Nanolocalized Nonlinear Electron Photoemission under Coherent Control

Similar documents
SUPPLEMENTARY INFORMATION

Supplementary Figure 1 Schematics of an optical pulse in a nonlinear medium. A Gaussian optical pulse propagates along z-axis in a nonlinear medium

Surface plasmon amplification by stimulated emission in nanolenses

Brief, Incomplete Summary of Some Literature on Ionization

Ultrafast nanoscience with ELI ALPS

Models for Time-Dependent Phenomena. I. Laser-matter interaction: atoms II. Laser-matter interaction: molecules III. Model systems and TDDFT

Models for Time-Dependent Phenomena

Slow Propagation, Anomalous Absorption, and Total External Reflection of Surface Plasmon Polaritons in Nanolayer Systems

Quantum Nanoplasmonics and the Spaser

Limits on Tunneling Theories of Strong-Field Ionization

Nanophysics: Main trends

SUPPLEMENTARY INFORMATION

Introduction to intense laser-matter interaction

Title: Ultrafast photocurrent measurement of the escape time of electrons and holes from

Optical Spectroscopy of Advanced Materials

Empirical formula for static field ionization rates of atoms and molecules by lasers in the barrier-suppression regime

Supplementary Materials

1 Mathematical description of ultrashort laser pulses

Abnormal pulse duration dependence of the ionization probability of Na atoms in intense laser fields

Γ f Σ z Z R

Theory of selective excitation in stimulated Raman scattering

Graphene for THz technology

Model for quantum efficiency of guided mode

Center of Quantum Matter, Peking University, Beijing , China.

Multi-cycle THz pulse generation in poled lithium niobate crystals

Provided for non-commercial research and educational use only. Not for reproduction, distribution or commercial use.

Simple strategy for enhancing terahertz emission from coherent longitudinal optical phonons using undoped GaAs/n-type GaAs epitaxial layer structures

Attosecond laser systems and applications

SUPPLEMENTARY INFORMATION

Femtosecond laser pulse nanolitography using STM tip

Taking cascaded plasmonic field enhancement to the ultimate limit in silver nanoparticle dimers S. Toroghi* a, P. G. Kik a,b

Coherently controlled femtosecond energy localization on nanoscale

PROCEEDINGS OF SPIE. Imaging carrier dynamics on the surface of the N-type silicon

Summary lecture IX. The electron-light Hamilton operator reads in second quantization

Coherent Electron Scattering Captured by an Attosecond Quantum Stroboscope

Sfb 658 Colloquium 11 May Part II. Introduction to Two-Photon-Photoemission (2PPE) Spectroscopy. Martin Wolf

Modeling the Local Response of Gain Media in Time-Domain

Instantaneous Multiphoton Ionization Rate and Initial Distribution of Electron Momenta. Abstract

Lecture on: Multiphoton Physics. Carsten Müller

Part II. Interaction with Single Atoms. Multiphoton Ionization Tunneling Ionization Ionization- Induced Defocusing High Harmonic Generation in Gases

Control of dispersion effects for resonant ultrashort pulses M. A. Bouchene, J. C. Delagnes

SUPPLEMENTARY INFORMATION

Mutual transparency of coherent laser beams through a terahertz-field-driven quantum well

Supplementary Information for. Vibrational Spectroscopy at Electrolyte Electrode Interfaces with Graphene Gratings

Nonlinear Optics (WiSe 2015/16) Lecture 12: January 15, 2016

Applications of field-enhanced near-field optical microscopy

Wavelength scaling of high-order harmonic yield from an optically prepared excited state atom

Nonresonant Transparency Channels of a Two-Barrier Nanosystem in an Electromagnetic Field with an Arbitrary Strength

Laserphysik. Prof. Yong Lei & Dr. Yang Xu. Fachgebiet Angewandte Nanophysik, Institut für Physik

Understanding Nanoplasmonics. Greg Sun University of Massachusetts Boston

Ultrafast nanophotonics - optical control of coherent electron -

Looking into the ultrafast dynamics of electrons

Chapter 3 Properties of Nanostructures

Remarks on the tunneling limit of strong-field photoionization

stabilized 10-fs lasers and their application to laser-based electron acceleration

Theory of optically thin emission line spectroscopy

Laser Control of Atom-Molecule Reaction: Application to Li +CH 4 Reaction

Electron detachment from negative ions in a short laser pulse

Fast proton bunch generation in the interaction of ultraintense laser pulses with high-density plasmas

