24.07.2014 Views

Thermal X-ray radiation (PDF) - SRON

Thermal X-ray radiation (PDF) - SRON

Thermal X-ray radiation (PDF) - SRON

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

High Energy Astrophysics<br />

Jelle Kaastra & Frank Verbunt<br />

January 12, 2010<br />

0


Chapter 3<br />

<strong>Thermal</strong> X-<strong>ray</strong> <strong>radiation</strong><br />

1


3.1 Introduction<br />

<strong>Thermal</strong> X-<strong>ray</strong> <strong>radiation</strong> is an important diagnostic tool for studying cosmic sources<br />

where high-energy processes are important. One can think of the hot corona of the<br />

Sun and of stars, solar and stellar flares, supernova remnants, cataclysmic variables,<br />

accretion disks in binary stars and around black holes (galactic and extragalactic),<br />

the diffuse interstellar medium of our Galaxy or external galaxies, the outer parts of<br />

active galactic nuclei (AGN), the hot intracluster medium, the diffuse intercluster<br />

medium. In all these cases there is thermal X-<strong>ray</strong> emission or absorption.<br />

We will see in this chapter that it is possible to derive many different physical<br />

parameters from an X-<strong>ray</strong> spectrum: temperature, density, chemical abundances,<br />

plasma age, degree of ionisation, irradiating continuum, geometry etc. In this chapter<br />

we focus on X-<strong>ray</strong> emission and absorption in optically thin plasma’s. For optically<br />

thick plasmas one needs to take account of the full <strong>radiation</strong> transport in<br />

order to understand these plasmas. In that case stellar atmosphere models become<br />

relevant. However, we will not treat those cases here and restrict our discussion to<br />

plasmas with τ 1.<br />

The power of high-resolution X-<strong>ray</strong> spectroscopy is shown in Fig. 3.1. This shows<br />

a simulated spectrum of Capella based on old low-resolution observations with the<br />

EXOSAT satellite. That the predictions work is shown in Fig. 3.2. This shows the<br />

success of the plasma emission codes developed at <strong>SRON</strong> by Rolf Mewe (1935–2004)<br />

and his colleagues.<br />

The strength of spectral analysis lies often in the details. While there are many<br />

spectral lines, some lines contain more information than others. Fig. 3.3 shows an<br />

example of this. It is a part of the spectrum of a flare star. The three lines shown –<br />

the famous Ovii triplet that we will encounter more often during this course – have<br />

an enormous diagnostic power.<br />

This triplet occurs in many other sources, with often totally different intensity<br />

ratio. See for example Fig. 3.4 and compare the intensities of the triplet with those<br />

in Fig. 3.3. While EQ Peg is in collisional ionisation equilibrium, NGC 1068 is in<br />

photoionisation equilibrium. These terms will be explained later.<br />

Up to now you only saw examples of emission spectra. A nice example of an<br />

absorption spectrum is shown in Fig. 3.5.<br />

In order to see thermal imprints on a spectrum sometimes requires the skills<br />

of a spectroscopist. See for example Fig. 3.6. There has been a fierce debate in<br />

the literature whether the sharp change in the spectrum near 18 Å is due to an<br />

absorption edge (lower flux to the left) or a broad emission line (higher flux to the<br />

right). Here we only note that this course is meant to train you as a spectroscopist<br />

so that you can judge such issues yourself.<br />

For the proper calculation of an X-<strong>ray</strong> spectrum, one should consider three different<br />

steps:<br />

1. the determination of the ionisation balance<br />

2. the determination of the emitted spectrum<br />

3. possible absorption of the photons on their way towards Earth<br />

However, before discussing these processes we will first give a short summary of<br />

atomic structure and radiative transitions, as these are needed to understand the<br />

basic processes.<br />

2


Figure 3.1: The predicted X-<strong>ray</strong> spectrum of Capella. The picture illustrates the<br />

remarkable increase in sensitivity and resolution afforded by the Reflection Grating<br />

Spectrometer (RGS) of XMM-Newton by comparing the predicted RGS spectrum<br />

of the star Capella with that obtained with the EXOSAT gratings (inset). The<br />

throughput is three orders of magnitude, and the spectral resolution more than<br />

a factor of 20, better than achieved with EXOSAT. ( c○ G. Branduardi-Raymont,<br />

www.mssl.ucl.ac.uk/www astro/rgs/rgs impact.html)<br />

3


Figure 3.2: The observed X-<strong>ray</strong> spectrum of Capella. The spectrum is taken with<br />

the Reflection Grating Spectrometer (RGS) of XMM-Newton. The plot shows the<br />

wealth of emission lines in this source. From Audard et al., A&A, 365, L329 (2001).<br />

4


Figure 3.3: The observed X-<strong>ray</strong> spectrum<br />

with the RGS of XMM-Newton<br />

of the flare star EQ Pegasi. The plot<br />

shows the so-called Ovii triplet, consisting<br />

of the resonance, intercombination<br />

and forbidden line. The relative<br />

line ratio’s of this triplet contain a<br />

wealth of physical information. Image<br />

courtesy of J. Schmitt, Hamburger<br />

Sternwarte, Germany and ESA.<br />

Figure 3.4: X-<strong>ray</strong> spectrum of the<br />

Seyfert 2 galaxy NGC 1068. From<br />

Kinkhabwala et al., ApJ, 575, 732<br />

(2002).<br />

Figure 3.5: X-<strong>ray</strong> spectrum of the Seyfert 1 galaxy IRAS 13349+2438, obtained<br />

with the RGS of XMM-Newton. From: Sako et al., A&A 365, L168.<br />

5


Figure 3.6: X-<strong>ray</strong> spectrum of the Seyfert 1 galaxy MCG 6-30-15, obtained with the<br />

RGS of XMM-Newton.<br />

6


3.2 Summary of spectroscopic notations<br />

3.2.1 Atoms or ions with one valence electron<br />

The electrons in an atom have orbits with discrete energy levels and quantum numbers.<br />

The valence electron or light-electron in an atom or charged ion is the electron<br />

in the outermost atomic shell which is responsible for the emission of light. The<br />

energy configuration can be described by four quantum numbers:<br />

n - principal quantum number, characteristic for the relevant shell. It takes discrete<br />

values n =1, 2, 3, . . .. An atomic shell consists of all electrons with the same<br />

value of n.<br />

l - quantum number of angular momentum of orbital motion of the electron: l =<br />

0, 1, 2, . . . n − 1.<br />

s - quantum number of spin of electron: s = ±1/2 (for l > 0) or s = 1/2 (for<br />

l = 0).<br />

j - quantum number of total angular momentum of electron: j = l±1/2 (for l > 0)<br />

or j = 1/2 (for l = 0).<br />

All angular momenta are measured in units of = h/2π (h is Planck’s constant).<br />

The magnitude of the angular momentum vectors is given by<br />

| ⃗ l| = √ l(l + 1) ≈ l (l ≫ 1).<br />

Note that ⃗j = ⃗ l + ⃗s.<br />

For hydrogenic orbits (Coulomb field) in the theory of Bohr, the semi-major<br />

axis r of an elliptical orbit in shell n is given by r = n 2 a 0 /Z, where Z is the<br />

nuclear charge and the Bohr radius a 0 equals 5.291772108 × 10 −11 m (∼0.529 Å).<br />

Note that a 0 = α/4πR ∞ with α the fine structure constant and R ∞ the Rydberg<br />

constant R ∞ = m e cα 2 /2h; R ∞ hc = 1 2 α2 m e c 2 is the Rydberg energy (13.6056923 eV,<br />

2.17987209 × 10 −18 J). Thus, heavy nuclei are more compact, and also the orbits<br />

with small n are more compact than orbits with large n.<br />

The principal quantum number n corresponds to the energy I n of the orbit.<br />

In the classical Bohr model for the hydrogen atom the energy I n = E H n −2 with<br />

E H = 13.6 eV the Rydberg energy. Thus, electrons with large value of n have small<br />

energies (and therefore are most easily lost from an atom by collisons).<br />

Further, the characteristic velocity of the electron in the Bohr model is given<br />

by v = αZc/n. Electrons withsmall n or in atoms with large Z have the highest<br />

velocities. These velocities are sometimes useful to see if an interaction is important<br />

or not. For instance, if a free electron or proton collides with an atom, it will have the<br />

longest and most intense interaction with those electrons that have similar orbital<br />

velocity as the velocity of the invader.<br />

The following notation for electron orbits is frequently used:<br />

l = 0 1 2 3 4 5 6 7 8 9<br />

s p d f g h i k l m electrons<br />

and further alphabetic (with no duplication of s and p, or course!). The characters<br />

s, p, d and f originate from sharp, principal, diffuse and fundamental series (the line<br />

7


series of alkaline spectra). The notation is complicated, but could have been worse<br />

if for example the Chinese alphabet would have been used.<br />

Example 3.1. Take the hydrogen atom. An electron in the ground state has n = 1<br />

and l = 0; this is called a 1s orbit. For n = 2, l = 0 we have a 2s orbit; for n = 2,<br />

l = 1 we have a 2p orbit. Which are the orbits and quantum numbers for n = 3?<br />

The spin quantum number s of an electron can take values s = ±1/2, and the<br />

combined total angular momentum j has a quantum number with values between<br />

l − 1/2 (for l > 0) and l + 1/2. Subshells are subdivided according to their j-value<br />

and are designated as nl j . Example: n = 2, l = 1, j = 3/2 is designated as 2p 3/2 .<br />

3.2.2 Atoms or ions with more than one electron<br />

Above we summarised the situation where only one electron plays a role in the<br />

formation of spectra. Now we consider a multi-electron system. For each of the<br />

electrons one may, in analogy with the single electron configuration, give the orbital<br />

and spin angular momentum of the electrons, like ⃗ l 1 , ⃗ l 2 , ⃗ l 3 , . . . and ⃗s 1 , ⃗s 2 , ⃗s 3 , . . .. Similar<br />

to the single electron system, we need to compose the total angular momentum ⃗ J<br />

from individual angular momenta. Now one needs to consider the coupling between<br />

the electrons. In general, two different types of coupling are used, the so-called jj<br />

coupling and the Russell-Saunders or LS coupling:<br />

1. jj coupling:<br />

⃗j i = l ⃗ i + ⃗s i<br />

∑<br />

⃗j i = J ⃗<br />

This coupling may occur in complicated energy configurations.<br />

i<br />

2. LS coupling: relevant for simple configurations, for example the light elements,<br />

alkaline and earth-alkaline elements (groups I and II of the periodic system).<br />

As the LS coupling is applicable in many situations, we will give a more detailed<br />

treatment here. The coupling is determined by first combining the individual orbital<br />

angular momenta and spins, and then adding the total angular momentum to the<br />

total spin:<br />

∑<br />

i<br />

⃗ li<br />

∑<br />

⃗s i<br />

i<br />

⃗L + ⃗ S =<br />

= ⃗ L<br />

= ⃗ S<br />

⃗ J<br />

In these equations, ⃗ L is the total orbital angular momentum, ⃗ S the total spin, and<br />

⃗J the total angular momentum.<br />

8


The corresponding quantum number L is an integer: L = 0, 1, 2, . . .. Further,<br />

S is integer (for an even number of electrons), or broken (for an odd number of<br />

electrons), because the individual ⃗s i must be either parallel or anti-parallel.<br />

Example 3.2. For helium, with two electrons, we have S = 0 or S = 1.<br />

Finally, J is also integer or broken, for an even or odd number of electrons,<br />

respectively. The allowed values for J are:<br />

J = |L − S|, |L − S| + 1, . . ., |L + S|<br />

Example 3.3. Below are a few illustrative cases:<br />

L S allowed values for J<br />

2 1 1,2,3<br />

3 1/2 5/2, 7/2<br />

1 0 1<br />

Similar to the single electron notation, there is a standard notation for the orbital<br />

angular momentum of the multi-electron configurations:<br />

L = 0 1 2 3 4 5 6 7 8 9<br />

S P D F G H I K L M levels<br />

Note the capitals in the notation!<br />

For L ≥ S there are r = 2S + 1 levels for a given value of L. These levels are<br />

only distinguished by a different magnetic interaction of ⃗ L and ⃗ S, and sometimes<br />

they have the same energy (so-called degeneration). Such a group with r levels are<br />

called a term with multiplicity r.<br />

Although for L < S the multiplicity is given by 2L + 1 ≠ r, even in those cases<br />

we call r the multiplicity.<br />

Example 3.4. In Helium or any other two electron system, for the situation with<br />

one 1s electron and another electron in the 2s orbit, the S = 1 state has L = 0<br />

(because both electrons have l = 0). Therefore, only J = 1 is allowed here, and the<br />

multiplicity of 1 is less than 2S + 1.<br />

The following notation for multiplicity is used:<br />

S = 0 1/2 1 3/2 2 5/2 3 7/2<br />

r = 1 2 3 4 5 6 7 8<br />

singlet doublet triplet quartet quintet sextet septet octet<br />

3.2.3 Building electron configurations<br />

A multi-electron atom or ion can be build by starting from a bare nucleus and adding<br />

subsequently electrons one by one. In doing this, we should take account of the Pauli<br />

principle, which states that no two identical electron states can occur. Therefore in a<br />

single atom there cannot exist two electrons with the same combination of quantum<br />

numbers, for example n, l, j and m j (for m j see §3.2.5). Electrons with the same<br />

value of n and l but with different value of j are called equivalent electrons.<br />

How is then such a complex ion build?<br />

9


• The first atomic shell (n = 1) has l = 0, and there are two possible orientations<br />

of the spin ⃗s, hence there are two 1s orbits. If there is one electron in this<br />

shell, the designation is 1s. If the shell is full, with two electrons, we write 1s 2<br />

for a closed shell.<br />

• The second shell (n = 2) has as one particular possibility l = 0, s = ±1/2,<br />

and these orbits are called 2s orbits. But however for n = 2 we also can have<br />

l = 1.<br />

• For l = 1, the orbital angular momentum can have three orientations, given<br />

by m l = −1, 0, 1, and again the electron spins can have two different values,<br />

s = ±1/2. This then leads to 6 different p-orbits, and since n = 2, these are<br />

2p orbits. The full n = 2 shell is now designated as 1s 2 2s 2 2p 6 . Frequently one<br />

omits the filled shells, so that for example 1s 2 2s 2 2p 5 is abbreviated as simply<br />

2p 5 .<br />

• The third closed shell is 1s 2 2s 2 2p 6 3s 2 3p 6 3d 10 , etc. There are 10 d-orbits<br />

because there we have l = 2 and hence m l = −2, −1, 0, 1, 2 which can be<br />

combined with two different electron spin values.<br />

• The sequence of filling subshells obeys the following rules: first fill orbits in<br />

order of increasing n + l, and where n + l has the same value, fill the orbits<br />

in order of increasing n. Therefore, the order of filling is 1s, 2s, 2p, 3s, 3p, 4s,<br />

3d, 4p, 5s, 4d, 5p, 6s, 4f, 5d, 6p, . . .. The 4f orbit is an inner orbit, that is<br />

filled only at relatively high nuclear charge (the rare earth elements).<br />

A closed shell has no net contribution to the angular momentum:<br />

∑<br />

⃗l i = 0<br />

i<br />

∑<br />

⃗s i = 0<br />

i<br />

∑<br />

⃗j i = 0<br />

i<br />

∑<br />

l i = even<br />

i<br />

∑<br />

s i = integer.<br />

Each given electron configuration corresponds to a certain group of terms which<br />

all have approximately similar energies (see also §3.4.3).<br />

3.2.4 Term notation<br />

In general, a term with quantum numbers LSJ is designated as r (L-symbol) J . We<br />

can illustrate this with an example.<br />

Example 3.5. Take the electron configuration 1s 2 2s 2 2p 6 3s 2 3p 2 , with L = 1,<br />

S = 1, J = 2. This is designated as 1s 2 2s 2 2p 6 3s 2 3p 2 3 P 2 , or shortly with omission<br />

of the closed shells, as 3p 2 3 P 2 (pronounce as ”three p two triplet p two”). This is<br />

a triplet term.<br />

In general we omit all shells from the designation that are spectroscopically<br />

irrelevant. Whenever the algebraic sum ∑ l i = even or odd, the term is even or<br />

i<br />

odd. For example, the doublet term 2 D is even, while the triplet term 3 P o is odd.<br />

Therefore sometimes the index o from ”odd” is added just like in the triplet P term<br />

above. In the table below we give as examples the ground state configuration of a<br />

few atoms.<br />

i<br />

10


Z Element configuration term<br />

1 H 1s 2 S 1/2<br />

2 He 1s 2 1 S 0<br />

3 Li 2s 2 S 1/2<br />

4 Be 2s 2 1 S 0<br />

5 B 2p 2 P 1/2<br />

6 C 2p 2 3 P 0<br />

All noble gases have a closed shell with L = 0, S = 0 and as a consequence their<br />

ground state is always 1 S 0 . The ground term of a shell that is filled only half is<br />

always an S term, for example for nitrogen 2p 3 4 S 3/2 .<br />

3.2.5 Statistical weight<br />

The statistical weight of an energy level with quantum number J is the number<br />

of directions that the vector ⃗ J can take with respect to some preferred direction,<br />

for example the magnetic field. This is also the allowed number of sublevels. The<br />

projection of ⃗ J on the preferred direction is called MJ (or m j for a single electron<br />

with angular momentum ⃗j). In the case of a magnetic field M J is the socalled<br />

magnetic quantum number (Zeeman effect). M J can take any value between<br />

−J, −J + 1,...,0,...,J − 1, J. This gives 2J + 1 possible orientations of ⃗ J with<br />

respect to the preferred direction. Therefore the statistical weight g J = 2J + 1.<br />

Example 3.6. The 3 D 2 level has g 2 = 5. Note that for 3 D we have L = 2,<br />

S = 1. The total statistical weight of the 3 D term (levels with J = 3, 2, and 1) is<br />

g = 7 + 5 + 3 = 15.<br />

In a strong magnetic field the vectors ⃗ L and ⃗ S are decoupled ( ⃗ J does not exist<br />

any more) and they are arranged independently with respect to ⃗ B. Instead of the<br />

4 quantum numbers n, L, J and M J (Zeeman effect), one then takes n, L, M L and<br />

M S , with M L and M S the projections of ⃗ L and ⃗ S on ⃗ B (Paschen Back effect). In that<br />

case we find for the 3 D term M L = −2, −1, 0, 1, 2, M S = −1, 0, 1 → g = 5 × 3 = 15.<br />

