Matter and Radiation at Extremes, 2020, 5 (6): 064201, Published Online: Nov. 24, 2020   

Dielectronic recombination in non-LTE plasmas

Author Affiliations
1 Sorbonne University, Faculty of Science and Engineering, UMR 7605, Case 128, 4 Place Jussieu, F-75252 Paris Cedex 05, France
2 LULI, Ecole Polytechnique, CNRS-CEA, Physique Atomique dans les Plasmas Denses (PAPD), Route de Saclay, F-91128 Palaiseau Cedex, France
3 Moscow Institute of Physics and Technology MIPT (National Research University), Dolgoprudnyi 141700, Russia
4 National Research Nuclear University—MEPhI, Department of Plasma Physics, Moscow 115409, Russia
5 National Research Center “Kurchatov Institute”, Moscow, Russia
6 P. N. Lebedev Physical Institute of the Russian Academy of Sciences, Moscow 119991, Russia
Abstract
Novel phenomena and methods related to dielectronic capture and dielectronic recombination are studied for non-local thermodynamic equilibrium (LTE) plasmas and for applications to non-LTE ionization balance. It is demonstrated that multichannel autoionization and radiative decay strongly suppress higher-order contributions to the total dielectronic recombination rates, which are overestimated by standard approaches by orders of magnitude. Excited-state coupling of dielectronic capture is shown to be much more important than ground-state contributions, and electron collisional excitation is also identified as a mechanism driving effective dielectronic recombination. A theoretical description of the effect of angular-momentum-changing collisions on dielectronic recombination is developed from an atomic kinetic point of view and is visualized with a simple analytical model. The perturbation of the autoionizing states due to electric fields is discussed with respect to ionization potential depression and perturbation of symmetry properties of autoionization matrix elements. The first steps in the development of statistical methods are presented and are realized in the framework of a local plasma frequency approach. Finally, the impact of collisional–radiative processes and atomic population kinetics on dielectronic recombination is critically discussed, and simple analytical formulas are presented.

1 I. INTRODUCTION

Atomic populations are of fundamental importance in a variety of areas in both pure and applied science. Examples include the equation of state in thermodynamics; absorption, emission, and scattering processes in matter; lasing; radiation transport; radiative cooling and energy loss; diagnostic and spectroscopic methods that employ the radiative properties of matter; astrophysics and planetary science; the physics of radiation sources; and fusion science and technology.1,2

In a plasma, several charge states usually exist simultaneously, and therefore the total radiation emission arises from excited states of different ionic charges. This indicates that not only are the populations of excited states and their excitation mechanisms relevant but so too is the ionization balance. In hot plasmas, while the excited-state populations are essentially driven by electron collisional excitation, the ionization balance depends strongly on ionization and recombination processes. As shown by Burgess3 in the context of an analysis of solar emission, electron collisional ionization and radiative recombination alone could not account for observations, and dielectronic recombination was proposed to explain a low level of ionization. In considering radiation balance, it is therefore important to take account of dielectronic recombination. Similarly, the concept of local thermodynamic equilibrium (LTE) depends not only on radiative and collisional rates (as in the traditional description), but on autoionization rates too.

Dielectronic recombination (DR) is the capture of an electron with simultaneous excitation of an atomic core and subsequent radiative stabilization of the core. Within the framework of a simplified model, DR can be viewed as a sequence involving dielectronic capture of a continuum electron into the state nl and subsequent radiative stabilization. The first step is the capture of an electron into the state nl with simultaneous excitation of the atomic core from state α0 to state α:After dielectronic capture, autoionization or radiative stabilization can take place:The radiative stabilization involves the core and the captured electron (spectator electron):

Here, A+(Z−1)** denotes a doubly excited state, A+(Z−1)* a singly excited state, and γ the ground state of the ion in charge state Z − 1, i.e., A+(Z−1)(γ) = A+(Z−1)(ground). The radiation emission of the core [the relation (1.3)] is known as dielectronic satellite emission and has properties that are of great importance for plasma temperature diagnostics, as revealed in the pioneering work by Gabriel.4 The study of dielectronic satellite emission has become a major aspect of the characterization of a variety of complex phenomena in nonequilibrium plasmas: determination of electron density by angular-momentum-changing collisions2,5–7 and by the Stark effect;8,9 characterization and measurement of hot electrons and classification of related instabilities;10–12 determination of Auger electron heating phenomena in X-ray free-electron laser (XFEL) interaction with solid matter;13 radiation field analysis via hollow ion X-ray emission;14,15 collisional phenomena induced by laser-produced plasma jets;16,17 impurity transport and charge-exchange phenomena with the neutral background in magnetic fusion plasmas;18,19 relaxation phenomena in fluctuating plasmas;20 disappearance of resonance line emission followed by accumulation of dielectronic satellite emission;21–24 ionization potential depression analysis via two-dimensional maps of hollow-ion X-ray emission in XFEL–solid matter interaction.25 It should be noted here that for diagnostic purposes, radiation transport of radiative transitions originating from autoionizing states should be avoided, because re-emission of the photons is considerably reduced owing to the high autoionization rates.16

Comparison of the relations (1.1) and (1.4) shows that effective recombination occurs because an ion A+Z is finally transformed into an ion A+(Z−1). As was demonstrated in Refs. 3 and 26, DR can even be the most important recombination process. As a rule, the DR rate is high for ions with a complex core that exhibits transitions between levels without any change in principal quantum number, i.e., transitions with Δn = 0, such as 2 s→2p transitions in lithium-like and more complex ions. It is obvious that, particularly for ions with a complex core, the number of angular momentum coupling possibilities for the whole series of captured electrons is enormous. Therefore, the corresponding atomic structure calculations that combine all the single contributions (1.1)(1.4) to give the total DR of each ion are numerically prohibitive, and this has manifested itself in a continuing controversy regarding the calculation of the ionic fractions.27–29

Moreover, the total DR rate is not an issue of atomic structure calculations alone, but also requires consideration of populations that are strongly out of LTE and consideration of the plasma electric microfield. Therefore, the theoretical determination of the total DR in a nonequilibrium plasma cannot be done within the framework of atomic structure calculations alone.

The present paper is therefore devoted to a critical analysis of methods for the determination of the total DR rate and to develop a new framework to address the challenges faced. In this respect, a rather surprising element is discovered that can be illustrated via the above-mentioned Δn = 0 transitions, e.g., 2s → 2p transitions in lithium-like and more complex ions. The transition energy ΔE = ℏωcore for Δn = 0 and Z ≫ 1 is of the order of ZRy, while the ionization energy is of the order of Z2Ry ≫ ΔE. The 2s → 2p transition is the main channel for both the autoionization and radiative decay. Since the energy E of the incident recombining electron is in all cases smaller than the excitation energy, this implies the following inequality [using Ry=12me(cα)2 and E=12meve2]:where η is the Coulomb parameter (note that the standard Coulomb parameter determines the possibility of a quasiclassical consideration). The condition (1.5) indicates the validity of the quasiclassical regime in which, for example, the spectrum from the quantum theory of bremsstrahlung30 coincides with the classical spectrum.31 Thus, an important part of the overall DR processes is described well by a quasiclassical approach. This opens up new ways to treat important phenomena occurring in plasmas other than purely quantum mechanical atomic structure calculations (e.g., the multiconfiguration Dirac–Fock method).

On the other hand, DR related to Δn = 1 transitions may require quantum mechanical approaches. A typical example is provided by the core transitions 1s → 2p in H-like, He-like, etc. ions. Here, the 1s → 2p transition is the main radiation channel, while the autoionization channel depends strongly on the state nl of the captured electron: for large quantum numbers n, we encounter autoionization according to 2pnl → 2s + e, while for small quantum numbers, the autoionization channel 2pnl → 1s + e is dominant. For the case 1s → 2p, the transition energy ΔE = ℏωcore is of the order of Z2Ry and thus of the same order as the ionization energy. Therefore, the Coulomb parameter is about η ≈ 1 and quantum mechanical calculations might be required.

The paper considers the most recent developments that allow a unique description of the relevant phenomena in DR in nonequilibrium plasmas. For a review of standard methods in the theory of DR, the reader is referred to Refs. 26 and 32–34. In Sec. II, general DR formulas are derived in terms of atomic structure and collisional–radiative decay probabilities. In Sec. III, quantum mechanical multichannel methods are developed and then in Sec. IV, excited-state coupling driven by electron collisional excitation is considered. In Sec. V, the impact of angular-momentum-changing collisions on DR is considered and visualized with a simple analytical approach. Plasma electric field effects are discussed in Sec. VI, where the impacts of ionization potential depression and of symmetry perturbations of autoionizing matrix elements are described. Finally, Sec. VII develops a local plasma frequency statistical approach to calculate DR rates for very complex ions.

2 II. AUTOIONIZATION, DIELECTRONIC CAPTURE, AND DIELECTRONIC RECOMBINATION

Let us recall the essence of the DR process. An incident electron with energy E excites an ion core with excitation energy ΔE = ℏωcore. In this case, if the energy E is smaller than ΔE, the electron is finally captured by the ion into a state with energy EfRyZeff2/nf2 (Zeff is the effective charge of the excited ion core and nf is the principal quantum number of the captured electron) obeying the conditionThis capture results in a doubly excited state of the ion; i.e., the ion core electron is excited with energy ΔE, while the captured electron occupies a highly excited level of the ion. This state of the ion can decay in two possible ways:

by relaxation of the ion core electron into the initial ground state with simultaneous ejection of the captured electron from the ion: this process is known as autoionization [compare with the relation (1.2)];

by radiative decay of the ion core electron, resulting in its return to the initial state after the emission of a photon of energy ℏωℏωcore = ΔE, whereas the captured electron remains bound to the ion [compare with the relation (1.3)].

For illustration, Fig. 1 shows the relevant energy–level diagram for the He-like 2l2l′ autoionizing levels from which the so-called Lyα satellites originate. The energy of the 2l2l′ levels is given approximately by E2l2l2Zeff2Ry/4=Zeff2Ry/2 (in the H-like approximation), which is about half of the ionization energy of the H-like ground state Zn2Ry (where Zn is the nuclear charge). The series limit of the autoionizing levels 2lnl′ is the first excited state 2l. Radiative decay (dielectronic satellite emission) from the 2l2l′ levels populates the singly excited levels 1s2l1,3L, from which further radiative decays proceed (e.g., the resonance line W = 1s2p1P1 − 1s2 1S0 and the intercombination line Y = 1s2p3P1 − 1s2 1S0) that finally populate the ground state 1s2 1S0.

Fig. 1. Energy-level diagram of the He-like autoionizing levels 2l2l′ and their associated radiative decays, so-called Lyα satellites. After radiative decay, the singly excited states 1s2l1,3L are formed, from which further radiative decay proceeds (e.g., the resonance and intercombination lines W and Y, respectively). Also indicated are the Li-like autoionizing levels 1s2l2l′.

下载图片 查看所有图片

The chain of processes of dielectronic capture (1s + e → 2l2l′), radiative decay to singly excited levels (2l2l′ − 1s2l + ), and radiative decay to the ground state (1s2l1,3L → 1s2 1S0 + ′) is called dielectronic recombination (the DR channel) because an effective recombination has taken place from the H-like ground state 1s2S1/2 to the He-like ground state 1s2 1S0.

