179
Investigations of counter propagating laser produced plasmas in a collision free system in the presence of a strong external magnetic field. Volume 1 of 1 Robert Alan David Grundy Doctor of Philosophy The University of York Department of Physics September 2003

Investigations of counter propagating laser produced ...phys35/documents/thesis/Rob Grundy/RADG... · Figure 6.1 Production of a magnetic field through use of a laser target. 142

  • Upload
    lyxuyen

  • View
    213

  • Download
    1

Embed Size (px)

Citation preview

Investigations of counter propagating laser

produced plasmas in a collision free system in the

presence of a strong external magnetic field.

Volume 1 of 1

Robert Alan David Grundy

Doctor of Philosophy

The University of York

Department of Physics

September 2003

2

2

Abstract

This thesis presents a beginning of an investigation into the formation of a collisionless

shock through the interaction of two rapidly expanding laser-produced plasmas

immersed in a strong magnetic field. The ultimate aim is to produce an experimental

simulation that can be scaled to be relevant to a 100-year-old Supernova remnant. This

requires the reproducible production of plasmas with low densities (~1018cm-3) and high

expansion velocities (~107cm/s) in a counter propagating geometry.

The chosen method of forming this simulation is the production of plasma by the direct

drive laser irradiation at 1014W/cm2 of 100nm thick solid targets. The experiments are

diagnosed primarily by optical probing techniques, the principal theme of this thesis.

These results show evolution of the plasma is in good agreement with both analytical

and numerical models.

However, the plasma produced by direct drive irradiation was found to be structured.

The observed structure is primarily a result of laser imprint, and experimental

techniques have been developed to overcome this using a spatially filtered pre-pulse.

With the ability to produce uniform, well characterised plasma we are able to

demonstrate that immersing thin foil plasma in a strong (~10T) magnetic field does not

affect the hydrodynamics of the plasma. When produced in a counter-propagating

geometry, two thin foil plasmas can be shown to interpenetrate in a collisionless

manner. Immersing this system in a strong magnetic field alters the interaction. Instead

of interpenetrating, the plasmas behave similarly to a collision-dominated counter-

propagating system.

The mechanism by which this occurs is still uncertain. However, if this change in

system behaviour is caused by the magnetic field penetrating into that plasma and

localising the particles on a scale smaller than that of binary collisions, then it may be

feasible to produce a scaled experimental simulation of a Supernova remnant using this

technique.

3

3

List of contents

ABSTRACT.................................................................................................................................. 2

LIST OF CONTENTS .................................................................................................................. 3

LIST OF ILLUSTRATIONS........................................................................................................ 8

ACKNOWLEDGEMENTS........................................................................................................ 13

AUTHOR’S DECLARATION................................................................................................... 15

CHAPTER 1 - INTRODUCTION.............................................................................................. 16

1.1 MOTIVATION.................................................................................................................. 16

1.2 BACKGROUND................................................................................................................ 17

1.2.1 Laboratory Astrophysics ....................................................................................... 17

1.2.2 Thin foil laser produced plasmas .......................................................................... 19

1.2.3 Colliding plasma experiments............................................................................... 20

1.3 CHAPTER OUTLINE......................................................................................................... 22

CHAPTER 2 - THEORY............................................................................................................ 24

2.1 INTRODUCTION .............................................................................................................. 24

2.2 PLASMA PARAMETERS ................................................................................................... 25

2.2.1 Collective effects ................................................................................................... 25

2.2.2 Single Particle Motions......................................................................................... 27

2.2.3 Plasma Oscillations .............................................................................................. 31

2.3 COLLISIONALITY............................................................................................................ 32

2.4 COLLISIONLESS SHOCKS ................................................................................................ 35

2.4.1 Shock formation .................................................................................................... 35

4

4

2.4.2 Shock conditions.................................................................................................... 36

2.4.3 Shock thickness...................................................................................................... 38

2.5 SCALING......................................................................................................................... 40

2.6 LASER-PLASMA INTERACTIONS...................................................................................... 44

2.7 CONCLUSIONS ................................................................................................................ 50

CHAPTER 3 - OPTICAL PROBES. .......................................................................................... 52

3.1 INTRODUCTION .............................................................................................................. 52

3.2 THEORY OF PROPAGATION OF LIGHT IN AN UNDER-DENSE PLASMA.............................. 53

3.2.1 The refractive index of a plasma ........................................................................... 54

3.2.2 Propagation of light in geometrical optics approximation ................................... 57

3.3 THE VULCAN LASER SYSTEM ........................................................................................ 58

3.4 PROBE BEAMS ................................................................................................................ 59

3.4.1 Utilised probe beam designs ................................................................................. 59

3.4.2 Temporal resolution requirements ........................................................................ 62

3.4.3 Wavelength selection............................................................................................. 64

3.4.4 Intensity selection.................................................................................................. 65

3.4.5 Wavefront quality .................................................................................................. 66

3.5 IMAGING SYSTEMS......................................................................................................... 67

