Sansone 2011

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

review articleS

Published online: 18 September 2011 | doi: 10.1038/nphoton.2011.167

High-energy attosecond light sources


Giuseppe Sansone1, Luca Poletto2 and Mauro Nisoli1*

The development of attosecond technology is one of the most significant achievements in the field of ultrafast optics over the
past decade. Since the first experimental demonstration of attosecond pulses just ten years ago, novel techniques have been
introduced for the generation, characterization and application of subfemtosecond pulses. The development of attosecond
tools is continuously triggering the introduction of new spectroscopic and measurement methods, which will offer the oppor-
tunity to investigate unexplored research areas with unprecedented time resolution. The wealth of ultrafast processes, which
can be investigated by taking advantage of attosecond temporal resolution, can be greatly extended by the development of
high-intensity attosecond sources. This Review covers a selection of recent advances in the field of attosecond technology, with
particular attention being given to the generation and application of high-energy attosecond pulses.

O
ver the past few decades, the ability to investigate ultra- plasmas, with particular emphasis on isolated attosecond pulses.
fast phenomena at shorter and shorter temporal scales has Perspective is given on the status and prospects of the applications
determined crucial progress in our understanding of the of high-energy attosecond pulses in research fields such as nonlin-
fundamental processes in matter. The 1980s saw femtosecond tech- ear attosecond metrology, electron correlation phenomena and the
nology revolutionize the field of time-resolved spectroscopy1, while measurement of ultrafast dynamics in atoms and molecules.
the past decade has brought remarkable developments in the field
of attosecond science2–5. Extreme-ultraviolet (XUV) coherent light High-energy attosecond pulses from gas harmonics
pulses with durations of a few tens of attoseconds are ideal tools Various schemes have been proposed and experimentally inves-
for investigating electronic dynamics on the atomic/molecular tigated for enhancing the conversion efficiency of the harmonic
scale. Such ultrashort pulses can be produced using high-order har- generation process, based on the use of high-energy, femtosecond
monic generation (HHG)6,7, which involves focusing high-intensity driving pulses focused in hollow-core fibres8,9 or cells10–13. A very
femtosecond pulses onto a gas medium3. Attosecond light bursts, powerful method for generating microjoule-energy XUV pulses
either in the form of isolated pulses or trains of pulses separated by is to use the loose-focusing geometry, in which a driving beam
half an optical cycle of the driving radiation, have been generated is focused in a long gas cell using long-focal-length optics14–16.
with conversion efficiencies in the range 10–4–10–8. Although many Favourable phase-matching conditions can be achieved in this
important applications of attosecond pulses have already been dem- configuration by using the positive dispersion of the neutral gas
onstrated, the widespread adoption of attosecond sources has so far medium at low pressure to compensate for the Gouy geomet-
been limited. This is due not only to the complexity of the laser tools ric shift, which is equivalent to providing a negative dispersion.
required to reliably generate attosecond pulses, but also to the low Driving intensities below the saturation intensity of the medium
photon flux characteristics of attosecond sources, particularly for are generally required to reduce the negative dispersion induced
producing isolated attosecond pulses. Although femtosecond meas- by the free-electron population. In a typical experimental set-up,
urement techniques are now standard methods in ultrafast optics, high-energy laser pulses are focused into a gas cell whose length,
extending such principles to the attosecond temporal domain is L, is a few centimetres. Long focal lengths (1–5  m) are used to
not straightforward. Novel experimental techniques to initiate and increase the excitation spot size and, consequently, the Rayleigh
probe electron dynamics on attosecond timescales are required. range. Driving pulse energy, beam spot size and gas pressure can be
Attosecond pump–probe spectroscopy has so far been demon- adjusted to maximize the harmonic conversion yield. XUV radia-
strated only partially and in a very small number of laboratories, tion reaching energies of 11.5 μJ at wavelengths of between 73.6 nm
owing to the high intensities required to produce observable two- and 42.6 nm (corresponding to harmonics between the 11th and
XUV-photon processes. The development of experimental methods 19th orders, respectively) has been generated in a 14-cm-long cell
for generating high-energy attosecond pulses could pave the way for filled with xenon at a pressure of 0.6 torr, by focusing 35 fs, 14 mJ
crucial progress in the field of attosecond science. driving pulses onto the gas cell using a lens with a focal length
This Review reports on recent advances in the field of attosecond of 5 m (ref. 17). The loose-focusing geometry can also be used to
technology for the generation of high-energy attosecond pulses. generate a low-divergence XUV beam, whose beam divergence is
Although the generation of microjoule-energy trains of attosecond proportional to L–1/2 (ref.  16). The temporal characteristics of an
pulses has been reported by various research groups, with impor- attosecond pulse train generated using this technique have been
tant applications to attosecond metrology based on nonlinear XUV measured by second-order interferometric autocorrelation16,18,19.
effects, the production, characterization and application of high- Optimization of the harmonic conversion efficiency can be also
energy isolated attosecond pulses is still under active investigation. achieved by employing quasi-phase-matching techniques, which
This Review analyses various techniques for the generation of high- prevent back-conversion of the harmonic generation process by
intensity attosecond pulses, both from gases and laser-produced applying a periodical correction to the phase mismatch after every

1
Department of Physics, Politecnico di Milano, National Research Council of Italy, Institute of Photonics and Nanotechnologies, Piazza L. da Vinci 32,
20133 Milano, Italy. 2National Research Council of Italy, Institute of Photonics and Nanotechnologies, Via Trasea 7, 35131 Padova, Italy.
*email: mauro.nisoli@fisi.polimi.it

nature photonics | VOL 5 | NOVEMBER 2011 | www.nature.com/naturephotonics 655


© 2011 Macmillan Publishers Limited. All rights reserved.
review articleS Nature photonics doi: 10.1038/nphoton.2011.167

