Open access peer-reviewed chapter

Stopping Power of Ions in a Magnetized Plasma: Binary Collision Formulatio

Written By

Hrachya B. Nersisyan, Günter Zwicknagel and Claude Deutsch

Reviewed: 12 April 2018 Published: 27 February 2019

DOI: 10.5772/intechopen.77213

From the Edited Volume

Plasma Science and Technology - Basic Fundamentals and Modern Applications

Edited by Haikel Jelassi and Djamel Benredjem

Chapter metrics overview

1,221 Chapter Downloads

View Full Metrics

Abstract

In this chapter, we investigate the stopping power of an ion in a magnetized electron plasma in a model of binary collisions (BCs) between ions and magnetized electrons, in which the two-body interaction is treated up to the second order as a perturbation to the helical motion of the electrons. This improved BC theory is uniformly valid for any strength of the magnetic field and is derived for two-body forces which are treated in Fourier space without specifying the interaction potential. The stopping power is explicitly calculated for a regularized and screened potential which is both of finite range and less singular than the Coulomb interaction at the origin. Closed expressions for the stopping power are derived for monoenergetic electrons, which are then folded with an isotropic Maxwell velocity distribution of the electrons. The accuracy and validity of the present model have been studied by comparisons with the classical trajectory Monte Carlo numerical simulations.

Keywords

  • ion stopping
  • magnetized plasma target
  • binary collisions

1. Introduction

There is an ongoing in the theory of interaction of charged particle beams with plasmas. Although most theoretical works have reported on the energy loss of ions in a plasma without magnetic field, the strongly magnetized case has not yet received as much attention as the field-free case. The energy loss of ion beams and the related processes in magnetized plasmas are important in many areas of physics such as transport, heating, magnetic confinement of thermonuclear plasmas, and astrophysics. The range of the related topics includes ultracold plasmas [1, 2], the cooling of heavy ion beams by electrons [3, 4, 5, 6, 7, 8, 9, 10, 11, 12], as well as many very dense systems involved in magnetized target fusions [11], or heavy ion inertial confinement fusion (ICF).

For a theoretical description of the energy loss of ions in a plasma, there exist some standard approaches. The dielectric linear response (LR) treatment considers the ion as a perturbation of the target plasma, and the stopping is caused by the polarization of the surrounding medium. It is generally valid if the ion couples weakly to the target. Since the early 1960s, a number of calculations of the stopping power (SP) within LR treatment in a magnetized plasma have been presented (see Refs. [13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37] and references therein). Alternatively, the stopping is calculated as a result of the energy transfers in successive binary collisions (BCs) between the ion and the electrons [37, 38, 39, 40, 41, 42, 43, 44, 45]. Here, it is necessary to consider appropriate approximations for the screening of the Coulomb potential by the plasma [8]. However, significant gaps between these approaches involve the ion stopping along magnetic field B and perpendicular to it. In particular, at high B values, the BC predicts a vanishingly parallel energy loss, which remains at variance with the nonzero LR one. Also, challenging BCLR discrepancies persist in the transverse direction, especially for vanishingly small ion projectile velocity vi when the friction coefficient contains an anomalous term diverging logarithmically at vi=0 [23, 24]. For calculation of the energy loss of an ion, two new alternative approaches have been recently suggested. One of these methods is specifically aimed at a low-velocity energy loss, which is expressed in terms of velocity-velocity correlation and, hence, to a diffusion coefficient [34]. Next, in Ref. [27] using the Bhatnagar-Gross-Krook approach based on the Boltzmann-Poisson equations for a collisional and magnetized classical plasma, the energy loss of an ion is studied through a LR approach, which is constructed such that it conserves particle number locally.

An alternative approach, particularly in the absence of any relevant experimental data, is to test various theoretical methods against comprehensive numerical simulations. This can be achieved by a particle-in-cell (PIC) simulation of the underlying nonlinear Vlasov-Poisson Equation [10, 31]. While the LR requires cutoffs to exclude hard collisions of close particles, the collectivity of the excitation can be taken into account in both LR and PIC approaches. In the complementary BC treatment, the stopping force has been calculated numerically by scattering statistical ensembles of magnetized electrons from the ions in the classical trajectory Monte Carlo (CTMC) method [7, 10, 37, 38, 39, 40, 41]. For a review we refer to a recent monograph [8] which summarizes all theoretical and numerical methods and approaches also discussing the ranges of their validity.

The very recent upheaval of successful experiments involving hot and dense plasmas in the presence of kilotesla magnetic fields (e.g., at ILE (Osaka), CELIA (Bordeaux), LULI (Palaiseau), LLNL (Livermore)) remaining nearly steady during 10–15 ns strongly motivates the fusion as well as the warm dense matter (WDM) communities to investigate adequate diagnostics for their dynamic properties. This opens indeed a novel perspective by allowing magnetic fields to play a much larger if not a central role both in ICF and WDM plasmas. In this context proton or any nonrelativistic ion stopping is likely to provide an option of choice for investigating genuine magnetization features such as anisotropy, when the electron plasma frequency turns significantly lower than the cyclotron one [46]. In addition, an experimental test of proton or alpha particle stopping in a magnetized plasma is currently envisioned (see, e.g., Ref. [46] for a preliminary discussion). The parameters at hand are a fully ionized hydrogen plasma with a density up to 1020cm3 and temperature between 1 and 100 eV. The steady magnetic field can be up to 45 T strong. A preliminary examination based on comparing electron Debye length with corresponding Larmor radius indicates that to experience a strong influence of the magnetic field, the electron density should be comparable with a few 1016cm3. We expect these endeavors to lead to the very first unambiguous and genuine identification of an experimental magnetic signature for nonrelativistic ion stopping in plasmas.

