Top Banner
Soft X-Ray Spectroscopy of Astrophysical Plasmas S.M. Kahn Columbia University, New York, USA 1 Introduction These lectures are intended to provide a review of the basic physics necessary for the interpretation of high resolution soft X-ray spectra of astrophysical sources. While many of the topics I discuss can be found at the requisite level of sophistication in standard textbooks on atomic physics and spec- troscopy (e.g. [1]), I have made an attempt to highlight those aspects which are especially important for X-ray transitions, and which are relevant at the characteristic temperatures and densities typically found in various types of X-ray emitting astrophysical plasmas. My emphasis is on discrete atomic transitions, which dominate the spectra of most cosmic sources in the soft X-ray band (100 eV E 10 keV). I do not discuss basic continuum processes like bremsstrahlung, synchrotron emission, and inverse Compton emission, as these are covered well in the usual texts used to introduce students to radiative processes in astrophysics (e.g. [2]). In general, I avoid long derivations, concentrating instead on the key phys- ical ideas that underlie the various formulas, and especially on the definition of terms that appear frequently in the atomic physics literature. The level is intended for advanced undergraduates and beginning graduate students with little or no background in X-ray spectroscopy. While I do assume a rudimen- tary familiarity with the basics of classical and quantum physics (typical of the preparation one would receive as an undergraduate physics major in an American university), the lectures are self-contained, and were designed to provide a suitable introduction to this field without the need for extensive consultation of other source materials. The organization is as follows: in the remainder of this initial chapter, I provide a brief introduction to the role of X-ray spectroscopy in astro- physics, and the physical conditions in various types of cosmic X-ray sources. Chapters 1 through 3 cover the essentials of atomic physics: classical and quantum radiation theory, atomic structure, and electron-ion collisional processes, respectively. In Chap. 4, I discuss the various types of equilib- ria that apply in astrophysical plasmas, and in Chap. 5, I provide a relatively brief review of the most important discrete-line spectral diagnostics that fall in the soft X-ray band. Chapter 6 includes a set of concluding remarks and some thoughts on where this field might be headed in the future.
80

Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Nov 10, 2020

Download

Documents

dariahiddleston
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopyof Astrophysical Plasmas

S.M. Kahn

Columbia University, New York, USA

1 Introduction

These lectures are intended to provide a review of the basic physics necessaryfor the interpretation of high resolution soft X-ray spectra of astrophysicalsources. While many of the topics I discuss can be found at the requisitelevel of sophistication in standard textbooks on atomic physics and spec-troscopy (e.g. [1]), I have made an attempt to highlight those aspects whichare especially important for X-ray transitions, and which are relevant at thecharacteristic temperatures and densities typically found in various types ofX-ray emitting astrophysical plasmas.

My emphasis is on discrete atomic transitions, which dominate the spectraof most cosmic sources in the soft X-ray band (100 eV ≤ E ≤ 10 keV). I do notdiscuss basic continuum processes like bremsstrahlung, synchrotron emission,and inverse Compton emission, as these are covered well in the usual textsused to introduce students to radiative processes in astrophysics (e.g. [2]).

In general, I avoid long derivations, concentrating instead on the key phys-ical ideas that underlie the various formulas, and especially on the definitionof terms that appear frequently in the atomic physics literature. The level isintended for advanced undergraduates and beginning graduate students withlittle or no background in X-ray spectroscopy. While I do assume a rudimen-tary familiarity with the basics of classical and quantum physics (typical ofthe preparation one would receive as an undergraduate physics major in anAmerican university), the lectures are self-contained, and were designed toprovide a suitable introduction to this field without the need for extensiveconsultation of other source materials.

The organization is as follows: in the remainder of this initial chapter,I provide a brief introduction to the role of X-ray spectroscopy in astro-physics, and the physical conditions in various types of cosmic X-ray sources.Chapters 1 through 3 cover the essentials of atomic physics: classical andquantum radiation theory, atomic structure, and electron-ion collisionalprocesses, respectively. In Chap. 4, I discuss the various types of equilib-ria that apply in astrophysical plasmas, and in Chap. 5, I provide a relativelybrief review of the most important discrete-line spectral diagnostics that fallin the soft X-ray band. Chapter 6 includes a set of concluding remarks andsome thoughts on where this field might be headed in the future.

Page 2: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

4 S.M. Kahn

1.1 The Role of X-Ray Spectroscopy in Astrophysics

X-ray astronomy is not a “new” field of research. Most practitioners dateits inception to the serendipitous detection of the very bright binary X-raysource, Scorpius X-1, in June of 1962 [3]. That momentous discovery provedthat cosmic systems could be copious X-ray emitters, and that observations inthe X-ray band could provide new insights into astrophysical phenomena thatcould not be gleaned from observations at longer wavelengths. In the ensuingforty years, this field has grown to become one of the major disciplines ofobservational astrophysics. Hundreds of thousands of discrete sources of X-ray emission have been detected, covering nearly all classes of astrophysicalsystems.

Until very recently, however, real X-ray spectra of astrophysical sources,with sufficient resolution and sensitivity to enable the investigation of indi-vidual atomic features, had been largely unavailable. This was principallydue to instrumental limitations. Since cosmic X-ray sources are exceedinglyfaint (typical fluxes for sources of interest are ∼10−3 phot cm−2 s−1 keV−1),early experiments required large area detectors with very high efficiency forphoton detection. Gas proportional counters were the instruments of choice.In the soft X-ray band, the spectral resolution achievable with such devicesis extremely limited: E/∆E ∼ few. While the data obtained with those ex-periments did provide some measure of the overall shapes of cosmic X-rayspectra, they could not be used to derive any real constraints on physicalconditions in source emission regions.

The situation improved significantly in the mid 1990’s with the launchof the ASCA Observatory. This was the first mission to incorporate charge-coupled device (CCD) detectors at the focus of an X-ray telescope. The en-ergy resolution of CCDs is roughly an order of magnitude better than thatachievable with proportional counters. That enabled the detection of broad“humps” in the spectra, which could loosely be identified with complexesof emission lines from particular ions. Yet detailed spectral constraints couldstill only be derived from model fits to the spectra – even CCD resolution wasinsufficient to allow for direct interpretation of the intensities of individualfeatures. Hence, the true power of spectroscopy had still not been realized.

Shortly before these lectures were delivered, however, the National Aero-nautics and Space Administration launched the Chandra X-ray Observa-tory (June 1999), and the European Space Agency launched the XMM-Newton Observatory (December 1999). These two magnificent space mis-sions both incorporate diffraction grating spectrometers, with resolving pow-ers E/∆E ≥ 200 across most of the soft X-ray band. They have collectivelyprovided the first high resolution X-ray spectra of a wealth of astrophysicalsources. This has created a revolution in this field, whose significance, even asof this writing two years later, is still continuing to be appreciated. In somecases, the data have provided striking confirmation of existing astrophysical

Page 3: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 5

models. In others, they have presented significant challenges to our basicunderstanding of the sources involved.

Why is soft X-ray spectroscopy an important tool for astronomy?There are several unique features of the soft X-ray band that play a role.

First, X-ray emitting gas is often the “key” component of the astrophysicalsystem. For many objects (e.g. elliptical galaxies, clusters of galaxies), the vir-ial temperature, kT ∼ GMmp/R, lies in the range 106–108 K, where most ofthe emission comes out at soft X-ray energies. In others (e.g. supernova rem-nants, binary sources), shocks heat gas into the same temperature regimes.Second, the conventional soft X-ray band (0.1–10 keV) is unusually rich indiscrete spectral features. The K-shell transitions of carbon through iron, andthe L-shell transitions of silicon through iron fall in this range. In contrastto other wave bands, all charge states are visible in a single X-ray spectrum.This makes the interpretation of the spectrum relatively unambiguous. Forexample, one can derive relative elemental abundances without invoking anyassumptions about the thermal state of the gas. Finally, because of the highradiative decay rates of X-ray transitions, astrophysical emitting plasmas aregenerally not in local thermodynamic equilibrium. This means that the de-tails of the observed spectra depend on the explicit mechanisms by which thelevels are populated. While that can occasionally lead to complications inthe interpretation of the data, it also implies that they are quite sensitive tophysical conditions in the source. Hence, X-ray spectra have high diagnosticutility.

Astrophysical X-ray spectroscopy can also be of interest as a probe offundamental physics issues in unusual environments. In particular, cosmicplasmas can achieve extremely low densities, ne < 10−3 cm−3, orders of mag-nitude below the densities found in the best vacuum obtainable in a labora-tory. At such low densities, radiative decays from very long-lived metastablelevels are important. In addition, the time scales for equilibration can bevery long in comparison to the lengths of our observations. This means thatfor some sources, the emitting plasmas appear “frozen” in non-equilibriumstates. Finally, given the vast physical scales characteristic of astronomicalsystems, we can find interesting examples of non-negligible optical depth forexotic absorption and scattering processes.

1.2 Characteristics of Cosmic X-ray Sources

An extensive review of the general science of X-ray astronomy is well beyondthe scope of these lectures. However, I believe it is useful, in this introduc-tory chapter, to provide a very brief accounting of physical conditions in thevarious types of cosmic X-ray sources we are studying with our spectroscopicexperiments. More complete discussions of all of these topics can be found ina series of conference proceedings that have appeared within the past year [4].

Page 4: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

6 S.M. Kahn

General introductions to X-ray astronomy, suitable for non-specialists, havebeen written recently by Schlegel [5] and Tucker & Tucker [6].

Late-Type Stars

X-ray emission from late-type stars (stars of spectral type F, G, K and M) isbelieved to be produced in coronae, tenuous collections of hot gas confined bymagnetic field lines above the stellar photospheres. The best known example,of course, is the solar corona, which was first detected in X-rays by a rocketexperiment in 1951. The X-ray luminosity of the quiescent solar corona is∼2 1027 erg s−1, which is only of order a part in a million of the total solarluminosity. The characteristic temperature is ∼2 106 K, and the characteristicelectron density is ∼109 cm−3. However, the Sun turns out be a rather weakX-ray source. More active late-type stars exhibit X-ray luminosities as high as1032 erg s−1, with temperatures ∼several 107 K, and densities that can reach1014 cm−3. Coronal plasmas are optically thin to photoelectric absorption,although line optical depths for the highest oscillator strength lines can begreater than unity. Most active stars exhibit flares, which can increase theluminosities by three to four orders of magnitude on time scales of minutes tohours. There are many issues associated with the formation and energizationof stellar coronae that are still poorly understood, making this an active areaof research.

Early-Type Stars

Massive early-type stars (spectral types O and B) do not possess the outerconvective zones believed to provide the dynamo necessary to generate stellarcoronae. On the other hand, these stars possess massive, radiatively drivenstellar winds, with mass loss rates ∼10−6 M per yr. X-ray emission fromthese systems is believed to arise in shocks in the wind, driven by inhomo-geneities resulting from both thermal and dynamical instabilities. Typical X-ray luminosities are ∼1031 erg s−1, with characteristic temperatures ∼ several106 K. Since the wind is dense (ne ≥ 1011 cm−3), and far from fully ionized,the overlying photoelectric opacity is significant. However, the shocks are be-lieved to be distributed throughout the wind, so the absorption structure canbe quite complex. Emission lines arising from this gas should exhibit velocitybroadening with characteristic velocity widths of several 103 km s−1.

Supernova Remnants

Supernovae are cataclysmic stellar explosions which drive high temperatureblast waves in the surrounding interstellar medium. There are essentially twovarieties. Type 2’s which result when massive stars exhaust their nuclear fueland implode, and Type 1a’s which result when white dwarf stars accrete mate-rial from their binary companions, causing their masses to exceed the critical

Page 5: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 7

“Chandrasekhar limit” (∼1.4 M) – the maximum allowable for hydrosta-tic stability. In both cases, ∼1050–1051 erg of kinetic energy is transferred tothe outer layers of the star, which expand into the neighboring environment.Shock waves form in both the stellar ejecta and the surrounding interstellargas, with initial temperatures ∼ few 108 K. These shocks radiate brightlyat X-ray energies for ≥ 104 yr. As the remnant expands, the temperaturedrops, roughly as the third power of the radius. The X-ray emitting gas canhave a range of densities, from 10−2 to 101 cm−3. At such low densities, thetime scale for ionization balance to be achieved can exceed the age of theremnant, implying that the plasma may be well out of equilibrium. Despitethe very large length-scales involved, the density is low enough that the gasis optically thin to both line and continuum radiation.

X-ray Binaries

Nearly half of all stars in the sky are in binaries, i.e. gravitationally boundtwo-star systems. Stars of higher mass evolve faster, eventually collapsing toform a “compact object” (white dwarf, neutron star, or black hole). Hence,binary systems can form where one member is compact, and the other isrelatively normal. If the binary separation is sufficiently short, these systemsexhibit mass transfer, wherein the normal star loses mass that is subsequentlyaccreted by the compact companion. Infall into the deep gravitational poten-tial well characteristic of a compact star, shocks the accreting material up tohigh temperatures, causing these systems to be copious X-ray emitters. Fora white dwarf, ∼10−4 of the rest mass energy of the accreting matter maybe released in the form of X-radiation. For a neutron star or black hole, thefraction can be much higher, approaching 20%, leading to X-ray luminositiesas high as 1038 erg s−1.

There are two possible modes of mass transfer. If the companion is anearly type star, it may have a significant stellar wind, some of which will begravitationally captured by the compact object. Such systems are called “highmass X-ray binaries” (HMXBs). On the other hand, if the companion has lowmass, but expands to fill the critical equipotential surface that connects tothe other star (the so-called Roche lobe), material can flow freely through aninner Lagrange point and fall inward to the compact star. This process iscalled Roche lobe overflow, and the resulting X-ray sources are called “lowmass X-ray binaries” (LMXBs). If the compact star is a white dwarf, insteadof a neutron star or black hole, the source is called a “cataclysmic variablestar” (CV).

The fate of the accreting material is not well understood, and probablyvaries from source to source. Since it is far easier to dissipate energy thanangular momentum, it is thought that the flow should settle into a thinaccretion disk, with the matter moving in near Keplerian orbits. Some form ofviscous interaction between neighboring “rings” allows angular momentum tobe transferred out, thereby enabling accretion to proceed, either continuously,

Page 6: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

8 S.M. Kahn

or episodically. Most of the X-radiation is released down near the surface ofthe compact star (or in the case of a black hole, near the event horizon).Because the material is nearly fully ionized, and the Compton depths arenon-negligible, the emergent flux is radiated primarily as a continuum, withcharacteristic photon energies of order a few keV. However, the transfer of thisintense continuum outward through the circumsource medium can generate awealth of discrete features. The irradiated environment is likely to be severelyphotoionized, with the energy density in the radiation field nearly four ordersof magnitude higher than the thermal kinetic energy of the gas.

Active Galactic Nuclei

The term “active galactic nucleus” (AGN) refers to an intense source ofradiation emanating from a compact nuclear region at the center of a galaxy.The first “quasi-stellar objects”, or quasars, were discovered in the early1960’s. Spectra of these sources indicated significant redshifts, implying largedistances, and thus very high luminosities, comparable to that of an entiregalaxy. In addition, observed short-term variations in the emission suggestedthat the emitting regions must be compact, with characteristic dimensionscomparable to the size of our solar system.

There is a rich variety of empirical phenomena associated with AGNs,leading to the definition of numerous “classes”, however it is generally be-lieved that most of these can be understood in terms of a grand unified model,wherein a supermassive accreting black hole is surrounded by an obscuringtorus of optically thick material, oriented in the equatorial plane. Accretiononto the black hole generates X-radiation, as well as relativistic jets alongthe spin axis. If our line of sight is oriented above the plane of the torus,we get a direct view of the black hole and the source is bright in X-rays.Such systems are called Seyfert 1 galaxies if they are radio-quiet, or Type 1quasars if they are radio-loud. If our line of sight is oriented along the plane,the central source is obscured, and the soft X-ray emission we see is mainlyreprocessed radiation emanating from the circumsource environment. Theseare called Seyfert 2 galaxies (radio-quiet), or Type 2 quasars (radio-loud).Finally, if our line of sight is oriented along the jet, the observed emission isgreatly enhanced by relativistic beaming. These systems are called BL Lacobjects, or more generally, “blazars”.

Our understanding of the accretion process in AGNs is even less well-developed than for X-ray binaries. However, it is believed that similar phys-ical processes must be involved. There is some evidence for the existenceof relativistically broadened X-ray emission lines in these systems, whichcould be produced in the inner most regions of the accretion disk aroundthe black hole. If this interpretation is correct, X-ray spectroscopy of AGNsmay provide us with one of our best observational handles on the physics ofultra-strong gravitational fields. For the Seyfert 2 systems, the obscuration

Page 7: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 9

of the central source affords a relatively “clean” view of the surrounding pho-toionized gas. Soft X-ray spectra of these systems are rich in discrete spectralfeatures.

Clusters of Galaxies

Clusters are massive collections of galaxies that have formed relatively re-cently via gravitational collapse as the universe has expanded. They aregravitationally bound systems, with most of the mass in the form of darkmatter that only interacts weakly (if at all) with ordinary baryonic matter.The richest, most evolved clusters contain hundreds of members, centered ona central dominant galaxy. The intracluster medium is filled with hot gas,in rough hydrostatic equilibrium with the dark matter gravitational poten-tial. Characteristic temperatures are in the range 107–108 K, so that the gasradiates mostly at X-ray energies. Electron densities are ∼10−3 cm−3, andtypical X-ray luminosities lie in the range 1043–1045 erg s−1. Even at suchlow densities, the cooling timescales appropriate to this gas are often signif-icantly less than the age of the system, especially at the cluster core. Thisleads to the expectation that gas should continually be cooling out of thismedium, perhaps eventually forming stars in the central galaxy. Curiously,however, recent X-ray spectra suggest a deficit of low temperature gas pre-dicted by this scenario. The intracluster media should be mostly opticallythin to continuum absorption, but can exhibit non-negligible optical depthfor scattering of bright emission lines.

2 Classical and Quantum Radiation Theory

2.1 Introduction

In this chapter, I review the essential components of classical and quantumradiation theory. I assume that most of the material will be very familiar tothe reader from undergraduate (and perhaps graduate) coursework in elec-trodynamics and elementary quantum mechanics. Nevertheless, I believe it isuseful to offer this quick review so that we have the relevant formulae readyat hand for reference in later discussions. You might expect that classical ra-diation theory should find very limited application in a discussion of discreteradiation from atoms, but as I will show, it does provide a quick means ofderiving order of magnitude estimates for a number of important processes.In addition, I find it pedagogically useful to discuss the classical and quantumformulae in a unified context. This is rarely done in textbooks, which makesit difficult to follow where and when quantum ideas are important.

In this, and all subsequent chapters, I utilize the CGS system of units.Although this is going out of fashion in most fields of physics (where SI unitshave indeed become standard), it is still common practice in astrophysics.

Page 8: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

10 S.M. Kahn

In addition, the fundamental equations of radiation theory take on a simpleand more elegant form in CGS units. In this system, the unit of charge is theesu, defined such that Coulomb’s Law for the attraction between two pointcharges, q1 and q2 is:

F (r) =q1q2r2

r , (1)

where r is the vector separation between them and r the unit vector pointingin the r direction. Thus, 1 (esu)2 = 1 erg cm. In this system, the electric andmagnetic fields, E and B, have the same units, usually expressed as gauss.

