• Photonics Research
  • Vol. 12, Issue 10, 2409 (2024)
Jiahao Dong1, Liang Xu1,6,*, Yiqi Fang2, Hongcheng Ni3..., Feng He4,5, Songlin Zhuang1 and Yi Liu1,5,7,*|Show fewer author(s)
Author Affiliations
  • 1Shanghai Key Laboratory of Modern Optical System, University of Shanghai for Science and Technology, Shanghai 200093, China
  • 2Fachbereich Physik, Universität Konstanz, 78464 Konstanz, Germany
  • 3State Key Laboratory of Precision Spectroscopy, East China Normal University, Shanghai 200241, China
  • 4Key Laboratory for Laser Plasmas (Ministry of Education) and School of Physics and Astronomy, Shanghai Jiao Tong University, Shanghai 200240, China
  • 5CAS Center for Excellence in Ultra-intense Laser Science, Shanghai 201800, China
  • 6e-mail: liangxu2021@usst.edu.cn
  • 7e-mail: yi.liu@usst.edu.cn
  • show less
    DOI: 10.1364/PRJ.528051 Cite this Article Set citation alerts
    Jiahao Dong, Liang Xu, Yiqi Fang, Hongcheng Ni, Feng He, Songlin Zhuang, Yi Liu, "Scheme for generation of spatiotemporal optical vortex attosecond pulse trains," Photonics Res. 12, 2409 (2024) Copy Citation Text show less

    Abstract

    The realization of spatiotemporal vortex structure of various physical fields with transverse orbital angular momentum (OAM) has attracted much attention and is expected to expand the research scope and open new opportunities in their respective fields. Here we present theoretically the first, to the best of our knowledge, study on the generation of attosecond pulse trains featuring a spatiotemporal optical vortex (STOV) structure by a two-color femtosecond light field, with each color carrying transverse OAM. Through careful optimization of relative phase and intensity ratio, we validate the efficient upconversion of the infrared pulse into its tens of order harmonics, showing that each harmonic preserves a corresponding intact topological charge. This unique characteristic enables the synthesis of an extreme ultraviolet attosecond pulse train with transverse OAM. In addition, we reveal that ionization depletion plays an outsize role therein. Our studies pave the way for the generation and utilization of light fields with STOV in the attosecond regime.

    1. INTRODUCTION

    Attosecond pulses, as a milestone technique of attosecond science, have led to great advances in the research communities of ultrafast optics and strong-field physics in recent years [1,2]. Three scientists achieved the key experimental methods to produce attosecond light pulses for the study of electron dynamics in matter, thus being awarded the 2023 Nobel Prize in Physics [3]. Attosecond pulses originate from high harmonic generation (HHG) essentially, which is well understood by the semiclassical three-step model [4,5]. When atoms or molecules are irradiated by an intense laser field, the outer electron first tunnels through the distorted Coulomb potential dressed by the laser field, then is accelerated and driven back by the remaining field, and finally may recombine to the initial bound state and emit a high-energy photon simultaneously, realizing upconversion of the driving laser frequency to the extreme ultraviolet (EUV) region. Thus, attosecond pulses and HHG respectively describe the time- and frequency-domain information of this coherent emission. From the quantum perspective, the underlying mechanism can be well revealed by the Lewenstein model or solving the time-dependent Schrödinger equation (TDSE) numerically [4,6,7].

    Since its advent, a lot of effort has been paid to control and shape this EUV radiation. At the beginning, researchers explored the effective methods to overcome technical hurdles from attosecond pulse trains (APTs) [8] to an isolated attosecond pulse (IAP) [9,10] and tailor its conventional properties, such as cutoff frequency [11,12], duration [1315], brightness [1620], harmonic odevity [2125], and polarization [2628]. Up to now, the highest photon energy based on HHG has reached up to 5.2 keV [12] and the shortest IAP has a duration of about 50 as [14,15].

    As a type of electromagnetic wave, attosecond pulses not only carry spin angular momentum, but also possess orbital angular momentum (OAM), which is structured light fields, called optical vortex [29]. Its Poynting vector circulates around a local axis parallel to the propagation direction, forming longitudinal OAM, which has potential technological applications in optical communication, chirality recognition, and micromanipulation [3033]. Then the manipulation of attosecond pulses entered the era of “vortex.” Attosecond optical vortices with longitudinal OAM from HHG have been reported [34,35] and gained much exciting progress in controlling the optical spin-orbit coupling [36] and generating EUV beams with time-varying OAM [37]. Just recently, free-space optical skyrmions based on HHG have been demonstrated theoretically [38].

    Lately, a newly discovered spatiotemporal vortex with phase and energy circulation in a space-time plane was demonstrated experimentally [39], which was theoretically proposed in 2012 [40]. It offers a new degree of freedom, transverse OAM, and has generated important advances in optics, terahertz, and acoustics fields [4148]. As a new form of structured light, its property of topological protection has shown promising application value in optical communication [49,50]. Meanwhile, its nonlinear effects and relativistic spatiotemporal optical vortex (STOV) HHG in plasma have recently gained increasing attention [51,52]. Significantly, Fang et al. turned their attention to HHG transverse OAMs and attempted to control them [53]. However, to the best of our knowledge, STOV attosecond pulses have not been explored yet, which would have a promising perspective for both attosecond science and extreme optics.