Motivated by these recent developments, our purpose is to investigate the SP of an ion moving in a magnetized plasma in a wide range of the value of a steady magnetic field. The present paper is based on our earlier studies in Refs. [8, 24, 44, 45] where the second-order energy transfers for individual collisions of electron-ion [8, 24, 44] of any two identical particles, like electron-electron [44], and finally of two gyrating arbitrary charged particles [45] have been calculated with the help of an improved BC treatment. This treatment is—unlike earlier approaches of, e.g., Refs. [9, 42]—valid for any strength of the magnetic field. As the first application of the theoretical BC model developed in Refs. [8, 24, 44, 45], we have calculated in Ref. [47] the cooling forces on the heavy ion beam interacting with a strongly magnetized and temperature anisotropic electron beam. It has been shown that there is a quite good overall agreement with both the CTMC numerical simulations and the experiments performed at the ESR storage ring at GSI [48, 49, 50].

In Section 2 we introduce briefly a perturbative binary collision formulation in terms of the binary force acting between an ion and a magnetized electron and derive general expressions for the second-order (with respect to the interaction potential) stopping power. In contrast to the previous investigations in Refs. [8, 24, 44, 45], we here consider the (macroscopic) stopping force which is obtained by integrating the binary force of an individual electron-ion interaction with respect to the impact parameter and the velocity distribution function of electrons. That is, the stopping force for monoenergetic electrons is folded with a velocity distribution. The resulting expressions involve all cyclotron harmonics of the electrons’ helical motion and are valid for any interaction potential and any strength of the magnetic field. In Section 2.4 we present explicit analytic expressions of this second-order stopping power for the specific case of a regularized and screened interaction potential [51, 52] which is both of finite range and less singular than the Coulomb interaction at the origin and which includes as limiting cases the Debye (i.e., screened) and the Coulomb potentials. For comparison of our expressions with previous approaches, we consider in Section 3 the corresponding asymptotic expressions for large and small ion velocities and strong and vanishing magnetic fields. The analytical expressions presented in Section 2.4 are evaluated numerically in Section 4 using parameters of the envisaged experiments on ion stopping [46]. In particular, we compare our approach with the CTMC simulations. The results are summarized and discussed in Section 5. The regularization parameter and the screening length involved in the interaction potential are briefly specified and discussed in Appendix A.

Advertisement

2. Theoretical model

2.1. Binary collision (BC) formulation

Let us consider two point charges with masses m,M and charges e,Ze, respectively, moving in a homogeneous magnetic field B=Bb. We assume that the particles interact with the potential Ze2Ur with e2=e2/4πϵ0, where ϵ0 is the permittivity of the vacuum and r=r1r2 is the relative coordinate of the colliding particles. For two isolated charged particles, this interaction is given by the Coulomb potential, i.e., UCr=1/r. In plasma applications UC is modified by many-body effects and the related screening and turns into an effective interaction. In general, this effective interaction, which is related to the wake field induced by a moving ion, is non-spherically symmetric and depends also on the ion velocity. For any BC treatment, however, this complicated ion-plasma interaction must be approximated by an effective two-particle interaction Ur. This effective interaction U may be modeled by a spherically symmetric Debye-like screened interaction uDr=er/λ/r with a screening length λ, given, e.g., by the Debye screening length λD (see, e.g., [16]), in case of low ion velocities and an effective velocity-dependent screening length λvi for larger ion velocities vi (see [53, 54, 55]). Further details on the choice of the effective interaction Ur are given in Ref. [47].

In the presence of an external magnetic field, the Lagrangian and the corresponding equations of particle motion cannot, in general, be separated into parts describing the relative motion and the motion of the center of mass (cm) [8]. However, in the case of heavy ions, i.e., Mm, the equations of motion can be simplified by treating the cm velocity vcm as constant and equal to the ion velocity vi, i.e., vcm=vi=const. Then, introducing the velocity correction through relations δvt=vetve0t=vtv0t, where vt=rt=vetvi is the relative electron-ion velocity ve0t and v0t=ṙ0t=ve0tvi are the unperturbed electron and relative velocities, respectively, the equation of relative motion turns into

r0t=R0+vrt+ausinωctb×ucosωct,E1
δv̇t+ωcδvt×b=Ze2mfrt.E2

Here, Ze2frtf=U/r is the force exerted by the ion on the electron, ωc=eB/m is the electron cyclotron frequency, and δvt0 at t. In Eq. (1) u=cosφsinφ is the unit vector perpendicular to the magnetic field; the angle φ is the initial phase of the electron’s helical motion; vr=vebvi is the relative velocity of the guiding center of the electrons, where ve and ve (with ve0) are the unperturbed components of the electron velocity parallel and perpendicular to b, respectively; and a=ve/ωc is the cyclotron radius. In Eq. (1), the quantities u and R0 are defined by the initial conditions. In Eq. (2) rt=retvit is the ion-electron relative coordinate.

2.2. The perturbative treatment

We seek an approximate solution of Eq. (2) in which the interaction force between the ion and electron is considered as a perturbation. Thus, we are looking for a solution of Eq. (2) for the variables r and v in a perturbative manner r=r0+r1+,v=v0+v1+, where r0t,v0t are the unperturbed ion-electron relative coordinate and velocity, respectively, and rnt,vntn=12 are the nth-order perturbations of rt and vt, which are proportional to Zn.

The parameter of smallness which justifies such kind of expansion can be read off from a dimensionless form of the equation of motion Eq. (2) by scaling lengths in units of the screening length λ, velocities in units of the initial relative velocity v0, and time in units of λ/v0. Then, it is seen (see Ref. [47] for details) that the perturbative treatment is essentially applicable in cases where Ze2/mv02λ<1, that is, when the (initial) kinetic energy of relative motion mv02/2, is large compared to the characteristic potential energy Ze2/λ in a screened Coulomb potential. Or, expressed in velocities, the initial relative velocity v0 must exceed the characteristic velocity vd=Ze2/1/2, that is, vd here demarcates the perturbative from the non-perturbative regime. If this condition is met not only for a single ion-electron collision but in the average over the electron distribution, e.g., by replacing v0 with the averaged initial ion-electron relative velocity v0, i.e., v0vd, we are in a regime of weak ion-target or, here, weak ion-electron coupling, which allows the use of perturbative treatments (besides BC also, e.g., linear response (LR)). For nonmagnetized electrons this is discussed in much detail in Refs. [53, 54]. Even though the particle trajectories are much more intricate in the presence of an external magnetic field, the given definitions and demarcations of coupling regimes are basically the same for magnetized electrons. That is, the applicability of a perturbative treatment is essentially related to the charge state Z of the ion and the typical range λ of the effective interaction, but not directly on the strength B of the magnetic field. The latter may affect the critical velocity vd only implicitly via a possible change of the effective screening length λ with B.