3.5.1 Resolution requirements........................................................................................ 69

3.5.2 Diffraction around target, diffraction limit........................................................... 71

3.5.3 Imaging for Interferometry.................................................................................... 73

3.5.4 Limitations due to refraction................................................................................. 75

3.6 CONCLUSIONS ................................................................................................................ 75

5

5

CHAPTER 4 - OPTICAL PROBE DIAGNOSTICS.................................................................. 76

4.1 INTRODUCTION .............................................................................................................. 76

4.2 INTERFEROMETRY.......................................................................................................... 76

4.2.1 Theory ................................................................................................................... 76

4.2.2 Interferometer Designs.......................................................................................... 79

4.2.3 Interferometery Analysis Technique...................................................................... 81

4.2.4 Abel inversion........................................................................................................ 85

4.3 SHADOWGRAPHY AND SCHLIEREN IMAGING................................................................. 87

4.3.1 Theory ................................................................................................................... 87

4.3.2 Schlieren Designs.................................................................................................. 88

4.3.3 Ray tracing analysis .............................................................................................. 89

4.3.4 Code design........................................................................................................... 90

4.3.5 Benchmarking the code. ........................................................................................ 91

4.3.6 Producing simulated schlieren images. ................................................................ 97

4.3.7 Analysis of the ray tracing .................................................................................... 98

4.4 POLARIMETRY.............................................................................................................. 100

4.4.1 Theory ................................................................................................................. 100

4.4.2 Polarimeter Designs............................................................................................ 104

4.4.3 Polarisation analysis code .................................................................................. 105

4.5 CONCLUSIONS .............................................................................................................. 108

CHAPTER 5 - PRODUCING PLASMA FROM A SINGLE THIN FOIL.............................. 110

5.1 INTRODUCTION ............................................................................................................ 110

6

6

5.2 EXPERIMENTAL TECHNIQUE ........................................................................................ 111

5.2.1 Introduction......................................................................................................... 111

5.2.2 Target Design and Manufacture ......................................................................... 111

5.2.3 Experimental Setup ............................................................................................. 112

5.3 RESULTS....................................................................................................................... 115

5.3.1 Expansion............................................................................................................ 115

5.3.2 Density Profile .................................................................................................... 117

5.3.3 Plasma non-uniformity........................................................................................ 120

5.3.4 Laser focussing conditions .................................................................................. 123

5.3.5 Target characterisation....................................................................................... 125

5.4 DISCUSSION.................................................................................................................. 128

5.4.1 Comparison of Expansion with models ............................................................... 128

5.4.2 Plasma non-uniformity........................................................................................ 130

5.4.2.1 Target structure ............................................................................................... 130

5.4.2.2 Laser focal spot intensity structure ................................................................. 132

5.4.3 Thermal smoothing.............................................................................................. 133

5.5 PLASMA SMOOTHING EXPERIMENT.............................................................................. 135

5.5.1 Experimental set up............................................................................................. 136

5.5.2 Results ................................................................................................................. 137

5.5.3 Discussion ........................................................................................................... 139

5.6 CONCLUSIONS .............................................................................................................. 139

CHAPTER 6 - COLLIDING MAGNETISED FOILS ............................................................. 140

6.1 INTRODUCTION ............................................................................................................ 140

7

7

6.2 PRODUCTION OF A MAGNETIC FIELD............................................................................ 140

6.2.1 Helmholtz coil laser target.................................................................................. 141

6.2.2 Pulsed power electromagnet ............................................................................... 146

6.3 SINGLE FOIL EXPANSION IN A MAGNETIC FIELD........................................................... 148

6.4 COUNTER PROPAGATING EXPLODING FOIL PLASMAS .................................................. 151

6.4.1 Counter propagating CH foils............................................................................. 152

6.4.2 Counter propagating Al foils............................................................................... 154

6.5 MAGNETISED COUNTER PROPAGATING PLASMAS........................................................ 156

6.6 DISCUSSION.................................................................................................................. 159

6.7 CONCLUSIONS .............................................................................................................. 164

CHAPTER 7 - CONCLUSIONS .............................................................................................. 166

7.1 CONCLUSIONS .............................................................................................................. 166

7.2 FURTHER WORK........................................................................................................... 168

REFERENCES ......................................................................................................................... 170

8

8

List of illustrations

Figure 2.1 Diagram of the temperature and density scales of different plasmas 44

Figure 2.2 Med103 simulations of the proposed colliding foils scheme 49

Figure 2.3 Scaling parameters for a 100 year old SNR. 50

Figure 3.1 Processes that occur during propagation of a laser through a plasma 53

Figure 3.2 Probe front-end layouts for each experiment 60

Figure 3.3 Schematic of a grating pair arranged for pulse compression 62

Figure 3.4 The effect of a mismatched gratings 63

Figure 3.5 Probe beam pulse shapes 64

Figure 3.6 Near field intensity pattern of the probe from the second experiment 66