a d fundamental radiation28 (Fig. 2). Using such a two-colour scheme


BS1
(known as double optical gating; DOG)29, the width of the lin-
ear polarization gate can be close to one optical cycle of the fun-
λ/4 BK7 damental radiation, as shown in Fig. 2. This has two advantages:
Quartz plate BS2 longer driving pulses with durations of up to ~12 fs can be used,
b and the ground-state population depletion on the leading edge of
λ/2 the driving pulses can be reduced30. Mashiko et al. reported that
using the DOG technique provides high XUV photon fluxes31
BBO BS3 (Fig. 1b). The researchers measured continuous XUV spectra with
e
a maximum XUV pulse energy from argon of ~6.5  nJ (immedi-
c Filament
Vacuum ately after the generating medium) at a conversion efficiency of
XUV
6  × 10–6 (ref. 31). The on-target pulse energy was estimated to be
Brewster of the order of 1 nJ, after taking into account the transmission of
window Noble gas the aluminium filter used to block the fundamental radiation and
reflections from focusing optics.
Figure 1 | Isolated attosecond pulses from gases. Various approaches for The extension of temporal gating techniques to high-energy driv-
the generation of isolated attosecond pulses have been experimentally ing pulses with durations of a few tens of femtoseconds and peak
demonstrated. a, Polarization gating. b, Double optical gating. powers up to the petawatt level is currently a very active research
c, Generalized double optical gating. d, Interferometric polarization gating. field. One of the main limiting factors is related to the depletion of
e, XUV generation from a laser filamentation in noble gases. BBO, β-barium the neutral-atom population on the leading edge of the driving field.
borate; BS, beamsplitter. The generalized DOG technique (Fig.  1c) is able to generate iso-
lated sub-150-as pulses from a sub-30-fs amplifier32. The idea of this
coherence length. A recent review by Popmintchev et al. compre- technique is to reduce the depletion of the neutral-atom population
hensively covers quasi-phase-matching methods in the harmonic by decreasing the ellipticity of the driving field, which is achieved
generation process20. A different quasi-phase-matching approach by adding a fused silica Brewster window to the DOG set-up33. In
— one that demonstrates the potential to enhance the photon flux 2007, Tzallas et al. proposed and implemented a polarization gating
in the soft-X-ray spectral region — is based on the use of a multi- scheme known as interferometric polarization gating (Fig. 1d)34,35,
gas-cell target21. This technique could be particularly useful for the in which an ellipticity-modulated pulse is produced from four indi-
loose-focusing geometry because of the long confocal parameter of vidually controlled, linearly polarized pulses of variable field ampli-
the excitation beam, thus allowing the separation between consecu- tude, relative polarization and delay. This method has been used
tive cells to be optimized. to generate a coherent continuum of radiation by loosely focusing
High-energy isolated attosecond pulses can be produced from 55 fs driving pulses with a peak power of 2 TW into a pulsed gas
gas harmonics in a variety of ways. This section discusses three jet. Particularly interesting is the high photon flux of the radia-
classes of generation techniques that are based on the use of suit- tion, whose spectral bandwidth supports isolated pulses with dura-
able temporal gating schemes. The technique of spectrally filtering tions down to ~260  as. In subsequent experiments, the energy of
the cut-off harmonics (known as amplitude gating) has been already the continuum was found to be around 20 nJ at photon energies of
analysed in various reviews3,4; pulses as short as 80 as have been pro- 30–70 eV, with an energy of ~10 nJ in the cut-off region36.
duced in neon (at a central photon energy of 80 eV) with a maxi-
mum on-target energy of 0.57 pJ (ref. 22). The main limiting factor
of this technique is the requirement of very short and energetic driv-
a
ing pulses, with durations below 5 fs, and a controlled electric field. ω T0 /2

Polarization gating. Modulating the polarization state of the driv-


ing field is a powerful method for confining the XUV generation
process to a single event23. This technique is based on the strong
dependence of the harmonic-generation process on the ellipticity
of the driving pulses, and therefore exploits an excitation field that
is elliptically polarized on the pulse’s leading and trailing edges but
is almost linearly polarized in a short temporal window around
the pulse peak24. A good estimation for the duration of driving b
T0
pulse required to generate efficiently isolated attosecond pulses is ω + 2ω
τ < 2.5 T0 (where T0 is the optical cycle of the driving field; ref. 25),
which is around 6.6 fs in the case of 800 nm excitation. The overall
conversion efficiency from the laser source to XUV light output
decreases as the laser pulse duration increases. In 2006, research-
ers generated isolated attosecond pulses using carrier–envelope
phase (CEP)-stabilized 5  fs, 0.2  mJ driving pulses (Fig.  1a)26,27.
Pulses as short as 130 as have been produced in argon, in the spec-
tral range of 25–50 eV, with energies of around 70 pJ. Table 1 lists
the relevant parameters of current attosecond sources. Adding a Figure 2 | Temporal gating. a,b, A multicycle driving field (blue curves)
second-harmonic field to the fundamental frequency can alleviate generates a train of attosecond pulses, separated by either half an optical
the requirements on driving pulse duration. For particular values cycle, T0/2, in the case of single-colour excitation (a), or a full optical
of the relative phase between the two laser field components, and cycle, T0, in the case of a two-colour excitation (b). Isolated attosecond
when the amplitude of the second-harmonic field is high enough, pulses can be generated using a temporal gate, which confines the XUV
the symmetry of the driving electric field breaks and the genera- generation process to a short window whose temporal extensions depends
tion of attosecond pulses takes place for every optical cycle of the on the separation between consecutive attosecond pulses.

656 nature photonics | VOL 5 | NOVEMBER 2011 | www.nature.com/naturephotonics


© 2011 Macmillan Publishers Limited. All rights reserved.
Nature photonics doi: 10.1038/nphoton.2011.167 review articleS
a b
Ionization rate Ionization rate

Path 1 Ψ=0 Path 1 Ψ = π/2

Time (cycles) Time (cycles)

Path 2 Path 2

c d

Total emission rate


1 Total emission rate
1
Emission rate (a.u.)

0.8

Emission rate (a.u.)


0.8
0.6
0.6
0.4
0.4
0.2
0.2
0
20 0
20
Pho 25
ton 30 Pho 25
ene Path 2 ton 30
rgy 35 ene Path 2
(eV 40 rgy 35
) Path 1 (eV 40
) Path 1

Figure 3 | Ionization gating: electron quantum path analysis. a,b, Ionization rate (green curves) calculated in xenon for the case of 5 fs driving pulses
(red curves), with a peak intensity of 2.5 × 1015 W cm–2, for two CEP values. The first two electron quantum paths are shown by dotted blue lines. c,d, XUV
emission rates associated with the first three electron quantum paths (black curves) and total XUV emission rate produced by coherent superposition of
all electron quantum paths (red curves) for two CEP values, calculated by using non-adiabatic saddle-point simulations.

Two-colour gating. Recent advances in the realization of high- without CEP control, when the two-colour field synthesis is prop-
energy optical parametric amplifiers in the near-infrared spectral erly optimized in terms of central wavelength of the near-infrared
region have allowed the investigation of HHG using long-wave- component and intensity ratio between the two colours. Moreover,
length femtosecond pulses. The maximum photon energy obtain- the two-colour scheme offers the major advantage of being able to
able through HHG in gases scales as Iλ2, where I and λ are the peak significantly reduce the ionization of the gas medium, thus allow-
intensity and central wavelength of the driving pulse, respectively. ing the use of efficient phase-matching techniques for the genera-
The harmonic cut-off can therefore be significantly extended by tion of intense attosecond pulses from a neutral gas12. By using the
increasing λ, at the expense of a reduced single-atom harmonic loose-focusing geometry, it is reasonable to expect the generation
yield, which scales as λ–6 (refs 37–39). Parametric sources, which are of microjoule-energy isolated attosecond pulses with conversion
tunable in the near-infrared, have been developed to extend XUV efficiencies of the order of 10–5. Such a scheme could therefore lead
emission up to kilo-electronvolt energies40,41. In 2007, Merdji and to new ways of generating isolated attosecond pulses for use in ter-
co-workers proposed a method for generating isolated attosecond awatt-class laser systems.
pulses by combining the fundamental field ω with its detuned sec-
ond harmonic at 2ω + δω (ref. 42). Ionization gating. An alternative approach for the generation of
In 2009, researchers reported an efficient XUV generation tech- high-energy isolated attosecond pulses is based on the use of driving
nique that combines an intense near-infrared pulse (λ = 1.45 μm, pulses with above-saturation intensity. Complete depletion of the
I  =  2  ×  1014  W  cm–2) with an intense detuned second-harmonic neutral-atom population on the leading edge of the driving pulses
pulse (λ = 0.8 μm, I = 8.5 × 1014 W cm–2) comprising only a few opti- confines the harmonic-generation process. To generate isolated
cal cycles43. This method was used to generate coherent continuous attosecond pulses, the atoms of the generating gas must be exposed
emission at energies of up to 160 eV in argon and 200 eV in neon. to a large non-adiabatic increase in the electric field of the leading
Importantly, the researchers observed that the harmonic conversion edge; such a condition can be easily achieved using intense, few-
efficiency for two-colour excitation was much larger than of a sin- optical-cycle pulses with controlled electric field. This process has
gle-colour near-infrared field. This suggests that the intense 0.8 μm been discussed in theoretical works by Cao et al.46 and Kim et al.47,
component of the driving field was responsible for the increase in which both describe the generation of isolated attosecond pulses by
conversion efficiency, thereby overcoming the major drawback of spectral selection of the high-frequency components in the XUV
using longer driving wavelengths. To relax the requirements on spectrum. From an experimental point of view, the influence of ion-
driving pulse duration, Takahashi and co-workers recently dem- ization dynamics on the generation of isolated attosecond pulses has
onstrated the generation of continuous XUV radiation around the been investigated for 8 fs (ref. 48) and 15 fs (ref. 49) driving pulses.
cut-off spectral region by mixing 30  fs, 0.8  μm pulses with 40  fs, In both cases, the enhancement of the confinement effect of the cut-
1.3 μm pulses44,45. Continuous XUV spectra can be generated, even off harmonics of the XUV spectrum to a single pulse was explained

