52
Coulomb acceleration of protons with a free-electron laser Prospects for nuclear fusion Elsa Caroline Souto Gonçalves de Abreu Dissertação para a obtenção do grau de mestre em Engenharia Física Tecnológica Júri Presidente: Prof. João Seixas Orientador: Profª. Marta Fajardo Vogais: Prof. Luís O. Silva Julho de 2007

Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

  • Upload
    others

  • View
    20

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

Coulomb acceleration of protons with a free-electron

laser

Prospects for nuclear fusion

Elsa Caroline Souto Gonçalves de Abreu

Dissertação para a obtenção do grau de mestre em

Engenharia Física Tecnológica

Júri

Presidente: Prof. João Seixas

Orientador: Profª. Marta Fajardo

Vogais: Prof. Luís O. Silva

Julho de 2007

Page 2: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

2

Acknowledgements

A big thank you to Nicusor Timneanu, my (co)supervisor here in Uppsala, for all the

help and the discussions, and to Marta Fajardo, my supervisor in Lisbon, who first made me

think of coming to Uppsala, and who helped me to make it happen. Thank you Magnus

Bergh, for the big help with Cretin, and thank you Janos Hajdu for comments and motivation.

The experiments at FLASH would not have been possible without the help of our

collaborators: Ryszard Sobierajski, Thomas Möller and Matthias Höner, the LLNL/Uppsala

collaboration, Björgvin Hjörvarsson, and Libor Juha. Thank you all.

I would also like to acknowledge the European project TUIXS for financial support

during my stay in Uppsala.

And last but not least, thank you all of you who, inside or outside the lab, in Sweden

or in Portugal, helped me enduring the winter and enjoying the summer, during this year in

Uppsala.

Page 3: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

3

Index

Abstract 4 Resumo 5 Figures and tables 6 Abbreviations 8 1 Introduction 9 2 Theory 11

2.1 Fusion reactions 11 2.2 Ionization mechanisms 13 2.3 Simulation models 14

2.3.1 Molecular dynamics simulations 14 2.3.2 Non-LTE plasma simulations 15

3 Experiments: background information 17

3.1 XFEL: working principle 17 3.2 Experiments on solids 19 3.3 Experiments on gas clusters 22

4 Results and analysis, for solids 24

4.1 Experiments on solids 24 4.1.1 General analysis 25 4.1.2 Light ions energy analysis 28 4.2 Simulations 34 4.2.1 Interaction of the FEL pulse with solid samples 35 4.3 Preliminary results on gas clusters 42

5 Discussion 45 6 Conclusions and outlook 47 References 48 Appendix A 50 Appendix B 51

Page 4: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

4

Abstract

X-ray free electron lasers (XFELs) are a new kind of light source, capable of

producing super-brilliant, ultra-short X-ray pulses with more than 10 orders of magnitude

higher peak brilliance than pulses from synchrotron sources. This extremely high peak

brilliance allows us to deposit huge amounts of energies into the sample over a very short

time, and will revolutionize many areas of physics, chemistry and biology in applications like

the imaging of single biomolecules, or the creation of highly ionized dense states of matter.

Molecular dynamics simulations, performed in Uppsala in the context of imaging

biomolecules, have predicted the ejection of protons with unexpectedly high energies, of the

order needed for fusion reactions. These protons are expelled from a nanometer-sized

sample in a Coulomb explosion, caused by interaction with an XFEL beam. This motivated

experiments at the FLASH facility, in Hamburg to study the interaction of the XFEL beam with

both solid and clustered gas samples, and to look for signatures of highly energetic protons or

deuterons, produced during the Coulomb explosion of samples rich in hydrogen or deuterium.

The energies detected, in the solid case, were high enough to support the hypothesis of

fusion. These results were analyzed and compared to predictions from a plasma code.

Improvements are still needed, mainly in an experimental point of view, and some of

them are suggested here; however the way is opened for a very exciting new area of

research related to XFELs.

Key words: X-ray free electron laser (XFEL), fusion, Coulomb Explosion, TOF (time-of-flight),

plasma simulations, Molecular Dynamics

Page 5: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

5

Resumo

Os lasers de electrões livres emitindo nos raios-x (XFELs) são uma nova fonte

luminosa, capaz de produzir impulsos ultra-curtos e extremamente brilhantes, na gama dos

raios-x, com uma intensidade 10 ordens de grandeza superior à dos sincrotrões. Enormes

quantidades de energia são depositadas nas amostras num espaço de tempo muito curto, o

que constitui uma revolução em diversas áreas da física, química e biologia, nomeadamente

para a obtenção de imagens de biomoléculas ou para a criação de estados altamente

ionizados na matéria.

Em simulações de Dinâmica Molecular, realizadas em Uppsala, foram observados

protões altamente energéticos, com capacidade para participar em reacções de fusão. Os

protões provêm da explosão, por efeito do potencial repulsivo de Coulomb, de uma amostra

nanométrica atingida pelo feixe do XFEL. Os resultados das simulações levaram à realização

de experiências no FLASH, em Hamburgo, com o intuito de procurar assinaturas deixadas

por protões ou deuterões altamente energéticos, produzidos durante a explosão de amostras

– sólidos ou clusters de gases – ricas em hidrogénio e deutério. As energias detectadas no

caso dos sólidos são suficientes para justificar a hipótese de fusão. Estes resultados foram

analisados e comparados com simulações feitas usando um código de plasma.

São ainda necessários melhoramentos, nomeadamente do ponto de vista

experimental, e alguns serão sugeridos neste estudo. No entanto, a porta está aberta para

uma nova e fascinante área de investigação relacionada com os XFELs.

Palavras-chave: Laser de electrões livres nos raios-x (XFEL), fusão, explosão de Coulomb,

TOF (time-of-flight: tempo de voo), código de plasma, Dinâmica Molecular

Page 6: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

6

Figures and tables

Figures

Figure 1: MD simulation of radiation-induced Coulomb explosion of a small protein molecule

(lysozyme) 9

Figure 2: High-energy protons leaving the molecule very early during the explosion. 9

Figure 3: Simple schemes of the photoabsorption and Auger emission processes. 14

Figure 4: Detailed scheme of the FLASH line, where the XFEL beam is produced 17

Figure 5: Scheme of an undulator 18

Figure 6: Radiation from a non-relativistic and relativistic accelerated electron 18

Figure 7: Microbunching process 19

Figure 8: Picture of the sample holder, with the samples used in September 2006 20

Figure 9: Left: picture of the inside of the chamber, with the sample holder mounted, although

not in the final position. Right: scheme of the geometry of the interaction 20

Figure 10: Picture and scheme of the Warsaw TOF detector, used during the September

2006 experiments at FLASH 21

Figure 11: Setup of the March 2007 gas experiments at FLASH and scheme of the Berlin

TOF detector used 22

Figure 12: TOF spectra for NbH, NbDH and PMMA 24

Figure 13: First step in the determination of the present ion species 26

Figure 14: Nomarski differential interference contrast microscope (x100) pictures taken from

PMMA and NbDH samples 28

Figure 15: Top: Direct TOF output for a single FEL shot on an Nb sample. Bottom:

Corresponding energy distribution for H+ ions and a multiple Gaussian fit 29

Figure 16: Top: Direct TOF output for a single FEL shot on an NbD sample. Bottom:

Corresponding energy distribution of the H+, D

+, D2

+ and Nb

+ peaks from fitted Gaussians 30

Figure 17: Intensity dependence of the three hydrogen subpeaks, in Nb, based on three FEL

shots 31

Figure 18: a: Intensity dependence of the kinetic energy for the three kinds of H+ ions in NbH

(green) and Nb (red). b: Intensity dependence of the kinetic energy for the H+, D

+ and D2

+ ions

in NbDH. 32

Figure 19: Intensity dependence of the three peaks (from left to right: H, C and O) of the

PMMA, based on three FEL shots 34

Figure 20: Simulated beam intensity dependence for Nb, at 21nm, with and without taking the

continuum lowering effect into account 36

Page 7: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

7

Figure 21: Top: simulated beam intensity dependence of the kinetic energy in PMMA. Bottom:

simulated dependence of the kinetic energy on the sample used, for fixed beam energy of

75!J. 37

Figure 22: Simulated ionization levels for Nb, at 21 nm, as calculated by Cretin, for a beam

energy of 20!J or 100!J 39

Figure 23: Dependence of the ion kinetic energy as a function of depth, for several beam

intensities, simulations performed on Nb. 39

Figure 24: Simulated electron temperature, average ionization and ion kinetic energy for Nb,

NbD, NbDH, NbH and PMMA, at 21nm, with a beam energy of 100!J. 40

Figure 25: Simulated absorption for Nb, NbD, NbDH, NbH and PMMA, at 21nm, with a beam

energy of 100 !J 41

Figure 26: Attenuation length for Nb and PMMA between 1 and 40 nm 42

Figure 27: Methane data for a backing pressure of 1.2 bar and 10 bar 43

Figure 28: Intensity depence of the H+ ions energy, for a backing pressure of 10 bar 44