3.2.6 The periodic system<br />

With the knowledge obtained before we now consider the periodic system. The<br />

table shows how the electronic subshells are being filled. Note also that there are<br />

sometimes small irregularities, such as for Mn.<br />

11


Z El 1s 2s 2p 3s 3p 3d 4s Ground term<br />

1 H 1 2 S 1/2<br />

2 He 2 1 S 0<br />

3 Li 2 1 2 S 1/2<br />

4 Be 2 2 1 S 0<br />

5 B 2 2 1 2 P 1/2<br />

6 C 2 2 2 3 P 0<br />

7 N 2 2 3 4 S 3/2<br />

8 O 2 2 4 3 P 2<br />

9 F 2 2 5 2 P 3/2<br />

10 Ne 2 2 6 1 S 0<br />

11 Na 2 2 6 1 2 S 1/2<br />

12 Mg 2 2 6 2 1 S 0<br />

13 Al 2 2 6 2 1 2 P 1/2<br />

14 Si 2 2 6 2 2 3 P 0<br />

15 P 2 2 6 2 3 4 S 3/2<br />

16 S 2 2 6 2 4 3 P 2<br />

17 Cl 2 2 6 2 5 2 P 3/2<br />

18 Ar 2 2 6 2 6 1 S 0<br />

19 K 2 2 6 2 6 1 2 S 1/2<br />

20 Ca 2 2 6 2 6 2 1 S 0<br />

21 Sc 2 2 6 2 6 1 2 2 D 3/2<br />

22 Ti 2 2 6 2 6 2 2 3 F 2<br />

23 V 2 2 6 2 6 3 2 4 F 3/2<br />

24 Cr 2 2 6 2 6 5 1 7 S 3<br />

25 Mn 2 2 6 2 6 5 2 6 S 5/2<br />

26 Fe 2 2 6 2 6 6 2 5 D 4<br />

27 Co 2 2 6 2 6 7 2 4 F 9/2<br />

28 Ni 2 2 6 2 6 8 2 3 F 4<br />

29 Cu 2 2 6 2 6 10 1 2 S 1/2<br />

30 Zn 2 2 6 2 6 10 2 1 S 0<br />

Atoms or ions with multiple electrons in most cases have their shells filled starting<br />

from low to high n and l. For example, neutral oxygen has 8 electrons, and the shells<br />

are filled like 1s 2 2s 2 2p 4 , where the superscripts denote the number of electrons in<br />

each shell. Ions or atoms with completely filled subshells (combining all allowed<br />

j-values) are particularly stable. Examples are the noble gases neutral helium, neon<br />

and argon, and more general all ions with 2, 10 or 18 electrons. In chemistry, it is<br />

common practice to designate ions by the number of electrons that they have lost,<br />

like O 2+ for doubly ionised oxygen. In astronomical spectroscopic practice, more<br />

often one starts to count with the neutral atom, such that O 2+ is designated as Oiii.<br />

As the atomic structure and the possible transitions of an ion depend primarily on<br />

the number of electrons it has, there are all kinds of scaling relationships along<br />

so-called iso-electronic sequences. All ions on an iso-electronic sequence have the<br />

same number of electrons; they differ only by the nuclear charge Z. Along such<br />

a sequence, energies, transitions or rates are often continuous functions of Z. A<br />

well known example is the famous 1s−2p triplet of lines in the helium iso-electronic<br />

sequence (2-electron systems).<br />

12


3.2.7 How to determine allowed terms<br />

In the example below we illustrate how it is possible to determine the allowed terms<br />

for a given electron configuration. We will determine the terms that may arise for a<br />

carbon atom described by the ground state configuration.<br />

The electron configuration is: 1s 2 2s 2 2p 2 . We have two closed subshells, 1s and<br />

2s, which both give 1 S 0 . We may therefore concentrate on the p-shell with two<br />

equivalent p-electrons. In this we include only states that are allowed with respect<br />

to the Pauli principle. Of course, we omit also states that are symmetric, i.e. a<br />

situation with electron 1 in state a and electron 2 in state b is the same as electron<br />

1 in state b and electron 2 in state a.<br />

The recipe is as follows:<br />

m l1 s 1 m l2 s 2 M L M S<br />

+1 + 0 + +1 +1 x<br />

+1 + −1 + 0 +1 x<br />

0 + −1 + −1 +1 x<br />

+1 + +1 − +2 0<br />

+1 + 0 − +1 0 x<br />

+1 − 0 + +1 0<br />

+1 + −1 − 0 0 x<br />

0 + 0 − 0 0<br />

−1 + +1 − 0 0<br />

−1 + 0 − −1 0 x<br />

−1 − 0 + −1 0<br />

−1 + −1 − −2 0<br />

+1 − 0 − +1 −1 x<br />

+1 − −1 − 0 −1 x<br />

−1 − 0 − −1 −1 x<br />

1. Choose the state with highest M S , in this case 1. This will also be the value<br />

of S for the term. Hence, the multiplicity defined by r = 2S + 1 = 3.<br />

2. Determine the highest M L for this M S . The value will also be the value of L<br />

for the term.<br />

3. Cross out one state function for each pair M L , M S which correspond to the<br />

possible quantum numbers M L = −L, −L + 1, . . ., 0, . . .,L − 1, L and M S =<br />

−S, −S + 1, . . ., 0, . . .,S − 1, S.<br />

4. Repeat the procedure stepwise for remaining functions.<br />

Application to the above case:<br />

1. Maximum M S = 1 ⇒ S = 1, thus, this is a triplet.<br />

2. The highest M L for M S = 1 is M L = 1 ⇒ L = 1, thus, it is a P state. The<br />

first term is thus 3 P.<br />

3. Cross out a state function for each set of combinations of M L = −1, 0, +1 with<br />

M S = −1, 0, +1, for instance the states labeled x in the table above. See table<br />

below for what is left over.<br />

13


m l1 s 1 m l2 s 2 M L M S<br />

+1 + +1 − +2 0 x<br />

+1 − 0 + +1 0 x<br />

0 + 0 − 0 0 x<br />

−1 + +1 − 0 0<br />

−1 − 0 + −1 0 x<br />

−1 + −1 − −2 0 x<br />

We see that there are only states left with M S = 0. A repetition of the steps<br />

1–3 gives S = 0 and L = 2, therefore we get a 1 D state. Next we cross out a state<br />

for each combination of M L = −2, −1, 0, +1, +2 with M S = 0, for instance those<br />

indicated with x above. See below what is left:<br />

m l1 s 1 m l2 s 2 M L M S<br />

−1 + +1 −1 0 0<br />

The remainder is clearly a 1 S state. Thus, we have shown that the allowed terms<br />

for the combination 2p 2 are given by 3 P, 1 D and 1 S.<br />

exercise 3.1. Show that the total statistical weight for these three configurations<br />

is 15. It is equal to the number of combinations that we had in our original table,<br />

as it should be.<br />

3.2.8 Allowed terms<br />

The allowed terms for some configurations are shown in Table 3.1 and Table 3.2.<br />

Table 3.1: Terms for non-equivalent electrons<br />

Configuration<br />

s s<br />

s p<br />

s d<br />

p p<br />

p d<br />

d d<br />

Terms<br />

1 S, 3 S<br />

1 P, 3 P<br />

1 D, 3 D<br />

1 S, 1 P, 1 D, 3 S, 3 P, 3 D<br />

1 P, 1 D, 1 F, 3 P, 3 D, 3 F<br />

1 S, 1 P, 1 D, 1 F, 1 G, 3 S, 3 P, 3 D, 3 F, 3 G<br />

14


Table 3.2: Terms for equivalent electrons<br />

Configuration<br />

s 2<br />

p 2<br />

p 3<br />

p 4<br />

p 5<br />

p 6<br />

d 2<br />

d 3<br />

d 4<br />

d 5<br />

Terms<br />

1 S<br />

1 S, 1 D, 3 P<br />

2 P, 2 D, 4 S<br />

1 S, 1 D, 3 P<br />

2 P<br />

1 S<br />

1 S, 1 D, 1 G, 3 P, 3 F<br />

2 P, 2 D(2), 2 F, 2 G, 2 H, 4 P, 4 F<br />

1S(2), 1 D(2), 1 F, 1 G(2), 1 I, 3 P(2), 3 D, 3 F(2), 3 G, 3 H, 5 D<br />

2 S, 2 P, 2 D(3), 2 F(2), 2 G(2), 2 H, 2 I, 4 P, 4 D, 4 F, 4 G, 6 S<br />

15


3.3 Energy levels<br />

We have seen the notation for the various electronic shells in an ion being designated<br />

as 1s, 2p etc. Except for this notation, there is yet another notation that is commonly<br />

being used in X-<strong>ray</strong> spectroscopy. Shells with principal quantum number n =1, 2,<br />

3, 4, 5, 6 and 7 are indicated with K, L, M, N, O, P, Q. A further subdivision is<br />

made starting from low values of l up to higher values of l and from low values of<br />

J up to higher values of J:<br />

1s 2s 2p 1/2 2p 3/2 3s 3p 1/2 3p 3/2 3d 3/2 3d 5/2 4s etc.<br />

K L I L II L III M I M II M III M IV M V N I<br />

The binding energy I of K-shell electrons in neutral atoms increases approximately<br />

as I ∼ Z 2 (see Table 3.3 and Figs. 3.7–3.8). Also for other shells the energy increases<br />

strongly with increasing nuclear charge Z.<br />

Figure 3.7: Energy levels of atomic<br />

subshells for neutral atoms.<br />

Figure 3.8: Energy levels of atomic<br />

subshells for neutral atoms.<br />

The binding energy of more highly ionised ions is in general larger than for neutral<br />

atoms (see Table 3.3). For example, neutral iron has a K-shell energy of 7.1 keV,<br />

Fexxvi has I = 9.3 keV. In general, if there is an ionised layer of gas between us<br />

and any X-<strong>ray</strong> source, the position of absorption edges tells you immediately which<br />

ions are present.<br />

It is common practice in X-<strong>ray</strong> spectroscopy to designate ions also according to<br />

the number of electrons, thus Fexxvi is hydrogen-like iron, Fexxv is helium-like,<br />

etc. The atomic physics for ions with the same number of electrons is very similar,<br />

and often the relevant parameters are smooth functions of Z, allowing interpolation<br />

along the so-called iso-electronic sequences.<br />

16


Table 3.3: Binding energies (in eV) of 1s electrons in the K-shell of free atoms and<br />

ions in the ground state<br />

Z el I II III IV V VI VII VIII IX X XI XII XIII<br />

1 H 13.6<br />

2 He 24.59 54.4<br />

3 Li 58 76 122<br />

4 Be 115 126 154 218<br />

5 B 192 206 221 259 340<br />

6 C 288 306 325 344 392 490<br />

7 N 403 426 448 471 494 552 667<br />

8 O 538 565 592 618 645 671 739 871<br />

9 F 694 724 754 785 815 846 876 954 1103<br />

10 Ne 870 903 937 971 1005 1039 1074 1108 1196 1362<br />

11 Na 1075 1101 1139 1177 1215 1253 1291 1329 1367 1465 1649<br />

12 Mg 1308 1333 1360 1402 1444 1486 1528 1570 1612 1654 1762 1963<br />

13 Al 1564 1590 1618 1646 1692 1738 1784 1830 1876 1922 1968 2086 2304<br />

14 Si 1844 1872 1901 1930 1959 2009 2059 2109 2160 2210 2260 2310 2438<br />

15 P 2148 2178 2208 2238 2269 2300 2354 2408 2462 2517 2571 2625 2679<br />

16 S 2476 2508 2540 2571 2603 2636 2668 2726 2785 2843 2902 2960 3018<br />

17 Cl 2829 2862 2896 2929 2962 2996 3030 3064 3126 3189 3252 3315 3377<br />

18 Ar 3206 3241 3276 3311 3346 3381 3417 3452 3488 3554 3621 3688 3755<br />

19 K 3610 3644 3681 3718 3754 3791 3828 3866 3902 3940 4010 4081 4152<br />

20 Ca 4041 4075 4110 4149 4187 4225 4264 4303 4342 4380 4420 4494 4569<br />

21 Sc 4494 4530 4567 4604 4644 4684 4724 4764 4805 4846 4886 4928 5006<br />

22 Ti 4970 5008 5047 5086 5125 5167 5209 5251 5292 5335 5378 5420 5464<br />

23 V 5470 5511 5551 5592 5633 5674 5717 5761 5805 5848 5893 5938 5982<br />

24 Cr 5995 6038 6080 6122 6165 6208 6250 6295 6341 6387 6432 6479 6526<br />

25 Mn 6544 6589 6633 6677 6721 6766 6810 6854 6901 6949 6997 7045 7094<br />

26 Fe 7117 7164 7210 7256 7301 7348 7394 7440 7486 7535 7585 7636 7686<br />

27 Co 7715 7763 7811 7859 7906 7954 8002 8050 8098 8146 8198 8250 8303<br />

28 Ni 8338 8387 8436 8486 8535 8585 8635 8685 8735 8785 8835 8889 8943<br />

29 Cu 8986 9035 9086 9137 9188 9240 9294 9344 9396 9448 9500 9552 9609<br />

30 Zn 9663 9708 9760 9813 9866 9919 9973 10028 10082 10136 10190 10244 10298<br />

Z el XIV XV XVI XVII XVIII XIX XX XXI XXII XXIII XXIV XXV XXVI<br />

14 Si 2673<br />

15 P 2817 3070<br />

16 S 3076 3224 3494<br />

17 Cl 3439 3501 3658 3946<br />

18 Ar 3821 3887 3953 4121 4426<br />

19 K 4223 4293 4363 4433 4611 4934<br />

20 Ca 4644 4719 4793 4867 4941 5129 5470<br />

21 Sc 5085 5164 5243 5321 5399 5477 5675 6034<br />

22 Ti 5546 5629 5712 5795 5877 5959 6041 6249 6626<br />

23 V 6028 6114 6201 6288 6375 6461 6547 6633 6851 7246<br />

24 Cr 6572 6621 6711 6802 6893 6984 7074 7164 7254 7482 7895<br />

25 Mn 7143 7191 7242 7336 7431 7526 7621 7715 7809 7903 8141 8572<br />

26 Fe 7737 7788 7838 7891 7989 8088 8187 8286 8384 8482 8580 8828 9278<br />

27 Co 8355 8408 8461 8513 8569 8671 8774 8877 8980 9082 9184 9286 9544<br />

28 Ni 8998 9053 9108 9163 9217 9275 9381 9488 9595 9702 9808 9914 10020<br />

29 Cu 9665 9722 9779 9836 9893 9949 10009 10119 10230 10341 10452 10562 10672<br />

30 Zn 10357 10415 10474 10533 10592 10651 10709 10772 10886 11001 11116 11231 11345<br />

17


3.4 Transitions between different levels<br />

3.4.1 Relative ordering of terms with the same configuration<br />

First we discuss the rules of Hund 1 for terms that correspond with one electron<br />

configuration in LS coupling:<br />

1. Of all terms with the same value of L the term with the highest multiplicity<br />

(i.e., maximum S, spin vectors ↑↑) has the lowest energy. For example triplet<br />

terms have a lower energy than the corresponding singlet terms.<br />

2. Of all terms with the same multiplicity the one with the highest L value (i.e.,<br />

orbital angular momenta ↑↑) has the lowest energy.<br />

3. Of all levels of one term the lowest energy is due to the state with the minimum<br />

value of J if the corresponding outer electron shell is filled with less than half of<br />

the maximum number of electrons, and due to the maximum value of J if the<br />

outer shell is filled with more than half of the maximum number of electrons.<br />

Example 3.7. Calcium in a 3dnp state has terms 3 F, 3 D, 3 P, 1 F, 1 D and 1 P; of<br />

all these terms, the 3 F term has the lowest energy.<br />

Example 3.8. See Fig. 3.9.<br />

Figure 3.9: Illustration of Hund’s rules and Landé’s interval rule.<br />

Furthermore, there is the interval rule of Landé 2 , which is strictly speaking only<br />

valid in LS coupling: the energy interval between two subsequent levels (J and J+1)<br />

in a multiplet is proportional to the highest quantum level (J + 1).<br />

Example 3.9. In a 4 D term (J = 1, 3<br />

2<br />

3:5:7.<br />

, 5<br />

and 7<br />

2 2 2<br />

) the separations have the ratio<br />

1 Named after Friedrich Hund (1896-1997)<br />

2 named after Alfred Landé (1888-1976)<br />

18


3.4.2 Selection rules<br />

The transition or combination between two terms corresponds to a spectral line.<br />

However, not all transitions between levels are possible. We will consider here only<br />

electric dipole <strong>radiation</strong>, which is the most important process. Magnetic dipole<br />

<strong>radiation</strong> is 5 orders of magnitude weaker and electric quadrapole <strong>radiation</strong> is 8<br />

orders of magnitude weaker.<br />

We start first with some general rules that are valid for both LS and JJ coupling.<br />

1. There is no restriction on n, ∆n = integer and can be ∆n = 0, 1, 2, 3, . . ..<br />

2. Rule of Laporte 3 : only transitions between even and odd terms and vice-versa<br />

are allowed: ∆( ∑ l i ) = ±1, ±3, . . .. Thus in principle there can be transitions<br />

with more than one electron involved, although these will be weaker. For<br />

example for the two-electron transition: ∆l 1 = ±1, ∆l 2 = δ (and reversed),<br />

with δ even. In most cases δ = −2, 0, 2 (1-electron jump: ∆l = ±1).<br />

3. ∆J = 0, ±1 and J = 0 cannot be combined with J = 0.<br />

4. ∆M J = 0, ±1 and M J = 0 cannot combine with M J = 0, in case ∆J = 0.<br />

For LS coupling only the additional rules apply:<br />

1. ∆L = 0, ±1<br />

2. ∆S = 0 (intercombinations are forbidden). This is for example strictly valid<br />

for helium (the singlet system does not combine with the triplet system), but<br />

not for Hg (deviation from LS coupling), and also not for higher ions.<br />

Finally for JJ coupling only the following condition applies:<br />

1. for 2 electron jumps: ∆j 1 = 0, ∆j 2 = 0, ±1 (and of course the other way<br />

around).<br />

3.4.3 Examples of combinations between terms.<br />

Because a single term can have more than one level, in general we talk about multiplet<br />

terms or simply multiplets. The combination between two multiplets gives rise<br />

to composite spectral lines; those composites are usually also called multiplets. The<br />

few rules that are applicable to combinations of lines have already been discussed<br />

above.<br />

It is in general also possible to make statements about the relative intensity of<br />

all lines belonging to a multiplet. The famous sum rule of Ornstein, Burger and<br />

