1. Introduction
We present theoretical aspects of high-harmonic generation (HHG) in this chapter. Harmonic generation is a nonlinear optical process in which the frequency of laser light is converted into its integer multiples. Harmonics of very high orders are generated from atoms and molecules exposed to intense (usually near-infrared) laser fields. Surprisingly, the spectrum from this process, high-harmonic generation, consists of a plateau where the harmonic intensity is nearly constant over many orders and a sharp cutoff (see Fig. 5).
The maximal harmonic photon energy
where
Rather than by the perturbation theory found in standard textbooks of quantum mechanics, many features of HHG can be intuitively and even quantitatively explained in terms of electron rescattering trajectories which represent the semiclassical three-step model and the quantum-mechanical Lewenstein model. Remarkably, various predictions of the three-step model are supported by more elaborate direct solution of the time-dependent Schrödinger equation (TDSE). In this chapter, we describe these models of HHG (the three-step model, the Lewenstein model, and the TDSE).
Subsequently, we present the control of the intensity and emission timing of high harmonics by the addition of xuv pulses and its application for isolated attosecond pulse generation.
2. Model of high-harmonic generation
2.1. Three Step Model (TSM)
Many features of HHG can be intuitively and even quantitatively explained by the semiclassical three-step model (Fig. 1)(Krause et al., 1992; Schafer et al., 1993; Corkum, 1993).
According to this model, in the first step, an electron is lifted to the continuum at the nuclear position with no kinetic energy through tunneling ionization (
Let us consider that the laser electric field
where
we obtain,
It is convenient to introduce the phase
and we also obtain, for the kinetic energy
One obtains the time (phase) of recombination
Figure 2 shows
For a given value of
The path
If (
One can show that the Fourier transform of Equation 10 takes nonzero values only at odd multiples of
In Fig. 3 we show an example of the harmonic field made up of the 9th, 11th, 13th, 15th, and 17th harmonic components. It indeed takes the form of Equation 10. In a similar manner, high harmonics are usually emitted as a train of bursts (
2.2. Lewenstein model
The discussion of the propagation in the preceding subsection is entirely classical. Lewenstein et al. (Lewenstein et al., 1994) developed an analytical, quantum theory of HHG, called
where
The contribution of all the excited bound states can be neglected.
The effect of the atomic potential on the motion of the continuum electron can be neglected.
The depletion of the ground state can be neglected.
Within this approximation, it can be shown (Lewenstein et al., 1994) that the time-dependent dipole moment
where p and
If we approximate the ground state by that of the hydrogenic atom,
Alternatively, if we assume that the ground-state wave function has the form,
with
In the spectral domain, Equation 12 is Fourier-transformed to,
Equation 12 has a physical interpretation pertinent to the three-step model:
The evaluation of Equation 17 involves a five-dimensional integral over p,
Using the solutions
where
The physical meaning of Equations 18-20 becomes clearer if we note that p + A(t) is nothing but the kinetic momentum v(t). Equation 18, rewritten as
Let us consider again that the laser electric field is given by Equation 2 and introduce θ = ω0t and k = pω0/E0. Then Equations 18-20 read as,
where γ is called the Keldysh parameter. If we replace I
slightly higher than the three-step model (Lewenstein et al., 1994). This can be understood qualitatively by the fact that there is a finite distance between the nucleus and the tunnel exit (Fig. 1); the electron which has returned to the position of the tunnel exit is further accelerated till it reaches the nuclear position. Except for the difference in E
2.3. Gaussian model
In the Gaussian model, we assume that the ground-state wave function has a form given by Equation 15. An appealing point of this model is that the dipole transition matrix element also takes a Gaussian form (Equation 16) and that one can evaluate the integral with respect to momentum in Equation 12 analytically, without explicitly invoking the notion of quantum paths. Thus, we obtain the formula for the dipole moment x(t) as,
where B(t, t’), C(t, t’), and D(t, t’) are given by,
The Gaussian model is also useful when one wants to account for the effect of the initial spatial width of the wave function within the framework of the Lewenstein model ( Ishikawa et al., 2009 b).
