ˆd = 1 2π. d(t)e iωt dt. (1)

Similar documents
Electrodynamics of Radiation Processes

Bremsstrahlung. Rybicki & Lightman Chapter 5. Free-free Emission Braking Radiation

Bremsstrahlung Radiation

1 Monday, October 31: Relativistic Charged Particles

1. Why photons? 2. Photons in a vacuum

Chapter 2 Bremsstrahlung and Black Body

Basics of Radiation Fields

Compton Scattering. hω 1 = hω 0 / [ 1 + ( hω 0 /mc 2 )(1 cos θ) ]. (1) In terms of wavelength it s even easier: λ 1 λ 0 = λ c (1 cos θ) (2)

Thermal Bremsstrahlung

de = j ν dvdωdtdν. (1)

Neutrinos, nonzero rest mass particles, and production of high energy photons Particle interactions

The Stellar Opacity. F ν = D U = 1 3 vl n = 1 3. and that, when integrated over all energies,

Astrophysical Radiation Processes

CHAPTER 27. Continuum Emission Mechanisms

The next three lectures will address interactions of charged particles with matter. In today s lecture, we will talk about energy transfer through

φ(ν)dν = 1. (1) We can define an average intensity over this profile, J =

2. NOTES ON RADIATIVE TRANSFER The specific intensity I ν

Retarded Potentials and Radiation

Recap Lecture + Thomson Scattering. Thermal radiation Blackbody radiation Bremsstrahlung radiation

Addition of Opacities and Absorption

1. Why photons? 2. Photons in a vacuum

The Larmor Formula (Chapters 18-19)

Outline. Today we will learn what is thermal radiation

Let s consider nonrelativistic electrons. A given electron follows Newton s law. m v = ee. (2)

Thermal Equilibrium in Nebulae 1. For an ionized nebula under steady conditions, heating and cooling processes that in

Special relativity and light RL 4.1, 4.9, 5.4, (6.7)

Lecture 14 (11/1/06) Charged-Particle Interactions: Stopping Power, Collisions and Ionization

1 Monday, November 7: Synchrotron Radiation for Beginners

Compton Scattering I. 1 Introduction

Particle acceleration and generation of high-energy photons

Line Broadening. φ(ν) = Γ/4π 2 (ν ν 0 ) 2 + (Γ/4π) 2, (3) where now Γ = γ +2ν col includes contributions from both natural broadening and collisions.

Radiative Processes in Flares I: Bremsstrahlung

Radiation processes and mechanisms in astrophysics I. R Subrahmanyan Notes on ATA lectures at UWA, Perth 18 May 2009

Statistical Equilibria: Saha Equation

Radiation from Charged Particle Interaction with Matter

Electromagnetic Spectra. AST443, Lecture 13 Stanimir Metchev

Chapter 2 Radiation of an Accelerated Charge

Bethe-Block. Stopping power of positive muons in copper vs βγ = p/mc. The slight dependence on M at highest energies through T max

Relativistic corrections of energy terms

1 Fundamentals. 1.1 Overview. 1.2 Units: Physics 704 Spring 2018

In today s lecture, we want to see what happens when we hit the target.

1 Monday, November 21: Inverse Compton Scattering

Ay Fall 2004 Lecture 6 (given by Tony Travouillon)

Atomic cross sections

Physics Oct A Quantum Harmonic Oscillator

Instability and different burning regimes

Passage of particles through matter

High-Energy Astrophysics

Physics Dec The Maxwell Velocity Distribution

_ int (x) = e ψ (x) γμ ψ(x) Aμ(x)

Quantum Statistics (2)

Notes on x-ray scattering - M. Le Tacon, B. Keimer (06/2015)

Radiative processes from energetic particles II: Gyromagnetic radiation

Atomic Structure and Processes

GEORGE B. RYBICKI, ALAN P. LIGHTMAN. Copyright W Y-VCH Verlag GmbH L Co. KCaA

If light travels past a system faster than the time scale for which the system evolves then t I ν = 0 and we have then