2.2 Overview of the Classical Equations

We start with the governing equations of electromagnetism, specificallyMaxwell’s equations:

∇ · E = 4π , ∇ · B = 0 , (2)

∇ × E = −1c

∂B

∂t, ∇ × B =

4πc

j +1c

∂E

∂t. (3)

which relate the spacetime derivatives of the electric and magnetic fields toeach other and to the charge and current densities of the medium, and j,respectively. In addition, the Lorentz force law:

f = E +1cj × B (4)

relates the force density on a charged volume, f , to the fields and the chargeand current densities. Due to the conservation of electric charge, and j obeya continuity equation:

∂t+ ∇ · j = 0 . (5)

It is useful to define also the scalar and vector potential functions, ϕ and A,respectively, which are related to the fields by:

B = ∇ × A , (6)

E = −∇ϕ− 1c

∂A

∂t. (7)

Equations (6) and (7) do not define ϕ and A uniquely. To make the definitionsunique, we need to further specify a gauge. For radiation theory, it is mostconvenient to adopt the Lorentz gauge:

∇ · A +1c

∂ϕ

∂t= 0 (8)

Substitution into Maxwell’s equations yield:

Page 9: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 11(∇2 − 1

c2∂2

∂t2

)ϕ = −4π , (9)(

∇2 − 1c2

∂2

∂t2

)A = −4π

cj . (10)

which relate the potentials to the charge and current densities. These equa-tions have solutions of the form:

ϕ(r, t) =∫

d3r′dt′(r′, t′)| r − r′ |δ [t′ − tr(r, t, r′)] , (11)

A(r, t) =∫

d3r′dt′j(r′, t′)| r − r′ |δ [t′ − tr(r, t, r′)] (12)

where tr is the retarded time, defined by:

tr ≡ t− | r − r′ |c

. (13)

Differentiation of the right-hand sides of (11), (12) according to (6), (7) yieldsthe electric and magnetic fields associated with arbitrary time-varying chargeand current distributions. We return to this shortly.

2.3 Electromagnetic Waves

In charge-free space, Maxwell’s equations (2), (3) give rise to wave equationsfor both the electric and magnetic fields:(

∇2 − 1c2

∂2

∂t2

)E = 0 , (14)(

∇2 − 1c2

∂2

∂t2

)B = 0 (15)

which have plane-wave solutions written in the form:

E = E0ei(k·r−ωt) , (16)

B = B0ei(k·r−ωt) (17)

where

ω = kc , (18)k · E0 = k · B0 = 0 , (19)

k × E0 =ω

cB0 , (20)

k × B0 = −ω

cE0 , (21)

k = E × B , (22)

Page 10: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

12 S.M. Kahn

the waves are transverse, and the fields have equal magnitudes. The first(18) requires that electromagnetic waves travel at the speed of light in thevacuum. The energy flux associated with electromagnetic waves is given bythe Poynting vector:

S =c

4πE × B (23)

which has units of erg cm−2 s−1 in the CGS system. For the plane-wave so-lutions (16), (17), (18)–(22) imply that the real part of the Poynting vectoris given by:

S(t) = kc

4π| E(t) |2 . (24)

The plane waves described above are monochromatic. Since the wave equa-tions are linear, however, arbitrary linear combinations of plane-waves alsoprovide solutions. In general, we are interested in the frequency dependenceof the radiation, which can be assessed by taking the Fourier transform ofthe electric field:

E(ω) ≡ 12π

∫ ∞

−∞E(t)eiωtdt . (25)

Parseval’s Theorem for Fourier transforms requires:∫ ∞

−∞| E(t) |2 dt = 2π

∫ ∞

−∞| E(ω) |2 dω = 4π

∫ ∞

0

| E(ω) |2 dω (26)

(Since E(t) is real, E(−ω) = E(ω)). The energy in the radiation per unitarea is given by:

dWdA

=∫ ∞

−∞S(t)dt = c

∫ ∞

0

| E(ω) |2 dω (27)

so the energy per unit area per unit frequency is:

dWdAdω

= c | E(ω) |2 . (28)

2.4 The Classical Multipole Expansion

As for the electric field, we can take the Fourier transform of the charge,current density and vector potential:

(r, ω) ≡ 12π

∫ ∞

−∞(r, t)eiωtdt , (29)

j(r, ω) ≡ 12π

∫ ∞

−∞j(r, t)eiωtdt , (30)

A(r, ω) ≡ 12π

∫ ∞

−∞A(r, t)eiωtdt . (31)

Page 11: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 13

Using (12) and (13) we get:

A(r, ω) =1c

∫d3r′ j(r′, ω)eik|r−r′|

| r − r′ | (32)

where k = ω/c. If we are interested in the character of the radiation far fromthe charge distribution, then | r || r′ |, so that | r − r′ |≈ r − n · r′, wheren is the unit vector pointing in the direction r. We thus obtain:

A(r, ω) ≈ eikr

rc

∫d3r′j(r′, ω)e−ik(n·r′) . (33)

The classical multipole expansion involves a Taylor expansion of the complexexponential inside the integral in (33), assuming that k(n · r′) 1. To seewhy this might be valid, note that:

k(n · r) ∼ Rω/c ∼ v/c (34)

where R is the characteristic dimension of the charge distribution, and v is acharacteristic velocity of the oscillating charge. Thus the multipole expansionis justified in the limit that the charge motions are non-relativistic. The lowestorder term is obtained by setting the complex exponential equal to unity. Weare then left with a simple integral of the Fourier transform of the currentdensity which can be rewritten in terms of the dipole moment of the chargedistribution, d(ω)∫

d3r′j(r′, ω) = −∫

d3r′(∇′ · j)r′ = −ikc∫

d3r′(r′, ω)r′

≡ −ikcd(ω) . (35)

We thus refer to this term as the electric dipole or (E1) term. Expressions forE and B in the electric dipole limit can be found by taking the appropriatederivatives of A. Then, converting back to the time domain, we obtain:

E(r, t) =1c2r

[n× (n× d(tr))] (36)

where d(tr) is the second time-derivative of the electric dipole moment eval-uated at the retarded time. The Poynting vector is:

S =c

4π| E |2 n =

| d |2 sin2θ

4πr2c3n (37)

Integrating over the surface of a sphere of radius r yields the total energyradiated per unit time:

dWdt

=∫

Sr2dΩ =23| d |2c3

. (38)

Page 12: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

14 S.M. Kahn

For a single, accelerating point charge, this reduces to the well-known Larmorformula:

dWdt

=23q2a2

c3(39)

where a is the acceleration of the charge.For non-relativistic motions, the electric dipole term will dominate when-

ever it is non-zero. If it is zero, however, the next highest term will be im-portant. In that case, the relevant integral in (33) is:∫

d3r′j(r′, ω)(n · r′) (40)

which can be “broken” into two parts:

j(r′, ω)(n · r′) =12

[j(n · r′) − r′(j · n)

]+

12

[j(n · r′) + r′(j · n)

](41)

The integral of the first term on the right-hand side of (41) can be shown tobe related to the magnetic dipole moment of the current distribution:

µ(ω) ≡ 12c

∫d3r′r′ × j(r′, ω) (42)

while the integral of the second term is related to the electric quadrupoletensor of the charge distribution:

Q(ω) ≡∫

d3r′[3r′r′ − r′2I

](r′, ω) (43)

where I is the identity tensor. The radiated power for the magnetic dipoleterm is:

dWdt

=23| µ |2c3

. (44)

For the electric quadrupole term, the integral over solid angle depends on theexplicit form of the quadrupole tensor, but the radiated power is proportional

to |...

Q|3 /c5. Note that for an oscillating charge distribution:

| d |∼ qRω2 , | µ | ∼ R(qvc

)ω2 (45)

and|...

Q| ∼ qR2ω3 . (46)

Taking Rω ∼ v, we find:

dWdt

∼ q2ω

R

(vc

)3

(E1) , (47)

dWdt

∼ q2ω

R

(vc

)5

(M1) , (48)

dWdt

∼ q2ω

R

(vc

)5

(E2) . (49)

Page 13: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 15

So the (M1) and (E2) terms are of the same order and are both down fromthe (E1) term by a factor ∼(v/c)2, where v is a characteristic velocity of thecharges.

2.5 The Classical Oscillator

An important application of classical radiation theory, and one that provesuseful in understanding radiation from atoms, is the classical harmonic os-cillator, in which the acceleration of the charge is given by:

x = −ω20x (50)

where x is the position of the oscillating charge, and ω0 is the oscillationfrequency. However, since an oscillating charge radiates energy, there mustbe a damping force associated with the radiation, which gradually reduces theamplitude of the oscillation to zero. This is called the “radiation reaction”.We can approximate it by noting that the power dissipated by the drag forcemust agree with the Larmor formula for the radiated energy. That yields:

F drag ≈ 23q2

c3...x≈ 2

3q2

c3ω2

0x (51)

so the equation of motion becomes:

x + Γ x + ω20x = 0 (52)

where

Γ =23q2ω2

0

mc3, (53)

and m is the mass of the charge. The solution has the form:

x(t) = x0e−Γt/2cos(ω0t+ ϕ) , (54)

and thus the Fourier transform of the electric dipole moment (d(t) = qx(t)),becomes:

| d(ω) |2 = q2(x0

)2 1(ω − ω0)2 + (Γ/2)2

. (55)

The radiated spectrum in the (E1) limit is given by:

dWdω

=8π3ω4

c3| d(ω) |2 ≈

(12k | x0 |2

)[Γ/2π

(ω − ω0)2 + (Γ/2)2

]. (56)

Here 1/2 k |x0 |2 (k ≡ mω20 is the “spring constant” of the oscillator) is

the initial energy of the oscillation, and the term in brackets describes aLorentzian line profile, centered at ω0, with width equal to Γ . Note that:

∆ν =12π

∆ω =4π3q2ν2

0

mc3(57)

Page 14: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

16 S.M. Kahn

and

∆λ =c

ν20

∆ν =4π3

q2

mc2(58)

is a constant, independent of frequency or wavelength. This is the classi-cal natural line width for electric dipole transitions. For an electron, ∆λ ≈1.2 10−4 A. For the soft X-ray transitions we are concerned with in this lec-ture, λ ≈ 1 − 100 A, so the natural width is nearly always a very smallcomponent of the line broadening. The time-averaged radiated power of theclassical oscillator is:

dWdt

=13q2ω4

0 | x0 |2c3

. (59)

Since the initial energy, W0 = 1/2 k |x0 |2 = 1/2 mω20 |x0 |2, the classical

radiative decay rate is given by:

Acl ≡ (dW/dt)W0

=23q2ω2

0

mc3(60)

which turns out to be equal to the damping constant, Γ . In terms of thelinear frequency:

Acl =8π2

3q2

mc3ν2 ≈ 2.5 10−22ν2 s−1 (61)

for an electron, where ν is in Hz. Note that in the X-ray band, where ν ≈ 1016–1018 Hz, radiative decay rates are extremely fast, ∼1010–1014 s−1. This hasan important effect on level populations for X-ray emitting plasmas, as wewill see later.

The discussion above pertains to spontaneous emission. To model inducedprocesses, like photoexcitation, we must consider driven oscillations, wherethere is an applied force due to an incoming wave, given by: F appl = qE0eiωt.The equation of motion is now that of a damped, driven harmonic oscillator.The time-averaged radiated power for this case becomes:

dWdt

=q4 | E0 |2

3m2c3ω4

(ω2 − ω20)2 + (Γω0)2

. (62)

Since the time-averaged incident energy flux in the wave is <S>= c/8π|E0|2,the cross-section for scattering is:

σ(ω) =dW/dt< S >

=8π3

q4

m2c4ω4

(ω − ω0)2 + (Γ/2)2. (63)

In the vicinity of line center: ω2 − ω20 ≈ 2ω0(ω − ω0), so this becomes:

σ(ω) =2π2q2

mc

Γ/2π(ω2 − ω2

0)2 + (Γω0)2(64)

Page 15: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 17

where we have used our earlier expression for Γ . The scattering cross-sectionagain has the Lorentzian line profile with width in angular frequency equalto Γ . Integrating over frequency yields:∫ ∞

0

σ(ω)dω = 2π∫ ∞

0

σ(ν)dν =2π2q2

mc(65)

so

σ(ν) =πq2

mcϕ(ν) . (66)

where ϕ(ν) is the normalized line shape (it may have other components asso-ciated with Doppler broadening, etc.). Note that the coefficient is independentof frequency. For an electron, it has the value: πe2/mc = 2.7 10−2 cm2 Hz.

2.6 Quantum Radiation Theory – Overview

We now turn to the quantum theory. There are two fundamental differencesbetween the classical and quantum treatments of the interaction betweenradiation and matter:

– In quantum mechanics, charge configurations are expressed in terms ofquantum “states”. Radiative interactions involve an exchange of energyand momentum, so they are associated with a change of state. The onlystationary quantum states are the eigenstates of the Hamiltonian, whichis the operator associated with the energy of the system. The rates for var-ious processes therefore involve quantum “matrix elements” of the form〈f | Hrad | i〉, where f represents the final state, i the initial state, andHrad is the perturbing Hamiltonian associated with the radiation field. Inthe classical picture, charges radiate when they are accelerated. Acceler-ation requires an external applied force, which can be identified with theperturbing Hamiltonian.

– In the quantum treatment, the radiation field is described in terms ofdiscrete particles or “photons”. The energy of an individual photon is E =ω = hν, where h is Planck’s constant, and has the value 6.626 10−27 erg s.The momentum of a photon is given by p = k = (ω/c)k, directed alongthe direction of propagation. Photons are spin 1 particles, and thereforethe emission or absorption of a photon changes the angular momentumof the system by one unit of .

The key rates and cross-sections for various radiative processes follow fromtime-dependent perturbation theory. We begin with the time-dependentSchroedinger equation:

H | ψ〉 = i∂ | ψ〉∂t

. (67)

The energy eigenstates satisfy:

H | ψE〉 = E | ψE〉 (68)

Page 16: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

18 S.M. Kahn

and therefore have a time dependence given by:

| ψE(t)〉 =| ψE〉e−iEt/ . (69)

Let the total Hamiltonian contain a dominant “unperturbed part” and asmall additional “perturbing part”:

H = H0 +H ′ (70)

and let | n〉 represent a complete set of energy eigenstates of H0. An arbitrarystate | ψ(t)〉 can be expanded in terms of these energy eigenstates:

| ψ(t)〉 =∑

n

an(t) | n〉e−iEnt/ . (71)

Substituting into (67) and taking the scalar product with a specific energyeigenstate 〈k | to both sides, then yields the differential equation:

i∂ak

∂t=∑

n

an〈k | H ′ | n〉eiωknt (72)

where ωkn ≡ (Ek − En)/. Here, we have used the fact that the energyeigenstates are orthonormal: 〈k | n〉 = δk,n. Suppose the system is initiallyin state “m”, so that ak(0) = δk,m. Then, to lowest order in the perturbingHamiltonian, the coefficients ak at some later time are given by:

ak(t) = (i)−1

∫ t

0

〈k | H ′(t) | m〉eiωkmtdt . (73)

For application to radiation theory, we are interested in perturbations whichare oscillatory in time:

H ′(t) = H ′e±iωt (74)

where ω is some angular frequency. Thus:

ak(t) = (i)−1〈k | H ′ | m〉∫ t

0

ei(ωkm±ω)tdt . (75)

The probability at time t that the system has made the transition from “m”to state “k” is given by | ak(t) |2. The transition rate, R, is thus given by:

R = limt→∞| ak(t) |2

t=

2π2

| 〈k | H ′ | m〉 |2 δ(ωkm ± ω) (76)

=2π

| 〈k | H ′ | m〉 |2 δ(Ek − Em ± ω) . (77)

This last expression indicates that the transition is possible only if the changeof state is accompanied by the emission or absorption of a single photon withenergy equal to the energy difference between the states. Note that this is afirst order perturbation result. Multi photon processes occur via higher orderterms in the perturbation expansion.

Page 17: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 19

2.7 The Radiation Hamiltonian

The appropriate Hamiltonian to use for the interaction between charged par-ticles and electromagnetic fields is derived from the formalism of classicalmechanics. Defining a Lagrangian of the form:

L =12mv2 − qϕ+

q

cA · v (78)

where ϕ and A are the classical scalar and vector potentials, respectively,and applying Lagrange’s Equation:

ddt

(∂L

∂r

)=∂L

∂r, (79)

we arrive at the desired Lorentz force law for the electromagnetic force on asingle charge:

F ≡ mv = qE +qv

c× B . (80)

The canonical momentum of the particle is defined by:

p ≡ ∂L

∂r= mv +

qA

c. (81)

The Hamiltonian is then:

H ≡ p · v − L =12mv2 + qϕ =

12m

(p − q

cA)2

+ qϕ . (82)

It is the canonical momentum p that we associate with the quantum mechan-ical operator (/i)∇. Substituting into (82) yields:

H = − 2

2m∇2 +

iq

mc(∇ · A) +

iq

mc(A · ∇) +

q2

2mc2A2 + qϕ . (83)

For an electromagnetic wave, ϕ = 0, and therefore, in the Lorentz gauge,∇ ·A = 0. The term involving A2 is small compared to the first order termsin A, so we ignore it. In addition, there may be a non-radiation potentialV (r), e.g. the binding potential of the atom. In that case:

H = − 2

2m∇2 + V (r) +

iq

mc(A · ∇) (84)

The first two terms on the right hand side are usually taken to be the unper-turbed Hamiltonian. The perturbing Hamiltonian associated with the inter-action with radiation is given by the third term. For a strictly monochromaticwave, we can write the vector potential in the form:

A(r, t) = Re[A0ei(k·r−ωt)

]=

12|A0|

[εei(k·r−ωt) + ε∗e−i(k·r−ωt)

](85)

Page 18: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

20 S.M. Kahn

where A = |A0| ε, and ε is the polarization vector of the wave. From (7) and(24), we find that the time-averaged Poynting vector is given by:

< S > =ω2

8πc| A0 |2 k (86)

Recall that <S > represents the energy flux of the radiation. If we think interms of discrete photons, the photon flux, dN/dAdt, is given by:

dNdtdA

=|< S >|

ω=

ω

8πc| A0 |2 . (87)

Note from (85) that the perturbing Hamiltonian in (84) has two pieces, oneproportional to e−iωt, and one proportional to e+iωt. The former leads to theabsorption of a photon (Ek = Em+ω), while the latter leads to the emissionof a photon (Ek = Em − ω). For a given set of initial and final states, onlyone of the two terms can satisfy energy conservation, so we can treat themseparately. The expression for the transition rate between initial state i andfinal state f is thus:

R =2π

q22

4m2c2| A0 |2

∣∣∣⟨f | e±ik·r ε(∗) · ∇ | i⟩∣∣∣2 δ(Ef − Ei ∓ ω) (88)

where the top sign corresponds to absorption (with ε in the matrix element)and the bottom sign corresponds to emission (with ε∗ in the matrix element).

