This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Adiabatic sheath model for beam-driven blowout plasma channels

Yulong Liu Institute of High Energy Physics, Chinese Academy of Sciences, Beijing 100049, China University of Chinese Academy of Sciences, Beijing 100049, China    Ming Zeng zengming@ihep.ac.cn Institute of High Energy Physics, Chinese Academy of Sciences, Beijing 100049, China University of Chinese Academy of Sciences, Beijing 100049, China    Lars Reichwein Institut für Theoretische Physik I, Heinrich-Heine-Universität Düsseldorf, D-40225 Düsseldorf, Germany Peter Grünberg Institut (PGI-6), Forschungszentrum Jülich, 52425 Jülich, Germany    Alexander Pukhov Institut für Theoretische Physik I, Heinrich-Heine-Universität Düsseldorf, D-40225 Düsseldorf, Germany
Abstract

In plasma wakefield accelerators, the structure of the blowout sheath is vital for the blowout radius and the electromagnetic field distribution inside the blowout. Previous theories assume artificial distribution functions for the sheath, which are either inaccurate or require prior knowledge of parameters. In this study, we develop an adiabatic sheath model based on force balancing, which leads to a self-consistent form of the sheath distribution. This model gives a better estimate of the blowout channel balancing radius than previous models.

Beam-driven plasma wakefield accelerators (PWFAs) have the capacity to accelerate particles to tens of GeV energies in meter-scale structures, making them competitive candidates for future high-energy accelerators that are used for next-generation particle colliders [1, 2, 3, 4, 5]. In a PWFA, when a high current, relativistic electron beam, called the drive beam, interacts with an underdense plasma, plasma electrons are completely radially expelled while the ions remain immobile due to their much larger mass, forming an electron-free ion channel along the propagation axis. The physics of plasma response in this regime is strongly nonlinear and the regime has been referred to as the blowout or bubble regime. In this regime, the acceleration (or deceleration) field is uniform and the focusing field is linear in regard to the radial offset from the axis [6, 7, 8, 9, 10].

On the boundary of the bubble, the expelled electrons form a narrow sheath, where the electron density and current profiles steepen with a large value and decay in a small thickness [11, 12, 13]. The sheath structure depends on the properties of the drive beam and is strongly correlated with the creation of the plasma blowout, the structure of wakefield inside the blowout, and the processes of particle injection into the wakefield [14, 15, 16].

Existing theories have used rectangular or exponential distributions to simplify the sheath shape [17, 18, 19, 20, 21, 22]. Dalichaouch et al. [23] proposed a multisheath model, which employs two sheath layers to describe the wakefield more accurately. However, these models require the prior knowledge of the sheath thickness, which cannot be obtained self-consistently. Recently, Golovanov et al. [24] have developed a blowout theory from the energy conservation point of view. They have assumed a δ\delta-function distribution for the blowout sheath, and found the evolution function for the blowout channel radius rδr_{\delta} as

A(rδ)rδd2rδdξ2+B(rδ)rδ2(drδdξ)2+Crδ2=Λ,A\left(r_{\delta}\right)r_{\delta}\frac{d^{2}r_{\delta}}{d\xi^{2}}+B\left(r_{\delta}\right)r^{2}_{\delta}\left(\frac{dr_{\delta}}{d\xi}\right)^{2}+Cr^{2}_{\delta}=\Lambda, (1)

where A=1+rδ2/4A=1+r_{\delta}^{2}/4, B=1/2+1/rδ2B=1/2+1/r_{\delta}^{2} and C=1/4C=1/4. ξ=ctz\xi=ct-z is the longitudinal co-moving coordinate, cc is the speed of light in vacuum, tt is time, and zz is the longitudinal coordinate. The drive term of the equation can be written as Λ=2I/IA\Lambda=2I/I_{A}, where II is the instant current of the beam, IA=4πϵ0mec3/e17kAI_{A}=4\pi\epsilon_{0}m_{e}c^{3}/e\approx 17\ \rm kA is the Alfvén current, ϵ0\epsilon_{0} is the vacuum permittivity, mem_{e} is the electron rest mass, and ee is the elementary charge [25, 26, 27]. If derivatives of rδr_{\delta} are assumed to be zero, one can obtain the balancing radius of the channel based on the δ\delta-sheath theory as

rδ0=2Λ.r_{\delta 0}=2\sqrt{\Lambda}. (2)

By contrast, the charge neutralization radius,

rn=2Λ,r_{n}=\sqrt{2\Lambda}, (3)

has been widely accepted as the balancing radius from the electrostatic point of view [28, 10, 26].

