de = j ν dvdωdtdν. (1)

Similar documents
1. Why photons? 2. Photons in a vacuum

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

1 Radiative transfer etc

2. NOTES ON RADIATIVE TRANSFER The specific intensity I ν

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

I ν. di ν. = α ν. = (ndads) σ ν da α ν. = nσ ν = ρκ ν

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

Chapter 2 Bremsstrahlung and Black Body

Outline. Today we will learn what is thermal radiation

Lecture 3: Emission and absorption

5. Light-matter interactions: Blackbody radiation

The term "black body" was introduced by Gustav Kirchhoff in The light emitted by a black body is called black-body radiation.

CHAPTER 26. Radiative Transfer

5. Light-matter interactions: Blackbody radiation

Spectroscopy Lecture 2

3 Some Radiation Basics

Interstellar Medium Physics

Electromagnetic Radiation.

Opacity and Optical Depth

Lecture 3: Specific Intensity, Flux and Optical Depth

Radiation in the Earth's Atmosphere. Part 1: Absorption and Emission by Atmospheric Gases

Sources of radiation

[09] Light and Matter (9/26/17)

Theory of optically thin emission line spectroscopy

Physics Oct A Quantum Harmonic Oscillator

Chapter 13. Phys 322 Lecture 34. Modern optics

ATMO/OPTI 656b Spring 2009

Stars AS4023: Stellar Atmospheres (13) Stellar Structure & Interiors (11)

Chemistry 795T. Lecture 7. Electromagnetic Spectrum Black body Radiation. NC State University

Chemistry 795T. Black body Radiation. The wavelength and the frequency. The electromagnetic spectrum. Lecture 7

ˆd = 1 2π. d(t)e iωt dt. (1)

Atomic Physics 3 ASTR 2110 Sarazin

1/30/11. Astro 300B: Jan. 26, Thermal radia+on and Thermal Equilibrium. Thermal Radia0on, and Thermodynamic Equilibrium

aka Light Properties of Light are simultaneously

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

CHAPTER 27. Continuum Emission Mechanisms

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)

What are Lasers? Light Amplification by Stimulated Emission of Radiation LASER Light emitted at very narrow wavelength bands (monochromatic) Light

PHAS3135 The Physics of Stars

Order of Magnitude Astrophysics - a.k.a. Astronomy 111. Photon Opacities in Matter

Problem 1. Hyperfine Emission from Neutral Hydrogen

Addition of Opacities and Absorption

ν is the frequency, h = ergs sec is Planck s constant h S = = x ergs sec 2 π the photon wavelength λ = c/ν

Blackbody radiation. Main Laws. Brightness temperature. 1. Concepts of a blackbody and thermodynamical equilibrium.

Phys 322 Lecture 34. Chapter 13. Modern optics. Note: 10 points will be given for attendance today and for the rest of the semester.

Astro 305 Lecture Notes Wayne Hu

FIBER OPTICS. Prof. R.K. Shevgaonkar. Department of Electrical Engineering. Indian Institute of Technology, Bombay. Lecture: 15. Optical Sources-LASER

Section 11.5 and Problem Radiative Transfer. from. Astronomy Methods A Physical Approach to Astronomical Observations Pages , 377

Astronomy The Nature of Light

Assignment 4 Solutions [Revision : 1.4]

MODERN OPTICS. P47 Optics: Unit 9

What is it good for? RT is a key part of remote sensing and climate modeling.

With certain caveats (described later) an object absorbs as effectively as it emits

Lasers PH 645/ OSE 645/ EE 613 Summer 2010 Section 1: T/Th 2:45-4:45 PM Engineering Building 240

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

Goal: The theory behind the electromagnetic radiation in remote sensing. 2.1 Maxwell Equations and Electromagnetic Waves

Chapter 1. From Classical to Quantum Mechanics

Physics Lecture 6

Relations between the Einstein coefficients

Astronomical Observations: Distance & Light 7/2/09. Astronomy 101

Infrared thermography

Astro 201 Radiative Processes Solution Set 1. by Roger O Brient and Eugene Chiang

Lecture Outline Chapter 30. Physics, 4 th Edition James S. Walker. Copyright 2010 Pearson Education, Inc.