Figure 3.7 First experimental imaging system layout 67

Figure 3.8 Second experimental imaging system 68

Figure 3.9 Third experimental imaging system. 69

Figure 3.10 Modulation Transfer functions for the three imaging 70

Figure 3.11 The comparison between theoretical and observed 72

Figure 3.12 OPD plots for the three imaging systems 74

Figure 4.1 Simulated interference pattern 78

Figure 4.2 Principle of a Wollaston prism based interferometer. 80

9

9

Figure 4.3 Schematic of an imaging Wollaston prism based interferometer 80

Figure 4.4 Principle of the vertical shearing interferometer 81

Figure 4.5 Rotating an interferogram to allow fringe extraction 82

Figure 4.6 The process of fringe extraction to produce a phase difference plot 83

Figure 4.7 Sample interferogram and extraction of the electron density 84

Figure 4.8 Broadband emission from a 200µm diameter Al dot 86

Figure 4.9 Contour plots of electron density data 86

Figure 4.10 Schlieren imaging schematic 88

Figure 4.11 Schlieren imaging arrangement for second experiment 89

Figure 4.12 Geometry of the ray tracing code 90

Figure 4.13 Spherical plasma geometry and radial electron density 92

Figure 4.14 Values of the calculated minimum radius compared with simulation 93

Figure 4.15 Variation of the impact parameter B with respect to the initial 94

Figure 4.16 Results of the modified ray tracing 96

Figure 4.17 Variation of the impact parameter for comparison Figure 4.15 97

Figure 4.18 Back propagation of rays onto a CCD at the assumed object plane 98

Figure 4.19 Forming a simulated schlieren image 99

Figure 4.20 Ray traced image overlaid on experimental schlieren data 99

Figure 4.21 Wollaston prism based polarisation analyser. 105

Figure 4.22 Statistical error in the calculation 107

Figure 4.23 Polarimeter testing for various input polarisation angles. 107

10

10

Figure 5.1 Design of the target holder used in all experiments 112

Figure 5.2 Generic plan of the target chamber layout for single foil experiments. 113

Figure 5.3 Focal spot imaging 115

Figure 5.4 Graph of plasma expansion, 116

Figure 5.5 Interferometric images of 100nm thick CH foils 118

Figure 5.6 Abel inverted electron density profiles 119

Figure 5.7 Schlieren images from the first experiment 121

Figure 5.8 Schlieren images taken during the second experiment. 121

Figure 5.9 Detail of high resolution schlieren 122

Figure 5.10 X-ray pinhole camera images 122

Figure 5.11 Normalised focal spot 123

Figure 5.12 2D FFT Power spectra 124

Figure 5.13 Electron micrographs of a CH foil and a Al foil on a Mylar holder 125

Figure 5.14 Electron micrographs of a CH foil and a Al foil on a Cu holder 126

Figure 5.15 High magnification electron micrograph of a CH foil a Cu holder 126

Figure 5.16 Laser interferogram of a CH foil mounted on a Cu holder 127

Figure 5.17 Graph of the analytical model interferometry measurements 128

Figure 5.18 Graph of Medusa simulation compared with experimental data 129

Figure 5.19 Schlieren Image lineouts across the schlieren images in Figure 5.9 131

Figure 5.20 Vertical lineouts across the schlieren images in Figure 5.7 132

11

11

Figure 5.21 Vertical Lineouts across the schlieren images in Figure 5.8 133

Figure 5.22 Experimental chamber layout for pre-pulse experiments 136

Figure 5.23 Evidence of pre-pulse smoothing 138

Figure 6.1 Production of a magnetic field through use of a laser target. 142

Figure 6.2 Photograph of the Helmholtz coil target geometry 143

Figure 6.3 Single turn search coil results from our first experiment 144

Figure 6.4 Interferogram of two foils and a Helmholtz coil taken 145

Figure 6.5 Schlieren data taken 500ps after foil irradiation. 145

Figure 6.6 The construction of the pulsed electromagnet 146

Figure 6.7 Polarogram of a glass slide inside the pulsed electromagnet 147

Figure 6.8 Evolution of the magnetic field in the electromagnet 148

Figure 6.9 Experimental chamber plan for magnetised single foil experiments. 149

Figure 6.10 Interferograms taken 750ps after target 150

Figure 6.11 Abel inverted thin CH foil electron density profiles 150

Figure 6.12 Experimental layout for counter-propagating thin foil experiments 152

Figure 6.13 Interferogram of the interaction of two counter propagating 153

Figure 6.14 Expansion of two CH foils compared to two single foils 153

Figure 6.15 Schlieren data with the ray-traced simulation overlaid 154

Figure 6.16 Interferogram of two Al foils taken 750ps after irradiation. 155

Figure 6.17 Expansion of two Al foils compared to two single foils 155

12

12

Figure 6.18 Arrangement for counter propagating experiments in a magnetic field 156