2.8 Bound-Free Absorption (Photoionization)

Consider first the application to bound-free absorption, where the initial stateof an electron is a bound state in an atom, and the final state is that of a freeparticle. To get the total transition rate, we must integrate over all possiblefinal states. For a free particle, the states are characterized by the momentumvector p. However, the uncertainty principle requires that a particle cannot belocalized in a 6-dimensional phase space cell smaller than d3rd3p = (2π)3.Therefore, the density of states for a free particle is given by:

(E)dE =V d3p

(2π)3=V m(2mE)1/2dEdΩ

(2π)3(89)

where V is the allowable volume for the free particle (it will drop out ofthe later expression), dΩ is a differential element of solid angle, and we haveassumed non-relativistic dynamics. The free particle final state of the chargecan be represented by:

ψf (r) = V −1/2eipf ·r/ (90)

where the coefficient has been introduced for normalization, i.e. 〈ψf | ψf 〉 = 1when the integration is performed over the allowable volume.

Page 19: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 21

Taking (2mEf )1/2 = mvf , and integrating over energy in (88), we obtain:

dR =14

q2

(2πc)2vf | A0 |2

∣∣∣⟨e−ipf ·r/ | eik·r ε · ∇ | ψi

⟩∣∣∣2 dΩ . (91)

The differential cross-section for this process is given by:

dσdΩ

=dR/dΩdNdtdA

=dR/dΩ

ω | A0 |2 /8πc(92)

=q2

2πc

νf

ω

∣∣∣⟨e−ipf ·r/ | eik·r ε · ∇ | ψi

⟩∣∣∣2 . (93)

Actually, this expression is an approximation to the real photoionizationcross-section because the liberated electron is not really “free” – it still feelsthe Coulomb attraction to the nucleus. A more accurate treatment woulduse a true continuum wave-function for the electron subject to the atomicpotential. We will come back to this later.

2.9 Bound-Bound Transitions

In the case of bound-bound transitions, which give rise to emission or ab-sorption lines, both the initial and final states are discrete. Equation (88)indicates that if the incoming wave is perfectly monochromatic, then thetransition rate will be infinite if ω = | Ef − Ei |, and zero otherwise. Toderive a meaningful cross-section, we must integrate over a finite spectrumof the incident radiation field. This is characterized by a continuum photonflux, dN/dtdAdω. Setting:

| A0 |2= 8πc

ω

dNdtdAdω

dω (94)

in (88) and integrating over frequency, yields:

Ri→f =4π2q2

m2cωif

dNdtdAdω

(ωif ) | ⟨f | e±ik·r ε∗ · ∇ | i⟩ |2 (95)

where ωif ≡ | Ef − Ei | /. Here again the (+) sign corresponds to ab-sorption and the (−) sign to emission. The emission case is actually inducedemission, since the transition rate is proportional to the incident flux. Becausethe radiation Hamiltonian operator is Hermitian, the rates for emission andabsorption are identical (with the appropriate reversal of initial and finalstates).

Dividing the transition rate in (95) by the continuum flux yields a quantitywith units of cm2 Hz, which is the cross-section integrated over frequency:

Ri→f

dN/dtdAdw=∫σ(ω)dω = 2π

∫σ(ν)dν . (96)

Page 20: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

22 S.M. Kahn

This yields: ∫σ(ν)dν =

(πq2

mc

)2

mωif| ⟨f | eik·r ε · ∇ | i⟩ |2 . (97)

Notice that the term within parentheses is the classical expression we hadearlier (65). The remainder of the right hand side is the “quantum correction”to the classical result, and is called the oscillator strength, usually denotedby the symbol f :

fi→f ≡ 2

mωif| ⟨f | eik·r ε · ∇ | i⟩ |2 . (98)

2.10 The Quantum Multipole Expansion

The matrix element which appears in (98) involves the complex exponentialfactor: eik·r. This is reminiscent of the classical expression (33) where wefound it useful to expand this expression as a Taylor expansion in k · r.The logic in the quantum calculation is the same: k · r ≈ v/c, where vis the characteristic velocity of oscillating charges in the system. For non-relativistic motions, this is a small parameter. In the lowest order limit, theelectric dipole (E1) limit, we set the complex exponential to unity. The matrixelement becomes:

〈f | ε · ∇ | i〉 =i

ε · 〈f | p | i〉 (99)

where p is the momentum operator. Using the commutation relation:[p2, r]

= −2ip = 2m[H0, r

], we can rewrite this in the form:

〈f | p | i〉 =mi

〈f | [H0, r

] | i〉=mi

(Ef − Ei)〈f | r | i〉 = imωif 〈f | r | i〉 . (100)

The (E1) expression for the oscillator strength is therefore:

fi→f =2mωif

| ε · 〈f | r | i〉 |2 (101)

Averaged over polarization directions, this becomes:

fi→f =23mωif

| 〈f | r | i〉 |2 . (102)

A simple set of operator manipulations shows that the (E1) oscillatorstrengths satisfy a sum rule (the Thomas-Reiche-Kuhn sum rule):∑

f

fi→f = Z (103)

Page 21: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 23

where Z is the number of bound electrons in the atom. This provides auseful limit on the oscillator strengths for highly excited transitions, whichare numerous and therefore unwieldy to calculate. The next term in themultipole expansion has the form 〈f | (k · r)(ε · p) | i〉, which, as in theclassical case, can be broken into two pieces:

1/2〈f | (k · r)(ε · p) − (k · p)(ε · r) | i〉+1/2〈f | (k · r)(ε · p) + (k · p)(ε · r) | i〉 . (104)

The first term can be rewritten as:

(k × ε) · (r × p) ∼ µ · B (105)

where µ is the magnetic dipole moment of the orbiting electron. This is themagnetic dipole term (M1). For atomic transitions, we need to include bothorbital and intrinsic spin contributions to the magnetic dipole moment. Thesecond term above gives rise to electric quadrupole (E2) transitions. Hereagain, (M1) and (E2) transitions are of the same order in v/c. The (E1) termalways dominates unless the matrix element of the position vector vanishesbetween the initial and final states. Transitions for which this is the case arecalled “electric dipole forbidden”, or simply “forbidden”. This condition givesrise to certain “selection rules” for (E1) transitions, which we discuss later inthe context of atomic structure. Transitions for which the expression in (98)vanishes to all orders in (k ·r) are called “strictly forbidden”. These can onlygo by two-photon decay.

2.11 Spontaneous Emission

The quantum theory summarized so far only works for induced transitions,where an external electromagnetic field is introduced as a perturbation. Thisis because the treatment is semi-classical, i.e. the radiation field is still mod-eled classically even though the radiating system is treated quantum mechan-ically. Spontaneous emission, in which a system in an excited state decayson its own by emitting a photon, does not occur in this picture because theinitial state involves no radiation field, so there is no perturbing Hamiltonian.The correct treatment of this process requires the quantization of the radi-ation field. That is straightforward, but too time-consuming to review here.However, another form of semi-classical argument can be invoked to derivewhat turns out to be the correct result. In the (E1) limit, our classical expres-sion for the radiated power is given by (38). For an oscillator at a particularfrequency:

| d |2= ω4(| d(ω) |2 + | d(−ω) |2) = 2ω4 | d(ω) |2 . (106)

Using (35), we can write this in terms of the integrated current density:

Page 22: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

24 S.M. Kahn

| d(ω) |2 =1ω2

| j0 |2 (107)

where j0 ≡ ∫ d3r′j(r′, ω) Thus:

dWdt

=43ω2

c3| j0 |2 . (108)

In quantum mechanics, the charge density for a point charge is = q | ψ(r) |2.From the continuity equation (5) and the time-dependent Schroedinger equa-tion (67), it can be shown that the current density must be given by:

j = − iq

2m[ψ∗∇ψ − ψ∇ψ∗] . (109)

An appropriate “quantization” of the classical expression (108) can thus beobtained by setting:

| j0 |2 =∣∣∣∣∫

d3r′(−iq

2m

)[ψ∗

f∇ψi − (∇ψf )∗ψi

]∣∣∣∣2

(110)

=q2

m2| 〈f | p | i〉 |2 (111)

=q2ωif

2mfi→f (112)

where fi→f is the electric dipole oscillator strength of (102). The resultingexpression for the decay rate is then:

Ai→f =1

ωif

dWdt

=23q2ω2

if

mc3fi→f (113)

Comparison with (60) shows that this is simply the expression for the radia-tive decay rate of the classical oscillator multiplied by the absorption oscilla-tor strength.

3 The Structure of Multi-Electron Atoms

3.1 Introduction

This chapter is devoted to the structure of multi-electron atoms. This is avast and complex subject and time limitations will unfortunately prevent mefrom going into any real depth on most of the topics I will cover. My mainfocus will be on defining the relevant terms and outlining the basic principlesand approximations which are used in modern atomic physics calculations. Iwill not discuss computational techniques or the specifics of particular codes.

Page 23: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 25

Once again, I assume that much of this material is familiar to the reader fromundergraduate and graduate courses in quantum mechanics.

The physics of atomic structure basically involves the solution of the time-independent Schroedinger equation:

Hψ = Eψ (114)

whereH is the Hamiltonian operator,E the energy and ψ is the wave-functionfor the electrons in the atom, usually expressed as a function of spatial andspin coordinates. For all but the simplest atoms, this equation is not analyt-ically solvable and various approximation techniques are required. The mostcommon, and most general is time-independent perturbation theory, in whichone writes the Hamiltonian in terms of two parts:

H = H0 +H1 (115)

a zeroth-order Hamiltonian H0, which is amenable to direct solution and anadditional perturbation H1 which has much smaller amplitude. In first orderperturbation theory, the corrections to the energy levels due to the presenceof the perturbation are given by:

∆E(1)n =

⟨ψ(0)

n | H1 | ψ(0)n

⟩(116)

where ψ(0)n is the zeroth-order wave-function associated with the n-th energy

level, En, and the corrections to the wave-functions are given by

∆ψ(1)n =

∑k =n

⟨ψ

(0)k | H1 | ψ(0)

n

⟩E

(0)k − E

(0)n

ψ(0)k . (117)

The zeroth-order wave-functions are orthonormal by construction and theperturbed wave-functions remain orthonormal to lowest order in H1.

Another approach which is frequently used for more complex atoms isthe Ritz variational method. Its utility follows from the fact that the ex-pectation value of the Hamiltonian with respect to an arbitrary normalizedwave-function ψ, 〈ψ | H | ψ〉, is a minimum when ψ is the ground stateeigenfunction of H. Even more generally, if the functional 〈ψ | H | ψ〉 is sta-tionary with respect to perturbations in ψ, then ψ must be an eigenfunctionof H. Typically, one uses this method by choosing a form for a trial wave-function characterized by a set of adjustable parameters and then minimizingthe expectation value of the Hamiltonian with respect to those parameters.

3.2 Hydrogen-like Ions

We will begin the discussion with a quick review of the structure of hydrogen-like ions or one-electron atoms. Hydrogen-like ions are important for a num-ber of reasons. First, in the non-relativistic limit, the time-independent

Page 24: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

26 S.M. Kahn

Schroedinger equation is exactly solvable so we can get analytic expressionsfor all important quantities. Second, the “hydrogenic approximation” is oftenuseful for orders of magnitude estimates of rates for important processes andfor simple scaling laws with the nuclear charge Z. Finally, hydrogen-like ionsare quite important contributors to the soft X-ray emission from astrophysi-cal plasmas. Indeed, the brightest lines are usually Lyman series transitionsfrom hydrogen-like oxygen, neon, silicon and other low-Z elements.

The non-relativistic Hamiltonian for a single electron in an attractivecentral potential is given by:

H =p2

2me− V (r) . (118)

Making the usual substitution: p = −i∇ we get the relevant form of (114):(−

2

2me∇2 − V (r)

)ψ(r) = Eψ(r) . (119)

It is convenient to use atomic units where the natural unit of length is theBohr radius: a0 ≡

2/me2 = 0.529 10−8 cm, and the natural unit of energyis twice the Rydberg constant: e2/a0 ≡ 2Ry = 27.2 eV = 4.36 10−11 erg. Inthese units, e = = m = 1.

Equation (119) then takes the form:(12∇2 + E + V (r)

)ψ(r) = 0 . (120)

Equation (120) is spherically symmetric, so it is useful to write it in sphericalcoordinates. A spherically symmetric Hamiltonian commutes with the totalangular momentum operator l = r × p, which implies that eigenstates of Hare also eigenstates of l2 and lz. In spherical coordinates (120) becomes:(

12

[1r2

(r∂

∂r

)(r∂

∂r

)+

1r

∂r− l2

r2

]+ E + V (r)

)ψ = 0 . (121)

The only dependence on the angular coordinates (ϑ, ϕ) in this expression isthe l2 term. That implies that the equation is separable and ψ can be writtenas a product of radial and angular parts:

ψ(r, ϑ, ϕ) ≡ R(r)r

Y (ϑ, ϕ) . (122)

The eigenfunctions of l2 and lz are called spherical harmonics and have theform:

Ylm(ϑ, ϕ) ≡[(l− | m |)!(l+ | m |)!

2l + 14π

]1/2

(−1)(m+|m|)/2P|m|l (cosϑ)eimϕ (123)

Page 25: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 27

where Pml is the associated Legendre Polynomial. The spherical harmonics

obey the eigenvalue equations:

l2Ylm(ϑ, ϕ) = l(l + 1)Ylm(ϑ, ϕ) (124)

lzYlm(ϑ, ϕ) = mYlm(ϑ, ϕ) (125)

where l and m are integers, with −l ≤ m ≤ l.After substitution of (122) into (121), we are left with the radial equation:(

12d2

dr2+

1r

d

dr− l(l + 1)

2r2+ E + V (r)

)R(r)r

= 0 . (126)

For bound-states, E < 0, the solutions are discrete and are characterized byan integer index n called the principal quantum number. Bound-state wave-functions are only obtained for n ≥ l + 1, so for a given principal quantumnumber, the only allowed angular momentum states are l = 0, 1, 2, . . . , n− 1.The radial eigenfunctions are thus characterized by the two indices n and l.

For the particular case of the Coulomb potential V (r) = Z/r, (126) isexactly solvable, and leads to the radial wave-functions:

Rnl(r) = −(Z(n− l − 1)!n2[(n+ l)!]3

)1/2

e−ρ/2ρl+1L2l+1n+1 (ρ) (127)

where ρ ≡ 2Zr/n and L2l+1n+1 (ρ) are associated Laguerre polynomials. The

energy eigenvalues, in atomic units, have the form:

En =−Z2

2n2(128)

and are independent of l. This is a unique property of the Coulomb potential.The probability density of finding the electron in the radial range r →

r + dr is given by R2nl(r). Plots of this function for a few low order orbitals

are given in Fig. 1. Several key features of these radial wave-functions areimmediately apparent from the plots. First, most of the charge is concentratedin a spherical shell of moderate thickness, whose radius increases with n.This is expected classically, i.e. smaller binding energy is associated withlarger orbits. Note that for a given n, the radius of this shell decreases withincreasing l. Again, this is in line with classical expectations. For a fixedenergy, smaller angular momentum implies an elliptical orbit with highereccentricity, in which the electron spends most of its time further away fromthe nucleus. Finally, note that as r goes to zero, the probability density goesto zero for all but the l = 0 states. Hence only these states are appreciablyaffected by nuclear interactions.

Since the energy only depends on n for hydrogen-like ions, there are ndegenerate l states for each value of n, and 2l + 1 degenerate m states foreach value of l. In addition, the electron is a spin 1/2 particle, so there are

Page 26: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

28 S.M. Kahn

Fig. 1. Probability density to find the electron as a function of r (from Rybickiand Lightman, Fig. 9.1)

two degenerate spin states for each spatial state. The total degeneracy oflevel n is therefore given by:

gn = 2n−1∑l=0

(2l + 1) = 2n2 (129)

3.3 Scaling with Nuclear Charge

It is useful, at this stage, to look at the scaling of various quantities withthe nuclear charge Z. First note that the energy levels scale like Z2, whichimplies that the frequencies of key transitions also scale like Z2. The Lyman-αor n = 2 → 1 transition, specifically, has photon energy given by:

ωKα = (10.2 eV)Z2 . (130)

Page 27: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 29

Note that this line falls in the soft X-ray band (0.1–10 keV) for Z = 3-31, whichincludes the abundant elements: C(Z = 6), N(Z = 7), O(Z = 8), Ne(Z = 10),Si(Z = 14), S(Z = 16), Ar(Z = 18), Ca(Z = 20) and Fe(Z = 26). The energyof this line is only slightly affected by the presence of additional electrons.So (130) gives a rough idea of the energies of all K-shell feature transitionsdown to n = 1, for these and other elements.

Transitions down to n = 2 are called L-shell transitions. For hydrogen-likeions, the brightest is the Balmer-α transition corresponding to n = 3 → 2,whose energy is given by:

ωLα = (1.89 eV)Z2 (131)

Note that the L-shell transitions for Fe fall close to 1 keV, in the center ofthe soft X-ray band. These are especially important for diagnostic purposes,as we will review in a subsequent chapter.

Equation (127) implies that the scaling of the radial wave-function islike Z−1. Specifically, the characteristic size of hydrogen-like ions is givenroughly by a0/Z, where a0 is the Bohr radius we defined earlier. Recall from(102) that the oscillator strength for an E1 transition is proportional toωij | 〈f | r | i〉 |2. This scales like Z2Z−2, and thus is independent of Z.The radiative decay rates for E1 transitions are proportional to ω2f , so theyscale like Z4.

The Coulomb potential for a hydrogen-like atom is proportional to 1/r soclassically, the electron orbit obeys the Virial theorem, i.e. the kinetic energyis −1/2 times the potential energy:

12mv2 =

Ze2

2r.

For the ground-state:r a0

Z,

and thus:

v (Z2e2

ma0

)1/2

= (Zα)c (132)

where α ≡ e2/c 1/137 is the fine structure constant. We saw earlierthat the expansion parameter for both the classical and quantum multipoleexpansion (k · r) ∼ v/c, where v is a characteristic velocity of the system.For atomic transitions, we see that this parameter is ∼Zα. The magneticdipole and electric quadrupole terms are thus ∼(Zα)2 times smaller thanelectric dipole terms, so they scale like Z6. For low-Z abundant elements (C,N, O), (Zα) is indeed a small parameter. However for Fe, it is ∼0.2, so higherorder multipole terms are non-negligible and can often be important in thespectrum.

Page 28: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

30 S.M. Kahn

3.4 Relativistic Corrections

The time independent Schroedinger equation as expressed in (119) assumesnon-relativistic dynamics. For relativistic charges, one must use the Diracequation instead. However, since v/c ∼ Zα, atomic electrons are only mildlyrelativistic, even for iron which is the highest Z abundant element. Thus,it is sufficient to use (119) and to treat relativistic corrections as a simpleperturbation to the atomic structure.

