You are on page 1of 6

PHYSICAL REVIEW A 78, 023413 2008

Attosecond-pulse-controlled high-order harmonic generation in ultrashort laser elds


Laboratory of Optical Physics, Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100080, China 2 Institute of Applied Physics and Computational Mathematics, Beijing 100088, China 3 Department of Physics, University of New Brunswick, P.O. Box 4400, Fredericton, New Brunswick, Canada E3B 5A3 4 Center for Theoretical Atomic and Molecular Physics, Academy of Fundamental and Interdisciplinary Sciences, Harbin Institute of Technology, Harbin 150080, China Received 5 March 2008; revised manuscript received 6 July 2008; published 21 August 2008 By solving the three-dimensional time-dependent Schrdinger equation, we study how to control the highorder harmonic generation HHG in a hydrogenlike neon, which is exposed to an extreme-ultraviolet xuv attosecond laser pulse followed by an infrared ir femtosecond pulse. The xuv pulse prepares the atom as a superposition of the ground state, the excited states, and the continuum. These three types of states provide different contributions to the high-order harmonics generated by the following ir pulse. In particular, the contributions from the excited states and the continuum extend the cutoff frequency of the harmonic spectrum. Moreover, the effect of the carrier-envelope phase of the ir pulse on the HHG has been studied. The structure of the laser-dressed atom can be imaged using this two-color scheme. DOI: 10.1103/PhysRevA.78.023413 PACS number s : 32.80.Rm, 42.50.Hz, 32.80.Fb
1

Bingbing Wang,1 J. Chen,2 Jie Liu,2 Zong-Chao Yan,3,4 and Panming Fu1

I. INTRODUCTION

Attophysics has become a hot area of research since an attosecond pulse or pulse train was predicted theoretically 17 and realized experimentally 811 . Attosecond pulse is a very effective and powerful tool to trace and control dynamic processes of atoms and molecules 12,13 , which can be realized in a strong laser eld through the so-called recollision mechanism 14 . According to this mechanism, after an electron in an atom has been ionized by a laser, the electron is still subjected to the laser eld and thus it oscillates. Then the electron may be driven back to the vicinity of its parent ion by the laser and recollide with its parent ion. The recollision process is the key to understanding many strong eld phenomena, such as high-order above threshold ionization ATI , high-order harmonic generation HHG , and nonsequential double ionization. It has been shown 1517 that control of a strong eld process can happen at the moment of ionization or recollision. Schafer and Gaarde 16 demonstrated the control of the HHG quantum path by an attosecond pulse train. They found that the low harmonic peaks can be controlled by adding an attosecond pulse train at different phases of the laser pulse. Faria et al. 17 performed an analysis on how HHG and ATI can be controlled by a time-delayed attosecond pulse train. With the development of the optical technology, much shorter infrared ir femtosecond pulse has now been realized 18 . Hence a new pathway to control the recollision process can be realized by controlling the carrier-envelope phase CEP of the laser pulse. It was reported that, when the pulse duration full width at half maximum FWHM is longer than three optical cycles, the harmonic spectrum is not affected by the CEP, and the cutoff frequency of the spectrum is a constant, which is 3.17U p + I p predicted by the classical three-step model, where U p is the ponderomotive energy and I p is the ionization threshold of the electron 19 . On the other hand, the spectrum varies with the CEP if the pulse
1050-2947/2008/78 2 /023413 6

duration is reduced to one optical cycle. In this case the cutoff frequency, which is not a constant, can still be well predicted by the maximum kinetic energy of the returning electron 20 . However, as the pulse duration becomes shorter than one optical cycle, which is the situation we shall investigate in this paper, the spectrum changes dramatically with the CEP, and the cutoff frequency can no longer be predicted by the maximum kinetic energy for some CEPs. This is because the maximum kinetic energy now does not correspond to an effective ionization process. We then study HHG in a hydrogenic neon by applying an extremeultraviolet xuv attosecond pulse before the ultrashort ir femtosecond pulse. The xuv pulse produces an atomic state which is a superposition of the atomic ground state, the excited states, and the continuum. For the rst time, we nd that the harmonic generation in the ir laser pulse can be utilized as a possible tool for imaging instantaneous structure of an atom. This paper is organized as follows. In Sec. II, we describe the numerical method for solving the time-dependent Schrdinger equation. The CEP dependence of the HHG in an ir femtosecond laser pulse is discussed in Sec. III. In Sec. IV, we control HHG by applying an xuv attosecond pulse before the ir femtosecond pulse. The extension of HHG cutoff is analyzed by using the classical three-step model, which shows that the electron with a nonzero initial position or initial velocity plays a crucial role in the HHG process. Section V is our conclusion.