nature photonics | VOL 5 | NOVEMBER 2011 | www.nature.com/naturephotonics 657


© 2011 Macmillan Publishers Limited. All rights reserved.
review articleS Nature photonics doi: 10.1038/nphoton.2011.167
a b c pulses is explained by considering the spatiotemporal dynamics of
the driving field during the filamentation process. In particular, as
pointed out in ref. 55, filamentation gives rise to the formation of
3TL
Time

high-intensity driving spikes with single-optical-cycle duration,


which are responsible for the generation of isolated attosecond
λ/5 λ/5 λ/5 pulses.
Normal space coordinates
High-energy attosecond pulses from laser-produced plasmas
The loose-focusing approach becomes experimentally unfeasible for
gaseous media if the energy of the excitation pulses is increased to a
few hundred millijoules. Harmonic generation from solid targets has
been suggested as an alternative technique for delivering high-energy
attosecond pulses up to the X-ray energy range. In this scheme, the
electric field of a high-intensity infrared pulse (I > 1016 W cm–2) ion-
izes the surface of a solid target to create an overdense plasma at the
boundary between the vacuum and the solid. The generation of XUV
radiation from such a laser-produced plasma has been analysed in
a recent review by Thaury and Quéré58. The interaction of the laser
with the plasma is governed by the normalized vector potential
a0 = eE/ωLmc, where E and ωL are the electric field and frequency of
the driving pulse, respectively, c is the velocity of light in vacuum,
and e and m are the electron charge and mass, respectively. Efficient
Figure 4 | Isolated attosecond pulses by harmonic generation from a
harmonic generation can be achieved for values of a0 close to or larger
solid surface. a–c, Motion of the plasma surface induced by excitation
than unity. Two generating mechanisms that can lead to HHG from
pulses with linear (a), circular (b) and time-dependent polarization (the
solid targets have so far been investigated: coherent wake emission59,
corresponding electric field, characterized by a short temporal window
which dominates when a0 ≤ 1, and the relativistic oscillating mirror
of linear polarization, is also shown) (c). Attosecond pulses, generated in
mechanism60, which dominates when a0 >> 1. Both cases cause the
the case of linear and time-dependent polarization, are shown in yellow.
generation of odd and even harmonics of the fundamental radiation.
By properly filtering the reflected radiation, a single attosecond pulse can
In contrast, HHG in gases using one-colour, multiple-optical-cycle
be isolated. λ and TL are driving laser wavelength and period, respectively.
excitation leads to only odd-order harmonics. The coherent wake
Figure reproduced with permission from ref. 67, © 2008 IOP.
emission process has analogies with harmonic generation in gases.
At each maximum of the electric field, electrons in the plasma cre-
by the very rapid loss of phase-matching occurring on the leading ated by the excitation pulse are ejected into the vacuum and then
edge of the excitation pulses. driven back towards the plasma surface. This gives rise to ultrafast
More recently, researchers generated sub-160-as isolated pulses by electron bunches that recollide with the plasma, where they trigger
exposing the sub-cycle ionization dynamics of a gas medium to high- high-frequency collective electron oscillations in their wake and thus
intensity, 5 fs pulses with controlled electric field50. The experiment cause the emission of attosecond pulses. The maximum frequency
achieved an on-target energy of 2.1 nJ for xenon and 1.4 nJ for argon, of the generated XUV spectrum is given by the maximum plasma
with an overall conversion efficiency of around 5 × 10–6 (ref. 50). The frequency of the laser-induced plasma, which typically corresponds
physical processes of the ionization gating scheme can be under- to harmonics between the 15th and 30th order of an 800 nm driv-
stood through a single-atom approach based on the non-adiabatic ing field. In 2009, Nomura  et  al. experimentally demonstrated that
saddle-point model51,52. In such a model, the Fourier transform of the XUV radiation produced by coherent wake emission consists of a
the atomic dipole moment responsible for harmonic generation can train of attosecond pulses by measuring the corresponding autocor-
be written as a coherent superposition of the contributions from the relation trace61. They estimated the XUV intensity in the interaction
different electron quantum paths. Figure 3c,d shows emission rates region to be approximately (0.5–1) × 1011 W cm–2.
associated with the electron trajectories displayed in Fig. 3a,b for two When a0 >> 1, the influence of the magnetic field on the electron
CEP values. Depending on the temporal evolution of the ionization motion cannot be neglected. Owing to the magnetic component of
rate (shown in Fig.  3a,b), either one or two quantum trajectories the Lorentz force ev × B (where v is the electron velocity and B is the
effectively contribute to the total emission rate, giving rise to either magnetic field of the excitation pulse) and also to the restoring force
one or two attosecond pulses, respectively. By using a non-adiabatic provided by the ions, the electrons move in an oscillatory motion
three-dimensional numerical propagation model53, we investigated around their equilibrium position with a velocity perpendicular to
the possibility of using the loose-focusing geometry to increase the the surface given by v⊥  ∝ a02(t)[1  +  sin(2ωLt)]. The incident laser
photon flux of attosecond pulses. In particular, we calculated how radiation is reflected by the oscillating electron layer, which acts as
attosecond pulses in a xenon gas cell evolve when the energy of the a plasma mirror; because of the relativistic Doppler effect, the spec-
5 fs driving pulses is increased while maintaining a constant excita- trum of the reflected radiation can extend up to the X-ray region,
tion intensity of 2.5  × 1015 W cm–2. Simulations show that the XUV depending on the value of a0 (ref. 62). Baeva et al. have shown that a
pulse energy could be increased up to the microjoule level while still more correct description of the harmonic-generation process from
preserving the excellent spatial characteristics of the XUV beam. solid surfaces considers the velocity of the electrons to be almost
Another promising technique for generating high-photon-flux always close to c (refs  63,64). In this ultrarelativistic regime, the
isolated attosecond pulses is based on the production of XUV radia- gamma factor of the plasma surface can be approximated as:
tion from laser filamentation in noble gases (Fig. 1e). This process
has been investigated extensively by Gaarde and co-workers54,55. p⊥2
γs = 1 +
From an experimental point of view, continuous XUV spectra sup- p||2
porting isolated attosecond pulses have been recently obtained from
35  fs infrared pulses undergoing filamentation in a semi-infinite where p|| and p⊥ are the components of the electron momentum (p)
argon gas cell56,57. In this case, the production of isolated attosecond parallel and perpendicular to the surface, respectively. The factor γS