Tables

Table 1: Cross section values for fusion reactions involving two hydrogen isotopes, at 10 keV

and 100 keV 12

Table 2: Ionization values for the ions relevant in the context of this experiment. 38

Page 8: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

8

Abbreviations

XFEL X-ray Free Electron Laser

FLASH Freie Elektronen Laser in Hamburg

DESY Deutsches Elektronen Synchrotron

MD Molecular Dynamics

LTE Local Thermodynamic Equilibrium

SASE Self Amplified Spontaneous Emission

LINAC Linear Accelerator

PMMA Polymethyl Methacrylate

TOF Time Of Flight

GMD Gas Monitor Detector

FWHM Full Width at Half Maximum

Page 9: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

9

1 Introduction

X-ray free electron lasers (XFELs) can reach much higher beam intensities at much

shorter wavelengths than usual lasers, and allow huge amounts of energy to be deposited

into a sample during a few femtoseconds. FLASH (“Freie Elektronen Laser in Hamburg”),

located at DESY (“Deutsches Elektronen Synchrotron”), in Hamburg, is the only operational

FEL reaching into the soft x-ray regime [1]. Hard X-ray FELs (down to 1.0Å predicted) are

being built in Stanford, California (LCLS – Linac Coherent Light Source) [2], and in Hamburg,

Germany (European XFEL) [3], and should be ready for operation in 2009 and 2013,

respectively. A third XFEL is being constructed in Japan (SCSS – Spring-8 Compact SASE

Source), and has reached saturation at 40nm wavelength during initial operations [4].

X-ray frequencies induce different processes in materials than optical frequencies, in

a laser-matter interaction. Inner shell ionization processes, such as photoemission and Auger

electron emission or fluorescence, dominate over outer shell processes; on the other hand,

optical nonlinearities, such as multiphoton absorption and free carrier absorption (inverse

Bremsstrahlung) lose part of their relevance, at x-ray frequencies [5]. Processes like Coulomb

Explosion are thus altered, and must be studied in this new context.

Figure 1: MD simulation of radiation-induced Coulomb explosion of a small protein molecule (lysozyme). White balls: H, Gray: C, Blue: N, Red: O, Yellow: S. Integrated X-ray intensity: 3.10

12(12keV) photons/100 nm diameter spot. The picture shows the protein exposed to a 2 fs FWHM

X-ray pulse, and its disintegration followed in time. The atomic positions in the first structures are practically identical at this pulse length due to an inertial delay in the explosion. Hydrogen ions and

highly ionized sulphurs are the first to escape the immediate vicinity of the protein. The encircled stage is shown in detail in Figure 2.

Figure 2: High-energy protons leaving the molecule very early during the explosion (encircled

stage from Figure 1). The energy of the fast escaping protons strongly depends on their original chemical environment. In the case of lysozyme, two protons shoot out at very high energies already at the beginning of all simulations (they are outside the picture frame in Figure 1). These protons are attached to two specific sulphur atoms in the native protein, and leave in the same direction in a reproducible manner.

Molecular Dynamics simulation of Coulomb explosion of lysozyme. Two high energy protons from the vicinity of specific sulphurs leave the molecule very early. XFEL: 3x1012 (12 keV) ph/100 nm diameter

Molecular Dynamics simulation of Coulomb explosion of lysozyme. Two high energy protons from the vicinity of specific sulphurs leave the molecule very early. XFEL: 3x1012 (12 keV) ph/100 nm diameter

Molecular Dynamics simulation of Coulomb explosion of lysozyme. Two high energy protons from the vicinity of specific sulphurs leave the molecule very early. XFEL: 3x1012 (12 keV) ph/100 nm diameter

Page 10: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

10

The use of XFELs has been proposed to achieve single-shot high-resolution images

of molecules [6], the final goal being to image macromolecules and living cells with atomic

resolution. Molecular dynamics (MD) simulations of the Coulomb explosion of a molecule [5,

7], presented in Figure 1, were made in the context of the imaging experiments. A sample is

hit by a 2 fs XFEL beam and loses its electrons. It then explodes, given the repulsive

Coulomb interaction between the remaining positive ions, but only after the end of the pulse,

i.e. an image taken during the illumination of the sample shows it unchanged. Figure 2

corresponds to the encircled stage in Figure 1, and shows the unexpected acceleration of

light ions, which follows the Coulomb explosion.

This observation set the possibility of accelerating protons or deuterons to KeV

energies, in a few femtoseconds, by means of the interaction of solid or clustered samples

with an XFEL beam. Deuterons accelerated to such energies are in good conditions to

undergo DD fusion reactions [5]. On a longer term, the aim will be to detect signatures of the

following reaction [8]:

D + D!3He + n , (1)

in the form of 2.45MeV neutrons and 0.82MeV alpha particles. The best candidates are

neutrons, neutral particles which can easily escape the interaction region, and which leave

with an energy characteristic of the reaction. Reaction (1) occurs along with a second one

D + D! T + p , (2)

with the same probability [8].

Intercluster fusion studies have been made, both theoretically [9] and experimentally

[9, 10] and so have intracluster fusion studies [11], on a theoretical level, based on the

interaction of ultra-intense optical/IR lasers with matter. However, in this case, the shift to x-

ray frequencies, keeping a high intensity and a pulse duration of a few femtoseconds, makes

the situation totally distinct, and opens the door to challenging new processes.

In the following sections, preliminary experiments are described, on the interaction of

an XFEL beam with solid samples and gas clusters. Very exciting results are presented and

analyzed, mainly in the case of the solids. Nevertheless, questions remain to be answered,

and improvements to be made, which are addressed later in this study.

Page 11: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

11

2 Theory

2.1 Fusion reactions

The aim of this study is to achieve fusion reactions, driven by the interaction between

an XFEL beam and a solid sample or a gas cluster. A brief summary of this kind of reactions

will therefore be given [8].

A nucleus of atomic number Z and mass number A has a mass whose value differs

from the sum of the masses of the protons and neutrons that constitute it by an amount

!m ,

given by:

!m = Zmp + (A " Z)mn "m , (3)

where

mp is the mass of the proton,

mn the mass of the neutron and

m the mass of the

nucleus.

!m is related to the nuclear binding energy B by

B = c2!m . (4)

The energy per nucleon,

B A , when plotted as a function of A, exhibits a maximum

for

A = 56. The existence of this maximum motivates fusion and fission reactions, since light

nuclei tend to fuse, and heavy nuclei to break, in order to reach a more stable configuration,

characterized by a higher value of

B A .

An important parameter in fusion reactions is the cross section

! , which has the

dimensions of an area, and measures the probability of occurrence of the reaction, per pair of

particles.

An average reactivity or reaction rate can be computed from the cross section

! and

the velocity

v of the incident nucleus (the target nuclei are assumed to be at rest), and is

given by:

!"v# = " (v) v f (v)dv0

$

% , (5)

where f is the velocity distribution of the incident particles.

Deuterium and tritium (two hydrogen isotopes) are generally used in fusion

experiments, instead of hydrogen, because of their higher cross-section at low temperatures.

Table 1 shows several cross section values for hydrogen isotopes fusion reactions, at

10KeV and 100KeV . Reaction b. is chosen for laboratory fusion experiments since its

cross section is higher than in reaction a., deuterium is easier to obtain than tritium, and both

products of the reaction are non-radioactive, and easy to detect (especially neutrons).

Page 12: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

12

Reaction !@10 keV

(barn)

!@100 keV

(barn)

!max

(barn)

!max

(KeV)

a.

p + p! D+ e+

+ "

3.6.10!26( )

4.4.10!25( )

b.

D+ D!3He + n

2.78.10!4

3.7.10!2

0.11 1750

c.

D+ D! T + p

2.81.10!4

3.3.10!2

0.096 1250

d.

D+ T!4He + n

2.72.10!2

3.43 5.0 64

e.

T + T!4He + 2n

7.90.10!4

3.4.10!2

0.16 1000

Table 1: Cross section values for fusion reactions involving two hydrogen isotopes, at 10 keV

and 100 keV. Also shown are the maximum value for the cross section, and the energy for which that

value is achieved. (Data from [8]) (

1barn =10!24cm

2)

Zweiback et al. [12] have derived the kinetic energy distribution from an exploding

cluster. Assuming that the cluster ionizes faster than it explodes, the potential energy of a

deuteron (charge e) on the surface of the cluster – which is the same as its kinetic energy

after the explosion – is given by Coulomb’s law:

E(r) =1

4!"0

eQcluster

r=

1

4!"0

1

re2 #q$n

4

3!r3 =

en#q$r2

3"0

,

i.e. Emax

=en!q"r

max

2

3#0

, (6)

where e is the elementary charge, Qcluster the charge of the cluster, n the particle density in

the cluster,

q the average charge of a particle,

rmax

the radius of the cluster, and

!0 is the

vacuum permittivity.