To lowest order, there are three contributions to the relativistic correc-tions:

H11 = −1

8p4

m3ec

2(133)

which is the lowest order correction to the kinetic energy,

H12 =

12m2

ec2

(1r

dV

dr

)l · s , (134)

the spin-orbit term, which represents the magnetic interaction between themagnetic dipole moment of the electron associated with its intrinsic spin andthe magnetic field that it sees as it orbits in the electric field of the nuclearcharge, and

H13 =

2

4m2ec

2

(dV

dr

)∂

∂r, (135)

the so-called Darwin term, which is a relativistic correction to the potentialenergy produced by the non-localizability of the electron associated with itsrest mass energy.

For the Coulomb potential in hydrogen-like atoms, a simple first orderperturbation theory calculation using zeroth-order wave-functions yields theenergy shift:

∆En = +En(Zα)2

n2

[n

(j + 1/2)− 3

4

](136)

where j is the eigenvalue associated with the total angular momentum –specifically j(j + 1)2 is the eigenvalue of j2, where j = l + s. The fact thatthe perturbed energies depend on j is a consequence of the spin-orbit term,which is proportional to the operator:

l · s =12(j2 − l2 − s2) . (137)

Ignoring the relativistic corrections, eigenfunctions of the Hamiltonian for aone-electron central potential are simultaneous eigenfunctions of H0, l2, lz,s2 and sz, so the states are characterized by the quantum numbers n, l, ml,s, ms. When the spin-orbit term is included however, lz and sz no longercommute with the Hamiltonian. The states are then characterized by n, l,s, j, mj . We will see shortly that this has important consequences for thespecification of the states in multi-electron atoms.

Page 29: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 31

3.5 The Central Field Approximationand Quantum Indistinguishability

When there is more than one electron in the atom, the Schroedinger equationacquires an additional term due to the electron-electron repulsion:⎛

⎝12

∑j

∇2j + E + Z

∑j

1rj

−∑i>j

1| ri − rj |

⎞⎠ψ(rj) = 0 (138)

where rj is the position coordinate of the jth electron, ∇j ≡ ∂/∂rj andthe sum is taken over all electrons. As indicated, the wave-function nowdepends on the set of all electron positions rj. Even for the case of justtwo electrons, (138) is impossible to solve analytically. The main problem isdue to the coupling of all of the individual rj ’s.

To make the problem tractable, some simplifying assumptions must bemade. The most common is called the central field approximation. We par-tially account for the effects of the electron-electron repulsion by modifyingthe central potential, and then treat the residual electron-electron repulsionas a perturbation. That is, we define a zeroth order Hamiltonian by:

H0 = −12

∑j

∇2j +∑

j

V (rj) (139)

and a perturbing Hamiltonian by:

H ′ =∑i>j

1| ri − rj | −

∑j

(Z

rj+ V (rj)

). (140)

Here V (r) takes the form of a screened Coulomb potential. Close to thenucleus,

V (r) → −Zr

+ C

where C is a constant. Far from the nucleus

V (r) → −(Z −N + 1)r

where N is the number of electrons in the atom. The constant C enters inbecause the outer electrons approximate a uniformly charged sphere wherethe electron is close to the nucleus, and the potential inside a uniformlycharged sphere is constant.

In the central field approximation, the zeroth order Hamiltonian given by(139) is the sum of single particle Hamiltonians, and thus the zeroth orderwave-functions can be written as the product of single particle wave-functions:

ψ(rj) = ψ1(r1)ψ2(r2) . . . ψN (rN ) (141)

Page 30: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

32 S.M. Kahn

where the individual ψj(rj) are solutions to the single electron Schroedingerequation: (

12∇2

j + E − V (rj))ψj(rj) = 0 (142)

and are individually characterized by the quantum numbers n, l, ml, s,ms. This would be sufficient if it were not for quantum indistinguishabil-ity. Because the atomic electrons form a system of identical particles andbecause they are fermions, the total wave-function must be anti-symmetricwith respect to particle interchange. We can construct such an anti-symmetricwave-function by forming the following linear combination of product wave-functions:

ψ(rj) =1√N !

∑P

(−1)Pψ1(rj1)ψ2(rj2) . . . ψN (jN ) . (143)

Here, in each term in the sum, the set of single-electron wave-functions isarranged in the same order, but the electron coordinates, rj1, rj2, . . . , rjN

have been arranged in a new order which is a permutation of the original set.The sum is taken over all possible permutations. For each permutation,

P represents the number of interchanges. Thus (−1)P = +1 for even permu-tations and −1 for odd permutations. The wave-function given by (143) isoften written in terms of what is called a Slater determinant:

ψ(rj) =1√N !

∣∣∣∣∣∣∣∣∣

ψ1(r1) ψ2(r1) . . . ψN (r1)ψ1(r2) ψ2(r2) . . . ψN (r2)

...ψ1(rN ) ψ2(rN ) . . . ψN (rN )

∣∣∣∣∣∣∣∣∣(144)

and is occasionally referred to as a determinantal wave-function. An impor-tant consequence of the anti-symmetrization is the Pauli Exclusion Princi-ple: “No two electrons can occupy the same individual quantum state”. Thiscan be seen to follow trivially from the Slater determinant. If two of thesingle particle wave-functions, ψi and ψj are identical then two columns inthe matrix are identical and the determinant vanishes. The Pauli exclusionprinciple implies that for multi-electron atoms, even the ground state mustinvolve electrons in the individual particle excited states. Recall that for prin-cipal quantum number n, there are 2n2 distinct spin and angular momentumstates. If there are more than two electrons in the atom, at least some mustbe in an n = 2 or higher level. If there are more than ten electrons, somemust be in an n = 3 or higher state.

The specification of the N individual particle quantum states for the setof N electrons is usually referred to as the configuration. The representationof the general wave-function ψ(rj) in terms of the Slater determinant issometimes called the single configuration approximation.

Page 31: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 33

3.6 Electron Exchange – Helium-like Atoms

A second important consequence of the anti-symmetrization of the wavefunc-tion is the existence of what are called electron exchange terms. These areadditional interaction terms which introduce spin dependence in the energylevels even when there is no explicit spin dependence in the Hamiltonian.

The key concepts are most simply illustrated by looking at the detailedlevel structure of helium-like atoms where there are two orbital electrons.The Hamiltonian for this system is:

H = −12∇2

1 −12∇2

2 −2r1

− 2r2

+1r12

(145)

where r12 ≡ | r1 − r2 |. The Hamiltonian is spin-independent, so the eigen-functions are functions only of the r1 and r2. However, because of theanti-symmetrization, there is a coupling to spin. Specifically, the total wave-function can be written in only one of the two forms:

ψ = ϕS(r1, r2)χA(ms1,ms2) (146)

orψ = ϕA(r1, r2)χS(ms1,ms2) . (147)

Here ϕ denotes the spatial component of the wave-function, while χ denotesthe spin component. The subscripts “S” and “A” indicate the symmetricand anti-symmetric combinations, respectively. Since the total wave-functionmust be anti-symmetric, one of the two must appear in a symmetric combi-nation while the other must be anti-symmetric.

The symmetric spin-state is the so-called triplet state, where the totalspin: s = s1 + s2 has eigenvalue s = 1. This state has three-fold degeneracy;the degenerate eigenstate can be written in the form:

| 1/2, 1/2〉 , ms = +1

1√2(| 1/2,−1/2〉+ | −1/2, 1/2〉) , ms = 0

| −1/2,−1/2〉 . ms = −1

Here the first index in each case is ms1 and the second index is ms2. Theanti-symmetric spin state is the singlet state, corresponding to s = 0. Thereis no degeneracy in this state. It can be written in the form:

1√2(| 1/2,−1/2〉− | −1/2, 1/2〉) . ms = 0

Invoking the central field approximation, we will treat the electron-electronrepulsion term as the perturbation. For simplicity, we will take the cen-tral potential to be the simple Coulomb potential of the nuclear charge:V (r) = −2/r. In that case the spatial part of the wave-function is the product

Page 32: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

34 S.M. Kahn

wave-function of hydrogen-like eigenfunctions. The symmetric combination is:

1√2(ϕ1(r1)ϕ2(r2) + ϕ2(r1)ϕ1(r2))

where ϕ1 and ϕ2 are each characterized by a particular choice of n, l, ml.The anti-symmetric combination is:

1√2(ϕ1(r1)ϕ2(r2) − ϕ2(r1)ϕ1(r2)) .

Now consider the ground state of the helium atom. Both of the electronsmust be in the lowest energy orbital, corresponding to n = 1, l = 0. Since thetwo electrons are in the same spatial state, the spatial wave-function must besymmetric. In that case, the spin wave-function is anti-symmetric, so this is asinglet state. In first order perturbation theory, the correction to the energylevel is given by:

∆E =⟨ψ

∣∣∣∣ 1r12

∣∣∣∣ψ⟩

=∫d3r1d

3r2 | ϕ10(r1) |2| ϕ10(r2) |2 1r12

. (148)

This expression has a simple classical interpretation: since | ϕ10(r1) |2 and| ϕ10(r2) |2 represent the probability density of finding the electrons at posi-tions r1 and r2, respectively, this is just the weighted average of the electro-static repulsion energy between them.

Next consider the first excited states. In this case, one of the electrons isin the n = 1, l = 0 orbital, while the other is in an n = 2, l = 0, 1 orbital. Inthis case, there are two possible spatial wave-functions:

1√2(ϕ10(r1)ϕ20(r2) + ϕ20(r1)ϕ10(r2))

which corresponds to the spin singlet, and

1√2(ϕ10(r1)ϕ20(r2) − ϕ20(r1)ϕ10(r2))

which corresponds to the spin triplet. The first order perturbation theorycorrection to the energy level now has two terms:

∆E =∫d3r1d

3r2 | ϕ10(r1) |2| ϕ20(r2) |2 1r12

±∫d3r1d

3r2ϕ∗10(r1)ϕ∗

20(r2)ϕ20(r1)ϕ10(r2)1r12

(149)

where the (+) sign applies to the spin singlet combination and the (−) sign ap-plies to the spin triplet. The first term has the same interpretation that we saw

Page 33: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 35

earlier; it is the weighted average of the electrostatic repulsion energy. How-ever, the second term is new. It appears because of the anti-symmetrizationof the wave-function and is generally referred to as the electron exchangeterm. It can be shown that the integral for this term is always positive, sothe triplet state has always lower energy. Thus the lowest excited state ofhelium-like atoms are spin triplet states.

A simple interpretation of the exchange energy is as follows: for a spintriplet combination, the spatial wave-function is anti-symmetric, so the Pauliexclusion principle requires that the electrons stay further apart. In thatcase, the electrostatic repulsion energy is reduced. For a spin singlet, theelectrons are closer together on average and the electrostatic repulsion energyis enhanced.

3.7 Approximation Techniques for Multi-Electron Atoms

For more complicated multi-electron atoms, the electron-electron interactionis a significant perturbation and some form of approximation scheme is re-quired to calculate wave-functions and energy levels. Within the context ofthe central field approximation, the simplest approach is to assume a centralV (r) which suitably accounts for the effects of electron shielding, and thento use this potential to calculate the single electron wave-functions which arethe basic ingredients for the Slater determinant wave-function appropriate tothe whole atom. Final wave-functions and energy levels are computed usingfirst order perturbation theory, with the perturbation given by (140).

An early candidate functional form for the central potential was theThomas-Fermi potential derived from a statistical treatment of the electroncloud as a gas of free-particle degenerate fermions at zero temperature. Thepotential is calculated classically from an assumed continuous charge densityρ(r) and the form of ρ(r) is adjusted so as to achieve a minimum in the total(kinetic plus potential) energies. This model yields moderately accurate en-ergy levels for the valence shells of multi-electron near-neutral atoms, wherethe semi-classical assumptions involved are most reliable.

A more modern, and more accurate approach is to assume a convenientanalytic form for the potential such as:

V (r) = −2r((N−1)e−α1r+α2re

−α2r+. . .+αN−1rke−αN r+Z−N+1) (150)

characterized by the adjustable set of parameters: α1, α2, . . . , αN . For a givenconfiguration, the values of the αi’s are determined by minimizing the totalenergy of the atom. This yields a unique form for the potential for eachelectron configuration. That is sufficient for calculating energy levels. How-ever, for the calculations of matrix elements (such as oscillator strengths), acommon potential must be chosen, or otherwise the wave-functions describ-ing initial and final states are not necessarily orthonormal. The parametric

Page 34: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

36 S.M. Kahn

potential method is computationally fast, and has been shown to yield rea-sonably accurate results, especially for highly charged ions, which are thedominant contributors to astrophysical X-ray spectra.

The most accurate conventional approach however is the Hartree-Fock orself-consistent field method. Here one takes a direct account of the dependenceof the individual electron wave-functions on one another, which is broughtabout by the electron-electron repulsion term. The governing equations canbe derived from the Ritz variational principle, i.e. using total wave-functions,ψ, constructed as Slater determinants of individual electron wave-functions,ϕi, we minimize the quantity 〈ψ | H | ψ〉 (where H is the total Hamiltonian)subject to the constraint that the individual wave-functions remain ortho-normal. This can be accomplished by introducing N Lagrange multipliers εi,such that:

δ(〈ψ | H | ψ〉 −∑

i

εi〈ϕi | ϕi〉) = 0 . (151)

The result is a set of N equations (the Hartree-Fock equations) which looklike Schroedinger equations, but with potentials that depends on the wave-function solutions:⎡

⎣−12∇2

i −Z

ri+∑j =i

∫d3rj

| ϕj(rj) |2| ri − rj |

⎤⎦ϕi(ri) −

∑j =i

δ(msi,msj)

×[∫

d3rj1

| ri − rj |ϕ∗j (rj)ϕi(rj)

]ϕj(ri) = εiϕiri . (152)

Here msi and msj are the eigenvalues of sz for the ith and jth orbitals inthe electron configuration, respectively. The first two terms on the left-handside of (152) are associated with the single particle Hamiltonian ignoring theelectron-electron interaction. The third term comes from the electron-electronrepulsion energy. The fourth term is due to the exchange energy. It is zerounless the two orbitals have the same spin (δ(msi,msj) = 1), so that thespatial part of the wave-function is anti-symmetric.

For a given set of trial wave-functions ϕ(0)i (r), the set of (152) can be

solved to yield a new set of wave-functions ϕ(1)i (r). This is repeated until it

converges, i.e. until the resulting set of eigenfunction solutions is “close” tothe trial set. The process yields a self-consistent potential for the electron-electron interaction which can then be used to calculate energy levels andmatrix elements.

Hartree-Fock calculations are generally time-consuming and unwieldy incomparison to the simpler parametric potential methods discussed earlier. Inaddition, the self-consistent potential is not always smooth and well-behavedwhich can complicate the calculation of relativistic corrections (134 and 135)that are important for highly charged ions.

Page 35: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 37

3.8 LS, jj and Intermediate Coupling

The Hamiltonian for the multi-electron atom as incorporated in (138) is rota-tionally invariant. In addition, it has no explicit spin dependence. This meansthat H must commute with the operators J , L and S:

[H,J ] = [H,L] = [H,S] = 0 (153)

where L is the total orbital angular momentum of all the electrons in theatom: L =

∑i li, S is the total spin angular momentum: S =

∑i si and J is

the total angular momentum: J = L + S. Hence, the eigenstates of H mustalso be eigenstates of J2, Jz, L2, Lz, S2 and Sz and will thus be characterizedby definite values of the corresponding eigenvalues: J , MJ , L, ML, S, MS ,in addition to the energy E.

However, in the central field approximation, we have constructed theeigenfunctions out of single-electron wave-functions, which are themselveseigenfunctions of l2, lz, s2, sz, and are thus characterized by the eigenvaluesl, ml, s, ms. The simple product wave-functions which comprise the Slaterdeterminant will be characterized by a set of definite eigenvalues l(i), m(i)

l ,s(i), m(i)

s for each of the electrons in the atom. But L2 does not commutewith the individual l(i)z operators and S2 does not commute with the individ-ual s(i)z . Hence these simple products cannot be eigenfunctions of the totalHamiltonian including the electron-electron repulsion.

Product states of definite L, ML, S, MS can however be generated by“coupling” individual product wave-functions into suitable superpositions.Here one uses the usual rules of angular momentum addition in quantummechanics, and the coefficients of the various terms are given by Clebsch-Gordan coefficients. One first couples the spatial wave-functions individuallyinto states of definite L2 and Lz and the spin wave-functions individuallyinto states of definite S2 and Sz. One couples their product together to yieldstates of definite J2 and Jz. This is called an LS coupling scheme or sometimesRussell-Saunders coupling.

The anti-symmetrization of the wave-function involves a superpositionover permutations of the electron coordinates. Coupling involves a super-position over different values of m(i)

l and m(i)s . In principle, one can anti-

symmetrize first and couple afterwards or couple first and anti-symmetrizeafterwards. In practice, the latter is usually easier. The calculation of thematrix elements using these anti-symmetrized, coupled wave-functions canbe quite complex if carried out by brute force. Fortunately, there is an ele-gant mathematical formalism known as Racah algebra – developed by Racahand Wigner in the 1940’s – which greatly simplifies the angular part of thesematrix elements.

The discussion above ignores the relativistic corrections covered in Sect. 3.4.In particular, the spin-orbit term (134) in the single electron Hamiltonian isproportional to the operator l ·s, which does not commute with lz and sz, but

Page 36: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

38 S.M. Kahn

does commute with j2 and jz. When this term is important, it is convenientto first couple the individual particle wave-functions into states of definitej(i), m(i)

j and then couple these states into states of definite J , MJ . This isknown as jj-coupling.

jj-coupling is formally incompatible with LS-coupling because states ofdefinite L2, Lz, S2, Sz are not characterized by definite values of j(i), m(i)

j .In practice, LS-coupling is preferred whenever the electron-electron repulsionterm dominates over the spin-orbit terms. This is especially true for low-Zatoms which are not highly ionized. jj-coupling would be preferred for high-Zatoms with only a few electrons. In cases where both electron-electron andspin-orbit terms are important, neither scheme is entirely appropriate. Inthat case, one chooses one or the other as the basis, and then diagonalizesthe “other” perturbing operator in this basis to achieve the appropriate su-perpositions. This is known as intermediate coupling. The final eigenstatesare then only characterized by definite values of J and MJ .

3.9 Spectroscopic Notation and Ground-State Configurations

In LS-coupling, a given electron configuration is specified by the quantumnumbers n(i), l(i), s(i) for each of the individual electrons and the total quan-tum numbers L, S, J , MJ for the atom as a whole. In the absence of anexternal field, the energy levels are degenerate in MJ so this is usually notincluded. In addition, all electrons have s = 1/2, so this too need not beindicated. Over the years, a notational scheme has become standard for des-ignating these configurations. Specifically, for a given nl “shell” the numberof electrons in that shell is indicated as an exponent. Recall that there are2(2l + 1) distinct states in such a shell , so the exponent cannot exceed thatnumber. For historical reasons, l is not indicated as an integer, but insteadas a letter, with the assignments:

l = 0 1 2 3 4 5 . . .symbol s p d f g h . . .

Thus the notation 3d24f indicates two electrons with principal quantum num-ber n = 3 and angular momentum l = 2 and one electron with n = 4 andl = 3.

For the total quantum numbers, the standard notation has the form

2S+1LJ .

Here again a letter is used in place of a number for L and the convention is thesame as that used for the individual l’s only with upper case letters insteadof lower case. Thus the designation 2D3/2 indicates a state with S = 1/2,L = 2 and J = 3/2.