7. FREE ELECTRON THEORY.

Theory of Electronic and Optical Properties of Nanostructures

Bound states of two particles confined to parallel two-dimensional layers and interacting via dipole-dipole or dipole-charge laws

Single Photon detectors

Quantum Condensed Matter Physics Lecture 12

Relaxation of femtosecond photoexcited electrons in a polar indirect band-gap semiconductor nanoparticle

Generating intense attosecond x-ray pulses using ultraviolet-laser-induced microbunching in electron beams. Abstract

Strong Field Quantum Control. CAMOS Spring Meeting 2012 o

Landau Bogolubov Energy Spectrum of Superconductors

Lecture 11: Introduction to diffraction of light

An Investigation of Benzene Using Ultrafast Laser Spectroscopy. Ryan Barnett. The Ohio State University

The Dielectric Function of a Metal ( Jellium )

The Photoelectric Effect

Graphene photodetectors with ultra-broadband and high responsivity at room temperature

SUPPLEMENTARY INFORMATION

Photoelectron Spectroscopy using High Order Harmonic Generation

Wavelength Effects on Strong-Field Single Electron Ionization

AMO physics with LCLS

What are Lasers? Light Amplification by Stimulated Emission of Radiation LASER Light emitted at very narrow wavelength bands (monochromatic) Light

Insights on Interfacial Structure, Dynamics and. Proton Transfer from Ultrafast Vibrational Sum. Frequency Generation Spectroscopy of the

Spectroscopies for Unoccupied States = Electrons

SUPPLEMENTARY INFORMATION

Free Electron Laser. Project report: Synchrotron radiation. Sadaf Jamil Rana

Spectroscopy at nanometer scale

Particle nature of light & Quantization

Dark pulses for resonant two-photon transitions

High-energy collision processes involving intense laser fields

ABSTRACT 1. INTRODUCTION

Plasmonics: Theory and Applications

Spaser, Plasmonic Amplification, and Loss Compensation

Lecture 9: Introduction to Diffraction of Light

QUANTUM THEORY OF LIGHT EECS 638/PHYS 542/AP609 FINAL EXAMINATION

Nonlinear Optics (NLO)

α particles, β particles, and γ rays. Measurements of the energy of the nuclear

Time resolved optical spectroscopy methods for organic photovoltaics. Enrico Da Como. Department of Physics, University of Bath

2013 CAP Prize Examination

MTLE-6120: Advanced Electronic Properties of Materials. Semiconductor p-n junction diodes. Reading: Kasap ,

Luminescence Process

Supplementary Figure 1: A potential scheme to electrically gate the graphene-based metamaterial. Here density. The voltage equals, where is the DC

Charge and Energy Transfer Dynamits in Molecular Systems

Surface Plasmon Amplification by Stimulated Emission of Radiation. By: Jonathan Massey-Allard Graham Zell Justin Lau

Transcription:

Nanolocalized Nonlinear Electron Photoemission under Coherent Control NANO LETTERS 2005 Vol. 5, No. 11 2325-2329 Mark I. Stockman* and Prabath Hewageegana Department of Physics and Astronomy, Georgia State UniVersity, Atlanta, Georgia 30303 Received September 22, 2005 ABSTRACT We theoretically show that two-photon coherent control yields electron photoemission from metal nanostructures that is localized in nanosize hot spots whose positions are controllable on a nanometer scale, in agreement with recent experiments. We propose to use silver V-shapes as tailored nanoantennas for which the position of the coherently controllable photoelectron emission hot spot can be deterministically predicted. We predict that the low-frequency, high-intensity (quasi-stationary) excitation of the photoemission leads to an exponentially high contrast of the coherent control. In various fields of nanoscience and nanotechnology, one of the key processes to achieve is controlled photoinduced injection of charges into nanoscopic regions of semiconductor or metal systems or vacuum. It is very important to perform this injection as an ultrafast process, on a femtosecond or subfemtosecond scale, and have a possibility to choose the injection nanosite dynamically with a nanometer-scale resolution. Such processes can be used, for instance, in sitespecific, time-resolved electron-excitation spectroscopy of molecules and nanoobjects; its technological applications are possible in superfast nano-optoelectronics to transfer signals from optical to electronic components. There are, however, major obstacles to implementation of such processes. One of them is that light cannot be focused to nanoscopic regions directly. Using the adiabatic transformation, it is possible to transfer the optical excitation energy and coherence to the nanoscale; 1 however, such a concentration and the electron photoemission caused by it are static in space. Coherent control has been employed to dynamically change spectrum of photoelectrons emitted in a two-photon process from the Cu(111) surface. 2 We have proposed to use the coherent control to dynamically concentrate the optical excitation energy in space and time on the nanometer-femtosecond scale. 3,4 For nonlinear photoprocesses, such a concentration is possible and the time integral is coherently controllable. 3 This phenomenon has been later observed for two-photon electron emission from random nanostructured metal systems using two-pulse, interferometric coherent control in combination with electron microscopy. 5 In this Letter, we pursue two related goals: (a) We show that the two-photon (interferometric) coherent control allows one to dynamically control electron emission from randomly rough surfaces, which is localized within a few nanometers; we also propose metal nanoantennas whose photoelectron Figure 1. (a) Example of electric field of excitation pulse (pω 0 ) 3 ev, τ ) 5 fs, T ) 8 fs). (b) Geometry of nanosystem in the cross section through the xz plane of symmetry. The units in x and z axes are are nanometers; the thickness of the system in the y direction is 4 nm. emission is predictable, highly localized, and coherently controllable. (b) We predict that the photoelectron emission from metal nanostructures in the strong field (quasi-stationary) regime allows for coherent control with extremely high contrast, suitable for nanoelectronics applications. Consider a specific variant of the coherent control where a femtosecond excitation pulse consists of two identical laser subpulses with a controllable temporal delay τ between them, as shown in Figure 1a, which excite a nanosystem of the type shown in Figure 1b. Such an excitation directly models the interferometric coherent control of experiments. 5 The physics of the coherent control in this case is very transparent. The first subpulse excites an oscillating polarization in the nanostructured system. Because different sites of a nanostructure have different eigenmodes localized at them, 6 and these eigenmodes generally have different frequencies, there will be phase shifts accumulated between them during the first subpulse and subsequent delay τ. When the second subpulse arrives, it will do positive work on those oscillations that are in phase with it enhancing them and depleting energy 10.1021/nl051895m CCC: $30.25 Published on Web 10/18/2005 2005 American Chemical Society

of the out-of-phase oscillations. By a proper choice of τ, any given hot spot of the polarization can be enhanced to its maximum or, alternatively, suppressed. The contrast of this control depends on the relaxation polarization time T p with respect to the typical delay τ that is on the order of or greater than the pulse length T. In turn, the minimum difference ω in eigenfrequencies that can be resolved (discriminated between) by this coherent control is ω J 1/τ 1/T p. Time T p depends on frequency ω and ranges from several femtoseconds in the blue region to 100 fs in the near-infrared (near-ir). 7-9 For ω J 1/T p, the control contrast is high; in the opposite case, it becomes low. Consider first an excitation pulse not too intense, so the response of the nanosystem to it is linear. We also assume that the size of the system is much less than the wavelength to use the quasi-static approximation (which only limits the spatial extension but fully allows for ultrafast processes). Then the local field potential φ(r,t), induced by the external (excitation) potential φ 0 (r,t), is determined by the retarded Green s function G r (r,r ;t) of the system 2,3 φ(r,t) ) φ 0 (r,t) - φ 0 (r,t ) 2 r 2 Gr (r,r ;t-t )d 3 r dt (1) For definiteness, we consider each of the two subpulses to have a Gaussian envelope with carrier frequency ω 0, duration T, linear z polarization, and amplitude E 0 φ 0 (r,t) )-ze 0 exp[-(t/t) 2 ] cos(ω 0 t) (2) Then the local electric field in the nanosystem is given by E(r,t) )- [φ(r,t-τ) + φ(r,t)] (3) where φ(r,t) is computed from eq 1 with φ 0 (r,t) given by eq 2. We assume that W f > pω 0 > W f /2, where W f is the metal work function. Then the electron emission is predominantly two photon. We consider nanosystems that are planar in the xz plane (cf. Figure 1b) with some small thickness b 4 nm in the y direction. We assume that only the electrons emitted in the positive y direction are detected. Under these conditions, the electron current density J(x,z) in the plane (xz) of the nanostructure, integrated over time, is proportional to J(x,z,τ) - dt - dy Θ(r)E(r,t) 2 where Θ is a characteristic function equal to 1 inside the metal nanostructure and zero otherwise. Note that the timeintegrated current J(x,z,τ) depends on the delay time, τ, between the subpulses as a parameter. We have performed numerical computations for an example of the random planar nanostructure shown in Figure 1b using eqs 1-4. We choose silver as the metal because it has the smallest optical losses of any natural metal in the (4) Figure 2. Density of the integrated two-photon electron emission current in the plane of the nanosystem shown in Figure 1 for pω 0 ) 3 ev; T ) 8 fs. The material is silver with a thickness in the y direction of 4 nm. The axis scales are in nanometers; the maximum current (in relative but consistent units) and the delay are shown above the corresponding plots. The color scale of the current density is shown to the right of the panels. visible and IR spectrum. 10 The results are shown in Figure 2. As one can see, for two delay times, τ ) 4.45 fs and τ ) 5.09 fs, that differ only by 0.64 fs (approximately half of optical period), the hot spot of electron emission (shown by the red) moves within 2 nm to the neighboring position (panels a and b), and then returns back in a fraction of the optical period (panel c). For a longer delay (panel d), this hot spot moves by 10 nm to a new position. Thus, it is possible to coherently control the position of the two-photon excitation of the electron emission within the entire nanostructure with a resolution of just a few nanometers. Note that this kinetics of hots spots blinking with the delay time, shown in Figure 2, is in excellent qualitative agreement with the experimental data. 5 The full dependence of the photocurrent on τ can be found in the Supporting Information. Also similar to the experiment, 5 there are two problems with this picture. First, this is a random nanostructure, and the position of any hot spot is completely random, which may not be desirable for various applications. Second, the contrast of the coherent control is not very high: the ratio of the maximum to minimum intensity of any hot spot as a function of τ does not exceed 10:1, which may not be enough for certain applications. The cause of this moderate contrast is certainly insufficiently long polarization relaxation time in the near-uv region. However, if the photon energy is significantly decreased to increase T p, then the two-photon electron emission from clean silver is impossible. This can be remedied in by using a coverage with a thin layer of an alkali metal to decrease W f, as is made industrially in photomultipliers. Also, one can use the photoemission not into vacuum, but into a semiconductor, with the corresponding electric bias. Below, we consider the photoemission in the near-ir, for pω 0 ) 1.25 ev. 2326 Nano Lett., Vol. 5, No. 11, 2005