The equation for the first-order velocity correction is obtained from Eq. (2) replacing on the right-hand side of the exact relative coordinate rt by r0t with the solutions v1t=ṙ1t and

r1t=Ze2mbQt+RebbQtQt+ib×Qt.E3

Here, we have introduced the following abbreviations:

Qt=tbfr0τtτ,Qt=1iωctfr0τeωctτ1E4

and have assumed that all corrections vanish at t.

2.3. Second-order stopping power

We now consider the interaction process of an individual ion with a homogeneous electron plasma described by a velocity distribution function fve and a density ne. We assume that the ion experiences independent binary collisions (BCs) with the electrons. The total stopping force, Fvi, acting on the ion is then obtained by multiplying the binary force Ze2frt by the element of the flux relative flux nevrd2sdt, integrating with respect to time and folding with velocity distribution of the electrons. The impact parameter s introduced here in the electron flux is defined by s=R0=R0nrnrR0 and is the component of R0 perpendicular to the relative velocity vector vr with nr=vr/vr. As can be inferred from Eq. (1), s represents the distance of the closest approach between the ion and the guiding center of the electron’s helical motion.

The resulting stopping power, Svi=viviFvi, then reads

Svi=Ze2nevidvefvevrd2svifrtdt,E5

which is an exact relation for uncorrelated BCs of the ion with electrons. We evaluate this expression within a systematic perturbative treatment (see Ref. [47] for more details). First, we introduce the two-particle interaction potential Ur, and the binary force fr is written using Fourier transformation in space. Furthermore, the factor eikrt in the Fourier transformed binary force is expanded in a perturbative manner as eikrteikr0t1+ikr1t, where r0t and r1t are the unperturbed and the first-order corrected relative coordinates (Eqs. (1) and (3)), respectively. Next, we consider only the second-order binary force f2 and the corresponding stopping force F2 with respect to the binary interaction since the averaged first-order force F1 (related to f1) vanishes due to symmetry reasons [8, 24, 44, 45, 47]. Within the second-order perturbative treatment, the stopping power can be represented as

Svi=Ze2nevidvefvevrd2sdkUkkvikr1teikr0tdt.E6

From Eq. (6) it is seen that the second-order stopping power is proportional to Z2. Inserting now Eqs. (1) and (3) into Eq. (6), assuming an axially symmetric velocity distribution fve=fveve, and performing the s integration, we then obtain

S=2π4Z2e4nemvidve0fvevevedve×dkUk2kvi0k2+k2sinωctωct×J02kasinωct2sinkvrtdtd,E7

where Jn is the Bessel function of the nth order; k=kb and k are the components of k parallel and transverse to b, respectively; and ve and ve are the electron velocity components parallel and transverse to b, respectively. This general expression (7) for the stopping power of an individual ion has been derived within second-order perturbation theory but without any restriction on the strength of the magnetic field B.

2.4. The SP for a regularized and screened coulomb potential

For an electron plasma with an isotropic Maxwell distribution, the velocity distribution relevant for the averaging in Eq. (7) is given by

fve=12π3/2vth3eve2/2vth2,E8

where the thermal velocity vth is related to the electron temperature by vth2=T/m (here, the temperature is measured in energy units). Inserting Eq. (8) into expression (7) and assuming now a spherically symmetric potential U=Uk yields after performing the velocity integrations (see Ref. [56]), the stopping power

Svi=8Z2e4nemωcvi2π440dk0U2kkdk0et22k2a2ek2a2(1costk2+k2sintttdt×kaicoskaitJ1kait+kaisinkaitJ0kait.E9

Here, we have introduced the thermal cyclotron radius of the electrons a=vth/ωc, and ai=vi/ωc, ai=vi/ωc, where vi and vi are the ion velocity components transverse and parallel to b, respectively. For the Coulomb interaction Uk=UCk, the full two-dimensional integration over the s-space results in a logarithmic divergence of the k integration in Eqs. (7) and (9). To cure this, cutoff parameters kmin and kmax must be introduced (see, e.g., Refs. [8, 24, 47] for details). These cutoffs are related to the screening of the interaction in a plasma target and the incorrect treatment of hard collisions in a classical perturbative approach. As an alternative implementation of this standard cutoff procedure, we here employ the regularized screened interaction Ur=URr=1er/λer/λ/r with the Fourier transform

URk=22π21k2+λ21k2+d2,E10

where d1=λ1+ƛ1. UR represents a Debye-like screened interaction UD (see Section 2.1) which is additionally regularized at the origin [51, 52] and thus removes the problems related to the Coulomb singularity in a classical picture and prevents particles (for Z>0) from falling into the center of the potential. The parameter λ related to this regularization is here considered as a given constant or as a function of the classical collision diameter [47].

Substituting the interaction potential (10) into Eq. (9) and performing the k integration, we arrive, after lengthy but straightforward calculations, at

Svi=4πZ2e4nemvth2v0dtt01expv2ζ2PtζΦΨtζ×P1tζ+sinαtαtP2(tζ)ζ21ζ2Gtζ,E11

where Ptζ=cos2ϑ+sin2ϑ/Gtζ and

P1tζ=2cos2ϑ+Ptζ12v2ζ2cos2ϑ,E12
P2tζ=2Gtζsin2ϑGtζ+Ptζ1v2ζ2sin2ϑGtζ.E13

Here, we have introduced the dimensionless quantities v=vi/2vth, α=ωcλ/vth. ϑ is the angle between b and vi, Ψtζ=t2/21ζ2/ζ2, Gtζ=Θtζ2+1ζ2, Θt=2αtsinαt22, and

Φz=ez+eϰ2z2ϰ211zezeϰ2z,E14

where ϰ=λ/d=1+λ/ƛ.

