ArticlePDF Available

Unusual under-threshold ionization of neon clusters studied by ion spectroscopy

IOP Publishing
Journal of Physics B Atomic Molecular and Optical Physics
Authors:

Abstract and Figures

We carried out time-of-flight mass spectrometry for neon clusters that were exposed to intense free electron laser pulses with the wavelength of 62 nm, which induce optical transition from the ground state (2s2 2p6) to an excited state (2s2 2p5nl ) in the Ne atoms. In contrast to Ne+ ions produced by two-photon absorption from isolated Ne atoms, the Ne+ ion yield from Ne clusters shows a linear dependence on the laser intensity (I). We discuss the ionization mechanisms which give the linear behaviour with respect to I and expected features in the electron emission spectrum.
This content is subject to copyright. Terms and conditions apply.
This content has been downloaded from IOPscience. Please scroll down to see the full text.
Download details:
IP Address: 94.154.24.1
This content was downloaded on 31/10/2013 at 21:27
Please note that terms and conditions apply.
Unusual under-threshold ionization of neon clusters studied by ion spectroscopy
View the table of contents for this issue, or go to the journal homepage for more
2013 J. Phys. B: At. Mol. Opt. Phys. 46 164023
(http://iopscience.iop.org/0953-4075/46/16/164023)
Home Search Collections Journals About Contact us My IOPscience
IOP PUBLISHING JOURNAL OF PHYSICS B: ATOMIC, MOLECULAR AND OPTICAL PHYSICS
J. Phys. B: At. Mol. Opt. Phys. 46 (2013) 164023 (5pp) doi:10.1088/0953-4075/46/16/164023
Unusual under-threshold ionization of
neon clusters studied by ion spectroscopy
K Nagaya1,2, A Sugishima1,2, H Iwayama1,2,HMurakami
1,2,MYao
1,2,
H Fukuzawa2,3, X-J Liu2,3,KMotomura
2,3,KUeda
2,3,NSaito
2,4,
LFoucar
2,3,5, A Rudenko2,6,MKurka
2,3,7,K-UK
¨
uhnel2,7, J Ullrich2,7,8,
A Czasch9,RD
¨
orner9,RFeifel
10, M Nagasono2, A Higashiya2,
M Yabashi2, T Ishikawa2, T Togashi2,11,HKimura
2,11 and H Ohashi2,11
1Department of Physics, Graduate School of Science, Kyoto University, Kyoto 606-8502, Japan
2RIKEN, XFEL Project Head Office, Hyogo 679-5148, Japan
3Institute of Multidisciplinary Research for Advanced Materials, Tohoku University, Sendai 980-8577,
Japan
4National Metrology Institute of Japan, AIST, Tsukuba 305-8568, Japan
5Max-Planck-Institut fuer Medizinische Forschung, Jahnstraße 29, D-69120 Heidelberg, Germany
6J R Macdonald Laboratory, Department of Physics, Kansas State University, Manhattan, KS 66506,
USA
7Max-Planck-Insitut f¨
ur Kernphysik, D-69117 Heidelberg, Germany
8Physikalisch-Technische Bundesanstalt, D-38116 Braunschweig, Germany
9Institut f¨
ur Kernphysik, Universit¨
at Frankfurt, D-60486 Frankfurt, Germany
10 Department of Physics and Astronomy, Uppsala University, SE-751 20 Uppsala, Sweden
11 Japan Synchrotron Radiation Research Institute, Hyogo 679-5198, Japan
E-mail: nagaya@scphys.kyoto-u.ac.jp
Received 27 February 2013, in final form 25 June 2013
Published 13 August 2013
Online at stacks.iop.org/JPhysB/46/164023
Abstract
We carried out time-of-flight mass spectrometry for neon clusters that were exposed to intense
free electron laser pulses with the wavelength of 62 nm, which induce optical transition from
the ground state (2s22p6) to an excited state (2s22p5nl ) in the Ne atoms. In contrast to Ne+
ions produced by two-photon absorption from isolated Ne atoms, the Ne+ion yield from Ne
clusters shows a linear dependence on the laser intensity (I). We discuss the ionization
mechanisms which give the linear behaviour with respect to Iand expected features in the
electron emission spectrum.
(Some figures may appear in colour only in the online journal)
1. Introduction
‘More is different’ is a famous phrase given by Anderson
[1], who proposed that, as the number of constituent atoms
increases, new phenomena could emerge. In this work, we
investigate ionization processes of clusters irradiated by free
electron laser (FEL) pulses. Here, ‘More is different’ has two
aspects: we can control not only the number of constituent
atoms but also the number of photons over a wide range.
Multi-photon absorption by a single atom, which leads to
the production of a highly charged ion, is a well-known
example for varying the number of photons, while the shift
of the ionization potential (IP) with cluster size is an example
for varying the number of atoms e.g. in the case of the
single-photon absorption. A particularly intriguing situation
is encountered when a cluster is exposed to photons whose
energy is lower than the IP. In this case, if any ionization
event (i.e. under-threshold ionization) takes place, it must be
an effect of ‘More is different’, either due to the number of
photons or to the number of atoms. Furthermore, if the photon
energy is tuned to the Rydberg states or exciton levels, various
auto-ionization processes such as an interatomic Coulombic
decay (ICD) [2,3] and other phenomena like the exciton–Mott
transition (EMT) [4] are expected to occur.
