Observations on the ponderomotive force

Similar documents
One-dimensional Schrödinger equation

4. Complex Oscillations

Linear and Nonlinear Oscillators (Lecture 2)

20. Alfven waves. ([3], p ; [1], p ; Chen, Sec.4.18, p ) We have considered two types of waves in plasma:

Equations of linear stellar oscillations

Mass on a Horizontal Spring

Problem 1: Lagrangians and Conserved Quantities. Consider the following action for a particle of mass m moving in one dimension

Physics 115/242 Comparison of methods for integrating the simple harmonic oscillator.

Laser-plasma particle accelerators. Zulfikar Najmudin

Damped harmonic motion

221A Lecture Notes Convergence of Perturbation Theory

13.1 Ion Acoustic Soliton and Shock Wave

A plane autonomous system is a pair of simultaneous first-order differential equations,

Answers to Problem Set Number MIT (Fall 2005).

Reminders: Show your work! As appropriate, include references on your submitted version. Write legibly!

Seminar 6: COUPLED HARMONIC OSCILLATORS

Chapter 16 - Waves. I m surfing the giant life wave. -William Shatner. David J. Starling Penn State Hazleton PHYS 213. Chapter 16 - Waves

Path integrals and the classical approximation 1 D. E. Soper 2 University of Oregon 14 November 2011

LAB 11: FREE, DAMPED, AND FORCED OSCILLATIONS

Lab 11 - Free, Damped, and Forced Oscillations

x(t+ δt) - x(t) = slope δt t+δt

8 Wavefunctions - Schrödinger s Equation

= k, (2) p = h λ. x o = f1/2 o a. +vt (4)

Part III. Interaction with Single Electrons - Plane Wave Orbits

The Schroedinger equation

The Wave Function. Chapter The Harmonic Wave Function

HRW 7e Chapter 2 Page 1 of 13

Lecture 4. 1 de Broglie wavelength and Galilean transformations 1. 2 Phase and Group Velocities 4. 3 Choosing the wavefunction for a free particle 6

Oscillations. Simple Harmonic Motion of a Mass on a Spring The equation of motion for a mass m is attached to a spring of constant k is

Vibrations and waves: revision. Martin Dove Queen Mary University of London

ENGI 3424 First Order ODEs Page 1-01

Energy Level Sets for the Morse Potential

Thursday, August 4, 2011

1. Types of Waves. There are three main types of waves:

Physics 115/242 The leapfrog method and other symplectic algorithms for integrating Newton s laws of motion

Lecture 41: Highlights

Introduction. Chapter Plasma: definitions

Damped Harmonic Oscillator

Double Spring Harmonic Oscillator Lab

Studies on Rayleigh Equation

Differential Equations

Math 216 Final Exam 24 April, 2017

Particle-in-Cell Simulations of Particle Acceleration. E. A. Startsev and C. J. McKinstrie. Laboratory for Laser Energetics, U.

12. Acceleration and Normalized Emittance *

CMPT 889: Lecture 2 Sinusoids, Complex Exponentials, Spectrum Representation

Lab 10: Harmonic Motion and the Pendulum

Chapter 15 - Oscillations

Energetic particle modes: from bump on tail to tokamak plasmas

. (70.1) r r. / r. Substituting, we have the following equation for f:

LAB 11: FREE, DAMPED, AND FORCED OSCILLATIONS

Computational project: Modelling a simple quadrupole mass spectrometer

Lecture Notes to Accompany. Scientific Computing An Introductory Survey. by Michael T. Heath. Chapter 9

Mechanics IV: Oscillations

Traveling Harmonic Waves

Wave Phenomena Physics 15c. Lecture 11 Dispersion

In which we count degrees of freedom and find the normal modes of a mess o masses and springs, which is a lovely model of a solid.

