06.01.2013 Views

Attosecond Control and Measurement: Lightwave Electronics

Attosecond Control and Measurement: Lightwave Electronics

Attosecond Control and Measurement: Lightwave Electronics

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

1 S C I E N T I F I C R E P O R T S<br />

1 . 3 A T T O S E C O N D A N D H I G H - F I E L D<br />

P H Y S I C S D I V I S I O N<br />

Ferenc Krausz<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 131


GROUP MEMBERS<br />

ASSOCIATED SCIENTIFIC STAFF<br />

- Prof. U. Kleineberg (LMU 1 ) - S. Becker<br />

- Dr. Jingquan Lin - A. Henig<br />

- M. Hofstetter - D. Kiefer<br />

- Prof. D. Habs - K. Golab<br />

(Max-Planck Fellow, LMU) - R. Weingartner<br />

- Dr. U. Schramm - M. Fuchs<br />

(until 9/2006)<br />

- Dr. J. Schreiber (Sabbatical from<br />

2/2008 till 1/2009)<br />

POSTDOCTORAl STAFF & GUEST RESEARChERS<br />

- MPQ:<br />

- M. Centurion - X. Gu - M. Schultze<br />

- Y. Deng - S. Karsch - M. Siebold<br />

- R. Ernstorfer - Zs. Major - G. Tsakiris<br />

- E. Fill - Y. Mikhailova - S. Trushin<br />

- M. Geissler - A. Marcinkevicius - L. Veisz<br />

(until 8/2008) (until 12/2006) - T. J. Wang<br />

- E. Goulielmakis - J. Meyer-ter-Vehn - T. Wittmann<br />

- F. Grüner - Y. Nomura - H.-C. Wu<br />

- LMU:<br />

- A. Apolonskiy - M. Uiberacker<br />

- J. Fülöp (until 1/2007) (until 4/2008)<br />

- S. Naumov (until 12/2006) - V. Yakovlev<br />

- V. Pervak - A. Sugita<br />

- J. Rauschenberger<br />

Ph.D CANDIDATES<br />

- MPQ:<br />

- I. Ahmad - S. Klingebiel - P. Reckenthäler<br />

- A. Buck - K. Kosma - S. Rykovanov<br />

- A. Fern<strong>and</strong>ez* - J. Li* - M. Schultze<br />

- J. Gagnon - C. Lin* - F. Tavella*<br />

- T. Ganz - Y. Nomura - A. Tronnier*<br />

- D. Herrmann - J. Osterhoff - A. Verhoef*<br />

- B. Horvath - A. Popp - C. W<strong>and</strong>t<br />

- N. Ishii* - I. Pupeza - M. Wen<br />

- LMU:<br />

- R. Hörlein - K. Schmid - C. Teisset<br />

- T. Metzger<br />

DIPlOMA CANDIDATES<br />

- MPQ:<br />

- S. Benavides* - I. Grguras* - W. Schweinberger<br />

- R. Graf* - B. Reiter* - R. Tautz<br />

- D. Grupe* - J. Schmidt*<br />

- LMU:<br />

- M. Kraus - B. Marx* - A. Wytrykus*<br />

- A. S. Kruber*<br />

1 LMU= Ludwig Maximilians Universität<br />

1 S C I E N T I F I C R E P O R T S<br />

TEChNICAl<br />

- MPQ:<br />

- A. Böswald - H. Haas - A. Raufer<br />

- G. Br<strong>and</strong>l - A. Horn - H.-P. Schönauer<br />

- M. Fischer<br />

- LMU:<br />

- S. Herbst<br />

MANAGEMENT, OUTREACh<br />

- MPQ: - LMU: - MAP 2 :<br />

- Th. Naeser - B. Ferus - K. Adler<br />

- B. Schütz - Ch. Hacken- - C. Reichelt<br />

- M. Wild berger<br />

KEY COllABORATORS<br />

- D. Attwood - K. Kompa - H. Schröder<br />

(Berkeley) (MPQ) (MPQ)<br />

- J. Barth - N. Kroo - A. Scrinzi<br />

(TUM) (Budapest) (TU Wien)<br />

- J. Burgdörfer - W. Leemans - D. Sutter<br />

(Vienna) (Berkeley) (Trumpf GmbH)<br />

- D. Charalambidis - R. Levine - M. Stockman<br />

(Crete) (Jerusalem) (Atlanta)<br />

- P. Feulner - M. Lezius - L. Strüder<br />

(TUM) (MPQ) (HLL)<br />

- I. Földes - D. Menzel - T. Taira<br />

(Budapest) (TUM) (IMS, Japan)<br />

- M. Geissler - R. Moshammer - V. Tarnetsky<br />

(Belfast) (Heidelberg) (Novosibirsk)<br />

- T. W. Hänsch - C. Nicolaides - A. V. Tikhonravov<br />

(MPQ, LMU) (Athens) & M. K. Trubetskov<br />

- D. Hoffmann - M. Nisoli (Moscow)<br />

(Aachen) (Milan) - A. Tünnermann<br />

- M. Hegelich - L. O´Silva (Jena)<br />

(Los Alamos) (Lissabon) - T. Udem<br />

- J. Hein - G. G. Paulus (MPQ)<br />

(Jena) (Jena) - J. Ullrich<br />

- S. Hooker - G. Pretzler (Heidelberg))<br />

(Oxford) (Düsseldorf) - M. Vrakking<br />

- M. Yu. Ivanov - F. Remacle (Amsterdam<br />

(Ottawa) (Liège) - M. Zepf<br />

(Belfast)<br />

JUNIOR RESEARCh GROUPS<br />

- <strong>Attosecond</strong> Dynamics - Dr. R. Kienberger:<br />

- Dr. A. Cavalieri - M. Fieß - W. Helml<br />

- Dr. G. Marcus - U. Graf - E. Magerl<br />

- <strong>Attosecond</strong> Imaging - Dr. M. Kling:<br />

- Dr. Th. Uphues - O. Herrwerth - I. Znakovskaya<br />

- Dr. S. Zherebtsov - A. Wirth<br />

* received degree during report period;<br />

2 MAP= Munich Centre for Advanced Photonics<br />

132 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008


INTRODUCTION<br />

The motion of electrons <strong>and</strong> light wave oscillations<br />

(being mutually the cause of each other) occur at an<br />

inconceivable pace, measured in billionths of a billionth<br />

of a second, in attoseconds (asec) [1]. By drawing on the<br />

interaction of electrons <strong>and</strong> light, attosecond science<br />

[2] aims at measuring, controlling, <strong>and</strong> exploiting<br />

these processes: electron motion <strong>and</strong> light waves. The<br />

symbiosis of light <strong>and</strong> electrons is now being extended<br />

from nature to technology.<br />

<strong>Control</strong>led few-cycle light waves [3] exert a force<br />

on electrons that is controlled <strong>and</strong> variable on the<br />

attosecond scale. Electron motion driven by this force has<br />

allowed the reproducible generation <strong>and</strong> measurement<br />

of isolated attosecond pulses [4,5]. These tools <strong>and</strong><br />

techniques gave birth to a new technology, which we<br />

dub lightwave electronics [6]. It allows control <strong>and</strong><br />

measurement of the atomic-scale motion of electrons<br />

<strong>and</strong> light field oscillations, just as microwave electronics<br />

permits control <strong>and</strong> measurement of nanometer-scale<br />

current <strong>and</strong> microwave fields. <strong>Lightwave</strong> electronics has<br />

spawned experimental attosecond science <strong>and</strong> advanced<br />

metrology to the electronic time scale. Steering electron<br />

motion in molecules [7], real-time observation of the<br />

electrons’ quantum transitions deep inside atoms [8],<br />

their escape via tunneling [9], <strong>and</strong> their atomic-scale<br />

transport in solids [10] demonstrate the power of the<br />

new technology.<br />

Confinement of light energy to within a few wave cycles<br />

is opening another field of physics, related to ultrastrong<br />

fields <strong>and</strong> forces. The power of laser light <strong>and</strong> the<br />

strength of its electromagnetic fields grow inversely with<br />

pulse duration <strong>and</strong> may access unprecedented levels.<br />

The resultant light forces ionize matter, drive electrons to<br />

near the velocity of light – producing relativistic electrons<br />

– <strong>and</strong> can even completely separate them from ions, thus<br />

generating giant electric fields in plasma waves. These<br />

light <strong>and</strong> plasma fields can accelerate electrons within<br />

micro- <strong>and</strong> milli-meters to energies acquirable only over<br />

tens to hundreds of meters in conventional accelerators<br />

[11-14]. <strong>Control</strong> of the emerging multi-million-volt<br />

electrons with ultra-strong fields of controlled light<br />

waveforms will extend attosecond control to relativistic<br />

electron motion <strong>and</strong> open a new branch of attosecond<br />

science: high-field attosecond physics. It may give rise to<br />

intense attosecond pulses [15] <strong>and</strong> ultra-bright sources<br />

of hard X-rays <strong>and</strong> other energetic particles.<br />

The mission of the Joint LMU-MPQ Laboratory for<br />

<strong>Attosecond</strong> Physics (LAP) affiliated with the MPQ <strong>and</strong><br />

the Department of Physics of the LMU is to develop<br />

<strong>and</strong> advance attosecond science, including lightwave<br />

electronics, for steering <strong>and</strong> probing the pico- to<br />

nanometre-scale motion of electrons bound to atomic<br />

structures, <strong>and</strong> high-field attosecond physics, for<br />

1 . 3 . 1 S U M M A R Y O F S C I E N T I F I C A C T I V I T I E S<br />

controlling with nano- to micrometre precision the<br />

trajectories of relativistic electrons, <strong>and</strong> demonstrate<br />

applications of the new technologies in interdisciplinary<br />

collaborations. The “critical mass” required for successful<br />

pursuit of these dem<strong>and</strong>ing long-term objectives is now<br />

emerging thanks to very substantial resources being<br />

devoted to the research activities in these areas by the<br />

Max-Planck Gesellschaft (MPG) <strong>and</strong> by the Ludwig-<br />

Maximilians-Universität (LMU).<br />

Whilst the MPG provides invaluable regular funding<br />

for running LAP as well as for several major projects<br />

including the Petawatt Field Synthesizer, the LMU is the<br />

host university of the Cluster of Excellence: Munich-<br />

Centre for Advanced Photonics – MAP (directors: D.<br />

Habs <strong>and</strong> F. Krausz) [16], which sponsors a number of<br />

LAP projects <strong>and</strong> supplies LAP with optical components<br />

of key importance from its newly-created MAP Service<br />

Center. Moreover, MAP has permitted the establishment<br />

of several new professorships: W3 (full, tenured): theory<br />

of relativistic light-matter interactions, W2 (associate,<br />

tenured track): high-intensity laser technology; W2:<br />

brilliant X-ray sources; W2: ultrafast spectroscopy on<br />

surfaces; W2: theory of microscopic electron dynamics.<br />

Furthermore the successor of Prof. Habs is now being<br />

appointed at the level of W3, several years before<br />

the retirement of Prof. Habs, in the area of laserbased<br />

particle acceleration <strong>and</strong> high-field science. We<br />

shall pursue LAP’s long-term research goals in close<br />

collaboration with these colleagues. Once all these<br />

groups are established, some 150-200 researchers at<br />

LAP <strong>and</strong> the associated MAP-groups will be active in<br />

the fields of research outlined above, forming a critical<br />

mass for efficient pursuit of LAP’s many ambitious <strong>and</strong><br />

dem<strong>and</strong>ing long-term goals.<br />

1.3.1 SUMMARY OF RESEARCh ACTIVITIES<br />

Our research is organized in three main areas:<br />

• advancing femtosecond technology towards<br />

higher average or peak powers, towards<br />

b<strong>and</strong>widths spanning an octave <strong>and</strong> beyond<br />

<strong>and</strong> towards stabilizing the carrier-envelope<br />

phase <strong>and</strong> thereby the waveform of laser<br />

pulses<br />

<strong>and</strong> with the improved femtosecond tools developing<br />

<strong>and</strong> advancing<br />

• lightwave electronics: attosecond control <strong>and</strong><br />

spectroscopy, for gaining insight into <strong>and</strong><br />

control over the motion of electrons in atomic<br />

structures<br />

<strong>and</strong><br />

• high-field attosecond science: relativistic<br />

electron control, for developing intense<br />

attosecond sources <strong>and</strong> ultra-brilliant sources<br />

of high-energy particles (X-ray photons,<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 133


1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

electrons <strong>and</strong> ions) for a range of applications<br />

in physics <strong>and</strong> life sciences.<br />

Each of these areas builds upon strong synergies<br />

between theoretical <strong>and</strong> experimental research <strong>and</strong> is<br />

based on a number of collaborations with groups from<br />

all over the world. The facilities of our LMU-MPQ joint<br />

laboratory are situated at MPQ <strong>and</strong> the LMU Physics<br />

Building at Coulombwall 1. LAP members working at<br />

both locations closely co-operate, forming a single<br />

coherent group with weekly meetings, discussions <strong>and</strong><br />

intense communications on a daily basis.<br />

1.3.1.1 PUShING ThE FRONTIERS OF FEMTOSECOND<br />

TEChNOlOGY<br />

<strong>Attosecond</strong> technology is based upon <strong>and</strong> has grown<br />

out of femtosecond technology. Improving femtosecond<br />

technology towards<br />

(i) b<strong>and</strong>widths approaching, reaching, <strong>and</strong><br />

exceeding an octave,<br />

(ii) intense pulses at higher (MHz) repetition<br />

rates, i.e. higher average powers,<br />

(iii) higher peak powers both at kilohertz as<br />

well as 10 Hertz repetition rates<br />

<strong>and</strong> endowing it with<br />

(iv) carrier-envelope phase stabilization<br />

constitute basic requirements for advancing attosecond<br />

technology.<br />

Advances (i) <strong>and</strong> (iv) enable researches to generate<br />

intense laser pulses with stabilized waveform. Strongfield<br />

(multi-photon-assisted) interaction of these<br />

waveform-stabilized pulses with matter has provided<br />

access to the value of their carrier-envelope phase<br />

[3], allowing the determination <strong>and</strong> thereby full<br />

control of their waveform. We dub the sources of<br />

these waveform-controlled few-cycle light pulses light<br />

waveform synthesizers <strong>and</strong> consider them as the major<br />

technological pillars of attosecond science. Increasing<br />

their b<strong>and</strong>width (i) improves attosecond metrology<br />

by generating shorter isolated attosecond pulses with<br />

higher efficiency (see next chapter). Increasing their<br />

pulse repetition rate into the MHz regime (ii) opens<br />

the door for endowing coincidence spectroscopy with<br />

attosecond temporal resolution. Last but not least,<br />

increasing their peak power to the multi-terawatt<br />

regime (iii) is the prerequisite for extending attosecond<br />

control to relativistic electrons.<br />

BROADBAND ChIRPED DIElECTRIC MUlTIlAYER<br />

MIRRORS<br />

constitute an enabling technology for all these<br />

developments: they are the only optical components<br />

capable of providing dispersion control<br />

(i) over b<strong>and</strong>widths required for shaping optical<br />

waveforms within the wave cycle,<br />

(ii) with extremely low loss required for high-averagepower<br />

MHz-rate femtosecond pulse generation<br />

in high-Q laser resonators <strong>and</strong> passive build-up<br />

cavities,<br />

(iii) with low amount of material withst<strong>and</strong>ing high<br />

intensities, making them ideal for tailoring,<br />

steering <strong>and</strong> focusing ultrabroad-b<strong>and</strong>, multiterawatt<br />

pulses.<br />

Therefore we continue – even some 15 years after its<br />

invention [18] – our efforts to advance this technology.<br />

In collaboration with A. V. Tikhonravov <strong>and</strong> M. K.<br />

Trubestkov from Moscow State University, Vladimir<br />

Pervak <strong>and</strong> Alex<strong>and</strong>er Apolonskiy have developed a<br />

family of dispersive mirrors for precision dispersion<br />

control over an octave or more in the visible (VIS), nearinfrared<br />

(NIR) spectral range for widespread use in LAP’s<br />

ultrafast laser “workhorses” (see next section).<br />

With MAP Service Center’s newly-installed electron-beam<br />

evaporation system, the team has recently succeeded<br />

in the first realization of a chirped multilayer dielectric<br />

mirror providing dispersion control over the spectral<br />

range of 300-900 nm <strong>and</strong> the first use of hafnium oxide<br />

in a chirped mirror [19]. Simultaneously, the chirped<br />

mirror technology is being extended to the mid IR<br />

(MIR) up to wavelengths of 3 μm. Precision dispersion<br />

metrology via white light interferometry is also being<br />

extended to the same spectral range (from MIR to UV).<br />

These developments allow extension of broadb<strong>and</strong><br />

dispersion control to a spectral range spanning several<br />

octaves, paving the way towards the generation<br />

of mono-cycle to sub-cycle optical waveforms <strong>and</strong><br />

arbitrary optical waveform synthesis, as well as towards<br />

frequency combs over the UV-VIS-NIR-MIR spectral<br />

range. Moreover, hafnium-oxide-based chirped mirrors<br />

will lead to high-damage-threshold broadb<strong>and</strong> optics<br />

for high-power applications.<br />

lIGhT WAVEFORM SYNThESIZERS<br />

constitute key tools for attosecond science. Thanks to<br />

ongoing advances in chirped multilayer technology, we<br />

can now reliably generate intense (sub-terawatt) sub-<br />

1.5-cycle laser pulses in the VIS/NIR spectral range at<br />

kHz repetition rates in several workhorse laser systems<br />

on a daily basis in our laboratory. Figure 1 shows the<br />

spectrum <strong>and</strong> interferometric autocorrelation recorded<br />

by Adrian Cavalieri <strong>and</strong> coworkers, featuring clean, ~3.5fs-duration<br />

(full width at half maximum) laser pulses<br />

carried at a wavelength of ~ 750 nm (field oscillation<br />

period ~ 2.5 fs) [6,20]. This performance has recently<br />

been reproduced by Sergei Trushin, Izhar Ahmad <strong>and</strong><br />

co-workers, <strong>and</strong> by Eleftherios Goulielmakis <strong>and</strong> Martin<br />

Schultze using other LAP laser systems, demonstrating<br />

134 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008


A<br />

B<br />

C<br />

D<br />

E<br />

Figure 1: (A) Photo of the chirped dielectric multilayer<br />

mirrors providing high reflectivity <strong>and</strong> dispersion<br />

control over the spectral range of 450 – 1000 nm.<br />

(B) Schematic of the hollow-fibre/chirped mirror<br />

compressor seeded with carrier-envelope-phasestabilized<br />

25-fs, 1-mJ, near-infrared laser pulses<br />

from a Ti:sapphire laser system. Spectrum (C) <strong>and</strong><br />

interferometric autocorrelation (D) of sub-3.5-fs laser<br />

pulses carried at ~ 750 nm. (E) Possible electric field<br />

waveforms of the generated waveform-controlled<br />

sub-1.5-cycle laser pulses.<br />

1 . 3 . 1 S U M M A R Y O F S C I E N T I F I C A C T I V I T I E S<br />

the maturity of the technology of near-single-cycle<br />

optical waveform generation. This progress had direct<br />

implication to attosecond technology as reviewed in the<br />

next chapter.<br />

The intense near-mono-cycle NIR pulses offer the<br />

potential for few-femtosecond pulse generation in the<br />

deep UV <strong>and</strong> VUV by direct harmonic conversion, or –<br />

with their different spectral components independently<br />

adjustable, in terms of both amplitude <strong>and</strong> phase – for<br />

the synthesis of a wide range of optical waveforms.<br />

Ulrich Graf <strong>and</strong> Ivanka Grguras, under the supervision<br />

of Eleftherios Goulielmakis, pursued these goals. Ulrich<br />

Graf has succeeded in generating sub-4-fs, microjoule<br />

pulses in the deep UV, at a carrier wavelength of about<br />

250 nm by third-harmonic generation in a high-pressure<br />

neon gas cell <strong>and</strong> is now working on the extension of<br />

the technique to higher (5 th <strong>and</strong> 7 th ) harmonics <strong>and</strong><br />

shorter pulse durations. Ivanka Grguras has set up a<br />

proto-typical 3-channel optical waveform synthesizer<br />

for the 400-1200nm spectral range using chirped,<br />

dichroic beam splitters designed, manufactured,<br />

<strong>and</strong> characterized by Vladimir Pervak <strong>and</strong> Alex<strong>and</strong>er<br />

Apolonskiy. Preliminary frequency-resolved optical gating<br />

measurements of the signals transmitted through the<br />

individual channels have yielded b<strong>and</strong>width-limited fewcycle<br />

pulses, characterization of the overall throughput<br />

is under way. The programmable optical waveforms,<br />

once available, are expected to allow versatile steering<br />

of the atomic-scale motion of electrons in a wide range<br />

of systems. For example, they provide a large degree<br />

of freedom for controlling the electron trajectories in<br />

attosecond pulse production by high-order harmonics,<br />

offering manipulation of the generation process for<br />

different purposes (e.g. fine tuning the photon energy<br />

or optimizing the efficiency of generation). As another<br />

example, in Figure 2C(iii) a high-frequency wavepacket<br />

preceds a more slowly oscillating intense waveform,<br />

A<br />

Figure 2: (A) Chirped-mirror-based optical waveform<br />

synthesizer. Generic scheme of a multi-channel<br />

optical waveform synthesizer based on wide-b<strong>and</strong><br />

chirped dichroic beamsplitters (DBS).<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 135


B<br />

C<br />

D<br />

1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

Figure 2: (B) Reflectivity curves of DBSs designed<br />

<strong>and</strong> manufactured for a prototypical three-channel<br />

synthesizer covering the spectral range of 400-<br />

1200 nm, which was recently generated from a<br />

neon-gas-filled capillary, see Fig. 1. Channel #1:<br />

400-550 nm, Channel #2: 550-800 nm, Channel<br />

#3: 800-1200 nm) designed to support b<strong>and</strong> pass<br />

reflectance with minimized additional spectral<br />

phase. (C) Representative waveforms that can be<br />

synthesized with this prototype optical waveform<br />

synthesizer, by adjusting amplitude <strong>and</strong> phase (delay)<br />

of the individual channels (6 adjustable “knobs”). (D)<br />

Top-view of the synthesizer illuminated with whitelight<br />

supercontinuum from the system shown in<br />

Figure 1. The scheme can be readily scaled to more<br />

independently-addressable channels (as indicated<br />

in panel A) <strong>and</strong> extended towards longer as well as<br />

shorter wavelength, where chirped multilayer optics<br />

are now becoming available.<br />

which may open the door for a new type of reaction<br />

control: the high-frequency (UV or VUV) prepulse<br />

selectively creates a localized electronic excitation at<br />

the selected site of the molecule <strong>and</strong> the subsequent<br />

synthesized control wave steers the wavepacket across<br />

the molecule towards a target bond which is to be<br />

destroyed to trigger a reaction.<br />

lIGhT WAVEFORM SYNThESIS AT lONGER<br />

WAVElENGThS<br />

affords promise for scaling high-order harmonic<br />

generation to shorter wavelengths, owing to the<br />

quadratic scaling of (maximum) harmonic photon<br />

energy with the wavelength of the driving light pulse.<br />

This prospect motivated us to extend waveformcontrolled<br />

light pulse generation to the MIR. To this end<br />

we have produced broadb<strong>and</strong> MIR light by differencefrequency<br />

generation from a few-cycle Ti:sapphire laser<br />

<strong>and</strong> amplified it by optical parametric amplification<br />

(OPA).<br />

136 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008<br />

A<br />

B<br />

Figure 3: (A) Third-harmonic frequency-resolved<br />

optical gating spectrogram of the millijoule-scale,<br />

waveform-controlled mid-IR few-cycle pulses<br />

generated by a broadb<strong>and</strong> optical parametric<br />

amplifier. (B) Temporal intensity profile <strong>and</strong> phase<br />

of the mid-IR pulse retrieved from the TH-FROG<br />

measurement, revealing a pulse duration of 15 fs.<br />

Inset: Measured spectrum of the pulse.


Building upon the preliminary work of Andrius Baltuska,<br />

Takao Fuji <strong>and</strong> Nobuhisa Ishii [21], persistent efforts<br />

of Xun Gu, Yunpei Deng <strong>and</strong> coworkers have recently<br />

led to the first intense waveform-controlled sub-twocycle<br />

light pulses in the MIR, at a carrier wavelength of<br />

2.1 μm (Figure 3).<br />

With the high-power diode-pumped disc-laser being<br />

developed by Thomas Metzger <strong>and</strong> Wolfram Helml<br />

as a new pump source for the MIR-OPA, we pursue<br />

the development of the first terawatt infrared light<br />

wave synthesizer, which we dubbed LWS-1 (for target<br />

parameters, see survey of projects <strong>and</strong> goals). We expect<br />

LWS-1 to allow us to significantly advance attosecond<br />

metrology <strong>and</strong> spectroscopy, both in terms of temporal<br />

resolution <strong>and</strong> wavelength coverage.<br />

lIGhT WAVEFORM SYNThESIS AT RElATIVISTIC<br />

INTENSITIES<br />

may allow attosecond control of the trajectories of<br />

relativistic electrons, which, in turn, holds promise for<br />

attosecond pulse generation with high intensity <strong>and</strong>/or at<br />

high (X-ray) photon energies <strong>and</strong> for the development of<br />

compact sources of brilliant X-rays. These prospects <strong>and</strong><br />

the expected applications, which we shall briefly discuss<br />

in Chapter 1.3.1.3, have motivated us to pursue scaling<br />

the technology of light waveform synthesis to relativistic<br />

intensity levels. In 2003, Andrius Marcinkevicius <strong>and</strong><br />

Franz Tavella started developing a 10-Hz near-infrared<br />

noncollinear optical parametric chirped-pulse amplifier<br />

(OPCPA), which we dubbed LWS-10, for producing<br />

few-cycle pulses with a peak power of about 10 TW.<br />

After the leave of A. M. in 2006, Laszlo Veisz took over<br />

the project <strong>and</strong> – in collaboration with Franz Tavella –<br />

recently demonstrated reliable sub-10-fs, 10-TW pulse<br />

generation from this system, constituting the world’s<br />

first multi-terawatt few-cycle light source [22], see<br />

Figure 4. Simultaneously, Tibor Wittmann <strong>and</strong> coworkers<br />

have managed to develop a single-shot carrier-envelope<br />

phase meter, which will allow to measure <strong>and</strong> control<br />

the phase <strong>and</strong> thereby the waveform of the intense fewcycle<br />

pulses delivered at a low repetition rate [23].<br />

At the same time, Stefan Karsch, together with Zsuzsanna<br />

Major, Mathias Siebold, Sergei Trushin <strong>and</strong> several PhD<br />

students work on the development of a petawattscale<br />

few-cycle light source, dubbed the Petawatt Field<br />

Synthesizer (PFS), in the framework of a project funded<br />

by the Max Planck Society. Once commissioned, PFS<br />

will constitute the world’s first high-power light source<br />

delivering<br />

• few-cycle, waveform-controlled light with<br />

petawatt-scale peak power<br />

• petawatt-scale pulses at a 10 Hz repetition<br />

rate from compact (laboratory-scale) system<br />

1 . 3 . 1 S U M M A R Y O F S C I E N T I F I C A C T I V I T I E S<br />

A<br />

B<br />

Figure 4: Light wave synthesizer - 10 (LWS-10):<br />

the world’s first 10-TW sub-10-fs light source. (A)<br />

Overview of the system occupying approximately<br />

10 m 2 on an optical table. For details see Ref.<br />

[22]. (B) Single-shot second-order autocorrelation<br />

trace routinely recorded during operation shows a<br />

FWHM correlation width of 10.7 fs, implying a pulse<br />

duration (FWHM) of 7.6 fs under the assumption of<br />

a Gaussian pulse shape.<br />

<strong>and</strong> expected to open a new chapter in attosecond <strong>and</strong><br />

high-field science. The concept PFS is considered as a<br />

key technology for the realization of the Extreme Light<br />

Infrastructure (ELI), a pan-European project aiming at<br />

the development of an exawatt-scale light source (1<br />

Exawatt = 10 18 W)*. Thanks to the recent successful<br />

demonstration of the two main technological pillars<br />

of the enterprise: (i) the feasibility of ultrashort-pulsedriven<br />

broadb<strong>and</strong> OPCPA by József Fülöp, Zsuzsanna<br />

Major, Stefan Karsch <strong>and</strong> coworkers, see Figure 5 [24],<br />

<strong>and</strong> (ii) the feasibility of scaling diode-pumped Yb:YAG<br />

amplifiers to the Joule energy regime <strong>and</strong> beyond, by<br />

Mathias Siebold, Marco Klingebiel, Christian W<strong>and</strong>t <strong>and</strong><br />

collaborators from the Univ. Jena [25,26], the second<br />

phase of the project was now approved <strong>and</strong> we hope<br />

that the system can be commissioned before the end<br />

of 2010.<br />

* ELI was recently included in the ESFRI roadmap for future large-scale infrastructures in<br />

Europe <strong>and</strong> its Preparatory Phase has recently been funded by the EU.<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 137


1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

Figure 5: Ultrashort-pulse-driven OPCPA test system using ATLAS. (A) Pulse-front tilt matching of pump <strong>and</strong><br />

signal pulses. Maximum gain b<strong>and</strong>width (B) <strong>and</strong> corresponding transform limited pulse duration (C). Spectrum<br />

transmitted by our chirped-mirror compressor (D) along with pulse reconstructed from SPIDER measurements (E).<br />

lIGhT WAVEFORM SYNThESIS AT MEGAhERTZ<br />

REPETITION RATES<br />

affords promise to produce attosecond pulses at MHz<br />

repetition rate, which, in turn, would open the door for<br />

combining attosecond spectroscopy with coincidence<br />

detection, a long-st<strong>and</strong>ing dream of researcher in the<br />

field of attosecond science. By inventing the concept of<br />

chirped-pulse oscillators (CPO) our group has previously<br />

shown how to scale the energy <strong>and</strong> peak power of<br />

femtosecond pulses from mode-locked oscillators by<br />

lowering their repetition rate from 50-100 MHz by an<br />

order of magnitude [27]. Recently, a collaborative effort<br />

of Akira Ozawa, Thomas Udem (from the group of Prof.<br />

T. Hänsch) <strong>and</strong> Jens Rauschenberger, Alma Fern<strong>and</strong>ez<br />

<strong>and</strong> Alex<strong>and</strong>er Apolonskiy from our group coupled<br />

sub-microjoule energy pulses from our Ti:sapphire<br />

chirped-pulse oscillator into a passive build-up cavity to<br />

generate high-order harmonics up to photon energies of<br />

~ 30 eV (wavelength ~ 40 nm) with average powers of<br />

about one microwatt [28]. Implementing the concept<br />

with few-cycle light, which we currently pursue, will<br />

allow extension of frequency combs to the soft-X-ray<br />

regime <strong>and</strong> attosecond pulses at MHz repetition rates<br />

for attosecond coincidence spectroscopy.<br />

1.3.1.2 lIGhTWAVE ElECTRONICS: ATTOSECOND<br />

CONTROl & SPECTROSCOPY<br />

At the time of the previous progress report, extreme<br />

ultraviolet (XUV) pulses of a duration of 250 as at a<br />

photon energy of about 100 eV, containing some 10 7<br />

photons/pulse <strong>and</strong> delivered at a 1 kHz repetition rate<br />

have represented the state of the art of attosecond<br />

technology [4]. In our newly-developed AS-1<br />

attosecond beamline [6], these sub-femtosecond pulses<br />

along with their few-cycle NIR drivers have served as<br />

a st<strong>and</strong>ard tool for attosecond real-time interrogation<br />

of atomic-scale electron dynamics over the past two<br />

years. Simultaneously, the improved NIR waveforms<br />

reported above have been used to push the frontiers of<br />

attosecond pulse generation <strong>and</strong> metrology.<br />

ATTOSECOND SPECTROSCOPY<br />

has – in its first implementation [8] – drawn on<br />

electrons leaving the atom after their excitation. Intraatomic<br />

motion has been explored by probing outgoing<br />

electrons. Probing the transient population of shakeup<br />

states by means of strong-field-induced tunnelling<br />

ionization offers another means of interrogating<br />

intra-atomic electron dynamics (Figure 6). If lightfield-induced<br />

electron tunnelling occurs as predicted<br />

by Keldysh some four decades ago, the probability of<br />

setting the electron free from these shake-up states<br />

increases in sub-femtosecond steps near the oscillation<br />

peaks of the NIR laser field, where the binding potential<br />

is suppressed. Matthias Uiberacker, Martin Schultze <strong>and</strong><br />

Thorsten Uphues, with collaborators from the group of<br />

Karl Kompa, from Bielefeld <strong>and</strong> Amsterdam, have now<br />

experimentally verified this long-st<strong>and</strong>ing theoretical<br />

prediction, the cornerstone of strong-field theories,<br />

by ionizing neon atoms with a 250-as X-ray pulse <strong>and</strong><br />

removing the shake-up electron from neon ions by a<br />

138 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008


Figure 6: Electronic excitation <strong>and</strong> relaxation in<br />

atoms, molecules <strong>and</strong> solids. Motion is initiated by<br />

an attosecond XUV pulse <strong>and</strong> measured by sampling<br />

the temporal evolution of outgoing electrons via<br />

attosecond streaking spectroscopy or by probing the<br />

transient population of bound states via attosecond<br />

tunnelling spectroscopy. Both methods are<br />

implemented with a waveform-controlled few-cycle<br />

NIR pulse synchronised with attosecond accuracy<br />

to the XUV trigger pulse. The labels +1 <strong>and</strong> +2<br />

indicated single <strong>and</strong> double ionization.<br />

time-delayed few-cycle NIR field [9]. The measurement<br />

has yielded a temporal ionization profile revealing<br />

sharp steps approximately half a laser cycle apart, just<br />

as predicted by Keldysh 40 years ago. Each sub-cycle<br />

ionization step was measured to last less than 380<br />

attoseconds, setting a corresponding upper limit on<br />

shake up <strong>and</strong> tunnelling.<br />

This sub-femtosecond ionization step offers an<br />

attosecond probe for interrogating the transient<br />

population dynamics of excited (bound) states in atoms,<br />

molecules or solids following a sudden excitation.<br />

We have termed this novel technique attosecond<br />

tunnelling spectroscopy <strong>and</strong> used it for probing multielectron<br />

dynamics such as shake up <strong>and</strong> single as well<br />

as cascaded Auger processes for the first time with<br />

attosecond resolution. In these experiments [9] we not<br />

only confirmed several results previously obtained by<br />

frequency-domain measurements but also shed light<br />

on dynamics that could not be accessed by frequencydomain<br />

techniques so far.<br />

1 . 3 . 1 S U M M A R Y O F S C I E N T I F I C A C T I V I T I E S<br />

Until recently, attosecond spectroscopy was restricted<br />

to isolated atoms in the gas phase. Can it possibly<br />

be extended to condensed-matter systems: to solids<br />

<strong>and</strong> surfaces? In an international collaborative effort,<br />

led by Adrian Cavalieri <strong>and</strong> coordinated by Reinhard<br />

Kienberger, with important contributions from Bielefeld,<br />

San Sebastian, <strong>and</strong> Vienna, we have been able to answer<br />

this question affirmatively, by attosecond streaking of<br />

photoelectrons liberated from a tungsten crystal by a<br />

300-as XUV pulse [10]. We have studied two different<br />

types of electrons, liberated from localized core states<br />

of the 4f b<strong>and</strong> <strong>and</strong> the delocalized states of the<br />

conduction b<strong>and</strong> near the Fermi energy. The recorded<br />

streaking spectrograms, the artistic representations of<br />

Figure 7: Artistic representation of streaking<br />

spectrograms of photoelectrons, released by a 300attosecond<br />

95-eV XUV pulse from a tungsten crystal<br />

<strong>and</strong> recorded by a 5-fs, 750-nm NIR laser pulse.<br />

“Electrons move in solids at very high speed –<br />

traversing atomic layers <strong>and</strong> interfaces within tens<br />

to hundreds of attoseconds. These time intervals<br />

will ultimately limit the speed of the electronics<br />

of the future. Physicists have now experimentally<br />

probed such electron dynamics in real time. The<br />

cover illustrates the first attosecond spectroscopic<br />

measurement in a solid, revealing a 110-attsecond<br />

difference in the travel time of two different types<br />

of electrons following photoexcitation in a tungsten<br />

crystal. The ability to time electrons moving in solids<br />

over merely a few interatomic distances makes it<br />

possible to probe the solid-state electronic processes<br />

occurring at the ultimate speed limit <strong>and</strong> thus helps<br />

to advance technologies such as computation, data<br />

storage <strong>and</strong> photovoltaics,which all rely on exquisite<br />

control of electron transport in ever smaller structures<br />

of solid matter. [Cover art: Barbara Ferus, MPQ/LAP]”<br />

Description of cover, from the 25 October 2007 issue<br />

of Nature, page ix.<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 139


1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

which are shown in Figure 7, indicate sub-femtosecond<br />

emission time for both types of electrons. Closer<br />

inspection has yielded a discernible temporal shift<br />

between the two streaking spectrograms: careful data<br />

analysis reveals a delay in emission of the core electrons<br />

with respect to the conduction-b<strong>and</strong> electrons of about<br />

110 attoseconds. This means that the former reach the<br />

surface – from the several-Angstrom depth which they<br />

can escape from – some 110 as later than the latter. The<br />

analysis of Pedro Echenique has revealed that only half<br />

of this delay can be accounted for by the lower kinetic<br />

energy of core electrons, the other half is attributed to<br />

a larger effective mass of these electrons. This proof-ofconcept<br />

study opens the door for direct time-domain<br />

access to a wide range of hyperfast electron processes<br />

in solids <strong>and</strong> surfaces, such as e.g. charge transfer in<br />

host-guest systems or within molecular assemblies on<br />

surfaces, charge screening <strong>and</strong> e-e scattering in metals,<br />

or collective motion in surface plasmons.<br />

After being successfully demonstrated in atoms <strong>and</strong><br />

solids, attosecond spectroscopy now remains to be<br />

extended to molecular systems. As a matter of fact,<br />

the valence electronic states are typically separated by<br />

one or more electronvolts, indicative of a hyperfast<br />

motion of electron wavepackets composed of such<br />

states. The few-femtosecond-duration broadb<strong>and</strong><br />

deep UV <strong>and</strong> VUV pulses reported in Chapter 1.3.1.1<br />

lend themselves to launching electron wavepackets<br />

on molecular orbitals with substantial probability,<br />

whilst synchronized attosecond XUV pulses would<br />

ideally suit for probing the unfolding electronic <strong>and</strong><br />

structural dynamics in molecules. Both of these tools<br />

are derived from waveform-controlled few-cycle NIR<br />

pulses, our basic tool, which ensures attosecond<br />

synchronism between the deep-UV/VUV <strong>and</strong> XUV<br />

pulses. In order that these tools become available for<br />

real-time observation of hyperfast electronic motion<br />

as well as complex, intertwined electronic <strong>and</strong> nuclear<br />

dynamics in molecules, Marcus Fieß, Wolfram Helml,<br />

<strong>and</strong> Eleftherios Goulielmakis – under the guidance of<br />

Reinhard Kienberger – have developed a next-generation<br />

attosecond beamline: AS-2, which accommodates two<br />

frequency-conversion assemblies, for attosecond XUV<br />

<strong>and</strong> few-femtosecond UV/VUV pulse generation. Both<br />

beams can be recombined on target with attosecond<br />

timing accuracy.<br />

Encouraged by the successful extension of attosecond<br />

spectroscopy to solids [10], a team led by Reinhard<br />

Kienberger: Elisabeth Magerl, Adrian Cavalieri, <strong>and</strong> Ralph<br />

Ernstorfer – in close cooperation with our MAP partners:<br />

Peter Feulner, Dietrich Menzel <strong>and</strong> Johannes Barth from<br />

the Technische Universität München, are developing<br />

the world’s first ultrahigh vacuum beamline for<br />

attosecond surface <strong>and</strong> solid-state spectroscopy: AS-3.<br />

We are hopeful that measurement in AS-2 <strong>and</strong> AS-3<br />

will make several contributions to the highlights of our<br />

forthcoming progress report.<br />

ADVANCING TIME-RESOlVED METROlOGY INTO<br />

ThE SUB-100-AS DOMAIN<br />

is indispensable for direct time-domain insight into<br />

electron correlations on the atomic scale. These govern<br />

or affect such fundamental processes as the intraatomic<br />

energy transfer between electrons, the response<br />

of atomic electron systems to external influence, or its<br />

rearrangement following the sudden loss of one or more<br />

electrons. Detailed insight into the electronic response<br />

of atomic-scale solid-state structures to strong external<br />

fields of infrared or visible light, e.g. via non-adiabatic<br />

tunneling is required for the development of solid-state<br />

lightwave electronics <strong>and</strong> will also rely on sub-100-as<br />

temporal resolution.<br />

By confining optical field ionization to a single, wellcontrolled<br />

light oscillation period, the sub-1.5-cycle NIR<br />

laser pulses described in Chapter 1.3.1.1 along with<br />

novel broadb<strong>and</strong> XUV multilayer mirrors developed<br />

by Michael Hofstetter <strong>and</strong> Ulf Kleineberg have recently<br />

permitted our team, Eleftherios Goulielmakis, Martin<br />

Schultze <strong>and</strong> coworkers, to generate isolated sub-100as<br />

pulses of XUV light [29], the measured XUV pulse<br />

profile <strong>and</strong> the NIR electric field waveform are shown<br />

in Figure 8. Detailed evaluation of the measurement<br />

has been performed with a new algorithm developed<br />

by Justin Gagnon <strong>and</strong> Vladislav Yakovlev [30]. The<br />

quasi-monocycle driving field benefits attosecond pulse<br />

generation in several ways. Abrupt onset of ionization<br />

within a single half cycle minimizes the backgroundfree<br />

electron density at the instant of harmonic pulse<br />

emission, minimizing thereby distortion of the driving<br />

wave <strong>and</strong> its dephasing with the generated harmonics<br />

during propagation. Hence coherent build-up of the<br />

harmonic emission over extended propagation is<br />

maximized. In addition, order-of-magnitude variation of<br />

the ionization probability between adjacent half-cycles<br />

creates unique conditions for single sub-100-as pulse<br />

emission without the need for sophisticated gating<br />

techniques.<br />

On the measurement side, improved resolution results<br />

from three (almost equally) important advances: (i)<br />

shorter XUV pulse duration, (ii) higher signal-to-noise<br />

ratio (S/N) <strong>and</strong>/or shorter overall measurement time due<br />

to the increased XUV photon flux, <strong>and</strong> (iii) the feasibility<br />

of stronger streaking before the onset of the NIR-fieldinduced<br />

ionization in attosecond streaking or enhanced<br />

S/N due to reduced number of tunnelling steps in<br />

attosecond tunnelling spectroscopy.<br />

Consequences of these favourable conditions include the<br />

routine generation of isolated sub-100-as XUV pulses<br />

(at a carrier photon energy of about 80 eV), for details<br />

140 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008


1 . 3 . 1 S U M M A R Y O F S C I E N T I F I C A C T I V I T I E S<br />

Figure 8: High-fidelity measurement of an isolated sub-100-attosecond pulse <strong>and</strong> its quasi-mono-cycle 3.3-fsduration<br />

(FWHM) driver laser waveform allows precision tests of basic models of strong-field interactions for the<br />

first time. Isolation of a single attosecond pulse means isolation of a single tunnelling <strong>and</strong> recollision event <strong>and</strong><br />

wavepacket motion between these the instants of ionization <strong>and</strong> recollision within an oscillation period of the<br />

driving laser field (shown in red in panel A) that is precisely known from an attosecond streaking measurement.<br />

A simple analysis of classical electron trajectories between ionization <strong>and</strong> recollision (black lines in panel A) in the<br />

framework of the semiclassical model of Paul Corkum implies a positive <strong>and</strong> negative chirp of the emitted XUV<br />

pulse for trajectories shorter <strong>and</strong> longer than the one resulting in highest-energy emission (green line in panel B).<br />

As an example of the power of high-fidelity attosecond metrology, we have calculated this intrinsic spectral chirp,<br />

i.e. the variation of group delay versus photon energy, carried by the attosecond pulse during its emergence, from<br />

the chirp measured by attosecond streaking <strong>and</strong> the known dispersion of the metal filter traversed by the pulse on<br />

its way from the source to the measurement. The degree of agreement of the measured chirp with the prediction<br />

of the semiclassical model is stunning, indicating that the emission from long trajectories is completely suppressed<br />

in the far field by spatial filtering, <strong>and</strong> secondly classical electron trajectory analysis provides a powerful tool for<br />

predicting the outcome of strong-field interactions.<br />

see Selected Reprint 11) “Single-cycle nonlinear optics”<br />

on pages 185 - 188, with a flux of greater than 10 11<br />

photons/s, exceeding by several orders of magnitude<br />

the flux of previously-reported sources of sub-fs light;<br />

a sub-10% measurement accuracy of the XUV pulse<br />

parameters <strong>and</strong> the NIR field evolution, suggesting that<br />

the temporal resolution of this pump-probe sampling<br />

system may approach or even reach the atomic unit of<br />

time (~ 24 as).<br />

OUR lONG-TERM OBJECTIVES<br />

<strong>Attosecond</strong> metrology includes the scaling of attosecond<br />

pulse generation towards higher photon energies<br />

<strong>and</strong> combination of high temporal with high spatial<br />

resolution in spectroscopy.<br />

Vladislav Yakovlev has analytically as well as numerically<br />

studied the generation of soft-X-ray harmonics with<br />

few-cycle waveforms of wavelength, λ L , varied from the<br />

near to the mid IR [31]. His studies predict that selfphase-matching<br />

becomes more efficient for longer<br />

wavelengths <strong>and</strong> under specific circumstances can<br />

provide ideal conditions for coherent growth of soft-<br />

X-ray harmonics over extended propagation lengths.<br />

Enhanced phase matching may thereby overcompensate<br />

the decrease in single-atom emission intensity for longer<br />

Figure 9: Spectral intensity distribution of soft-X-ray<br />

harmonics generated with a two-cycle laser pulse of<br />

a carrier wavelength ranging from 0.75 μm to 3 μm.<br />

λ L , <strong>and</strong> may result in orders of magnitude enhancement<br />

of the harmonic yield (see e.g. in Figure 9 harmonics in<br />

the 400 eV <strong>and</strong> 700 eV range for λ L = 2.1 μm <strong>and</strong><br />

λ L = 1.5 μm, respectively), as well as efficient filtering of<br />

isolated attosecond pulses by restricting phase matching<br />

to a limited spectral b<strong>and</strong> (in the example shown in<br />

Figure 8: 350-450nm <strong>and</strong> 600-700eV for λ L = 2.1 μm<br />

<strong>and</strong> λ L = 1.5 μm , respectively).<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 141