Introduction to Modern Physics NE 131 Physics for Nanotechnology Engineering

Monte Carlo Radiation Transfer I

THE ELECTROMAGNETIC SPECTRUM. (We will go into more detail later but we need to establish some basic understanding here)

Weighting Functions and Atmospheric Soundings: Part I

Lecture 2: Transfer Theory

Modern Physics (Lec. 1)

External (differential) quantum efficiency Number of additional photons emitted / number of additional electrons injected

The formation of stars and planets. Day 1, Topic 2: Radiation physics. Lecture by: C.P. Dullemond

Synchrotron Radiation II

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

Quantum Statistics (2)

Fundamental Stellar Parameters

Radiation Transport in a Gas

PHYS 231 Lecture Notes Week 3

Lecture 2 Solutions to the Transport Equation

Light carries energy. Lecture 5 Understand Light. Is light. Light as a Particle. ANSWER: Both.

INTRODUCTION Radiation differs from conduction and convection in that it does not require the presence of a material medium to take place.

FI 3103 Quantum Physics

ASTRONOMY 161. Introduction to Solar System Astronomy. Class 9

What are Lasers? Light Amplification by Stimulated Emission of Radiation LASER Light emitted at very narrow wavelength bands (monochromatic) Light

Astro 201 Radiative Processes Problem Set 6. Due in class.

Nearly everything that we know about stars comes from the photons they emit into space.

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

AST 301, Lecture 2. James Lattimer. Department of Physics & Astronomy 449 ESS Bldg. Stony Brook University. January 29, 2019

summary of last lecture

Radiative Transfer Plane-Parallel Frequency-Dependent

The Death of Classical Physics. The Rise of the Photon

Electromagnetic Spectra. AST443, Lecture 13 Stanimir Metchev

The Electromagnetic Spectrum

Electromagnetic Radiation.

is the minimum stopping potential for which the current between the plates reduces to zero.

AST 105 Intro Astronomy The Solar System. MIDTERM II: Tuesday, April 5 [covering Lectures 10 through 16]

Einstein s Approach to Planck s Law

CHAPTER 3 The Experimental Basis of Quantum Theory

2. Basic assumptions for stellar atmospheres

QM all started with - - The Spectrum of Blackbody Radiation

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.

Transcription:

Transfer Equation and Blackbodies Initial questions: There are sources in the centers of some galaxies that are extraordinarily bright in microwaves. What s going on? The brightest galaxies in the universe with distributed emission (i.e., we re not talking about active galactic nuclei) are ultraluminous infrared galaxies. Why infrared? We are now in a position to examine what happens to a beam of radiation as it goes through matter. In general, the things that affect the intensity of the beam are: Spontaneous emission. Matter along the path of the beam can spontaneously emit radiation (e.g., spontaneous transitions of atoms from excited states to lower energy states). This adds to the intensity. Stimulated emission. If radiation passes through matter, the matter can be stimulated to emit radiation of the same frequency and direction as the initial radiation. This is the principle behind lasers, and it, too, adds to the intensity. Absorption. Radiation can be absorbed by the matter. This takes away from the intensity. Note that both absorption and stimulated emission are proportional to the original intensity of the beam of matter. They are therefore commonly lumped together (with stimulated emission being negative absorption ), so we ll group both of them under the category of absorption. Scattering. The photons can be redirected and changed in frequency without being destroyed. Spontaneous emission The monochromatic spontaneous emission coefficient j ν is defined as the energy per time per volume per frequency interval emitted into a solid angle dω: de = j ν dvdωdtdν. (1) Note that this is independent of the intensity of the beam. In general one must consider anisotropic emission, especially when there is a global preferred direction (e.g., set by a magnetic field). However, in many situations the emission is nearly isotropic, either because the emission from individual particles/atoms/molecules is isotropic or because the preferred axes of individual particles etc. are distributed randomly. In that case, j ν = P ν /4π, where P ν is the total emitted power per volume per frequency. On occasion you ll see the power expressed as an emissivity ǫ ν, which is the energy per mass per time per frequency, so P ν = ρǫ ν where ρ is the density. Ask class: considering only spontaneous emission, what is the change in specific inten-

