Next Article in Journal
Integrated Scheduling of Picking and Distribution of Fresh Agricultural Products for Community Supported Agriculture Mode
Previous Article in Journal
Efficient Sequential and Parallel Prime Sieve Algorithms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Singly Resonant Multiphoton Processes Involving Autoionizing States in the Be-like CIII Ion

by
Viorica Stancalie
National Institute for Laser, Plasma and Radiation Physics, Atomistilor 409, P.O. Box MG-36, 77125 Magurele-Ilfov, Romania
Symmetry 2022, 14(12), 2528; https://doi.org/10.3390/sym14122528
Submission received: 17 October 2022 / Revised: 23 November 2022 / Accepted: 29 November 2022 / Published: 30 November 2022
(This article belongs to the Section Physics)

Abstract

:
In this paper, we investigate the applicability of different theories on the intensity-dependent ionization rate for C2+ atomic targets at different laser wavelengths (frequency) and at linear polarization. We use the analytical formulas and draw conclusions, from numerical comparison with the results from ab initio ‘two-state model’ R-matrix Floquet calculation, on their correct predictions of the ionization rate. The single-photon ionization has been studied in the vicinity of the 1s2 (2Po)2pns (1Po), n = 5–12 autoionizing resonances at non-perturbative laser intensity. The results obtained from Perelomov–Popov–Terent’ev and Ammosov–Delone–Krainov models are compared in a region away from resonance where the two-state model description is not as good. To quantify the deviation between theoretical models, we analyze the ratio between different data sets as functions of the Keldysh parameter. We conclude that the results obtained with the model of Perelemov–Popov–Terent’ev are the closest to the ab initio R-matrix Floquet calculation.

1. Introduction

With the advance of laser technologies and the impact on the development of atomic physics, the study of atoms in strong laser fields has received special attention recently. The first such studies were based on the one-electron model of the atom. Reviews can be found in Refs. [1,2]. However, the electron correlation effect that occurs in atoms with two or more electrons in intense laser fields [3,4] have recently drawn some attention, with particular interest in the excitation of autoionizing resonances of the atom. While the time dependent Schrodinger equation allows one to follow the dynamics of the electron correlation during the laser pulse [5], the Floquet ansatz (see Ref. [6] for a review) is the ideal candidate for the study, using the adiabatic approach, of the atomic structure in the laser field.
Experiments on the ionization of atomic targets by intense pulses [7,8,9,10,11] have also been performed. The autoionizing states in the continuum manifest themselves as asymmetric peaks in the photon absorption spectra due to the mixing between their configurations. This phenomenon is called the Fano interference. In most cases, control over the Fano interference [12] has been achieved by an interplay between an intense high-frequency pump and a low-frequency probe pulse [13,14,15]. In modern laser facilities, where strong laser fields are produced with a duration comparable to the lifetime of the autoionization state, single- and multi-photon processes can be studied using the ab initio approaches [16,17]. For low-intensity electric fields, the autoionization states are defined using perturbation theory [18].
The electron and photon interactions with Be-like ions, including C2+, have been the subject of many detailed experimental and theoretical works. Two ionization channels attached to a 2p ionic state of C3+ are relevant for photoionization, namely 2pns(1Po) and 2pnd(1Po). Besides their spontaneous decay, the C2+ 1s22pns (1P1o) and 1s22pnd (1P1o) Rydberg states are also unstable against autoionization. Based on the combination of the Floquet theory and the R-matrix numerical method, ‘the R-matrix Floquet’ theory and code (RMF) [19], an impressive number of multiphoton processes have been studied, and a wide variety of interesting resonance effects in multiphoton ionization [20,21,22] have been predicted. One of them is the occurrence of laser-induced degenerate state (LIDS) phenomenon. During LIDS, the energies and widths of two states are made identical at certain laser frequencies and intensities [23]. The existence of LIDS has been observed in the RMF calculation for multiphoton transitions in C2+ and Al9+ [24]. In this earlier work, we carried out ab initio RMF calculation using the ‘two state model’ of Latinne et al. [23] when the two excited Rydberg states, namely the 1s22sns (1Se) and 1s22pn’s(1Po), Δn = n′ − n = 0, are resonantly coupled by one photon. We investigated this process in detail in the case of the carbon ion [25] and found the existence of LIDS in the region in which Δn = n′ − n = 2. The numerical results obtained have shown a good agreement between the two-state model and the full RMF calculations near the resonance frequency, but further away from this the agreement is less. Several quasi-energy spectra of the above-mentioned excited Rydberg ion states in laser fields have been compiled in our earlier works.
The main goal of this work is to compare results on the intensity-dependence ionization rate obtained from three theoretical models. The motivation is twofold. First, the Be-like C2+ ion will be the test case for a future pump-probe experiment based on the use of a table-top XUV laser in combination with a storage ring at GSI [26]. Secondly, because the two-state model used in our earlier analysis does not include the electron correlation effect and the nearby resonances, it gives results that are less accurate compared to the full RMF calculation. Therefore, we proposed in this paper to find the closest analytical model that can be used as a starting point in the analysis of future experiments. In the present paper, we complete the previous data sets with a focus on the total (energy and angle-integrated) ionization rate of Be-like C ions, for the Keldysh parameter γ ~ 1 and γ >> 1, including the Coulomb-correction. We use the analytical formulas and draw conclusions, from numerical comparison with the ab initio RMF calculation results, on their correct predictions of the ionization rate in the case of the 1s2 (2Po)2pns (1Po) Rydberg states in Be-like C2+. Section 2 summarizes the earlier analysis and its achievements. Section 3 reviews the analytical ionization rate formulas and provides comparison between results from present and earlier ab initio calculations. In Section 4, we discuss the numerical results and present concluding remarks.
Except where indicated otherwise, atomic units are used throughout the paper