The aim is, therefore, to maximize n,

q and

rmax

. The use of an XFEL as a light

source (with short pulses) allows for a faster ionization of the sample, up to higher values,

which corresponds to an increase in

q , for given values of

rmax

and n; increasing

rmax

can

only be achieved by increasing the cluster size, while maintaining n, i.e. by optimizing the

parameters of cluster generation; the density n can be increased by compressing the clusters,

or by the use of solid samples. As described below, both approaches have been used, in the

experiments presented in this study, and will keep being used throughout the project. With

solid samples, high densities are attained much more easily than with gas clusters; on the

other hand, clusters are like “nano-labs” [13], which suffer very little from the influence of the

surroundings.

Ions located deeper inside the cluster will be characterized by a different potential

energy value, with a distribution such that

p(E)! E , for

E < Emax

, (7)

since p(r) = 4!r2n, r " r

max

Page 13: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

13

This distribution can be shifted towards

Emax

if heteroclusters are used, since heavier

atoms will serve as a catapult for lighter ones, which will have their kinetic energy increased.

[14]

2.2 Ionization mechanisms

The goal of the current work is to study the energies of ions ejected from a sample,

after its ionization by an XFEL beam. Some ionization mechanisms must therefore be referred

to, from a general point of view, before the discussion itself.

One of the main ionization mechanisms consists of the photoelectric effect, depicted

in Figure 3. An incident photon removes a bound electron from an atom. This electron will

leave with a kinetic energy

Ek corresponding to the difference between the incident photon

energy

E! and the ionization energy of the bound level the electron came from

En:

Ek= E! " En

= h# " En

, (8)

where h is the Planck constant and

! the photon frequency.

In the case of hard X-rays, a hole is left in a low-laying orbital in the electronic

structure of the ionized atom, which is therefore in an excited state. Two ways are possible, to

go back to the ground state. On the one hand, a higher energy electron can fall to the lower

empty energy level by emitting a photon, whose energy corresponds to the difference

between the two electronic levels, in a process known as fluorescence. On the other hand,

the energy lost by the electron that fills the vacancy can be transmitted to another electron,

known as Auger electron (Figure 3), which in turn leaves the atom. It will have a kinetic

energy

Ek

A corresponding to the difference between the decaying electron initial energy level

En and the Auger electron initial energy level

En

A.

Ek

A= E

n! E

n

A. (9)

If the ionization process is fast, as happens when one has energetic photons and

high beam intensities, a plasma is formed in the interaction region, where the electrons have,

at first, a higher temperature than the ions, and a lower mass. They will thus fly out first, and

leave behind a highly positively charged material. The positive charges will then repel each

other and be accelerated, in a so-called Coulomb explosion. When heavier and more highly

charged ions are also present, this repulsion can act like a catapult for light ions, such as

hydrogen and deuterium,

Page 14: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

14

Figure 3: Simple schemes of the photoabsorption (left) and Auger emission (right) processes.

During photoabsorption an incoming photon transfers its energy to a bound electron (from the inner

shell, in the case of hard x-rays), which leaves the atom. An electron coming from a higher orbital then

decays to fill the vacancy left by the photoelectron, and the gained energy can be used to remove an

electron – known as Auger electron – from an intermediate level (in the outer shell, in this case).

2.3 Simulation models

2.3.1 Molecular dynamics simulations

MD simulations have been performed, in an earlier stage, using the Gromacs code

[15].

Molecular dynamics aims at predicting macroscopic properties of systems, and follow

the kinetics of atomic and molecular processes, making use of realistic atomic models. More

generally, these simulations use as an initial input the potential of interaction, which depends

on atomic positions, and the positions r and velocities v of all atoms in the system. Different

forces are then computed: forces between non-bonded pairs (van der Waals and Coulomb

interactions), forces due to bonded interactions, and restraints (both for position and

distance). Finally, the atoms move according to Newton’s equations of motion, thus updating

the system configuration.

In spite of their accuracy in describing the “real” system, MD simulations generally do

not take into account complex processes such as collisional ionization, inverse

Bremsstrahlung, or continuum lowering, among others, which can be neglected in some

cases (for instance, inverse Bremsstrahlung at x-ray frequencies), but must be taken into

account for some energies and materials, when they are the more relevant.

Page 15: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

15

2.3.2 Non-LTE plasma simulations

Simulations were made using the code Cretin, developed at LLNL (Lawrence

Livermore National Laboratory), by Howard Scott [16]. Cretin is a radiation transfer plasma

code, which combines atomic kinetics and radiation transfer, and which can describe the

behavior of a non-LTE (non Local Thermodynamic Equilibrium) plasma. In each step it

updates the state of the system, based on initial input values of the density and temperature

of the plasma, and on the energy and intensity of the incident beam. The atomic populations

can be calculated either in a steady-state or a time-dependent mode.

Atomic kinetics is taken into account through a user-defined atomic model, and does

not include any radiation transfer or other non-local processes. A continuum-lowering model

is also available, which accounts for the effect of free electrons on the ionization potential of

the atom by shielding bound electronic levels, and therefore lowering the continuum level.

Radiation transfer is evaluated for continuum and line radiation, and couples back to atomic

kinetics. Spectral radiation is also calculated and allows the construction of detailed plasma

spectra. It is based on the plasma characteristics, but is not reused in the simulation.

This code can run in 0D to 3D. It includes a large number of switches and

parameters, which give it a wide versatility regarding the kind of processes to take into

account – such as continuum lowering, multiphoton ionization and time-dependence –, and

the specific values to assign to variables – such as the time step, the time between edits and

the minimum temperature – that are relevant for the description of the plasma or the

performance of the code.

Cretin can also be used as a postprocessor for a hydrodynamics code, using plasma

densities and temperatures as an input. One of its main applications lies in the study of ICF

(Inertial Confinement Fusion), where line emission from trace elements placed on the fusion

capsule is studied, in order to evaluate the extent of the hydrodynamics instabilities [17].

Cretin can also be used to study tokomak plasmas characteristics, in the context of MCF

(Magnetic Confinement Fusion), and to calculate the emission spectra of laser-produced

laboratory plasmas [17].

The code therefore corresponds to a static description, which does not include a

hydrodynamic expansion or a charge induced repulsion (leading to Coulomb explosion). In

our case, it was used to calculate the heating of the material – more specifically the electron

temperature – before the hydrodynamic motion occurs. During this time, the thermal energy of

the ions can be calculated from the electron-ion coupling coefficient. As opposed to what was

described in the examples above Cretin was thus used as a pre-processor rather than a post-

processor.

Combining these two approaches, or even trying to look at the problem through a

different perspective, with different simulation tools, can give a better insight of what is to be

expected from ongoing and future XFEL-based interaction experiments.

Page 16: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

16

3 Experiments: background information

3.1 XFEL: working principle

Free electron lasers (XFEL) differ from usual lasers by the lasing principle they use.

Instead of relying on atomic or molecular bond states to produce coherent radiation they use

electrons, accelerated to relativistic speeds, which radiate when going through an undulator,

with carefully chosen characteristics.

FLASH is a high gain XFEL, based on the “Self Amplified Spontaneous (or

Stimulated) Emission” (SASE) principle. In this case the undulator is five to ten times longer

than in a synchrotron and the electron beam has a lower emittance (i.e. a better collimation);

a good amplification can thus be obtained after a single passage through the undulator. The

problem of having no reflective mirrors available for short wavelengths is then solved, these

mirrors being necessary for creating an optical cavity resonator, when outside the SASE

regime [18].

In detail, an electron beam (where the electrons are randomly distributed inside a

gaussian shape) is first created by a laser-driven RF (radiofrequency) gun, and is then

accelerated to relativistic speeds, in a superconducting LINAC (“Linear Accelerator”). There,

at specific energies, the electron bunches are longitudinally compressed by magnetic fields,

as shown in in Figure 4.

Figure 4: Detailed scheme of the FLASH line, where the XFEL beam is produced. The main

features are the electron bunch generator (RF gun), the linear accelerator, the long undulator magnet,

and the final bending magnet, which separates the emitted radiation from the electron beam. (Source:

DESY, Hamburg) [19]

They then go into a long undulator, made of periodically spaced magnets, which produce

magnetic fields of alternating direction as can be seen in Figure 5. The Lorentz force

F!"= qv"!B!"(+ qE!") (10)

causes the (relativistic) electrons to wiggle – i.e. to be accelerated – and emit radiation in a

cone of

1

! N opening angle, illustrated in Figure 6, given the relativistic Doppler effect [19].