In this work, we develop a self-consistent model for the channel sheath using the adiabatic assumption. The sheath equations are examined in two cases with the help of particle-in-cell (PIC) simulations. We demonstrate that our model provides a more accurate description of blowout channel balancing, as its predicted channel radius falls between the electrostatic neutralization radius and that predicted by the δ\delta-sheath theory.

Throughout the paper, we adopt the plasma normalization units, in which charge is normalized to ee, mass to mem_{e}, velocity to cc, density to plasma density npn_{p}, time to ωp1\omega_{p}^{-1}, length to c/ωpc/\omega_{p}, electric field to mecωp/em_{e}c\omega_{p}/e, magnetic field to meωp/em_{e}\omega_{p}/e, electrostatic potential to mec2/em_{e}c^{2}/e and vector potential to mec/em_{e}c/e, where ωp=e2np/ϵ0me\omega_{p}=\sqrt{e^{2}n_{p}/\epsilon_{0}m_{e}} is the plasma frequency.

\begin{overpic}[width=208.13574pt]{plasma_density_and_psi.pdf} \put(33.3,27.55){ \leavevmode\hbox to147.11pt{\vbox to2pt{\pgfpicture\makeatletter\hbox{\hskip 1.0pt\lower-1.0pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{} {}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setdash{3.0pt,2.0pt}{0.0pt}\pgfsys@invoke{ }\color[rgb]{1,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{1,0,0}\pgfsys@color@rgb@stroke{1}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{1}{0}{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{1,0,0}\pgfsys@setlinewidth{2.0pt}\pgfsys@invoke{ }{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@lineto{145.10924pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}} \put(32.5,8.6){ \leavevmode\hbox to6pt{\vbox to91.91pt{\pgfpicture\makeatletter\hbox{\hskip 3.0pt\lower-1.0pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{} {}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setdash{3.0pt,2.0pt}{0.0pt}\pgfsys@invoke{ }\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\pgfsys@color@rgb@stroke{0}{0}{1}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{1}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{0,0,1}\pgfsys@setlinewidth{2.0pt}\pgfsys@invoke{ }{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@lineto{0.0pt}{89.91081pt}\pgfsys@stroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{{}}{}{{{}}{}{}{}{}{}{}{}{}}\pgfsys@beginscope\pgfsys@invoke{ }\color[rgb]{1,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{1,0,0}\pgfsys@color@rgb@stroke{1}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{1}{0}{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{1,0,0}\pgfsys@moveto{0.0pt}{46.23573pt}\pgfsys@moveto{3.0pt}{46.23573pt}\pgfsys@curveto{3.0pt}{47.89261pt}{1.65688pt}{49.23573pt}{0.0pt}{49.23573pt}\pgfsys@curveto{-1.65688pt}{49.23573pt}{-3.0pt}{47.89261pt}{-3.0pt}{46.23573pt}\pgfsys@curveto{-3.0pt}{44.57886pt}{-1.65688pt}{43.23573pt}{0.0pt}{43.23573pt}\pgfsys@curveto{1.65688pt}{43.23573pt}{3.0pt}{44.57886pt}{3.0pt}{46.23573pt}\pgfsys@closepath\pgfsys@moveto{0.0pt}{46.23573pt}\pgfsys@fill\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}} \end{overpic}
Figure 1: Illustration of the adiabatic sheath model in this work. The electron beam (orange color, beam head is not shown), with a linear density Λ(ξ)\Lambda\left(\xi\right), is moving to the left along the ξ\xi axis and drives the plasma wakefield in the blowout regime. The plasma electrons (gray region) are completely evacuated in the channel with the radius rc(ξ)r_{c}\left(\xi\right), leaving an immobile ion column (white region). As Λ\Lambda varies slowly with ξ\xi in our assumption, the transverse electromagnetic balance is achieved everywhere. The right two subplots show the transverse distribution of the pseudo-potential ψ\psi and the electron density nen_{e} at a certain ξ\xi denoted by the blue dashed line.

We consider a highly relativistic drive electron beam with density nb(ξ,r)n_{b}\left(\xi,r\right) that expels the plasma electrons to form an ion channel with a radius rc(ξ)r_{c}\left(\xi\right), where rr is the radial coordinate, and cylindrical symmetry is assumed. We assume that the drive beam is narrower than the channel, so that there is no drive charge outside the channel, or nb=0n_{b}=0 for r>rcr>r_{c}. Under these assumptions, the drive term can also be written as Λ(ξ)=0nbr𝑑r\Lambda\left(\xi\right)=\int_{0}^{\infty}n_{b}rdr, which has the physical meaning of the (normalized) linear density of the drive beam.