Dorgelo 4 states that the sum of the intensities of all lines in a multiplet that belong<br />

to the same lower or upper level is proportional to the statistical weight (g J = 2J+1)<br />

of that level.<br />

When we combine this with the selection rules it follows that<br />

1. The strongest transitions in a multiplet are those for which J and L change<br />

in the same way.<br />

2. Transitions corresponding to large values of J are stronger than those with<br />

smaller values of J.<br />

19


Figure 3.10: Transitions between a 2 D and 2 P term.<br />

Example 3.10. We combine a doublet D term with a doublet P term. In the<br />

spectrum, this gives a triplet (see Fig. 3.10):<br />

a: 2 P 3/2 – 2 D 3/2<br />

b: 2 P 3/2 – 2 D 5/2<br />

c: 2 P 1/2 – 2 D 3/2<br />

Note that the 2 P 1/2 – 2 D 5/2 transition is not allowed (why?). Applying the sum rule<br />

we find:<br />

I b : (I a + I c ) = 6:4<br />

(I a + I b ) : I c = 4:2<br />

and this implies<br />

I a : I b : I c = 1 : 9 : 5 (3.1)<br />

Figure 3.11: Illustration of the relative line intensities for the 2p 2 P − 3d 2 D triplet<br />

of Fexxiv.<br />

Example 3.11. We illustrate this transition with a practical example. Consider<br />

Fexxiv. This is Fe 23+ ; because an iron nucleus has Z = 26 protons, this ion has<br />

26−23 = 3 electrons. Hence, it has a similar electronic structure as lithium (Z = 3).<br />

For most transitions, the core of two 1s electrons can be ignored, and then only the<br />

outer valence electron is relevant. The ground state of this ion is a 2 S 1/2 state (why?),<br />

3 Named after Otto Laporte (1902-1971); paper in JOSA 11, 459 (1925)<br />

4 L.S. Ornstein & H.C. Burger, Z.Phys. 24, 41 (1924)<br />

20


ut of course the ion can also be in a higher state, for instance the valence electron<br />

can be in a 2p orbit or a 3d orbit. Suppose now that it is excited to a 3d orbit.<br />

This orbit has n = 3 and L = 2. The single electron produces S = 1/2. Therefore,<br />

the multiplicity r = 2S + 1 = 2, hence we have a 2 D term. Allowed values of J<br />

are 3/2 and 5/2. If the electron falls back to one of the n = 2 states, than it can<br />

be seen that only a transition to L = 1 is allowed (why?). It is easy to show that<br />

this is the 2p 2 P 1/2 , 2 P 3/2 doublet. Therefore, we will get the radiative transition<br />

triplet 2 P− 2 D as treated above. The three lines can be found easily in a simulated<br />

spectrum (Fig. 3.11). The lines are found at 11.029, 11.171 and 11.188 Å. The<br />

spectral region shows several other lines from other (mainly iron) ions, so knowing<br />

the predicted line ratios is helpfull in identifying these triplets. Verify the relative<br />

intensity of these lines.<br />

A more complex example is given by the following.<br />

Example 3.12. Transition from 3 D to 3 P. In the spectrum, this gives a sextet:<br />

a: 3 D 1 – 3 P 2<br />

b: 3 D 2 – 3 P 2<br />

c: 3 D 3 – 3 P 2<br />

d: 3 D 1 – 3 P 1<br />

e: 3 D 2 – 3 P 1<br />

f: 3 D 1 – 3 P 1<br />

Applying the sum rule we find:<br />

I c : (I b + I e ) : (I a + I d + I f ) = 7:5:3<br />

(I a + I b + I c ) : (I d + I e ) : I f = 5:3:1<br />

and this implies<br />

I a : I b : I c : I d : I e : I f = 1.2 : 17.9 : 100 : 17.9 : 53.6 : 23.8 (3.2)<br />

exercise 3.2. For the above example, make a plot similar to Fig. 3.10. Use Hund’s<br />

law to plot a realistic relative distance between the terms of the upper and lower<br />

multiplet.<br />

The spectroscopic notation is based on the LS-coupling, because that applies<br />

in the most simple cases. LS-coupling dominates whenever the distance between<br />

two multiplets (determined by the electrostatic interaction of the electrons) is much<br />

larger than the multiplet-splitting (= largest distance within one multiplet, caused<br />

by the spin-orbit interaction), which is the case for a low nuclear charge Z (the<br />

light elements). Because the spin-orbit interaction increases faster with Z than<br />

the electrostatic interaction, for the heavier elements in the end JJ-coupling will<br />

dominate, in particular for complicated term configurations, where there are many<br />

terms close together. In an excited state (with large principle quantum number<br />

n), where the electrostatic interaction decreases rapidly, already for lower values of<br />

Z JJ-coupling can occur. For example, for Pb, Sn and Ge for the ground term<br />

LS-coupling applies, while the higher excited states have JJ-coupling.<br />

21


3.5 Abundances of the elements<br />

With high spectral resolution and sensitivity, optical spectra of stars sometimes show<br />

spectral features from almost all elements of the Periodic Table, but in practice only<br />

a few of the most abundant elements show up in X-<strong>ray</strong> spectra of cosmic plasmas.<br />

In several situations the abundances of the chemical elements in an X-<strong>ray</strong> source are<br />

similar to (but not necessarily equal to) the abundances for the Sun or a constant<br />

fraction of that. There have been many significant changes over the last several<br />

years in the adopted set of standard cosmic abundances. A still often used set of<br />

abundances is that of Anders & Grevesse (1989), but a more recent one is the set of<br />

proto-Solar abundances of Lodders (2003), that we list in Table 3.4 for a few of the<br />

key elements.<br />

Table 3.4: Proto-Solar abundances for the 15 most common chemical elements.<br />

Abundances A are given with respect to hydrogen. Data from Lodders (2003).<br />

Element abundance Element abundance Element abundance<br />

H ≡ 1 Ne 89.1 × 10 −6 S 18.2 × 10 −6<br />

He 0.0954 Na 2.34 × 10 −6 Ar 4.17 × 10 −6<br />

C 288 × 10 −6 Mg 41.7 × 10 −6 Ca 2.57 × 10 −6<br />

N 79.4 × 10 −6 Al 3.47 × 10 −6 Fe 34.7 × 10 −6<br />

O 575 × 10 −6 Si 40.7 × 10 −6 Ni 1.95 × 10 −6<br />

In general, for absorption studies the strength of the lines is mainly determined<br />

by atomic parameters that do not vary much along an iso-electronic sequence, and<br />

the abundance of the element. Therefore, in the X-<strong>ray</strong> band the oxygen lines are the<br />

strongest absorption lines, as oxygen is the most abundant metal. The emissivity<br />

of ions in the X-<strong>ray</strong> band often increases with a strong power of the nuclear charge.<br />

For that reason, in many X-<strong>ray</strong> plasmas the strongest iron emission lines are often<br />

of similar strength to the strongest oxygen lines, despite the fact that the cosmic<br />

abundance of iron is only 6 % of the cosmic oxygen abundance.<br />

Without being complete or giving too many details, we give here a short list of<br />

situations where the abundances differ from proto-Solar, and where X-<strong>ray</strong> spectroscopic<br />

observations are key to determine these abundances:<br />

• Stellar coronae: here due to the (inverse) FIP effect (FIP = First Ionisation<br />

potential), abundances can be different from Proto-Solar. Elements with high<br />

FIP, for instance noble gases line Ne, remain longer neutral in the transition<br />

zones between photosphere and corona. Neutral particles are not tied to the<br />

magnetic field in stellar flares, hence segregration between neutral and ionised<br />

species is possible, leading to deviant abundances.<br />

• Some hot stars show strong effects of the CNO cycle, leading e.g. to enhnced<br />

nitrogen abundances in some O-stars like ζ Pup.<br />

• In several young supernova remnants (e.g. Cas A) one sees directly the freshly<br />

produced chemical elements lighten-up in X-<strong>ray</strong>s<br />

22


• In older supernova remnants, emission is dominated by swept-up Interstellar<br />

matter. This offers a unique opportunity to sample the ISM of for instance<br />

metal-poor galaxies like the LMC and SMC, by observing their old supernova<br />

remnants.<br />

• In clusters of galaxies, the hot gas is a fair sample of the chemical composition<br />

of the Universe as a whole, which differs from the Solar composition.<br />

• One of the poorest known environments is the cosmic web between clusters<br />

of galaxies. depending upon the galacitic evolution scenario of the galaxies<br />

in that web, one predicts models with low metallicity (say 0.1 time Solar), or<br />

models with strongly varying abundances, ranging from 0.001 to 1 times Solar.<br />

• Active Galactic Nuclei sample the composition of galactic cores, which due<br />

to higher stellar birth rates in those environments is higher than in the Solar<br />

neighbourhood.<br />

23


3.6 Basic processes<br />

In order to be able to calculate X-<strong>ray</strong> spectra, it is necessary to determine first the<br />

concentrations of the various ions. There are several ionisation and recombination<br />

processes that are important for determining these concentrations. Some of these<br />

processes also play directly a role in spectral line formation processes. We will<br />

first introduce the basic processes in some detail, then we will discuss how the ion<br />

concentrations can be determined, and then we return to the continuum and line<br />

formation processes. The relevant basic processes that we discuss here are:<br />

1. excitation processes:<br />

(a) collisional excitation<br />

(b) collisional de-excitation<br />

(c) radiative excitation<br />

2. Radiative transitions<br />

3. Auger processes and fluorescence<br />

4. ionisation processes:<br />

(a) collisional (direct) ionisation<br />

(b) Excitation-Autoionisation<br />

(c) Photoionisation<br />

(d) Compton ionisation<br />

5. recombination processes:<br />

(a) Radiative recombination<br />

(b) Dielectronic recombination<br />

6. charge transfer processes:<br />

(a) Charge Transfer ionisation<br />

(b) Charge Transfer recombination<br />

All these processes will be treated below.<br />

24


3.6.1 Collisional excitation<br />

A bound electron in an ion can be brought into a higher, excited energy level through<br />

a collision with a free electron or by absorption of a photon. The latter will be<br />

discussed in more detail in Sect. 3.10. Here we focus upon excitation by electrons.<br />

The cross section Q ij for excitation from level i to level j for this process can be<br />

conveniently parametrised by<br />

Q ij (U) = πa2 0<br />

w i<br />

E H Ω(U)<br />

E ij U , (3.3)<br />

where U = E/E ij with E ij the excitation energy from level i to j, E the energy of<br />

the exciting electron, E H the Rydberg energy (13.6 eV), a 0 the Bohr radius and w i<br />

the statistical weight of the lower level i. The dimensionless quantity Ω(U) is the socalled<br />

collision strength. For a given transition on an iso-electronic sequence, Ω(U)<br />

is not a strong function of the atomic number Z, or may be even almost independent<br />

of Z.<br />

Mewe (1972) introduced a convenient formula that can be used to describe most<br />

collision strengths, written here as follows:<br />

Ω(U) = A + B U + C U + 2D + F lnU, (3.4)<br />

2 U3 where A, B, C, D and F are parameters that differ for each transition. The expression<br />

can be integrated analytically over a Maxwellian electron distribution, and the<br />

result can be expressed in terms of exponential integrals.<br />

The Maxwell velocity distribution is given by<br />

f(v) = ( 2 v 2<br />

π )1/2 e −v2 /2ve, 2 (3.5)<br />

where v 2 e ≡ kT/m e and m e is the electron mass. Transforming this to energy units<br />

using E = 1 2 m ev 2 gives<br />

f(E) =<br />

v 3 e<br />

2E 0.5<br />

√ π(kT)<br />

3/2 e−E/kT , (3.6)<br />

exercise 3.3. Verify that<br />

∞∫<br />

0<br />

f(v)dv = 1. Also verify that<br />

∞∫<br />

0<br />

f(E)dE = 1.<br />

After averaging over a Maxwellian electron distribution the number of excitations<br />

can be determined as:<br />

∞∫<br />

√<br />

8πkT a 2 0<br />

S ij = Q ij (E)f(E)v(E)dE =<br />

E ∫∞<br />

H<br />

y 2 Ω(U)e −Uy dU, (3.7)<br />

m e w i E ij<br />

E 1<br />

ij<br />

where y ≡ E ij /kT.<br />

We now define the average collision strength ¯Ω(y) as:<br />

∫∞<br />

¯Ω(y) ≡ ye y<br />

1<br />

25<br />

Ω(U)e −Uy dU (3.8)


leading to<br />

Here the constant S 0 is given by<br />

S ij = S 0 ¯Ω(y)T −1/2 e −y , (3.9)<br />

S 0 ≡ √ 8π/m e k a 2 0 E H = h 2 (2πm e ) −3/2 k −1/2 ≃ 8.629 × 10 −12 K 0.5 m 3 s −1 (3.10)<br />

(using E H = 1 2 α2 m e c 2 and a 0 = h/2παm e c).<br />

For the special choice (3.4) it follows that<br />

¯Ω(y) = A + (By − Cy 2 + Dy 3 + F)e y E 1 (y) + (C + D)y − Dy 2 . (3.11)<br />

In the above, E 1 (x) is the exponential integral, defined by (see for example<br />

Abramowitz & Stegun, chapter 5):<br />

E 1 (x) ≡<br />

∫ ∞<br />

e −t<br />

x<br />

t<br />

dt. (3.12)<br />

It is possible to show that the asymptotic behaviour of E 1 (x) is given by:<br />

x ≪ 1 :<br />

x ≫ 1 :<br />

E 1 (x) = −γ − ln x −<br />

E 1 (x) = e−x<br />

x<br />

where γ = 0.5772156649 is the constant of Euler.<br />

With (3.13) and (3.14) we find (check!):<br />

∞∑<br />

n=0<br />

∞∑<br />

n=1<br />

(−x) n<br />

nn!<br />

(3.13)<br />

n!<br />

(−x) n (3.14)<br />

kT ≪ E ij , y ≫ 1 : S ij ≃ S 0 (A + B + C + 2D)T −1 2 e<br />

−E ij /kT<br />

(3.15)<br />

kT ≫ E ij , y ≪ 1 : S ij ≃ S 0 FT −1 2 ln(kT/Eij ) (3.16)<br />

For low T there are exponentially few excitations (only the tail of the Maxwellian has<br />

sufficient energy for excitation); for high T the number of excitations also approaches<br />

zero, because of the small collision cross section.<br />

exercise 3.4. Derive (3.15) and (3.16).<br />

Not all transitions have this asymptotic behaviour, however. For instance, socalled<br />

forbidden transitions have F = 0 and hence have much lower excitation rates<br />

at high energy. So-called spin-forbidden transitions even have A = B = F = 0.<br />

exercise 3.5. Derive the high-temperature limit equivalent of (3.16) for forbidden<br />

and spin-forbidden transitions.<br />

In most cases, the excited state is stable and the ion will decay back to the<br />

ground level by means of a radiative transition, either directly or through one or<br />

more steps via intermediate energy levels. Only in cases of high density or high<br />

<strong>radiation</strong> fields, collisional excitation or further radiative excitation to a higher level<br />

may become important, but for typical cluster and ISM conditions these processes<br />

are not important in most cases. Therefore, the excitation rate immediately gives<br />

the total emission line power.<br />

26


3.6.2 Collisional de-excitation<br />

A collision between an ion and a free electron can either bring an electron into a<br />

higher orbit or into a lower orbit. When a bound electron ends up in a lower energy<br />

orbit, the process is called collisional de-excitation. Of course, de-excitation can<br />

only occur when the ion is already in an excited state. Therefore, de-excitation is<br />

in practice important only for high density plasmas, where due to collisions higher<br />

levels are populated, or for so-called metastable levels (levels that have only a small<br />

probability to decay radiatively to a lower level).<br />

The rate coefficient S ′ ji of de-excitation j → i is related to S ij, the rate coefficient<br />

of the inverse excitation process i → j. The relation can be derived from the<br />

principles of detailed balance, and is given by:<br />

S ′ ji = w i<br />

w j<br />

e E ij/kT Sij . (3.17)<br />

27


3.6.3 Radiative transition probabilities<br />

It is well known from quantum mechanics that the occurence of radiative transitions<br />

from one upper level j to a lower level i cannot be predicted in advance. However,<br />

the transition probability can be calculated, and it is usualy denoted by A ji . It has<br />

units of s −1 . The transition probability is related to the dimensionless absorption<br />

oscillator strength f ij that we encounter later (see Sect. 3.10) as<br />

A ji = 4παw if ij E 2 ij<br />

w j m e c 2 h = 4.33 × 1013 s −1 w i<br />

w j<br />

f ij E 2 ij,keV, (3.18)<br />

where E ij is the energy difference between level i and j (the energy of the emitted<br />

line).<br />

28


• Auger channels<br />

[1s]µ K−MM(p)<br />

−→<br />

K−MM(s)<br />

−→<br />

K−LM(p)<br />

−→<br />

K−LM(s)<br />

−→<br />

K−LL<br />

−→<br />

⎧<br />

⎨<br />

⎩<br />

⎧<br />

⎨<br />

µ −2 + e −<br />

[3p]µ −1 + e −<br />

[3s] µ −1 + e −<br />

⎫<br />

⎬<br />

⎭<br />

[3p] 2 µ + e −<br />

[3s] [3p]µ + e −<br />

⎩<br />

[3s] 2 µ + e −<br />

{ }<br />

[2s] µ −1 + e −<br />

[2p] µ −1 + e −<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

⎧<br />

⎨<br />

⎩<br />

⎫<br />

⎬<br />

⎭<br />

(3.22)<br />

(3.23)<br />

(3.24)<br />

[2s] [3p]µ + e<br />

⎫⎪ −<br />

⎬<br />

[2p][3p] µ + e −<br />

[2s] [3s] µ + e −<br />

(3.25)<br />

⎪ ⎭<br />

[2p][3s]µ + e −<br />

[2s] 2 µ + e −<br />

[2s] [2p]µ + e −<br />

[2p] 2 µ + e −<br />

⎫<br />

⎬<br />

⎭<br />

(3.26)<br />

where the negative exponent of µ stands for the number of electrons that have<br />

been extracted from the outer subshells. In the radiative channels, forbidden and<br />

two-electron transitions have been excluded as it has been confirmed by calculation<br />

that they display very small transition probabilities (log A r < 11). Therefore, the<br />

two main photo-decay pathways are the result of the 2p → 1s and 3p → 1s single<br />

electron jumps that give rise respectively to the Kα (∼ λ1.94) and Kβ (∼ λ1.75)<br />

ar<strong>ray</strong>s.<br />

Let us focus now upon a useful example. If for example neutral Fe (Fei) looses a<br />