Page 17: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

17

Figure 5: Scheme of an undulator. The electron beam, accelerated to relativistic velocities, radiates

coherently when passing inside the undulator. At the end, a bending magnet deviates the electron

beam, and the emitted radiation alone goes into the experimental hall. (Source: DESY, Hamburg)

Figure 6: Radiation from a non-relativistic (left) and relativistic (right) accelerated electron. In the

first case the radiation is isotropic while in the second case it is emitted in a cone, because of the

relativistic Doppler effect.

The interaction between electron kinematics, which both creates and is affected by

the electromagnetic field, and the overall electromagnetic field, which has contributions from

the motion of the electrons and from the undulator magnets, results in a density modulation

and a microbunching of the electron beam (see Figure 7), with the same spatial period as the

undulator.

In this situation the electron microbunches have a longitudinal dimension shorter than

the wavelength of the radiation they emit, and can thus be described as a point-like electron

distribution, which radiates coherently, with a radiation power

P! given by [19]:

P! =N

2e

2

6"#0c

3! 4

˙ v 2, (11)

Page 18: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

18

where N is the number of electrons,

! the relativistic factor

! =1

1"# 2

=1

1" v c( )2

, and

˙ v the acceleration of the electrons.

This value is proportional to

N2, which is a large improvement compared to the case

of incoherent radiation from a spread bunch of electrons, where the radiated power scales

only with N [19]. The increase in the power radiated by the electron bunch, due to the

microbunching inside the undulator, is shown in Figure 7.

Figure 7: Microbunching process. The electron bunch experiences a spatial density modulation when

passing through the XFEL undulator. The power of the radiation emitted by the electrons increases

during the microbunching process (Source: DESY, Hamburg).

Finally, the electron beam is deflected into a dump (Figure 5), and the

electromagnetic radiation alone is directed to the beamlines, in the experimental hall. The

wavelength of the radiation depends on the electron beam energy and on the magnetic field

strength. An XFEL can therefore be tuned to a given frequency, depending on the application.

3.2 Experiments on solids

Experiments on solids were performed in September 2006, at FLASH.

The samples used are shown in Figure 8. They consisted of niobium rods, doped with

different amounts of hydrogen and deuterium, a tantalum plate doped with hydrogen, and a

PMMA (Polymethyl methacrylate, C5H8O2) plate. Niobium and tantalum samples were

prepared at the Material Physics group of the Physics Department, at Uppsala University, by

Gunnar Pálsson and Björgvin Hjörvarsson [20] (cf. Appendix A). The niobium samples have a

Page 19: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

19

concentration of dopants given by cH +D

= 0.86(mol!1) , with different relative amounts of

the two isotopes: cH= 0 and c

D= 0 ; c

H= c

D; c

H= 3c

D; c

D= 3c

H. The tantalum

sample has a concentration of H given by cH= 0.86(mol

!1) . The PMMA sample was

provided by the laboratory of Prof. Libor Juha, from Prague. All samples were kept in air

before entering the chamber, and therefore contain some extra amount of hydrogen, namely

in the case of “pure” Nb.

Figure 8: Picture of the sample holder, with the samples used in September 2006. The numbers

correspond to the order in which the samples were exposed to the FEL; A and B were not shot at.

In order to optimize the detection of emitted ions, the samples were mounted

perpendicularly to the detector. The sample mount was therefore bent by an angle of 35º ,

and rotated 20º away from the FEL beam direction, as shown in Figure 9.

Figure 9: Left: Picture of the inside of the chamber, with the sample holder mounted, although

not in the final position (cf. scheme on the right). Right: Scheme of the geometry of the

interaction. The TOF is mounted perpendicular to the samples, in order to maximize the detection

efficiency: the sample holder is bent by an angle of 35º relative to the vertical; the XFEL beam makes an

angle of 20º with the vertical plane, which contains the TOF direction.

Page 20: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

20

Both the chamber and the detector used for these experiments belong to Ryszard

Sobierajski’s team, in Warsaw. The detector consisted of an ion time-of-flight detector (TOF),

with the geometry presented in Figure 10.

Figure 10: Picture (left) and scheme (right) of the Warsaw TOF detector, used during the

September 2006 experiments at FLASH. For the experiments described here only the drift tube and

the MCP had some voltage applied.

The cones are used for ion acceleration, at the entrance of the TOF. They are

followed by a diaphragm and a tube diaphragm, used for beam focusing and direction

selection. The ions then travel inside a drift tube, at a constant potential, until they reach a

LED (“Low Energy Discriminator”), consisting of three golden grid electrodes, which cut out

ions that have a lower energy than the one applied to the grids. Finally, a microchannel plate

(MCP) detector, situated after the grids, records the signal.

In our case, neither the cones or the grids of the LED had any voltage applied to

them. As a consequence, no energy discrimination was possible for the ions reaching the

MCP detector.

A measure of the beam intensity could be obtained through the output of the GMD

(Gas Monitor Detector), located just before the XFEL shutter, at the entrance of the

experimental hall. This value had to be calibrated by an average value, obtained after

processing by the DOOCS software, at DESY.

As for the characteristics of the XFEL, during the experiment itself, a wavelength of

21nm , corresponding to 51.9eV photons, was used. The beam energy varied between 1

and 20µJ , and the pulse length between 10 and 20 fs . The focal spot had a 20µm

diameter (FWHM), i.e. a 314µm2 area. As a consequence, the beam intensity attained was

at the most 1014W cm

!2s!1

.

Page 21: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

21

The pressure measured in the chamber, during the experiment, was of about

10!6mbar = 10

!4Pa . For such vacuum conditions, there could be deposition of water on the

samples at a rate of approximately one monolayer per second [21].

The experiments were performed in single bunch mode, two consecutive shots being

about 2 s apart from each other. Each sample was shot at least 90 times, on at least 3

different spots. Only some of the available samples were used, namely Nb, NbD, NbH, NbDH

(relative concentration of 1, for H and D), TaH and PMMA. Data for TaH were later discarded,

since the output was mostly noise, and no clear peaks could be observed.

3.3 Experiments on gas clusters

Experiments on gas clusters were realized in March 2007, at DESY.

Two gases were used, for forming clusters: argon, in a first step, since rare gases

behavior is well known; and methane, the object under study, in a second step.

The chamber, the cluster generator, and the ion TOF detector were provided by the

cluster group at the Technischen Universität Berlin. The TOF has the geometry presented in

Figure 11.

Figure 11: Left: setup of the March 2007 gas experiments at FLASH; Right: scheme of the Berlin

TOF detector used. The XFEL and the cluster beam are parallel to the plates; the outgoing ions pass

through a small aperture in the grounded plate.

Page 22: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

22

The interaction between the XFEL and the clusters takes place between the two

plates. The ions then go through the energy discrimination grids, and finally reach the MCP

detector. In our case, no voltage was applied to the grids and therefore no energy

discrimination was possible for the ions.

Beam intensity measurements were made as for the solids experiments, by means of

the GMD located before the XFEL shutter of the experimental hall, whose signal was later

calibrated by other values, as explained for the case of the solids.

During these experiments, the XFEL was operating at a 13.5nm wavelength ( 92eV

photons). The pulse energy was on average 20µJ , but could go up to 40µJ , and its

duration was of about 10 fs ; the focusing achieved corresponded to an area of around

5µm2 (FWHM). The pulse intensity was therefore of about 4.10

16Wcm

!2s!1

.The pressure

inside the chamber was different from place to place: the minimum value was of about

10!5mbar , at the valve where the XFEL beam goes into the chamber; the highest was in the

interaction region, when gas clusters were injected in the chamber.

In order to study the influence of the gas backing pressure – the pressure in the gas

reservoir, before the cluster generating nozzle – in the cluster formation, it was set to both

300mbar and 7bar , in the case of Ar, and both 1bar and 10bar , in the case of methane.

A few thousand shots were recorded in each case, while the XFEL beam was operating in

multibunch mode, with a shooting frequency of 100Hz . Several hundreds of bunches were

sent every time, and data were recorded for the second one. At 10ms separation, it is

expected that all interactions are totally independent, i.e. the collected data should not have

any remnants from previous explosions.

Page 23: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

23

4 Results and analysis

4.1 Experiments on solids

Experimental data from solid samples clearly show the presence of light ions –

hydrogen, in all samples, and deuterium, in deuterated ones – leaving the interaction region

with very short times-of-flight. H+ ions always account for the most intense peak, together with

D+, but other species, with varying ionization levels, can also be observed and identified.

Typical TOF data for each solid sample are shown in Figure 12.

The first part of the analysis will deal with the identification of the observed ion peaks,

assuming constant energy among the species. In a second phase light ions will be studied in

more detail.

In the analysis only a part of the collected data were taken into account, in order to

create a homogeneous data set: three spots, shot 30 or 50 times, were kept for each sample.

The value of the intensity of the pulse, for each shot, was obtained from the GMD signal and

calibrated to the maximum value, for each spot. The data were then classified according to

their corresponding intensity, which made it possible to study the general evolution of the TOF

output with the pulse intensity.