Furthermore, we assume the radius of the ion channel changes adiabatically, which means (drc/rc)/dξ1\left(dr_{c}/r_{c}\right)/d\xi\ll 1, (dΛ/Λ)/dξ1\left(d\Lambda/\Lambda\right)/d\xi\ll 1, and ξr\partial_{\xi}\ll\partial_{r} for all field variables, as illustrated in Fig. 1. Under the cylindrical symmetry and the quasi-static approximation [29, 30], the pseudo-potential of the wakefield obeys the following Poisson-like equation [31, 10]

1rr(rrψ)=S,-\frac{1}{r}\frac{\partial}{\partial r}\left(r\frac{\partial}{\partial r}\psi\right)=S, (4)

where ψ=φAz\psi=\varphi-A_{z} is the pseudo-potential, φ\varphi is the electrostatic potential, AzA_{z} is the longitudinal component of the vector potential,

S=ρJz=1ne(1vz)S=\rho-J_{z}=1-n_{e}\left(1-v_{z}\right) (5)

is the source term, ρ\rho is the charge density, JzJ_{z} is the longitudinal component of the current density, nen_{e} is the plasma electron density, and vzv_{z} is the longitudinal velocity of the plasma electrons. Because ne=0n_{e}=0 for r<rcr<r_{c}, we can write the form of the pseudo-potential inside the blowout

ψ|r<rc=ψc+rc24r24,\left.\psi\right|_{r<r_{c}}=\psi_{c}+\frac{r_{c}^{2}}{4}-\frac{r^{2}}{4}, (6)

where

ψc=ψ|r=rc>0\psi_{c}=\left.\psi\right|_{r=r_{c}}>0 (7)

is to be determined. It can be derived that ψc=0\psi_{c}=0 with the δ\delta-function distribution, i.e. S=1H(rrc)rcδ(rrc)/2S=1-H\left(r-r_{c}\right)-r_{c}\delta\left(r-r_{c}\right)/2, where HH is the Heaviside step function, and δ\delta is the Dirac delta function. But in this work, SS is finite everywhere, so that ψc\psi_{c} is non-zero, and ψ/r\partial\psi/\partial r is continuous everywhere because of Eq. (4), which means

rψ|r=rc=rc2.\left.\frac{\partial}{\partial r}\psi\right|_{r=r_{c}}=-\frac{r_{c}}{2}. (8)

For a plasma electron which is at rest before the drive beam arrives, there is a constant of motion [30]

γvzγψ=1,\gamma-v_{z}\gamma-\psi=1, (9)

where γ=1/1vz2vr2\gamma=1/\sqrt{1-v_{z}^{2}-v_{r}^{2}} is the Lorentz factor, vzv_{z} and vrv_{r} are the longitudinal and radial components of the velocity of the electron, respectively. This equation can be solved under the adiabatic assumption, which means vr0v_{r}\approx 0, as

vz=21+(1+ψ)21.v_{z}=\frac{2}{1+\left(1+\psi\right)^{2}}-1. (10)

Based on the adiabatic assumption, the transverse electromagnetic field can be written using the Gauss’ Law and Stokes’ Theorem

Er\displaystyle E_{r} =\displaystyle= r2Λr1r0rne(r)r𝑑r,\displaystyle\frac{r}{2}-\frac{\Lambda}{r}-\frac{1}{r}\int_{0}^{r}n_{e}\left(r^{\prime}\right)r^{\prime}dr^{\prime}, (11)
Bθ\displaystyle B_{\theta} =\displaystyle= Λr1r0rne(r)vz(r)r𝑑r.\displaystyle-\frac{\Lambda}{r}-\frac{1}{r}\int_{0}^{r}n_{e}\left(r^{\prime}\right)v_{z}\left(r^{\prime}\right)r^{\prime}dr^{\prime}. (12)

The transverse force Fr=Er+vzBθF_{r}=-E_{r}+v_{z}B_{\theta} should be balanced for r>rcr>r_{c}, which means

0=rFr=\displaystyle 0=rF_{r}= r22+(1vz)Λ+0rne(r)r𝑑r\displaystyle-\frac{r^{2}}{2}+\left(1-v_{z}\right)\Lambda+\int_{0}^{r}n_{e}\left(r^{\prime}\right)r^{\prime}dr^{\prime} (13)
vz0rne(r)vz(r)r𝑑r.\displaystyle-v_{z}\int_{0}^{r}n_{e}\left(r^{\prime}\right)v_{z}\left(r^{\prime}\right)r^{\prime}dr^{\prime}.