Chapter 9 Uniform Circular Motion

Physics 351, Spring 2017, Homework #2. Due at start of class, Friday, January 27, 2017

Mechanical Waves. 3: Mechanical Waves (Chapter 16) Waves: Space and Time

Physics 486 Discussion 5 Piecewise Potentials

Coupled Oscillators Monday, 29 October 2012 normal mode Figure 1:

Relevant sections from AMATH 351 Course Notes (Wainwright): 1.3 Relevant sections from AMATH 351 Course Notes (Poulin and Ingalls): 1.1.

Solutions of a PT-symmetric Dimer with Constant Gain-loss

3 What You Should Know About Complex Numbers

Scientific Computing: An Introductory Survey

Continuum Limit and Fourier Series

L = 1 2 a(q) q2 V (q).

PHYSICS ADMISSIONS TEST Thursday, 2 November Time allowed: 2 hours

is conserved, calculating E both at θ = 0 and θ = π/2 we find that this happens for a value ω = ω given by: 2g

Damped Oscillation Solution

The Schrödinger Equation

Two-Body Problem. Central Potential. 1D Motion

Energy in Planetary Orbits

Physics 141 Energy 1 Page 1. Energy 1

Electric and Magnetic Forces in Lagrangian and Hamiltonian Formalism

Do not turn over until you are told to do so by the Invigilator.

Mechanics Lecture Notes

Robotics. Dynamics. Marc Toussaint U Stuttgart

FLUX OF VECTOR FIELD INTRODUCTION

Physics 106a, Caltech 4 December, Lecture 18: Examples on Rigid Body Dynamics. Rotating rectangle. Heavy symmetric top

Physics 115/242 The leapfrog method and other symplectic algorithms for integrating Newton s laws of motion

Weak interactions. Chapter 7

ENGI Partial Differentiation Page y f x

Solution Set Two. 1 Problem #1: Projectile Motion Cartesian Coordinates Polar Coordinates... 3

A Level. A Level Physics. Oscillations (Answers) AQA, Edexcel. Name: Total Marks: /30

CS 450 Numerical Analysis. Chapter 9: Initial Value Problems for Ordinary Differential Equations

1 (2n)! (-1)n (θ) 2n

Figure 1: Doing work on a block by pushing it across the floor.

06.lec Acceleration and Normalized Emittance *

General Exam Part II, Fall 1998 Quantum Mechanics Solutions

Chapter 4. Oscillatory Motion. 4.1 The Important Stuff Simple Harmonic Motion

Keble College - Hilary 2014 CP3&4: Mathematical methods I&II Tutorial 5 - Waves and normal modes II

Sinusoids. Amplitude and Magnitude. Phase and Period. CMPT 889: Lecture 2 Sinusoids, Complex Exponentials, Spectrum Representation

α(t) = ω 2 θ (t) κ I ω = g L L g T = 2π mgh rot com I rot

Chapter 15 Mechanical Waves

function independent dependent domain range graph of the function The Vertical Line Test

Donie O Brien Nigel Buttimore

Lagrangian and Hamiltonian Mechanics (Symon Chapter Nine)

Lecture 3 Dynamics 29

Methods of plasma description Particle motion in external fields

Transcription:

Observations on the ponderomotive force D.A. Burton a, R.A. Cairns b, B. Ersfeld c, A. Noble c, S. Yoffe c, and D.A. Jaroszynski c a University of Lancaster, Physics Department, Lancaster LA1 4YB, UK b University of St Andrews, School of Mathematics and Statistics, St Andrews KY16 9SS, UK c University of Strathclyde, Department of Physics, Glasgow G4 0NG, UK ABSTRACT The ponderomotive force is an important concept in plasma physics and, in particular, plays an important role in many aspects of the theory of laser plasma interactions including current concerns like wakefield acceleration and Raman amplification. The most familiar form of this gives a force on a charged particle that is proportional to the slowly varying gradient of the intensity of a high frequency electromagnetic field and directed down the intensity gradiant. For a field amplitude simply oscillating in time there is a simple derivation of this formula, but in the more general case of a travelling wave the problem is more difficult. Over the years there has been much work on this using Hamiltonian or Lagrangian averaging techniques, but little or no investigation of how well these theories work. Here we look at the very basic problem of a particle entering a region with a monotonically increasing electrostatic field amplitude and being reflected. We show that the equation of motion derived from a widely quoted ponderomotive potential only agrees with the numerically computed orbit within a restricted parameter range and that outside this range it shows features which are inconsistent with any ponderomotive potential quadratic in the field amplitude. Since the ponderomotive force plays a fundamental role in a variety of problems in plasma physics we think that it is important to point out that even in the simplest of configurations standard theories may not be accurate. Keywords: Laser plasma, ponderomotive force 1. INTRODUCTION The ponderomotive force, produced by a high frequency electromagnetic field whose intensity varies in space over a scale length long compared to the particle oscillation amplitude in the wave, is a fundamental concept in plasma physics that is widely used when considering laser-plasma interactions, but is by no means confined to this part of the subject. With an electric field of the form the familiar form of the ponderomotive force is E = E 0 (r) cos(ωt) (1) F = 1 4 mω 2 E2 0 (2) where q and m are the particle charge and mass respectively. This formula holds for both electrostatic and electromagnetic waves, but here we will be concerned only with longitudinal waves with the field and the spatial variation of its amplitude in the x direction, for which a simple derivation of the formula can be given as follows. Neglecting the spatial variation of the amplitude, the displacement of the particle from its mean position x 0 due to the wave is given by x x 0 = qe 0 cos(ωt) (3) mω2 and if we now take account of the spatial variation by making the approximation [ E(x, t) = E 0 (x 0 ) + (x x 0 ) de ] 0 cos(ωt), (4) dx

use (3) and take the time average we get (2). If we have a travelling wave with E = E 0 (x) cos(kx ωt) then the problem becomes more difficult since there is no simple solution for the particle displacement. One obvious way forward is to take the local solution x = x 0 + V t + ξ (5) with V an average velocity only varying over the long length scale, then neglect ξ in the argument of the cosine. This leads to q ξ = m(ω kv ) 2 E (6) and, if the procedure above is followed, an average force on the particle of 1 4 m(ω kv ) E0. 2 1 However, as 2 we shall see this is not a good approximation and later work concentrated on obtaining the force by more sophisticated Hamiltonian or Lagrangian methods. 2 5 These references are just a sample from a very extensive literature on the subject, from which we will concentrate particularly on the work of Bauer, Mulser and Steeb 4 since it gives an explicit formula relevant to our problem. The argument leading to their expression can be paraphrased as follows. With a displacement ξ, the resultant change in potential energy is -ξe and if we assume q ξ = m(ω kv ) 2 E 0 cos(kv t ωt) (7) E = E 0 cos(kv t ωt) (8) the average potential energy becomes 1 2 m(ω kv ) E 2 2 0 and the average kinetic energy due to the oscillatory motion ( is half of m qe 0 2. 2 m(ω kv )) The result is an average Lagrangian L = 1 2 mv 2 1 4 m(ω kv ) 2 E2 0. (9) The standard Lagrange equation is now used to obtain the variation over the long time and space scales, the result being m dv dt = q2 (ω kv )(ω 3kV ) de2 0 4m (ω kv ) 4 3 2 m k 2 E 2 dx. (10) 2 Clearly there are approximations involved in the calculation of the average potential and kinetic energy, the assumed particle displacement and velocity not being exact, so the question arises as to how well this formula works. Despite the fact that it is a simple matter to integrate the equation for the exact particle orbit in the field we describe, little or no work has, so far as we know, been done to verify the accuracy of the various expressions for the ponderomotive force that have appeared in the literature. We shall do this for the very basic case of a particle moving in an electrostatic travelling wave with a monotonic intensity profile and show that in certain parameter ranges the motion is not described by the above equation and, indeed, cannot be predicted from any averaged Lagrangian with a potential that is proportional to the square of the field intensity. This suggests that it is necessary, in these regimes, to consider ponderomotive potentials to higher order in the field amplitude, a problem we do not believe to have been considered in the previous literature. 2. NUMERICAL RESULTS Here we look at particle orbits described by the equation d 2 x dt 2 = q m E 0(x) cos(kx ωt) (11) with E 0 varying on a scale much longer than the amplitude of the rapid oscillations of the particle. For convenience we scale the variables with time in units of ω 1, E 0 (x) in terms of E 0 (0) and the length in units of X 0 = E 0 (0) q mω, 2 the oscillation amplitude of a particle with zero average velocity in the reference field. The scaled value of k is k = 2πX 0 λ (12)

