• Photonics Research
  • Vol. 8, Issue 6, 760 (2020)
Kaixuan Zhang1、3, Yizhu Zhang1、2、5、*, Xincheng Wang4, Tian-Min Yan1、6、*, and Y. H. Jiang1、3、4、7、*
Author Affiliations
  • 1Shanghai Advanced Research Institute, Chinese Academy of Sciences, Shanghai 201210, China
  • 2Center for Terahertz Waves and College of Precision Instrument and Optoelectronics Engineering, Key Laboratory of Opto-electronics Information and Technical Science, Ministry of Education, Tianjin University, Tianjin 300072, China
  • 3University of Chinese Academy of Sciences, Beijing 100049, China
  • 4ShanghaiTech University, Shanghai 201210, China
  • 5e-mail: zhangyz@sari.ac.cn
  • 6e-mail: yantm@sari.ac.cn
  • 7e-mail: jiangyh@sari.ac.cn
  • show less
    DOI: 10.1364/PRJ.377408 Cite this Article Set citation alerts
    Kaixuan Zhang, Yizhu Zhang, Xincheng Wang, Tian-Min Yan, Y. H. Jiang. Continuum electron giving birth to terahertz emission[J]. Photonics Research, 2020, 8(6): 760 Copy Citation Text show less

    Abstract

    The origin of terahertz (THz) generation in a gas-phase medium is still in controversy, although the THz sources have been applied across many disciplines. Herein, the THz generation in a dual-color field is investigated experimentally by precisely controlling the relative phase and polarization of dual-color lasers, where the accompanying third-harmonic generation is employed for in situ determination of the relative phase up to sub-wavelength accuracy. Joint studies with the strong approximation (SFA) theory reveal that the continuum-continuum (CC) transition within an escaped electron wave packet in the single atom gives birth to THz emission, without the necessity of considering the plasma effect. Meanwhile, we develop the analytic form from SFA-based CC description, which is able to reproduce and decompose the classical photocurrent model from the viewpoint of microscopic quantum theory, establishing the quantum-classical correspondence and bringing a novel insight into the mechanism of THz generation. Present studies leave open the possibility for probing the ultrafast dynamics of continuum electrons and a new dimension for the study of THz-related science and methodology.

    1. INTRODUCTION

    Terahertz (THz) wave generation (TWG) using dual-color femtosecond pulse, typically focusing 800 nm and 400 nm beams into a gas-phase medium, allows for the convenient and efficient access to moderately strong ultra-broadband THz pulses [1]. Although the approach is widely applied in several disciplines, the underlying generation mechanism is still an ongoing debate [2,3].

    The controversy is partially due to the ambiguity of the single-atom and plasma ensemble contributions. Due to the ambiguity, interpretations of TWG are rather distinct in their appearances, and the intrinsic physical pictures are completely different. From a plasma perspective, the photocurrent (PC) model formulates the emission process classically [4,5]. The propagation effects in plasma can modulate the spectra and polarization of the TWG [69]. From the single-atom perspective, the four-wave mixing (FWM), as in crystal nonlinear optics [1], explains the nonlinear THz emission based on the quantum perturbation theory. Based on the nonperturbative theory, the TWG has been numerically studied by solving the time-dependent Schrödinger equation [1013], accounting for the strong-field dynamics of a single atom. However, the time-dependent Schrödinger equation hardly provides a transparent physical picture. Comparatively, the strong-field approximation (SFA), which has been extensively developed to treat various strong-field phenomena, including above threshold ionization, high-order above-threshold ionization, multiple ionization, and high-harmonic generation (HHG), was also developed to provide a clear physical insight [14,15].

    Although the laser plasma has been considered the source of the TWG since the first observation from the laser–gas interaction [16], it is still unclear whether the plasma effect is a necessary ingredient. Similar debate arose in the early days when HHG was studied. Nowadays it is widely accepted that the HHG is a nonperturbative strong-field process dominated by the continuum-bound transition within a single atom or molecule, i.e., recombination of a released electron with its parent ion after ionization. Questions thus arise automatically. Is the THz radiation generated from the individual atom, analogous to HHG, or from the plasma ensemble effect? Can the strong-field TWG, analogous to other strong-field phenomena, be generalized to the strong-field theory framework, and can a unified theory for the TWG be established for integration of apparently distinct interpretations of FWM and the classical PC model?

    In this paper, the joint measurement between the TWG and third-harmonic generation (THG) is performed for in situ determination of phase delay zero of the dual-color fields, which allows us to select the most relevant mechanism in a direct comparison with different theoretical models. The dependence of the THz signal on the delay phase and relative polarization angles is measured, which shows remarkable consistencies with the continuum-continuum (CC) transition model. Our experimental and theoretical investigation confirms that the TWG is dominated by the CC transition within the escaped electronic wave packets, rather than the continuum-bound (CB) recollision, and the single-atom behavior is deconvoluted from the plasma ensemble. The schematic is illustrated in Fig. 1. Our work has manifold implications. From the theoretical aspect, the application scenario of the SFA is further expanded, bringing the TWG explicably under the framework of strong-field physics similar to the HHG. From application aspect, it implies that the TWG is still obtainable through the CC transition, even when the neutral atoms are fully depleted by the strong pump laser, showing the possibility to achieve intense THz fields by pumping the gas-phase medium with an extremely strong laser. Moreover, since the TWG is encoded by the time-dependent information of the continuum electron, it can be used as a spatial-temporal probe in the microscopic scale, complementary to HHG spectral line shape and photoelectron momentum distributions, to trace ultrafast dynamics of continuum electrons in atoms and molecules [17].

    Illustrative description of the TWG in dual-color fields and the mechanism. In panel (a), the dual-color laser ionizes the gas-phase medium and accelerates the electrons. According to the photocurrent model from the plasma perspective, the TWG originates from the rapidly varying macroscopic residual current as indicated by the giant arrow. However, the strong-field theory provides an alternative explanation from the microscopic aspect that each atom behaves as an individual THz emitter. As illustrated by the swarm of particles, each with an arrow indicates the THz radiation from a single atom. Panel (b) presents the radiation mechanisms under the strong-field approximation. After the photoelectron is released by the external light fields from the distorted atomic potential, it may recollide with its parent core, yielding the HHG; or within the continuum wave packet, the transition between continuum states of similar energies leads to the TWG. According to the derived SFA-CC, the involved continuum states are via different quantum paths as indicated by the white solid and dashed lines on the potential surface in (b). Panel (c) shows the origin of the two paths. The photoelectrons of initial state |Ψ(0)⟩ are released at different ionization times t′ and t′′ but arrive at the same intermediate momentum state k′(t′,t′′). The coherence between the continuum states yields the TWG.

    Figure 1.Illustrative description of the TWG in dual-color fields and the mechanism. In panel (a), the dual-color laser ionizes the gas-phase medium and accelerates the electrons. According to the photocurrent model from the plasma perspective, the TWG originates from the rapidly varying macroscopic residual current as indicated by the giant arrow. However, the strong-field theory provides an alternative explanation from the microscopic aspect that each atom behaves as an individual THz emitter. As illustrated by the swarm of particles, each with an arrow indicates the THz radiation from a single atom. Panel (b) presents the radiation mechanisms under the strong-field approximation. After the photoelectron is released by the external light fields from the distorted atomic potential, it may recollide with its parent core, yielding the HHG; or within the continuum wave packet, the transition between continuum states of similar energies leads to the TWG. According to the derived SFA-CC, the involved continuum states are via different quantum paths as indicated by the white solid and dashed lines on the potential surface in (b). Panel (c) shows the origin of the two paths. The photoelectrons of initial state |Ψ(0) are released at different ionization times t and t but arrive at the same intermediate momentum state k(t,t). The coherence between the continuum states yields the TWG.

    2. THEORY

    Quantum mechanically, the radiation is induced by the time-variant dipole moment d(t)=Ψ(t)|r^|Ψ(t)=Ψ0|U^(t0,t)r^U^(t,t0)|Ψ0 with the time-evolution operator U^ and the initial wave function |Ψ0. Using the Dyson series for U^, it is shown that d(t)=d(0)(t)+d(1)(t)+d(2)(t), including three components [18]. The first one, d(0)(t)=Ψ0(t)|r^|Ψ0(t), vanishes in the spherically symmetric system. The second term, d(1)(t)=ititdtΨ0(t)|r^U^(t,t)W^(t)|Ψ0(t)+c.c., describes the transition between CB states as presented by the upper panel of Fig. 1(b). The last term, d(2)(t)=titdttitdtΨ0(t)|W^(t)U^(t,t)r^U^(t,t)W^(t)|Ψ0(t), takes the form of a CC transition as presented by the lower panel of Fig. 1(b). Since the external light field is intense, the situation enters the scope of strong-field physics, and a natural choice to tackle the problem is SFA theory. Essentially, the SFA neglects the influence from the Coulomb potential of the ionic core. Hence, U^ can be substituted by U^(V), the evolution operator of the Volkov state, which is the eigenstate of an electron in the external light field alone, to simplify the further derivation. With the SFA, d(1)(t) is used to describe the HHG, which is essentially the widely used Lewenstein’s model of an illustrative interpretation: the atomic ionization is followed by the transition of the continuum electron back to the bound state and, more intuitively, the recollision of the released electron to its parent core, yielding the HHG. The contribution of d(2)(t) to the HHG is usually negligible [18], as only the “hard” recollision leads to photons of high energy, whereas d(2)(t) is the “soft” transition between continuum states, and the energy of the radiation photons is expected to be small. For the THz photons with small energy, the contribution from d(2)(t) should be considered, though it is rarely mentioned [14,15]. In this work, the TWG mechanism is investigated based on the analysis of d(2)(t), which is referred to as the SFA-based CC transition (SFA-CC).

    With the radiation given by the acceleration form d¨(t), we evaluate the emission field from the derived d(2)(t) (see Appendices A and C for details): E(t)d¨(2)(t)a1(t)+a2(t).The first term is a1(t)=E(t)titdttitdtη(t,t)W(t)W*(t)eiSk,Ip(t,t),where η(t,t)=[2πi(tt)]3/2 depicts the diffusion of the electronic wave packet and W(t)=μ[k+A(t)]·E(t) is the interaction of the electron with the incident light field E(t). Here k=k(t,t)[α(t)α(t)]/(tt), α(t)=dtA(t) is the excursion of the electron, and A(t) is the vector potential. The ionization rate is related to Sk,Ip(t,t)=ttdt{12[k+A(t)]2+Ip} with Ip the ionization energy. The presence of both W(t) and W(t) indicates the two electronic continuum states are involved, and eiSk,Ip(t,t) indicates the joint occurrences of these continuum states created by ionization. The process depicted by Eq. (2) is shown in Fig. 1(c). The TWG is the result of transition between continuum states, which share the same intermediate momentum k(t,t), while starting from different ionization times t and t. The second term of Eq. (1) emerges when the emission time t approaches the ionization time t: a2(t)=2Retitdtη(t,t)W(t)W*(t)[k+A(t)]eiSk,Ip(t,t),which is referred to as the temporal boundary term. Here k=k(t,t).

    3. EXPERIMENT

    The TWG in dual-color fields is determined by Eqs. (1)–(3). We apply these equations to examine the delay dependence and polarization dependence of the TWG in the measurement. Comparing the delay-dependent yields of the different-order harmonic generations is a commonly used method for in situ determination of the phase delay of the dual-color field. Here, the THG, a simple measurable physical characteristic, is measured to define the phase delay with sub-wavelength accuracy, which allows for comparison with the theoretical prediction. The femtosecond laser with pulse energy of 1.75  mJ and duration of 35  fs passes through a beta-barium borate (BBO) crystal, generating 800/400 nm two-color laser fields with an intensity ratio of 3:1. The two-color laser is then focused by a reflection mirror of 100 mm focal length to ionize the atmospheric air. The reflective mirror focuses light without chromatic aberration and thus promises the temporal and spatial overlap of the different wavelength lasers. The tight focus scheme is adapted to prevent the plasma filamentation around the focus. The propagation effect and the phase mismatching in the plasma filamentation is negligible [6,7] in our analysis.

    Throughout the measurement, the 2ω wave is kept s polarized, while the relative polarization angle θ and time delay τ between the two-color fields can be independently controlled. The time delay scheme is carefully designed in our measurement. The ordinary (o) axis of BBO crystal is always aligned with the polarization of the ω beam. Thus, the ω and 2ω beams are decomposed into orthogonal polarizations when propagating in the crystal. In previous experiments, the o axis of the BBO was aligned around 60° with respect to the polarization of the ω beam to ensure maximum TWG efficiency. The phase shift of the o ray and e ray of the ω beam in the BBO crystal [19] would induce the vague τ and θ of the ω and 2ω beams in the focus. Because of the inline design of the ω and 2ω beams, the fluctuation of τ can be passively suppressed with sub-wavelength accuracy. The vector of the emitted THz electric field ETHz(τ,θ) is recorded with the electro-optic sampling technique. More experimental details and the definition of observables can be referred from Appendix D. The components of peak-peak (PP) values along the orthogonal polarizations STHz,s(τ,θ) and STHz,p(τ,θ) extracted from ETHz(τ,θ) are presented in Fig. 2(a).

    Comparison of PP distributions, STHz,s(τ,θ) (upper row) and STHz,p(τ,θ) (lower row), respectively, for the s and p components of ETHz(t) obtained from (a) experiment and (b)–(d) theoretical models of the (b) SFA-CB, (c) SFA-CC, and (d) SPC.

    Figure 2.Comparison of PP distributions, STHz,s(τ,θ) (upper row) and STHz,p(τ,θ) (lower row), respectively, for the s and p components of ETHz(t) obtained from (a) experiment and (b)–(d) theoretical models of the (b) SFA-CB, (c) SFA-CC, and (d) SPC.

    Since it is nontrivial to precisely acquire the time delay τ between the two-color fields, a joint measurement of the intensities of the THG, I3rd(τ,θ), is performed. We notice that the THG emissions evaluated with all theoretical models present similar patterns in which the maximum of I3rd(τ,θ) appears at τ=0 (Appendix E). Hence, the time delay zero of the τ-dependent signals can be reliably determined by locating the maximum of I3rd(τ,θ). The phase shift (0.3π) of the τ-dependent TWG yield is not found in our measurement as in the previous studies [11]. The main reasons are the larger intensity ratio and total intensity of the 2ω and ω fields, diminishing the influence of the long-range Coulomb potential. Considering the experimental condition, our measurement shows consistency, with the exact solution based on the TDSE method in Ref. [20].

    In our measurement, the τ-dependent distributions STHz,s(p)(τ) [Fig. 2(a)] and I3rd(τ) (Appendix E) show the antiphase relationship in which the maximum TWG along τ coincides with the minimum THG. On one hand, the antiphase distribution clearly rules out the contribution from the CB transition d¨(1)(t) (Appendix B for details), since it contradicts the synchronized distributions of STHz,s(p)(τ) and I3rd(τ) as predicted by d¨(1)(t) [Fig. 2(b)]. The result also conflicts with the prediction of perturbative FWM, which predicts that the maxima of STHz,s(p)(τ) and I3rd(τ) are coincident. On the other hand, the PP values from the CC transition d¨(2)(t), as shown in Fig. 2(c), well reproduce the salient experimental characteristics, e.g., the decrease and revival of STHz,s(τ,θ) when θ is increasing, confirming the role of the single-atom ionization process in the TWG.

    The distribution is also evaluated with the PC model, where the TWG is determined by the time-variant plasma density ETHz(t)j(t)/t=e2N(t)E(t)/m. The transient electron density N(t) originates from the accumulated electrons from ionization, satisfying tN(t)=[NgN(t)]w(t),where Ng is the initial density of the air and w(t) is the ionization rate [21]. Thus, ETHz(t)Ng{1exp[dtw(t)]}. E(t) accounts for the residual current induced by the external field in the plasma. The expansion up to the first order of the exponent, i.e., ETHz(t)E(t)dtw(t),however, is essentially based on the single-atomic ionization, since Eq. (5) is the solution of tN(t)=Ngw(t), where the depletion of the neutral atoms in plasma, N(t)w(t), as appeared in Eq. (4), is neglected. Therefore, Eq. (5) is referred to as the single-atom PC (SPC) model (see Appendix F for comparison of results between PC and SPC). In Fig. 2(d), the distribution from the SPC model also shows good agreement with the experiment in Fig. 2(a). Comparing the measurement and theoretical results, the deviations only happen in the low-signal regime with low signal-to-noise ratio, where small errors would result in the uncertainty of the PP sign. However, the characteristics of STHz(τ,θ) are in fair agreement with the theory.

    4. DISCUSSION

    It is not a coincidence that all these models agree well with the experimental results. As is shown in the following, there exists a link among different theories. Obviously, the SFA-CC and the SPC share the similar form—the rate w(t) in Eq. (5) is simply substituted by w1(t;t) in Eq. (2) as defined by w1(t;t)=titdtη(t,t)W(t)W*(t)eiSk,Ip(t,t). Before showing the correspondence between w(t) and w1(t;t), we first examine the distribution of w1(t;t) versus the ionization time t at different emission instants t. As is presented in Fig. 3(a) and inset (c), w1(t;t) is nonvanishing only for t>t, as is restricted by the principle of causality that the ionization event should precede the emission. Despite the apparent dependence of w1(t;t) on t, the temporal distributions for t<t remain almost unaltered versus t, except around the boundary when t=t.

    w(t′;t) of the SFA-CC and the comparison with w(t′) of the SPC when θ=0° and τ=0.33 fs. In collinear dual-color laser fields, the SFA-CC derived w1(t′;t) and w2(t′;t) are shown in (a) and (b), respectively, with their detailed zoom-in around t′=0 in insets (c) and (d). The total contribution, w(t′;t)=w1(t′;t)+w2(t′;t), is presented in inset (e), showing that w2(t′;t) almost contributes at t′=t only. In (f), w(t′;t→∞) of the SFA-CC (solid line) is compared with w(t′) of the SPC (dashed line), showing the correspondence between the SFA-based quantum model and the semi-classical PC model.

    Figure 3.w(t;t) of the SFA-CC and the comparison with w(t) of the SPC when θ=0° and τ=0.33  fs. In collinear dual-color laser fields, the SFA-CC derived w1(t;t) and w2(t;t) are shown in (a) and (b), respectively, with their detailed zoom-in around t=0 in insets (c) and (d). The total contribution, w(t;t)=w1(t;t)+w2(t;t), is presented in inset (e), showing that w2(t;t) almost contributes at t=t only. In (f), w(t;t) of the SFA-CC (solid line) is compared with w(t) of the SPC (dashed line), showing the correspondence between the SFA-based quantum model and the semi-classical PC model.

    Besides w1(t;t) for a1, we can also define w2(t;t) from a2 so that the total contribution of the SFA-CC, w(t;t)=w1(t;t)+w2(t;t), is formally consistent with w(t) in Eq. (5). For the collinear dual-color laser fields, straightforwardly we have w2(t;t)=2Reη(t,t)W(t)W*(t)k+A(t)E(t)eiSk,Ip(t,t). The distribution of w2(t;t) is shown in Fig. 3(b) and inset (d). The contribution from w2(t;t) is almost negligible, except when tt; that is why it is referred to as the boundary term. It may influence some details of the TWG process, leading to subtle differences between the quantum mechanical SFA-CC and the semiclassical PC models. When t is sufficiently large, the distribution of w(t;t) is almost equivalent to w1(t;t).

    With the approximation that the contribution from w2(t;t) is negligible and w1(t;t) is roughly independent of t, w1(t;t) versus the ionization time t can be directly compared with w(t) as shown in Fig. 3(f). Comparing SFA-CC with the SPC model, if the emission time t is sufficiently distant from t, w1(t;t) versus t presents a similar distribution to w(t) of the SPC. In other words, w(t) can be considered the quasi-static limit of w1(t;t) restricted by the time-ordering t<t, even though w(t) is introduced from the view of macroscopic photoelectric current, while w1(t;t) is derived completely from single-atom-based microscopic process of strong-field ionization. In explaining the TWG, the SFA and (S)PC theories exhibit quantum-classical correspondence.

    Besides the formal similarities to the PC model, a1(t) in Eq. (2) explicitly shows the third-order dependence on the external electric field E(t) as presented by the perturbative third-order response in the FWM model. The response to the incident fields, as predicted in the FWM, can be verified by experimental observations of polarization- and intensity-dependent THz yields [19,22]. The conventional perturbative FWM susceptibility, however, is replaced by the nonperturbative transition dipole moment induced by strong fields. Thus, the SFA-CC is in concordance with the ionization-induced multiwave mixing [12], unifying the existent explanations, including the FWM and PC models.

    In conclusion, the joint measurement between the TWGs and THGs is implemented for in situ determination of the phase delay of the dual-color fields, allowing us to distinguish the different theoretical models. Our research exhibits that the underlying origin of the TWG resides within the scope of nonperturbative single-atomic strong-field processes, although the ensemble behavior of the laser-induced plasma may influence the radiation yields. In contrast to the HHG emitted by the CB transition of a recolliding electron, the TWG originates from the CC transition of a released electron after ionization. The TWG mechanism of the CC—the “soft” transition—beyond the HHG mechanism of “hard” recollision, is a complement to the radiation theory of strong-field physics. Meanwhile, it is shown that the classical PC model can be derived from the quantum SFA-CC method, bridging between the classical and quantum-mechanical interpretations. Also, the FWM can be reached from the SFA-CC by presenting the explicit third-order dependence on the electric field. Hence, the SFA-CC serves to unify the FWM and PC models, while it offers more microscopic details compared to the latter coarse-grained models. Our research of the TWG mechanism opens up the possibility to extract the ultrafast dynamics of continuum electrons from the ionization-induced THz emission.

    Acknowledgment

    Acknowledgment. We acknowledge the support from the Shanghai-XFEL Beamline Project (SBP) and Shanghai HIgh repetitioN rate XFEL and Extreme light facility(SHINE).

    APPENDIX A: TRANSITION DIPOLE MOMENT UNDER THE STRONG-FIELD APPROXIMATION

    The expected value of dipole moment is d(t)= Ψ(t)|r^|Ψ(t) = Ψ0(ti)|U^(ti,t)r^U^(t,ti)|Ψ0(ti) . With the time-evolution operator U^(t,ti) expanded by Dyson series U^(t,ti)=U^0(t,ti) ititdtU^(t,t)W^(t)U^0(t,ti), where U^0(t,ti) is the interaction-free time-evolution operator and W^ is the interaction operator, we find that d(t)=d(0)(t)+d(1)(t)+d(2)(t), where d(0)(t)= Ψ0(t)|r^|Ψ0(t) ,d(1)(t)= ( i)titdt Ψ0(t)|r^U^(t,t)W^(t)|Ψ0(t) +c.c.,d(2)(t)= titdt Ψ0(t)|W^(t)×titdtU^(t,t)r^U^(t,t)W^(t)|Ψ0(t) .The d(0)(t) vanishes in a spherically symmetric system. The d(1)(t), referred to as the CB transition dipole moment, depicts the coherent emission at time t induced by the transition of electrons from the continuum, which accumulates over all possible ionization events at time t, to the bound state. The d(1)(t) is widely recognized to dominate the HHG process. Similarly, the d(2)(t), referred to as the CC transition dipole moment, describes the coherent emission induced by the transition between the states of the continuum, which is often negligible for the HHG calculation. Here, however, we emphasize the role of d(2)(t) in TWG. In the following, the details of d(1)(t) and d(2)(t) as used to evaluate the emission are presented.

    APPENDIX B. d(1): CONTINUUM-BOUND TRANSITION

    The CB transition d(1)(t) is given by Eq. (A2). Under the SFA, which neglects the interaction between the photoelectron and the parent ion, the full time-evolution operator is substituted by one with the external light field only: U^(t,t)U^I(t,t)=dk|Ψk(V)(t) Ψk(V)(t)|, yielding d(1)(t)= ( i)titdt Ψ0(t)|r^U^I(t,t)W^(t)|Ψ0(t) +c.c..The Volkov state |Ψk(V)(t) =|k+A(t) e iSk(t) describes the free electron in the presence of a time-dependent electric field of vector potential A(t) with Sk(t)=12dt[k+A(t)]2. Assuming W(t)=μ[k+A(t)]·E(t), the interaction between the incident electric field E(t) and the electron of initial state |Ψ0(t) =|ψ0 e iE0t, it is shown that d(1)(t)=ititdt·dke iSk,Ip(t,t)μ*[k+A(t)](μ[k+A(t)]·E*(t))+c.c., where Sk,Ip(t,t)=ttdt(12[k+A(t)]2+Ip) with k the intermediate momentum and Ip= E0 the ionization energy.

    The integration over momentum can be approximated with the stationary phase, and the stationary point ks is the solution to the equation kSk,Ip(t,t)=!0. Therefore, we obtain the CB transition dipole d(1)(t)=ititdt[2πi(t t)]32e iSks,Ip(t,t)×μ*[ks+A(t)](μ[ks+A(t)]·E*(t))+c.c.,recovering the widely used Lewenstein model for the HHG.

    APPENDIX C. d(2): CONTINUUM-CONTINUUM TRANSITION

    Under the strong-field approximation, the substitution U^(t,t)U^I(t,t) yields d(2)(t) (+i)( i)titdt Ψ0(t)|W^(t) titdtU^I(t,t)r^U^I(t,t)W^(t)|Ψ0(t) .

    Applying the similar procedure as considered for the CB transition, the expansion arrives at d(2)(t)= (+i)( i)titdttitdtdk(μ[k+A(t)]×E(t))eiSk,Ip(t,t)×dk k+A(t)|r^|k+A(t) ×(μ*[k+A(t)]·E(t))e iSk,Ip(t,t).With k+A(t)|r^|k+A(t) =i kδ(k k) and applying integration by parts, the last integration over k reads { k i kSk,Ip(t,t)}(μ*[k+A(t)]·E(t))e iSk,Ip(t,t),yielding d(2)(t)= (+i)( i)(i)( 1)titdttitdtdk ×(μ[k+A(t)]·E(t)){ k i kSk,Ip(t,t)}×(μ*[k+A(t)]·E(t))eiSk,Ip(t,t).The integration over k can also be treated by the stationary phase approximation. Solving the saddle point equation kSk,Ip(t,t)=0, we obtain ksks(t,t)= [α(t) α(t)]/(t t) with the excursion of the oscillating electron α(t)=dtA(t). Within the curly bracket of Eq. (C3), the first term is negligible, and dipole is determined by the second term d(2)(t) titdttitdt[2πi(t t)]3/2(μ[ks+A(t)]×E(t))(μ*[ks+A(t)]·E(t))×[ks(t t)+α(t) α(t)]eiSks,Ip(t,t)after the substitution ksSks,Ip(t,t)=ks(t t)+α(t) α(t). The emission is given by the acceleration a=d¨(2)(t). Using the Leibniz integral rule, the second-order derivative of Eq. (C4) with respect to t is evaluated: d¨(2)(t)=E(t)titdttitdt[2πi(t t)]3/2×(μ[ks(t,t)+A(t)]·E(t))×(μ*[ks(t,t)+A(t)]·E(t))×eiSks(t,t),Ip(t,t) 2 Retitdt[2πi(t t)]3/2×(μ[ks(t,t)+A(t)]·E(t))×(μ*[ks(t,t)+A(t)]·E(t))×[ks(t,t)+A(t)]eiSks(t,t),Ip(t,t),showing the full form of d¨(2)(t)=a1(t)+a2(t) as presented by Eqs. (2) and (3) in the main text.

    In this work, considering the H(1s) initial state for simplicity, the dipole matrix element reads μ(k)= i272k/[π(k2+1)3], and Ip=0.5 a.u..

    APPENDIX D: EXPERIMENT

    The femtosecond amplifier (Libra, Coherent Inc.) delivers a laser pulse with 800 nm center wavelength and 35 fs pulse duration. The pulse with 1.75 mJ is guided into the experiment setup (Fig. 4). The beam with 96% pulse energy is reflected as the pump beam for the TWG, and the transmission beam is used as the probe beam for electro-optic sampling (EOS). The pump beam passes through a 200-μm type-I BBO crystal with the double-frequency efficiency of 23%. The polarization of the laser beam is horizontal (p polarized), and the optical axis of BBO is kept perpendicular to the laser polarization to obtain the maximum efficiency. The outcoming second harmonic beam is s polarized. The relative polarization θ between the fundamental ω and second-harmonic 2ω beams can be controlled by rotating the zero-order dual-wavelength wave plate (DWP), which acts as a half-wave plate for the ω beam and a full-wave plate for the 2ω beam. The definition of observables is schematically illustrated in Fig. 5. Controlling θ with a half-wave plate, instead of rotating BBO crystal, avoids the mixture of the polarization of o^ ray and e^ ray in the BBO crystal. The ellipticities of both 2ω and ω beams are better than 0.1 when DWP rotates. Throughout the measurement, the ω and 2ω beams can be considered as linearly polarized. The time delay τ of the dual-color fields can be varied by moving the BBO crystal along the propagation direction, because of different refractive indices at ω and 2ω in air. Because of the collinear propagation geometry, the fluctuation of τ can be passively suppressed within the sub-wavelength accuracy.

    Experimental setup. BS, beam splitter; CH, chopper; β-BBO, beta barium borate; DWP, dual-wavelength plate; PM, parabolic mirror; QWP, quarter-wave plate; GLP, Glan-laser polarizer; WP, Wollaston polarizer; BD, balanced detector.

    Figure 4.Experimental setup. BS, beam splitter; CH, chopper; β-BBO, beta barium borate; DWP, dual-wavelength plate; PM, parabolic mirror; QWP, quarter-wave plate; GLP, Glan-laser polarizer; WP, Wollaston polarizer; BD, balanced detector.

    Definition of observables. The 800 nm (ω) and 400 nm (2ω) beams collinearly propagate. The 2ω beam is always s polarized. The relative time delay τ and polarization angle θ are controlled in the measurement. Both the s- and p-polarized terahertz waves are detected.

    Figure 5.Definition of observables. The 800 nm (ω) and 400 nm (2ω) beams collinearly propagate. The 2ω beam is always s polarized. The relative time delay τ and polarization angle θ are controlled in the measurement. Both the s- and p-polarized terahertz waves are detected.

    The dual-color fields are focused by a silver parabolic mirror with an effective focal length of 100 mm. Atmospheric air is ionized for the TWG, simultaneously emitting THG. Here, we use a tightly focusing scheme to deliberately prevent the propagation effect in plasma. The THz waves are collected and collimated with a gold parabolic mirror with 100 mm focal length, and focused into 1 mm thick (110)-cut ZnTe crystal with a same parabolic mirror. A 500 μm thick polished silicon wafer reflects the residual laser, allowing the transmission of only the THz component. A pellicle beam splitter combines the THz pulse and the probe beam to implement the free-space EOS detection. The signal-to-noise ratio (SNR) of the terahertz time-domain waveforms is better than 100:1. A polarization-sensitive THz-EOS is employed in our experiment. A metal wire-grid THz polarizer filters out the orthogonally polarized components of TWG, and the ZnTe crystal is fixed at the special orientation, where the responses for s- and p-polarized THz components are the same [23].

    In the measurement, both s- and p-polarized THz electric fields ETHz(t) are recorded with the EOS method. The THz peak-peak (PP) amplitude is defined as STHz=±|max[ETHz(t)] min[ETHz(t)]|. The distributions of STHz,s(p)(τ,θ) are presented in Fig. 2(a) in the main text. The STHz,s(p)(τ,θ) are normalized to the maximum. The positive direction of STHz,s(τ,θ) and STHz,p(τ,θ) are defined when the maxima of THz waveforms appear along the positive direction of the s and p axes.

    The THG of 266 nm is coincidently measured with the TWG for the in situ determination of the absolute τ of the dual-color fields. The THG signals are reflected by the silicon wafer with residual ω and 2ω beams, and spectrally separated by a suprasil prism. The s and p components of THG signals are decomposed with a Glan-laser polarizer and collected into a fiber spectrometer. The THG signals are measured with 50 ms integration time, 10-time average in our measurement.

    APPENDIX E. DETERMINATION OF TIME DELAY ZERO BY JOINT MEASUREMENT OF THIRD-ORDER HARMONICS

    The dependence of TWG on time delay τ between the dual-color fields is critical to determine the TWG mechanisms. For instance, the TWG maximum is predicted to appear at τ=0.33 fs in the PC theory; however, it is predicated at τ=0 fs in the perturbative FWM theory. Only when the τ is precisely known, the electric field waveforms for the TWG can be determined for further comparison of the measured data with different theories. However, the precise τ is difficult to obtain, since it is nontrivial to directly monitor the electric fields in practical experiments.

    In our experiment, the joint measurement of TWG and THG is conducted. The THG yields along the s-polarization I3rd(τ,θ) are shown in Fig. 6(a). In addition, we have examined the I3rd(τ,θ) evaluated by different theories, including the CB, CC transitions, and SPC. All results, as shown in Figs. 6(b)6(d), predict the similar τ dependence that the maximum I3rd(τ) appears at τ=0 fs. The time delay zero of dual-color fields in the experiment can therefore be precisely determined by comparison with the theoretical results.

    Distribution of the THG along s-polarization I3rd(τ,θ) obtained from (a) the measurement, (b) d¨(1)(t) of SFA-CB, (c) d¨(2)(t) of SFA-CC, and (d) SPC.

    Figure 6.Distribution of the THG along s-polarization I3rd(τ,θ) obtained from (a) the measurement, (b) d¨(1)(t) of SFA-CB, (c) d¨(2)(t) of SFA-CC, and (d) SPC.

    APPENDIX F: PHOTOCURRENT AND SINGLE-ATOM PHOTOCURRENT MODEL

    In Eq. (5) in the main text, neglecting the neutral depletion is referred to as the SPC model. The STHz,s(p) from the SPC model is shown in Fig. 2 of the main text. Here, the STHz,s(p) from the traditional PC model is presented in Fig. 7 by comparison. It is shown that there is no significant deviation of the SPC from the PC, which involves the neutral depletion.

    (a) STHz,s(τ,θ) and (b) STHz,p(τ,θ) evaluated from the PC model.

    Figure 7.(a) STHz,s(τ,θ) and (b) STHz,p(τ,θ) evaluated from the PC model.

    (a) STHz,s(τ=0.33 fs,θ) and (b) STHz,s(τ,θ=90°) of the experiment (red star), PC (blue dashed), and SPC (magenta dotted) models.

    Figure 8.(a) STHz,s(τ=0.33  fs,θ) and (b) STHz,s(τ,θ=90°) of the experiment (red star), PC (blue dashed), and SPC (magenta dotted) models.

    References

    [1] D. J. Cook, R. M. Hochstrasser. Intense terahertz pulses by four-wave rectification in air. Opt. Lett., 25, 1210-1212(2000).

    [2] V. A. Andreeva, O. G. Kosareva, N. A. Panov, D. E. Shipilo, P. M. Solyankin, M. N. Esaulkov, P. G. de Alaiza Martnez, A. P. Shkurinov, V. A. Makarov, L. Bergé, S. L. Chin. Ultrabroad terahertz spectrum generation from an air-based filament plasma. Phys. Rev. Lett., 116, 063902(2016).

    [3] L.-L. Zhang, W.-M. Wang, T. Wu, R. Zhang, S.-J. Zhang, C.-L. Zhang, Y. Zhang, Z.-M. Sheng, X.-C. Zhang. Observation of terahertz radiation via the two-color laser scheme with uncommon frequency ratios. Phys. Rev. Lett., 119, 235001(2017).

    [4] K. Y. Kim, J. H. Glownia, A. J. Taylor, G. Rodriguez. Terahertz emission from ultrafast ionizing air in symmetry-broken laser fields. Opt. Express, 15, 4577-4584(2007).

    [5] K. Y. Kim, A. J. Taylor, J. H. Glownia, G. Rodriguez. Coherent control of terahertz supercontinuum generation in ultrafast laser-gas interactions. Nat. Photonics, 2, 605-609(2008).

    [6] Z. Zhang, Y. Chen, M. Chen, Z. Zhang, J. Yu, Z. Sheng, J. Zhang. Controllable terahertz radiation from a linear-dipole array formed by a two-color laser filament in air. Phys. Rev. Lett., 117, 243901(2016).

    [7] Z. Zhang, Y. Chen, S. Cui, F. He, M. Chen, Z. Zhang, J. Yu, L. Chen, Z. Sheng, J. Zhang. Manipulation of polarizations for broadband terahertz waves emitted from laser plasma filaments. Nat. Photonics, 12, 554-559(2018).

    [8] A. Nguyen, P. G. de Alaiza Martínez, J. Déchard, I. Thiele, I. Babushkin, S. Skupin, L. Bergé. Spectral dynamics of THz pulses generated by two-color laser filaments in air: the role of Kerr nonlinearities and pump wavelength. Opt. Express, 25, 4720-4740(2017).

    [9] A. Nguyen, P. G. de Alaiza Martnez, I. Thiele, S. Skupin, L. Bergé. THz field engineering in two-color femtosecond filaments using chirped and delayed laser pulses. New J. Phys., 20, 033026(2018).

    [10] N. Karpowicz, X.-C. Zhang. Coherent terahertz echo of tunnel ionization in gases. Phys. Rev. Lett., 102, 093001(2009).

    [11] D. Zhang, Z. Lü, C. Meng, X. Du, Z. Zhou, Z. Zhao, J. Yuan. Synchronizing terahertz wave generation with attosecond bursts. Phys. Rev. Lett., 109, 243002(2012).

    [12] V. A. Kostin, I. D. Laryushin, A. A. Silaev, N. V. Vvedenskii. Ionization-induced multiwave mixing: terahertz generation with two-color laser pulses of various frequency ratios. Phys. Rev. Lett., 117, 035003(2016).

    [13] C. Köhler, R. Guichard, E. Lorin, S. Chelkowski, A. D. Bandrauk, L. Bergé, S. Skupin. Saturation of the nonlinear refractive index in atomic gases. Phys. Rev. A, 87, 043811(2013).

    [14] Z. Zhou, D. Zhang, Z. Zhao, J. Yuan. Terahertz emission of atoms driven by ultrashort laser pulses. Phys. Rev. A, 79, 063413(2009).

    [15] T. Balčiūnas, D. Lorenc, M. Ivanov, O. Smirnova, A. M. Zheltikov, D. Dietze, K. Unterrainer, T. Rathje, G. G. Paulus, A. Baltuška, S. Haessler. CEP-stable tunable THz-emission originating from laser-waveform-controlled sub-cycle plasma-electron bursts. Opt. Express, 23, 15278-15289(2015).

    [16] H. Hamster, A. Sullivan, S. Gordon, W. White, R. W. Falcone. Subpicosecond, electromagnetic pulses from intense laser-plasma interaction. Phys. Rev. Lett., 71, 2725-2728(1993).

    [17] I. Babushkin, A. J. Galan, V. Vaičaitis, A. Husakou, F. Morales, A. Demircan, J. R. C. Andrade, U. Morgner, M. Ivanov. All-optical attoclock: accessing exahertz dynamics of optical tunnelling through terahertz emission(2018).

    [18] W. Becker, A. Lohr, M. Kleber, M. Lewenstein. A unified theory of high-harmonic generation: application to polarization properties of the harmonics. Phys. Rev. A, 56, 645-656(1997).

    [19] Y. Zhang, Y. Chen, S. Xu, H. Lian, M. Wang, W. Liu, S. L. Chin, G. Mu. Portraying polarization state of terahertz pulse generated by a two-color laser field in air. Opt. Lett., 34, 2841-2843(2009).

    [20] W. Chen, Y. Huang, C. Meng, J. Liu, Z. Zhou, D. Zhang, J. Yuan, Z. Zhao. Theoretical study of terahertz generation from atoms and aligned molecules driven by two-color laser fields. Phys. Rev. A, 92, 033410(2015).

    [21] M. Ammosov, N. Delone, V. Krainov. Tunnel ionization of complex atoms and of atomic ions in an alternating electromagnetic field. JETP, 64, 1191-1194(1986).

    [22] X. Xie, J. Dai, X.-C. Zhang. Coherent control of THz wave generation in ambient air. Phys. Rev. Lett., 96, 075005(2006).

    [23] P. Planken, H. Nienhuys, H. Bakker. Measurement and calculation of the orientation dependence of terahertz pulse detection in ZnTe. J. Opt. Soc. Am. B, 18, 313-317(2001).

    Kaixuan Zhang, Yizhu Zhang, Xincheng Wang, Tian-Min Yan, Y. H. Jiang. Continuum electron giving birth to terahertz emission[J]. Photonics Research, 2020, 8(6): 760
    Download Citation