For X-ray emitting astrophysical plasmas, we are mainly concerned withfew electron atoms, specifically K- and L-shell ions, isoelectronic with the

Page 37: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 39

neutral elements hydrogen through neon. Only a few key ideas are requiredto understand the ground configuration of such ions.

1. For a Coulomb potential, we have seen that the energy levels only dependon n not l. This is not true of the screened Coulomb potential appropriateto multi-electron atoms. The lower the angular momentum, the higher theprobability that the electron is close to the nucleus where it “sees” lessscreening of the nuclear charge and hence the lower the energy. The energytherefore increases strongly with n and/or l.

2. Because of the strong dependence on n and l, as electrons are added to anion, they continue to fill n, l “shells” until they are closed. A shell is closedwhen all of its magnetic spatial and spin orbitals are filled. A closed shelltherefore has J , L and S all equal to zero.

3. For a partially open shell, the state of highest S will have the lowestenergy. This is a consequence of the exchange energy, as we saw earlier.If S is maximal, the spin wave-function must be symmetric, which meansthat the spatial wave-function is anti-symmetric, and the electrons are onaverage further apart, thereby lowering their repulsion energy.

4. If the partially open shell is less than half-full, the lowest energy state willhave the lowest possible value of J . This is a consequence of the spin-orbitinteraction, which contributes positive energy that increases with J .

5. If the open shell is more than half-full, it is easier to think in terms ofthe electron “holes” rather than the electrons. These behave like positiveelectrons. Their spin-orbit contribution then has opposite sign. As a result,the lowest energy state has the highest possible J .

Using these rules, one can understand now the ground-states of hydrogen-likethrough neon-like ions have the following configurations:

H: 1s 2S1/2

He: 1s2 1S0

Li: 1s22s 2S1/2

Be: 1s22s2 1S0

B: 1s22s22p 2P1/2

C: 1s22s22p2 3P0

N: 1s22s22p3 4S3/2

O: 1s22s22p4 3P2

F: 1s22s22p5 2P3/2

Ne: 1s22s22p6 1S0

In cases of intermediate coupling, which is important for highly charged ions,it is sometimes useful to also indicate the j-values of the individual electrons.This is done by adding a subscript to the individual shell terms indicating thevalue of j. Since the spin-orbit interaction for an individual electron has thelowest energy for the lowest values of j, the lower j states are filled first. Thus,in this notation, the ground configuration of oxygen-like ions is representedby 1s22s22p2

1/22p23/2. Of course, for intermediate coupling, the L and S values

Page 38: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

40 S.M. Kahn

are not precisely defined. Typically, one lists the notation for the leading termin the LS expansion.

3.10 Configuration Interaction

In Sect. 3.5 we introduced the central field approximation and the associ-ated single configuration approximation, where the total wave-function iswritten as an anti-symmetrized product of single-electron wave-functions. Itshould be emphasized that this is an approximation – it is by no means clearthat the exact multi-electron eigenfunction of the total Hamiltonian is closeto a single configuration wave-function, i.e. to a single Slater determinant.When this is not true, we need to allow for configuration mixing, by form-ing multi-configuration superpositions derived from matrix elements of theHamiltonian. Codes which include these effects are called multi-configurationcalculations.

It is impractical of course to include a large number of configurations inconstructing the basis set. However, some guidance comes from the structureof the Hamiltonian. In LS-coupling, only configurations of common L, S, Jand parity need be included. In addition, since the Hamiltonian only containsterms involving one or two electrons, interactions can only occur betweenconfigurations that differ in at most two orbitals.

Configuration interaction tends to be strong between configurations whichare close in energy. For the highly charged ions important in X-ray emittingplasmas, the energy levels are more weakly dependent on l. Thus significantmixing can occur between configurations like 3s23pk and 3pk+2. In such cases,the identification of a particular transition with a set of upper and lowerconfigurations is not very meaningful.

3.11 Selection Rules for Radiative Transitions

The matrix elements which appear in the various terms in the multipoleexpansion for radiative transitions can vanish for particular choices of initialand final states. This gives rise to what are called selection rules for thevarious multipole transitions. Transitions which violate the selection rules arecalled forbidden, while those consistent with the selection rules are allowed.

First, consider electric dipole transitions. Here the matrix elements is〈f |ri〉, where r =

∑i ri. Since r is a sum of single electron operators, this

matrix element will vanish if the initial and final configurations differ by morethan one electron orbital. Hence, only single electron transitions are allowed.Second, note that r has odd parity. Thus initial and final states must haveopposite parity. Finally, since in spherical coordinates ri can be written asa superposition of the spherical harmonics with l = 1, it is easy to showthat this matrix element also vanishes unless ∆l = ±1 for the change in thesingle electron orbital. The essential selection rules are ∆l = ±1, ∆s = 0,∆L = 0,±1, ∆S = 0, ∆J = 0,±1, with J = 0 → 0 strictly forbidden.

Page 39: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 41

Second, for magnetic dipole transitions, the matrix element is 〈f | µ | i〉,where µ is the magnetic dipole moment. Including spin contributions, µ ∼L+2S = J +S. Since J commutes with H, 〈f | J | i〉 = 0, so we are only leftwith 〈f | S | i〉. This is a pure spin operator, so the net spatial configurationcannot change. Ignoring relativistic terms, S also commutes withH. However,the spin-orbit interaction introduces some mixing. The selection rules are∆S = 0,±1 (spin flip), ∆J = 0,±1, no J = 0 − 0, no parity change, nochange in configuration (i.e. ∆n = 0, ∆l = 0 for all electrons).

And third, for electric quadrupole transitions, the selection rules are: ∆l =0,±2, ∆L = 0,±1,±2, ∆J = 0,±1,±2, no J = 0 − 0, no change in parity.

When configuration interaction is important, these selection rules canappear to be violated because of mixing. That is, even if the dominant con-figurations in the initial and final states violate the selection rules, there maybe small admixtures in each case that do contribute to a non-zero matrixelement.

4 Electron-Ion Collisional Processes

4.1 Overview

In the previous two chapters, I have laid out the essential ingredients for thecalculation of radiative transitions rates between various energy levels and forthe atomic structure effects which give rise to the particular characteristics ofthose levels. To predict the emergent X-ray spectra of astrophysical plasmas,however, we also need to understand the details of how excited atomic levelsare populated. For the most part, that involves the study of electron-ion colli-sional processes in plasmas. This is also a rich and diverse field and it will notbe possible to do justice to the full complexity of this topic. My emphasis, asin the previous chapter, will be on the explication of key concepts, definitionof terms commonly used in the atomic physics literature and presentation ofsome quick back-of-the-envelope type calculations that enable us to deriverough estimates of the rate coefficients for these processes.

Each electron-ion collisional process is accompanied by a quantum me-chanical inverse, which can be viewed as the same process time-reversed. Notsurprisingly, the rates for direct and inverse processes involve common matrixelements, and are therefore related. The easiest way to derive these relationsis to resort to detailed balance arguments, i.e. to set the rates for direct andinverse processes equal in strict thermodynamic equilibrium. I will defer anextensive discussion of thermodynamic equilibrium to the next chapter, butwe will anticipate some important results from that discussion and utilizethem here.

There are essentially four key electron-ion collisional processes that areimportant for X-ray emitting plasmas. These are schematically illustrated

Page 40: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

42 S.M. Kahn

2

1

Collisional ionization 3-body recombination

Collisional deexcitationCollisional excitation

Fig. 2. The first two of the four key electron-ion collisional processes. The “inverse”process is on the right

in Figs. 2 and 3 where the “direct” process is depicted on the left and the“inverse” process on the right.

Collisional Excitation/Deexcitation

In collisional excitation, the interaction between a passing electron in a con-tinuum state and a bound electron in a discrete state results in the excitationof the bound electron to a higher energy discrete level. To conserve energy,the colliding electron gives up a fraction of its energy and thus “falls” into alower continuum state. The inverse process is collisional deexcitation, wherea passing electron interacting with an excited atom actually gains energy asa result of the collision.

Collisional Ionization/3-Body Recombination

Collisional ionization is similar to collisional excitation, except that in thiscase, the final state of the initially bound electron is also a continuum state.The inverse process is 3-body recombination. Here, two, initially free electronsinteract with the ion in the same collision. One of the two gets captured intoa bound discrete level, while the other carries off the excess energy in a highercontinuum state.

Page 41: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 43

Dielectronic Capture Autoionization

Photoionization

3

4

Radiative Recombination

Fig. 3. The last two of the four key electron-ion collisional processes

Radiative Recombination/Photoionization

In radiative recombination a free electron in a continuum state decays into abound discrete state through the emission of a photon. This is actually a formof spontaneous emission, similar to what we discussed for the radiative decaybetween two bound levels in Sect. 2.9. The inverse process is photoionization,or bound-free absorption, as discussed in Sect. 2.8.

Dielectronic Capture/Autoionization

Dielectronic capture is a resonant radiationless process in which the decayof an electron from a continuum state to a bound state is accompanied bythe elevation of a core electron into an excited state. The resulting atom isdoubly excited, and it has a total energy above the ionization potential ofthe initial ion. The inverse process is autoionization, where a doubly excitedatom decays via the emission of a weakly bound outer electron. If the coreexcitation is associated with a “hole”, in one of the orbitals of an inner shell,this process is usually called Auger decay.

In the remainder of this chapter, I will review each of these processes insomewhat more detail.

Page 42: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

44 S.M. Kahn

4.2 Collisional Excitation – Scattering Theory

Collisional excitation is essentially an example of inelastic scattering of anelectron off a complex atomic potential, and thus much of the formalismof quantum scattering theory can be applied to this process. Typically, oneexpresses the continuum wave-function at large distances from the atom asthe sum of an incident plane wave and an outgoing spherical wave:

ϕc(r)r→∞ A

[eiki·r + f(ϑ, ϕ)

eikf ·r

r

](154)

where ki is the initial momentum of the electron, 2k2

i /2m is its initialenergy and

2k2f/2m is its final energy. The flux in the wave is given by:

j(r) =

2mi[ϕ∗(∇ϕ) − (∇ϕ∗)ϕ] (155)

(see (109)). For the incident wave, this gives:

jin =ki

m| A |2 . (156)

For the outgoing wave:

jout · r =

2mi

[ϕ∗ ∂ϕ

∂r− ∂ϕ∗

∂rϕ

]

=kf

m

| A |2| f |2r2

. (157)

The number of scattered electrons in solid angle element dΩ is: (jout ·r)r2dΩ.Therefore, the differential cross-section for scattering is:

dΩ=

(jout · r)r2

jin=kf

ki| f(ϑ, ϕ) |2 , (158)

f is called the scattering amplitude.If we limit our consideration to single electron transitions, then the total

wave-function can be expressed in terms of product wave-functions for thecolliding electron and the bound transitioning electron. These are still identi-cal particles, so the total wave-function must be anti-symmetrized. Due to theexchange terms (see below), we get different answers for the singlet state andthe triplet state. Averaging over the four possible spin states, the differentialcross-section will then look like:

dΩ=kf

ki

[14| f+ |2 +

34| f− |2

](159)

where the (+) indicates a symmetric spatial wave-function and the (−) indi-cates an anti-symmetric spatial wave-function.

Page 43: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 45

The calculation of the scattering amplitude proceeds as follows: wewrite the total wave-function as the sum of anti-symmetrized product wave-functions for the initial and final states:

ψ =[ϕ±

ci(r1)ϕbi

(r2) ± ϕ±ci

(r2)ϕbi(r1)]

+[ϕ±

cf(r1)ϕbf

(r2) ± ϕ±cf

(r2)ϕbf(r1)]

(160)

where ϕ±ci,f

are the initial and final wave-functions for the colliding electronand ϕbi,f

are the initial and final wave-functions for the bound electron. ψmust satisfy the Schroedinger equation:[

−12∇2

1 −12∇2

2 + V (r1) + V (r2) +1r12

]ψ = Etotψ . (161)

Therefore, if we take a scalar product with ϕ∗bi

(r2) we must get:

∫d3r2ϕ

∗bi

(r2)[−1

2∇2

1 −12∇2

2 + V (r1) + V (r2) +1r12

− E

]ψ = 0 . (162)

But [−1

2∇2

2 + V (r2)]ϕbi

(r2) = Ebiϕbi

(r2) , (163)

and

Etot = Ebi+

2k2

i

2m. (164)

Substitution of (160) into (162) yields

(∇21 + k2

i

)ϕ±

ci(r1) = 2

[Vii(r1)ϕ±

ci(r1) + Vif (r1)ϕ±

cf(r1)]

±2[∫

d3r2Kii(r1, r2)ϕ±ci

(r2) +∫d3r2Kif (r1, r2)ϕ±

cf(r2)]

(165)

where

Vii ≡ V (r1) +⟨ϕbi

∣∣∣∣ 1r12

∣∣∣∣ϕbi

⟩(166)

Vif ≡⟨ϕbi

∣∣∣∣ 1r12

∣∣∣∣ϕbf

⟩(167)

Kii(r1, r2) ≡ ϕ∗bi

(r1)ϕbi(r2)[

1r12

− Etot −Ebi

](168)

Kif (r1, r2) ≡ ϕ∗bi

(r1)ϕbf(r2)[

1r12

− Etot − Ebi−Ebf

](169)

The terms involving the V ’s are the direct potential terms, the K’s are theexchange terms. A second similar equation can be obtained (with the i’s and

Page 44: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

46 S.M. Kahn

f’s reversed) by taking the scalar product with ϕ∗bf

in place of ϕ∗bi

in (162). Theresult is a set of two coupled equations which can be solved simultaneouslyfor ϕ±

ciand ϕ±

cfgiven expressions for ϕbi

and ϕbf. They are analogous to

the Hartree-Fock equations for a two electron atom. Once the continuumwave-functions are found, the scattering amplitudes can be computed andwe obtain the cross-section.

The exchange terms can be important at low collision energies, especiallyfor electric dipole forbidden transitions. At high energies, the continuumwave-functions, ϕci

and ϕcfoscillate strongly in comparison to the slowly

varying K-functions and so the integrals on the right-hand side of (165) tendto vanish.

This procedure is still an approximation since we have not allowed thecolliding electron to influence the bound-state wave-functions. One approachto correcting this is to include in the trial wave-function (160) other termsallowing for other proper collision channels, involving other sets of boundexcited states. That is called a close coupling calculation since it couplesin other states of the atom. It results in a much larger set of simultaneousequations, depending on how many channels are included.

At energies well above threshold, a much simpler calculation can be per-formed using the Born approximation. Here one assumes plane-wave wave-functions for both the initial and final continuum states. The transition ratecan be calculated from time-dependent perturbation theory (see (70)) takingthe electron-electron interaction as the perturbing potential:

R =2π

∣∣∣∣⟨f

∣∣∣∣ e2

| r1 − r2 |∣∣∣∣ i⟩∣∣∣∣

2

δ(Ef −Ei)

=∣∣∣∣2π 1

V 2

∫V

d3r1d3r2e

−ikf ·r2ϕ∗bf

(r1)e2

| r1 − r2 |ϕ∗bi

(r1)eiki·r2

∣∣∣∣2

× δ(Ef − Ei) (170)

where we have normalized the plane waves over a finite volume V .Because exchange effects were found to be small at higher energies, one

usually does not need to bother anti-symmetrizing the wave-function. Thetotal rate is found by summing over the trial states of the outgoing electronsso that the δ-function gets replaced by a density of states factor:

ρf =V

2π2

(2m2

)kf . (171)

The total rate thus scales like 1/V . However, the incident flux is given byvi/V in this picture, so the total cross-section is independent of the assumedvolume, as expected.

A similar, but somewhat improved calculation can be obtained using con-tinuum wave-functions of the Coulomb potential of the ion in place of theplane-waves. This is called the Coulomb-Born method. Even better yet is to

Page 45: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 47

use the continuum wave-functions derived from the effective central poten-tial V (r) of the atom. That is the distorted wave approach. Usually distortedwave radial wave-functions are calculated in a partial wave expansion, sum-ming over states of definite orbital angular momentum l. Close to threshold,the energy of the outgoing electron is low and only a small number of termsin the partial wave expansion need be kept. The maximum l required can beroughly estimated from classical considerations:

L ≈ pfa⇒ l ≈ kfa (172)

where a is the characteristic dimension of the atom. At high impact energies,many partial waves are required and the plane-wave Born approach providesa much simpler alternative.

The integral which appears in the plane-wave Born approximation (170),can be simplified using the Bethe integral:∫

d3rei∆·r

r=

4π∆2

(173)

which implies that the excitation cross-section is proportional to the squareof a matrix element given by:

1∆2

∫d3rϕ∗

bf(r)ei∆·rϕbi

(r) (174)

where ∆ ≡ ki − kf . Note that the expression in (174) can be approximatedby a multipole expansion:

ei∆·r ≈ 1 + i(∆ · r) + . . . (175)

entirely analogous to the multipole expansion invoked for radiative transitionsin Sect. 2.10. Here again, ∆ ·r ≈ k ·r ≈ v/c, so for non-relativistic electrons,only the lowest order non-vanishing term usually needs to be considered.We thus obtain selection rules for collisional excitation between bound lev-els which are identical to the selection rules for radiative transitions betweenthose levels. Therefore, transitions that are electric dipole forbidden also havelow cross-section for collisional excitation. The above argument, however, re-lies on the plane-wave Born calculations, ignoring exchange effects. Generally,exchange terms dominate the cross-section for higher order multipole transi-tions.

4.3 Collisional Excitation – Classical Estimate

The discussion in Sect. 4.2 provides a sketch of how accurate collisional ex-citation cross-sections are calculated using sophisticated atomic codes, butis not especially helpful for getting quick quantitative estimates of the mag-nitude of collisional excitation rates. For this, it is more useful to resort to

Page 46: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

48 S.M. Kahn

simple classical arguments. Imagine a passing electron interacting via theCoulomb force with one of the orbital electrons in the atom. The momentumtransfer to the bound electron is approximately:

∆p ≈∫ ∞

0

dtF (t) ≈ e2

b2

(2bv

)=

2e2

bv(176)

where b is the impact parameter of the colliding electron, and τ = 2b/v is thecharacteristic duration of the interaction. Thus, the energy transfer to thebound electron is:

∆E ≈ (∆p)2

2m≈ 2e4

mb2v2. (177)

The energy transfer must equal the energy of the excitation ∆E ≈ Emn,where we are considering a transition from initial state m to final state n.The cross-section at impact parameter b is σ ≈ πb2 so:

σmn ≈ 2πe4

mv2Emn=

πe4

EeEmn(178)

where Ee is the energy of the colliding electron. In atomic units:

σmn(Ee) ≈ 4πa20

EeEmn(179)

where a0 is the Bohr radius.It is traditional to express the cross-section in terms of a collision strength

Ωmn which is specific to the transition, but relatively independent of theelectron energy:

σmn(E) ≡ πa20

gmEeΩmn (180)

where gm is the degeneracy of the initial state. One thus sees that classicallyΩmn ≈ 4gm/Enm. The quantum mechanical treatment (for electric dipoletransitions) gives:

Ωmn

gm=

8π√3fmng

Emn(181)

where fmn is the dipole absorption oscillator strength for the transition, andg is a Gaunt factor which is ≈ 1 for ∆n = 0 transitions, and ≈ 0.2 for ∆n = 0transitions.