Figure 6.19 Interferogram of two CH foils immersed in a 7.5T magnetic field 157

Figure 6.20 Comparison of magnetised and non-magnetised colliding foil data 158

Figure 6.21 Interferometry reference channel image 158

Figure 6.22 Shadowgraph from a Helmholtz experiment 159

Figure 6.23 Comparison of Medusa simulation with the experimental profile 161

Figure 6.24 The expansion of two magnetised CH foils and two Al foils. 162

Figure 6.25 Abel inverted radial density gradient profile 164

13

13

Acknowledgements

Firstly, I would like to express my sincere thanks to Dr Nigel Woolsey for his guidance

and supervision of this thesis. I would also like to thank Professor Greg Tallents for his

supervision during the first years of my work. I am deeply grateful to Dr Cedric

Courtois, Dr Dave Chambers and Andrew Ash with whom it has been my great pleasure

to work with.

Special thanks also go to Yousef Abou-Ali and Dr Jalal Pechtehe for their contribution

to our experiments and companionship in our office and to Dr Stephen Tear for his

assistance and use of the Scanning Electron Microscope.

The experimental work would not have been possible without the help and goodwill of

the staff of the VULCAN laser at the Rutherford Appleton Laboratory, my thanks to

you all. In particular I would like to thank Margaret Notley, Rob Heathcote and Rob

Clarke for their assistance and advice when performing experiments. I would also thank

Dr John Collier for his assistance with designing optics and use of the Zeemax design

package.

I would also express my thanks to Dr Richard Dendy, Dr Per Hellander and Dr Ken

McClemments of the Culham Science Centre for their collaborative work with our

group and useful discussions. In addition, I would also express my thanks the Dr Paddy

Carrolan and Dr Neil Conway for their experimental collaboration and advice on

Zeeman splitting diagnostics.

For their collaboration in our final experiment, I would also like to thank Ben Lings and

Katarina Rosol’ankova of the University of Oxford.

Finally, I am grateful for the support and encouragement provided by my family and

friends, especially my partner Amber. Without you, this would not have been possible.

14

14

In memory of my father.

15

15

Author’s declaration

The work presented in this thesis is the product of the work of many people, the nature

of the laser-plasma experiments presented is such that they require a collaborative

research group using large-scale facilities. This not only includes the members of the

experimental teams, but also the facility staff and our collaborators.

Here I will outline the contribution I have made within the group. My primary role has

been that of an experimentalist, contributing to the planning and execution of our

experimental work. In the main part, this has been through the development and

utilisation of optical probing techniques and the associated diagnostics. I have carried

out all of the analysis of the data collected by these diagnostics and assisted in its

interpretation in conjunction with other diagnostics used. To this end, I have also been

responsible for the development of an iterative ray tracing code to simulate the optical

probe, and codes to simulate and analyse the diagnostic data. I have also carried out

investigations of the thermal smoothing techniques in thin foil plasma and of the

performance of our optical probe imaging.

The plasma simulation work quoted throughout the thesis (unless otherwise stated) is

the work of my supervisor, Dr Woolsey. The focal spot profile measurements in

Chapter 5 and the magnetic field calculations and Faraday cups measurements quoted in

Chapter 6 are the work of Dr Woolsey and Dr Courtois.

16

16

1 Chapter 1 - Introduction

1.1 Motivation

Astrophysical research has been traditionally based in theoretical modelling and

observational data. Traditionally the contribution of experimental physics to this

research has been in the main part through measuring the fundamental parameters of

atomic and molecular physics.

Developments in high-energy density physics experiments1 now enable us to produce

conditions in the laboratory that are relevant to astrophysical systems. This permits us to

perform laboratory astrophysics experiments to provide an accurate basis for theoretical

modelling such as equation of state2 and radiation transfer measurements3. These

provide the ‘inputs’ for astrophysical modelling. This also introduces the possibility of

producing a scaled4 reproduction of an astrophysical system in the laboratory,5 which

can be used to test the predictions of theories or the ‘outputs’ of astrophysical

modelling, and be compared with observational data.

Many different types of plasma production technique have been used, from wire-array

z-pinches6 to high power lasers1,7, covering a range of astrophysical systems from jet

formation6,8,9 to supernova remnants10.

One of the current unanswered questions in astrophysics is the source and acceleration

of cosmic rays. The observed spectra of cosmic rays may be produced by the

combination of two separate sources11. It is proposed that the production of cosmic rays

above an energy of 1019eV may be the result of acceleration by intergalactic magnetic

recombination events12. This type of system has been studied experimentally13 using

two spheromaks. These experiments have shown particles being significantly

accelerated by magnetic recombination.

Cosmic rays of energy lower than 1015 eV are widely assumed to by produced by

diffusive shock acceleration (DSA)14 across the shock front15 of a supernova16.

Characteristic X-ray synchrotron radiation of 1014eV fitting the theoretical spectra for

DSA has been observed17 in the vicinity of the supernova remnant SN1006. However,

the mechanism by which particles are ‘injected’ into the process of DSA is

17

17

unknown18,19. It has been proposed that the collisionless shock20 system of such a

supernova remnant required to simulate this phenomenon is within the reaches of

experimental possibility21.