Eq. (11) for the SP is the main result of the outlined BC treatment which will now be evaluated in the next sections.

Advertisement

3. Comparison with previous approaches

Previous theoretical expressions for the stopping power which have been extensively discussed by the plasma physics community (see, e.g., Refs. [3, 8] for reviews) basically concern the two limiting cases of vanishing and infinitely strong magnetic fields. We therefore investigate the present approach for these two cases, first for arbitrary interactions Uk and electron distributions fve as given by Eq. (7) and later for the more specific situation of the regularized interaction (10) and the velocity distribution (8) as given by Eq. (11).

3.1. General SP Eq. (7) at vanishing and infinitely strong magnetic fields

At vanishing magnetic field B0, sinωct/ωct1 and the argument of the Bessel function in Eq. (7) should be replaced by kvet. Then, denoting the second-order SP at vanishing magnetic field as S0 and assuming spherically symmetric potential with U=Uk, one obtains

S0vi=42π2Z2e4nemvi2U0vifveve2dve,E15

where U is the generalized Coulomb logarithm:

U=2π440U2kk3dk.E16

Employing the regularized and screened potential Uk given by Eq. (10), the generalized Coulomb logarithm is U=UR=Λϰ (see also Refs. [8, 24, 44, 45]), where

Λϰ=ϰ2+1ϰ21lnϰ1.E17

Taking the bare Coulomb interaction with Uk=UCk1/k2, Eq. (16) diverges logarithmically at k0 and k, and two cutoffs kmin=1/rmax and kmax=1/rmin must be introduced as discussed in Section 2.4. In this case the generalized Coulomb logarithm takes the standard form U=UC=lnkmax/kmin=lnrmax/rmin.

The asymptotic expression of Eq. (15) at high ion velocities can be easily derived using the normalization of the distribution function which results in

S0vi4πZ2e4nemvi2U.E18

At an infinitely strong magnetic field B, the term in Eq. (7) proportional to k2 and the argument of the Bessel function vanish since the cyclotron radius a0. In this limit, denoting the SP as Svi and assuming a spherically symmetric interaction potential, we arrive at

Svi=2πZ2e4nemUvisin2ϑ1v5vive2ve2+vi2fevedve.E19

The corresponding high-velocity asymptotic expression is given by

Svi=2πZ2e4nemvi2Usin2ϑ.E20

Eqs. (15) and (19) and their asymptotic expressions for high velocities in Eqs. (18) and (20), respectively, agree with the results derived by Derbenev and Skrinsky in Ref. [57] in case of the Coulomb interaction potential, i.e., with U=UC. Using instead a regularized interaction potential and thus the Coulomb logarithm, UR allows closed analytic expressions and converging integrals and avoids any introduction of lower and upper cutoffs “by hand” in order to restrict the domains of integration. Moreover, employing the bare Coulomb interaction may, as pointed out by Parkhomchuk [58], result in asymptotic expressions which essentially different from Eqs. (19) and (20), which is related to the divergent nature of the bare Coulomb interaction (see Ref. [47]).

3.2. Some limiting cases of Eq. (11)

Next, we discuss some asymptotic regimes of the SP (Eq. (11)) where the regularized interaction (Eq. (10)) and the isotropic velocity distribution (Eq. (8)) have been assumed. In the high-velocity limit where vi>ωcλvth, only small t contributes to the SP (Eq. (11)) due to the short time response of the electrons to the moving fast ion. In this limit we have sinαt/αt1. The remaining t integration can be performed explicitly. This integral is given by [47].

0dttΦΨtζΛϰ.E21

Here, the function Φz is determined by Eq. (14), and Λϰ is the generalized Coulomb logarithm (Eq. (17)). The remaining expressions do not depend on the magnetic field, i.e., ωc, as a natural consequence of the short time response of the magnetized electrons. In fact, sinαt/αt1 and Gtζ1 and the related t integration (Eq. (21)) are also valid for vanishing magnetic field α0. Integration by parts turns Eq. (11) into

S0=4πZ2e4nemvi2Λϰerfv2πvev2,E22

where erfz is the error function and v=vi/2vth is again the scaled ion velocity. The SP (Eq. (22)) is isotropic with respect to the ion velocity vi and represents the two limiting cases of high velocities at arbitrary magnetic field and arbitrary velocities at vanishing field. Of course, expression (22) can be also obtained by performing the remaining integration in the nonmagnetized SP (Eq. (15)) using the isotropic velocity distribution (Eq. (8)) and U=Λϰ.

A further increase of the ion velocity finally yields

S04πZ2e4nemvi2Λϰ,E23

which completely agrees with the asymptotic expression (18) in case of U=Λϰ. Inspecting Eq. (23) shows that the SP does not depend explicitly on the electron temperature T at sufficiently high velocities, while T may still be involved in the generalized Coulomb logarithm Λϰ.

At B0 and small velocities vi<vth, the SP (Eq. (22)) becomes

S04π2πZ2e4ne3mvth3viΛϰ.E24

Now, we consider the situation when the magnetic field is very strong and the electron cyclotron radius is the smallest length scale, ωcλvivth, and the SP is only weakly sensitive to the transverse electron velocities and, hence, is affected only by their longitudinal velocity spread. In this limit sinαt/αt0 and Gtζ1ζ2 are obtained from Eq. (11) after straightforward calculations:

S=4ππZ2e4nemvth2vΛϰ01ev2ζ2Pζζ22cos2ϑ+Pζ12v2ζ2cos2ϑ,E25

where Pζ=cos2ϑ+sin2ϑ/1ζ2.

After changing the variable ζ in Eq. (25) to x=ζPζ1/2 and some subsequent rearrangement, Eq. (25) can be expressed alternatively as

S=2πZ2e4nemvth2vΛϰsin2ϑev2x2x2dx1+x22xcosϑ)3/2.E26

Up to the definition of the Coulomb logarithm (i.e., U=Λϰ versus U=UC), the expressions are identical to those obtained by Pestrikov [59].