sity I ν after traveling a distance ds in a medium with spontaneous emission coefficient j ν? It is di ν = j ν ds. (2) This can be determined directly from the units. It also makes sense physically: it s positive, since intensity is being added; the higher j ν is, the more intensity is added; and the farther one travels in the medium, the more intensity is enhanced. Absorption (including stimulated emission) In an analogous way, we define the absorption coefficient α ν (dimensions cm 1 ), by the equation di ν = α ν I ν ds. (3) Note that, unlike spontaneous emission, absorption is proportional to the intensity of the beam. This is the intensity removed from the beam after traveling a distance ds. By convention, α ν is positive for net energy removed; note, however, that α ν < 0 if stimulated emission dominates. Phenomenologically, we can understand this if we imagine a number density n of absorbers of cross section σ ν at frequency ν. Ask class: based on an analysis of units, how should α ν be related to n and σ ν? The only combination that gives the required cm 1 is α ν = nσ ν. This can also be written α ν = ρκ ν, where κ ν is the opacity. This phenomenological picture is only valid if the distance between absorbers is much larger than σ ν and the absorbers are distributed randomly. This is almost always the case for astrophysical situations, but for a wild violation consider metals. Metals are good thermal and electric conductors because electrons can travel great distances in them before interacting. However, if you naively consider the cross section of each atom and multiply by the number density of atoms, you d conclude that the mean free path is tiny and so metals are terrible conductors! The loophole is that atoms in metals are distributed in highly regular lattices, and quantum analysis of the resulting periodic potential shows that electrons can travel large distances. Scattering has some subtleties, so we ll postpone dealing with that for the moment and consider only spontaneous emission and absorption. The Radiative Transfer Equation We can combine the effects of spontaneous emission and absorption into one equation telling us what happens to the specific intensity along a ray: di ν /ds = α ν I ν +j ν. (4) Ask class: what happens when there is no absorption? Then only the second term on the right side matters, and di ν /ds = j ν. Formally, this may be integrated to get I ν (s) = I ν (s 0 )+ s s 0 j ν (s)ds. (5)

Ask class: now, what happens when there is no spontaneous emission? Then only the first term contributes and the formal solution is [ s ] I ν (s) = I ν (s 0 )exp α ν ds. (6) s 0 Thus, if absorption dominates over stimulated emission (i.e., α ν > 0) the intensity decreases exponentially along the path. If stimulated emission dominates then the intensity increases exponentially, which is one reason why lasers can get to such high intensities. The pure absorption case is particularly simple if we use the optical depth, which we discussed in an earlier class. Ask class: remembering that the optical depth is basically the number of mean free paths traversed by the beam, how should it be related to α ν? It is just dτ ν = α ν ds, so I ν (s) = I ν (s 0 )exp( τ) in the pure absorption case. Using the optical depth as a new variable, it is convenient to divide the radiative transfer equation by α ν to get di ν /dτ ν = I ν +S ν, (7) where S ν j ν /α ν is called the source function (note that j ν and α ν are both local properties of the material, so S ν is as well). This equation may be solved formally to get I ν (τ ν ) = I ν (0)e τν + τν 0 e (τν τ ν) S(τ ν)dτ ν. (8) We can get this result qualitatively from the principle that radiation may be superposed linearly as follows. The initial radiation (intensity I ν ) is attenuated by a factor e τν. There is also radiation generated along the way, and it too is attenuated by the end, by the appropriate factor exp( τ ν ), where τ ν = τ ν τ ν for radiation generated at the position τ ν. These contributions add linearly. For example, consider a constant source function S ν. Then I ν (τ ν ) = I ν (0)e τν +S ν (1 e τν ) = S ν +e τν [I ν (0) S ν ] (9) Ask class: what does this mean for a very optically thick medium? High optical depth means τ ν 1, so the exponential factor goes to zero and I ν S ν. This is important: it means that at high optical depth, if S ν doesn t vary much then I ν approaches S ν. If you include scattering things get more difficult, because the source function depends on I ν at all directions through a given point. For example, consider coherent, isotropic scattering (i.e., the photon energy doesn t change but its direction is completely randomized by scattering). Then j ν = α ν (scatt)j ν, where α ν (scatt) is the scattering coefficient (similar to an absorption coefficient; R+L use σ ν for this but I don t want to confuse it with the

cross section σ) and J ν is the mean intensity within the emitting material. Dividing, S ν = j ν /α ν = J ν, so the transfer equation for pure scattering is di ν /ds = α ν (scatt)(i ν J ν ) (10) and the transfer equation including thermal emission, absorption, and coherent, isotropic scattering is di ν /ds = (α ν +α ν [scatt])(i ν S ν ), (11) where the source function is the weighted average of those for absorption and emission: S ν = [α ν B ν +α ν (scatt)j ν ]/[α ν +α ν (scatt)], (12) where here we use the source function for thermal emission B ν (see below). This is an integro-differential equation and although some progress may be made using the Eddington approximation (see section 1.8 in R+L; we won t focus on this), in practice one solves this numerically. Thermal Emission A special and important case is one in which the radiation and matter are in thermal equilibrium. This gives a universal and exact solution and was the problem that let to the first glimmerings of quantum mechanics. Still, in any given problem you need to consider carefully whether the radiation and matter are in thermal equilibrium or whether nonthermal processes are important. Let s imagine an enclosure at temperature T. We have radiation in the enclosure, and we wait a long time until equilibrium has been achieved. Note that there is no conservation law for photons, so they can be created or destroyed by many processes (equivalently, we say that photons have zero chemical potential). From thermodynamics, the specific intensity that we get must be independent of the shape of the enclosure. Otherwise, you could take two enclosures at the same temperature but with different shapes, and when you put them in contact energy would flow from one to the other, in violation of thermodynamics. Thus, the specific intensity in thermal equilibrium depends only on the temperature. This argument also shows that the specific intensity in equilibrium must be isotropic. Therefore, in thermal equilibrium. I ν = B ν (T) (13) If you now put thermally emitting matter at the same temperature T into the enclosure, the source function must be unchanged because we still have a blackbody enclosure at the temperature T. Therefore (Kirchoff s law), the source function in thermal equilibrium is the blackbody function S ν = B ν (T). (14)