1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

In another study, Vladislav Yakovlev has investigated the<br />

possibility of coherent superposition of laser-driven SXR<br />

harmonics in successive sources by one <strong>and</strong> the same<br />

laser pulse. His prediction of the feasibility of quasi-<br />

A<br />

B<br />

Figure 10: High-order harmonic generation from<br />

successive sources. (A) Schematic illustration of a<br />

sequence of harmonic sources formed by gas jets.<br />

The sources are arranged along the optical axis of<br />

a focused laser beam <strong>and</strong> pumped by one <strong>and</strong> the<br />

same laser pulse. In our first proof-of-concept study<br />

we have employed only two sources (depicted in<br />

black), further ones can be added for scaling the<br />

SXR harmonic yield in future experiments (depicted<br />

in grey). (B) Increasing the density of atomic dipole<br />

emitters tends to increase the coherent harmonic<br />

yield. However, increasing density causes the dipole<br />

oscillators to increasingly dephase along the laser<br />

propagation direction. In a single source, this leads to<br />

saturation of the harmonic yield at an atomic density<br />

corresponding to a backing pressure of ~ 40 mbar,<br />

red curve (calculated) <strong>and</strong> squares (measured).<br />

Splitting the generation medium into two equal<br />

sections <strong>and</strong> moving them apart so that the phase<br />

of the atomic dipole oscillations gets shifted by in<br />

the focused laser beam allows the atomic density to<br />

be increased by a factor of two (at backing pressure<br />

of ~ 80 mbar), leading to saturation at a factor of<br />

four higher harmonic intensities (blue curve <strong>and</strong><br />

diamonds).<br />

phase-matched SXR harmonic generation by a focused<br />

laser beam in a gas medium of modulated density<br />

has meanwhile been confirmed by our experimental<br />

observation of constructive <strong>and</strong> destructive interference<br />

between the harmonic signals originating from two<br />

successive sources, see Figure 10 [32]. Our proof-ofconcept<br />

study opens the prospect of enhancing the<br />

photon flux of SXR harmonic sources to levels enabling<br />

researchers to tackle a range of applications in physical<br />

as well as life sciences.<br />

Should attosecond pulse generation be scalable to<br />

photon energies of several kiloelectronvolts, attosecond<br />

X-ray diffraction might allow, one day, 4-dimensional<br />

microscopy with attosecond temporal <strong>and</strong> picometre<br />

spatial resolution. Drawing on the pioneering work of<br />

Ahmed Zewail, we also pursue this long-term goal by<br />

advancing ultrafast electron diffraction. To this end,<br />

Ernst Fill, Laszlo Veisz, <strong>and</strong> Sasha Apolonskiy have<br />

presented a novel concept of an electron gun for<br />

generating few-femtosecond- to sub-femtosecondduration<br />

electron bunches [33]. The basic idea is<br />

to utilize a DC acceleration stage combined with a<br />

microwave cavity, the time-dependent field of which<br />

generates an electron energy chirp for bunching at the<br />

target. To reduce space charge broadening, the number<br />

of electrons in the bunch is reduced <strong>and</strong> the gun is<br />

operated at a MHz repetition rate. In the absence of<br />

space charge (one electron per bunch), the duration<br />

of the generated electron wavepacket may potentially<br />

be shortened below 1 fs. The team around Ernst Fill,<br />

Martin Centurion <strong>and</strong> Peter Reckenthäler, have also<br />

devised a technique for measuring the bunch duration<br />

with potentially attosecond resolution [34] <strong>and</strong> have<br />

also performed preliminary time-resolved experiments<br />

with (multi-)electron bunches, which allowed them to<br />

image dynamics changes in the density distribution of a<br />

laser-induced plasma [35].<br />

1.3.1. 3 hIGh-FIElD ATTOSECOND SCIENCE: RElA-<br />

TIVISTIC ElECTRON CONTROl<br />

As described in Chapter 1.3.1.1, we pursue scaling of<br />

waveform-controlled few-cycle laser pulse generation to<br />

relativistic intensities in the expectation that controlling<br />

the motion of relativistic electrons with attosecond<br />

precision in time <strong>and</strong> nano- to micrometre precision in<br />

space paves the way towards scaling attosecond pulse<br />

generation to higher flux <strong>and</strong>/or photon energies <strong>and</strong><br />

may allow the development of compact, laboratory<br />

sources of brilliant X-rays. Simultaneously with our<br />

efforts aiming at laser development, in collaboration<br />

with Dieter Habs <strong>and</strong> his group, we are also conducting<br />

relevant proof-of-concept studies drawing on our<br />

laser sources currently available. Two routes are being<br />

pursued at MPQ: the development of a compact X-ray<br />

free electron laser seeded by laser-accelerated electron<br />

bunches <strong>and</strong> coherent harmonic XUV/X-ray pulse<br />

generated from relativistically-driven surfaces. In this<br />

142 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008


chapter we summarize first promising results of these<br />

proof-of-concept experiments.<br />

lASER-DRIVEN ElECTRON ACCElERATION<br />

affords promise for the development of compact,<br />

laboratory-sized X-ray free electron lasers. Simulations<br />

of our collaborators, Florian Grüner, Stefan Becker,<br />

Figure 11: Simulation results of Michael Geissler <strong>and</strong><br />

Jürgen Meyer-ter-Vehn showing the physical state<br />

of MPQ’s few-cycle-laser-driven plasma accelerator<br />

after the laser pulse has travelled a distance of<br />

130 μm inside the plasma. (A) Electron density<br />

(grey-scale false-colour plot) <strong>and</strong> the instantaneous<br />

laser intensity (rainbow-scale false-colour diagram).<br />

It is evident that the laser pulse fits within the<br />

bubble radius. The red line shows the longitudinal<br />

accelerating field caused by charge separation in the<br />

generated plasma wave (forward-directed field has<br />

positive value). The field strength is about 0.45 TV/m<br />

at the location of the bunch. (B) Lineouts of electron<br />

density <strong>and</strong> instantaneous laser intensity along the<br />

optical axis of the laser beam. The injected electron<br />

bunch (red) contains 4.5 pC charge <strong>and</strong> has an order<br />

of magnitude higher density than the background<br />

plasma. There is no direct interaction between the<br />

laser pulse <strong>and</strong> the electron bunch during its forward<br />

acceleration in the plasma “bubble”.<br />

1 . 3 . 1 S U M M A R Y O F S C I E N T I F I C A C T I V I T I E S<br />

Dieter Habs <strong>and</strong> coworkers, indicate that seeding lasergenerated,<br />

ultrahigh-peak-current, GeV-energy electron<br />

bunches into a compact, several-metre-long undulator<br />

may allow – one day – the realization of X-ray laser<br />

operation, producing ultra-brilliant, few-femtosecond<br />

X-ray pulses of several-Angström wavelength [36].<br />

Comprehensive 3-dimensional PIC simulations of Michael<br />

Geissler, Jürgen Meyer-ter-Vehn (MPQ) <strong>and</strong> Alex<strong>and</strong>er<br />

Pukhov (Düsseldorf) have predicted the feasibility of<br />

making electron bunches with the required parameters<br />

available with laser-driven plasma accelerators.<br />

A promising implementation relies on “broken” plasma<br />

waves driven by a laser pulse shorter than the half<br />

plasma oscillation period (Figure 11).<br />

In the absence of intense laser pulses of the required<br />

duration, first experiments drew on longer (several-<br />

10-fs) driver pulses. Under these circumstances,<br />

mono-energetic electron acceleration is preceded<br />

by a nonlinear interaction of the laser pulse with the<br />

Figure 12: Typical spectra of sub-10-fs-laser-driven<br />

mono-energetic electron beams free from thermal<br />

background. (A) Spectra with mean energies of 13.4,<br />

17.8 <strong>and</strong> 23 MeV, <strong>and</strong> relative energy spreads of<br />

11%, 4.3% <strong>and</strong> 5.7% (FWHM) respectively. (B) Results<br />

obtained with a spectrometer capable of accessing<br />

electron energies down to the non-relativistic<br />

regime, confirming the absence of an exponential<br />

background down to energies in the keV range. Here<br />

the mean energies <strong>and</strong> the relative energy spread<br />

are 4.1 <strong>and</strong> 9.7 MeV, <strong>and</strong> 14% <strong>and</strong> 9.5% (FWHM),<br />

respectively.<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 143


1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

Figure 13: Gigaelectronvolt-scale laser wake field accelerator at MPQ. G: GRENOUILLE pulse diagnostics, C1-C5:<br />

cameras; D1, D2: diodes; L1, L2: scintillating screens; S: spectrometer; W: wedge; OAP: off-axis parabola; CAP:<br />

capillary waveguide. The inset shows the incoming laser focus at the capillary entrance.<br />

relativistic plasma, which shortens its duration into the<br />

required domain. Longer-than-optimal driver pulses<br />

compromise efficiency as well as reproducibility, <strong>and</strong><br />

result in copious amounts of low-energy electrons<br />

accompanying the mono-energetic emission with an<br />

exponentially-decaying spectrum, forming a “thermal”<br />

background. By using the multi-terawatt, sub-10-fs<br />

pulses from LWS-10, Karl Schmid, Laszlo Veisz <strong>and</strong><br />

coworkers, in close cooperation with Ulrich Schramm,<br />

Dieter Habs <strong>and</strong> collaborators from the Univ. Düsseldorf<br />

have recently demonstrated the first electron accelerator<br />

based on high-density plasma waves driven with<br />

laser pulses that fit in one half of the plasma period<br />

(Figure 11) [14]. Direct “impulsive” excitation of the<br />

plasma wave permits mono-energetic electron<br />

acceleration virtually free from a thermal background<br />

for the first time. In our experiments, 5-terawatt, 8femtosecond<br />

laser pulses yield electron bunches up to<br />

energies of 25 MeV (Figure 12). The flux of low-energy<br />

electrons dramatically reduced as compared to earlier<br />

experiments also manifests itself in a strongly-reduced<br />

secondary radiation emerging from the accelerator <strong>and</strong><br />

offers the potential for enhancing efficiency <strong>and</strong> stability<br />

with more intense driver pulses.<br />

The electron energy can be further boosted by laser-driven<br />

plasma-wake-field acceleration (LWFA) in a discharge<br />

capillary to the gigaelectronvolt range, as demonstrated<br />

by Wim Leemans <strong>and</strong> coworkers at Berkeley in 2006.<br />

Jens Osterhoff, Antonia Popp, Zsuzsanna Major, Stefan<br />

Karsch <strong>and</strong> coworkers, in close co-operation with Dieter<br />

Habs <strong>and</strong> his group, have recently also demonstrated<br />

LWFA to the GeV frontier from a discharge capillary<br />

waveguide provided by Simon Hooker from Oxford<br />

University by driving the accelerator with 0.75 J, 40fs,<br />

800-nm pulses from ATLAS, MPQ’s multiterawatt<br />

Ti:Sa laser, see Figure 13 [12].<br />

Quasi-monoenergetic electron beams with energies as<br />

high as 500 MeV have been detected, with features<br />

reaching up to 1 GeV, albeit with large shot-to-shot<br />

fluctuations. The beam divergence <strong>and</strong> the pointing<br />

stability in this case are around or below 1 mrad <strong>and</strong><br />

8 mrad, respectively. These shot-to-shot electron-bunch<br />

parameter variations are greatly suppressed by utilizing<br />

the capillary as a gas-vessel without pre-ionization by an<br />

external electrical discharge. In this regime of operation<br />

quasi-monoenergetic electron beams of up to ~200<br />

MeV in energy have been observed (Figure 14). These<br />

beams emitted within a low-divergence cone of mrad<br />

FWHM display unprecedented shot-to-shot stability<br />

in energy (2.5% RMS), pointing (1.4 mrad RMS) <strong>and</strong><br />

charge (16% RMS) owing to a highly reproducible gasdensity<br />

profile within the interaction volume (Figure15)<br />

[13]. The excellent stability of bunch parameters affords<br />

promise for the potential ability of laser-accelerated<br />

electron bunches to meet the stringent criteria for<br />

seeding free electron lasers.<br />

INTENSE ATTOSECOND XUV/SXR PUlSES<br />

might allow triggering as well as probing ultrafast<br />

electron dynamics with attosecond pulses, i.e.<br />

attosecond XUV-pump/XUV-probe spectroscopy,<br />

which has not been possible so far. Relativistic laserplasma<br />

interactions have been identified as a promising<br />

approach to achieving this goal. Recent experiments<br />

confirmed that relativistically-driven overdense plasmas<br />

are able to convert infrared laser light into harmonic<br />

XUV radiation with unparalleled efficiency <strong>and</strong> that the<br />

generation technique is scalable towards hard X-rays.<br />

In a recent experiment (Figure 15) Yutaka Nomura,<br />

Rainer Hörlein, Sergey Rykovanov, George Tsakiris <strong>and</strong><br />

coworkers, in collaboration with the teams of Matt Zepf<br />

from Belfast <strong>and</strong> Dimitris Charalambidis from Heraklion,<br />

have recently succeeded in obtaining conclusive<br />

experimental evidence for the phases of the XUV<br />

harmonics emanating from the interaction that they are<br />

144 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008


1 . 3 . 1 S U M M A R Y O F S C I E N T I F I C A C T I V I T I E S<br />

Figure 14: (A) False-color images of 40 consecutive, spatially-dispersed electron beams on the fluorescence screen<br />

L2 behind the permanent dipole magnet (see Figure 13) recorded at a plasma density of n e ≈ 7.3 x 10 18 cm -3 .<br />

Each shot is normalized to its maximum value. (B) Five sample electron spectra from the same data set. (C) Five<br />

exemplary spectra for an electron density of n e ≈ 6.8 x 10 18 cm -3 . Ten consecutive fluorescence images obtained<br />

under those conditions are presented in the inset.<br />

synchronized to allow attosecond temporal bunching<br />

(Figure 16) [15]. Along with the previous findings about<br />

energy conversion <strong>and</strong> recent advances in high-power<br />

laser technology, our experiment demonstrates the<br />

feasibility of confining unprecedented amounts of light<br />

energy to within less than a femtosecond, initiating<br />

thus the symbiosis of relativistic high-field science <strong>and</strong><br />

attosecond science.<br />

1.3.1.4 VISIONS: SPACE-TIME IMAGING & BIO-<br />

MEDICAl APPlICATIONS<br />

<strong>Attosecond</strong> spectroscopy has enabled us to observe in<br />

real time how electrons undergo quantum transitions.<br />

However, in these experiments, we could not obtain<br />

information about how the electrons move in space.<br />

The atomic-scale density distribution of electrons can<br />

Figure 15: Fluorescence screen images of the radial intensity profile of the accelerated electron beam at the<br />

entrance of the dipole spectrometer (L1 in Figure 13). (A) Summed signal of 74 electron beams <strong>and</strong> their individual<br />

centers at n e ≈ 7.8 x 10 18 cm -3 . RMS shot-to-shot pointing fluctuations under these conditions are 1.4 mrad (yaxis)<br />

<strong>and</strong> 2.2 mrad (x-axis), respectively. (B) The signal of a single electron beam with a FWHM divergence of 1.6<br />

mrad. The crosses in the background originate from markings on the screens.<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 145


1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

Figure 16: Surface harmonic generation from relativistic<br />

plasmas <strong>and</strong> its temporal characterization.<br />

The p-polarized laser beam is focused onto a<br />

rotating disk-shaped target. The light reflected in<br />

the specular direction is re-collimated by a parabolic<br />

mirror <strong>and</strong> deflected towards a glass plate. The<br />

reflection off the glass plate close to Brewster’s<br />

angle (~57°) suppresses the p-polarized IR laser light<br />

while reflecting a substantial fraction (~5% in p-polarization)<br />

of the XUV harmonic emission. Passage<br />

through a 150-nm In filter selects harmonics greater<br />

than H8 <strong>and</strong> suppresses low-order harmonics as<br />

well as the residual laser light. The beam of selected<br />

harmonics (from H8 to H14) is focused by a spherical<br />

mirror split into two halves, serving as a focusing<br />

wave-front divider. The insets show data of a typical<br />

mass (upper) <strong>and</strong> harmonic (lower) spectrum.<br />

be resolved in space by conventional microscopic <strong>and</strong><br />

X-ray or electron diffraction techniques, but they can do<br />

so only if the system is at rest. Our long-term research<br />

goal is to combine sub-atomic resolution in space <strong>and</strong><br />

in time by developing sources of sub-femtosecond,<br />

Angström-wavelength electron <strong>and</strong> photon pulses. By<br />

recording a series of freeze-frame sub-femtosecond<br />

diffraction images of electron distributions by means<br />

of attosecond electron or X-ray diffraction, we will be<br />

able to observe atomic-scale electronic motion in matter<br />

with sub-atomic (picometre <strong>and</strong> attosecond) resolution<br />

in space <strong>and</strong> time. Once this dream comes true, we shall<br />

be able to make movies of all dynamics of the microcosm<br />

outside the atomic core.<br />

Our pursued laser-based high-energy particle <strong>and</strong><br />

photon sources also open up exciting prospects for<br />

applications in life sciences. Ultra-brilliant, femtosecond<br />

X-ray pulses from a future laser-driven XFEL may permit<br />

determination of the atomic structure of biomolecules<br />

without the need for crystallization by recording<br />

flash X-ray diffraction images of single molecules.<br />

Furthermore, collimated, laser-driven beams of mono-<br />

Figure 17. Nonlinear volume autocorrelation of the<br />

coherent XUV beam comprising harmonics H8 to<br />

H14. The red dots represent He + signal produced by<br />

two-photon ionization while the blue dots depict<br />

H 2 O + signal resulting from single-photon ionization.<br />

The H 2 O signal level relative to the He + signal is<br />

scaled arbitrarily. (A) Coarse scan over a delay interval<br />

corresponding to the laser pulse duration. A Gaussian<br />

pulse fit (green dashed line), yields the overall<br />

duration of the XUV emission of T XUV = 44±20 fs.<br />

(B) A fine scan taken near zero delay with a delay<br />

step size of 133 as, corresponding to about 20 data<br />

points per laser cycle. The dashed green line is a fit<br />

to the raw data of a sequence of Gaussian pulses<br />

of τ XUV = 0.9 fs in duration to the second-order XUV<br />

autocorrelation signal.<br />

energetic protons/ions <strong>and</strong> X-rays affords promise for<br />

opening a new chapter in cancer therapy <strong>and</strong> diagnosis,<br />

respectively. One of the main long-term objectives of<br />

the Munich-Centre for Advanced Photonics (MAP) [16]<br />

initiated <strong>and</strong> led by Dieter Habs <strong>and</strong> Ferenc Krausz is to<br />

bring together the requisite brain power <strong>and</strong> expertise<br />

for developing these novel photon-based ultra-brilliant<br />

electron/proton/ion/X-ray sources <strong>and</strong> demonstrating<br />

their suitability for atomic-resolution bioimaging,<br />

146 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008


high-resolution, low-dose cancer diagnostics <strong>and</strong> costeffective<br />

cancer therapy.<br />

REFERENCES<br />

[1] M. Hentschel et al.; <strong>Attosecond</strong> metrology. Nature<br />

414, 509 (2001).<br />

[2] P. B. Corkum <strong>and</strong> F. Krausz; <strong>Attosecond</strong> science.<br />

Nature Phys. 3, 381 (2007).<br />

[3] A. Baltuska et al.; <strong>Attosecond</strong> control of electronic<br />

processes by intense light fields. Nature 421, 611<br />

(2003).<br />

[4] R. Kienberger et al.; Atomic transient recorder.<br />

Nature 427, 817 (2004).<br />

[5] E. Goulielmakis et al.; Direct measurement of light<br />

waves. Science 305, 1267 (2004).<br />

[6] E. Goulielmakis et al.; <strong>Attosecond</strong> control <strong>and</strong><br />

measurement: lightwave electronics. Science 317, 769<br />

(2007).<br />

[7] M. F. Kling et al.; <strong>Control</strong> of electron localization in<br />

molecular dissociation. Science 312, 246 (2006).<br />

[8] M. Drescher et al.; Time-resolved atomic inner-shell<br />

spectroscopy. Nature 419, 803 (2002).<br />

[9] M. Uiberacker et al.; <strong>Attosecond</strong> real-time<br />

observation of electron tunnelling in atoms.<br />

Nature 446, 627 (2007).<br />

[10] A. L. Cavalieri et al.; <strong>Attosecond</strong> spectroscopy in<br />

condensed matter. Nature 449, 1029 (2007).<br />

[11] A. Pukhov <strong>and</strong> J. Meyer-ter-Vehn; Laser wake field<br />

acceleration: the highly non-linear broken-wave regime.<br />

Appl. Phys. B 74, 355 (2002).<br />

[12] S. Karsch et al.; GeV-scale electron acceleration in<br />

a gas-filled capillary discharge waveguide. New J. Phys.<br />

9 415 (2007).<br />

[13] J. Osterhoff et al.; Generation of Stable, Low-<br />

Divergence Electron Beams by Laser Wakefield<br />

Acceleration in a Steady-State-Flow Gas Cell. Submitted<br />

to Physical Review Letters. (2008).<br />

[14] K. Schmid et al.; Few-Cycle-Laser-Driven Electron<br />

Accelerator. Submitted to Nature Physics (2008).<br />

[15] Y. Nomura et al.; <strong>Attosecond</strong> phase locking of<br />

harmonics emitted from laser-produced plasmas.<br />

Submitted to Nature Physics (2008).<br />

[16] www.munich-photonics.de<br />

[17] G. Paulus et al.; <strong>Measurement</strong> of the phase of fewcycle<br />

laser pulses. Phys. Rev. Lett. 91, 253004 (2003).<br />

[18] R. Szipöcs, K. Ferencz, Ch. Spielmann, F. Krausz;<br />

Chirped multilayer coatings for broadb<strong>and</strong> dispersion<br />

control in femtosecond lasers. Opt. Lett. 19, 201<br />

(1994).<br />

[19] V. Pervak, F. Krausz, A. Apolonski; Dispersion<br />

control over the UV-VIS-NIR spectral range with HfO 2 /<br />

SiO 2 chirped dielectric multilayers. Opt. Lett. 32, 1183<br />

(2007).<br />

[20] A. L. Cavalieri et al.; Intense 1.5-cycle near infrared<br />

1 . 3 . 1 S U M M A R Y O F S C I E N T I F I C A C T I V I T I E S<br />

laser waveforms <strong>and</strong> their use for the generation of<br />

ultra-broadb<strong>and</strong> soft-x-ray harmonic continua. New J.<br />

Phys. 9, 242 (2007).<br />

[21] T. Fuji et al.; Parametric amplification of few-cycle<br />

carrier-envelope phase-stable pulses at 2.1 μm. Opt.<br />

Lett. 31, 1103 (2006).<br />

[22] F. Tavella et al.; Dispersion management for a sub-<br />

10-fs, 10 TW optical parametric chirped-pulse amplifier.<br />

Opt. Lett. 32, 2227 (2007).<br />

[23] T. Wittmann et al.; Single-shot measurement of<br />

the carrier-envelope phase. Submitted to Nature Phys.<br />

(2008)<br />

[24] J. A. Fülöp et al.; Short-pulse optical parametric<br />

chirped-pulse amplification for the generation of highpower<br />

few-cycle pulses, New J. Phys. 9, 438 (2007).<br />

[25] M. Siebold et al.; High-energy, diode-pumped,<br />

nanosecond Yb:YAG MOPA system. Opt. Exp. 16, 3675<br />

(2008).<br />

[26] Ch. W<strong>and</strong>t et al.; Generation of 220 mJ nanosecond<br />

pulses at a 10 Hz repetition rate with excellent beam<br />

quality in a diode-pumped Yb:YAG MOPA system. Opt.<br />

Lett. 33, 1111 (2008).<br />

[27] S. Naumov et al.; Approaching the microjoule<br />

frontier with femtosecond laser oscillators. New J. Phys.<br />

7, 216 (2005).<br />

[28] A. Ozawa et al.; High harmonic frequency combs<br />

for high resolution spectroscopy. Phy. Rev. Let. 100,<br />

253901 (2008).<br />

[29] E. Goulielmakis et al.; Single-cycle nonlinear optics.<br />

Science 320, 1614 (2008).<br />

[30] J. Gagnon, E. Goulielmakis, V. S. Yakovlev; The<br />

accurate FROG characterization of attosecond pulses<br />

from streaking measurements., Accepted for Publication<br />

in Appl. Phys. B (2008).<br />

[31] V. S. Yakovlev, M. Y. Ivanov, F. Krausz; Enhanced<br />

phase-matching for generation of soft X-ray harmonics<br />

<strong>and</strong> attosecond pulses in atomic gases. Opt. Exp. 15,<br />

No. 23, p. 153511-15364 (12 November 2007).<br />

[32] J. Seres et al.; Coherent superposition of laserdriven<br />

soft-X-ray harmonics from successive sources.<br />

Nature Phys. 3, 878 (2007).<br />

[33] E. Fill, L. Veisz, A. Apolonskiy, F. Krausz; Sub-fs<br />

electron pulses for ultrafast electron diffraction. New J.<br />

Phys. 8, 272 (2006); L.Veisz et al. Hybrid DC-AC electron<br />

gun for fs-electron pulse generation. New J. Phys. 9,<br />

451 (2007).<br />

[34] P. Reckenthaeler et al.; Proposed method for<br />

measuring the duration of electron pulses by attosecond<br />

streaking. Phys. Rev. A 77, 042902 (2008).<br />

[35] M. Centurion et al.; Picosecond electron<br />

deflectometry of optical-field ionized plasmas. Nature<br />

Phot. 2, 315 (2008).<br />

[36] F. Grüner et al.; Design considerations for table-top,<br />

laser-based VUV <strong>and</strong> X-ray free electron lasers, Appl.<br />

Phys. B 86, 431-435<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 147


JUNIOR RESEARCh GROUPS<br />

1.3.1.5 ATTOSECOND DYNAMICS<br />

Leader: Dr. R. Kienberger<br />

The main goal of the Junior Research Group is to<br />

investigate electronic processes in molecules <strong>and</strong> on<br />

solid surfaces on an attosecond timescale. Efforts have<br />

been undertaken to improve the necessary tools, i.e.<br />

phase-stabilized few-cycle laser pulses, ultrashort pulses<br />

in the ultraviolet (UV) <strong>and</strong> vacuum ultraviolet (VUV),<br />

<strong>and</strong> to develop attosecond beamlines for the envisaged<br />

experiments.<br />

IMPROVEMENT OF ThE lASER SYSTEM TO 1.5 CYClE<br />

lASER PUlSES<br />

The natural limit nature sets to the duration of a light<br />

pulse is one oscillation cycle of the carrier wavelength.<br />

The pulse duration is inversely proportional to the<br />

spectral width of the pulse, i.e. the broader the spectrum<br />

the shorter the pulse. In a chirped pulse laser amplifier,<br />

pulses are temporally stretched for the amplification<br />

<strong>and</strong> have to be compressed subsequently. During the<br />

compression, e.g. in prisms, the pulses normally lose<br />

spectral width due to parasitic effects like self-phase<br />

modulation (SPM). To overcome this problem, we<br />

developed a “hybrid” compressor consisting of prisms<br />

<strong>and</strong> additional chirped mirrors which take over the last<br />

part of the compression without the problem of SPM,<br />

leading to amplified laser pulses lasting only 18 fs.<br />

Sending theses pulses into a gas filled hollow core fiber<br />

<strong>and</strong> final compression resulted in 0.4 mJ laser pulses as<br />

short as 3.8 fs – only 1.5 cycles of the carrier wave [5, see<br />

reprint]. This is a new world record <strong>and</strong> made possible<br />

the generation of intense, ultrashort UV pulses.<br />

GENERATION OF UlTRAShORT UV AND VUV PUlSES<br />

Since the separation of quantum states which are<br />

relevant for electronic dynamics in molecules is on the<br />

order of 5 to 10 eV, ultrashort light pulses at this photon<br />

energy have to be generated. The UV/VUV generation<br />

was performed in a gas jet <strong>and</strong> a special “self diffraction<br />

FROG” (Frequency Resolved Optical Gating) device<br />

has been set up for the full characterization of the UV<br />

pulses.<br />

The spectra we were able to generate in the UV are<br />

extraordinarily broad (Figure 1) <strong>and</strong> could be compressed<br />

to 3.7 fs.<br />

ATTOSECOND TEChNOlOGY ON A SOlID SURFACE<br />

In a collaboration with the University of Bielefeld, a first<br />

attosecond experiment on a solid surface was performed<br />

1 . 3 AT T O S E C O N D A N D H I G H - F I E L D D I V I S I O N<br />

[2, see reprint]. We were able to measure a time delay<br />

between conduction b<strong>and</strong> <strong>and</strong> core level electrons in<br />

single-crystal tungsten as short as 110 as.<br />

Figure 1: Ultrabroad UV spectrum supporting 3 fs<br />

pulses<br />

REFERENCES<br />

[1] Remacle F., Kienberger R., Krausz F., Levine, R.D.;<br />

“On the feasibility of an ultrafast purely electronic<br />

reorganization in lithium hydride”, Chemical Physics<br />

338 (2-3): 342-347 ( 2007).<br />

[2] A. L. Cavalieri, N. Müller, Th. Uphues, V. Yakovlev,<br />

A. Baltuska, B. Horvath, B. Schmidt, L. Blümel, S.<br />

Hendel, P. M. Echenique, M. Drescher, U. Kleineberg,<br />

R. Kienberger, F. Krausz, U. Heinzmann; „<strong>Attosecond</strong><br />

time-resolved photoemission from solids”, Nature 449,<br />

1029 (2007).<br />

[3] Kienberger R, Uiberacker M, Kling MF, Krausz F.;<br />

„<strong>Attosecond</strong> physics comes of age: from tracing to<br />

steering electrons at sub-atomic scales” (invited); J.<br />

Mod. Opt. 54 (13-15), 1985-1998 (2007).<br />

[4] E. Goulielmakis, V. Yakovlev, A. L. Cavalieri, M.<br />

Uiberacker, V. Pervak, A. Apolonsky, R. Kienberger,<br />

U. Kleineberg, F. Krausz; “<strong>Attosecond</strong> <strong>Control</strong> <strong>and</strong><br />

<strong>Measurement</strong>: <strong>Lightwave</strong> <strong>Electronics</strong>”, Science 317,<br />

769 (2007).<br />

[5] A. L. Cavalieri, E. Goulielmakis, B. Horvath, W. Helml,<br />

M. Schultze, M. Fieß, V. Pervak, L. Veisz, V. Yakovlev,<br />

M. Uiberacker, A. Apolonski, F. Krausz, R. Kienberger;<br />

“Intense 1.5-cycle near infrared laser waveforms <strong>and</strong><br />

their use for the generation of ultrabroad-b<strong>and</strong> soft-Xray<br />

harmonic continua”, NJP 9, 242 (2007).<br />

[6] MB. Gaarde, M. Murakami, R. Kienberger; „Spatial<br />

separation of large dynamical blueshift <strong>and</strong> harmonic<br />

generation”, Phys. Rev. A 74 (5), 053401 (2006)<br />

[7] A. Scrinzi, M.Yu. Ivanov, R. Kienberger <strong>and</strong> D. M.<br />

Villeneuve; “<strong>Attosecond</strong> Physics” (invited); J. Phys. B: At.<br />

Mol. Opt. Phys. 39, R1 – R37 (2006).<br />

148 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008


1.3.1.6 ATTOSECOND IMAGING<br />

Leader: Dr. Matthias Kling<br />

With waveform-controlled laser fields to control<br />

electronic motion <strong>and</strong> (single) attosecond XUV pulses<br />

to probe electronic motion in real-time two powerful<br />

tools are available to explore electron dynamics in<br />

atoms <strong>and</strong> molecules. We have proposed to extend<br />

the application of these tools to studies of correlated<br />

<strong>and</strong> collective electron motion in nanosystems. One of<br />

our aims was to utilize imaging technologies as e.g.<br />

velocity-map imaging (VMI) to monitor the ultrafast<br />

electron dynamics <strong>and</strong> its control. As a demonstration<br />

that VMI can indeed be used to monitor the waveformcontrol<br />

of electron emission we have investigated<br />

atomic systems, where the obtained results have been<br />

compared to theoretical predictions (by e.g. solving<br />

the time-dependent Schrödinger equation (TDSE)).<br />

This control is applied to molecular systems, where the<br />

light-steering of binding electrons is used to coherently<br />

control a reaction. As a highlight, the waveform control<br />

of electron localization in molecular dissociation was<br />

time-resolved by an attosecond pump-probe experiment<br />

for the first time.<br />

The full spatio-temporal characterization of nanometerscale<br />

collective electron dynamics with attosecond<br />

temporal resolution was studied theoretically within<br />

the collaboration with Prof. Mark Stockman (GSU,<br />

Atlanta, USA). The original idea for studies on isolated<br />

nanoparticles was extended to photoelectron emission<br />

microscopy (PEEM) of nanosystems on surfaces, leading<br />

to a new technique dubbed “attosecond nanoplasmonic<br />

microscope”. Prof. Ulf Kleineberg (LMU) <strong>and</strong> his<br />

coworkers are currently building a UHV experiment<br />

for time-resolved Photoelectron Emission Microscopy,<br />

which will be capable for the first time to track electron<br />

dynamics in nanostructures on surfaces on a ~100<br />

fsec time scale <strong>and</strong> with 10 nm spacial resolution. This<br />

effort complements the development of the attosecond<br />

beamline AS 3.<br />

IMAGING OF SUB-FEMTOSECOND CONTROl OF<br />

ElECTRON DYNAMICS<br />

<strong>Control</strong> of the electric field waveform of laser light<br />

pulses has opened a new way to study <strong>and</strong> control<br />

electron dynamics in strong-field processes. We have<br />

utilized velocity-map imaging to study the angleresolved<br />

electron emission from Ar, Kr <strong>and</strong> Xe atoms<br />

with waveform-controlled few-cycle pulses. The data<br />

for Ar is compared to simulations using the strongfield<br />

approximation (SFA) <strong>and</strong> full time-dependent<br />

Schrödinger equation (TDSE) calculations. We find a<br />

pronounced asymmetry in both the energy <strong>and</strong> angular<br />

distributions of the electron emission that critically<br />

1 . 3 . 1 S U M M A R Y O F S C I E N T I F I C A C T I V I T I E S<br />

depends on the carrier-envelope phase of the laser field.<br />

The generally weak asymmetry of low-energy electrons<br />

can be explained by their deflection in the oscillating<br />

laser field that cancels the initial non-linear dependence<br />

of the ionization yield on the field strength. In the<br />

plateau region of the argon data, the asymmetry is<br />

larger <strong>and</strong> the phase of its oscillation with CEP shifts<br />

not only with the electron kinetic energy but also with<br />

the emission angle. A triangular pattern suggests that<br />

the observation of the asymmetry at a constant energy<br />

but two different emission angles can serve as a way to<br />

measure the phase CEP of the laser pulse without ±π<br />

ambiguity. In the case of xenon, significant asymmetry<br />

was observed even below 2 Up in particular for high<br />

emission angles around 60 degrees. This unique<br />

dependence indicates that VMI can also be used instead<br />

of Stereo-ATI detection to determine the CEP with high<br />

accuracy.<br />

A B<br />

C<br />

Figure 1: A) <strong>and</strong> B) show raw two-dimensional<br />

photoelectron images for above-threshold ionization<br />

of argon with 5fs laser pulses for phases of ϕ = 0<br />

<strong>and</strong> ϕ = π, measured by velocity map imaging. As<br />

seen in C), where the asymmetry of the emission is<br />

plotted (red = upwards, blue = downwards), the<br />

emission can be controlled with the laser phase.<br />

A pronounced shift in the phase of the asymmetry<br />

oscillation that is seen in the experimental data is<br />

reproduced well beyond 5 Up by both SFA <strong>and</strong> TDSE<br />

calculations. While a phase jump is visible for the SFA<br />

calculated asymmetries at around 5 Up, the difference<br />

between the TDSE calculated values <strong>and</strong> the experimental<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 149


data stays consistently below 0.1 rad <strong>and</strong> thus shows<br />

excellent agreement all the way down to ca. 2 U p (below<br />

this value, noise is dominating the curves due to the low<br />

amplitude of the asymmetry). We have demonstrated<br />

that velocity-map imaging is not only a powerful tool<br />

for measuring the full momentum distributions of ATI<br />

in rare gases but also provides a new way to determine<br />

the CEP of few-cycle laser pulses from the angular<br />

distribution of the emitted electrons.<br />

WAVEFORM CONTROl OF ElECTRON lOCAlIZATION<br />

IN MOlECUlES<br />

A previously explored single-color control scheme for<br />

attosecond electron localization in molecules is difficult<br />

to extend to systems with dissociation times much longer<br />

than the duration of the waveform-controlled fewcycle<br />

laser field (in previous experiments, 5 fs pulses at<br />

760 nm were used) as the field strength of the laser will<br />

be to low to coherently couple two electronic states<br />

close to the break-up of the molecule.<br />

As a demonstration for a suitable way to extend the<br />

control scheme, in collaboration with the groups of Anne<br />

L’Huillier (Lund), Marc Vrakking (Amolf), Franck Lepine<br />

(Lyon) <strong>and</strong> Mauro Nisoli (Milan), we have performed<br />

a two-color experiment, where an isolated attosecond<br />

laser pulse ionizes D <strong>and</strong> part of the ionized molecules<br />

2<br />

dissociate. The single attosecond pulse is generated by<br />

means of high-harmonic generation in Krypton using the<br />

polarization gating technique. In the photo-ionization<br />

+ by the attosecond pulse the repulsive 2pσ state is<br />

u<br />

populated when the photon-energy of the pulse is high<br />

enough (above ca. 29 eV). Interaction of this dissociating<br />

wave packet with a time delayed moderately strong IR<br />

field localizes the electron on the upper or lower D + ion.<br />

By varying the delay between the XUV pulse <strong>and</strong> the<br />

few-cycle IR pulse the asymmetry in the ejection of D +<br />

ions can be controlled with attosecond time resolution.<br />

This experiment may be viewed as a first example of<br />

the observation of attosecond time-resolved electron<br />

dynamics in molecular physics. As seen in Figure 2 the<br />

asymmetry shows a strong dependence on the kinetic<br />

energy of the D + fragment.<br />

A model solving the 1D time-dependent Schrödinger<br />

equation (TDSE) was used to confirm the interpretation<br />

<strong>and</strong> to give insight into the origin of the observed<br />

energy dependence. In the model a nuclear wave<br />

+ packet is projected on the repulsive 2pσ state <strong>and</strong><br />

u<br />

+ propagated on the bound 1sσ <strong>and</strong> the repulsive<br />

g<br />

+ 2pσ state of the molecular ion in the presence of a<br />

u<br />

few-cycle IR field. The IR field couples the two states<br />

which generates a coherent superposition where the<br />

electron can be localized on one or the other ion. Just as<br />

in the experiment the timing of the IR field determines<br />

the electron localization. The kinetic energy dependence<br />

1 . 3 AT T O S E C O N D A N D H I G H - F I E L D D I V I S I O N<br />

Figure 2: A) D + ion kinetic energy spectra from the<br />

dissociative ionization of D 2 versus the delay between<br />

the 300 as XUV pulse at 30 eV <strong>and</strong> a phase-stable<br />

6-fs IR pulse; B) Asymmetry of the D + ion emission<br />

along the laser polarization as a function of the D +<br />

kinetic energy <strong>and</strong> the delay between the XUV <strong>and</strong><br />

IR pulses.<br />

of the electron localization, which is observed in the<br />

experiment <strong>and</strong> the calculations, likely, originates from<br />

the initial spread of the wave packet.<br />

Electron transfer processes are ubiquitous in chemistry.<br />

The present combination of one- <strong>and</strong> two-color<br />

experiments, where electron localization is first<br />

controlled on attosecond timescales using the controlled<br />

waveform of a few-cycle IR laser pulse <strong>and</strong> then using<br />

the controlled delay between an isolated attosecond<br />

pulse <strong>and</strong> a few-cycle IR laser pulse, are first examples of<br />

strong-field control of chemical processes on attosecond<br />

timescales <strong>and</strong> of the direct observation of attosecond<br />

electron dynamics in molecules, paving the way towards<br />

attempts to observe <strong>and</strong> control electron dynamics in<br />

more complicated molecules <strong>and</strong> nanostructures. The<br />

results of these studies are currently written up.<br />

ATTOSECOND NANOPlASMONIC MICROSCOPY<br />

Nanoplasmonics deals with collective electronic<br />

dynamics on the surface of metal nanostructures, which<br />

arises as a result of excitations called surface plasmons.<br />

150 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008


This field, which has recently undergone rapid growth,<br />

could benefit applications such as computing <strong>and</strong><br />

information storage on the nanoscale, the ultrasensitive<br />

detection <strong>and</strong> spectroscopy of physical, chemical <strong>and</strong><br />

biological nanosized objects, <strong>and</strong> the development of<br />

optoelectronic devices. Because of their broad spectral<br />

b<strong>and</strong>width, surface plasmons undergo ultrafast<br />

dynamics with timescales as short as a few hundred<br />

attoseconds. So far, the spatiotemporal dynamics of<br />

optical fields localized on the nanoscale has been hidden<br />

from direct access in the real space <strong>and</strong> time domain.<br />

In collaboration with Prof. Mark Stockman we have<br />

proposed <strong>and</strong> developed a theory for an attosecond<br />

nanoplasmonic-field microscope. The main advantage<br />

of such a microscope is that it is non-invasive with<br />

respect to the nanoplasmonic fields. The principle of this<br />

microscope is based on the photoemission of electrons<br />

by an XUV attosecond pulse that is synchronized with<br />

a waveform-stabilized driving optical field. Information<br />

about the nanoplasmonic fields is imprinted in the<br />

energy of the XUV-emitted electrons owing to their<br />

acceleration in the instantaneous electrostatic potential<br />

of the surface plasmon oscillations excited by the optical<br />

field.<br />

The attosecond nanoplasmonic microscope will open up<br />

unique possibilities to directly study <strong>and</strong> control ultrafast<br />

photoprocesses in surface plasmonic nanosystems <strong>and</strong><br />

circuits. It images the local nanoplasmonic field in real<br />

space with nanometre-scale spatial resolution <strong>and</strong><br />

in real time with approx. 100 as temporal resolution.<br />

This approach will be especially important in ultrafast<br />

nanoplasmonic systems where very tight localization<br />

of optical fields occurs, for example, when a shaped<br />

pulse of radiation induces nanolocalized fields at a<br />

desired nanosite. The microscope could also be used<br />

to study various plasmon enhanced photoprocesses<br />

such as femtosecond photochemistry, light detection,<br />

<strong>and</strong> solar energy conversion where the processes of<br />

the energy transfer can be ultrafast. For example,<br />

with nanoplasmonic antennas, which couple together<br />

molecular <strong>and</strong> semiconductor systems with external<br />

fields, the ultrafast kinetics of the energy exchange<br />

can be studied through the corresponding kinetics of<br />

the nanoplasmonic fields. One of the most important<br />

potential uses of the attosecond nanoplasmonicfield<br />

microscope will be in the design <strong>and</strong> study of<br />

elements <strong>and</strong> devices for ultrafast <strong>and</strong> ultradense (at<br />

the nanoscale) optical <strong>and</strong> optoelectronic information<br />

processing <strong>and</strong> storage. To bring practical advantages<br />

over existing electronic <strong>and</strong> optoelectronic technology,<br />

such nanoscale devices must necessarily be ultrafast<br />

(with a subpicosecond or femtosecond response time).<br />

Examples of such devices include optical nanotransistors,<br />

memory cells, nanoplasmonic media for the mass storage<br />

of information, <strong>and</strong> nanoplasmonic interconnects<br />

1 . 3 . 1 S U M M A R Y O F S C I E N T I F I C A C T I V I T I E S<br />

where both the nanolocalization of energy <strong>and</strong> its<br />

ultrafast temporal kinetics are important. Currently, in<br />

collaboration with Prof. Ulf Kleineberg (LMU) we are<br />

preparing the setup <strong>and</strong> samples for first PEEM studies.<br />

Figure 3: After excitation of the collective electron<br />

motions (plasmons) in a nanoparticle assembly by an<br />

optical light pulse, the resulting plasmon dynamics<br />

can be probed with high spatial <strong>and</strong> temporal<br />

resolution by the generation of photoelectrons und<br />

their acceleration in the plasmonic fields. This figure<br />

shows calculated snapshots of the plasmonic fields<br />

at different time delays after excitation. The spatial<br />

resolution is given by the limit of the optics of the<br />

photoemission electron microscope (PEEM) that is<br />

used for the detection of the electrons (approx. 20 nm).<br />

The time resolution (approx. 100 attoseconds) is<br />

given by the duration of the XUV probe pulse. These<br />

theoretical calculations show for the first time, that<br />

the direct <strong>and</strong> non-invasive observation of rapidly<br />

evolving nanoplasmonic fields becomes possible (see<br />

the two snap shots in the lower part of the figure,<br />

which show the fields within a time interval of 1 fs<br />

(= 10 -15 s)).<br />

REFERENCES<br />

[1] M.F. Kling, <strong>and</strong> M.J.J. Vrakking; <strong>Attosecond</strong> Electron<br />

Dynamics, Annu. Rev. Phys. Chem. 59, 463-492<br />

(2008).<br />

[2] M.F. Kling, Ch. Siedschlag, I. Znakovskaya, A.J.<br />

Verhoef, S. Zherebtsov, F. Krausz, M. Lezius, <strong>and</strong> M.J.J.<br />

Vrakking; Strong-field control of electron localization<br />

during molecular dissociation, Mol.Phys. 106, 455-465<br />

(2008).<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 151


[3] M.F. Kling, J. Rauschenberger, A.J. Verhoef, E.<br />

Hasovic, T. Uphues, D.B. Milosevic, H.G. Muller, <strong>and</strong><br />

M.J.J. Vrakking; Imaging of carrier-envelope phase<br />

effects in above-threshold ionization with intense fewcycle<br />

laser fields, New J. Phys. 10, 025024 (2008).<br />

[4] Th. Uphues, M. Schultze, M.F. Kling, M. Uiberacker,<br />

S. Hendel, U. Heinzmann, N.M. Kabaschnik, <strong>and</strong> M.<br />

Drescher; Ion-charge-state chronoscopy of cascaded<br />

atomic Auger decay, New J. Phys. 10, 025009 (2008).<br />

[5] A. Gijsbertsen, W. Siu, M.F. Kling, P. Johnsson,<br />

P. Jansen, S. Stolte, <strong>and</strong> M.J.J. Vrakking; Direct<br />

Determination of the Sign of the NO Dipole Moment,<br />

Phys.Rev.Lett. 99, 213003 (2007)<br />

[6] R. Kienberger, M. Uiberacker, M.F. Kling, <strong>and</strong> F.<br />

Krausz; <strong>Attosecond</strong> physics comes of age: from tracing<br />

to steering electrons at sub-atomic scales, J.Mod.Opt.<br />

54, 1985-1998 (2007)<br />

[7] M.I. Stockman, M.F. Kling, U. Kleineberg, <strong>and</strong> F.<br />

Krausz; <strong>Attosecond</strong> nanoplasmonic field microscope,<br />

Nature Photonics 1, 539 (2007).<br />

[8] F. Lepine, M.F. Kling, Y. Ni, J. Khan, O. Ghafur, T.<br />

Martchenko, E. Gustafsson, P. Johnsson, K. Varju, T.<br />

Remetter, A. L’Huillier, M.J.J. Vrakking; Short XUV pulses<br />

to characterize field-free molecular alignment, J. Mod.<br />

Opt. 54, 953 (2007).<br />

[9] M. Uiberacker, Th. Uphues, M. Schultze, A.J.<br />

Verhoef, V. Yakovlev, M.F. Kling, J. Rauschenberger,<br />

N.M. Kabaschnik, H. Schröder, M. Lezius, K.L. Kompa,<br />

H.-G. Muller, M.J.J. Vrakking, S. Hendel, U. Kleineberg,<br />

U. Heinzmann, M. Drescher, F. Krausz; <strong>Attosecond</strong> realtime<br />

observation of electron tunneling in atoms, Nature<br />

446, 627-632 (2007)<br />

1 . 3 AT T O S E C O N D A N D H I G H - F I E L D D I V I S I O N<br />

152 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008


1.3.2 SURVEY OF ThE RESEARCh ACTIVITIES<br />

1 . 3 . 2 S U R V E Y O F T H E R E S E A R C H A C T I V I T I E S<br />

Pushing the frontiers of femtosecond technology<br />

Project coordinators: A. Apolonskiy, R. Kienberger, S. Karsch, L. Veisz<br />

Project Objectives Team<br />

Chirped multilayer<br />

dielectric mirrors<br />

Generation of terawattscale<br />

near-monocycle laser<br />

pulses with controlled<br />

waveform at several kHz<br />

repetition rate in the near<br />

infrared (LWS-N1) <strong>and</strong> in<br />

the mid-infrared (LWS-M1)<br />

Development of a highfield<br />

waveform synthesizer<br />

producing few-cycle nearinfrared<br />

laser pulses with a<br />

peak power of several 10<br />

terawatt <strong>and</strong> controlled<br />

waveform at a 10 Hz<br />

repetition rate (LWS-10)<br />

Development of the<br />

Petawatt Field Synthesizer<br />

(PFS)<br />

Design, manufacturing & characterization of broadb<strong>and</strong><br />

mirrors for optical waveform synthesis in the<br />

visible, infrared <strong>and</strong> ultraviolet spectral range<br />

LWS-N1: 4fs/2mJ/0.5TW@0.75μm&4kHz<br />

with Ti:sapphire MOPA + hollow-fibre/chirped-mirror<br />

compressor<br />

LWS-M1:<br />

12fs/10mJ/1TW@2.1μm&10kHz<br />

by parametric amplification pumped of by a sub-2ps,<br />

50-mJ, diode-pumped Yb:YAG disk laser<br />

Current status:<br />

8fs/70mJ/10TW@0.8μm&10Hz<br />

Target:<br />

5fs/600mJ/100TW@0.8μm&10Hz<br />

using an optical chirped-pulse amplifier chain<br />

pumped by a commercial picosecond Nd:YAG laser<br />

Target parameters:<br />

3J/~1PW@1.2μm&10Hz<br />

V. Pervak (PL)<br />

I. Grguras<br />

A. Apolonskiy<br />

External collaborators:<br />

A. V. Tikhonravov (Moscow)<br />

M. K. Trubetskov (Moscow)<br />

A. Cavalieri (PL: N1)<br />

X. Gu (PL: M1)<br />

Y. Deng<br />

W. Helml<br />

G. Marcus<br />

T. Metzger<br />

B. Horváth<br />

J. Li<br />

E. Magerl,<br />

R. Kienberger<br />

External collaborators:<br />

H. Giesen (Stuttgart)<br />

D. Sutter (Trumpf GmbH)<br />

L. Veisz<br />

A. Buck<br />

D. Herrmann<br />

K. Schmid<br />

F. Tavella<br />

T. Wittman<br />

A. Marcinkevicius (until<br />

12/2006)<br />

Zs. Major (PL)<br />

M. Siebold (PL)<br />

S. Trushin (PL)<br />

I. Ahmad<br />

S. Klingebiel<br />

Ch. W<strong>and</strong>t<br />

T.-J. Wang<br />

S. Karsch<br />

External collaborators:<br />

J. Hein (Jena)<br />

A. Tünnermann (Jena)<br />

PL = Project leader<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 153


1 . 3 AT T O S E C O N D A N D H I G H - F I E L D O H Y S I C S D I V I S I O N<br />

Pushing the frontiers of femtosecond technology (continued)<br />

Project coordinators: A. Apolonskiy, R. Kienberger, S. Karsch, L. Veisz<br />

Project Objectives Team<br />

Scaling femtosecond<br />

tech-nology towards<br />

multi-kW intra-cavity<br />

average power levels <strong>and</strong><br />

development of MHzrate<br />

XUV sources with<br />

milliwatt-scale average<br />

power<br />

Generation, measurement<br />

<strong>and</strong> applications of few-fs<br />

sub-relativistic electron<br />

bunches<br />

Millijoule-energy femtosecond laser pulses inside<br />

solid-state femtosecond laser oscillators <strong>and</strong> passive<br />

build-up cavities for intracavity<br />

production of UV/VUV/XUV/SXR light<br />

Several-10-keV, few-electron-bunches produced at<br />

MHz repetition rates in synchrony with MHz-rate,<br />

microjoule-energy few-cycle laser pulses for scaling<br />

time-resolved electron diffraction towards the 1femtosecond<br />

frontier<br />

lightwave electronics: attosecond control <strong>and</strong> metrology<br />

Project coordinators: R. Kienberger, U. Kleineberg, M. Kling<br />

Projects Objectives Team<br />

Chirped multilayer metallic<br />

mirrors<br />

<strong>Attosecond</strong> pulse<br />

generation from atomic<br />

harmonics<br />

Design, manufacturing & characterization of<br />

broadb<strong>and</strong> mirrors for attosecond XUV/SXR pulse<br />

technology (10 eV – 1000 eV)<br />

Scaling towards microjoule pulse energies or higher<br />

(several hundred to thous<strong>and</strong> electronvolt) photon<br />

energies by using LWS-10 <strong>and</strong> LWS-N1/LWS-M1<br />

as a driver, respectively, <strong>and</strong> exploiting quasi-phase<br />

matching schemes<br />

154 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008<br />

J. Rauschenberger (PL)<br />

R. Graf<br />

J. Pupeza<br />

A. Sugita<br />

C. Teisset<br />

A. Apolonskiy<br />

External collaborators:<br />

T. Udem & T. Hänsch (MPQ)<br />

D. Hoffmann (Aachen)<br />

A. Tünnermann (Jena)<br />

E. Fill (PL: Technology)<br />

P. Baum (PL: Applications)<br />

M. Centurion<br />

P. Reckenthäler<br />

L. Veisz<br />

A. Apolonskiy<br />

External collaborators:<br />

V. Tarnetsky (Novosibirsk)<br />

A. Zewail (Pasadena)<br />

U. Kleineberg (PL)<br />

M. Hofstetter<br />

J. Lin<br />

External collaborators:<br />

A. L. Aquila<br />

E. M. Gullikson &<br />

D. T. Attwood (Berkeley)<br />

G. Marcus (PL: LWS-10, M1)<br />

A. Cavalieri (PL: LWS-N1)<br />

D. Herrmann<br />

M. Hofstetter<br />

U. Kleineberg<br />

V. Yakovlev (theory)<br />

L. Veisz<br />

R. Kienberger<br />

External collaborators:<br />

D. Charalambidis (Heraklion)


1 . 3 . 2 S U R V E Y O F T H E R E S E A R C H A C T I V I T I E S<br />

lightwave electronics: attosecond control <strong>and</strong> metrology (continued)<br />