To increase the contrast of the coherent control, we turn to the regime of strong-field emission. Such an emission, considered in atomic physics by Keldysh, 11 occurs when the frequency is small enough and field strong enough so it can be treated as quasi-stationary: the corresponding value of Keldysh parameter γ ) (W f /U p ) 1/2 j 1, where U p ) m 2( ee mω 0 ) 2 Figure 3. (a-d) The same as in Figure 2 but for the V-shapes (geometry shown in the plots) with the angles of opening indicated; pω 0 ) 1.25 ev, and T ) 25 fs. (e) Integrated current density at the hot spots (relative units) as a function of delay τ at the openings of the two V-shapes, identified by color. In what follows, as a nanosystem we will use not a random planar system as above, but V-shapes with different angles: Figure 3 shows the geometry of two such V-shapes with the central angles of 28 and 51. By this we pursue a goal to have a completely reproducible nanoantenna photoelectron emitter with controllable properties. From Figure 3 we see that the hot spot of the photon emission for each of the V-shapes, depending on the delay τ, can localize either at the opening of a V-shape (a and d) or in its middle (b and c). This allows for the coherent control of the hot spot localization both within a given type of nanoantenna or between them. More insight into this controllability can be obtained from Figure 3e where we show the electron currents from the openings of these two V-shapes as functions of the delay. For relatively small delays, τ < 30 fs, these two dependencies oscillate practically in phase, and therefore no discrimination between them is possible. However, at longer delays, there is a phase shift accumulating for one relative to the other, which reaches approximately a half of the optical cycle for τ 80 fs, rendering the phase control very efficient. This is certainly due to the different frequencies of the eigenmodes that are localized at the openings of these two V-shapes. Note that the decay of the photocurrent with τ is certainly not monoexponential, which implies the involvement of several eigenmodes, as also was noticed in the experiment. 5 Note that the full dependence on τ can be found in Supporting Information as video files for the 28 and for the 51 V-shapes. is the electron quiver energy, m is the electron (effective) mass, and e is the elementary charge. There are two big advantages of a such a strong-field regime: first, the emission does not have a cutoff at a low frequency but, to the opposite, γ becomes small and the applicability of the quasi-stationary approximation improves; second, the electron emission for the fields, which is not too strong involves tunneling though a potential barrier and, consequently, the dependence on the parameters of the problem is exponentially strong. This leads to a potentially very efficient coherent control. Note that recently a multiphoton emission from a smooth metal surface has been used to determine the carrier-envelope phase of ultrashort laser pulses. 12 We follow the pioneering work of ref 13 and separate the electron emission process into two stages: an essentially quantum tunneling stage and the subsequent almost classical motion in the free space. For the tunneling stage, the probability per unit time of tunneling 14 integrated over the electron states at the Fermi surface is given by w(t) ) 8 p me ɛ 4 F W 2 f ν F exp( E W E y (t)) θ[-e y (t)]θ[-a y (t)] (5) E W ) 4 3ep (2mW 3 f ) 1/2 where θ[...] is the unit-step function, ɛ F is the Fermi energy, ν F is the electron state density at the Fermi surface, E y (t) is a component of the time-dependent local electric field normal to the surface of the system (our convention is that the positive value of E y (t) corresponds to the field directed away from the surface toward the free space, i.e., in the direction of the positive y axis), and the vector potential is chosen in the following gauge: A(t) ) c t E(t) dt, and c is the speed of light. Throughout, we set the electron charge to be -e. Note that we neglect the effects of the electron interaction with its electrostatic image that are not expected to radically change the predicted phase dependence that originates from the corresponding phase sensitivity of the local electric fields that are the cause of the tunneling process and themselves do not depend on it. After the tunneling, the electron velocity is given by W(t) ) e mc A(t) + v d (6) v d )- e mc A(t 0 ) Nano Lett., Vol. 5, No. 11, 2005 2327

