Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Third-order harmonic generation in a bi-chromatic elliptical laser field

Open Access Open Access

Abstract

The low-order harmonic generation induced by a strong laser field produces a bright, ultrashort, supercontinuum radiation ranging from the terahertz to ultraviolet band. By controlling the phase-delay and ellipticity of the bi-chromatic laser fields, the third harmonic generation is experimentally and theoretically investigated for elucidating the mechanism of the low-order harmonics. The third harmonic generation is found to be strongly suppressed in the counter-rotating bi-chromatic laser field due to the selection rule for harmonic emissions. The continuum-continuum transition in the strong field approximation is extended to explain the third harmonic generation as a function of the phase delay and ellipticity of the bi-chromatic laser fields. Compared with the semi-classical photocurrent model, the continuum-continuum transition on the basis of quantum-mechanical treatment achieves better agreement with the experimental observations. Our work indicates that the overlapping in continuum states via different quantum paths of a single electron plays a role in low-order harmonics generation under elliptical bi-chromatic laser fields.

© 2021 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Harmonic generation is a typical phenomena of extreme nonlinear optics when the strong laser field interacts with the matters. The high harmonic generation (HHG) delivers the extreme ultraviolet ultrashort pulse with high photon energy, which has been widely investigated for its intriguing nonperturbative quantum dynamics and enormous applications as emerging innovative light sources. Comparatively, the low harmonic generation (LHG) refers to the radiation whose photon energy is much lower than the ionization energy. Recently, the LHG has attracted tremendous research interests due to both application and theory meanings. From the application aspect, the LHG from both solid- and gas-phase medium is a reliable approach to generate the ultrashort supercontinuum pulse ranging from terahertz to ultraviolet and beyond [14], providing an advanced light source for ultrafast spectroscopy [57], subcycle pulse synthesis [8] and frequency-comb technology [911]. From the theory aspect, the LHG from solid-state medium is associated with interband and intraband electron dynamics, which is essential for understanding the strong-field electron dynamics and extreme nonlinear optics in solids [1216]. Thus, the systematical investigation of the LHG in simple atoms and molecules will provide a benchmark for a further understanding of the LHG in complex solid systems.

The physical mechanism behind the HHG can be well explained by the three-step model [17]. Firstly, the strong incident laser field ionizes the atom from the ground state to the continuum state; Then, the continuum electron is accelerated by the the laser field; Finally, the electron recombines with the parent ion, simultaneously emitting the HHG. Nevertheless, the LHG is only related to the first and second steps in three-step model and not necessarily associated with the recombination of electron with the parent ion. In the presence of sufficiently strong fields, the density of free electrons increases step-wise at the each laser half cycle due to tunnelling ionization, and the LHG comes from the acceleration of the transient photocurrent induced by the strong laser fields. This mechanism is also named Brunel model [18].

In the past decades, many efforts have been made to understand the LHG from gas-phase medium [16,1923]. Previous works exhibit that the LHG has been attribute to different mechanisms according to various laser parameters. In the regime of high-intensity laser field, the nonperturbative nonlinear response of free electrons dominates the LHG process, which can be described by Brunel model. Simultaneously, the terahertz radiation, considered as the 0th order harmonics in LHG, can also be successfully explained with Brunel model [2428]. In low-intensity regime, the nonlinear response of bound electron is responsible for the LHG which can be described with the perturbation theory [2931]. In general, it is challenging to accurately distinguish the two mechanisms [32]. Numerical simulation based on time-dependent Schr$\ddot {\mathrm {o}}$dinger equation (TDSE) indicates that the nonlinear response of continuum-state electron is superior to bound electron, dominating the LHG in high-intensity low-frequency fields [33].

In this letter, the third harmonic generation (THG) driven by the bi-chromatic laser field is theoretically and experimentally investigated as an exemplar of the LHG, and the continuum-continuum transition in strong field approximation (SFA) is used to explain the THG. Although the SFA model has well predicted terahertz radiation[3436], yet it is rarely applied to investigate the THG and LHG. Unlike the THG under the linearly-polarized bi-chromatic field [35], the SFA and photocurrent model are fairly distinguishable by introducing the ellipticity as a new experimental dimension. The SFA model on the basis of the quantum theory gives better predictions, indicating that the photocurrent model has limitation for the THG explanation.

2. Experimental setup

To control the THG and explore the mechanism of the THG, we prepared the bi-chromatic laser field with elliptical polarization and controllable relative phase. Our set-up is presented in Fig. 1(a). The Ti:sapphire laser with 1 kHz repetition rate delivers linearly polarized ($p$-polarized) 35 fs, 2.5 mJ per pulse, centred around 800 nm with a bandwidth of 40 nm. This fundamental wave (FW) passes through a 200 µm type-I $\beta$-barium borate (BBO) crystal, where a second-harmonic (SH) field of 400 nm is generated (conversion efficiency $\sim$32%). The co-propagating two-color driving fields are separated by a dichroic mirror (DM-2) into the two arms of a Michelson interferometer. The polarization of the two-color laser can be arbitrarily controlled by a pair of half- and quarter-wavelength retardation plates. For instance, the co-rotating or counter-rotating bi-chromatic field when the FW and SH are both circular polarization are shown in Fig. 1(b). The FW (1.2 mJ/pulse) and SH (0.56 mJ/pulse) driving lasers are combined with a DM-3 and focused into atmospheric air by using a lens with focal length of 10 cm. The THG beam emerges from the air plasma then passing through a optical low pass filter to block the pump laser beams. The intensity of the THG can be measured with a grating spectrometer, and the polarization is tested by a glan-laser polarizer before the grating spectrometer.

 figure: Fig. 1.

Fig. 1. (a) Schematics of the experimental set-up. DM: dichroic mirror; BBO: $\beta$-phase barium borate crystal; L: lens; GLP: glan-laser polarizer; Piezo: piezo-stage; (b) The co-rotating (green) or counter-rotating (purple) bi-chromatic field combined with circularly polarized FW (red) and SH (blue) beams.