The ICD in general is induced by a two-centre energy
transfer, and conventionally it is triggered by ionizing an
0953-4075/13/164023+05$33.00 1© 2013 IOP Publishing Ltd Printed in the UK & the USA
J. Phys. B: At. Mol. Opt. Phys. 46 (2013) 164023 K Nagaya et al
inner-valence electron. The ICD was proposed by Cederbaum
and has been verified experimentally in rare-gas clusters
[58] and molecular clusters [9,10]. Recently, Kuleff et al
[11] proposed a novel ICD mechanism, in which two electrons
are photo-excited from outer valence orbitals to the Rydberg
states and one of them is then emitted by using the relaxation
energy from the other. EMT is defined as an insulator-to-
metal transition, in which the electron–hole correlation plays
a crucial role: an exciton gas that was created by an optical
means can be transformed to an electron–hole plasma when
the exciton density is high enough to induce screening effects
by the overlap of the wavefunctions. In the rare-gas cluster,
the evolution of the Rydberg excited states to the excitons has
been investigated as a function of cluster size [12,13], but the
EMT has not been reported.
In this study, we adopted neon clusters with an average
size Nof 1000 atoms and exposed them to extreme ultraviolet
free electron laser (EUV-FEL) pulses with a wavelength of
62 nm, which corresponds to the optical transition from the
ground state (2s22p6) to an excited state (2s22p5nl )inthe
Ne atom [14]. We found that, whereas the Ne+ion yield from
the uncondensed Ne gas shows a quadratic dependence on
the laser intensity (I), indicating two-photon absorption, the
Ne+ion yield from Ne clusters shows a linear dependence on
the laser intensity. We discuss possible ionization mechanisms
which give rise to such a linear behaviour and predict expected
features in the electron emission spectrum.
2. Experiment
The experiments were performed at the SPring-8 Compact Self
Amplified Spontaneous Emission (SASE) Source (SCSS) test
accelerator in Japan [15]. Our experimental setup was almost
the same as the one reported in [16,17]. Briefly, the cluster
beam crossed the FEL beam at 45in the horizontal plane. The
photon energy was tuned to 20 eV (62 nm). The FEL beam was
partially blocked by a 1.5 mm wide horizontal beam stopper
before the ionization region, so that the unfocused beam did not
irradiate the cluster beam directly. The FEL beam was focused
back onto the cluster beam by a multi-layer focusing mirror
fabricated at the Lawrence Berkeley National Laboratory,
whichwasthesameasusedin[16]. Taking all the optical
elements (deflecting and focusing mirrors, etc) between the
radiation source point and the ionization volume into account,
we estimated the power density in the focus spot to be at
most 3×1014 Wcm
2at full power of the FEL, assuming a
diffraction limited focus size of 3 μm in diameter and a pulse
length of 30 fs [18]. The measured spectral fluctuation of the
FEL was 0.3 eV (FWHM) in this experiment.
The cluster beam was prepared by the adiabatic expansion
of a Ne gas through a pulsed 250 μm nozzle. The stagnation
pressure was 4.6 bar and the nozzle temperature was 80 K
[19]. The average cluster size Nwas estimated to be 1000
atoms according to scaling laws [20,21]. To avoid space–
charge effects due to the ionization of a background gas, the
pulsed gas jet was cut to 0.6 mm width and 0.4 mm height
with knife-edge slits and travelled to the focus spot located at
1.7 m downstream from the nozzle.
Time−of−flight [ s]
Intensity [counts/shot]
<GMD>
(a) 1.1
(b) 2.5
(c) 3.2
(d) 3.6
20Ne
22Ne
GMD peak height [a.u.]
counts
1.1 2.5 3.2
3.6
µ
mass [amu]
5.6 5.8 6 6.2
0
0.1
0.2
20 22
024
0
500
Figure 1. Ion TOF spectra of a Ne1000 cluster beam irradiated by
62 nm FEL pulses. The numbers in the figure denote the peak height
of the gas monitor detector (GMD). The inset shows the power
distribution in the GMD for the full and for one-third of the laser
power. Colour areas correspond to the laser power distributions for
each TOF spectrum.
We measured time-of-flight (TOF) spectra with our
momentum imaging spectrometer [16,17]. Fragment ions
were vertically extracted by a uniform electrostatic field. They
travelled through an extraction region (75 Vcm1electric field
strength, 40 mm in length), an acceleration region (110 V cm1
electric field strength, 52 mm in length) and a field-free region
(308 mm), and were finally detected by a microchannel-plate
(MCP) detector equipped with a three-layer delay-line anode
(RoentDek HEX120) [22].