where t 0 is the time of emergence of an electron in free space, which is assumed to occur by tunneling at zero velocity, 13 and v d is the drift velocity, which is the asymptotic (final) velocity of the electron reached after the end of the excitation pulse. Note that the θ[-e y (t)] factor in eq 5 ensures that the electron initially accelerates away from the metal surface, and θ[-a y (t)] guarantees that the drift velocity of the electron is likewise directed from the metal surface into the free space. An excitation pulse generates a bunch of electrons that drift away from the metal surface generating a current density J )-env d - w(t)dt (7) where n is the density of electrons in the emitting layer and V d is the component of the drift velocity normal to the metal surface and directed away from the metal. The electron photoemission current as given by eqs 5-7 exponentially depends on the time kinetics of the local optical electric field E(t) at every point of the metal nanosystem. To find this field in a general case for a strong excitation field E 0 would have been an extremely complicated, unrealistic task. However, there are properties of metal nanosystems that make an approximate solution possible. In the red to near-ir spectral region, the dielectric permittivity ɛ of noble metals is almost real, ɛ. Imɛ, and it is very large and negative, 10 -Reɛ. 1. Correspondingly, the normal field, which only causes the photoemission, can be strong enough outside the metal to slant the work function barrier sufficiently to produce the photoemission, but inside the metal the local fields may still be small enough to be treated by linear response theory. Also, the normal field in the metal is opposite in sign to (out of phase with) the normal field outside. Because for the emission to occur, the outside field force should be directed outward from the metal, then the inside field force is necessarily directed inward and cannot cause the photoemission. Consequently, the strong-field photoemission under such conditions can only occur from a very thin (on order of the Debye or Thomas-Fermi length, which is on a tenth of a nanometer scale for metals) layer at the surface where the local field E(t) istheexterior (outside of the metal) optical field at the surface. In accord with the above-given arguments, we use the same linear response Green function numerical approach of ref 4 as above to find the local field, E(r,t). The current density J(r) is computed by a numerical integration over the time in eq 7 at each spatial point. The results for the two V-shapes discussed above in this Letter are displayed in Figure 4. In this case, in a sharp contrast to Figure 3, the emission is concentrated at the tips of the V-shapes, as shown in panels a and b. This is certainly due to the fact that the strong field photoemission is excited by a nonscreened surface (external) optical field, while the two-photon emission is excited by the internal optical field in the metal, which are very differently distributed in space. We have found that the photoemission hot spots do not move from these tips for all delays, τ, between the two subpulses, but the intensity of the photocurrent does experience a very large modulation depending on τ(panels c and d). For relatively short delays Figure 4. For strong, quasi-static field, pω 0 ) 1.25 ev, T ) 25 fs, integrated photoemission current density for two V-shapes (with angles indicated in the panels) as a function of delay τ. (a and b) Distributions in space shown for τ yielding the maximum photocurrent. The spatial scales are in nanometers. (c and d) Timeintegrated current density at the hot spots as a function of delay (the full dependence and a blow-up for the larger delays). (τ e 30 fs), there is no appreciable phase difference between the two V-shapes (cf. the red and blue curves in panel c). However, for longer delays (panel d) the phase difference between them becomes very significant, comparable to π/2, for τ 50 fs. This difference can certainly yield a very high selectivity of excitation of one V-shape vs the other. 15 To summarize the major results of this Letter, we have developed a theory of localized optically nonlinear photoelectron emission from metal nanostructures under two-pulse (interferometric) coherent control. For two-photon emission from random metal nanostructures, we have shown that the coherent control allows one to move the hot spots of the electron emission within distances of 1-10 nm. However the contrast of the dependence on the control parameter (delay τ between the pulses) is not extremely high, and the positions of the hot spots are quasi-random. These findings are in a good qualitative agreement with experiments. 5 We have proposed silver V-shapes with different angles as engineered nanoantennas, for which the position of the hot spots and the delays required to move these hot spots along the nanostructures are deterministically predicted. To radically increase the spatial selectivity (contrast) of the coherent control, we have proposed to employ the high-field, lowfrequency (quasistationary) electron photoemission for which the dependence on the control parameter is exponentially strong. We have developed theory for such photoemission from metal nanostructures. As a result, we have shown that any of the two nano-v-shapes can be extremely selectively excited depending on the interpulse delay. Such systems can be used in different applications (nanospectroscopy, nanoelectronics, etc.) for nanolocalized, controllable photoinjection of electrons. Acknowledgment. This work was supported by grants from the Chemical Sciences, Biosciences and Geosciences Division of the Office of Basic Energy Sciences, Office of Science, U.S. Department of Energy, a grant from National 2328 Nano Lett., Vol. 5, No. 11, 2005