This also means that within the blackbody portion, I ν = B ν. Since blackbody radiation is homogeneous and isotropic, then from previous results we know that the pressure and energy density are related by p = u/3. Note that there is a useful distinction between blackbody radiation, where I ν = B ν, and thermal radiation, where S ν = B ν. Even if all the matter is radiating thermally (S ν = B ν ), you aren t guaranteed that I ν = B ν unless the medium is optically thick. We will now skip ahead a bit. Rybicki and Lightman give some interesting derivations about blackbody radiation on pages 17-21, from classical thermodynamics as well as quantum arguments. The quantum argument, which Planck used to derive the blackbody function (aka the Planck function), is particularly cute, and relies on photon modes being required to fit exactly (integral number of wavelengths) within a box, hence quantizing the photon energies. This was the first time that anyone had done that, and it marked a transition from classical to quantum theory that Einstein followed up five years later with the photoelectric effect. I recommend that you read the derivation, but let s focus now on properties of the function itself. Properties of the Planck Function Ablackbodyemitsaflux(energyperarea)ofσ SB T 4,whereσ SB = 5.67 10 5 ergcm 2 K 4 is the Stefan-Boltzmann constant. The Planck blackbody function is B ν (T) = or with wavelength λ = c/ν instead of ν as the primary variable, B λ (T) = Let s look at some limits of this expression. 2hν 3 /c 2 exp(hν/kt) 1, (15) 2hc 2 /λ 5 exp(hc/λkt) 1. (16) Low frequency, hν kt. The exponent is much less than unity, so B ν (T) 2ν 2 kt/c 2. Ask class: there is a fundamental constant missing here; what is it? The Planck constant h, of course! This leads to an important point that will allow you to check some equations. The Planck constant will appear if and only if quantization is important. In the same way, c appears if and only if special relativity is important, and G appears if and only if gravity is important. So why isn t quantization important here? In the low-frequency, or Rayleigh- Jeans, limit, there are many photons. We therefore have a classical description. However, it was noticed immediately that if this expression continued to be valid for arbitrarily high frequency the energy would diverge (this was called the ultraviolet catastrophe ). In the low-frequency limit the form of the Planck function (a power law, logarithmic slope 2 in frequency) is independent of the temperature.

High frequency, hν kt. Here the exponent is much greater than unity, so B ν (T) 2hν 3 /c 2 exp( hν/kt). Now the Planck constant does appear. It s because in this limit (the Wien limit), there are very few photons and hence their discrete nature matters. Here are some other interesting and important facts. First, there is a photon frequency at which B ν (T) peaks, and it s at hν max = 2.82kT, give or take. Incidentally, you could derive that the peak energy has to be proportional to kt simply by noticing that the only wayyoucangetsomethingwithdimensionsofenergyoutofh,c,k,andt isbykt! Scoreone for dimensional analysis. Moving on, although the peak frequency changes, B ν (T 1 ) > B ν (T 2 ) for all frequencies if T 1 > T 2. Finally, there are some characteristic temperatures that are defined in astrophysics to relate a given arbitrary spectrum to the blackbody spectrum. Brightness temperature. This is the temperature a blackbody would have to have to give the observed specific intensity at a given frequency: B ν (T b ) = I ν. This is especially common in radio astronomy, where because of the low frequencies you re usually in the Rayleigh-Jeans limit, so I ν = (2ν 2 /c 2 )kt b. Color temperature. This is the temperature of a blackbody that gives the same slope for the spectrum as the observed slope. This is useful whenever you don t know the luminosity, which is the case if you don t know the distance to the object (as an example). Effective temperature. This is the temperature of a blackbody that gives the same frequency-integrated intensity as the observed one, radiated at the source. That is, σ SB T 4 eff = F = I ν cosθdνdω. Recommended Rybicki and Lightman problem: 1.8