    Here, we demonstrated a theoretical scheme for the generation of a STOV APT from atom gas pumped by the two-color femtosecond light field composed of the fundamental wave (FW) and its third harmonic (TH), with both carrying a STOV structure. Different from a single-color excitation, we found that the employment of a weak TH field with optimized phase is crucial for upconversion of STOV from FW to the high-order harmonics. It was revealed that this optimized method can effectively suppress the contribution of the electron long trajectories to the harmonics generation and result in broadband harmonics carrying the STOV feature. With a properly truncated HHG spectrum, an APT carrying spatial-temporal OAM was achieved. Further studies revealed that the intensity ratio of the two pumping fields, which may lead to ionization depletion, also plays an important role for the robust formation of the STOV APT.

    2. THEORETICAL METHODS

    A linearly-polarized FW STOV pulse is described as [41] E(X,Y,Z,ωFWt)=E0N0(X2/wX2+ξ2/wξ2)|lFW|/2e(X2/wX2+Y2/wY2+ξ2/wξ2)ei(ωFWtkFWZlφf).As shown in Fig. 1, the laser electric field polarizes along the X axis and propagates along the Z axis. ξ=ctZ is the traveling longitudinal distance, and c is the light velocity. E0 is the peak amplitude, N0 is a normalization factor, ωFW is the angular frequency, and kFW denotes the wave number. wξ and wX refer to the pulse spatial widths, and φf=arctan(Xwξ/ξwX) is the azimuthal angle in the space-time plane with a topological charge lFW. Its Y component is trivial with a symmetrical Gaussian distribution. Hence, we focus only on electron dynamics in the XZ plane at Y=0 below. Since the range of electron motion contributing to HHG is much smaller than the spatial size of the laser focal point and wavelengths, a discrete spatial electric dipole approximation is adopted reasonably, where atoms at different X positions will experience different phase and intensity of the STOV field.

    Sketch of the discrete spatial electric dipole approximation. A STOV pumping pulse irradiates a blob of hydrogen atoms, and atoms at different X positions experience different phase and intensity of the driving field, then generating STOV high harmonics. (X,Y,Z) denotes the space coordinate system. (x,y,z) is the local atomic coordinate system, whose origin is located at (X,Y=0,Z=0).

    Figure 1.Sketch of the discrete spatial electric dipole approximation. A STOV pumping pulse irradiates a blob of hydrogen atoms, and atoms at different X positions experience different phase and intensity of the driving field, then generating STOV high harmonics. (X,Y,Z) denotes the space coordinate system. (x,y,z) is the local atomic coordinate system, whose origin is located at (X,Y=0,Z=0).

    Here, a one-dimensional TDSE describes the strong laser–atom interaction, whose velocity gauge is expressed as (atomic units are used throughout unless otherwise stated) itΨ(X,x,t)=[22+V0(x)iA(X,t)]Ψ(X,x,t),where V0(x)=1/x2+1.92 is a reduced potential energy for the hydrogen atom and x is the electron dynamic parameter. Such a reduced model is able to capture the main picture qualitatively since only free diffusion happens in the other two dimensions. The laser field consisting of FW and TH pulses with the same spatiotemporal envelope in the local atomic coordinate system is written as E(X,t)=E(X,ωFWt)+E(X,ωTHt+ϕ)=N0(X2/wX2+ξ2/wξ2)|lFW|/2e(X2/wX2+ξ2/wξ2)×[E0FWei(ωFWtlFWφf)+E0THei(ωTHtlTHφf+ϕ)],where ωTH=3ωFW, ϕ is the phase difference between both pulses, and A(X,t)=0tE(X,t)dt is the laser vector. To get APT, we first calculate the dipole acceleration by the Ehrenfest theorem [54]: a(X,t)=Ψ(X,x,t)|V0(x)xE(X,t)|Ψ(X,x,t).The TDSE is solved by the split-operator algorithm [55]. The initial state is numerically obtained via the imaginary time propagation [56]. A mask function cos1/6 is utilized in borders to prevent the unphysical reflections from boundaries [57].

    Then the X-position-dependent HHG spectra are obtained by the Fourier transform of a(X,t): S˜(X,ω)=|a˜(X,ω)|2=|a(X,t)eiωtdt|2.Finally, a pulse envelope in time domain can be synthesized via P(X,t)=|ω1ω2a˜(X,ω)eiωtdω|2,where ω1 and ω2 define the truncated frequency range. Equation (6) includes time-space dimension and thus gives the spatial-temporal structure of HHG and APT. Here HHG emission from atoms with the same X values is assumed to be perfectly in phase, and the HHG interference from different X-position atoms as well as the propagation effect is not considered.

    The simulation parameters are λFW=1600  nm, IFW=1×1014  W/cm2, λTH=533.3  nm, and ITH as well as the phase difference ϕ between both being variable. The topological charge is fixed as lFW=1 and lTH=3, respectively. The time and spatial steps are Δt=0.1 a.u. and Δx=0.2 a.u., respectively. The simulation box for the range of electronic motion is x=[1000,  1000] a.u., which is big enough to preserve the full HHG information. The transverse range of the laser focal point is X=[12λFW,12λFW] with 401 data points. The convergence has been tested by using a larger transverse box for the laser beam and finer spatial and time grids for the electron dynamics while almost identical harmonic spectra are obtained.

    3. RESULTS AND DISCUSSION

    A. Two-Color-Laser Scheme for Effectively Generating STOV APT

    To achieve a STOV APT, we need to produce HHG spectra extended to the EUV region with well-defined transverse OAMs. The first idea is adopting the FW STOV pulse with λFW=1600  nm as a pumping laser. The result is shown in Fig. 2(a1). At first glance, for any X position, odd and even harmonics are not well separated, which is very different from the common HHG spectra. In addition, there is no position-dependent frequency chirp for each harmonic. For example, the 35th harmonic pulse amplitude and phase distributions in the space-time plane are presented in Figs. 2(a2) and 2(a3). There is an indistinct intensity distribution and an irregular phase variation. Thus, harmonics in these HHG spectra do not possess STOV structure. We have systematically tested different pump laser wavelengths at various laser intensities ranging from 5×1013 to 1.25×1014  W/cm2 and always cannot find the possibility for generation of broadband harmonics carrying a clear STOV structure in the plateau region. We therefore concluded that harmonics pumped with the fundamental field carrying STOV cannot possess STOV structure.

