CHAPTER 22. Astrophysical Gases

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

Spectral Line Intensities - Boltzmann, Saha Eqs.

Interstellar Astrophysics Summary notes: Part 2

Theory of optically thin emission line spectroscopy

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

6. Stellar spectra. excitation and ionization, Saha s equation stellar spectral classification Balmer jump, H -

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

6. Stellar spectra. excitation and ionization, Saha s equation stellar spectral classification Balmer jump, H -

Equilibrium Properties of Matter and Radiation

Stellar Astrophysics: The Classification of Stellar Spectra

6. Stellar spectra. excitation and ionization, Saha s equation stellar spectral classification Balmer jump, H -

Ay Fall 2004 Lecture 6 (given by Tony Travouillon)

where n = (an integer) =

PHYS 231 Lecture Notes Week 3

Emitted Spectrum Summary of emission processes Emissivities for emission lines: - Collisionally excited lines - Recombination cascades Emissivities

CHAPTER 27. Continuum Emission Mechanisms

Lecture 4: Absorption and emission lines

7. Non-LTE basic concepts

Electron temperature is the temperature that describes, through Maxwell's law, the kinetic energy distribution of the free electrons.

(c) Sketch the ratio of electron to gas pressure for main sequence stars versus effective temperature. [1.5]

Saturation Absorption Spectroscopy of Rubidium Atom

X i t react. ~min i max i. R ij smallest. X j. Physical processes by characteristic timescale. largest. t diff ~ L2 D. t sound. ~ L a. t flow.

The early periodic table based on atomic weight. (Section 5.1) Lets review: What is a hydrogen atom? 1 electron * nucleus H 1 proton

Interstellar Medium Physics

Stellar atmospheres: an overview

The Classification of Stellar Spectra Chapter 8

II Light.

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

2. Basic Assumptions for Stellar Atmospheres

Lecture 3: Emission and absorption

Lecture 7: Molecular Transitions (2) Line radiation from molecular clouds to derive physical parameters

Photodissociation Regions Radiative Transfer. Dr. Thomas G. Bisbas

Photoionized Gas Ionization Equilibrium

RADIO SPECTRAL LINES. Nissim Kanekar National Centre for Radio Astrophysics, Pune

M.Phys., M.Math.Phys., M.Sc. MTP Radiative Processes in Astrophysics and High-Energy Astrophysics

Astronomy 421. Lecture 14: Stellar Atmospheres III

LECTURE NOTES. Ay/Ge 132 ATOMIC AND MOLECULAR PROCESSES IN ASTRONOMY AND PLANETARY SCIENCE. Geoffrey A. Blake. Fall term 2016 Caltech

Spectral Line Shapes. Line Contributions

The inverse process is recombination, and in equilibrium

The Birth Of Stars. How do stars form from the interstellar medium Where does star formation take place How do we induce star formation

Components of Galaxies Gas The Importance of Gas

Environment of the Radiation Field ...

2. Basic assumptions for stellar atmospheres

The Interstellar Medium

Astrochemistry and Molecular Astrophysics Paola Caselli

3. Stellar Atmospheres: Opacities

(c) (a) 3kT/2. Cascade

Chapter 39. Particles Behaving as Waves

Lecture 2: Formation of a Stellar Spectrum

Lecture 2 Interstellar Absorption Lines: Line Radiative Transfer

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

Redacted for Privacy (Major pro ssor)

Plasma Spectroscopy Inferences from Line Emission

5.111 Lecture Summary #5 Friday, September 12, 2014

Theory of Gas Discharge

7. Non-LTE basic concepts

Chapter 3C. 3-4C. Ionization

10.3 NMR Fundamentals

Outline. Today we will learn what is thermal radiation

Spectroscopy Lecture 2

Universo Primitivo (1º Semestre)

6. Interstellar Medium. Emission nebulae are diffuse patches of emission surrounding hot O and

Relations between the Einstein coefficients

Physics and Chemistry of the Interstellar Medium

11.1 Local Thermodynamic Equilibrium. 1. the electron and ion velocity distributions are Maxwellian,

Chapter 4. Spectroscopy. Dr. Tariq Al-Abdullah

Taking fingerprints of stars, galaxies, and interstellar gas clouds

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

Spontaneous Emission, Stimulated Emission, and Absorption

Astr 2310 Thurs. March 23, 2017 Today s Topics

Sharif University of Technology Physics Department. Modern Physics Spring 2016 Prof. Akhavan

10 May 2018 (Thursday) 2:30-4:30 pm

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

Some HI is in reasonably well defined clouds. Motions inside the cloud, and motion of the cloud will broaden and shift the observed lines!