K-shell electron by photoionisation, there is a hole in the K-shell. This state should<br />

formally be considered of course as a (core-excited) state of Feii, since now there are<br />

not 26 but 25 electrons left. This hole will be filled. For Fe, in approximately 30 %<br />

of all cases the hole is filled by a radiative transition from the L 2 or L 3 shell (2p),<br />

and in the other 70 % by an Auger transition, usually a K–L 2,3 L 2,3 transition (L 2,3<br />

means here either L 2 or L 3 , i.e. 2p). Therefore, in 30 % of all cases a single vacancy<br />

in the L 2,3 shells is produced, and in 70 % of all cases there are two vacancies in<br />

the L 2,3 shell. These L 2,3 vacancies are in turn filled in most cases by a L 2,3 –MM<br />

transition, etc. In this way one single photoionisation can release several electrons.<br />

For neutral iron one finds that a single K-shell ionisation leads to on average 5.7<br />

free electrons (including the first photo-electron). One L 1 -shell ionisation even leads<br />

to 5.8 electrons, an L 2,3 -shell ionisation to 3.3 electrons, a M 1 -shell ionisation to 2.9<br />

electrons and a M 2,3 -shell ionisation to 2 electrons.<br />

The fluorescence yield ω is defined as the probability of a radiative transition<br />

(fluorescence) to fill a vacancy in an inner shell of an atom or ion. It depends upon<br />

the nuclear charge (see Fig. 3.12).<br />

In general the fluorescence yield is increasing with nuclear charge, and largest for<br />

the innermost shells. For the most abundant elements ω L is small (for Fe < 0.5 %).<br />

For the labelling of fluorescent lines there is a distinct naming convention, invented<br />

by Manne Siegbahn 5 in the 1920’s. See Table 3.5. The notation is based<br />

5 Nobel Prize in Physics in 1924 for his discoveries and research in the field of X-<strong>ray</strong> spectroscopy<br />

30


Figure 3.12: Fluorescence yield as a<br />

function of atomic number Z.<br />

Figure 3.13: Probability of fluorescence,<br />

Coster-Kronig and Auger transitions<br />

for a vacancy in the L 1 subshell.<br />

Table 3.5: Siegbahn notation for inner shell transitions. The table gives the recommended<br />

notation by the IUPAC.<br />

Siegbahn IUPAC Siegbahn IUPAC Siegbahn IUPAC Siegbahn IUPAC<br />

Kα 1 K–L 3 Lα 1 L 3 –M 5 Lγ 1 L 2 –N 4 Mα 1 M 5 –N 7<br />

Kα 2 K–L 2 Lα 2 L 3 –M 4 Lγ 2 L 1 –N 2 Mα 2 M 5 –N 6<br />

Kβ 1 K–M 3 Lβ 1 L 2 –M 4 Lγ 3 L 1 –N 3 Mβ M 4 –N 6<br />

K I β 2 K–N 3 Lβ 2 L 3 –N 5 Lγ 4 L 1 –O 3 Mγ M 3 –N 5<br />

K II β 2 K–N 2 Lβ 3 L 1 –M 3 Lγ 4 ′ L 1 –O 2 Mζ<br />

Kβ 3 K–M 2 Lβ 4 L 1 –M 2 Lγ 5 L 2 –N 1<br />

M 4,5 –N 2,3<br />

K I β 4 K–N 5 Lβ 5 L 3 –O 4,5 Lγ 6 L 2 –O 4<br />

K II β 4 K–N 4 Lβ 6 L 3 –N 1 Lγ 8 L 2 –O 1<br />

Kβ 4x K–N 4 Lβ 7 L 3 –O 1 Lγ 8 ′ L 2 –N 6(7)<br />

K I β 5 K–M 5 Lβ 7 ′ L 3 –N 6,7 Lη L 2 –M 1<br />

K II β 5 K–M 4 Lβ 9 L 1 –M 5 Ll L 3 –M 1<br />

Lβ 10 L 1 –M 4 Ls L 3 –M 3<br />

Lβ 15 L 3 –N 4 Lt L 3 –M 2<br />

Lβ 17 L 2 –M 3 Lu L 3 –N 6,7<br />

Lv L 2 –N 6(7)<br />

upon the relative intensity of lines from different series. The system is still widely<br />

used but incomplete in particular for the M and N series. For this reason the IUPAC<br />

(International Union of Pure and Applied Chemistry) recommends another, more<br />

logical notation (see Table 3.5).<br />

The intensity ratio between Kα 1 and Kα 2 is 2:1. exercise 3.6. Argue why the<br />

above ratio is 2:1. In general Kβ is weaker than Kα. For neutral iron for example<br />

Kβ has about 12 % of the intensity of Kα.<br />

31


3.6.5 Collisional ionisation<br />

Collisional ionisation occurs when during the interaction of a free electron with an<br />

atom or ion the free electron transfers a part of its energy to one of the bound<br />

electrons, which is then able to escape from the ion. A necessary condition is that<br />

the kinetic energy E of the free electron must be larger than the binding energy I<br />

of the atomic shell from which the bound electron escapes.<br />

Q (m 2 )<br />

0 2×10 −26 4×10 −26 6×10 −26<br />

Fe I<br />

Fe XXV<br />

Fe XXVI<br />

10 100 1000<br />

Energy (keV)<br />

Figure 3.14: Illustration of the Lotz formula for three iron ions. Shown is the K-shell<br />

ionisation cross section. Note that Fei and Fexxv have two electrons in this shell,<br />

while Fexxvi only has one (therefore its cross section is half that of Fexxv).<br />

The Lotz formula<br />

There are several formulae that describe the effective cross section Q as a function<br />

of E. A formula that gives a correct order of magnitude estimate of the cross section<br />

σ of this process and that has the proper asymptotic behaviour (first calculated by<br />

Bethe and Born) is the formula of Lotz (1968):<br />

where<br />

and further<br />

uI 2 Q = C ln u, (3.27)<br />

u ≡ E/I (3.28)<br />

C = an s , (3.29)<br />

where n s is the number of electrons in the shell and the normalisation a = 4.5 ×<br />

10 −24 m 2 keV 2 . See Fig. 3.14 for an illustration of the shape and scaling of this<br />

formula.<br />

exercise 3.7. Show that for E → I, Q → 0 and in particular Q ∼ E−I. Determine<br />

the constant of proportionality.<br />

It is further evident that for E → ∞ the cross section Q ∼ (ln E)/E → 0. This<br />

shows that energetic electrons are not very efficient in ionising an atom or ion. They<br />

pass too fast.<br />

32


The formula of Younger<br />

While the Lotz formula in general gives the right order of magnitude for the cross<br />

section, it is not the most accurate and flexible formula for representation of the<br />

direct ionisation cross section Q. A better formula was obtained by Younger (1981):<br />

uI 2 Q = A(1 − 1 u ) + B(1 − 1 u )2 + C ln u + D ln u<br />

u . (3.30)<br />

The constants A, B, C and D differ per ion and shell. Note that the Lotz formula<br />

is a special case of this equation.<br />

Figure 3.15: Example of a fit to direct ionisation cross sections. From Arnaud &<br />

Rothenflug (1985).<br />

In general it is not always simple to calculate Q theoretically, and sometimes<br />

discrepancies between theory and observations remain. However, experimental data<br />

are also not always available, or sometimes the experiments suffer from ”contamination”,<br />

for instance because of contamination from so-called metastable levels. In<br />

general an uncertainty of 5 −10 % is not uncommon in the cross sections. The same<br />

holds by the way for many of the other processes that we describe elsewhere in this<br />

chapter. See Fig. 3.15 for an example of a fit to direct ionisation cross sections.<br />

Integration over an electron distribution<br />

You may have wondered why we have expressed the left hand side of eqns. (3.27)<br />

and (3.30) in the strange form uI 2 Q. The reason is that we need to average the<br />

effective cross section over a Maxwellian electron distribution:<br />

f(v) = n e ( 2 v 2<br />

π )1/2 e −v2 /2ve, 2 (3.31)<br />

where v 2 e ≡ kT/m e and m e is the electron mass. The total number of direct ionisations<br />

per unit volume per unit time is then given by:<br />

∫∞<br />

C DI = n i<br />

0<br />

v 3 e<br />

vf(v)Q(v)dv. (3.32)<br />

33


In this equation n i is the ion density of the ion that is being considered. It is easy<br />

to show (see exercise below) that this implies<br />

C DI =<br />

2√ ∫∞<br />

2n e n i<br />

√ (uI 2 Q)e −uy du, (3.33)<br />

πme (kT) 3<br />

1<br />

where<br />

y ≡ I/kT. (3.34)<br />

exercise 3.8. Show that by substituting E = 1 2 m ev 2 and u = E/I you arrive at<br />

eqn. (3.33).<br />

We see that the direct ionisation rate follows immediately from the Laplace<br />

transform of uI 2 Q.<br />

It can be shown that by substitution of the Lotz formula (3.27) into (3.33) the<br />

integral can be expressed as:<br />

C DI = 2√ 2n e n i C E<br />

√ 1 (y)<br />

, (3.35)<br />

πme (kT) 3 y<br />

where E 1 (x) is the exponential integral (see 3.12).<br />

We apply the asymptotic approximations (3.13) and (3.14) now to eqn. (3.35)<br />

and we see that<br />

kT ≪ I :<br />

C DI ≃ ( 2√ 2C<br />

√ πme<br />

) n en i<br />

√<br />

kTe<br />

−I/kT<br />

I 2 (3.36)<br />

and<br />

kT ≫ I :<br />

C DI ≃ ( 2√ 2C<br />

√ πme<br />

) n en i ln(kT/I)<br />

I √ kT<br />

(3.37)<br />

For low temperatures, the ionisation rate therefore goes exponentially to zero. This<br />

can be understood simply, because for low temperatures only the electrons from the<br />

exponential tail of the Maxwell distribution have sufficient energy to ionise the atom<br />

or ion. For higher temperatures the ionisation rate also approaches zero, but this<br />

time because the cross section at high energies is small.<br />

For each ion the direct ionisation rate per atomic shell can now be determined.<br />

The total rate follows immediately by adding the contributions from the different<br />

shells. Usually only the outermost two or three shells are important. That is because<br />

of the scaling with I −2 and I −1 in (3.36) and (3.37), respectively.<br />

34


3.6.6 Excitation-Autoionisation<br />

In the previous section we showed how a collision of a free electron with an ion can<br />

lead to ionisation. In principle, if the free electron has insufficient energy (E < I),<br />

there will be no ionisation. However, even in that case it is sometimes possible to<br />

ionise, but in a more complicated way. The process is as follows. The collision<br />

can bring one of the electrons in the outermost shells in a higher quantum level<br />

(excitation). In most cases, the ion will return to its ground level by a radiative<br />

transition. But in some cases the excited state is unstable, and a <strong>radiation</strong>less Auger<br />

transition can occur. Later on we will treat the Auger process in more detail, but<br />

shortly speaking the vacancy that is left behind by the excited electron is being filled<br />

by another electron from one of the outermost shells, while the excited electron is<br />

able to leave the ion (or a similar, process with the role of both electrons reversed).<br />

Because of energy conservation, this process only occurs if the excited electron<br />

comes from one of the inner shells (the ionisation potential barrier must be taken<br />

anyhow). The process is in particular important for Li-like and Na-like ions, and<br />

for several other individual atoms and ions.<br />

Figure 3.16: The Excitation-Autoionisation process for a Li-like ion.<br />

As an example we treat here Li-like ions (see Fig. 3.16). In that case the most<br />

important contribution comes from a 1s–2p excitation. The effective cross section<br />

Q for excitation from 1s to 2p followed by auto-ionisation can be approximated by<br />

(E H = 13.6 eV):<br />

Q = πa 2 0 2B<br />

Z2<br />

Z 2 eff (1s) E H<br />

E Ω H (3.38)<br />

in which a 0 is the Bohr radius, and Z eff = Z − 0.43 is the effective charge that<br />

a 1s electron experiences because of the presence of the other 1s electron and the<br />

2s electron in the Li-like ion. Just as before, E is the energy of the free electron<br />

that causes the excitation. Further, B is the so-called branching ratio. It gives<br />

the probability that after the excitation there will be an auto-ionisation (the other<br />

possibility was a radiative transition). The following approximation for B can be<br />

used:<br />

1<br />

B ≃<br />

(3.39)<br />

1 + (Z/17) 3<br />

Finally, the scaled hydrogenic collision strength is given by:<br />

Z 2 Ω H ≃ 2.22 lnǫ + 0.67 − 0.18<br />

ǫ<br />

35<br />

+ 1.20<br />

ǫ 2 (3.40)


where ǫ ≡ E/E ex and the excitation energy E ex ≡ I 1s − I 2p . Similar to the case of<br />

direct ionisation, it is straightforward to average the excitation-autoionisation cross<br />

section over a Maxwellian electron distribution.<br />

Figure 3.17: Collisional ionisation<br />

cross section for Ovi.<br />

Figure 3.18: Collisional ionisation<br />

cross section for Fexvi.<br />

In Figs. 3.17 and 3.18 we show the total collisional ionisation cross section (direct<br />

+ excitation-autoionisation) for an ion in the Li electronic sequence and the Na<br />

isoelectronic sequence. Clearly, for Fexvi the excitation autoionisation process is<br />

very important. The increasing importance of excitation autoionisation relative to<br />

direct ionisation is also illustrated in Fig. 3.19.<br />

Figure 3.19: Ratio between the Excitation-Autoionisation cross section to the Direct<br />

Ionisation cross section for Na-like ions. From Arnaud & Rothenflug (1985).<br />

36


3.6.7 Photoionisation<br />

Figure 3.20: Photoionisation cross section in barn (10 −28 m −2 ) for Fei (left) and<br />

Fexvi (right). The p and d states (dashed and dotted) have two lines each because<br />

of splitting into two sublevels with different j.<br />

This process is very similar to direct ionisation. The difference is that in the case<br />

of photoionisation a photon instead of an electron is causing the ionisation. Further,<br />

the effective cross section is in general different from that for direct ionisation. As<br />

an example Fig. 3.20 shows the cross section for neutral iron and Na-like iron. It<br />

is evident that for the more highly ionised iron the so-called ”edges” (ionisation<br />

potentials) are found at higher energies than for neutral iron. Contrary to the case<br />

of direct ionisation, the cross section at threshold for photoionisation is not zero. The<br />

effective cross section just above the edges sometimes changes rapidly (the so-called<br />

Cooper minima and maxima).<br />

Contrary to the direct ionisation, all the inner shells have now the largest cross<br />

section. For the K-shell one can approximate for E > I<br />

σ(E) ≃ σ 0 (I/E) 3 (3.41)<br />

where as before I is the edge energy. For a given ionising spectrum F(E) (photons<br />

per unit volume per unit energy) the total number of photoionsations follows as<br />

C PI = c<br />

For hydrogenlike ions one can write:<br />

∫ ∞<br />

0<br />

n i σ(E)F(E)dE. (3.42)<br />

σ PI = 64πng(E, n)αa2 0<br />

3 √ 3Z 2 ( I E )3 , (3.43)<br />

where n is the principal quantum number, α the fine structure constant and A 0 the<br />

Bohr radius. The gaunt factor g(E, n) is of order unity and varies only slowly. It<br />

has been calculated and tabulated by Karzas & Latter (1961). The above equation<br />

is also applicable to excited states of the atom, and is a good approximation for all<br />

excited atoms or ions where n is larger than the corresponding value for the valence<br />

electron.<br />

37


3.6.8 Compton ionisation<br />

Scattering of a photon on an electron generally leads to energy transfer from one<br />

of the particles to the other. In most cases only scattering on free electrons is<br />

considered. But Compton scattering also can occur on bound electrons. If the energy<br />

transfer from the photon to the electron is large enough, the ionisation potential can<br />

be overcome leading to ionisation. This is the Compton ionisation process.<br />

Figure 3.21: Compton ionisation cross section for neutral atoms of H, He, C, N, O,<br />

Ne and Fe.<br />

In the Thomson limit the differential cross section for Compton scattering is<br />

given by<br />

dσ<br />

dΩ = 3σ T<br />

16π (1 + cos2 θ), (3.44)<br />

with θ the scattering angle and σ T the Thomson cross section (6.65 × 10 −29 m −2 ).<br />

The energy transfer ∆E during the scattering is given by (E is the photon energy):<br />

∆E =<br />

E 2 (1 − cosθ)<br />

m e c 2 + E(1 − cosθ) . (3.45)<br />

Only those scatterings where ∆E > I contribute to the ionisation. This defines a<br />

critical angle θ c , given by:<br />

cosθ c = 1 −<br />

m ec 2 I<br />

E 2 − IE . (3.46)<br />

For E ≫ I we have σ(E) → σ T (all scatterings lead in that case to ionisation) and<br />

further for θ c → π we have σ(E) → 0. Because for most ions I ≪ m e c 2 , this last<br />

condition occurs for E ≃ √ Im e c 2 /2 ≫ I. See Fig. 3.21 for an example of some<br />

cross sections. In general, Compton ionisation is important if the ionising spectrum<br />

contains a significant hard X-<strong>ray</strong> contribution, for which the Compton cross section<br />

is larger than the steeply falling photoionisation cross section.<br />

One sees immediately that for E ≫ I we have σ(E) → σ T (all scatterings lead<br />

in that case to ionisation) and further that for θ c → π we have σ(E) → 0. Because<br />

for most ions I ≪ m e c 2 , this last condition occurs for E ≃ √ Im e c 2 /2 ≫ I. See<br />

Fig. 3.21 for an example of some cross sections.<br />

exercise 3.9. Estimate for which energy Compton ionisation becomes more important<br />

than photoionisation for neutral iron.<br />

38


3.6.9 Radiative recombination<br />

Radiative recombination is the reverse process of photoionisation. A free electron is<br />

captured by an ion while emitting a photon. The released <strong>radiation</strong> is the so-called<br />

free-bound continuum emission. It is relatively easy to show that there is a simple<br />

relation between the photoionisation cross section σ bf (E) and the recombination<br />

cross section σ fb , namely the Milne-relation:<br />

σ fb (v) = E2 g n σ bf (E)<br />

m e c 2 m e v 2 (3.47)<br />

where g n is the statistical weight of the quantum level into which the electron is captured<br />