2. One-Photon Ionization of the Excited 2pns 1Po States in Be-like C III Ion

In this section, we present the main achievements of our earlier RMF studies. The essential physical assumptions made when considering the process of one-photon ionization have been presented elsewhere. We recall here the main steps that are relevant for this work. Starting with the 1s22s (2S) and 1s22p (2P0) target states of the remaining Li-like target ion, the 1s22s2(1S) and 1s22p2(1S) states of Be-like ions were obtained by a configuration interaction expansion using the atomic structure code CIV3 [27]. RMATXII codes [28,29] were used to output energy values. To calculate the 1s2(2Se)2sns (1Se) and 1s2 (2Po)2pns (1Po), 1s2 (2Po)2pnd (1Po), ‘bound’ and ‘excited’ Rydberg states, respectively, we used the quantum defect theory and the R-matrix numerical method. More details on the calculation, including quantum defects and energies, are given in Refs. [24,25,30,31]. In these works, attention was focused on the generation of LIDS involving the 2sns 1Se ‘bound’ excited Rydberg states, where the principal quantum number ranges from 5 to 12 and from 9 to 11 for C2+ and Al9+, respectively. We have shown that, by employing a single photon, these states can be made degenerate with the 2pns (1Po) states, when Δn = 0 and Δn = 2 for C2+, and Δn = 0 for Al9+.
The resonances were described in the ‘two state model’ to obtain the physical quantities as follows: position and width of the bound and autoionizing state, laser frequency at LIDS, Fano parameter. The laser intensity was varied only slightly above the first ionization threshold so that the atom remained in the Floquet eigenstate connected to the initial state. Detailed results are presented in Refs. [24,25]. In the present work, we complete previous data sets for laser frequency ω >> |E00)|, where |E00)| is the binding energy of the state of the atom in the field, and in which α0 can take any value [32,33]. In this high-frequency regime, and for a linearly polarized field, RMF results show that as α0 increases, the binding energy decreases. In Figure 1a,b, our results are plotted on the Floquet lifetime as a function of intensity, associated to the 1s22p10s (1Po) and 1s22p6s(1Po) Rydberg state, respectively, in C2+.The numbers adjacent to the points give the corresponding value of α0, the classical amplitude of the oscillation of a free electron driven by the field.
The RMF calculations were performed for different laser frequencies and intensities. We used the laser field strength sets of data to estimate the Keldysh parameter at the barrier suppression ionization (BSI). This model occurs when the laser intensity is higher than the so-called barrier-suppression-ionization intensity, IBSI = Ip4/16 (with Ip being the ionization potential). Table 1 shows the driving laser frequencies and the corresponding laser intensity ranges used in our RMF calculation. The Keldysh parameter, γ, values at the barrier-suppression-ionization (BSI) intensity, IBSI, denoted as γBSI, and at the minimum laser intensity, Imin, denoted as γmax, are also presented in the last two columns, respectively.