Interstellar Medium Physics

Fundamental Interactions (Forces) of Nature

Radiative Processes in Astrophysics

Applied Nuclear Physics (Fall 2006) Lecture 19 (11/22/06) Gamma Interactions: Compton Scattering

Physics Nov Bose-Einstein Gases

3145 Topics in Theoretical Physics - radiation processes - Dr J Hatchell. Multiwavelength Milky Way

Radiation Processes. Black Body Radiation. Heino Falcke Radboud Universiteit Nijmegen. Contents:

Properties of Electromagnetic Radiation Chapter 5. What is light? What is a wave? Radiation carries information

1 Radiative transfer etc

Chapter 1: Useful definitions

Particle Interactions in Detectors

Plasma Processes. m v = ee. (2)

Particles and Waves Particles Waves

Interaction of Ionizing Radiation with Matter

3 Some Radiation Basics

University of Michigan Physics : Advanced Laboratory Notes on RADIOACTIVITY January 2007

Particle nature of light & Quantization

The Strange Physics of Nonabelian Plasmas

CHAPTER 26. Radiative Transfer

PHYS 352. Charged Particle Interactions with Matter. Intro: Cross Section. dn s. = F dω

The Nature of Light I: Electromagnetic Waves Spectra Kirchoff s Laws Temperature Blackbody radiation

Modal Analysis: What it is and is not Gerrit Visser

Final Exam - Solutions PHYS/ECE Fall 2011

Lecture: Scattering theory

Interaction of Particles and Matter

PHY-105: Nuclear Reactions in Stars (continued)

PHYSICS OF HOT DENSE PLASMAS

I. INTRODUCTION AND HISTORICAL PERSPECTIVE

Astrophysics Assignment; Kramers Opacity Law

Thermal radiation (a.k.a blackbody radiation) is the answer to the following simple question:

Some fundamentals. Statistical mechanics. The non-equilibrium ISM. = g u

Opacity and Optical Depth

Week 8: Stellar winds So far, we have been discussing stars as though they have constant masses throughout their lifetimes. On the other hand, toward

The Cosmic Microwave Background

Quantum physics. Anyone who is not shocked by the quantum theory has not understood it. Niels Bohr, Nobel Price in 1922 ( )

Units, limits, and symmetries

Physics 11b Lecture #24. Quantum Mechanics

PHYS 231 Lecture Notes Week 3

MITOCW watch?v=wr88_vzfcx4

Lecture 15 Notes: 07 / 26. The photoelectric effect and the particle nature of light

Radiative Processes in Astrophysics

Modern Physics notes Spring 2005 Paul Fendley Lecture 38

2. Passage of Radiation Through Matter

Synchrotron Radiation II

Transcription:

Bremsstrahlung Initial questions: How does the hot gas in galaxy clusters cool? What should we see in their inner portions, where the density is high? As in the last lecture, we re going to take a more in-depth look at a particular radiative process, in this case bremsstrahlung. The term means braking radiation, and occurs when an electron is accelerated by passage near an ion, and hence radiates. There is also an inverse process, which you get by running the film backward in time : a photon is absorbed by an electron that is moving in the Coulomb field of an ion. This is free-free absorption. Bremsstrahlung and free-free absorption are basic radiative processes that show up in manycontexts. Givenourlimitedtime, wewillonlybeabletotouchonafewoftheconcepts; read Rybicki and Lightman or Shu if you want more details. Let s make things easier for ourselves by starting with nonrelativistic bremsstrahlung. From our discussion of acceleration radiation, this means we ll use a dipole approximation. In such a case, this means the radiation is proportional to the second derivative of the dipole moment e i r i. Note that for interactions of identical particles (all electrons, or all protons, for example), the dipole moment is proportional to m i r i, which is the center of mass. Ask class: what can we conclude about bremsstrahlung from identical particles? There is no dipole radiation (although there can be quadrupole radiation), because the center of mass is stationary. We thus consider electron-ion bremsstrahlung. In principle, both are moving, so we have to include acceleration radiation from both electrons and ions. Ask class: is there a simplification that can help us out? Since ions are so much more massive than electrons, to a good approximation we can treat them as fixed, while the electrons move in their static Coulomb field. That s an approximation that appears over and over again in plasma physics. As with Compton scattering, we ll simplify things further by considering single-speed electrons at the start. As with lots of these basic processes, one can simply grind through with brute force, but in this case there are several important obstacles to overcome, so we ll go into this in more detail. First, let s assume that the electron is only slightly deflected from its path; that means that instead of having to compute the full trajectory, we can assume a straight line path and determine the radiation emitted as a result. See Figure 1 for the geometry of bremsstrahlung. Suppose the electron has charge e, and the ion has charge Ze. The impact parameter is b; that means that if the path were a perfect straight line, the closest the electron would come to the ion would be a distance b. The dipole moment is d = er, where R is the instantaneous location of the electron, so the second derivative is d = e v, where v is the electron velocity. Ask class: schematically, how would they compute the total energy

radiated over the electron s trajectory? More generally, however, we would like to figure out the energy radiated as a function of frequency. Here s a case where Fourier transforms can add some insight. Recall that the Fourier transform of a quantity, in this case d, is ˆd = 1 2π d(t)e iωt dt. (1) Thus the Fourier transform of d at some frequency ω is ω 2ˆd(ω) (think of taking the derivative inside the integral, where it acts on exp(iωt)). This gives ω 2ˆd(ω) = (e/2π) ve iωt dt. (2) In general this still might be a mess to evaluate, but let s get some insight by looking at extreme cases: large and small ω. We know that for most of the trajectory of the electron it s far from the ion, but it is close to its minimum distance over a time called the collision time, τ = b/v. That means the integral above is only really important for times between τ/2 and τ/2, give or take. Now, if ωτ 1, then the exponential oscillates many times, so the net contribution is small. If instead ωτ 1, then the exponent is close to 0 and thus the exponential is near unity. Thus, in our two limits: ˆd(ω) (e/2πω 2 ) v, ωτ 1 0, ωτ 1. (3) Here v is the change of velocity during the collision. Now let s think about how we could get this to a form where we have the energy emitted per frequency interval. We know that the energy per area per time is given by the Poynting flux dtda = c 4π E2 (t). (4) Thus the energy per area is this, integrated over time: /da = (c/4π) E2 (t)dt. Parseval s theorem says that the integral over all times of a quantity is related to the integral over all frequencies of the quantity s Fourier transform via E 2 (t)dt = 2π Ê(ω) 2 dω. (5) Here the vertical bars mean the magnitude of the complex quantity tells us that Ê(ω) 2 = Ê( ω) 2, so this gives /da = c 0 Ê(ω) 2 dω /dadω = c Ê(ω) 2. Ê(ω). E is real, which (6)