658 nature photonics | VOL 5 | NOVEMBER 2011 | www.nature.com/naturephotonics


© 2011 Macmillan Publishers Limited. All rights reserved.
Nature photonics doi: 10.1038/nphoton.2011.167 review articleS
a b
35 3 c
10
i
8 2pσu 2.5
30 νf

Potential energy (eV)


6
Kinetic energy (eV)

Intensity (a.u.)
4 25 d
iii
ii 1.5 ii
iii 3νf
2 20 i ×100
ii 5th 3rd 1st D + D+ 1
1 1sσg e 5νf
15 iii
i 5 11th D+D 0.5
6νf
0 ×5
0 0
−2 −1 0 1 2 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 0 2 4 6 8 10
Optical cycle Internuclear distance (Å) Harmonic order

Figure 5 | Attosecond nonlinear Fourier transform spectroscopy. a, Kinetic energy distribution of D+ ions as a function of the delay between a pair
of harmonic fields. The horizontal axis was scaled by the optical cycle of the fundamental radiation. b, Potential energy curves of D2. The one-photon
absorption of the 11th harmonic generates D2+ in the 1sσg state; the fifth, third and first harmonics excite D2+ ions to the 2pσu state, resulting in the
dissociation of D2+ into D+ and D. c, Fourier transform intensities of the kinetic energy temporal evolution of D+ ions as a function of the harmonic order for
the energy intervals (i), (ii) and (iii). Figure reproduced with permission from ref. 76, © 2010 APS.

is always of the order of unity, except for cases when the momentum XUV generation process is produced by the deformation of the
of the plasma mirror is perpendicular to the surface (p||(t–) = 0), plasma profile, which results in large changes in the local incidence
which leads to sharp features in the time evolution of γS. These angle (the angle between the original wavefront and the deformed
spikes are responsible for the rich harmonic content in the reflected reflecting surface) on a sub-cycle temporal scale. Moreover, in the
radiation. Using the ‘γS spikes’ model, it is possible to demonstrate relativistic λ3 regime, the intense and coherent relativistic electron
that the harmonic scaling law is given by I(ω) ∝ ω–8/3 up to the max- motion in the direction parallel to the wave vector of the reflected
imum frequency ωmax = (√8)γmax3ωL, which was experimentally con- wave can give rise to dramatic compression of the reflected pulses,
firmed by Dromey et al.65. The spectral brightness of the harmonic which allows reflection, deflection and compression to act together
radiation was in the range of 1022–1024 photons  s–1  mm–2  mrad–2 to generate isolated attosecond pulses. Particle-in-cell simulations
(0.1% bandwidth) for 4 nm and 17 nm, respectively, with estimated suggest that this technique could be used to produce isolated
energies of 0.07–0.7 mJ and 0.2–20 mJ, respectively. Harmonic gen- pulses with durations as low as 200 as and conversion efficiencies
eration of up to 3.8  keV from a CH target was reported in 2007, of up to 10%.
with harmonic energies of ~17 μJ at 1.4 keV and ~5 μJ at 3.1 keV in Recent progress in laser technology — in particular the devel-
1% bandwidth66. opment of laser systems based on optical parametric chirped-pulse
In analogy with harmonic generation in gases, surface plasma amplification69 — has opened the way to the generation of multi-
harmonics driven by multicycle linearly polarized pulses are emit- terawatt few-cycle light pulses70, which have potentially crucial
ted in trains of attosecond pulses. Isolated pulses can be attained applications in the field of attosecond science. On the basis of recent
by employing the polarization gating technique. In this case, the results, the development of laser systems with powers reaching sev-
conversion efficiency diminishes when the ellipticity of the driv- eral tens of terawatts and repetition rates in the kilohertz regime
ing pulse is increased, which can be understood by considering are within our reach. Such laser systems will be ideal drivers for the
the motion of the oscillating mirror67. For linear polarization, this efficient generation of attosecond pulses in the kilo-electronvolt
motion corresponds to a periodic oscillation every half-cycle of spectral range from laser-produced plasmas.
the fundamental radiation, thus providing a train of attosecond
pulses (Fig.  4a). For a circularly polarized incident pulse, the Applications of high-energy attosecond pulses
plasma mirror does not present strong oscillations but instead The generation of high-energy attosecond pulses poses a number
varies on the timescale of the laser envelope, according to the rela- of constraints in the optical design of XUV facilities. One problem
tion v⊥ ∝ a02(t). The envelope of the pulse determines a ‘smooth’ is that the high-energy driving beam must be removed after XUV
force on the electron plasma (Fig. 4b) and thus no harmonics are generation to avoid damage to the optics and the sample under
generated. The generation of isolated pulses (Fig.  4c) is possible investigation. The second important issue is using a design of XUV
using few-cycle pulses with a time-dependent polarization simi- focusing optics that achieves a high on-target beam intensity. In this
lar to the one described above. One-dimensional particle-in-cell section we will concentrate on a few applications of high-intensity
simulations have shown that the dependence of the harmonic attosecond pulses, including both theoretical proposals and partial
generation efficiency on the ellipticity is in turn dependent on experimental implementations.
the parameter a0; the higher the normalized vector potential, the
more sensitive the harmonics are to small deviations in ellipticity Attosecond XUV nonlinear processes. The production of high-
from the linear case. intensity isolated attosecond pulses is crucial for the investigation
It is worth mentioning that the efficient generation of attosec- of nonlinear processes in the XUV spectral region, particularly for
ond pulses has been predicted when high-energy, few-optical-cycle the development of attosecond metrology by XUV nonlinear opti-
laser pulses are focused onto a solid target at a spot size compara- cal methods. Watanabe and co-workers used two-photon above-
ble to the laser wavelength (λ). This achieves intensities of more threshold ionization to demonstrate the first autocorrelation and
than 1018 W cm–2 and contains the laser energy within a focal vol- frequency-resolved optical gating measurement of isolated ~0.9 fs
ume of ~λ3 (the λ3 regime). Particle-in-cell simulations show that, pulses by using a scandium/silicon multilayer mirror to select
in the λ3 regime, the radiation reflected by the overdense plasma is the ninth harmonic (28  eV) of the incident light71,72. In 1998, the
composed of a train of attosecond pulses, with each pulse deflected researchers used two-photon nonlinear absorption in helium to
to a specific angle68. As pointed out by Naumova et al., this efficient measure the first autocorrelation trace of the ninth harmonic of a

nature photonics | VOL 5 | NOVEMBER 2011 | www.nature.com/naturephotonics 659


© 2011 Macmillan Publishers Limited. All rights reserved.
review articleS Nature photonics doi: 10.1038/nphoton.2011.167
Such an interferometric technique forms the basis of the recently
r1
t = 0 fs
developed attosecond nonlinear Fourier transform spectroscopy
technique — a very powerful method for characterizing coupled
electronic–nuclear dynamics in molecules. The first experiments to
use this technique were performed in CO2 (ref. 75) and D2 (ref. 76).
In general, the interferometric trace of the attosecond pulse train
as a function of the delay between two pulse replicas, I(τ), can be
written as:
ω = 23.7 eV
r2

r1
I(τ) = ∑I
p, q
(τ)
p, q

t = 1.5 fs
where the single terms Ip,q can be expressed as:

∫ 2
Ip,q(τ) = σp,q [E2p+1(t − τ) + E2p+1(t)][E2q+1(t − τ) + E2q+1(t)] dt


r2
= σp,q [Fb + FIC(τ) + Fp(τ) + Fq(τ) + Fp−q(τ) + Fp+q(τ)] dt (1)
r1
t = 5.6 fs
The term σp,q represents the cross-section for the absorption of
one photon of the 2q + 1 harmonic and one photon of the 2p + 1
harmonic. The term Fb represents a constant background given by
the two-photon signal associated with the two pulses separately, and
does not depend on the relative delay τ. The term FIC(τ) gives the
intensity autocorrelation of the electric field; the terms Fp(τ), Fq(τ),
ω = 90 eV r2 Fp−q(τ) and Fp+q(τ) are functions oscillating at 2p + 1, 2q + 1, 2(p − q)
and 2(p + q + 1) times the fundamental frequency νf, respectively.
Figure 6 | Attosecond pump–probe experiment for investigating electron Fourier analysis of the measured signal allows the different Fourier
motion inside an atom. The first pump XUV pulse (central energy of components to be isolated and provides physical insight into the
23.7 eV) excites the ground state to the superposition 1s + 1snp. The radial nonlinear photoabsorption process (that is, the order of the har-
distributions in the (r1, r2) plane are shown before (t = 0 fs) and after monic fields involved in the photoabsorption process).
(t = 1.5 fs) the pump pulse. The wave packet evolves in time on a timescale Nonlinear Fourier transform spectroscopy has been used to
of a few femtoseconds, thus determining a ‘breathing mode’ of the radial retrieve information about the photodissociation of D2 molecules
density of the two electrons. This motion can be imaged by doubly ionizing by irradiating D2 molecules with an attosecond pulse train con-
the system using an intense attosecond pulse centred at 90 eV. After the taining harmonics from the first to the 19th order76. The D+ signal,
probe pulse, the radial density distribution shows that the electrons are at shown in Fig. 5a as a function of τ, exhibits periodic modulations
large radial distances of r1 and r2, which corresponds to double ionization. with different periods in three kinetic energy regions, (i), (ii) and
Figure reproduced with permission from ref. 79, © 2006 APS. (iii), shown in Fig. 5a,b. Analysis of the oscillation frequencies indi-
cates that the generation of D+ ions proceeds through the absorp-
tion of one photon (mainly from the 11th-order harmonic field),
Ti:sapphire laser (ref.  73). The same nonlinear process was used which ionizes the neutral molecules and projects the nuclear wave
in 2003 to achieve the first autocorrelation trace of a train of atto- packet onto the ground ionic state 1sσg, followed by the absorption
second pulses produced by phase-locked harmonics from the sev- of a second photon from the fifth, third and fundamental harmonic
enth to the 15th order74. In this case, the estimated XUV intensity fields, which excite the ion to the dissociative state 2pσu (Fig. 5b).
was 1011 W cm–2. The ion yield measured as a function of the delay For higher harmonic orders, the excitation takes place at shorter
between two XUV pulse replicas showed a clear attosecond struc- internuclear distances and the ion acquires a larger kinetic energy
ture characteristic of a train of attosecond pulses with an average on the dissociating curve, as shown in Fig. 5b. The contributions of
duration of 780 as. the different harmonic orders to the energy ranges (i), (ii) and (iii)
The Coulomb explosion of diatomic molecules has been used to can be easily recognized in the Fourier transform of the temporal
measure the interferometric autocorrelation trace of a train of atto- evolution of the ion kinetic energy (Fig. 5c–e). Indeed, the ion sig-
second pulses16,18,19. In particular, two-photon absorption in N2 was nals associated with the first (i), third (ii) and fifth (iii) harmonics
achieved by using an XUV field composed of harmonics ranging present a clear oscillation at frequencies of νf, 3νf and 5νf, respec-
from the ninth to the 19th order. Diatomic molecules were chosen tively, corresponding to the term Fp (or Fq) of equation (1). The dif-
instead of single atoms of similar ionization potential because dia- ference term Fp−q, which corresponds to components at 10νf, 8νf
tomic molecules exhibit a larger cross-section for XUV two-photon and 6νf, is also weakly visible in the Fourier transform of the signals
absorption. The loose-focusing geometry was used to generate the in Fig. 5c–e. The sum term Fp+q cannot be observed owing to the
attosecond pulse train, and the estimated intensity of the 11th-order experiment’s insufficient delay resolution. Nonlinear Fourier trans-
harmonic field at the focal point was 3  ×  1014  W  cm–2. The pulse form spectroscopy is a particularly powerful technique because
train was divided into two XUV pulse replica at a beamsplitter and only the nonlinear interactions, which are generally rather weak,
the N+ ion yield was subsequently integrated as a function of the are characterized by interference fringes that can be evidenced by
delay between the two. The resulting yield exhibited bunches with Fourier analysis of the experimental traces.
an average duration of 320  as and a periodicity twice that of the
driving field, corresponding to the interferometric autocorrelation XUV-pump/XUV-probe measurements. Recently, Skantzakis
trace of the train of attosecond pulses. and co-workers demonstrated the first XUV-pump/XUV-probe

660 nature photonics | VOL 5 | NOVEMBER 2011 | www.nature.com/naturephotonics


© 2011 Macmillan Publishers Limited. All rights reserved.
Nature photonics doi: 10.1038/nphoton.2011.167 review articleS
Table 1 | Relevant parameters of current attosecond sources based on HHG in gases.
Method Gas ħω (eV) τ (as) E Type Efficiency Reference
AG Neon 80 80 0.5 nJ IAP 1.7 × 10–6 22
PG Argon 35 130 70 pJ IAP 3.5 × 10–7 26,27
PG Neon 60 — 1 pJ IAP 5 × 10–9 26
DOG Argon 38 — 6.5 nJ IAP 6 × 10–6 31
DOG Neon 45 107 170 pJ IAP 1.5 × 10–7 31,85
IG Xenon 27 155 9 nJ IAP 2.6 × 10–5 50
IPG Xenon 50 <1,000 20 nJ IAP 2 × 10–5 36
LF Xenon 20 320 10 μJ APT 7 × 10–4 16
ħω is the central energy of the XUV spectrum, τ is the measured pulse duration and E is the pulse energy at the source (in the case of isolated pulses, E corresponds to the XUV energy in the spectral region
characterized by a continuous profile). The efficiency is defined as the ratio between the XUV pulse energy and the driving laser pulse energy. AG, amplitude gating; PG, polarization gating; DOG, double-optical
gating; IG, ionization gating; IPG, interferometric polarization gating; LF, loose-focusing; IAP, isolated attosecond pulses; and APT, attosecond pulse trains.