3. Strong Field Ionization Rates

In this section, we investigate the applicability of different theories on the intensity-dependent ionization rate to our atomic system. The theoretical treatment of ionization of an atom in the radiation field is based on the work of Keldysh [34]. The Keldysh parameter, γ = ω√2Ip/F (where Ip is the ionization potential and F and ω are the laser-field strength and frequency, respectively) characterizes the degree of adiabaticity of the motion of active electrons through the barrier. In the region of small frequencies (ω << ωt, where ωt is the so-called tunnelling frequency), in which adiabatic approximation is correct, the Keldysh parameter is γ = ω/ωt << 1. In this regime, only the initial and final states of the electron are significant, while the intermediate state places no rule. In this case, a simple ‘two state model’, similar to the one used in RMF calculations, can be applied. For energies of the emitted photon lower than the energy of the atomic state, and for laser strength lower than the atomic field strengths, the quasi-classic approximation becomes valid. Landau and Lifshitz [35] derived a formula for the ionization rate of hydrogen when the electron is in the ground state. The ionization rate is:
W ( t ) = 4 ( 2 I p ) 5 / 2 E ( t ) exp ( 2 ( 2 I p ) 3 / 2 3 E ( t ) )
where E(t) is the laser’s electric field. Other treatments based on the quasi-static limit have been derived by Keldysh and Ammosov–Delone–Krainov (ADK) [36,37]. According to Keldysh’s theory, photoionization in weak or strong field regimes, respectively, is controlled by the Keldysh parameter, the approach resulting in a rate, wK, expressed as the total ionization rate as a function of γ:
w K = Q ( γ , I p , ω ) exp ( ξ ( γ , I p , ω ) )
where Q (γ, Ip, ω) is a pre-exponential factor, and
ξ ( γ , I p , ω ) = 2 ω I p ( 1 + 1 2 γ 2 ) [ sinh 1 γ γ ( 1 + γ 2 ) 1 / 2 1 + 2 γ 2 ]
The pre-exponential factor in Equation (2) was obtained by Perelomov et al. [38,39] (the original PPT formula). For γ >> 1, Equations (2) and (3) lead to the scaling of the ionization rate depending on the intensity of the laser field and the minimum number of photons, K0, required for photoionization, IK0. When γ:<< 1, Equations (2) and (3) yield wK ∝ exp [−(4/3) (2)1/2Ip3/2E0−1 (1 − γ2/10) (see Equation (59) Ref. [36]). In this tunneling regime, these expressions describe the transmission of the electron’s wave functions through a triangular potential barrier formed by a rectangular potential step of height Ip and a dc electric field E0 [40].
The Keldysh, PPT, and ADK theories are for short-range potentials and a weak laser field. The ADK model is related to the Keldysh–Faisal–Reiss (KFR) formula [41,42] for the tunneling limit. For the ionization rate of complex atoms (ions) from a state with energy, E, with orbital quantum number, l, and its projection, m, the ADK formula is:
W A D K = C n * l 2 f ( l , m ) | I p | ( 3 F π ( 2 I p ) 3 / 2 ) 1 / 2 × ( 2 F ( 2 | I p | ) 3 / 2 ) 2 n * | m | 1 exp ( 2 ( 2 | I p | ) 3 / 2 3 F ) ,
where F is the electric field strength and
C n * l = ( 2 e n * ) n * ( 2 π n * ) 1 / 2 f ( l , m ) = ( 2 l + 1 ) ( l + | m | ) ! 2 | m | | m | ! ( l | m | ) !
In the above equations, n* = Z/(2Ip)−1/2, Z is the nuclear charge of the atomic residue, Ip is the ionization potential of the state, and e = 2.71828. The validity of the ADK rate is expected to be best for F << 1, n* >> 1 and a driving laser frequency of ω << |Ip|. The ADK ionization rate does not depend on the laser frequency, so it is used only for the instantaneous rate, which in turn is dominated by m = 0.
Based on the ‘two-state model’ and single photon ionization, Krainov [43,44] suggested an extension of ADK theory to incorporate BSI. The rate is:
W K r = 4 3 π n * F ( 2 F ) 1 / 3 ( 4 e ( | I p | ) 3 / 2 F n * ) 2 n * × 0 A i 2 ( x 2 + 2 | F | ( 2 F ) 3 / 2 ) x 2 d x
where Ai is the Airy function. This formula reduces to the ADK rate for a weak laser field (tunneling limit). In Figure 2 we plot, comparatively, the ionization rates versus laser intensity for ionization from the C2+ 1s22pns (1Po) Rydberg states, n = 6, 7, 9, 10. For each state there are two curves, one associated with the ADK model (black circles) and the other (red diamonds) associated with Krainov’s formula, respectively. Qualitatively, similar results are obtained from calculation. The ADK model underestimates the ionization rate by an order of magnitude compared to Krainov’s formula.
The original PPT ionization rate, including Coulomb correction [37], is expressed as a sum over partial rates:
W P P T ( F , ω ) = n ν w n ( F , ω )
where F and ω are the amplitude and frequency of the laser field, respectively. The partial rate, wn, is the probability of ionization with the absorption of n quanta, from a minimum value, ν, required to reach the effective ionization threshold. The PPT method can be applied under the condition F << F0, ω << ω0, with ω0 = k2/2 and F0 = k3 being the atomic quantities. In the case of ionization by a constant field and adiabatic approximation, the PPT formula for the probability of ionization is in the form (see Equation (4) in Ref. [37]:
w ( F , ω ) = ( 3 F / π F 0 ) 1 / 2 w s t a t ( F )
where F and ω are the external field strength and frequency, and F0 is the intra-atomic electric field strength. We have compared the PPT results for ionization by a constant field with the ab initio RMF numerical values. Figure 3 presents, as an example, the calculated rate of ionization from the 1s22p8s(1Po) Rydberg state by a constant field of intensity of I = 1·1012 W/cm2.
It has been mentioned that, when γ >> 1, scaling of the ionization rate with the incident laser-field intensity, I, is IK0 (with K0 = [Ip/ω]). The PPT calculation carried out for all atomic states considered in this work led to the value K0 = 1, confirming the number of Floquet blocks used in the RMF calculation. We exemplify this dependence in Figure 4 for the ionization from the 1s22p8s(1Po) Rydberg state at a constant field intensity of 1 TW/cm2.
The PPT formulas for w(F,ω) is applicable not only to the calculation of ionization from the ground state, but also from the excited states. Moreover, the PPT formula is applicable under the condition τ >> Tt (τ is the lifetime of the excited state, and Tt ~ 1/ωt), from which it follows that this condition is satisfied by the range of values of field intensities when our investigation is performed. The original PPT formula, with the Coulomb correction factor (2/Fn*3) used in Figure 3, gives less accurate results. A new expression of the Coulomb correction factor [45] in the form (2/Fn*3) (1 + 2e−1γ)−2n*3, e = 2. 718. (here F is the reduced electric field), which is applicable to any γ, led us to results close to the ab initio calculation for the corresponding laser intensity ranges used in our RMF calculation. In Figure 5a, we compare the results obtained with this ‘generalized PPT formula’ for the field of a plane-polarized wave (see Equation (54), Ref. [37]), the original PPT formula, and the results from RMF calculation. The graph refers to the calculated ionization rate vs. intensity at a driving laser frequency of ω = 0.3085 au for the 1s22p5s(1Po) Rydberg state.
To quantify the deviation between the PPT, ADK, and the RMF theoretical approaches, we used the ratio between the corresponding ionization rate for each two data sets, and the results are plotted versus γ in Figure 5b and Figure 5c, respectively. We conclude that, for our atomic system, the results obtained with the ‘generalized’ model of Perelemov–Popov–Terent’ev are the closest to the ab initio R-matrix Floquet calculation.
In Figure 6a, we compare the frequency dependence of the ionization rate for the 1s22p7s 1Po state, as outputs from the PPT, generalized PPT, and RMF calculations. In this figure, the RMF and generalized PPT calculations indicate similar rate values for the driving laser frequency (ω = 0.2982 au) close to the resonance frequency of 0.29885. The adiabaticity parameter is also indicated. The corresponding laser intensity (see Figure 6b) is 1010 W/cm2, which is close to the estimated LIDS intensity [24]. Both curves display a sharp variation with increasing laser intensity (Figure 6b), as the laser frequency exceeds the resonance frequency. We plot the ratio between the two data sets versus γ in Figure 6c. The ionization rate vs. laser intensity and γ ~ 1 is shown in Figure 6d.