2.4. Direct simulation of the time-dependent Schrödinger equation (TDSE)
The most straightforward way to investigate HHG based on the time-dependent Schrödinger equation 11 is to solve it numerically. Although such an idea might sound prohibitive at first, the TDSE simulations are indeed frequently used, with the rapid progress in computer technology. This approach provides us with exact numerical solutions, which are powerful especially when we face new phenomena for which we do not know a priori what kind of approximation is valid. We can also analyze the effects of the atomic Coulomb potential, which is not accounted for by the models in the preceding subsections. Here we briefly present the method developed by Kulander et al. (Kulander et al., 1992) for an atom initially in an s state. There are also other methods, such as the pseudo-spectral method (Tong & Chu, 1997) and those using the velocity gauge (Muller, 1999; Bauer & Koval, 2006).
Since we assume linear polarization in the z direction, the angular momentum selection rule tells us that the magnetic angular momentum remains m = 0. Then we can expand the wave function ψ(r, t) in spherical harmonics with m = 0,
At this stage, the problem of three dimensions in space physically has been reduced to two dimensions. By discretizing the radial wave function R
where the coefficients are given by,
Here, in order to account for the boundary condition at the origin properly, the Euler-Lagrange equations with a Lagrange-type functional (Kulander et al., 1992; Koonin et al., 1977),
has been discretized, instead of Equation 11 itself. c
Equations 35 and 36 can be integrated with respect to t by the alternating direction implicit (Peaceman-Rachford) scheme,
with Δt being the time step. This algorithm is accurate to the order of (Δt3), and approximately unitary. One can reduce the difference between the discretized and analytical wave function, by scaling the Coulomb potential by a few percent at the first grid point (Krause et al., 1992). We can obtain the harmonic spectrum by Fourier-transforming the dipole acceleration
V(r) is the bare Coulomb potential for a hydrogenic atom. Otherwise, we can employ a model potential (Muller & Kooiman, 1998) within the single-active electron approximation (SAE),
where Z denotes the atomic number. Parameters A, and B are chosen in such a way that they faithfully reproduce the eigenenergies of the ground and the first excited states. One can account for nonzero azimuthal quantum numbers by replacing a
In Fig. 5 we show an example of the calculated harmonic spectrum for a hydrogen atom irradiated by a Ti:Sapphire laser pulse with a wavelength of 800 nm (ω0 = 1.55eV) and a peak intensity of 1.6 × 1014W/cm2. The laser field E(t) has a form of E(t) = f (t) sinω0t, where the field envelope f (t) corresponds to a 8-cycle flat-top pulse with a half-cycle turn-on and turn-off. We can see that the spectrum has peaks at odd harmonic orders, as is experimentally observed, and the cutoff energy predicted by the cutoff law.
3. High-harmonic generation by an ultrashort laser pulse
Whereas in the previous section we considered the situation in which the laser has a constant intensity in time, virtually all the HHG experiments are performed with an ultrashort (a few to a few tens of fs) pulse. The state-of-the-art laser technology is approaching a single-cycle limit. The models in the preceding section can be applied to such situations without modification.
For completeness, the equations for the recombination time t and ionization time t’in the three-step model is obtained by replacing I
The canonical momentum is given by p = −A(t’).
In the Lewenstein model, any form of electric field E(t) can be, through Fourier transform, expanded with sine waves, defined in the complex plane. Thus the saddle-point equations 18-20 can be solved at least numerically.
In this subsection, let us consider HHG from a helium atom irradiated by au ultrashort laser pulse whose central wavelength is 800 nm, temporal profile is Gaussian with a full-width-at- half-maximum (FWHM) pulse duration T1/2 of 8 fs (1.5 cycles), and peak intensity of 5 × 1014W/cm2. There are two particular forms of electric field, as shown in Fig. 6,
and,
where the field envelope f (t) is given by,
In general, when the field takes a form of,
ϕ0 is call carrier-envelope phase (CEP). The CEP is zero and −π/2 for cos and sin pulses, respectively.
3.1. Cos pulse
Figure 7 (a) displays the real part of the recombination (t) and ionization (t’) times calculated with the saddle-point equations for the 1.5-cycle cos pulse. The recombination time from the three-step model, also shown in this figure, is close to the real part of the saddle-point solutions. By comparing this figure with the harmonic spectrum calculated with direct simulation of the TDSE (Fig. 7 (b)), we realize that the steps around 400 and 300 eV in the spectrum correspond to the cutoff of trajectory pairs C and D. Why does not a step (cutoff) for pair B appear? This is related to the field strength at time of ionization, indicated with vertical arrows in Fig. 7 (a). That for pair B (~ -5 fs) is smaller than those of pairs C (~ -2.5 fs) and D (~ 0 fs). Since the tunneling ionization rate (the first step of the three-step model) depends exponentially on intensity, the contribution from pair B is hidden by those from
pairs C and D. It is noteworthy that the trajectory pair C for the cutoff energy (~ 400 eV) is ionized not at the pulse peak but half cycle before it. Then the electron is accelerated efficiently by the subsequent pulse peak.