measurements of atomic coherence dynamics with a temporal probe pulse is required to achieve a substantial double-ionization
resolution of 1 fs (ref. 77). The XUV radiation, produced using the probability. The single-photon double-ionization process is based
interferometric polarization gating technique and characterized on the correlation between the two electrons and is therefore sen-
by a broadband continuum in the range of 15–25 eV, was focused sitive to the relative distance between them. When they are in
by a split spherical mirror into a xenon pulsed gas jet, which gener- close proximity, the electron correlation is relevant and facilitates
ated two pulse replicas. The XUV pump pulse used single-photon their ejection, thus maximizing the probability of double ioniza-
absorption to excite an electronic wave packet that was generated tion. When the electron distance is large, a minimum in double
by the coherent manifold of doubly excited states and inner-shell ionization occurs as a consequence of the reduced correlation. In
excited auto-ionizing states of xenon. The atomic dipole induced such a case, the probe photon is mostly absorbed by the core elec-
by the time-delayed XUV probe presented a phase-shift that was tron, which is promptly ejected by the atom. The second electron
determined by the temporal delay τ between the pump and probe is subsequently ejected by a shake-off process80. The total yield of
pulses. One would therefore expect to observe Ramsey fringes doubly ionized helium therefore oscillates according to the oscil-
resulting from quantum interference in the auto-ionization signal, lations of the inter-electron distance and to the correlated angu-
which gives information about the energy separation and lifetimes lar distribution of the two electrons as a function of the relative
of nearby states78. Indeed, the Fourier transform of the measured delay between the pump and probe pulses. In this respect, isolated
auto-ionizing signal as a function of τ reproduced the spectrum attosecond pulses are very efficient for imaging the evolution of
of the series of excited states. It is worth pointing out that in this the electron wave packet because they allow the correlation (and
experiment, the ionization was not a two-XUV-photon process77. therefore the distance) between the two electrons to be ‘frozen’ at
The use of attosecond-pump attosecond-probe schemes for the point of ionization.
investigating the dynamics of electron wave packets has been sug- Chu and Lin proposed the use of an XUV-pump/XUV-probe
gested in several theoretical studies. Hu and Collins79 analysed approach for investigating the time evolution of electronic wave
the possibility of exciting a coherent wave packet consisting of packets of auto-ionizing states in atoms81. Auto-ionization of Fano
a superposition of 1snp states (where n = 2, 3, 4...) in helium by resonances82 — a fundamental process that involves electron cor-
using an XUV pulse with a peak intensity of 6  × 1014 W cm–2, a relation — can be studied in a benchmark system such as helium,
duration of 1.5 fs and a central photon energy of 23.7 eV. After the using XUV radiation at 60 eV as the pump and an ultrashort syn-
excitation pulse, the radial distribution of the electron wave packet chronized infrared field as the probe83,84. However, as pointed out by
temporally oscillates in a ‘breathing’ motion. The periodic oscilla- Chu and Lin, the photoelectron spectra resulting from the interac-
tion of the radial distribution can be imaged by doubly ionizing tion of the auto-ionizing levels with the infrared field are often too
the atom using an intense 90 eV attosecond pulse. The evolution complex to deliver an intuitive picture of the outgoing wave packet.
of the electron wave packet during application of the pump–probe More direct characterization of the wave packet can be achieved by
can be described by the radial distribution P(r1, r2) of the two elec- using an ultrashort XUV pulse as the probe.
trons in the (r1, r2) plane, given by:
Conclusions
∫∫Φ(r , r ) dΩ dΩ
2
P(r1, r2) = Continuous progress in the development of laser sources for the
1 2 1 2
generation of femtosecond pulses and the exploitation of novel
techniques for the efficient generation of XUV pulses promises
where Φ(r1,  r2) is the two-electron wavefunction and the inte- to offer very important tools for the generation of attosecond
gration is performed over the solid angles Ω1 and Ω2. Before the pulses with energies ranging from a few tens of nanojoules to a
arrival of the pump pulse at t = 0 fs, the two electrons are in the few millijoules. In particular, in the case of joule-level driving
ground state 1s2 (Fig. 6a). Just after the pump pulse at t = 1.5 fs, one sources, the relativistic interaction of intense laser pulses with
electron is close to the nucleus while the other has a spread in the overdense plasma could offer a route for the generation of attosec-
radial distribution (Fig.  6b), which is indicative of a single-elec- ond pulses with unprecedented intensities, which might be use-
tron excitation process. At a later time (t = 5.1 fs), a probe pulse ful for the investigation of yet unexplored research topics such as
with a duration of 250 as, a peak intensity of 3  × 1015 W cm–2 and non-perturbative and relativistic inner-shell processes or exotic
a central photon energy of 90 eV is used to image the wave packet high-field physics in the XUV spectral region. Access to XUV
motion by doubly ionizing the system. The electron distribution at nonlinear optics using high-intensity attosecond pulses paves the
t = 5.6 fs indicates a substantial probability of finding both elec- way to the observation of novel classes of ultrafast phenomena,
trons at a large distance apart, which corresponds to a double- which will trigger further progress in the field of ultrafast optics
ionization process (Fig. 6c). The high intensity of the attosecond and spectroscopy.

nature photonics | VOL 5 | NOVEMBER 2011 | www.nature.com/naturephotonics 661


© 2011 Macmillan Publishers Limited. All rights reserved.
review articleS Nature photonics doi: 10.1038/nphoton.2011.167