II. NUMERICAL METHOD

Our theory is based on solving the three-dimensional time-dependent Schrdinger equation for a hydrogenlike Ne in a linearly polarized laser pulse atomic units are used throughout, unless otherwise stated as follows:
2008 The American Physical Society

023413-1

WANG et al.

PHYSICAL REVIEW A 78, 023413 2008

FIG. 1. Harmonic spectra for laser frequency = 0.057, peak intensity 1 1015 W / cm2, pulse duration T = 4 fs, and carrierenvelope phase = 0 a , 0.2 b , 0.5 c , and 0.8 d .

FIG. 2. Color online Kinetic energy solid curve of the returning electron and the population dashed curve of the ground state as a function of the ionization time for = 0 a , 0.2 b , 0.5 c , and 0.8 d . e Maximum value of part I solid square and part II open circle as a function of CEP.

r,t 1 = 2 t

Zeff +rE t r

r,t ,

where the effective nuclear charge is chosen to be Zeff = 1.2592 in order to reproduce the same ionization threshold as the ground state of Ne, and E t is the electric eld of the laser pulse. The above equation is integrated numerically in the spherical coordinate system by using the nite-difference technique 21 . The length of the integration grid is 400, the spatial step is 0.2, and the time increment is 0.01. To avoid reections of the wave packet from the boundaries, after each increment the wave function is multiplied by a factor of the cos1/8 mask function that varies from 1 to 0 over the range of r from 350 to 400 22 . The harmonic spectrum is obtained from the Fourier transform of the dipole operator in r,t . the velocity form D t = r , t

III. HHG BY AN ir FEMTOSECOND PULSE WITH DIFFERENT CARRIER-ENVELOPE PHASES

We rst consider HHG by an ir laser pulse which has duration FWHM shorter than one optical cycle. The electric eld of the laser pulse is E t = (0 , 0 , F t sin t + ), where F t = E0 sin2 t / T is the pulse envelope and is the CEP of the laser pulse. The laser frequency is = 0.057, the peak intensity is 1 1015 W / cm2, and the pulse duration is T = 4 fs, which corresponds to 1.5 optical cycles FWHM is 0.75 cycle . Figure 1 presents the harmonic spectra for the case of = 0 a , 0.2 b , 0.5 c , and 0.8 d . It shows that the harmonic strength decreases whereas the cutoff frequency increases as CEP increases. Specically, the cutoff frequency is 19 , 33 , 71 , and 117 for = 0, 0.2 , 0.5 , and 0.8 , respectively. The harmonic spectrum can be understood through examining the classical dynamics of the ionized electron. Figure 2 presents the kinetic energy Ekin solid curve of the returning electron, as well as the population dashed curve of the ground state, as a function of the ionization time t0 for = 0 a , 0.2 b , 0.5 c , and 0.8 d . For convenience, we

divide each energy curve into two parts, i.e., part I labeled as P I in the gures and part II labeled P II , as illustrated in Figs. 2 a 2 d , and the maximum values of part I and part II as a function of CEP are shown in Fig. 2 e . According to the semiclassical model 14 , the cutoff frequency of the harmonic spectrum can, in general, be predicted by the maximum kinetic energy that a returning electron can carry. Thus, for 0.3 , the cutoff frequency of a harmonic spectrum should correspond to the peak value in part I, while the cut0.3 corresponds to the peak value in off frequency for part II. However, it is found that in our case the cutoff fre0.3 corresponds to the peak value in part II quency for also, which is smaller than the value in part I. This can be understood easily by considering the ionization of the electron in the eld. As shown in Fig. 2, for the case of 0.3 the ionization of the initial state is very weak as t0 0.5 l, where l is the optical cycle of the ir laser. Hence the trajectories with t0 0.5 l will not contribute to the harmonic generation effectively. By contrast, the HHG originates mainly from the trajectories which start at t0 0.5 l , l because of their higher ionization rate. In other words, only the curve in Fig. 2 e , which corresponds to part II open circle can predict the cutoff frequencies of the harmonic spectra. Furthermore, since the peak of the part II curve in Figs. 2 a 2 d shifts towards a smaller value of t0 as the CEP increases, the corresponding ionization rate decreases with the CEP, leading to the decrease of the strength of the harmonic.
IV. HHG BY AN xuv ATTOSECOND PULSE FOLLOWED BY AN ir FEMTOSECOND PULSE