From the above consideration, and also remembering that harmonic emission occurs upon recombination, we can speculate the following:
The harmonics above 200 eV consists of a train of two pulses at t ≈ 1 (C) and 3.5 fs (D).
By extracting the spectral component above 300 eV, one obtains an isolated attosecond pulse at t ≈ 1 fs (C).
The emission from the short (long) trajectories are positively (negatively) chirped, i.e., the higher the harmonic order, the later (the earlier) the emission time. The chirp leads to temporal broadening of the pulse.
These are indeed confirmed by the TDSE simulation results shown in Fig. 8, for which the calculated dipole acceleration is Fourier transformed, then filtered in energy, and transformed back into the time domain to yield the temporal structure of the pulse radiated from the atom.
3.2. Sin pulse
Let us now turn to the sin pulse. The harmonic spectrum (Fig. 9 (b)) has two-step cutoffs at 370 and 390 eV associated with trajectory pairs C and B (Fig. 9 (a)), respectively. The latter (B) is much less intense, since the laser field is weaker at the ionization time for pair B than for pair C as can be seen from Fig. 9 (a), leading to smaller tunneling ionization rate. The inspection of Fig. 9 (a), similar to what we did in the previous subsection, suggests that, noting field strength at the time of ionization,
The harmonics above 100 eV consists of a train of two pulses at t ≈ 2.5 (C) and 5 fs (D). The contribution from pairs A and B are negligible due to small ionization rate.
By extracting the spectral component above 220 eV, one obtains an isolated attosecond pulse at t ≈ 2.5 fs (C), not double pulse (B and C).
As in the case of the cos pulse, the emission from the short (long) trajectories are positively (negatively) chirped. The chirp leads to temporal broadening of the pulse.
The second point indicates that broader harmonic spectrum is available for isolated attosecond pulse generation with the sin pulse than with the cos pulse, implying potentially shorter soft x-ray pulse generation if the intrinsic chirp can be compensated (Sansone et al., 2006). It also follows from the comparison between Figs. 7 and 9 that for the case of few-cycle lasers high-harmonic generation is sensitive to the CEP (Baltuška et al., 2003).
The examples discussed in this section stresses that the Lewenstein model and even the semiclassical three-step model, which is a good approximation to the former, are powerful tools to understand features in harmonic spectra and predict the temporal structure of generated pulse trains; the emission times in Figs. 8 and 10 could be predicted quantitatively well without TDSE simulations. One can calculate approximate harmonic spectra using the saddle-point analysis from Equation 21 or using the Gaussian model within the framework of SFA. On the other hand, the direct TDSE simulation is also a powerful tool to investigate quantitative details, especially the effects of excited levels and the atomic Coulomb potential (Schiessl et al., 2007; 2008; Ishikawa et al., 2009a).
4. High-harmonic generation controlled by an extreme ultraviolet pulse
In this section, we discuss how the addition of an intense extreme ultraviolet (xuv) pulse affects HHG (Ishikawa, 2003; 2004). Whereas the xuv pulse is not necessarily harmonics of the fundamental laser, let us first consider how a He+ ion behaves when subject to a fundamental laser pulse and an intense 27th or 13th harmonic pulse at the same time. Due to its high ionization potential (54.4 eV) He+ is not ionized by a single 27th and 13th harmonic photon. Here, we are interested especially in the effects of the simultaneous irradiation on harmonic photoemission and ionization. The fundamental laser pulse can hardly ionize He+ as we will see later. Although thanks to high ionization potential the harmonic spectrum from this ion would have higher cut-off energy than in the case of commonly used rare-gas atoms, the conversion efficiency is extremely low due to the small ionization probability. It is expected, however, that the addition of a Ti:Sapphire 27th or 13th harmonic facilitates ionization and photoemission, either through two-color frequency mixing or by assisting transition to the 2p or 2s levels. The direct numerical solution of the time-dependent Schrödinger equation shows in fact that the combination of fundamental laser and its 27th or 13th harmonic pulses dramatically enhance both high-order harmonic generation and ionization by many orders of magnitude.
