back to TopTop