In thermal plasmas, collisional excitation can be characterized by a ratecoefficient Cmn(T ), which is a function of electron temperature and is specificto the transition. The rate of collisional excitations for transition m to n perunit volume is given by nen

mi Cmn(T ) where ne is the free electron density and

nmi is the density of the relevant ion in state m. In terms of the cross-section:

Cmn(T ) =∫ ∞

v0

dvvf(v, T )σmn(v) (182)

Page 47: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 49

where v0 = (2Emn/m)1/2 is the threshold velocity for the transition andf(v, T ) is the Maxwellian velocity distribution appropriate to a thermalplasma:

f(v, T ) = 4π( m

2πkT

)3/2

v2e−mv2/2kT . (183)

The integration yields:

Cmn(T ) =(πa2

0

gm

)(2kTπme

)1/2(2RykT

)Ωmne

−Emn/kT

≈ 8.6 10−6

gmT−1/2Ωmne

−Emn/kT cm3/s (184)

where T is now in K.The inverse of collisional excitation is collisional deexcitation. The princi-

ple of detailed balance asserts that in thermodynamic equilibrium, the ratesfor a process and its inverse must be equal. The rate for collisional excita-tion is nen

mi Cmn(T ). The rate for collisional deexcitation is nen

ni Cnm(T ).

But in thermodynamic equilibrium, the level populations are related by thedegeneracies and the Boltzmann factor:

nni

nmi

=gn

gme−Emn/kT . (185)

Thus:

Cnm(T ) = Cmn(T )gm

gneEmn/kT

=8.6 10−6

gnT−1/2Ωnm cm3/s (186)

where Ωnm = Ωmn. Note that for isoelectronic sequences, Ωnm scales likeE−1

nm ∼ Z−2. In contrast, we saw earlier (Sect. 3.3) that radiative decay ratesscale like Z4. Thus, for X-ray emitting plasmas, whose spectra are domi-nated by higher Z ions, we need very high electron densities before collisionaldeexcitation competes with spontaneous radiative decay.

4.4 Collisional Ionization

Collisional ionization is essentially the same process as collisional excitationexcept that the final state of the initially bound electron is now also a contin-uum state. The general quantum formalism outlined in Sect. 4.2 can clearlybe applied to this case as well. With two continuum states in the final state,the square of the matrix element in (170) is proportional to 1/V 3 insteadof 1/V 2 but there are now two density of states factors instead of one, sothe final expression for the cross-section is still independent of the assumedvolume.

Page 48: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

50 S.M. Kahn

As in the case of collisional excitation, there is a simple classical calcula-tion that can be invoked to provide a rough estimate of the cross-section. Thisis originally due to Thomson and dates back to 1912 (before the discoveryof the electron!). Thomson calculated the energy transfer between two samecharges, assuming one is initially at rest:

∆E =E[

1 + E2b2

e2

] (187)

where E is the energy of the colliding electron and b is the impact parameter.Setting ∆E ≥ χ, where χ is the ionization potential of the atom, one findsb ≤ bc, where:

bc =e

E

(E

χ− 1)1/2

. (188)

The cross-section is thus given by:

σ = πb2c = πe21E2

(E

χ− 1)

=(4πa2

0

) 1E2

(E

χ− 1)

(189)

where the last expression is in atomic units, with E given in Rydbergs. Thisis a classical ionization cross-section per electron. It must be summed overall the electrons in the atom, using the appropriate χ value for each atomicshell and only including shells for which E ≥ χ.

This Thomson exchange cross-section provides a surprisingly good esti-mate of the true cross-section for E χ, but it gives a significant overesti-mate near threshold. This is due essentially to two effects:

1. The calculation ignores the initial binding energy of the target electron;2. It does not allow for the possibility that if too much energy is transfered,

the colliding electron itself becomes bound.

Hutchinson [7] suggests a simple modification that partially corrects for thesetwo effects:

σ = 4πa20

1E(E + E+)

(E

χ− 1)

(190)

where E+ is an adjustable parameter which is approximately a few times χ.The cross-section given in (190) can be integrated analytically over a

Maxwellian distribution (as in 182) to yield a rate coefficient. The resultinvolves an exponential integral, but Hutchinson shows that to a good ap-proximation one obtains:

C(T ) = 〈σv〉 = 4πa20

(8kTπm

)1/2Ry2

χ(χ+ E+)e−χ/kT

[1 − e−(χ+E+)/kT

]

≈ (8.5 10−8)[

Ry2

χ(χ+ E+)

] [kT

Ry

]1/2

e−χ/kT[1 − e−(χ+E+)/kT

]cm3s−1 .

(191)

Page 49: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 51

Similar (but not identical) formulae have been derived empirically from fitsto experimental data by Lotz [8] and others. These generally agree with oneanother to within a factor of two.

The inverse of collisional ionization is 3-body recombination. However,since this process involves the collision of two electrons with the atom in thesame interaction, it is usually only important at very high densities (ne ≥1019 cm−3), which rarely apply to X-ray emitting astrophysical plasmas.

4.5 Radiative Recombination

Radiative recombination involves the capture of a free electron, accompaniedby the emission of a photon with energy given by:

ωn = E + χn (192)

where E is the initial energy of the electron, and χn is the ionization potentialof the level into which the electron is captured. Since this is a radiativeprocess, it may be calculated using the techniques outlined in Chap. 2. Inparticular, we can get a quick semi-quantitative estimate of the cross-sectionfrom a classical treatment, where we view radiative recombination as a kind ofdiscrete limit of classical bremsstrahlung, the radiation emitted by an electronas it is accelerated in the Coulomb field of an ion. The energy emitted per unitfrequency per unit time per unit volume due to bremsstrahlung by electronsof velocity v is given by:

dW

dωdV dt=

16πe6

3√

3c3m2vneniZ

2g (193)

where ne is the electron density, ni is the ion density, Z is the charge on the ionand g is a Gaunt factor of order unity (see [2]). For radiative recombination,the final state of the electron is discrete, so the energy radiated must all comeout at a single frequency given by (192). We may thus write:

dWn

dV dt=

16πe6

3√

3c3m2vneniZ

2g(∆ωn) (194)

where (∆ωn) is the frequency difference between two neighboring shells.Adopting an “hydrogenic approximation” for the energy levels:

χn ≈ Z2Ry

n2(195)

∆ωn =2Z2Ry

n3≈ 2χn

n. (196)

We define a cross-section σn(v) by setting:

dWn

dV dt= nenivσn(v)ωn . (197)

Page 50: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

52 S.M. Kahn

Plugging in the relevant expressions from (192), (194) and (195) and solvingfor σnv yields:

σn(v)v =16πe6Z2g

3√

3c3m2

2χn

n(

12mv

2 + χn

) 1v. (198)

Finally, averaging over a Maxwellian velocity distribution yields a rate coef-ficient as a function of temperature:

α(T ) ≡ 〈σn(v)v〉 ≈ (5.2 10−14)gZ2(χn

kT

)3/2

eχn/kTEi

(χn

kT

)cm3s−1

(199)

where Ei(x) is the exponential integral.To get more accurate estimates from a quantum mechanical calculation,

it is usually easier to first calculate the photoionization cross-section and thenresort to a detailed balance argument to find the cross-section for radiativerecombination. Let σPI(ω) be the photon cross-section for photoionizationat frequency ω and let σRR(v) be the electron cross-section for radiativerecombination at electron velocity v. As we have seen, ω and v are relatedby energy conservation (192) with E = 1/2 mv2. Let ni be the density of theith ionic species and ni+1 be the density of the one higher ionization state.Then the rate of recombinations per unit volume in the velocity range v tov + dv is given by:

dRRR(v) = neσRR(v)vf(v)dvni+1 (200)

where f(v) is the Maxwellian electron distribution in velocity. The rate ofphotoionizations per unit volume in the frequency range ω to ω+dω is givenby:

dRPI(ω) =F (ω)dω

ωσPI(ω)ni

(1 − e−ω/kT

)(201)

where F (ω) is the energy flux per unit frequency in the radiation field. Inthermodynamic equilibrium, this is given by the expression:

F (ω) =ω3

π2c21

(eω/kT − 1)(202)

(see Sect. 5). The last factor which appears in (201) is a correction for stim-ulated emission – in thermodynamic equilibrium, there are always photon-induced radiative decays in addition to spontaneous radiative decays. Thus(201) gives a net photoabsorption rate. Using the expression we had earlierfor the Maxwellian distribution (183), and equating the rates in (200) and(201) yields:

σPI(ω)σRR(v)

=neni+1

ni

( m

2πkT

)3/2

v3e−(mv2/2−ω)/kT dv

dω. (203)

Page 51: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 53

But 1/2 mv2 − ω = −χ and dv/dω = (dω/dv)−1 = /mv. The ratio of thedensities is given by the Saha equation which we will introduce in the nextchapter:

neni+1

ni=

2gi+1

gi

(mkT

2π2

)3/2

e−χ/kT (204)

where gi+1 is the degeneracy of the final state of the more highly ionized ion,and gi is the degeneracy of the less ionized ion (see Sect. 5). Collecting termsyields:

σPI(ω)σRR(v)

=m2c2v2

2ω2

gi+1

gi(205)

which is called the Milne relation.The quantum mechanical calculation of photoionization cross-sections was

discussed in Sect. 2.8. For hydrogen-like ions, we can obtain an analyticalexpression. Averaging over l, the cross-section for ionization out of the nthshell is given by:

σn(ω) =64α33/2

Z4

n5

(Ry

ω

)3

πa20g (206)

[if ω > Z2Ry/n2 and is zero otherwise] where g is again a Gaunt factorof order unity. The ω−3 dependence is also typical of photoionization cross-sections of more complex atoms. The monochromatic emissivity (energy ra-diated per unit volume per unit frequency) associated with recombinationradiation is given by:

dW

dtdωdV= neni+1(ω)vf(v)σRR(v)

dv

dω=(

)1/2

nenigi

gi+1

×(

ω

χ

)3(χ2

mc2kT

)3/2

cσPI (ω) e−ω/kT eχ/kT . (207)

Notice that for σPI(ω) ∼ ω−3, the frequency dependence is essentially expo-nential above threshold.

4.6 Dielectronic Recombination and Autoionization

Dielectronic capture involves the capture of a free electron into a bound levelwith the accompanying excitation of a core electron. The resulting recom-bined atom is doubly excited. It can decay by autoionization, ejecting thecaptured electron back out into the continuum. In that case, there is no netchange in the level of ionization of the atom. However, the doubly excitedatom can also decay radiatively, thereby lowering its total energy below theionization potential of the recombined atom. When this occurs, the recom-bination is complete and the atom is left in a stable configuration with oneextra electron. The complete process – dielectronic capture followed by ra-diative decay is usually referred to as dielectronic recombination. This can be

Page 52: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

54 S.M. Kahn

a very important process in astrophysical plasmas, especially for ions, as weshall see shortly.

Let’s first consider the inverse process, autoionization. Its rate (from timedependent perturbation theory) is given by:

Aa =2π

∣∣∣∣⟨f

∣∣∣∣ e

| ri − rj |∣∣∣∣i⟩∣∣∣∣

2

(208)

where f and i represent the appropriate product wave-functions for the twoelectrons involved in the interaction in the initial and final states. Note thatin the final state, one of the electrons is in a continuum state. Since the con-tinuum states have wave-functions which are normalized to a delta-functionin energy, this wave-function has units of energy−1/2. Therefore, the squareof the matrix element has units of energy, not energy-squared, as one wouldotherwise expect. When divided by , it gives a finite rate.

The matrix element which appears in the autoionization decay rate (208)is the same matrix element one would use to calculate the configurationinteraction between the doubly bound level and the continuum level withequal energy. In some sense, autoionization is a consequence of configura-tion interaction. The diagonalized eigenstate of the perturbation is then asuperposition of the initial discrete state and a range of continuum states:

ψ = aψdiscrete +∫dEb(E)ψcontinuum(E) (209)

with the coefficient a and b(E) determined by the configuration interactionmatrix-element. It can be shown (see [1] pp. 526–535) that the width of thefunction b(E) is given roughly by Aa, as one would expect based on theenergy-time uncertainty principle. The autoionization process by assigning afinite lifetime to the doubly excited level, broadens this level into a narrowcontinuum whose width is inversely related to that lifetime.

The presence of the configuration interaction also gives rise to character-istic absorption line profiles for photoionization in the vicinity of autoionizingresonances. The continuum state can, of course, be reached by photoexcita-tion of a core electron. If there were no configuration interaction, these twoprocesses would be distinct and the photoabsorption spectrum would con-sist of a discrete absorption line on a photoionization continuum, as shownin Fig. 4, left panel. However, with configuration interaction, the final statewave-function is as given in (209), and we get interference between the twochannels. The photoabsorption spectrum in this case looks like Fig. 4, rightpanel, which is called a Beutler-Fano absorption profile. Such features are ex-pected in the extreme ultraviolet spectra of nearby white dwarf stars due tophotoabsorption by neutral helium in the intervening interstellar medium [9].The features so far observed have been associated with autoionizing reso-nances of neutral helium.

Page 53: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 55

ω

σω

h

σω

ωh

Fig. 4. Spectra without configuration interaction (left) and Beutler-Fano profile(right)

Note that using the simple Z-scaling arguments we invoked earlier, au-toionization decay rates are roughly independent of Z for isoelectronic se-quences. This is because the outgoing continuum wave-function is propor-tional to E−1/2 ∼ Z−1, while the perturbation Hamiltonian ∼r−1 ∼ Z+1.Thus, the matrix element is ∼Z0. This means that autoionization is extremelyimportant for low Z ions, but becomes less and less important in comparisonto radiative decay for high Z ions. We will return to this shortly.

We can derive a rate coefficient for dielectronic capture by resorting todetailed balance arguments. The process is resonant, so the cross-section isactually infinite at the velocity which satisfies energy conservation:

12mv2

c = E∗∗i − Ei+1 (210)

where E∗∗i is the energy of the doubly excited recombined ion, and Ei+1 is

the energy of the ground-state of the initial ion. That is:

σdc(v) = αdcδ(v − vc) (211)

where αdc has units of cm3s−1. If ni+1 is the density of i + 1 ions in theground state, then the rate of dielectronic captures per unit volume per unittime is given by:

Rdc =∫d3vneni+1vσdc(v)f(v) = 4πneni+1αdcv

3ce

−mv2c/kT

( m

2πkT

)3/2

(212)

where f(v) is the Maxwellian distribution given in (183). If n∗∗i is the density

of i ions in the doubly excited state, then the autoionization rate per unitvolume is:

Rauto = n∗∗i Aa . (213)

These rates must be equal in thermodynamic equilibrium. But, in thermody-namic equilibrium, the level populations are given by:

n∗∗i

ni=g∗∗i

gie−(E∗∗

i −Ei)/kT (214)

Page 54: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

56 S.M. Kahn

where ni, gi and Ei are the density, degeneracy and energy of the ith ionin the ground state (see Sect. 5), and the ionization structure neni+1/ni isgiven by the Saha equation:

neni+1

ni=

2gi+1

gi

(mkT

2π2

)3/2

e−χ/kT (215)

where χ = Ei+1 − Ei is the ionization potential for the ith ion. Collectingterms gives:

αdc =g∗∗i

2gi+12π2

(

mvc

)3

Aa . (216)

Not surprisingly, the temperature drops out since dielectronic capture andautoionization must be related by fundamental constants.

The dielectronic capture rate is obtained by plugging (216) back into(212):

Rdc = neni+1g∗∗i

2gi+1Aa

(h2

2πmkT

)3/2

e−mv2c/2kT . (217)

To get the dielectronic recombination rate, as opposed to dielectronic capturerate, we must multiply the expression in (217) by the probability that thedoubly excited atom stabilizes radiatively. Quite generally, this probabilityis given by the ratio of the sum of all radiative decay rates from the excitedstate to the sum of all radiative plus autoionizing decay rates:

Probability of stabilization =∑Ar∑

(Ar +Aa). (218)

Usually, however, there is only one dominant decay channel in each case,which involves the decay of the core excitation. Thus, the dielectronic ratecoefficient becomes:

Rdr ≈ neni+1g∗∗i

2gi+1

(h2

2πmkT

)3/2

e−mv2c/2kT

(AaAr

Aa +Ar

). (219)

The factor in parenthesis has a maximum when Aa = Ar. Hence, dielectronicrecombination is efficient when the rates for autoionization decay and radia-tive decay of the core excitation are approximately equal. Since Aa ∼ Z0 andAr ∼ Z4, this is primarily the case for high-Z ions.

We can get a further quantitative feel for how these rates compare byagain using a semi-classical treatment. Note that the dielectronic captureprocess is very similar to collisional excitation, except that the final state ofthe colliding electron is now a bound state rather than a continuum state.We should therefore be able to get a rough idea of the rate coefficient for thisprocess by extending our earlier classical treatment of collisional excitation toenergies below threshold. Recall that our earlier expression for the excitationcross-section was given by (180):

Page 55: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 57

σmn(E) ≡ πa20

gmEeΩmn . (220)

For capture into principal quantum number n, we can integrate this expres-sion over the velocity range between neighboring Rydberg levels to yield anestimate for αdc associated with this core excitation:

αdcij,n ≈ σij(δv) ≈ σij

2Z2Ryd

n3mv=πe4

giΩij

Z2Ryd

n3m2v3c

. (221)

But Ωij/gi = 2π/√

3fijg/Eij (181) and

fij =32mc3

e2

2

E2ij

Arij (222)

(Equation 113). Plugging these expressions in and equating αdcij,n from (221)

to αdc from (216) we obtain:

Aaij

Arij

=12√

3gi+1

g∗∗i

gZ2

n3

(Ryd

Eij

)3 1α3

. (223)

Note that since Eij ∼ Z2, this ratio scales like Z−4, as expected from ourearlier discussion. Taking all other features to be of order unity, with Eij ∼Z2Ryd, this ratio is found to be ∼Z−4α−3. Setting it equal to unity (formaximum dielectronic recombination efficiency) then implies Z ≈ 40. So wesee that dielectronic recombination becomes important only for the higher-Zelements, most notably iron.

5 Types of Equilibria

In most astrophysical settings, some form of equilibrium applies, in whichthere is a balance between competing processes, e.g. heating and cooling,ionization and recombination, excitation and deexcitation, etc. The natureof the equilibrium has a very important effect on the emergent spectrum.

There are three “systems” which may or may not equilibrate with oneanother:

– the kinetic distributions of the electrons and ions;– the atomic level populations;– the radiation field.

We say that we have strict thermodynamic equilibrium when all three systemsare characterized by statistical distributions at the same temperature T . Inparticular, for this case, the radiation field is characterized by the blackbodydistribution, so the spectrum is especially simple. For absolute equilibrium,the temperature T , must also be independent of spatial position within the

Page 56: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

58 S.M. Kahn

gas. However, as long as the scale length for temperature variations: T/ |∇T |is long compared to all relevant mean free paths for particle and photoninteractions, it is appropriate to talk about strict local thermodynamic equi-librium, where T = T (r).