The channel radius is determined by letting r=rcr=r_{c} in Eq. (13), as

rc=2(1vzc)Λ=2Λ1+1(1+ψc)2,r_{c}=\sqrt{2\left(1-v_{zc}\right)\Lambda}=2\sqrt{\frac{\Lambda}{1+\frac{1}{\left(1+\psi_{c}\right)^{2}}}}, (14)

where

vzc=vz|r=rc=21+(1+ψc)21.v_{zc}=\left.v_{z}\right|_{r=r_{c}}=\frac{2}{1+\left(1+\psi_{c}\right)^{2}}-1. (15)

We see that rcr_{c} reduces to rnr_{n} if ψc=0\psi_{c}=0, and to rδ0r_{\delta 0} if ψc\psi_{c}\rightarrow\infty. Then, we know rn<rc<rδ0r_{n}<r_{c}<r_{\delta 0} if ψc\psi_{c} is non-zero and finite, because rcr_{c} monotonically increases with ψc\psi_{c}.

The sheath equations defined by Eqs. (4), (10) and (13) are solvable using the left boundary condition Eq. (8) and the right boundary condition

limrψ=0,\lim_{r\to\infty}\psi=0, (16)

which is a natural requirement that the plasma disturbance is local.

We use the shooting method to numerically solve the equations. In other words, the numerical integral is performed from r=rcr=r_{c} to a sufficient large value of rr, while ψc\psi_{c} is predicted and corrected in iterations, so that Eq. (16) is satisfied to a certain accuracy. Practically, there is a critical value of ψc\psi_{c} for each rcr_{c}, which prevents divergence at large rr. The Python script for the numerical solution is attached as a Supplementary Material, and is maintained on GitHub [32]. The numerical solution of ψc\psi_{c} with different rcr_{c} is shown in Fig. 2. The polynomial fit,

ψc0.012rc2+0.363rc0.044,\psi_{c}\approx-0.012r_{c}^{2}+0.363r_{c}-0.044, (17)

is a good estimate for rc8r_{c}\lesssim 8 as shown in Fig. 2(a).

After ψc\psi_{c} is obtained, the electron density at the sheath boundary can be predicted. By taking derivative of Eq. (13), we obtain

ner(1vz2)=r+Λvzr+vzr0rne(r)vz(r)r𝑑r.n_{e}r\left(1-v_{z}^{2}\right)=r+\Lambda\frac{\partial v_{z}}{\partial r}+\frac{\partial v_{z}}{\partial r}\int_{0}^{r}n_{e}\left(r^{\prime}\right)v_{z}\left(r^{\prime}\right)r^{\prime}dr^{\prime}. (18)

At r=rcr=r_{c}, the integral is zero, the derivative of vzv_{z} can be determined by Eqs. (8) and (10), and Λ\Lambda can be written as the function of rcr_{c} and ψc\psi_{c} according to Eq. (14). As a result,

ne|r=rc=rc28(1ψ+1ψ3)+14(ψ+1ψ)2,\left.n_{e}\right|_{r=r_{c}}=\frac{r_{c}^{2}}{8}\left(\frac{1}{\psi_{*}}+\frac{1}{\psi_{*}^{3}}\right)+\frac{1}{4}\left(\psi_{*}+\frac{1}{\psi_{*}}\right)^{2}, (19)

where ψ=1+ψc\psi_{*}=1+\psi_{c}.

Next, we study the behavior of ψ\psi and ψc\psi_{c} in the limit ψc1\psi_{c}\ll 1. This also means rc1r_{c}\ll 1, Λ1\Lambda\ll 1, and ψ1\psi\ll 1. By expanding Eqs. (10) and (18) and keeping only the 1st order terms of ψ\psi, we have vzψv_{z}\approx-\psi and ne1n_{e}\approx 1. Thus, the linear form of Eq. (4) is obtained

2r2ψ+1rrψψ=0.\frac{\partial^{2}}{\partial r^{2}}\psi+\frac{1}{r}\frac{\partial}{\partial r}\psi-\psi=0. (20)

The solution satisfying the right boundary condition Eq. (16) is ψ=R(rc)K0(r)\psi=R\left(r_{c}\right)K_{0}\left(r\right), where R(rc)R\left(r_{c}\right) is a constant of rr to be determined, and K0(r)K_{0}\left(r\right) is the modified Bessel function of the second kind of order zero [33]. By also considering the left boundary condition Eq. (8), we may get the expression of R(rc)R\left(r_{c}\right) to the lowest order of rcr_{c}, and find

ψ|r>rc=rc22K0(r).\left.\psi\right|_{r>r_{c}}=\dfrac{r_{c}^{2}}{2}K_{0}\left(r\right). (21)