(for an empty shell this is g n = 2n 2 ). By averaging over a Maxwell distribution<br />

one gets the recombination-coefficient to level n:<br />

∫∞<br />

R n = n i n e<br />

Of course there is energy conservation, so E = 1 2 m ev 2 + I.<br />

0<br />

vf(v)σ fb (v)dv. (3.48)<br />

exercise 3.10. Demonstrate that for the photoionisation cross section (3.43) and<br />

for g = 1, constant and g n = 2n 2 :<br />

R n = 128√ 2πn 3 αa 2 0I 3 e I/kT n i n e<br />

3 √ 3m e kTZ 2 kTm e c 3 E 1 (I/kT) (3.49)<br />

With the asymptotic relations (3.13) and (3.14) it is seen that<br />

kT ≪ I : R n ∼ T −1 2 (3.50)<br />

kT ≫ I : R n ∼ ln(I/kT)T −3/2 (3.51)<br />

Therefore for T → 0 the recombination coefficient approaches infinity: a cool plasma<br />

is hard to ionise. For T → ∞ the recombination coefficient goes to zero, because of<br />

the Milne relation (v → ∞) and because of the sharp decrease of the photoionisation<br />

cross section for high energies.<br />

As a rough approximation we can use further that I ∼ (Z/n) 2 . Substituting<br />

this we find that for kT ≪ I (recombining plasma’s) R n ∼ n −1 , while for kT ≫ I<br />

(ionising plasma’s) R n ∼ n −3 . In recombining plasma’s therefore in particular many<br />

higher excited levels will be populated by the recombination, leading to significantly<br />

stronger line emission (see Sect. 3.9). On the other hand, in ionising plasmas (such<br />

as supernova remnants) recombination mainly occurs to the lowest levels. Note<br />

that for recombination to the ground level the approximation (3.49) cannot be used<br />

(the hydrogen limit), but instead one should use the exact photoionisation cross<br />

section of the valence electron. By adding over all values of n and applying an<br />

approximation Seaton (1959) found for the total radiative recombination rate α RR<br />

(in units of m −3 s −1 ):<br />

α RR ≡ ∑ n<br />

R n = 5.197 × 10 −20 Zλ 1/2 {0.4288 + 0.5 lnλ + 0.469λ −1/3 } (3.52)<br />

39


with λ ≡ E H Z 2 /kT and E H the Rydberg energy (13.6 eV). Note that this equation<br />

only holds for hydrogen-like ions. For other ions usually an analytical fit to numerical<br />

calculations is used:<br />

α RR ∼ T −η (3.53)<br />

where the approximation is strictly speaking only valid for T near the equilibrium<br />

concentration. The approximations (3.50) and (3.51) raise suspicion that for T → 0<br />

or T → ∞ (3.53) could be a poor choice.<br />

40


3.6.10 Dielectronic recombination<br />

This process is more or less the inverse of excitation-autoionisation. Now a free<br />

electron interacts with an ion, by which it is caught (quantum level n ′′ l ′′ ) but at the<br />

same time it excites an electron from (nl) → (n ′ l ′ ). The doubly excited state is in<br />

general not stable, and the ion will return to its original state by auto-ionisation.<br />

However there is also a possibility that one of the excited electrons (usually the<br />

electron that originally belonged to the ion) falls back by a radiative transition to<br />

the ground level, creating therefore a stable, albeit excited state (n ′′ l ′′ ) of the ion.<br />

In particular excitations with l ′ = l+1 contribute much to this process. In order to<br />

calculate this process, one should take account of many combinations (n ′ l ′ )(n ′′ l ′′ ).<br />

The final transition probability is often approximated by<br />

α DR = A<br />

T 3/2e−T 0/T (1 + Be −T 1/T ) (3.54)<br />

where A, B, T 0 and T 1 are adjustable parameters. Note that for T → ∞ the<br />

asymptotic behaviour is identical to the case of radiative recombination. For T → 0<br />

however, dielectronic recombination can be neglected; this is because the free electron<br />

has insufficient energy to excite a bound electron. Dielectronic recombination<br />

is a dominant process in the Solar corona, and also in other situations it is often<br />

very important.<br />

41


3.6.11 Charge transfer ionisation and recombination<br />

In most cases ionisation or recombination in collisionally ionised plasmas is caused<br />

by interactions of an ion with a free electron. At low temperatures (typically below<br />

10 5 K) also charge transfer reactions become important. During the interaction<br />

of two ions, an electron may be transferred from one ion to the other; it is usually<br />

captured in an excited state, and the excited ion is stabilised by one or more radiative<br />

transitions. As hydrogen and helium outnumber by at least a factor of 100 any other<br />

element, in practice only interactions between those elements and heavier nuclei are<br />

important. Reactions with Hi and Hei lead to recombination of the heavier ion,<br />

and reactions with Hii and Heii to ionisation.<br />

Figure 3.22: The effects of charge transfer reactions on the ionisation balance.<br />

We show the role of charge transfer processes on the ionisation balance in Fig. 3.22.<br />

The electron captured during charge transfer recombination of an oxygen ion<br />

(for instance Ovii, Oviii) is usually captured in an intermediate quantum state<br />

(principal quantum number n = 4 − 6). This leads to enhanced line emission from<br />

that level as compared to the emission from other principal levels, and this signature<br />

can be used to show the presence of charge transfer reactions. Another signature –<br />

actually a signature for all recombining plasmas – is of course the enhancement of<br />

the forbidden line relative to the resonance line in the Ovii triplet (Sect. 3.9).<br />

An important example is the charge transfer of highly charged ions from the<br />

Solar wind with the neutral or weakly ionised Geocorona. Whenever the Sun is<br />

more active, this process may produce enhanced thermal soft X-<strong>ray</strong> emission in<br />

addition to the steady foreground emission from our own Galaxy. See Bhardwaj et<br />

al. (2006) for a review of X-<strong>ray</strong>s from the Solar System. Care should be taken not<br />

to confuse this temporary enhanced Geocoronal emission with soft excess emission<br />

in a background astrophysical source.<br />

42


3.7 The ionisation balance<br />

In order to calculate the X-<strong>ray</strong> emission or absorption from a plasma, apart from the<br />

physical conditions also the ion concentrations must be known. These ion concentrations<br />

can be determined by solving the equations for ionisation balance (or in more<br />

complicated cases by solving the time-dependent equations). A basic ingredient in<br />

these equations are the ionisation and recombination rates, that we discussed in the<br />

previous section. Here we consider three of the most important cases: collisional<br />

ionisation equilibrium, non-equilibrium ionisation and photoionisation equilibrium.<br />

3.7.1 Collisional Ionisation Equilibrium (CIE)<br />

The simplest case is a plasma in collisional ionisation equilibrium (CIE). In this case<br />

one assumes that the plasma is optically thin for its own <strong>radiation</strong>, and that there<br />

is no external <strong>radiation</strong> field that affects the ionisation balance.<br />

Photo-ionisation and Compton ionisation therefore can be neglected in the case<br />

of CIE. This means that in general each ionisation leads to one additional free electron,<br />

because the direct ionisation and excitation-autoionisation processes are most<br />

efficient for the outermost atomic shells. The relevant ionisation processes are collisional<br />

ionisation and excitation-autoionisation, and the recombination processes<br />

are radiative recombination and dielectronic recombination. Apart from these processes,<br />

at low temperatures also charge transfer ionisation and recombination are<br />

important.<br />

We define R z as the total recombination rate of an ion with charge z to charge<br />

z −1, and I z as the total ionisation rate for charge z to z +1. Ionisation equilibrium<br />

then implies that the net change of ion concentrations n z should be zero:<br />

z > 0 : n z+1 R z+1 − n z R z + n z−1 I z−1 − n z I z = 0 (3.55)<br />

and in particular for z = 0 one has<br />

n 1 R 1 = n 0 I 0 (3.56)<br />

(a neutral atom cannot recombine further and it cannot be created by ionisation).<br />

Next an arbitrary value for n 0 is chosen, and (3.56) is solved:<br />

n 1 = n 0 (I 0 /R 1 ). (3.57)<br />

This is substituted into (3.55) which now can be solved. Using induction, it follows<br />

that<br />

n z+1 = n z (I z /R z+1 ). (3.58)<br />

Finally everything is normalised by demanding that<br />

Z∑<br />

n z = n element (3.59)<br />

z=0<br />

where n element is the total density of the element, determined by the total plasma<br />

density and the chemical abundances.<br />

Examples of plasmas in CIE are the Solar corona, coronae of stars, the hot<br />

intracluster medium, the diffuse Galactic ridge component. Fig. 3.23 shows the ion<br />

fractions as a function of temperature for two important elements.<br />

43


Figure 3.23: Ion concentration of oxygen ions (left panel) and iron ions (right panel)<br />

as a function of temperature in a plasma in Collisional Ionisation Equilibrium (CIE).<br />

Ions with completely filled shells are indicated with thick lines: the He-like ions Ovii<br />

and Fexxv, the Ne-like Fexvii and the Ar-like Feix; note that these ions are more<br />

prominent than their neighbours.<br />

3.7.2 Non-Equilibrium Ionisation (NEI)<br />

The second case that we discuss is non-equilibrium ionisation (NEI). This situation<br />

occurs when the physical conditions of the source, like the temperature, suddenly<br />

change. A shock, for example, can lead to an almost instantaneous rise in temperature.<br />

However, it takes a finite time for the plasma to respond to the temperature<br />

change, as ionisation balance must be recovered by collisions. Similar to the CIE<br />

case we assume that photoionisation can be neglected. For each element with nuclear<br />

charge Z we write:<br />

1<br />

n e (t)<br />

d<br />

⃗n(Z, t) = A(Z, T(t))⃗n(Z, t) (3.60)<br />

dt<br />

where ⃗n is a vector of length Z + 1 that contains the ion concentrations, and which<br />

is normalised according to Eqn. 3.59. The transition matrix A is a (Z +1) ×(Z +1)<br />

matrix given by<br />

⎛<br />

A =<br />

⎜<br />

⎝<br />

⎞<br />

−I 0 R 1 0 0 . . .<br />

I 0 −(I 1 + R 1 ) R 2 0<br />

0 I 1 . . . . . .<br />

.<br />

. .. . . . .<br />

. . . R Z−1 0<br />

⎟<br />

. . . 0 I Z−2 −(I Z−1 + R Z−1 ) R Z<br />

⎠<br />

. . . 0 I Z−1 −R Z<br />

We can write the equation in this form because both ionisations and recombinations<br />

are caused by collisions of electrons with ions. Therefore we have the uniform scaling<br />

with n e .<br />

In general, the set of equations (3.60) must be solved numerically. Because the<br />

characteristic time constants of the ions can have quite different values, solution<br />

through standard methods (for example Runge-Kutta integration) will take a lot<br />

44<br />

.


of computing time. Therefore one uses a trick. First, the eigenvalues (λ) of A<br />

are determined, and also the corresponding eigenvectors ⃗ V . These are put into a<br />

matrix V. We now transform the concentrations ⃗n to the eigenvector basis; the<br />

concentrations on the eigenvector basis will be denoted by ⃗m. By definition, we<br />

then have<br />

⃗n = V⃗m. (3.61)<br />

We now define a diagonal matrix Λ which has the eigenvalues on its diagonal. We<br />

then have<br />

AV = ΛV. (3.62)<br />

Using this we rewrite (3.60) as<br />

1 d<br />

V⃗m = ΛV⃗m. (3.63)<br />

n e dt<br />

For small steps in T the transition matrix A does not vary much. Therefore, for<br />

any time interval where T changes by only a small amount we can assume A and<br />

thereby Λ and V to be constant. In that case, one can interchange the order of<br />

matrix multiplication and differentiation in the left hand side of (3.63). It is then<br />

simple to solve this equation. One finds as a solution<br />

⃗m(t) = V −1 PV⃗m(t 1 ) (3.64)<br />

where the solution is valid for t ≥ t 1 and as long as T(t) − T(t 1 ) is small. Here P is<br />

a diagonal matrix with elements e λ∆u where<br />

∆u ≡<br />

∫ t<br />

t 1<br />

n e (t)dt. (3.65)<br />

For each interval in T one can now determine the eigenvectors and eigenvalues<br />

once and for all and store these for example in the computer memory. By systematically<br />

subdividing the range of T(t) in small intervals ∆T, one finds the final solution<br />

simply by applying repeatedly simple exponentiation and matrix multiplication.<br />

The time evolution of the plasma can be described in general well by the parameter<br />

∫<br />

U = n e dt. (3.66)<br />

The integral should be done over a co-moving mass element. Typically, for most<br />

ions equilibrium is reached for U ∼ 10 18 m −3 s. We should mention here, however,<br />

that the final outcome also depends on the temperature history T(t) of the mass<br />

element, but in most cases the situation is simplified to T(t) = constant.<br />

3.7.3 Photoionisation Equilibrium (PIE)<br />

The third case that we treat are the photoionised plasmas. Usually one assumes<br />

equilibrium (PIE), but there are of course also extensions to non-equilibrium photoionised<br />

situations. Apart from the same ionisation and recombination processes that<br />

play a role for plasmas in NEI and CIE, also photoionisation and Compton ionisation<br />

are relevant. Because of the multiple ionisations caused by Auger processes, the<br />

45


equation for the ionisation balance is not as simple as (3.55), because now one needs<br />

to couple more ions. Moreover, not all rates scale with the product of electron<br />

and ion density, but the balance equations also contain terms proportional to the<br />

product of ion density times photon density. In addition, apart from the equation<br />

for the ionisation balance, one needs to solve simultaneously an energy balance<br />

equation for the electrons. In this energy equation not only terms corresponding<br />

to ionisation and recombination processes play a role, but also several <strong>radiation</strong><br />

processes (Bremsstrahlung, line <strong>radiation</strong>) or Compton scattering. The equilibrium<br />

temperature must be determined in an iterative way. A classical paper describing<br />

such photoionised plasmas is Kallman & McG<strong>ray</strong> (1982).<br />

46


3.8 Continuum emission processes<br />

After we have determined the ionisation balance, and solved for the ion concentrations,<br />

it is possible to calculate the X-<strong>ray</strong> spectrum of a plasma. We distinguish<br />

continuum <strong>radiation</strong> and line <strong>radiation</strong>. Here we discuss briefly the continuum emission<br />

processes: Bremsstrahlung, free-bound emission and two-photon emission.<br />

3.8.1 Bremsstrahlung<br />

Bremsstrahlung is caused by a collision between a free electron and an ion. The<br />

emissivity ǫ ff (photonsm −3 s −1 J −1 ) can be written as:<br />

ǫ ff = 2√ 2ασ T cn e n i Zeff<br />

2 ( me c 2 ) 1<br />

2<br />

√ g ff e −E/kT , (3.67)<br />

3πE kT<br />

where α is the fine structure constant, σ T the Thomson cross section, n e and n i the<br />

electron and ion density, and E the energy of the emitted photon. The factor g ff is<br />

the so-called Gaunt factor and is a dimensionless quantity of order unity. Further,<br />

Z eff is the effective charge of the ion, defined as<br />

Z eff =<br />

( n<br />

2 )1<br />

r I r 2<br />

E H<br />

(3.68)<br />

where E H is the ionisation energy of hydrogen (13.6 eV), I r the ionisation potential of<br />

the ion after a recombination, and n r the corresponding principal quantum number.<br />

Example 3.14. For the Bremsstrahlung of Feii (Fe + ) one should take the ionisation<br />

potential of neutral iron (7.87 eV); the outermost electron for neutral iron has a 4s<br />

orbit, implying n r = 4 and hence for Feii we find Z eff = 3.04.<br />

Figure 3.24: The Gaunt factor for Bremsstrahlung. Note that ζ is Euler’s constant<br />

(1.781 . . .).<br />

It is also possible to write (3.67) as ǫ ff = P ff n e n i with<br />

P ff = 3.031 × 10−21 Z 2 eff g effe −E/kT<br />

E keV T 1/2<br />

keV<br />

47<br />

, (3.69)


where in this case P ff is in photons ×m 3 s −1 keV −1 and E keV is the energy in keV.<br />

The total amount of <strong>radiation</strong> produced by this process is given by<br />

W tot = 4.856 × 10−37 Wm 3<br />

√<br />

TkeV<br />

∫∞<br />

0<br />

Z 2 eff g ffe −E/kT dE keV . (3.70)<br />

From (3.67) we see immediately that the Bremsstrahlung spectrum (expressed in<br />

Wm −3 keV −1 ) is flat for E ≪ kT, and for E > kT it drops exponentially. In order<br />

to measure the temperature of a hot plasma, one needs to measure near E ≃ kT. The<br />

Gaunt factor g ff can be calculated analytically; there are both tables and asymptotic<br />

approximations available. In general, g ff depends on both E/kT and kT/Z eff . See<br />

Fig. 3.24.<br />

For a plasma (3.67) needs to be summed over all ions that are present in order<br />

to calculate the total amount of Bremsstrahlung <strong>radiation</strong>. For cosmic abundances,<br />

hydrogen and helium usually yield the largest contribution. Frequently, one defines<br />

an average Gaunt factor G ff by<br />

G ff = ∑ ( ni<br />

)<br />

Zeff,i 2 g eff,i . (3.71)<br />

n<br />

i e<br />

3.8.2 Free-bound emission<br />

Free-bound emission occurs during radiative recombination (Sect. 3.6.9).<br />

exercise 3.11. Why is there no free-bound emission for dielectronic recombination?<br />

The energy of the emitted photon is at least the ionisation energy of the recombined<br />

ion (for recombination to the ground level) or the ionisation energy that<br />

corresponds to the excited state (for recombination to higher levels). From the<br />

recombination rate (see Sect. 3.6.9) the free-bound emission is determined immediately:<br />