Back when we discussed acceleration and radiation, we found that E(t) = d(t) sinθ c 2 R 0 (7) for an angle Θ relative to the polarization direction and an average distance R 0 from the charges. Integrating over solid angle, and realizing that da = R 2 0dΩ, we get dω = 8πω4 3c 3 ˆd(ω) 2. (8) Plugging this into our bremsstrahlung formula, this means that the energy emitted during the whole encounter in a frequency interval dω is /dω (2e 2 /3πc 3 ) v 2, ωτ 1 0, ωτ 1. But how do we estimate v? Most of the change will be deflecting the electron, rather than speeding it up or slowing it down. That means we can consider just the change in velocity normal to the path: v = Ze2 m (9) bdt 2Ze2 = (b 2 +v 2 t 2 ) 3/2 mbv. (10) Putting it together, this means the energy radiated in a frequency interval dω, for electrons with impact parameter b, is (b)/dω = [8Z 2 e 6 /(3πc 3 m 2 v 2 b 2 )], = 0, b v/ω. b v/ω Here b v/ω is a restatement of ωτ 1. Now wait a second: isn t there a problem? This says that the power per frequency emitted for extremely small b diverges. Hm. Ask class: any ideas at this point what we can do? If not, let s plow on anyway. If we have lots of electrons (say, number density n e ) moving with a fixed speed v, then the flux of electrons (number per area per time) is n e v. The number per time that have an impact parameter between b and b + db is this flux times the area element, 2πbdb. If the number density of ions is n i, this means that the energy per frequency per volume per time is dωdvdt = n en i 2πv 0 (11) (b) bdb. (12) dω Yikes! Danger, Will Robinson! The integral in question is of 1/b, so we get a logarithm that diverges at both large b and small b. Ask class: what do we do? This is one manifestation of what we discussed last time: physically, power laws must be cut off at some limits. If, for example, we decide that the integral will really extend from b min to b max, we get dωdvdt = 16e6 3c 3 m 2 v n en i Z 2 ln(b max /b min ). (13)

You can do the usual checks: e = 0 implies no radiation, large m implies less acceleration and hence less radiation, and so on. You are in the presence of the famed Coulomb logarithm much spoken of in legend. This occurs a lot when Coulomb forces are involved, since the forcescalesas1/r 2 andtheareaelementscalesasr. Thefirstthingtonoteisthatlogarithms are blessedly insensitive to precise values; ln(10 16 ) is only twice ln(10 8 ), for example. That means that we don t have to be too precise in defining b max and b min to get a decent answer. But what is it that produces upper and lower cutoffs? Well, our approximation is invalid if b v/ω, so we can try b max = v/ω. What about the other limit? In our derivation we assumed that the electron was basically going in a straight line path. This will clearly be invalid when v v, which occurs when b = b min = 4Ze 2 /(πmv 2 ). Or, we could adopt a quantum approach: the classical calculation we ve done is invalid when b < h/mv, since that would imply x p <, where x b and p mv. Thus, we could also try b min = h/mv. Again, the insensitivity of the logarithm means we have latitude here. We will, of course, take the larger of the two possible values of b min. One reason we have to use these approximations is that we re trying to do a classical calculation of a quantum process. We can sort of adjust for that by writing the exact results in terms of a Gaunt factor g ff (v,ω): where dωdvdt = 16πe6 3 3c 3 m 2 v n en i Z 2 g ff (v,ω), (14) g ff (v,ω) = ( 3/π)ln(b max /b min ). (15) The Gaunt factor is normalized in this way so that typically its value is near unity. Note, by the way, that the expression has a 1/v in it. Ask class: does that mean that for a given frequency ω the power per volume diverges for electrons of arbitrarily low velocity? No! We have to realize that the energy of the photons comes from the kinetic energy of the electrons, so if 1 2 mv2 < hν then a photon of energy hν can t be created. There is thus another cutoff if you want the power at a particular frequency: only electrons in a certain velocity range contribute. This is called a photon discreteness effect. Now that we have an answer for a particular velocity, we can integrate over an electron velocity distribution to get bulk emission from a region. Consider a thermal distribution, meaning that electron velocities are apportioned according to a Maxwell-Boltzmann distribution. Rybicki and Lightman get the integrated emission for an electron temperature T, in erg s 1 cm 3 Hz 1 : ǫ ff ν = dvdtdν = 6.8 10 38 Z 2 n e n i T 1/2 e hν/kt ḡ ff, (16) where ḡ ff is the velocity-averaged Gaunt factor at temperature T.