The aim of this research reported here is to develop the capability to perform an

experimental simulation of a 100 year old supernova remnant, by use of a high power

laser. The laser is used to irradiate two thin foil plasmas in an opposing geometry

immersed in a strong magnetic field.

1.2 Background

The scope of this thesis primarily covers the production of plasma from thin foils in a

counter-propagating geometry and in the presence of a magnetic field. The setting for

this work is the field of laboratory astrophysics, which has already been touched upon in

the previous section. In this section, I will place the work presented in this thesis in the

context of the previous work that has been performed in these related fields. To this end,

this section is divided into three sections: laboratory astrophysical research relevant to

the production of a simulation of a 100 year old supernova remnant, laser-plasma

research into production of thin foil laser-plasmas, and research into the interactions of

plasmas in a colliding geometry.

1.2.1 Laboratory Astrophysics

The field of laboratory astrophysics using intense lasers has been reviewed by Rose7,

Ripin et al22, Remington et al23 and Takabe et al24. The specific field of collisionless

laboratory astrophysics has also been reviewed by Zakharov25.

The ultimate aim of the research project is to perform an experimental simulation of a

100 year old supernova remnant (SNR). This differs from previous scaled supernova

experiments reviewed by Drake26, since we are attempting only to produce simulation

snapshot of the SNR as opposed to modelling the evolution of the system (see Chapter

2, section 2.5). Previous experiments by Drake et al27,28 have modelled a young

supernova remnant, SN1987A, where the explosion of the supernova will collide with a

circumstellar ring. The supernova is simulated by the explosion by indirect drive (using

a gold hohlraum) of a plastic ‘plug’. The plasma formed by this method is allowed to

expand through a vacuum until it collides with a foam target, which models the

18

18

circumstellar ring. Experiments by Kane et al29 have also studied hydrodynamic

instabilities that play a critical role in the evolution of a core collapse supernovae, such

as SN1987A. In this stage of supernova remnant evolution the shock produced is

radiative, a system which has been studied by Shigemori et al30.

After the initial explosion of the supernova, the shock wave expands into the interstellar

medium (ISM)31. The interaction of the SNR shock with interstellar clouds has been

investigated by Klein et al32. The expansion of the SNR into the ISM is actually

comprised of two shocks16, the forward shock of the explosion and a reverse shock

travelling backwards through the ejecta. Although the reverse shock is propagating

towards the centre of the SNR, the expansion of the ejecta is much faster than the

reverse shock, which is therefore carried outwards. When the amount of material swept

up by the SNR shock is roughly of mass equivalent to that of the ejecta, then the reverse

shock is no longer carried outwards by the expansion, and propagates inwards. It is this

stage in the SNR evolution that we are interested in studying. The key feature of this

stage of SNR evolution is the formation of a magnetised, non-radiative collisionless

shock.

A collisionless shock is a shock where the thickness of the shock transition from the

upstream to downstream state (see Chapter 2) is less than the ion-ion binary collision

mean free path. As will be described in Chapter 2, collisionless shocks require a

dissipative mechanism, with a corresponding scale length, for this transition to occur.

Early collisionless shock experiments33 supported by simulations34 focussed on the

formation of an electrostatic collisionless shock, where the scale length of the shock is

the Debye length, λD. A shock thickness of 5λD was observed33 in a plasma where the

binary collision mean free path was 103λD. Laser-plasma experiments to produce

electrostatic collisionless shocks were performed by Koopman and Tidman35. A 15ns 3J

pulse from a ruby laser was used to irradiate a solid cluster target driving a spherical

expansion into an ambient background plasma. A shock thickness of 0.1cm was

observed compared to the mean free path of 10cm.

This type of collisionless interaction between counter-propagating ions was also studied

by Dean et al36 again using a laser-plasma produced from a solid target expanding into a

photo-ionised background gas. These results were found to agree with the prediction of

the interaction as a ion-ion two-stream instability37 in the presence of a self generated

19

19

magnetic field. However, there was debate over the validity of the collisionless nature

of the system38,39.

In a similar experiment by Cheung et al,40 collisionless coupling in inter-penetrating

plasmas was observed in the presence of a magnetic field. Experimental evidence shows

that the magnetic field introduces turbulent effects and produces features resembling

collisional results, which do not exist without the magnetic field.

Laser-plasma experiments by Bell et al41 formed a magnetised collisionless shock by

colliding a magnetised laser produced plasma with a solid obstacle. The plasma is

generated by laser ablation of a solid Carbon target producing a spherically expanding

plasma. The plasma at a density of ~1018cm-3 impinges on a spherical carbon obstacle in

a 10kG magnetic field. The ratio of thermal to magnetic pressure, the plasma β, is

around 300 in this case, and the mean free path is 1mm. A bow shock was observed

with a thickness of between 10 and 50µm, comparable to the 70µm electron Larmor

radius (see Chapter 2, section 2.2.2), and the 5µm electron skin depth (see Chapter 2,

section 2.4.3). However, this experiment was performed without attempting to scale the

experiment to match an astrophysical system.