Thus,

ψc=rc22K0(rc),\psi_{c}=\dfrac{r_{c}^{2}}{2}K_{0}\left(r_{c}\right), (22)

or ψcrc22(lnrc2+0.577)\psi_{c}\approx-\frac{r_{c}^{2}}{2}\left(\ln\frac{r_{c}}{2}+0.577...\right), where the second summand is the Euler-Mascheroni constant. We can see in Fig. 2(b) that Eq. (22) is satisfactory with the numerical results for rc0.28r_{c}\lesssim 0.28, while Eq. (17) is better for rc0.28r_{c}\gtrsim 0.28.

\begin{overpic}[width=209.4392pt]{psic_vs_rc_compact.pdf} \put(39.0,13.0){(a)} \put(90.0,13.0){(b)} \end{overpic}
Figure 2: Comparison of ψc\psi_{c} vs rcr_{c} obtained by the numerical integral of Eqs. (4), (10) and (13) using the boundary conditions Eqs. (8) and (16) (black squares), by the polynomial fit Eq. (17) (green line), by the linear theory Eq. (22) (yellow line), by PIC simulations in near-adiabatic cases (red dots) and in near-stationary cases (blue triangles). Both the large (a) and small (b) scales of rcr_{c} are plotted, in order to show the different asymptotic behavior of ψc\psi_{c} at the two ends.
\begin{overpic}[width=208.13574pt]{radius_theory_compare.pdf} \put(12.0,47.0){(a)} \put(12.0,33.0){(b)} \put(63.0,47.0){(c)} \put(63.0,33.0){(d)} \end{overpic}
Figure 3: The distribution of Λ\Lambda and slice plots of plasma electron density at y=0y=0 (gray) in simulations for (a) and (b) a near-adiabatic case, and (c) and (d) a near-stationary case. The numerical solution of the balancing radius based on the δ\delta-sheath theory rδr_{\delta} (yellow dashed line) and its balance rδ0r_{\delta 0} (red line), the neutralization radius rnr_{n} (blue line), the solution based on our theory rcr_{c} (orange dashed line), the numerical solution based on Lu’s model with β=1/rc+0.1\beta=1/r_{c}+0.1 (cyan dashed line) and the combined model with β\beta in Lu’s model replaced by Eq. (23) (violet dotted line) are plotted for the comparison.

Once ψc\psi_{c} is determined, the one-to-one mapping between Λ\Lambda and rcr_{c} can be obtained according to Eq. (14). For comparison, PIC simulations have been performed using the quasi-static code QuickPIC [34, 35]. In the simulations, we consider a 5 GeV drive electron beam, transversely positioned at x=y=0x=y=0 (i.e. the center of the simulation domain). It has a sufficiently small transverse size, while its longitudinal size is the same as the entire simulation domain. Two sets of simulations with different drive beam current profiles have been conducted. One is called the near-adiabatic profile, where Λ\Lambda increases linearly with a sufficiently small slope. The other is called the near-stationary profile, where Λ\Lambda increases from 0 to a plateau, and the increasing profile and length are designed so that the oscillation of the channel radius in the plateau is not severe.

In the near-adiabatic simulations, we have chosen the maximum Λmax=4\Lambda_{\max}=4 for the large rcr_{c}-scale shown in Fig. 2(a) and in Fig. 3(a), and Λmax=0.4\Lambda_{\max}=0.4 for the small rcr_{c}-scale shown in Fig. 2(b). The simulation domain has a size of 200 in ξ\xi-direction, which is the propagation direction, and 40 (10) in xx- and yy-direction for the Λmax=4\Lambda_{\max}=4 (0.4) case. The numbers of cells is chosen as 2048, 1024 and 1024, respectively. The near-adiabatic simulation results for ψc\psi_{c} vs. rcr_{c} are shown in Fig. 2, which are in good agreements with our model.

In the near-stationary simulations, Λ\Lambda increases from 0 to the plateau value Λ0\Lambda_{0} within 0ξ250\leq\xi\leq 25, and keeps the plateau value in the remaining of the simulation domain, as shown in Fig. 3(c). Cubic splines have been designed to smooth the transitions near ξ=0\xi=0 and 25. Here, the simulation domain has a longitudinal size of 48, while the transverse size is adjusted in accordance with Λ0\Lambda_{0}, such that ψ0\psi\rightarrow 0 at the simulation boundary is guaranteed (ψ/r0\partial\psi/\partial r\approx 0 at the boundary), and the transverse resolution is adequate to resolve the sheath. Again, the number of cells is chosen to be 2048×1024×10242048\times 1024\times 1024. Λ0\Lambda_{0} has been scanned from 2.5×1052.5\times 10^{-5} to 6.25 in the simulations. The results of ψc\psi_{c} vs. rcr_{c} are shown in Fig. 2, which are also in good agreements with our model.

