Lecture 3: Specific Intensity, Flux and Optical Depth

Similar documents
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

TRANSFER OF RADIATION

Opacity and Optical Depth

2. NOTES ON RADIATIVE TRANSFER The specific intensity I ν

The Radiative Transfer Equation

Interstellar Medium Physics

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

Outline. Today we will learn what is thermal radiation

Lecture 2: Transfer Theory

Lecture 2 Solutions to the Transport Equation

Astro 305 Lecture Notes Wayne Hu

Ay121 Problem Set 1. TA: Kathryn Plant. October Some parts are based on solutions by previous TAs Yuguang Chen and Rachel Theios.

CHAPTER 26. Radiative Transfer

Lecture 3: Emission and absorption

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

de = j ν dvdωdtdν. (1)

Fundamental Stellar Parameters

p(θ,φ,θ,φ) = we have: Thus:

SIMPLE RADIATIVE TRANSFER

point, corresponding to the area it cuts out: θ = (arc length s) / (radius of the circle r) in radians Babylonians:

Numerical Heat and Mass Transfer

1. Why photons? 2. Photons in a vacuum

12.815/12.816: RADIATIVE TRANSFER PROBLEM SET #1 SOLUTIONS

PHYS 390 Lecture 23 - Photon gas 23-1

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

Radiation in the atmosphere

Sources of radiation

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

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

Ay Fall 2004 Lecture 6 (given by Tony Travouillon)

Weighting Functions and Atmospheric Soundings: Part I

Spectroscopy Lecture 2

Stellar Atmospheres: Basic Processes and Equations

MCRT: L4 A Monte Carlo Scattering Code

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

Introduction to the School

1 CHAPTER 5 ABSORPTION, SCATTERING, EXTINCTION AND THE EQUATION OF TRANSFER

Bremsstrahlung Radiation

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

Lecture 6: Continuum Opacity and Stellar Atmospheres

1 Radiative transfer etc

3 Some Radiation Basics

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

Radiation Transport in a Gas

We start with a reminder of a few basic concepts in probability. Let x be a discrete random variable with some probability function p(x).

Kinds of Energy. Defining Energy is Hard! EXPLAIN: 1. Energy and Radiation. Conservation of Energy. Sco; Denning CSU ESMEI ATS 1

Chapter 3 Energy Balance and Temperature. Astro 9601

Assignment 4 Solutions [Revision : 1.4]

Addition of Opacities and Absorption

Beer-Lambert (cont.)

PHYS f: Problem set #0 Solutions

Chapter 3 Energy Balance and Temperature. Topics to be covered

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

aka Light Properties of Light are simultaneously

Atmospheric Sciences 321. Science of Climate. Lecture 6: Radiation Transfer

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

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

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

Ay 20 Basic Astronomy and the Galaxy Problem Set 2

Monday 9 September, :30-11:30 Class#03

Lecture 2: Transfer theory. Credit: Kees Dullemond

Mean Intensity. Same units as I ν : J/m 2 /s/hz/sr (ergs/cm 2 /s/hz/sr) Function of position (and time), but not direction

23 Astrophysics 23.5 Ionization of the Interstellar Gas near a Star

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

Spontaneous Emission, Stimulated Emission, and Absorption

Analysis of Scattering of Radiation in a Plane-Parallel Atmosphere. Stephanie M. Carney ES 299r May 23, 2007

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

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

Electromagnetic Spectra. AST443, Lecture 13 Stanimir Metchev

Astronomy 421. Lecture 13: Stellar Atmospheres II. Skip Sec 9.4 and radiation pressure gradient part of 9.3

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

1. Runaway Greenhouse

SISD Training Lectures in Spectroscopy

F = c where ^ı is a unit vector along the ray. The normal component is. Iν cos 2 θ. d dadt. dp normal (θ,φ) = dpcos θ = df ν

Limb Darkening. Limb Darkening. Limb Darkening. Limb Darkening. Empirical Limb Darkening. Betelgeuse. At centre see hotter gas than at edges

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

Basic Properties of Radiation, Atmospheres, and Oceans

Lecture Notes Prepared by Mike Foster Spring 2007

Plasma Processes. m v = ee. (2)

1. Theory of Radiative Transfer. An expert is someone who has made all the mistakes that can be made in a narrow field. - Niels Bohr.