Project coordinators: R. Kienberger, U. Kleineberg, M. Kling<br />

Projects Objectives Team<br />

Towards ultrawide-b<strong>and</strong><br />

waveform synthesis <strong>and</strong><br />

mono-cycle UV/VUV pulse<br />

generation<br />

<strong>Attosecond</strong> spectroscopy<br />

in isolated atoms <strong>and</strong><br />

nanoparticles<br />

<strong>Attosecond</strong> spectroscopy<br />

in molecules <strong>and</strong> solids<br />

Towards solid-state<br />

lightwave electronics<br />

Sub-cycle shaping of the electromagnetic field of<br />

light by Fourier-synthesis of intense (sub-terawatt)<br />

arbitrary waveforms in the NIR/VIS/UV spectral range<br />

& few-fs UV/VUV pulses<br />

Real-time observation of electron dynamics <strong>and</strong><br />

correlations in atoms <strong>and</strong> collective motion in<br />

nanoparticles by using isolated attosecond XUV<br />

pulses <strong>and</strong> few-cycle NIR waveforms for pump/<br />

probe spectroscopy <strong>and</strong> sophisticated electron <strong>and</strong><br />

ion detection techniques (including coincidence<br />

detection)<br />

Real-time observation of electron motion (individual<br />

as well as collective) in molecules (in the gas phase<br />

<strong>and</strong> on surfaces) <strong>and</strong> in solid-state structures by<br />

using isolated attosecond XUV pulses <strong>and</strong> few-cycle<br />

NIR/UV waveforms for pump/probe spectroscopy,<br />

also in combination with photo-electron emission<br />

microscopy (PEEM)<br />

Demonstration of controlling atomic-scale electron<br />

motion in solids with light fields<br />

E. Goulielmakis (PL)<br />

R. Ernstorfer<br />

U. Graf<br />

I. Grguras<br />

V. Pervak,<br />

A. Apolonskiy<br />

R. Kienberger<br />

M. Kling (PL)<br />

E. Goulielmakis<br />

O. Herrwerth<br />

U. Kleineberg<br />

A. Wirth<br />

S. Zherebtsov<br />

I. Znakovskaya<br />

External collaborators:<br />

K. Kompa<br />

M. Lezius &<br />

H. Schröder (MPQ)<br />

M. Vrakking (Amsterdam)<br />

R. Moshammer &<br />

J. Ullrich (Heidelberg)<br />

A. Cavalieri (PL: solids)<br />

M. Schultze (PL: molecules)<br />

J. Lin (PEEM)<br />

R. Ernstorfer<br />

M. Fieß<br />

E. Magerl<br />

R. Kienberger<br />

U. Kleineberg<br />

External collaborators:<br />

J. Barth<br />

P. Feulner &<br />

D. Menzel (TUM)<br />

K. Kompa &.<br />

M. Lezius (MPQ)<br />

R. Levine (Jerusalem)<br />

F. Remacle (Liège)<br />

R. Ernstorfer (PL)<br />

U. Graf<br />

E. Goulielmakis<br />

R. Kienberger<br />

External collaborators:<br />

J. Barth<br />

P. Feulner &<br />

D. Menzel (TUM)<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 155


1 . 3 AT T O S E C O N D A N D H I G H - F I E L D O H Y S I C S D I V I S I O N<br />

high-field attosecond science<br />

Project coordinators: F. Grüner, D. Habs, S. Karsch, G. Tsakiris, L. Veisz<br />

Projects Objectives Team<br />

Few-fs to sub-fs relativistic<br />

electron bunches<br />

Laser-plasma acceleration<br />

of electrons to superrelativistic<br />

(>>100 MeV)<br />

energies<br />

Reproducible generation of mono-energetic relativistic<br />

(1-100MeV) few-fs electron bunches in underdense<br />

plasmas driven by few-cycle laser pulses from LWS-10,<br />

sub-femtosecond bunch slicing by the inverse freeelectron<br />

laser mechanism, temporal characterization<br />

<strong>and</strong> applications for time-resolved experiments<br />

Reproducible generation of collimated, high-current,<br />

mono-energetic electron bunches of energies<br />

ranging from 100 MeV to several GeV by means of<br />

laser wake field acceleration (LWFA) for producing<br />

undulator radiation <strong>and</strong> free-electron lasing, <strong>and</strong><br />

other applications<br />

156 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008<br />

K. Schmid (few-fs)<br />

C. Sears (sub-fs)<br />

S. Becker<br />

A. Buck<br />

F. Grüner<br />

S. Karsch<br />

Y. Mikhailova (theory)<br />

D. Habs<br />

D. Herrmann<br />

R. Tautz<br />

L. Veisz<br />

External collaborators:<br />

W. Leemans (Berkeley)<br />

S. Karsch<br />

S. Becker<br />

M. Fuchs<br />

F. Grüner<br />

D. Habs<br />

Zs. Major<br />

B. Marx<br />

J. Osterhoff<br />

A. Popp,<br />

R. Weingartner<br />

M. Fuchs<br />

L. Veisz<br />

External collaborators:<br />

M. Geissler (Belfast)<br />

S. Hooker (Oxford)<br />

W. Leemans (Berkeley)


1 . 3 . 2 S U R V E Y O F T H E R E S E A R C H A C T I V I T I E S<br />

high-field attosecond science (continued)<br />

Project coordinators: F. Grüner, D. Habs, S. Karsch, G. Tsakiris, L. Veisz<br />

Project Objectives Team<br />

Towards a laser-based,<br />

compact X-ray freeelectron<br />

laser (XFEL)<br />

Towards intense<br />

attosecond XUV pulses<br />

Ion <strong>and</strong> X-ray beams from<br />

thin foils<br />

Optimization of LWFA <strong>and</strong> preparation of the<br />

accelerated electron beam for undulator seeding,<br />

demonstration of undulator radiation <strong>and</strong> FEL<br />

operation at XUV wavelengths <strong>and</strong> subsequently at<br />

successively shorter wavelengths<br />

Generation of intense single attosecond XUV pulses<br />

from few-cycle-driven relativistic plasmas at solid<br />

surfaces, their characterization <strong>and</strong> applications for<br />

XUV nonlinear optics <strong>and</strong> XUV-pump/XUV-probe<br />

spectroscopy<br />

Generation of mono-energetic ion beams from ultrathin<br />

diamond foils driven by ~ 100 TW laser pulses<br />

(from ATLAS) <strong>and</strong> of coherent X-rays by Thomson<br />

scattering from electron sheaths driven out of the foil<br />

by multi-TW laser pulses (from LWS-10)<br />

F. Grüner<br />

S. Becker<br />

M. Fuchs<br />

R. Weingartner<br />

D. Habs<br />

S. Karsch<br />

Zs. Major<br />

B. Marx<br />

J. Osterhoff<br />

A. Popp<br />

L. Veisz<br />

External collaborators:<br />

M. Geissler (Belfast)<br />

S. Hooker (Oxford)<br />

W. Leemans (Berkeley)<br />

G. Tsakiris<br />

A. Buck<br />

R. Hörlein<br />

S. Karsch<br />

Z. Major<br />

Y. Mikhailova<br />

Y. Nomura<br />

S. Rykovanov<br />

L. Veisz<br />

External collaborators:<br />

D. Charalambidis (Heraklion)<br />

M. Zepf (QU Belfast)<br />

M. Geissler (Qu Belfast)<br />

I. Foeldes (KFKI-Hungary)<br />

D. Habs<br />

A. Buck<br />

L. Veisz<br />

External collaborators:<br />

M. Hegelich (Los Alamos)<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 157


1.3.3 SElECTED REPRINTS<br />

1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

1) Design considerations for table-top, laser-based VUV<br />

<strong>and</strong> X-ray free electron lasers<br />

F. Grüner, S. Becker, U. Schramm, T. Eichner, M. Fuchs,<br />

R. Weingartner, D. Habs, J. Meyer-ter-Vehn, M. Geissler,<br />

M. Ferrario, L. Serafini, B. van der Geer, H. Backe, W.<br />

Lauth, S. Reiche,<br />

Applied Physics B 86, 431 (2007).<br />

MPQ Progress Report: page 160<br />

2) <strong>Attosecond</strong> real-time observation of electron tunnelling<br />

in atoms<br />

M. Uiberacker, Th. Uphues, M. Schultze, A. J. Verhoef,<br />

V. Yakovlev, M. F. Kling, J. Rauschenberger, N. M. Kabachnik,<br />

H. Schröder, M. Lezius, K. L. Kompa, H.-G.<br />

Muller, M. J. J. Vrakking, S. Hendel, U. Kleineberg, U.<br />

Heinzmann, M. Drescher, F. Krausz,<br />

Nature 446, 627 (2007).<br />

MPQ Progress Report: page 165<br />

3) Dispersion-control over the ultraviolet-visible-near-infrared<br />

spectral range with HfO 2 /SiO 2- chirped dielectric<br />

multilayers<br />

V. Pervak, F. Krausz, A. Apolonski,<br />

Optics Letters 32, 1183 (2007).<br />

MPQ Progress Report: page 171*<br />

4) Dispersion management for a sub-10-fs, 10-TW optical<br />

parametric chirped-pulse amplifier<br />

F. Tavella, Y. Nomura, L. Veisz, V. Pervak, A. Marcinkevicius,<br />

F. Krausz,<br />

Optics Letters 32, 2227 (2007).<br />

MPQ Progress Report: page 172*<br />

5) <strong>Attosecond</strong> control <strong>and</strong> measurement: lightwave<br />

electronics<br />

E. Goulielmakis, V. S. Yakovlev, A. L. Cavalieri, M. Uiberacker,<br />

V. Pervak, A. Apolonski, R. Kienberger, U. Kleineberg,<br />

F. Krausz,<br />

Science 317, 769 (2007).<br />

MPQ Progress Report: page 173<br />

6) Coherent superposition of laser-driven soft-X-ray harmonics<br />

from successive sources<br />

J. Seres, V. S. Yakovlev, E. Seres, Ch. Streli, P. Wobrauschek,<br />

Ch. Spielmann, F. Krausz,<br />

Nature Physics 3, 878 (2007).<br />

MPQ Progress Report: page 180*<br />

7) Enhanced phase-matching for generation of soft Xray<br />

harmonics <strong>and</strong> attosecond pulses in atomic gases<br />

V. S. Yakovlev, M. Ivanov, F. Krausz,<br />

Optics Express 15, 15351 (2007).<br />

Downloadable from http://www.opticsexpress.org/issue.cfm?volume=15&issue=23<br />

MPQ Progress Report: page 181*<br />

8) GeV-scale electron acceleration in a gas-filled capillary<br />

discharge waveguide<br />

S. Karsch, J. Osterhoff, A. Popp, T. P. Rowl<strong>and</strong>s-Rees, Zs.<br />

Major, M. Fuchs, B. Marx, R. Hörlein, K. Schmid, L. Veisz,<br />

S. Becker, U. Schramm, B. Hidding, G. Pretzler, D. Habs,<br />

F. Grüner, F. Krausz, S. M. Hooker,<br />

New Journal of Physics 9, 415 (2007).<br />

Downloadable from http://www.iop.org/EJ/abstract/1367-2630/9/11/415<br />

MPQ Progress Report: page 182*<br />

9) Hybrid dc–ac electron gun for fs-electron pulse generation<br />

L. Veisz, G. Kurkin, K. Chernov, V. Tarnetsky, A. Apolonski,<br />

F Krausz, E. Fill,<br />

New Journal of Physics 9, 451 (2007).<br />

Downloadable from http://www.iop.org/EJ/abstract/1367-2630/9/12/451<br />

MPQ Progress Report: page 183*<br />

10) Picosecond electron deflectometry of optical-field<br />

ionized plasmas<br />

M. Centurion, P. Reckenthaeler, S. A. Trushin, F. Krausz,<br />

E. E. Fill,<br />

Nature Photonics 2, 315 (2008).<br />

MPQ Progress Report: page 184*<br />

11) Single-cycle nonlinear optics<br />

E. Goulielmakis, M. Schultze, M. Hofstetter, V. S. Yakovlev,<br />

J. Gagnon, M. Uiberacker, A. L. Aquila, E. M.<br />

Gullikson, D. T. Attwood, R. Kienberger, F. Krausz, U.<br />

Kleineberg,<br />

Science 320, 1614 (2008)<br />

MPQ Progress Report: page 185<br />

12) Intense single attosecond pulses from surface harmonics<br />

using the polarization gating technique<br />

S G Rykovanov, M Geissler, J Meyer-ter-Vehn <strong>and</strong><br />

G D Tsakiris<br />

New Journal of Physics 10, 025025 (2007).<br />

Downloadable from http://www.iop.org/EJ/abstract/1367-2630/10/2/025025<br />

MPQ Progress Report: page 189*<br />

158 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008<br />

*) 1st page


JUNIOR RESEARCh GROUPS<br />

Dr. R. Kienberger<br />

1) Intense 1.5-cycle near infrared laser waveforms <strong>and</strong><br />

their use for the generation of ultra-broadb<strong>and</strong> soft-xray<br />

harmonic continua<br />

A. L. Cavalieri, E. Goulielmakis, B. Horvath, W. Helml, M.<br />

Schultze, M. Fieß, V. Pervak, L. Veisz, V. S. Yakovlev, M.<br />

Uiberacker, A. Apolonski, F. Krausz, R. Kienberger,<br />

New Journal of Physics 9, 242 (2007).<br />

Downloadable from http://www.iop.org/Select/abstract/1367-2630/9/7/242<br />

MPQ Progress Report: page 190*<br />

2) <strong>Attosecond</strong> spectroscopy in condensed matter<br />

A. L. Cavalieri, N. Müller, Th. Uphues, V. S. Yakovlev, A.<br />

Baltuska, B. Horvath, B. Schmidt, L. Blümel, R. Holzwarth,<br />

S. Hendel, M. Drescher, U. Kleineberg, P. M. Echenique,<br />

R. Kienberger, F. Krausz, U. Heinzmann,<br />

Nature 449, 1029 (2007).<br />

MPQ Progress Report: page 191<br />

Dr. M. Kling<br />

1) <strong>Attosecond</strong> nanoplasmonic-field microscope<br />

M. I. Stockman, M. F. Kling, U. Kleineberg, F. Krausz,<br />

Nature Photonics 1, 539 (2007).<br />

MPQ Progress Report: page 195<br />

1 . 3 . 3 S E L E C T E D R E P R I N T S<br />

*) 1st page<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 159


Appl. Phys. B 86, 431–435 (2007)<br />

DOI: 10.1007/s00340-006-2565-7<br />

f. grüner 1,�<br />

s. becker 1<br />

u. schramm 1<br />

t. eichner 1<br />

m. fuchs 1<br />

r. weingartner 1<br />

d. habs 1<br />

j. meyer-ter-vehn 2<br />

m. geissler 2<br />

m. ferrario 3<br />

l. serafini 3<br />

b. van der geer 4<br />

h. backe 5<br />

w. lauth 5<br />

s. reiche 6<br />

1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

Received: 12 December 2006<br />

Published online: 27 January 2007 • © Springer-Verlag 2007<br />

Applied Physics B<br />

Lasers <strong>and</strong> Optics<br />

Design considerations for table-top,<br />

laser-based VUV <strong>and</strong> X-ray free<br />

electron lasers<br />

ABSTRACT A recent breakthrough in laser-plasma accelerators, based upon ultrashort<br />

high-intensity lasers, demonstrated the generation of quasi-monoenergetic GeVelectrons.<br />

With future Petawatt lasers ultra-high beam currents of ∼ 100 kA in ∼ 10 fs<br />

can be expected, allowing for drastic reduction in the undulator length of free-electronlasers<br />

(FELs). We present a discussion of the key aspects of a table-top FEL design,<br />

including energy loss <strong>and</strong> chirps induced by space-charge <strong>and</strong> wakefields. These<br />

effects become important for an optimized table-top FEL operation. A first proof-ofprinciple<br />

VUV case is considered as well as a table-top X-ray-FEL which may also<br />

open a brilliant light source for new methods in clinical diagnostics.<br />

PACS 41.60.Cr; 52.38.Kd<br />

1 Introduction<br />

The pursuit of table-top FELs<br />

combines two rapidly developing fields:<br />

laser-plasma accelerators, where the<br />

generation of intense quasi-monoenergetic<br />

electron bunches up to the GeV<br />

range has been achieved [1–4], <strong>and</strong><br />

large-scale X-ray free-electron lasers<br />

(XFELs) that are expected to deliver<br />

photon beams with unprecedented peak<br />

brilliance. A prominent application of<br />

such FEL pulses is single molecule<br />

imaging [5]. The proposed laser-plasma<br />

accelerator-based FELs would not only<br />

allow a greater availability due to their<br />

smaller size <strong>and</strong> costs, but also offer new<br />

features, such as pulses synchronized<br />

to the phase-controlled few-cycle driver<br />

laser [6] for pump–probe experiments.<br />

Moreover, the X-ray energy of a tabletop<br />

XFEL can be as large as required for<br />

medical diagnostics (above 20 keV [7]).<br />

1 Department of Physics, Ludwig-Maximilians-Universität München, 85748 Garching, Germany<br />

2 Max-Planck Institute of Quantum Optics, 85748 Garching, Germany<br />

3 INFN, 00044 Frascati, Italy<br />

4 Pulsar Physics, 3762 Soest, Netherl<strong>and</strong>s<br />

5 Institute of Nuclear Physics, Mainz, Germany<br />

6 University of California Los Angeles, UCLA, Los Angeles, CA, USA<br />

The mechanism for the generation<br />

of intense (nC charge) quasi-monoenergetic<br />

electron pulses by laser-plasma<br />

accelerators requires an ultrashort, highintensity<br />

laser-pulse with a length<br />

shorter than the plasma wavelength<br />

(on the µm-scale corresponding to gas<br />

densities of 10 19 cm −3 ). Due to the<br />

ponderomotive force, plasma electrons<br />

are blown out transversely, leaving an<br />

electron-free zone – the bubble – behind<br />

the laser pulse [8]. These electrons<br />

return to the axis after half a plasma oscillation,<br />

thus determining the size of<br />

the bubble in the order of the plasma<br />

wavelength. Typically about 10 9 ...10 10<br />

electrons are captured into the bubble,<br />

as found both experimentally <strong>and</strong> from<br />

scaling laws of relevant bubble parameters<br />

within the framework of similarity<br />

arguments [9]. Due to the inertial positive<br />

ion background, these electrons<br />

experience a strong electrical field gra-<br />

� Fax: +49 89 2891 4072, E-mail: florian.gruener@physik.uni-muenchen.de<br />

dient of up to TV/m. Particle-in-cell<br />

(PIC) simulations [10] show that the<br />

bubble electrons form a stem that is geometrically<br />

considerably smaller than the<br />

bubble as seen in Fig. 1.<br />

For these dimensions, beam currents<br />

of the order of ∼ 100 kA (i.e. ∼ 1 nC<br />

charge within ∼ 10 fs) can be reached<br />

without the need for any bunch compressor.<br />

For preparation of the required<br />

gas densities, either supersonic gas-jets<br />

or capillaries [11] are used, the latter<br />

consisting of a ∼ 300 µm-thin gas-filled<br />

channel with a parabolic radial ion density<br />

profile generated by a second laser<br />

pulse or discharge. The advantage of<br />

the capillary is that the laser can be<br />

guided beyond the Rayleigh length allowing<br />

longer acceleration distances<br />

<strong>and</strong> hence higher electron energies. The<br />

very recent experiments by Leemans et<br />

al. [4], utilizing a 40-TW laser pulse<br />

of 38 fs duration <strong>and</strong> a gas density<br />

of 4.3 × 10 18 cm −3 , clearly show that<br />

1 GeV electron beams can now be produced<br />

with capillaries. However, the<br />

measured charge of 30 pC is significantly<br />

below our design goal of 1 nC.<br />

In order to further improve the resulting<br />

current, we propose the scheme described<br />

below. The number of bubble<br />

electrons can be increased when using<br />

plasmas with higher density such as in<br />

Fig. 1, because the feeding process is<br />

more efficient if more plasma electrons<br />

are present. An increased gas density<br />

requires shorter laser-pulses (sub-10fs),<br />

such that the entire laser pulse fits<br />

into one plasma period. In this case, the<br />

entire laser pulse energy can be used,<br />

160 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008<br />

Rapid communication


432 Applied Physics B – Lasers <strong>and</strong> Optics<br />

1 . 3 . 3 S E L E C T E D R E P R I N T S<br />

FIGURE 1 (Color online) Snapshot of electron density (in units of 10 20 cm −3 ) from PIC simulation.<br />

The geometrical size of the electron-free cavity (“bubble”) behind the laser spot corresponds to<br />

the plasma period, which is about 8 µm in this case (gas density 1.8 × 10 19 cm −3 ). The stem of the<br />

high-energy electrons is much shorter than the plasma period<br />

while longer pulses lose energy during<br />

self-shortening for entering the bubble<br />

regime. Therefore, in contrast to all previous<br />

laser-acceleration experiments,<br />

we propose to use a different driver<br />

laser, namely a ∼ 5 fs-Petawatt laser<br />

(like the Petawatt-field-synthesizer PFS<br />

at MPQ [12]). Own PIC simulations<br />

show that such ultrashort lasers can capture<br />

more than 1 nC charge inside the<br />

bubble. The smaller plasma period leads<br />

to smaller bubbles, hence shorter bubble<br />

stems <strong>and</strong> thus higher currents (1 nC<br />

in 10 fs). However, the expected final<br />

energy in the bubble regime is, according<br />

to the scaling laws, lower for shorter<br />

laser pulses. To overcome this limit, we<br />

also propose to use a capillary <strong>and</strong> suggest<br />

the following scenario therein: with<br />

a longitudinal plasma density gradient,<br />

the laser pulse can be forced to gradually<br />

increase its diameter, thus adiabatically<br />

turning from the pure bubble regime<br />

into the pure blow-out regime long<br />

before the laser is depleted. Figure 1<br />

shows a snapshot of a bubble around<br />

the transition from the bubble to the<br />

blow-out regime, where the electrons<br />

blown away from the laser <strong>and</strong> flowing<br />

around the bubble are so strongly<br />

deflected by the electric field of the captured<br />

bubble electrons, that electrons<br />

are no longer scattered into the bubble<br />

(which is not charge-neutralized yet).<br />

The number <strong>and</strong> absolute energy spread<br />

of the bubble electrons is thus frozen,<br />

but their energy is still increased due to<br />

the present bubble fields. The remaining<br />

energy of the laser allows maintaining<br />

of the bubble structure for the remaining<br />

distance inside the capillary, where<br />

the laser is then guided along to overcome<br />

the Rayleigh limit. It is only for<br />

this stage with increased laser beam<br />

size, that the capillary-induced guiding<br />

is relevant. Note that this scheme<br />

is a two-stage approach, but within one<br />

<strong>and</strong> the same capillary, where the laser<br />

turns adiabatically from one stage into<br />

the other. It can be expected that the energy<br />

spread for 100 MeV is about 1%<br />

(as confirmed by measurements [2]),<br />

but 0.1% for 1 GeV 1 . Normalized emittances<br />

are found both from simulations<br />

<strong>and</strong> experiments to range between<br />

0.1–1 mm mrad.<br />

1 Note that the detector resolution in the cited<br />

GeV-experiment did not suffice to resolve energy<br />

spreads below one percent.<br />

2 Table-top<br />

free-electron-lasers<br />

An FEL requires an undulator<br />

which is an arrangement of magnets<br />

with an alternating transverse magnetic<br />

field. Electrons in an undulator are<br />

forced on a sinusoidal trajectory <strong>and</strong><br />

can thus couple with a co-propagating<br />

radiation field. The induced energy<br />

modulation yields a current modulation<br />

from the dispersion of the undulator<br />

field. This modulation is called<br />

micro-bunching expressing the fact that<br />

the electrons are grouped into small<br />

bunches separated by a fixed distance.<br />

Therefore, electrons emit coherent radiation<br />

with a wavelength equal to<br />

the periodic length between the microbunches.<br />

In a self-amplification of spontaneous<br />

emission (SASE) FEL, there is<br />

no initial radiation field <strong>and</strong> the seed<br />

has to be built up by the spontaneous<br />

(incoherent) emission [13]. We present<br />

quantitative arguments, which are complemented<br />

by SASE FEL (GENESIS<br />

1.3 [14]) simulations. The gain length,<br />

which is the e-folding length of the exponential<br />

amplification of the radiation<br />

power, is<br />

Lgain,ideal = λu<br />

4π √ , (1)<br />

3�<br />

with undulator period λu <strong>and</strong> the basic<br />

scaling parameter �. This Pierce or FEL<br />

parameter describes the conversion efficiency<br />

from the electron beam power<br />

into the FEL radiation power <strong>and</strong> reads<br />

for the one-dimensional <strong>and</strong> ideal case<br />

(neglecting energy spread, emittance,<br />

diffraction, <strong>and</strong> time-dependence)<br />

[13, 15]<br />

� = 1<br />

� � � �<br />

2<br />

1/3<br />

I λu Au<br />

. (2)<br />

2γ IA 2πσx<br />

Here γ = Ebeam/mc2 is the electron<br />

beam energy, I the beam current, IA =<br />

17 kA the Alfven-current, σx the beam<br />

size <strong>and</strong> Au = au[J0(ζ) − J1(ζ)] (planar<br />

undulator), whereby a2 u = K 2 /2,<br />

K is the undulator parameter (K =<br />

0.93λu[cm]B0[T] <strong>and</strong> B0 the magnetic<br />

field strength on the undulator axis),<br />

ζ = a 2 u /(2(1 + a2 u<br />

)), <strong>and</strong> J are Bessel<br />

functions. In the presence of energy<br />

spread, emittance, <strong>and</strong> diffraction, a correction<br />

factor Λ is introduced for the<br />

gain length [15]: Lgain = Lgain,ideal(1 +<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 161


1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

GRÜNER et al. Design considerations for table-top, laser-based VUV <strong>and</strong> X-ray free electron lasers 433<br />

Λ). The saturation length Lsat is the<br />

length along the undulator at which<br />

maximum micro-bunching is reached.<br />

The power of the FEL radiation at this<br />

point is the saturation power Psat, while<br />

the shot noise power Pn is the power<br />

of the spontaneous undulator radiation<br />

from which the SASE FEL starts. The<br />

coupling factor α = 1/9 describes the<br />

efficiency at which the noise power couples<br />

to the FEL gain process. The saturation<br />

length reads<br />

� �<br />

Psat<br />

Lsat = Lgain ln , (3)<br />

αPn<br />

where the saturation power scales as<br />

� �2 1<br />

Psat ∼ (Iλu)<br />

1 + Λ<br />

4/3 . (4)<br />

The FEL-wavelength is, in case of a planar<br />

undulator,<br />

λ = λu<br />

2γ 2<br />

�<br />

1 +<br />

2 �<br />

K<br />

. (5)<br />

2<br />

Thus, for reaching a certain wavelength<br />

λ, a shorter undulator period λu allows<br />

the use of less energetic electrons. In the<br />

following Table 1 <strong>and</strong> Fig. 2, we compare<br />

the FLASH VUV FEL (DESY) in<br />

the so-called femtosecond mode [16]<br />

with our corresponding table-top VUV<br />

FEL as well as a table-top XFEL operating<br />

with 1.2 GeV electrons, as discussed<br />

below.<br />

The importance of the ultra-high<br />

current in the table-top case is evident.<br />

The smaller undulator period allows<br />

a smaller beam energy, hence decreasing<br />

the gain <strong>and</strong> saturation lengths. The<br />

Pierce parameter gives the upper limit<br />

of the acceptable energy spread σγ /γ .<br />

Thus, for compensating the relatively<br />

large energy spread of laser-plasma accelerators<br />

<strong>and</strong> for maintaining a large<br />

output power, a beam current significantly<br />

above ∼ 17 kA is m<strong>and</strong>atory for<br />

keeping � large <strong>and</strong> Λ small.<br />

2.1 Space charge effects<br />

Such ultra-high beam currents<br />

are subject to strong space-charge<br />

forces. After release into vacuum, the<br />

electron bunch starts exp<strong>and</strong>ing, for<br />

which there are two sources: (i) its<br />

own space-charge, driving a Coulombexplosion,<br />

i.e. (transverse) space-charge<br />

expansion <strong>and</strong> (longitudinal) debunching,<br />

<strong>and</strong> (ii) its initial divergence, which<br />

drives a linear transverse expansion.<br />

Such expansions transform potential energy<br />

into kinetic, hence changing the<br />

initial energy distribution <strong>and</strong> bunch<br />

form. First, we want to discuss an extreme<br />

case, that is, an upper limit of the<br />

charge that can be expected from a laserplasma<br />

accelerator at lower energies.<br />

For this study we take a beam energy of<br />

γ0 = 260, 1.25 nC charge, <strong>and</strong> a Gaussian<br />

bunch with sizes σx,y,z = 1 µm, corresponding<br />

to I = 150 kA, <strong>and</strong> an initial<br />

divergence of θ0 = 1 mrad. In the rest<br />

frame of the bunch, its length amounts<br />

Parameter FLASH (fs) TT-VUV-FEL TT-XFEL<br />

current 1.3 kA 50 kA 160 kA<br />

norm. emitt. 6 mm mrad 1 mm mrad 1 mm mrad<br />

beam size 170 µm 30 µm 30 µm<br />

energy 461.5 MeV 150 MeV 1.74 GeV<br />

energy spread 0.04% 0.5% 0.1%<br />

und. period 27.3 mm 5 mm 5 mm<br />

wavelength 30 nm 32 nm 0.25 nm<br />

Pierce par. 0.002 0.01 0.0015<br />

sat. length 19 m 0.8 m 5 m<br />

pulse length 55 fs 4 fs 4 fs<br />

sat. power 0.8 GW 2.0 GW 58 GW<br />

TABLE 1 Parameters for the comparison between the DESY femtosecond-mode Flash-VUV case<br />

<strong>and</strong> the corresponding table-top VUV FEL as well as a table-top X-ray FEL (rms values)<br />

162 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008<br />

FIGURE 2 Saturation lengths (3)<br />

<strong>and</strong> (inset) degradation factor Λ as<br />

a function of electron beam current<br />

I: DESY’s FLASH (dashed curve,<br />

triangle) <strong>and</strong> the table-top VUV scenario<br />

(solid line). The circles denote<br />

specific GENESIS runs<br />

to σ � z = σzγ0 = 260 µm. For such an<br />

aspect-ratio, the transverse Coulombexplosion<br />

dominates over the longitudinal<br />

one (GPT [17] simulations reveal<br />

that after a time period of 1 ps<br />

σ � x = 73 µm, while σ � z = 268 µm). Figure<br />

3 shows the electron distribution<br />

in the bunch rest frame from the GPT<br />

simulations.<br />

It can be seen that the transverse<br />

Coulomb explosion dominates the longitudinal<br />

one. We want to emphasize<br />

here that studies of space-charge effects<br />

of such ultra-dense relativistic bunches<br />

can be simulated most accurately only<br />

with codes (such as GPT <strong>and</strong> CSRtrack<br />

[18]) which utilize point-to-point<br />

interactions (a result also found in [19]).<br />

In contrast, other codes merely based<br />

upon Poisson-solvers cannot cope with<br />

large relative particle motions within the<br />

bunch rest frame in case of large initial<br />

energy spreads <strong>and</strong> Coulomb explosioninduced<br />

motion. In comparison with the<br />

Coulomb-driven explosion, the linear<br />

divergence-driven transverse expansion<br />

is even larger: after a distance of 1 cm<br />

in the laboratory frame, the beam size<br />

is increased to 10 µm, while the transverse<br />

Coulomb-explosion yields 6 µm<br />

only. Hence, even in the extreme case<br />

the transverse bunch expansion is dominated<br />

by the initial divergence <strong>and</strong> for<br />

lower charges this dominance increases.<br />

If γ0 is the bunch energy just before<br />

release into vacuum, γ � <strong>and</strong> β� the energy<br />

<strong>and</strong> velocity of a test electron in<br />

the bunch rest frame, γ <strong>and</strong> β ≈ 1 for<br />

its laboratory frame energy <strong>and</strong> velocity,<br />

then one can clearly distinguish<br />

between a transverse-dominated expansion<br />

<strong>and</strong> a longitudinal-dominated<br />

one: in the first case, where the electron<br />

moves in the rest frame with velocity<br />

±β� along transverse direction,<br />

γ = γ0γ � , while in the latter case, where


434 Applied Physics B – Lasers <strong>and</strong> Optics<br />

1 . 3 . 3 S E L E C T E D R E P R I N T S<br />

FIGURE 3 (Color online) Projected spatial electron distribution within the bunch rest frame in the<br />

extreme case, at τ = 0 s (left) <strong>and</strong> τ = 4 ps (right), from a GPT run. The color code indicates the electron<br />

energy (blue is initial kinetic energy, red is increased energy). The transverse expansion is clearly<br />

dominating<br />

the electron moves with velocity ±β �<br />

in the beam direction, γ = γ0γ � (1 ± β � ).<br />

Hence, in a purely transverse expansion<br />

all electrons gain kinetic energy,<br />

while in a purely longitudinal expansion,<br />

some electrons lose energy. The<br />

inset of Fig. 4 shows the absolute kinetic<br />

electron energy distribution along<br />

the bunch at a distance of z = 3 cm after<br />

its release into vacuum (in case of 0.6<br />

nC). This spectrum shows that the bunch<br />

under study mostly undergoes a transverse<br />

expansion. The maximum energy<br />

gain lies in the middle of the bunch,<br />

where the density is highest, i.e., where<br />

initially the highest potential energy was<br />

stored. That the space-charge driven<br />

transverse expansion is weaker than the<br />

divergence-driven one is also confirmed<br />

by the fact that (even in the extreme<br />

case) the divergence is increased only<br />

to 1.3 mrad. After passing a focusing<br />

system (a triplet of quadrupoles, positioned<br />

at z = 4 cm, having a length of<br />

about 10 cm), the Gaussian-like energy<br />

distribution is transformed into a quasilinear<br />

one, as depicted in Fig. 4. The<br />

reason is that in the resulting quasicollimated<br />

low-divergence beam faster<br />

electrons can catch up the slower ones.<br />

Note that after z = 3 cm (inset) the<br />

fastest electrons are symmetrically distributed<br />

around the middle of the bunch.<br />

These electrons start overtaking the<br />

slower ones, that is, they move forward<br />

within the bunch. If initially longitudinal<br />

debunching dominates, the<br />

fastest electrons are found in the head<br />

<strong>and</strong> the slowest ones in the tail of the<br />

bunch.<br />

A key parameter for SASE to occur<br />

is the energy spread σγ /γ , which must<br />

always be smaller than the Pierce parameter<br />

�. The relevant energy spread<br />

is taken over a bunch slice with the size<br />

of one cooperation length, i.e. the slippage<br />

length of the radiation along the<br />

bunch over one gain length. As seen in<br />

Fig. 4, a large fraction of the bunch fulfills<br />

σγ /γ < �. Due to the slippage of the<br />

radiation relative to the bunch, the linear<br />

energy chirp must be compensated,<br />

because the energy factor γ in (5) increases<br />

along the bunch. However, in<br />

case of short-period undulators the gap<br />

is correspondingly small (typically one<br />

third of the undulator period) <strong>and</strong> such<br />

small gaps lead to strong wakefields<br />

also causing an energy variation along<br />

the bunch [23, 24]. In case of ultrashort<br />

FIGURE 4 Energy vs. position<br />

(in co-moving frame) along the<br />

bunch before entering the undulator<br />

(GPT simulation): the<br />

Gaussian-like distribution after<br />

a drift of z = 3 cm behind the gasjet<br />

(inset), indicating a transversedominated<br />

bunch expansion, is<br />

stretched into a quasi-linear one<br />

behind the focusing system. The<br />

vertical lines show the slice energy<br />

spread σγ <strong>and</strong> the horizontal lines<br />

mark the initial energy γ0 = 260<br />

bunches, the characteristic length of the<br />

resistive wake potential is longer than<br />

the bunch length <strong>and</strong>, hence, the energy<br />

change is found to be negative<br />

with a linear few-percent variation along<br />

the bunch. This in turn implies that the<br />

bunch as a whole is decelerated during<br />

the passage of the undulator, whereby<br />

the head loses less energy than the tail.<br />

It turns out that the linear energy chirp<br />

induced by space-charge corresponds<br />

well with the wakefield-induced slowing<br />

down of the entire electron bunch.<br />

Therefore, the radiation slipping in the<br />

forward direction interacts with electrons<br />

of effectively constant energy, if<br />

by varying the gap the bunch energy<br />

loss is tuned with respect to the slippage.<br />

In other words, the space-charge<br />

effects <strong>and</strong> the wakefield effects cancel<br />

each other. We have found with simulations<br />

that for a first proof-of-principle of<br />

a table-top FEL the comparably low energy<br />

of 130–150 MeV is well suited to<br />

cope with the wakefields in the undulator.<br />

In this sense the table-top VUV case<br />

given in Table 1 can be regarded as an<br />

optimal test case.<br />

Coherent synchrotron radiation<br />

(CSR) [20] as another source for increasing<br />

the (slice) energy spread within<br />

the undulator was found to have negligible<br />

impact. This was shown with CSRtrack<br />

simulations which take into account<br />

that in our cases the bunches have<br />

a much larger transverse size (σx,y ∼<br />

30 µm) than length (σz ∼ 1 µm).<br />

2.2 Table-top XFEL<br />

So far we have discussed<br />

a proof-of-principle scenario at relatively<br />

low electron energies, where<br />

space-charge effects play a dominant<br />

role leading to a linear energy<br />

chirp. The situation of the proposed<br />

table-top X-FEL (TT-XFEL) is different.<br />

For reaching a wavelength of<br />

λ = 0.25 nm, an electron energy of<br />

1.74 GeV is needed in case of a period<br />

of λu = 5 mm. Space charge effects are<br />

much weaker here. For this less dem<strong>and</strong>ing<br />

situation we have confirmed<br />

with four different simulation codes<br />

(GPT, ASTRA [21], CSRtrack, HOM-<br />

DYN [22]) that above 1 GeV, with the<br />

same parameters as in the extreme case<br />

given above, Coulomb-explosion leads<br />

to a projected energy chirp of below<br />

0.3% <strong>and</strong> a bunch elongation with a fac-<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 163


1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

GRÜNER et al. Design considerations for table-top, laser-based VUV <strong>and</strong> X-ray free electron lasers 435<br />

tor below 1.1. Furthermore, CSR can<br />

be neglected in agreement with the 1D<br />

theory. Experimentally the most dem<strong>and</strong>ing<br />

constraint is that the Pierce<br />

parameter is one order of magnitude<br />

smaller than in the test case at 130 MeV,<br />

hence the electron (slice) energy spread<br />

should be as small as 0.1%. This goal<br />

seems to be within reach considering the<br />

bubble-transformation scenario mentioned<br />

above.<br />

Without the effect of wakefields<br />

GENESIS simulations have shown that<br />

this TT-XFEL scenario with an undulator<br />

length of only 5 m yields 8 × 10 11<br />

photons/bunch within ∼ 4 fs <strong>and</strong> 0.2%<br />

b<strong>and</strong>width, a divergence of 10 µrad, <strong>and</strong><br />

a beam size of 20 µm. However, the<br />

wakefields become the dominant degrading<br />

effect as the required undulator<br />

length is larger <strong>and</strong> the Pierce parameter<br />

reduced, hence also the tolerance with<br />

respect to variations in the electron energy.<br />

But since there is no initial spacecharge-induced<br />

energy chirp, one must<br />

find another method for compensating<br />

the wakefield-induced energy variation.<br />

A suitable method for compensating<br />

the wakefields for the TT-XFEL would<br />

be tapering, i.e., varying the undulator<br />

period along the undulator. Due to the<br />

fact that the undulator parameter K is<br />

smaller than unity, tapering via K, i.e.,<br />

by gap variation, could only be used as<br />

fine-tuning. Depending on the specific<br />

material properties of the undulator surface,<br />

our first wake calculations show<br />

that the bunch center loses over the entire<br />

5 m undulator length about 10%<br />

of its initial energy. Due to the linear<br />

wake field variation along the bunch<br />

only a fraction of it will undergo SASE.<br />

A detailed study will be further investigated<br />

in a future paper.<br />

2.3 Experimental status<br />

In order to demonstrate practical<br />

feasibility, we have built <strong>and</strong> tested<br />

a miniature focusing system <strong>and</strong> undulator<br />

consisting of permanent mag-<br />

nets. The focusing triplet consists of<br />

mini-quadrupoles with an aperture of<br />

just 5 mm, hence allowing for measured<br />

gradients of 530 T/m. Their focusing<br />

strength was measured with a 600 MeV<br />

electron beam (at Mainz Microtron facility<br />

MAMI, Mainz, Germany). A first<br />

test hybrid undulator with a period of<br />

only 5 mm <strong>and</strong> a peak magnetic field<br />

strength of ∼ 1 T has been built, <strong>and</strong><br />

produced, with a 855 MeV beam (also<br />

at MAMI), an undulator radiation spectrum<br />

as expected. These results will be<br />

published elsewhere.<br />

3 Conclusion<br />

We have shown by means<br />

of analytical estimates <strong>and</strong> SASE FEL<br />

simulations that laser-plasma accelerator-based<br />

FELs can only be operated<br />

with meter-scale undulators. The key<br />

parameter is the ultra-high electron peak<br />

current, which significantly reduces the<br />

gain length <strong>and</strong> increases the tolerance<br />

with respect to the energy spread <strong>and</strong><br />

emittance. The latter would also allow<br />

increasing of the TT-XFEL photon<br />

energy into the medically relevant<br />

range of 20–50 keV, because the limiting<br />

quantum fluctuations in the spontaneous<br />

undulator radiation, scale with<br />

γ 4 [25] <strong>and</strong> the required electron energy<br />

of the TT-XFEL is comparatively<br />

small.<br />

ACKNOWLEDGEMENTS We thank<br />

W. Decking, M. Dohlus, K. Flöttmann, T. Limberg,<br />

<strong>and</strong> J. Rossbach (all DESY) for fruitful<br />

discussions. Supported by Deutsche Forschungsgemeinschaft<br />

through the DFG-Cluster of Excellence<br />

Munich-Centre for Advanced Photonics.<br />

This work was also supported by DFG under Contract<br />

No. TR18.<br />

REFERENCES<br />

1 S.P.D. Mangles, C. Murphy, Z. Najmudin,<br />

A. Thomas, J. Collier, A. Dangor, E. Divall,<br />

P. Foster, J. Gallacher, C. Hooker,<br />

D. Jaroszynski, A. Langley, W. Mori, P. Norreys,<br />

F. Tsung, R. Viskup, B. Walton,<br />

K. Krushelnick, Nature 431, 535 (2004)<br />

164 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008<br />

2 C.G.R. Geddes, C. Toth, J. van Tilborg,<br />

E. Esarey, C. Schroeder, D. Bruhwiller,<br />

C. Nieter, J. Cary, W. Leemans, Nature 431,<br />

538 (2004)<br />

3 J. Faure, Y. Glinec, A. Pukhov, S. Kiselev,<br />

S. Gordienko, E. Lefebvre, J.-P. Rousseau,<br />

F. Burgy, V. Malka, Nature 431, 541<br />

(2004)<br />

4 W.P. Leemans, B. Nagler, A. Gonsalves,<br />

C. Toth, K. Nakamura, C. Geddes, E. Esarey,<br />

C. Schroeder, S. Hooker, Nature Phys. 2, 696<br />

(2006)<br />

5 R. Neutze, R. Wouts, D. van der Spoel,<br />

E. Weckert, J. Hajdu, Nature 406, 752<br />

(2000)<br />

6 A. Baltus, T. Udem, M. Uiberacker,<br />

M. Hentschel, E. Goulielmakis, C. Gohle,<br />

R. Holzwarth, V.S. Yakovlev, A. Scrinzi,<br />

T.W. Hänsch, F. Krausz, Nature 421, 611<br />

(2003)<br />

7 W. Thomlinson, P. Suortti, D. Chapman,<br />

Nucl. Instrum. Methods Phys. Res. A 543,<br />

288 (2005)<br />

8 A. Pukhov, J. Meyer-ter-Vehn, Appl. Phys. B<br />

74, 355 (2002)<br />

9 A. Pukhov, S. Gordienko, Philos. Trans. R.<br />

Soc. London A 364, 623 (2006)<br />

10 M. Geissler, J. Schreiber, J. Meyer-ter-Vehn,<br />

New J. Phys. 8, 186 (2006)<br />

11 D.J. Spence, A. Butler, S. Hooker, J. Opt.<br />

Soc. Am. B 20, 138 (2003)<br />

12 http://www.attoworld.de/research/PFS.html<br />

13 R. Bonifacio, C. Pellegrini, L.M. Narduci,<br />

Opt. Commun. 50, 373 (1984)<br />

14 S. Reiche, Nucl. Instrum. Methods Phys.<br />

Res. A 429, 243 (1999)<br />

15 M. Xie, Nucl. Instrum. Methods Phys. Res. A<br />

445, 59 (2000)<br />

16 Technical Design Report:<br />

http://www.hasylab.desy.de/facility/fel/vuv/<br />

ttf2_update/TESLA-FEL2002-01.pdf<br />

17 M.J. de Loos, S.B. van der Geer, Proc. 5th<br />

Eur. Part. Acc. Conf., Sitges, 1241 (1996)<br />

18 M. Dohlus, T. Limberg, Proc. FEL 2004<br />

Conf., 18-21<br />

19 G. Fubiana, J. Qiang, E. Esarey, W.P. Leemans,<br />

Phys. Rev. ST Accel. Beams 9, 064 402<br />

(2006)<br />

20 E. Saldin, E.A. Schneidmiller, M.V. Yurkov,<br />

Nucl. Instrum. Methods Phys. Res. A 417,<br />

158 (1998)<br />

21 K. Floettmann, Astra User’s Manual, 2000,<br />

http://www.desy.de/mpyflo/<br />

22 M. Ferrario, J.E. Clendenin, D.T. Palmer,<br />

J.B. Rosenzweig, L. Serafini, Proc. of the 2nd<br />

ICFA Adv. Acc. Workshop on “The Physics<br />

of High Brightness Beams”, UCLA, Nov.,<br />

1999; see also SLAC-PUB-8400<br />

23 A.W. Chao, Physics of Collective Beam Instabilities<br />

in High Energy Accelerators (Wiley,<br />

New York, 1993)<br />

24 K. Bane, SLAC-PUB-11829 (2006)<br />

25 M. S<strong>and</strong>s, Phys. Rev. 97, 470 (1955)


Vol 446 | 5 April 2007 |doi:10.1038/nature05648<br />

1 . 3 . 3 S E L E C T E D R E P R I N T S<br />

<strong>Attosecond</strong> real-time observation of<br />

electron tunnelling in atoms<br />

ARTICLES<br />

M. Uiberacker 1,2 , Th. Uphues 3 , M. Schultze 2 , A. J. Verhoef 2,4 , V. Yakovlev 1 , M. F. Kling 5 , J. Rauschenberger 1,2 ,<br />