The measurements were carried out with the full laser
power and with one-third of it. In the latter case, the intensity
was attenuated by transmitting the laser pulses through a gas
chamber filled with Ar. Since the pulses from the SASE
source fluctuate shot to shot, we measured the intensity
of each pulse by a gas monitor detector (GMD). At the
same time, the TOF spectrum was also measured for each
laser shot. Both signals were recorded by an eight-channel
digitizer (Acqiris DC282×2), and the timing signals were
extracted by a software constant fraction discriminator [23].
This procedure enabled us to deduce precise laser-power
dependences of the TOF spectra. In this study, we discuss
the FEL intensity dependence based on the GMD pulse height
which is proportional to the photon number in each FEL pulse
because of the uncertainty in estimating the focal size of FEL.
3. Results
In figure 1, spectrum (a) displays the TOF spectrum
accumulated during the experiment with one-third of the full
laser power. The corresponding power distribution is shown
as a black area in the inset of figure 1as a function of the
GMD peak height. For the one-third power, the weighted
average of the GMD peak height is 1.1. Since the power
2
J. Phys. B: At. Mol. Opt. Phys. 46 (2013) 164023 K Nagaya et al
0
0.1
0
0.05
0
0.02
0.04
0.06
20 22 24
0
0.02
0.04
<GMD>
3.2
1.1
exp
fit (total)
fit (component)
mass [amu]
2.5
3.6
<GMD>
<GMD>
<GMD>
Intensity [counts/shot]
(d)
(c)
(b)
(a)
Figure 2. Decomposition of the TOF spectrum into four Lorentzian
curves: 20Ne+from the uncondensed gas and from the cluster, and
22Ne+from the uncondensed gas and from the cluster.
distribution is much wider for the full power experiment, we
have divided the data into three parts: (b) below 3.0, (c) from
3.0 to 3.4 and (d) above 3.4 (green, blue and red areas). The
average peak height is 2.5 for (b), 3.2 for (c) and 3.6 for (d).
In each corresponding TOF spectrum, a prominent peak is
observed at the mass-to-charge ratio m/qof 20, and a smaller
one at m/q=22, corresponding to 20Ne+and its isotope
22Ne+, respectively. In addition, broad distributions centred at
m/q=20 and m/q=22 are also seen. The broad features
are due to the fragments from clusters, whereas the sharp
peaks correspond to ions from the uncondensed gas, because
large fragment energies resulting in a broad TOF distribution
can only be produced by Coulomb repulsion between ions
from the cluster. No multiply charged ions such as Ne2+were
observed.
We now examine the laser intensity dependence of the
ion yield to clarify the multiphoton ionization mechanism. For
this purpose, we decompose the TOF spectra into a sum of
two narrow and two wide Lorentzian curves, as displayed in
figure 2. The height and width parameters were determined by
a least-squares fitting. In the fitting procedure, the m/qregion
below 20 was eliminated, because the MCP signal in this region
was contaminated by H2O. We analysed the apparent ratio of
20Ne+to 22 Ne+for uncondensed atoms as a function of FEL
intensity and found a progressive deviation from the natural
abundance (about 9:1) with an increase of the laser power.
Thus, we use the data for 22Ne+hereafter. The benefit of using
the 22Ne+isotope for a quantitative analysis was demonstrated
in a previous paper [24].
In figure 3,the22 Ne+yield is plotted versus the laser
power. The directly measured GMD signal is shown on the
abscissa. The integrated intensity of the sharp peak exhibits a
quadratic dependence on the laser intensity, indicating that the
ionization of isolated atoms was due to two-photon absorption
as expected. In contrast, the ion yields from the clusters show a
linear dependence on the laser intensity. To check the influence
of the choice of how the data are divided, results of the other
such choices, in which the full power data are divided into two
parts, are displayed by diamonds in figure 3(a).
4. Discussion
These experimental results suggest that the ionization process
is qualitatively different between free neon atoms and neon
atoms embedded in clusters. Two-photon ionization was
observed for Ne atoms, while an unexpected linear dependence
of the ion yield on the FEL intensity was found for Ne clusters.
Although it is difficult to clarify the ionization process based
only on these results of ion spectrometry alone, we try here to
discuss the possible cluster ionization mechanisms and predict
the expected features of the electron emission spectrum in each
case.
024
0
1
GMD peak height [a.u.]
22Ne yield [a.u.]
atom
cluster
(a)
51
10−1
100
GMD peak height [a.u.]
22Ne yield [a.u.]
atom
cluster
(b)
Figure 3. (a) 22Ne+yield from uncondensed atoms (closed symbols) and from clusters (open symbols) as a function of the GMD peak
height. The intensities of these two components are re-normalized at the highest point of the GMD scale. The horizontal error bars denote
the standard deviation, and the vertical error bars denote the statistical errors, which are defined by the square root of the integrated number
of 22Ne+. The line guides a quadratic dependence on the laser power for atoms and a linear dependence for clusters. (b) Log–log plot of the
22Ne+yields. The lines are guides for the eyes the same as (a).