Figure 12a: TOF spectra for NbH. Nb and NbH spectra present the same characteristics,

since Nb samples also contained hydrogen. A typical spectrum for Nb, with a split H+ peak, is

shown on the top of Figure 15. The H+ peak can also be split in NbH.

Page 24: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

24

Figure 12b: TOF spectra for NbDH. Typical spectra for NbD are shown on Figure 13 and on

the top of Figure 16. NbD and NbDH spectra present the same characteristics, since NbD

samples also contained hydrogen.

Figure 12c: TOF spectra for PMMA.

4.1.1 General analysis

The analysis of the experimental data regarding solid samples was made using a

program developed by Florian Burmeister et al. [22], based on IgorPro, a widely used data

analysis and signal processing software [23]. This program was first developed for TOF data

acquisition and analysis. For our purposes only a small part of it is required, where direct

output data from the TOF (intensity vs. time) are converted to intensity vs. mass/charge plot.

Page 25: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

25

One needs to assume constant energy among the ions – which makes sense in a Coulomb

explosion, and for this time scale ( ! 10

!5s ) – and to know the mass/charge ratio for two

peaks, as shown in Figure 13.

Figure 13: First step in the determination of the present ion species. We assume the nature of two

peaks, which serve as an input to the fitting formula.

The mass/charge ratio for the other peaks, and thus a good guess about their nature,

is then obtained, by fitting the data to the following formula [22]:

t = ! +"m

q, (14)

where

! and

! are fitting parameters.

Before the analysis the origin of the plots has to be fixed so that the photon peak,

which appears together with the ion peaks, corresponds to

t = 0. The detected photons are

emitted during the interaction and can be used as a triggering signal for the time

measurements, since their flight time is negligible.

This analysis method carries a drawback, since it requires at least three reference

points. Two of those points are used for the time to charge/mass conversion, and the third

one for testing the validity of the initial assumption. Therefore, plots with only one or two

peaks could not be analyzed (this is a problem for Nb, for instance).

Although this analysis is not limited to the H+ and D

+ ions alone, which constitute the

main targets of this project, it is relevant when trying to define the best environment for future

experiments: it helps determining conditions for acceleration of light ions, and finding out how

the output features vary with different experimental parameters (sample composition, beam

intensity, etc.).

Page 26: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

26

It is observed that, in general, pulse intensity does not appear to have a very big

influence on the level of ionization of the heavy ions, in the studied energy range. Increasing

the intensity leads to a higher ion yield rather than create new ionization states, i.e., the ratio

between the intensity of the H+ peak – the highest – and the others decreases with increasing

pulse intensity.

In Nb containing samples, ionized states of Nb up to 5+ can be found (see Figure

12b). A comparison of the samples shows that Nb3+

and Nb2+

ions are quite rare, when

compared to Nb4+

and Nb5+

ions.

Furthermore, species with a mass/charge ratio of 2 in deuterium containing samples

(NbD and NbDH) are D+, and not H2

+. Otherwise, they would also show up in the Nb and NbH

data, which they do not (Figure 12a vs. Figure 12b or Figure 16).

A species with a mass/charge ratio of 4 is also observed in deuterium containing

samples. In the absence of more experimental diagnostics we assumed it to be D2+, created

during the flight time by recombination of two D+ ions. However, it is surprising not to see a

more enhanced HD+ yield, with a mass/charge ratio of 3, given the similar recombination

energies for D2 and HD. Also, since there was no energy discrimination between the ions

(Figure 10), the whole analysis was made assuming constant energy: would the energy be

different from peak to peak, what appear to be ions with a mass/charge ratio of 4 could be H+

or D+ ions with a different energy. Drawing a conclusion on the nature of the ions requires

further diagnostics.

Apart from these features, some peaks remain unassigned, which could be due to

impurities, located on or inside the samples used, in the vacuum chamber itself, or even on

the sample holder (supposing some shot might have been misplaced). Given the

mass/charge ratio of the unassigned peaks, and the experiment’s environment, both Si and

Fe seem to be “good candidates” for impurities.

After their exposure to the FEL beam, the samples were analyzed using a Nomarski

differential interference contrast microscope. Figure 14 shows craters, both on PMMA and

NbDH, which correspond to the areas exposed to the FEL beam. The diameter of the craters

is about 50µm , which is consistent to the diameter of 20µm (FWHM) assumed for the

focal spot, since the damaged area is always bigger than the beam itself, and each spot in the

pictures was shot at least 30 times. No further crater analysis has yet been pursued on the

samples used in our experiments, partially because they were not polished enough;

nevertheless, studies on ablation and crater formation have been successfully performed at

FLASH [24], and could be later used as a model.

Page 27: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

27

Figure 14: Nomarski differential interference contrast microscope (x100) pictures taken from

PMMA (left) and NbDH (right) samples. Craters are visible in both cases, with a diameter of about 50

!m. This is consistent with the size of the FEL beam focal spot, which was assumed to have a diameter

of 20 !m (FWHM).

4.1.2 Light ions energy analysis

The main interest of the current project lies on the behavior of light ions, and a

detailed analysis was therefore made on the hydrogen and deuterium peaks. Given the TOF

dimensions presented in Figure 10, it is possible to convert time-of-flight measurements, t, to

energy, En, for a given ion species, characterized by its mass m and charge q, through the

following calibration:

t = 106

m 2

c

d1+ d

4+ d

5

1000E+

d2+ d

3

1000 E + qVDrift( )

!

"##

$

%&&+

2m

c

d61000 E + qVMCP( )

1000qVMCP

'

(

)))

*

+

,,,

, (15)

where En is in eV and t in seconds.

This allows one to follow the evolution of, mainly, the H+ or D

+ ions’ energy, and of the

corresponding peaks inner structure – when there is any – with the pulse intensity.

Having converted the time-of-flight of the ions to kinetic energy, the most exciting

result is the fact that there are H+ ions with energies up to hundreds of eV and even 1keV ,

in all samples. Results for the H+ peak in Nb are presented in Figure 15.

Page 28: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

28

Figure 15: Top: Direct TOF output for a single FEL shot on an Nb sample. Bottom:

Corresponding energy distribution for H+ ions (from three different origins), fitted by three

gaussians. Three different contributions can thus be identified, for the H+ yield, which might account for

different chemical origins or ejection mechanisms.

The hydrogen peak is often split into several superimposed subpeaks, as seen in

Figure 15, which might correspond to several chemical origins or ejection mechanisms of the

H+ ions. This phenomenon is observed in all samples, but easier to identify in deuterium free

ones, since no other peak is present in that region of the time-of-flight axis. A first approach in

the description and understanding of these peaks’ nature is to fit three gaussians to the

multiple peak, in order to isolate the two or three possibly different hydrogen contributions.

The gaussians are later converted to an energy dependency, using Equation 15. It is then

possible to identify the energy range characteristic of each H+ ion kind, which spreads from a

few hundreds of eV , for the lowest energy one, until up to 1keV , for the most energetic.

Page 29: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

29

For deuterium containing samples, four light ion peaks were converted to an energy

dependency, in the way described above for the H+ peak in an Nb sample. As seen in Figure

16, light ions – H+, D

+ and D2

+ - all have the same energy distribution, centered on a few

hundreds of eV , and going up to 1keV for H+ ions. The D

+ ions are generally limited to

lower energies, corresponding to the main contribution in the split hydrogen peak. However,

deuterium ions in the tail of the energy distribution can reach energies up to 1keV .

Nb+ ions seem to have a different energy distribution than the lighter ions, in Figure

16. This might be due to the fact that the Nb peak contains several ionization states, while

only one was considered for the current time-of-flight to energy conversion.

Figure 16: Top: Direct TOF output for a single FEL shot on an NbD sample, with a gaussian fit to

the H+, D

+, D2

+ and Nb

+ peaks. Bottom: Corresponding energy distribution of the H

+, D

+, D2

+ and

Nb+ peaks. All light species have the same energy distribution, centered on hundreds of eV, but

reaching up to 1 keV. Nb+ ions have an energy shifted towards higher energy when compared to light

ions, which might indicate that several ionization states of Nb are present.

Page 30: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

30

A beam intensity dependence analysis was also made, based on data from the GMD,

as described above. Figures 17 and 18 depict one of the main conclusions of the present

study: there is an energy increase for the three H+ ions kinds – corresponding to the three

subpeaks –, which follows the increase in pulse intensity. The energy increase is obvious in

Figure 17. As for Figure 18a, there is a diminishing of the time-of-flight (i.e. an increase in

kinetic energy) in all plots, which correspond to the three H+ ions kinds, both in Nb and NbDH.

Furthermore, Figure 17 illustrates the different behavior of the H+ ions kinds with the

increase in pulse intensity, namely that some of them only become relevant after a given

intensity threshold. The existence of an intensity threshold is confirmed by Figure 18a, in

which data for the second and third hydrogen peaks are only present after a given intensity.