N. M. Kabachnik 3,6 , H. Schröder 2 , M. Lezius 2 , K. L. Kompa 2 , H.-G. Muller 5 , M. J. J. Vrakking 5 , S. Hendel 3 ,<br />

U. Kleineberg 1 , U. Heinzmann 3 , M. Drescher 7 & F. Krausz 1,2,4<br />

Atoms exposed to intense light lose one or more electrons <strong>and</strong> become ions. In strong fields, the process is predicted to occur<br />

via tunnelling through the binding potential that is suppressed by the light field near the peaks of its oscillations. Here we<br />

report the real-time observation of this most elementary step in strong-field interactions: light-induced electron tunnelling.<br />

The process is found to deplete atomic bound states in sharp steps lasting several hundred attoseconds. This suggests a new<br />

technique, attosecond tunnelling, for probing short-lived, transient states of atoms or molecules with high temporal<br />

resolution. The utility of attosecond tunnelling is demonstrated by capturing multi-electron excitation (shake-up) <strong>and</strong><br />

relaxation (cascaded Auger decay) processes with subfemtosecond resolution.<br />

With the invention of the laser in 1960 <strong>and</strong> subsequent advances in<br />

ultrashort light pulse generation, light fields with a strength of several<br />

volts per angstrom have become routinely available. They rival the<br />

fields acting on valence electrons in atomic systems, allowing their<br />

release from atoms, molecules <strong>and</strong> solids. These advances sparked a<br />

revolution in studying the interaction of electrons with light. The<br />

primary step in strong-field interactions is the liberation of electrons<br />

from their atomic bound state. The revolutionary theory of Keldysh1 <strong>and</strong> subsequent work2–6 suggested that a valence electron may escape<br />

by tunnelling through its atomic binding potential suppressed by the<br />

light field (Fig. 1a). If the dimensionless parameter:<br />

c ~ vL<br />

pffiffiffiffiffiffiffiffiffiffiffiffi<br />

2mWb<br />

ð1Þ<br />

jejE0 is less than one, under the assumption of "vL = Wb ionization is<br />

predicted to be confined to short intervals lasting a fraction of the half<br />

oscillation cycle of the light field (Fig. 1b). Here E0 <strong>and</strong> vL st<strong>and</strong> for<br />

the amplitude <strong>and</strong> angular frequency of the oscillations of the laser<br />

electric field EL(t) 5 E0e(t)cos(vLt 1 Q), with e(t) being the amplitude<br />

envelope function, <strong>and</strong> e, m <strong>and</strong> Wb the charge, mass <strong>and</strong> binding<br />

energy of the electron. Recent studies6 suggest that tunnelling<br />

remains the dominant ionization mechanism even for c substantially<br />

exceeding one, that is, under conditions when the potential barrier<br />

formed by the atomic binding potential <strong>and</strong> the ionizing light field<br />

varies during tunnelling (non-adiabatic regime).<br />

In this work we report what we believe is the first real-time observation<br />

of light-induced electron tunnelling. The observation of ionization<br />

occurring in subfemtosecond steps spaced by the half laser<br />

cycle up to values of the Keldysh parameter as high as three is in good<br />

agreement with analytic <strong>and</strong> numerical calculations. Our approach<br />

provides experimental access to all forms of optical field ionization—<br />

both adiabatic <strong>and</strong> nonadiabatic tunnelling, as well as barriersuppression<br />

ionization—<strong>and</strong> allows us to test models of these processes<br />

for the first time.<br />

Once the process of field ionization is fully understood, the technique<br />

of attosecond tunnelling will provide direct time-domain<br />

insight into a wide range of multi-electron dynamics <strong>and</strong> electron–<br />

electron interactions, ultimately with a resolution approaching<br />

the atomic unit of time (,24 as). We demonstrate this potential by<br />

probing shake-up <strong>and</strong> Auger cascade processes with subfemtosecond<br />

resolution.<br />

<strong>Attosecond</strong> probing of electron dynamics<br />

Figure 2 illustrates different options for attosecond sampling of electronic<br />

motion in atoms or molecules. A subfemtosecond extreme<br />

ultraviolet (XUV) pulse triggers the motion by exciting a valence<br />

or core electron (Fig. 2a, b). The unfolding excitation <strong>and</strong> relaxation<br />

processes (Fig. 2) could, in principle, be probed by a delayed replica<br />

of the pulse. However, the low flux of currently available subfemtosecond<br />

XUV pulses <strong>and</strong> the low two-photon transition probabilities<br />

in the XUV <strong>and</strong> X-ray regimes have thwarted this straightforward<br />

extension of conventional pump–probe techniques into the XUV–<br />

X-ray spectral range.<br />

A few-cycle wave of visible or near-infrared (NIR) light with controlled<br />

waveform 7 in combination with a highly nonlinear process<br />

may replace the subfemtosecond pulse either in probing or starting<br />

electron dynamics. This was first demonstrated by attosecond streaking<br />

8,9 : the strong-field interaction of a few-cycle light wave with free<br />

electrons released by a subfemtosecond XUV excitation pulse results<br />

in broadening <strong>and</strong> shifting of their final momentum distribution.<br />

Recording the streaked spectra of the emitted photo- <strong>and</strong> Auger<br />

electrons versus delay between the XUV pump <strong>and</strong> the few-cycle laser<br />

probe allowed us to retrieve the XUV pulse 10,11 <strong>and</strong> the laser field 12 as<br />

well as the inner-atomic relaxation dynamics 13 with subfemtosecond<br />

resolution.<br />

Here, we demonstrate that nonlinear interaction of the same light<br />

wave with bound electrons ionizes in subfemtosecond steps <strong>and</strong><br />

hence offers a means of probing intra-atomic <strong>and</strong> intra-molecular<br />

electron dynamics—including when no free electrons are released—<br />

by means of attosecond tunnelling. This approach relies on the fact<br />

that energetic photo-excitation as well as subsequent rearrangement<br />

1 Department für Physik, Ludwig-Maximilians-Universität, Am Coulombwall 1, 2 Max-Planck-Institut für Quantenoptik, Hans-Kopfermann-Strasse 1, D-85748 Garching, Germany.<br />

3 Fakultät für Physik, Universität Bielefeld, Universitätsstrasse 25, D-33615 Bielefeld, Germany. 4 Technische Universität Wien, Gusshausstrasse 27, A-1040 Vienna, Austria. 5 FOM-<br />

Instituut voor Atoom- en Molecuulfysica (AMOLF), Kruislaan 407, 1098 SJ, Amsterdam, The Netherl<strong>and</strong>s. 6 Institute of Nuclear Physics, Moscow State University, Moscow 119992,<br />

Russia. 7 Institut für Experimentalphysik, Universität Hamburg, Luruper Chaussee 149, D-22671 Hamburg, Germany.<br />

©2007 Nature Publishing Group<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 165<br />

627


1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

ARTICLES NATURE |Vol 446 | 5 April 2007<br />

via Auger decay is accompanied by transitions to unoccupied orbitals<br />

via ‘shake-up’ (Fig. 2d–f) 14–16 . Here, we use the term ‘shake-up’ in a<br />

broad sense, st<strong>and</strong>ing for all possible processes populating excited<br />

ionic states (including instantaneous <strong>and</strong> non-instantaneous ones),<br />

henceforth referred to as shake-up states (represented by levels 1, 2<br />

<strong>and</strong> 3 in the example of Fig. 2). These populations can be probed via<br />

optical field ionization by a strong, few-cycle NIR pulse of variable<br />

delay Dt, with respect to the subfemtosecond XUV excitation (Fig. 1c)<br />

by measuring the number of ions resulting from the XUV pump–NIR<br />

probe exposure as a function of Dt (Fig. 1d).<br />

Shake-up usually populates several quantum states in the valence<br />

b<strong>and</strong>, from which electrons can be freed by the NIR probe. Hence, the<br />

ion yield will constitute an integral signal, with contributions from all<br />

shake-up states up to a certain binding energy from which ionization<br />

is feasible for the intensity chosen. Population dynamics of individual<br />

states of significantly differing binding energy can be retrieved by<br />

pump–probe scans repeated at different NIR probe intensities <strong>and</strong>/<br />

or from the temporal separation of the depletion of the states in the<br />

same delay scan, as explained in Fig. 1d <strong>and</strong> demonstrated in Fig. 4.<br />

The ion yield constitutes an integral signal in a temporal sense, too.<br />

The shake-up states are exposed to the ionizing NIR field from the<br />

moment they have been populated until the end of the NIR pulse. The<br />

a<br />

W<br />

b<br />

0<br />

–W b<br />

Ionization<br />

rate<br />

e –<br />

0 x<br />

Core<br />

Time<br />

c<br />

XUV<br />

pump<br />

∆t (> 0)<br />

E L (t)=E 0 ε(t)cos(ω L t+ϕ)<br />

d<br />

Ionization<br />

yield<br />

T L /2<br />

Depletion of states<br />

with binding energy<br />

W 1 <strong>and</strong> W 2<br />

Delay time, ∆t<br />

beginning of this time interval within the NIR pulse is adjusted with<br />

the delay Dt between the XUV pump <strong>and</strong> the NIR probe. As the delay<br />

is scanned from large negative values (NIR probe first) to large positive<br />

values (XUV pump first) the measured ion yield (Fig. 1d) starts<br />

increasing at Dt , 0 owing to ionization on the trailing edge of the<br />

NIR pulse <strong>and</strong> continues to increase with increasing Dt because the<br />

XUV pump shifts towards the peak of the ionizing NIR probe <strong>and</strong> the<br />

shake-up states are exposed to ever higher NIR probe intensities.<br />

In the absence of Auger decay, shake-up excitation results from<br />

photoionization only (Fig. 2d). The time-dependent ionization<br />

dynamics sketched in Fig. 1d can then be traced by measuring the<br />

yield of doubly charged ions as a function of the delay between the<br />

XUV pump <strong>and</strong> the sampling NIR light field. The temporal ionization<br />

gradients sketched in Fig. 1d incorporate the finite duration of<br />

the XUV excitation, a possible delayed response of shake-up <strong>and</strong> the<br />

tunnelling dynamics. With a sufficiently rapid (=1 fs) excitation,<br />

these measurements can thus provide direct insight into the temporal<br />

evolution of shake-up (in the presence of a strong optical field) <strong>and</strong><br />

light-induced tunnelling. In what follows, first we prove this concept<br />

by exposing neon atoms to our subfemtosecond XUV pump <strong>and</strong> fewcycle<br />

NIR probe pulses (Fig. 3) <strong>and</strong> measuring the yield of the product<br />

of the pump–probe exposure, Ne 21 , versus Dt (Fig. 4). Then we<br />

extend the approach to probing electrons shaken up in xenon atoms<br />

during Auger decay. Our primary observables in this case are higher<br />

charged states, Xe N1 (for N . 2). Measuring their yield versus Dt<br />

displays the course of an Auger cascade.<br />

166 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008<br />

NIR<br />

probe<br />

Figure 1 | Strong-field ionization <strong>and</strong> pump-probe setting for its real-time<br />

observation. a, Exposing an atom to a strong NIR laser field will result in a<br />

modified potential (solid curve) composed of the Coulomb potential<br />

(dashed curve) <strong>and</strong> the time-dependent effective potential of the laser pulse.<br />

The laser is polarized along the x direction <strong>and</strong> W b is the binding energy of<br />

the electron. At sufficiently high laser field strengths the atomic binding<br />

potential is suppressed to a small barrier in the x or –x direction for the<br />

maxima <strong>and</strong> minima of the laser electric field, respectively, allowing optical<br />

tunnelling to become the dominant ionization mechanism. b, The highly<br />

nonlinear dependence of the tunnelling rate on the width of the potential<br />

barrier confines ionization to time intervals of very short duration near the<br />

field oscillation maxima. In the electric field of a few-cycle NIR laser pulse<br />

(thin line) ionization is predicted to be restricted to several subfemtosecond<br />

intervals (thick line). c, Concept for tracing optical-field ionization: a<br />

subfemtosecond XUV pulse generates ions in excited (shake-up) states, from<br />

which a time-delayed NIR few-cycle probe pulse liberates electrons to<br />

produce doubly charged ions. A delay Dt 5 0 is defined as the coincidence of<br />

the peaks of the envelopes of the NIR <strong>and</strong> XUV pulse. Dt . 0 implies that the<br />

peak of the XUV pulse envelope precedes that of the NIR pulse. d, Yield of<br />

doubly charged ions versus delay Dt between the XUV pump <strong>and</strong> the NIR<br />

probe, as predicted qualitatively for the case of electrons being prepared in<br />

two shake-up states of significantly differing binding energy (W 1 <strong>and</strong><br />

W 2 ? W 1) for liberation via strong-field ionization (see text for discussion).<br />

The state of low binding energy is depleted at low intensities, where multiphoton<br />

ionization dominates <strong>and</strong> hence sub-cycle structure in the ionization<br />

dynamics is absent. By contrast, the state of high binding energy is emptied<br />

by tunnelling, resulting in pronounced sub-half-cycle steps in the ionization<br />

profile.<br />

628<br />

Kinetic energy<br />

0<br />

Binding energy<br />

©2007 Nature Publishing Group<br />

+1 +1<br />

Valence Core-level<br />

photophotoemissionemission Probing by attosecond<br />

Streaking<br />

+2<br />

Auger<br />

decay<br />

1<br />

2<br />

+1 +1<br />

Tunnelling<br />

Photo-emission<br />

<strong>and</strong> shake-up<br />

a b c d e f<br />

3<br />

Unoccupied<br />

valence<br />

Occupied<br />

valence<br />

Core<br />

orbital<br />

+2 Final<br />

Auger charge<br />

decay <strong>and</strong> state<br />

shake-up<br />

Figure 2 | Probing electron dynamics in atoms, molecules or solids with<br />

attosecond sampling techniques. A subfemtosecond XUV pulse triggers the<br />

motion by inducing valence (process a) or core photoelectron emission<br />

(process b). The temporal evolution of photo- <strong>and</strong> Auger electron emission<br />

(process c) can be probed via attosecond streaking to retrieve the XUV pump<br />

pulse or the sampling NIR field <strong>and</strong> trace inner-shell relaxation dynamics.<br />

XUV photoexcitation as well as subsequent Auger decay processes are<br />

usually accompanied by shake-up of another electron to a previously<br />

unoccupied level (processes d, e <strong>and</strong> f). In this case the liberated electrons<br />

will be ejected at a reduced kinetic energy compared to the cases without<br />

shake-up processes a, b <strong>and</strong> c. The difference in energy is used for shaking up<br />

bound electrons (represented by curved black arrows to levels 1, 2 <strong>and</strong> 3).<br />

For sufficiently strong probing laser fields, the shake-up electrons can be<br />

liberated by tunnelling ionization. The temporal evolution of the tunnelling<br />

current will provide information about the inner-atomic electron dynamics<br />

that populates <strong>and</strong>/or depopulates the interrogated shake-up states <strong>and</strong> the<br />

duration of the process that have populated the levels on atto- <strong>and</strong><br />

femtosecond timescales. Note that the final charge state given in the figure is<br />

increased by attosecond tunnelling, while it remains unchanged in the case<br />

of attosecond streaking. The observable for streaking is the momentum<br />

distribution of liberated electrons, whereas in tunnelling it is the number of<br />

ions in different charge states.


Experimental set-up<br />

For a detailed description of the attosecond pump–probe apparatus,<br />

see the Supplementary Information. Briefly, the XUV pump originates<br />

from high-harmonic generation in a neon gas jet exposed to<br />

300-mJ, 750-nm waveform-controlled laser pulses with a duration of<br />

tL < (5.5 6 0.5) fs, at a repetition rate of 3 kHz. The collinear, linearly<br />

polarized XUV <strong>and</strong> NIR light beams are passed through several<br />

filters <strong>and</strong> reflected by a concentric double-mirror arrangement<br />

(Mo/Si multilayer: , 9 eV b<strong>and</strong>width at photon energy of 91 eV).<br />

The mirror introduces a variable delay between the XUV <strong>and</strong> the NIR<br />

fields, isolates a single or twin XUV pulse (depending on the carrierenvelope<br />

phase of the laser pulse) of tX < 250 as duration 11 by filtering<br />

the cut-off part of the harmonic emission spectrum <strong>and</strong> focuses<br />

both beams into a jet of atoms under scrutiny. The absolute delay<br />

between the XUV <strong>and</strong> the NIR signals is determined with an accuracy<br />

of better than 60.5 fs; for details see Supplementary Information.<br />

Ions created in the common focus of the two beams are detected<br />

by a time-of-flight ion spectrometer (reflectron) that combines<br />

high mass resolution (Dm/m < 10 23 ) with the capability of analysing<br />

Ne<br />

3s<br />

9.7%<br />

3p<br />

0.2%<br />

4p<br />

3d 4s<br />

4d<br />

1.0%<br />

12.3%<br />

2s –1<br />

2p –1<br />

1 . 3 . 3 S E L E C T E D R E P R I N T S<br />

NATURE | Vol 446 |5 April 2007 ARTICLES<br />

Energy (eV)<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

62.53 eV<br />

48.47 eV<br />

21.56 eV<br />

Shake-up satellites<br />

51.3%<br />

2p –2 25.0% nl<br />

XUV + IR (∆t > 0): 93.3%<br />

XUV: 95.2%<br />

Ne +<br />

1.9%<br />

( 1 S)<br />

( 1 D)<br />

( 3P) 2p –2<br />

6.7%<br />

4.8%<br />

Ne 2+<br />

Figure 3 | Energy levels <strong>and</strong> transitions in Ne 11 <strong>and</strong> Ne 21 ions relevant to<br />

this study. The plotted levels represent energies required to ionize <strong>and</strong><br />

possibly excite a neutral atom from its ground state. Absorption of photons<br />

of about 91-eV energy can produce singly or doubly charged ions with<br />

probabilities of 95.2% <strong>and</strong> 4.8%, respectively. Owing to shake-up a small<br />

fraction of the singly-charged ions is produced in 2p 22 nl configurations,<br />

where the probabilities of different channels are known from electron<br />

spectroscopy (see Supplementary Information). Each of these<br />

configurations, represented by a green box, consists of 2p 22 ( 3 P)nl,<br />

2p 22 ( 1 D)nl, <strong>and</strong> 2p 22 ( 1 S)nl states (the atomic terms in parentheses describe<br />

electrons not involved in the interaction). Our few-cycle laser field following<br />

the XUV pulse (Dt . 0) can remove electrons from these shake-up states,<br />

thus increasing the probability of double ionization to 6.7%. Depending on<br />

the initial 2p 22 nl state, the doubly ionized ion is left in one of the 3 P, 1 D or 1 S<br />

states. On its own, the laser pulse produces only singly charged ions (in<br />

configuration 2p 21 ) above the detection limit.<br />

particles within a micrometre-scale detection volume 17–19 . The<br />

target <strong>and</strong> background pressures were ,10 22 <strong>and</strong> ,10 28 mbar,<br />

respectively.<br />

Shake-up <strong>and</strong> tunnelling<br />

To probe shake-up <strong>and</strong> light-induced electron tunneling, we ionized<br />

neon atoms with our subfemtosecond XUV pulse. Figure 3 shows the<br />

level structure <strong>and</strong> transitions relevant to our experiments. The core<br />

shell was not accessed by our XUV photons, <strong>and</strong> hence Auger decay is<br />

absent 20 . The threshold energies for single <strong>and</strong> double ionization<br />

from the outer shell of Ne are 21.56 <strong>and</strong> 62.53 eV, respectively 21 .<br />

The XUV photons produced Ne 11 <strong>and</strong> Ne 21 ions with a ratio of<br />

a 1,000<br />

Ion counts<br />

b<br />

900<br />

800<br />

700<br />

Ionization yield (arbitrary units)<br />

©2007 Nature Publishing Group<br />

Ionization yield (arbitrary units)<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 167<br />

–1.0<br />

2p –2 4p<br />

380 as<br />

–0.5 0.0 0.5 1.0<br />

Delay time, ∆t (fs)<br />

2p –2 3d<br />

T L /2<br />

2p –2 3p<br />

2p –2 3s<br />

|E L |<br />

–14 –12 –10 –8 –6 –4 –2 0 2 4<br />

Delay time, ∆t (fs)<br />

Figure 4 | Ne 21 ion yield versus delay between the subfemtoscond XUV<br />

pump <strong>and</strong> the few-cycle NIR probe: experiment <strong>and</strong> modelling. The peak<br />

intensity of the NIR probe was (7 6 1) 3 10 13 W cm 22 . a, The experimental<br />

data were acquired from six delay scans repeated under the same<br />

experimental conditions. The signal was accumulated for 3 s at each delay<br />

setting. The squares <strong>and</strong> the error bars show the average <strong>and</strong> the st<strong>and</strong>ard<br />

error (s.e.m.) of the results of the six measurements. The thick red line shows<br />

the average of five adjacent data points; the thin grey line shows the same as<br />

the thick line but recorded with NIR probe pulses of r<strong>and</strong>om carrierenvelope<br />

phase. In the inset, squares, triangles, <strong>and</strong> circles depict an<br />

ionization step extracted from three different measurements normalized<br />

to give the same change in the ionization yield. The solid line shows an<br />

error-function fit to the data yielding a rise time of 380 as (full-width at<br />

half-maximum, FWHM, of the gaussian function derived from the error<br />

function). The zero of delay is set arbitrarily to the centre of the tunnelling<br />

process. b, Simulation of the pump–probe experiment based on the<br />

nonadiabatic theory of tunnel ionization 6 . The thin coloured lines show the<br />

calculated fractional ionization yields contributed by electrons liberated<br />

from different shake-up states. The thick blue line depicts the overall<br />

ionization rate obtained by totalling the fractional rates (<strong>and</strong> by adding the<br />

result to the background measured at large negative delays). The simulations<br />

were carried out for a gaussian 250-as XUV pulse <strong>and</strong> a gaussian 5.5-fs laser<br />

pulse with a peak intensity of 7 3 10 13 W cm 22 . The black solid curve<br />

represents the absolute value of the laser field. For discussion of the results,<br />

see text.<br />

629


1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

ARTICLES NATURE |Vol 446 | 5 April 2007<br />

(19.7 6 0.5):1 (with ,2,500 Ne 11 ions created per second), in good<br />

agreement with the results of synchrotron measurements 22 . A few per<br />

cent of the Ne 11 ions were promoted into 2p 22 nl (principal quantum<br />

number n: 3 or 4; quantum orbit l: s, p or d) configurations 23 (Fig. 3).<br />

These satellite states can only decay radiatively on a picosecond timescale.<br />

Double ionization by the NIR field was not observed at the intensity<br />

level chosen. The XUV-generated Ne 21 yield was therefore not<br />

affected by the NIR probe for Dt = 2t L but was significantly<br />

enhanced by the laser field for Dt approaching zero <strong>and</strong> becoming<br />

positive. The NIR-induced Ne 21 yield enhancement amounted to<br />

(40 6 4)% of the XUV-produced Ne 21 yield at a NIR peak intensity<br />

of (7 6 1) 3 10 13 W cm 22 . The absence of this enhancement for<br />

Dt = 2tL clearly indicates that the laser sets electrons free from states<br />

excited by the XUV pulse as sketched in Fig. 2d. The laser-induced<br />

change in the Ne 21 yield amounts to some 2% of the XUV-produced<br />

Ne 11 ions. This implies that a substantial fraction of the population<br />

of the 2p 22 nl shake-up satellites must have been depleted by field<br />

ionization.<br />

Figure 4a shows the number of Ne 21 ions detected as a function of<br />

delay Dt between the XUV pump <strong>and</strong> NIR probe. Figure 4b compares<br />

the prediction of the Yudin–Ivanov theory 6 (lines) with the experimental<br />

data (squares). In our modelling the shake-up states were<br />

populated instantly during XUV photoionization (for more details,<br />

see Supplementary Information). The calculations are in reasonable<br />

agreement with our measurements <strong>and</strong> reveal how the different<br />

shake-up states are depleted sequentially by laser-field ionization.<br />

The signal starts increasing at large negative delays owing to depletion<br />

of the 2p 22 4p state (relative population ,12%) <strong>and</strong> the 2p 22 3d state<br />

(,10%) at NIR intensity levels reached some 10 <strong>and</strong> 6 fs after the<br />

peak of the NIR probe (Dt , 210 <strong>and</strong> 26 fs), respectively. A more<br />

dramatic increase in the Ne 21 yield is observed as the delay<br />

approaches zero, as a consequence of the depletion of the most highly<br />

populated 2p 22 3p (,50%) <strong>and</strong> 2p 22 3s (,25%) states. In spite of<br />

their relatively high binding energy (,10 <strong>and</strong> 13 eV, respectively),<br />

these states are also depleted before zero delay, that is, by the field<br />

oscillation cycles comprised in the trailing edge of the pulse, leaving<br />

no room for increasing the Ne 21 yield with increasing Dt beyond 0.<br />

This main contribution to the Ne 21 yield emerges within approximately<br />

one <strong>and</strong> a half wave cycles of the NIR field, ,(3/2)TL 5 3p/vL,<br />

in several sharp steps that are spaced by ,TL/2; this clearly shows that<br />

field-induced tunnelling is the main cause of the observed increase in<br />

the Ne 21 yield. This conclusion is also supported by the disappearance<br />

of the steps in a pump–probe scan performed with a r<strong>and</strong>omly<br />

varying carrier-envelope phase of the NIR probe pulses (grey line in<br />

Fig. 4a).<br />

The main shake-up population (residing in the 2p 22 3p <strong>and</strong> 2p 22 3s<br />

states) is depleted at NIR intensities corresponding to a Keldysh<br />

parameter c of the order of three. Hence, our experiment verifies<br />

not only the existence of light-field-induced tunnelling, as predicted<br />

by Keldysh some four decades ago 1 , but also confirms the dominant<br />

role of this ionization mechanism up to c values substantially exceeding<br />

1, as predicted recently by Yudin <strong>and</strong> Ivanov 6 . This conclusion<br />

is also backed by numerical solutions of the time-dependent<br />

Schrödinger equation (see Supplementary Information).<br />

The steepness of the ionization steps <strong>and</strong> the dips preceding them<br />

in the measured data are not well reproduced by our model, which<br />

neglects the influence of electron–electron interactions <strong>and</strong> that of<br />

the strong NIR field on the XUV-induced transitions populating the<br />

shake-up states. Recent work 24 <strong>and</strong> our TDSE simulations (see Supplementary<br />

Information) indicate that the influence of the strong<br />

laser field on the XUV excitation process may (at least partially) be<br />

responsible for this discrepancy.<br />

We feel that these experiments afford profound insight into fundamental<br />

electronic processes such as tunnelling <strong>and</strong> shake-up by<br />

contrasting theoretical models with time-domain data. To exploit<br />

this potential both (1) accurate models of shake-up in the presence<br />

630<br />

of a strong laser field need to be developed <strong>and</strong> (2) the temporal<br />

resolution needs to be improved by using shorter XUV pulses 25<br />

<strong>and</strong> improving the signal-to-noise ratio as well as the accuracy of<br />

determining the zero of delay. These advances will allow determination<br />

of the attosecond temporal evolution of the light-field-induced<br />

tunnelling current <strong>and</strong> they will provide deep insight into the nature<br />

of the electron–electron interactions responsible for shake-up.<br />

Once models for shake-up <strong>and</strong> tunnelling have been tested <strong>and</strong><br />

verified with attosecond precision, the technique of attosecond tunnelling<br />

will provide direct time-domain insight into a wide range of<br />

multi-electron dynamics inside atoms <strong>and</strong> molecules by probing the<br />

transient population of excited valence states while these dynamics<br />

are unfolding. With improved signal-to-noise <strong>and</strong> XUV pulse duration,<br />

the temporal resolution may potentially approach the atomic<br />

unit of time (,24 as). At present, the observed rise time of the Ne 21<br />

yield of less than 400 as (which sets a corresponding upper limit on<br />

the time it takes the excited electronic states to become populated<br />

during XUV ionization <strong>and</strong> on tunneling; see inset in Fig. 4a) dictates<br />

the temporal resolution of our pump–probe approach. In the next<br />

section we demonstrate its applicability to probing intra-atomic<br />

multi-electron dynamics in real time.<br />

Photon<br />

energy<br />

5s –1<br />

3.3%<br />

5p –1<br />

9.7%<br />

Xe Xe + Xe 2+ Xe 3+ Xe 4+<br />

168 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008<br />

Energy (eV)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

}<br />

XUV -I<br />

©2007 Nature Publishing Group<br />

0<br />

8.9%<br />

Shake-up<br />

satellites<br />

A1<br />

4d –15p –1 5p<br />

nl<br />

–4 nl n′l′<br />

4d –1<br />

78.0%<br />

A1<br />

A1<br />

A1<br />

}<br />

β α<br />

5p –2<br />

γ<br />

A2<br />

A2<br />

NIR-I<br />

Energy (eV)<br />

NIR-DI<br />

XUV mirror reflectivity<br />

100<br />

90<br />

80<br />

5p –3<br />

5p –4<br />

Shake-up<br />

satellites<br />

8s,8p<br />

4d –15p –1 7s,7p<br />

5d,6p<br />

6p<br />

5d<br />

6s,5d<br />

nl<br />

Figure 5 | Energy levels <strong>and</strong> transitions in xenon ions relevant to the<br />

current study. The relative population of states in Xe 11 is given for an XUV<br />

excitation energy of 90 eV from ref. 30. The XUV light preferably ionizes<br />

from the 4d 21 shell, creating an inner-shell vacancy. A subsequent Auger<br />

process (green arrows labelled A1) will follow with 99% probability 26 <strong>and</strong> is<br />

predominantly decaying to Xe 21 in configuration 5p 22 . Some of the<br />

populated states, a <strong>and</strong> b (presumably 5s 21 5p 22 7p states 28,31 ) can further<br />

decay to Xe 31 via a second Auger process (green arrow labelled A2), leading<br />

to triply charged ions. The A1 process also populates states below the<br />

threshold for Xe 31 . These are denoted by c <strong>and</strong> represent 5s 21 5p 22 6p as well<br />

as 5p 23 nl configurations. The laser pulse can ionize these states (red arrow<br />

labelled NIR-I). Furthermore, a series of 4d 21 shake-up satellites in Xe 11 is<br />

also populated—with a small probability—by the XUV pulse. The inset<br />

shows the possible configurations 29 together with the XUV mirror<br />

reflectivity determining the XUV excitation spectrum. The satellites mainly<br />

decay to the a, b, c states, with a small fraction of the Xe 21 population<br />

ending up in 5p 24 nln9l9 configurations. These states are short-lived <strong>and</strong><br />

decay via the A2 process. Before this occurs, the nln9l9 electrons can be<br />

liberated by the laser field to yield Xe 41 in the 5p 24 state (red arrow labelled<br />

NIR-DI).


Multi-electron relaxation<br />

As a first application of the technique of attosecond tunnelling, we<br />

captured Auger cascade xenon atoms following excitation by a subfemtosecond<br />

XUV pulse. Figure 5 sketches the relevant energy levels<br />

<strong>and</strong> transitions. Energy-resolved synchrotron measurements have<br />

revealed that (1) the 91-eV XUV pulse will preferably liberate electrons<br />

from the 4d orbital 22 , (2) the vacancy decays by subsequent<br />

single (A1 in Fig. 5) <strong>and</strong> cascaded (A1 <strong>and</strong> A2 in Fig. 5) Auger<br />

processes, leading to Xe 21 <strong>and</strong> Xe 31 , respectively 26 , <strong>and</strong> (3) the lifetimes<br />

of the 4d3/2 <strong>and</strong> 4d5/2 holes are 6.3 6 0.2 <strong>and</strong> 5.9 6 0.2 fs,<br />

respectively 27 . These time-integral measurements have hitherto been<br />

able to set only a lower limit of 23 fs for the time constant of A2 (ref.<br />

27).<br />

To trace this dynamics in real time, we simultaneously recorded<br />

the number of Xe ions emerging in different charged states as a<br />

function of Dt. At the laser intensity of (7 6 1) 3 10 13 W cm 22 used<br />

in this experiment, the XUV-induced Xe 11 yield is buried in lasergenerated<br />

background, preventing delay-dependent effects from<br />

coming to light in the Xe 11 signal. This is not the case for higher<br />

charged states. With increasing delay, rapid exponential increase was<br />

observed in the Xe 31 signal near Dt 5 0 (Fig. 6b), concurrent with a<br />

significant decrease of the Xe 21 yield. The background in the Xe 31<br />

signal arises from the A1–A2 Auger cascade discussed above. From<br />

1 . 3 . 3 S E L E C T E D R E P R I N T S<br />

NATURE | Vol 446 |5 April 2007 ARTICLES<br />

a<br />

Ion counts<br />

b<br />

Ion counts (10 3 )<br />

40<br />

30<br />

20<br />

10<br />

0<br />

12<br />

10<br />

8<br />

τ A1 = 6.0 ± 0.7 fs<br />

τ A1<br />

τ A2 = 30.8 ± 1.4 fs<br />

0 50 100<br />

Delay time, ∆t (fs)<br />

150 200<br />

Xe 4+<br />

Xe 3+<br />

Figure 6 | Xe 41 <strong>and</strong> Xe 31 ion yields versus delay between the<br />

subfemtosecond XUV pump <strong>and</strong> the few-cycle NIR probe. The data have<br />

been compiled from the results of five delay scans repeated under the same<br />

experimental conditions. The signals were accumulated for 20 s at each delay<br />

setting. The squares <strong>and</strong> the error bars show the average <strong>and</strong> the s.e.m. of the<br />

results of these measurements. a, Assuming a sampling function identical to<br />

the ionization profile measured in the neon experiment (see solid line in<br />

Fig. 4a), a double exponential fit (see Supplementary Information) to the<br />

measured Xe 41 yield versus Dt (solid line) yields the Auger decay times<br />

tA1 5 6.0 6 0.7 fs <strong>and</strong> tA2 5 30.8 6 1.4 fs. b, With the temporal evolution of<br />

the A1 process acquired from the Xe 41 data <strong>and</strong> using the fact that the rise of<br />

the Xe 31 yield with t comprises both A1 <strong>and</strong> the laser-induced ionization<br />

process, we can obtain the temporal evolution of the laser-induced transition<br />

from that of the Xe 31 yield. The exponential rise from about 7,000 to 11,500<br />

counts is consistent with a laser-induced ionization time of 5.8 6 2.5 fs<br />

(FWHM). This is comparable to tL, indicating that low-order, one-photon<br />

<strong>and</strong>/or two-photon transitions may promote electrons from the c states of<br />

Xe 21 to the 5p 23 states of Xe 31 (see Fig. 5). Note that the fluctuations in the<br />

Xe 31 signal can be largely accounted for by XUV intensity variations, which<br />

can be efficiently eliminated by normalization to, for example, the Xe 21 ion<br />

yield (see Supplementary Information).<br />

©2007 Nature Publishing Group<br />

Fig. 5 we infer that the laser-induced increase in the Xe 31 yield is due<br />

to ionization from the c-states in Xe 21 , which cannot spontaneously<br />

decay into Xe 31 . This NIR-probe-induced transition (denoted by<br />

NIR-I in Fig. 5) yields Xe 31 in configuration 5p 23 . Because these<br />

are populated by the first Auger process, A1, the exponential increase<br />

in the Xe 31 signal is the convolution of the A1 decay <strong>and</strong> the NIRinduced<br />

ionization process. No decrease of the enhanced Xe 31 yield<br />

was observed up to our maximum delay of 300 fs, indicating that the<br />

lifetime of the c states was longer than 1 ps.<br />

Charge-states higher than Xe 31 cannot be created with the XUV<br />

pulse alone, because the XUV photon energy is below the threshold<br />

for Xe 41 production (,105 eV). However, with the probe switched<br />

on, we did observe Xe 41 ions for Dt . 2tL (Fig. 6a). The Xe 41 signal<br />

first grows within a few femtoseconds, followed by a longer decay.<br />

The Xe 41 ions are created by NIR-induced double ionization<br />

(denoted by NIR-DI in Fig. 5) from the intermediate doubly excited<br />

5p 24 nln9l9 (Fig. 5) <strong>and</strong>/or singly excited 5s 22 5p 21 nl (not shown in<br />

Fig. 5) states of Xe 21 . These states are populated by A1 from the<br />

satellites 4d 21 5p 21 nl of the 4d 21 state upon emission of low-energy<br />

electrons 28,29 <strong>and</strong> emptied by A2 to states of larger binding energy in<br />

Xe 31 , which cannot be reached by the NIR probe.<br />

From these results we conclude that the exponential rise <strong>and</strong> decay<br />

of the Xe 41 signal in Fig. 6a reveal the evolution of the A1 <strong>and</strong> A2<br />

Auger decays, respectively. The laser-induced double ionization may<br />

be either sequential or non-sequential (with the second step induced<br />

by recollision of the first electron). In either case, in the first step a<br />

binding energy similar to that of the most highly populated shake-up<br />

states in the neon experiment must be overcome by tunnel ionization.<br />

Hence, the probing (ionization) process may be assumed to be<br />

the same as measured in the neon experiment, see Fig. 4a. With this<br />

sampling function, fitting the result of a simple rate-equation analysis<br />

including the transitions A1 <strong>and</strong> A2 to the Xe 41 data shown in Fig. 6<br />

yields decay times of t A1 5 6.0 6 0.7 fs <strong>and</strong> t A2 5 30.8 6 1.4 fs for the<br />

A1 <strong>and</strong> A2 Auger processes, respectively. Both time constants are in<br />

accordance with the results of energy-resolved measurements 27 .<br />

Conclusions <strong>and</strong> outlook<br />

We have reported the observation of light-induced electron tunnelling<br />

from atoms in real time. Electrons are found to escape from their<br />

atomic binding potential within several subfemtosecond time intervals<br />

near the oscillation peaks of the ionizing few-cycle near-infrared<br />

laser field. Our results are in good agreement with the predictions of<br />

the theory Keldysh put forward four decades ago. The observed subfemtosecond<br />

ionization steps provide a powerful means of probing<br />

the transient population of short-lived valence electronic states in<br />

excited atoms or molecules, offering direct, time-domain access to a<br />

wide range of multi-electron dynamics unfolding on an attosecond<br />

to femtosecond timescale. Proof-of-principle attosecond tunnelling<br />

experiments in neon <strong>and</strong> xenon demonstrate this potential.<br />

Simultaneous implementation of attosecond tunnelling <strong>and</strong> attosecond<br />

streaking spectroscopy along with scaling of the techniques<br />

to higher photon energies <strong>and</strong> shorter X-ray pulse durations will<br />

provide unprecedented insight into the transient electronic states<br />

of matter.<br />

Received 2 November 2006; accepted 26 January 2007.<br />

1. Keldysh, L. V. Ionization in the field of a strong electromagnetic wave. Sov. Phys.<br />

JETP 20, 1307–1314 (1965).<br />

2. Faisal, F. H. M. Multiple absorption of laser photons by atoms. J. Phys. B 6,<br />

L89–L92 (1973).<br />

3. Reiss, H. R. Effect of an intense electromagnetic field on a weakly bound system.<br />

Phys. Rev. A 22, 1786–1813 (1980).<br />

4. Brabec, T. & Krausz, F. Intense few-cycle laser fields: frontiers of nonlinear optics.<br />

Rev. Mod. Phys. 72, 545–591 (2000).<br />

5. Scrinzi, A., Geissler, M. & Brabec, T. Ionization above the coulomb barrier. Phys.<br />

Rev. Lett. 83, 706–709 (1999).<br />

6. Yudin, G. L. & Ivanov, M. Yu. Nonadiabatic tunnel ionization: looking inside a laser<br />

cycle. Phys. Rev. A 64, 013409 (2001).<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 169<br />

631


1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

ARTICLES NATURE |Vol 446 | 5 April 2007<br />

7. Baltuska, A. et al. <strong>Attosecond</strong> control of electronic processes by intense light<br />

fields. Nature 421, 611–615 (2003).<br />

8. Itatani, J. et al. <strong>Attosecond</strong> streak camera. Phys. Rev. Lett. 88, 173903 (2002).<br />

9. Kitzler, M., Milosevic, N., Scrinzi, A., Krausz, F. & Brabec, T. Quantum theory of<br />

attosecond XUV pulse measurement by laser-dressed photoionization. Phys. Rev.<br />

Lett. 88, 173904 (2002).<br />

10. Hentschel, M. et al. <strong>Attosecond</strong> metrology. Nature 414, 509–513 (2001).<br />

11. Kienberger, R. et al. Atomic transient recorder. Nature 427, 817–821 (2004).<br />

12. Goulielmakis, E. et al. Direct measurement of light waves. Science 305, 1267–1269<br />

(2004).<br />

13. Drescher, M. et al. Time-resolved atomic inner-shell spectroscopy. Nature 419,<br />

783–787 (2002).<br />

14. Svensson, S., Eriksson, B., Martensson, N., Wendin, G. & Gelius, U. Electron shakeup<br />

<strong>and</strong> correlation satellites <strong>and</strong> continuum shake-off distributions in x-ray<br />

photoelectron spectra of the rare gas atoms. J. Electron Spectrosc. Related<br />

Phenomena 47, 327–384 (1988).<br />

15. Aksela, H., Aksela, S. & Kabachnik, N. Resonant <strong>and</strong> nonresonant Auger<br />

recombination. In VUV <strong>and</strong> Soft X-Ray Photoionization (eds Becker, U. & Shirley, D.<br />

A.) 401–440 (Plenum, New York, 1996).<br />

16. Istomin, A. Y., Manakov, N. L. & Starace, A. F. Perturbative analysis of the triply<br />

differential cross section <strong>and</strong> circular dichroism in photo-double-ionization of He.<br />

Phys. Rev. A 69, 032713 (2004).<br />

17. Schröder, H., Wagner, M., Kaesdorf, S. & Kompa, K. L. Surface-analysis by laser<br />

ionization. Ber. Bunsenges. Phys. Chem. 97, 1688–1692 (1993).<br />

18. Wagner, M. & Schröder, H. A novel 4 grid ion reflector for saturation of laser<br />

multiphoton ionization yields in a time-of-flight mass-spectrometer. Int. J. Mass<br />

Spectrom. 128, 31–45 (1993).<br />

19. Witzel, B., Schröder, H., Kaesdorf, S. & Kompa, K. L. Exact determination of<br />

spatially resolved ion concentrations in focused laser beams. Int. J. Mass Spectrom.<br />

172, 229–238 (1998).<br />

20. Larkins, F. P. Charge state dependence of x-ray <strong>and</strong> Auger electron emission<br />

spectra for rare-gas atoms—II. The neon atom. J. Phys. B 4, 14–19 (1971).<br />

21. National Institute of St<strong>and</strong>ards <strong>and</strong> Technology Physical Reference Data Æhttp://<br />

physics.nist.gov/PhysRefData/æ (1994).<br />

22. Holl<strong>and</strong>, D. M. P., Codling, K., West, J. B. & Marr, G. V. Multiple photoionization in<br />

the rare gases from threshold to 280 eV. J. Phys. B 12, 2465–2484 (1979).<br />

23. Becker, U. & Shirley, D. A. Partial Cross Sections <strong>and</strong> Angular Distributions. In<br />

VUV <strong>and</strong> Soft X-Ray Photoionization (eds Becker, U. & Shirley, D. A.) 135–173<br />

(Plenum, New York, 1996).<br />

632<br />

©2007 Nature Publishing Group<br />

24. Smirnova, O., Spanner, M. & Ivanov, M. Y. Coulomb <strong>and</strong> polarization effects in<br />

laser-assisted XUV ionization. J. Phys. B 39, 323–339 (2006).<br />

25. Sansone, G. et al. Isolated single-cycle attosecond pulses. Science 314, 443–446<br />

(2006).<br />

26. Kämmerling, B., Krässig, B. & Schmidt, V. Direct measurement for the decay<br />

probabilities of 4dj hole states in xenon by means of photoelectron-ion<br />

coincidences. J. Phys. B 25, 3621–3629 (1992).<br />

27. Penent, F., Palaudoux, J., Lablanquie, P. & Andric, L. Multielectron<br />

spectroscopy: the xenon 4d hole double Auger decay. Phys. Rev. Lett. 95, 083002<br />

(2005).<br />

28. Lablanquie, P. et al. Photoemission of threshold electrons in the vicinity of the<br />

xenon 4d hole: dynamics of Auger decay. J. Phys. B 35, 3265–3295 (2002).<br />

29. Hayaishi, T. et al. Manifestation of Kr 3d <strong>and</strong> Xe 4d conjugate<br />

shake-up satellites in threshold-electron spectra. Phys. Rev. A. 44, R2771–R2774<br />

(1991).<br />

30. Becker, U. et al. Subshell photoionization of Xe between 40 <strong>and</strong> 1000 eV. Phys.<br />

Rev. A 39, 3902–3911 (1989).<br />

31. Viefhaus, J. et al. Auger cascades versus direct double Auger: relaxation<br />

processes following photoionization of the Kr 3d <strong>and</strong> Xe 4d, 3d inner shells. J.<br />

Phys. B 38, 3885–3903 (2005).<br />

Supplementary Information is linked to the online version of the paper at<br />

www.nature.com/nature.<br />

Acknowledgements We thank A. F. Starace for discussions. We are grateful for<br />

financial support from the Volkswagenstiftung (Germany), the Marie Curie<br />

Research Training Network XTRA, Laserlab Europe, <strong>and</strong> a Marie Curie<br />

Intra-European Fellowship. F.K. acknowledges support from the FWF (Austria).<br />

The research of M.F.K. <strong>and</strong> M.J.J.V. is part of the research programme of the<br />

Stichting voor Fundamenteel Onderzoek der Materie, which is financially<br />

supported by the Nederl<strong>and</strong>se Organisatie voor Wetenschappelijk Onderzoek.<br />

This research was supported by the cluster of excellence Munich Centre for<br />

Advanced Photonics (www.munich-photonics.de).<br />

Author Contributions M.U., Th.U., M.S. <strong>and</strong> A.J.V. contributed equally to this work.<br />

Author Information Reprints <strong>and</strong> permissions information is available at<br />

www.nature.com/reprints. The authors declare no competing financial interests.<br />

Correspondence <strong>and</strong> requests for materials should be addressed to M.U.<br />

(matthias.uiberacker@mpq.mpg.de) or F.K. (ferenc.krausz@mpq.mpg.de).<br />

170 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008


1 . 3 . 3 S E L E C T E D R E P R I N T S<br />

Dispersion control over the<br />

ultraviolet–visible–near-infrared spectral range<br />

with HfO 2/SiO 2-chirped dielectric multilayers<br />

V. Pervak<br />

Max-Planck-Institute of Quantum Optics, Hans-Kopfermann-Strasse 1, D-85748 Garching, Germany<br />

F. Krausz<br />

Max-Planck-Institute of Quantum Optics, Hans-Kopfermann-Strasse 1, D-85748 Garching, Germany<br />

<strong>and</strong> Ludwig-Maximilians-Universitaet Muenchen, Am Coulombwall 1, D-85748 Garching, Germany<br />