Thus, the DR process as well as the photorecombination (PR) process result in the capture of an incident electron and its simultaneous photon emission. The difference is that the photon is emitted by the ion core electron in the DR process rather than by the incident electron as in the PR process. Note that the relationship between the PR and DR processes is analogous to that between standard bremsstrahlung and polarization bremsstrahlung.35

The DR rate can be calculated from the autoionization rate of a given atomic state with the help of the principle of detailed balance. The first step is the application of the principle of detailed balance to dielectronic capture, i.e.,where njZ is the atomic population of the autoionizing state, ΓjkZ,Z+1 is the autoionization rate from state j to a state k with population nkZ+1, and 〈DCkj is the dielectronic capture rate from state k to the upper state j. To solve Eq. (2.2) for the dielectronic capture rate, we need to specify the populations. For this purpose, we consider a system in thermodynamic equilibrium. In this case, the populations njZ and nkZ+1 are linked via the Saha–Boltzmann equation, because states j and k belong to different ionic states Z and Z + 1, respectively, i.e.,where gjZ and gkZ+1 are the statistical weights of states j and k, ne is the electron density, me is the electron mass, and Te is the electron temperature. The energy difference ΔEkjZ+1,Z is related to the so-called dielectronic capture energy EkjDC by (see also Fig. 1)where EkjDC is the energy of the Auger electron if the autoionizing state j decays via autoionization to state k. Combining Eqs. (2.2)(2.4), we find the general expression for the dielectronic capture rate:or, in convenient units (with ΓjkZ,Z+1 in s−1, and EkjDC and Te in eV),[note that Eq. (2.6) assumes a Maxwellian electron energy distribution function with temperature Te]. If Pj,grZ is the probability that the autoionizing state j of charge state Z decays to the ground state of the same charge state, then the quantity Pj,grZDCkj is called the dielectronic recombination rate coefficient (with units cm3 s−1) into state k via the intermediate state j:

Note that it is important to underline that the assumption of thermodynamic equilibrium to eliminate the populations of Eq. (2.2) with the help of Eq. (2.3) is only a convenient method to apply the principle of detailed balance, but does not mean that relations derived with the help of the detailed balance relations are only valid under the assumption of thermodynamic equilibrium. In fact, the dielectronic capture rate [Eq. (2.6)] can also be derived from purely quantum mechanical relations (micro-reversibility), providing a dielectronic capture rate that is also valid for an arbitrary electron energy distribution function F(E):20or, in convenient units (with ΓjkZ,Z+1 in s−1, and F(E), EkjDC, and Te in eV),

3 III. TOTAL RATES OF DIELECTRONIC RECOMBINATION: LOW-DENSITY APPROXIMATION

3.1 A. General considerations

In general, the probability Pj,grZ from Eq. (2.7) is a function of density and temperature, i.e.,Pj,grZ=Pj,grZ(ne,Te).The probability function (3.1) has to be determined from numerical calculations of a multilevel multi-charge-state atomic population kinetics that explicitly involves all necessary autoionizing states as “active levels” (“active” here means that the populations of the autoionizing levels are calculated in a similar way to the ground and singly excited states in the population kinetics). If collisions are negligible compared with spontaneous radiative decay rates as well as with autoionization rates, then the probability Pj,grZ can be approximated by the so-called satellite branching factors BjiZ:3Pj,grZiBjiZ=iAjiZlAjlZ+kΓjkZ,Z+1.That is to say, Eq. (3.2) is the low-density approximation of the probability Pj,grZ. In this case, the DR rate is given by the following approximate expression:DRkjZ+1,ZiBjiZ.DCkjZ+1,Z=iAjiZlAjlZ+kΓjkZ,Z+1DCkjZ+1,Z.With the help of Eq. (2.5), Eq. (3.3) can be written as follows:DRkjZ+1,Z12gkZ+12π2me3/2exp(EkjDC/kTe)(kTe)3/2igjZΓjkZ,Z+1AjiZlAjlZ+kΓjkZ,Z+1.The term in braces {⋯} is the so-called dielectronic satellite intensity factorQk,jiZ+1,Z=gjZΓjkZ,Z+1AjiZlAjlZ+kΓjkZ,Z+1,which has been calculated by the pioneering work in Refs. 36 and 37 with unprecedented precision via the multiconfiguration Z-expansion method. Therefore, under the assumptions made in Eq. (3.2), the DR rate is given by the sum of the dielectronic satellite intensity factors. One can see that the essential quantities that appear in Eqs. (3.4) and (3.5) to calculate the DR rate in the low-density approximation are the dielectronic capture energy, statistical weights, and radiative and autoionizing decay rates. These quantities can nowadays routinely be generated from atomic structure calculations.

However, as one can see, even for the simplest configurations 2lnl′, the numerical calculations are rather cumbersome because very large quantum numbers nl′ have to be involved to achieve convergence for the DR rates. For large quantum numbers, however, convergence is difficult to achieve in purely quantum numerical atomic structure calculations. Moreover, to obtain the total DR rate from H-like to He-like ions, one needs to invoke all possible intermediate states j = 3lnl′, 4lnl′, 5lnl′, ….

One can easily understand that for more complex configurations, the number of autoionizing states that need to be involved rapidly becomes numerically prohibitive for purely quantum mechanical numerical calculations. For practical applications, it is therefore mandatory to invoke additional methods, such as the Burgess approach,3,32 Vainshtein’s simplified quantum mechanical dielectronic recombination model,26,38 a quasiclassical approach,31,39 or a statistical approach.40,41

3.2 B. The Burgess approximation

To illuminate the essence of the various approximations currently employed, we consider first the most general expression for the total dielectronic recombination rate DRtotZ+1,Z. For the total rate, all DR rates DRkjZ+1,Z=Pj,grZDCkjZ+1,Z have to be summed with respect to the initial state k and also with respect to the intermediate states j, i.e.,DRtotZ+1,Z=kjDRkjZ+1,Z=kjPj,grZDCkjZ+1,Z.Because the probability Pj,grZ is a function of density and temperature [see Eq. (3.1)], it is very difficult to obtain general and closed formulas for the DR rate coefficient. Only in the low-density approximation, where Eq. (3.3) holds, can general formulas for the DR rate coefficients be obtained.

One of the most widely used general approximate empirical formulas in the framework of the approximation (3.2) is the so-called Burgess formula,3 in which it is assumed that the nl spectator electron is not interacting with the core and can be treated in the hydrogenic approximation and that the capture cross section averaged over the resonances can be obtained with the aid of the correspondence principle by extrapolating below threshold the partial cross section for the core excitation α0α:DRkjZ+1,Z:=DZ+1,Z(α0α,nl).For the total DR rate, we haveDRtotZ+1,Z:=DZ+1,Z=α0αnl=0n1DZ+1,Z(α0α,nl).For the simplest example of autoionizing states 2l2l′ outlined in Fig. 1, α0 = 1s and α1 = 2p, i.e., the transition α0α corresponds to the Lyα transition in H-like ions. For these configurations, DR into the ground state is the most important transition; i.e., there exists a single state k = α0 = 1s. Therefore, α0 coincides with the atomic ground state, and the sum over α0 can be suppressed:DZ+1,Zαnl=0n1DZ+1,Z(α0α,nl).The DR rate coefficient DZ+1,Z(α0α, nl) can then be expressed via the following analytical empirical expression:3DZ+1,Z(α0α,nl)=4.8×1011fα0αBdβ3/2eβχd(cm3/s),withβ=(z+1)2RykTe,χd=χ1+0.015z3(z+1)2,χ=ΔE(α0α)(z+1)2Ry.Here, z is the so-called spectroscopic symbol of the doubly excited ion after recombination, z = ZnNbound + 1, where Nbound is the number of bound electrons and Zne is the nuclear charge. If the first resonance transition is a Δn = 0 transition, then the branching factor Bd is given by the following fitting formula (the so-called Burgess–Mertz formula):32Bd=zχz2+13.41/211+0.105(z+1)χ+0.015(z+1)2χ2.For Δn ≠ 0 the fitting function for the branching factor Bd is different:32Bd=zχz2+13.41/20.51+0.210(z+1)χ+0.030(z+1)2χ2.Let us recall the meaning of the branching factor Bd: after dielectronic capture, a doubly excited state is formed that can decay via autoionization or radiative decay. For DR, only the radiative decays contribute finally to recombination, since autoionization returns the autoionizing state to the original state.

According to Eq. (3.10), α0 is the ground state, and therefore fα0α is the electric dipole absorption oscillator strength for the resonance transition α0α with transition energy ΔE(α0α) in eV. As the absorption oscillator strength decreases rapidly with increasing principal quantum number of the upper level, it is usually sufficient to consider only the first two α-terms in the sum in Eq. (3.9), and we finally obtain the desired expression for the total DR coefficient:DZ+1,ZDZ+1,Z(α0α1)+DZ+1,Z(α0α2),withDZ+1,Z(α0α):=nl=0n1DZ+1,Z(α0α,nl).Let us consider DR into neutral helium as an example (note that a single * indicates a singly excited state, while ** corresponds to a multiply excited state):He1+(1s)+eHe0+**(nlnl)He0+(1s2).For this example, α0 = 1s, α1 = 2p, α2 = 3p, …. Therefore, fα0α1 corresponds to the electric dipole absorption oscillator strength of the resonance line, namely, the H-like Lyα line of singly ionized helium, while fα0α2 corresponds to the Lyβ line. The oscillator strengths are f1s→2p = 0.4164 and f1s→3p = 0.079 14, and their transition energies are ΔE(1s → 2p) = 40.81 eV and ΔE(1s → 3p) = 48.37 eV. The spectroscopic symbol is z = 1 and Δn ≠ 0 [therefore, Eq. (3.15) applies]. As one can see, higher-n oscillator strengths make almost negligible contributions to the total DR rate (note that this has to be distinguished from the fact that high-n spectator electrons can make quite important contributions). From Eqs. (3.11)(3.13), we obtain β = 4Ry/kTe, Bd(1s → 2p) = 0.0825, Bd(1s → 3p) = 0.0846, χd(1s → 2p) = 0.747, and χd(1s → 3p) = 0.886. For the rate coefficients at kTe = Ry (β = 4), we obtain DHe1+,He0 +(1s2p)=1.65×1012cm3/s, and DHe1+,He0 +(1s3p)=3.2×1013cm3/s. This confirms that the leading terms for DR are indeed given by Eq. (3.16).

3.3 C. Quantum mechanical multichannel approach

Comparison of the results from Eqs. (3.10)(3.15) with more precise quantum mechanical calculations carried out with Vainshtein’s ATOM code42,43 show that the χd values are in quite good agreement, whereas the Bd values differ strongly. For the resonance transitions 1s → 2p and 1s → 3p, the Bd values obtained from Vainshtein’s simplified quantum mechanical multi-channel (QMMC) approach38,44 for the above example of helium are significantly different, namely, Bd,ref(1s → 2p) = 0.155, Bd,ref(1s → 3p) = 0.0144, χd,ref(1s → 2p) = 0.744, and χd,ref(1s → 3p) = 0.888, giving DrefHe1+,He0 +(1s2p)=3.10×1012cm3/s and DrefHe1+,He0 +(1s3p)=5.46×1014cm3/s, i.e., the Burgess formula underestimates the Bd value for α0α = 1s → 2p by a factor of 2 and overestimates the Bd value for α0α = 1s → 3p by a factor of 6. These are general observations: the precision of Eqs. (3.10)(3.15) is very difficult to estimate: it might be about a factor of 2 for the strongest resonance transition, but it might also deviate by orders of magnitude.

Understanding the large overestimate of the Bd value for the transition 1s → 3p is of particular importance and is related to the fact that the Burgess formulas take into account only one autoionizing channel. For example, the 3lnl′-configurations (which are related to the transition α0α2 = 1s → 3p in the above example) autoionize not only to the ground state but to excited states too:3lnl1s+eAuger2l+eAuger.Complex numerical multiconfiguration Hartree–Fock calculations show23,45 that the autoionization rates to the excited states “2l” are even more important than to the ground state “1s.” This reduces considerably the branching factor for DR [the Bd factor in Eq. (3.10)]. In fact, as one can see from Eq. (3.15), very similar branching factors are given for the transitions α0α1 = 1s → 2p and α0α2 = 1s → 3p because only one autoionizing channel is taken into consideration.