Plasma density slices at y=0y=0 for two of the above mentioned simulations are shown in Fig. 3(b) and 3(d). We compare the simulations with rδ0r_{\delta 0} obtained by Eq. (2), rnr_{n} obtained by Eq. (3) and rcr_{c} obtained by Eq. (14). We see that the channel radius in simulations best matches with rcr_{c}, while rnr_{n} is an underestimate and rδ0r_{\delta 0} is an overestimate.

Furthermore, the channel radius slightly oscillates around rcr_{c} in the near-stationary case because of the imperfect balancing, as shown in Fig. 3(d). This oscillation behavior can be numerically solved using Lu’s model [10, 17], which is similar to Eq. (1), but the parameters are A=1+[1/4+β/2+(rc/8)(dβ/drc)]rc2A=1+\left[1/4+\beta/2+\left(r_{c}/8\right)\left(d\beta/dr_{c}\right)\right]r_{c}^{2}, B=1/2+(3/4)β+(3rc/4)(dβ/drc)+(rc2/8)(d2β/drc2)B=1/2+\left(3/4\right)\beta+\left(3r_{c}/4\right)\left(d\beta/dr_{c}\right)+\left(r_{c}^{2}/8\right)\left(d^{2}\beta/dr_{c}^{2}\right) and C=[1+1/(1+βrc2/4)2]/4C=\left[1+1/\left(1+\beta r_{c}^{2}/4\right)^{2}\right]/4. The variable β\beta needs prior knowledge to the sheath thickness, and was recommended to be β=1/rc+0.1\beta=1/r_{c}+0.1. Although Lu’s model is in good agreement with the simulation in the short drive cases (drive beam is shorter than or comparable to the plasma wavelength), it does not predict the oscillation of rcr_{c} very well in the long drive cases as shown in Fig. 3(d).

To improve Lu’s model, we replace β\beta by

β=4ψcrc2,\beta=\frac{4\psi_{c}}{r_{c}^{2}}, (23)

where ψc\psi_{c} is expressed in a piecewise manner by Eq. (22) for rc<0.28r_{c}<0.28 and by Eq. (17) for rc>0.28r_{c}>0.28, and call it the combined model. The combined model has been applied to the near-stationary cases, and its numerical solutions have the best agreement with the simulations compared with other models, with one example shown as the violet dotted line in Fig. 3(d).

In conclusion, we have introduced a self-consistent model for the sheath structure of the blowout plasma channel driven by an electron beam based on the adiabatic assumption. The model consists of the equation for the pseudo-potential Eq. (4), the longitudinal velocity of the sheath electrons Eq. (10), and the transverse force balancing condition Eq. (13), which is solvable with the boundary conditions Eqs. (8) and (16). An analytical solution is obtained in the small-blowout limit as Eq. (21), and a general solution is retrieved numerically by the shooting method. The PIC simulations show that the radius obtained by Eq. (14) is a better estimate of the channel balancing radius than that from previous neutralization or δ\delta-sheath models.

It should be noted that the accuracy of our model is limited in short drive beam cases. Nevertheless, our model is very suitable for long drive beams. Especially, by combining our model with Lu’s model, we have obtained the most correct description of the oscillation of the channel radius. The better understanding of the pseudo-potential at the sheath boundary and inside the blowout for long drive cases is crucial for high transformer ratio and future high-energy PWFA studies [5, 36, 37].

Acknowledgements.
This work is supported by the Strategic Priority Research Program of the Chinese Academy of Sciences (Grant No. XDB0530000), the National Natural Science Foundation of China (Grant No. 12475159), and the BMBF project 05P24PF2 (Germany).