4. Discussion

In this article, we made use of our earlier reported works to extract those parameters that are important for the present investigation: the adiabaticity parameter values at the barrier suppression-ionization, and the maximum of the Keldysh parameter at the minimum laser intensity used in the ab initio calculation. The polarizability of these ions has also been previously reported [46]. Another reported result of this work is the lifetime of studied autoionizing Rydberg states as a function of laser intensity.
The present investigation revealed that the use of the PPT formula with Coulomb correction factor, valid for arbitrary values of γ, leads to values of ionization rates very close to those obtained from the ab initio calculation, near the resonance frequency. The closer the ratio is to 1, the better is the agreement between all these sets of data. This closeness of the results was to be expected considering the similarity of the hypotheses of both theoretical models: the single active electron approximation and ionization in a linearly polarized field. The fact that at intensities above the LIDS intensity this agreement is no longer valid is motivated by the essence of the LIDS phenomenon. At this laser intensity/frequency, the two states have identical energies and widths. At higher intensities, a real part of the energy remains adiabatically close to the real axis for one of these coupled states, while the autoionizing state has an imaginary part of the energy with a difference of zero. In addition, in the previous calculation, the LIDS frequency was close to the tuning frequency, given the small difference between the detuning and the ‘atomic’ frequency (which in this two state model is defined as the binding energy difference between the two states when the laser intensity goes to zero). From the preliminary study (not reported in this article) on another ion belonging to the beryllium isoelectronic sequence, we observed a similar behavior of ionization rates at laser intensities close to LIDS intensity. This study is in progress.
The ADK model has good agreement with the PPT for small γ values, but deviation becomes significant as γ approaches 1 and are of other orders of magnitude when γ is large.
Data sets are available from the author upon reasonable requests.