In particular, at ϑ=0 and ϑ=π/2 (i.e., when ion moves parallel or transverse to the magnetic field, respectively), Eq. (25) (or Eq. (26)) yields

S=4πZ2e4nemvth2Λϰvev2,E27
S=2πZ2e4nemvth2Λϰvev2/21+v2K0v22v2K1v22.E28

respectively, where Knz (with n=0,1) is the modified Bessel function. It is also constructive to obtain the angular averaged stopping power. From Eq. (25) one finds

S¯v=120πSvϑsinϑdϑ=4πZ2e4ne3mvi2Λϰerfv+2πvv2E1v2ev2,E29

where E1z is the exponential integral function.

In the high-velocity limit with ωcλvivth, the SP (Eq. (25)) becomes

S2πZ2e4nemvi2Λϰsin2ϑerfv2πvev2+4πv3ev2cos2ϑ.E30

With further increase of the ion velocity, we can then neglect the exponential term in Eq. (30), while erfv1 yields the asymptotic expression (Eq. (20)) (for U=Λϰ).

The SP given by Eq. (30) (or Eq. (20) with U=Λϰ) decays as the corresponding SP (Eq. (23)) like vi2 with the ion velocity. But here, the parallel SP (Eq. (27)) vanishes exponentially at ϑ=0 which is a consequence of the presence of a strong magnetic field, where the electrons move parallel to the magnetic field. If the ion moves also parallel to the field (i.e., ϑ=0), the averaged stopping force must vanish within the BC treatment for symmetry reasons.

Finally, we also investigate the case of small velocities at strong magnetic fields. Considering a small ion velocity v1 in Eq. (25), we arrive at

S=4πZ2e4nemvth2Λϰvsin2ϑln2vsinϑγ21+cos2ϑ,E31

where γ0.5772 is Euler’s constant. Now, it is seen that the SP, S, leads at low ion velocities v1 and for a nonzero ϑ to a term which behaves as vln1/v. Thus, the corresponding friction coefficient diverges logarithmically at small v. This is a quite unexpected behavior compared to the well-known linear velocity dependence without magnetic field (see asymptotic expressions above). Finally, at ϑ=0 the logarithmic term vanishes and the SP behaves as Sv.

Advertisement

4. Features of the SP (Eq. (11)) and comparison with CTMC simulations

In this section we study some general properties of the SP of individual ions resulting from the BC approach by evaluating Eq. (11) numerically. We consider the effect of the magnetic field on the SP at various temperatures of the plasma. The density ne1016cm3 and the temperatures T1eV, 10 or 100 eV of the electron plasma, are in the expected range of the envisaged experiments on proton or alpha particles stopping in a magnetized target plasma [46] (see corresponding Figures 13). As an example we choose proton projectile for our calculations. In all examples considered below, the regularization parameter ƛ0=1010mm thereby meets the condition ƛ0b00, i.e., ƛ0, and does not affect noticeably the SP (Eq. (11)) at low and medium velocities as shown in Appendix A (see also Ref. [47] for more details).

Figure 1.

The SP [in keV/cm] for protons as a function of the ion velocity vi [in units of vth] and for fixed plasma temperature T=1eV. The theoretical stopping power (Eq. (11)) is calculated for ƛ0=1010m (see appendix a for details) and for an electron plasma with ne=1016cm3 in a magnetic field of B=0 (black), 45 T (green), 200 T (blue), 103 T (red), 104 T (green), and B= (cyan). The angle ϑ between B and vi is ϑ=0 (left), ϑ=π/4 (center), and ϑ=π/2 (right). The CTMC results for B= case are shown by the filled circles.

Figure 2.

Same as in Figure 1 but for T=10eV. The SP is given in units eV/cm.

Figure 3.

Same as in Figure 1 but for T=100eV. The SP is given in units eV/cm.

For a BC description beyond the perturbative regime, a fully numerical treatment is required. In the present cases of interest, such a numerical evaluation of the SP is rather intricate but can be successfully implemented by classical trajectory Monte Carlo (CTMC) simulations [37, 38, 39, 40]. In the CTMC method, the trajectories for the ion-electron relative motion are calculated by a numerical integration of the equations of motion (Eq. (2)). The stopping force is then deduced by averaging over a large number (typically 105106) of trajectories employing a Monte Carlo sampling for the related initial conditions. For a more detailed description of the method, we refer to Refs. [8, 44, 45]. Both the analytic perturbative treatment and the non-perturbative numerical CTMC simulations are based on the same BC picture and use the same effective spherical screened interaction Ur. The following comparison of these both approaches thus essentially intends to check the validity and range of applicability of the perturbative approach as it has been outlined in the preceding sections.

Advertisement

5. Stopping profiles and ranges

5.1. General trends

The parameter analysis initiated on Figures 13 at ne=1016cm3 and T=110100eV is implemented for monitoring a possible experimental vindication through a fully ionized hydrogen plasma out of high-power laser beams available on facilities such as ELFIE (Ecole Polytechnique) or TITAN (Lawrence Livermore) [62]. The given adequately magnetized targets (in the 20–45 T range) would then be exposed to TNSA laser-produced proton beams out of the same facilities, in the hundred keV-MeV energy range [62].

Therefore, we are looking for the most conspicuous effect of the applied magnetized intensity B on the proton stopping.

Fixing ne and varying T (see Figures 13) display an ubiquitous and increasing anisotropy shared by the stopping profiles (SP) with increasing B and θ and angle between B and initial projectile velocity V.

Moreover, that anisotropy evolves only moderately between θ=π4 and π2.

Another significant feature is the extension to any B0 of the B=0 scaling neT. For instance, SP at ne=1012cm3 and T=1eV, at a given θ, is equivalent to that for ne=1014cm3 and T=100eV.

As expected, B effects impact essentially the low-velocity section (VVth, Vth = target electron thermal velocity) of the ion stopping profile. One can observe, increasing with B, a shift to the left of SP maxima, as shown in Figure 4 at B=45 tesla, for the profiles displayed in Figure 3, with θ-averaged SP remaining close to θ=π2.

Figure 4.

Same as in Figure 3 restricted to B=45T, featuring θ-dependent and θ-averaged SP in eV/cm.