Download Full Size | PDF

In order to suppress the sub-femtosecond time jitter of the two-color laser induced by the opto-mechanical components, the actively stabilized Michelson interferometer is adopted in the set-up. To stabilize the relative phase of the two-color laser, another continuous green laser (532 nm) co-propagates with the two-color lasers and interferes before the CCD camera [37]. The interference fringes are monitored by a CCD camera as a feedback signal. The real-time control of a mirror with a piezo actuator is employed to keep the interference fringes stable. The active feedback can efficiently reduce the influence from the airflow and the mechanical noise from the mirror, and minimize the relative phase fluctuations of two-color laser. After stabilization, the relative phase fluctuation in our system is smaller than $\pm 0.02\pi$ during data acquisition.

The time delay can be controlled by two approaches: An optical delay stage roughly compensates the relative time delay between the two-color laser pulses with femtosecond accuracy, and the sub-femtosecond phase delay is precisely tuned by moving the BBO crystal along the propagation direction. The relative phase delay is introduced by the difference between the refractive indices of the two-color fields in the air. When the distance $\Delta L$ between the BBO crystal and air plasma is changed, the relative time delay $\tau = \frac {(n_{ {\operatorname {SH}}} - n_{ {\operatorname {FW}}}) \cdot \Delta L}{c}$ is introduced with the sub-femtosecond accuracy, and the corresponding relative phase can be expressed as $\phi = 2 \phi _{ {\operatorname {SH}}} - \phi _{ {\operatorname {FW}}} = (2 k_{ {\operatorname {SH}}} - k_{ {\operatorname {FW}}}) \cdot \Delta L$, where $n$ is the refractive index of air and $k=2\pi /\lambda$.

3. Theoretical methods

In the photocurrent (PC) model, the THG $ {\boldsymbol {E}}_{ {\operatorname {THG}}} (t)$ comes from the time-variant photocurrent $ {\boldsymbol {j}}(t)$ induced by an external strong field $ {\boldsymbol {E}} (t)$ is,

$${\boldsymbol{E}}_{{\operatorname{THG}}} (t) \propto \frac{\partial {\boldsymbol{j}}(t)}{\partial t} = \frac{e_{}^2}{m} N (t) {\boldsymbol{E}} (t) .$$

The transient electron density $N (t)$ originates from accumulated electrons from tunneling ionization, satisfying

$$\partial N (t) / \partial t = [N_g - N (t)] w (t),$$
where $N_g$ is the initial density of atmosphere, and $w (t)$ is the static tunneling ionization rate [38,39], $w (t) = 4 \omega _a \left ( \frac {U_{N_2}}{U_H} \right )^{5 / 2} \frac {E_a}{E_L (t)} \exp \left [ - \frac {2 E_a}{3 E_L (t)} \left ( \frac {U_{N_2}}{U_H} \right )^{3 / 2} \right ]$, where $\omega _a = 4.13 \times 10^{16}$ s$^{- 1}$ is the atomic frequency, $E_a = 5.14 \times 10^9$ V/cm is the atomic field amplitude, $E_L (t) = \sqrt {E^2_x (t) + E_y^2 (t)}$ is the driving laser field, and $U_H = 13.6$ eV and $U_{N_2} = 15.6$ eV are ionization potentials of hydrogen atoms and nitrogen molecules, respectively. Thus,
$${\boldsymbol{E}}_{{\operatorname{THG}}} (t) \propto N_g \left( 1 - \exp \left[ - \int^t {\operatorname{dt}}' w (t') \right] \right) {\boldsymbol{E}} (t) ,$$
accounts for the residual current induced by the external field in air plasma. The Eq. (3) can be further simplified if its expansion is considered only up to the first order, which is essentially the solution when $N(t)$ in $\partial N (t) / \partial t = N_g w (t)$, i.e., where the depletion of the neutral atoms in plasma, $-N(t)w(t)$, as appeared in Eq. (2) is neglected, and such simplification is equivalent to the single atomic ionization,
$${\boldsymbol{E}}_{{\operatorname{THG}}} (t) \propto {\boldsymbol{E}} (t) \int^{t} w (t^{\prime}) {\operatorname{dt}}^{\prime}.$$

Therefore, Eq. (4) is referred to as the single-atom photocurrent (SPC) model.

Compared with the semi-classical SPC model, the THG can also be described under quantum mechanics treatment. The radiation is induced by the time variant dipole moment $ {\boldsymbol {d}} (t) = \langle \Psi (t) | \hat { {\boldsymbol {r}}} | \Psi (t) \rangle = \langle \Psi _0 (t) | \hat {U} (t_0, t) \hat {\mathbf {r}} \hat {U} (t, t_0) | \Psi _0 (t) \rangle$ with the time evolution operator $\hat {U}$ and the initial wave function $| \Psi _0 \rangle$. Using the Dyson series for $\hat {U}$, it is shown that the dipole moment includes three components, $ {\boldsymbol {d}} (t) = {\boldsymbol {d}}^{(0)} (t) + {\boldsymbol {d}}^{(1)} (t) + {\boldsymbol {d}}^{(2)} (t)$. The first term $ {\boldsymbol {d}}^{(0)} (t)$ vanishes in the spherically symmetric system. The second term $ {\boldsymbol {d}}^{(1)} (t) = i \int _{}^t {\operatorname {dt}}' \langle \Psi _0 (t) | \hat {\mathbf {r}} \hat {U} (t, t_{}') \hat {W} (t') | \Psi _0 (t') \rangle + c.c.$ describes the whole process of electron ionization, acceleration in two-color field and recollision to its parent core, corresponding to the CB transition, which are used to explain the HHG. The last term,