ǫ fb = ∑ i<br />

n e n i R r . (3.72)<br />

Also here it is possible to define an effective Gaunt factor G fb . Free-bound emission<br />

is in practice often very important. For example in CIE for kT = 0.1 keV, free-bound<br />

emission is the dominant continuum mechanism for E > 0.1 keV; for kT = 1 keV<br />

it dominates above 3 keV. For kT ≫ 1 keV Bremsstrahlung is always the most<br />

important mechanism, and for kT ≪ 0.1 keV free-bound emission dominates. See<br />

also Fig. 3.25.<br />

Of course, under conditions of photoionisation equilibrium free-bound emission<br />

is even more important, because there are more recombinations than in the CIE<br />

case (because T is lower, at comparable ionisation).<br />

3.8.3 Two photon emission<br />

This process is in particular important for hydrogen-like or helium-like ions. After<br />

a collision with a free electron, an electron from a bound 1s shell is excited to the<br />

2s shell. The quantum-mechanical selection rules do not allow that the 2s electron<br />

48


decays back to the 1s orbit by a radiative transition. Usually the ion will then be<br />

excited to a higher level by another collision, for example from 2s to 2p, and then it<br />

can decay radiatively back to the ground state (1s). However, if the density is very<br />

low (n e ≪ n e,crit , Eqn. 3.74–3.75), the probability for a second collision is very small<br />

and in that case two-photon emission can occur: the electron decays from the 2s<br />

orbit to the 1s orbit while emitting two photons. Energy conservation implies that<br />

the total energy of both photons should equal the energy difference between the 2s<br />

and 1s level (E 2phot = E 1s −E 2s ). From symmetry considerations it is clear that the<br />

spectrum must be symmetrical around E = 0.5E 2phot , and further that it should be<br />

zero for E = 0 and E = E 2phot . An empirical approximation for the shape of the<br />

spectrum is given by:<br />

√<br />

F(E) ∼ sin(πE/E 2phot ). (3.73)<br />

An approximation for the critical density below which two photon emission is important<br />

can be obtained from a comparison of the radiative and collisional rates from<br />

the upper (2s) level, and is given by Mewe, Gronenschild & van den Oord (1986):<br />

H − like : n e,crit = 7 × 10 9 m −3 Z 9.5 (3.74)<br />

He − like : n e,crit = 2 × 10 11 m −3 (Z − 1) 9.5 . (3.75)<br />

For example for carbon two photon emission is important for densities below 10 17 m −3 ,<br />

which is the case for many astrophysical applications. Also in this case one can determine<br />

an average Gaunt factor G 2phot by averaging over all ions. Two photon<br />

emission is important in practice for 0.5 kT 5 keV, and then in particular for<br />

the contributions of C, N and O between 0.2 and 0.6 keV. See also Fig. 3.25.<br />

3.9 Line <strong>radiation</strong><br />

Apart from continuum <strong>radiation</strong>, line <strong>radiation</strong> plays an important role for thermal<br />

plasmas. In some cases the flux, integrated over a broad energy band, can be<br />

completely dominated by line <strong>radiation</strong> (see Fig. 3.25). The production process of<br />

line <strong>radiation</strong> can be broken down in two basic steps: the excitation and the emission<br />

process.<br />

3.9.1 Line <strong>radiation</strong>: excitation<br />

Collisional excitation<br />

A bound electron in an ion can be brought into a higher, excited energy level through<br />

a collision with a free electron. Before we already discussed the case where the<br />

excited state is unstable, which sometimes can lead to auto-ionisation. In most<br />

cases, however, the excited state is stable and the ion will decay back to the ground<br />

level by means of a radiative transition, either directly or through one or more steps<br />

through intermediate energy levels.<br />

We have seen before that the excitation rate from level i to level j can be written<br />

as (Eq. 3.9):<br />

S ij = S 0 ¯Ω(Eij /kT)T −1/2 e −E ij/kT , (3.76)<br />

For sufficient low density every excitation is followed immediately by a radiative<br />

de-excitation to the ground level (or an intermediate level). Apart of a small<br />

49


Figure 3.25: Emission spectra of plasmas with solar abundances. The histogram<br />

indicates the total spectrum, including line <strong>radiation</strong>. The spectrum has been binned<br />

in order to show better the relative importance of line <strong>radiation</strong>. The thick solid<br />

line is the total continuum emission, the thin solid line the contribution due to<br />

Bremsstrahlung, the dashed line free-bound emission and the dotted line two-photon<br />

emission. Note the scaling with Ee E/kT along the y-axis.<br />

correction factor C 1 for cascades from higher excited levels to level k, and possibly<br />

a correction factor B for decay to intermediate levels or for autoionisation, (3.9)<br />

gives the line power from level k to the ground level. The factor B is the so-called<br />

branching ratio.<br />

Radiative recombination<br />

When a free electron is captured by an ion, continuum <strong>radiation</strong> is produced (freebound<br />

emission). The captured electron not always reaches the ground level immediately.<br />

We have seen before that in particular for cool plasma’s (kT ≪ I) the<br />

higher excited levels are frequently populated. In order to get to the ground state,<br />

one or more radiative transitions are required. Apart from cascade corrections from<br />

and to higher levels the recombination line <strong>radiation</strong> is essentially given by (3.49).<br />

A comparison of recombination with excitation tells that in particular for low temperatures<br />

(compared to the line energy) recombination <strong>radiation</strong> dominates, and for<br />

high temperatures excitation <strong>radiation</strong>.<br />

50


Inner-shell ionisation<br />

The third line <strong>radiation</strong> process is inner–shell ionisation. An electron from one of<br />

the inner shells can be removed from an ion not only by photoionisation, but also<br />

through a collision with electrons. This last process is in particular important under<br />

NEI conditions, for example in supernova remnants. There the plasma temperature<br />

can be much higher than the typical ionisation energy of the ions that are present.<br />

By means of fluorescence (discussed before) the hole in the shell can be filled again.<br />

Obviously, the line power is proportional to the ionisation cross section times the<br />

fluorescence yield.<br />

Dielectronic recombination<br />

Dielectronic recombination produces more than one line photon. Consider for example<br />

the dielectronic recombination of a He-like ion into a Li-like ion:<br />

e + 1s 2 → 1s2p3s → 1s 2 3s + hν 1 → 1s 2 2p + hν 2 → 1s 2 2s + hν 3 (3.77)<br />

The first arrow corresponds to the electron capture, the second arrow to the<br />

stabilising radiative transition 2p→1s and the third arrow to the radiative transition<br />

3s→2p of the captured electron. This last transition would have also occurred if the<br />

free electron was caught directly into the 3s shell by normal radiative recombination.<br />

Finally, the electron has to decay further to the ground level and this can go through<br />

the normal transitions in a Li-like ion (fourth arrow). This single recombination thus<br />

produces three line photons.<br />

Because of the presence of the extra electron in the higher orbit, the energy hν 1<br />

of the 2p→1s transition is slightly different from the energy in a normal He-like ion.<br />

The stabilising transition is therefore also called a satellite line. Because there are<br />

many different possibilities for the orbit of the captured electron, one usually finds<br />

a forest of such satellite lines surrounding the normal 2p→1s excitation line in the<br />

He-like ion (or analogously for other iso-electronic sequences). Fig. 3.26 gives an<br />

example of these satellite lines.<br />

3.9.2 Line <strong>radiation</strong>: emission<br />

It does not matter by whatever process the ion is brought into an excited state j,<br />

whenever it is in such a state it may decay back to the ground state or any other<br />

lower energy level i by emitting a photon. The probability per unit time that this<br />

occurs is given by the spontaneous transition probability A ij (units: s −1 ) which is a<br />

number that is different for each transition. The total line power P ij (photons per<br />

unit time and volume) is then given by<br />

P ij = A ij n j (3.78)<br />

where n j is the number density of ions in the excited state j. For the most simple<br />

case of excitation from the ground state g (rate S gj ) followed by spontaneous<br />

emission, one can simply approximate n g n e S gj = n j A gj . From this equation, the<br />

relative population n j /n g ≪ 1 is determined, and then using (3.78) the line flux<br />

is determined. In realistic situations, however, things are more complicated. First,<br />

51


Figure 3.26: Spectrum of a plasma in collisional ionisation equilibrium with kT =<br />

2 keV, near the Fe-K complex. Lines are labelled using the most common designations<br />

in this field. The Fexxv ”triplet” consists of the resonance line (w),<br />

intercombination line (actually split into x and y) and the forbidden line (z). All<br />

other lines are satellite lines. The labelled satellites are lines from Fexxiv, most of<br />

the lines with energy below the forbidden (z) line are from Fexxiii. The relative<br />

intensity of these satellites is a strong indicator for the physical conditions in the<br />

source.<br />

the excited state may also decay to other intermediate states if present, and also<br />

excitations or cascades from other levels may play a role. Furthermore, for high densities<br />

also collisional excitation or de-excitation to and from other levels becomes<br />

important. In general, one has to solve a set of equations for all energy levels of<br />

an ion where all relevant population and depopulation processes for that level are<br />

taken into account. For the resulting solution vector n j , the emitted line power is<br />

then simply given by Eqn. (3.78).<br />

Note that not all possible transitions between the different energy levels are<br />

allowed. There are strict quantum mechanical selection rules that govern which<br />

lines are allowed; see for instance Herzberg (1944) or Mee (1999). Sometimes there<br />

are higher order processes that still allow a forbidden transition to occur, albeit with<br />

much smaller transition probabilities A ij . But if the excited state j has no other<br />

(fast enough) way to decay, these forbidden lines occur and the lines can be quite<br />

strong, as their line power is essentially governed by the rate at which the ion is<br />

brought into its excited state j.<br />

One of the most well known groups of lines is the He-like 1s–2p triplet. Usually<br />

the strongest line is the resonance line, an allowed transition. The forbidden line<br />

has a similar strength as the resonance line, for the low density conditions in the<br />

ISM and intracluster medium, but it can be relatively enhanced in recombining<br />

plasmas, or relatively reduced in high density plasmas like stellar coronal loops. In<br />

between both lines is the intercombination line. In fact, this intercombination line<br />

is a doublet but for the lighter elements both components cannot be resolved. But<br />

see Fig. 3.26 for the case of iron.<br />

52


3.9.3 Line width<br />

For most X-<strong>ray</strong> spectral lines, the line profile of a line with energy E can be approximated<br />

with a Gaussian exp (−∆E 2 /2σ 2 ) with σ given by σ/E = σ v /c where the<br />

velocity dispersion is<br />

σ 2 v = σ2 t + kT i/m i . (3.79)<br />

Here T i is the ion temperature (not necessarily the same as the electron temperature),<br />

and σ t is the root mean squared turbulent velocity of the emitting medium. For large<br />

ion temperature, turbulent velocity or high spectral resolution this line width can<br />

be measured, but in most cases the lines are not resolved for CCD type spectra.<br />

3.9.4 Resonance scattering<br />

Resonance scattering is a process where a photon is absorbed by an atom and then<br />

re-emitted as a line photon of the same energy into a different direction. As for<br />

strong resonance lines (allowed transitions) the transition probabilities A ij are large,<br />

the time interval between absorption and emission is extremely short, and that is<br />

the reason why the process effectively can be regarded as a scattering process. We<br />

discuss the absorption properties in Sect. 3.10.3, and have already discussed the<br />

spontaneous emission in Sect. 3.9.2.<br />

Resonance scattering of X-<strong>ray</strong> photons is potentially important in the dense cores<br />

of some clusters of galaxies for a few transitions of the most abundant elements, as<br />

first shown by Gilfanov et al. (1987). The optical depth for scattering can be written<br />

conveniently as (cf. also Sect. 3.10.3):<br />

τ =<br />

4240 fN 24<br />

(<br />

ni<br />

n Z<br />

) (<br />

nZ<br />

n H<br />

) ( M<br />

T keV<br />

) 1/2<br />

E keV<br />

{<br />

1 + 0.0522Mv2 100<br />

T keV<br />

} 1/2<br />

, (3.80)<br />

where f is the absorption oscillator strength of the line (of order unity for the<br />

strongest spectral lines), E keV the energy in keV, N 24 the hydrogen column density<br />

in units of 10 24 m −2 , n i the number density of the ion, n Z the number density of the<br />

element, M the atomic weight of the ion, T keV the ion temperature in keV (assumed<br />

to be equal to the electron temperature) and v 100 the micro-turbulence velocity in<br />

units of 100 km/s. Resonance scattering in clusters causes the radial intensity profile<br />

on the sky of an emission line to become weaker in the cluster core and stronger in<br />

the outskirts, without destroying photons. By comparing the radial line profiles of<br />

lines with different optical depth, for instance the 1s−2p and 1s−3p lines of Ovii<br />

or Fexxv, one can determine the optical depth and hence constrain the amount of<br />

turbulence in the cluster.<br />

Another important application was proposed by Churazov et al. (2001). They<br />

show that for WHIM filaments the resonance line of the Ovii triplet can be enhanced<br />

significantly compared to the thermal emission of the filament due to resonant scattering<br />

of X-<strong>ray</strong> background photons on the filament. The ratio of this resonant line<br />

to the other lines of the triplet therefore can be used to estimate the column density<br />

of a filament.<br />

53


3.9.5 Some important line transitions<br />

In Tables 3.6–3.7 we list the 100 strongest emission lines under CIE conditions.<br />

Note that each line has its peak emissivity at a different temperature. In particular<br />

some of the H-like and He-like transitions are strong, and further the so-called Fe-L<br />

complex (lines from n = 2 in Li-like to Ne-like iron ions) is prominent. At longer<br />

wavelengths, the L-complex of Ne Mg, Si and S gives strong soft X-<strong>ray</strong> lines. At<br />

very short wavelengths, there are not many strong emission lines: between 6–7 keV,<br />

the Fe-K emission lines are the strongest spectral features.<br />

We conclude here with mentioning some important iron lines.<br />

First, in hydrogen-like ions the Lyman series (np→1s) and the Balmer series<br />

(n → 2) are important. For He-like ions in particular the 2 → 1 transitions are<br />

very prominent, and thee lines from the He-like ions are often stronger than for<br />

hydrogen-like ions. Also thee Ne-like ions often have a very pronounced spectrum.<br />

Some important lines from iron are mentioned in Table 3.8.<br />

54


Table 3.6: The strongest emission lines for a plasma with proto-solar abundances (Lodders 2003) in the X-<strong>ray</strong><br />

band 43 Å < λ < 100 Å. At longer wavelengths sometimes a few lines from the same multiplet have been added.<br />

All lines include unresolved dielectronic satellites. T max (K) is the temperature at which the emissivity peaks,<br />

Q max = P/(n en H ), with P the power per unit volume at T max, and Q max is in units of 10 −36 Wm 3 .<br />