Funding

This research was supported by the Institute of Atomic Physics, Romania, Project Number FAIR-01/2020: Atomic Interaction in Supercritical Field: Contribution to the GPAC E129 Experiment at ESR’.

Data Availability Statement

Not applicable.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Gavrila, M. (Ed.) Atoms in Intense Laser Fields; Academic Press: New York, NY, USA, 1992. [Google Scholar]
  2. Protopapas, M.; Kietel, C.H.; Knight, P.L. Atomic physics with super-high intensity lasers. Rep. Progr. Phys. 1997, 60, 389–486. [Google Scholar] [CrossRef]
  3. Lambropoulos, P.; Zoller, P. Autoionizing states in strong laser fields. Phys. Rev. A 1981, 24, 379–397. [Google Scholar] [CrossRef]
  4. Rzazewski, K.; Eberly, J.H. Confluence of Bound-Free Coherences in Laser-Induced Autoionization. Phys. Rev. Lett. 1981, 47, 408. [Google Scholar] [CrossRef]
  5. Watson, J.B.; Sanpera, A.; Lappas, D.G.; Burnett, K.H.; Knight, P.L. Nonsequential Double Ionization of Helium. Phys. Rev. Lett. 1997, 78, 1884–1888. [Google Scholar] [CrossRef] [Green Version]
  6. Chu, S.-I. Recent Developments in Semiclassical Floquet Theories for Intense-Field Multiphoton Processes. Adv. At. Mol. Phys. 1985, 21, 197–253. [Google Scholar]
  7. Laarmann, T.; de Castro, A.R.B.; Gürtler, P.; Laasch, W.; Schulz, J.; Wabnitz, H.; Möller, T. Interaction of argon clusters with high intense VUV -laser radiation: The role of electronic structure in the energy-deposition process. Phys. Rev. Lett. 2004, 92, 143401. [Google Scholar] [CrossRef] [Green Version]
  8. Moshammer, R.; Jiang, Y.; Foucar, L.; Rudenko, A.; Ergler, T.; Schröter, C.D.; Lüdemann, S.; Zrost, K.; Fischer, D.; Titze, J.; et al. Few-photon multiple ionization of Ne and Ar by strong free-electron -laser-pulses. Phys. Rev. Lett. 2007, 98, 203001. [Google Scholar] [CrossRef] [Green Version]
  9. Sansone, G.; Benedetti, E.; Calegari, F.; Vozzi, C.; Avaldi, L.; Flammini, R.; Poletto, L.; Villoresi, P.; Altucci, C.; Velotta, R.; et al. Isolated single-cycle attoseconde pulses. Science 2006, 314, 443–446. [Google Scholar] [CrossRef]
  10. Krausz, F.; Ivanov, M. Attosecond physics. Rev. Mod. Phys. 2009, 81, 163–234. [Google Scholar] [CrossRef] [Green Version]
  11. Fushitani, M. Applications of pump-probe spectroscopy. Annu. Rep. Sect. C 2008, 104, 272–297. [Google Scholar] [CrossRef]
  12. Fano, U. Effects of Configuration Interaction on Intensities and Phase Shifts. Phys. Rev. 1961, 124, 1866. [Google Scholar] [CrossRef]
  13. Chen, S.; Wu, M.; Gaarde, M.B.; Schafer, K.J. Quantum interference in attosecond transient absorption of laser-dressed helium atoms. Phys. Rev. A 2013, 87, 033408. [Google Scholar] [CrossRef] [Green Version]
  14. Chen, S.; Wu, M.; Gaarde, M.B.; Schafer, K.J. Laser-imposed phase in resonant absorption of an isolated attosecond pulse. Phys. Rev. A 2013, 88, 033409. [Google Scholar] [CrossRef]
  15. Chen, S.; Bell, M.J.; Beck, A.R.; Mashiko, H.; Wu, M.; Pfeiffer, A.N.; Gaarde, M.B.; Neumark, D.M.; Leone, S.R.; Schafer, K.J. Light-induced states in attosecond transient absorption spectra of laser-dressed helium. Phys. Rev. A 2012, 86, 063408. [Google Scholar] [CrossRef] [Green Version]
  16. Artemyev, A.N.; Cederbaum, L.S.; Demekhin, P.V. Impact of intense laser pulses on the autoionization dynamics of the 2s2p doubly excited state of He. Phys. Rev. A 2017, 96, 033410. [Google Scholar] [CrossRef] [Green Version]
  17. Artemyev, A.N.; Muller, A.D.; Hochstuhl, D.; Cederbaum, L.S.; Demekhin, P.V. Dynamic interference in the photoionization of He by coherent intense high-frequency laser pulses: Direct propagation of the two-electron wave packets on large spatial grids. Phys. Rev. A 2016, 93, 043418. [Google Scholar] [CrossRef] [Green Version]