    (1st column) Spatial-resolved HHG spectra driven by STOV pulses and retrieved spatiotemporal (2nd column) intensity and (3rd column) phase distributions of the 35th harmonic. (a) Only FW field with λFW=1600 nm; two-color STOV pulses with the phase difference (b) ϕ=0 and (c) ϕ=1.1π between FW and TH fields, respectively. The FW intensity is fixed as IFW=1×1014 W/cm2, and the TH intensity is ITH=9×1012 W/cm2. The black dashed line represents the ionization potential of H atom. o.c. is the FW period throughout.

    Figure 2.(1st column) Spatial-resolved HHG spectra driven by STOV pulses and retrieved spatiotemporal (2nd column) intensity and (3rd column) phase distributions of the 35th harmonic. (a) Only FW field with λFW=1600  nm; two-color STOV pulses with the phase difference (b) ϕ=0 and (c) ϕ=1.1π between FW and TH fields, respectively. The FW intensity is fixed as IFW=1×1014  W/cm2, and the TH intensity is ITH=9×1012  W/cm2. The black dashed line represents the ionization potential of H atom. o.c. is the FW period throughout.

    Diagrams of the time-frequency analyses of the HHG emission (X=0) generated by (a) single STOV pulse and (b), (c) two-color STOV pulses. (a)–(c) correspond to Figs. 2(a1)–2(c1), respectively.

    Figure 3.Diagrams of the time-frequency analyses of the HHG emission (X=0) generated by (a) single STOV pulse and (b), (c) two-color STOV pulses. (a)–(c) correspond to Figs. 2(a1)–2(c1), respectively.

    Inspired by the previous works on harmonic control with multiple color laser fields [18,53], we then added a TH laser pulse with ITH=9×1012  W/cm2 into the fundamental pump field. The new HHG spectra are shown in Figs. 2(b1) and 2(c1), corresponding to the relative phase difference ϕ=0 and 1.1π, respectively. As one can see, Fig. 2(b1) is similar to Fig. 2(a1), and no STOV harmonics are produced. In contrast, in Fig. 2(c1) each harmonic shows an unambiguous space-dependent frequency chirp. Furthermore, this chirp scales linearly with the order of the harmonics, and the number of interference dark fringes in each harmonic is equal to its order. Let us take the 35th harmonic as an example. Its intensity profile and phase distribution in the spatial-temporal domain are presented in Figs. 2(c2) and 2(c3). Its intensity shows a clear doughnutlike structure, while its phase manifests as a phase winding of 35×2π. We have confirmed that the same analysis holds for each harmonic in Fig. 2(c1). Therefore, it can be concluded that every harmonic in Fig. 2(c1) is a STOV light field with a topological charge ln=nl, where n is the harmonic order and l is the topological charge of the FW laser. We speculated that the microscopic mechanism responsible for the structures in Fig. 2(c) is that the frequency and X position are highly related in the nonlinear interaction between the STOV laser fields and atoms, making the FW transverse OAM l transfer into the nth-order high harmonic ones ln by n times. In other words, since the nth-order harmonic corresponds to the nth-order nonlinear effect, the complex exponent ei(ωFWtlFWϕf) recording the local phase information of the FW STOV field is transmitted to the nth-order harmonic after the nth-power execution, namely, ein(ωFWtlFWϕf).

    B. Revealing the Generating Mechanism

    Why does the two-color STOV field scheme with a definite phase difference ϕ=1.1π in Fig. 2(c) generate EUV high harmonics with clear STOV structures? To answer this question, we resort to the Gabor transform to trace the HHG bursts in the temporal domain [58]. Here we examine the representative space position X=0 case, and the results are shown in Fig. 3, where Figs. 3(a), 3(b), and 3(c), respectively, correspond to Figs. 2(a1), 2(b1), and 2(c1). As expected, the phase singularity of the driving field leads to a middle valley at t=0 in three panels. We first concentrate on Figs. 3(a) and 3(b). During per half cycle of the FW field, the long and short electron trajectories make nearly equivalent contributions to the HHG. Besides these common long and short trajectories, under careful observation, there are complex bursts in the lower energy region except for the first half cycle in Figs. 3(a) and 3(b), which complicates the time and frequency structure of HHG and leads to almost continuous HHG spectra in Figs. 2(a1) and 2(b1). In addition, there noticeably exists strong low-order HHG burst around t=0, where the laser field is almost zero. In contrast, Fig. 3(c) shows remarkably that the bursts induced by long trajectories almost disappear and only short trajectories dominate. Meanwhile, the complex lower energy bursts in Figs. 3(a) and 3(b) also vanish in Fig. 3(c). So, the X-dependent HHG spectra are discrete and each harmonic shows a certain STOV structure. Based on the above results, we speculate that the multiple scattering of the long electron trajectories and its suppression may be a critical factor of the formation of harmonics with clear STOV structure.