3
J. Phys. B: At. Mol. Opt. Phys. 46 (2013) 164023 K Nagaya et al
In the infrared spectral region, tunelling ionization
followed by plasma heating is known to give efficient cluster
ionization even if the photon energy is much smaller than the
IP [25,26]. In such a case, the ponderomotive energy should be
several electron volts or more, and the intensity dependence of
the ion yield would be nonlinear, contrary to what we observe
here. Therefore, we can safely ignore the ionization of the
cluster induced by the electric field of the laser.
The first candidate for the ionization mechanism of neon
clusters is the direct two-photon ionization of the constituent
atoms. In this case, the intensity dependence of the ion
yield should be quadratic, which is however inconsistent
with our experimental results. It should be noted that the
linear intensity dependence could be observed even in such
a two-photon process when the saturation of ionization is
reached for each constituent atom [16,27]. Such saturation
effects can however be excluded in this experiment, since
the simultaneously measured atomic ionization shows a
quadratic intensity dependence. Since the cross-section of
the two-photon ionization of the cluster constituent atoms is
expected to be of the same order of magnitude as that of an
isolated atom, the two-photon route could be excluded as an
ionization process of neon clusters based on the FEL intensity
dependence of the ion yield.
Thus, we have to search for alternative ionization
mechanisms in which the number of emitted ions is
proportional to the number of photon-absorbing atoms.
As a candidate of such an ionization process, a novel
ICD mechanism via resonant excited states is proposed by
Kuleff [11]. When many Rydberg excited atoms are formed
within a cluster, energy transport by the exchange of a virtual
photon becomes possible between the excited states of two
neighbouring atoms, and hence one atom can ionize another
atom utilizing this transition energy. In this case, the number of
generated ions is expected to be approximately one-half of the
number of excited atoms. Since the number of excited atoms
would be proportional to the FEL intensity, this type of ICD is
consistent with our experimental results. If this is the case, the
ICD electron should appear as a peak in the electron emission
spectrum.
Let us consider a situation where one single ion is created
in the cluster. The creation of this single ion can be either
due to the direct two-photon ionization of a single atom in the
cluster or due to the ICD following the sequential excitation
of two atoms by two photons. Since the binding potentials of
atoms in the cluster are distorted by the Coulombic potential
of a single neighbouring ion, the sequential excitation of other
atoms leads to the delocalization of excited electrons and to
the production of nano-plasma. This is the same effect as the
ionization barrier suppression found in the inner ionization of
the cluster [28]. In this case, the ion yield will linearly depend
on the FEL intensity too, since the number of ions is expected
to be proportional to the number of excited atoms. The electron
emission spectrum should be dominated by thermal electrons
from nano-plasma with an additional peak of ‘the first ionizing
atom’ at about 18 eV ( =2hνIP), independent of whether the
first electron emission is due to direct two-photon ionization
or double excitation ICD.
Now, if the multiple excitation takes place before the first
electron ejection as discussed above, the excited electrons will
be delocalized also because of the lowering of the inner IPs
[28]. As briefly noted in the introduction, the delocalization
of the excited electrons can be viewed as the EMT, which
has been intensively studied for bulk silicon [29]. When a
sufficient number of excitons are created in bulk silicon, it
is well known that the bound electrons are spontaneously
delocalized and transformed to a plasma state by the overlap
of the exciton wavefunctions. Also in this case of a neon
cluster, if the FEL irradiation achieves a high enough exciton
density, electrons will be liberated by the thermoelectronic
emission from the nano-plasma produced by EMT. Then, as in
the previous case, the electron emission spectrum is expected
to show an exponential component corresponding to thermal
electrons. The only difference from the previous case would be
the lack of a sharp peak corresponding to the direct two-photon
ionization of a single atom or a double excitation ICD.
5. Summary
We carried out time-of-flight mass spectrometry on neon
clusters that were exposed to intense free electron laser
(FEL) pulses with the wavelength of 62 nm, which induce
transitions from the ground state (2s22p6) to an excited state
(2s22p5(2P1/2,3/2) 3d) in a Ne atom. In contrast to isolated
Ne atoms, in which a Ne+ion is produced by two-photon
absorption, the Ne+ion yield from Ne clusters shows a linear
dependence on the laser intensity. We have considered possible
ionization mechanisms consistent with the experimentally
observed FEL intensity dependence of the ion yields and
discussed the expected electron emission spectrum for each
case. We believe that the additional information obtainable
from electron emission spectra will be helpful to decide the
underlying mechanism.
Acknowledgments
We thank Dr A I Kuleff, Dr Ph V Demekhin, Professor L S
Cederbaum and Professor U Saalmann for fruitful discussions.
We are grateful to the SCSS Test Accelerator Operation
Group at RIKEN for continuous support in the course of
the studies. We are also grateful to A Belkacem and the
optics group at the LBNL for fabricating the home-made
focusing mirror. This work was supported by the X-ray Free
Electron Laser Utilization Research Project of the Ministry of
Education, Culture, Sports, Science and Technology of Japan
(MEXT), by the grant-in-aid for the Global COE Program
‘The Next Generation of Physics, Spun from Universality and
Emergence’ from the MEXT, by the grants-in-aid (20310055,
21244062) from the Japan Society for the promotion of
Science (JSPS), by the IMRAM project and by the MPG
Advanced Study Group within CFEL. RD acknowledges
support by the DFG FOR 1789. RF thanks the Swedish
Research Council (VR) for financial support.