References

  • Chen et al. [1985] P. Chen, J. M. Dawson, R. W. Huff, and T. Katsouleas, Acceleration of electrons by the interaction of a bunched electron beam with a plasma, Phys. Rev. Lett. 55, 1537 (1985).
  • Blumenfeld et al. [2007] I. Blumenfeld, C. E. Clayton, F.-J. Decker, M. J. Hogan, C. Huang, R. Ischebeck, R. Iverson, C. Joshi, T. Katsouleas, N. Kirby, W. Lu, K. A. Marsh, W. B. Mori, P. Muggli, E. Oz, R. H. Siemann, D. Walz, and M. Zhou, Energy doubling of 42 gev electrons in a metre-scale plasma wakefield accelerator, Nature 44510.1038/nature05538 (2007).
  • Adolphsen and Hogan [2022] C. Adolphsen and M. Hogan, European strategy for particle physics – accelerator r&d roadmap, arXiv:2201.07895  (2022).
  • Foster et al. [2023] B. Foster, R. D’Arcy, and C. A. Lindstrøm, A hybrid, asymmetric, linear higgs factory based on plasma-wakefield and radio-frequency acceleration, New Journal of Physics 25, 093037 (2023).
  • Gao et al. [2024] J. Gao et al., Cepc technical design report: Accelerator, Radiation Detection Technology and Methods 810.1007/s41605-024-00463-y (2024).
  • Katsouleas [1986] T. Katsouleas, Physical mechanisms in the plasma wake-field accelerator, Phys. Rev. A 33, 2056 (1986).
  • Su et al. [1987] J. J. Su, T. Katsouleas, J. M. Dawson, P. Chen, M. Jones, and R. Keinigs, Stability of the driving bunch in the plasma wakefield accelerator, IEEE Transactions on Plasma Science 15, 192 (1987).
  • Rosenzweig et al. [1991] J. B. Rosenzweig, B. Breizman, T. Katsouleas, and J. J. Su, Acceleration and focusing of electrons in two-dimensional nonlinear plasma wake fields, Phys. Rev. A 44, R6189 (1991).
  • Barov and Rosenzweig [1994] N. Barov and J. B. Rosenzweig, Propagation of short electron pulses in underdense plasmas, Phys. Rev. E 49, 4407 (1994).
  • Lu et al. [2006a] W. Lu, C. Huang, M. Zhou, M. Tzoufras, F. S. Tsung, W. B. Mori, and T. Katsouleas, A nonlinear theory for multidimensional relativistic plasma wave wakefields, Physics of Plasmas 13, 056709 (2006a).
  • Rosenzweig [1987] J. B. Rosenzweig, Nonlinear plasma dynamics in the plasma wake-field accelerator, Phys. Rev. Lett. 58, 555 (1987).
  • Rosenzweig et al. [2004] J. B. Rosenzweig, N. Barov, M. C. Thompson, and R. B. Yoder, Energy loss of a high charge bunched electron beam in plasma: Simulations, scaling, and accelerating wakefields, Phys. Rev. ST Accel. Beams 7, 061302 (2004).
  • Kostyukov et al. [2004] I. Kostyukov, A. Pukhov, and S. Kiselev, Phenomenological theory of laser-plasma interaction in “bubble” regime, Physics of Plasmas 11, 5256 (2004).
  • Tzoufras et al. [2008] M. Tzoufras, W. Lu, F. S. Tsung, C. Huang, W. B. Mori, T. Katsouleas, J. Vieira, R. A. Fonseca, and L. O. Silva, Beam loading in the nonlinear regime of plasma-based acceleration, Phys. Rev. Lett. 101, 145002 (2008).
  • Kostyukov et al. [2009] I. Kostyukov, E. Nerush, A. Pukhov, and V. Seredov, Electron self-injection in multidimensional relativistic-plasma wake fields, Phys. Rev. Lett. 103, 175003 (2009).
  • Golovanov et al. [2016a] A. A. Golovanov, I. Y. Kostyukov, J. Thomas, and A. Pukhov, Beam loading in the bubble regime in plasmas with hollow channels, Physics of Plasmas 23, 093114 (2016a).
  • Lu et al. [2006b] W. Lu, C. Huang, M. Zhou, W. B. Mori, and T. Katsouleas, Nonlinear theory for relativistic plasma wakefields in the blowout regime, Phys. Rev. Lett. 96, 165002 (2006b).
  • Yi et al. [2013] S. A. Yi, V. Khudik, C. Siemon, and G. Shvets, Analytic model of electromagnetic fields around a plasma bubble in the blow-out regime, Physics of Plasmas 20, 013108 (2013).
  • Thomas et al. [2016] J. Thomas, I. Y. Kostyukov, J. Pronold, A. Golovanov, and A. Pukhov, Non-linear theory of a cavitated plasma wake in a plasma channel for special applications and control, Physics of Plasmas 23, 053108 (2016).