A. Apolonski<br />

Ludwig-Maximilians-Universitaet Muenchen, Am Coulombwall 1, D-85748 Garching, Germany<br />

<strong>and</strong> Institute of Automation <strong>and</strong> Electrometry, Russian Academy of Sciences, 630090 Novosibirsk, Russia<br />

Received December 21, 2006; accepted February 1, 2007;<br />

posted February 13, 2007 (Doc. ID 78368); published April 3, 2007<br />

We report the first realization, to the best of our knowledge, of a chirped multilayer dielectric mirror providing<br />

dispersion control over the spectral range of 300–900 nm <strong>and</strong> the first use of hafnium oxide in a<br />

chirped mirror. The technology opens the door to the reliable <strong>and</strong> reproducible generation of monocycle laser<br />

pulses in the blue–violet spectral range, will benefit the development of optical waveform <strong>and</strong> frequencycomb<br />

synthesizers over the ultraviolet–visible–near-infrared spectral range, <strong>and</strong> permits the development of<br />

ultrabroadb<strong>and</strong>-chirped multilayers for high-power applications. © 2007 Optical Society of America<br />

OCIS codes: 320.5520, 320.7160, 310.1620.<br />

Advancing ultrashort laser pulse generation to the<br />

limit set by the oscillation cycle of light has been pursued<br />

ever since the discovery of lasers. Pulses comprising<br />

an ever-decreasing number of wave cycles allow<br />

more efficient exploitation of nonlinear optical<br />

effects 1 with implications as striking as the generation<br />

of single subfemtosecond light pulses. 2 Moreover,<br />

the controlled superposition of light frequencies extending<br />

over more than one octave opens, along with<br />

carrier-envelope phase control, 3–5 the way to shaping<br />

the subcycle (i.e., subfemtosecond) evolution of light<br />

fields in laser pulses <strong>and</strong> thereby to a new way of<br />

quantum control based on the light-force-directed<br />

charge in atoms, molecules, or solids. 6 In this Letter,<br />

we present a chirped multilayer mirror offering high<br />

reflectivity <strong>and</strong> controlled group-delay dispersion<br />

(GDD) over some 1.5 octaves spanning from ultraviolet<br />

(UV) to near-infrared (NIR) frequencies. This<br />

technology may become instrumental for the development<br />

of future ultrawideb<strong>and</strong> optical waveform<br />

synthesizers.<br />

There have been several approaches to generating<br />

monocycle optical pulses: (i) phase-coherent superposition<br />

of pulses from different laser sources 7,8 ; (ii)<br />

phase-coherent synthesis of Raman sideb<strong>and</strong>s by exploiting<br />

vibrational or rotational transitions in<br />

molecules 9,10 ; (iii) multicolor optical parametric<br />

generation 11 ; <strong>and</strong> last but not least, (iv) by means of<br />

supercontinuum generation based on self- or crossphase<br />

modulation. 12–14 What all these techniques<br />

have in common is that they rely on some optical device<br />

that can be used to adjust the phase (<strong>and</strong> amplitude)<br />

of individual groups of frequencies independently.<br />

This has, so far, been implemented by<br />

May 1, 2007 / Vol. 32, No. 9 / OPTICS LETTERS 1183<br />

separating the spectral components of a broadb<strong>and</strong><br />

signal in space 15 <strong>and</strong> addressing the spectral channels<br />

in the Fourier plane by a spatial light modulator<br />

(SLM) based on liquid crystals 15 or acoustic waves. 16<br />

The concept can, in principle, be extended to b<strong>and</strong>widths<br />

exceeding one octave, 17 however, the UV–IR<br />

absorption <strong>and</strong> the low damage threshold of SLMs restricts<br />

the technology from being scaled to b<strong>and</strong>widths<br />

of several octaves <strong>and</strong> ever higher pulse energies.<br />

In this Letter, we demonstrate that chirped<br />

multilayer mirrors may provide a promising alternative<br />

for specific applications, such as monocycle pulse<br />

generation or wideb<strong>and</strong> frequency-comb generation.<br />

Chirped mirrors (CM) have been continuously improved<br />

ever since their invention in 1994. 18 Progress<br />

has been made in terms of b<strong>and</strong>width, losses, GDD,<br />

<strong>and</strong> the ability to compensate higher-order spectral<br />

phase errors introduced by optical components. 19–27<br />

As a result of the efforts of several groups, by the<br />

turn of the millennium, CM-based optical systems<br />

have been capable of controlling broadb<strong>and</strong> radiation<br />

over spectral ranges approaching an octave in the<br />

visible–NIR domain. 12,13,28–30<br />

Recently, we demonstrated a chirped multilayer<br />

mirror supporting sub-3 fs pulses carried at a wavelength<br />

of 600 nm. 31 This technological progress paves<br />

the way toward the generation of monocycle light<br />

pulses, wideb<strong>and</strong> optical waveform synthesis, <strong>and</strong><br />

frequency-comb generation from compact, userfriendly<br />

laser systems. Many applications motivate<br />

the extension of these capabilities into the UV spectral<br />

range. The dramatically increasing dispersion of<br />

optical materials toward the UV absorption b<strong>and</strong>s<br />

prevented chirped dielectric multilayer technology<br />

0146-9592/07/091183-3/$15.00 © 2007 Optical Society of America<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 171


1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

Dispersion management for a sub-10-fs, 10 TW<br />

optical parametric chirped-pulse amplifier<br />

Franz Tavella, 1,4 Yutaka Nomura, 1 Laszlo Veisz, 1 Vladimir Pervak, 2 Andrius Marcinkevi~ius, 1,3,5 <strong>and</strong><br />

Ferenc Krausz 1,2<br />

1 Max-Planck-Institut für Quantenoptik, Hans-Kopfermann-Strasse 1, 85748 Garching, Germany<br />

2 Department für Physik, Ludwig-Maximilians-Universität München, am Coulombwall 1, 85748 Garching, Germany<br />

3 Present affiliation, IMRA America Inc., 1044 Woodridge Avenue, Ann Arbor, Michigan 48105, USA<br />

4 franz.tavella@mpq.mpg.de<br />

5 amarcink@imra.com<br />

Received March 19, 2007; revised June 6, 2007; accepted June 11, 2007;<br />

posted June 11, 2007 (Doc. ID 81025); published July 24, 2007<br />

We report the amplification of three-cycle, 8.5 fs optical pulses in a near-infrared noncollinear optical parametric<br />

chirped-pulse amplifier (OPCPA) up to energies of 80 mJ. Improved dispersion management in the<br />

amplifier by means of a combination of reflection grisms <strong>and</strong> a chirped-mirror stretcher allowed us to recompress<br />

the amplified pulses to within 6% of their Fourier limit. The novel ultrabroad, ultraprecise dispersion<br />

control technology presented in this work opens the way to scaling multiterawatt technology to even<br />

shorter pulses by optimizing the OPCPA b<strong>and</strong>width. © 2007 Optical Society of America<br />

OCIS codes: 190.4970, 190.4410, 320.5520.<br />

High-peak-power few-cycle sources are of interest for<br />

a number of applications in nonlinear optics, highfield,<br />

<strong>and</strong> ultrafast science [1]. Few-cycle pulses not<br />

only offer high peak powers from compact systems<br />

(relative value) but also enable the emergence of entirely<br />

new technologies such as the generation <strong>and</strong><br />

measurement <strong>and</strong> spectroscopic applications of isolated<br />

attosecond pulses <strong>and</strong> steering the atomic-scale<br />

motion of electrons with controlled light fields [2–7]<br />

(absolute value). Noncollinear optical parametric amplification<br />

offers amplification over spectral ranges<br />

sufficiently broad for few-cycle pulse synthesis, but<br />

dispersion control during stretching <strong>and</strong> recompression<br />

has remained a major challenge because of the<br />

large b<strong>and</strong>width over which the dispersion needs to<br />

be compensated with high accuracy. This is one of the<br />

reasons why many terawatt-scale optical chirpedpulse<br />

amplifiers [8,9] are designed for higher energy<br />

amplification <strong>and</strong> for pulse durations longer than<br />

those potentially allowed by the amplification b<strong>and</strong>width.<br />

Only recently amplification <strong>and</strong> adaptive<br />

pulse compression of more than 100-THz b<strong>and</strong>widths<br />

to the few-cycle regime were demonstrated reaching<br />

the terawatt regime [10–12].<br />

In this Letter we report the implementation of an<br />

ultrabroadb<strong>and</strong> grism-pair stretcher capable of controlling<br />

group delay over a dynamic range of tens of<br />

picoseconds <strong>and</strong> a b<strong>and</strong>width exceeding 100 THz.<br />

This improvement led to what we believe to be the<br />

first sub-10-fs, 10 TW light source ever reported <strong>and</strong><br />

holds promise for further shortening of multiterawatt<br />

light pulses by means of optical parametric amplifier<br />

(OPA) b<strong>and</strong>width engineering.<br />

The schematic layout of the our novel stretchercompressor<br />

system is depicted in Fig. 1. The OPA<br />

chain is described in previous work [12]. Stretching is<br />

implemented by the negative dispersion of a pair of<br />

grisms [13]. These hybrid elements combining the<br />

dispersive effects of diffraction gratings <strong>and</strong> prisms<br />

were recently demonstrated to be capable of introduc-<br />

August 1, 2007 / Vol. 32, No. 15 / OPTICS LETTERS 2227<br />

ing near-linear dispersion with high throughput over<br />

a b<strong>and</strong>width of 60 nm in the near infrared [13–15].<br />

Here we demonstrate that they can also control optical<br />

delay efficiently over a b<strong>and</strong>width of 300 nm in<br />

the same spectral range. Our positive-dispersion<br />

compressor is composed of glass blocks <strong>and</strong> chirped<br />

mirrors (CM2). The latter are used for final compression<br />

to prevent excessive nonlinear effects in the<br />

glass compressors. Residual higher-order dispersion<br />

of the system is compensated by another pair of<br />

chirped mirrors (CM1) <strong>and</strong> a programmable acoustooptic<br />

dispersive filter (AODF), dubbed Dazzler [16].<br />

The grism stretcher is designed to compensate for<br />

the dispersion of (1) the glass compressor (160 mm of<br />

SF57 <strong>and</strong> 100 mm of fused silica), (2) the BBO crystals<br />

in the optical parametric chirped-pulse amplifier<br />

(OPCPA) chain (a total of 15-mm path length), <strong>and</strong><br />

(3) the acousto-optical filter (43 mm of TeO 2) at the<br />

central wavelengths of 850 nm of our amplifier chain.<br />

Fig. 1. Optical layout of the OPCPA experimental setup.<br />

0146-9592/07/152227-3/$15.00 © 2007 Optical Society of America<br />

172 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008


that is responsible for such quantum-mechanical<br />

phenomena as electron diffraction through crystals.<br />

Likewise, the bound electron state possesses a<br />

quantum phase, which can change sign from<br />

place to place in the molecule. The quantum<br />

phases of the returning electron wave interfere<br />

with the different parts of the molecular bound<br />

state wave function, creating attosecond modulation<br />

of the electronic structure that is recorded<br />

in the shape of the HHG spectrum (Fig. 2).<br />

The physics of HHG is still under active<br />

discussion, <strong>and</strong> this coherent scattering picture<br />

has many subtleties that are still unresolved.<br />

Nonetheless, the effect of molecular shape on<br />

the HHG spectrum is genuine <strong>and</strong> dramatic. For<br />

example, two simple molecules O2 <strong>and</strong> N2, are<br />

similar enough in their gross structure, each<br />

consisting of two atoms separated by about an<br />

angstrom, <strong>and</strong> a difference of only two electrons<br />

out of more than a dozen; but their highest<br />

occupied orbitals, the ones that will field-ionize<br />

in an intense laser field, have very different symmetries.<br />

HHG spectra reveal this difference, <strong>and</strong><br />

recent experiments have used the spectra to perform<br />

tomographic reconstructions of the wave<br />

function responsible for ionization.<br />

Conclusion<br />

Only a few years ago, the direct measurement of<br />

transient phenomena lasting less than an optical<br />

cycle seemed a supreme challenge. Now scien-<br />

REVIEW<br />

1 . 3 . 3 S E L E C T E D R E P R I N T S<br />

tists have brought together techniques from atomic<br />

<strong>and</strong> laser physics to make these measurements<br />

possible, if not yet routine. Attoscience is a new<br />

field, with the promise to reveal some of the<br />

fastest processes of chemistry <strong>and</strong> atomic physics,<br />

or to freeze motion that will allow us to view<br />

the structure of matter under extreme conditions.<br />

The subject is still dominated by attempts to<br />

underst<strong>and</strong> <strong>and</strong> improve the sources, <strong>and</strong> to interpret<br />

the data, <strong>and</strong> there are still frontiers in<br />

source development: For example, attosecond<br />

pulses at subangstrom wavelengths have not been<br />

demonstrated. There are even more challenges<br />

in the theory of attosecond molecular dynamics:<br />

One of the most difficult issues yet to be resolved<br />

in attoscience is how we should describe<br />

the structure <strong>and</strong> motion of electrons <strong>and</strong> nuclei<br />

in physical chemistry <strong>and</strong> molecular physics to<br />

accommodate processes that are so fast that cherished<br />

concepts such as potential energy surfaces<br />

or the Born Oppenheimer approximation are no<br />

longer valid. Despite these difficulties, the rapid<br />

progress in attoscience over the past few years is<br />

not abating, <strong>and</strong> we may anticipate new physical<br />

insights in the coming decade.<br />

References <strong>and</strong> Notes<br />

1. W.-M. Yao et al., J. Phys. G: Nucl. Part. Phys. 33,<br />

1 (2006); http://meetings.aps.org/link/BAPS.2007.APR.<br />

B2.1.<br />

2. D. E. Spence, P. N. Kean, W. Sibbett, Opt. Lett. 16, 42<br />

(1991).<br />

<strong>Attosecond</strong> <strong>Control</strong> <strong>and</strong> <strong>Measurement</strong>:<br />

<strong>Lightwave</strong> <strong>Electronics</strong><br />

E. Goulielmakis, 1 V. S. Yakovlev, 2 A. L. Cavalieri, 1 M. Uiberacker, 2 V. Pervak, 2 A. Apolonski, 2<br />

R. Kienberger, 1 U. Kleineberg, 2 F. Krausz 1,2 *<br />

Electrons emit light, carry electric current, <strong>and</strong> bind atoms together to form molecules. Insight into<br />

<strong>and</strong> control of their atomic-scale motion are the key to underst<strong>and</strong>ing the functioning of biological<br />

systems, developing efficient sources of x-ray light, <strong>and</strong> speeding up electronics. Capturing <strong>and</strong><br />

steering this electron motion require attosecond resolution <strong>and</strong> control, respectively (1 attosecond =<br />

10 −18 seconds). A recent revolution in technology has afforded these capabilities: <strong>Control</strong>led light waves<br />

can steer electrons inside <strong>and</strong> around atoms, marking the birth of lightwave electronics. Isolated<br />

attosecond pulses, well reproduced <strong>and</strong> fully characterized, demonstrate the power of the new<br />

technology. <strong>Control</strong>led few-cycle light waves <strong>and</strong> synchronized attosecond pulses constitute its key tools.<br />

We review the current state of lightwave electronics <strong>and</strong> highlight some future directions.<br />

Quantum mechanics predicts the characteristic<br />

time scale for the rapidity of<br />

microscopic dynamics as Dt ~ ħ/DW,<br />

where DW is the spacing between the<br />

relevant energy levels of the microscopic system<br />

<strong>and</strong> ħ is Planck’s constant. The milli–electron<br />

volt <strong>and</strong> multi–electron volt energy spacing of<br />

vibrational <strong>and</strong> electronic energy levels, respectively,<br />

imply that structural dynamics of mole-<br />

cules <strong>and</strong> solids as well as related chemical<br />

reactions <strong>and</strong> phase transitions evolve on a femtosecond<br />

time scale, whereas electronic motion on<br />

the atomic scale is to be clocked in attoseconds.<br />

Before the invention of the laser, the resolution<br />

of time-resolved spectroscopy was limited<br />

by the nanosecond duration of pulses of incoherent<br />

light. The laser <strong>and</strong> the successive technological<br />

developments for the generation <strong>and</strong><br />

3. M. Nisoli, S. De Silvestri, O. Svelto, Appl. Phys. Lett. 68,<br />

2793 (1996).<br />

4. C. P. Hauri et al., Appl. Phys. B 79, 673 (2004).<br />

5. R. L. Fork, C. H. Brito Cruz, P. C. Becker, C. V. Shank,<br />

Opt. Lett. 12, 483 (1987).<br />

6. E. Matsubara et al., J. Opt. Soc. Am. B 24, 985<br />

(2007).<br />

7. M. Hentschel et al., Nature 414, 509 (2001).<br />

8. Ahmed Zewail received the 1999 Nobel Prize in<br />

Chemistry for his studies of the transition states of<br />

chemical reactions using femtosecond spectroscopy.<br />

9. S. Augst et al., Phys. Rev. Lett. 63, 2212 (1989).<br />

10. P. B. Corkum, Phys. Rev. Lett. 71, 1994 (1993).<br />

11 X. F. Li et al., Phys. Rev. A. 39, 5751 (1989).<br />

12. J. L. Krause et al., Phys. Rev. A. 45, 4998 (1992).<br />

13. K. J. Schafer, K. C. Kul<strong>and</strong>er, Phys. Rev. Lett. 78, 638<br />

(1997).<br />

14. R. López-Martens et al., Phys. Rev. Lett. 94, 033001<br />

(2005).<br />

15. N. M. Naumova, J. A. Nees, G. A. Mourou, Phys. Plasmas<br />

12, 056707 (2005).<br />

16. A. A. Zholents, G. Penn, Phys. Rev. ST Accel. Beams 8,<br />

050704 (2005).<br />

17. M. Lein, N. Hay, R. Velotta, J. P. Marangos, P. L. Knight,<br />

Phys. Rev. A. 66, 023805 (2002).<br />

18. J. Itatani et al., Nature 432, 867 (2004).<br />

19. T. Kanai, S. Minemoto, H. Sakai, Nature 435, 470<br />

(2005).<br />

20. This paper was written at the Stanford PULSE Center,<br />

with support from the NSF <strong>and</strong> from the Stanford Linear<br />

Accelerator Center, a national laboratory operated by<br />

Stanford University on behalf of the U.S. Department of<br />

Energy, Office of Basic Energy Sciences. We acknowledge<br />

useful discussions with M. Guehr, who also provided Fig. 2<br />

for this article.<br />

10.1126/science.1142135<br />

SPECIALSECTION<br />

measurement of ultrashort laser pulses improved<br />

the resolving power of pump-probe<br />

spectroscopy from several nanoseconds to several<br />

femtoseconds (1). The birth of femtosecond<br />

technology permitted real-time observation of the<br />

breakage <strong>and</strong> formation of chemical bonds (2).<br />

We review the recent developments in the<br />

optical technology that have led to the breaking<br />

of the femtosecond barrier <strong>and</strong> provided real-time<br />

access to intra-atomic electron dynamics. Consequences<br />

include the observation of electronic<br />

motion deep inside (i.e., in inner shells of ) atoms<br />

(3) <strong>and</strong> its control in real time (4). We address the<br />

underlying physical concepts <strong>and</strong> highlight the<br />

current status as well as future prospects of attosecond<br />

technology (5).<br />

Femtosecond Technology: <strong>Control</strong> <strong>and</strong><br />

<strong>Measurement</strong> with the Amplitude <strong>and</strong><br />

Frequency of Light<br />

<strong>Control</strong> <strong>and</strong> measurement of dynamics are<br />

intertwined. Time-resolved measurement relies<br />

on a physical quantity varying in a controlled,<br />

reproducible fashion on the relevant time scale.<br />

1 Max-Planck-Institut für Quantenoptik (MPQ), Hans-<br />

Kopfermann-Straße 1, D-85748 Garching, Germany.<br />

2 Department für Physik, Ludwig-Maximilians-Universität,<br />

Am Coulombwall 1, D-85748 Garching, Germany.<br />

*To whom correspondence should be addressed. E-mail:<br />

krausz@lmu.de<br />

www.sciencemag.org SCIENCE VOL 317 10 AUGUST 2007 769<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 173<br />

on November 12, 2007<br />

www.sciencemag.org<br />

Downloaded from


770<br />

<strong>Attosecond</strong> Spectroscopy<br />

Femtosecond technology is the result of controlling<br />

the nonlinear polarization of matter with the<br />

amplitude of light. <strong>Control</strong>ling the absorption<br />

<strong>and</strong>/or refractive index in this way has yielded—<br />

along with group delay dispersion control—<br />

femtosecond pulses from laser oscillators by means<br />

of passive mode locking (1). <strong>Measurement</strong> of<br />

the pulses relies on the same physical effect:<br />

control of the nonlinear polarization of matter<br />

by a replica of the pulse in the presence of the<br />

pulse to be characterized (6, 7).<br />

The controlled, well-characterized evolution<br />

of the amplitude envelope <strong>and</strong> carrier-frequency<br />

sweep (chirp) of ultrashort laser pulses permits<br />

measurement (2) <strong>and</strong> control (8) of quantum<br />

transitions on a femtosecond time scale. Reliance<br />

on these cycle-averaged quantities implies that<br />

measurement resolution <strong>and</strong> control speed are<br />

ultimately limited by the carrier wave cycle. The<br />

carrier wave cycle period is about 3 fs in the near<br />

infrared, where low dispersion favors the generation<br />

of the shortest laser pulses.<br />

Toward <strong>Attosecond</strong> Technology:<br />

X-ray–Pump/X-ray–Probe Spectroscopy?<br />

Femtosecond measurement <strong>and</strong> control techniques<br />

utilizing nonlinear material response could,<br />

in principle, be extended into the attosecond time<br />

domain by using intense attosecond pulses of extreme<br />

ultraviolet (XUV) or x-ray light. Unfortunately,<br />

in these regions of the optical spectrum, the<br />

probability of two-photon absorption is prohibitively<br />

low. X-ray–pump/x-ray–probe spectroscopy<br />

<strong>and</strong> x-ray quantum control therefore rely on x-ray<br />

intensities that can be attained only with large-scale<br />

free-electron lasers (9). Even though these sources<br />

are expected to eventually deliver their radiation<br />

in subfemtosecond pulses (10) <strong>and</strong> XUV<br />

sources pumped by large-scale, high energy<br />

lasers have already pushed the frontiers of<br />

nonlinear optics to the range of several tens of<br />

electron volts (11, 12), proliferation of attosecond<br />

technology <strong>and</strong> its widespread applications<br />

call for another approach—one that relies on light<br />

sources suitable for small laboratories.<br />

An Alternative Route: Light-Induced<br />

Electronic Motion Within the Wave<br />

Cycle of Light<br />

The electric field of visible, near-infrared or<br />

infrared (henceforth, referred to collectively as<br />

NIR) laser light, E L(t), exerts a force on electrons<br />

that varies on a subfemtosecond scale. The use of<br />

this gradient for initiating <strong>and</strong> probing the<br />

subsequent dynamics with attosecond timing<br />

precision <strong>and</strong> resolution has led to the emergence<br />

of an attosecond technology that does not rely on<br />

the existence of intense x-ray pulses.<br />

According to theory, strong-field–induced<br />

ionization of atoms is confined to subfemtosecond<br />

intervals near the peaks of the oscillating<br />

NIR field (Fig. 1A), setting a subfemtosecond<br />

electron wave packet yfree(t) free; this prediction<br />

1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

was recently confirmed by real-time observation<br />

(13). The subfemtosecond ionization process,<br />

locked to the peak of NIR field oscillation with<br />

corresponding precision, may serve as a subfemtosecond<br />

“starter gun” for a wide range of<br />

dynamics (including electronic <strong>and</strong> subsequent<br />

nuclear motion <strong>and</strong> related dynamics of molecular<br />

structure).<br />

The energetic radiation concurrent with opticalfield<br />

ionization in a linearly polarized laser field<br />

(14) provides a means of triggering motion in a<br />

more gentle way, without applying an intense<br />

field. The broadb<strong>and</strong> XUV light building up<br />

during the propagation of the driving pulse<br />

through an ensemble of atoms was predicted to<br />

emerge in a subfemtosecond burst within every<br />

half optical cycle that is sufficiently intense for<br />

ionization (15). In this Review, we focus on these<br />

photonic tools of attosecond technology, noting<br />

that the recolliding electron wave packet itself<br />

can also be used to explore dynamics (16–18).<br />

The electronic (<strong>and</strong> concurrent) phenomena<br />

following the attosecond trigger can be tracked<br />

with comparable resolution with the use of the<br />

generating NIR field, the oscillations of which are<br />

synchronized with the trigger event (tunnel<br />

ionization, recollision, or XUV burst from recollision).<br />

Let us consider an electron ejected during<br />

the unfolding evolution of the system under<br />

scrutiny. In the presence of a linearly polarized<br />

NIR field, outgoing electrons collected along<br />

the direction of the laser electric field EL(t)<br />

(Fig. 1B) suffer a momentum change DpðtÞ ¼<br />

e∫ ∞ t ELðt′Þdt′ ¼ eALðtÞ. Here, e is the charge of<br />

the electron, t is the instant of its emission, <strong>and</strong> A L<br />

(t) is the vector potential of the laser field. This<br />

change varies monotonically within a half wave<br />

cycle of the laser field, mapping the temporal<br />

evolution of the emitted subfemtosecond electron<br />

wave packet to a corresponding final momentum<br />

<strong>and</strong> energy distribution of the emitted electrons.<br />

We recognize here a new embodiment of<br />

the basic concept of streak imaging, with the<br />

streaking controlled by the NIR field in this case.<br />

The electrons can gain or lose an energy of<br />

several electron volts depending on whether they<br />

have been emitted some hundred attoseconds<br />

sooner or later than the peak of the field cycle<br />

(Fig. 1B), endowing this light-field–driven streak<br />

camera with attosecond resolution (19, 20).<br />

The physical concepts <strong>and</strong> mechanisms outlined<br />

above have opened a realistic prospect of<br />

measuring <strong>and</strong> controlling electron dynamics on<br />

an attosecond time scale. However, full exploitation<br />

of the potential offered by this conceptual<br />

framework requires precise control over the applied<br />

laser field.<br />

Steering Electrons with the Electric<br />

Field of <strong>Control</strong>led Light Waves<br />

In a laser pulse comprising many field cycles, the<br />

subfemtosecond electron <strong>and</strong> photon bursts born<br />

in the ionizing field emerge every half cycle,<br />

10 AUGUST 2007 VOL 317 SCIENCE www.sciencemag.org<br />

resulting in a train of subfemtosecond bursts (21).<br />

Its spectrum consists of equidistant lines that<br />

represent high-order odd harmonics of the<br />

driving field (22). The subfemtosecond pulse<br />

train <strong>and</strong> its multicycle driver constitute powerful<br />

tools for controlling <strong>and</strong> tracking electron<br />

dynamics that terminate within the NIR field<br />

oscillation cycle (23–25). The characteristics<br />

of the individual bursts in the train can be difficult<br />

to control <strong>and</strong> measure. Extension of attosecond<br />

control <strong>and</strong> metrology to a wide range of<br />

electronic phenomena (including all those extending<br />

in time over more than half an optical cycle)<br />

calls for the isolation of one to several attosecond<br />

pulses <strong>and</strong> precise control of their properties (26).<br />

Isolation was predicted to be achievable by<br />

manipulating the polarization state (polarization<br />

gating) of the driving field (27) or shortening its<br />

duration to nearly a single cycle of the carrier<br />

wave (28). In fact, intense few-cycle laser pulses<br />

lasting several femtoseconds (1) led to the first<br />

observation of light lasting for less than 1 fs (29).<br />

Full control of the number <strong>and</strong> properties of bursts<br />

isolated, however, requires precise control of the<br />

driving electric-field waveform E L(t).<br />

Intense few-cycle laser pulses with controlled<br />

waveform (4)—along with careful b<strong>and</strong>pass filtering<br />

of the highest-energy (cutoff) harmonics—<br />

have indeed permitted the reproducible generation<br />

of single <strong>and</strong> twin subfemtosecond pulses with<br />

cosine- <strong>and</strong> the sine-shaped electric field waveforms,<br />

respectively, as well as their reliable measurement<br />

by means of attosecond streaking (30, 31)<br />

(fig. S1). The series of streaked electron spectra<br />

shown in Fig. 1C yields complete information<br />

about the driving laser field <strong>and</strong> allows determination<br />

of the temporal shape, duration, <strong>and</strong> a<br />

possible chirp of the emitted subfemtosecond<br />

XUV pulse (Fig. 1D) as well as the degree of<br />

its synchronism with its driver (31–33). These<br />

experiments reveal that few-cycle light with<br />

controlled waveform permits reproducible generation<br />

<strong>and</strong> complete characterization of subfemtosecond<br />

light. Owing to the attosecond<br />

synchronism between them, these tools allow<br />

attosecond metrology without the need for highintensity<br />

x-rays.<br />

<strong>Lightwave</strong> <strong>Electronics</strong>: Versatile<br />

Technology for <strong>Control</strong> <strong>and</strong> Chronoscopy<br />

on the Electronic Time Scale<br />

<strong>Control</strong>led light fields permit control of microscopic<br />

electric currents at the atomic scale just as<br />

synthesized microwave fields permit control of<br />

currents at the mesoscopic scale in semiconductor<br />

chips. By analogy to microwave electronics, we<br />

propose to name this new technology lightwave<br />

electronics. In marked contrast with previous<br />

quantum control through controlling transitions<br />

between quantum states with the amplitude <strong>and</strong><br />

frequency of light (8), lightwave electronics gives<br />

way to controlling dynamics directly by the force<br />

that the electric field of intense light exerts on<br />

174 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008<br />

on November 12, 2007<br />

www.sciencemag.org<br />

Downloaded from


electrons. This new approach is powerful in that<br />

(i) it provides a direct way of affecting the position<br />

<strong>and</strong> momentum of electrons <strong>and</strong> (ii) control is<br />

matched in speed to the electronic time scale.<br />

The first notable manifestations of the<br />

power of lightwave electronics include controlling<br />

subfemtosecond XUV emission (4, 25, 31–35);<br />

molecular dissociation (36); measuring subfemtosecond<br />

XUV <strong>and</strong> electron pulses (21, 31–33, 35);<br />

<strong>and</strong> imaging dynamic changes in molecular<br />

structure (16–18) by means of steering electron<br />

wave packets with light fields. Sub–wave cycle<br />

(i.e., subfemtosecond) control of electron wave<br />

packet motion can be accomplished by mixing<br />

multicycle fields (25, 35). However, full control<br />

over the entire system evolution from attosecond<br />

to femtosecond time scales—<strong>and</strong>, consequently,<br />

precision attosecond measurements—relies on<br />

Fig. 1. The birth <strong>and</strong> measurement of a subfemtosecond (XUV) light<br />

pulse. (A) The field of a femtosecond NIR laser pulse, EL(t), is able to<br />

suppress the Coulomb potential in atoms sufficiently to allow a valence<br />

electron to tunnel through the narrow barrier <strong>and</strong> release a subfemtosecond<br />

wave packet, y free(t), near the peaks of its most intense oscillations.<br />

The wave packet is subsequently removed from the vicinity of<br />

the atomic core <strong>and</strong> less than a period later pushed back by the reversed<br />

field (green trajectory). Upon recollision, it interferes with the boundstate<br />

portion of the electron wave function. This interference results in<br />

high-frequency oscillations, emitting broadb<strong>and</strong> XUV light. The highestfrequency<br />

portion, with intensity IX(t), is temporally confined to a small<br />

fraction of the optical period. (B) Concept of optical-field–driven streak<br />

imaging of electron emission from atoms. Electrons released by an XUV<br />

pulse parallel to the direction of electric field (red line) suffer a change in<br />

their initial momenta that is proportional to the vector potential of the<br />

field (black line) at the instant of release, mapping the intensity profile of<br />

1 . 3 . 3 S E L E C T E D R E P R I N T S<br />

optical waveforms that can be accurately controlled<br />

in shape <strong>and</strong> polarization state (4, 37)<br />

<strong>and</strong> confined to several cycles. Exploitation of<br />

the full potential of lightwave electronics for<br />

attosecond control <strong>and</strong> metrology requires<br />

waveform-controlled broadb<strong>and</strong> light.<br />

The attosecond streaking spectrogram shown<br />

in Fig. 1C yields the complete history of photoelectron<br />

emission induced by a subfemtosecond<br />

XUV pulse as well as of the streaking optical field<br />

(38). Similar spectrograms can also be recorded<br />

with secondary (Auger) electrons, which uncover<br />

the history of a quantum transition of an electron<br />

deep inside an atom in real time (3). <strong>Attosecond</strong><br />

streaking is not the only way of observing the<br />

temporal evolution of atomic-scale electron motion.<br />

Energetic excitation of atoms, molecules, or<br />

solids is often accompanied by the promotion of<br />

SPECIALSECTION<br />

valence electrons to excited states (referred to as a<br />

shake up). The transient population of these states<br />

provides information about the instantaneous<br />

(electronic) state of the system. This population<br />

can be probed by means of attosecond tunneling<br />

induced by the time-delayed waveform-controlled<br />

few-cycle laser field (13).<br />

In attosecond streaking <strong>and</strong> tunneling spectroscopy,<br />

the subfemtosecond XUV pulse serves<br />

as a pump <strong>and</strong> the controlled optical wave as the<br />

probe. Their roles can be interchanged: The optical<br />

field may—by means of tunnel ionization<br />

(13)—initiate an electronic process <strong>and</strong> control<br />

the unfolding dynamics (36), <strong>and</strong> the subfemtosecond<br />

XUV pulse may probe this process by, for<br />

example, photoelectron spectroscopy (39).<br />

<strong>Attosecond</strong> technology based on a controlled<br />

optical wave <strong>and</strong> a synchronized subfemtosecond<br />

the emitted electron <strong>and</strong> hence of the ionizing subfemtosecond XUV<br />

pulse into a corresponding final momentum <strong>and</strong> energy distribution of<br />

electrons. (C) Streaked spectra of photoelectrons released from neon<br />

atoms by a single subfemtosecond XUV pulse (ħw X ≈ 95 eV) recorded for<br />

a series of delays between the XUV pulse <strong>and</strong> NIR field (streaking<br />

spectrogram). The laser field causes only a moderate broadening<br />

(streaking) of the electron spectra; its main effect is the shift of the<br />

center of mass of the electron spectrum. In the limit of large initial<br />

kinetic energy of the electrons, this shift is linearly proportional to the<br />

vector potential of the field at the instant of impact of the XUV pulse (see<br />

white line). (D) Electric field of the NIR wave <strong>and</strong> intensity of the XUV<br />

pulse as retrieved from the streaking spectrogram shown in (C). The<br />

measurement confirms that the few-cycle waveform must be near cosineshaped<br />

to warrant the production of a single subfemtosecond pulse <strong>and</strong><br />

reveals a near-Fourier–limited XUV pulse of a duration of 250 attoseconds.<br />

arb. u., arbitrary units.<br />

www.sciencemag.org SCIENCE VOL 317 10 AUGUST 2007 771<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 175<br />

on November 12, 2007<br />

www.sciencemag.org<br />

Downloaded from


772<br />

<strong>Attosecond</strong> Spectroscopy<br />

pulse is both versatile <strong>and</strong> sufficiently compact<br />

<strong>and</strong> affordable to allow proliferation in small<br />

laboratories. The core infrastructure required is<br />

sketched in Fig. 2. The scope of the technology<br />

largely depends on the characteristics <strong>and</strong> flexibility<br />

of its key tools: synthesized fields of laser<br />

light <strong>and</strong> attosecond pulses synchronized to them.<br />

From <strong>Control</strong>led Light Waveforms Toward Optical<br />

Waveform Synthesis<br />

<strong>Lightwave</strong> electronics benefits from ever-broader<br />

optical b<strong>and</strong>width in several ways. Superposition<br />

of spectral components beyond an octave permits<br />

subcycle shaping of the light waveform<br />

<strong>and</strong> hence sculpting of the electric force on the<br />

electronic time scale for steering electrons in<br />

atomic systems. The increased b<strong>and</strong>width <strong>and</strong><br />

improved dispersion control also lead to shorter<br />

optical pulses, which allow advancing subfemtosecond<br />

XUV pulse generation<br />

in terms of all relevant parameters:<br />

duration, intensity, <strong>and</strong> photon<br />

energy.<br />

Intense few-cycle optical<br />

pulses (~5 fs) were first generated<br />

(40) by spectral broadening of<br />

femtosecond pulses (initially ~25<br />

fs) in a hollow-core waveguide,<br />

<strong>and</strong> the resultant supercontinuum<br />

was compressed by reflection off<br />

of chirped multilayer dielectric<br />

mirrors (41) (Fig. 3A). Combined<br />

with waveform control (4), the<br />

technology matured to become a<br />

workhorse for attosecond science.<br />

Its current state is represented by<br />

the results summarized in Fig. 3, B<br />

<strong>and</strong> C. The spectrum of near–20-fs,<br />

800-nm pulses is first broadened<br />

by self-phase modulation in a gasfilled<br />

capillary to a supercontin-<br />

uum reaching from 400 to 1000<br />

nm, followed by compression of<br />

the ~ 400-μJ pulses to ~3.5 fs<br />

with octave-spanning chirped mirrors<br />

(42). These pulses constitute<br />

the shortest controlled optical<br />

waveforms demonstrated to date.<br />

Figure 3D shows how a<br />

single subfemtosecond XUV<br />

pulse can emerge from an atom<br />

ionized by a linearly polarized<br />

few-cycle light field. The green<br />

lines depict the kinetic energy,<br />

W kin, with which different portions<br />

of the freed electron wave packet (y free in<br />

Fig. 1B) return to the core as a function of return<br />

time, determining the energy of the emitted<br />

harmonic photon, ħw = Wkin + Wb, where Wb is<br />

the binding energy of the electron. For the<br />

optimum (near-cosine) waveform (Fig. 3D), we<br />

can estimate—in the limit of ħw X >> W b—the<br />

b<strong>and</strong>width over which photons of the highest<br />

1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

energy (near the cutoff of the emitted spectrum<br />

depicted in violet) are emitted from one recollision<br />

only as<br />

DWsingle�pulse<br />

176 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008<br />

ℏwX<br />

≈ 0:6<br />

Tosc<br />

TFWHM<br />

2<br />

ð1Þ<br />

where T osc <strong>and</strong> T FWHM st<strong>and</strong> for the oscillation<br />

period (osc) of the carrier wave <strong>and</strong> the full width<br />

at half maximum (FWHM) duration of the<br />

laser pulse ionizing the atoms, respectively<br />

(43). Equation 1 underscores the importance of<br />

minimizing the number of cycles within the<br />

driver pulse. The intense sub–1.5 cycle (TFWHM<br />

< 1.5Tosc) laser pulses that can now be produced<br />

with commercially available ingredients of ultrafast<br />

laser technology (Fig. 3, A to C) constitute a<br />

promising tool for pushing the frontiers of<br />

attosecond-pulse technology.<br />

Advancing femtosecond technology to its<br />

ultimate limit is the key to extending the frontiers<br />

of both measuring <strong>and</strong> controlling dynamics with<br />

attosecond precision. Metrology benefits from<br />

broader spectral coverage <strong>and</strong> shorter duration of<br />

subfemtosecond pulses. The scope of control critically<br />

depends on the variety of waveforms, which<br />

is determined by the available b<strong>and</strong>width of intense<br />

coherent radiation. The shortest intense laser pulses<br />

also provide ideal conditions for generating the<br />

broadest supercontinua of coherent light, extending<br />

over more then three octaves (44). Advanced<br />

chirped multilayer mirror technology holds promise<br />

for versatile <strong>and</strong> scalable (multichannel) implementation<br />

of optical waveform synthesis. A<br />

prototypical three-channel synthesizer operational<br />

over the 1- to 3-eV b<strong>and</strong> indicates the vast variety<br />

of possibilities for steering electrons in atomic<br />

systems with synthesized optical waves (fig. S2).<br />

From Subfemtosecond Toward<br />

<strong>Attosecond</strong> Pulses<br />

Equation 1 suggests that femtosecond technology<br />

must approach the monocycle limit to push the<br />

frontiers of XUV pulse generation toward <strong>and</strong><br />

below 100 as. However, there are other options.<br />

Modulation of the polarization state of the ioniz-<br />

Fig. 2. Schematic of an experimental setup for attosecond-pulse generation <strong>and</strong> attosecond metrology as well as<br />

spectroscopy: the AS-1 attosecond beamline at MPQ. The intense, waveform-controlled few-cycle NIR laser pulse<br />

generates a subfemtosecond XUV pulse in the first interaction medium (jet of noble gas). The collinear XUV <strong>and</strong> NIR<br />

beams then propagate into a second vacuum chamber, where they are focused by a two-component XUV multilayer<br />

mirror into a gas target. The inner <strong>and</strong> outer part of the two-component mirror reflects <strong>and</strong> focuses the XUV <strong>and</strong> the<br />

(more divergent) NIR beam, respectively. By positioning the internal mirror with a nanometer-precision piezotranslator,<br />

the XUV pulse can be delayed with respect to the NIR pulse with attosecond accuracy. Analysis of the generated<br />

electrons <strong>and</strong>/or ions as a function of delay between the subfemtosecond XUV <strong>and</strong> the waveform-controlled NIR pulse<br />

permits characterization of the attosecond tools (Fig. 1) as well as real-time observation of atomic-scale electron<br />

dynamics in all forms of matter by means of attosecond streaking spectroscopy (3) <strong>and</strong> attosecond tunneling<br />

spectroscopy (13) with the same apparatus. In the former case, the final energy distribution of the outgoing electron is<br />

analyzed with a time-of-flight spectrometer; in the latter case, the observables measured as a function of delay are<br />

multiple-charged ions. ACF, auto-correlation function; CCD, charged-coupled device.<br />