Switching now attention to corresponding ranges, down to projectile at rest Ep=0, one witnesses on Figure 5 the counterpart of the above-noticed SP behavior.

Figure 5.

Ranges, down to zero energy pertaining to SP in Figure 3.

In a low projectile velocity VVth1, one gets the largest B effects and the smallest proton ranges attributed to the highest B. The fan of B ranges then merges on a given point, located between 10 keV and 100 keV at ne=1016cm3, and then inverts itself with increasing B featuring now increasing ranges. Moreover, the aperture of the fan of ranges increases steadily with θ.

Finally, it can be observed that for θ=0, the infinite magnetized range looks rather peculiar and reminiscent of the ion projectile gliding on BV [8, 34].

5.2. Specific trends

The projected experimental setup [62] could manage constant, static, and homogeneous B values up to 45 T. So, we are let to investigate ne range limits within which significant B effects can be observed.

Obviously, ne=1012cm3 is expected to show quantitatively larger B impact than 1018cm3.

Giving attention to proton ranges of T dependence in a low-density plasma ne=1012cm3 at T=1 and 100 eV, respectively, (Figure 6), one witnesses the smallest ranges for VVth1, increased by four orders of magnitude between 1 and 100 eV while remaining essentially unchanged for VVth1. Turning now to ne=1018cm3 at T=1eV, one can see that the given SP remains quasi-isotropic, hardly θ-dependent, except at extreme magnetization (B=). Discrepancies between B = 0 and 20 T remain visible only for VVth2. B= does not feature anymore the highest stopping when VVth1. Also, B=103 SP exhibits a few top wigglings. Upshifting T at 10 eV yields back ne=1018cm3 SPs very similar to these displayed on Figure 2 (ne=1016cm3, T=10eV) Figure 7.

Figure 6.

Proton ranges down to the rest of the target with ne=1012cm3, T=1, and 100 eV at θ=π2.

Figure 7.

Same as in Figure 1 for ne=1018cm3 and T=1eV with θ=0, π4, and π2.

Corresponding proton ranges (ne=1018cm3, T=1eV) are shown in Figure 8.

Figure 8.

Proton ranges in electron target ne=1012cm3 and T=1eV with θ=0 and π4.

Experimentally, accessible and very small ranges are thus documented for VVth1. Here, B=0 and 20 T data remain everywhere distinguishable.

5.3. Very-low-velocity proton slowing down

Up to now we limited our investigation to proton stopping by target electrons. In the very-low-velocity regime VVthi, the target protons can also contribute significantly as evidenced on Figure 9. This topic will be more thoroughly addressed in a separate presentation.

Figure 9.

Very-low-velocity proton slowing down on target protons at B=103T (upper straight line) contrasted to target electron stopping (any B, lower straight lines).

Advertisement

6. Summary

We developed and extensively used a kinetic approach based on a binary collision formulation and suitably regularized Coulomb interaction, to numerically document for any value of the applied magnetization B, the stopping of a proton projectile in a fully ionized hydrogen plasma target. Both ion projectile and target plasma parameters have been selected in order to fit a planned ion-plasma interaction experiment in the presence of an applied magnetic field B˜. It should be pointed out that we restricted the target plasma to its electron component. It therefore remains to include the target ion contribution to proton stopping [63], thus featuring a complete low-velocity ion slowing down.

More generally, we expect that the present investigation, experimentally geared as it is, could help to bridge a long-standing and persisting gap between theoretical speculations and experimental facts in the field of nonrelativistic ion stopping in magnetized target plasmas.

Advertisement

Acknowledgments

The work of H.B.N. has been supported by the State Committee of Science of the Armenian Ministry of Education and Science (Project No. 13-1C200). This work was supported by the Bundesministerium für Bildung und Forschung (BMBF) under contract 06ER9064.

Advertisement

Our results (Eq. (11)) were derived by using the screened interaction URr. As already mentioned, the use and the modeling of such an effective two-body interaction are a major but indispensable approximation for a BC treatment where the full ion-target interaction is replaced by an accumulation of isolated ion-electron collisions. The replacement of the complicated real non-spherically symmetric potential, like the wake fields as shown and discussed in Ref. [60], with a spherically symmetric one is, however, well motivated by earlier studies on a BC treatment at vanishing magnetic field (see Refs. [53, 54, 55]). It was shown by comparison with 3D self-consistent PIC simulations that the drag force from the real nonsymmetric potential induced by the moving ion can be well approximated by an BC treatment employing a symmetric Debye-like potential with an effective velocity-dependent screening length λvi. In these studies also a recipe was given how to derive the explicit form of λvi, which turned out to be not too much different from a dynamic screening length of the simple form λvi=λD1+vi/vth21/2. Here, λD=vth/ωp is the Debye screening length at vi=0, ωp is the electron plasma frequency, and vth is a thermal velocity of electrons. Although no systematic studies about the use of such an effective interaction with a screening length λvi have been made for ion stopping in a magnetized electron plasma, the replacement of the real interaction by a velocity-dependent spherical one should be a reasonable approximation also in this case. The introduced dynamical screening length λvi also implies the assumption of a weak perturbation of the electrons by the ion and linear screening where the screening length is independent of the ion charge Ze, which coincide with the regimes of perturbative BC (see, e.g., Ref. [54]). Therefore, we do not consider here possible nonlinear screening effects.

Next, we specify the parameter ƛ which is a measure of the softening of the interaction potential at short distances. As we discussed in the preceding sections, the regularization of the potential (Eq. (10)) guarantees the existence of the s integrations, but there remains the problem of treating accurately hard collisions. For a perturbative treatment, the change in relative velocity of the particles must be small compared to vr, and this condition is increasingly difficult to fulfill in the regime vr0. This suggests to enhance the softening of the potential near the origin of the smaller vr. Within the present perturbative treatment, we employ a dynamical regularization parameter ƛvi [44, 45], where ƛ2vi=Cb02vi+ƛ02 and b0vi=Ze2/mvi2+vth2. Here, b0 is the averaged distance of the closest approach of two charged particles in the absence of a magnetic field, and ƛ0 is some free parameter. In addition we also introduced C0.292 in ƛvi. In Refs. [44, 45], this parameter is deduced from the comparison of the second-order scattering cross sections with an exact asymptotic expression derived in Ref. [61] for the Yukawa type (i.e., with ƛ0) interaction potential. As we have shown in Refs. [44, 45] employing the dynamical parameter vi, the second-order cross sections for electron-electron and electron-ion collisions excellently agree with CTMC simulations at high velocities. Also, the free parameter ƛ0 is chosen such that ƛ0b00, where b00 is the distance b0vi at vi=0. From the definition of vi, it can be directly inferred that ƛ0 does not play any role at low velocities, while it somewhat affects the size of the stopping force at high velocities when b0viƛ0. More details on the parameter ƛ0 and its influence on the cooling force are discussed in Ref. [47].

