Angular momentum dependence in multiphoton ionization and attosecond time delays

Jakub Benda [email protected] Institute of Theoretical Physics, Faculty of Mathematics and Physics, Charles University, V Holešovičkách 2, 180 00, Prague 8, Czech Republic    Zdeněk Mašín Institute of Theoretical Physics, Faculty of Mathematics and Physics, Charles University, V Holešovičkách 2, 180 00, Prague 8, Czech Republic    Sreelakshmi Palakkal Universite Claude Bernard Lyon 1, CNRS, Institut Lumière Matière, UMR5306, F-69100, Villeurbanne, France    Franck Lépine Universite Claude Bernard Lyon 1, CNRS, Institut Lumière Matière, UMR5306, F-69100, Villeurbanne, France    Saikat Nandi Universite Claude Bernard Lyon 1, CNRS, Institut Lumière Matière, UMR5306, F-69100, Villeurbanne, France    Vincent Loriot Universite Claude Bernard Lyon 1, CNRS, Institut Lumière Matière, UMR5306, F-69100, Villeurbanne, France
(July 19, 2024)
Abstract

Attosecond ionization time-delays at photoelectron energies above typically 10 eV are usually interpreted using the so called asymptotic approximation as a sum of the atomic or molecular delays with a universal laser-induced contribution. Here, we employ a two-harmonic RABITT (Reconstruction of Attosecond Beating By Interference of Two-photon Transitions) configuration to isolate the multiphoton pathways and measure the ionization time delays as a function of the dressing field intensity. We show that the validity of the asymptotic theory can be extended to the threshold or to higher-order contributions by rigorously treating the angular-momentum dependence of the continuum-continuum transitions into universal and easily computable partial-wave-specific correction factors. Our asymptotic treatment is also valid for higher-order interfering amplitudes while significantly simplifying their evaluation and providing a transparent physical interpretation. The validity of the method for atomic and molecular targets in the vicinity of resonances, ionization thresholds, and for both the emission-integrated and angularly resolved signal is confirmed by comparison to ab initio calculations over a wide energy range.

preprint: APS/123-QED

I Introduction

An electron ionized by a single photon from an atom or a molecule scatters in the parent potential before reaching asymptotic distances. This affects the phase of the outgoing photoelectron, and can be semiclassically understood as a time delay in the scattering process as initially proposed by Wigner [1, 2]. Ionization time delay has recently attracted a substantial amount of interests from the attosecond community and is now investigated in molecules and quantum systems of increasing sizes [3, 4, 5].

The RABITT protocol is one of the common methods to characterize an attosecond pulse train [6] and to measure the ionization time delay with spectral resolution [7]. In the latter case it produces an intuitive signal with a close correspondence to the Wigner ionization time delay. However, the measurement needs a dressing field, that affects the observation. Different strategies have emerged to isolate the Wigner time delay from RABITT measurements such as performing the full two-photon calculation [8]. Up to date, though, the most widely used solution is based on a ‘universal’ dressing field delay through the so-called ‘continuum-continuum’ delay, τccsubscript𝜏𝑐𝑐\tau_{cc}italic_τ start_POSTSUBSCRIPT italic_c italic_c end_POSTSUBSCRIPT, introduced by Dahlström et al. [9]. The model has further been extended to molecules [10, 11, 12]. In addition to τccsubscript𝜏𝑐𝑐\tau_{cc}italic_τ start_POSTSUBSCRIPT italic_c italic_c end_POSTSUBSCRIPT, other correction terms have also to be considered, such as the ‘coupling delay’ τcouplsubscript𝜏coupl\tau_{\text{coupl}}italic_τ start_POSTSUBSCRIPT coupl end_POSTSUBSCRIPT or the ‘dipole-laser coupling delay’ τdLCsubscript𝜏𝑑𝐿𝐶\tau_{dLC}italic_τ start_POSTSUBSCRIPT italic_d italic_L italic_C end_POSTSUBSCRIPT [13, 14]. A common denominator of all these additional delays is that they are well defined only in the high kinetic energy limit. At low energies the problem becomes more complex due to electron correlation, which effectively intertwines the effects responsible for all these asymptotic contributions.

In this paper, we focus on the simplest RABITT configuration conceivable – one that combines a pair of adjacent odd harmonics of the fundamental field. The advantage of such arrangement is two-fold. First, by avoiding the complex cross-talk of different harmonics we make it possible to measure the outcome of the higher-order processes that are otherwise responsible for hard-to-analyze contamination of the pristine two-photon RABITT signal. Second, as discussed in the text, it allows reading off of the oscillation from the interference bands forming elsewhere than just in between the two harmonics. This can be advantageous for example in molecules, where closely spaced electronic states result in spectrograms with multiple overlapping sidebands arising from different interference pathways.

On the theory side, we model both the standard and the high-order RABITT delays using the time-independent molecular above-threshold multiphoton R-matrix method [15] at the leading-order level of the perturbation theory, as well as using fully time-dependent simulations using ‘R-matrix with time-dependence’ (RMT) [16]. Also—and crucially—we propose a simple and accurate analytical solution to treat the dynamics of the photoionized system driven by the infrared (IR) field. We include the interfering pathways of the RABITT method involving an arbitrary number of dressing field photons. Our approach is easy to implement, numerically cheap and provides accurate results even at low photoelectron energy and around resonances. Our method avoids the integration in the complex plane used e.g. in [17] and provides a computationally cheap and accurate alternative to more sophisticated multiphoton ab initio calculations [13] and fully time-dependent simulations [18, 19, 20]. A key property of the present approach is including the angular momentum dependency of the ionization amplitudes affected by the dressing field. Such a property has already been experimentally observed [21, 22] and discussed theoretically in numerical calculations [23, 24].

The article is organized as follows. Section II presents the RABITT interference pathways involving different orders of the perturbation theory. The Wigner time delay is derived and the asymptotic approximation of two-photon transitions is recalled. Our implementation of the partial-wave-resolved continuum-continuum contribution is introduced and generalized to arbitrary orders of interfering pathways for atoms and molecules. Section III gives numerical predictions for angle-integrated and angle-resolved RABITT delays. The results are illustrated with argon in detail and are compared with original measurements. The practical method to introduce IR transitions to partial wave-resolved extreme ultraviolet (XUV) photoionization amplitudes is provided in Section IV.

Additional application of our method are proposed in the supplementary material, such as the effective τccsubscript𝜏𝑐𝑐\tau_{cc}italic_τ start_POSTSUBSCRIPT italic_c italic_c end_POSTSUBSCRIPT extracted for noble gases, results for ionization from a molecular orbital of CO2 that features a shape resonance, and the time delays in LiH (polar molecule).

II Method

II.1 Interference pathways in RABITT

In a RABITT experiment, an attosecond pulse train is generated in a centrosymetric media that corresponds, in the spectral domain, to a comb of odd harmonics ΩΩ\Omegaroman_Ω reaching the XUV range, from a fundamental frequency ω𝜔\omegaitalic_ω (i.e. Ω=(2n+1)ωΩ2𝑛1𝜔\Omega=(2n+1)\omegaroman_Ω = ( 2 italic_n + 1 ) italic_ω, for integer n𝑛nitalic_n). Each harmonic above the ionization potential (Ip) can ionise the target and a photoelectron can be emitted into the continuum. The photoelectron spectrum is hence composed of several well localised peaks (main bands (MB)) where the ionization phase information is experimentally inaccessible. To be sensitive to such phases, the RABITT protocol proposes to dress the harmonic comb with a weak light pulse with a central frequency of ω𝜔\omegaitalic_ω interferometrically stabilised with the XUV pulse. The photoelectrons are hence redistributed into sidebands (SB) with an energy shift of ±ωplus-or-minus𝜔\pm\omega± italic_ω. As a function of the dressing pulse intensity, several ω𝜔\omegaitalic_ω-photons can be absorbed or emitted as shown in Fig. 1.

Refer to caption
Figure 1: Pathways considering only two consecutive harmonics of the comb (Ω13subscriptΩ13\Omega_{13}roman_Ω start_POSTSUBSCRIPT 13 end_POSTSUBSCRIPT and Ω15subscriptΩ15\Omega_{15}roman_Ω start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT) and a dressing field limited to (a) one, (b) two and (c) three photons absorbed or emitted reaching the sideband (SB), the ‘mainbands’ (MB) or the upper outer sideband (OSB1). Ip is the ionization potential.

To unambiguously identify the interference pathways, a set of only two following harmonics is considered all along the manuscript (Ω13subscriptΩ13\Omega_{13}roman_Ω start_POSTSUBSCRIPT 13 end_POSTSUBSCRIPT and Ω15subscriptΩ15\Omega_{15}roman_Ω start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT, i.e. separated by 2ω𝜔\omegaitalic_ω) theoretically and experimentally. The addition of a single ω𝜔\omegaitalic_ω-photon opens two quantum paths leading to the same final kinetic energy at SB14 that interfere (see Fig. 1(a)). By scanning the delay τ𝜏\tauitalic_τ between the pulses, the SB intensity oscillates as

I(τ)𝐼𝜏\displaystyle I(\tau)italic_I ( italic_τ ) =\displaystyle== 𝒜+cos(2ωτϕ2ω),𝒜2𝜔𝜏subscriptitalic-ϕ2𝜔\displaystyle\mathcal{A}+\mathcal{B}\cos\left(2\omega\tau-\phi_{2\omega}\right),caligraphic_A + caligraphic_B roman_cos ( 2 italic_ω italic_τ - italic_ϕ start_POSTSUBSCRIPT 2 italic_ω end_POSTSUBSCRIPT ) , (1)
=\displaystyle== 𝒜+cos(2ω(ττR)),𝒜2𝜔𝜏subscript𝜏𝑅\displaystyle\mathcal{A}+\mathcal{B}\cos\left(2\omega(\tau-\tau_{R})\right),caligraphic_A + caligraphic_B roman_cos ( 2 italic_ω ( italic_τ - italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ) ) ,

with 𝒜𝒜\mathcal{A}caligraphic_A being the baseline, \mathcal{B}caligraphic_B the amplitude and ϕ2ωsubscriptitalic-ϕ2𝜔\phi_{2\omega}italic_ϕ start_POSTSUBSCRIPT 2 italic_ω end_POSTSUBSCRIPT the phase of the oscillation at 2ω𝜔\omegaitalic_ω, and τR=ϕ2ω/(2ω)subscript𝜏𝑅subscriptitalic-ϕ2𝜔2𝜔\tau_{R}=\phi_{2\omega}/(2\omega)italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT = italic_ϕ start_POSTSUBSCRIPT 2 italic_ω end_POSTSUBSCRIPT / ( 2 italic_ω ) the RABITT delay. All of these quantities can be expressed using two-photon or higher-order dipole matrix elements between the initial neutral state and the final photoelectron state [25]. The information on the ionization time delay is stored within this RABITT delay, which is also affected by the dressing field.

Since the dressing pulse is weak, it does not ionize the target but only redistributes the photoelectrons. The increase in the population of a SB implies the depopulation of its surrounding MBs. Since the amplitude of the SB depends on τ𝜏\tauitalic_τ, the MBs reflect the complementary behavior i.e. they oscillate following Eq. 1 but in phase opposition. According to the perturbation theory, the MB oscillations are described by another interfering path. For instance, as shown in Fig. 1(b) for MB15 where the path (Ω15subscriptΩ15\Omega_{15}roman_Ω start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT) interferes with (Ω13+2ω(\Omega_{13}+2\omega( roman_Ω start_POSTSUBSCRIPT 13 end_POSTSUBSCRIPT + 2 italic_ω). The SB and MBs oscillation amplitudes are both linear with the dressing field intensity despite they involve different orders of the perturbation theory. This is experimentally illustrated on the left panels of Fig. 2 performed at low dressing field intensity with two isolated harmonics (see Section II of the Supplementary Material for experimental details). MB15 can be reached also by a beyond-leading-order transition (Ω15+ωω(\Omega_{15}+\omega-\omega( roman_Ω start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT + italic_ω - italic_ω), which combines with the plain one-photon transition driven by Ω15subscriptΩ15\Omega_{15}roman_Ω start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT. However, for weak dressing fields, such pathways are suppressed with respect to the lower-order transition and can be neglected. Generalization to strong-field processes (0.5greater-than-or-equivalent-toabsent0.5\gtrsim 0.5≳ 0.5 TW/cm2) is possible within the time-dependent approach or the strong-field approximation [26].

Refer to caption
Figure 2: (a, b) RABITT spectrograms obtained in argon by using a set of two isolated harmonics (Ω13subscriptΩ13\Omega_{13}roman_Ω start_POSTSUBSCRIPT 13 end_POSTSUBSCRIPT and Ω15subscriptΩ15\Omega_{15}roman_Ω start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT) for low (a) and high (b) dressing field intensity. (c,d) the corresponding Fourier transform showing a clear oscillation at 2ω𝜔\omegaitalic_ω (i.e. at 0.75 fs-1). The 2ω𝜔\omegaitalic_ω oscillation amplitude (e,f) and phase (g,h) in blue. In (g,h), the phase of the MB and OSB are also presented in orange with a π𝜋\piitalic_π-shift to be compared to the SB. Experimental details are in the supplementary material.

At low dressing field intensity, the first outer sideband (OSB1) appears through the pathway (Ω15+ωsubscriptΩ15𝜔\Omega_{15}+\omegaroman_Ω start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT + italic_ω) but it does not oscillate with τ𝜏\tauitalic_τ because of the lack of an interfering pathway (see Fig. 1(a) and Fig. 2(left)). At higher dressing field intensity the (Ω13+3ω)subscriptΩ133𝜔(\Omega_{13}+3\omega)( roman_Ω start_POSTSUBSCRIPT 13 end_POSTSUBSCRIPT + 3 italic_ω ) transition becomes non-negligible and interferes with (Ω15+ωsubscriptΩ15𝜔\Omega_{15}+\omegaroman_Ω start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT + italic_ω) as shown in see Fig. 1(c) and Fig. 2(right). OSB1 also oscillates with τ𝜏\tauitalic_τ following Eq. (1) in phase to the MB and in phase opposition with the SB. The SB-MB-OSB coupled oscillations are demonstrated under the soft photon approximation in the Supplementary Material and using Floquet formalism in [27]. Let’s notice that even at high dressing field intensity no 4ω𝜔\omegaitalic_ω oscillation (i.e. 1.5 fs-1 for an 800 nm fundamental field) is observed since no harmonics are separated by 4ω𝜔\omegaitalic_ω.

All such inteference pathways depend on the XUV ionization phases that carry the ionization time delay. This introduces a redundancy in the information that can be exploited to extract the value with high accuracy. However, the dressing field affects both the yield and angular distribution of the oscillation depending on the considered band. In the following, after expressing the Wigner time delay, a general expression is derived that accurately treats the influence of the dressing pulse on the SB, MB and OSBn𝑛nitalic_n.

II.2 The Wigner time delay

All along the analytical derivation, Hartree atomic units (=e=me=4πε0=1Planck-constant-over-2-pi𝑒subscript𝑚𝑒4𝜋subscript𝜀01\hbar=e=m_{e}=4\pi\varepsilon_{0}=1roman_ℏ = italic_e = italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 4 italic_π italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1) are used unless another units are explicitly indicated. The formalism is based on the laser-assisted photoionization perturbation theory of Dahlström et al. [9] with monochromatic fields. Here, only the necessary steps are recalled following the conventions defined in [14].