E λ − log log ion iso-el. lower upper<br />

(eV) (Å) Q max T max seq. level level<br />

126.18 98.260 1.35 5.82 Neviii Li 2p 2 P 3/2 3d 2 D 5/2<br />

126.37 98.115 1.65 5.82 Neviii Li 2p 2 P 1/2 3d 2 D 3/2<br />

127.16 97.502 1.29 5.75 Nevii Be 2s 1 S 0 3p 1 P 1<br />

128.57 96.437 1.61 5.47 Siv Ne 2p 1 S 0 3d 1 P 1<br />

129.85 95.483 1.14 5.73 Mgvi N 2p 4 S 3/2 3d 4 P 5/2,3/2,1/2<br />

132.01 93.923 1.46 6.80 Fexviii F 2s 2 P 3/2 2p 2 S 1/2<br />

140.68 88.130 1.21 5.75 Nevii Be 2p 3 P 1 4d 3 D 2,3<br />

140.74 88.092 1.40 5.82 Neviii Li 2s 2 S 1/2 3p 2 P 3/2<br />

147.67 83.959 0.98 5.86 Mgvii C 2p 3 P 3d 3 D, 1 D, 3 F<br />

148.01 83.766 1.77 5.86 Mgvii C 2p 3 P 2 3d 3 P 2<br />

148.56 83.457 1.41 5.90 Feix Ar 3p 1 S 0 4d 3 P 1<br />

149.15 83.128 1.69 5.69 Sivi F 2p 2 P 3/2 ( 3 P)3d 2 D 5/2<br />

149.38 83.000 1.48 5.74 Mgvi N 2p 4 S 3/2 4d 4 P 5/2,3/2,1/2<br />

150.41 82.430 1.49 5.91 Feix Ar 3p 1 S 0 4d 1 P 1<br />

154.02 80.501 1.75 5.70 Sivi F 2p 2 P 3/2 ( 1 D) 3d 2 D 5/2<br />

159.23 77.865 1.71 6.02 Fex Cl 3p 2 P 1/2 4d 2 D 5/2<br />

165.24 75.034 1.29 5.94 Mgviii B 2p 2 P 3/2 3d 2 D 5/2<br />

165.63 74.854 1.29 5.94 Mgviii B 2p 2 P 1/2 3d 2 D 3/2<br />

170.63 72.663 1.07 5.76 Svii Ne 2p 1 S 0 3s 3 P 1,2<br />

170.69 72.635 1.61 6.09 Fexi S 3p 3 P 2 4d 3 D 3<br />

171.46 72.311 1.56 6.00 Mgix Be 2p 1 P 1 3d 1 D 2<br />

171.80 72.166 1.68 6.08 Fexi S 3p 1 D 2 4d 1 F 3<br />

172.13 72.030 1.44 6.00 Mgix Be 2p 3 P 2,1 3s 3 S 1<br />

172.14 72.027 1.40 5.76 Svii Ne 2p 1 S 0 3s 1 P 1<br />

177.07 70.020 1.18 5.84 Sivii O 2p 3 P 3d 3 D, 3 P<br />

177.98 69.660 1.36 6.34 Fexv Mg 3p 1 P 1 4s 1 S 0<br />

177.99 69.658 1.57 5.96 Siviii N 2p 4 S 3/2 3s 4 P 5/2,3/2,1/2<br />

179.17 69.200 1.61 5.71 Sivi F 2p 2 P 4d 2 P, 2 D<br />

186.93 66.326 1.72 6.46 Fexvi Na 3d 2 D 4f 2 F<br />

194.58 63.719 1.70 6.45 Fexvi Na 3p 2 P 3/2 4s 2 S 1/2<br />

195.89 63.294 1.64 6.08 Mgx Li 2p 2 P 3/2 3d 2 D 5/2<br />

197.57 62.755 1.46 6.00 Mgix Be 2s 1 S 0 3p 1 P 1<br />

197.75 62.699 1.21 6.22 Fexiii Si 3p 3 P 1 4d 3 D 2<br />

198.84 62.354 1.17 6.22 Fexiii Si 3p 3 P 0 4d 3 D 1<br />

199.65 62.100 1.46 6.22 Fexiii Si 3p 3 P 1 4d 3 P 0<br />

200.49 61.841 1.29 6.07 Siix C 2p 3 P 2 3s 3 P 1<br />

203.09 61.050 1.06 5.96 Siviii N 2p 4 S 3/2 3d 4 P 5/2,3/2,1/2<br />

203.90 60.807 1.69 5.79 Svii Ne 2p 1 S 0 3d 3 D 1<br />

204.56 60.610 1.30 5.79 Svii Ne 2p 1 S 0 3d 1 P 1<br />

223.98 55.356 1.00 6.08 Siix C 2p 3 P 3d 3 D, 1 D, 3 F<br />

234.33 52.911 1.34 6.34 Fexv Mg 3s 1 S 0 4p 1 P 1<br />

237.06 52.300 1.61 6.22 Sixi Be 2p 1 P 1 3s 1 S 0<br />

238.43 52.000 1.44 5.97 Siviii N 2p 4 S 3/2 4d 4 P 5/2,3/2,1/2<br />

244.59 50.690 1.30 6.16 Six B 2p 2 P 3/2 3d 2 D 5/2<br />

245.37 50.530 1.30 6.16 Six B 2p 2 P 1/2 3d 2 D 3/2<br />

251.90 49.220 1.45 6.22 Sixi Be 2p 1 P 1 3d 1 D 2<br />

252.10 49.180 1.64 5.97 Arix Ne 2p 1 S 0 3s 3 P 1,2<br />

261.02 47.500 1.47 6.06 Six O 2p 3 P 3d 3 D, 3 P<br />

280.73 44.165 1.60 6.30 Sixii Li 2p 2 P 3/2 3d 2 D 5/2<br />

283.46 43.740 1.46 6.22 Sixi Be 2s 1 S 0 3p 1 P 1<br />

55


Table 3.7: As Table 3.6, but for λ < 43 Å.<br />

E λ − log log ion iso-el. lower upper<br />

(eV) (Å) Q max T max seq. level level<br />

291.52 42.530 1.32 6.18 Sx N 2p 4 S 3/2 3d 4 P 5/2,3/2,1/2<br />

298.97 41.470 1.31 5.97 Cv He 1s 1 S 0 2s 3 S 1 (f)<br />

303.07 40.910 1.75 6.29 Sixii Li 2s 2 S 1/2 3p 2 P 3/2<br />

307.88 40.270 1.27 5.98 Cv He 1s 1 S 0 2p 1 P 1 (r)<br />

315.48 39.300 1.37 6.28 Sxi C 2p 3 P 3d 3 D, 1 D, 3 F<br />

336.00 36.900 1.58 6.19 Sx N 2p 4 S 3/2 4d 4 P 5/2,3/2,1/2<br />

339.10 36.563 1.56 6.34 Sxii B 2p 2 P 3/2 3d 2 D 5/2<br />

340.63 36.398 1.56 6.34 Sxii B 2p 2 P 1/2 3d 2 D 3/2<br />

367.47 33.740 1.47 6.13 Cvi H 1s 2 S 1/2 2p 2 P 1/2 (Lyα)<br />

367.53 33.734 1.18 6.13 Cvi H 1s 2 S 1/2 2p 2 P 3/2 (Lyα)<br />

430.65 28.790 1.69 6.17 Nvi He 1s 1 S 0 2p 1 P 1 (r)<br />

500.36 24.779 1.68 6.32 Nvii H 1s 2 S 1/2 2p 2 P 3/2 (Lyα)<br />

560.98 22.101 0.86 6.32 Ovii He 1s 1 S 0 2s 3 S 1 (f)<br />

568.55 21.807 1.45 6.32 Ovii He 1s 1 S 0 2p 3 P 2,1 (i)<br />

573.95 21.602 0.71 6.33 Ovii He 1s 1 S 0 2p 1 P 1 (r)<br />

653.49 18.973 1.05 6.49 Oviii H 1s 2 S 1/2 2p 2 P 1/2 (Lyα)<br />

653.68 18.967 0.77 6.48 Oviii H 1s 2 S 1/2 2p 2 P 3/2 (Lyα)<br />

665.62 18.627 1.58 6.34 Ovii He 1s 1 S 0 3p 1 P 1<br />

725.05 17.100 0.87 6.73 Fexvii Ne 2p 1 S 0 3s 3 P 2<br />

726.97 17.055 0.79 6.73 Fexvii Ne 2p 1 S 0 3s 3 P 1<br />

738.88 16.780 0.87 6.73 Fexvii Ne 2p 1 S 0 3s 1 P 1<br />

771.14 16.078 1.37 6.84 Fexviii F 2p 2 P 3/2 3s 4 P 5/2<br />

774.61 16.006 1.55 6.50 Oviii H 1s 2 S 1/2 3p 2 P 1/2,3/2 (Lyβ)<br />

812.21 15.265 1.12 6.74 Fexvii Ne 2p 1 S 0 3d 3 D 1<br />

825.79 15.014 0.58 6.74 Fexvii Ne 2p 1 S 0 3d 1 P 1<br />

862.32 14.378 1.69 6.84 Fexviii F 2p 2 P 3/2 3d 2 D 5/2<br />

872.39 14.212 1.54 6.84 Fexviii F 2p 2 P 3/2 3d 2 S 1/2<br />

872.88 14.204 1.26 6.84 Fexviii F 2p 2 P 3/2 3d 2 D 5/2<br />

896.75 13.826 1.66 6.76 Fexvii Ne 2s 1 S 0 3p 1 P 1<br />

904.99 13.700 1.61 6.59 Neix He 1s 1 S 0 2s 3 S 1 (f)<br />

905.08 13.699 1.61 6.59 Neix He 1s 1 S 0 2s 3 S 1 (f)<br />

916.98 13.521 1.35 6.91 Fexix O 2p 3 P 2 3d 3 D 3<br />

917.93 13.507 1.68 6.91 Fexix O 2p 3 P 2 3d 3 P 2<br />

922.02 13.447 1.44 6.59 Neix He 1s 1 S 0 2p 1 P 1 (r)<br />

965.08 12.847 1.51 6.98 Fexx N 2p 4 S 3/2 3d 4 P 5/2<br />

966.59 12.827 1.44 6.98 Fexx N 2p 4 S 3/2 3d 4 P 3/2<br />

1009.2 12.286 1.12 7.04 Fexxi C 2p 3 P 0 3d 3 D 1<br />

1011.0 12.264 1.46 6.73 Fexvii Ne 2p 1 S 0 4d 3 D 1<br />

1021.5 12.137 1.77 6.77 Nex H 1s 2 S 1/2 2p 2 P 1/2 (Lyα)<br />

1022.0 12.132 1.49 6.76 Nex H 1s 2 S 1/2 2p 2 P 3/2 (Lyα)<br />

1022.6 12.124 1.39 6.73 Fexvii Ne 2p 1 S 0 4d 1 P 1<br />

1053.4 11.770 1.38 7.10 Fexxii B 2p 2 P 1/2 3d 2 D 3/2<br />

1056.0 11.741 1.51 7.18 Fexxiii Be 2p 1 P 1 3d 1 D 2<br />

1102.0 11.251 1.73 6.74 Fexvii Ne 2p 1 S 0 5d 3 D 1<br />

1352.1 9.170 1.66 6.81 Mgxi He 1s 1 S 0 2p 1 P 1 (r)<br />

1472.7 8.419 1.76 7.00 Mgxii H 1s 2 S 1/2 2p 2 P 3/2 (Lyα)<br />

1864.9 6.648 1.59 7.01 Sixiii He 1s 1 S 0 2p 1 P 1 (r)<br />

2005.9 6.181 1.72 7.21 Sixiv H 1s 2 S 1/2 2p 2 P 3/2 (Lyα)<br />

6698.6 1.851 1.43 7.84 Fexxv He 1s 1 S 0 2p 1 P 1 (r)<br />

6973.1 1.778 1.66 8.17 Fexxvi H 1s 2 S 1/2 2p 2 P 3/2 (Lyα)<br />

56


Table 3.8: Some important iron lines<br />

Ion E (keV) transition remark<br />

Fexxvi 6.97 2p–1s Lyα<br />

Fexxvi 8.21 3p–1s Lyβ<br />

Fexxvi 1.29 3p–2s Hα<br />

Fexxv 6.70 2p–1s 1 P 1 , resonance line<br />

Fexxv 6.67 2p–1s 3 P 2,1 , intercombination line<br />

Fexxv 6.63 2s–1s 3 S 1 , forbidden line<br />

Fexxv 7.90 3p–1s 1 P 1<br />

Fexxv 1.22 3p–2s 1 P 1<br />

Feii 6.40 2p–1s Kα (from ionisation of Fei)<br />

Feii 7.06 3p–1s ”Kβ” (from ionisation of Fei)<br />

Feii 0.79 3p–2s ”Lβ” (from ionisation of Fei)<br />

57


3.10 Absorption of X-<strong>ray</strong> <strong>radiation</strong><br />

3.10.1 Continuum versus line absorption<br />

X-<strong>ray</strong>s emitted by cosmic sources do not travel unattenuated to a distant observer.<br />

This is because intervening matter in the line of sight absorbs a part of the X-<br />

<strong>ray</strong>s. With low-resolution instruments, absorption can be studied only through<br />

the measurement of broad-band flux depressions caused by continuum absorption.<br />

However, at high spectral resolution also absorption lines can be studied, and in<br />

fact absorption lines offer more sensitive tools to detect weak intervening absorption<br />

systems. We illustrate this in Fig. 3.27. At an Oviii column density of 10 21 m −2 , the<br />

absorption edge has an optical depth of 1 %; for the same column density, the core<br />

of the Lyα line is already saturated and even for 100 times lower column density,<br />

the line core still has an optical depth of 5 %.<br />

Figure 3.27: Continuum (left) and Lyα (right) absorption spectrum for a layer<br />

consisting of Oviii ions with column densities as indicated.<br />

3.10.2 Continuum absorption<br />

Continuum absorption can be calculated simply from the photoionisation cross sections,<br />

that we discussed already in Sect. 3.6.7. The total continuum opacity τ cont<br />

can be written as<br />

τ cont (E) ≡ N H σ cont (E) = ∑ N i σ i (E) (3.81)<br />

i<br />

i.e, by averaging over the various ions i with column density N i . Accordingly,<br />

the continuum transmission T(E) of such a clump of matter can be written as<br />

T(E) = exp (−τ cont (E)). For a worked out example see also Sect. 3.10.6.<br />

3.10.3 Line absorption<br />

When light from a background source shines through a clump of matter, a part of<br />

the <strong>radiation</strong> can be absorbed. We discussed already the continuum absorption.<br />

The transmission in a spectral line at wavelength λ is given by<br />

T(λ) = e −τ(λ) (3.82)<br />

58


with<br />

τ(λ) = τ 0 ϕ(λ) (3.83)<br />

where ϕ(λ) is the line profile and τ 0 is the opacity at the line centre λ 0 , given by:<br />

τ 0 = αhλfN i<br />

2 √ 2πm e σ v<br />

. (3.84)<br />

Apart from the fine structure constant α and Planck’s constant h, the optical<br />

depth also depends on the properties of the absorber, namely the ionic column<br />

density N i and the velocity dispersion σ v . Furthermore, it depends on the oscillator<br />

strength f which is a dimensionless quantity that is different for each transition and<br />

is of order unity for the strongest transitions.<br />

In the simplest approximation, a Gaussian profile ϕ(λ) = exp −(λ−λ 0 ) 2 /b 2 ) can<br />

be adopted, corresponding to pure Doppler broadening for a thermal plasma. Here<br />

b = √ 2σ with σ the normal Gaussian root-mean-square width. The full width at<br />

half maximum of this profile is given by √ ln 256σ or approximately 2.35σ. One may<br />

even include a turbulent velocity σ t into the velocity width σ v , such that<br />

σ 2 v = σ 2 t + kT/m i (3.85)<br />

with m i the mass of the ion (we have tacitly changed here from wavelength to<br />

velocity units through the scaling ∆λ/λ 0 = ∆v/c).<br />

The equivalent width W of the line is calculated from<br />

W = λσ c<br />

∫ ∞<br />

[1 − exp (−τ 0 e −y2 /2 )]dy. (3.86)<br />

−∞<br />

For the simple case that τ 0 ≪ 1, the integral can be evaluated as<br />

τ 0 ≪ 1 :<br />

W = αhλ2 fN i<br />

2m e c<br />

= 1 2 α hν<br />

m e c 2 fλ2 N i . (3.87)<br />

A Gaussian line profile is only a good approximation when the Doppler width<br />

of the line is larger than the natural width of the line. The natural line profile for<br />

an absorption line is a Lorentz profile ϕ(λ) = 1/(1 + x 2 ) with x = 4π∆ν/A. Here<br />

∆ν is the frequency difference ν − ν 0 and A is the total transition probability from<br />

the upper level downwards to any level, including all radiative and Auger decay<br />

channels.<br />

Convolving the intrinsic Lorentz profile with the Gaussian Doppler profile due<br />

to the thermal or turbulent motion of the plasma, gives the well-known Voigt line<br />

profile<br />

ϕ = H(a, y) (3.88)<br />

where<br />

a = Aλ/4πb (3.89)<br />

and y = c∆λ/bλ. The dimensionless parameter a (not to be confused with the a in<br />

Eqn. 3.27) represents the relative importance of the Lorentzian term (∼ A) to the<br />

Gaussian term (∼ b). It should be noted here that formally for a > 0 the parameter<br />

τ 0 defined by (3.84) is not exactly the optical depth at line centre, but as long as<br />

a ≪ 1 it is a fair approximation. For a ≫ 1 it can be shown that H(a, 0) → 1/a √ π.<br />

59


3.10.4 Some important X-<strong>ray</strong> absorption lines<br />

There are several types of absorption lines. The most well-known are the normal<br />

strong resonance lines, which involve electrons from the outermost atomic shells.<br />

Examples are the 1s–2p line of Ovii at 21.60 Å, the Oviii Lyα doublet at 18.97 Å<br />

and the well known 2s–2p doublet of Ovi at 1032 and 1038 Å. See Richter et al.<br />

(2008) for an extensive discussion on these absorption lines in the WHIM.<br />

The other class are the inner-shell absorption lines. In this case the excited<br />

atom is often in an unstable state, with a large probability for the Auger process<br />

(Sect. 3.6.4). As a result, the parameter a entering the Voigt profile (Eqn. 3.89) is<br />

large and therefore the lines can have strong damping wings.<br />

In some cases the lines are isolated and well resolved, like the Ovii and Oviii<br />

1s–2p lines mentioned above. However, in other cases the lines may be strongly<br />

blended, like for the higher principal quantum number transitions of any ion in case<br />

of high column densities. Another important group of strongly blended lines are the<br />

inner-shell transitions of heavier elements like iron. They form so-called unresolved<br />

transition ar<strong>ray</strong>s (UTAs); the individual lines are no longer recognisable. The first<br />

detection of these UTAs in astrophysical sources was in a quasar (Sako et al. 2001).<br />

In Table 3.9 we list the 70 most important X-<strong>ray</strong> absorption lines for λ < 100 Å.<br />

The importance was determined by calculating the maximum ratio W/λ that these<br />

lines reach for any given temperature.<br />

Note the dominance of lines from all ionisation stages of oxygen in the 17–24 Å<br />

band. Furthermore, the Feix and Fexvii lines are the strongest iron features. These<br />

lines are weaker than the oxygen lines because of the lower abundance of iron; they<br />

are stronger than their neighbouring iron ions because of somewhat higher oscillator<br />

strengths and a relatively higher ion fraction (cf. Fig. 3.23).<br />

Note that the strength of these lines can depend on several other parameters,<br />

like turbulent broadening, deviations from CIE, saturation for higher column densities,<br />

etc., so some care should be taken when the strongest line for other physical<br />

conditions are sought.<br />

3.10.5 Curve of growth<br />

In most practical situations, X-<strong>ray</strong> absorption lines are unresolved or poorly resolved.<br />

As a consequence, the physical information about the source can only be retrieved<br />

from basic parameters such as the line centroid (for Doppler shifts) and equivalent<br />

width W (as a measure of the line strength).<br />

As outlined in Sect. 3.10.3, the equivalent width W of a spectral line from an<br />

ion i depends, apart from atomic parameters, only on the column density N i and<br />

the velocity broadening σ v . A useful diagram is the so-called curve of growth. Here<br />

we present it in the form of a curve giving W versus N i for a fixed value of σ v .<br />

The curve of growth has three regimes. For low column densities, the optical<br />

depth at line centre τ 0 is small, and W increases linearly with N i , independent of σ v<br />

(Eqn. 3.87). For increasing column density, τ 0 starts to exceed unity and then W<br />

increases only logarithmically with N i . Where this happens exactly, depends on σ v .<br />

In this intermediate logarithmic branch W depends both on σ v and N i , and hence by<br />

comparing the measured W with other lines from the same ion, both parameters can<br />

be estimated. At even higher column densities, the effective opacity increases faster<br />

(proportional to √ N i ), because of the Lorentzian line wings. Where this happens<br />

60


Table 3.9: The most important X-<strong>ray</strong> absorption lines with λ < 100 Å. Calculations<br />

are done for a plasma in CIE with proto-solar abundances (Lodders 2003) and only<br />

thermal broadening. The calculations are done for a hydrogen column density of<br />