To study the interaction of a He+ ion with a combined laser and xuv pulse, we solve the time-dependent Schrödinger equation in the length gauge,
where E(t) is the electric field of the pulse. Here we have assumed that the field is linearly polarized in the z-direction. To prevent reflection of the wave function from the grid boundary, after each time step the wave function is multiplied by a cos1/8 mask function (Krause et al., 1992) that varies from 1 to 0 over a width of 2/9 of the maximum radius at the outer radial boundary. The ionization yield is evaluated as the decrease of the norm of the wave function on the grid. The electric field E(t) is assumed to be given by,
with F
4.1. HHG enhancement
In Fig. 11 we show the harmonic photoemission spectrum from He+ for the case of simultaneous fundamental and 27th harmonic (H27) irradiation. The peak fundamental intensity I
photoemission spectrum from He+ for the case of simultaneous fundamental and H13 irradiation. Again the harmonic intensity is enhanced by more than ten orders of magnitude compared to the case of the laser pulse alone. We have varied the fundamental wavelength between 750 and 850 nm and found similar enhancement over the entire range.
4.2. Enhancement mechanism
The effects found in Figs. 11 and 12 can be qualitatively understood as follows. The H27 photon energy (41.85 eV) is close to the 1s-2p transition energy of 40.8 eV, and the H13 photon (20.15 eV) is nearly two-photon resonant with the 1s-2s transition. Moreover, the 2p and 2s levels are broadened due to laser-induced dynamic Stark effect. As a consequence, the H27 and H13 promote transition to a virtual state near these levels. Depending on the laser wavelength, resonant excitation of 2s or 2p levels, in which fundamental photons may be involved in addition to harmonic photons, also takes place. In fact, the 2s level is excited through two-color two-photon transition for the case of Fig. 11 at λ
4.3. Ionization enhancement
Let us now examine ionization probability. Table 1 summarizes the He2+ yield for each case of the fundamental pulse alone, the harmonic pulse alone, and the combined pulse. As can be expected from the discussion in the preceding subsection, the ionization probability by the combined pulse is by many orders of magnitude higher than that by the fundamental laser pulse alone. Especially dramatic enhancement is found in the case of the combined fundamental and H27 pulses: the He2+ yield is increased by orders of magnitude also with respect to the case of the H27 irradiation alone. This reflects the fact that field ionization from the 2s state by the fundamental pulse is much more efficient than two-photon ionization from the ground state by the H27 pulse. In Fig. 13, we plot the He2+ yield as a function of IH27 for a fixed fundamental intensity of 3 × 1014W/cm2. The ionization probability is linear in IH27 except for the saturation at IH27> 1013W/cm2. This is compatible with our view that an H27 photon promotes 1s-2s two-color two-photon transition, followed by field ionization.
Table 1. The He2+ yield for various combinations of a Gaussian fundamental and its 27th or 13th harmonic pulses with a duration (FWHM) of 10 fs and a peak intensity listed in the table. λ
Figure 14 shows the dependence of the He2+ yield on the peak intensity I
fundamental laser pulse is three-fold: to lift the electron in an excited (real or virtual) level to the continuum through optical-field ionization, to assist 1s-2s, 2p transitions through two-color excitation, and to induce dynamic Stark shift and broadening. The interplay of these three leads to a complicated behavior seen in Fig. 14.
4.4. Remarks
The mechanism of the HHG enhancement discussed above is that the xuv addition increases the field ionization rate by promoting transition to (real or virtual) excited states, from which ionization is much easier than from the ground state. Hence, the enhancement is effective only when the ionization rate by the fundamental pulse alone is not high enough. If HHG is already optimized by a sufficiently intense laser pulse, the xuv addition does not increase the HHG yield significantly.