with λ the wavelength. The scaled version of (11) is then d 2 x dt 2 = E 0(x) cos(kx t) (13) where, for convenience, we retain the same notation for the scaled variables. For the most part we take the scaled field to have an exponential dependence E 0 = exp( x a ) and unless otherwise stated a scaled value of a equal to 500. Our general conclusions are not sensitive to the exact profile used. We then look at the orbit of a particle starting with velocity 1 2 at large x, the behaviour of this according to the basic ponderomotive force equation (2) being reflection at x = 0 and outward motion just the reverse of the inward motion. If we take the force to be given simply by 1 4 m(ω kv ) E 2 2 0 in unscaled variables then there is a first integral 1 2 mω2 V 2 2 3 mωkv 3 + 1 4 mk2 V 4 + q2 4m E2 0 = const. (14) If this holds then it is clear that the asymptotic velocity in the outward direction is no longer the reverse of the initial inward velocity unless k = 0. However, numerical integration, for the scaled value of k below a critical value that we will discuss later, gives an exactly reversed velocity. This is given also by the equation (10), a result that follows from the fact that it is obtained from a time independent Lagrangian that has a constant of motion V L V L which tends to 1 2 mv 2 as E 0 0. For scaled values of k up to around 0.45 the formula (10) agrees well with the numerically averaged exact solution as can be seen from Figure 1. Figure 1. The average velocity from the numerical solution (red dots) and from the analytic formula (blue line) for k = 0.45 as a function of position. However, as we increase k the analytic formula runs into the problem that its denominator goes to zero and the equation of motion encounters a singularity. A simple indication of the point at which this happens can be found by noting that according to (10) dv dt 0 as V ω 3k, or 1 3k in scaled units, so if V starts below this value it can approach it but never cross it. On the other hand, by the argument already given, the asymptotic outward velocity must be 1 2 if the motion is described by a time independent Lagrangian. There is, therefore, a contradiction if the scaled value of k exceeds 2 3 0.47. For values of k above 2 3 there is a regime in which the exact solution still behaves smoothly with asymptotic velocity 1 2, suggesting that it can still be described by an averaged Lagrangian, but not that given by Bauer et al. There is a range in which, as shown in Figure 2, the averaged velocity decreases along part of the outward trajectory. The details of the averaging procedure are given in the Appendix.

Figure 2. The average velocity from the numerical solution for k = 0.6. Such behaviour can be given by (10), called uphill acceleration by Bauer et al., if the scaled V is between 1 3k and 1 k, but the average velocity can never cross from the uphill to downhill acceleration regimes in the way that the numerical solution does. In general, if we have any potential quadratic in the electric field, so that the Lagrangian is of the form, then the equation of motion is L = 1 2 mv 2 + F (V )E 0 (x) 2 (15) dv dt = V F F de0 2 m + F E0 2 dx. (16) With a monotonic intensity profile the acceleration can only change sign if the denominator changes sign, in which case the equation has a singularity, or if the numerator goes through zero. However, using the same argument already given for the Bauer et al. formula, the velocity can tend asymptotically to the value at which the numerator goes to zero if the intensity gradient is long enough, but never cross it. Another argument for the impossibility of a Lagrangian with a potential of this form is that the invariant V L V L is of the form 1 2 mv 2 + (V F F )E0. 2 For a given value of V this means that there must be a unique value of E0 2, while in our numerical solution the average velocity goes through the same value at different positions with different values of E0. 2 That there is still an invariant associated with some more complicated form of Lagrangian is suggested by comparing solutions with exponential and linear intensity dependence (the latter with E 0 = 1 x 1000 for x < 1000, zero otherwise) and plotting the relation between E 0 and V along the particle orbits, with the result shown in Figure 3. We are currently seeking to extend averaged Lagrangian theory to include effects of higher order in the field intensity.