The formation of a collisionless shock has been a recurring goal in laser-plasma

physics, and the concept of producing of a system suitable for scaling has inspired new

experimental proposals21 and the work presented in this thesis. The aim of the work

presented here is therefore at the forefront of research within this field.

1.2.2 Thin foil laser produced plasmas

The methodology used in attempting to achieve an experimental simulation of a 100

year old supernova remnant is the formation of plasma from thin foils using a high

power laser. This technique can be employed, as shown by Decoste et al42 to produce

plasmas with high expansion velocities (>107 cm/s). Analytical models of the

hydrodynamics of exploding thin foils are presented by London and Rosen43 and by

Helander et al44. These models are examined in Chapter 2, and are essentially

hydrodynamic treatments of the exploding thin foil expansion.

Historically, thin foils have been used experimentally for many purposes, from X-ray

laser production, e.g. Rosen45, to testing the initial behaviour of inertial confinement

20

20

fusion targets46 and for studies of instabilities in shock waves47. The plasma produced

by direct drive irradiation of thin foils is known to be susceptible to hydrodynamic

instability48,49,50 caused by seed perturbations in either the laser focus51 or in the target

surface. It has been shown by Obenschain et al,52 that the amount of non-uniformity

imprinted into the thin foil from a laser focus is a function of intensity, and by Gardner

and Bodner53, Cole et al54 and Glendinning et al55 that it also scales as a function of

laser wavelength. This susceptibility of thin foils to imprinting makes them ideal tests

for laser smoothing techniques such as induced spatial incoherence56 and smoothing by

spectral dispersion57.

The foils used in our experiments differ from those already discussed in this section by

virtue of their thickness. In the experiments noted previously, the thickness of the foil

material is typically a few microns, whereas the experiments reported in this thesis use

0.1µm thick foils.

1.2.3 Colliding plasma experiments

The interactions between two colliding laser produced plasmas have been studied in the

1970’s by Rumsby et al,58 where two plasmas were formed in adjacent positions on a

Carbon plate, separated by a fixed distance, d. The evolution of two plasmas separated

by d=10mm produces parameters consistent with a Mach 3 hydrodynamic shock

occurring at the interface where the plasmas collide. This results in an observed five-

fold increase in electron density when compared with the evolution of a single plasma.

In this case, the mean free path of the ions in the plasma is estimated to be 6mm, i.e.

shorter than the scale of the experiment. In a second experiment where the separation

distance d=40mm the mean free path of the ions is 400mm, greater than the scale

length, and a density increase of a factor of two is observed. This would correspond to

the addition of the densities of two interpenetrating plasmas.

Theoretical modelling by Berger et al59 using a two ion fluid model of colliding plasmas

shows that for the case of plasmas that are expected to produce a shock, there will be

some degree of interpenetration of the plasmas. This is expected to occur if the scale

length of the velocity gradient in the plasma is shorter than the mean free path. This

interpenetration will subsequently lead to stagnation.

21

21

Experiments to investigate this proposed degree of interpenetration in the collisions of

two laser ablated plasmas were performed by Bosch et al.60 The plasmas were formed

by the ablation of the inner surfaces of two face parallel disks at 1014 W/cm2.

Measurements of the degree of interpenetration were made possible through spatially

resolved spectroscopic measurements of emission from Aluminium (Al) and

Magnesium (Mg) targets in an opposing geometry. The results were found to be in

agreement with the model of Berger et al59, as opposed to a single fluid model.

This model was extended by Rambo and Denavit61 to a multi-fluid model, showing that

the degree of interpenetration can be interpreted in terms of the collisionality parameter

of the system. This collisionality parameter can be described as the ratio of mean free

path of the ions to the scale length of the system, and is discussed further in Chapter 2,

section 2.3. Two experimental schemes are explored: the expansion of plasma ablated

from the front surface of solid disks as used by Bosch et al60, and the interaction of

counter-propagating plasmas produced from the rear surfaces of a pair of thin foils.

These simulations imply that shock formation in a multi-fluid treatment will occur after

a period of soft stagnation, and the shock strengths will be reduced in comparison to the

predictions of a single-fluid model. Numerical simulations by Larroche62 have also

included a kinetic treatment of the period of stagnation. Comparisons between multi-

fluid and kinetic approaches to simulation by Rambo and Procassini63 show that both

treatments produce similar results. Simulations of counter propagating plastic (CH)

plasmas where the carbon and hydrogen components are treated as separate fluids61

show that in a colliding geometry there may be significant separation between the two

species.