10 AUGUST 2007 VOL 317 SCIENCE www.sciencemag.org<br />

ing NIR pulse (33, 37, 45) or of its amplitude by<br />

coherent addition of second harmonic radiation<br />

(46, 47) can efficiently increase DWsingle-pulse<br />

even by using several-cycle NIR driver pulses.<br />

Implementation of polarization gating with<br />

waveform-controlled, approximately two-cycle<br />

NIR pulses (37) has recently led to a spectacular<br />

achievement: the generation of isolated XUV<br />

on November 12, 2007<br />

www.sciencemag.org<br />

Downloaded from


pulses at ħwcarrier ≈ 36 eV with a b<strong>and</strong>width of<br />

DW single-pulse ≈ 15 eV. Along with dispersion<br />

control <strong>and</strong> trajectory selection (48), these isolated<br />

pulses resulted in near–single-cycle XUV pulses<br />

that had a duration of 130 as (33) (see the lowenergy<br />

streaking spectrogram in Fig. 4A).<br />

The increased DWsingle-pulse with several-cycle<br />

pulses comes, however, at the expense of efficiency<br />

because (i) only a small temporal fraction of the<br />

driver pulse participates in the generation process<br />

1 . 3 . 3 S E L E C T E D R E P R I N T S<br />

<strong>and</strong> (ii) the cycles preceding the generation moment<br />

pre-ionize the atoms, causing substantial<br />

depletion of the ground state before the XUV pulse<br />

can be emitted. The unprecedented confinement of<br />

electromagnetic energy into a single wellcontrolled<br />

oscillation of light in sub–1.5-cycle<br />

NIR pulses (Fig. 3) avoids these shortcomings <strong>and</strong><br />

offers several more benefits. Indeed, the high<br />

contrast between the wave crests at t = –Tosc/2 <strong>and</strong><br />

earlier ones in Fig. 3D permits—thanks to the<br />

SPECIALSECTION<br />

exponential scaling of ionization probability with<br />

the field strength—adjustment of the pulse intensity<br />

so as to allow the electron to survive with a<br />

high probability in its ground state until t = –T osc/2<br />

(in Fig. 3D) <strong>and</strong> then to be set free with a high<br />

probability near this instant through tunneling. The<br />

resulting wave packet, yfree (Fig. 1A), is launched<br />

with unprecedented amplitude. Upon its return to<br />

the core at about t ≈ Tosc/4, it interferes with the<br />

ground state portion of the wave function causing<br />

Fig. 3. Toward intense, monocycle optical pulses.<br />

(A) Hollow-fiber/chirped-mirror optical pulse compressor<br />

for the generation of intense, few-cycle<br />

laser pulses. (B) The spectrum of millijoule-energy,<br />

near–20-fs, phase-controlled near-infrared laser<br />

pulses delivered at a repetition rate of 3 kHz from<br />

a Ti:sapphire laser (Femtopower, Femtolasers<br />

GmbH) is first broadened to more than an octave<br />

(~450 to 950 nm) in a neon-filled capillary<br />

(pressure ~ 2 bar, length ~ 1 m, bore diameter<br />

250 mm). (C) The spectrally broadened pulses are<br />

compressed by octave-spanning chirped multilayer<br />

mirrors. The interferometric autocorrelation trace<br />

indicates a near–b<strong>and</strong>width-limited ~3.5-fs pulse<br />

carried at lL ≈ 720 nm. (D) Energies of XUV<br />

photons emitted from an atom exposed to a<br />

linearly polarized 3.5-fs, 720-nm Gaussian laser<br />

pulse that is sufficiently intense for efficient tunnel<br />

ionization. The photon energies (green lines) are<br />

depicted as a function of the return time of the<br />

electron to the atomic core <strong>and</strong> have been<br />

obtained by analyzing classical electron trajectories<br />

(43). The width of the energy b<strong>and</strong> within<br />

which only one recollision contributes to XUV<br />

emission determines the b<strong>and</strong>width DW single-pulse<br />

available for the generation of an isolated attosecond<br />

pulse. Based on the classical trajectory analysis,<br />

DWsingle-pulse is found to be maximum for a near–<br />

cosine-shaped waveform, E L(t) = E 0e(t)cos(w Lt +<br />

ϕ), with the carrier-envelope phase ϕ varying<br />

between p/12 <strong>and</strong> p/6 depending on the pulse<br />

shape <strong>and</strong> intensity. Here, E 0 <strong>and</strong> w L st<strong>and</strong> for the<br />

amplitude <strong>and</strong> angular frequency of the oscillations<br />

of the laser electric field, with e(t) being<br />

the amplitude envelope function. The table summarizes<br />

b<strong>and</strong>widths over which the sub–1.5-cycle<br />

pulses are predicted to support the emergence of<br />

an isolated attosecond pulse at different XUV<br />

photon energies. The estimated pulse durations<br />

derive from the conservative assumption that no<br />

spectral components are available for the pulse<br />

synthesis outside the given spectral window.<br />

www.sciencemag.org SCIENCE VOL 317 10 AUGUST 2007 773<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 177<br />

on November 12, 2007<br />

www.sciencemag.org<br />

Downloaded from


774<br />

<strong>Attosecond</strong> Spectroscopy<br />

high-frequency (soft–x-ray) dipole oscillations with<br />

unprecedented amplitude. Hence, the intense near–<br />

single-cycle waveforms appear to constitute ideal<br />

tools for the pursuit of powerful soft–x-ray pulses<br />

with pulse durations approaching the atomic unit<br />

of time (see table in Fig. 3D) <strong>and</strong> photon energies<br />

approaching the kilo–electron volt frontier.<br />

A first indication of the potential that near–<br />

single-cycle NIR pulses offer for attosecond-pulse<br />

generation is provided by the high-energy streaking<br />

spectrogram in Fig. 4A. It has been recorded<br />

with sub–two-cycle (near–4-fs) NIR (750-nm)<br />

pulses <strong>and</strong> XUV pulses filtered by a molybdensilicon<br />

chirped multilayer mirror with a highreflectivity<br />

b<strong>and</strong> of ~16 eV (FWHM) centered at<br />

~93 eV (Fig. 4B). The b<strong>and</strong>width (FWHM) of the<br />

b<strong>and</strong>pass filtered XUV light amounts to ~13 eV.<br />

The streaking spectrogram indicates isolated sub–<br />

Fig. 4. Current frontiers of attosecond technology.<br />

(A) <strong>Attosecond</strong> streaking spectrograms recorded<br />

with few-cycle NIR pulses (l L ≈ 750 nm). The lowenergy<br />

spectrogram (bottom) shows the carrier<br />

photon energy of an XUV pulse, ħwX ≈ 36 eV, with<br />

argon target atoms (courtesy of M. Nisoli). The<br />

high-energy spectrogram (top) shows ħw X ≈ 93 eV,<br />

with neon target atoms (49). (B) Computed<br />

reflectivity <strong>and</strong> group delay of the chirped Mo/Si<br />

multilayer mirror used for filtering <strong>and</strong> focusing the<br />

isolated sub–170-as, 93-eV pulses. (C) Computed<br />

reflectivity <strong>and</strong> group delay of an ultrabroadb<strong>and</strong><br />

chirped multilayer Mo/Si mirror designed for compensating<br />

the chirp carried by an attosecond pulse<br />

filtered near cutoff from short-trajectory emission<br />

(group delay dispersion ~ –0.007 fs 2 ) <strong>and</strong> for reflecting<br />

a b<strong>and</strong> sufficient for sub–100-as pulse<br />

generation (b<strong>and</strong>width > 35 eV). (D) Spectrum of<br />

high-order harmonics emitted from neon ionized<br />

by ~4-fs–duration NIR (~720-nm) pulses near<br />

cutoff as transmitted through a high-pass filter<br />

(150-nm palladium). The transmission of the Pd<br />

foil is shown by the red curve. The spectrum has<br />

been recorded with the carrier-envelope phase ϕ<br />

of the waveform-controlled NIR pulse set to<br />

provide the broadest continuum. Variation of ϕ<br />

reshapes the overall distribution of the continuum<br />

(42) rather than introducing a pronounced<br />

harmonic-like structure, as observed with severalcycle<br />

driver pulses for appropriate phase setting<br />

(dashed line). The smooth continuum with a<br />

b<strong>and</strong>width of >30 eV suggests the feasibility of<br />

sub–100-as XUV pulse generation without manipulation<br />

of the driving pulse (e.g., superimposing<br />

its second harmonic or gating its polarization).<br />

1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

170-as pulses that are near b<strong>and</strong>width-limited<br />

(49). The pulses contain more than 10 6 photons,<br />

which are delivered at a repetition rate of 3 kHz,<br />

implying a photon flux of >3 × 10 9 photons per<br />

second transported in a near–diffraction-limited<br />

beam. With the Mo/Si mirror, which has a 10-cm<br />

focal length, used in the MPQ attosecond beamline<br />

AS-1 (Fig. 2), this beam can be focused to a<br />

spot diameter of several micrometers.<br />

The recently demonstrated, waveformcontrolled<br />

sub–1.5-cycle NIR pulses are capable<br />

of producing DWsingle-pulse > 30 eV at ħwX ≈ 120<br />

eV (Fig. 4D), offering the potential for generating<br />

sub–100-as pulses at a wavelength of ~10 nm (42).<br />

To exploit this potential, researchers must develop<br />

ultrabroadb<strong>and</strong> soft–x-ray multilayer mirrors with<br />

precisely controlled group-delay (GD) dispersion.<br />

Figure 4C plots the reflectivity <strong>and</strong> GD of the first<br />

10 AUGUST 2007 VOL 317 SCIENCE www.sciencemag.org<br />

chirped multilayer mirror design that supports<br />

sub–100-as pulses. Chirped multilayer mirror<br />

technology (both metallic <strong>and</strong> dielectric) is likely<br />

to play as important a role in advancing attosecond<br />

technology as it has played in advancing femtosecond<br />

technology to its ultimate limits.<br />

Prospects<br />

New research tools allow scientists to seek answers<br />

to questions that, in the absence of a means<br />

of addressing them, have never been posed<br />

before. Gr<strong>and</strong> questions serve as a compass<br />

providing directions for research that promises<br />

the most extensive return. We pose a few<br />

questions across disparate areas of science, connected<br />

by the fundamental role that atomic-scale<br />

electronic motion plays in physical, chemical, <strong>and</strong><br />

biological processes.<br />

178 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008<br />

on November 12, 2007<br />

www.sciencemag.org<br />

Downloaded from


Atomic physics <strong>and</strong> x-ray science. How is the<br />

energy of an x-ray photon distributed between<br />

electrons upon its absorption by an atom? Can<br />

electronic transitions deep inside atoms be<br />

affected by controlled ultrastrong external fields<br />

rivaling in strength the internal Coulomb fields,<br />

e.g., for opening up novel routes to efficient, compact<br />

x-ray lasers?<br />

Physical chemistry, molecular biology, bioinformatics,<br />

<strong>and</strong> photovoltaics. Can controlled<br />

light fields offer a fundamentally new way of<br />

modifying the structure <strong>and</strong>/or composition of<br />

molecules by driving electron wave packets across<br />

molecules with synthesized optical fields? What<br />

are the microscopic mechanisms underlying biological<br />

information transport? Can charge-transfer<br />

in host-guest systems (e.g., dye-semiconductor<br />

assemblies) be exploited for developing solar cells<br />

with unprecedented efficiency?<br />

Information technology. Can electron-based<br />

information processing <strong>and</strong> storage be downscaled<br />

to atomic dimensions <strong>and</strong> sped up to the<br />

atomic time scale (i.e., to optical frequencies)?<br />

Can these ultimate limits be reached by exploiting<br />

electric interactions (electronics) or magnetic<br />

interactions (spintronics) or collective electron<br />

motion (plasmonics)? Which incarnation of lightwave<br />

electronics will be the ultimate electronbased<br />

information technology?<br />

The answers to these questions will rely on<br />

exploring <strong>and</strong> controlling the microscopic motion<br />

of electrons, on atomic scales in space <strong>and</strong> time.<br />

<strong>Attosecond</strong> technology now offers the tools for<br />

tackling these <strong>and</strong> many other exciting questions.<br />

The importance of the answers being sought will<br />

drive its proliferation.<br />

References <strong>and</strong> Notes<br />

1. T. Brabec, F. Krausz, Rev. Mod. Phys. 72, 545 (2000).<br />

2. A. H. Zewail, J. Phys. Chem. A 104, 5660 (2000).<br />

1 . 3 . 3 S E L E C T E D R E P R I N T S<br />

3. M. Drescher et al., Nature 419, 803 (2002).<br />

4. A. Baltuska et al., Nature 421, 611 (2003).<br />

5 For a comprehensive historical review, see P. Agostini,<br />

L. F. DiMauro, Rep. Prog. Phys. 67, 813 (2004).<br />

6. R. Trebino et al., Rev. Sci. Instrum. 68, 3277 (1997).<br />

7. C. Iaconis, I. A. Walmsley, IEEE J. Quantum Electron. 35,<br />

501 (1999).<br />

8. P. W. Brumer, M. Shapiro, Principles of the Quantum<br />

<strong>Control</strong> of Molecular Processes (Wiley, New York, 2003).<br />

9. W. Ackermann et al., Nature Phot. 1, 336 (2007).<br />

10. A. A. Zholents, W. M. Fawley, Phys. Rev. Lett. 92, 224801<br />

(2004).<br />

11. P. Tzallas, D. Charalambidis, N. A. Papadogiannis, K. Witte,<br />

G. Tsakiris, Nature 426, 267 (2003).<br />

12. T. Sekikawa, A. Kosuge, T. Kanai, S. Watanabe, Nature<br />

432, 605 (2004).<br />

13. M. Uiberacker et al., Nature 446, 627 (2007).<br />

14. P. B. Corkum, Phys. Rev. Lett. 71, 1994 (1993).<br />

15. P. Antoine, A. L’Huillier, M. Lewenstein, Phys. Rev. Lett.<br />

77, 1234 (1996).<br />

16. H. Niikura et al., Nature 417, 917 (2002).<br />

17. T. Kanai, S. Minemoto, H. Sakai, Nature 435, 470<br />

(2005).<br />

18. S. Baker et al., Science 312, 424 (2006).<br />

19. The first implementation of the basic concept of a<br />

light-field–driven streak camera has drawn on an<br />

orthogonal detection geometry (electrons collected<br />

along a direction orthogonal to the electric field vector<br />

of the streaking NIR field (29).<br />

20. The laser field needed to induce this change in electron<br />

energy is orders of magnitude less intense than that<br />

required for strong-field ionization. Hence, this<br />

streaking field has negligible influence on the atomic<br />

or molecular processes under study, unless its oscillations<br />

happen to be in resonance with a transition from an<br />

occupied to an unoccupied quantum state of the system.<br />

21. P. M. Paul et al., Science 292, 1689 (2001).<br />

22. M. Ferray et al., J. Phys. B 21, L31 (1988).<br />

23. K. J. Schafer, M. B. Gaarde, A. Heinrich, J. Biegert,<br />

U. Keller, Phys. Rev. Lett. 92, 023003 (2004).<br />

24. T. Remetter et al., Nature Phys. 2, 323 (2006).<br />

25. J. Mauritsson et al., Phys. Rev. Lett. 97, 013001 (2006).<br />

26. Experiments that can be performed with subfemtosecond<br />

pulse trains [including those described in (23–25)]<br />

would benefit from using a small number of wellcontrolled<br />

<strong>and</strong> characterized subfemtosecond pulses.<br />

27. P. B. Corkum, N. H. Burnett, M. Yu. Ivanov, Opt. Lett. 19,<br />

1870 (1994).<br />

REVIEW<br />

Harnessing <strong>Attosecond</strong> Science in the<br />

Quest for Coherent X-rays<br />

Henry Kapteyn, Oren Cohen, Ivan Christov, Margaret Murnane*<br />

Modern laser technology has revolutionized the sensitivity <strong>and</strong> precision of spectroscopy by<br />

providing coherent light in a spectrum spanning the infrared, visible, <strong>and</strong> ultraviolet wavelength<br />

regimes. However, the generation of shorter-wavelength coherent pulses in the x-ray region<br />

has proven much more challenging. The recent emergence of high harmonic generation techniques<br />

opens the door to this possibility. Here we review the new science that is enabled by an ability to<br />

manipulate <strong>and</strong> control electrons on attosecond time scales, ranging from new tabletop sources of<br />

coherent x-rays to an ability to follow complex electron dynamics in molecules <strong>and</strong> materials. We<br />

also explore the implications of these advances for the future of molecular structural<br />

characterization schemes that currently rely so heavily on scattering from incoherent x-ray sources.<br />

Next year, 2008, will mark the 50th anniversary<br />

of the revolutionary paper by<br />

Schawlow <strong>and</strong> Townes that proposed the<br />

laser (1). This paper extended concepts first used to<br />

demonstrate the maser in the microwave region of<br />

the spectrum into the visible spectrum. Soon after<br />

the laser was demonstrated, scientists discovered<br />

how to control laser light to generate extremely<br />

28. I. P. Christov, M. M. Murnane, H. C. Kapteyn, Phys. Rev.<br />

Lett. 78, 1251 (1997).<br />

29. M. Hentschel et al., Nature 414, 509 (2001).<br />

30. J. Itatani et al., Phys. Rev. Lett. 88, 173903 (2002).<br />

31. R. Kienberger et al., Nature 427, 817 (2004).<br />

32. E. Goulielmakis et al., Science 305, 1267 (2004).<br />

33. G. Sansone et al., Science 314, 443 (2006).<br />

34. I. P. Christov, R. Bartels, H. C. Kapteyn, M. M. Murnane,<br />

Phys. Rev. Lett. 86, 5458 (2001).<br />

35. N. Dudovich et al., Nature Phys. 2, 781 (2006).<br />

36. M. F. Kling et al., Science 312, 246 (2006).<br />

37. I. J. Sola et al., Nature Phys. 2, 319 (2006).<br />

38. F. Quéré, Y. Mairesse, J. Itatani, J. Mod. Opt. 52, 339 (2005).<br />

39. A. D. B<strong>and</strong>rauk, S. Chelkowski, H. S. Nguyen, Int. J.<br />

Quantum Chem. 100, 834 (2004).<br />

40. M. Nisoli et al., Opt. Lett. 22, 522 (1997).<br />

41. R. Szipöcs, K. Ferencz, C. Spielmann, F. Krausz, Opt. Lett.<br />

19, 201 (1994).<br />

42. A. L. Cavalieri et al., New J. Phys. 9, 242 (2007).<br />

43. The electron kinetic energies have been calculated in<br />

terms of the classical description of the center-of-mass<br />

motion of the freed electron wave packet (Fig. 2), under<br />

the assumption of a Gaussian pulse shape <strong>and</strong> by using<br />

the strong-field approximation. The factor of ~0.6 in<br />

Eq. 1 depends on the pulse shape <strong>and</strong> intensity but varies<br />

less than 15% in the relevant parameter range.<br />

44. N. Aközbek et al., New J. Phys. 8, 177 (2006).<br />

45. Z. Chang, Phys. Rev. A. 70, 043802 (2004).<br />

46. T. Pfeifer et al., Phys. Rev. Lett. 97, 163901 (2006).<br />

47. Y. Oishi, M. Kaku, A. Suda, F. Kannari, K. Midorikawa,<br />

Opt. Exp. 14, 7230 (2006).<br />

48. R. Lopez-Martens et al., Phys. Rev. Lett. 94, 033001 (2005).<br />

49. M. Schultze et al., New J. Phys. 9, 243 (2007).<br />

50. We apologize that many original research papers could not<br />

be cited because of space limitations. This work is<br />

supported by the Max-Planck-Society <strong>and</strong> the Deutsche<br />

Forschungsgemeinschaft through the DFG Cluster of<br />

Excellence Munich-Centre for Advanced Photonics<br />

(www.munich-photonics.de). E.G. acknowledges support<br />

from a Marie Curie Intra-European Fellowship. We thank B.<br />

Ferus, M. Hofstätter, B. Horvath, <strong>and</strong> M. Schultze for their<br />

support in the preparation of this manuscript.<br />

Supporting Online Material<br />

www.sciencemag.org/cgi/content/full/317/5839/769/DC1<br />

Figs. S1 <strong>and</strong> S2<br />

References<br />

10.1126/science.1142855<br />

SPECIALSECTION<br />

short nanosecond, picosecond, <strong>and</strong> even femtosecond<br />

pulses. Given the origin of the laser, it was<br />

also natural to attempt to generate coherent light at<br />

shorter <strong>and</strong> shorter wavelengths. However, this<br />

effort proved very challenging because of the punishing<br />

power scaling inherent in lasers. Basic physics<br />

dictates that the energy required to implement a<br />

laser scales roughly as 1/l 5 ; that is, a laser at a 10<br />

times shorter wavelength (l) requires ~100,000<br />

times the input power. Thus, the first x-ray lasers<br />

implemented in the 1980s used the building-sized<br />

Nova fusion laser at Lawrence Livermore National<br />

Laboratory as a power source to generate soft (relatively<br />

long-wavelength) x-rays. Since that initial<br />

x-ray laser, considerable progress has been made in<br />

downscaling the laser needed as the power source<br />

JILA <strong>and</strong> the National Science Foundation Center for<br />

Extreme Ultraviolet Science <strong>and</strong> Technology, University of<br />

Colorado at Boulder, Boulder, CO 80309–0440, USA.<br />

*To whom correspondence should be addressed. E-mail:<br />

murnane@jila.colorado.edu<br />

www.sciencemag.org SCIENCE VOL 317 10 AUGUST 2007 775<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 179<br />

on November 12, 2007<br />

www.sciencemag.org<br />

Downloaded from


ARTICLES<br />

1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

Coherent superposition of laser-driven<br />

soft-X-ray harmonics from successive<br />

sources<br />

J. SERES 1,2 , V. S. YAKOVLEV 3 , E. SERES 1,2 , CH. STRELI 4 , P. WOBRAUSCHEK 4 , CH. SPIELMANN 2<br />

AND F. KRAUSZ 3,5 *<br />

1 Institut für Photonik, Technische Universität Wien, A-1040 Wien, Austria<br />

2 Physikalisches Institut EP1, Universität Würzburg, D-97074 Würzburg, Germany<br />

3 Department für Physik, Ludwig-Maximilians-Universität München, D-85748 Garching, Germany<br />

4 Atominstitut der Österreichischen Universtitäten, Technische Universität Wien, A-1020 Wien, Austria<br />

5 Max-Planck-Institut für Quantenoptik, D-85748 Garching, Germany<br />

*e-mail: krausz@lmu.de<br />

Published online: 11 November 2007; doi:10.1038/nphys775<br />

High-order harmonic generation from atoms ionized by femtosecond laser pulses has been a promising approach for the development<br />

of coherent short-wavelength sources. However, the realization of a powerful harmonic X-ray source has been hampered by a phase<br />

velocity mismatch between the driving wave <strong>and</strong> its harmonics, limiting their coherent build-up to a short propagation length <strong>and</strong><br />

thereby compromising the efficiency of a single source. Here, we report coherent superposition of laser-driven soft-X-ray (SXR)<br />

harmonics, at wavelengths of 2–5 nm, generated in two successive sources by one <strong>and</strong> the same laser pulse. Observation of constructive<br />

<strong>and</strong> destructive interference suggests the feasibility of quasi-phase-matched SXR harmonic generation by a focused laser beam in a<br />

gas medium of modulated density. Our proof-of-concept study opens the prospect of enhancing the photon flux of SXR harmonic<br />

sources to levels enabling researchers to tackle a range of applications in physical as well as life sciences.<br />

The quest for powerful laboratory sources of coherent soft-X-ray<br />

(SXR) light has been ongoing since the discovery of the laser.<br />

So far, high-order harmonic generation from atoms ionized by<br />

ultrashort laser pulses 1,2 constitutes the only technique providing<br />

coherent short-wavelength radiation at any photon energy up<br />

to the kiloelectronvolt regime 3–11 . Laser-driven high-harmonic<br />

sources (henceforth briefly referred to as harmonic sources) up<br />

to photon energies of 100 eV are now in widespread use 12 <strong>and</strong><br />

have played a key role in pushing the frontiers of nonlinear<br />

optics into the extreme-ultraviolet region 13–17 <strong>and</strong> ultrafast science<br />

into the attosecond domain 18–23 . Above 100 eV, rapidly decreasing<br />

efficiency prevented harmonic sources from becoming useful for<br />

applications. The low harmonic conversion efficiency implies that<br />

most of the incident laser photons are transmitted through the<br />

harmonic source, offering the possibility of being reused for<br />

creating harmonics in successive sources <strong>and</strong>—by exploiting their<br />

coherence—adding them to increase the overall harmonic flux.<br />

Here, we demonstrate the feasibility of this approach.<br />

Efficient generation of high-order harmonics of intense laser<br />

light relies on a large number of ionizing atoms emitting<br />

high-frequency radiation with proper phase that allows coherent<br />

build-up of light on propagation through the generation medium.<br />

Free electrons, unavoidable concomitants of the generation<br />

process, tend to severely limit the propagation length over which<br />

phase-matching can be achieved (henceforth briefly referred to as<br />

the coherence length). Femtosecond laser pulses have permitted<br />

the generation of harmonics in the extreme-ultraviolet regime<br />

(10–100 eV) at sufficiently low levels of ionization, so that the<br />

coherence length could be extended beyond the absorption<br />

length 4,5 , leading to the production of microjoule-energy coherent<br />

extreme-ultraviolet pulses 6,7 .<br />

Generation of SXR harmonics (>100 eV) is favoured by<br />

lowered absorption but suffers from increased free-electron density<br />

implied by the higher laser intensities required. Hence, the<br />

efficiency of SXR harmonic sources is limited by dephasing of<br />

the atomic dipole oscillators driven at different positions in the<br />

generation medium 24 . The SXR harmonic yield has recently been<br />

improved by implementing quasi-phase-matching (QPM) in a<br />

gas-filled hollow-core fibre with a modulated inner diameter 8 or<br />

counter-propagating laser pulses 25,26 . At higher photon energies,<br />

subcycle modification of few-cycle driving fields has been identified<br />

as being beneficial for enhancing phase-matching 27 (referred to<br />

as non-adiabatic self-phase-matching), which allowed extension<br />

of harmonic generation beyond the kiloelectronvolt frontier 10,11 .<br />

Unfortunately the fluxes demonstrated so far are still insufficient<br />

for most applications.<br />

Here, we provide the first experimental evidence of the<br />

feasibility of enhancing the efficiency of SXR harmonic generation<br />

by coherent superposition of harmonics generated in successive<br />

sources traversed by the same pump laser beam (Fig. 1). SXR<br />

harmonic emission from a single source is maximized by using<br />

few-cycle driver pulses, which maximize both the single-atom<br />

emission intensity <strong>and</strong> the coherence length for SXR harmonics 24 ,<br />

<strong>and</strong> is limited by dephasing. The phase of the atomic dipole<br />

878 nature physics VOL 3 DECEMBER 2007 www.nature.com/naturephysics<br />

© 2007 Nature Publishing Group<br />

180 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008


1 . 3 . 3 S E L E C T E D R E P R I N T S<br />

Enhanced phase-matching for<br />

generation of soft X-ray harmonics <strong>and</strong><br />

attosecond pulses in atomic gases<br />

Vladislav S. Yakovlev 1∗ , Misha Ivanov 2 , Ferenc Krausz 1,3<br />

1 Department of Physics, Ludwig-Maximilians-Universität München, Am Coulombwall 1,<br />

D-85748 Garching, Germany<br />

2 National Research Council of Canada, M-23A, Ottawa, Ontario, Canada K1A OR6<br />

3 Max-Planck-Institut für Quantenoptik, Hans-Kopfermann-Strasse 1, D-85748 Garching,<br />

Germany<br />

∗ Corresponding author: Vladislav.Yakovlev@physik.uni-muenchen.de<br />

Abstract: We theoretically investigate the generation of high harmonics<br />

<strong>and</strong> attosecond pulses by mid-infrared (IR) driving fields. Conditions for<br />

coherent build-up of high harmonics are revisited. We show that the coherence<br />

length dictated by ionization-induced dephasing does not constitute<br />

an ultimate limitation to the coherent growth of soft X-ray (> 100 eV) harmonics<br />

driven by few-cycle mid-IR driving pulses: perfect phase-matching,<br />

similar to non-adiabatic self-phase-matching, can be achieved even without<br />

non-linear deformation of the driving pulse. Our trajectory-based analysis<br />

of phase-matching reveals several important advantages of using longer<br />

laser wavelengths: conversion efficiency can be improved by orders of<br />

magnitude, phase-matched build-up of harmonics can be achieved in a jet<br />

with a high gas pressure, <strong>and</strong> isolated attosecond pulses can be extracted<br />

from plateau harmonics.<br />

© 2007 Optical Society of America<br />

OCIS codes: (190.7110) Ultrafast nonlinear optics; (270.6620) Strong-field processes;<br />

999.9999 <strong>Attosecond</strong> science.<br />

References <strong>and</strong> links<br />

1. P. B. Corkum, F. Krausz, “<strong>Attosecond</strong> science,” Nat. Phys. 3, 381–387 (2007).<br />

2. M. Hentschel, R. Kienberger, Ch. Spielmann, G. A. Reider, N. Milosevic, T. Brabec, P. Corkum, U. Heinzmann,<br />

M. Drescher, F. Krausz, “<strong>Attosecond</strong> metrology,” Nature 414, 509–513 (2001).<br />

3. R. Kienberger, E. Goulielmakis, M. Uiberacker, A. Baltuˇska, V. Yakovlev , F. Bammer, A. Scrinzi, Th. Westerwalbesloh,<br />

U. Kleineberg, U. Heinzmann, M. Drescher, F. Krausz, “Atomic transient recorder,” Nature 427,<br />

817–821 (2004).<br />

4. G. Sansone, E. Benedetti, F. Calegari, C. Vozzi, L. Avaldi, R. Flammini, L. Poletto, P. Villoresi, C. Altucci, R.<br />

Velotta, S. Stagira, S. De Silvestri, M. Nisoli, “Isolated single-cycle attosecond pulses,” Science 314, 443–446<br />

(2006).<br />

5. J. Seres, E. Seres, A. J. Verhoef, G. Tempea, C. Streli, P. Wobrauschek, V. Yakovlev, A. Scrinzi, C. Spielmann,<br />

F. Krausz, “Source of coherent kiloelectronvolt X-rays,” Nature 433, 596 (2005).<br />

6. A. Gordon, F. X. Kärtner, “Scaling of keV HHG photon yield with drive wavelength,” Opt. Express 13, 2941–<br />

2947 (2005).<br />

7. J. Tate, T. Auguste, H. G. Muller, P. Salieres, P. Agostini, L. F. DiMauro, “Scaling of wave-packet dynamics in<br />

an intense midinfrared field,” Phys. Rev. Lett. 98, 013901 (2007).<br />

8. S. Gordienko, A. Pukhov, O. Shorokhov, T. Baeva, “Relativistic Doppler Effect: Universal Spectra <strong>and</strong> Zeptosecond<br />

Pulses”, Phys. Rev. Lett. 93, 115002 (2004).<br />

9. G. D. Tsakiris, K. Eidmann, J. Meyer-ter-Vehn, Ferenc Krausz, “Route to intense single attosecond pulses,” New<br />

J. Phys. 8, 1–20 (2006).<br />

#84689 - $15.00 USD Received 29 Jun 2007; revised 21 Sep 2007; accepted 25 Sep 2007; published 5 Nov 2007<br />

(C) 2007 OSA 12 November 2007 / Vol. 15, No. 23 / OPTICS EXPRESS 15351<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 181


1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

New Journal of Physics<br />

T h e o p e n – a c c e s s j o u r n a l f o r p h y s i c s<br />

GeV-scale electron acceleration in a gas-filled<br />

capillary discharge waveguide<br />

S Karsch 1,6 , J Osterhoff 1 , A Popp 1 , T P Rowl<strong>and</strong>s-Rees 2 ,<br />

Zs Major 1 , M Fuchs 1,3 , B Marx 1,3 , R Hörlein 1,3 , K Schmid 1,3 ,<br />

L Veisz 1 , S Becker 3 , U Schramm 4 , B Hidding 5 , G Pretzler 5 ,<br />

D Habs 3 , F Grüner 1 , F Krausz 1,3 <strong>and</strong> S M Hooker 2<br />

1 Max-Planck-Institut für Quantenoptik, Hans-Kopfermann-Str. 1,<br />

D-85748 Garching, Germany<br />

2 Clarendon Laboratory, University of Oxford, Parks Road,<br />

Oxford OX1 3PU, UK<br />

3 Sektion Physik der Ludwig-Maximilians-Universität München,<br />

Am Coulombwall 1, D-85748 Garching, Germany<br />

4 Forschungszentrum Dresden-Rossendorf, Bautzner L<strong>and</strong>str. 128,<br />

D-01328 Dresden, Germany<br />

5 Institut für Laser- und Plasmaphysik, Heinrich-Heine-Universität,<br />

Universitätsstr. 1, D-40225 Düsseldorf, Germany<br />

E-mail: stefan.karsch@mpq.mpg.de<br />

New Journal of Physics 9 (2007) 415<br />

Received 14 September 2007<br />

Published 23 November 2007<br />

Online at http://www.njp.org/<br />

doi:10.1088/1367-2630/9/11/415<br />

Abstract. We report experimental results on laser-driven electron acceleration<br />

with low divergence. The electron beam was generated by focussing 750 mJ,<br />

42 fs laser pulses into a gas-filled capillary discharge waveguide at electron<br />

densities in the range between 10 18 <strong>and</strong> 10 19 cm −3 . Quasi-monoenergetic electron<br />

bunches with energies as high as 500 MeV have been detected, with features<br />

reaching up to 1 GeV, albeit with large shot-to-shot fluctuations. A more stable<br />

regime with higher bunch charge (20–45 pC) <strong>and</strong> less energy (200–300 MeV)<br />

could also be observed. The beam divergence <strong>and</strong> the pointing stability are<br />

around or below 1 mrad <strong>and</strong> 8 mrad, respectively. These findings are consistent<br />

with self-injection of electrons into a breaking plasma wave.<br />

6 Author to whom any correspondence should be addressed.<br />

New Journal of Physics 9 (2007) 415 PII: S1367-2630(07)60033-0<br />

1367-2630/07/010415+11$30.00 © IOP Publishing Ltd <strong>and</strong> Deutsche Physikalische Gesellschaft<br />

182 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008


1 . 3 . 3 S E L E C T E D R E P R I N T S<br />

New Journal of Physics<br />

T h e o p e n – a c c e s s j o u r n a l f o r p h y s i c s<br />

Hybrid dc–ac electron gun for fs-electron<br />

pulse generation<br />

L Veisz 1,4 , G Kurkin 2 , K Chernov 2 , V Tarnetsky 2 , A Apolonski 3 ,<br />

F Krausz 1,3 <strong>and</strong> E Fill 1<br />

1 Max-Planck-Institut für Quantenoptik, Hans-Kopfermann-Strasse 1,<br />

D-85748 Garching, Germany<br />

2 Budker Institute for Nuclear Physics, 630090 Novosibirsk, Russia<br />

3 Department für Physik der Ludwig-Maximilians-Universität München,<br />

Am Coulombwall 1, D-85748 Garching, Germany<br />

E-mail: laszlo.veisz@mpq.mpg.de<br />

New Journal of Physics 9 (2007) 451<br />

Received 5 August 2007<br />

Published 20 December 2007<br />

Online at http://www.njp.org/<br />

doi:10.1088/1367-2630/9/12/451<br />

Abstract. We present a new concept of an electron gun for generating<br />

subrelativistic few-femtosecond (fs) electron pulses. The basic idea is to utilize<br />

a dc acceleration stage combined with an RF cavity, the ac field of which<br />

generates an electron energy chirp for bunching at the target. To reduce space<br />

charge (SC) broadening the number of electrons in the bunch is reduced <strong>and</strong><br />

the gun is operated at a megahertz (MHz) repetition rate for providing a high<br />

average number of electrons at the target. Simulations of the electron gun were<br />

carried out under the condition of no SC <strong>and</strong> with SC assuming various numbers<br />

of electrons in the bunch. Transversal effects such as defocusing after the dc<br />

extraction hole were also taken into account. A detailed analysis of the sensitivity<br />

of the pulse duration to various parameters was performed to test the realizability<br />

of the concept. Such electron pulses will allow significant advances in the<br />

field of ultrafast electron diffraction.<br />

4 Author to whom any correspondence should be addressed.<br />

New Journal of Physics 9 (2007) 451 PII: S1367-2630(07)56694-2<br />

1367-2630/07/010451+17$30.00 © IOP Publishing Ltd <strong>and</strong> Deutsche Physikalische Gesellschaft<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 183


1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

Picosecond electron deflectometry<br />

of optical-field ionized plasmas<br />

MARTIN CENTURION 1 * † , PETER RECKENTHAELER 1,2† , SERGEI A. TRUSHIN 1 , FERENC KRAUSZ 1,2<br />

AND ERNST E. FILL 1<br />

1 Max-Planck-Institut für Quantenoptik, Hans-Kopfermann-Strasse 1, D-85748 Garching, Germany<br />

2 Ludwig-Maximilians-Universität München, Am Coulombwall 1, D-85748 Garching, Germany<br />

† These authors contributed equally to this work.<br />

*e-mail: martin.centurion@mpq.mpg.de<br />

Published online: 20 April 2008; doi:10.1038/nphoton.2008.77<br />

Optical-field ionized plasmas are of great interest owing to their<br />

unique properties <strong>and</strong> the fact that they suit many applications,<br />

such as the study of nuclear fusion 1 , generation of energetic<br />

electrons 2–5 <strong>and</strong> ions 6,7 , X-ray emission 8,9 , X-ray lasers 10–12<br />

<strong>and</strong> extreme-UV attosecond pulse generation 13 . A detailed<br />

knowledge of the plasma dynamics can be critical for<br />

optimizing a given application. Here we demonstrate a method<br />

for real-time imaging of the electric-field distribution in<br />

optical-field ionized plasmas with ultrahigh temporal<br />

resolution, yielding information that is not accessible by other<br />

methods. The technique, based on electron deflectometry,<br />

yields images that reveal a positively charged core <strong>and</strong> a cloud<br />

of electrons exp<strong>and</strong>ing far beyond the Debye length.<br />

The parameters of an optical-field ionized (OFI) plasma can be<br />

varied over a wide range. The maximum ionization stage can be set<br />

by the intensity of the laser pulse generating the plasma 14 , the<br />

electron temperature can be controlled by applying linear or<br />

circular polarization 15 , <strong>and</strong>, by appropriately choosing the<br />

backing pressure of the nozzle, cluster plasmas 16 can be<br />

generated. Various methods have been applied to investigate the<br />

parameters of OFI plasmas. The electron temperature has been<br />

measured by Thomson scattering 17 . The ionization stage has been<br />

determined by ion spectrometry 18 <strong>and</strong> by recording the emission<br />

of X-rays from the plasma 11 . The time-resolved plasma density<br />

profile has been measured by optical interferometry <strong>and</strong><br />

holography 19,20 . Moiré deflectometry has been used to determine<br />

the density profile in the plasma channel <strong>and</strong> its lateral<br />

expansion 21 . Spectrometry of ions emitted from the plasma has<br />

yielded information on ion velocity <strong>and</strong> temperature 22 .<br />

Radiography using energetic (MeV) protons has been used to<br />

diagnose density perturbations <strong>and</strong> transient fields in highdensity<br />

plasmas with a temporal resolution of 100 ps (refs 23–25).<br />

Here, we demonstrate a new diagnostics technique for OFI<br />

plasmas with a temporal resolution of 2.7 ps <strong>and</strong> very high<br />

sensitivity. The technique is deflectometry using monoenergetic<br />

20-keV electron pulses, which are directed onto an OFI nitrogen<br />

plasma generated by a 50-fs Ti:sapphire laser pulse. The electrons<br />

are deflected by the fields resulting from charge separation, <strong>and</strong><br />

the resulting distortion of the electron beam yields time-resolved<br />

images of the plasma. Pump–probe experiments on the plasma<br />

evolution capture changes within a few picoseconds, with a<br />

spatial resolution of 30 mm. This direct time-resolved imaging of<br />

plasma fields reveals features not accessible to other methods.<br />

Such knowledge can help to improve the setting of parameters<br />

for optimizing particular applications. As an example, X-ray<br />

lasers <strong>and</strong> soft X-ray sources may greatly benefit from a<br />

better underst<strong>and</strong>ing of the dynamics of laser-generated<br />

plasmas. The new technique has the potential to lead to better<br />

control of plasma electron <strong>and</strong> ion accelerators <strong>and</strong> improve<br />

their features. The high sensitivity of electron deflectometry is<br />

based on the fact that even small charge imbalances within the<br />

plasma are observable as distortions in the spatial profile of<br />

the electron beam. The narrow energy spread <strong>and</strong> low emittance<br />

(1 ¼ 1.5 mrad mm) of the electron beam allows plasma fields<br />

below 1 � 10 6 V m 21 to be detected.<br />

An application of the method is illustrated in Fig. 1, which<br />

shows the evolution of a nitrogen OFI plasma. Initially, a small<br />

depleted region appears in the electron beam in the area of the<br />

laser focus. This hole exp<strong>and</strong>s for approximately 80 ps, after which<br />

a spot develops in the centre (Fig. 1e,f). The spot becomes brighter<br />

than the initial electron beam for a time delay T . 100 ps, <strong>and</strong> its<br />

intensity increases up to 200 ps. Beyond 200 ps (not shown in the<br />

figure) the brightness of the spot slowly decreases.<br />

Simultaneously with the depleted region, two bright lobes<br />

appear on each side of the plasma region, along the line of laser<br />

propagation (vertical in Figs 1 <strong>and</strong> 2). These lobes then move<br />

away from the focal region in opposite directions. The sidelobes<br />

reach the boundary of the electron beam after approximately<br />

100 ps (Fig. 1e) <strong>and</strong> then remain static. The pattern remained<br />

qualitatively unchanged until the maximum delay time in the<br />

experiment (T ¼ 300 ps), with only a slight decrease in the<br />

brightness of the main features.<br />

Figure 2 focuses on the initial stages of the plasma expansion.<br />

The duration of the electron pulses was decreased by reducing<br />

the number of electrons per pulse. Figure 2a–d shows the plasma<br />

evolution with a step size of �2.7 ps. The formation of an<br />

exp<strong>and</strong>ing feature is clearly visible after the first step. After only<br />

5.3 ps (Fig. 2c) there is already a well-defined boundary<br />

<strong>and</strong> deflected electrons appear in the shadow of the nozzle. For<br />

T . 5.3 ps (Fig. 2d–i) the hole exp<strong>and</strong>s <strong>and</strong> the number of<br />

electrons deflected out of the focal region increases. The sidelobes<br />

at the boundary of the plasma also appear at 5.3 ps (Fig. 2c),<br />

then increase in brightness <strong>and</strong> move away from the centre. At<br />

T ¼ 24 ps (Fig. 2g) the fraction of electrons ejected reaches its<br />

184 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008<br />

LETTERS<br />

nature photonics | VOL 2 | MAY 2008 | www.nature.com/naturephotonics 315<br />

© 2008 Nature Publishing Group


REPORTS<br />

1614<br />

1 . 3 . 3 S E L E C T E D R E P R I N T S<br />

Single-Cycle Nonlinear Optics<br />

E. Goulielmakis, 1 * M. Schultze, 1 M. Hofstetter, 2 V. S. Yakovlev, 2 J. Gagnon, 1<br />

M. Uiberacker, 2 A. L. Aquila, 3 E. M. Gullikson, 3 D. T. Attwood, 3 R. Kienberger, 1<br />

F. Krausz, 1,2 * U. Kleineberg 2 *<br />

Nonlinear optics plays a central role in the advancement of optical science <strong>and</strong> laser-based<br />

technologies. We report on the confinement of the nonlinear interaction of light with matter to a single<br />

wave cycle <strong>and</strong> demonstrate its utility for time-resolved <strong>and</strong> strong-field science. The electric field of<br />

3.3-femtosecond, 0.72-micron laser pulses with a controlled <strong>and</strong> measured waveform ionizes atoms<br />

near the crests of the central wave cycle, with ionization being virtually switched off outside this<br />

interval. Isolated sub-100-attosecond pulses of extreme ultraviolet light (photon energy ~ 80 electron<br />

volts), containing ~0.5 nanojoule of energy, emerge from the interaction with a conversion efficiency<br />

of ~10 –6 . These tools enable the study of the precision control of electron motion with light fields <strong>and</strong><br />

electron-electron interactions with a resolution approaching the atomic unit of time (~24 attoseconds).<br />

N<br />

onlinear electron-light interactions driven<br />

by strong light fields of controlled wave-<br />

form (1) have allowed for the control of<br />

electronic motion at light frequencies <strong>and</strong> the realtime<br />

observation of electron dynamics inside <strong>and</strong><br />

between atoms with ~100-as resolution (2–7).<br />

However, time-domain access to a number of<br />

fundamental processes, such as the intra-atomic<br />

energy transfer between electrons (resulting, for<br />

example, in shake-up) (8), the response of an<br />

atomic electron system to external influence (e.g.,<br />

to ionizing radiation) (9) <strong>and</strong> its rearrangement<br />

after the sudden loss of one or more electrons<br />

Fig. 1. Simulation of sub-femtosecond XUV emission from<br />

neon atoms ionized by a linearly polarized, sub-1.5-cycle, 720nm<br />

laser field. E0 <strong>and</strong> aL(t) are inferred from best agreement<br />

between the modeled (17) <strong>and</strong> measured (Fig. 2) spectra <strong>and</strong><br />

the streaking spectrogram (Fig. 3), respectively. The laser field<br />

liberates electrons near its most intense wave crests. (A)<br />

Classical trajectories of maximum return energy (left panels)<br />

<strong>and</strong> spectra of emerging XUV emission (right panels) are<br />

shown for waveforms consistent with a L(t) inferred from Fig. 3<br />

(correspondence established by colors <strong>and</strong> line style). The<br />

numbers in (i) to (iii) quantify, in units of 10 –4 , the ionization<br />

probability <strong>and</strong>, hence, the squared modulus of the amplitude of<br />

the electron wave packets launched. This amplitude substantially<br />

dictates the intensity of XUV emission upon recollision:<br />

Contrast ionization probability in (i) to (iii) with the corresponding<br />

emission intensities in (iv) to (vi). The pink “dottedline”<br />

emission is not visible in (v) because of the lower<br />

ionization probability (by two orders of magnitude) with<br />

respect to that resulting in the purple “solid-line” emission<br />

[see (ii)]. The gray dashed-<strong>and</strong>-dotted lines denote the b<strong>and</strong>pass<br />

used in our experiments (fig. S2). ϕ = 70° <strong>and</strong> ϕ = 135°<br />

yield highest contrast [see (B)] <strong>and</strong> highest XUV cutoff energy,<br />

respectively. arb.u., arbitrary units. (B) Contrast versus CE<br />

phase. Here, contrast is defined as the ratio of the energy of<br />

the main attosecond XUV pulse to the overall XUV emission<br />

energy transmitted through the b<strong>and</strong>pass [gray dashed-<strong>and</strong>dotted<br />

lines in (A)].<br />

(10), the charge transfer in biologically relevant<br />

molecules (11) <strong>and</strong> related changes in chemical<br />

reactivity (12) or because of nonadiabatic tunneling<br />

(13, 14), would require (or benefit from)<br />

an improved temporal resolution.<br />

We used waveform-controlled sub-1.5-cycle<br />

near-infrared (NIR) light to demonstrate the generation<br />

of robust, energetic, isolated sub-100-as<br />

pulses of extreme ultraviolet (XUV) radiation <strong>and</strong><br />

their precise temporal characterization. Photoionization<br />

confined to a single wave cycle results<br />

in observables (such as high-harmonic photons<br />

<strong>and</strong> electrons emitted by above-threshold ioniza-<br />

A<br />

Distance of electron from the core (Å)<br />

B<br />

Contrast<br />

20<br />

10<br />

0<br />

-10<br />

-20<br />

20<br />

10<br />

0<br />

-10<br />

-20<br />

20<br />

10<br />

0<br />

-10<br />

-20<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

(i)<br />

(ii)<br />

(iii)<br />

E L (t)<br />

E L (t)<br />

E L (t)<br />

0.49<br />

5.9<br />

36<br />

14<br />

48<br />

26<br />

45<br />

8.6<br />

0.3<br />

-3.75 0<br />

Time (fs)<br />

3.75<br />

tion) that can now be related to several distinguishable<br />

subcycle ionization events <strong>and</strong> subsequent<br />

electron trajectories with a known timing with<br />

respect to the driving field, whose strength <strong>and</strong><br />

temporal evolution is accurately known (3). These<br />

circumstances provide ideal conditions for testing<br />

models of strong-field control of electron motion<br />

<strong>and</strong> electron-electron interactions.<br />

The generation of attosecond pulses benefits<br />

from the abrupt onset of ionization within a single<br />

half-cycle, which minimizes the density of free<br />

electrons <strong>and</strong>, hence, the distortion of the driving<br />

wave <strong>and</strong> its dephasing with the generated harmonic<br />

wave. As a result, the coherent build-up of the<br />

harmonic emission over an extended propagation<br />

is maximized. In addition, the order-of-magnitude<br />

variation of the ionization probability between adjacent<br />

half-cycles creates unique conditions for<br />

single sub-100-as pulse emission without the need<br />

for sophisticated gating techniques (5, 15, 16).<br />

On the measurement side, improved resolution<br />

results from three provisions: (i) shorter<br />

1 Max-Planck-Institut für Quantenoptik, Hans-Kopfermann-<br />

Strasse 1, D-85748 Garching, Germany. 2 Department für<br />

Physik, Ludwig-Maximilians-Universität, Am Coulombwall<br />

1, D-85748 Garching. 3 Center for X-Ray Optics, Lawrence<br />

Berkeley National Laboratory, Berkeley, CA 94720, USA.<br />

*To whom correspondence should be addressed. E-mail: elgo@<br />

mpq.mpg.de (E.G.); krausz@lmu.de (F.K.); ulf.kleineberg@<br />

physik.uni-muenchen.de (U.K.)<br />

0<br />

40 60 80 100 120<br />

Energy (eV)<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 185<br />

XUV spectral intensity (arb.u.)<br />

1<br />

0<br />

1<br />

0<br />

1<br />

(iv)<br />

(v)<br />

(vi)<br />

ϕ = 0°<br />

ϕ = 70°<br />

ϕ = 135°<br />

0 20 40 60 80 100 120 140 160 180<br />

Carrier envelope phase, ϕ (deg)<br />

20 JUNE 2008 VOL 320 SCIENCE www.sciencemag.org<br />

on June 24, 2008<br />

www.sciencemag.org<br />

Downloaded from


Photoelectron energy (eV)<br />

70<br />

60<br />

50<br />

40<br />

A<br />

B<br />

70<br />

60<br />

50<br />

40<br />

1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

0 1<br />

20 40 60 80 100 120 140 160<br />

Carrier-envelope phase (deg)<br />

Electron counts (arb. u.)<br />

1<br />

C<br />

0<br />

1<br />

D<br />

0<br />

1<br />

E<br />

XUV pulse duration, (ii) improved signal-to-noise<br />

(S/N) ratio due to the increased XUV photon<br />

flux, <strong>and</strong> (iii) stronger streaking before the onset<br />

of the NIR field–induced ionization in attosecond<br />

streaking (2) or enhanced S/N ratio due to a<br />

reduced number of tunneling steps in attosecond<br />

tunneling spectroscopy (14).<br />

Figure 1 summarizes results of the modeling<br />

of the single-cycle interaction of ionizing NIR<br />

radiation with an ensemble of neon atoms (17). In<br />

Fig. 1A, the left panels plot possible NIR electric<br />

waveforms,ELðtÞ ¼ E0aLðtÞe −iðwLtþϕÞ þ cc<br />

(where cc st<strong>and</strong>s for complex conjugate) derived<br />

from our streaking measurements (as presented<br />

in the next sections) for different settings of the<br />

carrier-envelope (CE) phase, ϕ. Here, E0 is the<br />

peak electric-field strength, aL(t) is the normalized<br />

complex amplitude envelope, <strong>and</strong> wL is the<br />

carrier frequency. The probability of ionization<br />

outside the central cycle is more than two orders<br />

of magnitude lower than that at the field maximum<br />

<strong>and</strong> hence is negligible.<br />

The spectra of XUVemissions originating from<br />

the individual recollisions (18) are predicted to<br />

differ by tens of electron volts in cut-off energy <strong>and</strong><br />

by up to orders of magnitude in intensity as a consequence<br />

of the single-cycle nature of the driving<br />

field. The strong variation of emission energies <strong>and</strong><br />

intensities within a single wave cycle creates ideal<br />

conditions for isolated sub-100-as pulse generation.<br />

Indeed, filtering radiation with the b<strong>and</strong>pass<br />

depicted by the dashed-<strong>and</strong>-dotted line is predicted<br />

to isolate XUVradiation with more than 90% of its<br />

energy delivered in a single attosecond pulse for a<br />

range of CE phases as broad as 30° ≤ ϕ ≤ 90° (Fig.<br />

1B). In contrast, with few-cycle-driven harmonic<br />

generation resulting in isolated subfemtosecond<br />

pulses over only a relatively narrow range of the<br />

CE phase near ϕ ≈ 0° (3), single-cycle excitation<br />

appears to permit robust isolated attosecond pulses<br />

for a variety of driver waveforms, ranging from<br />

near-cosine– to sine-shaped ones, owing to the<br />

order-of-magnitude variation of the ionization<br />

probability within a single wave cycle.<br />

We used phase-controlled sub-1.5-cycle laser<br />

pulses carried at a wavelength of lL = 2pc/wL =<br />

720 nm (19) to generate XUV harmonics in a<br />

neon gas jet up to photon energies of ~110 eV<br />

(fig. S1). The emerging XUV pulse—following<br />

a spectral filtering through a b<strong>and</strong>pass (dashed<strong>and</strong>-dotted<br />

line in Fig. 1A) introduced by metal<br />

foils <strong>and</strong> a Mo/Si multilayer mirror (fig. S2)—<br />

subsequently propagates, along with its NIR driver<br />

wave, through a second jet of neon atoms in<br />

which the XUV pulse ionizes the atoms in the<br />

presence of the NIR field. The freed electrons<br />

with initial momenta directed along the electricfield<br />

vector of the linearly polarized NIR field are<br />

collected <strong>and</strong> analyzed with time-of-flight spectrometry<br />

(17).<br />

The variation of the measured photoelectron<br />

spectra versus CE phase shows good agreement<br />

with the predictions of our simulations (Fig. 2,<br />

A <strong>and</strong> B). Figure 2, C to E, shows plots of<br />

electron spectra corresponding to the CE phase<br />

186 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008<br />

ϕ = 70°<br />

ϕ = 130°<br />

ϕ = 170°<br />

0<br />

30 40 50 60 70 80<br />

Photoelectron energy (eV)<br />

Fig. 2. <strong>Control</strong> of b<strong>and</strong>pass-filtered XUV emission with the waveform of monocycle light. Measured<br />

(A) <strong>and</strong> simulated (B) (17) photoelectron spectra versus CE phase, with the delay increased in steps of<br />

~11° ( p /16 rad). (C to E) Spectra measured at the CE phase setting closest to the values selected in<br />

Fig. 1A. The zero of the CE phase scale in (A) was set to yield the best agreement with the modeled<br />

spectra in (B).<br />

Photoelectron energy ( eV)<br />

XUV intensity (arb.u.)<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

A B<br />

−4 −2 0<br />

Delay (fs)<br />

2 4<br />

1.0 C 1.0 D<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

τ x= 80 ±5 as<br />

-300 -200 -100 0 100 200 300<br />

Time (as)<br />

4<br />

3<br />

2<br />

1<br />

phase (rad)<br />

Photoelectron energy ( eV)<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

XUV spectral intensity (arb.u.)<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