Although the cases involving resonant transitions are highlighted above, the resonance with an excited state is not necessary for the enhancement. Figure 16 shows the harmonic spectra from He for λ = 1600nm with and without the XUV field (ω
4.5. Single attosecond pulse generation using the enhancement effect
The progress in the high-harmonic generation (HHG) technique has raised significant interest in the generation of attosecond pulses. As we have seen in Subsec. 2.1, the photoemission process is repeated every half-cycle of the laser optical field and produces an attosecond pulse train (Fig. 3). On the other hand, a single attosecond pulse (SAP), in particular, is critical for the study of the electron dynamics inside atoms (Krausz & Ivanov, 2009). If the driving laser pulse is sufficiently short that the effective HHG takes place only within one half-cycle, the cut-off region of the spectrum may become a continuum, corresponding to a single recollision (Figs. 8 and 10). The first SAPs (Baltuška et al., 2003) were obtained on this basis. Isolated attosecond pulses have also been realized by other methods such as ionization shutter (Sekikawa et al., 2004) and polarization gate (Corkum et al., 1994; Sansone et al., 2006), and two-color scheme (Pfeifer et al., 2006) has also been proposed.
The enhancement effect discussed in the preceding subsections provides a means to control HHG and new physical insights (Schafer et al., 2004; Ishikawa et al., 2009b). As an example, in this subsection, we present an alternative method of SAP generation using multi-cycle laser pulses, called attosecond enhancement gate for isolated pulse generation (AEGIS) (Ishikawa et al., 2007). Let us first describe AEGIS qualitatively (Figs. 17(a) and 18(a)). The underlying mechanism of the HHG enhancement by the addition of xuv pulses is that the seed pulse induces transition to real or virtual excited levels, facilitating optical-field ionization, the
first step of the three-step model. When the seed is composed of a train of attosecond pulses, this can be viewed as repeated attosecond enhancement gates. Let us consider that the fundamental pulse generating the seed harmonic pulse, referred to as the seed fundamental pulse hereafter, and the driving laser pulse which will be combined with the seed pulse have different wavelengths. For example, when the seed fundamental wavelength λ
Furthermore, the kinetic energy of the recombining electron and, thus, the emitted photon energy in the three-step model depends on the electron’s time of release, and that, in particular, an electron ionized by tunneling at ωt = ϕ
We now confirm this qualitative idea, using direct numerical solution of the time-dependent Schrödinger equation. The harmonic spectrum is calculated by Fourier transforming the dipole acceleration, then high-pass filtered as is done in experiments with multilayer mirrors, and transformed back into the time domain to yield the temporal structure of the pulse radiated from the atom.
Let us first consider the case where λ
is composed of the 11th to 19th harmonics. We use experimentally observed values (Takahashi et al., 2002) for the harmonic mixing ratio
Let us next turn to the case where λ
containing ca. 5 pulses, composed of the 15th to 23rd harmonics. The harmonic mixing ratio is
References
- 1.
Baltuška A. Udem Th. Uiberacker M. Hentschel M. Goulielmakis E. Gohle Ch. Holzwarth R. Yakovlev V. S. Scrinzi A. Hänsch T. W. Krausz F. 2003 Attosecond control of electronic processes by intense light fields. Nature421 6923 611 615 . - 2.
Bauer D. Koval P. 2006 Qprop: A Schrödinger-solver for intense laser-atom interaction. Comp. Phys. Comm.174 5 396 421 . - 3.
Corkum P. B. 1993 Plasma perspective on strong-field multiphoton ionization. Phys. Rev. Lett.71 13 1994 1997 . - 4.
Corkum P. B. Burnett N. H. Ivanov M. Y. 1994 Subfemtosecond pulses. Opt. Lett.19 22 1870 1872 . - 5.
Ishikawa K. 2003 Photoemission and ionization of He+ under simultaneous irradiation of fundamental laser and high-order harmonic pulses. 91(4): 043002. - 6.
Ishikawa K. L. 2004 Efficient photoemission and ionization of He+ by a combined fundamental laser and high-order harmonic pulse 70(1): 013412. - 7.
Ishikawa K. L. Takahashi E. J. Midorikawa K. 2007 Single-attosecond pulse generation using a seed harmonic pulse train. 75(2): 021801(R). - 8.
Ishikawa K. L. Schiessl K. Persson E. Burgdörfer J. 2009 Fine-scale oscillations in the wavelength and intensity dependence of high-order harmonic generation: Connection with channel closings. 79(3): 033411. - 9.
Ishikawa K. L. Takahashi E. J. Midorikawa K. 2009 Wavelength dependence of highorder harmonic generation with independently controlled ionization and ponderomotive energy. 80(1): 011807. - 10.