4. Energy, Power, and Photons

ATMO/OPTI 656b Spring 2009

Prof. Jeff Kenney Class 4 May 31, 2018

Monte Carlo Radiation Transfer I

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

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

The Electromagnetic Spectrum

Stellar Astrophysics: The Continuous Spectrum of Light

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

Components of Galaxies Gas The Importance of Gas

PHY-105: Nuclear Reactions in Stars (continued)

a few more introductory subjects : equilib. vs non-equil. ISM sources and sinks : matter replenishment, and exhaustion Galactic Energetics

Fundamentals on light scattering, absorption and thermal radiation, and its relation to the vector radiative transfer equation

The mathematics of scattering and absorption and emission

Characteristic temperatures

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

ASTR2050 Spring Please turn in your homework now! In this class we will discuss the Interstellar Medium:

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

1. Why photons? 2. Photons in a vacuum

Astr 323: Extragalactic Astronomy and Cosmology. Spring Quarter 2014, University of Washington, Željko Ivezić. Lecture 1:

Transcription:

Lecture 3: Specific Intensity, Flux and Optical Depth We begin a more detailed look at stellar atmospheres by defining the fundamental variable, which is called the Specific Intensity. It may be specified as a function of frequency, I ν, or of wavelength, I λ. It is basically the same as flux except limited to those photons headed in a particular direction (i.e. confined within a certain solid angle).

Specific intensity is a nearly complete description of the radiant energy field. (The only thing not included is information about polarization.) Otherwise, I describes how many photons of what energies are present and what directions they are headed. Everything else of interest (e.g. the flux through any particular area oriented any particular way, the radiation pressure, the radiant energy density, etc. can be derived from the intensity (notice I used its nickname here!) Let s get to know I by noting some of its properties. Since, I λ dλ = I λ = I ν dν dλ = I ν I ν dν to switch from frequency to wavelength use the following relations: c λ 2 = I ν ν 2 c The units of I ν are (ergs per sec per cm^2 per sterradian per Hz). I ν Note that is independent of distance. It is like a surface brightness... ie. magnitudes per square arc-sec. Surface brightness is not dependent on distance (neglecting cosmological effects!). Stars don t fade with distance, they get smaller...more later.

A very important quantity that can be derived from specific intensity is the flux through any surface. All we have to do is add up all the energy going through the surface, regardless of direction. Remember that what we mean by flux might be thought of as the net flux -- i.e. the difference between flux going one direction (normal to the chosen area) and going the opposite way (anti-normal). This follows immediately when we consider that the flux must be a weighted sum over the intensity, since photons parallel to our chosen area do not actually go through it and contribute zero to the flux. Some reflection (perhaps a decade will do) will convince you that the intensity needs to be weighted by the cosine of the angle between the normal to the area chosen and the direction of the rays. For now, accept it as the wisdom of those who have gone before you! F ν = I ν cosθdω = I ν cosθsinθdθdφ As a trivial example, we show that the flux in an isotropic radiation field is zero! F ν = I ν cosθdω =I ν 2π π π cosθsinθdθdφ =2πI ν cosθsinθdθ =.

Note on doing these types of integrals, which appear a lot in astrophysics. Use substitution µ = cosθ from which dµ = sinθdθ follows. π 1 [ ] µ 2 1 cosθsinθdθ = µdµ = = 2 As a less trivial and very important example, let s calculate the flux at the surface of a star, assuming that all of the intensity is outward directed (i.e. space is black!) and the outward bound intensity is independent of direction, ie. I ν I ν (θ, φ) π 2 F ν = I ν cosθdω =2πI ν cosθsinθdθ =2πI ν xdx = πi ν 1 The famous equation F = σt 4 follows by integrating the Planck function over all frequencies (see problem set)! The inverse square law follows for stars as follows: Since Intensity is independent of distance, and assuming that stars are very far away and nearly point sources, one can take cosθ = 1 and F ν = I ν Ω So, again note that stars get fainter with But Ω = πr2 d 2 so F d 2 distance because they subtend a smaller solid angle, not because they fade in any way. 1 1

In Section 9.1 of the text, the authors show you how to derive other quantities of interest from intensity. We don t need to repeat all the stuff here, but will highlight a few items: Mean intensity: Energy Density: J ν <I ν >= 1 u ν = 1 c I ν dω= 2π c I ν dω= 1 2 π π I ν (θ)sinθdθ I ν (θ)sinθdθ = c J ν Note: energy density is the ergs per cubic cm (or joules per cubic m) in the radiation field. The factor of c arises when one considers the volume in which light is confined. In time dt the light travels a distance ds=cdt. See book for more on this derivation. This quantity is sometimes called monochromatic energy density. To get the full radiant energy density (at all frequencies) integrate! In a blackbody in STE, I = J = B (the Planck function). In this case, we can calculate the energy density as: u = u ν dν = B ν (T )dν = 4σT 4 at 4 c c The new constant a is called the radiation constant

Radiation Pressure (see book for derivation): The radiation pressure can be considered the second moment of intensity, in the sense of a moment arm. The first moment is flux. The effect of photons on the net flux depends on their direction, so we weight photons according to the angle they make with the normal to the area considered. In the radiation pressure calculation, as we go to higher inclinations, not only are there fewer photons making it through da but each photon carries a diminished contribution to the normal force (i.e. the pressure). Again, the factor is cos (theta), so radiation pressure is the second moment of intensity. P = 1 c I ν cos 2 (θ)dωdν For an isotropic radiation field: P = 3c J νdν And, for a black body in STE, where J = B: P = 3c B νdν = 4σT 4 3c = 1 3 at 4 = 1 3 u

If the Universe were empty then the problem of radiative transfer would be easy and the solution of the equation would be I ν = I ν (). That is, intensity at the location of the observer would be the same as intensity at the location of the source. Intensity would be constant. However, the Universe is full of matter that will both absorb energy from the beam of light and add energy to it. A description of how the intensity changes as it moves through space (with matter in it) is given by the equation of radiative transfer. This may be thought of as the fundamental equation of stellar atmospheres and is one of the fundamental equations of all of astrophysics. Derivation is quite simple, because only two things can happen to light as it passes through a small volume of space -- the intensity is either reduced by absorption or scattering of radiation out of the beam, or it is increased by the emission of photons by matter in the volume or scattering of photons originally headed in other directions into the same direction as the beam is headed. Two parameters are needed to describe these two possibilities: an absorption coefficient (k) and an emission coefficient (j). Both, of course, are functions of frequency (or wavelength).

If we consider a small volume dv = da ds where ds is along the beam direction and da is perpendicular to it, then the amount of energy added to the beam in time dt and in frequency interval d(nu) and confined to within d(omega) of the beam direction in going from s to s+ds is given by Radiant energy added to beam = j ν dadsdtdνdω Likewise, the amount of radiant energy scattered from the beam or removed by absorption, is given by Radiant energy removed from beam = k ν I ν dadsdtdνdω Note that the amount of energy removed depends on how strong the beam is going in...if no energy comes in then none gets removed! On the other hand, the amount added to the beam is independent of incoming intensity. The only exception to this is laser emission (which sometimes happens in space!) Usually this effect is treated as negative absorption because it is emission which does depend on the level of the incoming radiation. We will not discuss laser emission any further in this course. The change in intensity is equal to the amount added - the amount removed: [I ν (s + ds) I ν (s)]dadtdνdω =[j ν k ν I ν ]dadsdtdνdω

Simplifying, and taking the limit as ds goes to zero, we have: di ν ds = j ν k ν I ν which is one form of the Equation of Radiative Transfer. A more common and useful form comes by dividing both sides of the equation by k and making two definitions, yielding: di ν dτ ν = S ν I ν where S ν j ν k ν is called the Source Function and dτ ν k ν ds is called the optical depth Some examples may help clarify these definitions and their point. Consider a gas in STE, then there is no change in intensity with position so the left hand side of the equation is zero and S ν = I ν = B ν The source function is also equal to the Planck function, at least approximately, in other situations, such as Local Thermodynamic Equilibrium (LTE) where the local kinetic temperature, T, is used for the assumption that S ν = B ν (T )

To get a feel for the meaning of optical depth, consider the transfer of radiation through a medium where the absorption and scattering (together often referred to as extinction ) is important, but the emission is not. An example of this is the propagation of star light through the interstellar medium. At optical wavelengths the ISM adds nothing of importance to the beam because the matter is so cold. However, the dust in space does cause extinction if the path length is long enough. In this case, the source function is zero, so the transfer equation can be written as: di ν = I ν or di ν = dτ ν dτ ν I ν This equation is easily solved by integrating both sides. It requires one boundary condition, which we take to be I = I() at tau =. Then, at any other optical depth (tau) we have I ν = I ν ()e τ ν This tells us that when we view a star through an optical depth of 1, its intensity is reduced by 1/e compared to what it would be if there were no extinction. So, we can think of optical depth as the number of e-folds that the light is attenuated. If tau << 1 the medium is called optically thin and there is little absorption. If tau >> 1 the medium is called optically thick and we can t see through it! It is opaque.

One can also relate the optical depth to the reduction of brightness (in magnitudes) suffered by the star due to the extinction. Using the magnitude equation we find the difference in magnitude between the extincted and unextincted star as: I ν m = 2.5log 1 I ν () = 2.5log 1e τ ν =1.86τ ν So, roughly speaking, the optical depth is about equal to the number of magnitudes of extinction suffered by the beam. If we think about it on a microscopic level, the attenuation (extinction) of light is due to its interaction with bits of matter -- atoms, electrons, molecules or larger stuff, collectively called dust or grains. While there can be many different kinds of particles present, lets consider for simplicity that there is only one kind, that there are n such particles per unit volume and that each particle has a cross-section for extinction equal to sigma. Then we may write that dτ ν = k ν ds = nσ ν ds If we integrate over a path length such that tau = 1 and call that path length s then we can identify s as the mean free path of a photon and the optical depth (tau) as the number of mean free paths to the source of the beam.

Clearly, s (mean free path) = 1 nσ ν Put another way, the mean free path is the distance through the medium that one has to go until the cross-sections of all the absorbing particles would block the entire beam (if they did not overlap each other). Because they DO overlap each other, some light does get through even when the path length through the medium is 1 mean free path. How much light? Well, the intensity is reduced by 1/e =.368, so only about 63% of the light is actually blocked. 37% makes it through!