Counter-propagating thin foil experiments using volumetric heating of the foils by X-

rays have been performed by Perry et al64. The evolution of counter-propagating Al and

Mg foils in parallel and angled configurations was studied using spectroscopic and x-

ray radiography techniques. The results for the parallel geometry when compared to

radiation hydrodynamic simulations show the peak densities reached during the

collision are lower than predicted, as suggested by Rambo and Denavit61.

The effects of varying the collisionality parameter of a counter propagating exploding

thin foil system was studied by Rancu et al65 and Chenais-Popovics et al66. In this work

~1µm thick Al and Mg foils were separated by 400, 600 or 900µm and irradiated at

22

22

either 3 or 6 x1013 W/cm2 at 527nm. This produced a variation in collisionality, as the

dominant term in describing the collisionality in this system is the relative flow velocity

of the colliding plasmas66. Variations from the quasi-collisionless to strongly collisional

regimes with the associated degree of interpenetration are observed.

The experiments reported in this thesis using counter propagating thin foil plasmas also

study the effect of changing the collisionality of the system through the use of different

foil materials. The innovative approach is to also attempt to alter the collisionality

through the effect of an external magnetic field on a CH plasma.

1.3 Chapter Outline

Chapter two describes the fundamental theoretical aspects of plasma physics relevant

to producing plasma by irradiating thin foils using a high power laser. The concepts of

the plasma state will be introduced. This will be discussed in the context of how the

parameters of laboratory produced plasmas can be relevant to an astrophysical plasma

through similarity and scaling arguments. This will be achieved by firstly deriving the

basic parameters that describe a plasma in terms of typical scale lengths and times and

the physics of a collisionless shock. Secondly, the fundamentals of the laser plasma

interaction will be introduced along with analytical models for thin foil plasma

evolution and the numerical simulations undertaken.

Chapter three will develop the theory of the propagation of light through plasma and

the approach of geometric optics. How this affects the design of an optical probe system

for studying low density plasma will be investigated, along with the design criteria for

an optical probe system. Examples of how the optical probe systems are utilised in our

experimental work are discussed. The development of the optical probe for specific

experimental requirements is also outlined.

Chapter four describes the diagnostics used to gain information from an optical probe

in terms of the theory behind the diagnostic, the experimental arrangements used and

the methods utilised to interpret the data. This includes the principal diagnostics of

interferometry, schlieren imaging, shadowgraphy and polarimetry and an iterative ray

tracing code to simulate the passage of light through a plasma. The ray tracing code is

used as a post process to our hydrodynamic numerical simulations.

23

23

Chapter five presents the experimental investigation into the production of a plasma

from a thin foil. This includes the effects of target surface structures and laser

imprinting on the uniformity of the plasma, the evolution of the density profile of the

plasma and a determination of the leading edge expansion velocity. The effect of a

spatially filtered pre-pulse in limiting the degree of laser imprinting is evaluated, and

the measurements of the plasma parameters are compared with the analytical models

and numerical simulations.

Chapter six extends the experimental study by the introduction of two variables, the

presence of a strong magnetic field perpendicular to the plasma expansion and the

presence of a second plasma in a collisionless counter propagating geometry. The

effects of these two variables on the evolution of the plasma are investigated

individually and when combined.

Chapter seven presents the conclusions and outlines suggested avenues for further

work.

24

24

2 Chapter 2 - Theory

2.1 Introduction

The statement that plasma is the state of matter that exists when a gas is heated

sufficiently to ionise is a useful analogy, but not entirely true, as at any time a gas will

have some ionised components due to it’s temperature distribution. A more precise

definition of what constitutes plasma is given by Chen67:

A plasma is a quasi-neutral gas of charged and neutral particles which exhibits

collective behaviour.

This implies that a plasma is a macroscopic body containing many particles which,

when considered as a whole, should have negligible overall charge. A second

implication is that as some of the particles are charged, there will be long range

Coulomb interactions between these charged particles. In addition, the motion of these

charged particles will generate magnetic fields and so affect the behaviour of all the

other charged particles in the plasma. This gives rise to collective behaviour in the

plasma.

This means that in plasma it is not necessary for binary collisions between particles to

occur in order for the plasma to exhibit fluid like behaviour. An example of a collision

free type of interaction68 is the formation of a collisionless shock, where the shock

transition takes place on a scale much shorter than the collisional mean free path of the

particles.

In this chapter, I will introduce the fundamental plasma parameters and discuss what is

meant by the collisionality of a plasma, and then describe the formation of a

collisionless shock. I will discuss the arguments for scaling laboratory produced

plasmas to be relevant to astrophysical systems, with a discussion of the range of

plasma parameters. The method of experimentally producing plasmas by direct drive

irradiation of thin foil targets is also described, along with the analytical models and

numerical simulations used to design experiments suitable for scaled collisionless shock

studies.

25

25

2.2 Plasma Parameters

In the definition of plasma given in section 2.1, we have noted that the system of

interest is a macroscopic collection of many particles where the affect of

electromagnetic forces on particles is important. The theoretical approach to describing

the behaviour of plasmas can be either a particle approach or a macroscopic approach.