References

  1. 1. Killian TC. Ultracold neutral plasmas. Science. 2007;316:705
  2. 2. Deutsch C, Nersisyan HB, Zwicknagel G. Equilibrium and transport in strongly coupled and magnetized ultracold plasmas. In: Shukla PK, Mendonça JT, Eliasson B, Resedes D, editors. Strongly Coupled Ultra-Cold and Quantum Plasmas. Vol. 1421. AIP Conf. Proc; 2012. p. 3
  3. 3. Sørensen AH, Bonderup E. Electron cooling. Nuclear Instruments & Methods. 1983;215:27
  4. 4. Poth H. Electron cooling, Theory, Experiment, Application. Physics Reports. 1990;196:135
  5. 5. Meshkov IN. Electron cooling, status and perspectives. Physics of Particles and Nuclei. 1994;25:631
  6. 6. Men’shikov LI. Electron cooling. Physics-Uspekhi. 2008;51:645
  7. 7. Zwicknagel G. Electron cooling of ions and antiprotons in traps. In: Nagaitsev S, Pasquinelli RJ, editors. Beam Cooling and Related Topics. Vol. 821. AIP Conf. Proc.; 2006. p. 513
  8. 8. Nersisyan HB, Toepffer C, Zwicknagel G. Interactions between Charged Particles in a Magnetic Field: A Theoretical Approach to Ion Stopping in Magnetized Plasmas. Heidelberg: Springer; 2007
  9. 9. Möllers B, Walter M, Zwicknagel G, Carli C, Toepffer C. Drag force on ions in magnetized electron plasmas. Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms. 2003;207:462
  10. 10. Mollers B, Toepffer C, Walter M, Zwicknagel G, Carli C, Nersisyan HB. Cooling of ions and antiprotons with magnetized electrons. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment. 2004;532:279
  11. 11. Cereceda C, Deutsch C, DePeretti M, Sabatier M, Nersisyan HB. Dielectric response function and stopping power of dens magnetized plasmas. Physics of Plasmas. 2000;7:2884
  12. 12. Cereceda C, DePeretti M, Deutsch C. Stopping power for arbitrary angle between test particle velocity and magnetic field. Physics of Plasmas. 2005;12:022102
  13. 13. Rostoker N, Rosenbluth MN. Test particle in a completely ionized plasma. Physics of Fluids. 1960;3:1
  14. 14. Rostoker N. Kinetic equation with a constant magnetic field. Physics of Fluids. 1960;3:922
  15. 15. Akhiezer IA, Eksp Z. Theory of the interaction of a charged particle with a plasma in a magnetic field. Teor. Fiz. 1961;40:954; Sov. Phys. JETP 1961;13:667
  16. 16. Akhiezer AI, Akhiezer IA, Polovin RV, Sitenko AG, Stepanov KN. Plasma Electrodynamics. 1st ed. Vol. 1. Pergamon: Oxford; 1975
  17. 17. Honda N, Aona O, Kihara T. Fluctuations in a plasma 3: Effect of a magnetic field on stopping power. Journal of the Physical Society of Japan. 1963;18:256
  18. 18. May RM, Cramer NF. Test ion energy loss in a plasma with a magnetic field. Physics of Fluids. 1970;13:1766
  19. 19. Boine-Frankenheim O, DAvanzo J. Stopping power of ions in a strongly magnetized plasma. Physics of Plasmas. 1996;3:792
  20. 20. Nersisyan HB. Stopping of charged particles in a magnetized classical plasma. Physical Review E. 1998;58:3686
  21. 21. Nersisyan HB. Stopping of heavy ions in a magnetized strongly coupled plasmas: high velocity limit. Contributions to Plasma Physics. 2005;45:46
  22. 22. Nersisyan HB, Deutsch C. Correlated fast ion stopping in magntized classical plasma. Physics Letters A. 1998;246:325
  23. 23. Nersisyan HB, Walter M, Zwicknagel G. Stopping power of ions in a magnetized two-temperature plasma. Physical Review E. 2000;61:7022
  24. 24. Nersisyan HB, Zwicknagel G, Toepffer C. Energy loss of ions in a magntized plasma:Conformity between linear response and binary collision treatment. Physical Review E. 2003;67:026411
  25. 25. Nersisyan HB. Ion stopping in a dense magnetized plasmas. Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms. 2003;205:276
  26. 26. Nersisyan HB, Zwicknagel G, Toepffer C. Conformity between linear response and binary collision treatment of an ion energy loss in a magnetized quantum plasma. European Physical Journal D: Atomic, Molecular, Optical and Plasma Physics. 2004;28:235
  27. 27. Nersisyan HB, Deutsch C, Das AK. Number-conserving linear response study of low velocity ion stopping in a collisional magnetized classical plasma. Physical Review E. 2011;83. DOI: 036403
  28. 28. Nersisyan HB, Deutsch C. Low velocity ion stopping in a a collisional plasma. European Physical Journal: Web of Conferences. 2013;59:05019
  29. 29. Nersisyan HB, Deutsch C. Energy loss of ions by electric field fluctuations in a magnetized plasma. Physical Review E. 2011;83:066409
  30. 30. Seele C, Zwicknagel G, Toepffer C, Reinhard P-G. Time-dependent stopping power and influence of an infinite magnetic field. Physical Review E. 1998;57:3368
  31. 31. Walter M, Toepffer C, Zwicknagel G. Stopping power of ions in a magnetized two-temperature plasmas. Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms. 2000;168:347
  32. 32. Walter M. Ph.D. thesis. Ion stopping in a magnetized plasma. University of Erlangen; 2002
  33. 33. Steinberg M, Ortner J. Energy loss of a charged particle in magnetized quantum plasma. Physical Review E. 2001;63:046401