4
J. Phys. B: At. Mol. Opt. Phys. 46 (2013) 164023 K Nagaya et al
References
[1] Anderson P W 1972 Science 177 393
[2] For a recent review, see Averbukh V et al 2011 J. Electron
Spectrosc. Relat. Phenom. 183 36
[3] Cederbaum L S, Zobeley J and Tarantelli F 1997 Phys. Rev.
Lett. 79 4778
[4] Mott N F 1968 Rev. Mod. Phys. 40 677
[5] Marburger S, Kugeler O, Hergenhahn U and M¨
oller T 2003
Phys. Rev. Lett. 90 203401
[6] ¨
Ohrwall G et al 2004 Phys. Rev. Lett. 93 173401
[7] Jahnke T et al 2004 Phys. Rev. Lett. 93 163401
[8] Ouchi T et al 2011 Phys. Rev. A83 053415
[9] Jahnke T et al 2010 Nature Phys. 6139
[10] Mucke M et al 2010 Nature Phys. 6143
[11] Kuleff A I, Gokhberg K, Kopelke S and Cederbaum L S 2010
Phys. Rev. Lett. 105 043004
[12] Stapelfeldt J, W¨
ormer J and M¨
oller T 1989 Phys. Rev. Lett.
62 98
[13] W¨
ormer J, Karnbach R, Joppien M and M¨
oller T 1996
J. Chem. Phys. 104 8269
[14] Chan W F, Cooper G, Guo X and Brion C E 1992 Phys. Rev. A
45 1420
[15] Shintake T et al 2008 Nature Photon. 2555
[16] Motomura K et al 2009 J. Phys. B: At. Mol. Opt. Phys.
42 221003
[17] Fukuzawa H et al 2009 J. Phys. B: At. Mol. Opt. Phys.
42 181001
[18] Moshammer R et al 2011 Opt. Express 19 21698
[19] Nagaya K et al 2010 J. Electron Spectrosc. Relat. Phenom.
181 125
[20] Hagena O F and Obert W 1972 J. Chem. Phys. 56 1793
[21] Karnbach R, Joppien M, Stapelfeldt J, W¨
ormer J and
M¨
oller T 1993 Rev. Sci. Instrum. 64 2838
[22] Jagutzki O et al 2002 IEEE Trans. Nucl. Sci. 49 2477
[23] Motomura K et al 2009 Nucl. Instrum. Methods
606 770
[24] Sugishima A et al 2012 Phys. Rev. A86 033203
[25] Ditmire T, Donnelly T, Rubenchik A M, Falcone R W
and Perry M D 1996 Phys. Rev. A53 3379
[26] Krainov V P and Smirnov M B 2002 Phys. Rep. 370 237
[27] Sorokin A A, Wellh¨
ofer M, Bobashev S V, Tiedtke K
and Richter M 2007 Phys. Rev. A75 051402
[28] Saalmann U and Rost J M 2003 Phys. Lett. 91 223401
[29] Shah J, Combescot M and Dayem A H 1977 Phys. Rev. Lett.
38 1497
5
... The intensity dependence of ionization by multiple below-threshold excitations was measured by ion spectroscopy. 306,515 This information, however, was not sufficient to make definite conclusions on the underlying processes. A study on He droplets at FERMI 516 at the photon energy of the 1s → 2p absorption maximum in clusters, 21.4 eV, also presented the electron spectra. ...
Article
Full-text available
Interatomic or intermolecular Coulombic decay (ICD) is a nonlocal electronic decay mechanism occurring in weakly bound matter. In an ICD process, energy released by electronic relaxation of an excited atom or molecule leads to ionization of a neighboring one via Coulombic electron interactions. ICD has been predicted theoretically in the mid nineties of the last century, and its existence has been confirmed experimentally approximately ten years later. Since then, a number of fundamental and applied aspects have been studied in this quickly growing field of research. This review provides an introduction to ICD and draws the connection to related energy transfer and ionization processes. The theoretical approaches for the description of ICD as well as the experimental techniques developed and employed for its investigation are described. The existing body of literature on experimental and theoretical studies of ICD processes in different atomic and molecular systems is reviewed.
... At SCSS, over the photon energy range 20 and 24 eV, with the FEL power density in the range 10 11 -10 14 W cm −2 , Nagaya and coworkers investigated extensively EUVFEL-induced dynamics of various rare gas clusters [389][390][391][392][393][394][395][396]. Under these experimental conditions they found (via ion spectrometry) frustration of direct cluster photoionization stemming from the strong Coulomb potential of the highly ionized cluster but no indications for heating mechanisms other than sequential photoionization of the individual atoms in the cluster [389,390,393]. ...