Taking fingerprints of stars, galaxies, and interstellar gas clouds. Absorption and emission from atoms, ions, and molecules

MIDSUMMER EXAMINATIONS 2001 PHYSICS, PHYSICS WITH ASTROPHYSICS PHYSICS WITH SPACE SCIENCE & TECHNOLOGY PHYSICS WITH MEDICAL PHYSICS

PHYS 633: Introduction to Stellar Astrophysics

2. Basic assumptions for stellar atmospheres

Physical Processes in Astrophysics

Photoionization Modelling of H II Region for Oxygen Ions

PHYS-1050 Hydrogen Atom Energy Levels Solutions Spring 2013

Gas 1: Molecular clouds

Atomic Structure and Processes

Lecture 6 - spectroscopy

Opacity. requirement (aim): radiative equilibrium: near surface: Opacity

23 Astrophysics 23.5 Ionization of the Interstellar Gas near a Star

Interstellar Medium and Star Birth

Atomic Physics 3 ASTR 2110 Sarazin

arxiv: v1 [astro-ph.sr] 14 May 2010

Collisionally Excited Spectral Lines (Cont d) Diffuse Universe -- C. L. Martin

Physics 1C Lecture 29B

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

Astrophysics of Gaseous Nebulae and Active Galactic Nuclei

Lecture 6: Continuum Opacity and Stellar Atmospheres

CHAPTER 26. Radiative Transfer

ASTRONOMY QUALIFYING EXAM August Possibly Useful Quantities

Atomic Structure & Radiative Transitions

Collisional Excitation and N-Level Atoms.

Chapter V: Interactions of neutrons with matter

Some recent work I. Cosmic microwave background, seeds of large scale structure (Planck) Formation and evolution of galaxies (Figure: Simpson et al.

Transcription:

CHAPTER 22 Astrophysical Gases Most of the baryonic matter in the Universe is in a gaseous state, made up of 75% Hydrogen (H), 25% Helium (He) and only small amounts of other elements (called metals ). Gases can be neutral, ionized, or partially ionized. The degree of ionization of any given element is specified by a Roman numeral after the element name. For example HI and HII refer to neutral and ionized hydrogen, respectively, while CI is neutral carbon, CII is singly-ionized carbon, and CIV is triply-ionized carbon. A gas that is highly (largely) ionized is called a plasma. Thermodynamic Equilibrium: a system is said to be in thermodynamic equilibrium (TE) when it is in - thermal equilibrium - mechanical equilibrium - radiative equilibrium - chemical equilibrium - statistical equilibrium Equilibrium means a state of balance. In a state of thermodynamic equilibrium, there are no net flows of matter or of energy, no phase changes, and no unbalanced potentials (or driving forces), within the system. A system that is in thermodynamic equilibrium experiences no changes when it is isolated from its surroundings. In a system in TE the energy in the radiation field is in equilibrium with the kinetic energy of the particles. If the system is isolated (to both matter and radiation), and in mechanical equilibrium, then over time TE will be established. For a gas in TE, the radiation temperature, T R, is equal to the kinetic temperature, T, is equal to the excitation temperature, T ex (see below for definitions). Since no photons are allowed to escape from a system in TE (this would correspond to energy loss, and thus violate TE), the photons that are produced in the gas (represented by T R ) therefore must be tightly coupled to the random motions of the particles (represented by 136

T). This coupling implies that, for bound states, the excitation and deexcitation of the atoms and ions must be dominated by collisions. In other words, the collision timescales must be shorter than the timescales associated with photon interactions or spontaneous de-excitations. If that were not the case, T could not remain equal to T R. Local Thermodynamic Equilibrium (LTE): True TE is rare (in almost all cases energy does escape the system in the form of radiation, i.e., the system cools), and often temperature gradients are present. A good, albeit imperfect, example of TE is the Universe as a whole prior to decoupling. Although true TE is rare, in many systems (stars, gaseous spheres, ISM), we may apply local TE (LTE), which implies that the gas is in TE, but only locally. In a system in LTE, there will typically be gradients in temperature, density, pressure, etc, but they are sufficiently small over the mean-free path of a gas particle. For stellar intereriors, the fact that radiation is locally trapped explains why it takes so long for photons to diffuse from the center (where they are created in nuclear reactions) to the surface (where they are emitted into space). In the case of the Sun, this timescale is of the order of 200,000 years. Thermal Equilibrium: Thermal equilibrium in a system for itself means that the temperature within the system is spatially and temporally uniform. Thermal equilibrium as a relation between the physical states of two bodies means that there is actual or implied thermal connection between them, through a path that is permeable only to heat, and that no energy is transferred through that path. For a gas in thermal equilibrium at some uniform temperature, T, and uniform number density, n, the number density of particles with speeds between v and v+dv is given by the Maxwell-Boltzmann (aka Maxwellian ) velocity distribution: ( ) 3/2 ) m n(v)dv = n exp( mv2 4πv 2 dv 2πk B T 2k B T where m is the mass of a gas particle. The temperature T is called the kinetic temperature, and is related to the mean-square particle speed, 137