It is very important for the practical application of DR rates in the modeling of ionization balance to explore in more detail the influence of the various multiple channels for Auger and radiative decay. Below, we perform QMMC calculations for DR in the simplified Vainshtein approach,38,42–44 and we fit the numerical results to an analytical expression in order to facilitate the application and allow direct comparison with the Burgess formulas:DZ+1,Z(α0α,nl)=108×m2l0+1Bdβ3/2eβχd(cm3/s),β=Z2RykTe,where Ry = 13.6057 eV, kTe is the electron temperature in eV, m is the number of equivalent electrons of state α0, Z is the charge of the ion where the core transition α0α takes place (e.g., for the 2lnl′ autoionizing states of He-like argon, the core transition is the 1s → 2p transition in H-like argon, Z = 18), and l0 is the corresponding orbital momentum of state α0. The physical meaning of the parameter χd is related to the fact that all contributions from the configurations with different spectator electrons nl have to be summed up to give the total DR rate with different energies [see Eq. (2.1)]. The parameter χd provides a fit to the numerical results to replace the sum of different energies in the best manner through an average energy parameter χdβ. Finally, the total sum is replaced by an average amplitude Bd to provide a simple analytical expression without the need for summation.

As a demonstration, let us consider the various mechanisms via a study of the DR related to the core transition 2s–4p. For example, the numerical calculations in the single-channel approximation for Be atoms give Bd(1- channel)(2s3p)=3.4×105, whereas Bd(1- channel)(2s4p)=1.6×105; i.e., 2s–4p transitions are only reduced by a factor of about two compared with 2s–3p transitions. Numerical calculations including multichannel decay provide an entirely different picture, with Bd(6- channel)(2s3p)=2.0×106, but Bd(6-channel)(2s4p)=3.5×107; i.e., the QMMC numerical calculations indicate that higher-order DR rates are strongly suppressed. This is a general observation that multichannel decay can reduce DR considerably and can even lead to a quite different interpretation of its importance.

Table I shows the numerical calculation for the DR Bd factors for single- and multichannel decay into Li-like ions for different orders and elements in comparison with the standard Burgess formula. One observes that the Burgess formula is in reasonable agreement with the numerical results for single-channel decay, although it differs by up to a factor of 3 in some cases. However, comparison with the numerical calculations using the QMMC approach reveals extremely large overestimates of DR obtained by the Burgess formula. In particular for light elements, the overestimation can be by as much as one or two orders of magnitude: for example, for the DR related to the autoionizing states 1s24lnl′ of Be, we have Bd(multichannel)(2s4p)=3.47×107 and Bd(Burgess)(2s4p)=1.10×105; i.e., the Burgess formula overestimates Bd by more than a factor of 30. It is therefore not recommended to calculate higher-order contributions to DR via the Burgess approach.

Table 1. Bd factors according to Eqs. (3.20) and (3.21) for DR into Li-like ions originating from the 1s2nlnl′ autoionizing levels, with Z = Zn − 2, m = 1, and l0 = 0. The numerical data show single- and multichannel approximations as well as the corresponding factors according to the Burgess approach (note that the different numerical coefficients and the oscillator strength in the original Burgess formula [Eq. (3.10)] compared with Eq. (3.20) have been included in the value for Bd(Burgess) to facilitate comparison of the different methods).

Element1s22lnl′: α0 = 1s22sα = 1s22p
Bd(1- channel)Bd(multichannel)Bd(Burgess)
Be8.09 × 10−51.34 × 10−4
C5.18 × 10−57.99 × 10−5
Mg1.34 × 10−51.94 × 10−5
Ar6.87 × 10−68.65 × 10−6
Fe4.02 × 10−64.88 × 10−6
Mo3.11 × 10−63.87 × 10−6
1s23lnl′: α0 = 1s22sα = 1s23p
Be3.44 × 10−51.97 × 10−62.88 × 10−5
C6.45 × 10−56.61 × 10−66.98 × 10−5
Mg6.43 × 10−52.57 × 10−56.96 × 10−5
Ar4.55 × 10−52.42 × 10−55.15 × 10−5
Fe2.61 × 10−51.57 × 10−53.54 × 10−5
Mo8.61 × 10−66.48 × 10−61.89 × 10−5
1s24lnl′: α0 = 1s22sα = 1s24p
Be1.60 × 10−53.47 × 10−71.10 × 10−5
C2.52 × 10−53.39 × 10−72.23 × 10−5
Mg2.06 × 10−51.30 × 10−61.87 × 10−5
Ar1.29 × 10−52.05 × 10−61.27 × 10−5
Fe6.54 × 10−62.00 × 10−68.01 × 10−6
Mo1.87 × 10−61.17 × 10−63.82 × 10−6

查看所有表

In Tables II–IV, we present the numerical results of the QMMC approach for H-, He-, and Li-like ions, which are the most important for K-shell spectroscopy. For ease of application, we have fitted all results to the simple analytical formula (3.20).

Table 2. Fitting coefficients according to Eqs. (3.20) and (3.21) for DR into H-like ions originating from the 2lnl′ and 3lnl′ autoionizing levels, with Z = Zn, m = 1, and l0 = 0. The numerical data include corrections for multiple decay channels (two channels for 2lnl′ and four channels for 3lnl′).

2lnl′: α0 = 1sα = 2p3lnl′: α0 = 1sα = 3p
ElementBdχdBdχd
He3.12 × 10−40.7445.48 × 10−60.888
Li3.72 × 10−40.7366.41 × 10−60.887
Be3.67 × 10−40.7276.53 × 10−60.885
B3.42 × 10−40.7186.47 × 10−60.883
C3.13 × 10−40.7096.32 × 10−60.881
N2.85 × 10−40.7006.31 × 10−60.879
O2.58 × 10−40.6915.92 × 10−60.877
F2.33 × 10−40.6825.70 × 10−60.874
Ne2.11 × 10−40.6735.48 × 10−60.872
Na1.90 × 10−40.6655.26 × 10−60.870
Mg1.72 × 10−40.6575.04 × 10−60.868
Al1.56 × 10−40.6494.84 × 10−60.866
Si1.41 × 10−40.6424.63 × 10−60.863
P1.27 × 10−40.6364.43 × 10−60.861
S1.15 × 10−40.6304.24 × 10−60.859
Cl1.05 × 10−40.6244.05 × 10−60.857
Ar9.50 × 10−50.6203.87 × 10−60.856
K8.61 × 10−50.6163.69 × 10−60.854
C7.82 × 10−50.6123.52 × 10−60.852
Sc7.09 × 10−50.6093.35 × 10−60.851
Ti6.45 × 10−50.6063.19 × 10−60.849
V5.85 × 10−50.6043.04 × 10−60.848
Cr5.33 × 10−50.6022.89 × 10−60.847
Mn4.85 × 10−50.6012.74 × 10−60.846
Fe4.42 × 10−50.5992.60 × 10−60.845
Co4.03 × 10−50.5982.47 × 10−60.844
Ni3.68 × 10−50.5982.34 × 10−60.843
Cu3.37 × 10−50.5972.22 × 10−60.842
Zn3.08 × 10−50.5972.10 × 10−60.842
Ga2.83 × 10−50.5961.99 × 10−60.842
Ge2.60 × 10−50.5961.88 × 10−60.841
As2.39 × 10−50.5961.78 × 10−60.841
Se2.20 × 10−50.5961.68 × 10−60.841
Br2.03 × 10−50.5961.59 × 10−60.841
Kr1.88 × 10−50.5961.50 × 10−60.841
Rb1.74 × 10−50.5971.42 × 10−60.841
Sr1.61 × 10−50.5971.34 × 10−60.842
Y1.50 × 10−50.5971.27 × 10−60.842
Zr1.39 × 10−50.5981.20 × 10−60.842
Nb1.30 × 10−50.5991.13 × 10−60.843
Mo1.21 × 10−50.5991.07 × 10−60.843

查看所有表

Table 3. Fitting coefficients according to Eqs. (3.20) and (3.21) for DR into He-like ions originating from the 1s2lnl′ and 1s3lnl′ autoionizing levels, with Z = Zn − 1, m = 2, and l0 = 0. The numerical data include corrections for multiple decay channels (two channels for 1s2lnl′ and four channels for 1s3lnl′).

1s2lnl′: α0 = 1s2α = 1s2p1s3lnl′: α0 = 1s2α = 1s3p
ElementBdχdBdχd
Li3.39 × 10−51.111.57 × 10−61.27
Be9.94 × 10−50.9612.12 × 10−61.14
B1.53 × 10−40.8912.51 × 10−61.07
C1.93 × 10−40.8482.98 × 10−61.03
N2.17 × 10−40.8183.40 × 10−61.00
O2.34 × 10−40.7953.92 × 10−60.983
F2.17 × 10−40.7754.23 × 10−60.967
Ne2.05 × 10−40.7574.50 × 10−60.956
Na1.88 × 10−40.7404.56 × 10−60.945
Mg1.72 × 10−40.7264.54 × 10−60.937
Al1.57 × 10−40.7134.47 × 10−60.929
Si1.43 × 10−40.7014.36 × 10−60.922
P1.30 × 10−40.6904.22 × 10−60.916
S1.18 × 10−40.6814.07 × 10−60.910
Cl1.07 × 10−40.6723.92 × 10−60.905
Ar9.72 × 10−50.6643.76 × 10−60.901
K8.83 × 10−50.6583.61 × 10−60.897
C8.02 × 10−50.6523.45 × 10−60.893
Sc7.28 × 10−50.6473.30 × 10−60.889
Ti6.62 × 10−50.6423.15 × 10−60.886
V6.02 × 10−50.6383.01 × 10−60.883
Cr5.47 × 10−50.6352.87 × 10−60.880
Mn4.98 × 10−50.6322.73 × 10−60.877
Fe4.54 × 10−50.6292.60 × 10−60.875
Co4.14 × 10−50.6272.47 × 10−60.873
Ni3.78 × 10−50.6252.35 × 10−60.871
Cu3.46 × 10−50.6232.23 × 10−60.869
Zn3.16 × 10−50.6222.11 × 10−60.868
Ga2.90 × 10−50.6202.00 × 10−60.867
Ge2.67 × 10−50.6191.90 × 10−60.865
As2.45 × 10−50.6191.80 × 10−60.864
Se2.26 × 10−50.6181.70 × 10−60.864
Br2.08 × 10−50.6171.61 × 10−60.863
Kr1.93 × 10−50.6171.52 × 10−60.862
Rb1.78 × 10−50.6161.44 × 10−60.862
Sr1.65 × 10−50.6161.36 × 10−60.861
Y1.53 × 10−50.6161.29 × 10−60.861
Zr1.43 × 10−50.6161.22 × 10−60.861
Nb1.33 × 10−50.6161.15 × 10−60.861
Mo1.24 × 10−50.6161.09 × 10−60.861

查看所有表

Table 4. Fitting coefficients according to Eqs. (3.20) and (3.21) for DR into Li-like ions originating from the 1s22lnl′ and 1s23lnl′ autoionizing levels, with Z = Zn − 2, m = 1, and l0 = 0. The numerical data include corrections for multiple decay channels (one channel for 1s22lnl′ and four channels for 1s23lnl′).