Article
Full-text available
This review is focused on free-electron lasers (FELs) in the hard to soft x-ray regime. The aim is to provide newcomers to the area with insights into: the basic physics of FELs, the qualities of the radiation they produce, the challenges of transmitting that radiation to end users and the diversity of current scientific applications. Initial consideration is given to FEL theory in order to provide the foundation for discussion of FEL output properties and the technical challenges of short-wavelength FELs. This is followed by an overview of existing x-ray FEL facilities, future facilities and FEL frontiers. To provide a context for information in the above sections, a detailed comparison of the photon pulse characteristics of FEL sources with those of other sources of high brightness x-rays is made. A brief summary of FEL beamline design and photon diagnostics then precedes an overview of FEL scientific applications. Recent highlights are covered in sections on structural biology, atomic and molecular physics, photochemistry, non-linear spectroscopy, shock physics, solid density plasmas. A short industrial perspective is also included to emphasise potential in this area.
... An example for a practical application is the characterization of heterogeneous structures and interfaces using spectral signatures attributed to ICD 15,16 . The high density of atomic excited states populated within a single cluster also results in collective autoionization processes in which several excited atoms are involved as predicted theoretically 17 and observed experimentally at a free electron laser (FEL) [18][19][20] . ...
Article
Full-text available
Irradiation of nanoscale clusters and large molecules with intense laser pulses transforms them into highly-excited non- equilibrium states. The dynamics of intense laser-cluster interaction is encoded in electron kinetic energy spectra, which contain signatures of direct photoelectron emission as well as emission of thermalized nanoplasma electrons. In this work we report on a so far not observed spectrally narrow bound state signature in the electron kinetic energy spectra from mixed Xe core - Ar shell clusters ionized by intense extreme-ultraviolet (XUV) pulses from a free-electron-laser. This signature is attributed to the correlated electronic decay (CED) process, in which an excited atom relaxes and the excess energy is used to ionize the same or another excited atom or a nanoplasma electron. By applying the terahertz field streaking principle we demonstrate that CED-electrons are emitted at least a few picoseconds after the ionizing XUV pulse has ended. Following the recent finding of CED in clusters ionized by intense near-infrared laser pulses, our observation of CED in the XUV range suggests that this process is of general relevance for the relaxation dynamics in laser produced nanoplasmas.
... Stimulated by the work of Kuleff et al. [26] and following the measurement of ionic fragmentation upon 2p → 3d multiple excitation (20.26 eV) of Ne clusters [30], Yase et al. [31] investigated the EUVFEL-induced electron emission from Ne clusters at SCSS (SPring-8 Compact SASE Source test accelerator, Japan) [28]. However, in the FEL intensity range they used (≥ 3 × 10 13 W=cm 2 ), no ICD processes could be identified. ...
Article
Full-text available
Ne clusters ( ∼ 5000 atoms ) were resonantly excited ( 2 p → 3 s ) by intense free electron laser (FEL) radiation at FERMI. Such multiply excited clusters can decay nonradiatively via energy exchange between at least two neighboring excited atoms. Benefiting from the precise tunability and narrow bandwidth of seeded FEL radiation, specific sites of the Ne clusters were probed. We found that the relaxation of cluster surface atoms proceeds via a sequence of interatomic or intermolecular Coulombic decay (ICD) processes while ICD of bulk atoms is additionally affected by the surrounding excited medium via inelastic electron scattering. For both cases, cluster excitations relax to atomic states prior to ICD, showing that this kind of ICD is rather slow (picosecond range). Controlling the average number of excitations per cluster via the FEL intensity allows a coarse tuning of the ICD rate.
Thesis
This work presents comparative measurements of different noble gas clusters doped with xenon (Xe), argon (Ar), calcium (Ca) and water (H2O) irradiated by intense mid-infrared and near-infrared femtosecond laser pulses. A new data acquisition method to correlate single shot velocity map imaging (VMI) of electrons and ion time of flight (TOF) measurements is introduced. With this method the analysis of the electron spectra of individual cluster explosions is possible. The photoelectron spectra and the peak maximal kinetic energy of electrons for the VMI images were analyzed. The maximal kinetic energy and the number of electrons produced by a single Coulomb explosion were correlated. For certain conditions, the process can be modeled as a uniform charge spherical cloud where the electron on the surface of the sphere determines the maximal kinetic energy resulting from the explosion. Under the same cluster size and doping conditions, the near-infrared (NIR) laser pulse drives the plasma process more efficiently than the mid-infrared (MIR) laser field. It was demonstrated that there is a laser intensity threshold necessary to start an optimal ionization of the cluster. The efficiency of the plasma formation has a strong relation to the cluster size and the doping level. Bigger droplets present a more efficient plasma ignition and the amount of detected electrons is enhanced. Neon clusters showed a similar ignition rate with a higher plasma heating efficiency than He nanodroplets under the MIR laser. Moreover, there exists an optimum doping level to efficiently start the plasma formation and the combination of certain dopant species appears to enhance the plasma process compared to one single dopant element. Using Xe and Ca doping, the replacement of a few Xe for Ca atoms increases the ignition probability by a factor of two in contrast to the ignition using pure Xe. Finally, it is shown that the number of cycles in the laser pulse plays a crucial role in the plasma ignition. At constant pulse energy, a longer pulse is more efficient in driving the process despite the losses in the peak intensity.