$${\boldsymbol{d}}^{(2)} (t) ={-} \int^t {\operatorname{dt}^{\prime\prime}} \int^t {\operatorname{dt}}' \langle \Psi_0 (t^{\prime\prime}) | \hat{W} (t^{\prime\prime}) \hat{U} (t^{\prime\prime}, t) \hat{\mathbf{r}} \hat{U} (t, t_{}') \hat{W} (t') | \Psi_0 (t') \rangle ,$$
takes the form of the CC transition. Since the external light field is intense, the situation enters the scope of the strong field physics and natural choice to tackle the problem is the strong field approximation (SFA) theory. Essentially, the SFA neglects the influence from Coulomb potential of the ionic core. Hence, $\hat {U}$ can be substituted by $\widehat {U^{}}^{(V)}$, the evolution operator of Volkov state which is the eigenstate of an electron in external light alone, to simplify the further derivation. With the radiation produced by the acceleration $\ddot { {\boldsymbol {d}}} (t)$, the THG induced by the CC transition can be evaluated by $\ddot { {\boldsymbol {d}}}^{(2)} (t)$, written as
$$\begin{aligned} {\boldsymbol{E}}_{{\operatorname{THG}}} (t) \propto \ddot{{\boldsymbol{d}}}^{(2)} (t) & = {\boldsymbol{E}} (t) \int_{}^t {\operatorname{dt}}' \int^t {\operatorname{dt}}^{\prime\prime} \eta (t', t^{\prime\prime}) W (t') W^{{\ast}} (t^{\prime\prime}) e^{{\operatorname{iS}}_{{\boldsymbol{k}}', I_{\operatorname{ion}} (t', t^{\prime\prime})}}\\ & \quad - 2{\operatorname{Re}} \int_{}^t {\operatorname{dt}}' \eta (t, t') W (t) W^{{\ast}} (t') [{\boldsymbol{k}}' +{\boldsymbol{A}} (t)] e^{{\operatorname{iS}}_{{\boldsymbol{k}}', I_{\operatorname{ion}}} (t, t')} , \end{aligned}$$
where $W (t) = {\boldsymbol {\mu }} [k + {\boldsymbol {A}} (t)] \cdot {\boldsymbol {E}} (t)$ is the interaction of the electron with the incident field, $\eta (t, t') = \left ( \frac {2 \pi }{i (t - t')} \right )^{3 / 2}$ depicts the diffusion of the electronic wave packet, and $ {\boldsymbol {k}}' = - [ {\boldsymbol {\alpha }} (t') - {\boldsymbol {\alpha }} (t'')] / [t' - t'']$, where $ {\boldsymbol {\alpha }} (t) = \int ^t {\operatorname {dt}}' {\boldsymbol {A}} (t')$ is the excursion of the electron. The ionization rate is related to $\operatorname {S}_{ {\boldsymbol {k}}', I_{\operatorname {ion}}} (t', t'') = \int _{t''}^{t'} {\operatorname {dt}}''' \left [ \frac {1}{2} ( {\boldsymbol {k}}+ {\boldsymbol {A}} (t'''))^2 + I_{\operatorname {ion}} \right ]$ with $I_{\operatorname {ion}}$ is the ionization energy of $\operatorname {N}_2$ . The presence of both $W (t')$ and $W (t'')$ indicates that the two electronic continuum states are involved, and $e^{ {\operatorname {iS}}_{ {\boldsymbol {k}}', I_{\operatorname {ion}}} (t', t'')}$ indicates the joint occurrences of these continuum states created by ionization. The first term of Eq. (6) describes that the THG is the result of transition between continuum states, which start from different ionization times $t'$ and $t''$. The second term emerges when the emission time $t$ approaches the ionization time $t'$, which is referred to as the temporal boundary term. The THG mechanism is investigated based on Eq. (6), which is referred to as the SFA based CC transition (SFA-CC).

4. Results

The electric components of the two-color fields take the form,

$$\begin{aligned} E_p (t) & =E_{\omega}^{} (t) \epsilon_{\omega}^1 \cos (\omega_{} t) + E_{2 \omega}^{} (t - \tau) \epsilon_{2 \omega}^1 \cos [2 \omega_{} (t - \tau) + \theta]\\ E_s (t) & =E_{\omega}^{} (t) \epsilon_{\omega}^2 \sin (\omega_{} t) \pm E_{2 \omega}^{} (t - \tau) \epsilon_{2 \omega}^2 \sin [2 \omega_{} (t - \tau) + \theta] , \end{aligned}$$
where $\theta$ is the initial phase of the SH, and $\tau$ is relative time delay between the FW and SH. The $\sin ^2$-envelop is used for the construction of the two-color field. $\epsilon _{\omega }^1 = \frac {1}{\sqrt {1 + \varepsilon _{\omega }^2}} , \epsilon _{\omega }^2 = \frac {\varepsilon _{\omega }}{\sqrt {1 + \varepsilon _{\omega }^2}}, \epsilon _{2 \omega }^1 = \frac {1}{\sqrt {1 + \varepsilon _{2 \omega }^2}}, \epsilon _{2 \omega }^2 = \frac {\varepsilon _{2 \omega }}{\sqrt {1 + \varepsilon _{2 \omega }^2}},$ where $\varepsilon _{\omega }$ and $\varepsilon _{2 \omega }$ are the ellipticities of the FW and SH, respectively. In the $E_s (t)$ of Eq. (7), the “+” and “$-$” indicate the co-rotating and counter-rotating two-color field, respectively.

The field amplitudes $E_\omega = 5.2\times 10^{8}$ $\rm {V/cm}$, $E_{2\omega } = 2.6\times 10^{8}$ $\rm {V/cm}$. Figure 2(a) shows that the SPC and SFA-CC models have very similar prediction, since the both of them come from the same physical origin, depicting the radiation induced by continuum electron. The intensity of the LHG has rapidly decreased with the increasing harmonic orders, and the plateau in high energy photon region is not found. It is easy to understand that the “hard” recollision predominantly contributes to high energy photons, while the CC transition mainly produces the low energy photons. The LHG yield in the counter-rotating bi-chromatic field is several orders of magnitude larger than that in the co-rotating field, which has been not experimentally validated yet. The 3$n$ order harmonics are suppressed in the counter-rotating field, and only the 3$n \pm$1 order harmonics can exist. The suppression of the 3$n$ order harmonics in the LHG is analogous to that in the HHG [4043], due to the triple rotational symmetry of two-color field and the selection rules [44,45]. The suppression of the 3$n$ order harmonics in the LHG has been experimentally validated in the terahertz region, i.e. 0th order harmonics of the LHG, and our measurement proves that the THG follows the same rule as well. The intensity ratio of the THG in co-rotating field to that in counter-rotating field is about 100:1, as shown in Fig. 2(b). The THG in co-rotating field is analyzed by a polarizer, Fig. 2(c) shows that the intensity of THG is almost unchanged with the rotation of glan-laser polarizer, which is consistent with the theoretical simulation.

 figure: Fig. 2.

Fig. 2. (a) The spectrum of low-order harmonics. The red and blue curves represent the harmonic spectra with the SFA-CC and SPC, respectively. The solid and dashed line represent the co-rotating and counter-rotating field, respectively. (b) Measured intensities of the THG in co-rotating and counter-rotating field. The THG yield in counter-rotating field is much lower than that in co-rotating field. (c) The intensity of the THG in co-rotating field along arbitrary polarization directions in the experiment (red dots) and theory (green line).

Download Full Size | PDF

In order to further study the mechanism of the THG, the THG yields $I$ as a function of the ellipticity of the SH $\varepsilon _{2 \omega }$ and relative time delay $\tau$ are measured when the FW is circular polarized ($\varepsilon _{\omega } = 1$), as shown in Fig. 3(a). The THG is normalized to the maximum value of its the $p$ polarization component. The $\varepsilon _{2\omega }$-dependent THG provides a new dimension of information, which is not accessible only with the $\tau$-dependent measurement. Firstly, it can be found that only at special $\varepsilon$ the intensity modulation along $\tau$ is clearly visible. The modulation of $I_p (\tau )$ is clearly observed at $\varepsilon _{2 \omega } \approx 0.25$, while the modulation of $I_s (\tau )$ is more clear at $\varepsilon _{2 \omega } \approx -0.4$. Although Our laser system is freely running, yet the phase of 2$\omega$ beam is passively locked to the phase of $\omega$ beam. In multi-cycle regime, the THG is not sensitive to the CEP of driving laser, whereas only the relative phase has critical impact on low harmonics. Secondly, the difference between the SPC and SFA-CC model is more distinct when introducing the $\varepsilon$ dimension, as shown in Fig. 3(b) and (c). For instance, the ratios between the maximum value of $I_s$ and $I_p$, $\max (I_s)/\max (I_p)$, predicted by the SPC and SFA-CC model are different: The SFA-CC predicts $\max (I_s)/\max (I_p)\approx 0.37$, while the prediction of the SPC is $\sim 0.13$. The measurement shows that the SFA-CC gives better prediction of $\max (I_s)/\max (I_p)$.

 figure: Fig. 3.

Fig. 3. The distribution of the THG with the phase delay $\tau$ and ellipticity of the SH $\varepsilon _{2\omega }$ when the FW is circular polarized ($\varepsilon _{\omega } = 1$). Upper row: The $p$-polarized $I_p (\tau , \varepsilon _{2\omega })$; Lower row: The $s$-polarized $I_s (\tau , \varepsilon _{2\omega })$. The $I_p$ and $I_s$ obtained from the measurement (a) and the SFA-CC model (b), the SPC model (c).

Download Full Size | PDF

Considering the THG distribution via the phase delay $\tau$ and the SH ellipticity $\varepsilon _{\omega }$ when the SH is circular polarized ($\varepsilon _{2 \omega } = 1$), as shown in Fig. 4, the $I_s (\tau , \varepsilon _{2 \omega })$ predicted by the SFA-CC [Fig. 4(b2)] and SPC [Fig. 4(c2)] show totally different patterns. Compared with the measurement [Fig. 4(a2)], the SFA-CC gives better agreement with the $\varepsilon _{\omega }$-dependent yield of the THG. The contrast of the THG modulation along the $\tau$ axis can not be fully reproduced by the theoretical models. There are several possible reasons as follows: Firstly, the grating spectrometer is an incoherence detection method, the incoherence component of the radiation may decrease the contrast of the modulation. Secondly, the Coulomb potential of the ionic core may have a nonnegligible impact on the THG [46]. Thirdly, other mechanisms, for instance the bound-bound transition [29,30,32,33], may contribute to the THG, which results in the background of the measurement.

 figure: Fig. 4.

Fig. 4. The distribution of the THG with the phase delay $\tau$ and ellipticity fo the FW $\varepsilon _{\omega }$ when the SH is circular polarization ($\varepsilon _{2 \omega } = 1$). Upper row: The $p$-polarized $I_p (\tau , \varepsilon _{2 \omega })$; Lower row: The $s$-polarized $I_s (\tau , \varepsilon _{2 \omega })$. The $I_p$ and $I_s$ obtained from the measurement(a) and theoretical models of the SFA-CC (b), the SPC (c).

Download Full Size | PDF

The Fig. 5 shows the THG intensity versus the intensity ratio of two color fields $I_{\omega } / I_{2 \omega }$. The intensity of the SH was fixed at 0.56 mJ/pulse, and the intensity of the FW was changed in the experiment. The experimental result shows the intensity of the THG increases rapidly with the increasing $I_{\omega } / I_{2 \omega }$. Comparatively, the SFA-CC well reproduces the experimental results, while the THG predicted by the SPC saturates at $I_{\omega } / I_{2 \omega } = 1.5$.

 figure: Fig. 5.

Fig. 5. The THG yield as a function of the ratio between the FW and SH intensities in co-rotating bi-chromatic field. The experiment (black scatter), the SFA-CC (red curve) and the SPC (blue curve) are presented.

Download Full Size | PDF

5. Discussions

Comparing the experiment and theory, we find that both the SPC and SFA-CC model can roughly replicate the measurement. Nevertheless, there are some discrepancies between the SPC and SFA-CC model in details, especially the dependence of ellipticity. In the following, comparing between the analytic formulas, the similarity and difference of two models are discussed.

In PC model, the depletion of the neutral atoms can be neglected, and the PC model can be simplified into the form of Eq. (4), referred to as the SPC model. As shown in Fig. 6, the approximation in SPC model does not significantly modify the THG yields $I_p(\varepsilon _{2\omega }=1,\tau = 0)$ and $I_s(\varepsilon _{2\omega }=1,\tau = 0)$.

 figure: Fig. 6.

Fig. 6. The THG yields $I_p(\varepsilon _{2\omega }=1,\tau = 0)$ and $I_s(\varepsilon _{2\omega }=1,\tau = 0)$ predicted by PC model, SPC model, $\boldsymbol {a}_1(t)$ and $\boldsymbol {a}_1(t) + \boldsymbol {a}_2(t)$ in SFA-CC model.

Download Full Size | PDF

In the SFA-CC model, the Eq. (6) can be written as $\ddot {\boldsymbol {d}}^{(2)} (t) = \boldsymbol {a}_1(t)+\boldsymbol {a}_2(t)$, where $\boldsymbol {a}_1(t)$ and $\boldsymbol {a}_2(t)$ correspond to the first and second terms in Eq. (6). The $\boldsymbol {a}_2(t)$ only contributes to radiation when the emission time $t$ approaches the ionization time $t'$, which can always be neglected due to the minor contribution [35]. The Fig. 6 shows that the $\boldsymbol {a}_1(t)$ dominates the THG, while $\boldsymbol {a}_2(t)$ can be safely neglected. Hence, the Eq. (6) can be approximated and rewritten as a new form,

$$\ddot{\boldsymbol{d}}^{(2)} (t) \approx \boldsymbol{a}_1(t) ={\boldsymbol{E}} (t) \int_{}^t {\operatorname{dt}}' \int^t {\operatorname{dt}}^{\prime\prime} \eta (t', t^{\prime\prime}) W (t') W^{{\ast}} (t^{\prime\prime}) e^{{\operatorname{iS}}_{{\boldsymbol{k}}', I_{\operatorname{ion}} (t', t^{\prime\prime})}}$$
$$={\boldsymbol{E}} (t) \int^t w' (t',t) {\operatorname{dt}},$$
where the $w'(t',t) = \int ^t {\operatorname {dt}}'' \eta (t', t'') W (t') W^{\ast } (t'') e^{ {\operatorname {iS}}_{ {\boldsymbol {k}}', I_{\operatorname {ion}} (t', t'')}}$. Thus, the Eq. (9) derived from the SFA-CC model and the Eq. (4) derived from the SPC model have analogous forms.

Since $w'(t',t)$ in Eq. (9) is roughly independent of $t$ [35], $w'(t',t)$ versus the ionization instant $t'$ can be directly compared with the $w(t')$ in Eq. (4).

As shown in Fig. 7 (a) and (b), $w(t')$ and $w'(t')$ have similar distributions along the ionization time $t'$, manifesting the correspondence between the semi-classical and quantum theories. Nevertheless, there are minute differences, leading to the discrepancies between SPC and FA-CC models. As shown in Fig. 7(c), $w'(t')$ gives apparently larger value than $w(t')$ at $\varepsilon _{\omega }=0$, resulting in different patterns predicted by the SPC and SFA model in Fig. 4 (b) and (c).

 figure: Fig. 7.

Fig. 7. The distributions of $w(t')$ in the SPC model (Eq. (4)) and $w'(t')$ in the SFA-CC model (Eq. (9)) when $\varepsilon _{2\omega }=1$ and $\tau = 0$ fs. (a) and (b), $w(t')$ and $w'(t',t\rightarrow \infty )$ versus the ellipticity of $\omega$ beam. (c), $w(t')$ and $w'(t',t\rightarrow \infty )$ at $\varepsilon _{\omega }=0$. Gray line, $|E(t)|=\sqrt {E_p^2+E_s^2}$; Green line, $w'(t', t\rightarrow \infty )$ in the SFA-CC; Red line, $w(t')$ in the SPC model.

Download Full Size | PDF

Both $w'(t')$ and $w(t')$ represent the weights of the electron released at ionization instant $t'$ on the radiation at emission time $t$, yet they have different physical meanings in the two models. In the SPC model, $w(t')$ is simply estimated by quasi-static tunneling rate under strong electric field. In contrast, $w'(t')$ in SFA-CC model derives from quantum mechanical treatment, which describes the overlapping between two quantum paths of a single electron released at different ionization instants $t'$ and $t''$, resulting in the coherence between the corresponding continuum states and yielding the radiation. The better agreement to the experimental results based on SFA-CC model indicates that the quantum effect may be significant in the THG.

6. Conclusion

In summary, the THG is experimentally and theoretically investigated by controlling the phase delay and ellipticity of the bi-chromatic laser field. The THG is found to be strong suppressed in the counter-rotating bi-chromatic fields, which can be explained with the selection rules of harmonic emissions. Furthermore, the dependency of the THG on the phase delay and ellipticity of bi-chromatic fields is experimentally acquired. The THG yields are periodically modulated along the phase delay, and highly dependent upon the ellipticity of the laser field. The THG as a function of the phase delay and ellipticity can be be roughly reproduced by both the photocurrent model and SFA-CC model, but the SFA-CC model gives better prediction of the elliptical dependency of the THG. We find that after formula simplification the photocurrent and SFA-CC model have similar analytic forms, but the overlapping between the continuum states via different quantum paths released at distinct ionization instants is not included in quasi-static ionization rate in photocurrent model, which leads to the slightly different prediction of elliptically-dependent THG. Our research provides a new perspective for further understanding of the mechanism of the LHG in gases and solid mediums.

Funding

National Natural Science Foundation of China (11827806, 11874368).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. D. Jang, R. M. Schwartz, D. Woodbury, J. Griff-McMahon, A. H. Younis, H. M. Milchberg, and K.-Y. Kim, “Efficient terahertz and Brunel harmonic generation from air plasma via mid-infrared coherent control,” Optica 6(10), 1338 (2019). [CrossRef]  

2. M. Kolesik, E. M. Wright, and J. V. Moloney, “Supercontinuum and third-harmonic generation accompanying optical filamentation as first-order scattering processes,” Opt. Lett. 32(19), 2816–2818 (2007). [CrossRef]  

3. A. V. Mitrofanov, A. A. Voronin, S. I. Mitryukovskiy, D. A. Sidorov-Biryukov, A. Pugzlys, G. Andriukaitis, T. Flory, E. A. Stepanov, A. B. Fedotov, A. Baltuska, and A. M. Zheltikov, “Mid-infrared-to-mid-ultraviolet supercontinuum enhanced by third-to-fifteenth odd harmonics,” Opt. Lett. 40(9), 2068 (2015). [CrossRef]  

4. A. V. Mitrofanov, D. A. Sidorov-Biryukov, M. M. Nazarov, A. A. Voronin, M. V. Rozhko, A. D. Shutov, S. V. Ryabchuk, E. E. Serebryannikov, A. B. Fedotov, and A. M. Zheltikov, “Ultraviolet-to-millimeter-band supercontinua driven by ultrashort mid-infrared laser pulses,” Optica 7(1), 15 (2020). [CrossRef]  

5. Y. Zhang, T.-M. Yan, and Y. Jiang, “Ultrafast Mapping of Coherent Dynamics and Density Matrix Reconstruction in a Terahertz-Assisted Laser Field,” Phys. Rev. Lett. 121(11), 113201 (2018). [CrossRef]  

6. H. Chen, X. Wen, J. Zhang, T. Wu, Y. Gong, X. Zhang, J. Yuan, C. Yi, J. Lou, P. M. Ajayan, W. Zhuang, G. Zhang, and J. Zheng, “Ultrafast formation of interlayer hot excitons in atomically thin MoS2/WS2 heterostructures,” Nat. Commun. 7(1), 12512 (2016). [CrossRef]  

7. U. Elu, L. Maidment, L. Vamos, F. Tani, D. Novoa, M. H. Frosz, V. Badikov, D. Badikov, V. Petrov, P. St. J. Russell, and J. Biegert, “Seven-octave high-brightness and carrier-envelope-phase-stable light source,” Nat. Photonics 15(4), 277–280 (2021). [CrossRef]  

8. E. Goulielmakis, V. S. Yakovlev, A. L. Cavalieri, M. Uiberacker, V. Pervak, A. Apolonski, R. Kienberger, U. Kleineberg, and F. Krausz, “Attosecond Control and Measurement: Lightwave Electronics,” Science 317(5839), 769–775 (2007). [CrossRef]  

9. S. Witte, “Deep-Ultraviolet Quantum Interference Metrology with Ultrashort Laser Pulses,” Science 307(5708), 400–403 (2005). [CrossRef]  

10. D. C. Yost, T. R. Schibli, J. Ye, J. L. Tate, J. Hostetter, M. B. Gaarde, and K. J. Schafer, “Vacuum-ultraviolet frequency combs from below-threshold harmonics,” Nat. Phys. 5(11), 815–820 (2009). [CrossRef]  

11. R. T. Zinkstok, S. Witte, W. Ubachs, W. Hogervorst, and K. S. Eikema, “Frequency comb laser spectroscopy in the vacuum-ultraviolet region,” Phys. Rev. A 73(6), 061801 (2006). [CrossRef]  

12. G. Ndabashimiye, S. Ghimire, M. Wu, D. A. Browne, K. J. Schafer, M. B. Gaarde, and D. A. Reis, “Solid-state harmonics beyond the atomic limit,” Nature 534(7608), 520–523 (2016). [CrossRef]  

13. S. Ghimire and D. A. Reis, “High-harmonic generation from solids,” Nat. Phys. 15(1), 10–16 (2019). [CrossRef]  

14. G. Vampa, T. J. Hammond, N. Thire, B. E. Schmidt, F. Legare, C. R. McDonald, T. Brabec, and P. B. Corkum, “Linking high harmonics from gases and solids,” Nature 522(7557), 462–464 (2015). [CrossRef]  

15. S. Chen, G. Li, F. Zeuner, W. H. Wong, E. Y. B. Pun, T. Zentgraf, K. W. Cheah, and S. Zhang, “Symmetry-Selective Third-Harmonic Generation from Plasmonic Metacrystals,” Phys. Rev. Lett. 113(3), 033901 (2014). [CrossRef]  

16. P. Jurgens, “Origin of strong-field-induced low-order harmonic generation in amorphous quartz,” Nat. Phys. 16(10), 1035–1039 (2020). [CrossRef]  

17. P. B. Corkum, “Plasma perspective on strong field multiphoton ionization,” Phys. Rev. Lett. 71(13), 1994–1997 (1993). [CrossRef]  

18. F. Brunel, “Harmonic generation due to plasma effects in a gas undergoing multiphoton ionization in the high-intensity limit,” Journal of the Optical Society of America B 7(4), 521 (1990). [CrossRef]  

19. A. Nath, J. A. Dharmadhikari, D. Mathur, and A. K. Dharmadhikari, “Enhanced third harmonic generation in air by two-colour ultrashort laser pulses,” Appl. Phys. B 122(9), 248 (2016). [CrossRef]  

20. S. Suntsov, D. Abdollahpour, D. G. Papazoglou, and S. Tzortzakis, “Filamentation-induced third-harmonic generation in air via plasma-enhanced third-order susceptibility,” Phys. Rev. A 81(3), 033817 (2010). [CrossRef]  

21. H. Xiong, H. Xu, Y. Fu, Y. Cheng, Z. Xu, and S. L. Chin, “Spectral evolution of angularly resolved third-order harmonic generation by infrared femtosecond laser-pulse filamentation in air,” Phys. Rev. A 77(4), 043802 (2008). [CrossRef]  

22. V. A. Kostin and N. V. Vvedenskii, “Mutual Enhancement of Brunel Harmonics,” JETP Lett. 110(7), 457–463 (2019). [CrossRef]  

23. G. Lambert, A. Andreev, J. Gautier, L. Giannessi, V. Malka, A. Petralia, S. Sebban, S. Stremoukhov, F. Tissandier, B. Vodungbo, and P. Zeitoun, “Spatial properties of odd and even low order harmonics generated in gas,” Sci. Rep. 5(1), 7786 (2015). [CrossRef]  

24. K.-Y. Kim, J. H. Glownia, A. J. Taylor, and G. Rodriguez, “Terahertz emission from ultrafast ionizing air in symmetry-broken laser fields,” Opt. Express 15(8), 4577 (2007). [CrossRef]  

25. V. Kostin, I. Laryushin, A. Silaev, and N. Vvedenskii, “Ionization-Induced Multiwave Mixing: Terahertz Generation with Two-Color Laser Pulses of Various Frequency Ratios,” Phys. Rev. Lett. 117(3), 035003 (2016). [CrossRef]  

26. A. A. Silaev, V. A. Kostin, I. D. Laryushin, and N. V. Vvedenskii, “Ionization Mechanism of the Generation of Tunable Ultrashort Pulses in the Mid-Infrared Range,” JETP Lett. 107(3), 151–156 (2018). [CrossRef]  

27. V. Kostin and N. Vvedenskii, “Generation of Few- and Subcycle Radiation in Midinfrared-to-Deep-Ultraviolet Range During Plasma Production by Multicolor Femtosecond Pulses,” Phys. Rev. Lett. 120(6), 065002 (2018). [CrossRef]  

28. D. Ma, L. Dong, M. Zhang, T. Wu, Y. Zhao, L. Zhang, and C. Zhang, “Enhancement of terahertz waves from two-color laser-field induced air plasma excited using a third-color femtosecond laser,” Opt. Express 28(14), 20598 (2020). [CrossRef]  

29. M. Bache, F. Eilenberger, and S. Minardi, “Higher-order kerr effect and harmonic cascading in gases,” Opt. Lett. 37(22), 4612–4614 (2012). [CrossRef]  

30. C. Rodríguez, Z. Sun, Z. Wang, and W. Rudolph, “Characterization of laser-induced air plasmas by third harmonic generation,” Opt. Express 19(17), 16115–16125 (2011). [CrossRef]  

31. H. Xu, W. Chu, Y. Liu, W. Liu, H. Xiong, Y. Fu, J. Yao, B. Zeng, J. Ni, S. L. Chin, Y. Cheng, and Z. Xu, “Third-harmonic generation in relative-phase-controlled two-color laser field,” Appl. Phys. B 104(4), 909–912 (2011). [CrossRef]  

32. U. Sapaev, A. Husakou, and J. Herrmann, “Combined action of the bound-electron nonlinearity and the tunnel-ionization current in low-order harmonic generation in noble gases,” Opt. Express 21(21), 25582 (2013). [CrossRef]  

33. E. Serebryannikov and A. Zheltikov, “Quantum and Semiclassical Physics behind Ultrafast Optical Nonlinearity in the Midinfrared: The Role of Ionization Dynamics within the Field Half Cycle,” Phys. Rev. Lett. 113(4), 043901 (2014). [CrossRef]  

34. Z. Zhou, D. Zhang, Z. Zhao, and J. Yuan, “Terahertz emission of atoms driven by ultrashort laser pulses,” Phys. Rev. A 79(6), 063413 (2009). [CrossRef]  

35. K. Zhang, Y. Zhang, X. Wang, T.-M. Yan, and Y. H. Jiang, “Continuum electron giving birth to terahertz emission,” Photonics Res. 8(6), 760 (2020). [CrossRef]  

36. K. Zhang, Y. Zhang, X. Wang, Z. Shen, T.-M. Yan, and Y. H. Jiang, “Experimental evidence for terahertz emission of continuum electrons in the dual-color laser field,” Opt. Lett. 45(7), 1838–1841 (2020). [CrossRef]  

37. C. Meng, W. Chen, X. Wang, Z. Lu, Y. Huang, J. Liu, D. Zhang, Z. Zhao, and J. Yuan, “Enhancement of terahertz radiation by using circularly polarized two-color laser fields,” Appl. Phys. Lett. 109(13), 131105 (2016). [CrossRef]  

38. P. B. Corkum, N. H. Burnett, and F. Brunel, “Above-threshold ionization in the long-wavelength limit,” Phys. Rev. Lett. 62(11), 1259–1262 (1989). [CrossRef]  

39. M. Kreb, T. Loffler, M. D. Thomson, R. Dorner, H. Gimpel, K. Zrost, T. Ergler, R. Moshammer, U. Morgner, J. Ullrich, and H. G. Roskos, “Determination of the carrier-envelope phase of few-cycle laser pulses with terahertz-emission spectroscopy,” Nat. Phys. 2(5), 327–331 (2006). [CrossRef]  

40. A. Fleischer, O. Kfir, T. Diskin, P. Sidorenko, and O. Cohen, “Spin angular momentum and tunable polarization in high-harmonic generation,” Nat. Photonics 8(7), 543–549 (2014). [CrossRef]  

41. N. Saito, P. Xia, F. Lu, T. Kanai, J. Itatani, and N. Ishii, “Observation of selection rules for circularly polarized fields in high-harmonic generation from a crystalline solid,” Optica 4(11), 1333 (2017). [CrossRef]  

42. O. Kfir, P. Grychtol, E. Turgut, R. Knut, D. Zusin, D. Popmintchev, T. Popmintchev, H. Nembach, J. M. Shaw, A. Fleischer, H. Kapteyn, M. Murnane, and O. Cohen, “Generation of bright phase-matched circularly-polarized extreme ultraviolet high harmonics,” Nat. Photonics 9(2), 99–105 (2015). [CrossRef]  

43. D. D. Hickstein, F. J. Dollar, P. Grychtol, J. L. Ellis, R. Knut, C. Hernandez-Garcia, D. Zusin, C. Gentry, J. M. Shaw, T. Fan, K. M. Dorney, A. Becker, A. Jaron-Becker, H. C. Kapteyn, M. M. Murnane, and C. G. Durfee, “Non-collinear generation of angularly isolated circularly polarized high harmonics,” Nat. Photonics 9(11), 743–750 (2015). [CrossRef]  

44. O. Neufeld, D. Podolsky, and O. Cohen, “Floquet group theory and its application to selection rules in harmonic generation,” Nat. Commun. 10(1), 405 (2019). [CrossRef]  

45. O. E. Alon, V. Averbukh, and N. Moiseyev, “Selection Rules for the High Harmonic Generation Spectra,” Phys. Rev. Lett. 80(17), 3743–3746 (1998). [CrossRef]  

46. W.-H. Xiong, J.-W. Geng, J.-Y. Tang, L.-Y. Peng, and Q. Gong, “Mechanisms of Below-Threshold Harmonic Generation in Atoms,” Phys. Rev. Lett. 112(23), 233001 (2014). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (7)

Fig. 1.
Fig. 1. (a) Schematics of the experimental set-up. DM: dichroic mirror; BBO: $\beta$-phase barium borate crystal; L: lens; GLP: glan-laser polarizer; Piezo: piezo-stage; (b) The co-rotating (green) or counter-rotating (purple) bi-chromatic field combined with circularly polarized FW (red) and SH (blue) beams.
Fig. 2.
Fig. 2. (a) The spectrum of low-order harmonics. The red and blue curves represent the harmonic spectra with the SFA-CC and SPC, respectively. The solid and dashed line represent the co-rotating and counter-rotating field, respectively. (b) Measured intensities of the THG in co-rotating and counter-rotating field. The THG yield in counter-rotating field is much lower than that in co-rotating field. (c) The intensity of the THG in co-rotating field along arbitrary polarization directions in the experiment (red dots) and theory (green line).
Fig. 3.
Fig. 3. The distribution of the THG with the phase delay $\tau$ and ellipticity of the SH $\varepsilon _{2\omega }$ when the FW is circular polarized ($\varepsilon _{\omega } = 1$). Upper row: The $p$-polarized $I_p (\tau , \varepsilon _{2\omega })$; Lower row: The $s$-polarized $I_s (\tau , \varepsilon _{2\omega })$. The $I_p$ and $I_s$ obtained from the measurement (a) and the SFA-CC model (b), the SPC model (c).
Fig. 4.
Fig. 4. The distribution of the THG with the phase delay $\tau$ and ellipticity fo the FW $\varepsilon _{\omega }$ when the SH is circular polarization ($\varepsilon _{2 \omega } = 1$). Upper row: The $p$-polarized $I_p (\tau , \varepsilon _{2 \omega })$; Lower row: The $s$-polarized $I_s (\tau , \varepsilon _{2 \omega })$. The $I_p$ and $I_s$ obtained from the measurement(a) and theoretical models of the SFA-CC (b), the SPC (c).
Fig. 5.
Fig. 5. The THG yield as a function of the ratio between the FW and SH intensities in co-rotating bi-chromatic field. The experiment (black scatter), the SFA-CC (red curve) and the SPC (blue curve) are presented.
Fig. 6.
Fig. 6. The THG yields $I_p(\varepsilon _{2\omega }=1,\tau = 0)$ and $I_s(\varepsilon _{2\omega }=1,\tau = 0)$ predicted by PC model, SPC model, $\boldsymbol {a}_1(t)$ and $\boldsymbol {a}_1(t) + \boldsymbol {a}_2(t)$ in SFA-CC model.
Fig. 7.
Fig. 7. The distributions of $w(t')$ in the SPC model (Eq. (4)) and $w'(t')$ in the SFA-CC model (Eq. (9)) when $\varepsilon _{2\omega }=1$ and $\tau = 0$ fs. (a) and (b), $w(t')$ and $w'(t',t\rightarrow \infty )$ versus the ellipticity of $\omega$ beam. (c), $w(t')$ and $w'(t',t\rightarrow \infty )$ at $\varepsilon _{\omega }=0$. Gray line, $|E(t)|=\sqrt {E_p^2+E_s^2}$; Green line, $w'(t', t\rightarrow \infty )$ in the SFA-CC; Red line, $w(t')$ in the SPC model.

Equations (9)

Equations on this page are rendered with MathJax. Learn more.

E THG ( t ) j ( t ) t = e 2 m N ( t ) E ( t ) .
N ( t ) / t = [ N g N ( t ) ] w ( t ) ,
E THG ( t ) N g ( 1 exp [ t dt w ( t ) ] ) E ( t ) ,
E THG ( t ) E ( t ) t w ( t ) dt .
d ( 2 ) ( t ) = t dt t dt Ψ 0 ( t ) | W ^ ( t ) U ^ ( t , t ) r ^ U ^ ( t , t ) W ^ ( t ) | Ψ 0 ( t ) ,
E THG ( t ) d ¨ ( 2 ) ( t ) = E ( t ) t dt t dt η ( t , t ) W ( t ) W ( t ) e iS k , I ion ( t , t ) 2 Re t dt η ( t , t ) W ( t ) W ( t ) [ k + A ( t ) ] e iS k , I ion ( t , t ) ,
E p ( t ) = E ω ( t ) ϵ ω 1 cos ( ω t ) + E 2 ω ( t τ ) ϵ 2 ω 1 cos [ 2 ω ( t τ ) + θ ] E s ( t ) = E ω ( t ) ϵ ω 2 sin ( ω t ) ± E 2 ω ( t τ ) ϵ 2 ω 2 sin [ 2 ω ( t τ ) + θ ] ,
d ¨ ( 2 ) ( t ) a 1 ( t ) = E ( t ) t dt t dt η ( t , t ) W ( t ) W ( t ) e iS k , I ion ( t , t )
= E ( t ) t w ( t , t ) dt ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.