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

Macroscopic effects in high-order harmonic generation – a focal-averaging method based on the integral solution of the wave equation

Open Access Open Access

Abstract

A macroscopic theory of high-order harmonic generation (HHG) is presented, which applies a focal-averaging method based on the integral solution of the wave equation. The macroscopic high-harmonic yield is the coherent superposition of the single-atom contributions of all atoms of the generating medium, which are positioned at different spatial points of the laser focus and exposed to the space-time-dependent laser pulse. The HHG spectrum obtained in our macroscopic simulations is qualitatively different from the one obtained using the microscopic or single-atom theory of HHG. Coherent intensity focal averaging, the simpler and more approximate of two methods we introduced, gives the spectrum which forms a declining plateau with the same cutoff position as that of the microscopic spectrum. The second, more precise method, which we call coherent spatio-temporal focal averaging, shows that it is possible, changing the macroscopic conditions, to obtain an observable peak in the harmonic spectrum at an energy much lower than the microscopic cutoff energy. Generally, the high-harmonic yield appears to be dominated by the contributions of laser-pulse spatio-temporal regions with lower intensities as well as by interference, so that the high-energy plateau and its sharp cutoff are quenched in the theoretical simulation and, presumably, in the experiment. The height and position of this peak strongly depend on the macroscopic conditions. We confirmed these findings by applying our macroscopic theory to simulate two recent experiments with mid-infrared laser fields, one with a linearly polarized field and the other one with a bicircular field.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

When we expose atoms to an intense laser field various nonlinear processes can occur [1]. One of the most important of these processes is the emission of extreme ultraviolet and soft x-ray radiation. This process was discovered in the seminal experimental works [2,3] and is named high-order harmonic generation (HHG). The energy of the emitted soft x-rays can be a few thousand times larger than the energy of the generating fundamental laser field photons. The HHG provides a table-top source of coherent and ultrashort soft x-ray radiation and has found many applications [410].

On the single-atom or microscopic level the HHG process is well understood. It is described by the semiclassical three-step model [1114]. In the first step the bound atomic electron tunnels through the atomic potential barrier which is lowered under the influence of the strong laser field. This electron is then driven by the laser field and, since the laser field is oscillatory and changes its sign, can return to the parent ion (second step). In the final third step, this electron recombines with the parent ion and a high harmonic photon is emitted. The maximal energy of this harmonic photon is equal to the ionization potential $I_p$ plus the maximal kinetic energy accumulated during the electron’s travel in the laser field from ionization to recombination [for a linearly polarized laser field this energy is $3.17U_p$, where $U_p=I/(4\omega ^{2})$ is the ponderomotive energy (cycle-averaged electron kinetic energy in the laser field; $I$ and $\omega$ are the laser field intensity and frequency, respectively)]. In the microscopic theory the intensity of the emitted high harmonics can be calculated using solutions of the time-dependent Schrödinger equation (see, for example, [7] and references therein). A simpler and less time consuming way to calculate the microscopic HHG spectra is to use the strong-field approximation, an approximate theory which neglects the influence of the atomic potential in the propagation step of the HHG process [14]. In our paper we will use a version of the strong-field approximation presented in [1].

Macroscopic effects in HHG are usually considered solving Maxwell’s wave equations for the fundamental and harmonic fields [410]. In the cited references the Maxwell wave equation is used to describe the macroscopic propagation and phase matching of the radiation. The microscopic quantities enter Maxwell’s wave equation as source terms. There are many different attempts to evaluate macroscopic effects in HHG. They can be divided into two groups. One is based on differential equations and the corresponding propagation. The three-dimensional propagation is described in detail in [8,9]. The relevant differential equations are solved using a Crank-Nicolson routine. The second group is based on integral solutions of the wave equation. We will use this approach. Such integral solutions were considered at the beginning of the development of the theory of macroscopic HHG [5] and were used to estimate the effect of phase matching (see also the more recent Refs. [15,16]; however, the HHG spectra are not presented in these references). It should be mentioned that in [17] a different propagation technique, based on the discrete dipole approximation and an integral solution of the wave equation, was used. Finally, in spite of the fact that macroscopic HHG is a well explored problem, its complexity and its dependence on many different parameters make this problem difficult and open for new discoveries. In this context, we mention here a recent study of the coherence in macroscopic HHG [18]. Our approach, which is based on integral solutions of the wave equation for the harmonic field, is particularly suitable for longer wavelengths and higher intensities of the driving laser field for which other computational methods may be too computer demanding.

In Sec. 2 we present formulas for the microscopic HHG theory and introduce our method of macroscopic focal averaging. Numerical results are presented in Sec. 3 and the conclusions are given in Sec. 4. We use atomic units.

2. Theory

2.1 Microscopic HHG

For a laser field defined in the $xy$ plane, spanned by the unit vectors $\hat {\mathbf {e}}_x$ and $\hat {\mathbf {e}}_y$, the $T$-matrix element is the vector $\mathbf {T}_n=|T_n|\mathbf {e}_n=T_{nx}\hat {\mathbf {e}}_x+T_{ny}\hat {\mathbf {e}}_y$, with $\mathbf {e}_n$ the complex unit polarization vector of the $n$th harmonic and $|T_n|=\mathbf {e}^{*}_n\cdot \mathbf {T}_n$. The vector $\mathbf {T}_n$ is the Fourier component of the time-dependent dipole matrix element between the initial and final laser-dressed atomic states. It can be approximated by [1,19]

$$\mathbf{T}_n=\int_0^{T} \frac{d t}{T}\sum_a\mathbf{d}_a(t)e^{i n\omega t},$$
with $T=2\pi /\omega$ the period and $\omega$ the fundamental frequency of the laser field. The sum in Eq. (1) is over all possible values $a$ of the atomic ground-state quantum number. The quantum-mechanical method of calculation of the vector $\mathbf {d}_a(t)$, based on the strong-field approximation, is described in [19] and references therein. It can be written as
$$\mathbf{d}_a(t)={-}i\int_{-\infty}^{t}dt_0\mathbf{m}^{*}_a(t)\left(\frac{2\pi}{i\tau}\right)^{3/2}e^{iS(t_0,t)}\mathbf{m}_a(t_0)\cdot\mathbf{E}(t_0),$$
where $\tau =t-t_0$, $\mathbf {m}_a(t)=\langle \mathbf {p}_s+\mathbf {A}(t)|\mathbf {r}|\psi _a\rangle$ is the dipole matrix element between the ground-state wave function $|\psi _a\rangle$ and the plane wave ket vector $|\mathbf {p}_s+\mathbf {A}(t)\rangle$, $\langle \mathbf {r}|\mathbf {q}\rangle =(2\pi )^{-3/2}e^{i\mathbf {q}\cdot \mathbf {r}}$, $\mathbf {p}_s=-\int _{t_0}^{t}dt'\mathbf {A}(t')/\tau$ is the stationary momentum, $\mathbf {E}(t)=-d\mathbf {A}(t)/dt$ is the electric field vector, and $S(t_0,t)=-{I_p}\tau -\int _{t_0}^{t}\left [\mathbf {p}_s+\mathbf {A}(t')\right ]^{2}/2$ is the action. The above result can be described by the (semi)classical three-step model [1114] (ionization followed by propagation and recombination with emission of a high harmonic). The matrix element $\mathbf {m}_a(t_0)$ corresponds to the ionization at the time $t_0$, the action is responsible for the propagation of the electron wave packet from the ionization time $t_0$ to the recombination time $t$ (with the wave-packet spreading factor proportional to $\tau ^{-3/2}$), the matrix element $\mathbf {m}^{*}_a(t)$ describes the recombination at the time $t$, and the emission of the $n$th harmonic is in accordance with Eq. (1).

The $n$th-harmonic intensity (power) is

$$P_n=\frac{(n\omega)^{4}}{2\pi c^{3}}\left|T_n\right|^{2},$$
while the $n$th-harmonic ellipticity can be calculated using the $T$-matrix element as described in [20].

2.2 Gaussian laser beam and the gas cell

In order to take into account the macroscopic effects we should define the macroscopic spatial dependence of the laser field. We suppose that the laser field can be described by a Gaussian beam, which depends on the spatial cylindrical coordinates $(\rho,z)$ [due to cylindrical symmetry the laser field does not depend on the azimuth angle $\varphi$ of cylindrical coordinates $\mathbf {r}_m =(\rho,\varphi,z)$; the subscript “$m$” stands for the medium in which high harmonics are generated]. The minimum beam waist is $w_0$ and the Rayleigh length is $z_0=\pi w_0^{2}/\lambda$ ($\lambda$ is the laser wavelength). For the peak intensity $I_0=E_0^{2}$, the laser intensity is [21]

$$I(\rho,z)=\frac{I_0}{1+z^{2}/z_0^{2}}\exp\left[-\frac{2\rho^{2}}{w_0^{2}(1+z^{2}/z_0^{2})}\right].$$

The corresponding electric field is

$$\begin{array}{rll} E(\rho,z,t)&=&|E(\rho,z)|\,\mathrm{Re}\left(e^{i[\varphi_E(\rho,z)-\omega t]}\right),\\ E(\rho,z)&=&\frac{E_0z_0}{z_0+iz}\exp\left[{-}i\frac{k\rho^{2}}{2(z_0+iz)}+ikz\right], \end{array}$$
with $|E(\rho,z)|^{2}=I(\rho,z)$ and the phase [2123]
$$\varphi_E(\rho,z)=kz-\psi_G(z)+k\rho^{2}/(2R),$$
where $k=|\mathbf {k}|$, $R(z)=z(1+z_0^{2}/z^{2})$ is the local radius of curvature, and $\psi _G(z)=\arctan (z/z_0)$ is the Gouy phase.

The above formula for the laser intensity can be generalized to include the laser pulse shape. For a Gaussian pulse with the temporal width $\tau _g$, we have $I(\rho,z,t_p)=I(\rho,z)f_p(t_p-|\mathbf {r}_m|/c)$ where $f_p(t)=\exp (-t^{2}/\tau _g^{2})$. In this case, the electric field is defined by $E(\rho,z,t_p,t)=|E(\rho,z,t_p)|\mathrm {Re}\left (\exp \{i[\varphi _E(\rho,z)-\omega t]\}\right )$, with $|E(\rho,z,t_p)|^{2}=I(\rho,z,t_p)$. In the experiment the pulse duration time is defined as FWHM (full width at the half maximum) of the intensity profile: $\tau _p=2\sqrt {\ln 2}\tau _g$.

In Fig. 1 we present the gas cell of the length $L$ (blue rectangle) and the focused Gaussian laser beam (area between the red curves). The focus F is at the point $z=0$, while the center of the gas cell is shifted by $\Delta z$ with respect to the focus ($\Delta z$ is negative for the presented example). The gas cell is positioned between the points $z_1=-L/2 +\Delta z$ and $z_2=L/2 +\Delta z$. We also indicated examples of a point $\mathbf {r}_m=(\rho,z)$ in the medium where a high harmonic is emitted and a point $\mathbf {r}_d=(\rho _d,z_d)$ in the detector plane D where high harmonics are detected. We suppose that the harmonic photons are detected in the far field, i.e., $z_d$ is much larger than the coordinates in the medium. Another important experimental parameter is the gas pressure $p$. It is not explicitly taken into account in our simulations. But we expect that the gas density increases linearly with the increase of this pressure. Higher is this density, more atomic centers are available for harmonic generation and we expect that the high harmonic yield increases proportionally to $p^{2}$. For higher pressures this dependence becomes nonlinear [24], or even constant [25], since various effects, like plasma effects, saturation, excitation, incoherent processes [26], etc., influence the HHG process.

 figure: Fig. 1.

Fig. 1. Schematic presentation of the gas cell, focused Gaussian beam, and the detector plane in the HHG experiment.

Download Full Size | PDF

2.3 Macroscopic effects

The $T$-matrix element $\mathbf {T}_n(I(\rho,z),\varphi _E(\rho,z))$ is calculated for the laser field $E(\rho,z,t)=|E(\rho,z)|\cos [\omega t-\varphi _E(\rho,z)]$ at a fixed point $(\rho,z)$ in the focal region of the laser Gaussian beam. Here we have taken into account that $\mathbf {T}_n$ depends on ${\mathbf {r}_m}$ via the local laser intensity $I(\rho,z)$ and phase $\varphi _E(\rho,z)$, Eqs. (4) and (6). According to Eqs. (1) and (2), the $T$-matrix element $\mathbf {T}_n(I,\varphi _E)$ is proportional to

$$\int_0^{T} dt\, e^{in\omega t}\int_{-\infty}^{t}dt_0 f(\omega t-\varphi_E,\omega t_0-\varphi_E,t-t_0),$$
which, after the substitution $\omega t'=\omega t -\varphi _E$ and $\omega t_0'=\omega t_0 -\varphi _E$, can be written as
$$e^{in\varphi_E}\int_{-\varphi_E/\omega}^{T-\varphi_E/\omega} dt'\, e^{in\omega t'} \int_{-\infty}^{t}dt_0' f(\omega t',\omega t_0',t'-t_0').$$

The subintegral function does not depend on the phase $\varphi _E$ and only the interval of integration over the recombination time is changed from $[0,T]$ to $[-\varphi _E,T-\varphi _E]$. In fact, the integration interval $[0,T]$ was obtained from the interval $(-\infty,\infty )$ using the $T$-periodicity of the subintegral functions. Therefore, we can approximate $\mathbf {T}_n(I,\varphi _E)$ with $e^{in\varphi _E}\mathbf {T}_n(I)$, where $\mathbf {T}_n(I)\equiv \mathbf {T}_n(I,\varphi _E=0)$.

It is generally accepted [49] that the macroscopic effects in HHG should be considered using classical electrodynamics, i.e., solving Maxwell’s wave equations for the fundamental and harmonic fields. Quantum mechanics comes into play only in the calculation of the (microscopic) time-dependent dipole moment which is related to the polarization of the medium (atomic gas) induced by the laser field. The $n$th harmonic field $\mathbf {\cal E}_n(\mathbf {r},t)$ satisfies the wave equation [4,5,15,16]

$$\mathbf{\nabla}^{2}\mathbf{\cal E}_n(\mathbf{r},t)-\frac{1}{n^{2}\omega^{2}}\frac{\partial^{2}}{\partial t^{2}}\left[K_n^{2}\mathbf{\cal E}_n(\mathbf{r},t)\right]=\frac{4\pi}{c^{2}}\frac{\partial^{2} \mathbf{\cal P}_n(\mathbf{r},t)}{\partial t^{2}},$$
where $\mathbf {K}_n$ is the harmonic field wave vector and $\mathbf {\cal P}_n(\mathbf {r},t)$ is the polarization of the medium. An integral solution of this equation is [27]
$$\mathbf{\cal E}_n(\mathbf{r},t)={-}\frac{1}{c^{2}}\int \frac{d\mathbf{r}'}{|\mathbf{r}-\mathbf{r}'|}\left[\frac{\partial^{2} \mathbf{\cal P}_n(\mathbf{r}',t')}{\partial t^{2}}\right]_\mathrm{ret},$$
where the square brackets $[\;]_\mathrm {ret}$ means that the time $t'$ is to be evaluated at the retarded time, $t'=t-|\mathbf {r}-\mathbf {r}'|/c$. We suppose that the time dependence of the harmonic field and polarization is harmonic: $\mathbf {\cal E}_n(\mathbf {r},t)=\mathbf {\cal E}_n(\mathbf {r};t)\exp (-in\omega t)/2+\mathrm {c.c.}$, $\mathbf {\cal P}_n(\mathbf {r}',t')=\mathbf {\cal P}_n(\mathbf {r}';t')\exp (-in\omega t')/2+\mathrm {c.c.}$ [with the time after the semicolon we denoted a slow time dependence which comes via the pulse shape factor $f_p(t')$, $t'=t-|\mathbf {r}-\mathbf {r}'|/c$]. Introducing this into (10) we obtain
$$\mathbf{\cal E}_n(\mathbf{r};t)=\left(\frac{n\omega}{c}\right)^{2} \int d\mathbf{r}' \frac{ e^{iK_n|\mathbf{r}-\mathbf{r}'|} }{|\mathbf{r}-\mathbf{r}'|} \left[\mathbf{\cal P}_n(\mathbf{r}';t')\right]_\mathrm{ret}.$$

We assume that the polarization $\mathbf {\cal P}_n$ is equal to the $T$-matrix element $\mathbf {T}_n(I,\varphi _E)\approx e^{in\varphi _E}\mathbf {T}_n(I)$.

We choose $\mathbf {r}=\mathbf {r}_d$ and $\mathbf {r}'=\mathbf {r}_m\equiv (\rho,\varphi,z)$, where $\mathbf {r}_d$ is the detector position (observation point) and $\mathbf {r}_m$ is a point in the generating medium. We consider two different levels of approximations. In the simpler one, which we call the coherent intensity focal averaging, we approximate $\exp (iK_n|\mathbf {r}-\mathbf {r}'|+in\varphi _E) /|\mathbf {r}-\mathbf {r}'|$ with a constant so that the subintegral function in Eq. (11) reduces to $\mathbf {T}_n(I)$. In the second one, which we call the coherent spatio-temporal focal averaging, we integrate over all times $t_p=t$ and calculate $\mathbf {\cal E}_{n,p}(\rho _d,z_d)=\int dt_p\mathbf {\cal E}_n(\mathbf {r}_d;t_p)$.

2.3.1 Coherent intensity focal averaging

We define the macroscopic coherent intensity focal-averaged $T$-matrix element as

$$\bar{\mathbf{T}}_n(I_0)=\int d\mathbf{r}_m \mathbf{T}_n(I)=2\pi\int_0^{\infty} \rho d\rho \int_{-\infty}^{\infty} dz \mathbf{T}_n(I(\rho,z)).$$

The factor $2\pi$ comes from the integral over the azimuthal angle, while the integrals over $\rho$ and $z$ can be replaced by the integral over the laser intensity, similarly as this has been done for the focal-averaging of the differential ionization yield in strong-field ionization [2830]. The result is

$$\bar{\mathbf{T}}_n(I_0)=\frac{\pi w_0^{2}z_0}{3}\int_0^{I_0}\frac{dI}{I}\left(2+\frac{I_0}{I}\right)\left(\frac{I_0}{I}-1\right)^{1/2}\mathbf{T}_n(I).$$

We have already used a similar formula for the calculation of the incoherent focal-averaged high-harmonic yield in Refs. [31,32].

In the case of the geometry presented in Fig. 1, the integral over $z$ in Eq. (12) is from $z_1$ to $z_2$, i.e., it is not from $-\infty$ to $+\infty$. In this case, in Eq. (12) we introduce $1=\int _0^{I_0}dI\,\delta (I-I(\rho,z))$ and make the substitutions $r=\rho ^{2}/w_0^{2}$ and $z'=z/z_0$. Then from Eq. (4) it follows that $-z'_{I}\le z'\le z'_{I}$, $z'_{I}=\sqrt {I_0/I-1}$. Therefore, the lower and upper limits of the integral over $z'$ are

$$z'_\mathrm{low}=\max\{{-}z'_{I},z_1/z_0\}, \quad z'_\mathrm{upp}=\min\{z'_{I},z_2/z_0\},$$
where $z_1=-L/2 +\Delta z$ and $z_2=L/2 +\Delta z$. The delta function can be written as $\delta (I-I(\rho,z))=\delta (h(r))=\delta (r-r_0)/|h'(r_0)|$, where $r_0$ is the solution of the equation $h(r)=I-I_0\exp [-2r/(1+z'^{2})]/(1+z'^{2})=0$, which is $r_0=- (1+z'^{2})\ln \left [(1+z'^{2})I/I_0\right ]/2$. Using $h'(r_0)=2I/(1+z'^{2})$ we obtain $\delta (I-I(\rho,z))=\delta (r-r_0)(1+z'^{2})/(2I)$. This delta function cancels the integral over $r$ so that the result is
$$\bar{\mathbf{T}}_n(I_0)=\frac{\pi w_0^{2}z_0}{2}\int_0^{I_0}\frac{dI}{I}\mathbf{T}_n(I)\int_{z'_\mathrm{low}}^{z'_\mathrm{upp}}dz'(1+z'^{2}),$$
where $\int dz'(1+z'^{2})=z'+z'^{3}/3$. For $z_1\rightarrow -\infty$ and $z_2\rightarrow \infty$ Eq. (15) reduces to Eq. (13).

Equations (12)–(15) can be generalized to include the laser pulse shape. We multiply the Gaussian laser beam intensity $I(\rho,z)$ with the factor $f_p(t_p-z/c)$, similarly as this has been done for the strong-field ionization with strong focusing in Refs. [33,34]. The result is

$$\bar{\mathbf{T}}_n(I_0,\tau_g)=2\pi\int_{-\infty}^{\infty} d t_p\int_0^{\infty} \rho d\rho \int_{-\infty}^{\infty} dz \mathbf{T}_n(I(\rho,z,t_p)).$$

After a derivation, analogous to that used to obtain Eq. (15), we get

$$\bar{\mathbf{T}}_n(I_0,\tau_g)=\pi w_0^{2}z_0\tau_g\int_0^{I_0}\frac{dI}{I}\mathbf{T}_n(I) g(I),\;\;\; g(I)=\int_{z'_\mathrm{low}}^{z'_\mathrm{upp}}dz'(1+z'^{2})\left[\ln\frac{I_0}{I(1+z'^{2})}\right]^{1/2}.$$

2.3.2 Coherent spatio-temporal focal averaging

In the case of coherent spatio-temporal focal averaging the harmonic field is given by

$$\mathbf{\cal E}_{n,p}(\mathbf{r}_d)= \left(\frac{n\omega}{c}\right)^{2} \int_{-\infty}^{\infty} dt_p \int d\mathbf{r}_m \frac{ e^{iK_n|\mathbf{r}_d-\mathbf{r}_m|}}{|\mathbf{r}_d-\mathbf{r}_m|} e^{in\varphi_E(\rho,z)} \mathbf{T}_n(I(\rho,z)f_p(t_p-|\mathbf{r}_d-\mathbf{r}_m|/c)).$$

In the far-field approximation we have $|{\mathbf {r}_m}-\mathbf {r}_d|\approx z_d-z+[\rho ^{2}+\rho _d^{2}-2\rho \rho _d\cos (\varphi -\varphi _d)]/[2(z_d-z)]$. We use this approximation for the term in the exponent, while for the denominator we take into account only the lowest order term, $|{\mathbf {r}_m}-\mathbf {r}_d|\approx z_d-z$. Since $\mathbf {T}_n$ does not depend on the azimuthal angle $\varphi$, the integral over $\varphi$ in Eq. (18) can be calculated analytically. The result is

$$\mathbf{\cal E}_{n,p}(\rho_d,z_d)= 2\pi\left(\frac{n\omega}{c}\right)^{2}e^{iK_nz_d}\int_0^{\infty} \rho d\rho \int_{-\infty}^{\infty} dz \frac{e^{{-}i\Delta K_n z}}{z_d-z} \int_{-\infty}^{\infty} dt_p\mathbf{T}_n(I(\rho,z,t_p)) J_0(\eta)e^{i\zeta},$$
where $I(\rho,z,t_p)=I(\rho,z)f_p(t_p-(z_d-z)/c)$, $\Delta K_n=K_n-nk$ is the phase mismatch [35], $J_0(\eta )$ is the zero-order Bessel function, $\eta =K_n\rho _d\rho /(z_d-z)$, and
$$\zeta=K_n\frac{\rho^{2}+\rho_d^{2}}{2(z_d-z)}-n\psi_G(z)+\frac{nk\rho^{2}z}{2(z^{2}+z_0^{2})}.$$

Next, in the so obtained equation we introduce $1=\int _0^{I_0}dI'\,\delta (I'-I(\rho,z,t_p))$ and make the substitutions $\rho '=\rho /w_0$, $z'=z/z_0$, and $t'=[t_p-(z_d-z)/c]/\tau _g$. We get a four-dimensional integral in the form

$$z_0 w_0^{2}\tau_g\int_{0}^{I_0}dI'\int_{z'_\mathrm{low}}^{z'_\mathrm{upp}}dz'\int_{0}^{\infty}\rho'd\rho' \int_{-\infty}^{\infty}dt'\delta(I'-I(\rho',z')f_p(\tau_g t')).$$

Here we used the fact that, in the case of the geometry presented in Fig. 1, the integral over $z$ is from $z_1$ to $z_2$, i.e., it is not from $-\infty$ to $+\infty$, so that the lower and upper limits of the integral over $z'$ are given by Eq. (14). Making another substitution: $\sqrt {2}\rho '/\sqrt {1+z'^{2}}=r\cos \beta$, $t'=r\sin \beta$, $d\rho ' dt'=\sqrt {(1+z'^{2})/2} \,rdr d\beta$, $r\in [0,\infty )$, $\beta \in [-\pi /2,\pi /2]$, we obtain

$$\mathbf{\cal E}_{n,p}(\rho_d,z_d)=\pi z_0w_0^{2}\tau_g\left(\frac{n\omega}{c}\right)^{2}\frac{e^{iK_nz_d}}{z_d}\int_{0}^{I_0}dI'\mathbf{T}_n(I')F_n(I'),$$
with
$$F_n(I')=\int_{z'_\mathrm{low}}^{z'_\mathrm{upp}}dz' g_n(z')\int_{0}^{\infty}dr r^{2} \delta(I'-I(r,z'))G_n(z',r),$$
where $I(r,z')=I_0e^{-r^{2}}/(1+z'^{2})$,
$$\begin{array}{rll} g_n(z')&=&\frac{(1+z'^{2})e^{ih_n(z')}}{1-\varepsilon_d z'}, \;\; h_n(z')={-}\Delta K_n z_0z'+\frac{K_n(kw_0\varepsilon_\rho)^{2}}{4k\varepsilon_d(1-\varepsilon_d z')}-n\arctan z',\\ G_n(z',r)&=&\int_{-\pi/2}^{\pi/2}d\beta\cos\beta J_0(a_n(z')r\cos\beta)e^{ib_n(z')r^{2}\cos^{2}\beta},\\ a_n(z')&=& \frac{K_n w_0 \varepsilon_\rho}{1-\varepsilon_d z'}\sqrt{\frac{1+z'^{2}}{2}},\;\; b_n(z')= \frac{n}{2}z'+\frac{K_n\varepsilon_d}{2k}\frac{1+z'^{2}}{1-\varepsilon_d z'}, \end{array}$$
and $\varepsilon _d=z_0/z_d$ and $\varepsilon _\rho =\rho _d/z_d$ are small parameters for the geometry of the HHG experiment. The integral over $r$ is cancelled with the delta function which satisfies the relation $\delta (f(r))=\delta (r-r_0)/|f'(r_0)|$, where $r_0$ is the solution of the equation $f(r)=0$. In our case, it is $r_0=\sqrt {\ln \frac {I_0}{I'(1+z'^{2})}}$, so that $f'(r_0)=2r_0 I'$ and $\int _0^{\infty }dr g(r)\delta (I'-I(r,z'))=g(r_0)/(2r_0I')$. Introducing this into (22) we get
$$F_n(I')= \frac{1}{2I'}\int_{z'_\mathrm{low}}^{z'_\mathrm{upp}}dz' g_n(z') r G_n(z',r),\;\; r= \sqrt{\ln\frac{I_0}{I'(1+z'^{2})}}.$$

The final result for the coherent spatio-temporal focal-averaged harmonic field is given by Eqs. (21), (23), and (24). Supposing an infinite gas cell, for $\varepsilon _d=0$, $\varepsilon _\rho =0$, $\Delta K_n=0$, we can approximate this by

$$\begin{array}{rll} F_n^{(0)}(I')&=&\frac{1}{I'}\int_{0}^{z'_{I'}}dz' \,r \,\mathrm{Re}\left[g_n^{(0)}(z')G_n^{(0)}(z',r)\right],\\ g_n^{(0)}(z')&=&(1+z'^{2})e^{{-}in\arctan z'},\;\; G_n^{(0)}(z',r)= 2\int_{0}^{\pi/2}d\beta\cos\beta\, e^{\frac{i}{2}nz'r^{2}\cos^{2}\beta}.\end{array}$$

3. Numerical results

3.1 Wavelength and intensity dependence of the macroscopic HHG spectra

In order to get an impression how the HHG spectra change with the change of the central wavelength and the peak intensity of the laser field, in Fig. 2 we present the spectra for the He atom (${I_p}=24.59$ eV) obtained using coherent intensity focal averaging for an infinitely long cell and a flat pulse, Eqs. (3) and (13). The laser field is linearly polarized having the peak intensity $I_0=9.4\times 10^{14}\;\mathrm {W/cm}^{2}$ and the wavelengths 800, 1300, 1600, and 2100 nm (upper panel), and the wavelength 800 nm and the peak intensities $I_0=9.4\times 10^{14}\;\mathrm {W/cm}^{2}$, $6\times 10^{14}\;\mathrm {W/cm}^{2}$, and $3\times 10^{14}\;\mathrm {W/cm}^{2}$ (lower panel). In all cases, after a maximum in the region $45-50$ eV, the spectrum has the form of a declining plateau, with a clear cutoff at the energy $3.173{{U_p}}_0+1.325{I_p}$. The plateau of the microscopic HHG spectrum has the same cutoff. However, it is rather flat (see Fig. 5). The reason for this discrepancy is the fact that HHG is a coherent process. Namely, for incoherent processes we sum the contributions of the yields (i.e., the probabilities, not the amplitudes) from all points in the laser focus and the corresponding plateau is flat. An example is the plateau in high-order above-threshold ionization [1,36], with the observed cutoff at $10{{U_p}}_0$. For HHG the coherent focal averaging leads to the declining plateaus as those presented in Fig. 2. We see that the harmonic intensity is higher for shorter wavelengths and higher intensities. The maximum harmonic photon energy scales as ${{U_p}}_0\propto I_0\lambda ^{2}$ in accordance with the mentioned cutoff law. In most experiments the high-energy photons in the upper part of the plateau region are below the detection threshold and this cutoff cannot be observed. For example, even in the early HHG experiments, it was found that the experimental cutoff energy is lower, approximately at ${I_p}+2{{U_p}}_0$ [37]. The detection threshold problem is particularly pronounced for longer wavelengths for which the yield is lower (compare the upper panel of Fig. 2). In the next two subsections we will show, applying our coherent spatio-temporal focal-averaging method, that it is possible, changing the macroscopic conditions, to obtain an observable peak in the spectrum which agrees with the experimental findings.

 figure: Fig. 2.

Fig. 2. Harmonic intensities for the He atom obtained using the coherent intensity focal averaging for an infinitely long cell, a flat laser pulse, and various wavelengths and peak intensities of the linearly polarized laser field. Upper panel: $I_0=9.4\times 10^{14}\;\mathrm {W/cm}^{2}$ and $\lambda =800$ nm, 1300 nm, 1600 nm, 2100 nm. Lower panel: $\lambda =800$ nm and $I_0=9.4\times 10^{14}\;\mathrm {W/cm}^{2}$, $6\times 10^{14}\;\mathrm {W/cm}^{2}$, $3\times 10^{14}\;\mathrm {W/cm}^{2}$.

Download Full Size | PDF

3.2 Examples of HHG by a linearly polarized field

As a check of our focal-averaging methods, we now calculate the HHG spectra of Ar and He atoms exposed to a linearly polarized laser field having the central wavelength 2100 nm and the peak intensity $I_0=4.1\times 10^{14}\;\mathrm {W/cm}^{2}$ for Ar and $I_0=9.4\times 10^{14}\;\mathrm {W/cm}^{2}$ for He. These are the parameters of the experiment [38]. In the experiment with Ar the laser beam was focused with an $f=750$ mm lens into a 6 mm long gas cell, which led to a beam with $w_0=80\;{\mathrm{\mu} \mathrm{m}}$, so that $z_0\approx 9.6$ mm. The above parameters for the experiment with He are $f=500$ mm, $w_0=53.5\;{\mathrm{\mu} \mathrm{m}}$, and $z_0\approx 4.3$ mm. For these parameters we estimate that the parameter $\varepsilon _d\le 0.01$. The estimated beam divergence in [38] is 1 mrad which corresponds to $\varepsilon _\rho \le 0.001$ so that we expect that the influence of this parameter can be neglected. Furthermore, we expect that the experimental conditions are such that phase matching is achieved and $\Delta K_n=0$.

The experimentally observed maximum harmonic photon energy for HHG by Ar atoms and the mentioned laser-field parameters was 170 eV in [38]. However, the (theoretical) cutoff law for high-order harmonic generation [14] gives $n_{\max }\omega =3.173{{U_p}}_0+1.325{I_p}=557$ eV, where the ponderomotive energy for the peak intensity $I_0=4.1\times 10^{14}\;\mathrm {W/cm}^{2}$ is ${{U_p}}_0=169$ eV and the ionization potential of Ar is ${I_p}=15.76$ eV. Since in the experiment there is no abrupt cutoff at 557 eV as expected from the microscopic theory, it is obvious that, for a successful simulation, the macroscopic effects should be taken into account.

 figure: Fig. 3.

Fig. 3. Comparison of the microscopic single-atom harmonic intensity (top curve) and the macroscopic harmonic intensities obtained using coherent intensity focal averaging (denoted as “intensity, $\infty$” for an infinitely long cell and a flat pulse) and using coherent spatio-temporal focal averaging for $\varepsilon _d=0$, $\varepsilon _\rho =0$, $\Delta K_n=0$, cell length $L=6$ mm, and its shift with respect to the laser focus $\Delta z=0$ mm and $\Delta z=-2.5$ mm, as denoted. HHG is by argon atoms and a linearly polarized laser field with the wavelength 2100 nm, the peak intensity $4.1\times 10^{14}\;\mathrm {W/cm}^{2}$, and the minimum beam waist $w_0=80\; \mu \mathrm {m}$.

Download Full Size | PDF

In Fig. 3 we compare, on a logarithmic scale, the results for the harmonic intensity as a function of the harmonic photon energy obtained using the microscopic theory (solid black line) and using coherent intensity focal averaging with an infinite cell, Eq. (13) (dashed green line). We see that the microscopic spectrum is rather flat having its cutoff really at 557 eV. On the other hand, the spectrum obtained using the coherent intensity focal averaging, with an infinite cell and for a flat pulse, exhibits a maximum near 35 eV and then decreases for many orders of magnitude, still having the cutoff at 557 eV. We also present the results obtained using coherent spatio-temporal focal averaging, Eqs. (21)–(24), for the minimum beam waist $w_0=80\; \mu \mathrm {m}$ and the shift of the gas cell, having the length $L=6$ mm, with respect to the laser focus $\Delta z=0$ mm (solid red line) and $\Delta z=-2.5$ mm (dotted blue line). All spectra exhibit a minimum of the harmonic yield at 367 eV. This minumum is a single-atom effect and related to the dipole matrix element with the $p$ ground state of the Ar atom. We have checked this by calculating the harmonic spectra (keeping the same laser parameters) for He atom, which has the $s$ ground state. In this case, the minimum at 367 eV does not appear and the spectrum is rather flat (see Fig. 5). The minimum near 367 eV should not be mixture with the Cooper minimum which appears for Ar atoms at much lower energies near 46 eV. The Cooper minimum does not appear in the strong-field approximation, used in the present paper, but it can be obtained in the low-frequency approximation considered in [31]. The low-energy part of the high-harmonic plateau, for energies below the mentioned minimum at 367 eV, is flat for the microscopic spectrum. However, for the macroscopic spectra, after a sharp decrease for the first few harmonics, the harmonic intensity increases, having a maximum in the region from 30 eV to 150 eV, depending on the geometry of the experiment, and then decreases toward the mentioned minimum at 367 eV. It seems that the macroscopic high-harmonic yield is affected by the destructive interference of contributions from various positions in the laser focus and is dominated by laser pulse regions with lower intensities. Instead of the flat plateau which is characteristic of the microscopic HHG, or the declining plateau in accordance with the coherent intensity focal averaging, a maximum is formed at lower energies; the position of this maximum depends on the macroscopic conditions. This is interesting since it allows manipulation of the harmonic yield by changing the macroscopic condition. Such kind of manipulation is not possible for incoherent processes such as above-threshold ionization. The constructive interference in macroscopic HHG leads to the formation of a maximum which is observed in the experiment. The position of this maximum strongly depends on the shift of the gas cell $\Delta z$.

 figure: Fig. 4.

Fig. 4. Harmonic intensities as functions of the harmonic photon energy, calculated using coherent spatio-temporal focal averaging, for HHG by argon atoms and the laser parameters as in Fig. 3. The parameter $\varepsilon _d$ is zero, except for the dashed green curve (denoted by a star) for which it is $\varepsilon _d=0.01$. The spectrum denoted by a double star (solid maroon line) is calculated for $\Delta z=-3.75$ mm taking into account the saturation effects.

Download Full Size | PDF

In Fig. 4 we present the results obtained using the coherent spatio-temporal focal averaging for various shifts $\Delta z$, as denoted in the legend. For $\Delta z=-2.5$ mm we also present the results for $\varepsilon _d=0.01$ (dashed green line denoted by a star). The results are now presented on a linear scale. We see that the harmonic photon energy for which the harmonic intensity has a maximum shifts from the lowest value of 60 eV for $\Delta z=0$ mm to 160 eV for $\Delta z=-5$ mm. The intensity is maximal for $\Delta z=-2.5$ mm. For the geometrical parameter $\varepsilon _d=z_0/z_d=0.01$ it is at 105 eV, while for $\varepsilon _d=0$ the maximum is slightly lower and is shifted to 110 eV. We conclude that the influence of the parameter $\varepsilon _d$ is small for this experiment. With a further shift of the focus to the negative values of $\Delta z$, the harmonic photon peak energy increases but the corresponding maximum harmonic intensity becomes lower. For $\Delta z=-3.75$ mm the peak is near 140 eV, which is very close to that observed in the experiment [38,39]. The harmonic peak energy is even higher for $\Delta z=-5$ mm: the peak is near 160 eV, but the harmonic intensity is ten times lower than that for $\Delta z=-2.5$ mm.

As the next example, in Fig. 5 we present our results for the He atom. The microscopic plateau is rather flat, with strong interference oscillations due to the used long wavelength and high intensity, and a sharp cutoff appears at 1260 eV. However, such high energies cannot be observed in the experiment since the macroscopic effects drastically change the spectrum. The spectrum obtained using the coherent intensity focal averaging exhibits a maximum near 50 eV and then continuously decreases by four orders of magnitude, having a sharp cutoff at the same energy as in the microscopic case. On the other hand, the results obtained using coherent spatio-temporal focal averaging show that it is possible to manipulate the spectrum changing the position of the focus, i.e., the shift $\Delta z$ (notice that the results in Fig. 5 are in arbitrary units and that the results for different $\Delta z$ are shifted down with respect to the remaining results for visual convenience). The spectra for $\Delta z=0$ mm (blue line), $-2.5$ mm (green line), and $-5$ mm (magenta line), exhibit the maxima at 310 eV, 410 eV, and 180 eV, respectively. These results show that it is possible, by changing the geometry of the experiment, to generate harmonic photons with energies above 400 eV, which is in accordance with the experiment [38]. However, for the used parameters it is not possible to reach the energy 1260 eV predicted by the microscopic theory.

 figure: Fig. 5.

Fig. 5. Comparison of the microscopic single-atom harmonic intensity (top curve) and the harmonic intensities obtained using coherent intensity focal averaging and coherent spatio-temporal focal averaging for $\varepsilon _d=0$, $\varepsilon _\rho =0$, and $\Delta K_n=0$. HHG is by helium atoms and a linearly polarized laser field with the wavelength 2100 nm, the peak intensity $9.4\times 10^{14}\;\mathrm {W/cm}^{2}$, and the minimum beam waist $w_0=53.5\; \mu \mathrm {m}$. The shift of the gas cell, having the length $L=6$ mm, with respect to the laser focus is $\Delta z=0$ mm, $-2.5$ mm, and $-5$ mm, as denoted.

Download Full Size | PDF

 figure: Fig. 6.

Fig. 6. Comparison of the microscopic single-atom harmonic intensity (top curves) and the harmonic intensities obtained using the coherent intensity and spatio-temporal focal averaging for $\varepsilon _d=0$, $\varepsilon _\rho =0$, and $\Delta K_n=0$. HHG is by neon atoms and a bicircular laser field with the component wavelengths $\lambda _1=2000$ nm and $\lambda _2=800$ nm and peak intensities $I_{10}=1.5\times 10^{14}\;\mathrm {W/cm}^{2}$ and $I_{20}=2.8\times 10^{14}\;\mathrm {W/cm}^{2}$. The minimum beam waist is $w_0=80\; \mu \mathrm {m}$ and the gas cell length is $L=6$ mm. For the coherent spatio-temporal focal averaging the shift of the gas cell with respect to the laser focus is $\Delta z=0$ mm and $\Delta z=-2.5$ mm, as denoted. The emitted harmonics are circularly polarized having the helicity $+1$ (red curves) or $-1$ (green curves).

Download Full Size | PDF

Saturation effects can be taken into account by multiplying the harmonic intensity with the probability that the electron remains bound in the volume element characterized by the intensity $I$ at the time $t$ during the laser pulse of duration $\tau _p$. Our calculations show that the He experiment is performed mostly outside of the saturation regime. However, for the Ar experiment we have complete saturation and only about 50% of the atoms survive up to the intensity of about $2\times 10^{14}\;\mathrm {W/cm}^{2}$. But our additional simulation shows that in this case the saturation effects only suppress the oscillations, which can be noticed in the spectra shown in Fig. 4. We confirmed this by calculating the spectra for the Ar experiment taking into account the saturation effects for the pulse duration of 30 fs (see the solid maroon line in Fig. 4).

3.3 Examples of HHG by a bicircular field

We illustrate the HHG spectra generated by a bicircular field using the example of Ne atoms (${I_p}=21.56$ eV) and recent experimental results [40]. The counter-rotating bicircular components have the wavelengths $\lambda _1=2000$ nm and $\lambda _2=800$ nm and the peak intensities $I_{10}=1.5\times 10^{14}\;\mathrm {W/cm}^{2}$ and $I_{20}=2.8\times 10^{14}\;\mathrm {W/cm}^{2}$ (these component intensities are in accordance with [41]; for the lower intensities mentioned in [40] the microscopic cutoff is at 82 eV).

The microscopic single-atom spectra generated by the bicircular field with the above-defined parameters are represented by the top curves in Fig. 6. They are in accordance with the results of Ref. [19]. If we define the fundamental wavelength $\omega$ such that it corresponds to the wavelength of 4000 nm, then our bicircular field consists of counter-rotating circularly polarized components of commensurate frequencies $\omega _1=2\omega$ and $\omega _2=5\omega$. Since the intensity of the $5\omega$ component is almost two times higher, the shape of the spectrum is as presented in Fig. 6: from 30 eV to 100 eV it is increasing while above 100 eV we have a flat plateau which finishes with the cutoff at 160 eV. There is an asymmetry in the emission of harmonics of different helicities: below 125 eV the harmonics with the helicity $+1$ are stronger, while for higher energies the harmonics with the helicity $-1$ prevail. This asymmetry is explained in [42,43] as a consequence of the $p$ ground state of Ne. The selection rules for the $2\omega$$5\omega$ bicircular field [19] imply that harmonics with the energy $(7q\pm 2)\omega =q(\omega _1+\omega _2)\pm \omega _1$, $q$ integer, with helicity $\pm 1$ are emitted. Therefore, the distance of the harmonic peaks for the harmonics of the same helicity is $\omega _1+\omega _2=7\omega$, while for adjacent harmonics of different helicity it is $\omega _2-\omega _1=3\omega$, which is in accordance with the experimental results [40].

As in the case of the linearly polarized field, the situation changes if the macroscopic effects are taken into account. If the simplest coherent intensity focal averaging method is applied, the spectrum is inclined and the harmonic intensity decreases with increasing harmonic photon energy. The change in the dominant helicity occurs at 120 eV, while the cutoff remains at 160 eV. If one applies the coherent spatio-temporal focal averaging, the spectrum further changes. For $\Delta z=0$ mm the spectrum exhibits a maximum in the region from 50 eV to 80 eV. For $\Delta z=-2.5$ mm this maximum is shifted to higher energies and is between 70 eV and 120 eV. The harmonics with negative helicity become stronger for energies higher than 115 eV. These results are in accordance with the experiment [40,41].

4. Conclusions

The microscopic quantum-mechanical $T$-matrix element, Eqs. (1) and (2), describes the emission of high harmonics from a single atom exposed to a laser field whose amplitude and phase depend on the position of this atom in the laser focus. For a realistic simulation of the high-order harmonic generation process, the macroscopic effects should be taken into account. The fields of all emitted high harmonics should be added coherently and the harmonic field should be calculated in the far field at the detector. In the present paper we have introduced various focal-averaging methods, based on an integral solution of the wave equation, and applied them to simulate two recent experiments [38,40].

In previous publications [31,32] we used a method of incoherent intensity focal averaging, in which the focal-averaged intensity was calculated integrating the harmonic intensity $P_n(I)$ over the laser intensity distribution in the focus. However, incoherent focal averaging is inadequate for the calculation of the harmonic ellipticity. In the present paper we have introduced coherent intensity focal averaging, Eqs. (12)–(17), which is a more approximative method [see the paragraph below Eq. (11)]. We have also presented a more precise method, the coherent spatio-temporal focal averaging, Eqs. (21)–(24), which takes into account the macroscopic conditions of the HHG experiment.

In comparison with the microscopic case, the macroscopic HHG spectrum is changed qualitatively: on a linear scale a peak in the spectrum appears at energies much lower than the microscopic cutoff energy. It seems that the high-harmonic yield is affected by the destructive interference of contributions from the various positions of the atoms in the laser focus and is dominated by the contributions of laser-pulse spatio-temporal regions with lower intensities. The microscopic high-energy plateau, with its sharp cutoff, becomes blurred or may not even be visible in the experimental spectrum. Our macroscopic simulations are consistent with these experimental results: the harmonic yield in the high-energy plateau and cutoff region is orders of magnitude lower than the yield of this low-energy peak. While for our simplest coherent intensity focal averaging the position on this peak does not change much with the laser intensity, wavelength, and macroscopic conditions, for our coherent spatio-temporal focal averaging both the position and the height of this peak strongly depend on the macroscopic conditions which we can simulate with our theory. In particular, the harmonic photon energy which corresponds to this peak can be increased few times by shifting the position of the gas cell with respect to the laser focus. We have illustrated this by presenting results for the parameters of recent experiments with a linearly polarized field and Ar and He atoms [38] and with a bicircular field and Ne atoms [40].

Acknowledgments

I gratefully acknowledge support by the Ministry for Science, Higher Education and Youth, Canton Sarajevo, Bosnia and Herzegovina, and by the Alexander von Humboldt Foundation. I thank Wilhelm Becker, Matthias Schnürer, and Anke Heilmann for useful discussions.

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. B. Milošević and F. Ehlotzky, “Scattering and reaction processes in powerful laser fields,” Adv. At. Mol. Opt. Phys. 49, 373–532 (2003). [CrossRef]  

2. A. McPherson, G. Gibson, H. Jara, U. Johann, T. S. Luk, I. A. McIntyre, K. Boyer, and C. K. Rhodes, “Studies of multiphoton production of vacuum-ultraviolet radiation in the rare gases,” J. Opt. Soc. Am. B 4(4), 595–601 (1987). [CrossRef]  

3. M. Ferray, A. L’Huillier, X. F. Li, L. A. Lompre, G. Mainfray, and C. Manus, “Multiple-harmonic conversion of 1064-nm radiation in rare gases,” J. Phys. B 21(3), L31–L35 (1988). [CrossRef]  

4. A. L’Huillier, K. J. Schafer, and K. C. Kulander, “Theoretical aspects of intense field harmonic generation,” J. Phys. B 24(15), 3315–3341 (1991). [CrossRef]  

5. A. L’Huillier, L.-A. Lompré, G. Mainfray, and C. Manus, “High-order harmonic generation in rare gases,” in Atoms in intense laser fields, M. Gavrila, ed., Supplement 1 to Adv. At. Mol. Opt. Phys. (Academic, Boston, 1992), pp. 139–206.

6. P. Salières, A. L’Huillier, P. Antoine, and M. Lewenstein, “Study of the spatial and temporal coherence of high-order harmonics,” Adv. At. Mol. Phys. 41, 83–142 (1999). [CrossRef]  

7. M. B. Gaarde, J. L. Tate, and K. J. Schafer, “Macroscopic aspects of attosecond pulse generation,” J. Phys. B 41(13), 132001 (2008). [CrossRef]  

8. F. Calegari, F. Ferrari, M. Lucchini, M. Negro, C. Vozzi, S. Stagira, G. Sansone, and M. Nisoli, “Principles and applications of attosecond technology,” Adv. At. Mol. Phys. 60, 371–413 (2011). [CrossRef]  

9. A.-T. Le, H. Wei, C. Jin, and C. D. Lin, “Tutorial: Strong-field approximation and its extension for high-order harmonic generation with mid-infrared lasers,” J. Phys. B 49(5), 053001 (2016). [CrossRef]  

10. V. V. Strelkov, V. T. Platonenko, A. F. Sterzhantov, and M. Yu. Ryabikin, “Attosecond electromagnetic pulses: generation, measurement, and application. Generation of high-order harmonics of an intense laser field for attosecond pulse production,” Phys. Usp. 59(5), 425–445 (2016). [CrossRef]  

11. J. L. Krause, K. J. Schafer, and K. C. Kulander, “High-order harmonic generation from atoms and ions in the high intensity regime,” Phys. Rev. Lett. 68(24), 3535–3538 (1992). [CrossRef]  

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

13. K. C. Kulander, K. J. Schafer, and J. L. Krause, “Dynamics of short-pulse excitation, ionization and atomic conversion,” in Super-intense laser-atom physics, B. Piraux, A. L’Huillier, and K. Rzążewski, eds., Vol. 316 of NATO Advanced Science Institutes Series, Series B: Physics (Plenum, New York, 1993), pp. 95–110.

14. M. Lewenstein, Ph. Balcou, M. Yu. Ivanov, A. L’Huillier, and P. B. Corkum, “Theory of high-harmonic generation by low-frequency laser fields,” Phys. Rev. A 49(3), 2117–2132 (1994). [CrossRef]  

15. C. F. R. Caron and R. M. Potvliege, “Phase matching and harmonic generation in Bessel-Gauss beam,” J. Opt. Soc. Am. B 15(3), 1096–1106 (1998). [CrossRef]  

16. C. F. R. Caron and R. M. Potvliege, “propBG: A program for calculating the intensity and the far-field profile of harmonics emitted by a dilute gas irradiated by an intense Gaussian or Bessel-Gauss laser beam,” Comput. Phys. Commun. 126(3), 269–293 (2000). [CrossRef]  

17. C. Hernández-García, J. A. Pérez-Hernández, J. Ramos, E. Conejero Jarque, L. Roso, and L. Plaja, “High-order harmonic propagation in gases within the discrete dipole approximation,” Phys. Rev. A 82(3), 033432 (2010). [CrossRef]  

18. B. Ghomashi, R. Reiff, and A. Becker, “Coherence in macroscopic high harmonic generation for spatial focal phase distributions of monochromatic and broadband Gaussian laser pulses,” Opt. Express 29(24), 40146–40160 (2021). [CrossRef]  

19. D. B. Milošević, “Circularly polarized high harmonics generated by a bicircular field from inert atomic gases in the p state: A tool for exploring chirality-sensitive processes,” Phys. Rev. A 92(4), 043827 (2015). [CrossRef]  

20. D. B. Milošević and W. Becker, “High-order harmonic generation by bi-elliptical orthogonally polarized two-color fields,” Phys. Rev. A 102(2), 023107 (2020). [CrossRef]  

21. A. E. Siegmann, Lasers (University Science Books, Sausalito, California, 1986).

22. M. Wegener, Extreme nonlinear optics (Springer, Berlin, 2005).

23. The phase $\varphi _E(\rho, z)=\int ^{\mathbf {r}_m}d\mathbf {r}_m'\cdot \mathbf {k}(\mathbf {r}_m')$ satisfies the relation $\mathbf {\nabla }\varphi _E(\rho, z)=\mathbf {k}(\rho, z)$, where the local wave vector k is perpendicular to the local wave front at the point (ρ, z) and is in the direction of the phase velocity. The group velocity, the wave’s energy flow, and the Poynting vector are in the direction of the z axis. The field (4) is monochromatic with the (constant) frequency $\omega =c\sqrt {k_x^{2}+k_y^{2}+k_z^{2}}$ and linearly polarized, $\mathbf {E}(\rho, z, t)=E(\rho, z, t)\hat {\mathbf {e}}_x$. Generalization to a polychromatic elliptically polarized laser field $\mathbf {E}(\rho, z, t)=E_x(\rho, z, t)\hat {\mathbf {e}}_x+E_y(\rho, z, t)\hat {\mathbf {e}}_y$ is straightforward.

24. A. Comby, S. Beaulieu, E. Constant, D. Descamps, S. Petit, and Y. Mairesse, “Absolute gas density profiling in high-order harmonic generation,” Opt. Express 26(5), 6001–6009 (2018). [CrossRef]  

25. V. S. Yakovlev, M. Ivanov, and F. Krausz, “Enhanced phase-matching for generation of soft X-ray harmonics and attosecond pulses in atomic gases,” Opt. Express 15(23), 15351–15364 (2007). [CrossRef]  

26. A. Čerkić and D. B. Milošević, “Focal averaging and incoherent scattering in laser-assisted radiative recombination and scattering processes,” Phys. Rev. A 75(1), 013412 (2007). [CrossRef]  

27. J. D. Jackson, Classical electrodynamics, 3rd ed., (John Wiley & Sons, New York, 1999), Ch. 6.4.

28. S. Augst, D. D. Meyerhofer, D. Strickland, and S. L. Chin, “Laser ionization of noble gases by Coulomb-barrier suppression,” J. Opt. Soc. Am. B 8(4), 858–867 (1991). [CrossRef]  

29. G. N. Gibson, R. R. Freeman, T. J. McIlrath, and H. G. Muller, “Excitation and ionization dynamics in short-pulse multiphoton ionization,” Phys. Rev. A 49(5), 3870–3874 (1994). [CrossRef]  

30. M. A. Walker, P. Hansch, and L. D. Van Woerkom, “Intensity-resolved multiphoton ionization: Circumventing spatial averaging,” Phys. Rev. A 57(2), R701–R704 (1998). [CrossRef]  

31. D. B. Milošević, “Low-frequency approximation for high-order harmonic generation by a bicircular laser field,” Phys. Rev. A 97(1), 013416 (2018). [CrossRef]  

32. D. B. Milošević, “Control of the helicity of high-order harmonics generated by bicircular laser fields,” Phys. Rev. A 98(3), 033405 (2018). [CrossRef]  

33. R. Kopold, W. Becker, M. Kleber, and G. G. Paulus, “Channel-closing effects in high-order above-threshold ionization and high-order harmonic generation,” J. Phys. B 35(2), 217–232 (2002). [CrossRef]  

34. D. B. Milošević, W. Becker, M. Okunishi, G. Prümper, K. Shimada, and K. Ueda, “Strong-field electron spectra of rare-gas atoms in the rescattering regime: enhanced spectral regions and a simulation of the experiment,” J. Phys. B 43(1), 015401 (2010). [CrossRef]  

35. The phase mismatch can be approximated by ΔKn = (nHn − 1)nk [5], where nHn is the (linear) refractive index at the frequency which is proportional to the atomic density and, therefore, increases with the increase of the gas pressure. Outside the medium it is ΔKn = 0. The above results can be generalized introducing the z-dependent wave vectors, i.e., replacing ΔKnz in the phase factor by the optical path integral $\int _{-\infty }^{z}dz'\Delta K_n (z')$. Furthermore, the absorption of the high harmonics in the medium can be taken into account introducing the imaginary part of the phase mismatch (κn) such that [5] $\Delta K_n z\rightarrow \int _{-\infty }^{z}dz'\Delta K_n (z')-i\int dz''\kappa _n(z'')$, with ΔKn(z) ≈ [nHn(z) − 1]nk(z).

36. W. Becker, F. Grasbon, R. Kopold, D. B. Milošević, G. G. Paulus, and H. Walther, “Above-threshold ionization: from classical features to quantum effects,” Adv. At. Mol. Opt. Phys. 48, 35–98 (2002). [CrossRef]  

37. A. L’Huillier, M. Lewenstein, P. Salières, Ph. Balcou, M. Yu. Ivanov, J. Larsson, and C. G. Wahlström, “High-order harmonic-generation cutoff,” Phys. Rev. A 48(5), R3433–R3436 (1993). [CrossRef]  

38. T. Feng, A. Heilmann, M. Bock, L. Ehrentraut, T. Witting, H. Yu, H. Stiel, S. Eisebitt, and M. Schnürer, “27W 2.1 μm OPCPA system for coherent soft X-ray generation operating at 10 kHz,” Opt. Express 28(6), 8724–8733 (2020). [CrossRef]  

39. The maximum at 140 eV in the experiment [38] is connected to the maximum transmission of the used silver filter (Anke Heilmann – private communication).

40. K. M. Dorney, T. Fan, Q. L. D. Nguyen, J. L. Ellis, D. D. Hickstein, N. Brooks, D. Zusin, C. Gentry, C. Hernández-García, H. C. Kapteyn, and M. M. Murnane, “Bright, single helicity, high harmonics driven by mid-infrared bicircular laser fields,” Opt. Express 29(23), 38119–38128 (2021). [CrossRef]  

41. T. Fan, P. Grychtol, R. Knut, C. Hernández-García, D. D. Hickstein, D. Zusin, C. Gentry, F. J. Dollar, C. A. Mancuso, C. Hogle, O. Kfir, D. Legut, K. Carva, J. L. Ellis, K. Dorney, C. Chen, O. Shpyrko, E. E. Fullerton, O. Cohen, P. M. Oppeneer, D. B. Milošević, A. Becker, A. A. Jaroń-Becker, T. Popmintchev, M. M. Murnane, and H. C. Kapteyn, “Generation of bright soft x-ray harmonics with circular polarization for x-ray magnetic circular dichroism, in Conference on lasers and electro-optics (OSA Technical Digest, Optical Society of America) CLEO: Science and innovations 2016, San Jose, California, United States, 5–10 June (2016) [ISBN: 978-1-943580-11-8, DOI: 10.1364/CLEOSI.2016.SF1I.1]; Tingting Fan (private communication).

42. D. B. Milošević, “Generation of elliptically polarized attosecond pulse trains,” Opt. Lett. 40(10), 2381–2384 (2015). [CrossRef]  

43. L. Medišauskas J. Wragg, H. W. van der Hart, and M. Yu. Ivanov, “Generating isolated elliptically polarized attosecond pulses using bichromatic counterrotating circularly polarized laser fields,” Phys. Rev. Lett. 115(15), 153001 (2015). [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 (6)

Fig. 1.
Fig. 1. Schematic presentation of the gas cell, focused Gaussian beam, and the detector plane in the HHG experiment.
Fig. 2.
Fig. 2. Harmonic intensities for the He atom obtained using the coherent intensity focal averaging for an infinitely long cell, a flat laser pulse, and various wavelengths and peak intensities of the linearly polarized laser field. Upper panel: $I_0=9.4\times 10^{14}\;\mathrm {W/cm}^{2}$ and $\lambda =800$ nm, 1300 nm, 1600 nm, 2100 nm. Lower panel: $\lambda =800$ nm and $I_0=9.4\times 10^{14}\;\mathrm {W/cm}^{2}$, $6\times 10^{14}\;\mathrm {W/cm}^{2}$, $3\times 10^{14}\;\mathrm {W/cm}^{2}$.
Fig. 3.
Fig. 3. Comparison of the microscopic single-atom harmonic intensity (top curve) and the macroscopic harmonic intensities obtained using coherent intensity focal averaging (denoted as “intensity, $\infty$” for an infinitely long cell and a flat pulse) and using coherent spatio-temporal focal averaging for $\varepsilon _d=0$, $\varepsilon _\rho =0$, $\Delta K_n=0$, cell length $L=6$ mm, and its shift with respect to the laser focus $\Delta z=0$ mm and $\Delta z=-2.5$ mm, as denoted. HHG is by argon atoms and a linearly polarized laser field with the wavelength 2100 nm, the peak intensity $4.1\times 10^{14}\;\mathrm {W/cm}^{2}$, and the minimum beam waist $w_0=80\; \mu \mathrm {m}$.
Fig. 4.
Fig. 4. Harmonic intensities as functions of the harmonic photon energy, calculated using coherent spatio-temporal focal averaging, for HHG by argon atoms and the laser parameters as in Fig. 3. The parameter $\varepsilon _d$ is zero, except for the dashed green curve (denoted by a star) for which it is $\varepsilon _d=0.01$. The spectrum denoted by a double star (solid maroon line) is calculated for $\Delta z=-3.75$ mm taking into account the saturation effects.
Fig. 5.
Fig. 5. Comparison of the microscopic single-atom harmonic intensity (top curve) and the harmonic intensities obtained using coherent intensity focal averaging and coherent spatio-temporal focal averaging for $\varepsilon _d=0$, $\varepsilon _\rho =0$, and $\Delta K_n=0$. HHG is by helium atoms and a linearly polarized laser field with the wavelength 2100 nm, the peak intensity $9.4\times 10^{14}\;\mathrm {W/cm}^{2}$, and the minimum beam waist $w_0=53.5\; \mu \mathrm {m}$. The shift of the gas cell, having the length $L=6$ mm, with respect to the laser focus is $\Delta z=0$ mm, $-2.5$ mm, and $-5$ mm, as denoted.
Fig. 6.
Fig. 6. Comparison of the microscopic single-atom harmonic intensity (top curves) and the harmonic intensities obtained using the coherent intensity and spatio-temporal focal averaging for $\varepsilon _d=0$, $\varepsilon _\rho =0$, and $\Delta K_n=0$. HHG is by neon atoms and a bicircular laser field with the component wavelengths $\lambda _1=2000$ nm and $\lambda _2=800$ nm and peak intensities $I_{10}=1.5\times 10^{14}\;\mathrm {W/cm}^{2}$ and $I_{20}=2.8\times 10^{14}\;\mathrm {W/cm}^{2}$. The minimum beam waist is $w_0=80\; \mu \mathrm {m}$ and the gas cell length is $L=6$ mm. For the coherent spatio-temporal focal averaging the shift of the gas cell with respect to the laser focus is $\Delta z=0$ mm and $\Delta z=-2.5$ mm, as denoted. The emitted harmonics are circularly polarized having the helicity $+1$ (red curves) or $-1$ (green curves).

Equations (26)

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

T n = 0 T d t T a d a ( t ) e i n ω t ,
d a ( t ) = i t d t 0 m a ( t ) ( 2 π i τ ) 3 / 2 e i S ( t 0 , t ) m a ( t 0 ) E ( t 0 ) ,
P n = ( n ω ) 4 2 π c 3 | T n | 2 ,
I ( ρ , z ) = I 0 1 + z 2 / z 0 2 exp [ 2 ρ 2 w 0 2 ( 1 + z 2 / z 0 2 ) ] .
E ( ρ , z , t ) = | E ( ρ , z ) | R e ( e i [ φ E ( ρ , z ) ω t ] ) , E ( ρ , z ) = E 0 z 0 z 0 + i z exp [ i k ρ 2 2 ( z 0 + i z ) + i k z ] ,
φ E ( ρ , z ) = k z ψ G ( z ) + k ρ 2 / ( 2 R ) ,
0 T d t e i n ω t t d t 0 f ( ω t φ E , ω t 0 φ E , t t 0 ) ,
e i n φ E φ E / ω T φ E / ω d t e i n ω t t d t 0 f ( ω t , ω t 0 , t t 0 ) .
2 E n ( r , t ) 1 n 2 ω 2 2 t 2 [ K n 2 E n ( r , t ) ] = 4 π c 2 2 P n ( r , t ) t 2 ,
E n ( r , t ) = 1 c 2 d r | r r | [ 2 P n ( r , t ) t 2 ] r e t ,
E n ( r ; t ) = ( n ω c ) 2 d r e i K n | r r | | r r | [ P n ( r ; t ) ] r e t .
T ¯ n ( I 0 ) = d r m T n ( I ) = 2 π 0 ρ d ρ d z T n ( I ( ρ , z ) ) .
T ¯ n ( I 0 ) = π w 0 2 z 0 3 0 I 0 d I I ( 2 + I 0 I ) ( I 0 I 1 ) 1 / 2 T n ( I ) .
z l o w = max { z I , z 1 / z 0 } , z u p p = min { z I , z 2 / z 0 } ,
T ¯ n ( I 0 ) = π w 0 2 z 0 2 0 I 0 d I I T n ( I ) z l o w z u p p d z ( 1 + z 2 ) ,
T ¯ n ( I 0 , τ g ) = 2 π d t p 0 ρ d ρ d z T n ( I ( ρ , z , t p ) ) .
T ¯ n ( I 0 , τ g ) = π w 0 2 z 0 τ g 0 I 0 d I I T n ( I ) g ( I ) , g ( I ) = z l o w z u p p d z ( 1 + z 2 ) [ ln I 0 I ( 1 + z 2 ) ] 1 / 2 .
E n , p ( r d ) = ( n ω c ) 2 d t p d r m e i K n | r d r m | | r d r m | e i n φ E ( ρ , z ) T n ( I ( ρ , z ) f p ( t p | r d r m | / c ) ) .
E n , p ( ρ d , z d ) = 2 π ( n ω c ) 2 e i K n z d 0 ρ d ρ d z e i Δ K n z z d z d t p T n ( I ( ρ , z , t p ) ) J 0 ( η ) e i ζ ,
ζ = K n ρ 2 + ρ d 2 2 ( z d z ) n ψ G ( z ) + n k ρ 2 z 2 ( z 2 + z 0 2 ) .
z 0 w 0 2 τ g 0 I 0 d I z l o w z u p p d z 0 ρ d ρ d t δ ( I I ( ρ , z ) f p ( τ g t ) ) .
E n , p ( ρ d , z d ) = π z 0 w 0 2 τ g ( n ω c ) 2 e i K n z d z d 0 I 0 d I T n ( I ) F n ( I ) ,
F n ( I ) = z l o w z u p p d z g n ( z ) 0 d r r 2 δ ( I I ( r , z ) ) G n ( z , r ) ,
g n ( z ) = ( 1 + z 2 ) e i h n ( z ) 1 ε d z , h n ( z ) = Δ K n z 0 z + K n ( k w 0 ε ρ ) 2 4 k ε d ( 1 ε d z ) n arctan z , G n ( z , r ) = π / 2 π / 2 d β cos β J 0 ( a n ( z ) r cos β ) e i b n ( z ) r 2 cos 2 β , a n ( z ) = K n w 0 ε ρ 1 ε d z 1 + z 2 2 , b n ( z ) = n 2 z + K n ε d 2 k 1 + z 2 1 ε d z ,
F n ( I ) = 1 2 I z l o w z u p p d z g n ( z ) r G n ( z , r ) , r = ln I 0 I ( 1 + z 2 ) .
F n ( 0 ) ( I ) = 1 I 0 z I d z r R e [ g n ( 0 ) ( z ) G n ( 0 ) ( z , r ) ] , g n ( 0 ) ( z ) = ( 1 + z 2 ) e i n arctan z , G n ( 0 ) ( z , r ) = 2 0 π / 2 d β cos β e i 2 n z r 2 cos 2 β .
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.