    To further reveal the physical origin behind Fig. 3, we trace electron classical trajectories via Newton’s equation in adiabatic approximation. The results are shown in Fig. 4, and other simulation details are included in Appendix A. The black dashed line describes the synthesized laser electric field. The curve color denotes the number of electron returns, and the thickness of the curve corresponds to the ionization rate with logarithmic scale. Both Figs. 4(a) and 4(b) show that the thicker lines possess more returns, which implies that the electron long trajectories born around the laser field peak contribute to the multiple scatterings responsible for low-energy structure and smear the HHG bursts in Figs. 3(a) and 3(b), leading to complex HHG spectra in Figs. 2(a1) and 2(b1). However, Fig. 4(c) shows that the additional weak TH field with a proper phase makes the combined field split into two tips. This synthesized field can restrain long trajectories with multiple scatterings and is conducive to the generation of EUV STOV, which can be understood well based on the three-step model. Ionization is solely dependent on the absolute value of the electric field, and the phase information of the STOV field is primarily imprinted in electron trajectories. The HHG plateau results from the contribution of not only first-returning trajectories but also higher-order scattering ones. The phase the latter accumulates does not synchronize with the former, thus resulting in the continuous HHG spectra. The earliest HHG bursts exclusively originate from the first return, while subsequent bursts are contributed considerably by multiple scattering channels with disordered phases. This explains why Figs. 2(a3) and 2(b3) gradually display irregular phase structures from left to right. Therefore, in order to obtain a clear STOV structure, it is essential to suppress the contribution of multiple scatterings.

    Classical electron trajectories driven by laser fields at X=0. (a)–(c) correspond to Figs. 2(a1)–2(c1), respectively. The blue, green, yellow, and red curves represent an electron trajectory with one, two, three, and four or more returns, respectively. The line thickness represents the relative magnitude of the ionization rate after logarithmic treatment. Note that only trajectories where electrons are able to return and the ionization rate is greater than 10% of the maximum ionization rate are shown, and about half of the laser field is shown by the black dashed line.

    Figure 4.Classical electron trajectories driven by laser fields at X=0. (a)–(c) correspond to Figs. 2(a1)–2(c1), respectively. The blue, green, yellow, and red curves represent an electron trajectory with one, two, three, and four or more returns, respectively. The line thickness represents the relative magnitude of the ionization rate after logarithmic treatment. Note that only trajectories where electrons are able to return and the ionization rate is greater than 10% of the maximum ionization rate are shown, and about half of the laser field is shown by the black dashed line.

    Since the relevant HHG spectra have been obtained, an APT can be retrieved by an inverse Fourier transform, as shown in Fig. 5. The spatiotemporal intensity distributions of APT in Figs. 5(a), 5(b), and 5(c), respectively, correspond to Figs. 2(a1), 2(b1), and 2(c1). The truncated frequency range in three panels is from 11ωFW to 39ωFW. As expected, there are no intensity zero points in central position in Figs. 5(a) and 5(b) since their HHG spectra have no STOV structure. In other words, spatiotemporal vortex APT cannot be produced in Figs. 5(a) and 5(b). Importantly, Fig. 5(c) gives a perfect hollow pulse train with an average topological charge l=19.5 at the spatiotemporal plane calculated by [53,59,60] l=2Im(ctEX*EX/X)dtdXEX*EXdtdX,where EX is the electric field intensity of the synthetic field. APT subpulse intensity distribution is extracted in Fig. 5(d). Its full width at half maximum (FWHM) indeed has reached attosecond scale, about 613 as.

    Retrieved spatiotemporal intensity distributions of APT in (a), (b), and (c) correspond to Figs. 2(a1), 2(b1), and 2(c1), respectively. (d) Temporal intensity profiles of a single attosecond pulse sampled from the black rectangle in (c). The red dotted curve represents a Gaussian fitting with an FWHM of 613 as.

    Figure 5.Retrieved spatiotemporal intensity distributions of APT in (a), (b), and (c) correspond to Figs. 2(a1), 2(b1), and 2(c1), respectively. (d) Temporal intensity profiles of a single attosecond pulse sampled from the black rectangle in (c). The red dotted curve represents a Gaussian fitting with an FWHM of 613 as.

    Since the X>0 part of the driving STOV pulse with lFW=1 has one more optical cycle than the X<0 part, APT shown in Fig. 5(c) exhibits a fork-shaped structure with a fork number N=2, where the pulse number in X>0 is two more than in X<0. The unique structure is also attributed to the interference of different harmonics with a specific topological charge interval Δl in time and space (for more explanations, see Appendix B). The aforementioned results manifest that our two-color strategy is an effective means of generating spatiotemporal vortex APT.

    C. Effect of Ionization Depletion

    We continue to explore spatially-resolved HHG spectrum by increasing the TH intensity. Figures 6(a1) and 6(a2) give the typical 35th harmonic HHG spectra with two laser intensities ITH=9×1012  W/cm2 and ITH=4.9×1013  W/cm2. Comparing the two panels, we find that a stronger TH laser significantly smears the fine interference patterns and thus destroys its due topological charge. As is known, the interference fringes are determined by the phase difference and amplitude ratio of subwaves. Since the phase difference between the FW and TH lasers remains unchanged in Figs. 6(a1) and 6(a2), the interference amplitude ratio, i.e., ionization probability, should undergo variation remarkably. So we trace the time-dependent ionization probability for Figs. 6(a1) and 6(a2), as shown in Fig. 6(b). The inset in Fig. 6(b) illustrates a STOV pulse traveling from left to right divided into two segments (Part A and Part B) in the time domain due to a phase singularity. When considering a ratio of ITH/IFW=0.09, the ion yield for both Part A and Part B is comparable, about 0.2. However, for the ITH/IFW=0.49 case, the ion generation probability for Part A is four times larger than that for Part B, resulting in a significant decrease in contrast of the interference fringes.