−4 −2 0<br />

Delay (fs)<br />

2 4<br />

φ″=(1.5 ± 0.2)×10 3 as 2<br />

40 50 60 70 80 90 100 110<br />

Photon energy (eV)<br />

Fig. 3. Sub-100-as XUV pulse retrieval. (A) Measured ATR spectrogram compiled from 126 energy<br />

spectra of photoelectrons launched by an XUV pulse with a b<strong>and</strong>width of ~28 eV (FWHM) <strong>and</strong> recorded<br />

at delay settings increased in steps of 80 as. Here, a positive delay corresponds to the XUV pulse<br />

arriving before the NIR pulse. The high flux of the XUV source allows this spectrogram to be recorded<br />

within ~30 min. (B) ATR spectrogram reconstructed after ~10 3 iterations of the FROG algorithm (17).<br />

(C) Retrieved temporal intensity profile <strong>and</strong> spectral phase of the XUV pulse. The intrinsic chirp of the<br />

XUV emission (Fig. 4B) is almost fully compensated by a 300-nm-thick Zr foil introduced into the XUV<br />

beam between the attosecond source <strong>and</strong> the ATR measurement. Arrows indicate the temporal FWHM of<br />

the XUV pulse. (D) XUV spectra evaluated from the measurement of the XUV-generated photoelectron<br />

spectrum in the absence of the NIR streaking field (blue dashed line) <strong>and</strong> from the ATR retrieval (blue<br />

solid line). The black dotted line indicates the retrieved spectral phase.<br />

0<br />

-3<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

phase (rad)<br />

REPORTS<br />

www.sciencemag.org SCIENCE VOL 320 20 JUNE 2008 1615<br />

on June 24, 2008<br />

www.sciencemag.org<br />

Downloaded from


REPORTS<br />

1616<br />

settings selected in Fig. 1A. Apart from a downshift<br />

by the ionization potential of neon (21.5 eV),<br />

they reveal close resemblance to the XUV spectra<br />

transmitted through the b<strong>and</strong>pass in Fig. 1A(v),<br />

(vi), <strong>and</strong> (iv), respectively. Figure 2C depicts the<br />

broadest filtered spectrum produced by a single<br />

recollision [full-line spectrum in Fig. 1A(v)].<br />

Emission from the same recollision dominates<br />

also in the spectrum shown in Fig. 2E, with this<br />

spectrum red-shifted <strong>and</strong> (upon transmission through<br />

the b<strong>and</strong>pass) correspondingly narrowed, as predicted<br />

in Fig. 1A(iv). The two humps of the<br />

spectrum plotted in Fig. 2D are indicative of contributions<br />

from two recollisions, in accordance<br />

with the “solid-line” <strong>and</strong> “dotted-line” contributions<br />

to the emission spectrum in Fig. 1A(vi).<br />

Before measuring the XUV pulse, we optimized<br />

the generation process by “fine-tuning”<br />

the NIR laser peak intensity to achieve the<br />

broadest possible XUV spectrum transmitted<br />

through the b<strong>and</strong>pass in the range of CE phase<br />

settings where the contrast is maximized (50° – 80°,<br />

according to Fig. 1B) to generate a clean single<br />

pulse with the shortest possible duration. For the<br />

temporal characterization of the generated XUV<br />

supercontinuum, we shined the NIR field into the<br />

neon atoms ionized by the XUV pulse to implement<br />

the atomic transient recorder (ATR)<br />

technique introduced in (2, 3). The NIR field<br />

boosts or decreases the momenta of the photoelectrons,<br />

depending on their instant of release<br />

within the 2.4-fs period of the laser field, resulting<br />

in broadened <strong>and</strong> shifted (streaked)<br />

spectra of the electrons’ final energy distribution.<br />

Figure 3A is a plot of a series of streaked spectra<br />

recorded versus delay between the XUV<strong>and</strong> NIR<br />

pulse, which we refer to as an ATR or streaking<br />

spectrogram. It is practically equivalent to the<br />

spectrogram obtained by frequency-resolved<br />

optical gating (FROG) (20), with the oscillating<br />

NIR field constituting an attosecond phase gate<br />

in the present case (21). As a consequence, a<br />

FROG retrieval algorithm (22) allows complete<br />

determination of both the (gated) XUV pulse <strong>and</strong><br />

the (gating) NIR laser field (17). The reconstructed<br />

ATR spectrogram is plotted in Fig. 3B<br />

<strong>and</strong> reveals excellent agreement with the measured<br />

one.<br />

The retrieved temporal intensity profile <strong>and</strong><br />

phase of the XUV pulse are shown in Fig. 3C.<br />

The pulse duration of t x = 80 ± 5 as is close to<br />

its transform limit of 75 as, with a small positive<br />

chirp of f′′ = (1.5 ± 0.2) × 10 3 as 2 being responsible<br />

for the deviation. As a further consistency<br />

check of the attosecond pulse retrieval, we<br />

compared the measured NIR field–free electron<br />

spectrum (dashed blue line in Fig. 3D) with the<br />

electron spectrum calculated from the retrieved<br />

attosecond pulse (solid line in Fig. 3D). Given<br />

that the pulse retrieval draws on streaked spectra<br />

that are strongly distorted with respect to the<br />

field-free one, the degree of agreement between<br />

the retrieved <strong>and</strong> directly measured spectrum<br />

provides yet another conclusive testimony of the<br />

reliability of the retrieved data.<br />

1 . 3 . 3 S E L E C T E D R E P R I N T S<br />

The ATR retrieval algorithm indicates the<br />

presence of a satellite pulse accompanying the<br />

main attosecond pulse, containing ~1% of the energy<br />

of the main pulse. This amount of satellite<br />

is consistent with the depth of the experimentally<br />

observed modulation in the XUV spectrum<br />

(Fig. 3D). However, this result is inconsistent<br />

with our numerical modeling, which predicts a<br />

satellite energy content of some 6 to 7% for the<br />

optimum range of CE phase settings (Fig. 1B).<br />

From analysis of the streaked spectra recorded<br />

at the maximum of the NIR electric field (fig.<br />

S3), where the momentum of the electrons released<br />

by the main attosecond pulse <strong>and</strong> its<br />

satellite is shifted in opposite directions, we<br />

inferred a relative satellite energy of ~8%, which<br />

is in good agreement with the prediction of our<br />

modeling. As a consequence, the fringe visibility<br />

in the XUV spectrum is lower than was<br />

implied by the relative amplitude of the satellite<br />

pulse. The discrepancy may originate from a<br />

temporal jitter between the main pulse <strong>and</strong> the<br />

satellite pulse <strong>and</strong>/or from a different spatial<br />

amplitude distribution of the beams transporting<br />

the emission from the adjacent recollision<br />

events. The important lesson from these findings<br />

is that the fringe visibility in the XUV spectrum<br />

does not allow a reliable determination of the<br />

A<br />

80 as<br />

XUV pulse envelope<br />

B<br />

Photon energy (eV)<br />

120<br />

100<br />

80<br />

60<br />

78<br />

52<br />

26<br />

Electric field (GV/m)<br />

energy carried by the satellite(s) accompanying<br />

the main attosecond pulse.<br />

The laser waveform evaluated from the ATR<br />

measurement is preserved between the location<br />

of the generation <strong>and</strong> measurement. Figure 4A<br />

illustrates the evaluated NIR waveform with<br />

electric-field amplitude corresponding to a peak<br />

intensity of I 0 ~ (5.8 ± 0.5) × 10 14 W/cm 2 , as<br />

evaluated from the cut-off of our XUV spectra.<br />

The pulse duration [full width at half maximum<br />

(FWHM)] of 3.3 fs is in good agreement with<br />

the results of previous interferometric autocorrelation<br />

measurements (19), <strong>and</strong> the evaluated<br />

CE phase of ~50° is consistent with the optimum<br />

contrast according to our modeling (Fig. 1B).<br />

Accurate knowledge of the attosecond XUV<br />

pulse parameters, the temporal evolution of the<br />

generating NIR field, <strong>and</strong> the emergence of the<br />

former from a single recollision permit one to<br />

perform precision tests of models of light-electron<br />

interactions underlying the ionization <strong>and</strong> subsequent<br />

attosecond pulse generation processes. As<br />

an example, we calculated the intrinsic spectral<br />

chirp (i.e., the variation of the group delay versus<br />

frequency) carried by the attosecond XUV pulse<br />

during its emergence from the chirp measured by<br />

the ATR <strong>and</strong> the known dispersion of the<br />

components traversed by the pulse on its way<br />

Laser electric field E L(t)<br />

Recolliding electron<br />

trajectories<br />

0<br />

-2 0 2<br />

4<br />

Time (fs)<br />

Short Long<br />

trajectories<br />

0.5 1.0 1.5 2.0 2.5 3.0<br />

Recollision time (fs)<br />

Fig. 4. (A) Retrieved electric field of the NIR laser pulse used for generating <strong>and</strong> measuring the<br />

attosecond XUV pulse shown in Fig. 3. The duration of the pulse (FWHM of the cycle-averaged intensity<br />

profile) is tL = 3.3 fs, <strong>and</strong> the CE phase is evaluated as ϕ ~ 50°. The classical electron trajectories<br />

responsible for the emission of the filtered attosecond pulse are calculated with the plotted electric field<br />

<strong>and</strong> shown in the same panel. The color coding indicates the final return energy of the electrons. (B)<br />

Energy of the recolliding electron plus ionization potential (which is equal to the emitted XUV photon<br />

energy) versus recollision time evaluated from the classical trajectory analysis (solid green line), <strong>and</strong><br />

emitted photon energy versus emission time (dashed purple line) inferred from the chirp of the measured<br />

attosecond pulse <strong>and</strong> the dispersion of the metal filter that the attosecond pulse passes through before the<br />

measurement. The basic idea for the graphical representation is borrowed from (29). Error bars indicate<br />

the uncertainty in the retrieved group delay.<br />

20 JUNE 2008 VOL 320 SCIENCE www.sciencemag.org<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 187<br />

60<br />

30<br />

Electron excursion (Å)<br />

on June 24, 2008<br />

www.sciencemag.org<br />

Downloaded from


from the source to the measurement (fig. S2). The<br />

result (purple dashed line in Fig. 4B) is compared<br />

with the intrinsic attosecond chirp (green solid<br />

line in Fig. 4B) calculated from a classical<br />

trajectory analysis (23, 24). There is a notable<br />

discrepancy at the high-energy components of<br />

the wave packet, possibly because of quantum<br />

effects near cutoff. Nevertheless, the agreement<br />

with the attochirp resulting from short trajectories<br />

is stunning in the main part of the spectrum,<br />

where the S/N ratio is excellent. This agreement<br />

indicates the powerfulness of semiclassical modeling<br />

of strong-field interactions (25, 26) <strong>and</strong> the<br />

negligible role of long trajectories in contributing<br />

to the XUV radiation in the far field (27).<br />

In a similar way, the confinement of interaction<br />

between the ionizing field <strong>and</strong> the atom to a<br />

single wave cycle will permit accurate quantitative<br />

tests of theories of strong-field ionization. The<br />

sub-100-as XUV pulses emerging from the interaction<br />

with a flux greater than 10 11 photons/s—<br />

along with their monocycle NIR driver wave—will<br />

push the resolution limit of attosecond spectroscopy<br />

to the atomic unit of time (~24 as) <strong>and</strong> allow<br />

for the real-time observation of electron correla-<br />

1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

tions, by means of streaking (6), tunneling (14),<br />

or absorption (28) spectroscopy.<br />

References <strong>and</strong> Notes<br />

1. A. Baltuska et al., Nature 421, 611 (2003).<br />

2. R. Kienberger et al., Nature 427, 817 (2004).<br />

3. E. Goulielmakis et al., Science 305, 1267 (2004).<br />

4. M. F. Kling et al., Science 312, 246 (2006).<br />

5. G. Sansone et al., Science 314, 443 (2006).<br />

6. A. L. Cavalieri et al., Nature 449, 1029 (2007).<br />

7. E. Goulielmakis et al., Science 317, 769 (2007).<br />

8. S. Svensson, B. Eriksson, N. Martensson, G. Wendin,<br />

U. J. Gelius, J. Electron Spectrosc. Relat. Phenom. 47,<br />

327 (1988).<br />

9. S. X. Hu, L. A. Collins, Phys. Rev. Lett. 96, 073004 (2006).<br />

10. J. Breidbach, L. S. Cederbaum, Phys. Rev. Lett. 94,<br />

033901 (2005).<br />

11. A. I. Kuleff, J. Breidbach, L. S. Cederbaum, J. Chem. Phys.<br />

123, 044111 (2005).<br />

12. F. Remacle, R. D. Levine, Proc. Natl. Acad. Sci. U.S.A.<br />

103, 6793 (2006).<br />

13. G. Yudin, M. Y. Ivanov, Phys. Rev. A 64, 035401 (2001).<br />

14. M. Uiberacker et al., Nature 446, 627 (2007).<br />

15. T. Pfeifer et al., Phys. Rev. Lett. 97, 163901 (2006).<br />

16. Y. Oishi et al., Opt. Express 14, 7230 (2006).<br />

17. See supporting material on Science Online.<br />

18. C. A. Haworth et al., Nature Phys. 3, 52 (2007).<br />

19. A. L. Cavalieri et al., New J. Phys. 9, 242 (2007).<br />

20. R. Trebino, D. J. Kane, J. Opt. Soc. Am. A 10, 1101 (1993).<br />

21. Y. Mairesse, F. Quéré, Phys. Rev. A 71, 011401(R) (2005).<br />

The Formation Conditions<br />

of Chondrules <strong>and</strong> Chondrites<br />

C. M. O’D. Alex<strong>and</strong>er, 1 * J. N. Grossman, 2 D. S. Ebel, 3 F. J. Ciesla 1<br />

Chondrules, which are roughly millimeter-sized silicate-rich spherules, dominate the most<br />

primitive meteorites, the chondrites. They formed as molten droplets <strong>and</strong>, judging from their<br />

abundances in chondrites, are the products of one of the most energetic processes that operated in<br />

the early inner solar system. The conditions <strong>and</strong> mechanism of chondrule formation remain<br />

poorly understood. Here we show that the abundance of the volatile element sodium remained<br />

relatively constant during chondrule formation. Prevention of the evaporation of sodium requires<br />

that chondrules formed in regions with much higher solid densities than predicted by known<br />

nebular concentration mechanisms. These regions would probably have been self-gravitating. Our<br />

model explains many other chemical characteristics of chondrules <strong>and</strong> also implies that chondrule<br />

<strong>and</strong> planetesimal formation were linked.<br />

C<br />

hondrules make up ~20 to 80 volume % of<br />

most chondrites <strong>and</strong> formed at peak temper-<br />

atures of ~1700 to 2100 K (1). Chondrules<br />

in the different chondrite groups have distinct physical<br />

<strong>and</strong> chemical properties (2), as well as distinct<br />

age ranges (3), indicating that they formed in relatively<br />

local environments via a process that operated<br />

at least periodically between ~1 <strong>and</strong> 4 million years<br />

after the formation of the solar system.<br />

It was long thought that individual chondrules<br />

behaved as chemically closed systems during their<br />

formation, inheriting their compositions from their<br />

precursors (4, 5). However, for likely cooling rates<br />

of 10 to 1000 K/hour (1) <strong>and</strong> at the low pressures<br />

1 Department of Terrestrial Magnetism, Carnegie Institution of<br />

Washington, Washington, DC 20015, USA. 2 U.S. Geological<br />

Survey, Reston, VA 20192, USA. 3 American Museum of<br />

Natural History, New York, NY 10024, USA.<br />

*To whom correspondence should be addressed. E-mail:<br />

alex<strong>and</strong>e@dtm.ciw.edu<br />

(total pressure ≈ 10 −6 to 10 −3 bars) of the solar<br />

protoplanetary disk (nebula), experiments (6–8),<br />

natural analogs (9, 10), <strong>and</strong> theoretical calculations<br />

(11, 12) all show that there should be extensive<br />

evaporation of major <strong>and</strong> minor elements, in<br />

the order S > Na, K > Fe > Si > Mg.<br />

Elemental fractionations in chondrules are<br />

generally a function of volatility (4, 5). If evaporation<br />

in the nebula produced the alkali metal <strong>and</strong><br />

Fe fractionations, the more volatile elements (such<br />

as S) should be entirely absent, which they are not.<br />

In addition, the fractionated elements should exhibit<br />

large <strong>and</strong> systematic isotopic fractionations,<br />

which they do not (13).<br />

Here we demonstrate that chondrules did indeed<br />

behave as essentially closed systems during<br />

melting, at least for elements with volatilities less<br />

than or equal to that of Na. We also propose a<br />

means of resolving the apparent conflict between<br />

this result <strong>and</strong> experimental <strong>and</strong> theoretical expec-<br />

22. D. J. Kane, G. Rodriguez, A. J. Taylor, T. S. Clement,<br />

J. Opt. Soc. Am. B 14, 935 (1997).<br />

23. P. Salières et al., Science 292, 902 (2001).<br />

24. V. S. Yakovlev, A. Scrinzi, Phys. Rev. Lett. 91, 153901<br />

(2003).<br />

25. P. B. Corkum, Phys. Rev. Lett. 71, 1994 (1993).<br />

26. K. J. Schafer, B. Yang, L. F. DiMauro, K. C. Kul<strong>and</strong>er,<br />

Phys. Rev. Lett. 70, 1599 (1993).<br />

27. R. López-Martens et al., Phys. Rev. Lett. 94, 033001<br />

(2005).<br />

28. Z.-H. Loh et al., Phys. Rev. Lett. 98, 143601 (2007).<br />

29. K. Varjú et al., Laser Phys. 15, 888 (2005).<br />

30. Supported by the Max Planck Society <strong>and</strong> the<br />

Deutsche Forschungsgemeinschaft Cluster of Excellence:<br />

Munich Centre for Advanced Photonics (www.munichphotonics.de).<br />

E.G. acknowledges a Marie-Curie<br />

fellowship (MEIF-CT-2005-02440) <strong>and</strong> a Marie-Curie<br />

Reintegration grant (MERG-CT-2007-208643). A.L.A.<br />

is supported by the NSF Extreme Ultraviolet Engineering<br />

Research Center. R.K. acknowledges support from<br />

the Sofia Kovalevskaya award of the Alex<strong>and</strong>er von<br />

Humboldt Foundation.<br />

Supporting Online Material<br />

www.sciencemag.org/cgi/content/320/5883/1614/DC1<br />

SOM Text<br />

Figs. S1 to S4<br />

References<br />

17 March 2008; accepted 27 May 2008<br />

10.1126/science.1157846<br />

tations that chondrules should have suffered considerable<br />

evaporation during formation. Our<br />

conclusions have implications for mechanisms of<br />

dust concentration in the solar nebula, for chondrule<br />

formation, <strong>and</strong> for planetesimal formation.<br />

Chondrules are dominated by olivine<br />

[(Mg,Fe)2SiO4], pyroxene [(Mg,Fe,Ca)SiO3],<br />

Fe-Ni metal, <strong>and</strong> quenched silicate melt (glass).<br />

Many of the more volatile elements (such as Na)<br />

can diffuse rapidly, particularly in melts <strong>and</strong><br />

glasses. Therefore, it is possible that volatiles were<br />

completely lost when chondrules melted, <strong>and</strong><br />

reentered the chondrules during cooling or even<br />

after solidification. However, Na clinopyroxene/<br />

glass ratios show that the Na contents of the final<br />

chondrule melts (now glass) had approximately<br />

their observed, relatively high abundances at<br />

temperatures of ~1600 to 1200 K (14–16).<br />

Calculations suggest that chondrule melts<br />

could have been stabilized in the nebula by substantially<br />

enriching solids (chondrule precursors<br />

or other dust) relative to gas (11, 12). This also<br />

substantially increases the condensation temperatures<br />

of even highly volatile elements such as S<br />

(11, 12). Even in solid-enriched systems, there is an<br />

initial phase of evaporation when a chondrule melts,<br />

but subsequent chondrule/gas re-equilibration would<br />

erase any isotopic fractionations (12). If the enrichment<br />

of solids is high, little evaporation may be<br />

needed to reach chondrule/gas equilibrium, <strong>and</strong><br />

the behavior of volatile elements during cooling<br />

would resemble closed-system behavior. However,<br />

even at a high total pressure of 10 −3 bars, with a<br />

solids enrichment of 1000 relative to the solar composition,<br />

all the Na would evaporate at near-liquidus<br />

temperatures, <strong>and</strong> substantial recondensation only<br />

begins well below 1600 K (11). Locally enriching<br />

chondrule-sized or smaller solids by 1000 times<br />

www.sciencemag.org SCIENCE VOL 320 20 JUNE 2008 1617<br />

188 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008<br />

REPORTS<br />

on June 24, 2008<br />

www.sciencemag.org<br />

Downloaded from


1 . 3 . 3 S E L E C T E D R E P R I N T S<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 189


1 . 3 AT T O S E C O N D A N D H I G H - F I E L D D I V I S I O N<br />

New Journal of Physics<br />

T h e o p e n – a c c e s s j o u r n a l f o r p h y s i c s<br />

Intense 1.5-cycle near infrared laser waveforms <strong>and</strong><br />

their use for the generation of ultra-broadb<strong>and</strong><br />

soft-x-ray harmonic continua<br />

A L Cavalieri 1,3 , E Goulielmakis 1 , B Horvath 1 , W Helml 1 ,<br />

M Schultze 1 , M Fieß 1 , V Pervak 2 , L Veisz 1 , V S Yakovlev 2 ,<br />

M Uiberacker 2 , A Apolonski 2 , F Krausz 1 <strong>and</strong> R Kienberger 1,3<br />

1 Max-Planck-Institut für Quantenoptik, Hans-Kopfermann-Str. 1,<br />

D-85748 Garching, Germany<br />

2 Department für Physik, Ludwig Maximilians Universität, Am Coulombwall 1,<br />

D-85748 Garching, Germany<br />

E-mail: adrian.cavalieri@mpq.mpg.de or reinhard.kienberger@mpq.mpg.de<br />

New Journal of Physics 9 (2007) 242<br />

Received 20 June 2007<br />

Published 31 July 2007<br />

Online at http://www.njp.org/<br />

doi:10.1088/1367-2630/9/7/242<br />

Abstract. We demonstrate sub-millijoule-energy, sub-4 fs-duration nearinfrared<br />

laser pulses with a controlled waveform comprised of approximately 1.5<br />

optical cycles within the full-width at half-maximum (FWHM) of their temporal<br />

intensity profile. We further demonstrate the utility of these pulses for producing<br />

high-order harmonic continua of unprecedented b<strong>and</strong>width at photon energies<br />

around 100 eV. Ultra-broadb<strong>and</strong> coherent continua extending from 90 eV to more<br />

than 130 eV with smooth spectral intensity distributions that exhibit dramatic,<br />

never-before-observed sensitivity to the carrier-envelope offset (CEO) phase of<br />

the driver laser pulse were generated. These results suggest the feasibility of<br />

sub-100-attosecond XUV pulse generation for attosecond spectroscopy in the<br />

100 eV range, <strong>and</strong> of a simple yet highly sensitive on-line CEO phase detector<br />

with sub-50-ms response time.<br />

3 Author to whom any correspondence should be addressed.<br />

New Journal of Physics 9 (2007) 242 PII: S1367-2630(07)53044-2<br />

1367-2630/07/010242+12$30.00 © IOP Publishing Ltd <strong>and</strong> Deutsche Physikalische Gesellschaft<br />

190 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008


Vol 449 | 25 October 2007 | doi:10.1038/nature06229<br />

1 . 3 . 3 S E L E C T E D R E P R I N T S<br />

LETTERS<br />

<strong>Attosecond</strong> spectroscopy in condensed matter<br />

A. L. Cavalieri 1 , N. Müller 2 , Th. Uphues 1,2 , V. S. Yakovlev 3 , A. Baltusˇka 1,4 , B. Horvath 1 , B. Schmidt 5 , L. Blümel 5 ,<br />

R. Holzwarth 5 , S. Hendel 2 , M. Drescher 6 , U. Kleineberg 3 , P. M. Echenique 7 , R. Kienberger 1 , F. Krausz 1,3<br />

& U. Heinzmann 2<br />

Comprehensive knowledge of the dynamic behaviour of electrons<br />

in condensed-matter systems is pertinent to the development of<br />

many modern technologies, such as semiconductor <strong>and</strong> molecular<br />

electronics, optoelectronics, information processing <strong>and</strong> photovoltaics.<br />

Yet it remains challenging to probe electronic processes,<br />

many of which take place in the attosecond (1 as 5 10 218 s) regime.<br />

In contrast, atomic motion occurs on the femtosecond (1 fs 5<br />

10 215 s) timescale <strong>and</strong> has been mapped in solids in real time 1,2<br />

using femtosecond X-ray sources 3 . Here we extend the attosecond<br />

techniques 4,5 previously used to study isolated atoms in the gas<br />

phase to observe electron motion in condensed-matter systems<br />

<strong>and</strong> on surfaces in real time. We demonstrate our ability to obtain<br />

direct time-domain access to charge dynamics with attosecond<br />

resolution by probing photoelectron emission from single-crystal<br />

tungsten. Our data reveal a delay of approximately 100 attoseconds<br />

between the emission of photoelectrons that originate<br />

from localized core states of the metal, <strong>and</strong> those that are freed<br />

from delocalized conduction-b<strong>and</strong> states. These results illustrate<br />

that attosecond metrology constitutes a powerful tool for exploring<br />

not only gas-phase systems, but also fundamental electronic<br />

processes occurring on the attosecond timescale in condensedmatter<br />

systems <strong>and</strong> on surfaces.<br />

Photoemission spectroscopy is based on the photoelectric effect,<br />

first explained by Einstein more than 100 years ago 6 . According to<br />

Einstein’s law, photoelectrons ejected from a metal surface by light<br />

will have a kinetic energy that depends on the incident photon energy<br />

<strong>and</strong> the electron’s original bound state energy. Photoelectron spectra<br />

will thus provide information about the electronic structure of the<br />

metal if well-characterized light sources are used 7 . Indeed, experiments<br />

using a broad range of photon energies have now been performed<br />

to determine the steady-state electronic properties of many<br />

bulk materials, thin films, <strong>and</strong> surfaces. The photoemission process<br />

itself involves three steps: excitation, transport, <strong>and</strong> ultimately escape<br />

of the photoelectron through the surface 8 . Here, in a proof-ofprinciple<br />

experiment, we combine this spectroscopy with attosecond<br />

temporal resolution to obtain time-domain insight into the electron<br />

transport stage of the photoemission process. The measurements<br />

represent (to our knowledge) the first direct attosecond timeresolved<br />

observation of electron transport in a condensed-matter<br />

system, <strong>and</strong> we expect that they will trigger other experimental<br />

research into the dynamics of processes that have attracted interest<br />

in solid-state <strong>and</strong> surface science. Such processes include charge<br />

transfer 9,10 , charge screening 11 , image charge creation <strong>and</strong> decay 12 ,<br />

electron–electron scattering 13 , <strong>and</strong> collective electronic motion 14 .<br />

Time-resolved photoemission spectroscopy was originally implemented<br />

in the picosecond (1 ps 5 10 212 s) to femtosecond regime,<br />

using first visible 15–17 <strong>and</strong> then extreme ultraviolet (XUV) 18,19 radiation.<br />

These experiments utilize one light pulse to trigger the<br />

dynamics, followed by a second light pulse to induce photoemission<br />

<strong>and</strong> thereby probe the transient state. Experiments using the laserassisted<br />

photoelectric effect have been carried out 20–22 ; but the XUV<br />

photoemission lasted over several wave cycles of the coincident nearinfrared<br />

(NIR) light, limiting the time resolution to .10 fs. To overcome<br />

this limitation, we use single sub-femtosecond XUV pulses 4,5<br />

for pumping, <strong>and</strong> coincident NIR waveform-controlled few-cycle<br />

laser pulses 23 as a probe 5 . The XUV pulse triggers the photoemission<br />

process, with only those photoelectron wave packets initiated from<br />

the uppermost atomic layers escaping without inelastic collision.<br />

Delayable<br />

XUV mirror<br />

To preparation<br />

chamber<br />

Tungsten<br />

crystal<br />

Zirconium foil,150 nm,<br />

on pellicle<br />

1 Max-Planck-Institut für Quantenoptik, Hans-Kopfermann-Str. 1, D-85748 Garching, Germany. 2 Fakultät für Physik, Universität Bielefeld, D-33615 Bielefeld, Germany. 3 Department<br />

für Physik, Ludwig-Maximilians-Universität, Am Coulombwall 1, D-85748 Garching, Germany. 4 Institut für Photonik, Technische Universität Wien, Gußhausstr. 27, A-1040 Wien,<br />

Austria. 5 Menlo Systems GmbH, Am Klopferspitz 19, D-82152 Martinsried, Germany. 6 Institut für Experimentalphysik, Universität Hamburg, Luruper Chaussee 149, D-22761<br />

Hamburg, Germany. 7 Dpto. Fisica de Materiales UPV/EHU, Centro Mixto CSIC-UPV/EHU <strong>and</strong> Donostia International Physics Center (DPIC), Paseo Manual de Lardizabal 4, 20018<br />

San Sebastian, Spain.<br />

Silver<br />

mirror<br />

©2007 Nature Publishing Group<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 191<br />

e –<br />

TOF<br />

Neon-filled<br />

tube<br />

Figure 1 | Experimental set-up. Waveform-controlled, ,5-fs, 750-nm, 400-mJ<br />

laser pulses are focused with a mirror of 500-mm focal length into a ,2-mmdiameter<br />

tube filled with neon to generate XUV radiation by high-harmonic<br />

generation. The collinear XUV <strong>and</strong> NIR beams co-propagate towards a cored<br />

two-part mirror in the measurement chamber that is maintained under<br />

ultrahigh vacuum in a 1-m-length differential vacuum pumping stage, until<br />

the beams are separated by a pellicle/zirconium foil assembly. The zirconium<br />

foil transmits the XUV but blocks the NIR. The XUV radiation, indicated by<br />

the blue beam, is incident on a 6 eV (full-width at half-maximum FWHM)<br />

broad multilayer b<strong>and</strong>-pass reflector centred at ,91 eV, which is mounted on<br />

a piezo-electric delay stage. With a proper XUV spectrum, the multilayer<br />

mirror reflects <strong>and</strong> focuses ,300-as (FWHM) XUV pulses. The NIR pulse,<br />

indicated by the violet beam, is reflected by a stationary (silver) outer annular<br />

mirror confocal with the inner mirror (f 5 12.5 cm). Both pulses are focused<br />

onto the (110) surface of a tungsten single crystal that is mounted on a<br />

manipulator to control the angle of incidence. The manipulator is also used to<br />

retract the crystal into a preparatory chamber for cleaning. Resultant XUVinduced<br />

photoemission, which is detected by the time-of-flight spectrometer<br />

(TOF), is streaked by the coincident NIR laser pulse.<br />

1029


1 . 3 AT T O S E C O N D A N D H I G H - F I E L D D I V I S I O N<br />

LETTERS NATURE |Vol 449 | 25 October 2007<br />

Photoexcited electron wave packets propagate through the material<br />

in upper conduction b<strong>and</strong>s, ultimately leaving the surface with an<br />

average kinetic energy determined by the XUV photon energy, the<br />

initial binding energy, <strong>and</strong> the material work function. An attosecond<br />

transient recorder (ATR), previously developed <strong>and</strong> used in gasphase<br />

experiments 5 , is used to observe the emitted photoelectron<br />

wave packet. In this scheme, the photoelectron momentum is further<br />

influenced by the electric field of a coincident few-cycle NIR laser<br />

pulse, giving rise to a ‘streaked’ final momentum distribution 24,25 .<br />

An ATR spectrogram is compiled by measuring a series of streaked<br />

photoelectron spectra with a time-of-flight detector, recorded as a<br />

function of time delay between the XUV pump <strong>and</strong> NIR streaking<br />

field. Important characteristics of the emitted electron wave packets,<br />

including their duration <strong>and</strong> frequency sweep, or ‘chirp’, can be determined<br />

from the spectra 24,25 . If measured for two or more different types<br />

of electrons, the complete ATR spectrograms can also yield relative<br />

timing information about the arrival of the wave packets on the surface,<br />

because the streaking effect is negligible until the electrons emerge<br />

from the surface (see Methods). The resolution of the ATR depends on<br />

the duration of the XUV excitation, the gradient of the streaking NIR<br />

field, <strong>and</strong> the signal-to-noise ratio in the photoelectron spectra.<br />

In comparison to experiments performed on isolated atoms, ATR<br />

measurements in condensed-matter systems are more complicated<br />

because the photoelectron wave packets can be released from energy<br />

a<br />

Kinetic energy (eV) Kinetic energy (eV)<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

8 7 6 5 4 3 2 1 0<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

Raw spectrum (1)<br />

ATI subtracted <strong>and</strong> smoothed (1)<br />

Streaked spectrum (2)<br />

Conduction-b<strong>and</strong><br />

photoemission<br />

4f-states<br />

photoemission<br />

c d<br />

Measured streaked spectrum<br />

Reconstructed spectrum<br />

30<br />

8 7 6 5 4 3 2 1 0<br />

Intensity (a.u.)<br />

Figure 2 | <strong>Attosecond</strong> time-resolved photoemission spectra.<br />

a, Photoelectron spectra collected at two different relative delays. The<br />

spectrum recorded at the delay indicated by the white dashed line labelled<br />

‘(1)’ in b is far from the zero of delay as defined by the overlap of the<br />

maximums of the NIR <strong>and</strong> XUV pulse envelopes. This spectrum shows<br />

pronounced peaks corresponding to the 4f-state <strong>and</strong> conduction-b<strong>and</strong><br />

photoemission. The blue line shows the raw spectrum as recorded by the<br />

time-of-flight detector, <strong>and</strong> the red line shows the corresponding spectrum<br />

after subtraction of NIR-induced ATI background <strong>and</strong> numerical<br />

smoothing. The 4f photoemission is peaked near 56 eV. The conductionb<strong>and</strong><br />

photoemission is peaked near 83 eV, owing to a high density of d-b<strong>and</strong><br />

conducting states just below the Fermi energy. Ef denotes the kinetic energy<br />

of a photoelectron excited from the Fermi energy level. The other spectrum<br />

in a (displayed only after ATI subtraction <strong>and</strong> smoothing) was recorded near<br />

zero delay, as indicated by the white dashed line labelled ‘(2)’. At this delay,<br />

1030<br />

E f<br />

b<strong>and</strong>s containing many distinct states rather than from a single,<br />

isolated energy level. Unoccupied conduction-b<strong>and</strong> states just above<br />

the Fermi energy (defined by the highest occupied energy level in the<br />

absence of thermal excitation) might become populated by singlephoton<br />

absorption of the leading edge of the NIR probe field prior to<br />

XUV photoemission, which is an unwanted complication. In contrast<br />

to conduction-b<strong>and</strong> states, the localized 4f core states of tungsten<br />

are deeply bound <strong>and</strong> fully populated. Therefore, these states are<br />

unsusceptible to this potential influence of the streaking field, <strong>and</strong><br />

constitute an ideal test case for proof of the extension of attosecond<br />

metrology to solids.<br />

As a further challenge, above-threshold ionization (ATI) by the<br />

streaking field can, in condensed matter, generate energetic photoelectrons,<br />

obscuring detection of XUV-induced photoelectrons. ATI<br />

is favoured by the low work function of metals (as compared to the<br />

relatively high ionization potential of isolated atoms), which limits<br />

the intensity of the applied streaking field to levels far below those<br />

that can be used in gas-phase experiments.<br />

As indicated by the diagram of the experimental set-up in Fig. 1,<br />

streaked photoemission spectra from a tungsten (110) crystal surface<br />

were recorded by collecting electrons within a narrow cone aligned<br />

perpendicularly to the surface. The relative delay between the XUV<br />

pulse <strong>and</strong> the NIR waveform-controlled streaking field was varied in<br />

300-as steps in a sequence chosen to minimize systematic error, with<br />

–6 –4 –2 0<br />

192 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008<br />

b<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

(1)<br />

–6<br />

©2007 Nature Publishing Group<br />

(2)<br />

(2)<br />

–4 –2 0<br />

Relative delay (fs)<br />

2 4<br />

2 4<br />

the XUV pulse peak coincides with the NIR field maximum on our target.<br />

Consequently, the NIR vector potential crosses zero, giving rise to the<br />

strongest streaking <strong>and</strong> a clear broadening of the photoemission peaks is<br />

observed. b, The full ATI subtracted spectrogram for both the 4f-states <strong>and</strong><br />

conduction-b<strong>and</strong> photoemission. The streaking waveform (vector potential)<br />

is evident in both spectrograms, proving the extension of attosecond<br />

metrology to condensed-matter systems. We can expect that the energy<br />

modulation of the conduction-b<strong>and</strong> peak should be a little larger than that of<br />

the 4f peak owing to their initial photoelectron kinetic energy. The<br />

amplitude of the spectral shift has a square-root dependence on initial<br />

kinetic energy (KE) 24 , yielding an expected ratio in spectral shift of<br />

p ffiffiffiffiffiffiffiffiffiffiffiffiffi.<br />

p ffiffiffiffiffiffiffiffiffi<br />

KEcond KE4f


spectra integrated for 60 s at each delay. (For a brief description of the<br />

experiment, see Methods Summary; full details regarding set-up,<br />

measurements <strong>and</strong> data analysis are provided as Supplementary<br />

Information.) Characteristic spectra obtained with our system are<br />

shown in Fig. 2a, indicating that emission from the conduction b<strong>and</strong><br />

occurs at a kinetic energy of ,83 eV while emission from the localized<br />

4f states occurs at ,56 eV. At kinetic energies significantly<br />

below the 4f peak, the measured spectrum is due to NIR-induced<br />

ATI photoelectrons <strong>and</strong> XUV-generated photoelectrons that have<br />

undergone inelastic scattering.<br />

The two distinct background components were distinguished by<br />

recording an additional photoelectron spectrum without the NIR<br />

streaking field. The ATI component was subsequently subtracted<br />

from the measured data (see Supplementary Information). This subtraction<br />

is illustrated for a fixed delay in Fig. 2a, <strong>and</strong> was performed at<br />

each of the delay steps, resulting in the full spectrogram presented in<br />

Fig. 2b. Here, a positive relative delay corresponds to the XUV pulse<br />

arriving earlier with respect to the streaking field at the surface. Both<br />

the 4f <strong>and</strong> conduction-b<strong>and</strong> photoemission exhibit a pronounced<br />

periodic upshift <strong>and</strong> downshift in energy as a function of relative<br />

delay <strong>and</strong>, as in previous gas-phase experiments, the spectrogram<br />

reveals the waveform (vector potential) of the streaking field 5,26,27 .<br />

Our ability to resolve the field oscillation indicates that the photoemission<br />

from the 4f core states <strong>and</strong> from the conduction b<strong>and</strong> is subfemtosecond<br />

in duration, <strong>and</strong> proves that attosecond metrology has<br />

been successfully extended to condensed-matter systems.<br />

Further examination reveals that the 4f spectrogram is shifted<br />

along the delay coordinate with respect to the conduction-b<strong>and</strong><br />

spectrogram. This effect is readily apparent upon inspection of the<br />

smoothed spectrograms that are obtained by interpolation of the<br />

measured data <strong>and</strong> shown in Fig. 3a. We quantify the temporal shift<br />

in the measured data by evaluating, for each delay step, the centre-ofmass<br />

(COM) of the spectral regions spanning the 4f <strong>and</strong> conductionb<strong>and</strong><br />

peaks that cover the energy intervals 47–66 eV <strong>and</strong> 66–110 eV,<br />

respectively. Characterizing the periodic motion of the peaks through<br />

their COM requires no assumptions or fitting parameters, yet yields<br />

timing information that is invariant to fluctuations in the instantaneous<br />

laser parameters. The approach is also relatively insensitive to<br />

inelastic scattered background photoelectrons, which could not be<br />

subtracted from our measurements. As a result, the COM accurately<br />

describes the streaking-induced time-dependence of the energy shift<br />

of the 4f <strong>and</strong> conduction-b<strong>and</strong> peaks, as shown in Fig. 3b.<br />

By comparing the COM trajectories of the 4f <strong>and</strong> conduction b<strong>and</strong><br />

at the seven zero-crossings of the vector potential, we obtain seven<br />

independent measurements of their relative timing. This yields a<br />

temporal shift of Dt 5 110 6 70 as between the ATR spectrograms<br />

of the conduction-b<strong>and</strong> <strong>and</strong> 4f photoelectrons. (The error estimate<br />

results from a straightforward extrapolation of the error in calculating<br />

the COM; see Supplementary Information.) This shift or delay<br />

was observed in different independent measurements made at different<br />

locations on the tungsten sample, with the results corroborating<br />

the above value of Dt to within the measurement error. We note<br />

that the rather large error associated with our Dt value could be most<br />

effectively reduced in future measurements by using higher XUV<br />

photon energies <strong>and</strong> fluxes.<br />

The shift between the two spectrograms indicates that, on average,<br />

photoelectrons originating from the localized 4f states emerge from<br />

the tungsten surface approximately 100 as later than those originating<br />

from the delocalized conduction b<strong>and</strong>—even though the<br />

photoemission process for both types of electrons is initiated simultaneously<br />

by the same XUV pulse. The delay effect thus occurs during<br />

transport of the excited photoelectrons to the surface, illustrating<br />

that our technique provides a means to directly observe features of<br />

electron wave packet propagation towards the surface with attosecond<br />

precision.<br />

By adapting a quantum mechanical model used in previous gasphase<br />

streaking experiments 28 , we are able to reconstruct the measured<br />

1 . 3 . 3 S E L E C T E D R E P R I N T S<br />

NATURE | Vol 449 |25 October 2007 LETTERS<br />

spectra <strong>and</strong> spectrograms. The modelling of the streaking experiment<br />

requires some assumptions, leaving several parameters (such as duration<br />

of the electron wave packets, their chirp, <strong>and</strong> their emission time)<br />

for optimization. Figure 2c <strong>and</strong> d shows the reconstructions that best<br />

agree with experiment. These were obtained for wave packets with a<br />

duration of ,300 as (full-width at half-maximum, FWHM) <strong>and</strong><br />

assuming a delay of ,100 as between the emission times of the electron<br />

wave packets, which supports the conclusions drawn from the COM<br />

analysis.<br />

Our measurements also indicate that electron wave packets<br />

launched from both the localized 4f <strong>and</strong> delocalized conduction-b<strong>and</strong><br />

states are nearly undistorted on propagation to the surface. To explain<br />

the observed delay, we consider the group velocities for the two different<br />

photoelectron wave packets travelling in the solid. The crucial point<br />

is that after absorption of an XUV photon, the electron is excited<br />

into an upper conduction b<strong>and</strong> region that depends on the electron’s<br />

Energy shift (eV) 4f-states kinetic energy (eV)<br />

65<br />

63<br />

61<br />

59<br />

57<br />

55<br />

1<br />

0.5<br />

0<br />

–0.5<br />

–6<br />

–1 –6<br />

©2007 Nature Publishing Group<br />

a<br />

b<br />

4f states<br />

Cond. b<strong>and</strong><br />

∆t = 110 ± 70 as<br />

–4 –2 0 2 4<br />

Relative delay (fs)<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 193<br />

∆t<br />

–4 –2 0 2 4<br />

Relative delay (fs)<br />

91<br />

89<br />

87<br />

85<br />

83<br />

Cond.-b<strong>and</strong> kinetic<br />

energy (eV)<br />

Figure 3 | Evidence of delayed photoemission. a, The 4f <strong>and</strong> conductionb<strong>and</strong><br />

spectrograms, following cubic-spline interpolation of the measured data<br />

(but without background subtraction). The spectral region between ,65 eV<br />

<strong>and</strong> ,83 eV has been omitted to more easily compare the edges of the 4f <strong>and</strong><br />

conduction-b<strong>and</strong> peaks. A small shift in the relative delay is evident, as<br />

indicated by the white dashed lines through the fringes, <strong>and</strong> can be seen at<br />

each fringe maximum <strong>and</strong> minimum. Quantification of the shift of the 4f with<br />

respect to the conduction-b<strong>and</strong> spectrogram is made by COM analysis, <strong>and</strong> is<br />

summarized in b. The energy intervals, within which the COMs were<br />

calculated, are 47–66 eV for the 4f photoemission peak <strong>and</strong> 66–110 eV for the<br />

conduction-b<strong>and</strong> photoemission peak. Vertical error bars (61 s.d.) are<br />

calculated from noise in the measured spectra (see Supplementary<br />

Information for details). For ease of visual comparison, the COM energy shift<br />

of the 4f spectral region was scaled by a factor of 2.5, to offset the stabilizing<br />

effect of the background plateau underneath the 4f peak (see Supplementary<br />

Information), in order to illuminate the ,100 as delay in emission. Rescaling<br />

these COM data points along the energy axis cannot influence the measured<br />

delay. The COM data points were fitted with a damped sinusoid, which<br />

corresponds to the NIR streaking field, to guide the eye.<br />

1031


1 . 3 AT T O S E C O N D A N D H I G H - F I E L D D I V I S I O N<br />

LETTERS NATURE |Vol 449 | 25 October 2007<br />

initial binding energy <strong>and</strong> the XUV photon energy. If the relationship<br />

between momentum <strong>and</strong> energy of such an electron were that of a free<br />

electron, the ratio of velocities of the conduction-b<strong>and</strong> <strong>and</strong> 4f electrons<br />

travelling with energies of 85 eV <strong>and</strong> 58 eV, respectively, would be a<br />

factor of ,1.2 (energy is defined with respect to the Fermi energy inside<br />

the material). However, elastic interactions with atoms in the crystal<br />

lattice, which give rise to electronic b<strong>and</strong> structure, modify the<br />

momentum–energy relationship. This implies that in tungsten for<br />

XUV photon energies of ,91 eV, the mean velocity of the conductionb<strong>and</strong><br />

photoelectrons is approximately twice that of the 4f photoelectrons<br />

(see Supplementary Information). We also note that due to<br />

their longer inelastic mean free paths 29 , the slow 4f photoelectrons<br />

originate, on average, from ,1 A ˚ deeper in the tungsten crystal than<br />

the fast conduction-b<strong>and</strong> photoelectrons. On the basis of these<br />

considerations, we estimated the absolute delay between the initial<br />

excitation of a photoelectron <strong>and</strong> its escape through the surface to<br />

be ,60 as <strong>and</strong> ,150 as for the conduction <strong>and</strong> 4f photoelectrons,<br />

respectively. These values suggest a relative delay between photoelectron<br />

emissions from the surface of ,90 as, which is in good<br />

agreement with the observed delay in the emission of the 4f <strong>and</strong><br />

conduction-b<strong>and</strong> photoelectrons.<br />

In summary, our observations demonstrate the successful<br />

extension of attosecond metrology to condensed-matter systems.<br />

Although the current experimental apparatus provides access to only<br />

the relative group delay in electron wave packet propagation, future<br />

measurements of absolute emission delays may be feasible using the<br />

same methods, for example by direct comparison with gas-phase<br />

ATR data. At this point, attosecond photoemission spectroscopy<br />

presents a clear path toward ultimately uncovering the intermediate<br />

processes leading to ejection of a photoelectron. This information<br />

might shed new light on data previously obtained with conventional<br />

time-integral photoemission spectroscopy <strong>and</strong> allow for accurate<br />

description of charge dynamics on the electronic timescale in both<br />

condensed matter <strong>and</strong> on surfaces.<br />

METHODS SUMMARY<br />

A 1-kHz-repetition-rate, waveform-controlled, few-cycle, ,5-fs, 400-mJ, 750-nm<br />

Ti:sapphire laser system is the front-end of our apparatus <strong>and</strong> is used in combination<br />

with proper spectral filtering to efficiently generate isolated attosecond<br />

pulses of XUV radiation by high-harmonic generation. As shown in Fig. 1, the<br />

XUV <strong>and</strong> NIR pulses co-propagate towards a two-part focusing mirror at near<br />

normal incidence. The inner component is a Mo/Si multilayer mirror <strong>and</strong> reflects<br />

the XUV radiation over a b<strong>and</strong>width of ,6 eV (FWHM) centred at ,91 eV,<br />

supporting 300-as transform-limited pulses 30 . The XUV mirror is mounted on a<br />

translation stage, providing a precise delay between the XUV pump <strong>and</strong> the NIR<br />

streaking pulse. The temporal resolution that can be achieved in pump-probe<br />

experiments using these pulses <strong>and</strong> this apparatus is expected to be a small<br />

fraction of the pulse half-width because XUV pulses generated with similar<br />

spectra <strong>and</strong> multilayer optics have previously been fully characterized <strong>and</strong><br />

observed to be gaussian 24 .<br />

The tungsten surface must be sufficiently free of contamination to minimize<br />

photoelectron scattering on emission. Therefore, the measurement chamber is<br />

maintained under ultrahigh vacuum conditions with typical background pressures<br />

of , 10 29 mbar, which suppresses the accumulation of contaminates on<br />

the crystal surface to a level permitting a full ATR spectrogram to be recorded<br />

without interruption.<br />

In our application of the ATR, detection of electrons occurs in the direction<br />

normal to the tungsten crystal surface, <strong>and</strong> the NIR streaking field is incident on<br />

the tungsten (110) crystal surface near Brewster’s angle (,75u). For this angle of<br />

incidence, owing to the refractive index of tungsten, photoelectron wave packets<br />

are not efficiently accelerated in the direction of observation until they emerge<br />

from the surface. Even though the streaking field penetrates the tungsten crystal,<br />

inside the material the electric field component along the surface normal is<br />

weaker by a factor of approximately 16, allowing us to neglect streaking effects<br />

until the electron wave packets emerge from the surface. The absence of effective<br />

streaking until the photoelectron emerges from the surface is generally the case<br />

for solids, allowing us to time processes occurring within the material.<br />

Additional, detailed description of the experimental apparatus, measurement<br />

technique, <strong>and</strong> data analysis is provided in Supplementary Information.<br />