The different behavior of the H+ ions kinds with the increase in pulse intensity supports the

hypothesis of a different chemical origin or ejection mechanism for the hydrogen ions.

Figure 17: Intensity dependence of the three hydrogen subpeaks, in Nb, based on three FEL

shots. The three kinds of H+ ions experience an increase in kinetic energy with increasing pulse

intensity. The energy increase is different for each kind of ion, suggesting several distinct ejection

mechanisms or chemical origins.

Page 31: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

31

Figure 18a: Intensity dependence of the time-of-flight for the three kinds of H+ ions (illustrated in

Figure 17) in NbH (green) and Nb (red). The ions energy increases (the tof decreases) with the pulse

intensity, for all kinds of H+ ions. Furthermore, there are thresholds in the intensity, corresponding to the

onset of the second and third H+ kind contribution.

Page 32: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

32

Figure 18b: Intensity dependence of the time-of-flight for the H+, D

+ and D2

+ ions in NbDH. An

increase in the kinetics energy (a decrease in the tof) of the ions follows the increase in beam intensity.

While Figures 15 to 17 correspond to single shot results, Figure 18 is built on data coming from many

shots. It therefore proves the consistency of the collected data among each other.

Figure 18b shows a similar evolution for the three independent peaks featured in the

NbDH sample, corresponding to H+, D

+, and D2

+. The three ions are accelerated with

increasing beam intensity, which can be deducted from the clear decrease – at least for H+

and D+ data – in the time-of-flight measurements.

Figure 19 shows data from a PMMA sample where the hydrogen, carbon and oxygen

peaks can be easily identified. It is clear that the beam intensity increase is followed by an

Page 33: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

33

energy increase, for all three kinds of ions. It must be noted that Nb+ ions in Nb-containing

samples also present an increase in energy for increasing beam intensity, although that is not

so striking in the plots, since the peak is always rather small.

Figure 19: Intensity dependence of the three peaks (from left to right: H, C and O) of the PMMA,

based on three FEL shots. The kinetic energy of all three species increases (the flight time diminishes)

with the beam intensity. Also, higher ionization states become more relevant, for higher beam intensity,

and contribute for a broadening of the carbon and oxygen peaks (since H+ cannot be ionized more).

Furthermore, higher ionization states become more relevant for higher beam

intensity: the highly ionized states change the shape of the carbon and oxygen peaks, as can

be seen in Figure 19. The hydrogen ions cannot be further ionized, thus their lower time-of-

flight, for higher intensities, is only due to a kinetic energy increase.

4.2 Simulations

The plasma simulation software package Cretin [16] was used for simulating the

initial plasma heating after an XFEL-solid interaction, which lasts for about 1ps and is

characterized by Te! Ti , where Te is the electron temperature and Ti the ion

temperature. Each simulation describes the 50 fs period right after the interaction, during

which the hydrodynamics expansion is governed by the electron thermal energy, due to a

high Te , and by the ions mass, which behaves like an inertial break. Given this information,

one can calculate the approximate expansion velocity of the system

vexp , assuming it to be

equal for all ions, by [25]:

Page 34: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

34

vexp =Z! K

BTe

M, (16)

where

! =3

5 is the adiabatic coefficient and

KB

the Boltzmann’s constant. M and Z are the

average mass and ionization of the ions, calculated by weighting the mass and ionization of

each atom by its stoichiometric coefficient in the material. This approach is no longer valid

once the plasma reaches an equilibrium between electrons and ions, about 1ps after the

interaction has occurred.

The expansion kinetic energy of the system can be calculated as well, making use of

the average ionization of the ions, which is an output of the simulation.

Eexp =1

2Mvexp

2=1

2Z! K

BTe

. (17)

This expansion energy is, however, characteristic of the system as a whole, and

cannot be applied to the individual ions. Very fast ions are therefore not modeled in this initial

thermal heating stage.

Both 0d and 1d simulations were made for the materials (Nb, NbD, NbH, NbDH,

PMMA) used during the solid-XFEL interaction experiments, in September 2006, at DESY.

The input parameters used were kept as close as possible to the experimental ones,

meaning a wavelength of 21nm (i.e. 59.1eV photons), a pulse duration of 15 fs , and a

focal area of 314µm2 (focal spot with D = 20µm ). The beam energy was varied from

1µJ to 100µJ , which corresponds to a variation in beam intensity between

2.12.1013W cm

!2s!1

and 2.12.1015W cm

!2s!1

. The aim here was not only to reproduce

the experimental conditions (1! 20µJ range), but also to study the effect of an increase in

the beam intensity. This increase was achieved increasing the beam energy ( 20 !100µJ

range), but it could also have been done by decreasing the focal spot area or the pulse

duration, given the formula below:

BeamIntensity =BeamPower

FocalArea=BeamEnergy PulseDuration

FocalArea. (18)

4.2.1 Interaction of the FEL pulse with solid samples

In a first approach, 0d simulations were made, which are expected to model the

heating at the surface of the sample, since no depth information is included.

The main conclusion of this calculation is the fact that the highest achieved energy is

around 0.1keV , for the Nb sample, at 100µJ beam energy, when continuum lowering is

taken into account, as can be seen in Figure 20. This energy corresponds to the final thermal

Page 35: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

35

energy of the ions, and is about an order of magnitude lower than the experimental values

described in the previous section.

Figure 20: Simulated beam intensity dependence for Nb, at 21nm, with (bottom) and without (top)

taking the continuum lowering effect into account. It is clear that an increase in the beam intensity

corresponds to an increase in the kinetic energy of the ions ejected from the sample. Furthermore,

taking the continuum lowering effect into account leads to higher kinetic energies of the ions, especially

in the early plasma heating phase.

Kinetic energy values are always higher for the case with continuum lowering, which

is as expected since the effect of the continuum lowering is to speed up ionization, and

therefore ions leave the sample sooner, and with larger kinetic energy. Figure 20 shows that

the simulations appear to be somehow unstable for some beam energies/intensities.

The top of Figure 21 shows the ion kinetic energy evolution with time, for all the

intensities considered, in the case of PMMA. It is clear that the ions kinetic energy increases

Page 36: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

36

with the beam intensity. This result can be very relevant and useful for later experiments,

although it only concerns the initial thermal expansion phase of the explosion process. Similar

plots, for the other samples, can be found in Figure B1, in Appendix B. They all present the

same kind of increase in the kinetic energy, following the increase of pulse intensity, which

shows that this behavior is independent of the material considered.

Figure 21: Top: Simulated beam intensity dependence of the kinetic energy in PMMA. Bottom:

Simulated dependence of the kinetic energy on the sample used, for fixed beam energy of 75!J.

Doing a similar kind of plot, but this time for a fixed intensity – 75µJ was chosen as

an example, and is shown in the bottom part of Figure 21 –, and for all materials, one can see

that the lowest ion kinetic energies are achieved for PMMA, then NbDH, NbD, NbH and finally

Nb. A complete set of results can be found in Figure B2, in Appendix B, which confirms the

generality of the conclusion. It is interesting to note that the NbD and NbH curves overlap (for

Page 37: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

37

higher intensities), as is to be expected given the similarity between hydrogen and deuterium

masses.

Ionization E (eV) 1st 2nd 3rd 4th 5th 6th

H 13.6

D 13.6

C 11.3 24.4 47.9 64.5 (392.1)

O 13.6 35.1 55.0 77.4 (113.9)

Nb 6.76 14.0 25.0 38.3 50.6 (102.2)

Table 2: Ionization values for the ions relevant in the context of this experiment. The incoming

photons have an energy of 59.1eV, meaning that ionization states with energies higher than that cannot

be achieved through direct photoionization (since multiphoton ionization and inverse Bremstrahlung

have a small effect, at these wavelengths, the ionization state is limited by the ionization potential and

the photon energy).

As can be seen in Table 2, Nb is ionizable up to 5 times, by 59.1eV photons. 0d

simulations were run for the Nb sample, at beam energies of 20µJ and 100µJ , to test

how much the ions would be ionized, in the initial plasma heating process. The results

obtained are shown in Figure 22.

The situation closest to the experiment is the 20µJ case, and there is a larger

fraction of Nb3+

, Nb4+

, and even Nb2+

, than of the others, whereas in the experiments Nb2+

and Nb3+

are rather rare. Nb2+

and Nb3+

could recombine during the flight time, to become

Nb+, more than Nb

4+ and Nb

5+ do. On the other hand, results for simulations made for a beam

energy of 100µJ are closest to the experimental ones, with more Nb4+

and Nb5+

than other

kinds of ions. The change is most likely due to the fact that a higher beam intensity enhances

higher ionization levels, as noticed in the experimental results. However, both simulations

show a small contribution of the Nb+ ion yield, which does not seem to be the case in the

experiments.