    Spatially-resolved 35th harmonic spectra, ITH=9×1012 W/cm2 for (a1) and ITH=4.9×1013 W/cm2 for (a2). (b) Time-dependent ionization probability for (a) calculated via TDSE. (c1) and (c2) are reconstructed HHG spectra by the quantum-orbit model. The intensity ratios ITH/IFW are taken to be 0.09 and 0.49, respectively. IFW=1×1014 W/cm2 is fixed in all panels. The X=0 position is considered in (b) and (c).

    Figure 6.Spatially-resolved 35th harmonic spectra, ITH=9×1012  W/cm2 for (a1) and ITH=4.9×1013  W/cm2 for (a2). (b) Time-dependent ionization probability for (a) calculated via TDSE. (c1) and (c2) are reconstructed HHG spectra by the quantum-orbit model. The intensity ratios ITH/IFW are taken to be 0.09 and 0.49, respectively. IFW=1×1014  W/cm2 is fixed in all panels. The X=0 position is considered in (b) and (c).

    In order to confirm our deduction further, we calculate the HHG spectrum based on the quantum-orbit model in the strong-field approximation framework [53] (for more details, see Appendix C), as shown in Figs. 6(c1) and 6(c2). As one sees, Figs. 6(a1) and 6(a2) are well reproduced in Figs. 6(c1) and 6(c2), respectively. The above results suggest that intensity is a key parameter in generating STOV APT and ionization depletion may hinder HHG STOV formation because the target system cannot experience the intact STOV information of the driving laser.

    4. CONCLUSION

    We theoretically and numerically show that a spatiotemporal vortex APT can be produced from atom gas excited by a two-color femtosecond laser pulse with STOV. Our analyses revealed that the synthesized pump laser field can be optimized by the relative phase and intensity ratio, so that FW STOV information can be robustly transferred to each of the high-order harmonics by suppressing electron long trajectories, achieving an APT carrying an averaged topological charge of 19.5, with each sub-cycle optical burst 613 as.

    The transverse OAM of light opens up an entirely new dimension, empowering us to shape light pulses spatiotemporally [61]. With the rapid development of generating intense STOV laser pulses and characterizing STOV structure [62], our scheme will be realized experimentally in the coming future. Analogous to the streaking method [63], it is well expected that superposing the STOV APT onto an intense long-wavelength Gaussian or STOV pulse will help us gain a profound understanding of the physical mechanisms underlying the interaction between matter and the transverse OAM of light [64,65]. Last but not least, our research sheds light on how STOV pulses influence electron trajectories and light emission during HHG processes, laying the groundwork for the generation of IAP with transverse OAM and bridging the two communities of STOV and attosecond science.

    Acknowledgment

    Acknowledgment. Authors are grateful to Qiwen Zhan for the helpful and enlightening discussions.

    APPENDIX A: DETAILS FOR SEMI-CLASSICAL ELECTRON TRAJECTORIES SIMULATIONS

    To focus on the electron trajectories that contribute most to HHG spectrum, we first employ the Ammosov–Delone–Krainov (ADK) model, based on adiabatic approximation [66], to calculate the time-dependent ionization rate of hydrogen atom induced by different driving fields: AADK(t)=4|E(t)|exp[2|3E(t)|].

    Because our numerical model adopts the one-dimensional time-dependent Schrödinger equation (1D TDSE), it actually overestimates the ionization rate compared with the three-dimensional case. This means that if we apply the laser parameters in the main text directly to the ADK model, very few electrons will be ionized. Therefore, we adjust the Eq. (A1) by multiplying a coefficient K to make the calculated ionization probability comparable to the 1D TDSE. The coefficient K=9.0, 5.0, and 18.3 for only the FW, the two-color strategy with phase difference ϕ=0, and ϕ=1.1π, respectively.

    Next, we track ionized electron dynamics by using Newtonian mechanics. The electron initial velocity at the tunnel exit is set as zero. The number of times that an ionized electron returns to the nucleus is determined by counting the number of times that its displacement curve crosses x=0.

    APPENDIX B: FORK STRUCTURE OF STOV APT

    Here we assume the harmonics involved in synthesizing APT have the same spatiotemporal envelope Ast and intensity. Then the complex amplitude of the harmonic can be expressed as Ej=Astexp[i(ωjtljφf)], where φf is the azimuthal angle in the Xt plane. The intensity distribution of the APT obtained by synthesizing all STOV harmonics can be expressed as I=|j=1qEj|2=Ast2{q+2n=1q1m=n+1qcos[(ωmωn)t(lmln)φf]}.

    In order to match the conditions in the main text, we assume that the frequency difference between adjacent harmonics is ωj+1ωj=2ωFW and the topological charge difference is lj+1lj=2. Thus, the intensity distribution is I=qAst2+2Ast2{(q1)cos(2ωFWt2φf)+(q2)cos[2(2ωFWt2φf)]++2cos[(q2)(2ωFWt2φf)]+cos[(q1)(2ωFWt2φf)]}.We define Ω=2ωFWt2φf and rewrite Eq. (B2) as I=qAst2+2Ast2{(q1)cosΩ+(q2)cos2Ω++2cos[(q2)Ω]+cos[(q1)Ω]}.