Koonin S. E. Davies K. T. R. Maruhn-Rezwani V. Feldmeier H. Krieger S. J. Negele J. W. 1977 Time-dependent Hartree-Fock calculations for 16O + 16O and 40Ca + 40Ca reactions. Phys. Rev. C15 4 1359 1374 . - 11.
Krause J. L. Schafer K. J. Kulander K. C. 1992 High-order harmonic generation from atoms and ions in the high intensity regime. Phys. Rev. Lett.68 24 3535 3538 . - 12.
Krausz F. Krausz F. 2009 Attosecond physics. Rev. Mod. Phys.81 1 163 234 . - 13.
Kulander K. C. Schafer K. J. Krause J. L. 1992 Time-dependent studies of multiphoton processes, In: , Gavrila, M. (Ed.),247 300 , Academic Press,012003901 New York. - 14.
Lewenstein M. Balcou Ph. Ivanov M. Yu L’Huillier A. Corkum P. B. 1994 Theory of high-harmonic generation by low-frequency laser fields. Phys. Rev. A49 3 2117 2132 . - 15.
Muller H. G. Kooiman F. C. 1998 Bunching and focusing of tunneling wave packets in enhancement of high-order above-threshold ionization. Phys. Rev. Lett.81 6 1207 1210 . - 16.
Muller H. G. 1999 An efficient propagation scheme for the time-dependent Schrödinger equation in the velocity gauge. Laser Phys.9 1 138 148 . - 17.
Nabekawa Y. Shimizu T. Okino T. Furusawa K. Hasegawa H. Yamanouchi K. Midorikawa 2006 Conclusive evidence of an attosecond pulse train observed with the mode-resolved autocorrelation technique. 96(8): 083901. - 18.
Pfeifer T. Gallmann L. Abel M. J. Neumark D. M. Leone S. R. 2006 Single attosecond pulse generation in the multicycle-driver regime by adding a weak second-harmonic field. Opt. Lett.31 7 975 977 . - 19.
Salières P. Carré B. Le Déroff L. Grasbon F. Paulus G. G. Walther H. Kopold R. Becker W. Milošević D. B. Sanpera A. Lewenstein M. 2001 Feynman’s path-integral approach for intense-laser-atom interactions. Science292 5518 902 905 . - 20.
Sansone G. Benedetti E. Calegari F. Vozzi C. Avaldi L. Flammini R. Poletto L. Villoresi P. Altucci C. Velotta R. Stagira S. De Silvestri S. Nisoli M. 2006 Isolated single-cycle attosecond pulses. Science314 5798 443 446 . - 21.
Schafer K. J. Yang B. Di Mauro L. F. Kulander K. C. 1993 Above threshold ionization beyond the high harmonic cutoff. Phys. Rev. Lett.70 11 1599 1602 . - 22.
Schafer K. J. Gaarde M. B. Heinrich A. Biegert J. Keller U. 2004 Strong field quantum path control using attosecond pulse trains. 92(2): 023003. - 23.
Schiessl K. Ishikawa K. L. Persson E. Burgdörfer J. 2007 Quantum path interference in the wavelength dependence of high-harmonic generation. 99(25): 253903. - 24.
Schiessl K. Ishikawa K. L. Persson E. Burgdörfer J. 2008 Wavelength dependence of high-harmonic generation from ultrashort pulses. J. Mod. Opt.55 16 2617 2630 . - 25.
Sekikawa T. Kosuge A. Kanai T. Watanabe S. 2004 Nonlinear optics in the extreme ultraviolet. Nature432 7017 605 608 . - 26.
Takahashi E. J. Nabekawa Y. Otsuka T. Obara M. Midorikawa K. 2002 Generation of highly coherent submicrojoule soft x rays by high-order harmonics. 66(2): 021802(R). - 27.
Takahashi E. J. Kanai T. Ishikawa K. L. Nabekawa Y. Midorikawa K. 2007 Dramatic enhancement of high-order harmonic generation. 99(5): 053904. - 28.
Tong X. M. Chu S. I. 1997 Theoretical study of multiple high-order harmonic generation by intense ultrashort pulsed laser fields: A new generalized pseudospectral timedependent method. 217(2-3): 119-130. - 29.
Watson J. B. Sanpera A. Chen X. Burnett K. 1996 Harmonic generation from a coherent superposition of states. 53(4): R1962 -R1965.