From the above results it becomes evident that one can control the harmonic generation by adding an xuv attosecond pulse before the ir femtosecond pulse so that the ionization of the atom changes. The combined electric eld is E t = (0 , 0 , F t d sin t d + + Fas t sin t ). Here, d is the time delay between the xuv and ir pulses, is the fre-

023413-2

ATTOSECOND-PULSE-CONTROLLED HIGH-ORDER

PHYSICAL REVIEW A 78, 023413 2008 TABLE I. Population of an excited state with angular momentum l = 1 after the xuv attosecond pulse of = 21.6 eV. n Population 9.3 2 103 4.6 3 103 2.4 4 103 1.3 5 103 8.5 6 104

FIG. 3. Color online Harmonic spectrum in the combined eld with = 21.6 eV solid curve . The CEP of the ir pulse is = 0 a , 0.2 b , 0.3 c , and 0.4 d . The dashed curve is the HHG spectrum with no xuv pulse.

= 29.4 eV Fig. 4 . Specically, the spectrum presents a plateau with cutoff at 77 , 65 , and 49 followed by a hump at 137 , 130 , and 115 when the CEP is 0.2 , 0.3 , and 0.4 , respectively. As will be discussed later, the hump in the harmonic spectrum presents the image of the wave packet prole at the recollision time, where the wave packet is generated by the xuv pulse through direct ionization. The harmonic spectra in the combined eld can be understood by analyzing effects of the xuv attosecond pulse on the HHG process. After the interaction of the atom with the xuv attosecond pulse, the atomic wave function can be written as: t =a
g

quency of the xuv photon, and Fas t = Eas sin2 t / is the envelope of the xuv pulse with the xuv pulse duration. The ir pulse is the same as that used in Sec. III, and the time delay is set to be d = , so that the xuv pulse is just ahead of the ir pulse and there is no overlap between these two pulses. The peak intensity of the xuv pulse is 1 1014 W / cm2, and the pulse duration is = 960 attoseconds. Figures 3 and 4 present the harmonic spectra in the combined eld with = 21.6 eV and 29.4 eV, respectively, where the photon energy of xuv is larger than the ionization potential I p = 21.56 eV. The CEP of the ir pulse is = 0 a , 0.2 b , 0.3 c , and 0.4 d . For comparison, HHG spectra with no xuv pulse are also presented in both gures dashed curves . When = 21.6 eV Fig. 3 , the spectrum displays a stretched plateau with cutoff at 137 , 139 , 199 , and 99 for being 0, 0.2 , 0.3 , and 0.4 , respectively. On the other hand, the spectrum shows a complex structure when

+
i

bi

d3 pc p

p,

FIG. 4. Color online Harmonic spectrum in the combined eld with = 29.4 eV solid curve . The CEP of the ir pulse is = 0 a , 0.2 b , 0.3 c , and 0.4 d . The dashed curve is the HHG spectrum with no xuv pulse. The dash-dotted curve in d corresponds to the spectrum of = 35 eV.

which includes three parts: the ground state g, the excited states i, and the continuum p. Thus, harmonics generated by the following ir pulse originate from three sources: the electron in the ground state denoted as source S1 hereafter ; the electron in the excited states source S2 ; and the electron in the continuum due to the direct ionization by the xuv pulse source S3 . As shown in Figs. 3 and 4, the low parts of the spectra for all CEP cases agree quite well with that without the XUV pulse, indicating that these harmonics are due to the contribution from S1. Now, we consider the remaining high-order parts of the spectra, which are caused by sources S2 and S3. To understand these results, in the following we will analyze the HHG processes from sources S2 and S3 separately from a semiclassical viewpoint. Let us rst consider the contribution from source S2. As one can expect, the xuv attosecond pulse will excite atoms to many excited states, which have low ionization thresholds and large orbital radii. When an electron is ionized from one of these states by the ir eld, the initial distance between the electron and its parent ion is the radius of the excited state, which scales as rn n2, where n is the principal quantum number of the excited state. Since the initial distance is not zero, the electron may bring larger kinetic energy when it returns, and therefore contribute to HHG with higher-order harmonics 23 . The efciency of the HHG from the excited state depends on the population of the state. Table I presents the population of the excited states with angular momentum l = 1 for the case of = 21.6 eV. It is found that the population of the excited state decreases with n. Specically, the population of the state n = 2 is more than one order of magnitude larger than that of n = 6. In our calculations, we nd that the cutoff frequency of the harmonic spectrum can be described by Emax + I p, where Emax is the maximum kinetic energy of the returning electron which is ionized from the state n = 5. This indicates that the excited state with n 6 has too small a population to contribute to the HHG effectively.