Article
Full-text available
The ionization dynamics of helium droplets irradiated by intense, femtosecond extreme ultraviolet (XUV) pulses is investigated in detail by photoelectron spectroscopy. Helium droplets are resonantly excited to atomic-like 2p states with a photon energy of 21.5 eV and autoionize by interatomic Coulombic decay (ICD). A complex evolution of the electron spectra as a function of droplet size (250 to 106 He atoms per droplet) and XUV intensity (109–1012 W cm−2) is observed, ranging from narrow atomic-like peaks that are due to binary autoionization, to an unstructured feature characteristic of electron emission from a nanoplasma. The experimental results are analyzed and interpreted with the help of a numerical simulation based on rate equations taking into account all relevant processes—multi-step ionization, electronic relaxation, ICD, secondary inelastic collisions, desorption of electronically excited atoms, and collective autoionization (CAI).
Article
The fragmentation of doubly and triply charged mercury clusters is theoretically studied to analyze an experiment performed by Katakuse’s group at Osaka University [T. Satoh et al., J. Mass Spectrom. Soc. Jpn. 51, 391 (2003)]. The fission barrier is calculated using a liquid-drop model proposed by Echt et al. In the decay of doubly charged clusters, the barrier height is found to take the minimum value for nearly symmetric fission. On the other hand, in the decay of triply charged clusters, the barrier is the lowest for strongly asymmetric fission. These results well explain the product size distribution observed in the experiment. The appearance size for multiply charged clusters measured in the experiment is found to be the size where the fission barrier is equal to the monomer evaporation energy. These findings provide evidence that small mercury clusters behave like van der Waals clusters in the process of fragmentation.
Chapter
Clusters and nanocrystals constitute intermediates between molecules and condensed matter. Due to their finite size, clusters have a wide spectrum of applications ranging from building blocks for novel materials to model systems for fundamental investigations about light – matter interactions. Short-wavelength radiation from synchrotron radiation sources and free-electron lasers (FELs) allows the detailed investigation of their geometric, electronic, and magnetic structure as well as dynamical processes. Conversely, clusters can serve as idealized sample systems for the development of new experimental techniques and pioneering experiments with novel x-ray sources. The chapter starts with a brief introduction to cluster physics, followed by a comprehensive overview of research performed at synchrotron light sources on van der Waals, metal, and semiconductor clusters. With the advent of short-wavelength FELs, a new research field in the x-ray peak intensity regime has opened. Experiments on single clusters, such as x-ray imaging and tracing ultrafast dynamics, now become possible.
Article
Full-text available
At the transition from the gas to the liquid phase of water, a wealth of new phenomena emerge, which are absent for isolated H2O molecules. Many of those are important for the existence of life, for astrophysics and atmospheric science. In particular, the response to electronic excitation changes completely as more degrees of freedom become available. Here we report the direct observation of an ultrafast transfer of energy across the hydrogen bridge in (H2O)2 (a so-called water dimer). This intermolecular coulombic decay leads to an ejection of a low-energy electron from the molecular neighbour of the initially excited molecule. We observe that this decay is faster than the proton transfer that is usually a prominent pathway in the case of electronic excitation of small water clusters and leads to dissociation of the water dimer into two H2O + ions. As electrons of low energy (∼0.7-20 eV) have recently been found to efficiently break-up DNA constituents, the observed decay channel might contribute as a source of electrons that can cause radiation damage in biological matter.
Article
Full-text available
We have investigated multiple ionization of atomic argon by extreme-ultraviolet light pulses (62 nm, 100 fs in width, <2 × 1014 W cm−2) at the free-electron laser facility in Japan, and observed highly charged ions with the charge state up to +6. The measured laser power dependence of the highly charged ions indicates that the multiple ionization proceeds via the sequential stripping of electrons.
Article
In sharp contrast to molecules, electronic states of clusters with an excited intermediate-shell electron can efficiently decay via an intermolecular Coulombic mechanism. Explicit examples are presented using large scale ab initio propagator calculations. The mechanism is illustrated and its generality is stressed.
Article
Measurements of luminescence spectra in pure Si demonstrate the dissociation of excitons with increasing excitation intensity. The density at which the dissociation begins and its temperature variation agree well with Mott's predictions. We have also determined a phase diagram of Si, including the effects of electron-hole-liquid and Mott transition.
Article
At the soft-x-ray free-electron laser FLASH in Hamburg, we have studied multiphoton ionization on neon and helium by ion mass-to-charge spectroscopy. The experiments were performed in a focused beam at 42.8 and 38.4 eV photon energy and irradiance levels up to 1014 W∕cm2. Direct, sequential, and resonant two-, three-, and four-photon excitations were investigated by quantitative measurements as a function of the absolute photon intensity. The atomic and ionic photoionization cross sections derived indicate a clear dominance of sequential compared to direct multiphoton processes.