10 24 m −2 , and we list equivalent widths for the temperature T max (in K) where it<br />

reaches a maximum. Equivalent widths are calculated for the listed line only, not<br />

taking into account blending by other lines. For saturated, overlapping lines, like<br />

the Oviii Lyα doublet near 18.97 Å, the combined equivalent width can be smaller<br />

than the sum of the individual equivalent widths.<br />

λ ion W max log λ ion W max log<br />

(Å) (mÅ) T max (Å) (mÅ) T max<br />

9.169 Mgxi 1.7 6.52 31.287 N i 5.5 4.04<br />

13.447 Neix 4.6 6.38 33.426 C v 7.0 5.76<br />

13.814 Nevii 2.2 5.72 33.734 C vi 12.9 6.05<br />

15.014 Fexvii 5.4 6.67 33.740 C vi 10.3 6.03<br />

15.265 Fexvii 2.2 6.63 34.973 C v 10.3 5.82<br />

15.316 Fexv 2.3 6.32 39.240 S xi 6.5 6.25<br />

16.006 O viii 2.6 6.37 40.268 C v 17.0 5.89<br />

16.510 Feix 5.6 5.81 40.940 C iv 7.2 5.02<br />

16.773 Feix 3.0 5.81 41.420 C iv 14.9 5.03<br />

17.396 O vii 2.8 6.06 42.543 S x 6.3 6.14<br />

17.768 O vii 4.4 6.11 50.524 Six 11.0 6.14<br />

18.629 O vii 6.9 6.19 51.807 S vii 7.6 5.68<br />

18.967 O viii 8.7 6.41 55.305 Siix 14.0 6.06<br />

18.973 O viii 6.5 6.39 60.161 S vii 12.2 5.71<br />

19.924 O v 3.2 5.41 61.019 Siviii 11.9 5.92<br />

21.602 O vii 12.1 6.27 61.070 Siviii 13.2 5.93<br />

22.006 O vi 5.1 5.48 62.751 Mgix 11.1 5.99<br />

22.008 O vi 4.1 5.48 68.148 Sivii 10.1 5.78<br />

22.370 O v 5.4 5.41 69.664 Sivii 10.9 5.78<br />

22.571 O iv 3.6 5.22 70.027 Sivii 11.3 5.78<br />

22.739 O iv 5.1 5.24 73.123 Sivii 10.7 5.78<br />

22.741 O iv 7.2 5.23 74.858 Mgviii 17.1 5.93<br />

22.777 O iv 13.8 5.22 80.449 Sivi 13.0 5.62<br />

22.978 O iii 9.4 4.95 80.577 Sivi 12.7 5.61<br />

23.049 O iii 7.1 4.96 82.430 Feix 14.0 5.91<br />

23.109 O iii 15.1 4.95 83.128 Sivi 14.2 5.63<br />

23.350 O ii 7.5 4.48 83.511 Mgvii 14.2 5.82<br />

23.351 O ii 12.6 4.48 83.910 Mgvii 19.8 5.85<br />

23.352 O ii 16.5 4.48 88.079 Neviii 15.2 5.80<br />

23.510 O i 7.2 4.00 95.385 Mgvi 15.5 5.67<br />

23.511 O i 15.0 4.00 95.421 Mgvi 18.7 5.69<br />

24.779 N vii 4.6 6.20 95.483 Mgvi 20.5 5.70<br />

24.900 N vi 4.0 5.90 96.440 Siv 16.8 5.33<br />

28.465 C vi 4.6 6.01 97.495 Nevii 22.8 5.74<br />

28.787 N vi 9.8 6.01 98.131 Nevi 16.7 5.65<br />

depends on the Voigt parameter a (Eqn. 3.89). In this range of densities, W does<br />

not depend any more on σ v .<br />

In Fig. 3.28 we show a few characteristic examples. The examples of Oi and<br />

Ovi illustrate the higher W for UV lines as compared to X-<strong>ray</strong> lines (cf. Eqn. 3.87)<br />

as well as the fact that inner shell transitions (in this case the X-<strong>ray</strong> lines) have<br />

61


Figure 3.28: Equivalent width versus column density for a few selected oxygen<br />

absorption lines. The curves have been calculated for Gaussian velocity dispersions<br />

σ = b/ √ 2 of 10, 50 and 250 kms −1 , from bottom to top for each line. Different<br />

spectral lines are indicated with different line styles.<br />

larger a-values and hence reach sooner the square root branch. The example of<br />

Ovii shows how different lines from the same ion have different equivalent width<br />

ratios depending on σ v , hence offer an opportunity to determine that parameter.<br />

The last frame shows the (non)similarties for similar transitions in different ions.<br />

The innershell nature of the transition in Ovi yields a higher a value and hence an<br />

earlier onset of the square root branch.<br />

3.10.6 Galactic foreground absorption<br />

All <strong>radiation</strong> from X-<strong>ray</strong> sources external to our own Galaxy has to pass through<br />

the interstellar medium of our Galaxy, and the intensity is reduced by a factor of<br />

e −τ(E) with the optical depth τ(E) = ∑ i σ i(E) ∫ n i (l)dl, with the summation over<br />

all relevant ions i and the integration over the line of sight dl. The absorption<br />

cross section σ i (E) is often taken to be simply the (continuum) photoionisation<br />

cross section, but for high-resolution spectra it is necessary to include also the line<br />

opacity, and for very large column densities also other processes such as Compton<br />

ionisation or Thomson scattering.<br />

For a cool, neutral plasma with cosmic abundances one often writes<br />

τ = σ eff (E)N H (3.90)<br />

62


where the hydrogen column density N H ≡ ∫ n H dx. In σ eff (E) all contributions to<br />

the absorption from all elements as well as their abundances are taken into account.<br />

Figure 3.29: Left panel: Neutral interstellar absorption cross section per hydrogen<br />

atom, scaled with E 3 . The most important edges with associated absorption lines<br />

are indicated, as well as the onset of Thompson scattering around 30 keV. Right<br />

panel: Contribution of the various elements to the total absorption cross section as<br />

a function of energy, for solar abundances.<br />

The relative contribution of the elements is also made clear in Fig. 3.29. Below<br />

0.28 keV (the carbon edge) hydrogen and helium dominate, while above 0.5 keV in<br />

particular oxygen is important. At the highest energies, above 7.1 keV, iron is the<br />

main opacity source.<br />

Figure 3.30: Left panel: Column density for which the optical depth becomes unity.<br />

Right panel: Distance for which τ = 1 for three characteristic densities.<br />

Yet another way to represent these data is to plot the column density or distance<br />

for which the optical depth becomes unity (Fig. 3.30). This figure shows that in<br />

particular at the lowest energies X-<strong>ray</strong>s are most strongly absorbed. The visibility<br />

in that region is thus very limited. Below 0.2 keV it is extremely hard to look outside<br />

our Milky Way.<br />

63


exercise 3.12. The hydrogen column density to the Galactic pole is about 10 24 m −2 .<br />

How large is the transmission for photons of 0.1, 0.2, 1 and 10 keV?<br />

Complexity of the ISM<br />

Figure 3.31: Left panel: Simulated 100 ks absorption spectrum as observed with the<br />

Explorer of Diffuse Emission and Gamma-<strong>ray</strong> Burst Explosions (EDGE), a mission<br />

proposed for ESA’s Cosmic Vision program. The parameters of the simulated source<br />

are similar to those of 4U 1820−303 (Yao et al. 2006). The plot shows the residuals<br />

of the simulated spectrum if the absorption lines in the model are ignored. Several<br />

characteristic absorption features of both neutral and ionised gas are indicated.<br />

Right panel: Simulated spectrum for the X-<strong>ray</strong> binary 4U 1820−303 for 100 ks with<br />

the WFS instrument of EDGE. The simulation was done for all absorbing oxygen in<br />

its pure atomic state, the models plotted with different line styles show cases where<br />

half of the oxygen is bound in CO, water ice or olivine. Note the effective shift of<br />

the absorption edge and the different fine structure near the edge. All of this is well<br />

resolved by EDGE, allowing a determination of the molecular composition of dust<br />

in the line of sight towards this source. The absorption line at 0.574 is due to highly<br />

ionised Ovii.<br />

The interstellar medium is by no means a homogeneous, neutral, atomic gas. In<br />

fact, it is a collection of regions all with different physical state and composition.<br />

This affects its X-<strong>ray</strong> opacity. We briefly discuss here some of the most important<br />

features.<br />

The ISM contains cold gas (< 50 K), warm, neutral or lowly ionised gas (6000<br />

− 10000 K) as well as hotter gas (a few million K). We have seen before that for<br />

ions the absorption edges shift to higher energies for higher ionisation. Thus, with<br />

instruments of sufficient spectral resolution, the degree of ionisation of the ISM can<br />

be deduced from the relative intensities of the absorption edges. Cases with such<br />

high column densities are rare, however (and only occur in some AGN outflows), and<br />

for the bulk of the ISM the column density of the ionised ISM is low enough that only<br />

the narrow absorption lines are visible (see Fig. 3.27). These lines are only visible<br />

when high spectral resolution is used (see Fig. 3.31). It is important to recognise<br />

these lines, as they should not be confused with absorption line from within the<br />

64


source itself. Fortunately, with high spectral resolution the cosmologically redshifted<br />

absorption lines from the X-<strong>ray</strong> source are separated well from the foreground hot<br />

ISM absorption lines. Only for lines from the Local Group this is not possible, and<br />

the situation is here complicated as the expected temperature range for the diffuse<br />

gas within the Local Group is similar to the temperature of the hot ISM.<br />

Another ISM component is dust. A significant fraction of some atoms can be<br />

bound in dust grains with varying sizes, as shown below (from Wilms et al. 2000):<br />

H He C N O Ne Mg Si S Ar Ca Fe Ni<br />

0 0 0.5 0 0.4 0 0.8 0.9 0.4 0 0.997 0.7 0.96<br />

The numbers represent the fraction of the atoms that are bound in dust grains.<br />

Noble gases like Ne and Ar are chemically inert hence are generally not bound in<br />

dust grains, but other elements like Ca exist predominantly in dust grains. Dust has<br />

significantly different X-<strong>ray</strong> properties compared to gas or hot plasma. First, due to<br />

the chemical binding, energy levels are broadened significantly or even absent. For<br />

example, for oxygen in most bound forms (like H 2 O) the remaining two vacancies in<br />

the 2p shell are effectively filled by the two electrons from the other bound atom(s)<br />

in the molecule. Therefore, the strong 1s–2p absorption line at 23.51 Å (527 eV)<br />

is not allowed when the oxygen is bound in dust or molecules, because there is<br />

no vacancy in the 2p shell. Transitions to higher shells such as the 3p shell are<br />

possible, however, but these are blurred significantly and often shifted due to the<br />

interactions in the molecule. Each constituent has its own fine structure near the<br />

K-edge (Fig. 3.31b). This fine structure offers therefore the opportunity to study<br />

the (true) chemical composition of the dust, but it should be said that the details of<br />

the edges in different important compounds are not always (accurately) known, and<br />

can differ depending on the state: for example water, crystalline and amorphous ice<br />

all have different characteristics. On the other hand, the large scale edge structure,<br />

in particular when observed at low spectral resolution, is not so much affected. For<br />

sufficiently high column densities of dust, self-shielding within the grains should be<br />

taken into account, and this reduces the average opacity per atom.<br />

Finally, we mention here that dust also causes scattering of X-<strong>ray</strong>s. This is in<br />

particular important for higher column densities. For example, for the Crab nebula<br />

(N H = 3.2×10 25 m −2 ), at an energy of 1 keV about 10 % of all photons are scattered<br />

in a halo, of which the widest tails have been observed out to a radius of at least<br />

half a degree; the scattered fraction increases with increasing wavelength.<br />

3.10.7 Absorption from ionised gas<br />

The absorption properties become different when the plasma is ionised. Fig. 3.32<br />

gives the effective cross section for plasmas in CIE.<br />

For 10 4 K, σ is essentially at the neutral limit. For 2 × 10 4 K hydrogen is fully<br />

ionised, and therefore σ for energies of a few hundred eV is decreasing; at 10 5 K also<br />

He is ionised and σ is in particular small below the C-edge. Around the O-edge (the<br />

region between 0.5–1 keV) we see that by increasing the ionisation of oxygen the<br />

edge shifts from 533 eV (Oi) to 870 eV (Oviii). Below the O-edge σ becomes very<br />

small for increasing temperatures, because the most important absorbers (H, He)<br />

are fully ionised. But unless oxygen is fully ionised (3 × 10 6 K) σ above the oxygen<br />

K-edge is decreasing only slowly. Because of this the contrast around the K-edge of<br />

65


Figure 3.32: The optical depth per hydrogen atom times E 3 for a plasma in CIE.<br />

Figure 3.33: The optical depth per hydrogen atom times E 3 for a plasma in PIE.<br />

oxygen increases for increasing T. Similar features are seen around the K-edge of<br />

Si (∼ 2 keV), S (∼ 2.5 keV) and in particular Fe (7–8.5 keV) between 10 5 K and<br />

3 × 10 8 K.<br />

Therefore we see that for T > 10 7 K the plasma is almost transparent below<br />

2 keV. between 10 7 and 10 8 K, σ between 2-6 keV is an order of magnitude smaller<br />

than for neutral plasma’s, while above the K-edge of iron the effective cross section<br />

hardly changes, unless T > 10 8 K.<br />

For plasmas in photoionisation one finds qualitatively a similar behaviour (Fig. 3.33).<br />

The parameter Ξ is the so-called ionisation parameter:<br />

Ξ ≡ L/4πcr 2 n H kT. (3.91)<br />

Essentially, it represents the ratio between <strong>radiation</strong> pressure and gas pressure. In<br />

general, Ξ depends on the precise form of the incoming X-<strong>ray</strong> spectrum. For a<br />

power law spectrum with energy index 1, or photon index 2 (dF/dE ∼ E −1 ) the<br />

equilibrium solution for the temperature as a function of Ξ is shown in Fig. 3.34.<br />

66


Figure 3.35: Stability for the ionisation<br />

equilibrium.<br />

Figure 3.34: Equilibrium temperature<br />

versus ionisation parameter.<br />

Arrows indicate the points<br />

that are shown in Fig. 3.33. From<br />

Krolik & Kallman 1984.<br />

Note that in general for a single value of Ξ more temperature solutions are possible<br />

(in particular for Ξ close to 10). Only those solutions where dT/dΞ > 0 are stable:<br />

in that case a small increase in temperature brings the material to a region where<br />

cooling is stronger than heating and hence the temperature will decrease again (see<br />

Fig. 3.35). In the opposite case temperature variations will become stronger, hence<br />

instability.<br />

We see that for Ξ < 0.03 the absorption properties are almost identical to those<br />

for neutral matter. Further, there are significant differences with the CIE case: in<br />

photoionisation equilibrium a temperature of 10 5 K is already sufficient to reduce<br />

σ eff to the same level as for CIE at 10 6 K. Plasmas in photoionisation equilibrium<br />

therefore are much more ”transparent” than CIE plasmas.<br />

67


3.11 The He-like triplets<br />

Now you have seen most of the important <strong>radiation</strong> mechanisms, it is time to have<br />

a closer look to the He-like triplets. This triplet has very important applications,<br />

and therefore is important to understand.<br />

First, we speak about a triplet, but in fact there are four spectral lines involved.<br />

Two of these lines often have very similar wavelengths, hence are sometimes considered<br />

to be one spectral line.<br />

Figure 3.36: Simplified level scheme for helium-like ions. w (or r), x; y (or i), and<br />

z (or f): resonance, intercombination, and forbidden lines, respectively. Full upward<br />

arrows: collisional excitation transitions, broken arrows: radiative transitions<br />

(including photo-excitation from n = 2 3 S 1 to n = 2 3 P 0,1,2 levels, and 2-photon continuum<br />

from n = 2 1 S 0 to the ground level), and thick skew arrows: recombination<br />

(radiative and dielectronic) plus cascade processes. From Porquet et al. 2001.<br />

These lines correspond to transitions between the n = 2 shell and the n = 1<br />

ground-state shell (see Fig. 3.36). Gabriel & Jordan (1969) were the first to show<br />

that the line ratios can be used to estimate temperature and electron density:<br />

R(n e ) =<br />

z<br />

x + y<br />

(also designated as f/i), (3.92)<br />

G(T e ) = z + x + y (also designated as (f + i)/r). (3.93)<br />

w<br />

In the low-density limit, all n = 2 states are populated directly, or via upperlevel<br />

radiative cascades by electron impact from the He-like ground state and/or<br />

by (radiative and dielectronic) recombination of H-like ions. These states decay<br />

radiatively directly or by cascades to the ground level. The relative intensities of<br />

the three intense lines are then independent of density. As n e increases from the<br />

low-density limit, some of these states (1s2s 3 S 1 and 1 S 0 ) are depleted by collisions<br />

to the nearby states where n crit C ∼ A, with C being the collisional coefficient rate,<br />

68


Table 3.10: Wavelengths in Å of the triplet lines for some elements<br />

Ion resonance intercombination intercombination forbidden<br />

w x y z<br />

Cv 40.268 40.728 40.731 41.472<br />

Nvi 28.787 29.082 29.084 29.534<br />

Ovii 21.602 21.804 21.807 22.101<br />

Neix 13.447 13.550 13.553 13.699<br />

Mgxi 9.1685 9.2279 9.2310 9.3141<br />

Sixiii 6.6477 6.6848 6.6879 6.7400<br />

Fexxv 1.8505 1.8554 1.8595 1.8682<br />

A being the radiative transition probability from n = 2 to n = 1 (ground state),<br />

and n crit being the critical density. Collisional excitation depopulates first the 1s2s<br />

3 S 1 level (upper level of the forbidden line) to the 1s2p 3 P 0,1,2 levels (upper levels<br />

of the intercombination lines). The intensity of the forbidden line decreases while<br />

those of the intercombination lines increase, hence implying a reduction of the ratio<br />

R (according to Eq. 3.92), over approximately two or three decades of density. For<br />

much higher densities, 1s2s 1 S 0 is also depopulated to 1s2p 1 P 1 , and the resonance<br />

line becomes sensitive to the density (this has been nicely illustrated by Gabriel &<br />

Jordan (1972) in their Fig. 4.6.9).<br />

As the critical density for each ion is different, using more than one triplet allows<br />

one to cover a broad range of densities.<br />

However, there are other ways to modify the triplet ratio. If the source has a<br />

strong UV <strong>radiation</strong> field, the upper level of the forbidden line can be photo-excited<br />

to the upper level of the intercombination lines, and therefore a strong UV field will<br />

also decrease the R ratio.<br />

Further, we note that the triplet is also an excellent indicator of the ionisation<br />

balance. For collisionally ionised plasmas, the resonance line has a similar intensity<br />

as the sum of the forbidden and intercombination lines. However, for a photoionised<br />

plasma, in general these last two lines contain most power. This is because<br />

recombination is proportional to the statistical weight of the level towards which<br />

the recombination occurs. While the upper level of the resonance line (a singlet)<br />

has w = 3, while that of the intercombination line has w = 9 and the forbidden<br />

line w = 3. Hence, as a rough approximation G = (9 + 3)/3 = 4 for a photoionised<br />

plasma.<br />

Finally, because the oscillator strength for the intercombination and forbidden<br />

line are much smaller than for the resonance line, if one observes the triplet in absorption,<br />

one usually only sees the resonance line. Only for high nuclear charge (like<br />

iron, Z = 26) the intercombination line also starts contributing to the absorption.<br />

69

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!