    As seen in Eq. (B3), the intensity distribution of the synthesized field consists of a DC component and q1 cosine components, which oscillate at integer multiples of Ω. It is obvious that Eq. (B3) has a local maximum value only when the phase of all the cosine components is uniformly zero. In addition, the coefficient of the cosine component decreases as its oscillation frequency increases. Therefore, the intensity distribution characteristics of the APT mainly depend on the cosine component with the lowest frequency, i.e., cos(2ωFWt2φf), which oscillates at frequency of 2ωFW and possesses a fork-shaped structure with a fork number N=2 in the Xt plane, as shown in Fig. 7. As one sees, the intensity distribution in Fig. 4(c) in the main text is consistent well with those shown in Fig. 7. The latter is analogous to the result of a vortex beam with a longitudinal topological charge l=2 interfering with a plane wave.

    Distribution of cos(2ωt−2φf) in the X–t plane, where φf is the azimuthal angle in the space-time plane.

    Figure 7.Distribution of cos(2ωt2φf) in the Xt plane, where φf is the azimuthal angle in the space-time plane.

    In other words, the fork-shaped structure of the APT depends on the topological charge difference between adjacent harmonics, i.e., N=Δl=lj+1lj. For example, when the topological charges of all harmonics are l=1, the APT will no longer have a fork-shaped structure (namely, the fork number is zero). In this case, each attosecond pulse within the APT is parallel to the XY plane, with a region of almost zero intensity near X=0 and t=0.

    APPENDIX C: SFA QUANTUM-ORBIT MODEL BASED ON IMPROVED THREE-STEP MODEL

    Here, we refer to the quantum-orbit model developed by Fang et al. [53] and extend it to analyze the influence of the ground state depletion on the structure of STOV harmonic spectrum. Based on the three-step model described by strong-field approximation (SFA) [67], the dipole moment of HHG consists of the contributions of ionization (Aion), propagation (Aprop), and recombination (Arec), namely, D(ω)+AionApropAreceiωtdt.

    In Figs. 3 and 4 in the main text, we have demonstrated that electron multiple collisions are suppressed significantly by the optimized field. This means that we only need to consider the contribution of the first return of electrons, which is similar to the situation near the cut-off region discussed by Fang et al. [53]. However, the HHG spectra we discussed are in the plateau region, and the electron trajectories considered are not limited to those that ionize at some particular time. Therefore, we calculate the real first return time tR corresponding to any given real time of birth tB using the improved three-step model [67]. In the model, the time of birth is defined as the moment when the instantaneous velocity of an electron ionized with canonical momentum ps is zero for the first time, i.e., k(tB)=ps+A(tB)=0, where k is the momentum of the electron and A(t)=tE(t)dt is the vector potential of the laser field, which corresponds to the classical case where electrons are ionized at zero initial velocity. The time of birth tB is considered to be the starting point of the electron trajectory, and the return time tR is determined by solving tBtR[A(τ)A(tB)]dτ=0.Thanks to the improved three-step model, we can achieve relatively accurate results while circumventing the complex ionization time and return time discussed in saddle point approximation [67].

    Then, we use Eq. (A1) to consider the ionization rate and ground state depletion for the H atom. Similar to the discussion above, since 1D TDSE overestimates the ionization rate, we multiply the ionization rate by a factor K to correct it. Therefore, the contribution of ionization is Aion(tB)=KAADK(tB)exp[tBKAADK(t)dt],where K=18.3 and 6.3 for ITH=9×1012  W/cm2 (namely balanceable ionization) and ITH=4.9×1013  W/cm2 (namely excessive ionization), respectively.

    The electron propagation process considering the quantum diffusion effect can be expressed as Aprop(tB,tR)=(2πitRtBiε)32exp[iS(tB,tR)],where ε is an arbitrary small positive constant and the phase factor S(tB,tR)=tBtR{12[A(t)A(tB)]2+Ip}dt is the quasi-classical action.

    The contribution of recombination is calculated by Arec(tR)=d*[p(tR)],where d(p)={i[5/(4πIp)]3/45p/(4Ip)}exp[5p2/(8Ip)] is the dipole matrix element for the transition from ground state to free state, approximated by a Gaussian form, Ip is the ionization potential, and p(tR)=A(tR)A(tB) is the return momentum of electron.

    Finally, substituting Eqs. (C3)–(C5) into Eq. (C1), the expression for the amplitude of each quantum orbit at X=0 is analytically given by D(ω)Σ(tB,tR)(2πitRtBiε)32KAADK(tB)exp[tBKAADK(t)dt]×exp[iS(tB,tR)+iωt]d*[p(tR)].In Eq. (C6), the harmonic dipole is considered to be the result of a weighted summation of the contributions of all trajectories based on the ionization rate and ground state depletion.

    References

    [1] P. Agostini, L. F. DiMauro. The physics of attosecond light pulses. Rep. Prog. Phys., 67, 813-855(2004).

    [2] F. Krausz, M. Ivanov. Attosecond physics. Rev. Mod. Phys., 81, 163-234(2009).

    [3] https://www.nobelprize.org/prizes/physics. https://www.nobelprize.org/prizes/physics

    [4] J. L. Krause, K. J. Schafer, K. C. Kulander. High-order harmonic generation from atoms and ions in the high intensity regime. Phys. Rev. Lett., 68, 3535-3538(1992).

    [5] P. B. Corkum. Plasma perspective on strong field multiphoton ionization. Phys. Rev. Lett., 71, 1994-1997(1993).

    [6] M. Lewenstein, P. Balcou, M. Y. Ivanov. Theory of high-harmonic generation by low-frequency laser fields. Phys. Rev. A, 49, 2117-2132(1994).

    [7] J. H. Eberly, Q. Su, J. Javanainen. Nonlinear light scattering accompanying multiphoton ionization. Phys. Rev. Lett., 62, 881-884(1989).

    [8] P. M. Paul, E. S. Toma, P. Breger. Observation of a train of attosecond pulses from high harmonic generation. Science, 292, 1689-1692(2001).