Article
Using momentum-resolved electron-ion multicoincidence spectroscopy, we have investigated interatomic Coulombic decay (ICD) in the heteronuclear NeAr dimer following Ne 1s Auger decay. The measured intensity ratio for the three ICD transitions Ne2+(2s-12p-1 1P)Ar to Ne2+(2p-2 1S)–Ar+(3p-1), Ne2+(2s-12p-1 1P)Ar to Ne2+(2p-2 1D)-Ar+(3p-1), and Ne2+(2s-12p-1 3P)Ar to Ne2+(2p-2 3P)-Ar+(3p-1) reasonably agree with predictions. The kinetic energy release distribution for the fragmentation to Ne2+(2p-2 1D)-Ar+(3p-1) after the ICD transition from singlet Ne2+(2s-12p-1 1P)Ar state, which is a mirror image of the kinetic energy distribution of the emitted ICD electrons, suggests that the corresponding ICD rate is roughly two times lower than predicted by ab initio calculations.
Article
The multiple ionization of Ar-core–Ne-shell clusters in intense extreme-ultraviolet laser pulses (λ∼62 nm) from the free-electron laser in Japan was investigated utilizing a momentum imaging technique. The Ar composition dependence of the kinetic energies and the yields of the fragment ions give evidence for charge transfer from the Ar core to the Ne shell. We have extended the uniformly charged sphere model originally applied to pristine clusters [ Islam et al. Phys. Rev. A 73 041201 (2006)] to the core-shell heterogeneous clusters to estimate the amounts of charge and energy transfers.
Article
The evolution of excitonic energy levels (Wannier and Frenkel type) is investigated for ArN clusters in the range N=200–106 using fluorescence excitation spectroscopy. In the case of Wannier excitons, a pronounced blue shift of the absorption bands relative to the position in the infinite solid is observed. As a consequence of the lower dimensionality, the shift of the transition energy of surface excitons is considerably smaller than that of the bulk states of clusters. The evolution with size is discussed within several theoretical models for exciton confinement. In addition, model calculations are performed for bulk excitons which give good quantitative agreement with the experimental results. In the case of n=1 Frenkel or intermediate type excitons, there are blue and red shifts observed. The spectral shift of (3p→4s) and deep valence (3s→4p) excitations differs considerably. From the shift of the transition energies the exciton mass of the (3p→4s) exciton is derived. © 1996 American Institute of Physics.
Article
To study homogeneous condensation in an expanding nozzle flow, cluster beams were sampled from the core of the flow field and transferred into a vacuum chamber for further analysis. Sonic and hypersonic nozzles with throat diameters 0.015 cm ≤ d ≤ 0.15 cm were used. Source temperature was varied between 120 ≤ T0 ≤ 450°K, source pressure between 100 ≤ p0 ≤ 12 000 torr. Test gases were the rare gases (except He), N2, and CO2. The size of the clusters (=microdroplets or ‐crystals) and the intensity of the cluster beam was measured with a through‐flow ionization detector with retarding potential system to get the mass‐to‐charge distribution of the cluster ions. The mean cluster size varied between 102 and 104 atoms∕cluster. The mean cluster size remained almost constant with increasing T0 if p0 was increased simultaneously according to the isentropic relation p0T0γ/(1−γ) = const. Considering the various types of cluster‐growth reactions one expects to get cluster beams with the same size, if p0 and T0 fall within the narrow range between the isentrope p0T0γ/(1−γ) = const and the line for equal bimolecular processes, p0T0(1.5γ−1)/(1−γ) = const. The experiments confirm this result. The same model predicts that a decrease of nozzle throat diameter d can be compensated by an increase of source pressure p0 such that p0dq = const with 0<q<1. The experimental scaling law for constant cluster size gives q=0.8 for argon and q=0.6 for CO2. Comparing different gases, the same cluster size was obtained for the rare gases if they were in corresponding states prior to expansion and if the reduced nozzle scale was the same. This confirms the model of ``corresponding jets'' which extends the thermodynamic principle of corresponding states to real gas effects in a time‐dependent system like a nozzle flow, and which applies equally to condensation in slow and fast expansions, including the transition to molecular flow.
Article
Single-pass free-electron lasers based on self-amplified spontaneous emission1, 2, 3, 4 are enabling the generation of laser light at ever shorter wavelengths, including extreme ultraviolet5, soft X-rays and even hard X-rays6, 7, 8. A typical X-ray free-electron laser is a few kilometres in length and requires an electron-beam energy higher than 10 GeV (refs 6, 8). If such light sources are to become accessible to more researchers, a significant reduction in scale is desirable Here, we report observations of brilliant extreme-ultraviolet radiation from a 55-m-long compact self-amplified spontaneous-emission source, which combines short-period undulators with a high-quality electron source operating at a low acceleration energy of 250 MeV. The radiation power reaches saturation at wavelengths ranging from 51 to 61 nm with a maximum pulse energy of 30 J. The ultralow emittance (0.6 mm mrad) of the electron beam from a CeB6 thermionic cathode9 is barely degraded by a multiple-stage bunch compression system that dramatically enhances the beam current from 1 to 300 A. This achievement expands the potential for generating X-ray free-electron laser radiation with a compact 2-GeV machine.