023413-3

WANG et al.

PHYSICAL REVIEW A 78, 023413 2008

FIG. 5. Color online a Time prole of the harmonics generated by the combined eld with = 21.6 eV and = 0. b Time prole of the harmonics generated by the ir eld with = 0. c The corresponding classical kinetic energy as a function of the emission time. d Time prole of the harmonics generated by the combined eld with = 21.6 eV and = 0.4 . e Time prole of the harmonics generated by the ir eld with = 0.4 . f The corresponding classical kinetic energy as a function of the emission time. The quantum results are plotted in a logarithmic scale.

We then consider the HHG from source S3. After absorbing an xuv photon, which has energy larger than the ionization potential, an electron may be ionized directly from the atom with nonzero velocity 24 . The electron is then driven back by the ir pulse and recombines with its parent ion through emitting a harmonic photon. Here, the nonzero initial velocity plays an important role for characterizing the harmonic generation. Specically, the initial velocity of the electron is determined by the central frequency and the FWHM duration / 2 of the xuv pulse 15 . The frequency bandwidth is given by = 2 / , thus the velocity is between
vmin = 2

Ip

and
vmax = 2

Ip

I p . Therefore, an elecwith central velocity vmid = 2 tron wave packet which has an initial distribution in the momentum space is produced by the xuv pulse. For some CEP,

the prole of the wave packet at the recollision time can be imaged by the harmonic spectrum when the xuv photon energy is larger than the atomic ionization potential. The contributions of sources S2 and S3 to HHG can be claried further by performing the wavelet time-frequency analysis of the spectral and temporal structures of HHG 25 . We rst consider the case that the xuv photon energy is = 21.6 eV. Figures 5 a and 5 d present the time prole of the harmonics generated by the combined eld with = 0 and 0.4 , respectively. For comparison, Figs. 5 b and 5 e are the corresponding time prole of the harmonics with no xuv, while Figs. 5 c and 5 f are the corresponding classical kinetic energy of the returning electron which is ionized from the different initial states as a function of the returning time. Considering = 0, Fig. 5 b shows that, without the xuv pulse, the harmonic photon at the cutoff, which is near the 19th order dashed line in Fig. 1 a , emits at the time of approximately 1.6 optical cycles. With the application of the xuv pulse, the cutoff extends to about the 137th order and harmonic photons in the cutoff region emit at approximately 1.3 optical cycles. The time prole of the harmonics gener-

023413-4

ATTOSECOND-PULSE-CONTROLLED HIGH-ORDER

PHYSICAL REVIEW A 78, 023413 2008

FIG. 6. Color online a Time prole of the harmonics generated by the combined eld with = 29.4 eV and = 0. b The corresponding classical kinetic energy as a function of the emission time. c Time prole of the harmonics generated by the combined eld with = 29.4 eV and = 0.4 . d The corresponding classical kinetic energy as a function of the emission time. The quantum results are plotted in a logarithmic scale.

ated by the combined eld can be understood by calculating the classical kinetic energy of the returning electron which is ionized from different initial states Fig. 5 c . For source S2, four orbital radii with n = 2 5 are chosen as the initial distance of the ionized electron, while for source S3, the initial velocity is from vmin = 0 to vmax = 0.323 with vmid = 0.05. As shown in Figs. 5 a and 5 c , the classical model can predict both the cutoff and its emission time quite well. Furthermore, since both sources S2 and S3 contribute to the harmonics at the cutoff region, interference between them leads to the oscillatory structure in the spectrum see Fig. 3 a . Now, we consider the case of = 0.4 . Similar to the case of = 0, the application of the xuv pulse extends the cutoff from the 59th order, which emits at a time of approximately 1.55 optical cycles Fig. 5 e , to the 99th harmonic order with emission time at approximately 1.25 optical cycles Fig. 5 d . Furthermore, from the classical calculations Fig. 5 f , it is found that both sources S2 and S3 contribute to the harmonics when the harmonic is below the 61st order, thus, interference fringes appear in this region solid curve in Fig. 3 d . By contrast, the harmonic spectrum exhibits a smooth curve above the 61st order. Now, we consider the case when = 29.4 eV. Figures 6 a and 6 c present the time prole of the harmonics gen-