1032<br />

©2007 Nature Publishing Group<br />

Received 20 June; accepted 3 September 2007.<br />

1. Reis, D. A. & Lindenberg, A. M. in Light Scattering in Solids IX (eds Cardona, M. &<br />

Merlin, R.) 371–422 (Topics in Applied Physics 108, Springer, Berlin, 2007).<br />

2. Fritz, D. M. et al. Ultrafast bond softening in bismuth: Mapping a solid’s<br />

interatomic potential with X-rays. Science 315, 633–636 (2007).<br />

3. Pfeifer, T., Spielmann, C. & Gerber, G. Femto-second X-ray science. Rep. Prog.<br />

Phys. 69, 443–505 (2006).<br />

4. Hentschel, M. et al. <strong>Attosecond</strong> metrology. Nature 414, 509–513 (2001).<br />

5. Kienberger, R. et al. Atomic transient recorder. Nature 427, 817–821 (2004).<br />

6. Einstein, A. Über einen die Erzeugung und Verw<strong>and</strong>lung des Lichts betreffenden<br />

heuristischen Gesichtspunkt. Ann. Phys. 17, 132–148 (1905).<br />

7. Siegbahn, K. Electron-spectroscopy — outlook. J. Electron Spectrosc. Relat. Phenom.<br />

5, 3–97 (1974).<br />

8. Berglund, C. N. & Spicer, W. E. Photoemission studies of copper <strong>and</strong> silver: theory.<br />

Phys. Rev. 136, A1030–A1044 (1964).<br />

9. Brühwiler, P. A., Karis, O. & Martensson, N. Charge-transfer dynamics studied<br />

using resonant core spectroscopies. Rev. Mod. Phys. 74, 703–740 (2002).<br />

10. Föhlisch, A. et al. Direct observation of electron dynamics in the attosecond<br />

domain. Nature 436, 373–376 (2005).<br />

11. Borisov, A., Sánchez-Portal, D., Díez-Muino, R. & Echenique, P. M. Dimensionality<br />

effects in time-dependent screening. Chem. Phys. Lett. 387, 132–137 (2004).<br />

12. Huber, R. et al. How many-particle interactions develop after ultrafast excitation<br />

of an electron-hole plasma. Nature 414, 286–289 (2001).<br />

13. Haight, R. Electron dynamics at surfaces. Surf. Sci. Rep. 21, 277–325 (1995).<br />

14. Cavalleri, A. et al. Tracking the motion of charges in a terahertz light field by<br />

femtosecond X-ray diffraction. Nature 442, 664–666 (2006).<br />

15. Yen, R. et al. Picosecond laser interaction with metallic zirconium. Appl. Phys. Lett.<br />

40, 185–187 (1982).<br />

16. Höfer, U. et al. Time-resolved coherent photoelectron spectroscopy of quantized<br />

electronic states on metal surfaces. Science 277, 1480–1482 (1997).<br />

17. Petek, H. & Ogawa, S. Femtosecond time-resolved two-photon photoemission<br />

studies of electron dynamics in metals. Prog. Surf. Sci. 56, 239–310 (1997).<br />

18. Haight, R. & Peale, D. R. Tunable photoemission with harmonics of subpicosecond<br />

lasers. Rev. Sci. Instrum. 65, 1853–1857 (1994).<br />

19. Siffalovic, P. et al. Laser-based apparatus for extended ultraviolet femtosecond<br />

time-resolved photoemission spectroscopy. Rev. Sci. Instrum. 72, 30–35 (2001).<br />

20. Schins, J. M. et al. Observation of laser-assisted Auger decay in argon. Phys. Rev.<br />

Lett. 73, 2180–2183 (1994).<br />

21. Glover, T. E., Schoenlein, R. W., Chin, A. H. & Shank, C. V. Observation of laser<br />

assisted photoelectric effect <strong>and</strong> femtosecond high order harmonic radiation.<br />

Phys. Rev. Lett. 73, 2180–2183 (1994).<br />

22. Miaja-Avila, L. et al. Laser-assisted photoelectric effect from surfaces. Phys. Rev.<br />

Lett. 97, 113604 (2006).<br />

23. Baltuska, A. et al. <strong>Attosecond</strong> control of electronic processes by intense light<br />

fields. Nature 422, 611–615 (2003).<br />

24. Quéré, F., Mairesse, Y. & Itatani, J. Temporal characterization of attosecond XUV<br />

fields. J. Mod. Opt. 52, 339–360 (2005).<br />

25. Yakovlev, V., Bammer, F. & Scrinzi, A. <strong>Attosecond</strong> streaking measurements.<br />

J. Mod. Opt. 52, 395–410 (2005).<br />

26. Goulielmakis, E. et al. Direct measurement of light waves. Science 305, 1267–1269<br />

(2004).<br />

27. Sansone, G. et al. Isolated single-cycle attosecond pulses. Science 314, 443–446<br />

(2006).<br />

28. Kitzler, M., Milosevic, N., Scrinzi, A., Krausz, F. & Brabec, T. Quantum theory of<br />

attosecond XUV pulse measurement by laser dressed photoionization. Phys. Rev.<br />

Lett. 88, 173904 (2002).<br />

29. Tanuma, S., Powell, C. J. & Penn, D. R. Calculations of electron inelastic mean free<br />

paths. 2. Data for 27 elements over the 50–2000-eV range. Surf. Interface Anal. 17,<br />

911–926 (1991).<br />

30. Wonisch, A. et al. Design, fabrication, <strong>and</strong> analysis of chirped multilayer mirrors<br />

for reflection of extreme-ultraviolet attosecond pulses. Appl. Opt. 45, 4147–4156<br />

(2006).<br />

Supplementary Information is linked to the online version of the paper at<br />

www.nature.com/nature.<br />

Acknowledgements We thank W. Hachmann for expeditious preparation of the<br />

XUV multilayer optical substrate. We acknowledge partial financial support by the<br />

Deutsche Forschungsgemeinschaft through the DFG Cluster of Excellence Munich<br />

Centre for Advanced Photonics, <strong>and</strong> through the SFB 613, <strong>and</strong> by the Volkswagen<br />

Stiftung Germany, <strong>and</strong> by the EURYI scheme award. P.M.E. acknowledges support<br />

from the Basque <strong>and</strong> Spanish Governments. R.K. acknowledges a fellowship from<br />

the Austrian Academy of Sciences <strong>and</strong> additional support from the Sofja<br />

Kovalevskaja Award of the Alex<strong>and</strong>er von Humboldt Foundation. The apparatus to<br />

generate attosecond pulses was constructed at Technische Universität Wien,<br />

thanks to the support of the FWF.<br />

Author Information Reprints <strong>and</strong> permissions information is available at<br />

www.nature.com/reprints. Correspondence <strong>and</strong> requests for materials should be<br />

addressed to A.L.C. (adrian.cavalieri@mpq.mpg.de) or F.K. (krausz@lmu.de) or<br />

U.H. (uheinzm@physik.uni-bielefeld.de).<br />

194 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008


1 . 3 . 3 S E L E C T E D R E P R I N T S<br />

<strong>Attosecond</strong> nanoplasmonic-field<br />

microscope<br />

MARK I. STOCKMAN1,2 *, MATTHIAS F. KLING2 , ULF KLEINEBERG3 AND FERENC KRAUSZ2,3 *<br />

1Department of Physics <strong>and</strong> Astronomy, Georgia State University, Atlanta, Georgia 30303, USA<br />

2Max-Planck-Institut für Quantenoptik, Hans-Kopfermann-Straße 1, D-85748 Garching, Germany<br />

3Ludwig-Maximilians-Universität München, Department für Physik, Am Coulombwall 1, D-85748 Garching, Germany<br />

*e-mail: mstockman@gsu.edu; ferenc.krausz@mpq.mpg.de<br />

Published online: 3 September 2007; doi:10.1038/nphoton.2007.169<br />

Nanoplasmonics deals with collective electronic dynamics on the surface of metal nanostructures, which arises as a result of<br />

excitations called surface plasmons. This field, which has recently undergone rapid growth, could benefit applications such as<br />

computing <strong>and</strong> information storage on the nanoscale, the ultrasensitive detection <strong>and</strong> spectroscopy of physical, chemical <strong>and</strong><br />

biological nanosized objects, <strong>and</strong> the development of optoelectronic devices. Because of their broad spectral b<strong>and</strong>width, surface<br />

plasmons undergo ultrafast dynamics with timescales as short as a few hundred attoseconds. So far, the spatiotemporal dynamics<br />

of optical fields localized on the nanoscale has been hidden from direct access in the real space <strong>and</strong> time domain. Here, we<br />

propose an approach that will, for the first time, provide direct, non-invasive access to the nanoplasmonic collective dynamics,<br />

with nanometre-scale spatial resolution <strong>and</strong> temporal resolution on the order of 100 attoseconds. The method, which combines<br />

photoelectron emission microscopy <strong>and</strong> attosecond streaking spectroscopy, offers a valuable way of probing nanolocalized optical<br />

fields that will be interesting both from a fundamental point of view <strong>and</strong> in light of the existing <strong>and</strong> potential applications<br />

of nanoplasmonics.<br />

Recently, significant attention has been devoted to the development<br />

of attosecond science <strong>and</strong> technology, in particular regarding pulse<br />

generation, physical-system excitation, detection, spectroscopy <strong>and</strong><br />

electron-motion control on the attosecond timescale 1–13 . In<br />

nanoscience, which is another rapidly evolving field, an important<br />

issue is the study <strong>and</strong> exploitation of effects that are both ultrafast<br />

<strong>and</strong> localized on the nanoscale. The localization length of surface<br />

plasmon eigenmodes in nanoplasmonic systems is determined by<br />

the size of the constituent nanoparticles <strong>and</strong> can be on the order<br />

of several nanometres 14 . Furthermore, the relaxation rate of the<br />

surface plasmon polarization is in the 10–100 fs range across the<br />

plasmonic spectrum, allowing coherent control of nanoscale<br />

energy localization with femtosecond laser light 15–25 . Importantly,<br />

collective motion in nanoplasmonic systems unfolds on much<br />

shorter attosecond timescales, as defined by the inverse spectral<br />

b<strong>and</strong>width of the plasmonic resonant region. In this paper we<br />

propose an approach that enables direct measurement of the<br />

spatiotemporal dynamics of nanolocalized optical fields with<br />

100-as temporal resolution <strong>and</strong> nanometre-scale spatial resolution.<br />

Nanoplasmonic fields could be used in place of electrons in<br />

ultrafast computation <strong>and</strong> information storage on the nanoscale,<br />

where the fields act as the medium that carries, processes <strong>and</strong><br />

stores information.<br />

The proposed attosecond nanoplasmonic-field microscope<br />

combines two modern techniques: photoelectron emission<br />

microscopy <strong>and</strong> attosecond streaking spectroscopy 26 . A plasmonic<br />

nanostructure is excited by an intense, waveform-controlled field<br />

in the optical spectral range that drives collective electron<br />

oscillations (the quantum of which is called a surface plasmon).<br />

This generates optical fields localized on the nanometre scale,<br />

which we will refer to as nanolocalized optical fields. An<br />

ARTICLES<br />

attosecond extreme ultraviolet (XUV) pulse, which is produced<br />

from <strong>and</strong> synchronized with the driving optical pulse, is then<br />

sent to the system. This XUV pulse produces photoelectrons that,<br />

owing to their large energy <strong>and</strong> short emission time (determined<br />

by the attosecond-pulse duration), escape from the nanometresized<br />

regions of local electric fields enhanced by plasmon<br />

resonances within a fraction of the oscillation period of the<br />

driven plasmonic field. The ultrashort escape time implies a final<br />

energy change of the emitted photoelectrons that is proportional<br />

to the local electric potential at the surface at the instant of<br />

electron release. This is in sharp contrast to previous attosecond<br />

streaking experiments performed in macroscopic volumes of gasphase<br />

media, where the electron escape times are longer than the<br />

optical period; consequently, the change in electron energy<br />

probes the vector potential of the optical field 26 . For<br />

nanoplasmonic systems, the imaging of the emitted XUVinduced<br />

photoelectrons by an energy-resolving photoelectron<br />

emission microscope (PEEM) probes the electric-field potential<br />

at the surface as a function of the XUV-pulse incidence time <strong>and</strong><br />

the position at the surface with an attosecond temporal <strong>and</strong><br />

nanometre-scale spatial resolution.<br />

Schematically our approach is illustrated in Fig. 1. Similarly to<br />

ref. 27, it is based on the use of attosecond XUV pulses that are<br />

synchronized with the waveform-controlled optical fields 5 , which<br />

are used here to drive the nanolocalized excitations of a<br />

plasmonic nanostructure. A metal (plasmonic) nanosystem is<br />

driven by an optical ultrashort laser pulse with frequency within<br />

the range of the plasmonic resonances (from near UV to near<br />

infrared, NIR). The wavelength of the optical radiation is orders<br />

of magnitude greater than the characteristic size of the<br />

nanosystem, which we assume to be on the order of tens of<br />

nature photonics | VOL 1 | SEPTEMBER 2007 | www.nature.com/naturephotonics 539<br />

© 2007 Nature Publishing Group<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 195


ARTICLES<br />

XUV pulse<br />

Optical pulse<br />

1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

Local plasmonic field<br />

PEEM<br />

Photoelectrons<br />

Time<br />

Figure 1 Schematic of the system <strong>and</strong> photoprocesses. The nanosystem is<br />

shown in a plane denoted in light blue. Instantaneous local fields, which are<br />

excited by the optical pulse, are shown as a three-dimensional plot. The local<br />

optical field at a point of the maximum field (‘hottest spot’) is shown as a<br />

function of time by a red waveform, enhanced with respect to the excitation<br />

field. The application of an XUV pulse is shown by a violet waveform that is<br />

temporally delayed with respect to the excitation field. The XUV excitation<br />

causes the emission of photoelectrons shown by the blue arrows, which are<br />

accelerated by the local plasmonic potential. They are detected with spatial <strong>and</strong><br />

energy resolution by a PEEM.<br />

nanometres. The plasmonic nanosystem responds to the almost<br />

uniform field of the optical excitation with ‘hot spots’, where the<br />

local fields are enhanced with respect to the excitation field by a<br />

factor that depends on the quality of the localized surface<br />

plasmon resonances <strong>and</strong> can be as high as a few hundred for<br />

silver nanosystems. These hot spots are indicated as the peaks of<br />

the local field amplitude in Fig. 1. For the highest of these<br />

peaks (‘hottest spot’), we show (as the red waveform) the<br />

temporal dynamics of the local electric field. Similar to direct<br />

measurements of the optical field’s instantaneous magnitude 27 ,<br />

an attosecond XUV pulse is incident on the system<br />

synchronously with the optical excitation waveform. It causes the<br />

photoemission of electrons whose energy is high enough to select<br />

them from the background of above-threshold ionization (ATI)<br />

<strong>and</strong> multiphoton emission created by the optical field.<br />

These XUV electrons are accelerated by the local plasmonic<br />

electric fields, which define their energies. Imaging such a system<br />

in a PEEM with energy resolution will allow one to see the hot<br />

spots <strong>and</strong> find the instantaneous magnitude of the plasmonic<br />

field at the site <strong>and</strong> time of the emission. As the emission instant<br />

is defined by the XUV pulse, this would allow one to measure<br />

the complete spatiotemporal dynamics of the local fields. The<br />

corresponding spatial resolution is defined solely by the electron<br />

optics of the PEEM (the de Broglie wavelength of the XUV<br />

540<br />

© 2007 Nature Publishing Group<br />

electrons is much smaller <strong>and</strong> is not a limiting factor) <strong>and</strong> can<br />

realistically be on the order of nanometres. The temporal<br />

resolution is determined by the duration of the attosecond pulse<br />

<strong>and</strong> the time of flight of the photoelectrons through the localfield<br />

region, which can be on the order of (a few) hundred<br />

attoseconds. Very importantly, this proposed approach to the<br />

visualization of the attosecond–nanometre dynamics is noninvasive<br />

with respect to the plasmonic fields because of the low<br />

power of the XUV pulses.<br />

ELECTRON ACCELERATION BY NANOPLASMONIC FIELDS<br />

Different regimes of the electron emission by an XUV pulse will<br />

now be discussed to identify those realistic for our conditions<br />

<strong>and</strong> conducive to our goals. The escape velocity for an electron<br />

can be estimated as ve ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

p<br />

2ðhvXUV 2 W f Þ=m,<br />

where m is the<br />

electron mass, vXUV is the frequency of the XUV pulse, Wf is the<br />

metal workfunction <strong>and</strong> h is the reduced Planck’s constant.<br />

The electron escape time from a region of local field of size b is<br />

te ¼ b/ve. The most important case for our purposes is the one<br />

to be called a regime of instantaneous acceleration, when an<br />

electron leaves the local-field region much faster than this field<br />

oscillates in time, that is,<br />

196 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008<br />

t e � T ð1Þ<br />

where T ¼ 2p/v is a characteristic period of the optical excitation,<br />

<strong>and</strong> v is the excitation-pulse carrier frequency.<br />

Under this condition, an electron is driven by a nearly frozen,<br />

instantaneous local electrostatic potential f(r, t XUV), where t XUV<br />

is the emission time defined with a sufficient precision by the<br />

incidence time of the XUV pulse. The final kinetic energy of an<br />

electron, which is measured by the PEEM, is found from energy<br />

conservation,<br />

E XUVðr; t XUVÞ ¼ hv XUV � W f þ efðr; t XUVÞ ð2Þ<br />

where e is the electron charge <strong>and</strong> r the emission point. In sharp<br />

contrast to attosecond streaking in the gas phase 26 , this energy<br />

does not depend on the initial momentum of the photoelectron<br />

or its subsequent flight trajectory.<br />

Estimates for the currently available XUV pulse parameters<br />

are chosen as follows 28 : pulse duration t p ¼ 170 as <strong>and</strong> photon<br />

energy hv XUV ¼ 91 eV. In such a case, assuming the workfunction<br />

W f ¼ 5 eV, we obtain the escape velocity v e ¼ 5 � 10 8 cm s 21 <strong>and</strong><br />

the escape time for the localization distance b ¼ 1 nm as t e ¼ 180<br />

as. Considering an NIR driving pulse with a characteristic period<br />

T � 3 fs, we see that the instantaneous-regime condition (1) is<br />

well satisfied for an optical field localization length up to several<br />

nanometres. Temporal resolution t r of the attosecond plasmonicfield<br />

microscope is determined by both flight time t e through<br />

the region of nanolocalized optical fields <strong>and</strong> the duration t p<br />

of the XUV pulse itself, t r � t e þ t p. For an NIR driving field <strong>and</strong><br />

b . 3 nm, this resolution is sufficient, t r � T. For visible excitation<br />

<strong>and</strong> plasmonic frequencies, sufficient temporal resolution can be<br />

achieved in the future with attosecond pulses delivered with<br />

photon energy hv XUV of several hundred eV to a keV <strong>and</strong> duration<br />

below 100 as.<br />

In the instantaneous regime (1) <strong>and</strong> (2), we can collect a large<br />

number of electrons leaving the surface without sacrificing the<br />

temporal resolution provided by the attosecond streaking, owing<br />

to the fact that the electron energy shift is independent of the<br />

emission angle. This is in sharp contrast to the regime of<br />

the multiple electron oscillations induced by the optical field<br />

nature photonics | VOL 1 | SEPTEMBER 2007 | www.nature.com/naturephotonics


(which, for brevity, will be referred to as the ‘oscillatory regime’).<br />

The oscillatory regime was dominant in the earlier experiments<br />

with the gas phase 27,29 . In this regime, the dwelling (escape) time<br />

of an emitted electron in the region of the streaking optical field<br />

is large enough, t e & T. In that case, the motion is governed not<br />

by the scalar potential but by the vector potential A(t), <strong>and</strong> the<br />

electron velocity v undergoes many oscillations 26 . As a result, the<br />

field-dependent part of the final electron energy E XUV is<br />

dominated by a term DE XUV ¼ (e/mc)v 0A(t XUV) that depends on<br />

the direction of emission velocity v 0. In energy-resolved<br />

microscopy, this would lead to a drastic decrease in the intensity<br />

of the spectral features <strong>and</strong> their smearing-out.<br />

The oscillatory regime of the electron acceleration also took<br />

place in the experiments 30,31 where electrons were emitted from a<br />

plasmonic metal surface in a multiphoton process excited by an<br />

NIR field. After the emission, these electrons are accelerated by<br />

local fields of the surface plasmon polaritons (SPPs). The major<br />

difference between these experiments <strong>and</strong> our proposed<br />

attosecond microscope arises from the fact that the electron<br />

emission energy of refs 30 <strong>and</strong> 31 is much lower, <strong>and</strong> the field<br />

localization radius characteristic of SPPs is much larger than in<br />

the present case, b � 1 mm. Thus, in those experiments, the<br />

Fermi edge of the emitted electron distribution was significantly<br />

smeared out.<br />

In contrast, in our proposed approach, E XUV of equation (2)<br />

does not depend on the direction of emission or a specific<br />

trajectory of the electron motion. This will allow electrons to be<br />

collected in a wide solid angle of emission, which is of utmost<br />

importance in providing a good signal-to-noise ratio, without<br />

sacrificing the energy resolution (<strong>and</strong>, consequently, the accuracy<br />

of the plasmonic potential measurement). It will be shown in the<br />

following that for the moderate optical intensities assumed, the<br />

shift of the electron energy due to the acceleration in the local<br />

field DE XUV ¼ ef is on the order of 10 eV <strong>and</strong> should be easily<br />

measurable. Thus, this instantaneous regime is ideal for the<br />

direct measurement of the attosecond–nanometre dynamics of<br />

the local plasmonic potential <strong>and</strong> is the most desirable. However,<br />

at present it can only be achieved for strongly localized<br />

plasmonic fields whose extension should not exceed a few<br />

nanometres. This restriction applies only until attosecond<br />

pulses become available at much higher photon energies (�1 keV<br />

<strong>and</strong> beyond).<br />

CALCULATIONS AND RESULTS<br />

PLANAR METAL NANOSTRUCTURES IN THE ATTOSECOND PLASMONIC-FIELD MICROSCOPE<br />

Concentrating on this most important case of the instantaneous<br />

acceleration described by equations (1) <strong>and</strong> (2), we start with a<br />

model of a silver 32 r<strong>and</strong>om planar composite (whose geometry<br />

will be described in the next paragraph). We apply an s-polarized<br />

(along the z-direction), waveform-controlled 5 , 5.5-fs optical pulse<br />

as shown in Fig. 2a. We have computed the electric potential<br />

f(r, t) using the quasi-static spectral-expansion Green’s function<br />

method 14,15,33 . The excitation intensity is kept at a moderate level<br />

of I ¼ 10 GW cm 22 to ensure non-damaging conditions of the<br />

excitation. The carrier frequency of the optical (NIR) pulse is<br />

chosen to be 1.55 eV (corresponding to an 800-nm vacuum<br />

wavelength). The temporal kinetics of the local field at a site<br />

where it reaches its global maximum is displayed in Fig. 2b. It<br />

shows the initial period of the driven oscillations for t , 20 fs,<br />

where the response closely reproduces, albeit with a delay, the<br />

5-fs excitation pulse, which reflects the b<strong>and</strong>width of this<br />

plasmonic system. At longer times, after the end of the excitation<br />

pulse, the free-induction evolution shows the interference<br />

1 . 3 . 3 S E L E C T E D R E P R I N T S<br />

Excitation field (a.u.)<br />

Hot spot field (a.u.)<br />

1<br />

0<br />

–1<br />

20<br />

0<br />

–20<br />

100 200<br />

t (fs)<br />

100 200<br />

t (fs)<br />

ARTICLES<br />

Figure 2 Excitation field <strong>and</strong> kinetics of the local field at a hot spot.<br />

a, Excitation field as a function of time. b, Local field at the position of the<br />

maximum (‘hottest spot’) in Fig. 3b as a function of time. The field magnitude is<br />

shown relative to the amplitude of the excitation pulse, which is set to 1.<br />

The vertical arrow denotes the oscillation period within which an XUV pulse is<br />

applied to probe the local field.<br />

beatings of several plasmonic eigenmodes. The maximum<br />

enhancement in this hot spot is Q � 30.<br />

The r<strong>and</strong>om planar composite is generated as a collection of<br />

uncorrelated silver cubes (monomers) positioned on a plane in<br />

vacuum, which is illustrated in Fig. 3a for a monomer size of 4 nm.<br />

As is characteristic of plasmonic nanosystems, there are hot spots<br />

of local fields induced by the optical excitation. The local field<br />

dynamics is shown in Fig. 2b for the ‘hottest spot’, where the local<br />

field reaches its global maximum in space <strong>and</strong> time. We assume<br />

that an XUV attosecond pulse is incident at this system delayed<br />

with respect to the waveform-controlled 5 driving field in such a<br />

way that it probes the system close to the instance of the local field<br />

maximum (indicated by an arrow in Fig. 2b). According to<br />

equation (2), we compute the energy shift of an electron emitted by<br />

such an XUV pulse, which is due to the acceleration in the local<br />

plasmonic fields, as DE XUV(r, t XUV) ¼ ef(r, t XUV). We assume that<br />

these photoelectrons are spatially resolved by a PEEM <strong>and</strong> show in<br />

Fig. 3b–f a series of the electron energy distributions with an<br />

interval of Dt XUV � 200–400 as during a half period of the driving<br />

field. The distributions are shown as a three-dimensional map in<br />

panel 3b <strong>and</strong> as topographic maps in panels c–f. Even for the<br />

moderate excitation intensities used, the energy shift jDE XUVj in<br />

hot spots of the plasmonic potential is rather large (�10 eV) <strong>and</strong>,<br />

consequently, relatively easily measurable. There is a pronounced<br />

nanometre–attosecond kinetics of the electron energy observed in<br />

these distributions, with sharp hot spots indicative of those of the<br />

local fields. These hot spots are concentrated around the edges <strong>and</strong><br />

voids of the metal nanostructure as is a general trend in<br />

nanoplasmonics. (See Supplementary Information for more details,<br />

nature photonics | VOL 1 | SEPTEMBER 2007 | www.nature.com/naturephotonics 541<br />

© 2007 Nature Publishing Group<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 197


ARTICLES<br />

1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

z (nm)<br />

z (nm)<br />

z (nm)<br />

60<br />

40<br />

20<br />

0<br />

0<br />

60<br />

40<br />

20<br />

60<br />

40<br />

20<br />

20 40 60<br />

x (nm)<br />

t XUV = 16.87 fs<br />

20 40 60<br />

x (nm)<br />

t XUV = 17.64 fs<br />

20 40 60<br />

x (nm)<br />

in the movie <strong>Attosecond</strong>_Electron_Energy_tm5.5_130_180_.2mov<br />

of the XUV electron energy distribution for a longer, 30-fs interval<br />

of the evolution.)<br />

Note that after leaving the region of the enhanced local fields,<br />

the electron’s motion is oscillatory <strong>and</strong> defined by the<br />

vector potential of the excitation field in the empty<br />

space. The maximum change of the electron energy due to<br />

this motion is given by a leading term (hv XUV 2 W f)j, where<br />

j � p {8pe 2 I(t XUV)/[mcv 2 (hv XUV 2 W f)]} is a dimensionless<br />

parameter (c is the speed of light in a vacuum); it is assumed<br />

that j � 1. In fact, for I ¼ 10 GW cm 22 , j � 5 � 10 23 ; thus, this<br />

energy change in the free space is small enough <strong>and</strong> can be<br />

safely neglected.<br />

XUV ELECTRON ENERGIES FOR NANOSHELLS<br />

It may also be useful to carry out the first experiments without<br />

spatial resolution by studying the energy distribution of the<br />

XUV-emitted electrons. As an example, we consider in such a<br />

case a much simpler nanoplasmonic system such as metal<br />

nanoshells 34 . Such nanoshells are frequency tunable by adjusting<br />

their aspect ratio A (the ratio of the inner to outer shell radius).<br />

Assuming small shell radius (R . 10 nm), we can use the quasielectrostatic<br />

approximation where the corresponding solutions<br />

are obtained in a simple analytical form.<br />

Consider for instance a silver nanoshell of R ¼ 2.5 nm on a core<br />

with a permittivity of 1 ¼ 10, which is resonant to 800 nm<br />

radiation for A ¼ 0.831. For such a case, the energy shift of the<br />

198 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008<br />

0<br />

12<br />

–12<br />

ΔE XUV (eV)<br />

60<br />

z (nm)<br />

40<br />

z (nm)<br />

z (nm)<br />

20<br />

60<br />

40<br />

20<br />

60<br />

40<br />

20<br />

t XUV = 16.68 fs<br />

20<br />

40<br />

x (nm)<br />

t XUV = 17.25 fs<br />

60<br />

0<br />

5<br />

–5<br />

–10<br />

10<br />

ΔE XUV (eV)<br />

20 40 60<br />

x (nm)<br />

t XUV = 18.03 fs<br />

20 40 60<br />

x (nm)<br />

Figure 3 Topography of a nanosystem <strong>and</strong> spatiotemporal kinetics of the local field potential as detected by the attosecond plasmonic-field microscope.<br />

Intensity of the excitation optical pulse I ¼ 10 GW cm 22 , <strong>and</strong> photon energy hv ¼ 1.55 eV. a, Topography of the nanosystem: r<strong>and</strong>om planar composite consisting of<br />

4 nm� 4 nm � 4 nm silver cubes arranged on a plane with fill factor 0.5. This composite is smoothed to within 2 nm to improve numerical precision.<br />

b–f, Distributions of the energy shift DE XUV of electrons emitted by the XUV pulse in the plane of this nanostructure shown for different moments t XUV (as indicated in<br />

the panels) of the XUV pulse incidence within the half cycle of local field oscillations. b, A three-dimensional map. c–f, Topographic colour maps showing details of<br />

the nanometre–attosecond spatiotemporal kinetics.<br />

542<br />

© 2007 Nature Publishing Group<br />

nature photonics | VOL 1 | SEPTEMBER 2007 | www.nature.com/naturephotonics


electrons is shown in Fig. 4, as a function of both the phase delay w<br />

between the driving NIR field <strong>and</strong> the incident attosecond pulse,<br />

<strong>and</strong> the azimuthal angle u of the electron-emission point (for<br />

I ¼ 10 GW cm 22 <strong>and</strong> hv ¼ 1.55 eV). To avoid any confusion, we<br />

emphasize that the electron energy does not depend on the<br />

direction of the electron velocity but only on the azimuth of a<br />

point at the nanoshell surface from which the emission took place.<br />

In the exact resonance for Fig. 4a, the maximum shift is<br />

jDE XUVj � 10 eV, on the same order as for the r<strong>and</strong>om planar<br />

composite above. The maximum is reached for w ¼ p/2, as<br />

expected for resonant excitation. The nanoshell resonance is very<br />

sharp; therefore, a small change of A to a value of 0.829 detunes<br />

the resonance <strong>and</strong> leads to a significant reduction in the<br />

magnitude of the effect <strong>and</strong> a shift in its phase dependence<br />

(compare with Fig. 4b).<br />

DISCUSSION AND CONCLUSIONS<br />

ΔE XUV (eV)<br />

10<br />

0<br />

–10<br />

0<br />

2<br />

The state-of-the-art of attosecond technology for the attosecond<br />

plasmonic-field microscope proposed in this paper is represented<br />

by t p ¼ 170 as, hv XUV ¼ 93 eV XUV pulses containing &�10 6<br />

photons <strong>and</strong> delivered at a 3-kHz repetition rate 28,35 . As has been<br />

demonstrated earlier, such XUV pulses can be focused to a spot<br />

diameter of 2 mm (ref. 36).<br />

The signal-to-noise ratio of the proposed attosecond<br />

nanoplasmonic-field microscope is limited by the electron<br />

currents that can be obtained from nanoscopic surface areas for<br />

the presently available intensities of the XUV sources 28,35 . In<br />

Materials <strong>and</strong> Methods the XUV photoelectron current emitted<br />

per 1-nm 2 area of the nanosystem surface is estimated. The<br />

number of photoelectrons per unit area of the nanosystem<br />

surface <strong>and</strong> unit time of observation (photoelectron fluence) is<br />

j � 15 nm 22 s 21 for gold <strong>and</strong> j � 10 nm 22 s 21 for silver. For the<br />

spatial resolution of PEEM available at present, r � 10 nm, this<br />

yields the electron flux from the area of resolution as<br />

N 0 ¼ pr 2 j � 4,500 s 21 for gold <strong>and</strong> N 0 � 3,000 s 21 for silver.<br />

Such electron counts should not pose any problem for detection<br />

with PEEM.<br />

Future steps towards the experimental implementation of the<br />

attosecond nanoplasmonic field microscope will use ultrashort<br />

pulsed XUV radiation with photon energies exceeding<br />

hv XUV ¼ 100 eV. At present, single-attosecond XUV pulses are<br />

obtained by high-harmonic generation driven by a few-cycle<br />

φ<br />

4<br />

1 . 3 . 3 S E L E C T E D R E P R I N T S<br />

A = 0.831<br />

6<br />

0<br />

1<br />

2<br />

θ<br />

3<br />

ΔE XUV (eV)<br />

5<br />

0<br />

–5<br />

0<br />

2<br />

waveform-stabilized Ti:sapphire laser pulse (800 nm, 5 fs) in a Ne<br />

gas jet target. Selecting a broadb<strong>and</strong> (�9 eV full-width at-halfmaximum<br />

(FWHM), centre frequency 91 eV) spectrum from the<br />

harmonic plateau range close to its cut-off by means of a Mo/Si<br />

multilayer-coated XUV mirror, single-attosecond XUV pulses have<br />

been extracted <strong>and</strong> pulse duration down to 250 as FWHM has been<br />

experimentally verified 7 . Improvements (broader-b<strong>and</strong> mirrors)<br />

have yielded pulses shorter than 200 as. This offers a temporal<br />

resolution of at least 100 as. The excitation optical pulse <strong>and</strong> the<br />

probing XUV pulse are refocused to the sample by means of an<br />

interferometric double-mirror configuration allowing for controlled,<br />

variable time delays with 10-as steps. The spatially <strong>and</strong> energyresolved<br />

detection of photoelectrons is achieved using a PEEM that<br />

is equipped with a flight drift tube followed by a two-dimensional<br />

delay-line detector. The spatial resolution of this time-resolved<br />

PEEM is limited by the electron optics to 20 nm. Silver <strong>and</strong> gold<br />

samples have been grown (see Supplementary Information for<br />

details, Experimental_Steps.pdf).<br />

To conclude, we have proposed <strong>and</strong> developed a theory for an<br />

attosecond nanoplasmonic-field microscope. The main advantage<br />

of such a microscope is that it is non-invasive with respect to the<br />

nanoplasmonic fields. The principle of this microscope is based<br />

on the photoemission of electrons by an XUV attosecond pulse<br />

that is synchronized with a waveform-stabilized driving optical<br />

field. Information about the nanoplasmonic fields is imprinted in<br />

the energy of the XUV-emitted electrons owing to their<br />

acceleration in the instantaneous electrostatic potential of the<br />

surface plasmon oscillations excited by the optical field.<br />

Our proposed microscope will open up unique possibilities to<br />

directly study <strong>and</strong> control ultrafast photoprocesses in surface<br />

plasmonic nanosystems <strong>and</strong> circuits. It images the local<br />

nanoplasmonic field in real space with nanometre-scale spatial<br />

resolution <strong>and</strong> in real time with �100 as temporal resolution.<br />

This approach will be especially important in ultrafast<br />

nanoplasmonic systems where very tight localization of optical<br />

fields occurs, for example, when a shaped pulse of radiation<br />

induces nanolocalized fields at a desired nanosite. The<br />

microscope could also be used to study various plasmonenhanced<br />

photoprocesses such as femtosecond photochemistry,<br />

light detection, <strong>and</strong> solar energy conversion where the processes<br />

of the energy transfer can be ultrafast. For example, with<br />

nanoplasmonic antennas, which couple together molecular <strong>and</strong><br />

semiconductor systems with external fields, the ultrafast kinetics<br />

Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008 199<br />

φ<br />

4<br />

A = 0.829<br />

6<br />

0<br />

1<br />

2<br />

θ<br />

ARTICLES<br />

Figure 4 Energy shift of electrons emitted from the surface of silver nanoshells as a function of the azimuthal angle u of the emission point <strong>and</strong> the phase<br />

w of the delay between the driving optical radiation <strong>and</strong> the probing attosecond XUV pulse. The intensity of the excitation optical pulse is I ¼ 10 GW cm 22 ,<br />

the photon energy is hv ¼ 1.55 eV <strong>and</strong> the outer radius of the nanoshells is R ¼ 2.5 nm. a,b, Data for the nanoshells with aspect ratio A ¼ 0.831 (a) <strong>and</strong> for<br />

A ¼ 0.829 (b).<br />

nature photonics | VOL 1 | SEPTEMBER 2007 | www.nature.com/naturephotonics 543<br />

© 2007 Nature Publishing Group<br />

3


ARTICLES<br />

1 . 3 AT T O S E C O N D A N D H I G H - F I E L D P H Y S I C S D I V I S I O N<br />

of the energy exchange can be studied through the corresponding<br />

kinetics of the nanoplasmonic fields.<br />

One of the most important potential uses of the attosecond<br />

nanoplasmonic-field microscope will be in the design <strong>and</strong> study<br />

of elements <strong>and</strong> devices for ultrafast <strong>and</strong> ultradense (at the<br />

nanoscale) optical <strong>and</strong> optoelectronic information processing <strong>and</strong><br />

storage. To bring practical advantages over existing electronic <strong>and</strong><br />

optoelectronic technology, such nanoscale devices must<br />

necessarily be ultrafast (with a subpicosecond or femtosecond<br />

response time). Examples of such devices include optical<br />

nanotransistors, memory cells, nanoplasmonic media for the<br />

mass storage of information, <strong>and</strong> nanoplasmonic interconnects<br />

where both the nanolocalization of energy <strong>and</strong> its ultrafast<br />

temporal kinetics are important.<br />

MATERIALS AND METHODS<br />

Here we estimate the XUV electron photocurrent that can be obtained per 1-nm 2<br />

area of the nanosystem metal surface. We adopt the values of the XUV photon flux<br />

for state-of-the-art XUV pulses 28,35 : J XUV � 10 9 s 21 (the flux averaged over time).<br />

The XUV pulses can be focused 36 to a spot with a radius R � 1 mm. This yields an<br />

XUV photon fluence I X � 8 � 10 16 cm 22 s 21 . For hv XUV ¼ 91 eV photon energy,<br />

the photoionization cross-section per atom for silver is s a ¼ 7.1 Mb (ref. 37) <strong>and</strong><br />

for gold is s a ¼ 10.9 Mb (refs 37, 38).<br />

Consider emission from an area a ¼ 1 nm 2 . For electron energies �100 eV,<br />

the elastic escape depth (from which electrons come without losing their energy<br />

to collisions) is h ¼ 0.5 nm (ref. 39). The total photoemission cross-section from<br />

atoms within this elastic depth <strong>and</strong> area a is s tot ¼ rahs a, where r is the number<br />

density of atoms. From this we obtain the expected values of XUV photoelectron<br />

fluence j � 15 nm 22 s 21 for gold <strong>and</strong> j � 10 nm 22 s 21 for silver.<br />

Received 30 March 2007; accepted 23 July 2007; published 3 September 2007.<br />

References<br />

1. Paul, P. M. et al. Observation of a train of attosecond pulses from high harmonic generation. Science<br />

292, 1689–1692 (2001).<br />

2. Hentschel, M. et al. <strong>Attosecond</strong> metrology. Nature 414, 509–513 (2001).<br />

3. Niikura, H. et al. Sub-laser-cycle electron pulses for probing molecular dynamics. Nature 417,<br />

917–922 (2002).<br />

4. Drescher, M. et al. Time-resolved atomic inner-shell spectroscopy. Nature 419, 803–807 (2002).<br />

5. Baltuska, A. et al. <strong>Attosecond</strong> control of electronic processes by intense light fields. Nature 421,<br />

611–615 (2003).<br />

6. Niikura, H. et al. Probing molecular dynamics with attosecond resolution using correlated wave<br />

packet pairs. Nature 421, 826–829 (2003).<br />

7. Kienberger, R. et al. Atomic transient recorder. Nature 427, 817–821 (2004).<br />

8. Sekikawa, T., Kosuge, A., Kanai, T. & Watanabe, S. Nonlinear optics in the extreme ultraviolet. Nature<br />

432, 605–608 (2004).<br />

9. Lopez-Martens, R. et al. Amplitude <strong>and</strong> phase control of attosecond light pulses. Phys. Rev. Lett. 94,<br />

033001 (2005).<br />

10. Baker, S. et al. Probing proton dynamics in molecules on an attosecond time scale. Science 312,<br />

424–427 (2006).<br />

11. Dudovich, N. et al. Measuring <strong>and</strong> controlling the birth of attosecond XUV pulses. Nature Phys. 2,<br />

781–786 (2006).<br />

12. Sansone, G. et al. Isolated single-cycle attosecond pulses. Science 314, 443–446 (2006).<br />

13. Kling, M. F. et al. <strong>Control</strong> of electron localization in molecular dissociation. Science 312,<br />

246–248 (2006).<br />

544<br />

© 2007 Nature Publishing Group<br />

14. Stockman, M. I., Faleev, S. V. & Bergman, D. J. Localization versus delocalization of surface plasmons<br />

in nanosystems: Can one state have both characteristics? Phys. Rev. Lett. 87, 167401 (2001).<br />

15. Stockman, M. I., Faleev, S. V. & Bergman, D. J. Coherent control of femtosecond energy localization<br />

in nanosystems. Phys. Rev. Lett. 88, 067402 (2002).<br />

16. Bergman, D. J. & Stockman, M. I. Surface plasmon amplification by stimulated emission of<br />

radiation: Quantum generation of coherent surface plasmons in nanosystems. Phys. Rev. Lett. 90,<br />

027402 (2003).<br />

17. Lehmann, J. et al. Surface plasmon dynamics in silver nanoparticles studied by femtosecond timeresolved<br />

photoemission. Phys. Rev. Lett. 85, 2921–2924 (2000).<br />

18. Zentgraf, T., Christ, A., Kuhl, J. & Giessen, H. Tailoring the ultrafast dephasing of quasiparticles in<br />

metallic photonic crystals. Phys. Rev. Lett. 93, 243901 (2004).<br />

19. Kubo, A. et al. Femtosecond imaging of surface plasmon dynamics in a nanostructured silver film.<br />

Nano Lett. 5, 1123–1127 (2005).<br />

20. Stockman, M. I. & Hewageegana, P. Nanolocalized nonlinear electron photoemission under coherent<br />

control. Nano Lett. 5, 2325–2329 (2005).<br />

21. Brixner, T. et al. Quantum control by ultrafast polarization shaping. Phys. Rev. Lett. 92,<br />

208301 (2004).<br />

22. Brixner, T., d. Abajo, F. J. G., Schneider, J. & Pfeiffer, W. Nanoscopic ultrafast space–time-resolved<br />

spectroscopy. Phys. Rev. Lett. 95, 093901 (2005).<br />

23. Sukharev, M. & Seideman, T. Phase <strong>and</strong> polarization control as a route to plasmonic nanodevices.<br />

Nano Lett. 6, 715–719 (2006).<br />

24. Aeschlimann, M. et al. Adaptive subwavelength control of nano-optical fields. Nature 446,<br />

301–304 (2007).<br />

25. Pelton, M., Liu, M. Z., Park, S., Scherer, N. F. & Guyot-Sionnest, P. Ultrafast resonant optical<br />

scattering from single gold nanorods: Large nonlinearities <strong>and</strong> plasmon saturation. Phys. Rev. B 73,<br />

155419 (2006).<br />

26. Corkum, P. B. & Krausz, F. <strong>Attosecond</strong> science. Nature Phys. 3, 381–387 (2007).<br />

27. Goulielmakis, E. et al. Direct measurement of light waves. Science 305, 1267–1269 (2004).<br />

28. Schultze, M. et al. Powerful 170-attosecond XUV pulses generated with few-cycle laser pulses <strong>and</strong><br />

broadb<strong>and</strong> multilayer optics. New J. Phys. 9, 243 (2007).<br />

29. Drescher, M. et al. X-ray pulses approaching the attosecond frontier. Science 291, 1923–1927 (2001).<br />

30. Kupersztych, J., Monchicourt, P. & Raynaud, M. Ponderomotive acceleration of photoelectrons in<br />

surface-plasmon-assisted multiphoton photoelectric emission. Phys. Rev. Lett. 86,<br />

5180–5183 (2001).<br />

31. Kupersztych, J. & Raynaud, M. Anomalous multiphoton photoelectric effect in ultrashort time<br />

scales. Phys. Rev. Lett. 95, 147401 (2005).<br />

32. Johnson, P. B. & Christy, R. W. Optical constants of noble metals. Phys. Rev. B 6, 4370–4379 (1972).<br />

33. Stockman, M. I., Bergman, D. J. & Kobayashi, T. Coherent control of nanoscale localization of<br />

ultrafast optical excitation in nanosystems. Phys. Rev. B 69, 054202–10 (2004).<br />

34. Nehl, C. L. et al. Scattering spectra of single gold nanoshells. Nano Lett. 4, 2355–2359 (2004).<br />

35. Goulielmakis, E. et al. <strong>Attosecond</strong> control <strong>and</strong> measurement: <strong>Lightwave</strong> electronics. Science<br />

(in the press).<br />

36. Schnurer, M. et al. Guiding <strong>and</strong> high-harmonic generation of sub-10-fs pulses in hollow-core fibers<br />

at 10(15) w/cm 2 . Appl. Phys. B 67, 263–266 (1998).<br />

37. Henke, B. L., Lee, P., Tanaka, T. J., Shimabukuro, R. L. & Fujikawa, B. K. Low-energy X-ray<br />

interaction coefficients: Photoabsorption, scattering, <strong>and</strong> reflection. Atomic Data <strong>and</strong> Nuclear Data<br />

Tables 27, 1–131 (1982).<br />

38. Manson, S. T. in Photoemission in Solids Vol. 1, (eds Cardona, M. & Ley, L.) 135–163 (Springer,<br />

Berlin, New York, 1978).<br />

39. Tanuma, S., Powell, C. J. & Penn, D. R. Calculations of electron inelastic mean free paths. ii. Data for<br />

27 elements over the 50–2000 eV range. Surf. Interface Anal. 17, 911–926 (1991).<br />

Acknowledgements<br />

The work of M.I.S. is supported by grants from the Chemical Sciences, Biosciences <strong>and</strong> Geosciences<br />

Division of the Office of Basic Energy Sciences, Office of Science, US Department of Energy, a grant CHE-<br />

0507147 from NSF, <strong>and</strong> a grant from the US-Israel BSF. M.I.S.’s work at the Max-Planck-Institute for<br />

Quantum Optics (Garching, Germany) was supported by a Research Stipend of the Max Planck Society.<br />

The work of M.F.K., U.K., <strong>and</strong> F.K. was partially supported by the German Science Foundation (DFG)<br />

through the Cluster of Excellence Munich Center for Advanced Photonics. M.F.K. acknowledges support<br />

by an EU reintegration grant <strong>and</strong> the DFG Emmy–Noether program. MIS acknowledges helpful<br />

discussions with S. Manson regarding photoelectron cross-sections <strong>and</strong> with P. Corkum regarding<br />

charging of the surfaces.<br />

Correspondence <strong>and</strong> requests for materials should be addressed to M.I.S. or F.K.<br />

Supplementary information accompanies this paper on www.nature.com/naturephotonics.<br />

Competing financial interests<br />

The authors declare no competing financial interests.<br />

Reprints <strong>and</strong> permission information is available online at http://npg.nature.com/reprints<strong>and</strong>permissions/<br />

nature photonics | VOL 1 | SEPTEMBER 2007 | www.nature.com/naturephotonics<br />

200 Max-Planck-Institut für Quantenoptik • Progress Report 2007/2008

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!