Science foundation, and a grant from the US-Israel Binational Science Foundation. Supporting Information Available: Video clips showing the dependence of the photocurrent on τ. This material is available free of charge via the Internet at http://pubs.acs.org. References (1) Stockman, M. I. Phys. ReV. Lett. 2004, 93, 137404-1-4. (2) Ogawa, S.; Nagano, H.; Petek, H.; Heberle, A. P. Phys. ReV. Lett. 1997, 78, 1339-1342. (3) Stockman, M. I.; Faleev, S. V.; Bergman, D. J. Phys. ReV. Lett. 2002, 88, 067402-1-4. (4) Stockman, M. I.; Bergman, D. J.; Kobayashi, T. Phys. ReV. B2004, 69, 054202-10. (5) Kubo, A.; Onda, K.; Petek, H.; Sun, Z.; Jung, Y. S.; Kim, H. K. Nano Lett. 2005, 5, 1123-1127. (6) Stockman, M. I.; Faleev, S. V.; Bergman, D. J. Phys. ReV. Lett. 2001, 87, 167401-1-4. (7) Stietz, F.; Bosbach, J.; Wenzel, T.; Vartanyan, T.; Goldmann, A.; Trägger, F. Phys. ReV. Lett. 2000, 84, 5644-5677. (8) Bosbach, J.; Hendrich, C.; Stietz, F.; Vartanyan, T.; Trager, F. Phys. ReV. Lett. 2002, 89, 257404-1-4. (9) Zentgraf, T.; Christ, A.; Kuhl, J.; Giessen, H. Phys. ReV. Lett. 2004, 93, 243901-1-4. (10) Johnson, P. B.; Christy, R. W. Phys. ReV. B1972, 6, 4370-4379. (11) Keldysh, L. V. J. Exptl. Theor. Phys. (USSR) 1965, 20, 1307-1314. Zh. Eksp. Teor. Fiz. 1964, 47, 1945-1957. (12) Apolonski, A.; Dombi, P.; Paulus, G. G.; Kakehata, M.; Holzwarth, R.; Udem, T.; Lemell, C.; Torizuka, K.; Burgdoerfer, J.; Hansch, T. W.; Krausz, F. Phys. ReV. Lett. 2004, 92, 073902-1-4. (13) Corkum, P. B.; Burnett, N. H.; Brunel, F. Phys. ReV. Lett. 1989, 62, 1259-1262. (14) Landau, L. D.; Lifshitz, E. M. Quantum Mechanics: Non-RelatiVistic Theory; Pergamon Press: Oxford and New York, 1965. (15) The relatively smaller value of current for τ J 30 fs is not of principal importance since it can always be compensated by a moderate (not more than by a factor of 2) increase of the excitation intensity NL051895M Nano Lett., Vol. 5, No. 11, 2005 2329