Home Search Collections Journals About Contact us My IOPscience Keldysh theory of strong field ionization: history, applications, difficulties and perspectives This content has been downloaded from IOPscience. Please scroll down to see the full text. 2014 J. Phys. B: At. Mol. Opt. Phys. 47 204001 (http://iopscience.iop.org/0953-4075/47/20/204001) View the table of contents for this issue, or go to the journal homepage for more Download details: IP Address: 176.9.124.142 This content was downloaded on 15/10/2014 at 08:30 Please note that terms and conditions apply. Journal of Physics B: Atomic, Molecular and Optical Physics J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 (35pp) doi:10.1088/0953-4075/47/20/204001 Review Article Keldysh theory of strong field ionization: history, applications, difficulties and perspectives S V Popruzhenko National Research Nuclear University MEPhI, Kashirskoe shosse 31, 115409, Moscow, Russian Federation E-mail: [email protected] Received 4 April 2014, revised 16 June 2014 Accepted for publication 30 June 2014 Published 8 October 2014 Abstract The history and current status of the Keldysh theory of strong field ionization are reviewed. The focus is on the fundamentals of the theory, its most important applications and those aspects which still raise difficulties and remain under discussion. The Keldysh theory is compared with other nonperturbative analytic methods of strong field atomic physics and its important generalizations are discussed. Among the difficulties, the gauge invariance problem, the tunneling time concept, the conditions of applicability and the application of the theory to ionization of systems more complex than atoms, including molecules and dielectrics, are considered. Possible prospects for the future development of the theory are also discussed. Keywords: intense laser fields, nonlinear ionization, Keldysh theory, optical tunneling, strong field approximation, coulomb effects, simple man model (Some figures may appear in colour only in the online journal) 1. Introduction theory and several productive generalizations of the original Keldysh approach. Second, despite the very broad and longterm use of the theory, several important conceptual questions remain open and are being actively debated in the literature. What is the difference between the Keldysh theory, the KFR theory and the strong field approximation (SFA)? How is the Keldysh model connected to the simple-man model (SMM) of ionization? Why is the theory, in general, gauge-noninvariant and how can a noninvariant theory be used for the calculation of observables? Whatis the accuracy of the theory and which parameters define its applicability conditions? What is the physical interpretation of the Keldysh tunneling time; can this quantity be measured in experiments? Is the Keldysh theory actually applicable to the description of the ionization of dielectrics and other spatially extended systems? These questions are answered in different ways in the literature and some of them do not yet have any answer. Here we try to sort such conceptual questions into two groups: those with a known correct answer and those that still remain unclear. In This review paper is dedicated to the 50th anniversary of Keldyshʼs seminal work [1], where a nonperturbative approach to the description of the nonlinear ionization of atoms and dielectrics by intense electromagnetic fields was pioneered. Now known as the Keldysh theory (often also referred to as the Keldysh model, the Keldysh ionization ansatz or the Keldysh theory of optical tunneling) or the Keldysh–Faisal–Reiss (KFR) [2, 3] theory, it is being routinely used for the description of multiquantum processes induced by intense laser radiation. Apart from giving a historic overview of the theory, this paper has two aims. First, the development of the method proposed by Keldysh has generated a bulk of literature which is difficult to navigate, particularly for those researchers beginning their work in the field. Therefore, a sort of manual to guide researchers through the theory and its applications would be useful. Having this purpose in mind, we describe here the main results of the 0953-4075/14/204001+35$33.00 1 © 2014 IOP Publishing Ltd Printed in the UK J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article the first case we will formulate the solution, in the second we will present the most essential controversial viewpoints. In order to keep the length of this paper reasonably limited, we leave out those aspects of the theory which have already received detailed and clear consideration in the literature. This includes relativistic and Lorentz ionization, spin effects (see reviews [4, 5] and references therein) and ionization by ultrashort laser pulses, reviewed in [4, 6]. Ionization of molecules is considered only briefly in connection with the gauge invariance problem. We will also not discuss recollision phenomena, i.e. high-energy above-threshold ionization, generation of high-order harmonics and nonsequential ionization. For overview of this currently thriving area of strong field physics, we direct the reader to the reviews [7–10]. For routine use of the Keldysh theory, one needs reliably checked analytic expressions for photoelectron momentum distributions, angular distributions, spectra and rates calculated for standard cases (including monochromatic fields, elliptical polarization, ionization from states with an arbitrary angular momentum, etc). For the most frequently used cases we have tried to collect such expressions in this review. More formulas can be found in the early papers by Nikishov and Ritus [11, 12], Perelomov, Popov and Terentʼev [13–15], in the pivotal paper by Reiss [3] (where, in particular, properties of the generalized Bessel functions are discussed in detail), in the paper by Grybakin and Kuchiev [16] and the review by Popov [4]. For the reader interested in the theory of ionization beyond the Keldysh approximation as well as in physics of other strong field phenomena, including both experiment and theory, we recommend the recent review [17] and the books [18–22]. This review is organized as follows. Section 2.1 is devoted to the formulation of the fundamentals of the Keldysh theory, including the nonlinear photoionization matrix element and general expressions for photoelectron momentum distributions. In section 2.2 we explain the terminology and introduce the dimensionless parameters which are important for further discussion. In section 3 a brief historic overview is given. Sections 4 and 5 review the current status of the theory. Subsection 4.1 contains a derivation of the matrix element. This derivation allows one to formulate the conditions of applicability (section 4.2) and to clarify the relationship between the Keldysh, KFR and SFA approaches (section 4.3). In section 5 we formulate the theory in terms of complex trajectories, achieving this via application of the saddle-point method (SPM; section 5.1), and introduce the imaginary time method (ITM; section 5.2). Section 6 is devoted to extensions of the Keldysh ionization model including the Coulomb–Volkov approximation (CVA; section 6.1), the Coulomb-corrected SFA (section 6.2) and the analytic R-matrix (ARM) method (section 6.3). In section 7 we discuss other efficient nonperturbative methods which can be reduced to the Keldysh theory or viewed as complementary. Section 8 is devoted to difficulties, including the problem of gauge invariance (section 8.1), the concept of the tunneling time and relation to classical models of ionization (section 8.2) and application of the Keldysh model to ionization of spatially extended systems (section 8.3). The final section contains a brief conclusion and outlook. We use the CGS (centimeter–gram–second) system in the following section, and atomic units = e = m = 1 with e and m being the absolute value of the electron charge and the electron mass, respectively, throughout the rest of the paper. 2. The basic equations of the Keldysh theory 2.1. Ionization amplitude and spectra According to the Keldysh ansatz, the transition probability amplitude between an atomic bound state and a continuum state specified by the value of the photoelectron momentum p measured at the detector is given by MK (p) = − +∞ i ∫−∞ Ψp Vint (t ) Ψ0 dt . (1) This expression is not present in [1], but its form can be deduced unambiguously from equation (8) of that paper. Here and below the subscript K denotes values defined within the Keldysh model; Ψ0 = ψ0 (r)e iIp t is the bound state wave function unperturbed by the laser field and having the ionization potential Ip, Ψp is the Volkov function [23–25] corresponding to the electron canonical momentum equal to p and Vint is the electron field interaction operator. In the nonrelativistic regime, the interaction operator and the Volkov function can be used in the dipole approximation by neglecting the spatial dependence and the magnetic component of the wave. An explicit form of these functions is determined by the gauge chosen for description of the electromagnetic wave. The most commonly used are the length gauge when Vint (r , t ) = e E (t ) r , Ψp (r , t ) = 1 exp (2π ) 3 2 − m 2 t (2) { ∫−∞ v2p ( i ⎡ ⎣ m vp (t ) r ⎤⎫ t ′ dt ′⎥⎬ ⎦⎭ ) (3) and the velocity gauge Vint (r , t ) = − Ψp (r , t ) = ie e2 2 A (t ) + A (t ), m 2m ⎧i ⎡ 1 m exp ⎨ ⎢ pr − 32 ⎣ ⎩ 2 (2π ) t (4) ⎤⎫ ∫−∞ v2p (t′)dt′⎥⎦ ⎬⎭.(5) ˙ (t ) is the electric field strength of the laser Here E (t ) = −A wave and A(t ) is the respective vector potential, both are spatially homogeneous in the dipole approximation and vp (t ) = (p + eA (t ) ) m (6) is the time-dependent electron velocity. In the field of a plane electromagnetic wave, p is conserved (in the nonrelativistic regime) and equal to the average (drift) electron momentum in the laser field and to the asymptotic momentum the electron has after the field is turned off. If the amplitude of ionization M (p) is known (not necessarily in the Keldysh approximation) the differential probability to find the photoelectron in the elementary volume 2 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article d3p near the momentum p is given by dW (p) = M (p) 2d3p . absorption of an integer number n of photons (7) ( W= ∫ 2 3 M (p) d p . n × R (p)dεp dOp, εp = MK (p, τ ) τ τ →∞ MK (p, τ ) = − i ∫0 R (p) = 2 n ⩾ Nmin = ⎡⎣ I p + UP ( d3p , τ Ψp Vint (t ) Ψ0 dt . ⎛ iT ⎡ p 2 ⎤⎞ = Ψp Vint Ψ0 (t ) exp ⎜⎜ ⎢ + I p + UP ⎥⎟⎟ ⎦⎠ ⎝ ⎣2m 2 = (10) pn2 T ≡ εp + UP. (11) E0 (cos ωt , − ξ sin ωt , 0). ω Kn (p) = − (12) For the ponderomotive energy, (11) gives UP = e2E 02 ( 1 + ξ ), 2 4 mω2 (13) ∑ k =−∞ Kn = 2π i +∞ 2iπkx e = ∑ δ (n − x ) (17) (18) = nω − I p − UP. (19) i ∫0 T dt Ψp Vint (t ) Ψ0 . (20) p = pn For a monochromatic field with arbitrary polarization, the amplitude (20) and the respective momentum distributions can be expressed via Bessel functions. Assuming the vector potential is in the form (12) and using the velocity gauge representation (4), (5), one obtains (after taking the time integral in (20) by parts) for a linearly polarized field, ξ = 0, where − 1 ⩽ ξ ⩽ 1 is the field ellipticity; ξ = 0 corresponds to linear and ξ = ±1 to right (left) circular polarization. Using (10) the time integral in (9) can be presented as a sum over the laser periods. Applying the relation +∞ ω⎤⎦ + 1, Equations(15)–(19) are model-independent and apply in any theory describing ionization in quasimonochromatic fields. Within the Keldysh model, the partial amplitude in (16) is given by The first term in the r.h.s. of (11) is the drift energy and the second is the ponderomotive (quiver) energy. In a monochromatic field elliptically polarized in the (x,y) plane A(t ) = ) with pn defined from the energy conservation law 2m T (16) dwn (n) = R ( p = pn n) ≡ R n (n) dOp 2 p e + A2 2m 2m 2 where Nmin is the minimal number of photons required for ionization, i.e. the ionization threshold. Correspondingly, peaks with n > Nmin are called above-threshold peaks and the regime of ionization when these peaks are present in the spectrum is known as above-threshold ionization (ATI). Angular distribution in a given peak is obtained by integration (15) over energy (9) where UP is the average quiver electron energy in the field. Using (6), we obtain for kinetic energy averaged over the laser period: 1 = (p + eA(t ))2 2m mω2p (p) 2π where dOp is the solid angle along the momentum direction and (p) is a contribution into the amplitude of ionization from a single laser period1. Photoelectron peaks in (15) exist for Ψp Vint Ψ0 (t + T ) T (15) with In a periodic field the integrand in (9) possesses the property: mv2 2 p2 2m (8) These two expressions are meaningful for laser pulses of finite duration, particularly for few-cycle pulses. For sufficiently long pulses containing a large number of optical periods—so that its electromagnetic field is close to a periodical function of time: E (t + T ) = E (t )—it is physically more appropriate to use probabilities per time unit (rates) instead of time-integrated values (7) and (8). The differential rate of ionization is defined as a limit dw K = lim ) dw (p) = ∑δ εp + I p + UP − nω This defines the momentum distribution of photoelectrons. The total probability of ionization (14) I p + p2 2 m ω ⎛ e E 0 p UP ⎞ ⎟. ψ0 (p) Jn ⎜ , ⎝ mω2 2ω ⎠ (21) n =−∞ 1 Note that the discreteness of the photoelectron spectrum (15) is a consequence of the periodicity of the classical external laser field and does not require for its explanation a concept of photons, in contrast to what is commonly stated. one can extract a delta-function responsible for energy conservation, so that the momentum distribution takes the form of a set of infinitely narrow peaks, each corresponding to the 3 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article Here the Fourier–Bessel expansion is used introduced in [1] and is known as the Keldysh parameter +∞ ∑ eiz sin φ = γ= Jk (z)eikφ 2mI p ω eE 0. (29) k =−∞ It can be interpreted as a ratio of the characteristic atomic momentum ℏκ = 2mI p to the field induced momentum pF = eE0 ω or, alternatively, as a ratio of the time the electron takes to cover the distance b 0 = I p eE0 moving with the atomic velocity vat = 2I p m . The value b0 is the width of a static barrier created by the field E0, so that τ = b 0 vat = γ ω is the time of flight under the barrier or the Keldysh tunneling time. The tunneling time concept in the Keldysh theory is discussed in section 8.2. The limit γ ≪ 1 is known in the literature as the tunneling domain. This characterization means that ionization proceeds similarly to the case of a static electric field when the electron tunnels through a time-independent potential barrier. The tunneling limit is also often mentioned as a quasistatic or adiabatic regime of ionization, underlining the fact that in this limit time can be treated as a parameter. As we will see below, in the tunneling limit, accurate expressions for the rate of ionization differ from those for the static field by small corrections ∼γ2. The very existence of a correct static limit is a standard benchmark for a theory of strong field ionization. The opposite limit γ ≫ 1 is known as the multiphoton ionization regime. Over the years the terminology has been evolving, so that identical or similar terms are sometimes used for the description of different ionization regimes. Namely, the tunneling limit, γ ≪ 1, is also often mentioned as a low-frequency limit which is not the same thing, according to (28). On the other hand, as soon as condition (28) is satisfied, many photons are needed to ionize the atom. It is, however, not appropriate to call this situation ‘multiphoton ionization’ because this term is reserved for the more specific case of γ ≫ 1. Instead, when (28) is satisfied, we will specify this as the multiquantum regime, in contrast to the few-quantum, K0 ∼ 1, or the single quantum, K0 < 1, regimes. Note that the value of the multiquantum parameter itself is insufficient for full characterization of an ionization regime, as the ionization problem is two-parametric [3]. From the parameters of the field and atom, Ip, E0 and ω, more than two dimensionless combinations can be constructed. Apart from K0 and γ there are another two frequently used in the theory. One is the ratio of the laser electric field amplitude to the characteristic electric field Ech of the respective atomic level and the generalized Bessel function is introduced [3, 26–28] +∞ Jn (a , b) = ∑ Jn + 2k (a) Jk (b). (22) k =−∞ For circularly polarized fields, ξ = ±1, the expression for the amplitude is simpler Kn = 2π i I p + p2 2 m ω ⎛ eE 0 p ⎞ ⎟. ψ0 (p) Jn ⎜ ⎝ mω2 ⎠ (23) Equations (21) and (23) contain the Fourier transform of the bound state wave function ψ0 (p) = 1 (2π )3 2 ∫ d3r e−ipr ψ0 (r). (24) For arbitrary ellipticity, the amplitude still has the form (21) with a more complex function instead of Jn (a , b ) (see, e.g. equation (45) in [29]). In the length gauge (2), the Fourier transform of the bound state wave function contains the time-dependent velocity vp (t ) in the argument, instead of p . As a result, the generalized Bessel expansion does not apply, excluding the special case when the electron is bound by the zero-range potential (ZRP) usually defined in 3D space as U (r ) = ∂ 2 π 2 δ (r ) r , κ = κm ∂r 2I p m (25) and the bound state wave function has the form2 κ e−κr . 2π r ψ0 ( r ) = (26) Then ψ0 (q) = κ π 3/2 1 κ + q2 2 (27) 2 and (I p + v 2p 2 m ) ψ0 (vp ) = (I p + p 2 2 m ) ψ0 (p), so that the amplitude calculated in the length gauge coincides exactly with (21). Thus for a system bound by zero-range forces, the Keldysh theory is gauge-invariant. There is further discussion of the gauge dependence in section 8.1. 2.2. Useful definitions and dimensionless parameters F = E 0 E ch, The Keldysh theory is usually applied to the description of nonlinear ionization in low-frequency electromagnetic fields. The term ‘low-frequency field’ means that the ionization potential Ip of an atom is essentially larger than the photon energy ω , i.e. the multiquantum parameter is large, K0 = I p ω ≫ 1. Another 2 independent dimensionless where the characteristic field is defined as E ch = (28) parameter (30) 32 ⎛ κ ⎞3 m2 ⎛ 2I p ⎞ ⎜ ⎟ = ⎜ ⎟ Ea , e ⎝ m ⎠ ⎝ κ0 ⎠ (31) where Ea = m2e 5 4 ≈ 5.14 · 109 V cm−1 is the atomic with electric field unity and κ0 = 1 a B a B = 2 me2 = 5.29 · 10−9 cm is the Bohr radius. Another parameter is the ratio of the ponderomotive energy (13) to the was In ZRP there is a unique bound state with l=0 [30]. 4 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article photon energy (with the factor (1 + ξ 2 ) 4 omitted) zF = e2E 02 was obtained in the form equivalent to (20) and evaluated by the SPM for small photoelectron momenta corresponding to the maximum of the ionization probability. An expression for the total ionization rate was obtained with exponential accuracy. Using the notations of section 2.2 one can present this result in the form (32) mω3 known as the strong field parameter [3]. Only two parameters are independent, so that another two can be correspondingly expressed as, for example F = 1 2K0 γ , z F = 8F K03 2 2 = 2K0 γ . w K ∼ exp { −2K0 fK (γ ) }, (33) ⎛ 1 ⎞ fK (γ ) = ⎜ 1 + 2 ⎟ arcsinhγ − ⎝ 2γ ⎠ 3. Brief history 1 + γ2 2γ , (34) where the function fK is known as the Keldysh function and the Keldysh parameter γ is defined by (29). A pre-exponential factor was also found in [1], but it does not reproduce the known static-field asymptotic [40] because the Coulomb interaction between the photoelectron and the atomic core is disregarded. This circumstance is clearly formulated in [1]. At the same time, detachment of negative ions [41] and ionization of atoms [42] by a static electric field was considered using a different method and simple expressions for ionization rates were obtained. In the limit γ → 0 the result of Keldysh agrees with the formulas of [41, 42] with exponential accuracy. Soon after the publication of Keldyshʼs paper,his result was generalized in several important cases. Namely,photoelectron spectra and total ionization rates were calculated in a closed analytic form for linear [13] and arbitrary elliptical polarization of laser light [14],the effect of quantum interference in photoelectron spectra was considered [13] and the method was extended into the relativistic domain [11]. In [12, 15] the problem of describing the Coulomb interaction was addressed and in [15] a correct expression for the rate of ionization was found in the tunneling limit, γ ≪ 1. For derivation of the respective Coulomb correction, the ITM was developed [15, 43]. It allows the interpretation of the Keldysh theory in terms of classical electron trajectories propagating in complex time and space. This interpretation appeared to be extremely useful for the development of new analytic methods in the theory of strong field ionization,in particular, to describe the Coulomb interaction and treatment of relativistic ionization including spin effects (see the reviews [4, 44, 45]). Although the ITM is already expounded in reviews [4, 44], in order to make this paper self-contained for the reader, we describe the ITM procedure in section 5.2. Thus, by the end of 1960s the Keldysh ionization ansatz was substantially explored. However, it soon became clear that the application of the results of [1, 11–15] to descriptions of experimental data is essentially restricted. The Keldysh theory allows the calculation of ionization rates and photoelectron momentum distributions consisting of ATI peaks separated by the photon energy. At that time momentum distributions were not available to be measured because at intensities of 1010 – 1011 W cm −2 even the second ATI maximum was too low to be detected. Thus the only measurable values predicted by the theory were angular distributions and total ionization rates. However, angular distributions were affected by Soon after the invention of lasers in 1960 the level of intensity sufficient for observation of highly nonlinear optical effects in atoms, molecules and solids was reached. The first observations of laser-induced breakdown of gases was already reported in 1963 [31, 32], although it was not clear if space charge effects played a role or not. The first unambiguous observations of nonlinear ionization of atomic xenon via a seven-photon absorption [33] and of the molecule H2 via a nine-photon absorption of radiation emitted by a ruby laser [34] were reported in 1965 by Delone and coworkers. The theoretical work of the pre-laser epoch was restricted by the application of the lowest-order perturbation theory to the description of processes involving two photons (see e.g. [35, 36]). In two papers [26, 37] published in 1962, Reiss suggested applying Volkov functions for the description of the electron continuum dressed by an intense electromagnetic field. This idea opened the way for the development of nonperturbative theories of strong field phenomena. In 1964 the basis of relativistic Volkov functions was used by Nikishov and Ritus for the description of elementary quantum processes in the field of a plane electromagnetic wave [38]. For scattering problems (e.g. annihilation of electron–positron pairs and the inverse process of pair creation, nonlinear Thompson scattering, etc), the use of a complete set of Volkov waves instead of bare plane waves allowed the development a theory where the classical field of a plane electromagnetic wave is accounted for exactly, while interaction with the spontaneous radiation field is treated perturbatively. In contrast, in the case of ionization, when a bound atomic state is involved, there is no rigorous method to describe the field of a strong electromagnetic wave. The easiest case is a system subjected to circularly polarized light. In the reference frame rotating with the field frequency, the Hamiltonian becomes time-independent [39]. For a system bound by the zero-range force this allows the calculation of the width of the quasistationary state developed from the initial bound state, i.e. the rate of ionization. For a general case of an atom and an arbitrary polarized field this method does not apply, so an essentially nonperturbative ansatz is needed. The ansatz expressed by (1) was proposed in the work by Keldysh, where he calculated the probability of ionization of the ground state of hydrogen in a linearly polarized monochromatic field with the interaction operator in the length gauge form (2). The partial ionization amplitude 5 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article that the photoelectron energy εf observed at a detector does not depend on the laser intensity ponderomotive scattering (see below in this Section) and the rate requires, for its correct calculation, the Coulomb interaction to be accounted for. The latter was performed in the tunneling limit [15], while the available data corresponded to a deeply multiphoton regime of ionization. For example, in early experiments [33, 34], the peak intensity was about 1011 W cm−2, so that γ ≃ 30 . These inconsistent theoretical and experimental limitations were the reason why the Keldysh theory was not extensively used in 1970s and was essentially even forgotten. In this period an important contribution to the theory was made by the work of Faisal [2], where the nonlinear ionization amplitude (1) was derived under more general assumptions than those of [1] (see the discussion in section 4.2). In 1980 Reiss [3] developed an Smatrix formalism based on the approximation of the exact continuum state by the Volkov function. Using the velocity gauge(4), he obtained an expression for the amplitude of ionization virtually identical to that of Faisal (see equations (21) and (22) of [2] and equations (26) and (42) of [3]). In [3], the first systematic study of ionization spectra in different regimes and an extended investigation of the properties of the generalized Bessel functions were given. Simultaneously, the theoretical discussion of problems inherent to the theory, including its gauge-noninvariance, also reduced interest in the Keldysh model in this period (see, e.g., a paradox discussed in [46] and its resolution in [47]). Only in 1979 was the first observation of a photoelectron spectrum consisting of more than one peak—i.e. an ATI spectrum—reported [48]3. In contrast to the energy conservation law given by (15), where the positions of the ATI peaks are intensity-dependent due to the presence of the ponderomotive energy, the positions of the photoelectron peaks observed in [48] and the subsequent experiments of that time did not depend on the laser intensity. The reason for this was the relatively long duration of laser pulses. Equation (15) assumes that the laser field is a plane electromagnetic wave, so the electric field amplitude does not depend on the coordinate perpendicular to the propagation direction of the pulse. In a real experiment, a laser pulse is focused and the ponderomotive energy depends on the lateral coordinate dropping down from its maximum at the focus center to zero at its edge. If the pulse lasts long enough, so that the electron can travel out of the laser focus before the field is off, it experiences ponderomotive acceleration and its canonic momentum p is no longer constant [50, 51]. In the limit τ ≫ R p, ε f = nω − I p . At parameters typical for the ionization experiments of the 1970s and early 1980s, τ ≃ 10−8 – 10−10 s, v ≃ 108 cm s−1 and R ≃ 10−3 cm, one obtains that inequality (35) is amply satisfied, so that the intensity-independent positions of ATI peaks are not surprising. The equations of section 2 assume the limit opposite to (35), i.e. short laser pulses. This discrepancy, although not physically fundamental, additionally postponed the time when the Keldysh theory was accepted overall for descriptions of strong field ionization. At the same time, the application of long laser pulses in strong field experiments has stimulated the development of the theory of ponderomotive forces (see [52, 53] for reviews including ponderomotive forces in the relativistic regime). Since the mid−1980s, after the invention of the chirped pulse amplification method [54], femtosecond laser pulses became routinely used in experiments. Pulses with a duration of less than 1ps are short in the sense of (35). Simultaneously to the shortening of pulse duration, peak laser intensity has increased enormously. The strong field regime z F ≫ 1 and even the regime of tunneling γ ≪ 1 were first achieved with CO2 lasers [55] and later with infrared lasers with a wavelength of approximately a micron. Since then, the Keldysh theory and closely related approaches have been frequently used for the interpretation of experimental data. The theoretical and experimental achievements of this period are reviewed, e.g., in [56, 57]. Fast experimental progress offered higher and higher intensities, so that investigations of nonlinear ionization moved from the multiphoton to the intermediate regime γ ≃ 1. Here the Keldysh model provided a commonly accepted picture of the phenomenon: a typical ionization spectrum (the spectra shown in the figures of sections 7.1 and 8.1 serve as good examples) extends up to several ponderomotive energies containing ∼ z F ≫ 1 ATI peaks; highly charged ionic states are produced via the sequential ionization mechanism, when electrons are being detached from an atom one by one. The slopes of ionization rates plotted versus intensity are determined mostly by the intensity dependence of the Keldysh function (34),although the tunneling Coulomb correction to the prefactor is also important for the quantitative description of data. Finally, with the further development of laser technique and photoelectron diagnostics, new features of photoionization spectra have been observed which are beyond the standard Keldysh ionization model. They can be sorted in three groups. First, there are the effects of atomic structure completely disregarded in the model. Freeman resonances with excited states can serve as an example [58]. Such resonances generate a fine intensity-independent structure for ATI peaks. Theoretical approaches describing the resonant structure of ATI spectra beyond the Keldysh model are reviewed, e.g., in [56]. Second, significant effects are induced by the Coulomb interaction between the photoelectron and the parent ion. Coulomb effects are determined mostly by the ion residual (35) where τ is the pulse duration and R is the laser focal spot radius, the final photoelectron momentum p f satisfies energy conservation p f2 p2 + UP = , 2 2 which means, in combination with energy conservation (15), 3 It is interesting to note that one of the phenomena directly connected to ATI, namely nonsequential double ionization, was observed for the first time earlier, in 1975 [49]. 6 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article charge and the laser pulse parameters but otherwise are species-independent. Asymmetric photoelectron angular distributions in elliptically polarized fields [59] present a good example of a qualitative Coulomb effect. Third, a variety of effects are generated due to laser-driven recollisions of the photoelectron with its parent ion [60, 61]. Understanding of the rescattering mechanism, including its role in the generation of the high-energy photoelectron plateau, high-order harmonics, nonsequential double ionization and its potential for probing atomic and molecular structure, has made an enormous impact in strong field physics over the past twenty years. Theoretical work of this late period has shown that the Keldysh model can be regularly extended to include rescattering effects [62–64]. This extension is currently known as the improved SFA or, alternatively, as the SFA with rescattering. Nonperturbative inclusion of Coulomb effects is less straightforward but also possible. Here, considerable progress (reviewed in section 6) has been achieved over the past decade. The extension of the Keldysh theory toward inclusion of the Coulomb interaction is mostly based on the technique of Coulomb corrections to the phase of the amplitude (1). Evaluation of these corrections employs the ITM. In an early paper by Perelomov and Popov [15], the ITM was applied to the calculation of the total ionization rate of atoms in the tunneling limit, γ ≪ 1. It was shown there that the Coulomb field enhances the rate of ionization, typically by several orders in magnitude—the effect has been reliably proven in experiments [65]. Later this result was generalized to arbitrary values of the Keldysh parameter [66]. Apart from the overall enhancement of the total ionization rate, the technique of Coulomb corrections allows the description of several effects accessible for experimental observation, including Coulomb asymmetry in elliptically polarized fields [59, 67, 68], cusps and double-hump structures [69–71], low-energy structures [72–76] and side lobes [77] in the momentum spectra of photoelectrons. Inclusion of atomic structure effects into the theory seems to be most problematic. The only atomic information encoded in the Keldysh ionization amplitude (1) is the value of the ionization potential and the symmetry of the initial bound state, therefore the role of excited bound states including possible multiphoton resonances is fully ignored. The application of few-cycle laser pulses with a broad spectrum, which has become routine over the past 10–15 years, makes the influence of the atomic structure in strong field ionization less important, because in a spectrally broad field transient resonances are essentially suppressed. Thus the Keldysh theory is more appropriate for the description of ionization in short pulses than in quasimonochromatic fields. The Keldysh theory without Coulomb corrections is quantitatively accurate for systems bound by short-range (SR) forces. Therefore observation of the strong field photodetachment of negative ions offers a strict test of the model [78]. The measurement of above-threshold detachment (ATD) of negative ions requires mid-infrared femtosecond lasers (in order to achieve the strong field regime, z F ≫ 1, γ ≃ 1 at low intensities) with high repetition rates (for accumulation of good statistics with low-density targets). Such measurements became possible over the past 10 –15 years when high-resolution ATD spectra of H−, Br − and F− were recorded [79–83]. Comparisons of the data with calculations by Grybakin and Kuchiev made using the Keldysh theory [16], demonstrated a very good quantitative agreement for the low-energy part of the spectrum, while for the high-energy part an equally good agreement was achieved when rescattering was taken into account. Another efficient test of the Keldysh model justifying its quantitative applicability for the description of negative ions photodetachment employs comparisons with predictions of the method of quasistationary quasienergy states (QQES), which provides a highly accurate solution of the ATD problem (see the discussion in section 7.1). Currently, the Keldysh theory is a cornerstone for the theoretical apparatus of strong field physics. Modern experiments performed with short intense laser pulses provide conditions under which the theory is applicable and can serve as a reliable zeroth-order approximation for the investigation of high field phenomena. 4. The Keldysh theory and closely related approaches In this section we discuss how the Keldysh ionization ansatz relates to another standard theoretical tool of strong field physics, the SFA. In order to formulate the applicability conditions of the theory, we start from a derivation of the Keldysh ionization amplitude. 4.1. Derivation of the ionization amplitude In [1] the probability distribution (given by equations (8)–(15) therein) is postulated—the reason why the theory is often called ‘the Keldysh ionization ansatz’. It was shown later [13] how this result can be rigorously derived. Below we sketch this derivation. Consider an electron bound by a potential U(r) and subject to a time-dependent electric field E(t ).4 The timedependent Schrödinger equation (TDSE) then reads i ⎤ ⎡ 1 ∂ ψ (r , t ) = ⎢ − Δ + U (r ) + E (t ) r ⎥ ψ (r , t ) ⎦ ⎣ 2 ∂t (36) where the length gauge representation (2) for the interaction operator is used. It has to be solved with the initial condition ψ (r , t → − ∞) = ψ0 (r)e iI p t , (37) where ψ0 (r) is the initial atomic bound state. After the laser field turns off, the wave function can be presented in the form ψ (r , t → + ∞) = ψb (r , t ) + ψout (r , t ), (38) where the first term in the r.h.s describes the bound part of the 4 As we consider the problem nonrelativistically, the dipole approximation applies, so the magnetic field of the wave can be neglected and the electric field can be considered space-independent. 7 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article wave function and the second the detached electron which goes away from the atom. At r → ∞, the bound part ψb vanishes while ψout takes the form of an outgoing spherical wave. It can be expanded in plane waves; the respective coefficients determine the amplitudes of ionization. Equation (36) with initial condition (37) can be presented in the integral form : +∞ ψ (r , t ) = ∫−∞ dt1 S 0 (p , t ) = η ( t − t1) − (2π ) 3 1 2 t ∫t M (p) = 1 (40) = { + 1 2 (41) t p t 1 ⎤⎫ v 2p t ′ dt ′⎥ ⎬× ⎦⎭ () ×U ( r1) ψ0 ( r1)e iI p t1 . (42) At large distances and for t → ∞, when the laser field is off it takes the form ψ (r , t ) = 1 (2π )3 2 ∫ d3peipr−ip t 2M (p). 2 (43) Comparing (42) and (43), one obtains for the amplitude M (p) = −i (2π ) 3 2 × +∞ ∫−∞ dt1e−iS0( p, t1) ∫ d3r1e−iv ( t ) r U ( r1) ψ0 ( r1), p 1 1 ⎫ ⎤ p 1 1 0 1 ∞ ∫−∞ dt1 ∫ d3r1e−iv ( t ) r −iS ( p,t ) E ( t1) r1ψ0 ( r1) , p 1 1 0 1 1. Relatively weak laser fields, F ≪ 1. This limitation means that the application of the Keldysh theory to the description of ionization in the barrier suppression or superintense, F ⩾ 1, regimes is not justified. It is shown by comparisons with results of numerical calculations and asymptotic expansions for the widths of atomic levels in a static electric field, that the rate of ionization calculated from the Keldysh model (as well as from the KFR and SFA), even with Coulomb corrections included, already deviates considerably from exact results for F ≃ 0.1, and this deviation grows with increasing reduced field F. In this domain the SFA overestimates the rate and predicts the effect of stabilization which is not confirmed by other methods based on summation of series of the perturbation theory and a 1 N expansion (see the discussion in [84] including figure 1 therein). On the other hand, the restriction F ≪ 1 is actually not severe. Ionization usually saturates during a few optical periods when the reduced field approaches F ≃ 0.1 or even earlier (this can easily be estimated using the expressions for the rate of ionization given in section 5.1 and 6.2.3). Therefore, it is rather difficult to achieve the regime F ≃ 1 in experiments, as most of the atoms will be ionized at lower intensities at the front edge of the laser pulse. 2. SR potential. This condition is of principal importance, because otherwise the spatial integral in (39) forms at large distances from the atom where the wave function differs substantially from the unperturbed bound state function. ∫ d3pe iv (t) r ∫−∞ dt1 ∫ d3r1 exp {− i ⎡⎣ vp ( t1) r1 ∫t ⎧⎡ ∫ d3r1 ⎨⎩ ⎢⎣ 12 Δ1 − i ∂∂t1 ⎥⎦ e−iv ( t ) r −iS ( p,t ) ⎬⎭ ψ0 ( r1) The derivation above allows the formulation the applicability conditions of the Keldysh ionization model. These are: t −i (2π ) 3 −i (2π )3 2 ∞ ∫−∞ dt1 4.2. Applicability conditions ∫ d3p exp i ⎡⎣ vp (t ) r − vp ( t1) r1 () (45) p i.e. the Keldysh amplitude (1) in the length gauge. ∫ d3pΨp* ( r1, t1) Ψp (r, t )= with η (t ) = ∫−∞ δ (x )dx being the Heaviside step function. Assume now that (i) the electric field amplitude is small compared to the atomic field, i.e. the reduced field (30) is small, F ≪ 1, and (ii) U(r) is an SR potential of the radius rc, so that U = 0 for r > rc [13]. The second assumption guarantees that only a small vicinity of the atom contributes to the spatial integral in (39). Then, under the first assumption, the atomic wave function remains only slightly disturbed by the laser field at r ⩽ rc , and we can replace the exact wave function ψ (r, t ) by its unperturbed limit (37). Then ψ (r , t ) = −i (2π ) 3 2 × ⎤⎫ v2p t ′ dt ′⎥⎬ ⎦⎭ ⎤ 2 ( ) + I ⎥⎦ dt′. Applying the self-conjugated operators Δ and i∂ ∂t1 onto the left part of the integrand in (44) one obtains with the condition GV (r, t; r1, t1 ) = 0 at t < t1. It expresses via the Volkov functions (2) = −i ⎡1 ⎢ vp t ′ ⎣2 ⎛1 ∂ ⎞ U ( r1) ψ0 ( r1)e iI p t1 = ⎜ Δ1 + i ⎟ ψ0 ( r1)e iI p t1 . ∂t1 ⎠ ⎝2 where GV is the retarded Greenʼs function in a homogeneous electric field, satisfying the equation GV ( r , t ; r1, t1) = −iη ( t − t1) +∞ Amplitude (44) coincides with (1). Indeed, the Schrödinger equation for the unperturbed bound state reads ∫ d3r1GV ( r, t; r1, t1) U ( r1) ψ ( r1, t1),(39) ⎤ ⎡ ∂ 1 ⎢⎣ i + Δ − E (t ) r⎥⎦ GV ( r , t ; r1, t1) 2 ∂t = δ ( t − t1) δ ( r − r1), ∫t (44) There are no special limitations on the laser frequency, in contrast to a widely accepted viewpoint that the Keldysh with the phase 8 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article theory only works in the multiquantum, K0 ≫ 1, or even only in the tunneling, γ ≪ 1, regimes. For example, for an SR potential supporting unique s-state amplitudes, (1) and (44) reproduce the first-order perturbation theory limit exactly. In a weak field the exponential factor can be replaced by the fieldfree one exp { − iS0 } → exp {i(I p + p 2 2) t}. Then integrating in (44) twice by parts with ψ0 from (26) and (27) we obtain for the amplitude 2 κ M (p) = 2 ∞ ∫−∞ pE (t )ei(I +p 2 2 p 2 2) t dt , in [2]) U (r − a (t )) → U (r ) . This simplification is valid either for an SR potential or when the quiver amplitude is small compared to the characteristic size of the atomic initial bound state. The latter is, however, satisfied only for sufficiently weak fields: even at intensities of about 1013W cm−2 and wavelengths of 1μm the quiver amplitude a = E0 ω2 ≃ 5 a.u. Thus a more general expression (47) reduces in practice to (21). In the paper by Reiss [3], the probability amplitude was defined as (46) π (κ + p ) which coincides with the result of the first-order perturbation theory employing exact continuum states in the ZRP 1 ψp (r) = (2π )3 2 ⎡ 1 eipr ⎤ ⎥, ⎢ eipr − r κ + ip ⎦ ⎣ κ= +∞ (S − 1)if = −i (48) where Vint is the interaction operator in the velocity gauge (4) and Ψ f( − ) is the exact final state wave function which contains both the effect of the laser field and of the binding potential and has an asymptotic behavior ∼ exp {ipr} at r → ∞. The central approximation of Reissʼs paper is that this exact function can be replaced by the Volkov function (5). It reduces the amplitude (48) to (1) with accuracy to a prefactor, again because of the different gauge used. This is also clearly stated in [3]. Thus, the differences between the results of Keldysh [1], Faisal [2] and Reiss [3] are: (i) that the Keldysh version uses the length gauge form of the interaction operator, while Faisal and Reiss employ the velocity gauge form and (ii) that Faisalʼs version contains a principal possibility to go beyond the Volkov wave approximation for the final state (equations (14) and (17) in [2]), although this possibility was not explored in the actual calculations. Using the velocity gauge allows a direct relativistic generalization of the theory, while the length gauge approach in the form (2) is usually restricted by the dipole approximation5. In the nonrelativistic domain the two gauges give formally different results, but this difference is not of physical significance: for the ZRP both gauges lead to the same expression, while for the Coulomb potential the preexponential factor is again incorrect in both cases (see section 8.1). One of the arguments in favour of the SFA form of the ionization amplitude is that its derivation does not require the two assumptions made to derive the Keldysh ionization amplitude (see the previous subsection). It can, however, be seen that the same approximations are indirectly required. Indeed, replacement of the exact wavefunction by the Volkov solution is unambiguously justified only for SR binding potentials. A common argument, that with increasing laser field amplitude the role of an atomic potential will sooner or later become negligible, does not apply to the Coulomb potential because of its long-range nature: even in a very strong oscillating field with a zero time-averaged value, the Coulomb force generates a net accumulating effect. This mechanism can be understood well on the level of classical mechanics. As is shown in section 6 using the language of trajectories, the Coulomb correction to the complex phase of 4.3. Keldysh theory, KFR and SFA In the literature, the analytic theory of strong field ionization is associated with several closely related approaches. Along with the term ‘Keldysh theory’, KFR theory and the SFA are frequently used. Here we discuss briefly the relationship between these three methods. To this end we compare ionization amplitude (1) with the respective amplitudes derived by Faisal [2] and Reiss [3]. In the work by Faisal the expression (equation (10) therein) for the amplitude of ionization differs by the gauge choice (velocity instead of length). In addition, the amplitude of [2] is determined for an arbitrary final state, which is not necessarily a plane wave. So, instead of (21), the n-photon amplitude of ionization in a linearly polarized field has the form [2] nω − UP ω × ψi (s), Ψ (f − ) Vint (t ) Ψi dt 2I p . Thus the original Keldysh approximation is restricted by SR potentials and applicable in a wide range of frequencies under the condition F ≪ 1. Extensions of the theory making it applicable to the quantitative description of the ionization dynamics of atoms introduce additional restrictions on the field parameters which will be discussed in sections 6 and 9. if , n = ∫−∞ ⎛E s U ⎞ ∫ d3 sψ *f (s) Jn ⎜⎝ ω02 , 2ωP ⎟⎠ (47) where ψi, f (s) are the Fourier transforms (24) of the initial and final atomic wave functions. Taking a plane wave for the final state ψf, so that ψ f (s) = δ (p − s) we obtain (21) (with accuracy to the factor 2π i ω which is due to the different definitions of the transition amplitude here and in [2]) or, equivalently, equation (21) of [2]. The amplitude (47) is obtained under the assumption the quiver amplitude of the electron in the laser field, a (t ) = ∫ dt′A (t′), can be omitted in the Kramers–Henneberger transformation (see [85, 86] and references therein for a review of the Kramers–Henneberger method in application to the ionization problem) of the binding potential (equations (5)–(8) 5 It is also possible to adopt the length gauge for relativistic calculations [87]. 9 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article the ionization amplitude, although it becomes smaller with increasing laser field amplitude, remains numerically large at any reasonable intensity. Apart from the KFR and SFA, the Perelomov–Popov–Terenʼev (PPT) and Ammosov–Delone– Karinov (ADK) models are also often mentioned in the literature. In [13, 14] expressions for the total ionization rate and momentum distributions of photoelectrons were derived from the Keldysh model for the case of an initial state in an SR potential with the arbitrary azimuthal and magnetic quantum numbers l and m. In the paper by Perelomov and Popov [15] the expression for the total ionization rate was generalized to the case of atoms, i.e. with the Coulomb interaction accounted for, but only in the tunneling limit, γ ⩽ 1. This result extends to low-frequency fields the staticfield rate derived by Smirnov and Chibisov [42]. These two formulas of PPT (one for negative ions, another for atoms) are now frequently used for the calculation of ionization rates in the field of intense laser radiation and are often mentioned as PPT rates. An identical expression for the total rate of ionization of atomic levels as derived by Perelomov and Popov was later obtained in [88] and is known as the ADK rate. For a more detailed description of the relation between the PPT and the ADK rates, we direct the reader to the review [4]. 5.1. Saddle-point approximation If the phase S0 in (45) is numerically large, the integrand in (1), (44) oscillates fast in time and the SPM [91] can be applied. After spatial integration, the amplitude (1) takes the form +∞ MK (p) = ∫−∞ dtP (p, t )e−iS0(p, t ) , (49) where the phase S0 (p, t ) is given by (45) and P is a gaugedependent prefactor P (p , t ) = − i (2π )3 2 ∫ d3r e−isrVint (t ) ψ0 (r) (50) with s = p or s = vp (t ) in the velocity and the length gauge, respectively. The phase in (49) can be scaled using the dimensionless parameters (28) and (32) zF 2 S 0 ( p , φ) = − K 0 φ + ∫φ 2 +∞ (q + a(φ′)) dφ′, (51) where we introduce dimensionless time φ = ωt , momentum q = p pF and vector potential a = A pF with pF = E0 ω . Saddle points tsα satisfy the equation S0′ ≡ 5. Representation in terms of trajectories ∂S 0 2 = 0, → γ 2 + ( q + a( ts)) = 0. ∂t (52) Clearly, the solutions of (52) are always complex, ts = t0 + iτ0 . If the prefactor (50) is not singular at t = ts , application of the SPM is straightforward and gives for the amplitude The explicit analytic expressions given in section 2.1 are useful for the numerical calculation of photoionization spectra in quasimonochromatic laser pulses. They have been extensively applied for the theoretical investigation of different ionization regimes as well as for the analysis of experimental data (see, e.g., [3, 89, 90] and references therein). It is, however, more common to use approximations for the amplitudes (1) and (20) and the respective probabilities obtained by the SPM. There are three good reasons for this. First, application of the SPM allows one to obtain much simpler analytic expressions than given by (21) and (23). With compact saddle-point formulas, the qualitative investigation of the main features of photoelectron spectra is straightforward. Second, the Bessel expansions assume a quasimonochromatic field and cannot be applied in the case of the now routinely used few-cycle laser pulses where pulse shape effects are significant. Third, the saddle-point result allows an interpretation in terms of complex classical trajectories satisfying Newtonʼs equation in complex time and space. This interpretation has appeared to be exceptionally fruitful for generalization of the theory and for heuristic prediction of new phenomena. In addition, the domains of the parameters where the Keldysh theory itself and its saddlepoint version are applicable essentially overlap, so the saddlepoint analysis almost does not narrow the applicability of the theory, and the respective SPM results are usually quantitatively accurate, if the Keldysh approximation applies in principle. MK (p) ≈ ∑ α 2π iS0′′ ( tsα ) P ( p, tsα )e−iS0( p, tsα ) , (53) where the sum is taken over all the solutions of (52) in the upper complex half-plane. The reason for choosing an integration contour which always goes through the upper halfplane is the sign of the ionization potential, I p > 0 . For the symmetric solutions of (52) with τ0 < 0 , the imaginary part of the phase at a saddle point is positive, Im S0 (p, ts ) > 0 , so that the corresponding contributions are exponentially large. If the prefactor is singular at t = ts , a modified version of the SPM should be applied [6, 16]. The respective expressions are given in appendix A. As we show in section 8.1, for practical calculations only one special case is of interest, namely, when the initial state wave function corresponds to an (l,m)-state in an SR potential and the length gauge is used. Then the amplitude takes the form [16] (see appendix A) MK (p) ≈ ∑ ( p, tsα )e−iS ( p,t 0 sα ) , α = 2Cκl iκ S0′′ ( tsα ) Ylm ( n v), (54) where Ylm is a spherical function and its complex argument n v = vp (ts ) κ is a ‘unit’ vector along the complex velocity taken at t = ts . 10 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article Below we consider the simplest and simultaneously frequently used case of a linearly polarized monochromatic field E (t ) = E 0 cos ωt ( ) ωts = arcsin qx ± i γ 2 + q⊥2 , { { −2K ⎡⎣ b (γ ) q 0 ( 1 b 2 (γ ) = 1 arcsinhγ γ2 2 x × 1 + ( − 1)n + l + m cos 2qx z F 1 + γ 2 )}. (58) 1 3 3 γ− γ + ..., 3 10 1 1 3 3 b2 ≈ − γ + γ + ..., γ 6 40 (56) + b2 (γ ) q⊥2 ⎤⎦ ⎤ ⎥, 2 ⎥ 1+γ ⎦ γ with asymptotics b1 ≈ where qx and q⊥ are the dimensionless momentum components parallel and perpendicular to the linear polarization direction. For elliptically polarized fields (except the special cases of emission along and perpendicular to the major polarization axis considered in [92]), bichromatic fields and few-cycle pulses, let alone more complex cases, there are, in general, no closed analytic expressions for saddle points, although the numeric solution of the saddle-point equation (52) presents no difficulty. Analysis of saddle-point trajectories in the complex time plane (see examples in [6, 92]) shows that they never merge, so that the SPM is always applicable in its simplest form when the expansion of the phase is truncated after the quadratic term proportional to the second derivative of the action (45). This is in contrast with the case of rescattering and high harmonic generation (HHG), when a pair of saddle points merge at the classical boundary (e.g. at ε ≈ 10UP for backscattering in a linearly polarized field and at Ω ≈ 3.17UP + I p for HHG with Ω being the harmonic photon frequency) and the SPM requires a modification [93, 94]. The fact that stationary points do not merge means, in particular, that there is no cut-off in the direct ionization spectra, although the ‘classical direct ionization cut-off’ at ε ≈ 2UP is often mentioned in the literature. Inspecting a typical direct photoelectron spectrum (see e.g. the spectra shown in the figures of sections 7.1 and 8.1) calculated from (53), one notices that nothing special happens at this energy. The slope of the spectrum changes abruptly only when the probability of rescattering exceeds that of direct ionization. The position of this point depends on the relative magnitude of rescattering and is not universal. For linear polarization it is usually between 2 and 5 UP. In the field (55) the imaginary part of (56) is minimal for p = 0. This corresponds to the maximum of probability. The spectrum drops quickly with increasing the momentum. Expanding (56) in a series with respect to qx , q⊥, one obtains a simple analytic expression for the momentum distribution (18) [4, 13, 16] dwn = w0 qq⊥2 m exp dOp 1 γ2 (55) when there are two solutions for (52) per laser period in the upper complex half-plane ⎡ ⎢ arcsinhγ − ⎢ ⎣ b1 (γ ) = b1 ≈ γ ≪ 1, (59) 1 ln (2γ ) ,γ≫1 [ln (2γ ) − 1] , b2 ≈ γ2 γ2 (60) and w0 = Cκ2l ωγ −2 m π 2 2 1+γ ( 2 l + 1 (l + m ) ! exp { −2K0 fK (γ ) }, 2 2 m m ! ) (l − m ) ! (61) Cκl is an asymptotic coefficient of the bound state wave function (A.2) and the Keldysh function fK (γ ) is given by (34). The second term in the parentheses of (57) describes interference between contributions of the two stationary points. The presence of interference effects in spectra of strong field ionization was noted for the first time in [13]. Integrating the distribution (57) over the angle and omitting the interference term, one obtains the rate of ionization with absorption of n laser photons and the total rate of ionization for the (l,m)-state in an SR potential 4Cκ2l β1 2 ⎛ 2 ⎞ ⎜ ⎟ πK03 2 ⎝ zF ⎠ |m| wSRn = I p 2 l + 1 (l + m ) ! (n − n th)|m| 2 (l − m ) ! m 2 m !) ( × e−2K0 fK (γ ) − 2c1(γ )( n − n th ) ( ) β [ n − n th ] , ∞ wSR = ∑ wSRn, (62) n = Nmin β = 2γ 1 + γ 2 , c1 (γ ) = γ 2b1 (γ ), 2 2 x the function (x ) = e−x ∫ (1 − t 2 /x 2 )|m|et dt turns to the 0 Dawson integral [95] for m = 0, n th = K0 (1 + γ −2 2) is the ionization threshold and Nmin = [n th ] + 1 is the minimal number of photons required for ionization (17). In the tunneling limit, this sum can be replaced by the integral which gives a simple tunneling formula for the rate of ionization from an SR potential [4, 13] where 3F (2 l + 1)(l + m ) ! m + 1 F π 22 m m ! (l − m ) ! ⎧ γ2 ⎞ ⎫ 2 ⎛ × exp ⎨ − ⎜1 − ⎟ ⎬. 10 ⎠ ⎭ ⎩ 3F ⎝ wSR (γ → 0) = 2I p Cκ2l } (57) ⎪ ⎪ ⎪ ⎪ (63) Expressions for momentum distributions and rates in circularly and elliptically polarized fields and for arbitrary values of photoelectron momenta can be found in [4, 84, 96–98]. Here 11 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article motion, in contrast to the trajectories used in Feynman representation of quantum mechanics [103]. For ionization from an SR potential ITM gives no new result, but provides an appealing ‘almost’ classical picture. The trajectory representation is, however, very efficient in accounting for semiclassical effects beyond the Keldysh approximation. This includes, in particular, the Coulomb interaction and spin effects. In order to apply the ITM for the calculation of the Coulomb-corrected ionization amplitude, it is useful to rewrite the action via the Lagrangian function 5.2. Representation via classical trajectories and the ITM The saddle-point result of the previous section can be presented in terms of classical trajectories propagating in complex time and space. To this end we note that the value −S0 (p) in (53), (54) is a classical action expressed as a function of the detection time td and final momentum p (in order to introduce the time instant td when the electron arrives at a detector we replace in the upper integration limit in (45), td → +∞). The corresponding electron trajectory is: rp (t ) = ∂S 0 = ∂p ∫t t () vp t ′ dt ′ s = p ( t − ts ) + ∫t t () A t ′ dt ′. s (64) r¨p (t ) = −E (t ), ∫t − ipτ0 + i E0 ω2 2 ω = Im rpm (t ) = 0. ∫t td s dt ′ 1 2 v p + v˙p rp − I p 2 } ≡F0 ⎡⎣ rp (t ), td ⎤⎦ + W0 ⎡⎣ rp (t ), td ⎤⎦ , (65) dt ′= (70) where W0 is the reduced action expressed via the Lagrange function (66) W0 ⎡⎣ rp (t ), td ⎤⎦ = ∫t = ∫t td s { ( ) ( ) } dt ′ 1 2 v p − E t ′ rp t ′ − I p 2 td L 0dt + ε0 ( td − ts ). (71) s with ε0 = −I p being the initial electron energy and td F0 ⎡⎣ rp (t ), td ⎤⎦ = −rpvp = −rp ( td ) p. (72) ts (67) Interpretation of the function (71) as a reduced action is connected to the fact that the latter is defined in classical mechanics as W = S + εt [99, 100]6. The function (72) is important because the trajectory rp (t ) has, in general, an imaginary part. The ITM representation of the Keldysh ionization amplitude reads then as where the last two terms in the r.h.s. give a constant imaginary part. There is no physical controversy here, as a trajectory is not an observable in the considered problem. The most probable final momentum pm corresponds to the minimum of the imaginary part of the action d ( Im S0 ) dp + ts } { 1 2 vp + Ip 2 td ( cos ωt − cos ωt0 cosh ωτ0 ) sin ωt0 sinh ωτ0, s = − rpvp The detection time td is real. Then the electron velocity vp (td ) = p + A (td ) at the detector is real too. However, the trajectory (64) can have an imaginary part even in real time. For example, in the linear field (55) the trajectory (64) is (assuming time t is real) E0 { td − S0 ( p , t d ) = − with initial and final conditions rp ( ts ) = 0, v2p ( ts ) = −2I p ≡ − κ 2, vp ( td ) = p. (69) Then It starts from the origin at t = ts with an initial velocity vp (ts ) = p + A (ts ) which is a purely imaginary value, according to the saddle-point equation (52). Thus the exponential part of the transition amplitude is defined by a classical action calculated along a trajectory which satisfies Newtonʼs equation in the laser field rp (t ) = p ( t − t0 ) + 1 2 v p − E (t ) rp. 2 L0 = MK ( p, td ) ≈ (68) ∑ ( p, tsα )e iW ⎡⎣ r 0 ⎤ ⎡ ⎤ pα, td ⎦ + iF0⎣ rpα, td ⎦ (73) α p = pm with the prefactor defined in equation (54). Under proper calculations, the dependence of the amplitude (73) on the detection time td reduces to an inessential phase factor. Concluding this section, note that the ITM is conceptually close to the Landau–Dykhne method [40, 104] developed for the calculation of semiclassical transition amplitudes and matrix elements. This means that the most probable trajectory is real in real time. Equations (64)–(66) introduce the ITM in the theory of strong field ionization [15, 43, 44]. In contrast to standard classical mechanics where Newtonʼs equation is supplied by 2N initial conditions (N is the dimensionality of the problem, N = 3 in the considered case), the ITM equations require 2N + 1 conditions (66). The reason is that in the ITM the initial time instant ts can not be arbitrary chosen, but is found from the saddle-point equation. These trajectories are often referred to as ‘quantum trajectories’ or ‘quantum orbits’ [101, 102]. We find the term ‘classical complex trajectories’ to be more appropriate, for it underlines that they do satisfy the classical equations of 6 This interpretation is, however, partially misleading. Reduced action usually appears in descriptions of conservative systems where it makes sense to extract the term εt containing the integral of motion ε. In our case energy is not conserved. An alternative interpretation treats W0 as a sum of two actions. One is the atomic action in the absence of the laser field the electron has before it is detached at t = ts . Another is the Volkov action. The author is grateful to B M Karnakov for his explanation of this vague point. 12 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article 6. Modifications calculated or postulated. Attempts to develop a regular procedure for the calculation of such a solution face tremendous difficulties. Therefore a heuristic approach looks admissible. Let us assume that the laser field distorts the Coulomb continuum adiabatically, so that the quantum state ‘follows’ the instant value of the electron velocity. Then a continuum Coulomb state Ψk( − ) [40] In its original formulation, the Keldysh theory does not account for the presence of excited bound atomic states and the electron–ion interaction in the continuum. Both aspects are particularly important for atoms and molecules (in contrast to negative ions, where the final state interaction is less significant although also causes observable effects). Therefore, modifications of the Keldysh theory mostly attempt to account for these two features inherent in atomic systems. A rigorous theory should incorporate both factors on the same footing, for they have a common origin. In practice, due to the approximate and nonperturbative nature of theoretical approaches, bound states and the Coulomb interaction in the continuum are usually treated using different methods. Currently, much greater progress has been achieved in describing the final state interaction than in the incorporation of bound states into the theory. In the simplest way, the final state interaction can be considered perturbatively in the first Born approximation. The wave function (42) describes photoelectrons emitted by the atom without a further interaction with the residual ion. This mechanism of ionization described by the Keldysh model is known as direct ionization. Considering the electron–ion interaction as a small perturbation, one can calculate a correction to the wave function (42) induced by scattering of direct photoelectrons driven by the laser field on the parent ion. This is the basic idea of the rescattering or recollsion scenario of strong field ionization [61, 62]. The respective rescattering correction to the amplitude (1) reeds ∫ M1 (p) = − d3q Ψ k( − ) (r) = +∞ MCV (p) = − i ∫−∞ dt1 Ψ p(CV ) Vint (t ) Ψ0 dt . ⎛ i Ψ p(CV ) (r, t ) = Ψ p( − ) (r) exp ⎜ iA(t )r − ⎝ 2 (76) t ⎞ ∫−∞ v 2p (t′)dt′⎟⎠. (77) It was proposed for description of strong field processes in [106, 107] and is known in the literature as the Coulomb–Volkov (CV) wave function. The corresponding expression (76) for the ionization amplitude was adopted by Dorr, Shakeshaft and Potvliege [108, 109] and by Basile with coauthors [110]. Since then it has been widely used in the theory of strong field ionization and laser-assisted bremsstrachlung. We will refer this approach as the CVA. The first advance of the CVA was made through its application in the description of the Coulomb asymmetry of photoelectron distributions in elliptically polarized fields [110, 111]. A simple examination of the symmetry properties of the Keldysh ionization amplitude (1) shows that in monochromatic fields, angular distributions possess four-fold symmetry, independent of the field polarization, while experiments show that in elliptically polarized fields angular distributions only obey inverse symmetry [59, 112]. This symmetry violation was, from the beginning, attributed to the influence of the Coulomb field on the outgoing electron and is now known as Coulomb asymmetry [67] (see also the discussion in sections 6.2 and 6.3). Later on the CVA technique was applied in the calculation of photoelectron momentum distributions in linearly polarized fields [113]. It was shown there that the CVA provides a significant improvement in the description of photoelectron momentum distributions compared to the SFA, particularly for low energies and relatively low intensities when γ ⩾ 1. Typical results of CVA calculations are shown in figure 1 where two-dimensional (in the (k z, k ρ ) plane where kz is the parallel and k ρ is the transversal with respect to the field polarization component of the photoelectron momentum) and transversal momentum distributions, calculated using the standard Keldysh approach, the CVA and by solving the TDSE numerically, are presented. It is clearly seen from the dt Ψp U Ψq (t ) Ψq Vint Ψ0 ( t1). ∫−∞ Here the notation Ψp(CV ) is used for the function t × (75) where 1F1 is the confluent hypergeometric function, can be adopted for approximate description of the laser-distorted continuum by replacing k → vp (t ). Instead of (1) one then obtains +∞ ∫−∞ ⎛ 1 i⎞ eπ 2k Γ ⎜ 1 + ⎟ 32 ⎝ k⎠ (2π ) ⎛ ⎞ i × eikr 1F1 ⎜ − , 1, − i(kr + kr) ⎟ , ⎝ ⎠ k (74) It describes the high-energy part of photoionization spectra (see [7, 9] for reviews of the SFA with rescattering). For negative ions the amplitude (74) does indeed give a small correction to the total probability of ionization (although in spectra it dominates at high energies), so that the perturbative account is justified and yields qualitatively correct results (see also sections 7.1 and 8.1). In the presence of the Coulomb interaction, the perturbative correction to the probability appears in the direct part of the spectrum comparable to or even dominating the Keldysh probability. Even then the SFA with rescattering is able to reproduce qualitatively some effects in direct ionization spectra (see, e.g. [105] where the low-energy structure is described by this method). However, for the development of a consistent theory of ionization in the presence of the Coulomb field nonperturbative methods are obviously required. Here we describe several significant modifications of the theory made in the nonperturbative spirit. 6.1. Coulomb–Volkov approximation (CVA) In order to improve the Keldysh amplitude (1) one could try to find a continuum wave function approximately describing both the laser and the Coulomb field. Such a function can be 13 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article Figure 1. Left panel: photoelectron momentum distributions in cylindrical coordinates (k z, k ρ ) calculated for the case of ionization of hydrogen by a six-cycle laser pulse (of duration τ = 151 a.u.) with the carrier frequency ω = 0.25 a.u. and the field amplitude E0 = 0.05 a.u. obtained from the SFA, the CVA and from exact numerical solution of the TDSE. Right panel: normalized transversal momentum distributions calculated from the SFA (dashed line), the CVA (red solid line) and the TDSE (black solid line) for the following parameters: E0 = 0.05, ω = 0.25 a.u. (a), E0 = 0.0377, ω = 0.05 a.u. (b) and E0 = 0.053, ω = 0.05 a.u. (c). Corresponding values of the Keldysh parameter are shown in the figures. Reprinted with permission from [113]. Copyright (2008) by the American Physical Society. plots that for low values of photoelectron momenta k ⩽ 0.3, the application of the CVA yields a very good quantitative agreement with the TDSE, while the SFA result even disagrees qualitatively. The transversal distributions in the right panel clearly show the cusp observed in experiments and not reproduced by the SFA [70, 71]. Agreement between the CVA and the TDSE becomes less accurate for lower frequencies and higher intensities, i.e. with decreasing γ (see, e.g. figures 4 and 5 in [113]). Also, the CVA does not appear to give any visible improvement for higher photoelectron energies (in the direct part of the ionization spectrum) and does not reproduce the rescattering plateau. The main problem of the CVA approach is that the function (77) is not a solution of any form of the Schrödinger equation, so that there is no regular way to estimate the accuracy of the approximation and formulate conditions restricting its applicability. As a consequence, there is no clear answer to the question why some Coulomb effects are quantitatively reproduced while other do not even appear on a qualitative level. Some insight into the applicability conditions of the CVA was achieved by Kornev and Zon [114, 115] who examined the accuracy of the CV functions using a nonstationary generalization of the Siegert theorem [116]. The latter simply employs the fact that for exact solutions of the Schrödinger equation the matrix elements of momentum and coordinate are connected as f p i = ∂t f r i fulfilled with reasonable accuracy. This observation agrees with the above-discussed results of [113] where the accuracy of the result is higher for higher frequencies. However, to the best of our knowledge, there is no qualitative understanding of these features of the CV functions. The question of the applicability of these functions in the description of strong field processes remains open. It could probably be solved by applying the saddle-point analysis to the amplitude (76). 6.2. CCSFA Another simple and physically transparent approach which allows the approximate inclusion of the Coulomb field into the theory of nonlinear ionization is based on the ITM. Instead of correcting the final state wave function one can modify the transition amplitude, which is technically much easier. Indeed, if the amplitude of ionization can be expressed via classical complex trajectories (as shown in section 5.2), no knowledge of the full wave function is needed; it is sufficient to find all the trajectories corresponding to desirable initial and final conditions and satisfying Newtonʼs equation in the presence of the laser and the Coulomb fields. Simplicity of the ITM has allowed to extend it on arbitrary values of γ and to apply for description of several experimentally observed features of strong field ionization. 6.2.1. The CCSFA algorithm. As described in section 5.2, the (78) ITM allows one to represent the ionization amplitude in the form (73) where the action in the exponent is calculated along classical trajectories satisfying Newtonʼs equation (65) for an electron in the laser field with the initial conditions (66). If For two CV functions this relation is in general not satisfied. It was shown in [114, 115], however, that in relatively weak and high-frequency fields, E0 < ω , the condition (78) is 14 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article there is an extra force acting on the electron, one can account for it by correcting the functions W0 and F0. When this force is relatively small, the respective trajectories and action can be found perturbatively r (p, t ) = r0 (p, t ) + r1(p, t ) + ..., W (p) = W0 (p) + W1 (p) + .... Then the Coulomb-corrected ionization amplitude is α (79) td W= ( ) ∫ F ( p, t ) ≈ F ( p′, t ) ± iν (80) d . } td dt , F = −rv . (85) ts′ 0 td ts Z dt , r0 (p, t ) (86) d Here Z Z ≡ κ 2I p ν= (87) is the effective principle quantum number, ν = 0 in the case of negative ions. The term W0 (p′, td ) is the Coulomb-free action calculated along the new trajectory and the second term is the Coulomb action calculated along the Coulomb-free trajectory. The correction to ±iν is a constant close in absolute value to unity. The Coulomb integral WC = ∫t td s Z dt r0 (p, t ) (81) which assumes that the electron is already far from the atom, so that the Coulomb force is small with respect to the laser force, but the time passed after ts is still a small fraction of the laser period. Then the Coulomb-free trajectory (64) can be expanded in the series with respect to t − ts : (82) (iii) Photoelectron momentum is no longer conserved, so that its value p′ entering the saddle-point equation (52) is different from the one measured at a detector. This initial drift momentum is to be found from r0 (t ) ≈ ( p + A ( ts ) ) ( t − ts ) − 1 E ( ts )( t − ts)2 , 2 r 20 (t ) ≈ − κ 2( t − ts)2 − ( p + A ( ts ) ) E ( ts )( t − ts)3 . v ( td ) = p′ + v1( td ) = p. (88) is logarithmically divergent at the lower limit when r0 (p, t → ts ) → 0 and requires regularization. This is performed by replacing ts → t* (see figure 2, right panel) and matching the result of integration to the asymptotic of the atomic bound state wave function [15, 44, 118]. The matching point t* satisfies the condition νγ ≪ ω2 ts − t * 2 ≪ 1. (89) 2 K0 1 + γ with the correction r1 to be found from Newtonʼs equation r03 { 1 2 Z v − E (t ) r + − I p 2 r W ( p, td ) ≈ W0 p′, td + vp (t ) = p + A (t ) → v (p, t ) = v0 (p, t ) + v1(p, t ) , r¨1 = − s′α Taking into account that the trajectories (81) are found perturbatively, we should only keep contributions linear with respect to the charge Z in (84). The first-order expansion of the action (85) yields (see appendix B) (i) For ionization of a level with quantum numbers (l,m) and ionization potential Ip we calculate the respective SR Keldysh amplitude and present it in the form (73). For each final momentum p , there are several (typically two from each laser period) Coulomb-free trajectories to be found from (65) with initial conditions (66). (ii) A Coulomb-free trajectory is replaced by a corrected one: Z r0 ∫t ′ s so the effect of these corrections on the photoelectron momentum distribution can be significant. As we will see below this is exactly the case. In order to develop the perturbation theory we need to express the action as a function of coordinate, because this coordinate will then be corrected. This is the reason why action (71) but not (45) should be used as a zero-order approximation. Appendix B illustrates that starting from the form (45) one obtains a meaningless result. The procedure of calculating the Coulomb-corrected ionization amplitude is the following [66, 117, 118]: rp (t ) → r (p, t ) = r0 (p, t ) + r1(p, t ) (p′, t ) exp (iW [r (p, t) ] + iF [r (p, t) ])(84) with Here we introduce a new notation r0 (p, t ) ≡ rp (t ) for a trajectory (64) in the laser field. This approach is known as a perturbation theory for the action. For applicability, it requires that the corrections W1 and F1 are small in absolute value compared to the leading terms. At the same time, a correction to the action can be numerically large: W0 ≫ W1 ≫ 1, ∑ M (p) ≈ (83) (90) (91) Replacing in (71), (88) the lower integration limit by t* we obtain the following asymptotic (iv) A new saddle point ts′ = ts (p′) is calculated from the same saddle-point equation (52) with p′ (found from (83)) instead of p. (v) The Coulomb potential energy UC = −Z r is added to action (71). W ( t * → ts ) ≈ κ 2t * + iν ln ω ( t * − ts ) + ... (92) It follows from (91) that the distance between the electron and the atom at the matching time instant is 15 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article Figure 2. Left panel: complex time plane with the saddle point ts = t0 + iτ0 and different integration paths connecting ts and td shown by a solid and two dashed lines. Right panel: an integration path (dashed line) corresponding to the matching at t*. The circle characterizes the radius |ts − t * | of the matching region. r * ≡ r0 ( p, t * ) = iκ ( t * − ts ). (93) g= The phase of the atomic wave function in this part of space, κ|r * | ≫ 1, is given by its semiclassical asymptotic Wat ( r * ) = ∫ r* p (r )dr ≈ iκr * − iν ln κr * (94) () (i) At the upper integration limit td → ∞ the action (86) grows as ⎧ ⎫ Z iν ⎪ ⎨ ⎬ dt. (95) + ⎪ 2 t − ts ⎪ ⎩ r0 (p, t ) ⎭ td ⎪ 2 p′ Z W∼ td + ln td 2 p′ A similar procedure allows one to obtain a regularized expression for the Coulomb correction v1, by matching a solution of (81) with a trajectory in the Coulomb field of the atomic core calculated in the domain r ≃ r * [120]. The regularized result has the form (for the linearly polarized field (55)) v1 = ⎧ ⎫ Z r (t ) ⎨ − 30 + f ( t − t s ) ⎬ dt s r0 (t ) ⎩ ⎭ + iZF g (p, γ ) ln ( td ), ∫t (100) For a fixed value of the final momentum p the Coulombcorrected momenta p′α corresponding to different trajectories indexed by α are different. As a result, a phase difference between any two contributions becomes proportional to the time of observation td which is apparently unphysical. This effect results from the perturbative expansion of the action: it is clear that the exact action (85) does not suffer this problem. Therefore, although the Coulomb-corrected trajectories can be found perturbatively, the action must be calculated from (85) without expansions, particularly if interference effects are of interest. The actual calculations of [68, 75, 117] have been performed by solving the full Newtonʼs equation numerically. (ii) Coulomb-free and corrected trajectories are, in general, complex: velocity vp (t ) is real for real t, while the td (96) where ⎧ q+a g ( q, φs ) ⎫ ⎬. f = iZF ⎨ − 2 ( t − ts ) ⎭ ⎩ ω( t − ts) (99) Equations (95)–(99) define the action in (85), (86) and thus the ionization amplitude. In practice, calculations along the described algorithm encounter four serious difficulties: is finite. The regularized action (88) can be presented in a form convenient for calculations [15] s () v ( ts ) = p + A ts′ − v1 ≡ p′ + A ts′ . ⎛ ω⎞ Wat ( r * ) + W ( t * ) ≈ κ 2ts + iν ln ⎜ 2 ⎟ ⎝ iκ ⎠ ∫t (98) Here q = p pF , a = A(ts ) pF and pF = E0 ω and the unit vector e along the polarization direction. The integral in (96) converges and its value does not depend on the integration path provided the latter does not intersect cuts of the function r0 r03. The obtained value of v1 corrects the initial complex photoelectron velocity so that with p (r ) = 2( − I p + Z r ) . Comparing (92) and (94) we find that the terms containing t* and r* do cancel, so that the action WC = −iν ln ( 2iK0ω t d ) + ⎤ 1⎡ 3 ⎢ e + 2 (q + a) (e (q + a) ) ⎥ cos ωts 2⎣ ⎦ γ (97) and 16 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article Figure 3. Ionization rate (in atomic units) versus laser intensity (in units W cm−2) calculated from (108) (solid line), without the correction (106) (dashed line) and from the tunneling formula (rate (63) with the correction Q1) (dotted line). The gray line in the left panel shows the rate with only Q1 accounted for. Rates obtained numerically using the Floquet method [124] and from the solution of the TDSE [125] are shown using dots and triangles, respectively. Left panel: hydrogen ground state in the field of a Nd laser with λ = 1064 nm. The correction QC ≈ 103. Right panel: Xe17+ (4p0-electron with I p ≈ 434 eV) in the field of an XUV laser with λ = 13.3 nm, corresponding to the parameters of the experiment [126]. The correction is QC ≈ 10 9 . The values of the Keldysh parameter are shown at the upper horizontal axis. Reprinted with permission from [118]. trajectory rp (t ) has a constant imaginary part (67). Complex coordinates in the r.h.s. of Newtonʼs equation make numerical solution difficult. In practice, in most cases the imaginary part of a trajectory can be omitted without introducing a significant error, for the following reason: the most probable trajectory is always real, therefore imaginary parts of trajectories corresponding to relatively high probabilities of ionization remain small. Most of the calculations along the CCSFA (including those of [68, 75, 77, 117, 118]) have ignored these imaginary parts. Account of the complexness of trajectories can also be important at the initial stage of motion corresponding to the vertical part of the integration path of figure 6(a), known as sub-barrier motion (see also section 8.2). In particular, it has been shown in [121] that this account corrects the phase difference between the two dominant trajectories by π for emission of photoelectrons along the polarization direction. This shifts the interference pattern replacing minima by maxima and vice versa, in agreement with the results of ab initio numerical TDSE solutions. (iii) Unlike the laser-induced action (71), the the full actions (85) and(86) are not analytic functions in the complex time plane. An analytic function r 20 (p, t ) has, in general, an infinite number of first-order zeros, which generate branching points and cuts of the functions 1 r0 and r0 r03. A proper integration path in (95) and (96) must connect ts and td without intersecting the cuts. In a linearly polarized monochromatic field roots of the equation r 20 ( tn ) = 0 always is. Then the Coulomb potential energy has no branching points but only first-order poles. The presence of even a small lateral momentum component converts these poles into pairs of branching points which move away one from another with increasing p⊥. Cuts are defined as lines in the complex time plane where r 20 < 0 . Thus, integrating between ts and td one should ensure that no cuts are intersected [120]. The choice of integration paths and their relation to the position of the exit point from the barrier is discussed in section 8.2. (iv) The Coulomb field generates new trajectories. In a linearly polarized field, and with no Coulomb interaction accounted for, there are two trajectories per laser period bringing the electron to a given final momentum p . They are known in the literature as the short and the long orbit. When the Coulomb field is accounted for, these two Coulomb-free trajectories survive, experiencing only a smooth deformation (see, e.g. Figure 3 of [117] for illustration). In addition, another two trajectories can emerge, where the lateral momentum changes its sign once. The new classes of trajectories have their classical cut-offs, i.e. they exist only in some part of the momentum space. The boundary where they disappear can be analytically found for a model case [76]. As is typical for classical boundaries, the density of trajectories has an integrable divergence there, the classical probability is divergent as well and the probability calculated quantum mechanically reaches a sharp maximum. This effect generates the recently observed low-energy structure [72, 74–76]. In the above-described perturbative algorithm of calculation, these new trajectories are missing. In order to surmount this difficulty one should use the ‘shooting’ method instead of a perturbative calculation [75, 77]. The method works as follows: in the (101) come in pairs which merge for electron momenta parallel to the polarization axis. In this special case tn are secondorder zeros of the function r20 like the saddle point ts 17 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article momentum space a sufficiently large number (up to several million) of photoelectron momenta are randomly selected, so that they cover the momentum space with sufficient density. For each momentum all the corresponding Coulomb-free trajectories and initial conditions (66) are determined. With these initial conditions new trajectories are calculated solving the full Newtonʼs equation and the respective new final momenta are determined. The final momentum space is covered by a grid whose cell size is adjusted in such a way that about ten or more new trajectories arrive in each cell on average. The momentum distribution is then calculated as ∫0 τ dτ ′ cosh τ0 − cosh τ ′ ⎡ ( cosh τ0 + 1) ( cosh τ − 1) ⎤ 2 ⎥ arctanh ⎢ = ⎥⎦ ⎢⎣ sinh τ0 sinh τ0 sinh τ we arrive at an explicit result ⎛2⎞ Im ΔW = − ν ln ⎜ ⎟ + ν (ln 2γ − 1) . ⎝F⎠ This correction leads to appearance of the Coulomb factor in the ionization rate 2 w ( pmn ) ∼ ∑ke iWk + iFk , QC = e−2 Im ΔW = (102) k 6.2.2. Examples of applications. Total ionization rate. Coulomb corrections change the imaginary part of the action and therefore the total rate of ionization. In order to calculate a correction to the rate it is sufficient to consider the most probable trajectory. In the monochromatic linearly polarized field this trajectory corresponds to p = 0 . A perturbative correction to the action (86) reads () ΔW ≡ W (p = 0) − W0 (p = 0) = W0 p′ − W0 (0) ∫t td s ⎧ ⎫ ⎪ Z iν ⎪ ⎨ ⎬ dt . (103) + 2 ⎪ t − ts ⎪ ⎩ r0 (p = 0, t ) ⎭ 2ν Here p′ is the initial drift momentum found from (83). It cannot be calculated analytically for arbitrary γ, while for γ ≫ 1 using the Kapitza averaging method [99, 123] one obtains [66, 118] p′2 ≈ 2ων ≪ I p. ⎛ 2 ⎞2ν ⎛ 2γ ⎞−2ν ⎜ ⎟ ⎜ ⎟ ≡ Q1Q2 . ⎝F⎠ ⎝ e ⎠ (106) Here F is the reduced field (30). Its value, typical for experiments on nonlinear ionization, is F ≃ 10−2 ÷ 10−1. As a result the correction Q1 is always numerically large and enhances the rate of ionization by several orders in magnitude —this effect is well documented in experiments [4, 65]. This correction originates from the presence of the Coulomb potential energy in the action and was first calculated by Perelomov and Popov [15]. The respective expression is valid at arbitrary γ. In the tunneling limit the effect can be interpreted as a consequence of the lowering of the potential barrier the electron tunnels through. The correction Q2 calculated in [66] originates from the Coulomb deceleration of the photoelectron: in order to arrive to the detector with the minimal possible momentum p = 0 , the electron has to start having a momentum along the field polarization satisfying (104). The price to pay for this is the reduction of the ionization probability by the factor Q2 ≪ 1. The respective expression is derived assuming the multiphoton regime. This result can be extrapolated into the tunneling domain by replacing γ → γ + e 2 in (106). Although in the multiphoton limit Q2 ≪ 1, the net correction remains numerically large: where pmn is the momentum corresponding to the center of the (m,n)th cell (for a two-dimensional distribution), the sum is taken over all trajectories rk (t ) arriving in this cell. A detailed description of the shooting algorithm is given in [75, 122]. With these four extensions the CCSFA algorithm is ready for application. − iν ln ( 2iK 0 ω td ) + (105) QC ≈ ( 2eK0) , γ ≫ 1 (107) The result (106) allows one to obtain a quantitatively correct expression for the total ionization rate of atoms [66, 118]: (104) w = QC wSR, (108) Then from (57), (58) where wSR is the ionization rate of a level with the same Ip, l and m in an SR potential (62). In the tunneling limit (108) reproduces the well known PPT rate [13, 15]. The high quantitative accuracy of the rate (108) has been checked by comparing it with results of exact numerical TDSE solutions in the single-electron approximation. Two examples of such comparisons are given in figure 3. Coulomb asymmetry in elliptically polarized fields. Violation of the four-fold symmetry of angular distributions in a field with elliptical polarization is an example of a qualitative Coulomb effect. For a sufficiently long laser pulse such that the carrier-envelope phase effects are not important, the Keldysh amplitude (1) predicts angular distributions symmetric with respect to both the major and the minor ⎛ ⎞2 p′ Im W0 p′ − W0 (0) = K0 b1 (γ ) ⎜⎜ ⎟⎟ ≈ ν (ln 2γ − 1) . p ⎝ F⎠ ( () ) The Coulomb-free trajectory (64) corresponding p = 0 has the form r0 (p = 0, t ) = ωts = E0 ω2 ( sin (ωt ) − sin ( ωts ) ), π π + iarcsinhγ ≡ + iωτ0. 2 2 A contribution into the imaginary part of the integral comes only from the vertical part of the integration path shown on figure 6(a). Using the integral 18 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article Figure 4. Normalized angular distributions in the polarization plane, evaluated by different methods: the standard Keldysh approximation (green), the CCSFA (red) and the ab initio TDSE solution (black) for the ground state of hydrogen (a),(b), neon (c) and argon (d). In the latter case the data recorded in [112] are shown using black circles while the TDSE result is not shown. The laser intensity, frequency, ellipticity and the photoelectron energy are, respectively: (a) 1.0 · 1014 W cm−2, 1.55 eV, 0.5 and 8.2 eV, (b) 1.6 · 1014 W cm−2, 3.1 eV, 0.5 and 2.6 eV, (c) 2.0 · 1014 W cm−2, 1.55 eV, 0.36 and 7.1 eV(d) 0.6 · 1014 W cm −2 , 1.55 eV, 0.36 and 2.8 eV. For neon and argon the ionization probability is averaged over the magnetic quantum numbers m = 0, ± 1 within the p-shell. The orientation of the polarization ellipse is shown in the insert with the rotation direction of the electric field vector indicated by an arrow. Reprinted with permission from [117]. Copyright © 2008 Taylor & Francis. polarization axes. Experimental data clearly show that the distributions possess only central symmetry [59, 112]. According to the Keldysh theory or the SFA, at intermediate ellipticity (say, 0.2 < ξ < 0.8) the most probable photoelectron momentum is directed along the minor polarization axis [14]. In the tunneling regime this can easily be explained: the electron appears (with the highest probability when the field is maximal) at the tunnel exit with a zero velocity, vp (tm ) = 0. At this time instant tm the vector potential is minimal in absolute value and directed along the minor axis. Thus the momentum p = vp − A(tm ) coincides with the vector potential. In elliptically polarized fields, realization of the CCSFA algorithm is particularly simple, because only one trajectory gives a dominant contribution to the probability, so that one needs not care about a proper account of interference and the shooting algorithm is also not required: a calculation can be performed in a straightforward manner along (84)–(99). The results shown in figure 4 demonstrate good quantitative agreement with exact TDSE calculations. Agreement with the data is acceptable, but less accurate. This discrepancy can be attributed either to multielectron effects or to uncertainties in experimental parameters (mostly in the laser intensity). Coulomb-induced trajectories. With the Coulomb field accounted for, new complex classical trajectories corresponding to a given value of the final momentum emerge. In a linearly polarized field it can be up to four trajectories per cycle, instead of the two Coulomb-free trajectories. Figure 5 shows an example. All trajectories can be grouped into four types. The first two types are the so-called short and long trajectories existing in the Keldysh theory. The short trajectory (type I) fulfills the condition bz pz > 0 and px px′ > 0 , where bz is a z-projection of the tunnel exit position (see section 8.2 for a definition of this value) and pz is the photoelectron momentum projection on the polarization axis. This means that the ejected electron moves from the tunnel exit directly towards the detector. Type II is a long trajectory obeying bz pz < 0 and px px′ > 0 . Here, the electron starts from the tunnel exit that points away from the detector but ends up with a parallel momentum in the opposite direction because of the drift it acquires from the laser field at the time of ionization. The lateral initial momentum px′ of type I and type 19 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article Figure 5. (a) Partial photoelectron momentum distributions in the (pz , px ) plane (with z chosen along the linear polarization axis) due to trajectories of types I–IV calculating by the shooting method [75, 122]. A caustic forming the LES is seen on the panel III. (b) Dominant trajectories in the zx plane contributing to the final momentum p = ( − 0.22, 0.1) close to the LES position. Open circles indicate the end of the laser pulse. The scales in x and z are different. Reprinted with permission from [75]. Copyright (2010) by the American Physical Society. II trajectories is already in the same direction as the final momentum at the detector. This is always the case in a Coulomb-free theory where p = p′. Type III is classified via bz pz < 0 and px px′ < 0 . The corresponding electrons start on the opposite tunnel exit, as type II electrons do, but also have an initial lateral momentum pointing in the ‘wrong’ direction. Due to the Coulomb force their lateral momentum is ultimately reversed. The remaining class of type IV trajectories obey bz pz > 0 and px px′ < 0 , i.e. although the tunnel exit already points towards the detector, the initially incorrect lateral momentum is reversed by the Coulomb field. The new classes of trajectories generate at least two significant effects in photoelectron momentum distributions. First, trajectories of type III form a caustic (a sort of classical boundary, where the density of trajectories diverges) which induces the low-energy structure [75]. Second, interference between type III from one side and types I and II from another, modulates momentum distributions forming the structures known as side lobes [77]. Coulomb field [127, 128]. The method combines two techniques. The first is based on the partitioning of the coordinate space into inner and outer regions, such that the electron dynamics is dominated by the electron–ion interaction in the first, and by the laser field in the second region. The ARM method allows one to formulate equations for the two (inner and outer) wave functions matched at the boundary [128]. The second technique known as EVA [127] allows one to construct an approximate Coulomb-corrected Greenʼs function in the outer region needed to define the amplitude of ionization. Both the ARM and EVA methods explore ideas similar to those used in CCSFA, but independently developed and differently realized. The perturbation theory for the action and the matching with the atomic wave function described in section 6.2 are close (but not identical) to the eikonal calculation of the Coulomb correction and to the space partitioning, respectively. Here we describe the algorithm which combines the ARM and EVA and illustrate its application using several examples. 6.3.1. Ionization amplitude in the EVA and ARM approximation. Within the ARM the coordinate space is 6.3. Eikonal–Volkov approximation (EVA) and the analytical Rmatrix (ARM) method partitioned into inner (+) and the outer (-) regions separated by a spherical boundary of radius a centered at the atom. The objective is to find an approximate wave function (in contrast to the CCSFA where the ionization amplitude is constructed Another method of the inclusion of Coulomb effects suggests a regular procedure for calculation of an approximate wave function in the presence of a strong laser field and the 20 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article without knowledge of the Coulomb-corrected wave function) developing from the initial bound state Ψ0 (r, t ) under the action of a laser pulse. The Hamiltonian operator H=− Δ + U (r ) + Vint (r , t ) 2 obtained, if we account for the Coulomb interaction correcting the phase of the Volkov function. In the outer region the electron does not approach the nucleus and the Coulomb field can be considered to be small compared to the laser field. In addition, the laser-induced velocity (6) of the electron remains relatively high most of the time, so that electron trajectories are smoothly deformed by Coulomb attraction. Under such conditions the Coulomb correction to the phase of the wave function can be evaluated in the eikonal approximation along a trajectory found in the absence of the Coulomb field [127, 128] (109) is not Hermitian if only part of the position space is considered. Hermiticity can be restored by adding the Bloch operator [129] to the inner and outer Hamiltonians: ⎛ d b⎞ + ⎟ (110) H → H (±) = H + L(±) , L(±) = ± δ (r − a) ⎜ ⎝ dr r⎠ ΨEVA (p, r , t ) = Ψp (r , t )eiWC (r, p, t ) , with b being an arbitrary constant which escapes the final result. The time-dependent Schrödinger equation (36) then reads ( i ∂ ∂t − H ) Ψ ( r , t ) = − L (±) (±) Ψ (r , t ) r=a . where Ψp (r, t ) is the Volkov function (3) and the Coulomb phase is given by (111) WC (r , p, t ) = Because of the delta-function entering the Bloch operator, the wave function in the r.h.s. of (111) is taken at the boundary r = a. Then, using the respective Greenʼs function one can present these equations in the integral form (see (39)) ∫t td U [ rL (r , p, t , t′) ] dt′ . t ( ) × L(±) Ψ r′, t ′ . rL (r , p, t , t′) = r + (112) MARM (p) = − i t ∫−∞ dt′ ∫ d3r1G (−) ( r, t, r1, t1) × L(−) ψ0 ( r1) r=a e−iε0 t1 . ∫t t′ vp ( t1)dt1 (116) and is shifted by r with respect to the trajectory (64) of the ITM. The wave function (114) with the phase (115) is known in the literature as the EVA solution [127, 128]. The EVA functions can be used for construction of the outer Greenʼs function in (113). Projecting Ψ1( − ) (r, td → +∞) onto a plane wave with the momentum p , one obtains for the amplitude of ionization in the ARM approximation r=a Equations (112) can be approximately solved by applying an iteration procedure. Replacing the inner wave function by the one of the unperturbed bound state, Ψ0( + ) (r, t ) = ψ0 (r)e−iε0 t (where the ground state energy ε0 may include a complex Stark shift) we obtain for the first approximation in the outer region Ψ1( − ) (r , t ) = − (115) The argument of the Coulomb potential U (r) in (115) is a trajectory in the laser field which starts at the time instant t at the spatial point r : ∫−∞ dt′ ∫ d3r′G (±) (r, t, r′, t′) Ψ (±) (r , t ) = − (114) td td ∫−∞ dt ∫ d3rΨp* (r, t )e−i ∫ t ⎛ d b⎞ × δ (r − a ) ⎜ + ⎟ Ψin (r , t ). ⎝ dr r⎠ (113) Substituting this Ψ1( − ) into (112) for the inner part and using there the atomic Greenʼs function, one obtains the first correction Ψ1( + ) to the atomic wave function in the inner region, and so on. If the radius of the inner region is sufficiently small, the series resulting from this iterative procedure is an expansion with respect to the number of hard recollisions experienced by the electron: the function Ψ1( − ) corresponds to no recollision (direct electrons), Ψ2( − ) describes electrons experiencing a single hard recollision, etc. After a hard recollision an electron usually acquires a large drift momentum, so a second hard recollision is very unlikely, let alone a third. Thus the series should converge quickly. In order to perform calculations within this algorithm, an analytic expression for the outer Greenʼs function G (−) (r, t , r1, t1 ) is required. The simplest realization corresponding to the Volkov Greenʼs function (41) will reproduce the ZRP version of the Coulomb-free SFA for direct ionization and (if the next terms Ψ1( + ) and Ψ2( − ) are found) rescattering. A more accurate (and desirable) result can be ⎡ ⎤ U ⎢ rL t ′ ⎥ dt ′ ⎣ ⎦ () (117) Because of the delta-function, spatial integration is confined to the sphere of radius a. This radius should be small enough that the Coulomb and laser fields are of the same order there. Under this condition, integration over the R-matrix sphere and application of the SPM yield the optimal trajectory rL which is identical to the Coulomb-free trajectory of ITM (64). The amplitude (117) has to be compared to the Keldysh (1) and the CCSFA (84) amplitudes. Omitting the Coulomb integral in the phase and replacing the inner wavefunction by the unperturbed initial atomic function ψ0 (r)e iIp t we obtain the amplitude (44) for the ZRP. With the Coulomb field accounted for, the amplitude (117) is different from the CCSFA result (84). Indeed, the Coulomb integral WC is introduced into the phase before the application of the SPM and the phase itself is calculated along trajectories unperturbed by the laser field, in contrast to CCSFA where the trajectories are also corrected. Exact numerical evaluation of the time integral (117) is cumbersome, while the saddle-point analysis remains efficient if the Coulomb field is considered as a perturbation. The saddle-point equation, which now 21 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article includes the Coulomb term, reads − for the full Coulomb factor QC may appear different from (106)), but this has not been attempted yet. Within the method, the Coulomb action can be calculated for arbitrary values of the final momentum p with a result identical to (95). Trajectories corresponding to arbitrary final momenta are complex, so the action WC influences both the probability of ionization and the interference pattern of photoionization spectra. In order to calculate the Coulomb action along a complex trajectory, an analytic continuation of the Coulomb potential in the form 1 ∂ ( S0 + WC ) = 2 v2p ( ts ) + I p + U (a) ∂t + ∫t td U vp (t′)dt′ = 0. (118) s The last two terms are the Coulomb potential energy at the boundary separating the inner and the outer regions and the work done by the Coulomb field along the Coulomb-free trajectory. The Coulomb phase is matched to that of the atomic wave function using a technique similar to that described in section 6.2.1. After this, (118) is solved perturbatively, taking a Coulomb-free solution for the zeroth-order approximation. As a result, the stationary point ts (p) takes the form U (r ) = − Z r = − Z r 2 has been used [121, 132, 133]. A very important advantage of the method, compared to the CCSFA, it that it allows a straightforward generalization to many-electron systems. As the outer wave function is expressed via the inner one, a multiparticle structure of the ts (p) = ts0 (p) + Δts (p), (119) latter can be accounted for in the amplitude of ionization. This allows the study of multichannel ionization of complex atoms where the Coulomb correction is given by [130, 131] and molecules. The corresponding formalism is based on the introduction of direct and indirect (driven by electron–elecW ′ ( t s0 ) Δts (p) = − C′′ . (120) tron correlations) ionization amplitudes for each final ionic S0 ( t s0 ) state and appears to be rather involved technically , so we This value is, in general, complex and its real part was direct the interested reader to [134]. A rather unexpected interpreted as an ionization delay appearing due to decelera- result of the calculations performed there is that in the tion of the electron during its sub-barrier motion [130, 131] nonadiabatic regime γ > 1 correlation-driven channels of ionization of molecules can considerably dominate over the (see also the discussion in section 8.2). These equations summarize the application of the ARM direct channel. In connection to this observation, ab initio and EVA methods to the problem of strong field ionization. numerical solutions of the TDSE for lithium in a strong laser The logic of this theory is close to that of the CCSFA. field [135] show that core excitations are only important at Differences originate mostly from the slightly different relatively high frequencies, K0 ≃ 1, when ionization proceeds in the few-photon regime, while in the multiquantum regime mathematical realizations of the same idea, namely: the ionization dynamics remain essentially single-electron, if (i) After matching, the Coulomb phases WC of both methods the laser frequency does not hit a resonance. Whether or not coincide, but not the corrections to the final photoelec- this qualitative difference reflects the difference between an tron momentum. atom and a molecule or between the applied theoretical (ii) The Coulomb corrections to the stationary point given by approaches, remains an open question. ts′ = ts (p′) in the CCSFA and by (120) are also different, As was already discussed in section 6.2.2, a natural although both are linear with respect to the atomic application for Coulomb-corrected theories is ionization in residual charge Z. intense elliptically polarized fields, where the Coulomb (iii) The ARM method allows one to incorporate rescattering asymmetry is a pronounced qualitative effect. In short pulses, into the Coulomb-corrected theory, while the CCSFA in there is no qualitative difference between elliptical (with ξ ≃ 1) its present form only describes direct ionization. and circular polarization, because of the pulse envelope effect. The predictions of the CCSFA and ARM methods seem Recently, a series of high-precision experiments has been to be close but not identical. If one of the two versions of the performed where ionizations of atoms by short, near circularly theory suffers logical inconsistencies or both do not precisely polarized pulses have been used for verification of the so-called attoclock setup [136–140]. The attoclock idea is, put simply, achieve the ultimate goal, there remains an open question. the following: in a circularly polarized field the direction of the 6.3.2. Applications. The high quantitative accuracy of the most probable photoelectron momentum is determined by the ARM and EVA methods has now been proven by several vector potential of the wave at the instant of ionization tm , applications. First, in the simplest realization, when the when the field is maximal (see also the discussion in Coulomb action is only calculated along the most probable section 6.2.2 and 8.2). Therefore, measuring a photoelectron trajectory, the method naturally reproduces the Coulomb momentum distribution in such a field, one can (provided the factor Q1 or, in other words, the rate of ionization obtained by time dependence of the electric field vector E(t ) is known) Perelomov and Popov [15]. Calculating the corrections to the extract the instant of ionization with a sub-cycle (attosecond) stationary point (120) one should obtain the factors Q2 and accuracy. In order to properly perform such an extraction one, QC from (106) and (107) (to be more precise, due to however, needs to account for the Coulomb distortion of quantitative differences between the AMR and EVA methods photoelectron trajectories. Both the CCSFA and the ARM on the one hand and the CCSFA on the other, an expression methods are perfectly suited for such calculations. The first has 22 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article the eigenvalue problem not yet been applied to the problem, while application of the second has shown an excellent agreement with the results of numerical solutions of the TDSE [130, 131] performed by three different methods: relative deviations in the value of the offset angle calculated analytically from the ARM method and numerically do not exceed 2%, while the predictions of different numerical methods also differ one from another at this level. Moreover, it has been shown that just a 0.1% deviation from the peak intensity shifts the results to a value exceeding the difference between them. ⎤ ⎡ ∂ Δ − U (r ) − E (t ) r⎥ Φϵ (r , t ) = 0 ⎢⎣ i + ϵ + ⎦ 2 ∂t (121) has to be found. Here ϵ = ϵ′ − iϵ′′ is a complex quasienergy. Its imaginary part determines the total ionization rate, Γ = 2ϵ′′. For an SR potential U(r) this equation can be solved by matching a solution for a free electron in the laser field E(t ) valid at r > rc , where rc is the effective radius of the potential (rc = 0 for the ZRP), with a laser-free solution at r < rc . The approximation of the field-free wave function at r < rc is similar to the one made in the Keldysh model but not exactly identical to it. The difference which allows an increase of the accuracy of the solution, compared to the Keldysh ionization amplitude, consists in the matching of the inner and the outer wave functions at r = rc (similarly to the ARM method of section 6.3). For the ZRP, the atomic force acts at the singular point r = 0 where it is divergent, then the respective solution of (121) becomes exact. For an SR potential with rc ≠ 0 (modeling bound states of negative ions with nonzero angular momentum), the solution is approximate but still highly accurate. The procedure of calculation of the wave function in the two spatial regions and their matching at r = rc is described in [29]. When Φϵ (r, t ) is found, its Fourier expansion (at r → ∞) determines the amplitude of ionization with absorption of n photons. For a system bound by zero-range forces and subject to a linearly polarized field 7. Generalized and complementary approaches In the theory of strong field ionization there are several efficient analytic approaches which combine relative simplicity with quantitative accuracy and stay conceptually close to the Keldysh method. Here we sketch two analytic methods known for their successful application to the problem. The method of QQES [29, 141, 142] provides an exact solution of the photodetachment problem for the ZRP and a quantitatively accurate solution for an SR potential treated within the effective range approximation. As far as detachment of negative ions is concerned, the QQES method includes Keldysh theory as a limiting case. It is instructive to trace how the Keldysh result follows from this more general approach. Another method, developed by Berson [143, 144], is based on the semiclassical description of electron motion in the Coulomb field of the atom and perturbatively accounts for laserinduced corrections to the electron action. It can be viewed as an approach complementary to Coulomb-corrected versions of the Keldysh theory. ∞ n (p) = in ∑ ( − 1)k fk k =−∞ ∞ × ⎛ E 0 k n cos θ ⎞ ⎛ zF ⎞ ⎟J ⎟, n + 2 s − 2k ⎜ ⎝ ⎠ 8 ⎠ ω2 ∑ Js ⎜⎝ s =−∞ (122) where θ is the angle between the photoelectron momentum p and the laser polarization and k n = 2(ϵ + nω − UP ) is the complex quantity which turns into the absolute value of the photoelectron momentum when we take ϵ = − I p . Coefficients fk determine the Fourier expansion of the function Φϵ (r, t ) at r → 0 7.1. The quasistationary quasienergy states (QQES) method The concept of quasienergy states (QES) of a quantum system subject to a periodic external field is based on the Floquet theorem [145] and was introduced by Shirley [146], Zeldovich [147] and Ritus [148]. The term quasienergy (introduced by Zeldovich and Ritus) relates to a new quantum number— the Floquet index—which is conserved in a periodic field. There is an obvious analogy with conservation of quasimomentum in a spatially periodic potential. In order to apply the QES formalism to the problem of laser-induced ionization, a continuum of quasienergies has to be considered, because there are no true bound states in the system. This complicates calculations enormously. To avoid this difficulty, a generalization of the QES formalism making it suitable for treatment of problems involving bound–free transitions was suggested in the mid−1970s [149, 150] by introduction of complex quasienergies and the respective QQES. For a complete description of the method we direct the reader to the review [29]. According to the QQES method, in order to describe ionization of a bound state in the potential U(r) a solution of Φϵ (r → 0, t ) = ⎛1 ⎞ ⎜ − 1⎟ fε (t ), fε (t ) = ⎝r ⎠ ∞ ∑ fk e−2ikωt . (123) k=0 Absence of odd Fourier components in the series (123) reflects a specific symmetry of the wave function at the origin: equivalent wave packets of the detached electron return there two times per laser period. Generalizations on arbitrary elliptical polarization and an SR potential are described in [29, 151]. In order to obtain the result of the Keldysh theory from (122) we replace the quasienergy by its field-free value −I p , then kn in (122) coincides with pn from (19). This means that we disregard the Stark shift and the width of the bound state. Second, the Fourier coefficients fk are approximated by fk = δ k,0 2π , (124) i.e. 2ω-periodic oscillations of the wave function at the origin 23 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article Figure 6. Left panel: photoelectron spectra along the polarization direction as functions of electron energy En scaled by Up for H− (s-state; thin lines) and F− (p-state; thick lines) for E0 = 0.2835 and ω = 0.203 (with the photon energy scaled by the ionization potential Ip and the electric field amplitude by (2I p )3 2 ). Solid lines: exact QQES results; dashed lines: Keldysh theory. Solid circles mark the positions of ATI peaks. The rates are shown in scaled units (s.u.): results for F − are multiplied by the factor 6.24 in order that the maximum rates are the same for both ions. From [151]. Copyright (2003) by APS. Right panel: photoelectron spectra along the polarization direction for ionization of Br − in a laser field with the wavelength 1300 nm and intensity 6.5 · 1013 W cm−2. The data are shown by circles, including several points with error bars; the Keldysh result is shown by the dashed line, the SFA result with rescattering by the solid line. The theoretical distributions are normalized to the data at the maximum of the signal. Reprinted with permission from [83]. Copyright (2010) by the American Physical Society. corresponding to ionization in a field with relatively low intensity and high frequency, was developed by Berson [143, 144]. The term in the action describing interaction between the electron and the laser wave is induced by the photoelectron wave packets returning to the core each laser half-period are neglected. Thus approximation (124) does not describe rescattering7. After these simplifications the amplitude (122) turns (21) with accuracy into a factor reflecting different definitions of the ionization amplitudes. The left panel of figure 6 shows comparisons between the photodetachment spectra of H− and F− ions calculated using the QQES approach and the Keldysh model. As expected, the Keldysh result does not reproduce the rescattering plateau which dominates in the spectrum for energies ε > (4 ÷ 5) UP . For lower energies, where the direct ionization mechanism dominates, the agreement is excellent. The right panel of this figure shows a comparison of the photodetachment spectra of Br − calculated along the Keldysh model with the data recorder in [83]. This comparison proves that for SR potentials, Keldysh theory yields quantitatively accurate results. For a review of the further development of the QQES approach, including its generalizations for atoms and molecules, we direct the reader to [152–156] and references therein. Wint = − ∫ E (t ) rC (t )dt (125) with rC (t ) being the electron trajectory in the atom. Using classical trajectories for the description of atomic motion in bound states requires semiclassical conditions, therefore (125) assumes a highly excited (Rydberg) initial state. The amplitude of ionization with absorption of n laser photons Mn ∼ ∫ dt ei[W int(t ) + nωt ] ∼ Jn (B), (126) where Jn is the Bessel function and B has to be found from (125). To this end the method suggested by Kramers for calculation of single-photon ionization [157] can be used. In a Rydberg state the electron moves almost along a parabola. Taking a parabolic trajectory from a parametric solution of the respective classical problem [99] ⎛ 1 ⎞2 1⎛ 1 ⎞2 xC (η) = ⎜ l + ⎟ η , zC (η) = ⎜ l + ⎟ 1 − η2 , ⎝ 2⎠ 2⎝ 2⎠ 3 ⎛ 2 1⎛ 1⎞ η ⎞ t = ⎜l + ⎟ η ⎜1 + ⎟ ⎝ ⎠ 2 2 3⎠ ⎝ ( 7.2. Semiclassical theory of ionization The EVA and CCSFA approaches are based on a perturbative account of the Coulomb interaction in the classical action of a photoelectron in the field of a strong laser wave. In both methods the laser-induced action (45) is taken as a zeroth approximation. It makes sense to consider the opposite limit when the Coulomb field is treated as the main interaction, while the effect of the laser field is accounted for within the perturbation theory for the action. Such an approach, ) one can calculate the integral (125) and extract the coefficient B at sin ωt . For a linearly polarized field with the polarization direction specified by angles θ and φ respective to the axes z and x this calculation gives [143] ⎛ 2 ⎞5 3 B = πE 0 ⎜ ⎟ Ai′ 2 (u) cos2θ + uAi2 (u) sin2θ cos2φ (127) ⎝ω⎠ 7 An account of the Fourier components fk with k ⩽ (3.17UP + I p ) 2ω reproduces the standard rescattering plateau predicted by the generalized SFA. where u = (l + 1 2)2 (ω 2)2 3 and Ai(u) is the Airy function. 24 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article Note that the ionization potential I p ≪ 1 does not enter this expression. It is assumed that B ≪ 1, so the Bessel function can be replaced by its expansion for small arguments. Then, after integration over the angles θ and φ and summation over the orbital quantum number l ⩽ N − 1 (N ≫ 1 is the principal quantum number, I p = 1 2N 2 ) one obtains for the rate of n-photon ionization [143] wn ( E 0, ω , N ) = E 02n Tn This is not so for the Coulomb potential, as its Fourier transform 〈p|U |q〉 present in the amplitude (74) diverges for small momentum transfers, |p − q| ≪ p. As a result, at low photoelectron energies εp ⩽ UP the amplitude M1 is not a correction. A gauge transformation leads to some exchange between the terms MK and M1. Adopting upper indexes l and v for the length and the velocity gauge, respectively, we can write (128) N 5ω(10n + 2) 3 M l = MKl + M1l = eiΦM v = eiΦ ( MKv + M1v ), where Tn is an n-dependent number. The respective expression for circular polarization can be found in [143]. For n = 1 this result recovers the Kramers’ formula for single-photon ionization. However, at n ≫ 1 there is no matching to the multiphoton limit of the CCSFA. The latter follows from (108) at γ ≫ 1 [118]: but it is generally MKl ≠ eiΦMKv , E 02n( 2eK0) ω2n + 2ν − 3 2 . M1l ≠ eiΦM1v . Here Φ is a phase generated by the transformation, | exp (iΦ ) | = 1. For an SR potential |M1 | ≪ |MK | for direct ionization, so it cannot change MK significantly on transformation, as a result the theory is only weakly gauge-noninvariant, while for the Coulomb potential it is strongly noninvariant. Although the gauge invariance problem in strong field ionization has been addressed in a number of papers (see [159–164] and references therein), there are still a lot of misunderstandings and erroneous statements. Here we consider three viewpoints commonly accepted in the literature: 2ν wn ( E 0, ω , ν ) ≈ T˜n (130) (129) Here T˜n is another n-dependent coefficient [118]. The power dependence of the rates (128) and (129) on the laser frequency is essentially different. The reason for this disagreement is that the applicability domains of the two theories do not overlap. Equation (125) implies that the characteristic size of the electron orbit in the atom is larger than the quiver amplitude in the laser field. The technique of Coulomb corrections used in the CCSFA assumes the opposite condition: the size of the initial bound state is negligible compare to all other spatial scales of the problem. The amplitude (127) is proportional to the parameter E0 ω5 3 which is supposed to be small in the theory, indicating that it is applicable in fields with low intensities and relatively high frequencies. The same parameter appears in semiclassical calculations of WKB matrix elements in the Coulomb field [56, 158]. In this context, the border between semiclassical methods, in essence close to the one described above and strong field Keldysh-like theories, is defined by the condition E0 ≃ ω5 3. (i) The problem of gauge invariance is simply misundersood. Under a proper treatment of the full Hamiltonian of the system ‘atom+field’, the velocity and the length gauge versions of the SFA ionization amplitude are equivalent. (ii) There is a preferable gauge, which has to be used in calculations. (iii) The theory is essentially noninvariant and none of the of the interaction operator forms (2), (4) provide a correct result. An essential modification of the theory is needed in order to restore the property of gauge invariance. Below we comment these statements and argue in favor of the third, although the second one is also partially correct. The first statement is fully erroneous. Under a certain choice of partition of the full Hamiltonian of the system, the amplitude of ionization (1) appears gaugeinvariant [160, 165]. For example, in the velocity gauge the Hamiltonian of the problem can be presented in the forms 8. Difficulties and open questions 8.1. Gauge problem The probability distributions following from the amplitude (1) depend on the gauge chosen for description of the electron–laser coupling. Exact gauge invariance is only achieved for the special case of the ZRP (see section 2.1). There is nothing surprising in the fact that a nonperturbative approximate theory does not possess the property of gauge invariance. The Keldysh amplitude MK differs from the exact unknown amplitude M by a term M1 (also unknown) which is in general not small in absolute value compared to MK. In order to estimate the magnitude of M1 we may approximate it by the amplitude of rescattering (74). From what we know about rescattering, if the atomic potential U(r) is an SR, then |M1 | 2 ≪ |MK | 2 for the direct part of the ionization spectrum. H= 1 (p + A (t ) )2 + U (r ) = Hi + Vi = H f + V f , (131) 2 where the partial Hamiltonians Hi, f and Vi, f can be chosen differently for the initial and the final Hilbert space, for example Hi = H − E (t ) r , Vi = E (t ) r (132) and Hf = 1 (p + A (t ) )2 , V f = U (r ). 2 (133) With this partition we obtain for the eigenfunctions of the 25 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article initial unperturbed Hamiltonian Hi ψ˜ i (r , t ) = ψi (r)e−iεi t − iA(t ) r , that the statements (i) and (ii) are in general erroneous. For nonzero frequencies there is no analytic benchmark, however the results of the SFA or the Keldysh model in any gauge can be compared to numerical solutions of the TDSE. Such comparisons, presented widely in the literature, show that in the case of atoms (not to say molecules and more complex systems), the results of the SFA agree with exact solutions only qualitatively, while some features of momentum distributions are not reproduced at all (see the discussion in section 6). Note that there are features of ionization spectra stable with respect to the gauge transformations. This includes, for example, the absence of every second ATI peak for lateral emission in a linearly polarized field [166, 167] or a smooth spectrum with the maximum around UP in a flied with circular polarization [90]. On the other hand, the predictions of the Keldysh theory in the length gauge agree quantitatively with exact calculations and data for negative ions (see an example given on figure 2). Combining these observations together we can formulate the following conclusions: (134) where ψi are the eigenfunctions of the atomic Hamiltonian. Taking initial states in the form (134) one obtains the amplitude of ionization (1) in the length gauge. From the other side, staring from the length gauge Hamiltonian and using the partition Hi = p2 2 + U (r ), Vi = E (t ) r , H f = p2 2 + E (t ) r , Vi = U (r ) (135) one again obtains the length gauge form of the amplitude. This argumentation ‘proves’ that the theory is gauge-invariant (on proper partitioning) and that the correct form is given by MlK. Thus, this treatment of the gauge problem combines the viewpoints (i) and (ii). In our opinion, the described arguments do not solve the problem, because the choice of the gauge is replaced there by the choice of the partition. Indeed, there is no difficulty in showing that the Hamiltonian Hi from (132) does not contain any interaction with the laser field and therefore is equivalent to the atomic Hamiltonian obtained from the latter using a unitary transformation. There is an infinite number of such Hamiltonians, so that by choosing different partitions one will arrive at different forms of the ionization amplitude. Thus the technique of partitioning simply redefines the problem without solving it. It can be unambiguously shown that for atoms both the velocity and the length forms yield incorrect (and different) results. To this end we consider the static limit γ → 0 and calculate total ionization rates using the saddle-point expressions of section 5 and appendix A for the ground state of hydrogen (ν = 1, l = m = 0 ). In the velocity gauge there is no divergency in the prefactor P and one obtains from (53) after integration over photoelectron momenta wv = Cv 4 I p ⎛ ω ⎞ −2 3F ⎜⎜ ⎟⎟ e . F 9/2 ⎝ I p ⎠ (i) When the Coulomb interaction is present, the Keldysh model and the SFA do not possess gauge invariance and neither the length (2) nor the velocity gauge (4) give a correct result. (ii) However, in the saddle-point expression for the amplitude, only the prefactor (p, ts ) is gauge-dependent, while the exponential factor e−iS0(p, ts ) is invariant. As the exponential factor is much more sensitive to atom and laser parameters than the prefactor, incorrectness of the latter usually causes secondary-order effects. This explains why, being gauge-noninvariant, the Keldysh theory can be used for reliable calculations of the photoionization spectra of atoms. (iii) In order to describe the ionization of atoms correctly, one must not apply the amplitude (1) for a corresponding atomic state in any gauge. Instead, the respective amplitude for a state with the same I p, l, m in an SR potential should be calculated and then corrected applying the techniques described in section 6. (136) The same calculation performed in the length gauge involves (A.8) and gives w l = Cl I p F e−2 3F . The next question which now arises naturally is: what gauge should be used for the calculation of the respective amplitude in an SR potential (if it is not the ZRP). The results of the numerical study [159] show that the length gauge (2) should be chosen. Indeed, comparisons with exact TDSE solutions demonstrate that with increasing radius rc of an SR potential, the disagreement between them and predictions of the velocity gauge version of SFA grows fast, so that already for rc = 2 a.u. the velocity gauge gives an essentially erroneous result, while the length gauge version of the SFA stays in good agreement with the numerical data8. This conclusion can be supported by comparisons with multiple experimental results on detachment of negative ions. The example shown in figure 7 also provides a clear argument in favor of the length (137) Here Cv, l are numeric coefficients. The quasistatic limit for the ionization rate of hydrogen in a linearly polarized field follows from (108) for v = 1, l = m = 0 and γ = 0 : wst = 8I p 3 −2 3F e . πF (138) It differs from the rate of ionization in a constant electric field [40] by the factor 3F π appearing due to the averaging over the laser period. Comparing these three results we find that the prefactors in (136) and (137) are essentially incorrect (the velocity gauge result even turns to zero as ω4!). This constrains one to conclude that none of the gauges provide a quantitatively correct description in the case of atoms. This example proves 8 As expected from the general analysis above, with a further increase of the effective potential size the length gauge result becomes different from the exact solution as well. 26 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article emphasized however, that such arguments can only make sense within an approximate theory. Another, rather mathematical argument in favor of the length gauge for SR potentials was found in [151]: the QQES approach is explicitly gauge-invariant, however, the matching procedure necessary for actual calculation of the eigenfunction Φϵ (r, t ) can only be performed in the length gauge representation. 8.2. Tunnel exit, Keldysh tunneling time and the simple-man model (SMM) An invaluable insight into strong field physics including recollision has been made by the SMM. The scheme of this semiclassical method [61, 168–170] is remarkably clear: the process of ionization is considered as a sequence of either two or three steps. The first step is quantum-mechanical tunneling of the electron through the potential barrier created by the laser field. The tunneling probability can be calculated at any time instant using the quasistatic rate (63) corrected by the Coulomb factor (106) (i.e. the tunneling limit of the PPT rate) where the electric field E(t0 ) at the time instant t0 is substituted. The electron appears at the end of the barrier, the point known as the tunnel exit. Considering the laser field to be quasistatic and ignoring the Coulomb distortion of the barrier, we obtain for the exit point: b 0 ≡ b (γ → 0 ) = − I p E ( t0 ) E 2 ( t0 ) (139) The time interval the electron needs to cover this distance moving with the atomic velocity vat = 2I p Figure 7. Comparison between the photoelectron spectra recorded for the case of photodetachment of F− in a linearly polarized field of the wavelength 1400 nm and intensity 1.7 · 1013 W cm−2 for emission along (circles) and perpendicular (crosses) to the polarization axis. The predictions of the SFA with rescattering are shown by solid and dashed lines correspondingly. The upper panel shows the theoretical results obtained in the length gauge and the lower panel in the velocity gauge. The experimental data are identical for both panels. Reprinted with permission from [82]. Copyright (2007) by the American Physical Society. Δt K ≃ b0 vat ≃ γ ω (140) is known as the Keldysh tunneling time. In the tunneling limit it is equal to the imaginary part τ0 of the saddle point ts. This value gives an estimate of the time necessary for establishing of the static-field ionization rate when the electric field turns on instantly [174]. The second step is a classical motion of the photoelectron from the tunnel exit to a detector. The third step comes into play if during this motion the electron revisits the parent ion—a phenomenon known as recollision. A recollision can result in elastic scattering, recombination with the emission of a high harmonic photon or inelastic scattering with excitation or further ionization of the parent ion. This simple model allows one to investigate the kinematics of strong field ionization and recollision and to predict the positions of classical cut-offs of rescattering spectra and spectra of high-order harmonics [61, 63, 169]. The model is particularly flexible for generalizations, namely for inclusion of the Coulomb force into the classical kinematics of the photoelectron after its emergence at the tunnel exit, Coulomb correction of the tunnel exit position and inclusion of the dipole part of the electron–ion interaction. A number of effects in the spectra of direct and rescattered photoelectrons have been described applying the simple-man approach (see, e.g. [69, 71, 74, 136, 138–140, 171, 172] and references therein). It is apparent from analysis of the quantum orbit representations of the SFA [101] or, equivalently, the ITM versions gauge. Thus the viewpoint (ii) is partially correct: for SR potentials there is indeed a preferable choice of gauge. Why does the length gauge work better than the velocity gauge? The reason for this can be understood through the saddlepoint analysis [16]: in the length gauge the Fourier transform of the bound state wave function is singular at the saddle point. Therefore, in this case we need not know the behavior of the atomic wave function in the whole space, but only at distances r ≃ 1 κ where it is given by the asymptotic (A.2). In contrast, the velocity gauge form involves a p -dependent transform of the wave function which, for its correct calculation, requires knowledge of the wave function in the whole position space, including the region where it is essentially perturbed by the laser field. The qualitatively erroneous factor (ω I p )4 in (136) appears from the momentum dependence of the ground state Fourier transform. In addition, as is noted in [159], in the length gauge the electron does not interact with the laser field prior to ionization, because the operator of momentum present in the atomic Hamiltonian is equivalent in this case to the field-free kinetic momentum. It should be 27 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article of the Keldysh theory that the SSM follows from these quantum-mechanical approaches in the limit of tunneling, γ ≪ 1. Here we address two questions: (1) how should one modify the SMM equations in order to make them applicable at arbitrary γ and (2) are the concepts of the barrier exit point and the tunneling time (as they appear in the Keldysh theory) physically meaningful? The answer to the first question is straightforward. As the SMM is based on the same physical assumptions as the Keldysh model (with the Coulomb field included if necessary), correct classical modeling must deal with real parts of complex classical trajectories9. Therefore the initial time instant for classical photoelectron motion is given by the real part of the stationary point satisfying (52), t0 = Re ts . The tunnel exit point is a real part of the electron position at this time instant, b = Re [r0 (p, t0 )] given by (64). The initial photoelectron velocity at the tunnel exit is given by v0 = p + A(t0 ), etc. In a linearly polarized monochromatic field E (t ) = E 0 sin ωt one obtains E0 sin ωt0 ( cosh ωτ0 − 1) , ω2 v0 ≡ v ( t0 ) = p + A ( t0 ) = v⊥ + ev b ( t0 ) = − = v⊥ − E0 cos ωt0 ( cosh ωτ0 − 1) , ω The second question is more difficult. Both values, b and ts, emerge in the Keldysh theory after application of the SPM for calculation of the time integral and associate with the photoelectron trajectory (64). It is usually assumed that the integration path connecting ts and td is chosen as shown in figure 2(a) by a solid curve: a vertical segment connecting the stationary point ts with its real part t0 is followed by a horizontal segment along the real time axis. With such a path, a seemingly natural separation into the ‘sub-barrier’ (t ∈ [t0 + iτ0, t0 ]) and ‘after-barrier’ motion (t ∈ [t0, td ]) is introduced. Within this picture, the SMM accounts for the electron dynamics along the second shoulder of the path, while the sub-barrier part is reduced to an instant emergence of the electron at the tunnel exit with a given probability. There are several alternative treatments of the sub-barrier dynamics present in the literature, some of them involve interpretations of the Keldysh tunneling time τ0 as a real interval of the physical time the electron spends between the ‘moment of ionization’ and the time instant t0 when it starts its classical motion. These concepts considering t0 and b as physically meaningful and even indirectly measurable values can, however, be seriously criticized. Indeed, the trajectory (64) and the respective actions (45) and (71) are analytic functions in the whole complex time plane. This means that the saddlepoint result depends only on the initial and final points ts and td but not on the form of an integration path connecting these points. Any path different from the standard (a few examples are shown on figure 2) can be chosen. As the tunnel exit position and the time t0 when the electron arrives at the real time axis depend on the path, this invariance of the Keldysh ionization amplitude with respect to the path deformation means that both values b and t0 have no physical meaning in the sense that there can be no experiment whose results would allow concluding even indirectly that these values are equal to this and that. This does not contradict the fact that the introduction of such values into the theory can be useful for calculations or for development of pictorial interpretations. The situation becomes complicated (however, without a principle change) when the Coulomb interaction is accounted for via corrections to the classical action (as described in section 6.2 and 6.3). The full action (85) is no longer an analytic function, as the functions 1 r0 (p, t ) and 1 r03 (p, t ) have branching points in the complex time plane. In this case the integration path is not arbitrary but should be chosen so that it does not intersect cuts [120]. Note, that the Coulomb corrections to the action (95) and to the photoelectron velocity (96) are presented in a form independent on the integration path (provided it does not intersect cuts) and therefore they do not depend on the position of the tunnel exit as well. A question naturally arises: how can classical simpleman calculations based on the use of a physically meaningless (as shown above) concept of the tunnel exit yield good agreement with experimental data and numerical results? Indeed, within the SMM, Coulomb-corrected trajectories are (141) (142) where e is a unit vector along the polarization direction. These values are momentum-dependent (via the time of start t0 connected to the final momentum). In the tunneling limit, ωτ0 ≪ 1, (141) gives (139) and the longitudinal initial velocity vanishes as γ2, so that this justifies the common approach where it is set equal to zero. For γ ≃ 1, both the tunnel exit position and the initial velocity differ considerably from their tunneling values. In the multiphoton limit γ ≫ 1 the exit (141) becomes field-independent, b = 2I p ω. At a fixed field amplitude the barrier gets narrower when ω and γ grow, although the Keldysh tunneling time remains constant. Longitudinal initial velocity v|| is, as follows from (142), an unambiguously defined function of the start time t0. No longitudinal velocity spread at the instant of ionization takes place. Therefore, calculations where this spread is heuristically added introduce an error instead of an inprovement. In contrast, transversal spread is inherent to the classical model due to the fact that v⊥ does not depend on t0. Equations (141) and (142) can easily be generalized to arbitrary polarization of the laser pulse; then the transversal initial velocity v⊥ is perpendicular to the electric field vector E(t0 ). Summarizing this part, when γ is not small, the initial conditions (141) and (142) following from the trajectory representation of the Keldysh model differ substantially from their tunneling values adopted in the SMM. Already at γ ≃ 1, the difference can have a much greater influence on the photoelectron classical kinematics than the Coulomb correction to the tunnel exit position and the dipole correction to the electron–ion interaction. 9 Neglecting imaginary parts of the ionization times and trajectories, which is typical for classical models, makes them less accurate compared to the Keldysh model and the SFA (see [173] in this context). 28 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article usually found by solving Newtonʼs equation in two fields, restrictions in the application of the Keldysh model for the description of molecules in intense laser pulses. These Zr r¨ = − E (t ) − 3 (143) restrictions already become apparent for the simplest case of a r diatomic molecule with frozen nuclei. As discussed in section 8.1, the gauge dependence of the ˙ with initial conditions r (t0 ) = b , r (t0 ) = v0 with v0 taken matrix element leads to an erroneous pre-exponential factor in from (142) or v0 = 0 . Obviously the final photoelectron momentum found in this calculation depends on the values of the rate of ionization. For atoms, a practical recipe for b and t0. However, if the result is correct it can not be obtaining a quantitatively correct result consists in the use of b -dependent. In order to clarify this seeming controversy we the length gauge version in an SR potential with a subsequent compare the barrier width with the distance between the Coulomb correction of the exponent. Simple arguments show electron and the ion at the matching time instant t* used in the that in the application to the ionization of a diatomic moleCoulomb-corrected theory of section 6.2. Combining (89), cule, the length gauge version of the SFA will immediately (93) and (141) we find that, in the tunneling regime, the two lead to a contradictory result [181]. This argumentation is length parameters, the spatial matching point r* and the bar- particularly transparent for a molecule with a large internuclear separation, 2a ≫ 1 κ . In this case the ground state rier width b, obey the same inequality wave function in a two-center potential 1 ≪ κr *, κb ≪ 1 γ 2F , (144) U (r) = U0 (r − a) + U0 (r + a) can be approximately preso the tunnel exit can be taken as a starting point for the sented as a superposition of the two atomic functions, each electronʼs motion, as well as any other spatial point from the located at its center, vicinity of a circle, shown in figure 2(b). In other words, in (145) ψ0 = A ⎡⎣ ψ0 (r − a , t ) ± ψ0 (r + a , t ) ⎤⎦ . the tunneling regime solutions of (143) are weakly sensitive to the value of the initial coordinate. In the multiphoton Using this initial state for calculation of the Keldysh ampliregime, the relation is different, r * ≪ E0 ω2 and b ≫ E0 ω2 tude with the interaction operator (2) and the Volkov function (the diagram in figure 2(b) corresponds to this limit). Then the (3), we find that the ionization potentials of two contributions integrals in (88) and (96) contain a substantial ‘sub-barrier’ are effectively shifted by ± E (t ) a , because the saddle-point contribution which is neglected in simple-man calculations. equations read The results of the classical modeling along (143) become 1 (146) ( p + A( ts))2 ± E (t ) a + I p = 0. essentially dependent on the choice of b and, generally 2 speaking, incorrect—a fact which has been widely noticed in the literature and received various interpretations. Thus, the This result is apparently unphysical. The problem is solved by results of the Coulomb-corrected Keldysh theory, although dressing the bound states according to the rule they do not depend on the width of the barrier, appear virψ0 (r ± a) → ψ0 (r ± a)e∓iA(t ) a . (147) tually identical to those of the SMM in the tunneling regime. For γ ⩾ 1 predictions of the two approaches can differ With this ansatz the saddle-point equation takes its standard considerably. form, and the amplitude reads Closing this subsection we have to emphasize that the M (p) = A e−ipa ± e+ipa MK (p), (148) conclusion made here about the unphysical nature of the tunnel exit and the tunneling time as they appear in Keldysh theory does not mean that the problem of time delays in where MK (p) is the single-center contribution given by (1). strong field processes has no physical content. Experimental The factor in parenthesis describes to some extent the orienand theoretical investigations of time delays in ionization tation dependence of the ionization rate [181]. Comparisons have raised great interest. The discussion of this problem with the data also support the dressed length gauge verbeyond the Keldysh approximation requires a separate sion [182]. The method of dressed wave functions is, however, review. For the problem of delays in ionization and in limited by the case where the internuclear separation is large, quantum and classical systems in general, we direct the reader so the wave functions of the two centers almost do not to [175–177]. overlap. If this condition is not satisfied, neither of two versions of the SFA applies: the dressing recipe of the length 8.3. Ionization of molecules and dielectrics gauge is not valid and the velocity gauge version, though The physics of molecules driven by intense laser pulses is seemingly not problematic, gives an incorrect result, as it does currently one of the pivotal parts of strong field science. The for atoms. The predictions of the two gauges diverge even ionization dynamics of even the simplest H2+ molecule is more than for the case of atoms [183]. Thus in the current qualitatively much richer than that of an atom, due to the status of the theory, a direct application of the Keldysh matrix effects of the orientation of the molecule with respect to the element for the description of ionization in molecules is laser field and of nuclear motion. The literature devoted to questionable. In the tunneling domain, quasistatic ionization models strong field molecular phenomena is enormous. For overviews of this research field we direct the reader to [178–180]. accounting for the structure of molecular orbitals are widely The only question we address here is related to the principal used in calculations instead of the Keldysh approach. Another ( 29 ) J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article possibility for a theoretical breakthrough is connected to the application of the ZRP model with two or more centers. For a two-center negative molecular ion, this model was introduced in [184] and later developed in application to a circularly polarized field [185]. Recently, it was shown [156] that the QQES approach described in section 7.1 can be generalized to the case of two ZRP in an arbitrary polarized laser field. Within the QQES method the problem becomes exactly solvable and therefore does not suffer the above-described difficulty of the ionization potential shift. This makes the method applicable at arbitrary internuclear separations. In [156], general equations determining the complex quasienery were derived and studied in the low-intensity limit. The question of how the Keldysh limit emerges from this theory (as it does for a single ZRP, see section 7.1) is not yet solved. This work is currently in progress. The situation becomes even more complicated when ionization of macroscopic systems is considered. In the work by Keldysh [1], his method was applied to the description of inner ionization in dielectrics, i.e. a nonlinear excitation of electrons from the valence to the conductive band. The formalism is similar to the one for atoms with the following difference: now both the initial and final states are approximated by Bloch waves where the quasimomentum p is replaced by the time-dependent velocity vp (t ) = p + A (t ) dressing procedure can be used to eliminate unphysical terms from the saddle-point equation. For solids with normal density, where the interatomic distance is comparable to the atom size, this representation is no longer possible and the dressing method does not apply. On the other hand, because the Bloch functions are spatially delocalized, ionization described as a transition between the states (149) looks more similar to a free–free transition in the presence of an external laser field. In a free space (which corresponds to the choice up (r) = const ) the gauge invariance property would be restored. It is unclear, however, how much a periodical spatial dependence of the Bloch function would influence this property. Investigation of simple model examples (not yet performed to the best of our knowledge) where the spatial matrix element can be explicitly calculated in both gauges could be extremely useful for understanding this currently open question. In practice, only the tunneling limit of the theory when the rate of excitation is determined by the standard tunneling exponent [186] is reliably checked. Note that the dispersion law (150) has the same form as in relativistic mechanics of free particles. This formal similarity allows one to apply the mathematical methods developed in the relativistic theory of ionization and pair production in superintense laser fields to the theory of laser-driven dielectrics. This interesting analogy has been pointed out in [4]. In contrast to atoms, inner ionization of dielectrics is tricky for a careful experimental investigation because the interaction of exited electrons accelerated by the laser wave with the lattice causes additional effects. Only relatively recently, thanks to the common use of femtosecond lasers, direct investigation of multiphoton inner ionization of crystals has become possible (see, e.g. [187, 188] for details and references). Studies of the laser–crystal interaction reveal a number of new phenomena including the orientation dependence of nonlinear ionization, the development of avalanches in the femtosecond regime, etc. For their description a theory going far beyond the matrix element (1) with the functions (149) is required. Ψ pc, v (r , t ) = u vcp,(tv) (r) exp i ⎡⎣ vp (t ) r { − ∫ t ( ( )) ε c, v vp t ′ ⎤⎫ dt ′⎥ ⎬ . ⎦⎭ (149) Here upc, v are periodical Bloch functions for the conductive and the valence band and ε c, v are the respective energies approximated by a simple dispersion law ε (p) ≈ Δ 1 + p2 mΔ (150) with m and Δ being the effective electron mass and the forbidden band width, respectively10. In contrast to the Volkov function, (149) is not an exact solution to the Schrödinger equation so this approximation is rather similar to the CV ansatz discussed in section 6.1. In combination with the fact that explicit forms of the Bloch functions up (r) are rather complicated even for model cases, this makes the question of gauge dependence harder than it is for molecules. On the one hand, for solids with a relatively large interatomic separation the wave functions of the valence and the conductive bands can be approximately represented as superpositions of individual atomic functions each localized around the respective atom. Then a problem similar to the one discussed above for the case of molecules appears: in the amplitude of ionization which will have a form of the sum of partial amplitudes for all possible pairs of atoms, each amplitude will carry the factor E (t )(R i − R j ) in the exponent (in the length gauge representation). Here R i, j are the positions of the respective atoms. Then a generalization of the 9. Conclusions and outlook Perhaps the secret of the enduring 50 year success of the Keldysh theory is that it precisely fulfills the criterion of ‘making things as simple as possible, but not simpler’. Although a number of modifications and generalizations of the model have been developed over the past few decades, some of which are described in this paper, none of them— apart from the SFA with rescattering and the corresponding simple-man recollision scenario—have achieved the exceptional combination of simplicity and universality inherent to the Keldysh ionization ansatz. It is remarkable that even the gauge was properly chosen, although this became apparent only recently. In its current form, with the technique of Coulomb corrections, the Keldysh theory describes the ionization of negative ions and atoms with a quantitative accuracy. For the 10 Note that this approximation does not account for the anisotropy of the electron effective mass. 30 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article V Frolov and O Smirnova who read the manuscript in the preparatory stage and suggested a number of improvements. The work was supported by the President Program for Support of Young Russian Scientists and Leading Research Schools (grant no. MD-5838.2013.2) and the Russian Foundation for Basic Research (project no. 13–02-00372). Coulomb-free version of the theory, which implies an SR electron–atom interaction, the only restriction of its applicability is the relative smallness of the reduced field strength, F ≪ 1. Precisely this restriction often ignored in applications. When the Coulomb field is taken into account, additional limitation emerge, including K0 ≫ 1. Besides, it is required that the size of the atomic bound state is small compared to the electron quiver amplitude in the laser field—a restriction making the theory invalid for the description of ionization from highly excited Rydberg states. In the application to the ground state this restriction means, according to (89), that K0 ≫ ν . The Keldysh parameter can meanwhile be formally arbitrary. Looking to the future, we try to outline possible prospects for the further development of the theory. There are several conceptually simple (although important for applications) improvements waiting to be realized. This includes extension of the Coulomb-corrected version of the theory into the relativistic domain (a correction to the total ionization rate and a spin correction at relativistic conditions have already been found [4]), trajectory analysis of the CV version of the theory, examination of yet unexplained features of ATI, like the recently observed very-low-energy structure [189], further application to ionization in multimode fields including XUV +IR schemes with free-electron lasers, and more. These tasks relate to a routine extension of the theory. There are, however, at least three directions for fundamental improvements. They include: Appendix A. Saddle-point expression for the prefactor A saddle-point calculation of the amplitude (1) involves two steps. First, evaluation of the spatial integral (50) gives a preexponential time- and momentum-dependent factor. In both gauges, taking first the time integral in (1) by parts and then integrating over the spatial coordinates we obtain ⎛ s2 ⎞ P (p, t ) = i ⎜ I p + ⎟ ψ0 (s). 2⎠ ⎝ (A.1) Here s = p (velocity gauge) or s = vp (t ) (length gauge). For an SR potential with the radius rc the bound state wave function at r > rc is ψκ, l, m (r) ≈ 2Cκl κ 3 2(κr )ν − 1e−κr Ylm (n), (A.2) where κ = 2I p , ν = Z κ is the effective principal quantum number, Z is the charge of the atomic residual (Z = 0, ν = 0 for negative ions, Z = 1 for neutral atoms, etc) and Ylm is the spherical function with n = r r . A dimensionless asymptotic coefficient Cκl can be taken from tables (see, e.g. table 1 in [4]) or calculated from the Hartree approximate formula [194] (Cκl = 1 for hydrogen and 1/ 2 for the ZRP) • The development of a fully gauge-invariant version of the Keldysh ionization model. This would make the theory applicable to the description of the ionization of molecules, nanoparticles and bulk dielectrics to the same extent as is currently achieved for atoms. • Incorporation of bound states. Currently this is only performed for a trivial case when the initial state is a superposition of two bound atomic states. This extension would allow a consistent theoretical description of intermediate resonances, of ionization from Rydberg states and of recently discovered frustrated tunneling ionization [190]. • The development of a many-body version of the theory allowing the inclusion of correlation effects on an equal footing with the effect of a strong electromagnetic wave. The very first steps towards this goal were made by the theory of collective tunneling [191–193] and by the ARM method [134]. Cκ2l ≈ 22ν− 2 , x != Γ (x + 1). ν(ν + l) ! (ν − l − 1)! (A.3) The Fourier transform of (A.2) reads [13] ψκlm (s) = Cκl 2ν + 3/2κ 2ν + 1 2ν! Ylm ( n s ), π s2 + κ 2 ν+1 ( ) (A.4) where n s is the unit vector along s . Second, the integral is taken by the SPM +∞ ∫−∞ None of these tasks is currently being explored and there is no clear understanding of how one could do this, though a solution to any of them will give new life to the theory. P (p, t )e−iS0(p, t ) dt ≡ ∑ ( p, tsα )e−iS ( p,t 0 sα ) . (A.5) α For ν = 0 the prefactor (A.1) is regular at the saddle point v 2p (ts ) = − κ 2 and application of the SPM gives immediately (54). When ν ≠ 0 , the prefactor is singular at the saddle point. A correct saddle-point calculation requires expanding the photoelectron velocity in the denominator of (A.4) Acknowledgements The author acknowledges numerous discussions with D Bauer, C F Faria, A M Fedotov, M Ivanov, B M Karnakov, N L Manakov, V D Mur, V S Popov, H R Reiss and B A Zon. He is particularly grateful to W Becker, S P Goreslavsky, M v2p (t ) ≈ −κ 2 − 2vp ( ts ) E ( ts )( t − ts ) = − κ 2 − 2S0′′ ( ts )( t − ts ) . 31 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article This leads to the integral Next, expanding the Coulomb-free action +∞ Iν = ∫−∞ e−iS0(p, t ) dt . ( t − t s )ν Using the formula ( ) and taking into account that 1 1 = xν Γ (ν ) ∫0 +∞ ∂S 0 = ∂p dyy ν − 1e−xy , − 2π i 2iS0″( ts) S0″ ( ts ) Γ (ν 2) 2 Γ (ν ) ( ν 2 −iS ( p, t ) 0 s ) e ∫t td (p + A (t ) ) dt = r0 ( td ) − r0 ( ts ), s ∂S 0 ( p, t = ts ) = 0 ∂t s one obtains for the integral (A.6) in the saddle-point approximation: Iν ≈ ⎞ ∂S 0 ∂S 0 ⎛ p′ − p + ⎜ts′ − ts ⎟ ∂p ∂t s ⎝ ⎠ ( ) S0 p′, td ≈ S0 ( p, td ) + (A.6) (B.4) we obtain . (A.7) ( ) S0 p′, td ≈ S0 ( p, td ) − Δp [ r0 ( td ) − r0 ( ts ) ] . (B.5) Instead of (53) we obtain for the amplitude of ionization [16] Collecting all the terms and taking into account that v1(td ) = Δp we obtain: MK (p) ≈ Cκl i1/2 − 3ν /223ν /2 + 1 κ 2ν + 1Γ (ν 2 + 1) × ∑S0′′( tsα)−1 2 − ν 2 Ylm ( n α)e−iS0( p, ts ) , S ( p, td ) ≈ S0 ( p, td ) + r0 ( ts ) [ Δp − v1( ts ) ] . (A.8) (B.6) α The correction v1 diverges at t → ts , therefore the limit has to be taken. Substituting the expansion (90) into the equation of motion (82) we obtain with n α = ∓ ivp (tsα ) κ being a complex ‘unit’ vector along the saddle-point velocity vp (ts ) [16]. It is easy to check that for ν = 0 this formula recovers the SR case (54). v1(t ) ≈ Appendix B. Perturbative calculation of Coulomb corrections with Hamiltonian and Lagrangian forms of action r0 (p, t ) v1(p, t ) ∫t td s ( ) where r0 (p, t ) is the Coulomb-free trajectory (64), v0 = r˙0 and v1 are to be found from (81). Integrating by parts one obtains ∫t s td − ts ∫t td s td r0 v˙1dt = r0 v1 + ts ∫t td s = ± iZ 2 I p ≡ ± iν , (B.8) (B.9) This result means that the momentum distribution remains unaffected, while the total rate changes by the fieldindependent factor exp ( ± 2ν ) ∼ 1. This result contradicts the whole bulk of our knowledge about Coulomb effects in strong field ionization, including the Coulomb-corrected ionization rates proven by comparisons with static-field results. Consider now the same calculation made for the Lagrangian form of the action (71), (72). To shorten equations we introduce W˜ = W + F . Substituting there the same Coulomb-corrected trajectories we obtain (B.1) ⎧1 ⎫ Z ⎨ v(p, t )2 − + Ip⎬ dt ≈ S0 p′, td ⎩2 ⎭ r (p , t ) td ⎧ Z⎫ ⎨ v0 v1 − ⎬ dt + (B.2) ts ⎩ r0 ⎭ v0 v1dt = r0 v1 t → ts S ( p , t d ) ≈ S 0 ( p , t d ) ± iν . ∫ td (B.7) ) where ν is given by (87). Finally the Coulomb-corrected action (in the linear approximation with respect to Z) is Then S ( p, td ) = ( Then Here we calculate Coulomb corrections to the laser-induced action presented in the Hamiltonian (45) and the Lagrangian (71) forms, assuming that the Coulomb perturbation is small, so that the first order of the perturbation theory with respect to the atomic core charge Z is sufficient: Δp = p − p′ ∼ Z , v1 ∼ Z , Δts = ts − ts′ ∼ Z . Z p + A ( ts ) , t → t s. t − t s − 2I 3 2 p ( ) W˜ ( p, td ) ≈ W˜ 0 p′, td + Z dt . (B.3) r0 − rpv1 ttds 32 ∫t − r1vp td s td ts . ⎧ Z⎫ ⎨ v0 v1 + E (t ) r1 + ⎬ dt r0 ⎭ ⎩ (B.10) J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article Integration by parts gives ∫t td td v0 v1dt = v0 r1 s − ts − ∫t ∫t [17] Agostini P and DiMauro L F 2012 Adv. At. Mol. Opt. Phys. 61 117 [18] Rapoport L P, Zon B A and Manakov N L 1978 Theory of Multiphoton Processes in Atoms (Moscow: Atomizdat) (in Russian) [19] Faisal F H M 1987 Theory of Multiphoton Processes (New York: Plenum Press) [20] Fedorov M V 1998 Atomic and Free Electrons in a Strong Light Field (Singapore: World Scientific) [21] Delone N B and Krainov V P 1985 Atoms in Strong Light Fields (Berlin: Springer) Delone N B and Krainov V P 2001 Multiphoton Processes in Atoms (Berlin: Springer) [22] Mulser P and Bauer D 2010 High-Power Laser-Matter Interaction (Berlin: Springer) [23] Gordon W 1926 Z. Phys. 40 117 [24] Wolkow D M 1935 Z. Phys. 94 250 Volkov D M 1937 Zh. Eksp. Teor. Fiz. 7 1286 [25] Berestetskii V B, Lifshitz E M and Pitaevskii L P 1982 Quantum Electrodynamics 2nd edn (Oxford: Pergamon) [26] Reiss H R 1962 J. Math. Phys. 3 59 [27] Reiss H R 1992 Prog. Quantum Electron. 16 1 [28] Reiss H R and Krainov V P 2003 J. Phys. A: Math. Gen. 36 5575 [29] Manakov N L, Frolov M V, Borca B and Starace A F 2003 J. Phys. B: At. Mol. Opt. Phys. 36 R49 [30] Demkov Yu N and Ostrovsky V N 1988 Zero-Range Potentials and Their Applications in Atomic Physics (New York: Plenum) [31] Dammon E K and Tomlinson R G 1963 Appl. Opt. 2 546 [32] Meyerhand R and Haught A F 1963 Phys. Rev. Lett. 11 401 [33] Voronov G S and Delone N B 1965 JETP Lett. 1 42 [34] Voronov G S, Delone G A, Delone N B and Kudrevatova O V 1965 JETP Lett. 1 377 [35] Göppert-Mayer M 1931 Ann. Phys. 9 273 [36] Breit G and Wheeler J A 1934 Phys. Rev. 46 1087 [37] Reiss H R 1962 J. Math. Phys. 3 387 [38] Nikishov A I and Ritus V I 1964 Zh. Eksp. Teor. Fiz. 46 776 Nikishov A I and Ritus V I 1964 Sov. Phys.-JETP 19 529 (Engl. transl.) [39] Bunkin V F and Prokhorov A M 1964 Zh. Eksp. Teor. Fiz. 46 1090 Bunkin V F and Prokhorov A M 1964 Sov. Phys. –JETP 19 739 (Engl. transl.) [40] Landau L D and Lifshits E M 1963 Quantum Mechanics 2nd edn (Moscow: Fizmatlit) Landau L D and Lifshits E M 1965 Quantum Mechanics 2nd edn (Oxford: Pergamon) (Engl. transl.) [41] Demkov Yu N and Drukarev G F 1964 Zh. Eksp. Teor. Fiz. 47 918 Demkov Yu N and Drukarev G F 1965 Sov. Phys. –JETP 20 614 (Engl. transl.) [42] Smirnov B M and Chibisov M I 1965 Zh. Eksp. Teor. Fiz. 49 841 Smirnov B M and Chibisov M I 1965 Sov. Phys. –JETP 22 585 (Engl. transl.) [43] Popov V S, Kuznetsov V P and Perelomov A M 1967 Zh. Eksp. Teor. Fiz. 53 331 Popov V S, Kuznetsov V P and Perelomov A M 1968 Sov. Phys.-JETP 26 222 (Engl. transl.) [44] Popov V S 2005 Phys. At. Nuclei 68 686 [45] Karnakov B M, Mur V D, Popov V S and Popruzhenko S V 2011 JETP Lett. 93 256 [46] Neto H S A and Davidovich L 1984 Phys. Rev. Lett. 53 2238 [47] Becker W, Davidovich L and McIver J K 1994 Phys. Rev. A 49 1131 [48] Agostini P, Fabre F, Mainfray G, Petite G and Rahman N K 1979 Phys. Rev. Lett. 42 1127 td td v˙ 0 r1dt = v0 r1 s ts td E (t ) r1dt . (B.11) s Therefore, in the linear approximation ( ) W˜ ( p, td ) ≈ W˜ 0 p′, td + ∫t td s Z d t ± iν . r (p , t ) (B.12) This result looks physically meaningful. The first term is the Coulomb-free action calculated along the new trajectory while the second is the Coulomb action calculated along the Coulomb-free trajectory. The third term is identical to that of (B.9) and gives a constant (field- and momentum-independent) factor slightly correcting the total ionization probability. The two principally different results (B.9) and (B.12) obtained by calculating corrections to the same Coulomb-free classical action demonstrate the fact that the perturbation theory in the action is essentially sensitive to the choice of independent variables which are different for the actions (45) and (71). References [1] Keldysh L V 1964 Zh. Eksp. Teor. Fiz. 47 1945 Keldysh L V 1965 Sov. Phys. –JETP 20 1307 (Engl. transl.) [2] Faisal F H M 1973 J. Phys. B: At. Mol. Opt. Phys. 6 L89 [3] Reiss H R 1980 Phys. Rev. A 22 1786 [4] Popov V S 2004 Usp. Fiz. Nauk 147 921 Popov V S 2004 Phys. Usp. 47 855 [5] Di Piazza A, Müller C, Hatsagortsyan C Z and Keitel C H 2012 Rev. Mod. Phys. 84 1177 [6] Milošević D B, Paulus G G, Bauer D and Becker W 2006 J. Phys. B: At. Mol. Opt. Phys. 39 R203 [7] Becker W, Grasbon F, Kopold R, Milošević D B, Paulus G G and Walther H 2002 Adv. At. Mol. Opt. Phys. 48 35 [8] Agostini P and DiMauro L F 2004 Rep. Prog. Phys. 67 813 [9] Becker A and Faisal F H M 2005 J. Phys. B: At. Mol. Opt. Phys. 38 R1 [10] Krausz F and Ivanov M 2009 Rev. Mod. Phys. 81 163 [11] Nikishov A I and Ritus V I 1966 Zh. Eksp. Teor. Fiz. 50 255 Nikishov A I and Ritus V I 1966 Sov. Phys. –JETP 23 168 (Engl. transl.) [12] Nikishov A I and Ritus V I 1967 Zh. Eksp. Teor. Fiz. 52 223 Nikishov A I and Ritus V I 1967 Sov. Phys. –JETP 25 145 (Engl. transl.) [13] Perelomov A M, Popov V S and Terentʼev M V 1966 Zh. Eksp. Teor. Fiz. 50 1393 Perelomov A M, Popov V S and Terentʼev M V 1966 Sov. Phys. –JETP 23 924 (Engl. transl.) [14] Perelomov A M, Popov V S and Terentʼev M V 1966 Zh. Eksp. Teor. Fiz. 51 309 Perelomov A M, Popov V S and Terentʼev M V 1967 Sov. Phys. –JETP 24 207 (Engl. transl.) [15] Perelomov A M and Popov V S 1967 Zh. Eksp. Teor. Fiz. 52 514 Perelomov A M and Popov V S 1967 Sov. Phys. –JETP 25 336 (Engl. transl.) [16] Gribakin G F and Kuchiev M Yu 1997 Phys. Rev. A 55 3760 33 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 Review Article [49] Zapesochny I P and Suran V V 1975 Pis. Zh. Tekh. Fiz 1 973 [50] Gaponov A V and Miller M A 1958 Zh. Eksp. Teor. Fiz. 34 242 Gaponov A V and Miller M A 1958 Sov. Phys. JETP 7 168 (Engl. transl.) [51] Kibble T W B 1966 Phys. Rev. A 150 1060 [52] Goreslavsky S P, Narozhny N B and Yakovlev V P 1989 J. Opt. Soc. Am. B 6 1752 [53] Goreslavsky S P and Narozhny N B 1995 J. Nonlinear Opt. Phys. Mat. 4 799 [54] Strickland D and Mourou G 1985 Opt. Commun. 56 219 [55] Chin S, Yergeau F and Lavigne P 1985 J. Opt. Soc. Am. B18 L213 [56] Delone N B and Fedorov M V 1989 Phys.–Usp 158 215 [57] Mainfray G and Manus G 1991 Rep. Prog. Phys. 54 1333 [58] Freeman R R, Bucksbaum P H, Milchberg H, Darack S, Schumacher D and Geusic M E 1987 Phys. Rev. Lett. 59 1092 [59] Bashkansky M, Bucksbaum P H and Schumacher D W 1988 Phys. Rev. Lett. 60 2458 [60] Kuchiev N Yu 1987 Pis. Zh. Eksp. Teor. Fiz. 45 319 Kuchiev N Yu 1987 JETP Lett. 45 404 (Engl. transl.) [61] Corkum B P 1993 Phys. Rev. Lett. 71 1994 [62] Becker W, Lohr A and Kleber M 1994 J. Phys. B: At. Mol. Opt. Phys. 27 L325 [63] Lewenstein M, Balcou Ph, Yu Ivanov M, Huillier A and Corkum P B 1994 Phys. Rev. A 49 2117 [64] Becker W, Long S and McIver J K 1994 Phys. Rev. A 50 1540 [65] Chin S L, Rolland C, Corkum P B and Kelly P 1988 Phys. Rev. Lett. 61 153 [66] Popruzhenko S V, Mur V D, Popov V S and Bauer D 2008 Phys. Rev. Lett. 101 193003 [67] Goreslavski S P, Paulus G G, Popruzhenko S V and Shvetsov-Shilovski N I 2004 Phys. Rev. Lett. 93 233002 [68] Popruzhenko S V, Paulus G G and Bauer D 2008 Phys. Rev. A 77 053409 [69] Chen J and Nam C H 2002 Phys. Rev. A 66 053415 [70] Moshammer R et al 2003 Phys. Rev. Lett. 91 113002 [71] Dimitriou K I, Arbó D G, Yoshida S, Persson E and Burgdörfer J 2004 Phys. Rev. A 70 061401 [72] Blaga C I et al 2009 Nature Phys. 5 335 [73] Quan W et al 2009 Phys. Rev. Lett. 103 093001 [74] Liu C and Hatsagortsyan K Z 2010 Phys. Rev. Lett. 105 113003 [75] Yan T-M, Popruzhenko S V, Vrakking M J J and Bauer D 2010 Phys. Rev. Lett. 105 253002 [76] Kästner A, Saalmann U and Rost J M 2012 Phys. Rev. Lett. 108 033201 [77] Huismans Y et al 2011 Science 331 61 [78] Becker W, Long S and McIver J K 1990 Phys. Rev. A 42 R4416 [79] Yu Kiyan I and Helm H 2003 Phys. Rev. Lett. 90 183001 [80] Reichle R, Helm H and Yu Kiyan I 2003 Phys. Rev. A 68 063404 [81] Bergues B, Ni Y, Helm H and Yu Kiyan I 2005 Phys. Rev. Lett. 95 263002 [82] Bergues B, Ansari Z, Hanstorp D and Kiyan I Yu 2007 Phys. Rev. A 75 063415 [83] Gazibegović-Busuladzĭć A, Milošević D B, Becker W, Bergues B, Hultgren H and Kiyan I Yu 2010 Phys. Rev. Lett. 104 103004 [84] Popov V S 2000 Zh. Eksp. Teor. Fiz. 118 56 Popov V S 2000 JETP 91 48 (Engl. transl.) [85] Popov A M, Tikhonova O V and Volkova E A 1999 J. Phys. B: At. Mol. Opt. Phys. 32 3331 [86] Morales F, Richter M, Patchkovskii S and Smirnova O 2011 Proc. Natl Acad. Sci. USA 108 16906 [87] Klaiber M, Hatsagortsyan Karen Z and Keitel Christoph H 2006 Phys. Rev. A 73 053411 [88] Ammosov M V, Delone N B and Krainov V P 1986 Zh. Eksp. Teor. Fiz. 91 2008 Ammosov M V, Delone N B and Krainov V P 1986 Sov. Phys. –JETP 64 1191 (Engl. transl.) [89] Reiss H R 1996 Phys. Rev. A 54 R1765 [90] Reiss H R 2007 Phys. Rev. A 75 R031404 [91] Morse P M and Feshbach H 1953 Methods of Theoretical Physics (New York: McGraw-Hill) chapter 4.6 [92] Paulus G G et al 1998 Phys. Rev. Lett. 80 484 [93] Goreslavski S P and Popruzhenko S V 1999 J. Phys. B: At. Mol. Opt. Phys. 32 L531 [94] Figueira de Morisson Faria C, Schomerus H and Becker W 2002 Phys. Rev. A 66 043413 [95] Abramowitz M and Stegun I A (ed) 1965 Handbook of Mathematical Functions, Graphs, and Mathematical Tables (New York: Dover) chapter 7.1.16 [96] Goreslavskii S P and Popruzhenko S V 1996 Zh. Eksp. Teor. Fiz. 110 1200 Goreslavskii S P and Popruzhenko S V 1996 JETP 83 661 (Engl. transl.) [97] Mur V D, Popruzhenko S V and Popov V S 2001 Zh. Eksp. Teor. Fiz. 119 893 Mur V D, Popruzhenko S V and Popov V S 2001 JETP 92 777 (Engl. transl.) [98] Goreslavskii S P, Popruzhenko S V, Shvetsov-Shilovski N I and Shcherbachev O V 2005 Zh. Eksp. Teor. Fiz. 127 27 Goreslavskii S P, Popruzhenko S V, Shvetsov-Shilovski N I and Shcherbachev O V 2005 JETP 100 22 (Engl. transl.) [99] Landau L D and Lifshitz E M 1976 Mechanics (Oxford: Pergamon) [100] Goldstein H 1964 Classical Mechanics (Reading, MA: Addison-Wesley) [101] Kopold R et al 2000 Opt. Commun. 178 39 [102] Saliéres P et al 2001 Science 292 902 [103] Feynman R P and Hibbs A R 1965 Quantum Mechanics and Path Integrals (New York: McGraw-Hill) [104] Dykhne A M 1960 Zh. Eksp. Teor. Fiz. 38 570 Dykhne A M 1960 Sov. Phys. –JETP 11 411 (Engl. transl.) Dykhne A M 1961 Zh. Eksp. Teor. Fiz. 41 1324 Dykhne A M 1961 Sov. Phys. –JETP 14 941 (Engl. transl.) [105] Guo L et al 2013 Phys. Rev. Lett. 110 013001 [106] Jain M and Tzoar N 1978 Phys. Rev. A 18 538 [107] Cavalieri P, Ferrante G and Leone C 1980 J. Phys. B: At. Mol. Opt. Phys. 13 4495 [108] Dorr M and Shakeshaft R 1987 Phys. Rev. A 36 421 [109] Shakeshaft R and Potvliege 1987 Phys. Rev. A 36 5478 [110] Basile S, Trombetta F and Ferrante G 1988 Phys. Rev. Lett. 61 2435 [111] Jaron A, Kaminski J Z and Ehlotzky F 1999 Opt. Commun. 163 115 [112] Paulus G G, Grasbon F, Dreischuh A, Walther H, Kopold R and Becker W 2000 Phys. Rev. Lett. 84 3791 [113] Arbó D G, Miraglia J E, Gravielle M S, Schiessl K, Persson E and Burgdörfer J 2008 Phys. Rev. A 77 013401 [114] Kornev A S and Zon B A 2002 Laser Phys. 12 795 [115] Kornev A S and Zon B A 2002 J. Phys. B: At. Mol. Opt. Phys. 35 2451 [116] Bohr A and Mottelson B 1969 Nuclear Structure (New York: Benjamin) [117] Popruzhenko S V and Bauer D 2008 J. Mod. Opt 55 2573 [118] Popruzhenko S V, Mur V D, Popov V S and Bauer D 2009 Zh. Eksp. Teor. Fiz. 135 1092 Popruzhenko S B, Mur V D, Popov V S and Bauer D 2009 JETP 108 947 (Engl. transl.) [119] Baz A I, Zeldovich Ya B and Perelomov A M 1971 Scattering, Reactions and Decays in Nonrelativistic Quantum Mechanics (Moscow: Nauka) 34 J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 204001 [120] [121] [122] [123] [124] [125] [126] [127] [128] [129] [130] [131] [132] [133] [134] [135] [136] [137] [138] [139] [140] [141] [142] [143] [144] [145] [146] [147] [148] [149] [150] [151] [152] [153] [154] [155] [156] [157] Review Article [158] Delone N B, Goreslavsky S P and Krainov V P 1982 J. Phys. B: At. Mol. Phys. 15 L421 Delone N B, Goreslavsky S P and Krainov V P 1982 J. Phys. B: At. Mol. Phys. 16 2369 [159] Bauer D, Milošević D B and Becker W 2005 Phys. Rev. A 72 023415 [160] Faisal F H M 2007 J. Phys. B: At. Mol. Opt. Phys. 40 F145 [161] Reiss H R 2008 Phys. Rev. Lett. 101 043002 [162] Reiss H R 2009 Phys. Rev. A 82 023418 [163] Reiss H R 2013 Phys. Rev. A 87 033421 [164] Bandrauk A D, Fillion-Gourdeau F and Lorin E 2013 J. Phys. B: At. Mol. Opt. Phys. 46 153001 [165] Vanne Y V and Saenz A 2009 Phys. Rev. A 79 023421 [166] Korneev Ph A et al 2012 Phys. Rev. Lett. 108 223601 [167] Korneev Ph A et al 2012 New J. Phys. 14 055019 [168] Kulander K C, Schafer K J and Krause K L 1993 Superintense Laser-Atom Physics (NATO Advanced Studies Institute, Series B: Physics vol 316) ed B Piraux, A L Huillier and K Rzazewski (New York: Plenum) p 95 [169] Paulus G G, Becker W, Nicklich W and Walther H 1994 J. Phys. B: At. Mol. Opt. Phys. 27 L703 [170] Paulus G G, Becker W and Walther H 1995 Phys. Rev. A 52 4043 [171] Shvetsov-Shilovski N I, Dimitrovski D and Madsen L B 2012 Phys. Rev. A 85 023428 [172] Shvetsov-Shilovski N I, Dimitrovski D and Madsen L B 2013 Phys. Rev. A 87 013427 [173] Zhao J and Lein M 2013 Phys. Rev. Lett. 111 043901 [174] McDonald C R, Orlando G, Vampa G and Brabec T 2013 Phys. Rev. Lett. 111 090405 [175] Kheifets A S and Ivanov I A 2010 Phys. Rev. Lett. 105 233002 [176] Dahlström J M, LʼHuillier A and Maquet A 2012 J. Phys. B: At. Mol. Opt. Phys. 45 183001 [177] Kolomeitsev E E and Voskresensky D N 2013 J. Phys. G: Nucl. Part. Phys. 40 113101 [178] Bandrauk A 1994 Molecules in Laser Fields (New York: Dekker) [179] Posthumus J H 2004 Rev. Mod. Phys. 67 623 [180] Stapelfeld H and Seideman T Rev. Mod. Phys. 75 543 [181] Becker W, Chen J, Chen S G and Milošević D B 2007 Phys. Rev. A 76 033403 [182] Busuladźić M and Milošević D B 2010 Phys. Rev. A 82 015401 [183] Milošević D B 2006 Phys. Rev. A 74 063404 [184] Smirnov B M and Firsov O B 1964 Zh. Eksp. Teor. Fiz. 47 232 Smirnov B M and Firsov O B 1965 Sov. Phys. –JETP 20 156 (Engl. transl.) [185] Kristić P S, Milošević D B and Janev R K 1991 Phys. Rev. A 44 3089 [186] Zener C 1934 Proc. R. Soc. A 145 523 [187] Gertsvolf M et al 2008 Phys. Rev. Lett. 101 243001 [188] Rajeev P P, Gertsvolf M, Corkum P B and Rayner D M 2009 Phys. Rev. Lett. 102 083001 [189] Wu C Y et al 2012 Phys. Rev. Lett. 109 043001 [190] Nubbemeyer T, Gorling K, Saenz A, Eichmann U and Sandner W 2008 Phys. Rev. Lett. 101 23300 [191] Zon B A 1999 Zh. Eksp. Teor. Fiz. 116 410 Zon B A 1999 Sov. Phys. –JETP 89 219 (Engl. transl.) [192] Eichmann U, Dörr M, Maeda H, Becker W and Sandner W 2000 Phys. Rev. Lett. 84 3550 [193] Kornev A S, Tulenko E B and Zon B A 2009 Phys. Rev. A 79 063405 [194] Hartree D R 1927 Proc. Camb. Phil. Soc 24 89 Baz A I, Zeldovich Ya B and Perelomov A M 1976 Scattering, Reactions and Decays in Nonrelativistic Quantum Mechanics (Jerusalem: Israel Program for Scientific Translations) (Engl. transl.) Popruzhenko S V 2014 Zh. Eksp. Teor. Fiz. 145 664 Popruzhenko S V 2014 JETP 118 580 (Engl. transl.) Yan T-M and Bauer D 2012 Phys. Rev. A 86 053403 Yan T-M, Popruzhenko S V and Bauer D 2013 Progress in Ultrafast Intense Laser Science (Springer Series in Chemical Physics vol 104) (Berlin: Springer) p 1 Kapitsa P L 1951 Zh. Eksp. Teor. Fiz. 21 588 Kapitsa P L 1951 Usp. Fiz. Nauk 44 7 (Engl. transl.) Potvliege R M 1998 Comput. Phys. Commun. 114 42 Bauer D and Koval P 2006 Comput. Phys. Commun. 174 396 Sorokin A A et al 2007 Phys. Rev. Lett. 99 213002 Smirnova O, Spanner M and Ivanov M 2008 Phys. Rev. A 77 033407 Torlina L and Smirnova O 2012 Phys. Rev. A 86 043408 Bloch C 1957 Nucl. Phys. 4 503 Torlina L et al 2014 arXiv:1402.5620 Torlina L, Morales F, Muller H G and Smirnova S 2014 J. Phys. B: At. Mol. Opt. Phys. 47 204021 Kaushal J and Smirnova O 2013 Phys. Rev. A 88 013421 Torlina L, Kaushal J and Smirnova O 2013 Phys. Rev. A 88 053403 Torlina L, Ivanov M, Walters Z B and Smirnova O 2012 Phys. Rev. A 86 043409 Rapp J and Bauer D 2014 Phys. Rev. A 89 033401 Eckle P et al 2008 Science 322 1525 Eckle P et al 2008 Nature Phys. 4 565 Pfeiffer A N et al 2011 Nature Phys. 8 76 Pfeiffer A N et al 2012 Phys. Rev. Lett. 109 083002 Pfeiffer A N, Cirelli C, Landsmani A S and Keller U 2013 Chem. Phys. 414 84 Manakov N L and Fainshtein A G 1980 Zh. Eksp. Teor. Fiz. 79 751 Manakov N L and Fainshtein A G 1980 Sov. Phys. JETP 52 382 Manakov N L, Ovsiannikov V D and Rapoport L P 1986 Phys. Rep. 141 320 Berson I 1981 Phys. Lett. A 84 364 Bersons I Ya 1982 Zh. Eksp. Teor. Fiz. 83 1276 Bersons I Ya 1982 Sov. Phys. JETP 56 731 (Engl. transl.) Floquet G 1883 Ann. Ecole Norm. Sup. 12 47 Shirley J H 1965 Phys. Rev. B 138 979 Zelʼdovich Ya B 1966 Zh. Eksp. Teor. Fiz. 51 1492 Zelʼdovich Ya B 1967 Sov. Phys. JETP 24 1006 (Engl. transl.) Ritus V I 1966 Zh. Eksp. Teor. Fiz. 51 1544 Ritus V I 1967 Sov. Phys. JETP 25 1027 (Engl. transl.) Manakov N L and Rapoport L P 1975 Zh. Eksp. Teor. Fiz. 69 842 Manakov N L and Rapoport L P 1975 Sov. Phys. –JETP 42 430 (Engl. transl.) Berson I J 1975 J. Phys. B: At. Mol. Opt. Phys. 8 3078 Frolov M V, Manakov N L, Pronin E A and Starace A F 2003 Phys. Rev. Lett. 91 053003 Borca B, Frolov M V, Manakov N L and Starace A F 2002 Phys. Rev. Lett. 88 193001 Frolov M V, Manakov N L and Starace A F 2008 Phys. Rev. Lett. 100 173001 Frolov M V, Manakov N L, Sarantseva T S and Starace A F 2011 Phys. Rev. A 83 043416 Frolov M V et al 2012 Phys. Rev. Lett. 108 213002 Borzunov S V, Frolov M V, Ivanov M Yu, Manakov N L, Marmo S S and Starace A F 2013 Phys. Rev. A 88 033410 Kramers H A 1923 Phil. Mag. 46 836 35
© Copyright 2024