Notethatthisisarather flat spectrumforhν < kt,andthatitrollsoverexponentially for higher energies (see Figure 2). That high-energy cutoff comes from photon discreteness and the lack of too many high-energy electrons. Ask class: given this spectrum, what is the rough frequency range where most of the energy will come out? Somewhere in the kt range. Note that if the electron distribution is nonthermal, one has to do different integrals. If we integrate this emission over all frequencies, we get a total emission dvdt = 1.4 10 27 T 1/2 n e n i Z 2 ḡ B (17) (in cgs units), where ḡ B is the frequency and velocity averaged Gaunt factor, usually in the range 1.1 to 1.5. Now, we ve discussed this in terms of an emission process. But there must therefore be a corresponding absorption process; in this case, free-free absorption. We can get a free-free absorption coefficient from Kirchoff s law, and we find that α ff ν = 3.7 10 8 cm 1 T 1/2 Z 2 n e n i ν 3( 1 e hν/kt) ḡ ff. (18) Ask class: why is there a 1 e hν/kt factor in there? It corrects for stimulated emission. The Rosseland mean of the free-free absorption coefficient is α ff R = 1.7 10 25 T 7/2 Z 2 n e n i ḡ R, (19) where quantities are measured in cgs units, and ḡ R is weighted as in the Rosseland mean and is of order unity. The particular form T 7/2 for the Rosseland mean, or ν 3 for the frequency dependence, is called a Kramers opacity and occurs for bound-free as well as free-free. The preceding was all done for nonrelativistic electrons. In certain very high-energy situations, though, the electron speeds can be relativistic. What is to be done? A cool way of approaching this problem is through the method of virtual quanta. First, transform into a frame in which the electron is stationary and the ion is moving. The moving ion produces a pulse of electric field, which Compton scatters off of the electron. The radiation that emerges is the bremsstrahlung radiation as seen in this new frame. You can then Lorentz transform back into the lab frame where the electron is moving to get the radiation rate. Stuff like this is common in quantum electrodynamics. One can often find symmetries between apparently unrelated processes. For example, consider Compton scattering. A photon hits an electron, and bounces off; this means that if one plots the interaction on a time axis, then a photon and electron converge, then diverge. It happens, though, that a positron can be treated as an electron moving backward in time (that s Feynman s insight). This means that if you look at Compton scattering sideways, two photons converge and produce an electron-positron pair, or the pair annihilates to produce two photons. Thus, there is an essential relation between Compton scattering and pair annihilation/production.

Enough of this diversion. When one puts in the relativistic effects, a reasonable fitting formula in cgs units is dvdt = 1.4 10 27 T 1/2 Z 2 n e n i ḡ B (1+4.4 10 10 T). (20) One final set of comments. We have now discussed in some more detail two processes: Compton scattering and bremsstrahlung. As observers, we are often presented with a spectrum and we d like to know the fundamental process that created it. In principle it may not sound too bad: look at the characteristics of the spectrum, then just identify it! In reality, though, if only continuum processes are operating (these vary only slowly with frequency), it can be difficult to do unique identification. Is your spectrum due to Comptonization, or is it the sum of many blackbodies of different temperatures? Be careful in these circumstances. Many, many people have a tendency to use the following approach: (1) assume some form or mechanism (e.g., bremsstrahlung), (2) derive parameters from the spectral fit (e.g., temperature or density), (3) assign great meaning to those derived parameters. The fact is that the derived parameters can vary significantly given different fits. A good fit does not guarantee physical meaning! If one has lines or edges it s usually easier, but beware of continuum fits. Recommended Rybicki and Lightman problem: 5.1

Fig. 1. Geometry of bremsstrahlung, from Wikipedia. An electron of initial energy E 1 and speed v 1 as seen in the reference frame of the positive ion is deflected by the ion, radiates a photon, and ends up with energy E 2 < E 1 and speed v 2 < v 1 as seen in the original frame. The inverse process, in which an electron absorbs a photon as it goes by an ion, is free-free absorption.

Fig. 2. Representation of the spectrum of bremsstrahlung, from http://www.astro.utu.fi/ cflynn/astroii/brems.gif. The energy per volume per time per frequency is nearly constant until hν > kt, after which it rolls over exponentially.