References 35. Charalambidis, D. et al. Exploring intense attosecond pulses. New J. Phys. 10,
1. Zewail, A. H. Femtochemistry: Atomic-Scale Dynamics of the Chemical Bond. 025018 (2008).
J. Phys. Chem. A 104, 5660–5694 (2000). 36. Skantzakis, E., Tzallas, P., Kruse, J., Kalpouzos, C. & Charalambidis, D.
2. Agostini, P. & DiMauro, L. F. The physics of attosecond light pulses. Rep. Prog. Coherent continuum extreme ultraviolet radiation in the sub100nJ range
Phys. 67, 813–855 (2004). generated by a high-power many-cycle laser field. Opt. Lett. 34,
3. Corkum, P. B. & Krausz, F. Attosecond science. Nature Phys. 3, 381–387 (2007). 1732–1734 (2009).
4. Krausz, F. & Ivanov, M. Attosecond physics. Rev. Mod. Phys. 81, 37. Tate, J. et al. Scaling of wave-packet dynamics in an intense midinfrared field.
163–234 (2009). Phys. Rev. Lett. 98, 013901 (2007).
5. Nisoli, M. & Sansone, G. New frontiers in attosecond science. Prog. Quant. 38. Colosimo, P. et al. Scaling strong-field interactions towards the classical limit.
Electron. 33, 17–59 (2009). Nature Phys. 4, 386–389 (2008).
6. Li, X. F., L’Huillier, A., Ferray, M., Lompré, L. A. & Mainfray, G. Multiple- 39. Shiner, A. D. et al. Wavelength scaling of high harmonic generation efficiency.
harmonic generation in rare gases at high laser intensity. Phys. Rev. A 39, Phys. Rev. Lett. 103, 073902 (2009).
5751–5761 (1989). 40. Takahashi, E., Kanai, T., Nabekawa, Y. & Midorikawa, K. 10 mJ class
7. McPherson, A. et al. Studies of multiphoton production of vacuum-ultraviolet femtosecond optical parametric amplifier for generating soft X-ray harmonics.
radiation in the rare gases. J. Opt. Soc. Am. B 4, 595–601 (1987). Appl. Phys. Lett. 93, 041111 (2008).
8. Rundquist, A. et al. Phase-matched generation of coherent soft X-rays. Science 41. Popmintchev, T. et al. Extended phase matching of high harmonics driven by
280, 1412–1415 (1998). mid-infrared light. Opt. Lett. 33, 2128–2130 (2008).
9. Bartels, R. et al. Shaped-pulse optimization of coherent soft-X-rays. Nature 406, 42. Merdji, H. et al. Isolated attosecond pulses using a detuned second-harmonic
164–166 (2000). field. Opt. Lett. 32, 3134–3136 (2007).
10. Schnürer, R. et al. Absorption-limited generation of coherent ultrashort soft-X- 43. Calegari, F. et al. Efficient continuum generation exceeding 200 eV by intense
Ray pulses. Phys. Rev. Lett. 83, 722–725 (1999). ultrashort two-color driver. Opt. Lett. 34, 3125–3127 (2009).
11. Hergott, J.F. et al. Extreme-ultraviolet high-order harmonic pulses in the 44. Takahashi, E. J., Lan, P., Mücke, O. D., Nabekawa, Y. & Midorikawa, K. Infrared
microjoule range. Phys. Rev. A 66, 021801(R) (2002). two-color multicycle laser field synthesis for generating an intense attosecond
12. Takahashi, E., Nabekawa, Y., Otsuka, T., Obara, M. & Midorikawa, K. pulse. Phys. Rev. Lett. 104, 233901 (2010).
Generation of highly coherent submicrojoule soft X-rays by high-order 45. Lan, P., Takahashi, E. J. & Midorikawa, K. Optimization of infrared two-color
harmonics. Phys. Rev. A 66, 021802 (2002). multicycle field synthesis for intenseisolatedattosecond-pulse generation. Phys.
13. Kazamias, S. et al. Global optimization of high harmonic generation. Phys. Rev. Rev. A 82, 053413 (2010).
Lett. 90, 193901 (2003). 46. Cao, W., Lu, P., Lan, P., Wang, X. & Yang, G. Single-attosecond pulse generation
14. Constant, E. et al. Optimizing high harmonic generation in absorbing gases: with an intense multicycle driving pulses. Phys. Rev. A 74, 063821 (2006).
model and experiment. Phys. Rev. Lett. 82, 1668–1671 (1999). 47. Kim, K. T., Kim, C. M., Baik, M. G., Umesh, G. & Nam, C. H. Single
15. Ditmire, T., Crane, J. K., Nguyen, H., DaSilva, L. B. & Perry, M. D. Energy-yield sub50attosecond pulse generation from chirp-compensated harmonic radiation
and conversion efficiency measurements of high-order harmonic radiation. using material dispersion. Phys. Rev. A 69, 051805(R) (2004).
Phys. Rev. A 51, R902–R905 (1995). 48. Abel, M. J. et al. Isolated attosecond pulses from ionization gating of high-
16. Midorikawa, K., Nabekawa, Y. & Suda, A. XUV multiphoton processes with harmonic emission. Chem. Phys. 366, 9–14 (2009).
intense high-order harmonics. Prog. Quant. Electron. 32, 43–88 (2008). 49. Thomann, I. et al. Characterizing isolated attosecond pulses from hollow-core
17. Takahashi, E. J., Hasegawa, H. & Midorikawa, K. Generation of 10-mJ coherent waveguides using multi-cycle driving pulses. Opt. Express 17,
extreme-ultraviolet light by use of high-order harmonics. Opt. Lett. 27, 4611–4633 (2009).
1920–1922 (2002). 50. Ferrari, F. et al. High-energy isolated attosecond pulses generated by above-
18. Nabekawa, Y. et al. Interferometric autocorrelation of an attosecond pulse train saturation few-cycle fields. Nature Photon. 4, 875–879 (2010).
in the single-cycle regime. Phys. Rev. Lett. 97, 153904 (2006). 51. Lewenstein, M., Balcou, Ph., Ivanov, M. Y., L’Huillier, A. & Corkum, P. B.
19. Shimizu, T. et al. Observation and analysis of an interferometric Theory of high-harmonic generation by low-frequency laser pulses. Phys. Rev.
autocorrelation trace of an attosecond pulse train. Phys. Rev. A 75, A 49, 2117–2132 (1994).
033817 (2007). 52. Sansone, G., Vozzi, C., Stagira, S. & Nisoli, M. Nonadiabatic quantum path
20. Popmintchev, T., Chen, M.C., Arpin, P., Murnane, M. M. & Kapteyn, H. C. analysis of high-order harmonic generation: role of the carrier-envelope phase
The attosecond nonlinear optics of bright coherent X-ray generation. Nature on short and long paths. Phys. Rev. A. 70, 013411 (2004).
Photon. 4, 822–832 (2010). 53. Priori, E. et al. Nonadiabatic three-dimensional model of high-order harmonic
21. Seres, J. et al. Coherent superposition of laser-driven soft-X-ray harmonics generation in the fewopticalcycle regime. Phys. Rev. A 61, 63801 (2000).
from successive sources. Nature Phys. 3, 878–883 (2007). 54. Chakraborty, H. S., Gaarde, M. B. & Couairon, A. Single attosecond pulses
22. Goulielmakis, E. et al. Single-cycle nonlinear optics. Science 320, from high harmonics driven by self-compressed filaments. Opt. Lett. 31,
1614–1617 (2008). 3662–3664 (2006).
23. Corkum, P. B., Burnett, N. H. & Ivanov, M. Y. Subfemtosecond pulses. Opt. Lett. 55. Gaarde, M. B. & Couairon, A. Intensity spikes in laser filamentation:
19, 1870–1872 (1994). diagnostics and application. Phys. Rev. Lett. 103, 043901 (2009).
24. Tcherbakoff, O., Mével, E., Descamps, D., Plumridge, J. & Constant, E. Time- 56. Steingrube, D. S. et al. Generation of high-order harmonics with ultra-short
gated high-order harmonic generation. Phys. Rev. A 68, 043804 (2003). pulses from filamentation. Opt. Express 17, 16177–16182 (2009).
25. Strelkov, V. et al. Single attosecond pulse production with an ellipticity- 57. Steingrube, D. S. et al. High-order harmonic generation directly from a
modulated driving IR pulse. J. Phys. B 38, L161–L167 (2005). filament. New J. Phys. 13, 043022 (2011).
26. Sola, I. J. et al. Controlling attosecond electron dynamics by phase-stabilized 58. Thaury, C. & Quéré, F. High-order harmonic and attosecond pulse generation
polarization gating. Nature Phys. 2, 319–322 (2006). on plasma mirrors: basic mechanisms. J. Phys. B 43, 213001 (2010).
27. Sansone, G. et al. Isolated single-cycle attosecond pulses. Science 314, 59. Quéré, F. et al. Coherent wake emission of high-order harmonics from
443–446 (2006). overdense plasmas. Phys. Rev. Lett. 96, 125004 (2006).
28. Mauritsson, J. et al. Attosecond pulse trains generated using two color laser 60. Lichters, R., MeyerterVehn, J. & Pukhov, A. Short-pulse laser harmonics from
fields. Phys. Rev. Lett. 97, 013001 (2006). oscillating plasma surfaces driven at relativistic intensity. Phys. Plasmas 3,
29. Chang, Z. Controlling attosecond pulse generation with a double optical gating. 3425–3437 (1996).
Phys. Rev. A 76, 051403(R) (2007). 61. Nomura, Y. et al. Attosecond phase locking of harmonics emitted from laser-
30. Mashiko, H. et al. Double optical gating of high-order harmonic generation produced plasmas. Nature Phys. 5, 124–128 (2009).
with carrier-envelope phase stabilized lasers. Phys. Rev. Lett. 100, 62. Tsakiris, G. D., Eidmann, K., MeyerterVehn, J. & Krausz, F. Route to intense
103906 (2008). single attosecond pulses. New J. Phys. 8, 19 (2006).
31. Mashiko, H., Gilbertson, S., Li, C., Moon, E. & Zenghu, C. Optimizing the 63. Baeva, T., Gordienko, S. & Pukhov, A. Theory of high-order harmonic
photon flux of double optical gated high-order harmonic spectra. Phys. Rev. A generation in relativistic laser interaction with overdense plasma. Phys. Rev. E
77, 063423 (2008). 74, 046404 (2006).
32. Feng, X. et al. Generation of isolated attosecond pulses with 20 to 28 64. Baeva, T., Gordienko, S. & Pukhov, A. Relativistic plasma control for single
femtosecond lasers. Phys. Rev. Lett. 103, 183901 (2009). attosecond X-ray burst generation. Phys. Rev. E 74, 065401(R) (2006).
33. Gilbertson, S. et al. Isolated attosecond pulse generation using multicycle pulses 65. Dromey, B. et al. High harmonic generation in the relativistic limit. Nature
directly from a laser amplifier. Phys. Rev. A 81, 043810 (2010). Phys. 2, 456–459 (2006).
34. Tzallas, P. et al. Generation of intense continuum extreme-ultraviolet radiation 66. Dromey, B. et al. Bright multi-keV harmonic generation from relativistically
by many-cycle laser fields. Nature Phys. 3, 846–850 (2007). oscillating plasma surfaces. Phys. Rev. Lett. 99, 085001 (2007).