Figure 3. The relation between the electric field amplitude and the average velocity for the exponential (red circles) and linear (blue triangles) profiles, both with k = 0.6. 3. CONCLUSIONS We have looked at a very basic problem concerning ponderomotive force, the orbit of a single particle being reflected by an electrostatic wave with a monotonic intensity profile, comparing the numerical solution of the equation of motion with the predictions of averaged Lagrangian theory. While we have concentrated on a particular theory of the ponderomotive force given by Bauer et al., mainly because it gives an explicit formula for the force relevant to our problem, the averaged ponderomotive potential is the same as that quoted by other authors using a variety of methods. Our work shows that for small enough values of the wavenumber the analytic theory and the numerical solution agree well, but that there are parameter ranges where the analytic formula encounters a singularity. In some cases the orbit of a particle goes from a region with acceleration down the wave intensity gradient to one where the acceleration is in the opposite direction. While the analytic theory due to Bauer et al. can give such uphill acceleration, we show that it cannot give rise to a transition between downhill and uphill acceleration along a single particle orbit. Indeed, we show that no Lagrangian with an average potential which is quadratic in the electric field amplitude can produce such a transition. This leads us to the conclusion that standard theories must, in some regimes, be extended to include higher order terms in the field amplitude, a problem on which we are currently working. The ponderomotive force is a fundamental idea, widely applied in plasma physics, so we believe that it is of interest to note that even in a very basic configuration there are circumstances in which widely used theoretical models do not apply. 4. APPENDIX: NUMERICAL METHODS The numerical calculations were carried out using Mathcad 15, but could easily be reproduced on Maple, Mathematica or any other mathematical software with ordinary differential equation solvers. Both Adams and adaptive Runge-Kutta routines were used and shown to give consistent results. In scaled units throughout, the particle was started from x = 4000 and followed for 10 4 time steps, sufficient to follow the particle through reflection and outwards to a sufficient distance for the electric field to be negligible and allowing us to test the symmetry between outward and inward asymptotic speeds. Only part of the orbit near the reflection point is plotted, since this is the region of most interest. Along the particle path, 10 5 data points were taken so that, if necessary, it could be followed on the scale of the fast oscillations. The average velocity was calculated from a simple average of the instantaneous velocities at 500 data points on either side of the point of interest. This made the residual effects of the fast oscillations small, while being over a small enough region to represent a local average. Changing the width of the averaging interval, but still within the range of a few hundred points on each side, made no discernable difference to the plots.

ACKNOWLEDGMENTS This work was supported by the UK Engineering and Physical Sciences Research Council grant EP/N028694/1 Lab in a bubble and the UK Science and Technology Facilities Council grant ST/G008248/1. All of the figures were produced using Mathcad 15 on a standard desktop computer, and the data in the figures can be entirely generated using the methods specified in the paper. REFERENCES [1] A. A. Vedenov, A. V. Gordeev and L. I. Rudakov, Plasma Phys. 9, 719 (1967) [2] J.R. Cary and A.N. Kaufman, Phys. Rev. Lett. 39, 402 (1977) [3] G.W. Kentwell, J. Plasma Phys. 34, 289 (1985) [4] D. Bauer, P. Mulser and W.H. Steeb, Phys. Rev. Lett. 75, 4622 (1995) [5] I.Y.Dodin and N.J.Fisch, Phys. Rev. E 77, 036402 (2008)