erated by the combined eld with = 0 and 0.4 , respectively, while Figs. 6 b and 6 d are the corresponding classical kinetic energy as a function of the returning time. Here again, the cutoff of the spectrum extends to higher harmonic orders by the combined eld. On the other hand, when = 0 the cutoff frequency for = 29.4 eV, which is 80 Figs. 6 a and 4 a , is much lower than that for = 21.6 eV, which is 137 Figs. 5 a and 3 a . This can be understood as follows. From the classical calculation Fig. 6 b , source S3 contributes only to the lower harmonics when = 29.4 eV, thus, the harmonics near the cutoff are mainly from source S2. HHG from source S2 depends on the populations of the excited states, which are excited by the lower part of the frequency spectrum of the xuv pulse. As increases the spectral density of this part decreases, leading to the decrease of the excited state population. Now, we consider the case with = 0.4 . It is found from the classical kinetic energy of the returning electron Fig. 6 d that the contribution from source S3 is now separated from the contributions from sources S1 and S2. Thus, the hump after the plateau Fig. 4 d represents the prole of the wave packet in the momentum space at the recollision time, where the wave packet is generated by the xuv pulse through direct ionization. To conrm this, we present the harmonic spec-

023413-5

WANG et al.

PHYSICAL REVIEW A 78, 023413 2008

trum when = 35 eV and = 0.4 dashed curve in Fig. 3 d . It is found that because the initial momentum distribution of the wave packet for = 35 eV is narrower than that of = 29.4 eV, a narrower hump appears between 90 and 120 .
V. CONCLUSIONS

In this paper, we have demonstrated through a hydrogenlike neon that, by using an xuv attosecond pulse followed by an ir femtosecond pulse, one can control the HHG process by changing the CEP of the ir pulse. Specically, the xuv pulse prepares the atom as a superposition of the ground state, the excited states, and the continuum. Therefore, according to the different initial states, the contribution to HHG can be divided into three parts. The electron which is ionized from the excited state has a nonzero initial distance from its parent ion, while the electron which is ionized directly by the xuv pulse has a nonzero velocity. Both of these electrons can

obtain higher kinetic energy from the following ir pulse, and thus, emit harmonic photons with higher energy. We also nd that the prole of the wave packet at the recollision time, where the wave packet is generated by the xuv pulse through direct ionization, can be imaged with this two-color scheme when the xuv photon energy is larger than the atomic ionization potential. This result may shed light on the study of the dynamic process of an atom in an ultrashort strong laser eld.
ACKNOWLEDGMENTS

This research was supported by the National Natural Science Foundation of China under Grant Nos. 60478031, 60778009, 10574019, and 1072552, and by the 973 Research Project under Grant Nos. 2006CB806003, 2006CB921400, and 2007CB814800. Z.C.Y. was supported by NSERC of Canada. B.W. thanks Jiangbin Gong, A. Scrinzi, and P. Villoresi for helpful discussions.