The one-photon ionization amplitude

Tfi(1)=2πiΨf𝒌()|𝜺^XUV𝑫|Ψi=2πidfi(1)(𝒌)superscriptsubscript𝑇fi12𝜋𝑖quantum-operator-productsuperscriptsubscriptΨf𝒌subscript^𝜺XUV𝑫subscriptΨi2𝜋𝑖superscriptsubscript𝑑fi1𝒌\displaystyle T_{\textrm{fi}}^{(1)}=2\pi i\langle\Psi_{\textrm{f}\bm{k}}^{(-)}% |\hat{\bm{\varepsilon}}_{\text{XUV}}\cdot\bm{D}|\Psi_{\textrm{i}}\rangle=2\pi id% _{\textrm{fi}}^{(1)}(\bm{k})italic_T start_POSTSUBSCRIPT fi end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT = 2 italic_π italic_i ⟨ roman_Ψ start_POSTSUBSCRIPT f bold_italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( - ) end_POSTSUPERSCRIPT | over^ start_ARG bold_italic_ε end_ARG start_POSTSUBSCRIPT XUV end_POSTSUBSCRIPT ⋅ bold_italic_D | roman_Ψ start_POSTSUBSCRIPT i end_POSTSUBSCRIPT ⟩ = 2 italic_π italic_i italic_d start_POSTSUBSCRIPT fi end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( bold_italic_k ) (2)

from the initial bound state ΨisubscriptΨi\Psi_{\textrm{i}}roman_Ψ start_POSTSUBSCRIPT i end_POSTSUBSCRIPT to a final continuum state Ψf𝒌()superscriptsubscriptΨf𝒌\Psi_{\textrm{f}\bm{k}}^{(-)}roman_Ψ start_POSTSUBSCRIPT f bold_italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( - ) end_POSTSUPERSCRIPT for field with the polarization vector 𝜺^XUVsubscript^𝜺XUV\hat{\bm{\varepsilon}}_{\text{XUV}}over^ start_ARG bold_italic_ε end_ARG start_POSTSUBSCRIPT XUV end_POSTSUBSCRIPT is expressed in terms of the one-photon ionization dipole matrix element

dfi(1)(𝒌)=lmdflm,i(1)(k)Ylm(𝒌^),superscriptsubscript𝑑fi1𝒌subscript𝑙𝑚superscriptsubscript𝑑f𝑙𝑚i1𝑘superscriptsubscript𝑌𝑙𝑚^𝒌\displaystyle d_{\textrm{fi}}^{(1)}(\bm{k})=\sum_{lm}d_{\textrm{f}lm,\textrm{i% }}^{(1)}(k)Y_{l}^{m}(\hat{\bm{k}})\,,italic_d start_POSTSUBSCRIPT fi end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( bold_italic_k ) = ∑ start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT italic_d start_POSTSUBSCRIPT f italic_l italic_m , i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_k ) italic_Y start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT ( over^ start_ARG bold_italic_k end_ARG ) , (3)

where 𝑫𝑫\bm{D}bold_italic_D is the operator of the total electronic dipole moment, 𝒌𝒌\bm{k}bold_italic_k is the photoelectron momentum, and Ylmsuperscriptsubscript𝑌𝑙𝑚Y_{l}^{m}italic_Y start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT are the spherical harmonics. After a single ionization by an XUV radiation, and at large distances of the ejected photoelectron from the molecule, the state of the system can be asymptotically written as the partial wave channel expansion [15]

Ψi+Ω(+)(𝒓)1rιλμaιλμ,i(1)Hλ+(Zκι,κιr)Yλμ(𝒓)Φι,superscriptsubscriptΨiΩ𝒓1𝑟subscript𝜄𝜆𝜇superscriptsubscript𝑎𝜄𝜆𝜇i1superscriptsubscript𝐻𝜆𝑍subscript𝜅𝜄subscript𝜅𝜄𝑟superscriptsubscript𝑌𝜆𝜇𝒓subscriptΦ𝜄\Psi_{\textrm{i}+\Omega}^{(+)}(\bm{r})\rightarrow\frac{1}{r}\sum_{\iota\lambda% \mu}a_{\iota\lambda\mu,\text{i}}^{(1)}H_{\lambda}^{+}(-\tfrac{Z}{\kappa_{\iota% }},\kappa_{\iota}r)Y_{\lambda}^{\mu}(\bm{r})\Phi_{\iota}\,,roman_Ψ start_POSTSUBSCRIPT i + roman_Ω end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( + ) end_POSTSUPERSCRIPT ( bold_italic_r ) → divide start_ARG 1 end_ARG start_ARG italic_r end_ARG ∑ start_POSTSUBSCRIPT italic_ι italic_λ italic_μ end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_ι italic_λ italic_μ , i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT italic_H start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( - divide start_ARG italic_Z end_ARG start_ARG italic_κ start_POSTSUBSCRIPT italic_ι end_POSTSUBSCRIPT end_ARG , italic_κ start_POSTSUBSCRIPT italic_ι end_POSTSUBSCRIPT italic_r ) italic_Y start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT ( bold_italic_r ) roman_Φ start_POSTSUBSCRIPT italic_ι end_POSTSUBSCRIPT , (4)

with Hl±(η,ρ)superscriptsubscript𝐻𝑙plus-or-minus𝜂𝜌H_{l}^{\pm}(\eta,\rho)italic_H start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ± end_POSTSUPERSCRIPT ( italic_η , italic_ρ ) a travelling Coulomb wave function, denoted in this paper also as Hl±(r)superscriptsubscript𝐻𝑙plus-or-minus𝑟H_{l}^{\pm}(r)italic_H start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ± end_POSTSUPERSCRIPT ( italic_r ) for brevity. Asymptotically, its behavior is that of a Coulomb-corrected plane wave,

Hλ+exp[i(κιr+Zκιln2κιrπλ2+σλ(κι))].superscriptsubscript𝐻𝜆𝑖subscript𝜅𝜄𝑟𝑍subscript𝜅𝜄2subscript𝜅𝜄𝑟𝜋𝜆2subscript𝜎𝜆subscript𝜅𝜄\displaystyle H_{\lambda}^{+}\rightarrow\exp\left[i\left(\kappa_{\iota}r+\frac% {Z}{\kappa_{\iota}}\ln 2\kappa_{\iota}r-\frac{\pi\lambda}{2}+\sigma_{\lambda}(% \kappa_{\iota})\right)\right]\,.italic_H start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT → roman_exp [ italic_i ( italic_κ start_POSTSUBSCRIPT italic_ι end_POSTSUBSCRIPT italic_r + divide start_ARG italic_Z end_ARG start_ARG italic_κ start_POSTSUBSCRIPT italic_ι end_POSTSUBSCRIPT end_ARG roman_ln 2 italic_κ start_POSTSUBSCRIPT italic_ι end_POSTSUBSCRIPT italic_r - divide start_ARG italic_π italic_λ end_ARG start_ARG 2 end_ARG + italic_σ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_κ start_POSTSUBSCRIPT italic_ι end_POSTSUBSCRIPT ) ) ] . (5)

The symbol ΦιsubscriptΦ𝜄\Phi_{\iota}roman_Φ start_POSTSUBSCRIPT italic_ι end_POSTSUBSCRIPT in Eq. (4) stands for the electronic state of the residual ion coupled to the partial wave λμ𝜆𝜇\lambda\muitalic_λ italic_μ, and σλ(κ)=argΓ(λ+1Zi/κ)subscript𝜎𝜆𝜅Γ𝜆1𝑍𝑖𝜅\sigma_{\lambda}(\kappa)=\arg\Gamma(\lambda+1-Zi/\kappa)italic_σ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_κ ) = roman_arg roman_Γ ( italic_λ + 1 - italic_Z italic_i / italic_κ ) is the Coulomb phase for a center with the residual charge Z𝑍Zitalic_Z. In the equations only the coordinates of the photoelectron are explicitly labeled. The expansion coefficient aιλμ,i(1)superscriptsubscript𝑎𝜄𝜆𝜇i1a_{\iota\lambda\mu,\text{i}}^{(1)}italic_a start_POSTSUBSCRIPT italic_ι italic_λ italic_μ , i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT in Eq. (4) is proportional to a specific partial wave component of the one-photon transition dipole [13],

aιλμ,i(1)superscriptsubscript𝑎𝜄𝜆𝜇i1\displaystyle a_{\iota\lambda\mu,\text{i}}^{(1)}italic_a start_POSTSUBSCRIPT italic_ι italic_λ italic_μ , i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT =2πiλeiσλYλμ(𝜿^ι)Ψι𝜿ι()|VXUV|Ψid2𝜿^ιabsent2𝜋superscript𝑖𝜆superscript𝑒𝑖subscript𝜎𝜆superscriptsubscript𝑌𝜆𝜇subscript^𝜿𝜄quantum-operator-productsuperscriptsubscriptΨ𝜄subscript𝜿𝜄subscript𝑉XUVsubscriptΨisuperscriptd2subscript^𝜿𝜄\displaystyle=-2\pi\int i^{\lambda}e^{-i\sigma_{\lambda}}Y_{\lambda}^{\mu*}(% \hat{\bm{\kappa}}_{\iota})\langle\Psi_{\iota\bm{\kappa}_{\iota}}^{(-)}|V_{% \text{XUV}}|\Psi_{\textrm{i}}\rangle\mathrm{d}^{2}\hat{\bm{\kappa}}_{\iota}= - 2 italic_π ∫ italic_i start_POSTSUPERSCRIPT italic_λ end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_σ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_Y start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_μ ∗ end_POSTSUPERSCRIPT ( over^ start_ARG bold_italic_κ end_ARG start_POSTSUBSCRIPT italic_ι end_POSTSUBSCRIPT ) ⟨ roman_Ψ start_POSTSUBSCRIPT italic_ι bold_italic_κ start_POSTSUBSCRIPT italic_ι end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( - ) end_POSTSUPERSCRIPT | italic_V start_POSTSUBSCRIPT XUV end_POSTSUBSCRIPT | roman_Ψ start_POSTSUBSCRIPT i end_POSTSUBSCRIPT ⟩ roman_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over^ start_ARG bold_italic_κ end_ARG start_POSTSUBSCRIPT italic_ι end_POSTSUBSCRIPT
=2πκιiλeiσλdιλμ,i(1)(κι).absent2𝜋subscript𝜅𝜄superscript𝑖𝜆superscript𝑒𝑖subscript𝜎𝜆superscriptsubscript𝑑𝜄𝜆𝜇i1subscript𝜅𝜄\displaystyle=-\sqrt{\frac{2\pi}{\kappa_{\iota}}}i^{\lambda}e^{-i\sigma_{% \lambda}}d_{\iota\lambda\mu,\textrm{i}}^{(1)}(\kappa_{\iota})\,.= - square-root start_ARG divide start_ARG 2 italic_π end_ARG start_ARG italic_κ start_POSTSUBSCRIPT italic_ι end_POSTSUBSCRIPT end_ARG end_ARG italic_i start_POSTSUPERSCRIPT italic_λ end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_σ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_ι italic_λ italic_μ , i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_κ start_POSTSUBSCRIPT italic_ι end_POSTSUBSCRIPT ) . (6)

Here VXUV=(𝑫ion+𝒓)𝜺^XUVsubscript𝑉XUVsubscript𝑫ion𝒓subscript^𝜺XUVV_{\text{XUV}}=(\bm{D}_{\text{ion}}+\bm{r})\cdot\hat{\bm{\varepsilon}}_{\text{% XUV}}italic_V start_POSTSUBSCRIPT XUV end_POSTSUBSCRIPT = ( bold_italic_D start_POSTSUBSCRIPT ion end_POSTSUBSCRIPT + bold_italic_r ) ⋅ over^ start_ARG bold_italic_ε end_ARG start_POSTSUBSCRIPT XUV end_POSTSUBSCRIPT stands for the projection of the total electronic dipole operator along the polarization 𝜺^XUVsubscript^𝜺XUV\hat{\bm{\varepsilon}}_{\text{XUV}}over^ start_ARG bold_italic_ε end_ARG start_POSTSUBSCRIPT XUV end_POSTSUBSCRIPT of the XUV field.

Following the Wigner-theory [1], the ionization time delay of a given partial wave lm𝑙𝑚lmitalic_l italic_m for an electron energy Ek=k2/2=ΩIpsubscript𝐸𝑘superscript𝑘22ΩIpE_{k}=k^{2}/2=\Omega-\text{Ip}italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 = roman_Ω - Ip corresponds to the derivative of the phase of an XUV transition dipole with respect to energy,

τW,lm(Ek)=Ωarg(dflm,i(1)(Ω)).subscript𝜏𝑊𝑙𝑚subscript𝐸𝑘Ωsubscriptsuperscript𝑑1f𝑙𝑚iΩ\tau_{W,lm}(E_{k})=\frac{\partial}{\partial\Omega}\arg\left(d^{(1)}_{\textrm{f% }lm,\textrm{i}}(\Omega)\right).italic_τ start_POSTSUBSCRIPT italic_W , italic_l italic_m end_POSTSUBSCRIPT ( italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) = divide start_ARG ∂ end_ARG start_ARG ∂ roman_Ω end_ARG roman_arg ( italic_d start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT f italic_l italic_m , i end_POSTSUBSCRIPT ( roman_Ω ) ) . (7)

In most cases, the final state corresponds to a partial-wave mixture and the ionization time delay depends on the emission angle. Within the RABITT framework the differentiation step becomes the energy gap between two adjacent odd harmonics (Ω2ωΩ2𝜔\partial\Omega\rightarrow 2\omega∂ roman_Ω → 2 italic_ω). For oriented emission of a photoelectron at energy Ek=k2/2subscript𝐸𝑘superscript𝑘22E_{k}=k^{2}/2italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2, one then recovers the ‘atomic’ one-photon delay

τ1(𝒌)=12ωarg(dfi(1)(κ+𝒌^)dfi(1)(κ𝒌^)),subscript𝜏1𝒌12𝜔superscriptsubscript𝑑fi1subscript𝜅^𝒌superscriptsubscript𝑑fi1subscript𝜅^𝒌\tau_{1}(\bm{k})=\frac{1}{2\omega}\arg\left(d_{\textrm{fi}}^{(1)*}(\kappa_{+}% \hat{\bm{k}})d_{\textrm{fi}}^{(1)}(\kappa_{-}\hat{\bm{k}})\right)\,,italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( bold_italic_k ) = divide start_ARG 1 end_ARG start_ARG 2 italic_ω end_ARG roman_arg ( italic_d start_POSTSUBSCRIPT fi end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) ∗ end_POSTSUPERSCRIPT ( italic_κ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT over^ start_ARG bold_italic_k end_ARG ) italic_d start_POSTSUBSCRIPT fi end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_κ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT over^ start_ARG bold_italic_k end_ARG ) ) , (8)

where 𝒌^^𝒌\hat{\bm{k}}over^ start_ARG bold_italic_k end_ARG is a unit vector, and κ±=2(Ekω)subscript𝜅plus-or-minus2minus-or-plussubscript𝐸𝑘𝜔\kappa_{\pm}=\sqrt{2(E_{k}\mp\omega)}italic_κ start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT = square-root start_ARG 2 ( italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ∓ italic_ω ) end_ARG is the intermediate momentum in the (Ω<+ω)subscriptΩ𝜔(\Omega_{<}+\omega)( roman_Ω start_POSTSUBSCRIPT < end_POSTSUBSCRIPT + italic_ω ) and (Ω>ω)subscriptΩ𝜔(\Omega_{>}-\omega)( roman_Ω start_POSTSUBSCRIPT > end_POSTSUBSCRIPT - italic_ω ) pathways, respectively. This quantity represents a finite difference approximation to an effective Wigner delay and corresponds to the expected value to extract from the RABITT protocol. All along the manuscript τ1subscript𝜏1\tau_{1}italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is taken as the reference for oriented emission. For emission-integrated and orientation-averaged RABITT signal we use instead the reference formula

τ1(k)=12ωarglmlmablm|n^an^b|lmsuperscriptsubscript𝜏1𝑘12𝜔subscriptsuperscript𝑙superscript𝑚𝑙𝑚𝑎𝑏quantum-operator-productsuperscript𝑙superscript𝑚subscript^𝑛𝑎subscript^𝑛𝑏𝑙𝑚\displaystyle\tau_{1}^{\prime}(k)=\frac{1}{2\omega}\arg\sum_{l^{\prime}m^{% \prime}lmab}\langle l^{\prime}m^{\prime}|\hat{n}_{a}\hat{n}_{b}|lm\rangleitalic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_k ) = divide start_ARG 1 end_ARG start_ARG 2 italic_ω end_ARG roman_arg ∑ start_POSTSUBSCRIPT italic_l start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_m start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_l italic_m italic_a italic_b end_POSTSUBSCRIPT ⟨ italic_l start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_m start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT | italic_l italic_m ⟩
×ε^aε^bdflm,i(1)(κ+)dflm,i(1)(κ)d2𝜺^,\displaystyle\times\int\hat{\varepsilon}_{a}\hat{\varepsilon}_{b}d_{\textrm{f}% l^{\prime}m^{\prime},\textrm{i}}^{(1)*}(\kappa_{+})d_{\textrm{f}lm,\textrm{i}}% ^{(1)}(\kappa_{-})\mathrm{d}^{2}\hat{\bm{\varepsilon}}\,,× ∫ over^ start_ARG italic_ε end_ARG start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT over^ start_ARG italic_ε end_ARG start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT italic_d start_POSTSUBSCRIPT f italic_l start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_m start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) ∗ end_POSTSUPERSCRIPT ( italic_κ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ) italic_d start_POSTSUBSCRIPT f italic_l italic_m , i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_κ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT ) roman_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over^ start_ARG bold_italic_ε end_ARG , (9)

which corresponds to the ‘molecular’ delay [10]; 𝜺^^𝜺\hat{\bm{\varepsilon}}over^ start_ARG bold_italic_ε end_ARG denotes the common linear polarization vector of XUV and IR.