  • Golovanov et al. [2016b] A. Golovanov, I. Kostyukov, A. Pukhov, and J. Thomas, Generalised model of a sheath of a plasma bubble excited by a short laser pulse or by a relativistic electron bunch in transversely inhomogeneous plasma, Quantum Electronics 46, 295 (2016b).
  • Golovanov et al. [2017] A. A. Golovanov, I. Y. Kostyukov, J. Thomas, and A. Pukhov, Analytic model for electromagnetic fields in the bubble regime of plasma wakefield in non-uniform plasmas, Physics of Plasmas 24, 103104 (2017).
  • Golovanov et al. [2021] A. A. Golovanov, I. Y. Kostyukov, L. Reichwein, J. Thomas, and A. Pukhov, Excitation of strongly nonlinear plasma wakefield by electron bunches, Plasma Physics and Controlled Fusion 63, 085004 (2021).
  • Dalichaouch et al. [2021] T. N. Dalichaouch, X. L. Xu, A. Tableman, F. Li, F. S. Tsung, and W. B. Mori, A multi-sheath model for highly nonlinear plasma wakefields, Physics of Plasmas 28, 063103 (2021).
  • Golovanov et al. [2023] A. Golovanov, I. Y. Kostyukov, A. Pukhov, and V. Malka, Energy-conserving theory of the blowout regime of plasma wakefield, Phys. Rev. Lett. 130, 105001 (2023).
  • Whittum et al. [1991a] D. H. Whittum, A. M. Sessler, and V. K. Neil, Transverse resistive wall instability in the two-beam accelerator, Phys. Rev. A 43, 294 (1991a).
  • Huang et al. [2007] C. Huang, W. Lu, M. Zhou, C. E. Clayton, C. Joshi, W. B. Mori, P. Muggli, S. Deng, E. Oz, T. Katsouleas, M. J. Hogan, I. Blumenfeld, F. J. Decker, R. Ischebeck, R. H. Iverson, N. A. Kirby, and D. Walz, Hosing instability in the blow-out regime for plasma-wakefield acceleration, Phys. Rev. Lett. 99, 255001 (2007).
  • Martinez de la Ossa et al. [2015] A. Martinez de la Ossa, T. J. Mehrling, L. Schaper, M. J. V. Streeter, and J. Osterhoff, Wakefield-induced ionization injection in beam-driven plasma accelerators, Physics of Plasmas 22, 093107 (2015).
  • Geraci and Whittum [2000] A. A. Geraci and D. H. Whittum, Transverse dynamics of a relativistic electron beam in an underdense plasma channel, Physics of Plasmas 7, 3431 (2000).
  • Whittum [1997] D. H. Whittum, Transverse two-stream instability of a beam with a Bennett profile, Physics of Plasmas 4, 1154 (1997).
  • Mora and Antonsen [1997] P. Mora and T. M. Antonsen, Jr., Kinetic modeling of intense, short laser pulses propagating in tenuous plasmas, Physics of Plasmas 4, 217 (1997).
  • Whittum et al. [1991b] D. H. Whittum, W. M. Sharp, S. S. Yu, M. Lampe, and G. Joyce, Electron-hose instability in the ion-focused regime, Phys. Rev. Lett. 67, 991 (1991b).
  • Zeng [2024] M. Zeng, https://github.com/mingzeng8/adiabatic_sheath_solver (2024), the git repository for the numerical solver of the adiabatic sheath model.
  • Jeffrey and Dai [2008] A. Jeffrey and H.-H. Dai, Handbook of Mathematical Formulas and integrals (Academic Press, Elsevier, 2008) 4th edition.
  • Huang et al. [2006] C. Huang, V. K. Decyk, C. Ren, M. Zhou, W. Lu, W. B. Mori, J. H. Cooley, T. M. Antonsen, and T. C. Katsouleas, Quickpic: A highly efficient particle-in-cell code for modeling wakefield acceleration in plasmas, J. Comput. Phys. 217, 658 (2006).
  • An et al. [2013] W. An, V. K. Decyk, W. B. Mori, and T. M. Antonsen, An improved iteration loop for the three dimensional quasi-static particle-in-cell algorithm: Quickpic, Journal of Computational Physics 250, 165 (2013).
  • Lu et al. [2010] W. Lu, W. An, C. Huang, C. Joshi, W. Mori, M. Hogan, T. Raubenheimer, and A. Seryi, High Transformer Ratio PWFA for Application on XFELs, in Particle Accelerator Conference (PAC 09) (2010) p. WE6RFP098.
  • Loisch et al. [2018] G. Loisch, G. Asova, P. Boonpornprasert, R. Brinkmann, Y. Chen, J. Engel, J. Good, M. Gross, F. Grüner, H. Huck, D. Kalantaryan, M. Krasilnikov, O. Lishilin, A. M. de la Ossa, T. J. Mehrling, D. Melkumyan, A. Oppelt, J. Osterhoff, H. Qian, Y. Renier, F. Stephan, C. Tenholt, V. Wohlfarth, and Q. Zhao, Observation of high transformer ratio plasma wakefield acceleration, Phys. Rev. Lett. 121, 064801 (2018).