1s22lnl′: α0 = 1s22sα = 1s22p1s23lnl′: α0 = 1s22sα = 1s23p
ElementBdχdBdχd
Be8.09 × 10−50.057 11.97 × 10−60.197
B6.86 × 10−50.040 02.85 × 10−60.173
C5.18 × 10−50.030 66.61 × 10−60.161
N3.95 × 10−50.024 81.06 × 10−50.153
O3.09 × 10−50.020 71.47 × 10−50.149
F2.47 × 10−50.017 91.85 × 10−50.145
Ne2.02 × 10−50.015 62.17 × 10−50.142
Na1.69 × 10−50.013 92.41 × 10−50.140
Mg1.43 × 10−50.012 62.57 × 10−50.138
Al1.23 × 10−50.011 52.67 × 10−50.136
Si1.07 × 10−50.010 52.71 × 10−50.135
P9.43 × 10−60.009 812.69 × 10−50.133
S8.41 × 10−60.009 142.60 × 10−50.131
Cl7.57 × 10−60.008 582.53 × 10−50.130
Ar6.87 × 10−60.008 092.42 × 10−50.128
K6.25 × 10−60.007 722.31 × 10−50.127
C5.76 × 10−60.007 362.19 × 10−50.126
Sc5.35 × 10−60.007 042.09 × 10−50.124
Ti5.00 × 10−60.006 771.97 × 10−50.123
V4.67 × 10−60.006 581.86 × 10−50.122
Cr4.42 × 10−60.006 371.76 × 10−50.120
Mn4.20 × 10−60.006 201.66 × 10−50.119
Fe4.02 × 10−60.006 051.57 × 10−50.118
Co3.86 × 10−60.005 921.48 × 10−50.117
Ni3.72 × 10−60.005 811.40 × 10−50.116
Cu3.61 × 10−60.005 711.32 × 10−50.115
Zn3.51 × 10−60.005 641.25 × 10−50.114
Ga3.42 × 10−60.005 581.18 × 10−50.113
Ge3.35 × 10−60.005 531.11 × 10−50.112
As3.25 × 10−60.005 561.05 × 10−50.111
Se3.20 × 10−60.005 549.96 × 10−60.110
Br3.20 × 10−60.005 469.43 × 10−60.109
Kr3.17 × 10−60.005 468.92 × 10−60.108
Rb3.15 × 10−60.005 478.45 × 10−60.107
Sr3.13 × 10−60.005 488.01 × 10−60.106
Y3.12 × 10−60.005 517.59 × 10−60.105
Zr3.11 × 10−60.005 547.20 × 10−60.105
Nb3.11 × 10−60.005 586.83 × 10−60.104
Mo3.11 × 10−60.005 636.48 × 10−60.103

查看所有表

Table II presents the results of a numerical calculation of the total DR rate into H-like ions for the core transitions 1s–2p and 1s–3p for all elements from He (Z = 2) up to Mo (Z = 42) and the corresponding fitting parameters according to Eq. (3.20). It can be seen that for low-Z elements, the DR related to the core transition 1s–2p is dominant, whereas for large Z, the relative contribution of the DR with the core transition 1s–3p increases. The Burgess formula provides amplitudes Bd that are about a factor of 3 smaller than the present numerical calculations.

For the 3lnl′ states, the Burgess formula considerably overestimates the DR rate because it does not take into account multichannel radiative and Auger decays. This is of particular importance for low-Z elements. For example, for C, the single-channel approximation gives Bd = 6.75 × 10−5, whereas the four-channel approximation gives Bd = 6.32 × 10−6, i.e., a reduction by a factor of 10. Multichannel decay is much less important for higher Z: for example, for Fe, Bd = 5.13 × 10−6, whereas the four-channel approximation gives Bd = 2.60 × 10−6.

Table III presents numerical results for the total DR rate into He-like ions for the core transitions 1s–2p and 1s–3p for all elements from He (Z = 2) up to Mo (Z = 42) and the corresponding fitting parameters according to Eq. (3.20). It can be seen that for low-Z elements, the DR related to the core transition 1s–2p is dominant, whereas for large Z, the relative contribution of the DR with the core transition 1s–3p increases. The Burgess formula gives amplitudes Bd that are about a factor of 3 smaller than the present numerical calculations. For the 1s3lnl′ states, the Burgess formula considerably overestimates the DR rate because it does not take into account multichannel radiative and Auger decays. This is of particular importance for low-Z elements. For example, for C, the single-channel approximation gives Bd = 6.76 × 10−5, whereas the four-channel approximation gives Bd = 2.98 × 10−6, i.e., a reduction by a factor of 20. Multichannel decay is much less important for higher Z: for example, for Fe, Bd = 5.34 × 10−6, whereas the four-channel approximation gives Bd = 2.60 × 10−6.

Table IV presents numerical results for DR into Li-like ions related to the core transition 2s–2p, i.e., a Δn = 0 transition. Therefore, the fitting parameter χd is rather small and the associated exponential factor for DR does not vary much. In addition, the configurations 1s22lnl′ are only autoionizing for rather high principal quantum numbers. This is quite different to the DR related to the core transition 2s–3p, where the states are autoionizing for rather low quantum numbers nl and the temperature dependence is very different owing to an order-of-magnitude difference in the χd factor.

Unlike DR into H- and He-like ions (Tables II and III), the DR related to the n = 3 core transition is very important compared with the 2s–2p related recombination. For this reason, the temperature dependence of the total recombination rate (which is the sum of the DR rates related to the 2s–2p, 2s–3p, …, core transitions) is complex and differs qualitatively from the rates of DR into H- and He-like ions, which are dominated by a single exponential factor.

4 IV. EXCITED-STATE COUPLING OF DIELECTRONIC RECOMBINATION IN DENSE PLASMAS

Table V shows the DR rates related to the excited states 1s22p of Li-like ions. It can be seen from a comparison of the numerical data in Tables IV and V that the contribution from the excited states is even more important than that from the ground state. For example, for Be, Bd(2s − 3p) = 1.97 × 10−6, whereas Bd(2p − 3d) = 1.78 × 10−4 and Bd(2p − 4d) = 1.88 × 10−5. This means that the excited-state contribution is up to two orders of magnitude larger than the ground-state contribution. Therefore, even for rather moderate densities with small populations of the excited states, their contribution to DR can be important.

Table 5. Fitting coefficients according to Eqs. (3.20) and (3.21) for DR into excited states of Li-like ions originating from the 1s23lnl′ and 1s24lnl′ autoionizing levels, with Z = Zn − 2, m = 1, and l0 = 1. The numerical data include corrections for multiple decay channels (three channels for 1s23lnl′ and six channels for 1s24lnl′).

1s23lnl′: α0 = 1s22pα = 1s23d1s24lnl′: α0 = 1s22pα = 1s24d
ElementBdχdBdχd
Be1.78 × 10−40.1401.88 × 10−50.190
B2.99 × 10−40.1372.01 × 10−50.189
C3.74 × 10−40.1352.04 × 10−50.188
N4.44 × 10−40.1332.18 × 10−50.187
O5.15 × 10−40.1312.35 × 10−50.187
F5.52 × 10−40.1302.53 × 10−50.186
Ne5.65 × 10−40.1282.67 × 10−50.185
Na5.76 × 10−40.1272.88 × 10−50.181
Mg5.73 × 10−40.1253.28 × 10−50.174
Al5.61 × 10−40.1243.32 × 10−50.172
Si5.39 × 10−40.1223.33 × 10−50.171
P5.19 × 10−40.1203.48 × 10−50.167
S4.96 × 10−40.1193.46 × 10−50.165
Cl4.71 × 10−40.1173.44 × 10−50.164
Ar4.48 × 10−40.1153.41 × 10−50.163
K4.25 × 10−40.1143.38 × 10−50.161
C4.04 × 10−40.1123.34 × 10−50.160
Sc3.83 × 10−40.1103.30 × 10−50.159
Ti3.64 × 10−40.1093.25 × 10−50.158
V3.45 × 10−40.1073.20 × 10−50.157
Cr3.27 × 10−40.1053.14 × 10−50.156
Mn3.11 × 10−40.1043.08 × 10−50.156
Fe2.95 × 10−40.1023.02 × 10−50.155
Co2.80 × 10−40.1012.95 × 10−50.154
Ni2.66 × 10−40.09922.88 × 10−50.154
Cu2.53 × 10−40.09782.80 × 10−50.153
Zn2.40 × 10−40.09642.72 × 10−50.153
Ga2.28 × 10−40.09512.64 × 10−50.153
Ge2.17 × 10−40.09392.56 × 10−50.152
As2.06 × 10−40.09272.47 × 10−50.152
Se1.96 × 10−40.09162.39 × 10−50.152
Br1.86 × 10−40.09052.30 × 10−50.152
Kr1.77 × 10−40.08952.22 × 10−50.152
Rb1.68 × 10−40.08852.14 × 10−50.152
Sr1.60 × 10−40.08762.05 × 10−50.152
Y1.52 × 10−40.08671.97 × 10−50.152
Zr1.45 × 10−40.08591.89 × 10−50.152
Nb1.38 × 10−40.08511.82 × 10−50.152
Mo1.31 × 10−40.08441.74 × 10−50.152

查看所有表

Particular important cases are encountered if the first excited states are related to Δn = 0 radiative transitions. Because these transition probabilities are orders of magnitude lower than those for Δn > 0 transitions, Boltzmann populations with respect to the ground state are already achieved for rather low electron densities. For example, for Be, at densities of about 1015 cm−3, the population of the excited state 1s22p is more important than that of the ground state 1s22s.46 Therefore, all excited states of beryllium (e.g., for tokamaks at typical divertor densities) make larger contributions than the ground state.

The excited-state contribution could even be important at very low densities if the excited states are metastable states. Therefore, for heavy ions, where we encounter excited states that are close to ground states, related either by a dipole-allowed radiative transition or by multipole transitions, DR is extremely complex even at rather low densities. This is the main reason why, to date, ionic balance calculations of heavy elements differ strongly from one method to another and why DR remains an active field of research and of considerable interest for a number of applications (nuclear fusion, astrophysics, radiation sources, spectroscopic diagnostics, etc.).

In conclusion, the excited-state contribution is driven by atomic kinetics that can have a much greater impact than any other complicated effects related to ground-state contributions. This points again to the great practical importance of approximate methods, including the quasiclassical approach, that provide the possibility of obtaining numerical data even for large quantum numbers (which can be quite important for DR).

We underline that the excited-state contributions to DR up to high quantum numbers for the corresponding core transitions may exceed the ground-state contribution by many orders of magnitude. It is for this reason that it is essential to include excited-state contributions as much as possible, even if these are based on atomic structure calculations of limited precision, rather than attempting to improve via sophisticated atomic structure calculations the simplest core-transition-related DR rates while ignoring higher-order and excited-state contributions.

5 V. ANGULAR-MOMENTUM-CHANGING COLLISIONS

A strict consideration of angular-momentum-changing collisions requires a very extended atomic-level system that includes all details of the autoionizing states in order to treat properly the collisional redistribution of populations. We restrict ourselves here to a discussion of principles with the help of the most frequently employed formula for DR and proceed from dielectronic capture from channel k and with radiative transition ji [see also Eqs. (3.3)(3.5)]:Let us now consider a simple illustrative example, namely the Lyα dielectronic 2l2l′ satellites of He-like ions and depict two levels, one that has very large autoionization rate and one that has a negligible one. For the first case, we consider the level j=2p21D2, k=1s2S1/2 and the radiative transition j=2p21D2i=1s2p1P1. Atomic structure calculations for carbon (Zn = 6) deliver37

For the second case, we consider the autoionizing configuration j=2p23P1, k=1s2S1/2 and the radiative transition j=2p23P1i=1s2p3P2. Atomic structure calculations (again for Zn = 6) deliverfrom which it follows that Qk,jiZ+1,Z=0.

Assuming a two-level system where only dielectronic capture and angular momentum changing collisions (characterized by the rate coefficient Cjj) contribute, the atomic populations njZ and njZ are given bywherewith q = j, j′. In the absence of collisions, Eqs. (5.1)(5.4) becomewhere the superscript “(0)” indicates the low-density case. As can be seen from Eq. (5.5), the low-density dielectronic intensity satellite factor [Eq. (3.5)] is reproduced from Eq. (5.2) if the angular-momentum-changing collisions are negligible, i.e., if

Note that for very closely spaced levels, ion–ion collisions might also be of importance.

To understand the effect of angular-momentum-changing collisions on the total DR rate, we need to consider the sum for the two levels, i.e.,where the subscript “coll” for the single DR rates DRcollk,jiZ+1,Z and DRcollk,jiZ+1,Z indicates that these rates include collisional processes. This has to be distinguished from Eq. (5.1), which is a low-density approximation. It is of principal interest to understand the change in DR due to collisions with reference to the low-density case, i.e., we consider the ratioThe collisional DR rates cannot be determined from relations like Eq. (5.5) but need to be determined directly from the populations, i.e.,because the product of the level population with the radiative decay is the rate at which the excited state decays to the ground state, which is equivalent to DR [note that the usual branching ratios that appear in formulas like Eq. (5.1) are already included via the equilibrium population] if the right-hand sides of Eqs. (5.2) and (5.3) are driven by dielectronic capture and angular-momentum-changing collisions between the autoionizing levels under consideration. Combining Eqs. (5.7) and (5.8), we obtaini.e.,Because EkjDCEkjDC, we have for the population ratio in the low-density case (for the example given above)since ΓjkZ,Z+1ΓjkZ,Z+1 and ΓjkZ,Z+1lAjlZ (see the above example). Therefore, population is essentially transferred by angular-momentum-changing collisions from level j′ to level j, but not vice versa. Let us now consider the above example with the populations given by Eqs. (5.2) and (5.3):andEquations (5.2), (5.3), and (5.8) indicate that for autoionizing levels with very large autoionization rates, the populations are close to the low-density case. Equation (5.4) therefore corresponds to the low-density case [Eq. (5.5)], i.e.,Substituting Eqs. (5.11), (5.12), and .(5.14) into Eq. (5.9), we obtainBecause gjCjjgjCjj for closely spaced levels, Eq. (5.15) takes the formIf lAjlZneCjj, then the term in parentheses in Eq. (5.16) is about one-half, and the relation indicates that the total DR rate is enhanced (i.e., DRcolltotZ+1,Z/DRtotZ+1,Z>1) owing to angular-momentum-changing collisions. This can be understood in a transparent qualitative picture: for the level j′ with high autoionization rate, the dielectronic capture is high and, owing to the large autoionization rate, the branching factor for radiative deexcitation is small. If, however, a certain fraction of population is collisionally transferred to another level before autoionization and radiative decay disintegrate the upper level j′, then level j is effectively populated by collisions from j′ → j (because the population of level j is small since dielectronic capture is insignificant owing to a small autoionization rate). The transferred population, however, has a very favorable branching factor for level j compared with level j′. In the above example,for level j, whereasfor level j′. Therefore, the transferred population is more effectively transferred to the ground state to finally contribute to DR.

The impact of angular-momentum-changing collisions on the total DR rate is difficult to observe in dense plasmas, because the impact of such collisions is only indirect, namely, via the change in ionization balance, where other recombination processes (e.g., three-body recombination and radiative recombination) also come into play. However, angular-momentum-changing collisions can be directly observed via the collisional induced change in the spectral distribution of the corresponding dielectronic satellite transitions.

Figure 2 demonstrates this effect via He-like Lyα transitions and their associated satellite transitions 2lnl′ → 1s2l + hν. Numerical calculations of the spectral distribution have been carried out using the MARIA suite of codes,10,47–49 taking into account an extended level structure, with LSJ-split levels of different ionization stages for ground, single, and multiple excited states being simultaneously included. Strong density effects are indicated by red arrows in Fig. 2. Not only do the 2l2l′ satellites show strong density effects near λ ≈ 0.853 nm, but so do the 2l3l′ satellites near λ ≈ 0.847 nm. The density sensitivity of the 2l3l′ satellites starts at lower densities, because the collisional rates between the 2l3l′ configurations are in general larger than those for the 2l2l′ configurations [the collisional rates C(2lnl′–2lnl″) increase with principal quantum number n, the corresponding radiative rates A(2lnl″–1snl″) are almost independent of n, and the autoionization rates Γ(2lnl″–1s) decrease with n]. Also indicated in Fig. 2 is the so-called “blue satellite” emission located on the blue wing of the Lyα resonance line. These satellite transitions exhibit negative screening7 that is due to the strong effect of angular-momentum coupling (F states). [Note that the term “negative screening” arises from use of the Bohr formulato match the actual energy E via a screening constant σ: in cases where the match can only be obtained for effective charges Zeff = Znσ, the screening constant is negative; see also Ref. 12.] As can be seen from Fig. 2, angular-momentum-changing collisions have little effect on blue satellites (because their autoionization rates are quite large).

Fig. 2. MARIA simulations of dielectronic satellite emission near Lyα of H-like Mg ions for different values of the electron density at kTe = 100 eV. The red arrows indicate the rises in intensity of particular satellite transitions with increasing density. Satellites indicated in blue have effective negative screening due to strong angular-momentum coupling effects.

下载图片 查看所有图片

In dense-plasma spectroscopy, the perturbation of the spectral dielectronic satellite distribution due to angular-momentum-changing collisions is employed for density diagnostics.5–7 Note that although the satellite intensity shows variations with temperature too, density determination via dielectronic satellite transitions is essentially not a two-parameter problem: the temperature determination relies on the intensities of satellite transitions (with high autoionization rates) relative to the resonance line, while the density determination is based on the relative intensities of dielectronic satellites among the transitions themselves (with low and high autoionization rates). The density diagnostic is usually rather independent of the electron temperature, because all cross sections of angular-momentum-changing collisions are typically in the Born limit (the energy differences are typically much smaller than the electron temperature).

6 VI. ELECTRIC FIELD EFFECTS

6.1 A. Ionization potential depression of spectator electron orbitals

Atomic population kinetics of gases and plasmas has been applied very successfully to the study of low-density environments, where atoms and ions are essentially field-free. As the density increases, however, the free-atom model breaks down, resulting in a perturbation of the atomic energy levels. This perturbation manifests itself essentially in a broadening and a shift. Such perturbations can be observed in high-resolution spectroscopic experiments via analysis of the line broadening, the line shift, and the disappearance of the line emission corresponding to the ionization potential depression (IPD) of the upper level. The IPD is of great interest for applications in thermodynamics and also for the understanding of various radiative properties, such as emission, absorption, and scattering (for further reading on current developments in IPD research with spectroscopic precision, the interested reader is referred to Refs. 50 and 51 and references therein).

The fundamental origin of the perturbation of energy levels is the plasma electric microfield, which in turn also limits the number of bound states. Electric field ionization starts at a critical field strength Fcrit given by39,52Fcrit6.8×108Zeff3nF4[V/cm],where Zeff is the effective ion charge and nF is the principal quantum number at which field ionization starts. To estimate the limited number of quantum states that effectively take part in the recombination process, we identify the critical field strength Fcrit with the mean field strength, which can be expressed in terms of the Holtsmark normal mean field strength F0 (ZP is the perturber charge and NP is the perturber density),52aFmean3.41F0=3.4×4π4152/3Ryea0ZPNP2/31.3×106ZPNP2/3[cm3][V/cm],and we identify the principal quantum number nF with the maximum principal quantum number of the spectator electron nl of the autoionizing configuration, i.e.,nFnmaxspectator4.8×103Zeff3/4ZP1/4NP1/6(cm3).nmaxspectator limits the contribution of high-n spectator electrons according to Eq. (3.17) [and therefore effectively limits the Bd values in Eqs. (3.14), (3.15), and .(3.20)]. Equation (6.3) is useful to understand the various effects that limit DR (apart from the electric field effect, kinetic effects, for example, may likewise limit the effective rate coefficient; see below). Note, however, that Eq. (6.3) itself provides only a rough estimate of the principal quantum number of the spectator electrons due to the electric field effect, since (a) it neglects the fact that usually up to five perturber charge states are present in a dense plasma, (b) the field strength distribution is only taken into account via the mean field value, and (c) charged-particle correlations that influence the field strength distribution itself are ignored.

Effective DR is not only limited by the maximum principal quantum number nmaxspectator from Eq. (6.3), but also by collisional disintegration of the autoionizing level: in fact, in a dense plasma, not only does the high density of charged particles result in a high microfield, but the high density also implies a significant rate of collisional processes. For autoionizing states, we need to consider two stages: a first stage involving collisional processes associated with core relaxation and a second stage following core relaxation. Concerning the first stage, we need to compare the radiative stabilization rate of the autoionizing configuration (which is the origin of effective recombination) with the collisional ionization rate of the spectator electron (which is the origin of collisional disintegration of the autoionizing level). Therefore, effective DR requiresArad(αα0)neIn,where Arad(αα0) is the radiative stabilization rate from Eqs. (3.10) and (3.20) (note that this rate is distinctly different from the radiative transition rate of the spectator electron itself), ne is the electron density, and In is the electron collisional ionization rate coefficient of the spectator electron with principal quantum number n. Note that for autoionizing states, the stabilization rate is almost independent of the high-n spectator electron, which is distinctly different from the scaling law for singly excited states (approximately A ∝ 1/n3). If the radiative stabilization is associated with a change in principal quantum number, then the H-like approximation can be used to estimate Arad(αα0):Aradαα0,Δn>01.57×1010Zeff4nα0nα3(nα2nα02)[s1],where Zeff is the effective ion charge. If Arad(αα0) is associated with a transition without change in principal quantum number, i.e., Δn = 0, then numerical Hartree–Fock calculations including angular-momentum coupling are required to obtain reasonable estimates of the transition probability. Numerical calculations indicate that for Li-like and Be-like ions, transition probabilities and oscillator strengths can be roughly estimated from the expressionsArad(1s22p2P1s22s2S)1.2×108(Zn3)(s1),f(1s22s2S1s22p2P)1.2/(Zn2),Arad(1s22s2p1P1s22s21S)8×108(Zn3)[s1],f(1s22s21S1s22s2p1P)4.5/(Zn2).The ionization rate of the spectator electron can be estimated from the following expression (note that there are many competing formulas available in the literature,26 but for the present estimates, the simple Lotz-like formula appears to provide a good balance between complexity and precision):In6×108RyEn3/2βneβnln1+0.562+1.4βnβn(1+1.4βn)[cm3/s],where Ry = 13.6056 eV andβn=EnkTe.En is the ionization potential of the spectator electron, which can be approximated byEn(Zeff1)2Ryn2.

It is important to point out that the condition (6.4) is necessary for DR to be effective, but it is not sufficient. After radiative stabilization αα0 of the core, we need to radiatively stabilize the spectator electron nl to the ground state “gr” (if it disappears in the continuum, no recombination is encountered), i.e., we must haveArad(nlgr)neIn.For spectator electrons, Δn = 0 transitions are not relevant and an expression similar to Eq. (6.5) describes all cases of practical interest:Arad(nlgr,Δn>0)1.57×1010(Zeff1)4ngrn3(n2ngr2)[s1].Whether condition (6.4) or (6.9) is more stringent depends on the type of core transition. Let us consider as an example a magnetically confined deuterium fusion plasma containing Li-like iron impurities. Assuming ZP = 2, NP = 1013 cm−3, ne = 1014 cm−3, and kTe = 10 keV, we obtain from Eq. (6.3) with Zeff = 24 a maximum principal quantum number of spectator electrons of about nmaxspectator300. For Li-like iron, DR is associated with the core transition α0α = 1s22s2S → 1s22p2P with Arad(1s22p2P → 1s22s2S) ≈ 3.5 × 109 s−1 and f(1s22s2S → 1s22p2P) ≈ 0.066. As the relevant principal quantum numbers are of the order of 100, we haveβn=EnkTe=(241)2×13.6104n2104,and we can approximate the βn-dependent terms in Eq. (6.6) roughly by 0.1 and solve Eq. (6.4) for the critical principal quantum number of the spectator electron ncritspectator [note that the critical principal quantum number is obtained from Eq. (6.4) with equality sign]: ncritspectator(7×1021/ne)1/3400.

Let us now investigate the condition (6.9). For Li-like iron, ngr = 2 and Eq. (6.10) provides roughly Arad(nl → gr, Δn > 0) ≈ 2 × 1015/n5 [s−1]. Taking into account the n dependence of Arad(nl → gr, Δn > 0), we can solve Eq. (6.9) for the critical principal quantum number ncritspectator [note that the critical principal quantum number is obtained from Eq. (6.10) with equality sign] of the spectator electron: ncritspectator(4×1023/ne)1/816. Therefore, in the above example, the condition (6.10) is much more stringent than conditions (6.3) and (6.4).

Let us now consider the above example for parameters typical of inertial fusion plasmas. We assume that iron is employed as a diagnostic tracer element in a compressed dense plasma. Assuming ZP = 2, NP = 1023 cm−3, ne = 1025 cm−3, and kTe = 10 keV, we obtain from Eq. (6.3) with Zeff = 24 a maximum principal quantum number of the spectator electron of about nmaxspectator6. As before, α0α = 1s22s2S → 1s22p2P with Arad(1s22p2P → 1s22s2S) ≈ 3.5 × 109 s−1 and f(1s22s2S → 1s22p2P) ≈ 0.066. As the relevant principal quantum numbers are of the order of 10, we haveβn=EnkTe=(241)2×13.6104n2102,and we can approximate the βn-dependent terms in Eq. (6.6) roughly by 0.4 and solve Eq. (6.4) for the critical principal quantum number of the spectator electron ncritspectator [note that the critical principal quantum number is obtained from Eq. (6.4) with equality sign]: ncritspectator(2×1021/ne)1/3<1. Therefore, DR associated with the Li-like core transition α0α = 1s22s2S → 1s22p2P is ineffective. The essential physical reason is related to the low transition probability for the Δn = 0 core transition. Therefore, total recombination rates are effectively related to core transitions involving the K shell because, in this case, the radiative decay rates are very large owing to the strong Z scaling [see Eq. (6.5)]: assuming Zeff = 24 as before, but taking nα0 = 1 and nα = 2 in Eq. (6.5), we have Arad(αα0, Δn > 0) ≈ 1.1 × 1014 s−1, i.e., a value five orders of magnitude higher compared with the Δn = 0 core transition discussed above. In this case, K-electron involvement in a core transition allows a few principal quantum numbers of the order of 1 to survive and to contribute effectively to DR.

What is the conclusion from the above estimates? They indicate that it might not be appropriate to adopt a purely atomic structure point of view to obtain convergence of the sum of DR rates over a large range of principal quantum numbers of the spectator electron nl in Eq. (3.17): owing to microfields and collisional–radiative competition, large principal quantum numbers might not effectively contribute to the total DR rate. Therefore, total DR rates calculated with rather approximate methods (quasiclassical methods or Vainshtein’s simplified QMMC method) but taking into account the plasma microfield and collisional processes might be more accurate than purely adopting sophisticated atomic structure calculations.

In view of these results, we now address the influence of the plasma microfield itself on the autoionization rates.

6.2 B. Perturbed autoionization rates

The influence of the electric field on the autoionization and corresponding DR rates was first studied in Refs. 53–55 in the context of the simplest atomic system of He-like autoionizing states 2l2l′. It was realized that forbidden autoionizing processes (i.e., forbidden in the LS-coupling scheme) become allowed by electric field mixing of autoionizing bound-state wavefunctions. The allowed autoionization width is given by the first-order transition rateΓ(dc)=2πdVc2δ(EdEc),where V is the electrostatic interaction. Because V is a scalar operator, the autoionization vanishes unless there are available adjacent continuum states c with the same angular momentum and parity as those of the discrete levels d.32 Because of the absence of even-parity P states below the second ionization threshold, the 2p23P-state of He-like ions is metastable against autoionization decay. In the presence of perturbing electric fields, however, autoionization of the state a = 2p23P may occur by a second-order process involving a field-induced transition to the nearby autoionizing state d = 2s2p3P. In a quasistatic ion field, the field-induced autoionization rate is given byΓ(ac)=2π|daQEddVc(EaEd)+i(Γd+Ad)/2|2δ(EaEc),where Q is the electric dipole moment operator, and Γd and Ad are respectively the autoionization and radiative widths of the state d. Therefore, the first-order contribution from the field-induced transition decays directly into the nonresonant continuum c = 1sεp3P.

It should be noted that for practical applications, not only field-induced transitions have to be considered, but intermediate coupling, configuration, and magnetic interactions too. In particular for highly charged ions, these “non-electric-field effects” may make a considerable contribution to the forbidden autoionization width, as is demonstrated by the results in Table VI, which have been calculated using the FAC code.56 In addition, the Breit interaction56a induces an autoionization rate for the 2p23P1 state.

Table 6. Field-free autoionization decay rates (s−1) including intermediate coupling, configuration, and magnetic interaction.

StateZn = 3Zn = 6Zn = 13Zn = 18Zn = 26Zn = 42
2p21S08.4 × 10105.1 × 10121.3 × 10131.9 × 10133.4 × 10137.0 × 1013
2p21D21.5 × 10142.5 × 10143.1 × 10143.1 × 10142.3 × 10142.1 × 1014
2p23P02.9 × 1072.3 × 1092.3 × 10111.2 × 10123.7 × 10122.8 × 1012
2p23P1000000
2p23P1 with Breit interaction2.6 × 1076.8 × 1081.9 × 10107.2 × 10103.2 × 10112.2 × 1012
2p23P21.1 × 1093.1 × 10103.0 × 10122.1 × 10131.1 × 10141.5 × 1014

查看所有表

Table VI also illustrates the general effect that if the nuclear charge increases, then the autoionization widths are more evenly distributed over the levels. Therefore, electric field effects are best studied for low-Z elements. The table also demonstrates that autoionization rates are strongly dependent on LSJ quantum numbers: therefore, simple summations over l quantum numbers might be a quite inappropriate way to simplify complex atomic structures in kinetics.48

From the relationship between the corresponding capture and autoionization rates, it follows that the electric field can induce DR through normally inaccessible high-angular-momentum states that have large statistical weights.54,55 In fact, in a plasma, the angular momentum l is no longer a good quantum number, because the presence of an electric field destroys the spherical symmetry. However, the projection LZ of L, which generates the magnetic quantum number m, defined with respect to the direction of the electric field, remains a good quantum number. For nonzero quantum numbers m, this results in a twofold degeneracy of the outer electrons in addition to the twofold degeneracy due to spin. The appropriate transformation of the field-free substates l has the formnλm=l=mn1nlmnlmnλm,where the electric quantum number λ, which replaces l in the presence of the electric field, can take integer values in the range λ = 0, …, n − |m| − 1. Calculations54,55,57,58 demonstrate that the dependence of the autoionization rates on the quantum number λ is rather smooth, in contrast to the field-free case, where the autoionization rates decrease rapidly with quantum number l. For this reason, dielectronic capture increases in the presence of an electric field, because it is proportional to the autoionization rate and the statistical weight: DCk,jZ+1,ZgjZΓjkZ,Z+1. Because the DR rate is proportional to dielectronic capture rate [see Eq. (2.7)], this results in a considerable increase in the total DR rate. For example, for the autoionizing states 1s22pnl in Be-like Fe22+, an approximately threefold increase in the DR rate was found even for densities as low as 1014 cm−3.54 This dramatic increase at rather low densities is connected in particular with the fact that for the 1s22pnl configuration, the resonance spontaneous transition probability 2s–2p is not very large and high-n states have autoionization rates larger than radiative decay rates for n quantum numbers up to about 100. Consequently, high-n states contribute considerably to the DR rate. As high-n states are likewise strongly affected by rather small electric fields, a considerable impact on the total recombination rate is encountered even for rather low plasma densities (being of importance for typical densities of the solar corona or magnetic fusion plasmas).

Interaction with an electric field makes atomic structure calculations extremely complex, and it is difficult to derive general conclusions. However, it has been demonstrated57,58 that the quasiclassical approach combined with a transformation to parabolic quantum numbers [Eq. (6.13)] gives results that are in surprisingly good agreement with those of extremely complex numerical calculations.59 Moreover, the quasiclassical approach combined with the transformation to parabolic quantum numbers59a,59b enables the derivation of a closed-form expression for the autoionization rate in an electric field:Γ(n,λ,m)=lminlmaxP(nl;λm)Γ(nl)dl,withlmin2=12{[(n1)2+m2λ2][(n1)2+m2λ2]24(n1)2m2}andlmax2=12{[(n1)2+m2λ2]+[(n1)2+m2λ2]24(n1)2m2},where Γ(nl) is the standard autoionization rate in spherical polar coordinates (which is independent of m owing to spherical symmetry) and P(nl; λm) is a joint probability (with normalization equal to unity) for the appearance of spherical (nl) and parabolic (nλm) quantum numbers that can be expressed in terms of Clebsch–Gordan coefficients. For large quantum numbers and the condition m < l « n (the quasiclassical limit of Clebsch–Gordan coefficients that is of practical interest), the joint probability can be approximated by58P(nl;λm)1π2l(l2lmin2)(lmax2l2).Substituting quasiclassical values for the autoionization rate31,32,36,38,42Γ(nl) into Eq. (6.14) and using Eq. (6.13), we obtain an autoionization rate in parabolic quantum numbers expressed in terms of universal functions (t = l/leff, leff=(3Z2/ω)1/3):Γ(n,λ,m)=fijπn3I(tmin,tmax),I(tmin,tmax)2lmax3Z2ω2/3Ylmin(ω/3Z2)1/3,Y(x)0.284exp(2x3),where fij is the oscillator strength of the core transition with charge Z (e.g., the oscillator strength corresponding to the transition 1s–2p in H-like Al for the He-like 2lnl′ satellites, Z = 13). The expressions (6.14)(6.20) demonstrate similarly a broad distribution over the electric quantum number λ that finally results in an increase of the DR rate.

7 VII. THE LOCAL PLASMA FREQUENCY APPROACH TO DIELECTRONIC RECOMBINATION

As discussed already, DR is the most effective recombination channel in electron–heavy-ion collisions. Owing to the complex electronic structure of multielectron ions, providing a proper account of all necessary channels is a very difficult task, in particular for open-shell configurations. In addition, in dense plasmas, dielectronic capture might effectively proceed from excited states (see also Sec. IV), thus considerably increasing the number of quantum channels for dielectronic capture. Moreover, in heavy ions, numerous metastable states may play the role of excited states even in rather low-density plasmas, thereby increasing the numerical complexity of fully quantum calculations considerably. At present, DR of heavy ions is still a matter of controversy and is one of the main sources of discrepancy between different methods of calculation for radiation loss and ionic charge state distributions. It is therefore of great interest to develop different methods for the calculation of the DR rate in heavy ions that permit more general studies, including analysis of scaling laws. Below, we develop a twofold statistical approach that is realized by a combination of the statistical theory of atoms60–65 with the local plasma frequency approximation.40,41,66

Let us start from Eq. (3.4) and rewrite the formula for the total DR rate aswhere kTe is the electronic temperature in eV, gi and gf are the statistical weights of the initial and final states of the atomic core, A is the radiative transition probability inside the core, Γ is the autoionization decay rate of an excited atomic energy level, ℏω is the transition energy inside the core, Zi is the ion charge, a0 is the Bohr radius, and n and l are the principal and orbital quantum numbers, respectively, of the captured electron. The radiative decay rate is expressed simply in terms of the oscillator strength fij for the transition inside the core:where c is the speed of light. To obtain an expression for the autoionization decay rate Γ(n, l), we use a relationship between the decay rate Γ(n, l) and the partial electron excitation cross section σex(n, l) in the semiclassical representation. The quantities Γ(n, l) and σex(n, l) describe mutually inverse processes, so the relationship between them can be obtained from the detailed balance between ions XZi+1 and XZi. Thus, we obtainThe electron excitation cross section in the semiclassical approximation takes the formwhere the function G(u) is given bywhere K1/2 and K3/2 are the Macdonald functions (modified Bessel functions of the second kind). Taking into account that the essential values of the argument of the function G(u) are never close to zero, it is possible to replace G(u) by its asymptotic expansion:With these approximations, the autoionization decay rate takes the formThe sum of the absorption oscillator strengths satisfies the Thomas–Reiche–Kuhn sum rule, i.e.,(note that Ne is the number of electrons, while ne is the electron density and n the principal quantum number). In the statistical model, the oscillator strengths are expressed in terms of the atomic electron density ne(r, q, Zn), and the statistical sum rule is given byThe application of the semiclassical statistical model to the general formula (7.1) for the total DR is achieved by using the relationshipsandAfter all the substitutions, we obtain for the DR rateswith Te[a.u.] = Te[eV]/27.21 and t = n/n1, where n1 is the minimum possible quantum number. n1 is the lowest level at which electron capture is possible and corresponds to an energy of an incident electron Ei, given bythat is equal to zero, i.e.,from which it follows thatIn the framework of the Thomas–Fermi model,60–65 the electron density distribution of a particular element and charge state is given bywithwhere Zn is the nuclear charge, Z is the ion charge, q characterizes the degree of ionization, and rTF is the Thomas–Fermi radius. The Thomas–Fermi function φ(x, q) can be approximated by the Sommerfeld method61,62,67 as an exact particular solution of the Thomas–Fermi differential equation:withThe reduced radius x0(q) is determined from the boundary conditionIn a high-temperature plasma, i.e., when the degree of ionization q = Z/Zn is not too low, the reduced radius can be approximated byNote that the use of the Thomas–Fermi model described above, which ignores exchange corrections, is quite appropriate within the framework of the approximations of the statistical DR model itself (see also the further discussion below). The ionization energy of an atom or ion is then given byAs can be seen from Eq. (7.31), the hydrogenic approximation Zn2Ry of the ionization potential of an ion with charge Zn is corrected via the Thomas–Fermi electron density distribution, which depends on the nuclear and ionic charges [the factor in braces {⋯} in Eq. (7.31)]. A comparison of the ionization energies obtained from Eq. (7.31) with the results of detailed Hartree–Fock calculations shows reasonable agreement for heavy elements over a wide range of degrees of ionization.66 Note that more accurate descriptions of the ionization potentials can certainly be obtained from a direct fit to the vast number of ionization potentials that are known as functions of Z and Zn:68Many modifications of the Thomas–Fermi model have been proposed with the aim of including shell structure, obtaining improved ionization energies, and, in particular, approaching the Hartree–Fock results for the electron density distribution. In further developments to improve the statistical approach, however, one must not lose sight of the requirement that the fundamental equations of the statistical model of atoms, including the various corrections terms, should not be too complicated, in particular no more complicated than the basic equations of the quantum mechanical many-body approximation (e.g., the multiconfiguration Hartree–Fock methods). One must always bear in mind that the statistical theory of atoms is only a rough approximation of the quantum atom and that its advantage is its extreme simplicity both in structure and application to determine the electron and potential distributions of atoms, to derive elementary processes in collisional–radiative regimes, to shed light on detailed atomic structure calculations (in particular for heavy atoms), and, in particular, to derive general scaling laws that could hardly be obtained otherwise. It is this practical philosophy that we adopt when we consider Eqs. (7.19)(7.31) for the statistical framework of the atom/ion and its realization via the local plasma frequency.

In the simplest version of the statistical model, the atomic density, excitation energies, and oscillator strengths do not depend on the orbital momentum quantum number l. If we average the branching factor over orbital momentum l, i.e.,then we obtain for the total DR ratewhere the function A(x, l) is given by Eqs. (7.14) and (7.15). Instead of averaging over the branching factor, we may investigate averaging the autoionization decay rate Γ(n, l) from Eq. (7.7) over the orbital quantum number, i.e.,For the corresponding total DR rate, we then obtain

For heavy ions, the quantum mechanical level-by-level calculations are very complex and have so far been carried out mainly for closed-shell configurations. Only recently have open-shell configurations also been considered.69,70 In open-shell configurations (e.g., the open 4p, 4d, and 4f shells, or even higher ones such as the 5p, 5d, 5f, and 5g shells), excitation–autoionization channels are very complex, and the overall completeness of quantum mechanical level-by-level calculations should still be considered with care. Analysis shows that order-of-magnitude disagreements can be expected at low temperatures, while at high temperatures, different level-by-level quantum mechanical models differ by about a factor of 2, and the Burgess–Mertz approach32 may deviate by many orders of magnitude and also gives an entirely inadequate temperature dependence, as demonstrated by more detailed calculations.71

Below, we compare the different approaches with detailed quantum mechanical level-by-level calculations of the DR rates. Figure 3 shows the total DR rates of xenon Xe26+ and gold Au51+ (the Ni-like 3s23p63d10 configuration into which dielectronic capture proceeds) calculated with the l-averaged statistical model from Eqs. (7.33)(7.35) that employs the Thomas–Fermi model of Eqs. (7.19)(7.31), the Burgess–Mertz formula from Eqs. (3.10)(3.15), and the quantum level-by-level calculations from Ref. 71.

Fig. 3. Comparison of the l-averaged statistical approach with the Burgess and quantum level-by-level calculations for the Ni-like sequence 3s23p63d10 of Xe26+ and Au51+.

下载图片 查看所有图片

The statistical model compares quite well (within a factor of two) over a very large temperature interval until very low temperatures, while the Burgess approach entirely fails to describe the total DR rate of heavy ions. Similar observations are made for other isoelectronic sequences. Figure 4 shows a comparison of the results from the different approaches for the DR rates of Sr-like (4s24p64d2) and Zn-like (4s2) tungsten W36+ and W44+, respectively.70

Fig. 4. Comparison of the l-averaged statistical approach with the Burgess and quantum level-by-level calculations for the Sr-like sequence 4s24p64d2 of W36+ and the Zn-like sequence 4s2 of tungsten W44+.

下载图片 查看所有图片

It is particularly impressive that the statistical model provides a rather good approximation of the total DR rate in the low-temperature region that is numerically exceedingly difficult to treat by fully quantum mechanical level-by-level calculations. Thus, the statistical model in its simplest version seems to provide even the possibility of estimating the order-of-magnitude correctness of very complex quantum level-by-level calculations. Moreover, it should be remembered that currently even the most sophisticated quantum level-by-level calculations70 have been performed only in the low-density limit (the coronal model: a low-density limit in which three-body recombination may be entirely neglected), where the branching factors are entirely determined by radiative and autoionization decay rates while dielectronic capture proceeds from the respective ground states of the various charge states only. In high-density plasmas, however, as discussed above, collisional depopulation is due to electron collisional ionization or collisional transfer to other levels. In addition, excited states are highly populated, and very efficient channels of DR may proceed from these. This may entirely change the properties of the total DR rate, because dielectronic capture into excited states can be even more important than the corresponding capture to the ground state. This effect has been explicitly confirmed by high-resolution X-ray spectroscopy of dense laser-produced plasmas, where it has been shown that DR into excited states can exceed by many orders of magnitude the corresponding DR into ground states.23,45 For high-Z elements and open M, N, and O shells, excited states might be highly populated even at rather moderate electron densities. Therefore, all current detailed quantum level-by-level calculations to determine the DR rate have to be considered with care for each particular application. In this respect, the properties and the innovation potential of the statistical model look very advantageous for the determination of total DR rates for heavy elements.

Finally, it should be noted that the inclusion of more levels in the detailed quantum mechanical level-by-level calculations may not necessarily result only in an increase in the DR rate, but can also lead to a decrease, as discussed in Sec. III. Therefore, at present, the simple statistical method as presented above compares quite well with other available much more complex methods of calculation and has the advantages of generality and ease of application. In addition, there is much room for improvement to the statistical model via improvements to the Thomas–Fermi model (ionization energies, l-quantum-number dependence, adopting the Vlasov approach instead of the local plasma frequency etc.).

8 VIII. CONCLUSION

Dielectronic recombination (DR) can be cast as a product of dielectronic capture and a probability for radiative stabilization of the excited core followed by radiative decay of the spectator electron to the ground state. In the limiting case of negligible collisions compared with radiative decay, the total DR rate is a function of atomic structure constants only; i.e., the DR rate is a function of radiative decays, Auger rates, energy levels, statistical weights, and temperature. The quantum mechanical multichannel coupling (QMMC) approach demonstrates that the Burgess theory (including the Burgess–Mertz formulas) of DR may overestimate higher-order contributions by orders of magnitude. Collisional processes are identified to have multiple impacts on DR: (a) electron collisional excitation drives excited-state couplings that are often even more important than ground-state contributions; (b) angular-momentum-changing collisions between autoionizing states change the effective core relaxation and induce DR rates that are dependent on density—an effect that manifests itself in a perturbation of the spectral distribution of dielectronic satellite spectra; (c) collisional ionization of the spectator electron reduces high-n contributions to DR if ionizations are more frequent than radiative stabilizations of the core and the spectator electron. The plasma microfield strongly influences high-n DR contributions via ionization potential depression and perturbations of the spherical symmetry of autoionization matrix elements. Approximate calculations of DR rates, such as the quasiclassical method and Vainshtein’s approximate QMMC approach (which allows treatment of very large quantum numbers without convergence problems) combined with atomic population kinetic effects (excited-state-driven DR, angular-momentum-changing collisions, ionization potential depression, and collisional ionization of the spectator electron before radiative relaxation) are identified as providing effectively higher precision for the total DR rate than pure atomic structure calculations (even though the latter are more sophisticated). Finally, the first steps have been undertaken in a statistical approach to DR that is based on the local plasma frequency approximation rather than on standard atomic structure calculations. Quite good agreement with the most advanced quantum mechanical calculations so far available have been obtained, opening up a new field of activity for the plasma atom approach. Moreover, as DR has an impact on all ionization balance calculations, an accurate treatment of DR is mandatory and therefore remains an active and important field of research in atomic physics. On the other hand, spectroscopic diagnostics based on dielectronic satellite transitions provide a unique characterization of very complex phenomena and therefore contribute considerably to advances in many different areas of atomic and plasma physics.

References

[1] H. R.Griem, Principles of Plasma Spectroscopy (Cambridge University Press, New York, 1997).

[2] F. B.Rosmej, V. A.Astapenko, and V. S.Lisitsa, Plasma Atomic Physics (Springer, 2020).

[3] A.Burgess, “Dielectronic recombination and the temperature of the solar corona,” Astrophys. J.139, 776 (1964); 10.1086/147813A.Burgess, “Dielectronic recombination and the temperature of the solar corona,” 141, 1588 (1965).

[4] A. H. Gabriel. Dielectronic satellite spectra for highly-charge helium-like ion lines. Mon. Not. R. Astron. Soc., 1972, 160: 99.

[5] I. Yu. Skobelev, V. A. Vinogradov, E. A. Yukov. Effect of collisions on the intensities of the dielectronic satellites of resonance lines of hydrogenlike ions. Sov. Phys. JETP, 1977, 45: 925.

[6] M. Blaha, V. L. Jacobs. Effects of angular-momentum-changing collisions on dielectronic satellite spectra. Phys. Rev. A, 1980, 21: 525.

[7] J. Abdallah, F. B. Rosmej. Blue satellite structure near Heα and Heβ and redistribution of level populations. Phys. Lett. A, 1998, 245: 548.

[8] C. F. Hooper, V. L. Jacobs, L. A. Woltz. Effects of electric microfields on argon dielectronic satellite spectra in laser-produced plasmas. Phys. Rev. A, 1991, 44: 1281.

[9] A. Calisti, E. Galtier, F. B. Rosmej, et al.. Interference effects and Stark broadening in XUV intra-shell transitions in aluminum under conditions of intense XUV free electron laser irradiation. Phys. Rev. A, 2013, 87: 033422.

[10] F. B. Rosmej. Hot electron x-ray diagnostics. J. Phys. B: At., Mol. Opt. Phys., 1997, 30: L819.

[11] S. H. Glenzer, R. W. Lee, F. B. Rosmej. Measurements of suprathermal electrons in hohlraum plasmas with x-ray spectroscopy. Phys. Rev. Lett., 1998, 81: 365.

[12] A. Colaitis, O. Renner, M. Smid, et al.. Characterization of suprathermal electrons inside a laser accelerated plasma via highly-resolved Kα emission. Nat. Commun., 2019, 10: 4212.

[13] E. Galtier, D. Riley, F. B. Rosmej, et al.. Decay of crystaline order and equilibration during solid-to-plasma transition induced by 20-fs microfocused 92 eV free electron laser pulses. Phys. Rev. Lett., 2011, 106: 164801.

[14] R. W. Lee, F. B. Rosmej. Hollow ion emission driven by pulsed x-ray radiation fields. Europhys. Lett., 2007, 77: 24001.

[15] J. Abdallah, J. Colgan, A. Y. Faenov, et al.. Exotic dense-matter states pumped by a relativistic laser plasma in the radiation-dominated regime. Phys. Rev. Lett., 2013, 110: 125001.

[16] R. C. Elton, H. R. Griem, F. B. Rosmej, et al.. Investigation of charge exchange induced formation of two electron satellite transitions in dense laser produced plasmas. Phys. Rev. E, 2002, 66: 056402.

[17] V. S. Lisitsa, F. B. Rosmej, R. Schott. Charge exchange driven X-ray emission from highly ionized plasma jets. Europhys. Lett., 2006, 76: 815.

[18] V. S. Lisitsa, F. B. Rosmej. A self-consistent method for the determination of neutral density from X-ray impurity spectra. Phys. Lett. A, 1998, 244: 401.

[19] V. S. Lisitsa, D. Reiter, F. B. Rosmej. Influence of charge exchange processes on X-ray spectra in TEXTOR tokamak plasmas: Experimental and theoretical investigation. Plasma Phys. Controlled Fusion, 1999, 41: 191.

[20] V. S. Lisitsa, F. B. Rosmej. Non-equilibrium radiative properties in fluctuating plasmas. Plasma Phys. Rep., 2011, 37: 521.

[21] A. Y. Faenov, F. B. Rosmej. New innershell phenomena from Rydberg series of highly charged ions. Phys. Scr., 1997, T73: 106.

[22] A. Y. Faenov, T. A. Pikuz, F. B. Rosmej. Inner-shell satellite transitions in dense short pulse plasmas. J. Quant. Spectrosc. Radiat. Transfer, 1997, 58: 859.

[23] A. Y. Faenov, T. A. Pikuz, F. B. Rosmej. Line formation of high intensity Heβ-Rydberg dielectronic satellites 1s3lnl′ in laser produced plasmas. J. Phys. B: At., Mol. Opt. Phys., 1998, 31: L921.

[24] E. Krouský, O. Renner, F. B. Rosmej. Observation of H-like Al Lyα disappearance in dense cold laser produced plasmas. Appl. Phys. Lett., 2001, 79: 177.

[25] B. Deschaud, O. Peyrusse, F. B. Rosmej. Simulation of XFEL induced fluorescence spectra of hollow ions and studies of dense plasma effects. Phys. Plasmas, 2020, 27: 063303.

[26] I. I.Sobelman and L. A.Vainshtein, Excitation of Atomic Spectra (Alpha Science, 2006).

[27] C. Bowen, R. Florido, J. G. Rubiano. Review of the 4th NLTE code comparison workshop. High Energy Density Phys., 2007, 3: 225.

[28] C. Bowen, H.-K. Chung, C. J. Fontes. Comparison and analysis of collisional-radiative models at the NLTE-7 workshop. High Energy Density Phys., 2013, 9: 645.

[29] J. Colgan, C. F. Fontes, H. Zhang. Collisional-radiative modeling of tungsten at temperatures of 1200–2400 eV. Atoms, 2015, 3: 76.

[30] A.Sommerfeld, Atombau und Spektrallinien (Harri Deutsch, 1978), Vol. II.

[31] V. I. Kogan, A. B. Kukushkin, V. S. Lisitsa. Kramers electrodynamics and electron-atomic radiative collisional processes. Phys. Rep., 1992, 213: 1.

[32] R. D.Cowan, The Theory of Atomic Structure and Spectra (California University Press, 1981).

[33] Handbook of Atomic, Molecular, and Optical Physics, edited by G. W. F.Drake (Springer, 2006).

[34] A.Pradhan and S. N.Nahar, Atomic Astrophysics and Spectroscopy (Cambridge University Press, Cambridge, 2011).

[35] V. A.Astapenko, Polarization Bremsstrahlung on Atoms, Plasmas, Nanostructures and Solids (Springer, 2013).

[36] U. I. Safronova, L. A. Vainshtein. Wavelengths and transition probabilities of satellites to resonance lines of H- and He-like ions. At. Data Nucl. Data Tables, 1978, 21: 49.

[37] F. F. Goryaev, A. M. Urnov, L. A. Vainshtein. Atomic data for doubly-excited states 2lnl′ of He-like and 1s2lnl′ of Li-like ions with Z=6-36 and n=2,3. At. Data Nucl. Data Tables, 2017, 113: 117.

[38] I. L. Beigman, B. N. Chichkov, L. A. Vainshtein. Dielectronic recombination. J. Exp. Theor. Phys., 1981, 53: 490.

[39] V. S.Lisitsa, Atoms in Plasmas (Springer, 1994).

[40] D. S. Leontyev, V. S. Lisitsa. Statistical model of dielectronic recombination of heavy ions in plasmas. Contrib. Plasma Phys., 2016, 56: 846.

[41] A. V. Demura, D. S. Leont’iev, V. S. Lisitsa. Statistical dielectronic recombination rates for multielectron ions in plasma. J. Exp. Theor. Phys., 2017, 125: 663.

[42] V. P.Shevelko and L. A.Vainshtein, Atomic Physics for Hot Plasmas (IOP Publishing, Bristol, 1993).

[43] L. A.Vainshtein and V. P.Shevelko, Program ATOM, Preprint No. 43, Lebedev Physical Institute, Moscow1996.

[44] L. A.Vainshtein, Proc. P. N. Lebedev Inst.119, 3 (1980).

[45] F.Petitdemange and F. B.Rosmej, “Dielectronic satellites and Auger electron heating: Irradiation of solids by intense XUV-free electron laser radiation,” in New Trends in Atomic & Molecular Physics: Advanced Technological Applications, edited by M.Mohan (Springer, 2013), Vol. 76, pp. 91114, ISBN: 978-3-642-38166-9.

[46] F. B. Rosmej. Diagnostic properties of Be-like and Li-like satellites in dense transient plasmas under the action of highly energetic electrons. J. Quant. Spectrosc. Radiat. Transfer, 1994, 51: 319.

[47] F. B. Rosmej. A new type of analytical model for complex radiation emission of hollow ions in fusion and laser produced plasmas. Europhys. Lett., 2001, 55: 472.

[48] F. B. Rosmej. An alternative method to determine atomic radiation. Europhys. Lett., 2006, 76: 1081.

[49] F. B.Rosmej, “X-ray emission spectroscopy and diagnostics of non-equilibrium fusion and laser produced plasmas,” in Highly Charged Ion Spectroscopic Research, edited by Y.Zou and R.Hutton (Taylor and Francis, 2012), pp. 267341, ISBN: 9781420079043.

[50] V. A. Astapenko, X. Li, F. B. Rosmej. An analytical plasma screening potential based on the self-consistent-field ion-sphere model. Phys. Plasmas, 2019, 26: 033301.

[51] X. Li, F. B. Rosmej. Analytical approach to level delocalization and line shifts in finite temperature dense plasmas. Phys. Lett. A, 2020, 384: 126478.

[52] H. A.Bethe and E. E.Salpeter, Quantum Mechanics of One- and Two-Electron Atoms (Plenum Publishing, New York, 1977);J. D.Hey, “On the role of atomic metastability in the production of Balmer line radiation from cold atomic hydrogen, deuterium and hydrogenic ion impurities in fusion edge plasmas,” J. Phys. B: At., Mol. Opt. Phys.45, 065701 (2012).

[53] J. Davis, V. L. Jacobs. Effects of plasma microfields on radiative transitions from atomic levels above the ionization threshold. Phys. Rev. A, 1975, 12: 2017.

[54] J. Davis, V. L. Jacobs, P. C. Kepple. Enhancement of dielectronic recombination by plasma electric microfields. Phys. Rev. Lett., 1976, 37: 1390.

[55] J. Davis, V. L. Jacobs. Properties of Rydberg autoionizing states in electric field. Phys. Rev. A, 1979, 19: 776.

[56] M. F.Gu, “The flexible atomic code FAC,” Can. J. Phys.86(5), 675 (2008);10.1139/p07-197I. P.Grant and N. C.Pyper, “Breit interaction in multi-configuration relativistic atomic calculations,” J. Phys. B: A., Mol. Phys.9, 761 (1976).

[57] L. A. Bureyeva, T. Kato, V. S. Lisitsa. Quasiclassical representation of autoionization decay reates in parabolic coordinates. J. Phys. B: At., Mol. Opt. Phys., 2001, 34: 3909.

[58] L. A. Bureyeva, T. Kato, V. S. Lisitsa, et al.. Quasiclassical theory of dielectronic recombination in plasmas. Phys. Rev. A, 2002, 65: 032702.

[59] F.Robicheaux and M. S.Pindzola, “Enhanced dielectronic recombination in crossed electric and magnetic fields,” Phys. Rev. Lett.79, 2237 (1997);10.1103/physrevlett.79.2237J. D.Hey, “On the Runge-Lenz-Pauli vector operator as an aid to the calculations of atomic processes in laboratory and astrophysical plasmas,” J. Phys. B: At., Mol. Opt. Phys.48, 185701 (2015);10.1088/0953-4075/48/18/185701J. D.Hey, “On the use of the axially symmetric paraboloidal coordinate system in deriving some properties of Stark states of hydrogenic atomc and ions,” J. Phys. A: Math. Theor.52, 045203 (2019).

[60] P. Gombas. Erweiterung der statistischen theroy des atoms. Z. Phys., 1943, 121: 523.

[61] P.Gombas, Die statistische theorie des Atoms und ihre Anwendungen (Springer-Verlag, Wien, 1949).

[62] P. Gombás. Present state of the statistical theory of atoms. Rev. Mod. Phys., 1963, 35: 512.

[63] C. Deutsch, P. Fromy, G. Maynard. Thomas-Fermi-like and average atom models for dense and hot matter. Phys. Plasmas, 1996, 3: 714.

[64] E. H. Lieb, B. Simon. The Thomas-Fermi theory of atoms, molecules and solids. Adv. Math., 1977, 23: 22.

[65] G. Kemister, S. Nordholm. A radially restricted Thomas-Fermi theory for atoms. J. Chem. Phys., 1982, 76: 5043.

[66] A. V. Demura, M. B. Kadomtsev, V. S. Lisitsa. Universal statistical approach to radiative and collisional processes with multielectron ions in plasmas. High Energy Density Phys., 2015, 15: 49.

[67] A. Sommerfeld. Integrazione asintotica dell’equazione differentiale di Thomas–Fermi. Rend. R. Accad. Lincei, 1932, 15: 293.

[68] V. D. Kirillow, B. A. Trubnikov, S. A. Trushin. Role of impurities in anomalous plasma resistance. Sov. J. Plasma Phys., 1975, 1: 117.

[69] C. P. Balance, S. D. Loch, M. S. Pindzola, et al.. Dielectronic recombination of W35+. J. Phys. B: At., Mol. Opt. Phys., 2010, 43: 205201.

[70] Y. Fu, X. Ma, Z. Wu. Electronic impact excitation and dielectronic recombination of highly charged tungsten ions. Atoms, 2015, 3: 474.

[71] E. Behar, P. Mandelbaum, J. L. Schwob. Dielectronic recombination rate coefficients for highly-ionized Ni-like atoms. Phys. Rev. A, 1996, 54: 3070.

F. B. Rosmej, V. A. Astapenko, V. S. Lisitsa, L. A. Vainshtein. Dielectronic recombination in non-LTE plasmas[J]. Matter and Radiation at Extremes, 2020, 5(6): 064201.

本文已被 1 篇论文引用
被引统计数据来源于中国光学期刊网
引用该论文: TXT   |   EndNote

相关论文

加载中...

关于本站 Cookie 的使用提示

中国光学期刊网使用基于 cookie 的技术来更好地为您提供各项服务,点击此处了解我们的隐私策略。 如您需继续使用本网站,请您授权我们使用本地 cookie 来保存部分信息。
全站搜索
您最值得信赖的光电行业旗舰网络服务平台!