II.3 The asymptotic theory

In this section, we review the widely used asymptotic theory that separates the ionization time delay from the measurement-induced time shift due to the dressing field. The original formulation has been developed for atoms by Dahlström et al. [9], extended to molecules by Baykusheva and Wörner [10], and its application to systems with coupled channels was further discussed [28, 13].

The two-photon ionization amplitude is calculated as the dipole transition between the intermediate state and a proper final stationary photoionization state Ψf𝒌()superscriptsubscriptΨf𝒌\Psi_{\textrm{f}\bm{k}}^{(-)}roman_Ψ start_POSTSUBSCRIPT f bold_italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( - ) end_POSTSUPERSCRIPT given by the boundary condition [29]

Ψf𝒌()(𝒓)lmileiσl(k)Ylm(𝒌^)gλμFflm,gλμ()(r)Yλμ(𝒓^)Φg,superscriptsubscriptΨf𝒌𝒓subscript𝑙𝑚superscript𝑖𝑙superscript𝑒𝑖subscript𝜎𝑙𝑘superscriptsubscript𝑌𝑙𝑚^𝒌subscript𝑔𝜆𝜇superscriptsubscript𝐹f𝑙𝑚𝑔𝜆𝜇𝑟superscriptsubscript𝑌𝜆𝜇^𝒓subscriptΦ𝑔\displaystyle\Psi_{\textrm{f}\bm{k}}^{(-)}(\bm{r})\rightarrow\sum_{lm}i^{l}e^{% -i\sigma_{l}(k)}Y_{l}^{m*}(\hat{\bm{k}})\sum_{g\lambda\mu}F_{\textrm{f}lm,g% \lambda\mu}^{(-)}(r)Y_{\lambda}^{\mu}(\hat{\bm{r}})\Phi_{g}\,,roman_Ψ start_POSTSUBSCRIPT f bold_italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( - ) end_POSTSUPERSCRIPT ( bold_italic_r ) → ∑ start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT italic_i start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_σ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( italic_k ) end_POSTSUPERSCRIPT italic_Y start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m ∗ end_POSTSUPERSCRIPT ( over^ start_ARG bold_italic_k end_ARG ) ∑ start_POSTSUBSCRIPT italic_g italic_λ italic_μ end_POSTSUBSCRIPT italic_F start_POSTSUBSCRIPT f italic_l italic_m , italic_g italic_λ italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( - ) end_POSTSUPERSCRIPT ( italic_r ) italic_Y start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT ( over^ start_ARG bold_italic_r end_ARG ) roman_Φ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ,
Fflm,gλμ()(r)=ir2πk(Hl+δgλμflmHlSgλμflm).superscriptsubscript𝐹f𝑙𝑚𝑔𝜆𝜇𝑟𝑖𝑟2𝜋𝑘superscriptsubscript𝐻𝑙subscriptsuperscript𝛿f𝑙𝑚𝑔𝜆𝜇superscriptsubscript𝐻𝑙superscriptsubscript𝑆𝑔𝜆𝜇f𝑙𝑚\displaystyle F_{\textrm{f}lm,g\lambda\mu}^{(-)}(r)=\frac{-i}{r\sqrt{2\pi k}}(% H_{l}^{+}\delta^{\textrm{f}lm}_{g\lambda\mu}-H_{l}^{-}S_{g\lambda\mu}^{\textrm% {f}lm*})\,.italic_F start_POSTSUBSCRIPT f italic_l italic_m , italic_g italic_λ italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( - ) end_POSTSUPERSCRIPT ( italic_r ) = divide start_ARG - italic_i end_ARG start_ARG italic_r square-root start_ARG 2 italic_π italic_k end_ARG end_ARG ( italic_H start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT italic_δ start_POSTSUPERSCRIPT f italic_l italic_m end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_g italic_λ italic_μ end_POSTSUBSCRIPT - italic_H start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT italic_S start_POSTSUBSCRIPT italic_g italic_λ italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT f italic_l italic_m ∗ end_POSTSUPERSCRIPT ) . (10)

For pure one-electron Coulomb problem the S𝑆Sitalic_S-matrix is trivial, the formula applies to all distances, and it simplifies to the regular Coulomb wave Fl(η,ρ)subscript𝐹𝑙𝜂𝜌F_{l}(\eta,\rho)italic_F start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( italic_η , italic_ρ ) in each partial wave channel:

Ψf𝒌()(𝒓)=lmileiσl(k)Ylm(𝒌^)Ylm(𝒓^)Fflm()(r)Φf,superscriptsubscriptΨf𝒌𝒓subscript𝑙𝑚superscript𝑖𝑙superscript𝑒𝑖subscript𝜎𝑙𝑘superscriptsubscript𝑌𝑙𝑚^𝒌superscriptsubscript𝑌𝑙𝑚^𝒓superscriptsubscript𝐹f𝑙𝑚𝑟subscriptΦf\displaystyle\Psi_{\textrm{f}\bm{k}}^{(-)}(\bm{r})=\sum_{lm}i^{l}e^{-i\sigma_{% l}(k)}Y_{l}^{m*}(\hat{\bm{k}})Y_{l}^{m}(\hat{\bm{r}})F_{\textrm{f}lm}^{(-)}(r)% \Phi_{\textrm{f}}\,,roman_Ψ start_POSTSUBSCRIPT f bold_italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( - ) end_POSTSUPERSCRIPT ( bold_italic_r ) = ∑ start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT italic_i start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_σ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( italic_k ) end_POSTSUPERSCRIPT italic_Y start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m ∗ end_POSTSUPERSCRIPT ( over^ start_ARG bold_italic_k end_ARG ) italic_Y start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT ( over^ start_ARG bold_italic_r end_ARG ) italic_F start_POSTSUBSCRIPT f italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( - ) end_POSTSUPERSCRIPT ( italic_r ) roman_Φ start_POSTSUBSCRIPT f end_POSTSUBSCRIPT ,
Fflm()(r)=1r2πkFl(Zk,kr).superscriptsubscript𝐹f𝑙𝑚𝑟1𝑟2𝜋𝑘subscript𝐹𝑙𝑍𝑘𝑘𝑟\displaystyle F_{\textrm{f}lm}^{(-)}(r)=\frac{1}{r}\sqrt{\frac{2}{\pi k}}F_{l}% (-\tfrac{Z}{k},kr)\,.italic_F start_POSTSUBSCRIPT f italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( - ) end_POSTSUPERSCRIPT ( italic_r ) = divide start_ARG 1 end_ARG start_ARG italic_r end_ARG square-root start_ARG divide start_ARG 2 end_ARG start_ARG italic_π italic_k end_ARG end_ARG italic_F start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( - divide start_ARG italic_Z end_ARG start_ARG italic_k end_ARG , italic_k italic_r ) . (11)

The asymptotic approximation assumes validity of the asymptotic forms of the wavefunctions Eq. (4) and (10) throughout the whole radial range including the origin and neglects the non-diagonal, S𝑆Sitalic_S-matrix-dependent second term in Eq. (10).

Another possibility, employed in this work, is to replace the asymptotic form of the final-state wavefunction Eq. (10) with an exact hydrogenic solution Eq. (11). This will prove advantageous later, because the regular Coulomb function from Eq. (11) has the correct angular-momentum-dependent asymptotics rl+1superscript𝑟𝑙1r^{l+1}italic_r start_POSTSUPERSCRIPT italic_l + 1 end_POSTSUPERSCRIPT at r0𝑟0r\rightarrow 0italic_r → 0 and does not diverge as opposed to the usual choice Eq. (10).

Ultimately, the resulting two-photon amplitude is written as [30]

Tfi(2)(𝒌)superscriptsubscript𝑇fi2𝒌\displaystyle T_{\textrm{fi}}^{(2)}(\bm{k})italic_T start_POSTSUBSCRIPT fi end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ( bold_italic_k ) =2πiΨf𝒌()|VIR|Ψi+Ω(+)absent2𝜋𝑖quantum-operator-productsuperscriptsubscriptΨf𝒌subscript𝑉IRsuperscriptsubscriptΨiΩ\displaystyle=-2\pi i\langle\Psi_{\textrm{f}\bm{k}}^{(-)}|V_{\text{IR}}|\Psi_{% \textrm{i}+\Omega}^{(+)}\rangle= - 2 italic_π italic_i ⟨ roman_Ψ start_POSTSUBSCRIPT f bold_italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( - ) end_POSTSUPERSCRIPT | italic_V start_POSTSUBSCRIPT IR end_POSTSUBSCRIPT | roman_Ψ start_POSTSUBSCRIPT i + roman_Ω end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( + ) end_POSTSUPERSCRIPT ⟩
=lmTfi,lm(2)(k)Ylm(𝒌^),absentsubscript𝑙𝑚superscriptsubscript𝑇fi𝑙𝑚2𝑘superscriptsubscript𝑌𝑙𝑚^𝒌\displaystyle=\sum_{lm}T_{\textrm{fi},lm}^{(2)}(k)Y_{l}^{m}(\hat{\bm{k}})\,,= ∑ start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT fi , italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ( italic_k ) italic_Y start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT ( over^ start_ARG bold_italic_k end_ARG ) , (12)

where each partial wave channel contribution

Tfi,lm(2)(k)p[Tfi,lmp,pws(2)(k)+Tfi,lmp,ion(2)(k)],superscriptsubscript𝑇fi𝑙𝑚2𝑘subscript𝑝delimited-[]superscriptsubscript𝑇fi𝑙𝑚𝑝pws2𝑘superscriptsubscript𝑇fi𝑙𝑚𝑝ion2𝑘T_{\textrm{fi},lm}^{(2)}(k)\approx\sum_{p}[T_{\textrm{fi},lmp,\text{pws}}^{(2)% }(k)+T_{\textrm{fi},lmp,\text{ion}}^{(2)}(k)],italic_T start_POSTSUBSCRIPT fi , italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ( italic_k ) ≈ ∑ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT [ italic_T start_POSTSUBSCRIPT fi , italic_l italic_m italic_p , pws end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ( italic_k ) + italic_T start_POSTSUBSCRIPT fi , italic_l italic_m italic_p , ion end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ( italic_k ) ] , (13)

arises in a free-free or ion-ion transition:

Tfi,lmp,pws(2)(k)superscriptsubscript𝑇fi𝑙𝑚𝑝pws2𝑘\displaystyle T_{\textrm{fi},lmp,\text{pws}}^{(2)}(k)italic_T start_POSTSUBSCRIPT fi , italic_l italic_m italic_p , pws end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ( italic_k ) =2πilm|𝜺^IR𝒓^|λpμpδfιpabsent2𝜋𝑖quantum-operator-product𝑙𝑚subscript^𝜺IR^𝒓subscript𝜆𝑝subscript𝜇𝑝superscriptsubscript𝛿fsubscript𝜄𝑝\displaystyle=-2\pi i\langle lm|\hat{\bm{\varepsilon}}_{\text{IR}}\cdot\hat{% \bm{r}}|\lambda_{p}\mu_{p}\rangle\delta_{\textrm{f}}^{\iota_{p}}= - 2 italic_π italic_i ⟨ italic_l italic_m | over^ start_ARG bold_italic_ε end_ARG start_POSTSUBSCRIPT IR end_POSTSUBSCRIPT ⋅ over^ start_ARG bold_italic_r end_ARG | italic_λ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ italic_δ start_POSTSUBSCRIPT f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ι start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_POSTSUPERSCRIPT
×Aκpλpkl(1)dιpλpμp,i(1)(κp),absentsuperscriptsubscript𝐴subscript𝜅𝑝subscript𝜆𝑝𝑘𝑙1superscriptsubscript𝑑subscript𝜄𝑝subscript𝜆𝑝subscript𝜇𝑝i1subscript𝜅𝑝\displaystyle\qquad\times A_{\kappa_{p}\lambda_{p}kl}^{(1)}d_{\iota_{p}\lambda% _{p}\mu_{p},\textrm{i}}^{(1)}(\kappa_{p})\,,× italic_A start_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_ι start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT , i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_κ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) , (14)
Tfi,lmp,ion(2)(k)superscriptsubscript𝑇fi𝑙𝑚𝑝ion2𝑘\displaystyle T_{\textrm{fi},lmp,\text{ion}}^{(2)}(k)italic_T start_POSTSUBSCRIPT fi , italic_l italic_m italic_p , ion end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ( italic_k ) =2πiΦf|𝜺^IR𝑫ion|Φιpδlλpδmμpabsent2𝜋𝑖quantum-operator-productsubscriptΦfsubscript^𝜺IRsubscript𝑫ionsubscriptΦsubscript𝜄𝑝superscriptsubscript𝛿𝑙subscript𝜆𝑝superscriptsubscript𝛿𝑚subscript𝜇𝑝\displaystyle=-2\pi i\langle\Phi_{\textrm{f}}|\hat{\bm{\varepsilon}}_{\text{IR% }}\cdot\bm{D}_{\text{ion}}|\Phi_{\iota_{p}}\rangle\delta_{l}^{\lambda_{p}}% \delta_{m}^{\mu_{p}}= - 2 italic_π italic_i ⟨ roman_Φ start_POSTSUBSCRIPT f end_POSTSUBSCRIPT | over^ start_ARG bold_italic_ε end_ARG start_POSTSUBSCRIPT IR end_POSTSUBSCRIPT ⋅ bold_italic_D start_POSTSUBSCRIPT ion end_POSTSUBSCRIPT | roman_Φ start_POSTSUBSCRIPT italic_ι start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_POSTSUBSCRIPT ⟩ italic_δ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_λ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_δ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_μ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_POSTSUPERSCRIPT
×Aκpλpkl(0)dιpλpμp,i(1)(κp).absentsuperscriptsubscript𝐴subscript𝜅𝑝subscript𝜆𝑝𝑘𝑙0superscriptsubscript𝑑subscript𝜄𝑝subscript𝜆𝑝subscript𝜇𝑝i1subscript𝜅𝑝\displaystyle\qquad\times A_{\kappa_{p}\lambda_{p}kl}^{(0)}d_{\iota_{p}\lambda% _{p}\mu_{p},\textrm{i}}^{(1)}(\kappa_{p})\,.× italic_A start_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_ι start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT , i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_κ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) . (15)

The continuum-continuum integral Aκλkl(s)superscriptsubscript𝐴𝜅𝜆𝑘𝑙𝑠A_{\kappa\lambda kl}^{(s)}italic_A start_POSTSUBSCRIPT italic_κ italic_λ italic_k italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT is