v 2, according to 1 2 m v2 = 3 2 k BT The most probable speed of the Maxwell-Boltzmann distribution is v mp = 2kB T/m, while the mean speed is v mean = 8k B T/m. Setting up a Maxwellian velocity distribution requires many elastic collisions. In the limit where most collisions are elastic, the system will typically very quickly equilibrate to thermal equilibrium. However, depending on the temperature of the gas, collisions can also be inelastic: examples of the latter are collisional excitations, in which part of the kinetic energy of the particle is used to excite its target to an excited state (i.e., the kinetic energy is now temporarily stored as a potential energy). If the de-excitation is collisional, the energy is given back to the kinetic energy of the gas (i.e., no photon is emitted). However, if the de-excitation is spontaneous or via stimulated emission, a photon is emitted. If the gas is optically thin to the emitted photon, the energy will escape the gas. The net outcome of the collisional excitation is then one of dissipation, i.e., cooling (kinetic energy of the gas is being radiated away). In the optically thick limit, the photon will be absorbed by another atom (or ion). If the system is in (L)TE, the radiation temperature of these photons being emitted and absorpbed is equal to the kinetic temperature of the gas. Although in LTE a subset of the collisions are inelastic, this subset is typically small, and one may still use the Maxwell-Boltzmann distribution to characterize the velocities of the gas particles. Mechanical Equilibrium: A system is said to be in mechanical equilibrium if (i) the vector sum of all external forces is zero and (ii) the sum of the moments of all external forces about any line is zero. Radiative Equilibrium: A system is said to be in radiative equilibrium if any two randomly selected subsystems of it exchange by radiation equal amounts of heat with each other. Chemical Equilibrium: In a chemical reaction, chemical equilibrium is 138

the state in which both reactants and products are present at concentrations which have no further tendency to change with time. Usually, this state results when the forward reaction proceeds at the same rate as the reverse reaction. The reaction rates of the forward and reverse reactions are generally not zero but, being equal, there are no net changes in the concentrations of the reactant and product. Statistical Equilibrium: A system is said to be in statistical equilibrium if the level populations of its constituent atoms and ions do not change with time (i.e., if the transition rate into any given level equals the rate out). For a system in statistical equilibrium, the population of states is described by the Boltzmann law for level populations: N j = g ( j exp hν ) ij N i g i k B T Here N i is the number of atoms in which electrons are in energy level i (here i reflects the principal quantum number, with i = 1 refering to the ground state), ν ij isthefrequencycorrespondingtotheenergydifference E ij = hν ij oftheenergylevels, histheplanckconstant, andg i isthestatistical weight of population i. Statistical weight: the statistical weight (aka degeneracy ) indicates the number of states at a given principal quantum number, n. In quantum mechanics, a total offour quantum numbers is needed todescribe the stateof an electron: the principle quantum number, n, the orbital angular momentum quantum number, l, the magnetic quantum number m l, and the electron spin quantum number, m s. For hydrogen l can take on the values 0,1,...,n 1, the quantum number m l can take on the values l, l+1, 1,0,1,...,l, and m s can take on the values +1/2 ( up ) and 1/2 ( down ). Hence, g n = 2n 2. Excitation Temperature: The above Boltzmann law defines the excitationtemperature, T ex, asthetemperaturewhich, whenputintotheboltzman law for level populations, results in the observed ratio of N j /N i. For a gas in (local) TE, all levels in an atom can be described by the same T ex, which is 139