Page 38: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

38

Figure 22: Simulated ionization levels for Nb, at 21 nm, as calculated by Cretin, for a beam

energy of 20!J (top) or 100!J (bottom). Ionization states of Nb up to 5+ are formed during the

interaction, while the Nb atoms disappear almost entirely.

In a second stage results were obtained for 1d simulations that allow one to draw

conclusions on the depth evolution of the ion kinetic energy by running the simulation in

several zones, at different depths.

Comparing the results obtained for these simulations to the ones obtained above, for

0d simulations, one sees that the single zone considered in the 0d simulations corresponds to

the surface of the material, in 1d simulations. The results are even improved in the 1d case

since the instabilities in the output disappear, when continuum lowering is included.

Figure 23: Simulated dependence of the ion kinetic energy as a function of depth, for several

beam intensities, on Nb.

Page 39: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

39

As shown in Figure 23, the kinetic energy increase with the beam intensity,, for the

Nb sample, is not to restricted to the surface layer alone, rather it is common to all zones,

independently of how deep they are situated in the material. Furthermore, deeper lying zones

contribute with lower kinetic energy ions. In conclusion, higher beam intensity is translated

into energetic ions in deeper zones within the sample.

It is worth noticing, from Figure 24, that the PMMA contributes with lower energy ions

from zones close to the surface, but that this behavior changes when one goes deeper inside

the sample. Ion energies in niobium-containing samples are relevant until about 20nm inside

the sample, whereas in PMMA they stay relatively high until a depth of about 70nm .

Also shown in Figure 24 are the electron temperature, whose evolution with depth is

equivalent within Nb-containing samples, although rather different when compared to PMMA,

and the average ionization level of the ions, which differs even among Nb-containing

samples.

Page 40: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

40

Figure 24: Simulated electron temperature (top), average ionization (center) and ion kinetic

energy (bottom) for Nb, NbD, NbDH, NbH and PMMA, at 21nm, with a beam energy of 100!J. The

energy is higher at the surface for Nb-containing samples compared to PMMA. When going deeper

inside the material, the energy drops quickly for Nb samples, but stays high for a longer depth, for

PMMA.

The different behavior of PMMA and Nb-containing samples with increasing depth is

reflected in the absorption of the material, as can be seen in Figure 25.

Figure 25: Simulated absorption for Nb, NbD, NbDH, NbH and PMMA, at 21nm, with a beam

energy of 100 !J. The absorption presents the same behavior with respect to Nb-containing samples

and PMMA, as was described for the kinetic energy.

Comparing the results with data from the Center for x-ray optics of Lawrence

Berkeley National Labs [26], one sees that the characteristic depths in both kinds of samples

correspond to the attenuation length of the material at 21nm , which is 18nm for Nb – with

very little variation when H or D is added – and 75nm for PMMA. Figure 26 shows the

Page 41: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

41

attenuation length for Nb and PMMA over a range of wavelengths. To have a similar xray

penetration strength for both materials, a wavelength of about 12 nm would be required.

Figure 26: Attenuation length for Nb (top) and PMMA (bottom) between 1 and 40 nm. For our

experiments a wavelength of 21 nm was used, corresponding to an 18 nm attenuation length for Nb and

a 75 nm attenuation length for PMMA. [26]

4.3 Preliminary results on gas clusters

The near-solid density of gas clusters makes them a very interesting object to study.

One would expect clusters to have a behavior similar to that of solids, namely to present signs

of light ions acceleration, as was reported in [14].

Page 42: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

42

Figure 27: Methane data for a backing pressure of 1.2 bar (top) and 10 bar (bottom). Only the

higher backing pressure leads to cluster formation. There is no significant difference in the spectra

arising from the presence of clusters rather than gas.

Experimental results are presented in Figure 27, for two values of the backing

pressure. Only the higher backing pressure leads to cluster formation. Figure 27 shows no

significant difference between the gas and the cluster spectra, namely regarding the light ions

(H+ and H2

+) yield.

Page 43: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

43

Figure 28: Intensity dependence of the H+ ions energy, for a 10 bar backing pressure.

Figure 28 shows no decrease in the H+ ions time-of-flight, i.e. no increase in their

energy, for higher beam intensity. The result is surprising, since an ion energy increase was

observed to follow the beam intensity increase for all ions, in the case of solid samples,

Gas experiments results do not show any visible light ions energy increase due to the

presence of clusters or to an increase in the beam intensity, which would be expected given

the results obtained for solids, and those presented for clusters in the literature [14]. However,

the absence of fast ions in the spectra could be explained by a poor adaptation of the TOF to

energies of the order of 1KeV or higher, or by its geometry, namely the size of the slit the

ions have to go through to reach the MCP detector (shown in Figure 11).

Page 44: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

44

5 Discussion

The present study describes the first measurements of light ions acceleration during

an XFEL-solid interaction, performed at the only SASE-based XFEL available at this point,

FLASH.

During the interaction of the XFEL beam with solid samples, H+, D

+, and what was

assumed to be D2+ ions were detected, with energies of up to 1KeV . It was noticed that

there are three different contributions to the H+ ions yield, reflected in the spectra as a split

hydrogen peak. Furthermore, the energy of the ions was seen to increase for increasing

beam intensity. The energy increase occurs differently for different ions, and also for each

kind of H+ ions. The beam intensity increase also contributes to a larger yield of higher

charged ion species. Measurements of the methane clusters interaction with the XFEL beam

have shown no light ions acceleration compared to the case when methane gas was used.

No light ions energy increase was observed for increasing beam intensity.

Simulations made for solids, with the radiation transfer plasma code Cretin, estimate

the ion energy after the initial heating phase to be ten times lower than the experimental one.

The ion energy increase, following an increase in the beam intensity, is also modeled.

The ions energy increase with the beam intensity was predicted by MD simulations

[5]. Higher beam intensity means larger number of photons per unit time and per unit area on

the sample; this leads to an easier ionization of the samples, up to a higher ionization level. At

soft x-ray energies, photoelectric effect is one of the main ionization processes, which makes

the number of photons a very relevant parameter. Increasing the ionization efficiency can be

achieved by decreasing the focal spot area or the pulse duration, i.e. concentrating the

photons in time or space, or by increasing the beam energy, i.e. increasing the number of

photons in the beam.

We also measured that the energy distribution of the ejected protons has a complex

structure. This led to the hypothesis that these ions could have different origins, namely they

could come from different chemical environments, or be accelerated by different ejection

mechanisms. Several such mechanisms are discussed below.

As all the samples contain hydrogen, this constitutes the main origin for H+ ions

ejected after the interaction with the XFEL beam. The hydrogen in the samples is either

located in the interstitials of the lattice structure of the niobium, or as a part of the molecule

itself, for PMMA. The pure Nb sample, although not injected with hydrogen, still contains

some, absorbed in the sample from the atmosphere.

On the other hand, some water molecules might be present on the sample, not only

during the first shot in a given spot, but also in subsequent shots: the water layer could rebuild

to up to 2 monolayers between each shot, given the pressure conditions (10!6mbar ) and the

Page 45: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

45

2 s shooting time delay [21]. However, the sample heats up during the interaction, and it is

unclear at this point how much the water molecules contribute to the observed H+ signal.

With the aim of understanding the mechanisms responsible for the ejection of these

energetic ions, simulations were made with a plasma code, which calculates the thermal

contribution of the initial plasma heating phase for the final energy of the ions. This

contribution is found to be smaller than the values observed experimentally. In our

experiments, the faster hydrogen ions have kinetic energies of up to 1keV (Figure 16),

whereas the simulations report ions of 0.1keV at the most (Figure 20). Only 10% of the

energy of the ions can thus be explained by thermal heating in the initial phase of the

interaction. The remaining acceleration must be explained by other mechanisms, such as

Coulomb acceleration, in which the ions are further accelerated by charge repulsion from the

positively charged core of the sample.

Furthermore, one should notice the difference in the ionization states of Nb ions,

between the experiments and the simulations. A probable explanation could come from

processes such as ion recombination, which occur after the initial thermal heating described

by Cretin, and have an influence in the ionization state of the ions.

It is difficult to have a good estimate of the contribution of the Coulomb explosion,

through simulations. Cretin does not include this kind of mechanism, and MD simulations,

performed for hard x-rays interactions, might not be entirely accurate in this energy range.

Other simulation methods, namely particle-in-cell [11], would be required, in order to have a

good estimate of the Coulomb acceleration contribution to the final energy of the ions. This

would confirm whether this is the only effect that should be considered, or if other acceleration

mechanisms should be taken into account – such as space charge effects, as described by

Mora in [27] – which should be relevant in ionized materials.

Another considered hypothesis was a possible contribution of the water layer, present

at the surface of the samples, in the acceleration of the ions. The result of simulations

including few water zones on the surface of the solid sample, show no increase of the kinetic

energy of the ions due to the presence of the water.

