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

Similar documents
de = j ν dvdωdtdν. (1)

Quantum Statistics (2)

Relations between the Einstein coefficients

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.

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

1. Why photons? 2. Photons in a vacuum

Outline. Today we will learn what is thermal radiation

Addition of Opacities and Absorption

1 Radiative transfer etc

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

Physics Oct A Quantum Harmonic Oscillator

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

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

Lecture Outline: Spectroscopy (Ch. 4)

Advanced Optical Communications Prof. R. K. Shevgaonkar Department of Electrical Engineering Indian Institute of Technology, Bombay

Sources of radiation

2. NOTES ON RADIATIVE TRANSFER The specific intensity I ν

ATMO/OPTI 656b Spring 2009

MODERN OPTICS. P47 Optics: Unit 9

[10] Spectroscopy (9/28/17)

Lecture 3: Emission and absorption

Atomic Physics 3 ASTR 2110 Sarazin

Lecture 3: Specific Intensity, Flux and Optical Depth

Example: model a star using a two layer model: Radiation starts from the inner layer as blackbody radiation at temperature T in. T out.

Ay Fall 2004 Lecture 6 (given by Tony Travouillon)

Chapter 13. Phys 322 Lecture 34. Modern optics

Semiconductor Optoelectronics Prof. M. R. Shenoy Department of physics Indian Institute of Technology, Delhi

Lecture 4: Absorption and emission lines

A system of two lenses is achromatic when the separation between them is

Statistical Equilibria: Saha Equation

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)

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

Synchrotron Radiation II

Light and Atoms

ˆd = 1 2π. d(t)e iωt dt. (1)

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

AST1100 Lecture Notes

Lecture 2 Line Radiative Transfer for the ISM

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

Physics Lab #2: Spectroscopy

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

PH104 Lab 1 Light and Matter Pre-lab

PHYS 172: Modern Mechanics Fall 2009

Quantum Electronics/Laser Physics Chapter 4 Line Shapes and Line Widths

5. Light-matter interactions: Blackbody radiation

Radiation Transport in a Gas

Equivalent Width Abundance Analysis In Moog

ASTRO 114 Lecture Okay. What we re going to discuss today are what we call radiation laws. We ve

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

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

EFFECTIVE PHOTON HYPOTHESIS, SELF FOCUSING OF LASER BEAMS AND SUPER FLUID

Chapter 26: Cosmology

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

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

Solving with Absolute Value

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

Big Bang, Black Holes, No Math

23 Astrophysics 23.5 Ionization of the Interstellar Gas near a Star

Units, limits, and symmetries

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

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

Phys 344 Ch 7 Lecture 6 April 6 th, HW29 51,52 HW30 58, 63 HW31 66, 74. w ε w. π ε ( ) ( )

Light and Atoms. ASTR 1120 General Astronomy: Stars & Galaxies. ASTR 1120 General Astronomy: Stars & Galaxies !ATH REVIEW: #AST CLASS: "OMEWORK #1

Light and Matter(LC)

Physics 161 Homework 3 - Solutions Wednesday September 21, 2011

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

Astro-2: History of the Universe. Lecture 12; May

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

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

PHYS 231 Lecture Notes Week 3

Dept. of Physics, MIT Manipal 1

5. Light-matter interactions: Blackbody radiation

Plasma Processes. m v = ee. (2)

Lecture 21: Lasers, Schrödinger s Cat, Atoms, Molecules, Solids, etc. Review and Examples. Lecture 21, p 1

Cosmology Dark Energy Models ASTR 2120 Sarazin

The Electromagnetic Spectrum

Einstein s Approach to Planck s Law

Chapter 27: The Early Universe

CHAPTER 26. Radiative Transfer

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

What can laser light do for (or to) me?

Spectroscopy Lecture 2

ASTRONOMY 161. Introduction to Solar System Astronomy. Class 9

Today. Announcements. Big Bang theory cont d Introduction to black holes

Laser Types Two main types depending on time operation Continuous Wave (CW) Pulsed operation Pulsed is easier, CW more useful

Quantum electronics. Nobel Lecture, December 11, 1964 A.M. P ROCHOROV

22. Lasers. Stimulated Emission: Gain. Population Inversion. Rate equation analysis. Two-level, three-level, and four-level systems

ASTRONOMY 103: THE EVOLVING UNIVERSE. Lecture 4 COSMIC CHEMISTRY Substitute Lecturer: Paul Sell

4. Energy, Power, and Photons

Physics 161 Homework 3 Wednesday September 21, 2011

Topics Covered in Chapter. Light and Other Electromagnetic Radiation. A Subatomic Interlude II. A Subatomic Interlude. A Subatomic Interlude III

Light and Other Electromagnetic Radiation

This is the third of three lectures on cavity theory.

Chapter 2 Bremsstrahlung and Black Body

Descriptive Statistics (And a little bit on rounding and significant digits)

Electrodynamics of Radiation Processes

The Basics of Light. Sunrise from the Space Shuttle, STS-47 mission. The Basics of Light

Review Chap. 18: Particle Physics

Physics Nov Bose-Einstein Gases

Transcription:

Ask about final Saturday, December 14 (avoids day of ASTR 100 final, Andy Harris final). Decided: final is 1 PM, Dec 14. Rate Equations and Detailed Balance Blackbodies arise if the optical depth is big enough. What does that mean? It means that all processes are in equilibrium with their inverses, so there is no net change in the chemical composition, ionization fraction, or whatever. If you can establish that this is the case in a particular situation, you are justified in assuming a blackbody. Even if it isn t exactly a blackbody, you can show in many cases (e.g., in stellar interiors) that in a local region (say, within one mean free path) your matter is close to equilibrium, so you assume LTE: Local Thermodynamic Equilibrium. This simplifies things greatly, so if you can show that LTE applies you can heave a sigh of relief. Ask class: can they think of astronomical situations in which LTE is a bad approximation? The interstellar medium is a good example; it s optically thin, and there are all sorts of nonequilibrium processes. In that case you can t assume equilibrium and must instead go to the fundamental rate equations. A side note: in countless situations involving radiative processes and other astrophysics it is a good idea to determine if your system is in equilibrium in one way or another. If it is, you are allowed drastic simplifications. For example, Ask class: is the Sun in equilibrium? We d have to say what that means. The free-fall time for the Sun would be about an hour if nothing stopped it, but the Sun has been shining away for billions of years, so it s in dynamic equilibrium. Look for this kind of equilibrium, but always check first, otherwise you could mislead yourself a lot! Okay, back to rate equations. In principle you would think that you would have multiple unrelated rates for a given process. For example, consider a two-level atom: ground state (level 1) and excited state (level 2). There s an absorption rate (1 2), a spontaneous emission rate (2 1), and a stimulated emission rate (also 2 1). We know from Kirchoff s law (j ν = α ν B ν ) that for a thermal emitter at least there is some relation between emission and absorption, so we might suspect there is a microscopic relation between the two. In fact, Einstein (1916) showed that all three rates are related to each other whether or not the system is in thermal equilibrium! This is called the principle of detailed balance, and it simplifies the analysis of rates considerably. Let s follow Einstein s treatment of a two-level atom. Let A 21, the Einstein A coefficient, be the rate of spontaneous emission. This is the transition probability per unit time that an atom currently in level 2 will go to level 1. Now let s assume that the atom is in the midst of a radiation field with a mean intensity

at frequency ν of J ν = (1/4π) I ν dω. Note that the difference in energies between level 2 and level 1 can t be defined with infinite precision, because of the uncertainty principle. Let us therefore be careful and define a line profile φ(ν) that is sharply peaked at the line center ν 0 and is normalized so that φ(ν)dν = 1. (1) 0 We can define an average intensity over this profile, J = 0 J ν φ(ν)dν. Note that if J ν varies slowly over the line profile (which it almost always does), this practically reduces to J ν0, but it s nice to be careful. Ask class: what effects can the external radiation field produce on the atom? Absorption or stimulated emission. Now, we d like to assume that the absorption rate and stimulated emission rate are both proportional to J, i.e., they are linearly proportional to intensity. As always, though, we want to be sure we know when this simplifying assumption can become invalid. This is a tough one, though: Ask class: when might this be incorrect? The linear relation could break down if the radiation is self-interacting, which might happen at extremely high intensities or with high photon energies (when processes such as pair production enter in). In almost all applications this is insignificant, but if you re analyzing a very high energy environment (e.g., gamma-ray bursts) you should be cautious. Anyway, let s apply that assumption. Then we have a transition rate of B 12 J for absorption and B 21 J for stimulated emission, where B12 and B 21 are the Einstein B coefficients. How are the three coefficients related to each other? We ll start by analyzing the situation in thermodynamic equilibrium. For this analysis we assume that the upper level has a statistical weight of g 2 and the lower level has a statistical weight of g 1. We also assume that in equilibrium there are n 2 atoms in level 2 and n 1 in level 1. Ask class: what relation does equilibrium imply? It implies that the rate from level 1 to 2 equals the rate from level 2 to 1, or n 1 B 12 J = n2 A 21 + n 2 B 21 J. (2) Therefore, A 21 /B 21 J = (n 1 /n 2 )(B 12 /B 21 ) 1. (3) In equilibrium, the level populations n 1 and n 2 are given by the Boltzmann distribution, which says that the number is proportional to the statistical weight times exp( E/kT), where E is the energy of the level. We re taking a ratio, so the zero point of the energy doesn t matter and therefore J = A 21 /B 21 (g 1 B 12 /g 2 B 21 ) exp(hν 0 /kt) 1. (4)

Here hν 0 is the energy difference between the levels, with level 2 assumed to have higher energy than level 1. We also know, however, that in thermodynamic equilibrium J ν = B ν, the Planck function. Since the Planck function varies slowly over a sharp line profile, J = B ν. This expression must hold for all temperatures and for all frequencies ν 0. The only way this can happen is if the Einstein relations hold: g 1 B 12 = g 2 B 21 A 21 = (2hν 3 /c 3 )B 21. (5) Now let s sit back and meditate on what we ve just shown. You might think that we have only shown that certain relations hold in thermal equilibrium. But actually this is much more general. Consider: none of the coefficients A 21, B 12, and B 21 can depend on the external radiation field. A 21 is a spontaneous emission coefficient, so it doesn t interact with the external field at all. Also, since we have assumed that the rates of absorption and stimulated emission are linearly proportional to the mean intensity, the B coefficients are also independent of any properties of the radiation field. Think of this another way: the B coefficients only depend on a narrow range of frequencies near ν 0, so they don t know about the full spectrum. Thus, it can t be relevant whether the full spectrum is a blackbody or not, or even what the intensity is. As a result, the Einstein relations always hold given our assumption that the mean intensity enters linearly. Let s develop our intuition a bit more with a couple of problems from the book. Ask class: what result do we expect if stimulated emission is ignored? This isn t obvious, but it s good to take a guess anyway so we can learn more after doing the problem. Again using the trick of thinking first about thermal equilibrium, we now have Using the Boltzmann distribution this implies n 1 B 12 J = n2 A 21. (6) J = (g 2 A 21 /g 1 B 12 ) exp( hν 0 /kt). (7) Ask class: can this equal the Planck function with A 21 and B 12 independent of temperature? Nope, the functional form is wrong. However, we do find that if A 21 /B 12 = (2hν 3 0/c 3 )(g 1 /g 2 ) then J equals the Planck function in the Wien limit hν 0 kt. This says that stimulated emission is negligible in the Wien limit. Ask class: why is this true? It s because the Wien limit is where the number of photons in the radiation field drops off exponentially. Since stimulated emission is proportional to the number of photons whereas spontaneous emission is independent of the number of photons, there will come a point when stimulated emission can be ignored. Now another example. Suppose that the atom interacts with a neutrino field instead of a photon field. As we ll discuss more next time, a neutrino is a fermion (a photon is a

boson), so that two or more neutrinos cannot occupy the same state. As a result, instead of stimulated emission you d have suppressed emission: In addition, in thermal equilibrium the mean intensity is n 1 B 12 J = n2 A 21 n 2 B 21 J. (8) J ν = 2hν 3 /c 3 exp(hν/kt) + 1. (9) Ask class: if we carry through the same analysis as before, what would they guess would be the neutrino Einstein relations? As you can verify directly, you get exactly the same expressions as we had before! This is a dramatic indication that the Einstein relations are properties of the atomic physics and have nothing whatsoever to do with the properties of an external radiation field! Essentially, bringing in the field is just a convenient way to get the answer, but it isn t necessary. One comment, before we go on: from the quantum mechanical standpoint, the natural processes are absorption and stimulated emission. Both are, essentially, the result of an interaction of the atom with an external field. The strange one is spontaneous emission. Why should an isolated atom decide to radiate? In quantum electrodynamics, the answer is that even in a vacuum the atom is interaction with something, in this case vacuum fluctuations. You can therefore think of it as interacting with virtual photons instead of real photons, but at its base it is the same thing. One can actually observe this in the laboratory, by shutting an atom in a box smaller than the wavelength of the light it emits, so that the vacuum fluctuations are inhibited (Kleppner, D. 1981, Phys. Rev. Lett., 47, 233 for the theory) or enhanced in a resonant cavity (Goy et al. 1983, Phys. Rev. Lett., 50, 1903 for an experiment). We can express the emission and absorption coefficients using the Einstein coefficients. Let s assume that the line profile for emission is the same as it is for absorption (a very good assumption most of the time). Then, as derived in the book, the emission coefficient is and the absorption coefficient is j ν = (hν 0 /4π)n 2 A 21 φ(ν) (10) α ν = (hν/4π)φ(ν)(n 1 B 12 n 2 B 21 ). (11) Note that again stimulated emission is acting like a negative absorption. For the record (and again from the book, equation 1.6), the transfer equation then becomes and the source function is di ν /ds = (hν/4π)(n 1 B 12 n 2 B 21 )φ(ν)i ν + (hν/4π)n 2 A 21 φ(ν) (12) S ν = n 2 A 21 /(n 1 B 12 n 2 B 21 ) = (2hν 3 /c 2 )/ ( g 2 n 1 g 1 n 2 1 ) 1 (13).

Note that the A coefficient has different units from the B coefficients (the latter, multiplied by I ν, give the units of A 21 ). Also, note that it is pointless to memorize equations such as this. You can get an idea of what the answer should be by asking yourself how the intensity should change. Clearly, the intensity will increase if there is spontaneous emission or stimulated emission, and decrease if there is absorption. Also, remember that absorption and stimulated emission are proportional to I ν whereas spontaneous emission isn t. That should allow you to piece together most of the important parts. If the matter is in thermal equilibrium with itself, but not necessarily with the radiation, we have local thermodynamic equilibrium and n 1 /n 2 = (g 1 /g 2 ) exp(hν/kt). You can verify that this makes S ν = B ν. In all other cases (nonthermal emission), this equality does not hold. A particularly cool way in which matter can be out of thermal equilibrium is in a laser or maser (the same phenomena but different wavelengths; however, masers are found in space whereas lasers tend to be laboratory creations). Let s divert with these a bit. Suppose for simplicity that g 1 = g 2 (equal multiplicity of the two states). Then B 12 = B 21. Ask class: if n 1 > n 2, which dominates, absorption or stimulated emission? Absorption. This is the normal case, because E 2 > E 1 so in thermal equilibrium (or even most forms of non-thermal distributions) the lower-energy state is more populated. Ask class: but what if n 2 > n 1? Then the intensity increases exponentially along the path of the beam. Moreover, although it isn t clear from our discussion here, bosons (e.g., photons) like to occupy the same quantum state, so the direction and phase of the emitted photon is the same as that of the original photon. This means tremendous intensity! Just for fun, let s think about what conditions might allow a laser/maser. Ask class: what are the requirements? We need an inverted population, which has higher energy than the normal population, meaning we need some supply of energy. Another, trickier, requirement is that the upper state must persist for a while (otherwise spontaneous emission will reduce the number of atoms in the upper state). In practice this often means that the level structure has to be complicated: excitation can occur to another state, say state 3, which decays to state 2. State 2 is then metastable, meaning it has a long time before it will decay. In the lab it takes lots of special preparation to get all this to work, mainly because the number densities (of gas, liquid, or solid) are high enough that thermal equilibrium (specifically a normal population) can be established rapidly. In space, the number densities are much less and therefore it is a lot easier for inverted populations to exist. The result is that masers are common in astronomy: around very young stars, around old stars, in the centers of galaxies, and so on. The tiny size of masing regions, plus their high intensity and extreme frequency coherence, makes them remarkably good probes of astrophysical conditions.