    [9] M. Hentschel, R. Kienberger, C. Spielmann. Attosecond metrology. Nature, 414, 509-513(2001).

    [10] M. Chini, K. Zhao, Z. Chang. The generation, characterization and applications of broadband isolated attosecond pulses. Nat. Photonics, 8, 178-186(2014).

    [11] T. Popmintchev, M.-C. Chen, D. Popmintchev. Bright coherent ultrahigh harmonics in the keV X-ray regime from mid-infrared femtosecond lasers. Science, 336, 1287-1291(2012).

    [12] J. Gao, J. Wu, Z. Lou. High-order harmonic generation in an X-ray range from laser-induced multivalent ions of noble gas. Optica, 9, 1003-1008(2022).

    [13] G. Sansone, E. Benedetti, F. Calegari. Isolated single-cycle attosecond pulses. Science, 314, 443-446(2006).

    [14] J. Li, X. Ren, Y. Yin. 53-attosecond X-ray pulses reach the carbon K-edge. Nat. Commun., 8, 186(2017).

    [15] T. Gaumnitz, A. Jain, Y. Pertot. Streaking of 43-attosecond soft-X-ray pulses generated by a passively CEP-stable mid-infrared driver. Opt. Express, 25, 27506-27518(2017).

    [16] S. Kazamias, D. Douillet, F. Weihe. Global optimization of high harmonic generation. Phys. Rev. Lett., 90, 193901(2003).

    [17] E. J. Takahashi, T. Kanai, K. L. Ishikawa. Coherent water window X-ray by phase-matched high-order harmonic generation in neutral media. Phys. Rev. Lett., 101, 253901(2008).

    [18] C. Jin, G. Wang, H. Wei. Waveforms for optimal sub-keV high-order harmonics with synthesized two- or three-colour laser fields. Nat. Commun., 5, 4003(2014).

    [19] B. Xue, Y. Tamaru, Y. Fu. Fully stabilized multi-TW optical waveform synthesizer: toward gigawatt isolated attosecond pulses. Sci. Adv., 6, eaay2802(2020).

    [20] Y. Fu, K. Nishimura, R. Shao. High efficiency ultrafast water window harmonic generation for single-shot soft X-ray spectroscopy. Commun. Phys., 3, 92(2020).

    [21] P. M. Kraus, A. Rupenyan, H. J. Wörner. High-harmonic spectroscopy of oriented OCS molecules: emission of even and odd harmonics. Phys. Rev. Lett., 109, 233903(2012).

    [22] E. Frumker, C. T. Hebeisen, N. Kajumba. Oriented rotational wave-packet dynamics studies via high harmonic generation. Phys. Rev. Lett., 109, 113901(2012).

    [23] D. Shafir, Y. Mairesse, D. M. Villeneuve. Atomic wavefunctions probed through strong-field light-matter interaction. Nat. Phys., 5, 412-416(2009).

    [24] H. Niikura, N. Dudovich, D. M. Villeneuve. Mapping molecular orbital symmetry on high-order harmonic generation spectrum using two-color laser fields. Phys. Rev. Lett., 105, 053003(2010).

    [25] H. Hu, N. Li, P. Liu. Pure even harmonic generation from oriented CO in linearly polarized laser fields. Phys. Rev. Lett., 119, 173201(2017).

    [26] O. Kfir, P. Grychtol, E. Turgut. Generation of bright phase-matched circularly-polarized extreme ultraviolet high harmonics. Nat. Photonics, 9, 99-105(2015).

    [27] P.-C. Huang, C. Hernández-García, J.-T. Huang. Polarization control of isolated high-harmonic pulses. Nat. Photonics, 12, 349-354(2018).

    [28] W. Li, F. He. Generation and application of a polarization-skewed attosecond pulse train. Phys. Rev. A, 104, 063114(2021).

    [29] L. Allen, M. W. Beijersbergen, R. J. C. Spreeuw. Orbital angular momentum of light and the transformation of Laguerre-Gaussian laser modes. Phys. Rev. A, 45, 8185-8189(1992).

    [30] Y. Shen, X. Wang, Z. Xie. Optical vortices 30 years on: OAM manipulation from topological charge to multiple singularities. Light Sci. Appl., 8, 90(2019).

    [31] N. Bozinovic, Y. Yue, Y. Ren. Terabit-scale orbital angular momentum mode division multiplexing in fibers. Science, 340, 1545-1548(2013).

    [32] G. C. G. Berkhout, M. W. Beijersbergen. Method for probing the orbital angular momentum of optical vortices in electromagnetic waves from astronomical objects. Phys. Rev. Lett., 101, 100801(2008).

    [33] J. Gao, X. Zhang, Y. Wang. Structured air lasing of N2+. Commun. Phys., 6, 97(2023).

    [34] M. Zürch, C. Kern, P. Hansinger. Strong-field physics with singular light beams. Nat. Phys., 8, 743-746(2012).

    [35] C. Hernández-García, A. Picón, J. San Román. Attosecond extreme ultraviolet vortices from high-order harmonic generation. Phys. Rev. Lett., 111, 083602(2013).

    [36] K. M. Dorney, L. Rego, N. J. Brooks. Controlling the polarization and vortex charge of attosecond high-harmonic beams via simultaneous spin-orbit momentum conservation. Nat. Photonics, 13, 123-130(2019).

    [37] L. Rego, K. M. Dorney, N. J. Brooks. Generation of extreme-ultraviolet beams with time-varying orbital angular momentum. Science, 364, eaaw9486(2019).