  34. 34. Deutsch C, Popoff R. Low ion-velocity slowing down in a strongly magnetized plasma target. Physical Review E. 2008;78:056405
  35. 35. Deutsch C, Popoff R. Low ion-velocity slowing down in a classical plasma. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment. 2009;606:212
  36. 36. Deutsch C, Zwicknagel G, Bret A. Ultracold plasmas: A paradigm for strongly coupled and classical electron fluids. Journal of Plasma Physics. 2009;75:799
  37. 37. Zwicknagel G. Theory and Simulation of the Interaction of Ions with Plasmas: Nonlinear Stopping, Ion-Ion Correlation Effects and Collisions of Ions with Magnetized Electrons, Habilitation Treatise. University of Erlangen; 2000. [http://www.opus.ub.uni-erlangen.de/opus/volltexte/2008/913/]
  38. 38. Zwicknagel G. Ion-electron collisions in a homogeneous magnetic field. In: Bollinger JB, Spencer RL, Davidson RC, editors. Non-Neutral Plasma Physics III. Vol. 498. AIP Conf. Proc.: 1999. p. 469
  39. 39. Zwicknagel G. Nonlinear energy loss of highly charged heavy ions. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment. 2000;441:44
  40. 40. Zwicknagel G, Toepffer C. Energy loss of ions by collisions with magnetized electrons. In: Anderegg F, Schweikhard L, Driscoll CF, editors. Non-Neutral Plasma Physics IV. Vol. 606. AIP Conf. Proc.; 2002. p. 499
  41. 41. Zwicknagel G. Trapped charged particles and fundamental interactions. In: Blaum K, Herfurth F, editors. Trapped Charged Particles and Fundamental Interactions, Lecture Notes in Physics. Vol. 749. Berlin: Springer-Verlag; 2008. pp. 69-96
  42. 42. Toepffer C. plaScattering of magnetized electrons by ions. Physical Review A. 2002;66:022714
  43. 43. Nersisyan HB, Zwicknagel G. Conformity between linear response and binary collision treatments of an ion energy loss in a magnetized quantum plasma. Journal of Physics A: Mathematical and General. 2004;37:11073
  44. 44. Nersisyan HB, Zwicknagel G. Binary collisions of charged particles in a magnetic field. Physical Review E. 2009;79:066405
  45. 45. Nersisyan HB, Zwicknagel G. Energy transfer in binary collisions of two gyrating charged particles in a magnetic field. Physics of Plasmas. 2010;17:082314
  46. 46. Deutsch C, Nersisyan HB, Bendib A. Diagnostics of arbitrary magnetized and dense plasmas of ICF/WDM concern. In: Proceedings of the 20th International Symposium on Heavy- Ion Inertial Fusion, HIF 2014, Lanzhou, China, 11–15 August, 2014. DOI: http://dx.doi.org/10.1017/S0263034615000919 in Laser Part. Beams 2015
  47. 47. Nersisyan HB, Zwicknagel G. Cooling force on ions in a magnetized electron plasma. Physical Review Accelerators and Beams. 2013;16:074201
  48. 48. Winkler T. Ph.D. thesis; Electron cooling of highly charged ions. University Heidelberg; 1996
  49. 49. Winkler T et al. Electron cooling force measurements for highly charged ions in the ESR. Hyperfine Interactions. 1996;99:277
  50. 50. Winkler T, Beckert K, Bosch F, Eickhoff H, Franzke B, Nolden F, Reich H, Schlitt B, Steck M. Electron cooling of highly chargd ions in the ESR. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment. 1997;391:12
  51. 51. Kelbg G. Theorie des quanten plasmas. Annalen der Physik (Berlin). 1963;467:219
  52. 52. Deutsch C. Nodal expansion in a real matter plasma. Physics Letters A. 1977;60:317
  53. 53. Zwicknagel G. Theory and simulation of heavy ion stopping in plasma. Laser and Particle Beams. 2009;27:399
  54. 54. Zwicknagel G. Nonlinear energy loss of heavy ions in plasmas. Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms. 2002;197:22
  55. 55. Zwicknagel G, Toepffer C, Reinhard P-G. Stopping of heavy ions in plasmas at strong coupling. Physics Reports. 1999;309:117; 1999;314:671(E)
  56. 56. Gradshteyn IS, Ryzhik IM. Tables of Integrals, Series and Products. 2nd ed. New York: Academic; 1980
  57. 57. Derbenev YS, Skrinsky AN. The effect of a magnetic field on electron cooling. Particle Accelerator. 1978;8:235
  58. 58. Parkhomchuk VV. In: Poth H, editor. Proceedings of the Workshop on Electron Cooling and Related Applications (ECOOL84, 1984). Karlsruhe: KfK; 1985. p. 71; KfK Report No. 3846
  59. 59. Pestrikov DV. Longitudinal cooling force due to magnetized electrons. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment. 2005;554:13
  60. 60. Peter T, Meyer-ter-Vehn J. Energy loss of heavy ions in dense plasmas 1. Linear and nonlinear Vlasov theory for the stopping power. Physical Review A. 1991;43:1998
  61. 61. Hahn H, Mason EA, Smith FJ. Quantum transport cross sections for ionized gases. Physics of Fluids. 1971;14:278
  62. 62. Chen S, Fuchs J. Ecole Polytechnique. Private Communication; July 2014
  63. 63. Tashev B, Baimbetov F, Deutsch C, Fromy P. Low velocity ion stopping in binary ionic mixtures. Physics of Plasmas. 2008;15:102701

Written By

Hrachya B. Nersisyan, Günter Zwicknagel and Claude Deutsch

Reviewed: 12 April 2018 Published: 27 February 2019