  18. Henneberger, W.C. Perturbation method for atoms in intense light beams. Phys. Rev. Lett. 1968, 21, 838–841. [Google Scholar] [CrossRef]
  19. Burke, P.G.; Francken, P.; Joachain, C.J. R-matrix Floquet theory of multiphoton processes. J. Phys. B At. Mol. Opt. Phys. 1991, 24, 761–790. [Google Scholar] [CrossRef]
  20. Purvis, J.; Dörr, M.; Terao-Dunseath, M.; Joachain, C.J.; Burke, P.G.; Noble, C.J. Multiphoton ionization of H- and He in intense laser fields. Phys. Rev. Lett. 1993, 71, 3943–3947. [Google Scholar] [CrossRef] [PubMed]
  21. Dörr, M.; Purvis, J.; Terao-Dunseath, M.; Burke, P.G.; Noble, C.J. R-matrix Floquet theory of multiphoton processes. V. Multiphoton detachment of the negative hydrogen ion. J. Phys. B At. Mol. Opt. Phys. 1995, 28, 4481–4500. [Google Scholar] [CrossRef]
  22. Fearnside, A.S. Intensity dependence of resonances profiles in multiphoton partial detachment rates of H. J. Phys. B At. Mol. Opt. Phys. 1998, 31, 275–288. [Google Scholar] [CrossRef]
  23. Latinne, O.; Klistra, N.J.; Dörr, M.; Purvis, J.; Terao-Dunsheath, M.; Joachain, C.J.; Burke, P.G.; Noble, C.J. Laser-Induced Degeneracies Involving Autoionizing States in Complex Atoms. Phys. Rev. Lett. 1995, 74, 46–49. [Google Scholar] [CrossRef] [PubMed]
  24. Stancalie, V. 1s22pns(1Po) autoionizing levels in Be-like Al and C ions. Phys. Plasmas 2005, 12, 043301. [Google Scholar] [CrossRef]
  25. Stancalie, V. Complements to nonperturbative treatment of radiative damping effect in dielectronic recombination: Δn = 2 transition in C IV. Phys. Plasmas 2005, 12, 100705. [Google Scholar] [CrossRef]
  26. Rothhardt, J.; Bilal, M.; Beerwerth, R.; Volotka, A.V.; Hilbert, V.; Sttohlker, T.; Fritzsche, S.; Limpert, J. Ultrashort-lived Excited States in Be-like ions. X-ray Spectrom. 2019, 49, 165–168. [Google Scholar] [CrossRef]
  27. Hibbert, A. CIV3-A general program to calculate configuration interaction wave functions and electric-dipole oscillator strengths. Comput. Phys. Commun. 1975, 9, 141–172. [Google Scholar] [CrossRef]
  28. Burke, V.M. A new method for calculating angular integrals in electron-atom scattering. Comput. Phys. Commun. 1998, 114, 210–224. [Google Scholar] [CrossRef]
  29. Berrington, K.A.; Eissner, W.B.; Norrington, P.H. RMATRXI: Belfast atomic R-matrix codes. Comput. Phys. Commun. 1995, 92, 290–420. [Google Scholar] [CrossRef]
  30. Stancalie, V. State selective photo-recombination cross sections in Be-like C and Al ions. Eur. Phys. J. D 2013, 67, 223. [Google Scholar] [CrossRef]
  31. Stancalie, V. Photoionization dynamics of C2+ ion in Rydberg states. Eur. Phys. J. D 2014, 68, 349. [Google Scholar] [CrossRef]
  32. Pont, M.; Gavrila, M. Stabilization of atomic hydrogen in superintense, high-frequency laser fields of circular polarization. Phys. Rev. Lett. 1990, 65, 2362–2365. [Google Scholar] [CrossRef] [PubMed]
  33. Gavrila, M.; Kaminski, J.Z. Free-free transitions in intense high-frequency laser fields. Phys. Rev. Lett. 1984, 52, 614–616. [Google Scholar] [CrossRef]
  34. Keldysh, L.V. Ionization in the field of a strong electromagnetic wave. Sov. Phys. JETP 1964, 47, 1945. [Google Scholar]
  35. Landau, L.; Lifchitz, E. Mécanique Quantique, Théorie Non Relativiste, 2nd ed.; Èditions Mir: Moscow, Russia, 1967; pp. 319–327. [Google Scholar]
  36. Ammosov, M.V.; Delone, N.B.; Krainov, V.P. Tunnel ionization of complex atoms and of atomic ions in an alternating electromagnetic field. Zh. Eksp. Teor. Fiz. 1986, 64, 1191–1194. [Google Scholar]
  37. Ilkov, F.A.; Decker, J.E.; Chin, S.L. Ionization of atoms in the tunneling regime with experimental evidence using Hg atoms. J. Phys. B At. Mol. Opt. Phys. 1992, 25, 4005–4020. [Google Scholar] [CrossRef] [Green Version]