The results from preliminary experiments performed on methane are surprising when

compared to data from the solids experiment or from the literature [14]. The incapacity of the

TOF to measure energies of the order of 1KeV , or the small size of the slit the ions have to

go through to reach the MCP detector (see Figure 11) could come as explanations for the

discrepancy among the results.

Page 46: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

46

6 Conclusions and outlook

This study represents the beginning of a larger project, whose aim is to analyze and

define conditions for achieving fusion through the interaction of an XFEL beam with both solid

samples and gas clusters. The first experiments performed at FLASH in this context have

yielded very exciting results, in respect to solid samples, through the observation of H+ – and

some D+ – ions with energies up to 1KeV .

We have observed considerable proton emissions in which the energy of the emitted

protons is very high (many hundreds of eV to KeV), namely higher than expected from

simulations made with an atomic physics code, Cretin. Furthermore, theoretical models

grounded on Molecular Dynamics predict different energies for protons coming from different

chemical environments (including protons originating from adsorbed waters or hydrogens on

the surface). We, therefore, expected to see split or multiple proton peaks in the TOF signals,

and they were indeed present in the results. The models also predict that the energy of

protons coming from different chemical environments will change differently with the FEL

pulse energy. We have observed this and have seen a larger proportion of faster protons at

higher beam intensities, in solid samples. Moreover, we expected to see more than one

mechanism behind proton ejection, and the gaussians fitted to the data from solid samples

confirmed such an expectation. Finally, methane cluster studies are an ongoing project,

where experimental details need to be worked out.

These promising results led to the allocation of dedicated beamtime for this project

during the next run at FLASH, starting in the end of 2007, and to the development of

collaborations, necessary to cover the broadness of the experiments and of the analysis

required. Several improvements are necessary, namely regarding the solid and gas samples

– chosen materials, amount of hydrogen and deuterium, polishing of the surface, in the case

of the solids –, the available diagnostics – type of detectors and their adaptation to the

experiment –, and the overall characteristics of the beam – focusing, wavelength, stability –,

some of which can, and will, be addressed in the next experiments. Apart from these

improvements in the experiments, one must both develop tools for analyzing the data, and

compare the outcoming results with simulation models (based on Molecular Dynamics or

particle-in-cell) that are better adapted to this particular matter.

Eventually, XFELs will start operating at hard x-rays wavelengths, opening new

possibilities for all kinds of research, among which fusion related projects.

Page 47: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

47

References

[1] http://flash.desy.de/

[2] http://www-ssrl.slac.stanford.edu/lcls/

[3] http://xfel.desy.de/

[4] http://www.spring8.or.jp/en/

[5] Carlson L.-A. (2006) On the feasibility of nuclear fusion experiments with XUV and X-

ray free-electron lasers – Master Thesis, at Uppsala University.

[6] Neutze R. et al. (2000) Potential for biomolecular imaging with femtosecond X-ray

pulses Nature 406 752-757

[7] X-ray induced Coulomb Explosion and nuclear fusion - experimental proposal II-

20060283 EC for beamtime at FLASH, accepted in January 2007

[8] Atzeni S., Meyer-Ter-Vehn J. (2004) The physics of inertial fusion – Beam plasma

interaction, hydrodynamics, hot dense matter, Clarendon Press – Oxford

[9] Ditmire T. et al. (1998) Explosion of atomic clusters heated by high-intensity

femtosecond laser pulses Physical Review A 57 369-382

[10] Ditmire T. et al. (1999) Nuclear fusion from explosion of femtosecond laser-heated

deuterium clusters Nature 398 489-492

Wabnitz H. et al. (2002) Multiple ionization of atom clusters by intense soft X-rays

from a free-electron laser Nature 420 482-485

[11] Peano F., Fonseca R. A., Silva L. O. (2005) Physical Review Letters 94 033401

Peano F., Fonseca R. A., Martins J. L., Silva L. O. (2006) Controlled shock shells and

intracluster fusion reactions in the explosion of large clusters Physical Review A 73 053202

[12] Zweiback J. et al. (2000) Nuclear fusion driven by Coulomb explosions of large

deuterium clusters Physical Review Letters 84 2634

[13] http://www.physik.tu-berlin.de/cluster, for the expression “nano-labs”

[14] Last I. and Jortner J. (2001) Nuclear fusion induced by Coulomb explosion of

heteronuclear clusters Physical Review Letters 87 033401

Grillon G. et al. (2002) Deuterium-Deuterium fusion dynamics in low-density

molecular cluster jets irradiated by intense ultrafast laser pulses Physical Review Letters 89

065005

Madison K. W. et al. (2004) Fusion neutron and ion emission from deuterium and

deuterated methane cluster plasmas Physics of Plasmas 11 270-277

[15] Berendsen H. J. C., van der Spoel D., van Drunen R. (1995) GROMACS: A

message-passing parallel molecular dynamics implementation Computer Physics

Communications 91 43-56

Lindhal E., Hess B. A., van der Spoel D. (2001) Gromacs 3.0: A package for

molecular simulation and trajectory analysis Journal of Molecular Modeling 7 306-317

Page 48: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

48

Van der Spoel et al. (2005) GROMACS: fast, flexible and free Journal of

Computational Chemistry 26 1701-1718

[16] Scott H. A. (2001) Cretin – a radiative transfer capability for laboratory plasmas

Journal of Quantitative Spectroscopy and Radiative Transfer 71 689-701

[17] Langer S. H. et al. (2003) Comparisons of line emission from 2- and 3-dimensional

simulations of ICF capsules to experiments Journal of Quantitative Spectroscopy and

Radiative Transfer 81 275-286

[18] Castro P. et al (2000) Overview of SASE experiments XX International Linac

Conference, Monterey, California

[19] Rossbach J. (2004) Linac based free-electron laser – TESLA-FEL Report 2004-08

[20] Pálsson G. K. (2006) Reportlet on hydrogen concentration determination of NbHx and

TaHx bulk samples

[21] http://en.wikipedia.org/wiki/Monolayer

[22] Burmeister F. et al. (2003) Using Igor Pro as the user interface together with a FAST

TDC card for PEPIPICO data acquisition: manual and overview

http://www.fastcomtec.com/applications/mass-spectrometry.html

[23] www.wavemetrics.com

[24] Hau-Riege S. P. et al. (2007) Damage threshold of inorganic solids under free-

electron laser irradiation at 32.5 nm wavelength Applied Physics Letters 90 173128

[25] Attwood D. (2000) Soft x-rays and extreme ultraviolet radiation - Principles and

applications Cambridge University Press

[26] http://www-cxro.lbl.gov/index.php?content=/tools.html

[27] Mora P. (2003) Plasma expansion into a vacuum Physical Review Letters 90 185002

Mora P. (2005) Thin-foil expansion into a vacuum Physical Review E 72 056401

Page 49: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

49

Appendix A

Nb and Ta containing samples used during the September experiment at FLASH

were analyzed at the Material Physics group of the Physics Department, at Uppsala

University, using High Resolution X-Ray Diffraction (HRXRD) [20]. This method, combined

with precision density measurements, gives the concentration of both H and D isotopes in the

sample. HRXRD results consist of shifts in Bragg reflection peaks, which can be translated

into a measure of the variation of the lattice parameter

!a , and, therefore, of the volume of

expansion of the sample, through

!V

V=a0

3" (a

0+ !a)

3

a0

3. (12)

Density measurements determine the specific volume change

!v

". The H (or D)

concentration

cH

(or

cD

) then follows:

!V

V= c

H

!v

". (13)

Page 50: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

50

Appendix B

Detailed simulation plots showing the dependence of ion kinetic energy with sample

composition and beam intensity are presented below. All of them assume continuum

lowering.

Figure B1 shows the dependence of the ion kinetic energy time evolution with the

beam intensity, for all samples. The kinetic energy increases with the pulse intensity, for all

materials.

Figure B1: Variation of the ion kinetic energy with the beam intensity, for each of the five

samples (Nb, NbD, NbDH, NbH, PMMA, from left to right, and top to bottom).

Page 51: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

51

Figure B2 shows the dependence of the ion kinetic energy time evolution with the

sample, for all the intensities considered in the simulation. The lowest ion kinetic energies are

achieved for PMMA, then NbDH, NbD, NbH and finally Nb, for all intensities.

It is interesting to note that the NbD and NbH curves sometimes overlap (for higher

intensities), which makes sense given the similarity between hydrogen and deuterium

masses.

Page 52: Coulomb acceleration of protons with a free-electron laser · X-ray free electron lasers (XFELs) can reach much higher beam intensities at much shorter wavelengths than usual lasers,

52

Figure B2: Variation of the ion kinetic energy with the sample, for each of the ten beam

intensities considered (1 !J, 2 !J, 5 !J, 10 !J, 15 !J, 20 !J, 25 !J, 50 !J, 75 !J, 100 !J, from left to

right, and top to bottom).