1 G. Farkas and C. Toth, Phys. Lett. A 168, 447 1992 . 2 Ph. Antoine, A. LHuillier, and M. Lewenstein, Phys. Rev. Lett. 77, 1234 1996 . 3 I. P. Christov, M. M. Murnane, and H. C. Kapteyn, Phys. Rev. Lett. 78, 1251 1997 . 4 D. B. Milosevic and W. Becker, Phys. Rev. A 62, 011403 R 2000 . 5 N. Milosevic, A. Scrinzi, and T. Brabec, Phys. Rev. Lett. 88, 093905 2002 ; M. Kitzler, N. Milosevic, A. Scrinzi, F. Krausz, and T. Brabec, ibid. 88, 173904 2002 . 6 K. T. Kim, C. M. Kim, M. G. Baik, G. Umesh, and C. H. Nam, Phys. Rev. A 69, 051805 R 2004 . 7 A. A. Zholents and W. M. Fawley, Phys. Rev. Lett. 92, 224801 2004 . 8 M. Haentschel, R. Kienberger, Ch. Spielmann, G. A. Reider, N. Milosevic, T. Brabec, P. B. Corkum, U. Heinzmann, M. Drescher, and F. Krausz, Nature London 414, 511 2001 . 9 P. M. Paul, E. S. Toma, P. Breger, G. Mullot, F. Aug, Ph. Balcou, H. G. Muller, and P. Agostini, Science 292, 1689 2001 . 10 R. Lopez-Martens, K. Varju, P. Johnsson, J. Mauritsson, Y. Mairesse, P. Salieres, M. B. Gaarde, K. J. Schafer, A. Persson, S. Svanberg, C. G. Wahlstrom, and A. LHuillier, Phys. Rev. Lett. 94, 033001 2005 ; J. Mauritsson, P. Johnsson, E. Gustafsson, A. LHuillier, K. J. Schafer, and M. B. Gaarde, ibid. 97, 013001 2006 . 11 Y. Nabekawa, T. Shimizu, T. Okino, K. Furusawa, H. Hasegawa, K. Yamanouchi, and K. Midorikawa, Phys. Rev. Lett. 96, 083901 2006 . 12 H. Niikura, D. M. Villeneuve, and P. B. Corkum, Phys. Rev. Lett. 94, 083003 2005 ; M. Wickenhauser, J. Burgdorfer, F. Krausz, and M. Drescher, ibid. 94, 023002 2005 ; A. Nazarkin, ibid. 97, 163904 2006 ; G. L. Yudin, A. D. Bandrauk, and P. B. Corkum, ibid. 96, 063002 2006 ; S. X. Hu and L. A. Collins, ibid. 96, 073004 2006 ; M. Forre, J. P. Hansen, L. Kocbach, S. Selsto, and L. B. Madsen, ibid. 97, 043601 2006 .

13 F. Lindner, M. G. Schatzel, H. Walther, A. Baltuska, E. Goulielmakis, F. Krausz, D. B. Milosevic, D. Bauer, W. Becker, and G. G. Paulus, Phys. Rev. Lett. 95, 040401 2005 . 14 P. B. Corkum, Phys. Rev. Lett. 71, 1994 1993 ; K. C. Kulander, K. J. Schafer, and J. L. Krause, in Super-Intense LaserAtom Physics, edited by K. Rzazewski Plenum, New York, 1993 , Vol. 316, p. 95. 15 A. D. Bandrauk, S. Chelkowski, and N. H. Shon, Phys. Rev. Lett. 89, 283903 2002 ; Appl. Phys. B: Lasers Opt. 77, 337 2003 ; Phys. Rev. A 68, 041802 R 2003 . 16 K. J. Schafer, M. B. Gaarde, A. Heinrich, J. Biegert, and U. Keller, Phys. Rev. Lett. 92, 023003 2004 . 17 C. Figueira de Morisson Faria, P. Salieres, P. Villain, and M. Lewenstein, Phys. Rev. A 74, 053416 2006 . 18 B. I. Greene, J. F. Federici, D. R. Dykaar, R. R. Jones, and P. H. Bucksbaum, Appl. Phys. Lett. 59, 893 1991 ; D. You, R. R. Jones, D. R. Dykaar, and P. H. Bucksbaum, Opt. Lett. 18, 290 1993 ; E. Goulielmakis, V. S. Yakovlev, A. L. Cavalieri, M. Uiberacker, V. Pervak, A. Apolonski, R. Kienberger, U. Kleineberg, and F. Krausz, Science 317, 769 2007 . 19 G. Sansone, C. Vozzi, S. Stagira, M. Pascolini, L. Poletto, P. Villoresi, G. Tondello, S. De Silvestri, and M. Nisoli, Phys. Rev. Lett. 92, 113904 2004 . 20 A. deBohan, P. Antoine, D. B. Milosevic, and B. Piraux, Phys. Rev. Lett. 81, 1837 1998 . 21 K. C. Kulander, K. J. Schafer, and J. L. Krause, Atoms in Intense Laser Fields, edited by M. Gavrila Academic, New York, 1992 . 22 X. M. Tong and S.-I. Chu, Chem. Phys. 217, 119 1997 ; B. Wang, T. Cheng, X. Li, P. Fu, S. Chen, and J. Liu, Phys. Rev. A 72, 063412 2005 . 23 B. Wang, Z.-C. Yan, and P. Fu unpublished . 24 P. Johnsson, J. Mauritsson, T. Remetter, A. LHuillier, and K. J. Schafer, Phys. Rev. Lett. 99, 233001 2007 . 25 P. Antoine, B. Piraux, and A. Maquet, Phys. Rev. A 51, R1750 1995 ; C. Figueira de Morisson Faria, M. Dorr, and W. Sandner, ibid. 55, 3961 1997 .

023413-6

You might also like