  38. Perelomov, A.M.; Popov, V.S.; Terent’ev, M.V. Ionization of atoms in alternating electric field I. Sov. Phys. JETP 1966, 23, 924–934. [Google Scholar]
  39. Perelomov, A.M.; Popov, V.S. Ionization of atoms in alternating electric field III. Sov. Phys. JETP 1967, 25, 336–343. [Google Scholar]
  40. Zheltikov, A.M. Keldysh parameter, photoionization adiabaticity, and the tunneling time. Phys. Rev. A 2016, 94, 043412. [Google Scholar] [CrossRef] [Green Version]
  41. Faisal, F.H.M. Multiple absortion of laser photons by atoms. J. Phys. B At. Mol. Phys. 1973, 6, L89–L92. [Google Scholar] [CrossRef]
  42. Reiss, H.R. Effect of an intense electromagnetic field on a weakly bound system. Phys. Rev. A 1980, 22, 1786–1813. [Google Scholar] [CrossRef]
  43. Krainov, V.P. Ionization rates and energy and angular distributions at the barrier-suppression ionization of complex atoms and atomic ions. J. Opt. Soc. Am. B 1997, 14, 425–431. [Google Scholar] [CrossRef]
  44. Delone, N.B.; Krainov, V.P. Tunneling and barrier-suppression ionization of atoms and ions in a laser radiation field. Phys. Usp. 1998, 41, 469–474. [Google Scholar] [CrossRef]
  45. Popruzhenko, S.V.; Mur, V.D.; Popov, V.S.; Bauer, D. Strong field ionization rate for arbitrary laser frequencies. Phys. Rev. Lett. 2008, 101, 193003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Stancalie, V. Static and dynamic polarizability for C2+ in Rydberg states. AIP Adv. 2015, 5, 077186. [Google Scholar] [CrossRef]
Figure 1. The Floquet lifetime (fs) as a function of laser intensity (au) associated with: (a) 1s22p10s 1Po, (b) 1s22p6s 1Po Rydberg state, respectively. The numbers adjacent to the points give the corresponding value of α0, the classical amplitude of the oscillation of a free electron driven by the field.
Figure 1. The Floquet lifetime (fs) as a function of laser intensity (au) associated with: (a) 1s22p10s 1Po, (b) 1s22p6s 1Po Rydberg state, respectively. The numbers adjacent to the points give the corresponding value of α0, the classical amplitude of the oscillation of a free electron driven by the field.
Symmetry 14 02528 g001
Figure 2. Calculated ionization rate (in atomic units, au) of the 1s22pns (1Po), n = 6, 7, 9, 10, Rydberg states of C2+ as a function of laser intensity (in TW/cm2). There are two curves associated with each state, black and red, corresponding to the ADK and Krainov formulas; n = 10 (circles), n = 9 (diamond), n = 7 (triangle up), n = 6 (triangle down), respectively.
Figure 2. Calculated ionization rate (in atomic units, au) of the 1s22pns (1Po), n = 6, 7, 9, 10, Rydberg states of C2+ as a function of laser intensity (in TW/cm2). There are two curves associated with each state, black and red, corresponding to the ADK and Krainov formulas; n = 10 (circles), n = 9 (diamond), n = 7 (triangle up), n = 6 (triangle down), respectively.
Symmetry 14 02528 g002
Figure 3. Frequency (ω, in au) dependence of the ionization rate of 1s22p8s(1Po) Rydberg state: ab initio RMF Floquet (red dotted line), PPT formula (black dotted line), PPT generalized formula (green dotted line). The adiabaticity parameter, γ << 1, is also indicated in the graph.
Figure 3. Frequency (ω, in au) dependence of the ionization rate of 1s22p8s(1Po) Rydberg state: ab initio RMF Floquet (red dotted line), PPT formula (black dotted line), PPT generalized formula (green dotted line). The adiabaticity parameter, γ << 1, is also indicated in the graph.
Symmetry 14 02528 g003
Figure 4. Intensity dependence as IK0 of the ionization rate from 1s22p8s(1Po) Rydberg state. Field intensity is 1·1012 W/cm2.
Figure 4. Intensity dependence as IK0 of the ionization rate from 1s22p8s(1Po) Rydberg state. Field intensity is 1·1012 W/cm2.