662 nature photonics | VOL 5 | NOVEMBER 2011 | www.nature.com/naturephotonics


© 2011 Macmillan Publishers Limited. All rights reserved.
Nature photonics doi: 10.1038/nphoton.2011.167 review articleS
67. Rykovanov, S. G., Geissler, M., MeyerterVehn, J. & Tsakiris, G. D. Intense 78. Cavalieri, S., Eramo, R., Materazzi, M., Corsi, C. & Bellini, M. Ramsey-type
single attosecond pulses from surface harmonics using the polarization gating spectroscopy with high-order harmonics. Phys. Rev. Lett. 89, 133002 (2002).
technique. New J. Phys. 10, 025025 (2008). 79. Hu, S. X. & Collins, L. A. Attosecond pump probe: exploring ultrafast electron
68. Naumova, N. M., Nees, J. A., Sokolov, I. V., Hou, B. & Mourou, G. A. motion inside an atom. Phys. Rev. Lett. 96, 073004 (2006).
Relativistic generation of isolated attosecond pulses in a λ3 focal volume. Phys. 80. Wehlitz, R. et al. Electron-energy and angular distributions in the double
Rev. Lett. 92, 063902 (2004). photoionization of helium. Phys. Rev. Lett. 67, 3764–3767 (1991).
69. Dubietis, A., Jonušauskas, F. & Piskarskas, A. Powerful femtosecond pulse 81. Chu, W.C. & Lin, C. D. Theory of ultrafast autoionization dynamics of Fano
generation by chirped and stretched pulse parametric amplification in BBO resonances. Phys. Rev. A 82, 053415 (2010).
crystal. Opt. Commun. 88, 437–440 (1992). 82. Fano, U. Effects of configuration interaction on intensities and phase shifts.
70. Herrmann, D. et al. Generation of subthreecycle, 16 TW light pulses by using Phys. Rev. 124, 1866–1878 (1961).
noncollinear optical parametric chirped-pulse amplification. Opt. Lett. 34, 83. Wickenhauser, M., Burgdörfer, J., Krausz, F. & Drescher, M. Time resolved
2459–2461 (2009). Fano resonances. Phys. Rev. Lett. 94, 023002 (2005).
71. Sekikawa, T., Kosuge, A., Kanai, T. & Watanabe, S. Nonlinear optics in the 84. Argenti, L. & Lindroth, E. Ionization branching ratio control with a resonance
extreme ultraviolet. Nature 432, 605–608 (2004). attosecond clock. Phys. Rev. Lett. 105, 053002 (2010).
72. Kosuge, A. et al. Frequency-resolved optical gating of isolated attosecond pulses 85. Mashiko, H. et al. Extreme ultraviolet supercontinuua supporting pulse
in the extreme ultraviolet. Phys. Rev. Lett. 97, 263901, (2006). durations of less than one atomic unit of time. Opt. Lett. 34,
73. Kobayashi, Y., Sekikawa, T., Nabekawa, Y. & Watanabe, S. 27-fs extreme 3337–3339 (2009).
ultraviolet pulse generation by high-order harmonics. Opt. Lett. 23, 64–66 (1998).
74. Tzallas, P., Charalambidis, D., Papadogiannis, N. A., Witte, K. & Tsakiris, G. D. Acknowledgements
Direct observation of attosecond light bunching. Nature 426, 267–271 (2003). The research leading to the results presented in this Review received funding from the
75. Okino, T. et al. Attosecond nonlinear Fourier transformation spectroscopy of European Research Council under the European Community’s Seventh Framework
CO2 in extreme ultraviolet wavelength region. J. Chem. Phys. 129, Programme (FP7/2007-2013)/ERC grant agreement n.227355-ELYCHE. The authors
161103 (2008). acknowledge financial support from the Italian Ministry of Research (FIRB-IDEAS
76. Furukawa, Y. et al. Nonlinear Fourier-transform spectroscopy of D2 using high- RBID08CRXK), support from the European Union under contract n.228334 JRA-
order harmonic radiation. Phys. Rev. A 82, 013421 (2010). ALADIN (Laserlab Europe II) and from MC-RTN ATTOFEL (FP7-238362). G.S.
77. Skantzakis, E. et al. Tracking autoionizingwavepacket dynamics at the 1-fs acknowledges financial support from the Alexander von Humboldt Foundation
temporal scale. Phys. Rev. Lett. 105, 043902 (2010). (project Tirinto).

nature photonics | VOL 5 | NOVEMBER 2011 | www.nature.com/naturephotonics 663


© 2011 Macmillan Publishers Limited. All rights reserved.

You might also like