Aκλkl(s)superscriptsubscript𝐴𝜅𝜆𝑘𝑙𝑠\displaystyle A_{\kappa\lambda kl}^{(s)}italic_A start_POSTSUBSCRIPT italic_κ italic_λ italic_k italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT =2κkiλleiσliσλabsent2𝜅𝑘superscript𝑖𝜆𝑙superscript𝑒𝑖subscript𝜎𝑙𝑖subscript𝜎𝜆\displaystyle=-\frac{2}{\sqrt{\kappa k}}i^{\lambda-l}e^{i\sigma_{l}-i\sigma_{% \lambda}}= - divide start_ARG 2 end_ARG start_ARG square-root start_ARG italic_κ italic_k end_ARG end_ARG italic_i start_POSTSUPERSCRIPT italic_λ - italic_l end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_σ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT - italic_i italic_σ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT
×0Fl(Zk,kr)rsHλ+(Zκ,κr)dr.\displaystyle\quad\times\int_{0}^{\infty}F_{l}(-\tfrac{Z}{k},kr)r^{s}H_{% \lambda}^{+}(-\tfrac{Z}{\kappa},\kappa r)\mathrm{d}r\,.× ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_F start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( - divide start_ARG italic_Z end_ARG start_ARG italic_k end_ARG , italic_k italic_r ) italic_r start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT italic_H start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( - divide start_ARG italic_Z end_ARG start_ARG italic_κ end_ARG , italic_κ italic_r ) roman_d italic_r . (16)

While the partial-wave mixing angular integrals in Eqs. (14) and (15) have always been an integral ingredient of the asymptotic approximation and were discussed at length by Baykusheva and Wörner [10], the radial integration in Eq. (16) has always been strongly approximated. The integrand is usually simplified by neglecting the first term of Fl=(Hl+Hl)/2isubscript𝐹𝑙superscriptsubscript𝐻𝑙superscriptsubscript𝐻𝑙2𝑖F_{l}=(H_{l}^{+}-H_{l}^{-})/2iitalic_F start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT = ( italic_H start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT - italic_H start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) / 2 italic_i and replacing both remaining Coulomb-Hankel functions with their long-range asymptotics (5). In other words, it corresponds to an approximation Aκλkl(s)Aκk(s)superscriptsubscript𝐴𝜅𝜆𝑘𝑙𝑠superscriptsubscript𝐴𝜅𝑘𝑠A_{\kappa\lambda kl}^{(s)}\approx A_{\kappa k}^{(s)}italic_A start_POSTSUBSCRIPT italic_κ italic_λ italic_k italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT ≈ italic_A start_POSTSUBSCRIPT italic_κ italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT that cancels the sensitivity to the intermediate and final partial waves, and leads to the expression

Aκλkl(s)Aκk(s)=iseZπ/2κ+Zπ/2kκk(κk)s+1Γ(s+1+ZiκZik)(κk)Zi/κZi/ksuperscriptsubscript𝐴𝜅𝜆𝑘𝑙𝑠superscriptsubscript𝐴𝜅𝑘𝑠superscript𝑖𝑠superscript𝑒𝑍𝜋2𝜅𝑍𝜋2𝑘𝜅𝑘superscript𝜅𝑘𝑠1Γ𝑠1𝑍𝑖𝜅𝑍𝑖𝑘superscript𝜅𝑘𝑍𝑖𝜅𝑍𝑖𝑘\displaystyle A_{\kappa\lambda kl}^{(s)}\rightarrow A_{\kappa k}^{(s)}=i^{s}% \frac{e^{-Z\pi/2\kappa+Z\pi/2k}}{\sqrt{\kappa k}(\kappa-k)^{s+1}}\frac{\Gamma(% s+1+\tfrac{Zi}{\kappa}-\tfrac{Zi}{k})}{(\kappa-k)^{Zi/\kappa-Zi/k}}italic_A start_POSTSUBSCRIPT italic_κ italic_λ italic_k italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT → italic_A start_POSTSUBSCRIPT italic_κ italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT = italic_i start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT divide start_ARG italic_e start_POSTSUPERSCRIPT - italic_Z italic_π / 2 italic_κ + italic_Z italic_π / 2 italic_k end_POSTSUPERSCRIPT end_ARG start_ARG square-root start_ARG italic_κ italic_k end_ARG ( italic_κ - italic_k ) start_POSTSUPERSCRIPT italic_s + 1 end_POSTSUPERSCRIPT end_ARG divide start_ARG roman_Γ ( italic_s + 1 + divide start_ARG italic_Z italic_i end_ARG start_ARG italic_κ end_ARG - divide start_ARG italic_Z italic_i end_ARG start_ARG italic_k end_ARG ) end_ARG start_ARG ( italic_κ - italic_k ) start_POSTSUPERSCRIPT italic_Z italic_i / italic_κ - italic_Z italic_i / italic_k end_POSTSUPERSCRIPT end_ARG
×(2κ)Zi/κ(2k)Zi/k[1+δs1Zi2(κ2+k2)(κk)1+Zi/κZi/k],absentsuperscript2𝜅𝑍𝑖𝜅superscript2𝑘𝑍𝑖𝑘delimited-[]1subscript𝛿𝑠1𝑍𝑖2superscript𝜅2superscript𝑘2𝜅𝑘1𝑍𝑖𝜅𝑍𝑖𝑘\displaystyle\times\frac{(2\kappa)^{Zi/\kappa}}{(2k)^{Zi/k}}\left[1+\delta_{s1% }\frac{Zi}{2}\frac{(\kappa^{-2}+k^{-2})(\kappa-k)}{1+Zi/\kappa-Zi/k}\right]\,,× divide start_ARG ( 2 italic_κ ) start_POSTSUPERSCRIPT italic_Z italic_i / italic_κ end_POSTSUPERSCRIPT end_ARG start_ARG ( 2 italic_k ) start_POSTSUPERSCRIPT italic_Z italic_i / italic_k end_POSTSUPERSCRIPT end_ARG [ 1 + italic_δ start_POSTSUBSCRIPT italic_s 1 end_POSTSUBSCRIPT divide start_ARG italic_Z italic_i end_ARG start_ARG 2 end_ARG divide start_ARG ( italic_κ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT + italic_k start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ) ( italic_κ - italic_k ) end_ARG start_ARG 1 + italic_Z italic_i / italic_κ - italic_Z italic_i / italic_k end_ARG ] , (17)

that can be possibly further refined by making other ad hoc corrections (‘regularized’ variant).

In absence of ion-ion transitions the replacement of the Coulomb-Hankel functions H±superscript𝐻plus-or-minusH^{\pm}italic_H start_POSTSUPERSCRIPT ± end_POSTSUPERSCRIPT by exponentials removes all partial-wave dependence from the infinite integral in Eq. (16). This allows to treat the factor as a ‘universal’ correction independent of the specific system and partial wave,

Tfi(2)(𝒌)2πiAκk(1)dfi(1)(κ𝒌^),superscriptsubscript𝑇fi2𝒌2𝜋𝑖superscriptsubscript𝐴𝜅𝑘1superscriptsubscript𝑑fi1𝜅^𝒌T_{\textrm{fi}}^{(2)}(\bm{k})\approx-2\pi iA_{\kappa k}^{(1)}d_{\textrm{fi}}^{% (1)}(\kappa\hat{\bm{k}})\,,italic_T start_POSTSUBSCRIPT fi end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ( bold_italic_k ) ≈ - 2 italic_π italic_i italic_A start_POSTSUBSCRIPT italic_κ italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT fi end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_κ over^ start_ARG bold_italic_k end_ARG ) , (18)

giving rise, ultimately, to the separability of the RABITT delay into the universal direction-independent continuum-continuum delay τccsubscript𝜏𝑐𝑐\tau_{cc}italic_τ start_POSTSUBSCRIPT italic_c italic_c end_POSTSUBSCRIPT and the one-photon (Wigner-like) delay τ1subscript𝜏1\tau_{1}italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT,

τR(𝒌)subscript𝜏𝑅𝒌\displaystyle\tau_{R}(\bm{k})italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( bold_italic_k ) =12ωarg(Tfi,+(2)Tfi,(2))absent12𝜔superscriptsubscript𝑇fi2superscriptsubscript𝑇fi2\displaystyle=\frac{1}{2\omega}\arg(T_{\textrm{fi},+}^{(2)*}T_{\textrm{fi},-}^% {(2)})= divide start_ARG 1 end_ARG start_ARG 2 italic_ω end_ARG roman_arg ( italic_T start_POSTSUBSCRIPT fi , + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) ∗ end_POSTSUPERSCRIPT italic_T start_POSTSUBSCRIPT fi , - end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT )
12ω[arg(Aκ+k(1)Aκk(1))+arg(dfi(1)(𝒌^κ+)dfi(1)(𝒌^κ))]absent12𝜔delimited-[]superscriptsubscript𝐴subscript𝜅𝑘1superscriptsubscript𝐴subscript𝜅𝑘1superscriptsubscript𝑑fi1^𝒌subscript𝜅superscriptsubscript𝑑fi1^𝒌subscript𝜅\displaystyle\approx\frac{1}{2\omega}\left[\arg(A_{\kappa_{+}k}^{(1)*}A_{% \kappa_{-}k}^{(1)})+\arg(d_{\textrm{fi}}^{(1)*}(\hat{\bm{k}}\kappa_{+})d_{% \textrm{fi}}^{(1)}(\hat{\bm{k}}\kappa_{-}))\right]≈ divide start_ARG 1 end_ARG start_ARG 2 italic_ω end_ARG [ roman_arg ( italic_A start_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) ∗ end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ) + roman_arg ( italic_d start_POSTSUBSCRIPT fi end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) ∗ end_POSTSUPERSCRIPT ( over^ start_ARG bold_italic_k end_ARG italic_κ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ) italic_d start_POSTSUBSCRIPT fi end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( over^ start_ARG bold_italic_k end_ARG italic_κ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT ) ) ]
τR(𝒌)subscript𝜏𝑅𝒌\displaystyle\tau_{R}(\bm{k})italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( bold_italic_k ) τcc(k)+τ1(𝒌).absentsubscript𝜏𝑐𝑐𝑘subscript𝜏1𝒌\displaystyle\approx\tau_{cc}(k)+\tau_{1}(\bm{k})\,.≈ italic_τ start_POSTSUBSCRIPT italic_c italic_c end_POSTSUBSCRIPT ( italic_k ) + italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( bold_italic_k ) . (19)

This decomposition in two independent steps is illustrated in Fig. 3(a) where the dressing pulse affects all photoelectron partial waves equally.

Refer to caption
Figure 3: (a) ’Universal’ dressing approach where the XUV and dressing photons behave independently. Partial wave resolved approach (this work) where (b) one and (c) two ω𝜔\omegaitalic_ω-photons transitions are considered.

II.4 Partial wave dependent IR contribution

The crucial point of our work is to include the coupling of the photoelectron angular momentum to the dressing field. We do this by evaluating the continuum-continuum transition amplitude Eq. (16) exactly. Use of the exact partial-wave specific correction factors Aκλklsubscript𝐴𝜅𝜆𝑘𝑙A_{\kappa\lambda kl}italic_A start_POSTSUBSCRIPT italic_κ italic_λ italic_k italic_l end_POSTSUBSCRIPT introduces a qualitative change: it generally prevents the factorization of the amplitudes given by Eq. (18) and the separability of the delays given by Eq. (19). Only in simple systems, where there is only one relevant residual ion state (f) and one intermediate partial wave (λ,μ)subscript𝜆subscript𝜇(\lambda_{\circ},\mu_{\circ})( italic_λ start_POSTSUBSCRIPT ∘ end_POSTSUBSCRIPT , italic_μ start_POSTSUBSCRIPT ∘ end_POSTSUBSCRIPT ), one can still achieve some factorization, as then the summation over the contributions of the intermediate partial waves in Eq. (13) reduces to a single term and the ion-ion transition does not contribute,

Tfi(2)(𝒌)superscriptsubscript𝑇fiabsent2𝒌\displaystyle T_{\textrm{f${}_{\circ}$i}}^{\circ{}(2)}(\bm{k})italic_T start_POSTSUBSCRIPT f i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ ( 2 ) end_POSTSUPERSCRIPT ( bold_italic_k ) =2πi[lmAκλkl(1)lm|𝜺^𝒓^|λμYlm(𝒌^)]absent2𝜋𝑖delimited-[]subscript𝑙𝑚superscriptsubscript𝐴𝜅subscript𝜆𝑘𝑙1quantum-operator-product𝑙𝑚^𝜺^𝒓subscript𝜆subscript𝜇superscriptsubscript𝑌𝑙𝑚^𝒌\displaystyle=-2\pi i\left[\sum_{lm}A_{\kappa\lambda_{\circ}kl}^{(1)}\langle lm% |\hat{\bm{\varepsilon}}\cdot\hat{\bm{r}}|\lambda_{\circ}\mu_{\circ}\rangle Y_{% l}^{m}(\hat{\bm{k}})\right]= - 2 italic_π italic_i [ ∑ start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT italic_κ italic_λ start_POSTSUBSCRIPT ∘ end_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ⟨ italic_l italic_m | over^ start_ARG bold_italic_ε end_ARG ⋅ over^ start_ARG bold_italic_r end_ARG | italic_λ start_POSTSUBSCRIPT ∘ end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT ∘ end_POSTSUBSCRIPT ⟩ italic_Y start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT ( over^ start_ARG bold_italic_k end_ARG ) ]
×df,λμ,i(1)(κ).absentsuperscriptsubscript𝑑subscriptfsubscript𝜆subscript𝜇i1𝜅\displaystyle\qquad\times d_{\textrm{f}_{\circ},\lambda_{\circ}\mu_{\circ},% \textrm{i}}^{(1)}(\kappa)\,.× italic_d start_POSTSUBSCRIPT f start_POSTSUBSCRIPT ∘ end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT ∘ end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT ∘ end_POSTSUBSCRIPT , i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_κ ) . (20)

This factorization leads to separation of the RABITT delay to

τR(𝒌)τcc(𝒌)+τ1(k).superscriptsubscript𝜏𝑅𝒌subscript𝜏𝑐𝑐𝒌superscriptsubscript𝜏1𝑘\tau_{R}^{\circ}(\bm{k})\approx\tau_{cc}(\bm{k})+\tau_{1}^{\circ}(k)\,.italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT ( bold_italic_k ) ≈ italic_τ start_POSTSUBSCRIPT italic_c italic_c end_POSTSUBSCRIPT ( bold_italic_k ) + italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT ( italic_k ) . (21)

In contrast to the conventional situation in Eq. (19), here the one-photon delay τ1superscriptsubscript𝜏1\tau_{1}^{\circ}italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT is angularly-independent because the intermediate wavefunction consists of a single partial wave only, but it becomes angularly-dependent through the continuum-continuum delay τccsubscript𝜏𝑐𝑐\tau_{cc}italic_τ start_POSTSUBSCRIPT italic_c italic_c end_POSTSUBSCRIPT. In less trivial systems the combination is more complex and becomes difficult to interpret as a whole. Instead, one has to work with individual partial waves composing the total ionization signal as given by Eqs. (13)–(15).

To sum up, the asymptotic approximation used to analyze the RABITT experiments can be qualitatively significantly improved by calculating the Aκλklsubscript𝐴𝜅𝜆𝑘𝑙A_{\kappa\lambda kl}italic_A start_POSTSUBSCRIPT italic_κ italic_λ italic_k italic_l end_POSTSUBSCRIPT coefficients more accurately. This naturally leads to distinct photoionization delays accumulated by individual partial waves in the free-free transition. Consequently, such approach allows improved interpretation of angular dependence and angular averaging of RABITT oscillation, where multiple non-equivalent partial waves mix and interact. Also, neither use of ab initio time-dependent method, nor of the second order perturbation theory are needed to obtain the amplitudes: the only input data are the one-photon ionization amplitudes.

While there is no closed-form expression for the value of the integral in Eq. (16), its numerical evaluation can be done at rather low computational cost. Additionally, it is a universal coefficient that can be tabulated once and for all and used for any target. In our implementation the integration range is divided into two regions as follows:

  • The region r<r1𝑟subscript𝑟1r<r_{1}italic_r < italic_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT before some asymptotic distance r1subscript𝑟1r_{1}italic_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is integrated using Levin quadrature [31], which is well-suited for highly oscillatory functions and has proved invaluable for the R-matrix multi-photon method [13].

  • The asymptotic region r1<r<+subscript𝑟1𝑟r_{1}<r<+\inftyitalic_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < italic_r < + ∞ is treated by regularization using a decreasing exponential, expansion of both Coulomb functions in asymptotic series and by integrating their product term by term using asymptotic integration by parts [32, 15].

The parameter r1subscript𝑟1r_{1}italic_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is subject to convergence checking, but we have found the value of r1=100subscript𝑟1100r_{1}=100italic_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 100 a.u. to give good agreement with direct second-order perturbation theory calculations demonstrated in Section III. Additionally, the special case of the radial integral in Eq. (16) for the ion-ion transitions, with q=0𝑞0q=0italic_q = 0 and l=λ𝑙𝜆l=\lambdaitalic_l = italic_λ, can be also calculated analytically and expressed in a closed form [33]:

limc0+0Fl(Zk,kr)Hl+(Zκ,κr)ecrdr=subscript𝑐superscript0superscriptsubscript0subscript𝐹𝑙𝑍𝑘𝑘𝑟superscriptsubscript𝐻𝑙𝑍𝜅𝜅𝑟superscripte𝑐𝑟differential-d𝑟absent\displaystyle\lim_{c\rightarrow 0^{+}}\int_{0}^{\infty}F_{l}(-\tfrac{Z}{k},kr)% H_{l}^{+}(-\tfrac{Z}{\kappa},\kappa r)\mathrm{e}^{-cr}\mathrm{d}r=roman_lim start_POSTSUBSCRIPT italic_c → 0 start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_F start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( - divide start_ARG italic_Z end_ARG start_ARG italic_k end_ARG , italic_k italic_r ) italic_H start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( - divide start_ARG italic_Z end_ARG start_ARG italic_κ end_ARG , italic_κ italic_r ) roman_e start_POSTSUPERSCRIPT - italic_c italic_r end_POSTSUPERSCRIPT roman_d italic_r =
kk2κ2(kκ)leZπ/2κ+Zπ/2k|Γ(l+1Zi/k)Γ(l+1Zi/κ)|.𝑘superscript𝑘2superscript𝜅2superscript𝑘𝜅𝑙superscript𝑒𝑍𝜋2𝜅𝑍𝜋2𝑘Γ𝑙1𝑍𝑖𝑘Γ𝑙1𝑍𝑖𝜅\displaystyle\frac{k}{k^{2}-\kappa^{2}}\left(\frac{k}{\kappa}\right)^{l}e^{-Z% \pi/2\kappa+Z\pi/2k}\left|\frac{\Gamma(l+1-Zi/k)}{\Gamma(l+1-Zi/\kappa)}\right|.divide start_ARG italic_k end_ARG start_ARG italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_κ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( divide start_ARG italic_k end_ARG start_ARG italic_κ end_ARG ) start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_Z italic_π / 2 italic_κ + italic_Z italic_π / 2 italic_k end_POSTSUPERSCRIPT | divide start_ARG roman_Γ ( italic_l + 1 - italic_Z italic_i / italic_k ) end_ARG start_ARG roman_Γ ( italic_l + 1 - italic_Z italic_i / italic_κ ) end_ARG | . (22)

Here the imaginary component of Hl+superscriptsubscript𝐻𝑙H_{l}^{+}italic_H start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT does not contribute due to orthogonality of the regular Coulomb waves and the only contribution comes from the real part of Hl+superscriptsubscript𝐻𝑙H_{l}^{+}italic_H start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT. The formula has been obtained from the general one [33] using the known limiting behaviour of Coulomb functions at r0𝑟0r\rightarrow 0italic_r → 0. Substituting this identity into Eq. (16) leads to

Aκlkl(0)=2k2κ2kl+1/2eZπ/2kΓ(l+1Zi/k)κl+1/2eZπ/2κΓ(l+1Zi/κ).superscriptsubscript𝐴𝜅𝑙𝑘𝑙02superscript𝑘2superscript𝜅2superscript𝑘𝑙12superscript𝑒𝑍𝜋2𝑘Γ𝑙1𝑍𝑖𝑘superscript𝜅𝑙12superscript𝑒𝑍𝜋2𝜅Γ𝑙1𝑍𝑖𝜅A_{\kappa lkl}^{(0)}=\frac{2}{k^{2}-\kappa^{2}}\frac{k^{l+1/2}e^{Z\pi/2k}% \Gamma(l+1-Zi/k)}{\kappa^{l+1/2}e^{Z\pi/2\kappa}\Gamma(l+1-Zi/\kappa)}\,.italic_A start_POSTSUBSCRIPT italic_κ italic_l italic_k italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT = divide start_ARG 2 end_ARG start_ARG italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_κ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG divide start_ARG italic_k start_POSTSUPERSCRIPT italic_l + 1 / 2 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_Z italic_π / 2 italic_k end_POSTSUPERSCRIPT roman_Γ ( italic_l + 1 - italic_Z italic_i / italic_k ) end_ARG start_ARG italic_κ start_POSTSUPERSCRIPT italic_l + 1 / 2 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_Z italic_π / 2 italic_κ end_POSTSUPERSCRIPT roman_Γ ( italic_l + 1 - italic_Z italic_i / italic_κ ) end_ARG . (23)

Fig. 4 compares the calculated values of the coefficients Aκλkl(1)superscriptsubscript𝐴𝜅𝜆𝑘𝑙1A_{\kappa\lambda kl}^{(1)}italic_A start_POSTSUBSCRIPT italic_κ italic_λ italic_k italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT to the traditional ‘long-range’ asymptotic approximation Aκk(1)superscriptsubscript𝐴𝜅𝑘1A_{\kappa k}^{(1)}italic_A start_POSTSUBSCRIPT italic_κ italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT for some angular momentum transitions. At high energies the two approximations converge, but they differ significantly at low photoelectron energies. The evaluated Coulomb integrals for the plotted and many other transitions are included in the Supplementary Material.

Refer to caption
Figure 4: Values of Aκλkl(1)superscriptsubscript𝐴𝜅𝜆𝑘𝑙1A_{\kappa\lambda kl}^{(1)}italic_A start_POSTSUBSCRIPT italic_κ italic_λ italic_k italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT for absorption of an 800 nm quantum, compared to the ‘long-range’ variant of the traditional asymptotic approximation (Aκk(1)superscriptsubscript𝐴𝜅𝑘1A_{\kappa k}^{(1)}italic_A start_POSTSUBSCRIPT italic_κ italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT, solid black). Left and right panels show the real and imaginary part, respectively, for selected λl𝜆𝑙\lambda\rightarrow litalic_λ → italic_l angular momentum transitions as labelled in the key of the individual panels.

II.5 Higher-order interference pathways

The intermediate state Ψi+Ω(+)superscriptsubscriptΨiΩ\Psi_{\textrm{i}+\Omega}^{(+)}roman_Ψ start_POSTSUBSCRIPT i + roman_Ω end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( + ) end_POSTSUPERSCRIPT, that represents the state of the system after absorption of an ΩΩ\Omegaroman_Ω-photon, can be used to obtain further intermediate states corresponding to additional ω𝜔\omegaitalic_ω-photon absorptions or emissions. For the case with no ion-ion transitions, this has been proposed by Bharti et al. [34]. Additional ω𝜔\omegaitalic_ω absorption can be included in the time-independent description by means of solution of the time-independent Schrödinger equation. The first iteration, leading to intermediate state that interacted once with ΩΩ\Omegaroman_Ω and once with ω𝜔\omegaitalic_ω is

(Ei+Ω+ωH)Ψi+Ω+ω(+)=VIRΨi+ω(+).subscript𝐸iΩ𝜔𝐻superscriptsubscriptΨiΩ𝜔subscript𝑉IRsuperscriptsubscriptΨi𝜔(E_{\textrm{i}}+\Omega+\omega-H)\Psi_{\textrm{i}+\Omega+\omega}^{(+)}=V_{\text% {IR}}\Psi_{\textrm{i}+\omega}^{(+)}\,.( italic_E start_POSTSUBSCRIPT i end_POSTSUBSCRIPT + roman_Ω + italic_ω - italic_H ) roman_Ψ start_POSTSUBSCRIPT i + roman_Ω + italic_ω end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( + ) end_POSTSUPERSCRIPT = italic_V start_POSTSUBSCRIPT IR end_POSTSUBSCRIPT roman_Ψ start_POSTSUBSCRIPT i + italic_ω end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( + ) end_POSTSUPERSCRIPT . (24)

Within an approximation of uncoupled ionization channels, this equation can be solved by application of the hydrogenic Coulomb-Green’s function for the photoelectron,

G(+)(𝒓,𝒓)superscript𝐺𝒓superscript𝒓bold-′\displaystyle G^{(+)}(\bm{r},\bm{r^{\prime}})italic_G start_POSTSUPERSCRIPT ( + ) end_POSTSUPERSCRIPT ( bold_italic_r , bold_italic_r start_POSTSUPERSCRIPT bold_′ end_POSTSUPERSCRIPT ) =𝒓|1EkH+i0|𝒓absentquantum-operator-product𝒓1subscript𝐸𝑘𝐻𝑖0superscript𝒓\displaystyle=\langle\bm{r}|\frac{1}{E_{k}-H+i0}|\bm{r}^{\prime}\rangle= ⟨ bold_italic_r | divide start_ARG 1 end_ARG start_ARG italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_H + italic_i 0 end_ARG | bold_italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩
=1rrlmgl(+)(r,r)Ylm(𝒓^)Ylm(𝒓^),absent1𝑟superscript𝑟subscript𝑙𝑚superscriptsubscript𝑔𝑙𝑟superscript𝑟subscript𝑌𝑙𝑚^𝒓subscript𝑌𝑙𝑚superscriptsuperscript^𝒓\displaystyle=\frac{1}{rr^{\prime}}\sum_{lm}g_{l}^{(+)}(r,r^{\prime})Y_{lm}(% \hat{\bm{r}})Y_{lm}(\hat{\bm{r}}^{\prime})^{*}\,,= divide start_ARG 1 end_ARG start_ARG italic_r italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT italic_g start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( + ) end_POSTSUPERSCRIPT ( italic_r , italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_Y start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ( over^ start_ARG bold_italic_r end_ARG ) italic_Y start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ( over^ start_ARG bold_italic_r end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ,

to all channels. The radial part of G𝐺Gitalic_G is

gl(+)(r,r)=2kFl(r<)Hl+(r>).superscriptsubscript𝑔𝑙𝑟superscript𝑟2𝑘subscript𝐹𝑙subscript𝑟superscriptsubscript𝐻𝑙subscript𝑟g_{l}^{(+)}(r,r^{\prime})=-\frac{2}{k}F_{l}(r_{<})H_{l}^{+}(r_{>})\,.italic_g start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( + ) end_POSTSUPERSCRIPT ( italic_r , italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = - divide start_ARG 2 end_ARG start_ARG italic_k end_ARG italic_F start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( italic_r start_POSTSUBSCRIPT < end_POSTSUBSCRIPT ) italic_H start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_r start_POSTSUBSCRIPT > end_POSTSUBSCRIPT ) . (25)

The resulting second intermediate state for total energy Etot=Ei+Ω+ω=ef+k2/2subscript𝐸totsubscript𝐸iΩ𝜔subscript𝑒fsuperscript𝑘22E_{\text{tot}}=E_{\textrm{i}}+\Omega+\omega=e_{\textrm{f}}+k^{2}/2italic_E start_POSTSUBSCRIPT tot end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT i end_POSTSUBSCRIPT + roman_Ω + italic_ω = italic_e start_POSTSUBSCRIPT f end_POSTSUBSCRIPT + italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 is

Ψi+Ω+ω(+)(𝒓)superscriptsubscriptΨiΩ𝜔𝒓\displaystyle\Psi_{\textrm{i}+\Omega+\omega}^{(+)}(\bm{r})roman_Ψ start_POSTSUBSCRIPT i + roman_Ω + italic_ω end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( + ) end_POSTSUPERSCRIPT ( bold_italic_r ) =G(+)(𝒓,𝒓)VIR(𝒓)Ψi+Ω(+)(𝒓)d3𝒓,absentsuperscript𝐺𝒓superscript𝒓subscript𝑉IRsuperscript𝒓superscriptsubscriptΨiΩsuperscript𝒓superscriptd3superscript𝒓\displaystyle=\int G^{(+)}(\bm{r},\bm{r}^{\prime})V_{\text{IR}}(\bm{r}^{\prime% })\Psi_{\textrm{i}+\Omega}^{(+)}(\bm{r}^{\prime})\mathrm{d}^{3}\bm{r}^{\prime},= ∫ italic_G start_POSTSUPERSCRIPT ( + ) end_POSTSUPERSCRIPT ( bold_italic_r , bold_italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_V start_POSTSUBSCRIPT IR end_POSTSUBSCRIPT ( bold_italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) roman_Ψ start_POSTSUBSCRIPT i + roman_Ω end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( + ) end_POSTSUPERSCRIPT ( bold_italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) roman_d start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT bold_italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ,
=2klmYlm(𝒓^)Φflm|𝜺^IR𝒓^|λμafλμ,i(1)absent2𝑘subscript𝑙𝑚subscript𝑌𝑙𝑚^𝒓subscriptΦfquantum-operator-product𝑙𝑚subscript^𝜺IR^𝒓𝜆𝜇superscriptsubscript𝑎f𝜆𝜇i1\displaystyle=-\frac{2}{k}\sum_{lm}Y_{lm}(\hat{\bm{r}})\Phi_{\textrm{f}}% \langle lm|\hat{\bm{\varepsilon}}_{\text{IR}}\cdot\hat{\bm{r}}|\lambda\mu% \rangle a_{\textrm{f}\lambda\mu,\text{i}}^{(1)}= - divide start_ARG 2 end_ARG start_ARG italic_k end_ARG ∑ start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT italic_Y start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ( over^ start_ARG bold_italic_r end_ARG ) roman_Φ start_POSTSUBSCRIPT f end_POSTSUBSCRIPT ⟨ italic_l italic_m | over^ start_ARG bold_italic_ε end_ARG start_POSTSUBSCRIPT IR end_POSTSUBSCRIPT ⋅ over^ start_ARG bold_italic_r end_ARG | italic_λ italic_μ ⟩ italic_a start_POSTSUBSCRIPT f italic_λ italic_μ , i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT
×0Fl(r<)Hl+(r>)rHλ+(r)dr.\displaystyle\quad\times\int_{0}^{\infty}F_{l}(r_{<})H_{l}^{+}(r_{>})r^{\prime% }H_{\lambda}^{+}(r^{\prime})\mathrm{d}r^{\prime}\,.× ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_F start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( italic_r start_POSTSUBSCRIPT < end_POSTSUBSCRIPT ) italic_H start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_r start_POSTSUBSCRIPT > end_POSTSUBSCRIPT ) italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_H start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) roman_d italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT . (26)

This can be inserted into the formula for the leading-order perturbation theory transition amplitude,

Tfi(3)(𝒌)=2πiΨf𝒌()|VIR|Ψi+Ω+ω(+),superscriptsubscript𝑇fi3𝒌2𝜋𝑖quantum-operator-productsuperscriptsubscriptΨf𝒌subscript𝑉IRsuperscriptsubscriptΨiΩ𝜔\displaystyle T_{\textrm{fi}}^{(3)}(\bm{k})=2\pi i\langle\Psi_{\textrm{f}\bm{k% }}^{(-)}|V_{\text{IR}}|\Psi_{\textrm{i}+\Omega+\omega}^{(+)}\rangle\,,italic_T start_POSTSUBSCRIPT fi end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( bold_italic_k ) = 2 italic_π italic_i ⟨ roman_Ψ start_POSTSUBSCRIPT f bold_italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( - ) end_POSTSUPERSCRIPT | italic_V start_POSTSUBSCRIPT IR end_POSTSUBSCRIPT | roman_Ψ start_POSTSUBSCRIPT i + roman_Ω + italic_ω end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( + ) end_POSTSUPERSCRIPT ⟩ , (27)

and simplified by keeping only such radial integrals that contain Coulomb-Hankel functions of complementary signs. The resulting expression for the third-order amplitude then features triangular radial integrals of the kind

0FL(r)rHl+(r)0rFl(r)rHλ+(r)drdr.superscriptsubscript0subscript𝐹𝐿𝑟𝑟superscriptsubscript𝐻𝑙𝑟superscriptsubscript0𝑟subscript𝐹𝑙superscript𝑟superscript𝑟superscriptsubscript𝐻𝜆superscript𝑟differential-dsuperscript𝑟differential-d𝑟\displaystyle\int_{0}^{\infty}F_{L}(r)rH_{l}^{+}(r)\int_{0}^{r}F_{l}(r^{\prime% })r^{\prime}H_{\lambda}^{+}(r^{\prime})\mathrm{d}r^{\prime}\mathrm{d}r\,.∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_F start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_r ) italic_r italic_H start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_r ) ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT italic_F start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_H start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) roman_d italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT roman_d italic_r . (28)

This method can be further generalized to even higher orders shown in Fig. 1, and to ion-ion transitions by accounting also for the ion-ion transition term 𝑫ion𝜺^subscript𝑫ion^𝜺\bm{D}_{\text{ion}}\cdot\hat{\bm{\varepsilon}}bold_italic_D start_POSTSUBSCRIPT ion end_POSTSUBSCRIPT ⋅ over^ start_ARG bold_italic_ε end_ARG in VIRsubscript𝑉IRV_{\text{IR}}italic_V start_POSTSUBSCRIPT IR end_POSTSUBSCRIPT that has been disregarded in Eq. (26) for simplicity. Each additional absorption or emission of an ω𝜔\omegaitalic_ω-photon thus adds another layer of nested integration. In [34] these radial integrals were separated into individual factors, essentially by fixed identification of r𝑟ritalic_r and rsuperscript𝑟r^{\prime}italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT with r<subscript𝑟r_{<}italic_r start_POSTSUBSCRIPT < end_POSTSUBSCRIPT and r>subscript𝑟r_{>}italic_r start_POSTSUBSCRIPT > end_POSTSUBSCRIPT, respectively, in Eq. (25). This is motivated by a similar factorization implicitly taking place in the standard RABITT,

0FL(r)rHl+(r)0rFl(r)rPi(r)drdrsuperscriptsubscript0subscript𝐹𝐿𝑟𝑟superscriptsubscript𝐻𝑙𝑟superscriptsubscript0𝑟subscript𝐹𝑙superscript𝑟superscript𝑟subscript𝑃isuperscript𝑟differential-dsuperscript𝑟differential-d𝑟\displaystyle\int_{0}^{\infty}F_{L}(r)rH_{l}^{+}(r)\int_{0}^{r}F_{l}(r^{\prime% })r^{\prime}P_{\text{i}}(r^{\prime})\mathrm{d}r^{\prime}\mathrm{d}r∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_F start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_r ) italic_r italic_H start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_r ) ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT italic_F start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_P start_POSTSUBSCRIPT i end_POSTSUBSCRIPT ( italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) roman_d italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT roman_d italic_r
0FL(r)rHl+(r)drAκlkL(1)0Fl(r)rPi(r)drdl,i(1),absentsubscriptsuperscriptsubscript0subscript𝐹𝐿𝑟𝑟superscriptsubscript𝐻𝑙𝑟differential-dsuperscript𝑟absentsuperscriptsubscript𝐴𝜅𝑙𝑘𝐿1subscriptsuperscriptsubscript0subscript𝐹𝑙superscript𝑟superscript𝑟subscript𝑃isuperscript𝑟differential-dsuperscript𝑟absentsuperscriptsubscript𝑑𝑙i1\displaystyle\approx\underbrace{\int_{0}^{\infty}F_{L}(r)rH_{l}^{+}(r)\mathrm{% d}r^{\prime}}_{\rightarrow A_{\kappa lkL}^{(1)}}\underbrace{\int_{0}^{\infty}F% _{l}(r^{\prime})r^{\prime}P_{\text{i}}(r^{\prime})\mathrm{d}r^{\prime}}_{% \rightarrow d_{l,\text{i}}^{(1)}}\,,≈ under⏟ start_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_F start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_r ) italic_r italic_H start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_r ) roman_d italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG start_POSTSUBSCRIPT → italic_A start_POSTSUBSCRIPT italic_κ italic_l italic_k italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT end_POSTSUBSCRIPT under⏟ start_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_F start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_P start_POSTSUBSCRIPT i end_POSTSUBSCRIPT ( italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) roman_d italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG start_POSTSUBSCRIPT → italic_d start_POSTSUBSCRIPT italic_l , i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT end_POSTSUBSCRIPT , (29)

where Pi(r)subscript𝑃i𝑟P_{\text{i}}(r)italic_P start_POSTSUBSCRIPT i end_POSTSUBSCRIPT ( italic_r ) represents radial part of the initial (Dyson) orbital. Due to the short range of Pi(r)subscript𝑃i𝑟P_{\text{i}}(r)italic_P start_POSTSUBSCRIPT i end_POSTSUBSCRIPT ( italic_r ), the factorization (29) is meaningful, because the extension of the integration domain contains effectively zeros. However, when Eq. (28) is factorized in the same way, the integration domain is doubled even though the integrand is non-zero in the newly added integration space. This has to be compensated by an appropriate combinatoric factor. For an n𝑛nitalic_n-dimensional integral this factor is (n!)1superscript𝑛1(n!)^{-1}( italic_n ! ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, giving

0f10r1f20rn1fndr1drn1n!j=1n0fjdrj.superscriptsubscript0subscript𝑓1superscriptsubscript0subscript𝑟1subscript𝑓2superscriptsubscript0subscript𝑟𝑛1subscript𝑓𝑛differential-dsubscript𝑟1differential-dsubscript𝑟𝑛1𝑛superscriptsubscriptproduct𝑗1𝑛superscriptsubscript0subscript𝑓𝑗differential-dsubscript𝑟𝑗\displaystyle\int\limits_{0}^{\infty}f_{1}\int\limits_{0}^{r_{1}}f_{2}\dots% \int\limits_{0}^{r_{n-1}}f_{n}\mathrm{d}r_{1}\dots\mathrm{d}r_{n}\approx\frac{% 1}{n!}\prod_{j=1}^{n}\int\limits_{0}^{\infty}f_{j}\mathrm{d}r_{j}\,.∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT … ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT roman_d italic_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT … roman_d italic_r start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ≈ divide start_ARG 1 end_ARG start_ARG italic_n ! end_ARG ∏ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT roman_d italic_r start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT . (30)

This factorial factor is missing in the derivation [34].

Consequently, even amplitudes for higher-order transitions can be approximated in a factorized way. However, every possible photoelectron angular momentum pathway (e.g., λpλqlsubscript𝜆𝑝superscriptsubscript𝜆𝑞𝑙\lambda_{p}\rightarrow\lambda_{q}^{\prime}\rightarrow litalic_λ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT → italic_λ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT → italic_l in case of (Ω+2ωΩ2𝜔\Omega+2\omegaroman_Ω + 2 italic_ω) interactions) now needs a different set of factors representing the individual ω𝜔\omegaitalic_ω-photon absorptions. For pure free-free transitions (in absence of ion-ion transitions) in (Ω+2ωΩ2𝜔\Omega+2\omegaroman_Ω + 2 italic_ω) process this gives the second intermediate state

Ψi+Ω+ω(+)(𝒓)1rpqaιpλpμp,i(1)λqμq|𝜺^IR𝒓^|λpμpYλqμq(𝒓^)ΦιpsuperscriptsubscriptΨiΩ𝜔𝒓1𝑟subscript𝑝𝑞superscriptsubscript𝑎subscript𝜄𝑝subscript𝜆𝑝subscript𝜇𝑝i1quantum-operator-productsuperscriptsubscript𝜆𝑞superscriptsubscript𝜇𝑞subscript^𝜺IR^𝒓subscript𝜆𝑝subscript𝜇𝑝superscriptsubscript𝑌superscriptsubscript𝜆𝑞superscriptsubscript𝜇𝑞^𝒓subscriptΦsubscript𝜄𝑝\displaystyle\Psi_{\textrm{i}+\Omega+\omega}^{(+)}(\bm{r})\approx\frac{1}{r}% \sum_{pq}a_{\iota_{p}\lambda_{p}\mu_{p},\text{i}}^{(1)}\langle\lambda_{q}^{% \prime}\mu_{q}^{\prime}|\hat{\bm{\varepsilon}}_{\text{IR}}\cdot\hat{\bm{r}}|% \lambda_{p}\mu_{p}\rangle Y_{\lambda_{q}^{\prime}}^{\mu_{q}^{\prime}}(\hat{\bm% {r}})\Phi_{\iota_{p}}roman_Ψ start_POSTSUBSCRIPT i + roman_Ω + italic_ω end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( + ) end_POSTSUPERSCRIPT ( bold_italic_r ) ≈ divide start_ARG 1 end_ARG start_ARG italic_r end_ARG ∑ start_POSTSUBSCRIPT italic_p italic_q end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_ι start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT , i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ⟨ italic_λ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_μ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | over^ start_ARG bold_italic_ε end_ARG start_POSTSUBSCRIPT IR end_POSTSUBSCRIPT ⋅ over^ start_ARG bold_italic_r end_ARG | italic_λ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ italic_Y start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_μ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ( over^ start_ARG bold_italic_r end_ARG ) roman_Φ start_POSTSUBSCRIPT italic_ι start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_POSTSUBSCRIPT
×(2κq)Hλq+(r)0rFλq(r)rHλp+(r)dr,absent2superscriptsubscript𝜅𝑞superscriptsubscript𝐻superscriptsubscript𝜆𝑞𝑟superscriptsubscript0𝑟subscript𝐹superscriptsubscript𝜆𝑞superscript𝑟superscript𝑟superscriptsubscript𝐻subscript𝜆𝑝superscript𝑟differential-dsuperscript𝑟\displaystyle\times\left(-\frac{2}{\kappa_{q}^{\prime}}\right)H_{\lambda_{q}^{% \prime}}^{+}(r)\int_{0}^{r}F_{\lambda_{q}^{\prime}}(r^{\prime})r^{\prime}H_{% \lambda_{p}}^{+}(r^{\prime})\mathrm{d}r^{\prime},× ( - divide start_ARG 2 end_ARG start_ARG italic_κ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG ) italic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_r ) ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT italic_F start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) roman_d italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , (31)

and the leading-order perturbation amplitude

Tfi(3)(𝒌)superscriptsubscript𝑇fi3𝒌\displaystyle T_{\textrm{fi}}^{(3)}(\bm{k})italic_T start_POSTSUBSCRIPT fi end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( bold_italic_k ) 12!2πilmpqAκpλpκqλq(1)Aκqλqkl(1)absent122𝜋𝑖subscript𝑙𝑚𝑝𝑞superscriptsubscript𝐴subscript𝜅𝑝subscript𝜆𝑝superscriptsubscript𝜅𝑞superscriptsubscript𝜆𝑞1superscriptsubscript𝐴superscriptsubscript𝜅𝑞superscriptsubscript𝜆𝑞𝑘𝑙1\displaystyle\approx\frac{1}{2!}2\pi i\sum_{lmpq}A_{\kappa_{p}\lambda_{p}% \kappa_{q}^{\prime}\lambda_{q}^{\prime}}^{(1)}A_{\kappa_{q}^{\prime}\lambda_{q% }^{\prime}kl}^{(1)}≈ divide start_ARG 1 end_ARG start_ARG 2 ! end_ARG 2 italic_π italic_i ∑ start_POSTSUBSCRIPT italic_l italic_m italic_p italic_q end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_λ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_λ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_k italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT
×lm|𝜺^IR𝒓^|λqμqλqμq|𝜺^IR𝒓^|λpμpabsentquantum-operator-product𝑙𝑚subscript^𝜺IR^𝒓superscriptsubscript𝜆𝑞superscriptsubscript𝜇𝑞quantum-operator-productsuperscriptsubscript𝜆𝑞superscriptsubscript𝜇𝑞subscript^𝜺IR^𝒓subscript𝜆𝑝subscript𝜇𝑝\displaystyle\quad\times\langle lm|\hat{\bm{\varepsilon}}_{\text{IR}}\cdot\hat% {\bm{r}}|\lambda_{q}^{\prime}\mu_{q}^{\prime}\rangle\langle\lambda_{q}^{\prime% }\mu_{q}^{\prime}|\hat{\bm{\varepsilon}}_{\text{IR}}\cdot\hat{\bm{r}}|\lambda_% {p}\mu_{p}\rangle× ⟨ italic_l italic_m | over^ start_ARG bold_italic_ε end_ARG start_POSTSUBSCRIPT IR end_POSTSUBSCRIPT ⋅ over^ start_ARG bold_italic_r end_ARG | italic_λ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_μ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ ⟨ italic_λ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_μ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | over^ start_ARG bold_italic_ε end_ARG start_POSTSUBSCRIPT IR end_POSTSUBSCRIPT ⋅ over^ start_ARG bold_italic_r end_ARG | italic_λ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩
×dfλpμp,i(1)(κp)Ylm(𝒌^).absentsuperscriptsubscript𝑑fsubscript𝜆𝑝subscript𝜇𝑝i1subscript𝜅𝑝superscriptsubscript𝑌𝑙𝑚^𝒌\displaystyle\quad\times d_{\textrm{f}\lambda_{p}\mu_{p},\textrm{i}}^{(1)}(% \kappa_{p})Y_{l}^{m}(\hat{\bm{k}})\,.× italic_d start_POSTSUBSCRIPT f italic_λ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT , i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_κ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) italic_Y start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT ( over^ start_ARG bold_italic_k end_ARG ) . (32)

The agreement between the asymptotic approximation and a full multi-photon calculation for higher orders is compared in Fig. 5. The calculations use a non-relativistic static exchange model of argon, where only the magnetic sublevel m=0𝑚0m=0italic_m = 0 of the 3p3𝑝3p3 italic_p shell is made available for ionization. The polarization and emission directions point along the z𝑧zitalic_z axis in this calculation. The ab initio multi-photon ionization matrix elements used were obtained from UKRmol+ [35]. The approximate amplitudes agree very well with the multi-photon ones, including the magnitude, pointing to the importance of the permutation factor discussed above.

Refer to caption
Figure 5: Multi-photon (Ω+nωΩ𝑛𝜔\Omega+n\omegaroman_Ω + italic_n italic_ω) ionization amplitudes of argon in static exchange model for emission along the polarization axis of the 800 nm dressing field. Magnitudes (left panels) and phases (right panels) are plotted for (from top to bottom) n=0,1,2,3𝑛0123n=0,1,2,3italic_n = 0 , 1 , 2 , 3. Solid lines: Amplitudes from the leading-order perturbation theory. Dashed lines: Partial-wave asymptotic approximation using first-order amplitudes.

III Results

To illustrate the performance of the proposed method, we present results using the two harmonics RABITT protocol that isolates the interferences and high orders are observed up to 3ω𝜔\omegaitalic_ω dressing photons. The theoretical results are supported by measurements in argon with angular resolution. Ionization time delays along the polarization axis and angularly integrated RABITT delays are shown in at low photoelectron energy where the asymptotic approximations is becoming crucial and also calculated at larger photon energy. Additional results, including angular distribution in argon, as well as delays for other atomic gases and in molecules demonstrating utility of the partial wave asymptotic theory all the way down to the boundary of the under-threshold RABITT area [36] are provided in the Supplementary Material.

The calculation, performed with UKRmol+ [35] in D2hsubscript𝐷2D_{2h}italic_D start_POSTSUBSCRIPT 2 italic_h end_POSTSUBSCRIPT point group as the largest available group, is based on a static-exchange model with Hartree-Fock (HF) orbitals of neutral argon obtained in Psi4 [37] from cc-pVDZ basis. It involves three symmetry components B3usubscript𝐵3𝑢B_{3u}italic_B start_POSTSUBSCRIPT 3 italic_u end_POSTSUBSCRIPT, B2usubscript𝐵2𝑢B_{2u}italic_B start_POSTSUBSCRIPT 2 italic_u end_POSTSUBSCRIPT and B1usubscript𝐵1𝑢B_{1u}italic_B start_POSTSUBSCRIPT 1 italic_u end_POSTSUBSCRIPT of the highly symmetric ground state of Ar(3p1)+{}^{+}(3p^{-1})start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT ( 3 italic_p start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) and radial continuum basis set consisting of B-splines spanning the radial range from the origin to the R-matrix radius ra=150a0subscript𝑟𝑎150subscript𝑎0r_{a}=150\,a_{0}italic_r start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT = 150 italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (Bohr radii). The wavelength of the dressing field is assumed λ=800𝜆800\lambda=800italic_λ = 800 nm. The two-photon RABITT delay τRsubscript𝜏𝑅\tau_{R}italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT is calculated by the molecular multi-photon above-threshold ionization method [13]. The asymptotic approximation to the latter is obtained from Eqs. (13) and (14), reusing the one-photon XUV-only dipoles. The continuum-continuum delay τccsubscript𝜏𝑐𝑐\tau_{cc}italic_τ start_POSTSUBSCRIPT italic_c italic_c end_POSTSUBSCRIPT, where used for reference, is evaluated from the ‘long-range’ analytic expression [9]. The comparison with RMT [16] is performed with the same HF model of the atom. The calculation involved a combination of three fields with cos2 envelopes: 16-cycle dressing field with central wavelength 800 nm and peak intensity 1011superscript101110^{11}10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT W/cm2, and the harmonic pair (Ω13subscriptΩ13\Omega_{13}roman_Ω start_POSTSUBSCRIPT 13 end_POSTSUBSCRIPT and Ω15subscriptΩ15\Omega_{15}roman_Ω start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT) each composed by 240-cycle XUV fields with peak intensity 1010superscript101010^{10}10 start_POSTSUPERSCRIPT 10 end_POSTSUPERSCRIPT W/cm2. Both pulses have a duration of approximately 20 fs full width at half maximum (FWHM). The wavefunction was time-propagated for 60 fs in a simulation domain of radius rmax4600a0subscript𝑟max4600subscript𝑎0r_{\text{max}}\approx 4600~{}a_{0}italic_r start_POSTSUBSCRIPT max end_POSTSUBSCRIPT ≈ 4600 italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Fourier filtering is applied to the simulated photoelectron spectra to extract the parameters of Eq. (1).

Angularly resolved interference signal is the most sensitive probe of the relative magnitudes and phases of individual contributing partial waves; it contains enough information for reconstruction of phases of the individual partial waves in two-photon ionization [38].

Refer to caption
Figure 6: Angularly resolved RABITT oscillation amplitude \mathcal{B}caligraphic_B (a,c,e) and delay τRsubscript𝜏𝑅\tau_{R}italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT (b,d,f) in attoseconds extracted from two-harmonic (Ω13subscriptΩ13\Omega_{13}roman_Ω start_POSTSUBSCRIPT 13 end_POSTSUBSCRIPT and Ω15subscriptΩ15\Omega_{15}roman_Ω start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT) RABITT in argon 3p3𝑝3p3 italic_p for the channels SB14 (a,b), MB15 and OSB16 as a function of the angle ϑitalic-ϑ\varthetaitalic_ϑ between the polarization and photoelectron emission axis. Red solid curves mark the multiphoton results, dashed black represents the partial-wave asymptotic approximation, green with circles is our measurement, green shaded area its experimental uncertainty at 1σ𝜎\sigmaitalic_σ. The light dash-dot line marks the calculation from [19]. Triangles [19], empty circles and diamonds [12] and squares [39] mark earlier measurements at SB14. The oscillation amplitudes are normalized to one, except for the black dashed curve in the right half of panels a,c,e, where it is scaled by the same factor as the multiphoton (solid) result. The delays are shifted to zero along the polarization axis for better comparison.

Figure 6 compares the angularly resolved RABITT in SB, MB and OSB1 for ioinization from the (complete) 3p3𝑝3p3 italic_p-shell. The oscillation amplitudes are given by |T<T>|proportional-tosuperscriptsubscript𝑇subscript𝑇\mathcal{B}\propto|T_{<}^{*}T_{>}|caligraphic_B ∝ | italic_T start_POSTSUBSCRIPT < end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_T start_POSTSUBSCRIPT > end_POSTSUBSCRIPT |, and the delay by τR=arg(T<T>)/2ωsubscript𝜏𝑅superscriptsubscript𝑇subscript𝑇2𝜔\tau_{R}=\arg(T_{<}^{*}T_{>})/2\omegaitalic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT = roman_arg ( italic_T start_POSTSUBSCRIPT < end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_T start_POSTSUBSCRIPT > end_POSTSUBSCRIPT ) / 2 italic_ω [25]. Our static exchange model is compared with our measurements and also with other results available in the literature [19, 12, 39] on the SB14. It appears that the angularly resolved τR(ϑ)subscript𝜏𝑅italic-ϑ\tau_{R}(\vartheta)italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_ϑ ) remains close to zero up to ϑsimilar-toitalic-ϑabsent\vartheta\simitalic_ϑ ∼45 and rapidly decreases to large negative values for angles closer to perpendicular emission. The distributions of τR(ϑ)subscript𝜏𝑅italic-ϑ\tau_{R}(\vartheta)italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_ϑ ) are not exactly the same for the SB, MB and OSB1. The partial waves involved in SBs and MBs do not have the same parities (see Fig. 3) and cannot reproduce the exact same angular distribution (except if τRsubscript𝜏𝑅\tau_{R}italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT is isotropic). The OSBn𝑛nitalic_n involves higher angular quantum numbers and concentrates the oscillation along the polarization axis.

Refer to caption
Figure 7: Comparison of the ionization time delay determined from single XUV photon ionisation (solid line, τ1+τccLRsubscript𝜏1superscriptsubscript𝜏𝑐𝑐𝐿𝑅\tau_{1}+\tau_{cc}^{LR}italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_τ start_POSTSUBSCRIPT italic_c italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L italic_R end_POSTSUPERSCRIPT), and the predicted oscillation phase along the polarisation axis for the SB, MB and OSB (broken lines, τRsubscript𝜏𝑅\tau_{R}italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT) at a given final photoelectron energy, all from time-independent lowest order perturbation theory. Connected circles (labeled ‘RMT (two-harmonic)‘) represent delays extracted from six separate non-perturbative time-dependent simulations of ‘two-harmonic RABITT’, each giving rise to a specific sideband (labelled atop) and to the surrounding bands. The two methods are in very close agreement. Empty triangles show a time-dependent simulation with a full comb attosecond pulse train. Emission (a,c) along polarization axis and (b,d) angularly-integrated. Panels (a) and (b) show a subsets of (c) and (d) with partial-wave asymptotic results included as lines with squares and compared to experiment. The experimental points were uniformly shifted to match the calculation for oriented emission at SB14.

Figures 6(a,c,e) show that the angularly resolved oscillation amplitude \mathcal{B}caligraphic_B is almost the same for SB and MB despite a different arrangement of dressing photons (see Fig. 1(a,b)) and gets narrower with the number of dressing photons involved, such as for OSB1. This can be understood using the asymptotic approximation, even when the angular momentum dependence of the free-free transition coefficient is neglected. As there are no ion-ion transitions in the static exchange model of argon, the amplitude of a higher-order RABITT ionization for oriented emission can be written as

Tfi(n+1)(𝒌)(𝜺^IR𝒌^)nAκn1k(1)Aκκ1(1)dfi(1)(𝒌^κ).similar-tosuperscriptsubscript𝑇fi𝑛1𝒌superscriptsubscript^𝜺IR^𝒌𝑛superscriptsubscript𝐴subscript𝜅𝑛1𝑘1superscriptsubscript𝐴𝜅subscript𝜅11superscriptsubscriptdfi1^𝒌𝜅T_{\textrm{fi}}^{(n+1)}(\bm{k})\sim(\hat{\bm{\varepsilon}}_{\text{IR}}\cdot% \hat{\bm{k}})^{n}A_{\kappa_{n-1}k}^{(1)}\dots A_{\kappa\kappa_{1}}^{(1)}% \mathrm{d}_{\textrm{fi}}^{(1)}(\hat{\bm{k}}\kappa)\,.italic_T start_POSTSUBSCRIPT fi end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n + 1 ) end_POSTSUPERSCRIPT ( bold_italic_k ) ∼ ( over^ start_ARG bold_italic_ε end_ARG start_POSTSUBSCRIPT IR end_POSTSUBSCRIPT ⋅ over^ start_ARG bold_italic_k end_ARG ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT … italic_A start_POSTSUBSCRIPT italic_κ italic_κ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT roman_d start_POSTSUBSCRIPT fi end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( over^ start_ARG bold_italic_k end_ARG italic_κ ) . (33)

This high-energy limit suggests that the RABITT signal

(𝒌)|Tfi(p)(𝒌)Tfi(q)(𝒌)|,similar-to𝒌superscriptsubscript𝑇fi𝑝𝒌superscriptsubscript𝑇fi𝑞𝒌\mathcal{B}(\bm{k})\sim|T_{\textrm{fi}}^{(p)*}(\bm{k})T_{\textrm{fi}}^{(q)}(% \bm{k})|\,,caligraphic_B ( bold_italic_k ) ∼ | italic_T start_POSTSUBSCRIPT fi end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_p ) ∗ end_POSTSUPERSCRIPT ( bold_italic_k ) italic_T start_POSTSUBSCRIPT fi end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_q ) end_POSTSUPERSCRIPT ( bold_italic_k ) | , (34)

where p𝑝pitalic_p and q𝑞qitalic_q represent the number of involved photon in each path, acquires an additional emission-angle-dependent factor cos2ϑ=(𝜺^IR𝒌^)2superscript2italic-ϑsuperscriptsubscript^𝜺IR^𝒌2\cos^{2}\vartheta=(\hat{\bm{\varepsilon}}_{\text{IR}}\cdot\hat{\bm{k}})^{2}roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ϑ = ( over^ start_ARG bold_italic_ε end_ARG start_POSTSUBSCRIPT IR end_POSTSUBSCRIPT ⋅ over^ start_ARG bold_italic_k end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for each absorbed ω𝜔\omegaitalic_ω-photon. This factor is responsible for making the distribution more peaked along the polarization axis. One can ascribe this cosine factor to a simple geometrical effect of projection of the electric strength vector along the photoelectron emission direction. In other words, the leaving photoelectron is, asymptotically, affected only by a parallel component of the dressing field.

As established by Bharti et al. [34], if the approximation given by Eq. (33) was accurate, the coefficients Akk(1)superscriptsubscript𝐴𝑘superscript𝑘1A_{kk^{\prime}}^{(1)}italic_A start_POSTSUBSCRIPT italic_k italic_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT in each of the two RABITT ionization pathways would combine to the universal continuum-continuum delays in such a way that the measured RABITT delay would be the same when extracted from any of the bands or sidebands arising from the same pair of harmonics. The only difference predicted by the standard asymptotic theory would be an overall phase factor ip+qsuperscript𝑖𝑝𝑞i^{-p+q}italic_i start_POSTSUPERSCRIPT - italic_p + italic_q end_POSTSUPERSCRIPT arising from unbalanced number of i𝑖iitalic_i factors in the products of the Akk(1)superscriptsubscript𝐴𝑘superscript𝑘1A_{kk^{\prime}}^{(1)}italic_A start_POSTSUBSCRIPT italic_k italic_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT coefficients present in each pathway, see Eq. (17). This factor is responsible for the π𝜋\piitalic_π-shift of the other bands with respect to the central sideband. However, both the theory and the experiment confirm that distinct bands exhibit slightly different angular distribution of the delays, particularly at low energies. This points to insufficiency of the standard asymptotic approximation, Eq. (33), and to a need for the more complete, partial-wave-sensitive theory.

Fig. 7 presents simulated time delays along the polarization axis (a and c) and integrated angularly (b and d) on the low photon energy range of interest (a,b) and over a wider energy range (c,d). The angularly integrated case corresponds to the widest available RABITT measurements performed with a simple electron spectrometer such as magnetic bottle. The delays obtained from the emission-integrated ionization signal is described as,

I(τ)𝐼𝜏\displaystyle I(\tau)italic_I ( italic_τ ) Tfi(p)(𝒌)Tfi(q)(𝒌)d2𝒌^similar-toabsentsuperscriptsubscript𝑇fi𝑝𝒌superscriptsubscript𝑇fi𝑞𝒌superscript𝑑2^𝒌\displaystyle\sim\int T_{\text{fi}}^{(p)*}(\bm{k})T_{\text{fi}}^{(q)}(\bm{k})d% ^{2}\hat{\bm{k}}∼ ∫ italic_T start_POSTSUBSCRIPT fi end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_p ) ∗ end_POSTSUPERSCRIPT ( bold_italic_k ) italic_T start_POSTSUBSCRIPT fi end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_q ) end_POSTSUPERSCRIPT ( bold_italic_k ) italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over^ start_ARG bold_italic_k end_ARG
(𝒌)e2iωτR(𝒌)d2𝒌^.similar-toabsent𝒌superscript𝑒2𝑖𝜔subscript𝜏𝑅𝒌superscript𝑑2^𝒌\displaystyle\sim\int\mathcal{B}(\bm{k})e^{2i\omega\tau_{R}(\bm{k})}d^{2}\hat{% \bm{k}}\,.∼ ∫ caligraphic_B ( bold_italic_k ) italic_e start_POSTSUPERSCRIPT 2 italic_i italic_ω italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( bold_italic_k ) end_POSTSUPERSCRIPT italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over^ start_ARG bold_italic_k end_ARG . (35)

Along the polarization axis, Fig. 7(a), the predicted τRsubscript𝜏𝑅\tau_{R}italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT is almost the same for the SB, MB and OSB specially above 10 eV. In practice, it corresponds to a π𝜋\piitalic_π-shift in Eq. (1) as predicted under the soft photon approximation (see Supplementary Material). The very good agreement between predicted delays extracted from different orders of RABITT is caused by the nearly identical Aκλklsubscript𝐴𝜅𝜆𝑘𝑙A_{\kappa\lambda kl}italic_A start_POSTSUBSCRIPT italic_κ italic_λ italic_k italic_l end_POSTSUBSCRIPT factors pertaining to different partial waves at energies greater than 10 eV, together with absence of nodes in relevant spherical harmonics for the axial emission. Consequently, the multiphoton amplitude can be factorized and the RABITT delay is then well-separable into the asymptotic components according to Eq. (33).

However, the non-trivial angular dependence of (ϑ)italic-ϑ\mathcal{B}(\vartheta)caligraphic_B ( italic_ϑ ) and τR(ϑ)subscript𝜏𝑅italic-ϑ\tau_{R}(\vartheta)italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_ϑ ), which is also different for the individual higher-order RABITT pathways, is responsible for slight differences in the integrated time delay extracted from the individual bands, Fig. 7(b). Such differences fall in the order of tens of attoseconds for argon in the range under consideration.

In general, the delays extracted from the mainbands are very similar to those obtained from the central sidebands even in the emission-integrated case, but this is not always the case with delays obtained from the outer sideband. Such behaviour is easy to explain within the standard asymptotic theory and it is the consequence of the angular integrals appearing in construction of higher orders. The phase ϕ2ω=2ωτRsubscriptitalic-ϕ2𝜔2𝜔subscript𝜏𝑅\phi_{2\omega}=2\omega\tau_{R}italic_ϕ start_POSTSUBSCRIPT 2 italic_ω end_POSTSUBSCRIPT = 2 italic_ω italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT of the emission-integrated RABITT signal in the central sideband is

ϕ2ωSBsuperscriptsubscriptitalic-ϕ2𝜔SB\displaystyle\phi_{2\omega}^{\text{SB}}italic_ϕ start_POSTSUBSCRIPT 2 italic_ω end_POSTSUBSCRIPT start_POSTSUPERSCRIPT SB end_POSTSUPERSCRIPT arglmlmlm|(𝜺^IR𝒓^)2|lmdflm,i(1)(κ+)dflm,i(1)(κ)absentsubscriptsuperscript𝑙superscript𝑚𝑙𝑚quantum-operator-productsuperscript𝑙superscript𝑚superscriptsubscript^𝜺IR^𝒓2𝑙𝑚superscriptsubscript𝑑fsuperscript𝑙superscript𝑚i1subscript𝜅superscriptsubscript𝑑f𝑙𝑚i1subscript𝜅\displaystyle\approx\arg\sum_{l^{\prime}m^{\prime}lm}\langle l^{\prime}m^{% \prime}|(\hat{\bm{\varepsilon}}_{\text{IR}}\cdot\hat{\bm{r}})^{2}|lm\rangle d_% {\text{f}l^{\prime}m^{\prime},\text{i}}^{(1)*}(\kappa_{+})d_{\text{f}lm,\text{% i}}^{(1)}(\kappa_{-})≈ roman_arg ∑ start_POSTSUBSCRIPT italic_l start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_m start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_l italic_m end_POSTSUBSCRIPT ⟨ italic_l start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_m start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | ( over^ start_ARG bold_italic_ε end_ARG start_POSTSUBSCRIPT IR end_POSTSUBSCRIPT ⋅ over^ start_ARG bold_italic_r end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | italic_l italic_m ⟩ italic_d start_POSTSUBSCRIPT f italic_l start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_m start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) ∗ end_POSTSUPERSCRIPT ( italic_κ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ) italic_d start_POSTSUBSCRIPT f italic_l italic_m , i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_κ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT )
+argAκ+k(1)Aκk(1),superscriptsubscript𝐴subscript𝜅𝑘1superscriptsubscript𝐴subscript𝜅𝑘1\displaystyle+\arg A_{\kappa_{+}k}^{(1)*}A_{\kappa_{-}k}^{(1)}\,,+ roman_arg italic_A start_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) ∗ end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT , (36)

while for the upper mainband (MB>) one gets

ϕ2ωMB>superscriptsubscriptitalic-ϕ2𝜔subscriptMB\displaystyle\phi_{2\omega}^{\text{MB}_{>}}italic_ϕ start_POSTSUBSCRIPT 2 italic_ω end_POSTSUBSCRIPT start_POSTSUPERSCRIPT MB start_POSTSUBSCRIPT > end_POSTSUBSCRIPT end_POSTSUPERSCRIPT arglmlmlm|(𝜺^IR𝒓^)2|lmdflm,i(1)(κ+)dflm,i(1)(κ)absentsubscriptsuperscript𝑙superscript𝑚𝑙𝑚quantum-operator-productsuperscript𝑙superscript𝑚superscriptsubscript^𝜺IR^𝒓2𝑙𝑚superscriptsubscript𝑑fsuperscript𝑙superscript𝑚i1subscript𝜅superscriptsubscript𝑑f𝑙𝑚i1subscript𝜅\displaystyle\approx\arg\sum_{l^{\prime}m^{\prime}lm}\langle l^{\prime}m^{% \prime}|(\hat{\bm{\varepsilon}}_{\text{IR}}\cdot\hat{\bm{r}})^{2}|lm\rangle d_% {\text{f}l^{\prime}m^{\prime},\text{i}}^{(1)*}(\kappa_{+})d_{\text{f}lm,\text{% i}}^{(1)}(\kappa_{-})≈ roman_arg ∑ start_POSTSUBSCRIPT italic_l start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_m start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_l italic_m end_POSTSUBSCRIPT ⟨ italic_l start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_m start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | ( over^ start_ARG bold_italic_ε end_ARG start_POSTSUBSCRIPT IR end_POSTSUBSCRIPT ⋅ over^ start_ARG bold_italic_r end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | italic_l italic_m ⟩ italic_d start_POSTSUBSCRIPT f italic_l start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_m start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) ∗ end_POSTSUPERSCRIPT ( italic_κ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ) italic_d start_POSTSUBSCRIPT f italic_l italic_m , i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_κ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT )
+argAκ+k(1)Akκ(1),superscriptsubscript𝐴subscript𝜅𝑘1superscriptsubscript𝐴𝑘subscript𝜅1\displaystyle+\arg A_{\kappa_{+}k}^{(1)*}A_{k\kappa_{-}}^{(1)*}\,,+ roman_arg italic_A start_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) ∗ end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_k italic_κ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) ∗ end_POSTSUPERSCRIPT , (37)

which differs only in the continuum-continuum factors and these are identical, up to a sign factor, due to complex-conjugation properties of Eq. (17). In contrast, the phase of emission-integrated OSB1 results in

ϕ2ωOSB1superscriptsubscriptitalic-ϕ2𝜔OSB1\displaystyle\phi_{2\omega}^{\text{OSB1}}italic_ϕ start_POSTSUBSCRIPT 2 italic_ω end_POSTSUBSCRIPT start_POSTSUPERSCRIPT OSB1 end_POSTSUPERSCRIPT arglmlmlm|(𝜺^IR𝒓^)4|lmdflm,i(1)(κ+)dflm,i(1)(κ)absentsubscriptsuperscript𝑙superscript𝑚𝑙𝑚quantum-operator-productsuperscript𝑙superscript𝑚superscriptsubscript^𝜺IR^𝒓4𝑙𝑚superscriptsubscript𝑑fsuperscript𝑙superscript𝑚i1subscript𝜅superscriptsubscript𝑑f𝑙𝑚i1subscript𝜅\displaystyle\approx\arg\sum_{l^{\prime}m^{\prime}lm}\langle l^{\prime}m^{% \prime}|(\hat{\bm{\varepsilon}}_{\text{IR}}\cdot\hat{\bm{r}})^{4}|lm\rangle d_% {\text{f}l^{\prime}m^{\prime},\text{i}}^{(1)*}(\kappa_{+})d_{\text{f}lm,\text{% i}}^{(1)}(\kappa_{-})≈ roman_arg ∑ start_POSTSUBSCRIPT italic_l start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_m start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_l italic_m end_POSTSUBSCRIPT ⟨ italic_l start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_m start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | ( over^ start_ARG bold_italic_ε end_ARG start_POSTSUBSCRIPT IR end_POSTSUBSCRIPT ⋅ over^ start_ARG bold_italic_r end_ARG ) start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT | italic_l italic_m ⟩ italic_d start_POSTSUBSCRIPT f italic_l start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_m start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) ∗ end_POSTSUPERSCRIPT ( italic_κ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ) italic_d start_POSTSUBSCRIPT f italic_l italic_m , i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_κ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT )
+argAκ+κ(1)Aκκ(1)Aκk(1)Aκk(1),superscriptsubscript𝐴subscript𝜅𝜅1superscriptsubscript𝐴𝜅subscript𝜅1superscriptsubscript𝐴subscript𝜅𝑘1superscriptsubscript𝐴subscript𝜅𝑘1\displaystyle+\arg A_{\kappa_{+}\kappa}^{(1)*}A_{\kappa\kappa_{-}}^{(1)*}A_{% \kappa_{-}k}^{(1)*}A_{\kappa_{-}k}^{(1)}\,,+ roman_arg italic_A start_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT italic_κ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) ∗ end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_κ italic_κ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) ∗ end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) ∗ end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT , (38)

which, at the level of Eq. (17), yields the same continuum-continuum phase (second term of Eq. (38)), again, because the last two A𝐴Aitalic_A-factors cancel in phase, but the one-photon part (first term of Eq. (38)) is different in Eqs. (36) or (37). In energy regions dominated by a single partial wave (d𝑑ditalic_d-wave in the present case around SB22) the partial wave mixing by the angular integral is unimportant and the delays obtained from SB (Eq. (37)) and OSB1 (Eq. (38)) are in agreement.

Only for the case of the lowest sideband SB14 and for both the axial case and the angle integrated case the neighboring inner and outer sidebands do not reproduce the delay given by the central sideband. This is caused by the strong energy dependence of the free-free transitions close to the threshold, while our asymptotic theory reproduces this effect accurately. The excellent agreement with our experiment confirms this effect.

IV Practical implementation

In practice, reconstruction of higher-order ionization amplitudes (at the leading order of perturbation) follows the scheme indicated in Fig. 3(c) and Eqs. (12)–(15). Specifically, in absence of permanent dipole moments and of near-resonant dipole transitions within the residual ion the application of the partial-wave asymptotic approach reduces to the following steps:

  • 1.

    The partial-wave-resolved one-photon XUV ionization amplitudes T~±,fi,lm(1):=2πidflm,i(1)(κ±)assignsuperscriptsubscript~𝑇plus-or-minusfi𝑙𝑚12𝜋𝑖superscriptsubscript𝑑f𝑙𝑚i1subscript𝜅plus-or-minus\tilde{T}_{\pm,\text{fi},lm}^{(1)}:=2\pi id_{\text{f}lm,\text{i}}^{(1)}(\kappa% _{\pm})over~ start_ARG italic_T end_ARG start_POSTSUBSCRIPT ± , fi , italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT := 2 italic_π italic_i italic_d start_POSTSUBSCRIPT f italic_l italic_m , i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_κ start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ) defined in Eq. (3) are calculated for intermediate photoelectron momenta κ±subscript𝜅plus-or-minus\kappa_{\pm}italic_κ start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT.

  • 2.

    All partial-wave amplitudes are multiplied by the angular coupling coefficient and by the continuum integral coefficient describing the free-free transition,

    T~±,fi,lm(n+1)=λμlm|𝜺^IR𝒓^|λμAκ±λkl(1)T~±,fi,λμ(n),superscriptsubscript~𝑇plus-or-minusfi𝑙𝑚𝑛1subscript𝜆𝜇quantum-operator-product𝑙𝑚subscript^𝜺IR^𝒓𝜆𝜇superscriptsubscript𝐴subscript𝜅plus-or-minus𝜆𝑘𝑙1superscriptsubscript~𝑇plus-or-minusfi𝜆𝜇𝑛\tilde{T}_{\pm,\text{fi},lm}^{(n+1)}=-\sum_{\lambda\mu}\langle lm|\hat{\bm{% \varepsilon}}_{\text{IR}}\cdot\hat{\bm{r}}|\lambda\mu\rangle A_{\kappa_{\pm}% \lambda kl}^{(1)}\tilde{T}_{\pm,\text{fi},\lambda\mu}^{(n)}\,,over~ start_ARG italic_T end_ARG start_POSTSUBSCRIPT ± , fi , italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n + 1 ) end_POSTSUPERSCRIPT = - ∑ start_POSTSUBSCRIPT italic_λ italic_μ end_POSTSUBSCRIPT ⟨ italic_l italic_m | over^ start_ARG bold_italic_ε end_ARG start_POSTSUBSCRIPT IR end_POSTSUBSCRIPT ⋅ over^ start_ARG bold_italic_r end_ARG | italic_λ italic_μ ⟩ italic_A start_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT italic_λ italic_k italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT over~ start_ARG italic_T end_ARG start_POSTSUBSCRIPT ± , fi , italic_λ italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT , (39)

    to produce amplitudes for the higher order at the photoelectron momentum k=(κ±2±2ω)1/2𝑘superscriptplus-or-minussuperscriptsubscript𝜅plus-or-minus22𝜔12k=(\kappa_{\pm}^{2}\pm 2\omega)^{1/2}italic_k = ( italic_κ start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ± 2 italic_ω ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT; cf. Eq. (C9) of [13]. For a given polarization 𝜺^IRsubscript^𝜺IR\hat{\bm{\varepsilon}}_{\text{IR}}over^ start_ARG bold_italic_ε end_ARG start_POSTSUBSCRIPT IR end_POSTSUBSCRIPT the angular integral in Eq. (39) can be expressed using Clebsch-Gordan coefficients. The radial integral Aκ±λkl(1)superscriptsubscript𝐴subscript𝜅plus-or-minus𝜆𝑘𝑙1A_{\kappa_{\pm}\lambda kl}^{(1)}italic_A start_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT italic_λ italic_k italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT is plotted in Fig. 4 and its precomputed values for l10𝑙10l\leq 10italic_l ≤ 10 and k2/2150superscript𝑘22150k^{2}/2\leq 150italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 ≤ 150 eV are available in the Supplementary Material. If needed, this step is repeated for each additional absorbed IR photon with a correspondingly increasing (or decreasing) final photoelectron momentum.

  • 3.

    The final multiphoton partial-wave amplitudes are multiplied by the combinatoric factor,

    T±,fi,lm(n)=1(n1)!T~±,fi,lm(n),superscriptsubscript𝑇plus-or-minusfi𝑙𝑚𝑛1𝑛1superscriptsubscript~𝑇plus-or-minusfi𝑙𝑚𝑛T_{\pm,\text{fi},lm}^{(n)}=\frac{1}{(n-1)!}\tilde{T}_{\pm,\text{fi},lm}^{(n)}\,,italic_T start_POSTSUBSCRIPT ± , fi , italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG ( italic_n - 1 ) ! end_ARG over~ start_ARG italic_T end_ARG start_POSTSUBSCRIPT ± , fi , italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT , (40)

    and used to evaluate the observables of interest, for instance the higher-order emission-integrated RABITT delay

    τR=12ωarglmT+,fi,lm(p)T,fi,lm(q).subscript𝜏𝑅12𝜔subscript𝑙𝑚superscriptsubscript𝑇fi𝑙𝑚𝑝superscriptsubscript𝑇fi𝑙𝑚𝑞\tau_{R}=\frac{1}{2\omega}\arg\sum_{lm}T_{+,\text{fi},lm}^{(p)*}T_{-,\text{fi}% ,lm}^{(q)}\,.italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 italic_ω end_ARG roman_arg ∑ start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT + , fi , italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_p ) ∗ end_POSTSUPERSCRIPT italic_T start_POSTSUBSCRIPT - , fi , italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_q ) end_POSTSUPERSCRIPT . (41)

If the residual ion has a permanent dipole or supports dipole transitions, the second step needs to be supplemented with analogical contribution from the ion-ion terms, Eqs. (15) and (23).

V conclusion

In this work, we provide a numerically cheap and accurate solution to take into account the dressing photons in a RABITT experiment. It builds on the well-accepted asymptotic approximation theory, but represents the interaction of the photoelectron with IR field separately for each partial wave. This allows us to take into account the different ionization phase acquired by distinct partial waves. Unequal phases from the IR absorption then affect reconstruction of the angularly resolved RABITT oscillation, where multiple partial waves interfere, as well as orientation-unresolved measurements around Cooper minima and similar interference structures. The method can be applied to the emerging RABITT variants where the dressing field corresponds to the energy difference between the harmonics [40, 41], the third [42] or the fourth [20].

We implemented a two-harmonic RABITT to isolate the interference pathways and observe the higher orders free of overlap between the contributions. This allows the disentaglement of various ionization pathways as well as investigation of individual multi-photon interferences in higher-order RABITT processes. The method is particularly promising for applications involving high density of electronic states such as in molecules. We have shown that even in the emission-integrated case the delays obtained from MBs are very accurately in phase opposition to the central SB. This is sometimes referred to as the ‘rule of thumb’ [18]. The OSBs are similar in this regard, but they are affected slightly differently by the integration over emission directions if multiple partial waves contribute comparably, which may affect the extracted delays close to Cooper minima and similar interference structures.

We have shown that the method gives valid results in argon even at very low energies, suggesting that it is particularly the effect of the angular momentum barrier that is the most lacking in the traditional asymptotic formalism of laser-assisted ionization. We also discussed the higher-order RABITT schemes, where a similar partial-wave asymptotic factorization can be performed for the higher-order photoionization amplitudes. This also gives, asymptotically, results that match the full leading-order perturbation theory amplitudes, but at low energies the accuracy of the approximate treatment is reduced. Our method is straightforward and provides similar results compared to time-dependent simulations and the time-independent leading-order perturbation method.

Finally, the method is compatible with any computational approach that provides one-photon ionization amplitudes in the partial-wave decomposition. The method is very accurate for settings where electron correlation in the continuum is negligible. When it is not, the photoelectron in the vicinity of the molecule exchanges energy with multiple ionic states and does not have a well-defined energy. Consequently, approximating the short-range part of the photoelectron wavefunction with a pure Coulomb wave in Eq. (16) is no longer justified. Such a situation may arise for example in case of the strongly dipole-coupled B and C states of CO+2superscriptsubscriptabsent2{}_{2}^{+}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT [13] and in the core-excited shape resonance in C state which couples to hundreds of ionic states [43, 44].

Acknowledgements.
This work has been supported by the Charles University Research Centre program No. UNCE/24/SCI/016. Computational resources were provided by the e-INFRA CZ project (ID:90254), supported by the Ministry of Education, Youth and Sports of the Czech Republic. Zdeněk Mašín and Jakub Benda acknowledge the support of the Czech Science Foundation (Grant No. 20-15548Y). We acknowledge support from the CNRS, ANR-21-CE30-0052 ‘FAUST’, the Fédération de recherche André Marie Ampère and the European COST Action AttoChem (CA18222). We acknowledge Evan Langloÿs for experimental support.

References