    [38] Y. Fang, Y. Liu. Generation and control of extreme ultraviolet free-space optical skyrmions with high harmonic generation. Adv. Photonics Nexus, 2, 046009(2023).

    [39] S. W. Hancock, S. Zahedpour, A. Goffin. Free-space propagation of spatiotemporal optical vortices. Optica, 6, 1547-1553(2019).

    [40] K. Y. Bliokh, F. Nori. Spatiotemporal vortex beams and angular momentum. Phys. Rev. A, 86, 033824(2012).

    [41] N. Jhajj, I. Larkin, E. W. Rosenthal. Spatiotemporal optical vortices. Phys. Rev. X, 6, 031037(2016).

    [42] A. Chong, C. Wan, J. Chen. Generation of spatiotemporal optical vortices with controllable transverse orbital angular momentum. Nat. Photonics, 14, 350-354(2020).

    [43] A. P. Sukhorukov, V. V. Yangirova. Spatio-temporal vortices: properties, generation and recording. Proc. SPIE, 5949, 594906(2005).

    [44] K. Y. Bliokh. Spatiotemporal vortex pulses: angular momenta and spin-orbit interaction. Phys. Rev. Lett., 126, 243601(2021).

    [45] H. Ge, S. Liu, X.-Y. Xu. Spatiotemporal acoustic vortex beams with transverse orbital angular momentum. Phys. Rev. Lett., 131, 014001(2023).

    [46] S. Wang, Y. Bai, N. Li. Generation of terahertz spatiotemporal optical vortices with frequency-dependent orbital angular momentum. Opt. Express, 31, 16267-16280(2023).

    [47] K. Y. Bliokh. Orbital angular momentum of optical, acoustic, and quantum-mechanical spatiotemporal vortex pulses. Phys. Rev. A, 107, L031501(2023).

    [48] F. Sun, W. Wang, H. Dong. Generation of isolated attosecond electron sheet via relativistic spatiotemporal optical manipulation. Phys. Rev. Res., 6, 013075(2024).

    [49] Q. Cao, Z. Chen, C. Zhang. Propagation of transverse photonic orbital angular momentum through few-mode fiber. Adv. Photonics, 15, 036002(2023).

    [50] S. Huang, Z. Li, J. Li. Spatiotemporal vortex strings. Sci. Adv., 10, eadn6206(2024).

    [51] G. Gui, N. J. Brooks, H. C. Kapteyn. Second-harmonic generation and the conservation of spatiotemporal orbital angular momentum of light. Nat. Photonics, 15, 608-613(2021).

    [52] Z.-Y. Chen, R. Hu, S. Zhang. Relativistic high-order harmonic generation of spatiotemporal optical vortices. Phys. Rev. A, 106, 013516(2022).

    [53] Y. Fang, S. Lu, Y. Liu. Controlling photon transverse orbital angular momentum in high harmonic generation. Phys. Rev. Lett., 127, 273901(2021).

    [54] K. Burnett, V. C. Reed, J. Cooper. Calculation of the background emitted during high-harmonic generation. Phys. Rev. A, 45, 3347-3349(1992).

    [55] M. D. Feit, J. A. Fleck, A. Steiger. Solution of the Schrödinger equation by a spectral method. J. Comput. Phys., 47, 412-433(1982).

    [56] R. Kosloff, H. Tal-Ezer. A direct relaxation method for calculating eigenfunctions and eigenvalues of the schrödinger equation on a grid. Chem. Phys. Lett., 127, 223-230(1986).

    [57] F. He, C. Ruiz, A. Becker. Absorbing boundaries in numerical solutions of the time-dependent Schrödinger equation on a grid using exterior complex scaling. Phys. Rev. A, 75, 053407(2007).

    [58] P. Antoine, B. Piraux, A. Maquet. Time profile of harmonics generated by a single atom in a strong electromagnetic field. Phys. Rev. A, 51, R1750-R1753(1995).

    [59] Y. Fang, M. Han, P. Ge. Photoelectronic mapping of the spin-orbit interaction of intense light fields. Nat. Photonics, 15, 115-120(2021).

    [60] S. W. Hancock, S. Zahedpour, H. M. Milchberg. Mode structure and orbital angular momentum of spatiotemporal optical vortex pulses. Phys. Rev. Lett., 127, 193901(2021).

    [61] Q. Zhan. Spatiotemporal sculpturing of light: a tutorial. Adv. Opt. Photonics, 16, 163-228(2024).

    [62] S. Huang, P. Wang, X. Shen. Diffraction properties of light with transverse orbital angular momentum. Optica, 9, 469-472(2022).

    [63] J. Itatani, F. Quéré, G. L. Yudin. Attosecond streak camera. Phys. Rev. Lett., 88, 173903(2002).

    [64] K. J. Schafer, M. B. Gaarde, A. Heinrich. Strong field quantum path control using attosecond pulse trains. Phys. Rev. Lett., 92, 023003(2004).

    [65] D. Azoury, M. Krüger, G. Orenstein. Self-probing spectroscopy of XUV photo-ionization dynamics in atoms subjected to a strong-field environment. Nat. Commun., 8, 1453(2017).

    [66] Z. Chang. Fundamentals of Attosecond Optics(2011).

    [67] T. Schultz, M. Vrakking. Attosecond and XUV Physics: Ultrafast Dynamics and Spectroscopy(2014).

    Jiahao Dong, Liang Xu, Yiqi Fang, Hongcheng Ni, Feng He, Songlin Zhuang, Yi Liu, "Scheme for generation of spatiotemporal optical vortex attosecond pulse trains," Photonics Res. 12, 2409 (2024)
    Download Citation