Symmetry 14 02528 g004
Figure 5. (a) Calculated ionization rate from the C2+ 1s22p5s(1Po) Rydberg state, as a function of intensity at a driving laser frequency of ω = 0.3085 au: RMF calculation (green dotted line); PPT calculation with generalized version of Coulomb correction factor in Ref. [45] (red solid line); PPT calculation (black solid line) with original version of Coulomb correction factor (see text). (b) The ratio between theoretical PPT and RMF values vs. γ; (c) The ratio between theoretical, ADK, and RMF values, vs. γ.
Figure 5. (a) Calculated ionization rate from the C2+ 1s22p5s(1Po) Rydberg state, as a function of intensity at a driving laser frequency of ω = 0.3085 au: RMF calculation (green dotted line); PPT calculation with generalized version of Coulomb correction factor in Ref. [45] (red solid line); PPT calculation (black solid line) with original version of Coulomb correction factor (see text). (b) The ratio between theoretical PPT and RMF values vs. γ; (c) The ratio between theoretical, ADK, and RMF values, vs. γ.
Symmetry 14 02528 g005
Figure 6. (a) Frequency (in au) dependence of the ionization rate (in au.) at 1010 W/cm2 for the 1s22p7s 1Po Rydberg state; PPT (black circles); PPT generalized formula (blue circles); RMF calculation (green circles); (b) Ionization rate (in au) vs. laser intensity (in TW/cm2) for the 1s22p7s 1Po Rydberg state at a driving laser frequency of ω = 0.2985 au; (PPT (red dotted line), RMF (green dotted line); (c) The ratio between theoretical (PPT and RMF) values versus γ. (d) The generalized PPT formula for the ionization rate vs. laser intensity and γ ~ 1.
Figure 6. (a) Frequency (in au) dependence of the ionization rate (in au.) at 1010 W/cm2 for the 1s22p7s 1Po Rydberg state; PPT (black circles); PPT generalized formula (blue circles); RMF calculation (green circles); (b) Ionization rate (in au) vs. laser intensity (in TW/cm2) for the 1s22p7s 1Po Rydberg state at a driving laser frequency of ω = 0.2985 au; (PPT (red dotted line), RMF (green dotted line); (c) The ratio between theoretical (PPT and RMF) values versus γ. (d) The generalized PPT formula for the ionization rate vs. laser intensity and γ ~ 1.
Symmetry 14 02528 g006
Table 1. Driving laser frequency in C2+, ωtune (in atomic units, au), intensity I (W/cm2) ranges (Imin-Imax) of theoretical RMF data. The γ values at the barrier-suppression-ionization (BSI) intensity IBSI and at the minimum laser intensity, Imin, are denoted as γBSI, and γmax, respectively.
Table 1. Driving laser frequency in C2+, ωtune (in atomic units, au), intensity I (W/cm2) ranges (Imin-Imax) of theoretical RMF data. The γ values at the barrier-suppression-ionization (BSI) intensity IBSI and at the minimum laser intensity, Imin, are denoted as γBSI, and γmax, respectively.
1s22pns (1Po)
n
ωtune
(au)
Imin
(TW/cm2)
Imax
(TW/cm2)
γBSIγmax
50.30890.110.00.04250.2629
60.297850.4510.00.07050.0949
70.298850.011.00.11550.5459
80.298770.011.00.17570.4759
90.298620.0030.90.25330.7685
100.298490.0050.20.35050.5328
110.298390.0030.10.46990.6227
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Stancalie, V. Singly Resonant Multiphoton Processes Involving Autoionizing States in the Be-like CIII Ion. Symmetry 2022, 14, 2528. https://doi.org/10.3390/sym14122528

AMA Style

Stancalie V. Singly Resonant Multiphoton Processes Involving Autoionizing States in the Be-like CIII Ion. Symmetry. 2022; 14(12):2528. https://doi.org/10.3390/sym14122528

Chicago/Turabian Style

Stancalie, Viorica. 2022. "Singly Resonant Multiphoton Processes Involving Autoionizing States in the Be-like CIII Ion" Symmetry 14, no. 12: 2528. https://doi.org/10.3390/sym14122528

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop