This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Relativistic calculation of non-dipole effects in high harmonic generation

I. A. Ivanov1 igorivanov@ibs.re.kr    Kyung Taec Kim1,2 kyungtaec@gist.ac.kr 1 Center for Relativistic Laser Science, Institute for Basic Science (IBS), Gwangju 61005, Republic of Korea 2Department of Physics and Photon Science, GIST, Gwangju 61005, Korea
Abstract

We present results of relativistic calculations of even order harmonic generation from various atomic targets. The even order harmonics appear due to the relativistic non-dipole effects. We take these relativistic effects into account by using an approach based on the solution of the time-dependent Dirac equation. The spectra of the non-dipole even harmonics look qualitatively similar to the spectra of the dipole harmonics obeying the same classical cutoff rule. The temporal dynamics of the formation of the non-dipole harmonics is, however, distinctly different from the process of dipole harmonics formation. Even order harmonics emission is strongly suppressed at the beginning of the laser pulse, and the emission times of the non-dipole harmonics are shifted with respect to the bursts of the dipole emission. These features are partly explained by a simple modification of the classical three-step model which takes into account selection rules governing the emission of harmonic photons.

pacs:
32.80.Rm 32.80.Fb 42.50.Hz

I Introduction

One can expect relativistic effects to play important role in the dynamics of the processes of atomic or molecular interactions with strong laser pulses for laser intensities over 101810^{18} W/cm2 Reiss (1998), when, with increasing ponderomotive energy, electron velocity can approach speed of light in the vacuum. It has been realized since the pioneering paper by Reiss Reiss (1990), however, that relativistic effects may reveal themselves even for moderately intense (101310^{13}-101410^{14} W/cm2) low frequency infrared (IR) laser fields. For instance, even for the IR laser fields of intensity of the order of 101310^{13} W/cm2, the relativistic effects are visible in the photo-electron spectra Ludwig et al. (2014a); Chelkowski et al. (2014, 2015); I.A.Ivanov et al. (2016); Popov et al. (2006); Klaiber and Hatsagortsyan (2014); Yakaboylu et al. (2015) in the tunneling regime of ionization, characterized by the values γ1\gamma\lesssim 1, where γ=ω2|Ip/E0\gamma=\omega\sqrt{2|I_{p}}/E_{0} is the Keldysh parameter Keldysh (1965), and ω\omega, E0E_{0} and IpI_{p} are the field frequency, field strength and ionization potential of the target system expressed in atomic units. These relativistic non-dipole effects are due to the influence of the magnetic field component of the laser pulse which induces a non-negligible momentum transfer to the photoelectrons Smeenk et al. (2011a); Ludwig et al. (2014b). Alternatively, if we prefer the photon picture of light, one might say that an IR photon carries small momentum, but a large number of the photons participating in the process of the tunneling ionization Krausz and Ivanov (2009) deliver non-negligible momentum to the ionized electron Chelkowski and Bandrauk (2018); Chelkowski et al. (2017).

The momentum delivered by the photons to the photo-electron was measured experimentally under the typical parameters of the tunneling ionization regime Smeenk et al. (2011b). This momentum manifests itself, on average, as a shift of the photo-electron momentum distributions (PMD) in the pulse propagation direction. More detailed picture, which emerges as a result of the complex interplay of the magnetic and Coulomb forces, includes the so-called direct electrons which never recollide with the parent ion and are driven in the direction of the laser photon momentum, and the slow electrons which experience recollisions and may acquire momentum opposite to the photon momentum Chelkowski et al. (2015).

Theoretical study of these effects clearly necessitates methods which go beyond the commonly used non-relativistic dipole approximation. A number of theoretical procedures allowing to consider the relativistic non-dipole effects have been described in the literature, including the relativistic strong-field approximation Reiss (2013); Yakaboylu et al. (2013); Klaiber and Hatsagortsyan (2014); Yakaboylu et al. (2015), time-dependent Schrödinger equation (TDSE) with non-dipole corrections Chelkowski and Bandrauk (2018); Chelkowski et al. (2017); Chelkowski and Bandrauk (2017); I.A.Ivanov et al. (2016), an approach based on the non-dipole strong-field-approximation Hamiltonian Jensen et al. (2020), and the time-dependent Dirac equation (TDDE) Telnov and Chu (2020); Selstø et al. (2009); I.A.Ivanov (2015); Ivanov (2017).

The non-dipole effects manifest themselves as well in other processes occurring when atoms or molecules interact with laser fields. The process which will interest us in the present work is the process of the High Harmonic Generation (HHG). The non-dipole effects are known to produce several modifications in the HHG spectra. It was found Zhu and Wang (2016) that the non-dipole interactions lead to decrease of harmonic intensity and shift of odd order harmonics in the spectra. A detailed investigation of the effect of the pulse magnetic field on harmonic spectra was reported in Potvliege et al. (2000); Kylstra et al. (2001); Chirilă et al. (2002). It was found Potvliege et al. (2000) that the non-dipole magnetic field effects result in the emission of photons polarized along the propagation direction which, for the laser pulse wavelength of 800800 nm and intensity of the order of 5×10155\times 10^{15} W/cm2, is several orders of magnitude weaker than the photon emission polarized parallel to the driving pulse polarization direction. For stronger pulses with the intensities of the order of 101710^{17} W/cm2, the magnetic field effects start playing crucial role Chirilă et al. (2002). Electron drift in the laser propagation direction due to the magnetic-field component of the laser pulse prevents recollisions, and hence, as one could expect on the basis of the picture provided by the celebrated three-step model of HHG Lewenstein et al. (1994); Corkum (1993), leads to the decrease of the harmonic emission.

Perhaps one of the most striking manifestations of the non-dipole effects is appearance of even order harmonics in the HHG spectra de Aldana and Roso (2002); Bandrauk and Lu (2006); Mu-Xue et al. (2020), presenting an example of a relatively small perturbation producing not only relatively minor quantitative modifications of the spectra, but introducing a qualitative change: harmonics with frequencies forbidden in the dipole approximation. The appearance of the even order harmonics can be understood as a result of the break-up of the well-known symmetry which the electron trajectories responsible for the emission of the harmonic photons exhibit in the dipole approximation Lewenstein et al. (1994). Magnetic field effects break this symmetry, and thus make possible generation of even order harmonics. These harmonics were studied theoretically in Bandrauk and Lu (2006), using perturbative treatment of the non-dipole effects.

In the present paper we report a systematic theoretical study of the non-dipole effects, in particular generation of the even order harmonics, from various atomic targets. We use the TDDE as our main calculational tool, basing on the previously developed procedure for the numerical solution of the time-dependent Dirac equation I.A.Ivanov (2015); Ivanov (2017). The approach based on the TDDE provides a complete non-perturbative description of the non-dipole, as well as other relativistic effects.

Atomic units with =1\hbar=1, e=1e=1, m=1m=1, and c137.036c\approx 137.036 (here ee and mm are charge and mass of the electron, cc- speed of light) are used throughout the paper.

II Theory

II.1 Numerical solution to the time-dependent Dirac equation

We solve the TDDE:

iΨ(r,t)t=H^Ψ(r,t)i\frac{\partial\Psi(r,t)}{\partial t}=\hat{H}\Psi(r,t) (1)

following the procedure we described in I.A.Ivanov (2015); Ivanov (2017), which we briefly recapitulate below for the readers convenience. In Eq. (1) Ψ(r,t)\Psi(r,t) is a four-component bispinor and the Hamiltonian operator has a form:

H^=H^atom+H^int,\hat{H}=\hat{H}_{\rm atom}+\hat{H}_{\rm int}\ , (2)

with:

H^atom=c𝜶𝒑^+c2(βI)+IV(r),\hat{H}_{\rm atom}=c{\bm{\alpha}}\cdot{\hat{\bm{p}}}+c^{2}(\beta-I)+I\ V(r)\ , (3)

and

H^int=c𝜶𝑨^,\hat{H}_{\rm int}=c{\bm{\alpha}}\cdot{\hat{\bm{A}}}\ , (4)

In Eq. (3):

𝜶=(𝟎𝝈𝝈𝟎)\displaystyle{\bm{\alpha}}=\left(\begin{array}[]{cc}{\bm{0}}&{\bm{\sigma}}\\ {\bm{\sigma}}&{\bm{0}}\\ \end{array}\right), β=(𝑰𝟎𝟎𝑰)\displaystyle\beta=\left(\begin{array}[]{cc}{\bm{I}}&{\bm{0}}\\ {\bm{0}}&-{\bm{I}}\\ \end{array}\right), I=(𝑰𝟎𝟎𝑰)\displaystyle I=\left(\begin{array}[]{cc}{\bm{I}}&{\bm{0}}\\ {\bm{0}}&{\bm{I}}\\ \end{array}\right), 𝝈{\bm{\sigma}} are Pauli matrices, 𝟎{\bm{0}} and 𝑰{\bm{I}} are 2×22\times 2 null and identity matrices, V(r)V(r) is the atomic potential and c=137.036c=137.036- the speed of light. We subtracted from the field-free atomic Hamiltonian (3) the constant term Ic2Ic^{2} corresponding to the rest mass energy of the electron.

We use a laser pulse linearly polarized in zz- and propagating in xx- directions. The vector potential of the pulse is defined in terms of the pulse electric field:

𝑨(x,t)=𝒆z^0uE(τ)𝑑τ,\bm{A}(x,t)=-\hat{{\bm{e}}_{z}}\int\limits_{0}^{u}E(\tau)\ d\tau\ , (5)

where u=tx/cu=t-x/c. At any given point in space the pulse has a finite duration T1T_{1} so that E(τ)E(\tau) in Eq. (5) is non-zero only for 0<τ<T10<\tau<T_{1}. As targets, we will consider below a model atom with a short range (SR) Yukawa-type potential V(r)=1.903er/rV(r)=-1.903e^{-r}/r, hydrogen atom, and helium atom described by means of an effective potential Sarsa et al. (2003). The target atom is initially in the ground ss- state |ϕ0|\phi_{0}\rangle with the ionization potential (IP) of 0.5 a.u. for the hydrogen and Yukawa atoms and IP of 0.902 a.u. for the He atom.

The solution to Eq. (1) is expanded as a series in the basis bispinors:

Ψ(𝐫,t)=jl=j±1/2M=jjΨjlM(𝒓,t),\Psi(\mathbf{r},t)=\sum\limits_{j\atop l=j\pm 1/2}\sum\limits_{M=-j}^{j}\Psi_{jlM}({\bm{r}},t), (6)

where:

ΨjlM(𝒓,t)=(gjlM(r,t)ΩjlM(𝒏)fjlM(r,t)ΩjlM(𝒏)),\Psi_{jlM}({\bm{r}},t)=\left(\begin{array}[]{c}g_{jlM}(r,t)\Omega_{jlM}({\bm{n}})\\ f_{jlM}(r,t)\Omega_{jl^{\prime}M}({\bm{n}})\\ \end{array}\right), (7)

and the two-component spherical spinors are defined as ΩjlM(𝒏)=(ClM121212jMYl,M12(𝒏)ClM+121212jMYl,M+12(𝒏))\displaystyle\Omega_{jlM}({\bm{n}})=\left(\begin{array}[]{c}C^{jM}_{l\ M-{1\over 2}{1\over 2}{1\over 2}}Y_{l,M-{1\over 2}}({\bm{n}})\\ C^{jM}_{l\ M+{1\over 2}{1\over 2}-{1\over 2}}Y_{l,M+{1\over 2}}({\bm{n}})\end{array}\right), (here Clm12μjMC^{jM}_{lm{1\over 2}\mu} are the Clebsch-Gordan coefficients, Ylm(𝒏)Y_{lm}({\bm{n}})- spherical harmonics, and 𝒏=𝒓/r{\bm{n}}={\bm{r}}/r). Parameters ll and ll^{\prime} in Eq. (6) must satisfy the relation l+l=2jl+l^{\prime}=2j.

To take into account the non-dipole effects due to the spatial dependence of the laser fields, vector potential (5) is expanded in a series of spherical harmonics at every time-step of the integration procedure. Substituting expansion (6) and expansion for the vector potential in the TDDE (1), and using well-known properties of spherical spinors Akhiezer and Berestetskii (1965); Lifshitz and Berestetskii (1982), one obtains a system of coupled differential equations for the radial functions gjlM(r,t)g_{jlM}(r,t) and fjlM(r,t)f_{jlM}(r,t) in Eq. (7). This system has been solved using a relativistic generalization of the well-known matrix iteration method (MIM) Nurhuda and Faisal (1999), which we described in detail in I.A.Ivanov (2015).

Appropriate choice of the propagation technique is essential, as the Dirac equation, as it is well-known, possesses some properties which are absent in the case of the non-relativistic wave-equation. These properties are due to the presence of the continuum of the negative energy states in the Dirac Hamiltonian which makes the Dirac Hamiltonian unbounded from below. One problem which this fact entails is the well-known problem of the collapse to the negative energies continuum Hill and Krauthauser (1994), which may manifest itself when basis set methods are used to construct approximations to the bound states of the Dirac Hamiltonian Hill and Krauthauser (1994). We avoid this problem, since we do not rely on the basis set methods. Initial state of the system is prepared in our calculation by solving numerically the eigenvalue equation for the field-free Dirac Hamiltonian employing shooting method. A related problem is the so-called Zitterbewegung problem Dirac (1964). Presence of a superposition of the states with positive and negative energies implies that a solution to the TDDE should exhibit very fast oscillations with characteristic frequencies of the order of c2c^{2}. Such oscillations are indeed present and we can reproduce them in the framework of our numerical procedure by using sufficiently small integration time-step Δ\Delta I.A.Ivanov (2015). We had to use the time-step Δ\Delta of the order of 10610^{-6} a.u. in I.A.Ivanov (2015) to reproduce these oscillations. Use of such small values for Δ\Delta, if it were imperative, would make any practical calculations impossible, of course. Fortunately, one can bypass this problem by using an appropriate time-propagation technique. We discussed this issue in greater detail in Ivanov and Kim (2015); I.A.Ivanov (2015). For readers convenience we present a core of the argument below. From the purely numerical point of view, presence of the fast oscillating terms in a system of the ordinary differential equations (ODE) gives us an example of a numerically stiff system of ODE, i.e. a system in which vastly different time-scales are present. To solve such a system of ODE we must use a stable integration method Shampine (1994), which ensures that while the numerical solution does not reproduce very fast oscillations, it describes accurately the overall behavior of the true solution. The integration procedure that we use provides such a stability. We can illustrate this point using a simple example of a stiff system of two ODE:

i𝒚˙=𝑨𝒚,i\dot{\bm{y}}={\bm{A}}\cdot{\bm{y}}, (8)

with Hermitian matrix 𝑨=diag(λ1(t),λ2(t)){\bm{A}}=diag(\lambda_{1}(t),\lambda_{2}(t)). To mimic the problem at hand let us assume that λ1\lambda_{1} is of order of 11, while λ2\lambda_{2} has large negative value on the interval of time that we consider. Short-time propagator in the MIM method is a unitary Crank-Nicholson (CN) propagator Goldberg et al. (1967), which relates solution vectors 𝒚n+1=𝒚(tn+1){\bm{y}}_{n+1}={\bm{y}}(t_{n+1}) and 𝒚n=𝒚(tn){\bm{y}}_{n}={\bm{y}}(t_{n}) at times tnt_{n} and tn+1=tn+Δt_{n+1}=t_{n}+\Delta as follows:

𝒚n+1=1iΔ2𝑨(tn+1/2)1+iΔ2𝑨(tn+1/2)𝒚n,{{\bm{y}}}_{n+1}={1-{i\Delta\over 2}{\bm{A}}(t_{n+1/2})\over 1+{i\Delta\over 2}{\bm{A}}(t_{n+1/2})}{{\bm{y}}}_{n}\ , (9)

where tn+1/2=tn+Δ/2t_{n+1/2}=t_{n}+\Delta/2. One can see from Eq. (9) that if at the nn-th step of the propagation the second component of the vector 𝒚{\bm{y}} acquires a numerical error δyn(2)\delta y^{(2)}_{n}, the unitarity of the CN propagation matrix in Eq. (9) makes this error remain bounded for m>nm>n.

Spatial variables in the coupled differential equations for the radial functions gjlM(r,t)g_{jlM}(r,t) and fjlM(r,t)f_{jlM}(r,t) were discretized on a grid with the step size δr=0.05\delta r=0.05 a.u., the radial variable was restricted to an interval (0,Rmax)(0,R_{\rm max}), with Rmax=400R_{\rm max}=400 a.u., and angular momenta jj up to 70 were included in the expansion (6) in the calculations below. The propagation time-step Δ\Delta was 0.050.05 a.u. Before proceeding to the description of the results of this calculation, it is instructive, however, to discuss an alternative treatment of the non-dipole effects based on the leading order perturbation theory (LOPT) expansion, as it provides a more transparent physical picture of the non-dipole effects than the complete Dirac equation. LOPT calculation described below was also used as an accuracy test for our solution to the TDDE.

II.2 LOPT treatment of the non-dipole effects.

We are interested in a LOPT solution to the TDDE considering the non-dipole effects as relativistic corrections.

The leading order relativistic corrections describing the non-dipole effects in atom-field interaction can be obtained by expanding the minimal coupling atom-field interaction Hamiltonian Lambropoulos and Petrosyan (2007); Sobelman (1972); Chelkowski et al. (2015) in the velocity gauge:

H^intmin(t)=𝒑^𝑨(𝒓,t)+𝑨^2(𝒓,t)2\hat{H}^{\rm min}_{\rm int}(t)=\hat{\bm{p}}\cdot{\bm{A}}({\bm{r}},t)+{\hat{\bm{A}}^{2}({\bm{r}},t)\over 2} (10)

in powers of c1c^{-1} I.A.Ivanov et al. (2016):

H^min(t)=p^zA(t)+v^zxE(t)c+A2(t)2+O(c2),\hat{H}_{\rm min}(t)=\hat{p}_{z}A(t)+{\hat{v}_{z}xE(t)\over c}+{A^{2}(t)\over 2}+O(c^{-2})\ , (11)

where E(t)=A(t)t\displaystyle E(t)=-{\partial A(t)\over\partial t} is the electric field of the pulse, and the velocity operator 𝒗^=𝒑^+𝑨(t)\displaystyle\hat{\bm{v}}=\hat{\bm{p}}+{\bm{A}}(t) has been introduced. The last term on the r.h.s. of Eq. (11) is a function of time only and can be removed by a unitary transformation of the wave-function.

Including spin effects in the interaction Hamiltonian is not necessary, if we are interested in the effects of the leading order in powers of c1c^{-1} Chelkowski et al. (2015); Zhu and Wang (2016). The fact that the spin degrees of freedom can be neglected in the leading order of the c1c^{-1} expansion, can be understood using the semi-classical picture of the spin effects, in which additional force due to the presence of the spin degrees of freedom, acting on the electron, is 𝑭=Um{\bm{F}}=-\nabla U_{m}, where Um=𝝁𝑯U_{m}=-{\bm{\mu}}\cdot{\bm{H}}, energy of the spin-magnetic field interaction. Here 𝑯{\bm{H}} is the magnetic field and 𝝁{\bm{\mu}} is electron’s magnetic moment related to the expectation value of electron’s spin 𝝁=2𝑺/c{\bm{\mu}}=-2{\bm{S}}/c. Spatial gradient of 𝑯{\bm{H}} introduces an additional factor of c1c^{-1}, making contribution of the force 𝑭\bm{F} an effect of higher order in c1c^{-1}. As for the relativistic corrections to the field-free atomic Hamiltonian, the so-called Breit-Pauli Hamiltonian Sobelman (1972), it adds terms of the order of c2c^{-2} to the non-relativistic atomic Hamiltonian. We do not have, therefore, to include these corrections in the LOPT treatment. To the leading order in powers of the c1c^{-1}-expansion, the dynamics of the system can thus be described by the time-dependent Schrödinger equation (TDSE):

iΨ(𝒓,t)t=(H^atom+H^d(t)+H^nd(t))Ψ(𝒓,t),i{\partial\Psi({\bm{r}},t)\over\partial t}=\left(\hat{H}_{\rm atom}+\hat{H}_{\rm d}(t)+\hat{H}_{\rm nd}(t)\right)\Psi({\bm{r}},t)\ , (12)

where

H^atom=𝒑^22+V(r)\hat{H}_{\rm atom}={\hat{\bm{p}}^{2}\over 2}+V(r) (13)

is atomic field-free Hamiltonian,

H^d(t)=p^zA(t)\hat{H}_{\rm d}(t)=\hat{p}_{z}A(t) (14)

is the dipole part of the atom-field interaction and

H^nd(t)=v^zxE(t)c\hat{H}_{\rm nd}(t)={\hat{v}_{z}xE(t)\over c} (15)

is the non-dipole part of the atom-field interaction containing the effects of the order of c1c^{-1}.

It is easy to check that the LOPT solution to the equation (12), with the non-dipole term (15) considered as a perturbation, can be written as:

ΨLOPT(𝒓,t)=Ψd(𝒓,t)+Ψnd(1)(𝒓,t),\Psi^{\rm LOPT}({\bm{r}},t)=\Psi_{\rm d}({\bm{r}},t)+\Psi^{(1)}_{\rm nd}({\bm{r}},t)\ , (16)

where the LOPT non-dipole correction is given by the expression:

Ψnd(1)(𝒓,t)=i0tU^d(t,τ)H^nd(τ)Ψd(𝒓,τ)𝑑τ.\Psi^{(1)}_{\rm nd}({\bm{r}},t)=-i\int\limits_{0}^{t}\hat{U}_{\rm d}(t,\tau)\hat{H}_{\rm nd}(\tau)\Psi_{\rm d}({\bm{r}},\tau)\ d\tau\ . (17)

As can be seen from Eq. (15) for the operator H^nd\hat{H}_{\rm nd} this correction is of the order of c1c^{-1}. In Eq. (16) and Eq. (17) Ψd(𝒓,t)\Psi_{\rm d}({\bm{r}},t) is the zero-order solution to the non-relativistic TDSE taking into account only the dipole part of the atom-field interaction, U^d(t,τ)\hat{U}_{\rm d}(t,\tau) is the evolution operator describing evolution of the system driven by the non-relativistic dipole Hamiltonian. U^d(t,τ)\hat{U}_{\rm d}(t,\tau) satisfies the operator equation:

iU^d(t,τ)t=(H^atom+H^d(t))U^d(t,τ),i{\partial\hat{U}_{\rm d}(t,\tau)\over\partial t}=\left(\hat{H}_{\rm atom}+\hat{H}_{\rm d}(t)\right)\hat{U}_{\rm d}(t,\tau)\ , (18)

and the initial condition U^d(τ,τ)=I^\hat{U}_{\rm d}(\tau,\tau)=\hat{I}. In practice, we need not solve the operator equation (18). All we have to do to compute the expression under the integral on the r.h.s of Eq. (16) for given τ\tau and tt, is to propagate first the initial state wave-function on the interval (0,τ)(0,\tau) using the non-relativistic TDSE with the Hamiltonian (14), obtaining thus a state vector Ψd(τ)\Psi_{\rm d}(\tau). We act than on this vector with the operator H^nd(τ)\hat{H}_{\rm nd}(\tau) and propagate it further in time till the moment tt. The non-relativistic TDSE was solved using the well-tested numerical procedure described in Ivanov (2014).

II.3 Calculation of electron velocity and HHG spectra

Once the solution to the TDDE (1) is obtained, expectations value of the electron velocity can be obtained as Avetissian et al. (2011):

𝒗(t)=cΨ(t)|𝜶|Ψ(t).{\bm{v}}(t)=c\langle\Psi(t)|{\bm{\alpha}}|\Psi(t)\rangle\ . (19)

Harmonic spectra can then be calculated using the usual semi-classical approach, in which the spectral intensity of the harmonic emission can be expressed in terms of the Fourier transform of electron’s velocity:

Sa(Ω)|0T1va(t)W(t)eiΩt𝑑t|2.S_{a}(\Omega)\propto\left|\int\limits_{0}^{T_{1}}v_{a}(t)W(t)e^{i\Omega t}\ dt\right|^{2}\ . (20)

where va(t)v_{a}(t) is either xx- or zz- component of the electron velocity for the non-dipole and dipole harmonic intensities Sx(Ω)S_{x}(\Omega) and Sz(Ω)S_{z}(\Omega), respectively. In the velocity form for the harmonics intensity which we use here, we do not need to introduce additional powers of harmonic frequency, which would be present had we used length or acceleration forms Baggesen and Madsen (2011). The factor W(t)W(t) in Eq. (20) is the window function Reiff et al. (2020), for which we employ the Hann form: W(t)=sin2(πtT1)\displaystyle W(t)=\sin^{2}{\left(\pi t\over T_{1}\right)}.

The most noticeable effects which the relativistic non-dipole corrections produce are appearance of harmonic photons polarized in the laser propagation direction Potvliege et al. (2000); Kylstra et al. (2001); Chirilă et al. (2002) and appearance of even order harmonics in the HHG spectra Mishra et al. (2012); Mu-Xue et al. (2020). The LOPT picture allows to explain these features transparently. Substituting the expression Eq. (16) for the LOPT wave-function into the matrix element:

ΨLOPT(t)|𝒗^|ΨLOPT(t)𝒙^vx(t)+𝒚^vy(t)+𝒛^vz(t),\langle\Psi^{\rm LOPT}(t)|\hat{\bm{v}}|\Psi^{\rm LOPT}(t)\rangle\approx\hat{\bm{x}}v_{x}(t)+\hat{\bm{y}}v_{y}(t)+\hat{\bm{z}}v_{z}(t)\ , (21)

defining the leading order contributions to the expectation value of electron velocity, one obtains:

vz(t)=Ψd(t)|v^z|Ψd(t).v_{z}(t)=\langle\Psi_{\rm d}(t)|\hat{v}_{z}|\Psi_{\rm d}(t)\rangle\ . (22)

For the geometry we use, the evolution operator U^d(t,τ)\hat{U}_{\rm d}(t,\tau) commutes with l^z\hat{l}_{z}- the zz- component of the angular momentum, i.e., it is a conserved quantity for the quantum evolution driven by the dipole Hamiltonian (13) and (14). l^z{\hat{l}}_{z}, therefore, has a definite value lz=0l_{z}=0 in the state described by the wave-function Ψd(t)\Psi_{\rm d}(t), and the matrix element Ψd(t)|v^x|Ψd(t)\displaystyle\langle\Psi_{\rm d}(t)|\hat{v}_{x}|\Psi_{\rm d}(t)\rangle vanishes because of the well-known dipole selection rules Sobelman (1972). Leading order contribution to vx(t)v_{x}(t), is, therefore, of the order of c1c^{-1}, and is given by the expression:

vx(t)\displaystyle v_{x}(t) =\displaystyle= Ψd(t)|v^x|Ψnd(1)+Ψnd(1)(t)|v^x|Ψd\displaystyle\langle\Psi_{\rm d}(t)|\hat{v}_{x}|\Psi^{(1)}_{\rm nd}\rangle+\langle\Psi^{(1)}_{\rm nd}(t)|\hat{v}_{x}|\Psi_{\rm d}\rangle
=\displaystyle= 2ReΨd(t)|v^x|Ψnd(1)\displaystyle 2{\rm Re}\langle\Psi_{\rm d}(t)|\hat{v}_{x}|\Psi^{(1)}_{\rm nd}\rangle
=\displaystyle= 2Im(0tΨd(t)|p^xU^d(t,τ)H^nd(τ)|Ψd(τ)𝑑τ).\displaystyle 2{\rm Im}\left(\int\limits_{0}^{t}\langle\Psi_{\rm d}(t)|\hat{p}_{x}\hat{U}_{\rm d}(t,\tau)\hat{H}_{\rm nd}(\tau)|\Psi_{\rm d}(\tau)\rangle\ d\tau\right)\ .

In the last line of Eq. (LABEL:velx) we used expression (17) for Ψnd(1)\Psi^{(1)}_{\rm nd}. The same dipole selection rules Sobelman (1972) and the structure of Eq. (17) ensure that the contribution of the order of c1c^{-1} to vy(t)v_{y}(t) is zero. The leading contribution of the non-dipole effects is, therefore, non-zero only for the xx-component of the electron velocity. Orientation of the dipole velocity due to this relativistic contribution results, thus, in the appearance of the harmonic photons polarized in the propagation direction in accordance with the observations made in Potvliege et al. (2000); Kylstra et al. (2001); Chirilă et al. (2002).

As we mentioned above, the appearance of the even order harmonics can be understood as a result of violation of the symmetry of the electron trajectories responsible for the emission of harmonic photons in the dipole approximation Lewenstein et al. (1994). From the LOPT perspective this effect can be explained as follows. As one can see from Eq. (14) and Eq. (15), the dipole interaction operator (14) has odd parity, i.e. it couples states of different parities, while the non-dipole operator (15) has even parity. Employing a somewhat lousy language, we might say that the presence of these two atom-field interaction Hamiltonians can be described as the presence of two kinds of photons: the ”dipole” photons and the ”non-dipole” photons, whose emission and absorption are governed by the operators (14) and (15), respectively. Using these notions and the LOPT expression for vx(t)v_{x}(t) in Eq. (LABEL:velx), contribution of the non-dipole interaction to the formation of the NN-th harmonic can be described as absorption of N1N-1 ”dipole” photons and one ”non-dipole” photon, with subsequent recombination to atomic ground state accompanied by emission of a harmonic photon with frequency NωN\omega. Using the informal terminology which we adopted, one might say that the emitted harmonic photon is of the ”dipole” nature since spontaneous emission satisfies the dipole selection rules. Conservation of the total parity for the combined system of atom and the ”dipole” and the ”non-dipole” photons implies then that NN must necessarily be even.

Besides providing a simple physical picture of the appearance of even harmonics, the LOPT approach which we described above, can be used as a test of the accuracy of our solution to the TDDE. To do such a test we performed calculations of the expectation values of electron velocity using TDDE and LOPT approaches for the cosine-pulse form shown in Fig. 1, with the vector potential in Eq. (5) given by the equation: 𝑨(x,t)=𝒆zE0ωsin2(πuT1)sinωu\displaystyle{\bm{A}}(x,t)=-{\bm{e}}_{z}{E_{0}\over\omega}\sin^{2}{\left(\pi u\over T_{1}\right)}\sin{\omega u} where ω=0.057\omega=0.057 a.u., E0=0.0534E_{0}=0.0534 a.u., u=tx/cu=t-x/c. A comparison of the TDDE results obtained using Eq. (19) and the LOPT results obtained using Eq. (LABEL:velx) for the xx- component of electron velocity is shown in Fig. 2. The results of the LOPT treatment prove to be virtually identical to the results of the TDDE calculation which is not surprising given that the relativistic corrections could be expected to be small for the field parameters we consider.

III Results

We report below results which we obtained from our TDDE calculations for dipole Sz(Ω)S_{z}(\Omega) and non-dipole Sx(Ω)S_{x}(\Omega) harmonic intensities for different targets. HHG spectra were obtained by computing electron velocity as prescribed by Eq. (19) and using Eq. (20) to compute harmonic intensities. Calculations were performed using the sine waveform shown in Fig. 1 with the electric field given by the equation: E(u)=E0sin2(πuT1)sinωu\displaystyle E(u)=E_{0}\sin^{2}{\left(\pi u\over T_{1}\right)}\sin{\omega u}. We report below results for the base frequencies ω=0.114\omega=0.114 a.u. (wavelength of 400400 nm) and ω=0.057\omega=0.057 a.u. (wavelength of 800800 nm).

In Fig. 3 we show HHG spectra that we obtained for the driving pulse wavelength λ=400\lambda=400 nm and different field strengths for various targets. Fig. 3 shows both dipole Sz(Ω)S_{z}(\Omega) and non-dipole Sx(Ω)S_{x}(\Omega) harmonic intensities. The vertical lines in the Figures show positions of the classical cutoffs given by the well-known 3.17Up+Ip3.17U_{p}+I_{p} (here Up=E02/4ω2U_{p}=E_{0}^{2}/4\omega^{2} and IpI_{p} are ponderomotive and ionization energies respectively) rule of the three-step model Lewenstein et al. (1994); Corkum (1993). In Fig. 4 we zoom on the parts of the harmonic spectra more closely to demonstrate the presence of odd and even harmonics in the dipole and non-dipole spectra respectively.

Quite expectedly, behavior of the dipole intensity Sz(Ω)S_{z}(\Omega) shown in Fig. 3 agrees very well with the three-step model predictions, exhibiting a sharp drop in magnitude after reaching the classical cutoff. The non-dipole Sx(Ω)S_{x}(\Omega) spectra mimic this behavior very closely. This may be not surprising if we make use again of the LOPT picture of formation of the non-dipole harmonics we presented above, which relied on the notions of ’dipole’ and ’non-dipole’ photons with operators describing their interactions with an atom given by Eq. (14) and Eq. (15), respectively. We remind, that in the framework of this picture the NN-th non-dipole harmonic is produced as a result of the absorption of N1N-1 ”dipole” photons and one ”non-dipole” photon. As far as the harmonic spectra are concerned, the mechanism responsible for the formation of the non-dipole harmonic emission differs thus from the mechanism of the emission of the dipole harmonics only in the replacement of one ’dipole’ photon with a ’non-dipole’ one. This replacement leads to the replacement of the odd order harmonics in the spectra by the even order ones and results in an overall drop in magnitude in the harmonic spectra due to the presence of the additional factor of c1c^{-1} in the non-dipole interaction operator (15).

The energy and parity conservation considerations which lead us to the general conclusions about the character of the non-dipole spectra do not tell us anything about temporal dynamics of the formation of the non-dipole harmonics. We can have a glimpse of this temporal dynamics by analyzing Gabor transforms Gabor (1946) of dipole and non-dipole velocities:

Ta(Ω,t)=0T1va(τ)Φ(t,τ,Ω)𝑑τ,T_{a}(\Omega,t)=\int\limits_{0}^{T_{1}}v_{a}(\tau)\Phi^{*}(t,\tau,\Omega)d\tau\ , (24)

where Φ(t,τ,Ω)=exp{iΩτ(tτ)2/2(x0T)2}\displaystyle\Phi(t,\tau,\Omega)=\exp{\left\{i\Omega\tau-(t-\tau)^{2}/2(x_{0}T)^{2}\right\}}, parameter x0x_{0} determines resolution in the temporal domain, and TT is an optical cycle of the laser field. Gabor transform, as well as closely related wavelet transform, allows us to take a look simultaneously at both time and frequency domains, and allows to determine, in particular, when different harmonics are emitted Antoine et al. (1995); Tang et al. (2000); Wang et al. (2005). We used x0=0.1x_{0}=0.1 in the calculations below. This value of x0x_{0} gives us rather poor resolution in the frequency domain, but high resolution in the time domain, which is of interest to us presently.

The absolute values |Ta(Ω,t)||T_{a}(\Omega,t)| for both dipole and non-dipole velocities are shown in Fig. 5 and Fig. 6 for the SR Yukawa and hydrogen atoms. One can see that, dynamically, formation of dipole and non-dipole harmonics proceeds quite differently. For both Yukawa and hydrogen atoms systems emission of the non-dipole harmonics is strongly suppressed at the early stages of pulse development, and emission times for the non-dipole harmonics are shifted with respect to the dipole radiation bursts. Such behavior could be anticipated by looking at Fig. 2 which shows that xx- component of the velocity starts actually respond to the field only for times approaching the midpoint of the pulse. The reason for this could be traced back to the character of the fully quantum expression for the velocity component vxv_{x} in the second LOPT equation (LABEL:velx), with time integration on the right-hand side of this equation smoothing out high frequency oscillations. To elucidate this issue further we performed a simple classical calculation of the emitted photon energy as a function of the recombination time using the physical picture provided by the three-step model. We assume that electron is ionized at the moment of time tiont_{ion} and returns to the parent ion at the moment of time trett_{ret}, emitting a harmonic photon with energy Eret+IpE_{ret}+I_{p}. As is usually assumed in the three-step model calculations, we consider only the effect of the external field (5) on the electron motion, neglecting completely ionic potential. The only difference of our calculation and the traditional three-step model analysis of the harmonic emission, is that we take into account effect of the Lorentz force due to the magnetic field of the pulse. We simulate electron motion in a plane (which is the (x,z)(x,z)- plane for the geometry we employ), solving the set of the classical Newton equations, which for the fields configuration, geometry and atomic units system we employ, can be written as:

x¨\displaystyle\ddot{x} =\displaystyle= vzcE(t)\displaystyle-{{v}_{z}\over c}E(t)
z¨\displaystyle\ddot{z} =\displaystyle= E(t)+vxcE(t)\displaystyle-E(t)+{{v}_{x}\over c}E(t)
.\displaystyle\ . (25)

Following the prescription of the traditional three-step model we solve equations (25) with zero initial conditions imposed at the ionization time: vx(tion)=vz(tion)=0v_{x}(t_{ion})=v_{z}(t_{ion})=0 and x(tion)=z(tion)=0x(t_{ion})=z(t_{ion})=0. We assume that the electron trajectory returns to the origin, if at the moment of time trett_{ret}, zz-coordinate of the electron trajectory changes sign.

Fig. 7(a) shows results of such a simulation, which qualitatively agree with the dynamics of the dipole harmonics emission shown in Fig. 5 and Fig. 6, with bursts of harmonics emission occurring every half cycle of the laser pulse. To be able to apply this classical analysis to the emission of the non-dipole harmonics we must, however, introduce one essentially quantum ingredient in the model described by the classical equations(25). Emission of the non-dipole radiation differs from the emission of the dipole harmonics in one important aspect. For the geometry we employ, the dipole harmonics photon emission process satisfies selection rule ΔM=0\Delta M=0, where MM is the zz- projection of the electron angular momentum. On the other hand, emission of the non-dipole harmonic photon, as can be seen from the LOPT analysis we presented above, must satisfy selection rule ΔM=±1\Delta M=\pm 1. This means that for the ground ss- state that we consider, non-dipole radiation can be emitted only by electrons with non-zero angular momentum. We can incorporate this fact in our classical model by introducing a filter parameter ff in the simulations, and considering only those returning trajectories for which at the moment of time trett_{ret} squared classical angular momentum value exceeds the threshold value set by the filter parameter ff. Results of such calculations are shown in Fig. 7(b-d) for different values of the filter parameter ff. One can see that by increasing the value of the filter parameter, we make the classical picture in Fig. 7 look more like the Gabor transform results shown in Fig. 5 and Fig. 6. In particular, Fig. 7(c-d) show the absence of the non-dipole harmonics emission during the first two cycles of the laser pulse, the feature which is also demonstrated by the quantum analysis based on the Gabor transform in Fig. 5 and Fig. 6. Applying non-zero filter parameter does not change, however, the maximum energy EretE_{ret} of the returning electron, which explains why non-dipole harmonic emission spectra exhibit essentially the same cutoffs as the dipole harmonic emission spectra. This simple classical picture of the formation of the non-dipole harmonics, which takes as quantum ingredient only the requirement that the electron angular momentum on the returning trajectories should exceed certain threshold value, agrees, thus, qualitatively with the fully quantum picture.

We also performed TDDE calculations for the pulse base frequency ω=0.057\omega=0.057 a.u. (corresponding to the wavelength of 800800 nm). In Fig. 8 and Fig. 9 we show harmonic spectra we obtain from TDDE for the SR Yukawa and hydrogen atoms. Fig. 10 shows results of the analysis of the temporal dynamics of the harmonic formation based on the Gabor transform (24). These Figures show essentially the same picture as the results we presented above for the driving pulse wavelength of 400400 nm. The spectra of the non-dipole harmonics follow closely the classical dipole cutoff rule, and differ in this respect from the dipole emission spectra only in their intensity. Temporal pictures of the harmonics formation in the dipole and the non-dipole cases are, however, totally different. The main difference is, just as in the case of the driving pulse wavelength of 400400 nm, the absence of the harmonic emission at the early stages of the pulse development, the feature which we explained above using the results of the classical calculations shown in Fig. 7(c,d).

The factor which is responsible for the difference in intensity between the dipole and non-dipole harmonics is the additional factor of c1c^{-1} which, as one can see from Eq. (LABEL:velx) and Eq. (15), is present in the LOPT formula for the xx-component of the velocity. The presence of this factor in vxv_{x} leads to a dampening factor of c2c^{-2} in the expression for the non-dipole harmonics intensity. It is rather difficult to obtain a more detailed insight about relative magnitude of the dipole and non-dipole harmonic intensities form the cumbersome LOPT expressions Eq. (22) and Eq. (LABEL:velx). One can, however, obtain a simple estimate using the reasoning based not on the Schrödinger picture that we have used so far, but on the equivalent Heisenberg picture of the quantum mechanics (QM). In the latter, we remind, the operators evolve in time, while the state vectors do not. We obtain, of course, the same expectation values for all physical observables in both pictures.

In the Heisenberg picture time-evolution of the operators 𝒓^(t){\hat{\bm{r}}}(t) and 𝒑^(t){\hat{\bm{p}}}(t) is described by the equations Landau and Lifshitz (1977): i𝒓^˙=[𝒓^,H^]\displaystyle i{\dot{\hat{\bm{r}}}}=[\hat{{\bm{r}}},\hat{H}], i𝒑^˙=[𝒑^,H^]\displaystyle i{\dot{\hat{\bm{p}}}}=[{\hat{\bm{p}}},\hat{H}], where the Hamiltonian operator in our problem is H^=H^atom+H^d(t)+H^nd(t)\hat{H}=\hat{H}_{\rm atom}+\hat{H}_{\rm d}(t)+\hat{H}_{\rm nd}(t), with H^atom\hat{H}_{\rm atom}, H^d(t)\hat{H}_{\rm d}(t) and H^nd(t)\hat{H}_{\rm nd}(t) given by Eq. (13), Eq. (14), and Eq. (15), respectively. Calculating the commutators, one obtains the following equations of motion:

x^˙\displaystyle\dot{\hat{x}} =\displaystyle= p^x\displaystyle\hat{p}_{x}
p^˙x\displaystyle\dot{\hat{p}}_{x} =\displaystyle= i[p^x,V^]v^zcE(t)\displaystyle-i[{\hat{p}}_{x},\hat{V}]-{\hat{v}_{z}\over c}E(t)
z^˙\displaystyle\dot{\hat{z}} =\displaystyle= p^z+A(t)+x^cE(t)\displaystyle\hat{p}_{z}+A(t)+{{\hat{x}}\over c}E(t)
p^˙z\displaystyle\dot{\hat{p}}_{z} =\displaystyle= i[p^z,V^],\displaystyle-i[{\hat{p}}_{z},\hat{V}]\ ,

where V^\hat{V} is atomic potential operator, v^z=p^z+A(t)\hat{v}_{z}=\hat{p}_{z}+A(t), A(t)A(t) and E(t)E(t) are the vector potential and the electric field of the pulse. Eq. (LABEL:he) is the quantum-mechanical analogue of the classical equations describing electron motion in the potential VV in presence of the Lorenz force. It contains the same physical information and is, therefore, equivalent to the LOPT equations Eq. (22) and Eq. (LABEL:velx), but it provides a more clear physical picture and can be used as a starting point for making simplifying assumptions.

From the first two equations (LABEL:he) one obtains:

x^¨=i[p^x,V^]v^zcE(t),\ddot{\hat{x}}=-i[{\hat{p}}_{x},\hat{V}]-{\hat{v}_{z}\over c}E(t)\ , (27)

We will make an assumption that one can omit the commutator [p^x,V^][{\hat{p}}_{x},\hat{V}] in Eq. (27). Some justification for this operation can be provided in the case of the SR Yukawa atom, when potential function V(𝒓)V({\bm{r}}) is effectively zero everywhere excepting a small neighborhood of the atom. We obtain then from Eq. (27) a relation for the expectation values of the electron acceleration ax=ϕ0|x^¨|ϕ0\displaystyle a_{x}=\langle\phi_{0}|\ddot{\hat{x}}|\phi_{0}\rangle and velocity vz=ϕ0|v^z|ϕ0\displaystyle v_{z}=\langle\phi_{0}|{\hat{v}}_{z}|\phi_{0}\rangle:

ax=vzcE(t),a_{x}=-{v_{z}\over c}E(t)\ , (28)

where |ϕ0|\phi_{0}\rangle is the initial atomic state, which does not evolve in time in the Heisenberg picture. Assuming further that E(t)E(t) is a monochromatic wave: E(t)=E0cosωtE(t)=E_{0}\cos{\omega t} and calculating Fourier transforms of both sides of Eq. (28), we obtain a relation between the Fourier transforms v~x(Ω)=vx(t)eiΩt𝑑t\displaystyle\tilde{v}_{x}(\Omega)=\int v_{x}(t)e^{i\Omega t}\ dt and v~z(Ω)=vz(t)eiΩt𝑑t\displaystyle\tilde{v}_{z}(\Omega)=\int v_{z}(t)e^{i\Omega t}\ dt:

iΩv~x(Ω)=E02c(v~z(Ω+ω)+v~z(Ωω)),-i\Omega\tilde{v}_{x}(\Omega)={E_{0}\over 2c}\left(\tilde{v}_{z}(\Omega+\omega)+\tilde{v}_{z}(\Omega-\omega)\right)\ , (29)

from which, using the fact that for any complex numbers z1z_{1}, z2z_{2}: |z1+z2|2(|z1|+|z2|)2\displaystyle|z_{1}+z_{2}|^{2}\leq(|z_{1}|+|z_{2}|)^{2}, we obtain an inequality:

Ω2Sx(Ω)E024c2(Sz(Ω+ω)+Sz(Ωω))2\Omega^{2}S_{x}(\Omega)\leq{E_{0}^{2}\over 4c^{2}}\large(\sqrt{S_{z}(\Omega+\omega)}+\sqrt{S_{z}(\Omega-\omega)}\large)^{2} (30)

We see from Eq. (30) that for Ω>ω\Omega>\omega we have:

R(Ω)=4c2ω2E02Sx(Ω)(Sz(Ω+ω)+Sz(Ωω))21.R(\Omega)={4c^{2}\omega^{2}\over E_{0}^{2}}{S_{x}(\Omega)\over\large(\sqrt{S_{z}(\Omega+\omega)}+\sqrt{S_{z}(\Omega-\omega)}\large)^{2}}\leq 1\ . (31)

Introducing the magnitude A0=E0/ωA_{0}=E_{0}/\omega of the pulse vector potential, we can rewrite inequality (31) as:

Sx(Ω)(Sz(Ω+ω)+Sz(Ωω))2A024c2{S_{x}(\Omega)\over\large(\sqrt{S_{z}(\Omega+\omega)}+\sqrt{S_{z}(\Omega-\omega)}\large)^{2}}\leq{A_{0}^{2}\over 4c^{2}} (32)

The ratio R(Ω)R(\Omega) defined in Eq. (31) is shown in Fig. 11(a) for the SR Yukawa potential and various pulse parameters. Of course, we cannot expect Eq. (31) to provide a rigorous upper bound since deriving it we neglected atomic potential in Eq. (27), which constitutes a rather drastic approximation. As one can see from Fig. 11(a) inequality (31) can indeed be violated. One can see, nevertheless, that Eq. (31), and consequently Eq. (32) provide reasonably accurate estimates of the relative magnitude of the intensities of the dipole and non-dipole harmonics.

While the non-zero expectation value vxv_{x} and appearance of the non-dipole harmonics is an entirely relativistic phenomena, the non-dipole effects also modify slightly the velocity component vzv_{z}. The magnitude of this effect is of the order of c2c^{-2}. This can be most easily seen from the Heisenberg equations of motion (LABEL:he). The equation for vz(t)v_{z}(t) (the third of the equations (LABEL:he)) contains the term x^E(t)/c{\hat{x}}E(t)/c on the right-hand side. Since the expectation value of xx is itself of the order of c1c^{-1}, the resulting effect on vz(t)v_{z}(t) is of the order of c2c^{-2}, which will produce a relativistic correction of the order of c2c^{-2} for the dipole harmonic intensity. We may expect, therefore, that the normalized difference:

ΔSz(Ω)Sz(Ω)=Sz(Ω)Sznr(Ω)Sznr(Ω),{\Delta S_{z}(\Omega)\over S_{z}(\Omega)}={S_{z}(\Omega)-S^{\rm nr}_{z}(\Omega)\over S^{\rm nr}_{z}(\Omega)}\ , (33)

where Sz(Ω)S_{z}(\Omega) is the dipole harmonics intensity obtained in the present TDDE calculation and Sznr(Ω)S^{\rm nr}_{z}(\Omega) is the result of the non-relativistic TDSE calculation, should be of the order of c2c^{-2}. i.e., we may expect ΔSz(Ω)/Sz(Ω)104\Delta S_{z}(\Omega)/S_{z}(\Omega)\sim 10^{-4}. That this is indeed the case can be seen from Fig. 11(b), where we show results of the TDDE and TDSE calculations performed for the same pulse parameters for the Yukawa atom.

The analysis based on the Heisenberg equations of motion (LABEL:he) also allows to give a simple explanation for the behavior of vx(t)v_{x}(t) shown in Fig. 2, where the xx- component of the electron velocity starts responding to the field only for the times approaching the midpoint of the pulse. Integrating Eq. (28) we obtain for the expectation value vx=ϕ0|v^x|ϕ0\displaystyle v_{x}=\langle\phi_{0}|{\hat{v}}_{x}|\phi_{0}\rangle (assuming that it has zero value at t=0t=0):

vx(t)=1c0tvz(τ)E(τ)𝑑τ.v_{x}(t)=-{1\over c}\int\limits_{0}^{t}v_{z}(\tau)E(\tau)\ d\tau\ . (34)

We could have obtained the same equation by integrating the first of the set of the classical equations (25), which is not surprising given the great formal similarity between the classical mechanics and the QM in the Heisenberg picture. We show in Fig. 12(a) the expectation value vz(t)v_{z}(t) obtained in the LOPT calculation for the cosine pulse with E0=0.0534E_{0}=0.0534 a.u. and ω=0.057\omega=0.057 a.u. We show only the LOPT result. Just as in the case of vx(t)v_{x}(t), shown in Fig. 2, the TDDE and LOPT results for vz(t)v_{z}(t) differ very slightly. In Fig. 12(b) we show the LOPT expectation value vx(t)v_{x}(t), as well as the estimate for vx(t)v_{x}(t) that we obtain if we substitute the LOPT value for vz(τ)v_{z}(\tau) under the integral sign in Eq. (34). One can see that the estimate thus obtained reproduces fairly well the general behavior of vx(t)v_{x}(t). In particular, it reproduces the feature that we mentioned above: the xx-component of the velocity begins deviating from zero appreciably only for the times approaching the midpoint of the pulse. We remind that effect of the atomic potential on the motion in the xx-direction was neglected in the Heisenberg equation of motion (28) which we used to obtain the estimate (34). The fact that the estimate (34) reproduces qualitative behavior of the xx-component of electron velocity shown in Fig. 2 tells us, therefore, that this behavior is a result of the interplay of the motion in xx- and zz- directions which are mutually interconnected due to presence of the Lorentz force.

IV Conclusion

We have presented results of the relativistic calculations of even harmonic generation from various atomic targets. Our approach was based on the numerical solution of the TDDE. The HHG spectra of the non-dipole even order harmonics were found to look qualitatively similar to the spectra of the dipole harmonics, obeying the same classical cutoff rules. The temporal formation of the non-dipole harmonics, however, was found to be quite different. The results of the Gabor transform analysis show that formation of the non-dipole harmonics is strongly suppressed at the beginning of the laser pulse, and bursts of the non-dipole radiation are shifted in time with respect to the bursts of the dipole emission. These features are partly explained by a simple generalization of the classical three-step model, which takes into account the selection rules governing emission of harmonic photons. We modeled the effect of these selection rules by using a filter parameter, which selects the trajectories with angular momentum exceeding a certain threshold value at the recollision time.

For the field parameters we considered the relativistic effects are still relatively weak and could be described perturbatively. LOPT provides, as we have seen, an adequate description of the non-dipole effects responsible for the even order harmonics emission. Use of the TDDE, however, is technically simpler than the calculations based on the LOPT, and opens the perspective of making an excursion into the truly relativistic domain in the future. We relied, therefore, on the TDDE-based approach in the present work. The present approach can also be generalized relatively easily to include some quantum electrodynamical (QED) effects, such as the vacuum polarization effects, or the QED strong Coulomb field radiative corrections, which can be taken into account by using effective potentials such as the Uehling potential Uehling (1935a, b) or the radiative potential proposed in Flambaum and Ginges (2005). The procedure we apply to solve the Dirac equation can also be used to study the process of electron-positron pair production (PP) in strong electromagnetic fields, which occurs when field strength reaches the characteristic Schwinger field strength of 1.3×10161.3\times 10^{16} V/cm. The process of PP in both homogeneous and inhomogeneous electric fields has received considerable interest in the literature Gies and Klingmüller (2005). Theoretical treatment of PP in the semiclassical approximation relies on a solution of the TDDE for a given field configuration Mocken et al. (2010). Our procedure might prove useful for this purpose, especially in the case of the spatially inhomogeneous field, which has been found to play an important role in the PP Gies and Klingmüller (2005); Kohlfürst (2020).

The numerical procedure we employ can be relatively easily generalized for the case of the many-electron relativistic Hamiltonians used in the quantum chemistry calculations Nakai (2021); Liu (2020). Use of the representation of the wave-function analogous to the expansion (6) would be, of course, impractical for systems with more than one electron if we want to use such expansions to represent the wave-function in the whole space. One may use, however, the idea of the RR-matrix approach, which separates the coordinate space in the inner region, where a suitable basis set representation can be used to represent many-electron wave-functions and the outer region, where one has to concentrate on the description of a single electron motion, for which the finite difference method might be better suited. Such a strategy has been implemented with success in the framework of the so-called R-Matrix incorporating Time method (RMT) Lysaght et al. (2012) which allows to solve the non-relativistic TDSE for many-electron systems. One can use a similar approach in the relativistic case, relying on the results of the stationary quantum chemistry calculations Nakai (2021); Liu (2020) for the description of the inner region, where many-electron effects are important, and using the present procedure to solve the TDDE describing electron propagation in the outer region.

Acknowledgments

This work was supported by the Institute for Basic Science grant (IBS-R012-D1) and the National Research Foundation of Korea (NRF), grant funded by the Korea government (MIST) (No. 2022R1A2C3006025). Computational works for this research were performed on the IBS Supercomputer Aleph in the IBS Research Solution Center.

Refer to caption
Figure 1: (Color online) Pulse shapes E(t)E(t) employed in the calculations.
Refer to caption
Figure 2: (Color online) Expectation value of the xx- component of the electron velocity as a function of time obtained in TDDE and LOPT calculations Cosine pulse with E0=0.0534E_{0}=0.0534 a.u., ω=0.057\omega=0.057 a.u. has been used in the calculation.
Refer to caption
Figure 3: (Color online) Dipole (dash green) and non-dipole (red solid) harmonic intensities for the pulse wavelength λ=400\lambda=400 nm for the SR Yukawa, hydrogen and He atoms. Vertical dash lines show cutoff positions.
Refer to caption
Figure 4: (Color online) Dipole (dash green) and non-dipole (red solid) harmonic intensities for the pulse wavelength λ=400\lambda=400 nm for harmonics with orders n20n\leq 20 for the SR Yukawa and hydrogen atoms. Vertical solid and dash lines show positions of odd and even harmonics, respectively.
Refer to caption
Figure 5: (Color online) Gabor transform |T(Ω,t)||T(\Omega,t)| for the pulse wavelength λ=400\lambda=400 nm and different field strengths for the SR Yukawa atom.
Refer to caption
Figure 6: (Color online) Gabor transform |T(Ω,t)||T(\Omega,t)| for the pulse wavelength λ=400\lambda=400 nm and different field strengths for the hydrogen atom.
Refer to caption
Figure 7: (Color online) Classical calculations of emitted photon energy as function of return time for the dipole harmonic radiation (a) and non-dipole harmonic radiation (b-d) with different filter parameters.
Refer to caption
Figure 8: (Color online) Dipole (dash green) and non-dipole (red solid) harmonic intensities for the pulse wavelength λ=800\lambda=800 nm. for the SR Yukawa and hydrogen atoms. Vertical dash lines show cutoff positions.
Refer to caption
Figure 9: (Color online) Dipole (dash green) and non-dipole (red solid) harmonic intensities for the pulse wavelength λ=800\lambda=800 nm for harmonics with orders 40n6040\leq n\leq 60 for the SR Yukawa and hydrogen atoms. Vertical solid and dash lines show positions of odd and even harmonics, respectively.
Refer to caption
Figure 10: (Color online) Gabor transform |T(Ω,t)||T(\Omega,t)| for the pulse wavelength λ=800\lambda=800 nm for the SR Yukawa and hydrogen atoms.
a)             b)
Figure 11: (Color online) (a) Estimate (31) for the ratio R(Ω)R(\Omega). (b) Normalized difference (Sz(Ω)Sznr(Ω))/Sznr(Ω)\left(S_{z}(\Omega)-S^{\rm nr}_{z}(\Omega)\right)/S^{\rm nr}_{z}(\Omega) of the TDDE and TDSE calculations for the dipole harmonic intensities.
a)             b)
Figure 12: (Color online) (a) LOPT expectation value vz(t)v_{z}(t). (b) vx(t)v_{x}(t) obtained in the LOPT calculation and using Eq. (34). Cosine pulse with E0=0.0534E_{0}=0.0534 a.u., ω=0.057\omega=0.057 a.u. has been used in the calculations.

References

  • Reiss (1998) H. R. Reiss, Opt. Express 2, 261 (1998).
  • Reiss (1990) H. R. Reiss, J. Opt. Soc. Am. B 7, 574 (1990).
  • Ludwig et al. (2014a) A. Ludwig, J. Maurer, B. W. Mayer, C. R. Phillips, L. Gallmann, and U. Keller, Phys. Rev. Lett. 113, 243001 (2014a).
  • Chelkowski et al. (2014) S. Chelkowski, A. D. Bandrauk, and P. B. Corkum, Phys. Rev. Lett. 113, 263005 (2014).
  • Chelkowski et al. (2015) S. Chelkowski, A. D. Bandrauk, and P. B. Corkum, Phys. Rev. A 92, 051401(R) (2015).
  • I.A.Ivanov et al. (2016) I.A.Ivanov, J. Dubau, and K. T. Kim, Phys. Rev. A 94, 033405 (2016).
  • Popov et al. (2006) V. S. Popov, B. M. Karnakov, V. D. Mur, and S. G. Pozdnyakov, Sov. Phys. -JETP 102, 760 (2006).
  • Klaiber and Hatsagortsyan (2014) M. Klaiber and K. Z. Hatsagortsyan, Phys. Rev. A 90, 063416 (2014).
  • Yakaboylu et al. (2015) E. Yakaboylu, M. Klaiber, and K. Z. Hatsagortsyan, Phys. Rev. A 91, 063407 (2015).
  • Keldysh (1965) L. V. Keldysh, Sov. Phys. -JETP 20, 1307 (1965).
  • Smeenk et al. (2011a) C. T. L. Smeenk, L. Arissian, B. Zhou, A. Mysyrowicz, D. M. Villeneuve, A. Staudte, and P. B. Corkum, Phys. Rev. Lett. 106, 193002 (2011a).
  • Ludwig et al. (2014b) A. Ludwig, J. Maurer, B. W. Mayer, C. R. Phillips, L. Gallmann, and U. Keller, Phys. Rev. Lett. 113, 243001 (2014b).
  • Krausz and Ivanov (2009) F. Krausz and M. Ivanov, Rev. Mod. Phys. 81, 163 (2009).
  • Chelkowski and Bandrauk (2018) S. Chelkowski and A. D. Bandrauk, Phys. Rev. A 97, 053401 (2018).
  • Chelkowski et al. (2017) S. Chelkowski, A. D. Bandrauk, and P. B. Corkum, Phys. Rev. A 95, 053402 (2017).
  • Smeenk et al. (2011b) C. T. L. Smeenk, L. Arissian, B. Zhou, A. Mysyrowicz, D. M. Villeneuve, A. Staudte, and P. B. Corkum, Phys. Rev. Lett. 106, 193002 (2011b).
  • Reiss (2013) H. R. Reiss, Phys. Rev. A 87, 033421 (2013).
  • Yakaboylu et al. (2013) E. Yakaboylu, M. Klaiber, H. Bauke, K. Z. Hatsagortsyan, and C. H. Keitel, Phys. Rev. A 88, 063421 (2013).
  • Chelkowski and Bandrauk (2017) S. Chelkowski and A. D. Bandrauk, Molecular Physics 115, 1971 (2017).
  • Jensen et al. (2020) S. V. B. Jensen, M. M. Lund, and L. B. Madsen, Phys. Rev. A 101, 043408 (2020).
  • Telnov and Chu (2020) D. A. Telnov and S.-I. Chu, Phys. Rev. A 102, 063109 (2020).
  • Selstø et al. (2009) S. Selstø, E. Lindroth, and J. Bengtsson, Phys. Rev. A 79, 043418 (2009).
  • I.A.Ivanov (2015) I.A.Ivanov, Phys. Rev. A 91, 043410 (2015).
  • Ivanov (2017) I. A. Ivanov, Phys. Rev. A 96, 013419 (2017).
  • Zhu and Wang (2016) X. Zhu and Z. Wang, Optics Communications 365, 125 (2016).
  • Potvliege et al. (2000) R. M. Potvliege, N. J. Kylstra, and C. J. Joachain, J. Phys. B 33, L743 (2000).
  • Kylstra et al. (2001) N. J. Kylstra, R. M. Potvliege, and C. J. Joachain, J. Phys. B 34, L55 (2001).
  • Chirilă et al. (2002) C. C. Chirilă, N. J. Kylstra, R. M. Potvliege, and C. J. Joachain, Phys. Rev. A 66, 063411 (2002).
  • Lewenstein et al. (1994) M. Lewenstein, P. Balcou, M. Y. Ivanov, A. L’Huillier, and P. B. Corkum, Phys. Rev. A 49, 2117 (1994).
  • Corkum (1993) P. B. Corkum, Phys. Rev. Lett. 71, 1994 (1993).
  • de Aldana and Roso (2002) J. R. V. de Aldana and L. Roso, J. Phys. B 35, 1633 (2002).
  • Bandrauk and Lu (2006) A. D. Bandrauk and H. Z. Lu, Phys. Rev. A 73, 013412 (2006).
  • Mu-Xue et al. (2020) W. Mu-Xue, C. Si-Ge, L. Hao, and P. Liang-You, Chinese Physics B 29, 013302 (2020).
  • Sarsa et al. (2003) A. Sarsa, F. J. Gálvez, and E. Buendia, J. Phys. B 36, 4393 (2003).
  • Akhiezer and Berestetskii (1965) A. Akhiezer and V. Berestetskii, Quantum Electrodynamics (John Wiley & Sons, 1965).
  • Lifshitz and Berestetskii (1982) E. M. Lifshitz and V. B. Berestetskii, Quantum Electrodynamics (Pergamon Press, 1982).
  • Nurhuda and Faisal (1999) M. Nurhuda and F. H. M. Faisal, Phys. Rev. A 60, 3125 (1999).
  • Hill and Krauthauser (1994) R. N. Hill and C. Krauthauser, Phys. Rev. Lett. 72, 2151 (1994).
  • Dirac (1964) P. A. M. Dirac, The Principles of Quantum Mechanics (McGrow-Hill, New York, 1964).
  • Ivanov and Kim (2015) I. A. Ivanov and K. T. Kim, Phys. Rev. A 92, 053418 (2015).
  • Shampine (1994) L. Shampine, Numerical Solution of Ordinary Differential Equations (Chapman &\& Hall, New York, NY, 1994).
  • Goldberg et al. (1967) A. Goldberg, H. M. Schey, and J. L. Scwartz, Am. J. Phys. 35, 177 (1967).
  • Lambropoulos and Petrosyan (2007) P. Lambropoulos and D. Petrosyan, Fundamentals of Quantum Optics and Quantum Information (Springer-Verlag, Berlin, 2007).
  • Sobelman (1972) I. I. Sobelman, Introduction to the Theory of Atomic Spectra (Pergamon Press, New York, 1972).
  • Ivanov (2014) I. A. Ivanov, Phys. Rev. A 90, 013418 (2014).
  • Avetissian et al. (2011) H. K. Avetissian, A. G. Markossian, and G. F. Mkrtchian, arXiv: Atomic Physics (2011).
  • Baggesen and Madsen (2011) J. C. Baggesen and L. B. Madsen, J. Phys. B 44, 115601 (2011).
  • Reiff et al. (2020) R. Reiff, T. Joyce, A. Jaroń-Becker, and A. Becker, J. Phys. Commun 4, 065011 (2020).
  • Mishra et al. (2012) R. Mishra, D. Kalita, and A. Gupta, Eur. Phys. J. D 66, 169 (2012).
  • Gabor (1946) D. Gabor, J. Inst. Electr. Eng. 93, 429 (1946).
  • Antoine et al. (1995) P. Antoine, B. Piraux, and A. Maquet, Phys. Rev. A 51, R1750 (1995).
  • Tang et al. (2000) Y. Y. Tang, L. H. Yang, J. Liu, and H. Ma, Wavelet Theory and Its Application to Pattern Recognition (World Scientific, 2000).
  • Wang et al. (2005) B. Wang, T. Cheng, X. Li, P. Fu, S. Chen, and J. Liu, Phys. Rev. A 72, 063412 (2005).
  • Landau and Lifshitz (1977) L. D. Landau and E. M. Lifshitz, Quantum Mechanics (Pergamon Press, New York, 1977).
  • Uehling (1935a) E. A. Uehling, Phys. Rev. 48, 55 (1935a).
  • Uehling (1935b) E. A. Uehling, Physical Review 48, 55 (1935b).
  • Flambaum and Ginges (2005) V. V. Flambaum and J. S. M. Ginges, Phys. Rev. A 72, 052115 (2005).
  • Gies and Klingmüller (2005) H. Gies and K. Klingmüller, Phys. Rev. D 72, 065001 (2005).
  • Mocken et al. (2010) G. R. Mocken, M. Ruf, C. Müller, and C. H. Keitel, Phys. Rev. A 81, 022122 (2010).
  • Kohlfürst (2020) C. Kohlfürst, Phys. Rev. D 101, 096003 (2020).
  • Nakai (2021) H. Nakai, Bull. Chem. Soc. Jpn. 94, 1664 (2021).
  • Liu (2020) W. Liu, J. Chem. Phys. 152, 180901 (2020).
  • Lysaght et al. (2012) M. A. Lysaght, L. R. Moore, L. A. A. Nikolopoulos, J. S. Parker, H. W. van der Hart, and K. T. Taylor, Journal of Physics: Conference Series 388, 012027 (2012).