also equal to the kinetic temperature, T, and the radiation temperature, T R. Under non-lteconditions, eachpair ofenergy levels canhave adifferent T ex. Radiation Temperature: the radiation temperature, or T R, is a temperature that specifies the energy density in the radiation field, u γ, according to u γ = a r T 4 R. In TE T R = T, while for a Black Body, T R is the temperature that appears in the Planck function that describes the specific intensity of the radiation field (see Chapter 21 for more details). Partition Function: Using the above Boltzmann law, it is straightforward to show that N n N = g ( n U exp E ) n k B T where N = N i is the total number of atoms (of all energy levels) and E n is the energy difference between state n and the ground-state, and U = g i e E i/k B T is called the partition function. (Note: the summation is over all principal quantum numbers up to some maximum n max, which is required to prevent divergence; see Irwin 2007 or Rybicki & Lightmann 1979 for details). Saha equation: the Saha equation expresses the ratios of atoms/ions in different ionization states. In particular, the number of atoms/ions in the (K+1) th ionization state versus those in the K th ionization state is given by N K+1 N K = 2 n e U K+1 U k ( ) 2πme k B T exp h 2 ( χ K k B T where U n is the partition function of the n th ionization state, n e is the electron number density, and χ K is the energy required to remove an electron from the ground state of the K th ionization state. Note that unlike the Boltzmann equation, the Saha equation for the ionization fractions has a dependence on electron density. This reflects that if there is a higher density of free electrons, there is a greater probability that an electron will recombine with the ion, lowering the ionization state of the gas. 140 )

Figure 18: Illustration of the energy levels of the hydrogen atom and some of the most important transitions. Hydrogen Since Hydrogen is the most common element in the Universe, it is important to have some understanding of its structure. Fig. 14 shows the energy levels of the hydrogen atoms and some if its most important transitions. The Balmer transitions have wavelengths that fall in the optical, and are therefore well known to (optical) astronomers. The Lyman lines typically fall in the UV, and can only be observed from space (or for highredshift objects, for which the rest-frame UV is redshifted into the optical). Using the Saha equation, we can compute the ionization fraction for hydrogen as function of temperature and electron density: N HII N HI = 2.41 10 15T3/2 n e exp ) ( 1.58 105 wherewehaveusedthatχ HI = 13.6eV,U HI = 2andU HII = 1(i.e., theionized hydrogen atom is just a free proton and only exists in a single state). We can also compute, using the Boltzmann law, the ratio of hydrogen atoms in the first excited state compared to those in the ground state. The latter shows that the temperature must exceed 30,000K in order for there to be an appreciable (10 percent) number of hydrogen atoms in the first excited T 141

state. However, at such high temperatures, one typically has that most of the hydrogen will be ionized (unless the electron density is unrealistically high). This somewhat unintuitive result arises from the fact that there are many more possible states available for a free electron than for a bound electron in the first excited state. In conclusion, neutral hydrogen in LTE will have virtually all its atoms in the ground state. If densities are sufficiently low, which is typically the case in the ISM, the collisionalexcitationrate(whichscaleswithn 2 e)islowerthanthespontaneous de-excitation rate. If that is the case, the hydrogen gas is no longer in LTE, and once again, virtually all neutral hydrogen will find itself in the ground state. Neutral hydrogen in the ISM is in the ground-state and is typically NOT in LTE. One important consequence of the fact that neutral hydrogen is basically always observed in the ground state, is that observations of hydrogen emission lines (Lyman, Balmer, Paschen, etc) indicates that the hydrogen must be ionized; the lines arise from recombinations, and are therefore called recombination lines. Note that Balmer lines are often observed in a absorption(for example, Balmer lines are evident in a spectrum of the Sun). This implies that their must be hydrogen present in the first excited state, which seems at odds with the conclusions reached above. A small fraction of excited hydrogen atoms, though, can still produce strong absorption lines, simply because hydrogen is so abundant. 21cm line emission: The ground state of hydrogen is split into two hyperfine states due to the two possible orientations of the proton and electron spins: The state in which the spins of proton and electron are aligned has slightly higher energy than the one in which they are anti-aligned. The energy difference between these two hyperfine states corresponds to a photon with a wavelength of 21cm (which falls in the radio). The excitation temperature of this spin-flip transition is called the spin temperature. Since, for typical interstellar medium conditions, the spin-flip transition is collisionally induced (i.e., the rate for spontaneous de-excitation is extremely low), the spin-temperature is typically equal to the kinetic temperate of the hydrogen gas. The 21cm line is an important emission line to probe the distribution (and temperature) of neutral hydrogen gas in the Universe. 142