The more common term, local thermodynamic equilibrium (LTE) usuallyapplies to the situation where the particle distributions and level populationsare in equilibrium, but the radiation field is not, i.e. the scale lengths of thesystem are not sufficient to trap emitted photons and enforce thermalization.

5.1 Properties of LTE

In LTE, the population of a given energy level is proportional to the degen-eracy in that level and a Maxwell-Boltzmann factor e−E/kT . This gives riseto:

The Maxwellian velocity distribution for free particles

n(v)dvn

= 4πv2( m

2πkT

)3/2

e−mv22kT dv , (224)

The Maxwell-Boltzmann distribution for level populations

nzj

nz=

gzj

Uz(T )e−

Ezj −Ez

0kT , (225)

where Uz(T ) is the partition function:

Uz(T ) =∑

j

gzj e

−Ezj −Ez

0kT (226)

and The Saha equation for the ionization balance

nenz+1

nz=

2Uz+1(T )Uz(T )

(2πmkT )3/2

h3e−

χz

kT . (227)

The definition of Uz(T ) can be problematic. For example, for H-like atoms

gn = 2n2 (228)

e−En−E0

kT = e−z2Ry

kT (1− 1n2 ) (229)

⇒ Uz(T ) → ∞ ; (230)

we must truncate the expansion at some high Rydberg level. This is usuallya function of the particle density, due to the effects of neighboring charges.

In LTE, the prediction of the emergent spectrum requires the solution ofthe radiative transfer equations

Page 57: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 59

dIνdτν

= −Iν + Sν (231)

Sν =jνkν

(232)

dτν = kνds (233)

Here, Iν is the specific intensity of the radiation field, jν is the emissivity ofthe gas, and kν is the opacity, all of which are functions of the position alongthe path of propagation s. Sν is called the source function.

For discrete lines:

jnm =hνnm

4πgmnnA

nmϕ(ν) (234)

knm = gmnmσmn(ν) − gnnnσmn(ν) (235)

But from radiation theory, we found:

Anm =8π2

3e2

mc3fnmν

2 (236)

σmn(ν) = σnm(ν) =13πe2

mcfnm (237)

and, relating the level populations using the Maxwell-Boltzmann distribution(225), we get:

Snm =jnm

kmn=

2hν3nm

c21

(ehνnm/kT − 1)= Bνnm

(T ) (238)

which is the blackbody function evaluated at the frequency of the transitionνnm!

Looking inward to an optically thick medium at constant temperature,(231) implies:

Iν(τν) = Bν(T )(1 − e−τν ) (239)

The line intensities are “limited” to the blackbody intensity evaluated at thelocal temperature.

For the approximation of LTE to hold, we need the rates for collisionaldeexcitation of discrete levels to be comparable to the rates for spontaneousradiative decay:

neCnm(T ) ∼ Anm (240)

⇒ ne ∼ 9 1019T1/2K (δE)3keV cm−3 (241)

In astrophysical settings, such high densities are only reached in the at-mospheres of compact objects like white dwarfs and neutron stars.

When the assumption of LTE is invalid, the calculation of the emergentspectrum can be much more complicated. In general, we have to explicitly

Page 58: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

60 S.M. Kahn

log Iν

log ν

Fig. 5. An illustration of the limitation of line intensities to the blackbody intensityfor cases where LTE holds

account for all microphysical processes that feed and deplete the individualquantum levels. The most general, time-dependent equations are of the form:

dnzi

dt= −nz

i

∑j

Rij +∑z′,k

nz′k Rki (242)

where the R’s represent the rates for collisional and photon interactions cou-pling levels within the same charge state and in neighboring charge states.

5.2 Coronal Equilibrium

Equation (242) is difficult to solve because of the requirement for inclusion ofsuch a large array of diverse processes. Therefore, it is useful to adopt someapproximations, applicable to particular cases. One of the most importantsets of approximations applies to the case of coronal equilibrium, sometimesalso referred to as collisional ionization equilibrium.

There are three basic assumptions underlying this limit:

– Excitation and ionization are dominated by electron-ion collisions. Deex-citation is dominated by spontaneous radiative decay.

– Densities are low enough so that atoms are always in their ground states.– The radiation field has a negligible effect on the atomic populations, and

the plasma is optically thin, so photoabsorption and scattering can beignored.

Sources of applicability for these assumptions include: stellar coronae, theshocked gas of older supernova remnants, and the intracluster media of galaxyclusters.

The charge state distribution in coronal equilibrium is determined by abalance of collisional ionization and radiative and dielectronic recombination:

Page 59: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 61

dnz

dt= −nenz(Cz + αz) + nenz+1αz+1 + nenz−1Cz−1 (243)

Here Cz represents the rate coefficient for collisional ionization (see Sect. 5.4),and αz represents the combined RR + DR rate coefficient for recombina-tion (Sects. 4.5 and 4.6, respectively). Note that the characteristic timescalesfor equilibrium to be established are ∼(neC)−1 or ∼(neα)−1. These can belarger than 103 yr for ne ≤ 1 cm−3, as found in young supernova remnants.Since this age exceeds the age of the remnant (for the most recent super-novae), the shocked gas that we observe for these cases may still be ionizing,and the charge balance may be far from equilibrium. A similar situation canbe found during weak flares in stellar coronae. Here the electron density iscloser to ne ∼ 1010 cm−3, so the equilibration time is of order a few seconds,comparable in some cases to the duration of the flare.

However, if equilibrium is established, so that the left-hand side of (243)vanishes, the electron density ne, drops out of the equation, and the resultingsteady-state ionization structure becomes a function only of temperature.This turns out to be also true of the discrete spectrum. Specifically, since weare assuming that the atoms are “always” in the ground state, the populationsof upper levels are given by the ratio of collisional excitation rates from theground level, to the spontaneous radiative decay rates back down:

n2 =nen1γ12(T )

A21, (244)

and the line emissivities become:

ε21 = nen1γ21(T )E12 , (245)

where γ12(T ) is the collisional excitation coefficient (Sect. 5.3), and E12 isthe energy of the transition. The density of the ion in the ground state isgiven by n1 = AelemfZ(T )nH , where Aelem is the abundance of the elementrelative to hydrogen, and fZ(T ) is the steady-state ion fraction, as discussedabove. It is useful to define a line power for the transition: P21 = ε21/n

2e. We

thus get:

P21(T ) =(nH

ne

)Aelemfz(T )γ12(T )E12 (246)

which is typically expressed in units of erg cm3s−1.Actually, the “two-state” model discussed above is too simple, since im-

portant contributions to upper level populations can also come from ground-state excitations to higher levels, which then radiatively decay to intermediatestates. However, even these more complicated “channels” can still be incorpo-rated via the definition of more general, effective excitation rate coefficientsthat include these terms. A number of coronal equilibrium “spectral synthe-sis” codes have been developed over the years to provide these line powercalculations, and some are in widespread use in the community. The largest

Page 60: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

62 S.M. Kahn

residual uncertainties in these codes generally involve the treatment of theDR rates, and the completeness of the line lists.

For an intermediate charge state, the ion fraction, fZ , peaks in temper-ature at some particular value. The excitation rate coefficient, γ, generallyincreases across the range of temperatures where the ion exists in appreciableabundance. Therefore, the line power, P , exhibits a peak at a temperature of-ten called the temperature of formation, Tf . The presence of a particular linein the spectrum implies the existence of plasma at or near the temperatureof formation for that line. The modulation of line powers by the temperaturedependence of the ion fraction thus gives us a crude temperature diagnostic.

The measured line flux for a collisional plasma is given by:

F21 =e−NHσ(E21)

4πd2

∫dV dTn2

e(T, V )P21(T ) (247)

∼ e−NHσ(E21)

4πd2P21(Tf )

∫dV n2

e(Tf ) (248)

where e−NHσE21 is the attenuation factor through the interstellar and circum-source media, and d is the distance to the source. The integral that remainsin (247) is called the volume emission measure, V EM(Tf ). As indicated, itis a function of temperature. For an assumed set of abundances, and a givencolumn density, NH , the shape of the emergent spectrum for a coronal plasmais given completely by the shape of the volume emission measure distribution.

5.3 X-Ray Photoionization Equilibrium

A quite different set of approximations applies to the case of photoionizationequilibrium, where the presence of an intense continuum radiation field has asignificant effect on the ionization and thermal structure of the surroundinggas. The electrons are generally too cool to excite prominent X-ray lines inthis case, and excited levels are instead populated by direct recombination, byradiative cascades following recombination onto higher levels, and by directphotoexcitation from the continuum.

These conditions are typically found in the circumsource media ofaccretion-powered sources, such as X-ray binaries and active galactic nuclei.For example, in the accreting gas surrounding an X-ray binary, the energydensity in the continuum radiation field is given by:

Uγ ∼ L

4πR2c∼ 3.7 104 erg cm−3 (249)

where we have taken L ∼ 1038 erg s−1, and R ∼ 1011 cm. In contrast, thethermal energy density in the electron distribution is given by:

Ue ∼ 32nekT ∼ 2.4 erg cm−3 (250)

Page 61: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 63

Fig. 6. The power radiated (ε/n2e) of a cosmic abundance plasma as a function of

temperature in coronal equilibrium. The contributions of the individual elementsare indicated. Line radiation dominates at temperatures below 107K

for typical values of the electron density and temperature, ne ∼ 1012 cm−3,kT ∼ 10 eV.

In photoionization equilibrium, the ionization structure is determined bythe balance between photoionization and recombination.

nz

∫ ∞

0

dEFE

Eσz(E) = nenz+1αz+1(T ) (251)

where FE is the differential continuum flux, in units of erg cm−2 s−1 keV−1,σz(E) is the photoelectric cross-section as a function of energy (Sect. 3.8),and αz+1(T ) is the recombination coefficient, again including both RR andDR contributions. The equilibrium temperature is determined by the solutionof the equation of energy balance, where the rate of energy injection is dueto photoelectric heating, and the rate of energy loss is due to radiation:

∑elem,z

nz,elem

∫ ∞

0

dEFE

Eσz,elem(E)

(E − Ez,elem

thresh

)

=∑

elem,z

nenz,elemΛz,elem(T ) (252)

Page 62: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

64 S.M. Kahn

In the optically thin limit: FE = L4πR2 f(E), where f(E) is a normalized func-

tion containing the details of the spectral shape of the irradiating continuum.In addition, we can write nz,elem = AelemfznH , and ne = µenH , where µe,the mean number of electrons per hydrogen atom, is only a weak function ofgas parameters. Therefore “environment specific” factors are all embodied ina single quantity

ξ =L

nR2(253)

which is usually referred to as the ionization parameter. Given the specifica-tion of this ionization parameter, the self-consistent solution of the ionizationand energy balance equations yield the fz(ξ) values for all the elements, andT (ξ). A variety of codes are in widespread use to calculate these quantities.

Plots of the ionization structure of iron as a function of temperature forconditions of coronal equilibrium and photoionization equilibrium are shownin Fig. 7. Two important features are immediately apparent from this figure:

– First, the “dominance of closed shells” is much less obvious in the caseof photoionization equilibrium. Given the big jump in ionization poten-tial following the removal of all the electrons in a closed shell, the closedshell charge states (e.g. Ne-like and He-like) dominate over a wide rangeof temperature for a plasma in coronal equilibrium. However, for a pho-toionized plasma, photoionization out of inner shells (L-shell and K-shell)plays a significant role for the hard irradiating spectra characteristic ofaccretion-powered sources. This process is essentially unaffected by theremoval of outer valence electrons, eliminating any important distinctionbetween open shell and closed shell charge states.

– Second, the gas is significantly “overionized” relative to the electron tem-perature in a photoionized plasmas. For example, Ne-like iron (FeXVII)peaks at kTe = 10 eV in the photoionized case, while for the coronalplasma Ne-like iron peaks at kTe = 400 eV.

The significantly different temperatures appropriate to a given charge statefor coronal and photoionized plasmas lead to several important characteristicdifferences in the emergent X-ray spectra. For a coronal plasma, kT ∼ χ, theionization potential of the ion, and δE, the characteristic energies of the lineexcitations. The lines are formed primarily via collisional excitation fromthe ground state. The brightest lines are E1 transitions, or those “fed” byE1 transitions. In a photoionized plasma, kT χ and δE, so the electronshave insufficient energy to collisionally excite X-ray lines. Instead, lines areformed mostly by radiative cascades following recombination. Recombinationflux tends to distribute evenly among all the available levels. Hence, thebrightest lines tend to come from ions with the fewest states in the upperlevel configuration (e.g. K-shell ions). In addition, the cascades “rain” intothe lowest lying excited levels. Therefore, lines from these levels are usuallyquite bright. Often, these are higher order multipole transitions, with lowcollisional coupling strengths to the ground.

Page 63: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 65

Fig. 7. Plots of the ionization structure of iron as a function of temperature forcoronal equilibrium (top), and photoionization equilibrium (bottom). The elementsymbols refer to the isoelectronic charge state of iron, e.g. the curve labeled O refersto oxygen-like Fe (figure courtesy of Masao Sako)

However, the most useful spectroscopic diagnostics for distinguishing coro-nal equilibrium from photoionization equilibrium are the narrow radiativerecombination continua (RRC’s) expected for the latter case. In Sect. 4.5, wefound that RRC’s are described by

dW

dtdωdV∼(

ω

χ

)3

σPI(ω)(

χ2

mc2kT

)3/2

eχ/kT e−ω/kT (254)

For a coronal plasma, kT ∼ χ ∼ ω. The RRC’s are broad and do not havehigh contrast relative to the accompanying bremsstrahlung continuum. Onthe other hand, in a photoionized plasma, kT χ and ω. For this case, theRRCs are strong and fall off steeply with increasing energy. They resemble“lines” at moderate resolution. The relative width of this feature is a good

Page 64: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

66 S.M. Kahn

Fig. 8. Plots of characteristic emergent soft X-ray spectra for conditions appro-priate to a coronal plasma top and an X-ray photoionized plasma bottom. Notethat the coronal spectrum is more “rich”, due to the greater prominence of theFe L complex in that case. The photoionized spectrum is dominated by lines fromlower-Z K-shell elements, and by low temperature radiative recombination continua(figure courtesy of Masao Sako)

temperature diagnostic, and, if the width is larger than predicted, can signalthe presence of extra sources of heating in the gas.

This is illustrated in Fig. 9, which shows the predicted spectrum of neon ina photoionized plasma for electron temperatures of both 10 eV and 50 eV. Theformer is the expected temperature for these charge states, if photoelectricheating provides the only form of energy injection in the gas. The latter mightapply if there are other sources of heating which contribute. As can be seen,the discrete line spectra look very similar for the two cases. However, the RRC(near 9 A) is much broader and less pronounced at the higher temperature.

With the launches of the grating spectrometers on the Chandra and XMM-Newton observatories, we now have clear detections of these features in manysources. A particular dramatic case is illustrated in Fig. 10, which shows thespectrum of the bright Seyfert 2 galaxy NGC 1068, as obtained with the re-flection grating spectrometer on XMM-Newton [10] As can be seen, the spec-trum is rich in emission lines, especially H-like and He-like lines of carbon,nitrogen, oxygen, and neon. The RRC’s from most of these species are la-beled in the figure. They are narrow, indicating a low electron temperatureof a few eV, characteristic of a photoionized plasma. In NGC 1068, the soft

Page 65: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 67

Fig. 9. Plots of the expected spectra of H-like and He-like neon in photoionizedplasmas with electron temperatures of 10 eV top, and 50 eV bottom, but with similarion fractions. Note the differences in the RRC’s for the two cases (figure courtesyof Masao Sako)

X-ray spectrum is produced in an ionization cone, which is irradiated by anintense X-ray continuum emanating from a central obscured nucleus.

5.4 Thermal Instability in Photoionized Plasmas

It has been known for many years that X-ray photoionized plasmas can bethermally unstable in certain regions of ionization parameter space. Typically,this is represented by means of an “S-curve”, a plot of the temperature, de-rived by solving the equation of energy balance (252), versus an ionizationparameter Ξ = F/neT ∼ ξ/T . An example is shown in Fig. 11. On thecurve itself, the heating rate is equal to the cooling rate, so the gas is inthermal balance. To the right, heating dominates over cooling, as indicated,while to the left, cooling dominates over heating. On branches of the curvewhich have positive slope in this figure, the gas is thermally stable. Smallperturbations upward in temperature increase the cooling, while small per-turbations downward in temperature increase the heating. However, on thebranches which have negative slope, the gas is thermally unstable. A smallperturbation upward in temperature increases the heating, causing furthertemperature rise, while a small perturbation downward increases the cool-ing. Many different calculations of these effects exist in the literature, and

Page 66: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

68 S.M. Kahn

Fig. 10. XMM-Newton reflection grating spectrum of the prototypical Seyfert 2galaxy NGC 1068 [10]. Features of H-like and He-like ions from carbon to silicon, aswell as significant emission due to Fe L-shell transitions, dominate the spectrum ofits active nucleus. Bright, narrow RRC’s point unambiguously to the predominanceof recombination in a photoionized plasma. Strong higher order Rydberg transitions(np → 1s) are also present, implying the presence of photoexcitation as well

the resulting S-curves show a lot of variations, even for similar assumptions.However, most show some degree of thermal instability in similar regions of(Ξ, T )-space.

The thermal instability has important spectroscopic implications. Growthrates are ∼kcs where k is the wave number, and cs is the sound speed, upuntil a maximum value of k, the inverse of the so-called “Field length”, wherethey saturate due to the increasing importance of thermal conduction. Themedium is expected to “break” into multiple stable phases, which can coex-ist in pressure and ionization equilibrium. Gas in an unstable phase shouldquickly disappear, unless it is replenished on a timescale comparable to theinverse of the growth rate. We do not expect to see emission lines character-istic of ionization parameters in the unstable regimes.

The instability arises because of ionization through various atomic shells,which acts as a type of phase transition. The criterion for instability is:(

∂(C −H)∂T

< 0 (255)

Page 67: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 69

Fig. 11. The phase diagram for a photoionized gas with cosmic abundances irra-diated by a 10 keV bremsstrahlung spectrum (figure from [11])

where C represents the complete set of cooling processes, and H representsthe complete set of heating processes. Continuum and bound-state processescontribute to both C and H, but the latter dominate in the region of insta-bility. To see the effect of ionization, it is useful to group charge-states for agiven atomic shell, e.g. Fe L, Si K etc., but to also distinguish between twotypes: “X-ray ions”, such as Fe L, O K, Si K, Fe K, in which χ ∼ keV kTe,and “EUV ions”, such as Fe M, O L, He K, in which χ ≤ 100 eV ≤ kTe.

For the X-ray ions, the primary heating contribution is due to the photo-electric effect:

H = niζPE < ε > (256)

where ζPE is the photoionization rate per ion, and <ε > is the mean energyreleased in the photoelectron. The primary cooling contribution is due toradiative recombination:

C = neni+1αR(Te)kTe . (257)

Because the gas is in ionization balance, the photoionization rate must beequal to the recombination rate:

niζPE = neni+1αR (258)

In addition, <ε >∼ χ (kTe), so H C. As the ionization parameter isincreased, so that we ionize through an atomic shell, both H and C initiallyrise and then fall. One finds that this shell contributes a negative term tothe partial derivative in (256), during the rise and a positive term during thefall. Thus, each atomic shell contributes both an unstable and a stable lobe.

For the EUV ions, the same analysis holds, but in this case: kTe <ε >, so that C H, and the contribution is positive during the rise and

Page 68: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

70 S.M. Kahn

negative during the fall. The net thermal stability is determined by the sumof the contributions from all of these atomic shells. The situation can bequite complex, because the stable and unstable lobes contributed by thedifferent elements occur at different temperatures. One finds that there are“near cancellations”, which makes the total stability quite sensitive to detailsrelated to the elemental abundances and the shape of the ionizing spectrum.This can be beneficial, because we can exploit this sensitivity to derive strongconstraints on physical conditions in the gas, if the signatures of thermalinstability are visible in the spectra.

6 Discrete Line Diagnostics

The relative prominence of various emission line features in cosmic X-rayspectra is determined principally by the abundances of the different elements,and the locations of the K- and L-shell complexes associated with these ele-ments within the X-ray band. Scaling from the H-like isoelectronic sequence,the energies of the K-shell features are given roughly by:

EK ∼ (10 eV)Z2 , (259)

while the energies of the L-shell features are approximately:

EL ∼ (1.5 eV)Z2 . (260)

If we define the conventional soft X-ray band to cover the range 100 eV ≤E ≤ 10 keV, we see that it includes the K-shell features of beryllium (Z = 4)through gallium (Z = 31), and the L-shell features of oxygen (Z = 8) throughthallium (Z = 81).

A plot of standard cosmic abundances as a function of atomic numberappears in Fig. 12. Several features should be noted:

– The abundances drop precipitously with increasing Z above carbon (Z = 6).The abundances of lithium, beryllium, and boron (Z = 3, 4, and 5, re-spectively) are especially low.

– In general, elements with even values of Z have considerably higher abun-dances than elements with odd values of Z. This is a consequence of theimportance of α-chain reactions, in the production of the heavier elementsduring the late stages of stellar evolution.

– There is a very prominent abundance peak at iron (Z = 26) in the higherZ-range. This is a consequence of nuclear stability. 56Fe has the highestbinding energy per nucleon of any nucleus. Fusion reactions that pro-duce lower Z elements are exothermic, while above iron, fusion reactionsbecome endothermic.

Given these considerations, the most significant K-shell complexes in cosmicX-ray spectra are due to C, N, O, Ne, Mg, Si, S, Ar, Ca, Fe, and Ni, while the

Page 69: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 71

Fig. 12. A plot of the standard cosmic abundance of the elements as a function ofatomic number Z (figure courtesy of Masao Sako)

most significant L-shell complexes are associated with Si, S, Ar, Ca, Fe, andNi. It is one of the major strengths of cosmic X-ray spectroscopy that such awide range of elements and charge states is measured in a single wavelengthband.

6.1 Lyman Series Transitions in H-like Ions

At the characteristic temperatures of X-ray emitting plasmas, the low-Z abun-dant elements are often found in their H-like charge states. The most promi-nent emission lines are the Lyman series transitions:

Ly α1: 1s-2p 2P3/2; Ly α2: 1s-2p 2P1/2;Ly β1: 1s-3p 2P3/2; Ly β2: 1s-3p 2P1/2;Ly γ1: 1s-4p 2P3/2; Ly γ2: 1s-4p 2P1/2

...

The ratio of the line intensities for the two transitions in each case is givenroughly by the degeneracy factors, e.g.:

Ly α1/Ly α2 ∼ 2(3/2 + 1)2(1/2) + 1

= 2 .

Recall that the splitting is:

Page 70: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

72 S.M. Kahn

∆En,j = En(Zα)2

n2

[n

j + 1/2− 3/4

](261)

∆E1,2

E∼ (Zα)2

2n(262)

so these are barely resolvable, especially at low Z.These lines are usually quite bright, and are therefore good for abundance

and velocity determinations. Examples are shown in Fig. 13, which displaysthe XMM-Newton reflection grating spectrum of the supernova remnant SNR1E0102-72.3 in the Small Magellanic Cloud [12]. This young core collapseremnant is an oxygen-rich Type 1b SNR akin to Cas A [13], so the spectrum isdominated by lines of elements produced by α-burning reactions. The Lymanseries lines (α through γ) of H-like C, N, Ne, and Mg are clearly visible inthe spectrum, as marked in the figure.

Despite their prominence in astrophysical X-ray spectra, Lyman seriestransitions have rather limited utility as density and temperature diagnos-tics. Lines in this series are all produced through electric dipole transitions,so the radiative decay rates are high, and the collisional couplings are negli-gible. In addition, because of the n−2 dependence of the H-like energy levels

Fig. 13. The XMM-Newton reflection grating spectrum of SNR 1E0102-72.3 from[12]. For clarity, the spectrum is shown in both linear (top) and logarithmic (bottom)units. H-like and He-like emission lines from carbon to silicon are present with somesignificant emission from Fe L transitions as well

Page 71: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 73

(261), the upper levels for the different transitions in the series are close in en-ergy, so the Boltzmann factor in the excitation rates varies only slightly fromtransition to transition in the temperature range where the H-like ion is thedominant species (see Fig. 14). At the very low temperatures characteristic

Fig. 14. Plots of the ratio of higher series Lyman line intensities to the Lymanα line intensity as a function of temperature in O VIII, for both coronal plasmas(top), and photoionized plasmas (bottom)

Page 72: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

74 S.M. Kahn

of photoionized plasmas, Lyman series lines are formed by radiative cascadesassociated with radiative recombination. The line ratios produced by theseprocesses are somewhat different than those associated with collisional exci-tation in collisional plasmas. This is apparent from Fig. 14, where it can beseen that the Ly β to Ly α ratio for O VIII is ∼0.11 for a coronal plasma,and ∼0.14 for a photoionized plasma. Similar enhancements are found forthe higher series line ratios as well.

6.2 He-like Transitions

He-like K-shell lines are among the most important of all in the soft X-ray band. Since the He-like charge state is a tight “closed shell”, this isthe dominant ion species over a wide range in temperature, particularly incoronal plasmas. In addition, as explained below, these lines exhibit strongsensitivity to electron density, temperature, and ionization conditions in theemitting plasma.

The most important K-shell He-like transitions are as follows:

W : 1s2 1S0 – 1s2p 1P1

X: 1s2 1S0 – 1s2p 3P2

Y : 1s2 1S0 – 1s2p 3P1

Z: 1s2 1S0 – 1s2p 3S1

W is an electric dipole transition, also called the resonance transition, andis sometimes designated with the symbol r. X and Y are the so-called in-tercombination lines. These are usually blended (especially for the lower-Zelements), and are collectively designated with the symbol i. Z is the forbid-den line, often designated by the symbol f . It is a relativistic magnetic dipoletransition, with a very low radiative decay rate.

The temperature sensitivity of these lines arises as follows [14–16]: SinceW is an electric dipole transition, the collision strength for collisional excita-tion of this line includes important contributions from higher order terms inthe partial wave expansion, and thus continues to increase with energy abovethreshold. By contrast, X and Z are electric dipole forbidden. The dominantterm in the excitation collision strength for these transitions involves elec-tron exchange. Therefore, their excitation collision strengths drop off stronglywith energy above threshold, whereas Y remains relatively constant. As a re-sult, the line ratio: G = (X + Y + Z)/W is a decreasing function of electrontemperature.

The density sensitivity comes from the fact that the 3S1 level can becollisionally excited to the 3P levels. At high electron density, that processsuccessfully competes with radiative decay of the forbidden line. Therefore,the ratio R = Z/(X + Y ) drops off above a critical density, nc. The criticaldensity depends strongly on Z. For C V, nc ∼ 109 cm−3, while for Si XIII,nc ∼ 1013 cm−3.

Page 73: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 75

However, the R-ratio can also be affected by the presence of a significantultraviolet radiation field [14]. In particular, the 3S1 level can be photoexcitedto the 3P levels, prior to radiative decay, if there is sufficient ultravioletintensity at the energy of the relevant transitions. That leads to suppression ofthe forbidden line and enhancement of the intercombination lines, mimickingthe effects of high electron density.

These dependences are illustrated in Figs. 16 and 15, which shows the He-like spectra of oxygen, nitrogen, and carbon for two stellar coronal sources,Procyon and Capella, as measured with the Chandra low energy transmis-sion grating spectrometer [17]. The corona of Procyon is cooler than that ofCapella. As can be seen, the resonance lines are consequently less intense forProcyon, in comparison to both the intercombination and forbidden lines.Note that the forbidden line of carbon is also comparatively suppressed forProcyon in relation to the intercombination line. While this looks like a den-sity effect, it is actually due to the ultraviolet radiation field from this star.Procyon is an F star, with a relatively high UV flux.

In photoionized plasmas, the excited levels for He-like ions are fed directlyby recombination and also by radiative cascades following recombination ontohigher levels. The forbidden line is most intense, since most of the cascadesfrom high-n, high-l (high-J) levels land on the lowest lying 1s2s(J = 1) level,which produces the forbidden line. This is illustrated in Fig. 17, and can alsobe seen in the spectrum of NGC 1068 shown in Fig. 10 for both the He-likeoxygen lines near 22 A, and the He-like nitrogen lines near 29 A.

6.3 Iron L-Shell Transitions

Since iron is the most abundant high-Z element, its L-shell spectrum playsa crucial role in astrophysical X-ray spectroscopy. As a result of their higherionization potentials, the iron L-shell ions contribute significant line emissioneven when the lower-Z elements are full stripped. For collisionally ionizedplasmas, this complex samples a wide range in temperature (0.2–2 keV). Inaddition, the L-shell spectrum is very “rich”, and there is significant diag-nostic sensitivity.

The brightest iron L-shell lines are of the form:

2s22pk − 2s22pk−13d2s22pk − 2s22pk−13s2s22pk − 2s2pk3p

The 2p− 3d lines generally have the highest oscillator strength. The linepositions are a strong function of charge state. Thus, the ionization struc-ture is easily discernible, which provides a simple, abundance-independentconstraint on the temperature distribution.

Page 74: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

76 S.M. Kahn

Fig. 15. He-like complexes for O, N, and C from the coronal star Procyon, asmeasured with the Chandra low energy transmission grating spectrometer (From[17])

Page 75: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 77

Fig. 16. He-like complexes for O, N, and C from the coronal star Capella, asmeasured with the Chandra low energy transmission grating spectrometer (From[17])

Page 76: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

78 S.M. Kahn

Fig. 17. Calculated He-like emission line spectra of oxygen, magnesium, and siliconfor photoionization equilibrium top and coronal equilibrium bottom plasmas. Notethe prominence of the forbidden lines in the case of the photoionized plasmas (figurecourtesy of Masao Sako)

This is illustrated in Fig. 18, which shows the iron L spectrum of Capella,as observed with the Chandra high energy transmission grating spectrometer.Plotted below the measured data are the calculated contributions from eachof the individual charge states, ranging from Na-like iron (Fe XVI) to Be-likeiron (Fe XXII). Note the relatively clean separation between the L-shell com-plexes from each of these ions, allowing for relatively easy decomposition ofthe spectrum, even with only moderate resolution. The density sensitivity ofthe iron L complex arises from the fact that the intermediate iron L chargestates (e.g. N-like and C-like) possess a number of low lying metastable lev-els associated with n = 2 → n′ = 2 excitations. These can be populatedcollisionally, leading to new “seed” states for 2 → 3 excitations, followed by3 → 2 radiative decays. Such density diagnostics turn on at electron densities∼1013 cm−3.

6.4 The Iron K-Shell Complex

The iron K complex is relatively isolated in the spectrum at energies ∼6 −7 keV, where even non-dispersive detectors have moderate spectral resolution.Thus, iron K lines were the first discrete atomic features unambiguouslydetected for cosmic X-ray sources.

An important contributor to iron K emission, especially for accretion-powered sources, is due to fluorescence from cold material in the vicinity ofa bright X-ray continuum. Fluorescence involves a radiative decay follow-ing inner shell photoionization, i.e. a transition of the form 1s22s22pk−1nl−1s2s22pknl. The excited level, in this case, can also decay via autoionization

Page 77: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 79

Fig. 18. The spectrum of Capella obtained with Chandra high energy transmis-sion grating spectrum, compared with a calculated spectrum showing the separatecontributions of each of the iron L charge states (From [19])

by ejecting one of the outer electrons in the valence shell. This latter processdominates for low-Z elements. However, since radiative decay rates scalelike Z4, and autoionization decay rates scale like Z0, the fluorescence yieldbecomes appreciable for a high-Z element like iron. The near-neutral iron Kfluorescence line falls at 6.4 keV, easily distinguishable from the He-like linesnear 6.7 keV, and the Lyman α line at 7.1 keV.

The iron K complex also exhibits new features due to the relative im-portance of dielectronic recombination. DR leads to Li-like “satellites” toHe-like K-lines: 1s2pnl − 1s2nl. These satellites are shifted down in energy.Higher n implies a smaller shift, and is associated with a higher energy ofthe recombining electron. Therefore, the satellite spectrum is temperaturesensitive (cf. [20]).

At astrophysical densities, all atoms are in the ground state. Most ofthe satellite lines cannot be produced by collisional excitation of Li-like iron(e.g. 1s2p2 − 1s22p). They come purely from DR on He-like atoms. How-ever, other lines terminate in the ground configuration of the Li-like ion (e.g.1s2s2p− 1s22s). These can be produced by both collisional excitation of Li-like atoms, and DR on He-like atoms. Hence, the line ratios for these varioustransitions provide an independent measure of the charge balance. Analysisof the Fe K He-like spectrum thus provides independent constraints on the

Page 78: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

80 S.M. Kahn

electron temperature and the level of ionization, and is ideal for investigatingdepartures from ionization equilibrium.

7 Concluding Remarks

As a field, astrophysical X-ray spectroscopy is still in its infancy. While thegrating spectrometers on Chandra and XMM-Newton have already showeredus with fascinating results on a wide variety of diverse sources, most of thedata have not been completely reduced, and many sources bright enough toprovide reasonable spectra have still not yet been observed. A much largerpopulation of interesting sources are too faint for these instruments, butshould be amenable for study with the more sensitive experiments plannedfor future missions such as Constellation-X and XEUS.

The complete analysis of all of these observations will require a greaterlevel of spectroscopic sophistication than most X-ray astronomers are accus-tomed to. In the past, we have had the luxury of fitting relatively simple“canned” spectral models to low resolution, low statistics data. As the qual-ity of our spectra improves, these more familiar techniques no longer suffice.Some would prefer to ignore the complications, and continue to work only onthe faintest sources where the paucity of photons precludes worrying aboutspectral details. I have even heard some argue that we should not attempt tobuild higher resolution spectroscopic instruments, because the data they willacquire will be too difficult to interpret. I find this view to be very unscientific.We will always benefit by better instruments and better data.

In these lectures, I have tried to provide a synopsis of the kinds of issuesX-ray astronomers must consider in analyzing their spectroscopic data. Butthis is by no means a “user manual”. There are no simple codes that will takeproper account of all relevant processes, and provide a neat set of “results” atthe push of a button. We will all have to continue to learn as we go along. Thefirst data sets we have obtained have already pointed to holes in our existingatomic databases, and in our understanding of particular excitation processes.To make progress, we must complement our data analysis activities withdirect involvement in laboratory astrophysics experimentation, and atomiccalculation. Astronomers must become spectroscopists, and spectroscopistsmust become astronomers. This is how real progress will emerge.

Acknowledgments

I am indebted to a number of key individuals for helping me to finally makethese lecture notes available for publication. First, I would like to thank Pas-cal Favre of the Integral Science Data Centre, for his tremendous assistancewith the preparation of the manuscript. Second, I would like to thank my stu-dents and colleagues at Columbia: Ehud Behar, Jean Cottam, Mingfeng Gu,Ali Kinkhabwala, Maurice Leutenegger, Frits Paerels, John Peterson, Masao

Page 79: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Soft X-Ray Spectroscopy of Astrophysical Plasmas 81

Sako, and Daniel Savin for help with the figures, editing the text, and for con-tributing many of the ideas that are contained within. I have also benefitedfrom numerous conversations with current and previous collaborators, mostnotably Peter Beiersdorfer and Duane Liedahl at the Lawrence LivermoreNational Laboratory, and Bert Brinkman, Jelle Kaastra, and Rolf Mewe ofSRON, Utrecht. Finally, I would like to thank my hosts for the Saas Fee pro-gram: Manuel Gudel and Roland Walter, for inviting me to Les Diableretsand allowing me to participate in this distinguished lecture series.

References

1. Cowan, R., 1981, The Theory of Atomic Structure and Spectra, Los Alamosseries in Basic and Applied Science, University of California Press, Berkeley,CA

2. Rybicki, G. B., and Lightman, A. P., 1979, Radiative Processes in Astrophysics,Wiley, New York, 1979

3. Giacconi, R., Gursky, H., Paolini, F., et al., 1962, Phys. Rev. Lett., 9, 4394. Blandford, R., Fabian, A., Pounds, K., 2003, X-Ray Astronomy in the New

Millennium, Cambridge University Press5. Schlegel, E. M., 2002, The Restless Universe: Understanding X-Ray Astronomy

in the Age of Chandra and Newton. Oxford University Press6. Tucker, W., Tucker, K., 2001, Revealing the Universe: the Making of the Chan-

dra X-ray Observatory, Harvard University Press, Cambridge, MA7. Hutchinson, I. H. 1987, Principles of plasma diagnostics, Cambridge University

Press8. Lotz, W. 1967, ApJS, 14, 2079. Rumph, T., Bowyer, S., and Vennes, S., 1994, AJ, 107, 2108

10. Kinkhabwala, A., Sako, M., Behar, E., et al., 2002, ApJ, 575, 73211. Hess, C. J., Kahn, S. M., & Paerels, F. B. S., 1997, ApJ, 478, 9412. Rasmussen, A. P., Behar, E., Kahn, S. M., et al., 2001, A&A, 365, 23113. Blair, W.P., Morse, J. A., Raymond, J. C., et al., 2000, ApJ, 537, 66714. Gabriel, A. H., and Jordan, C., 1969, MNRAS, 145, 24115. Pradhan, A. K., 1982, ApJ, 263, 47716. Porquet, D., Mewe, R., Dubau, J., et al., 2001, A&A, 376, 111317. Ness, J.-U., Mewe, R., Schmitt, J. H. M. M., et al., 2001, A&A, 367, 28218. Kahn, S. M., Leutenegger, M. A., Cottam, J., et al., 2001, A&A, 365, 31219. Behar, E., Cottam, J., and Kahn, S., 2001, ApJ, 548, 96620. Dubau, J., Volonte, S., 1980, Reports on Progress in Physics, vol. 43, 199

Page 80: Soft X-Ray Spectroscopy of Astrophysical Plasmasuser.astro.columbia.edu/~jules/C3273_10/skahn.pdf · physical conditions in the source. Hence, X-ray spectra have high diagnostic utility.

Peter von Ballmoos