A macroscopic treatment of plasma as a continuous conducting fluid, where quantities

are a function of space and time, gives rise to the concept of magneto-hydrodynamics

(MHD). The plasma can be treated either as a single fluid, or as separate fluids of

electrons and different species of ions. These models enable simulation of large systems

where changes occur over long periods. However, this is at the expense of the detail in

the underlying physics causing the fluid like behaviour. By contrast, many forms of

plasma behaviour can be interpreted by a consideration of the effects of fields on a

single particle, giving rise to orbit theory. In the microscopic regime between these two

extremes, the approach of kinetic theory with a distribution function in space, time and

velocity can be used. This model encompasses more of the physics, but can become

computationally expensive.

Experimentally we most often measure macroscopic properties, such as the density,

which can be predicted by a MHD or hydrodynamic approach as is discussed later in

this chapter.

A full derivation of each of these approaches is beyond the scope of this chapter, and

there are many texts covering the field.69,70,71,72 In this section, I will introduce some of

the fundamental plasma parameters, mainly through particle orbit theory. This is due to

the simplicity of the arguments involved and the essential concepts introduced. In

addition, the particle orbit treatment is most valid for high-energy particles in low-

density plasma with an external magnetic field, the situation most applicable to our

experiments.

2.2.1 Collective effects

Collective behaviour is what distinguishes a plasma from a weakly ionised gas. In

essence, internal electromagnetic forces, as opposed to collisions, govern the behaviour

of particles in the plasma.

26

26

For the internal electromagnetic forces to dominate then there must be a shielding of

perturbations in the electric field to screen the bulk of the plasma. This also maintains

the quasi-neutrality of the system. The thickness of such a shielding effect, known as

Debye screening, can be derived as follows73.

If a test particle of charge q is placed in a neutral plasma where every ion is singly

ionised and there are an equal number of electrons and ions in thermal equilibrium, then

the distribution function, f, for both the electrons and ions is a Maxwell – Boltzmann74

distribution for each species denoted by a subscript j:

⎟⎟⎠

⎞⎜⎜⎝

⎛−=

Tkq

Tkmnf

B

j

Bjj

φ2

exp),(2vvx [2.1]

Where n is the number density, m is the mass, v is the velocity, kB is the Boltzman

constant, T is the temperature and q is the charge. This leads to a charge density given

by:

⎟⎟⎠

⎞⎜⎜⎝

⎛−=∑ Tk

qnq

B

jjj j

)(exp)(

rr

φρ [2.2]

Where φ(r) is the unknown potential function of the test particle. The potential function

must satisfy Poisson’s equation, given by:

)(1

0

2 rρε

φ −=∇ [2.3]

Where ε0 is the permittivity of vacuum. Therefore, in one dimension φ(r) must satisfy

1

21

0

22

3

<<

=

Tke

Tken

drdr

drd

r

B

B

φ

φε

φ

[2.4]

Solving for the solution where r → ∞ gives

⎟⎟⎠

⎞⎜⎜⎝

⎛ −=D

rrA

λφ exp [2.5]

27

27

Where λD is termed as the Debye length, the distance beyond which the potential of the

test particle is exponentially attenuated:

202en

TkBD

ελ = [2.6]

The effectiveness of this form of shielding can be evaluated by considering the number

of particles in a sphere of radius λD, termed ND. If there is a local change in potential,

this will be screened over the Debye length if the following condition is satisfied:

134 3 >>= DD nN πλ [2.7]

The inverse of ND, termed g, is often referred to as the plasma parameter75. A low value

of g implies that the plasma consists of a large number of particles, which interact

simultaneously but weakly. Conversely, a high value of g implies there are a small

number of collectively interacting particles with a strong interaction.

2.2.2 Single Particle Motions

As noted in the previous section the behaviour of particles in an electromagnetic field

govern the characteristics of plasmas. As such, it is useful to consider the behaviour of

individual particles in electric and magnetic fields. Such a treatment illustrates that the

behaviour of test particles can be described in terms of a Larmor radius and the gyro-

frequency. The full discussion of this topic can be found in the literature,76,77,78,79

however the main points of relevance will be discussed here.

If we consider a particle of charge q, mass m, and velocity v in the presence of an

electric field, E, we have the result:

Ev qdtdm = [2.8]

Simply that the particle accelerates linearly in the direction of the electric field. If the

field is purely magnetic, and of magnitude and direction B, the force on the particle is

[ ]Bvv ×= qdtdm [2.9]

28

28

If the velocity is split into components parallel (vII) and perpendicular (v⊥) to the

direction of the magnetic field the following results can be obtained.

0=

= ⊥⊥

dtdv

Bvdt

dvqm

II

[2.10]

The motion perpendicular to the field is an orbital motion of frequency ωc, known as the

gyro-frequency, Larmor frequency or cyclotron frequency.

mqB

c =ω [2.11]

When combined with the test particle travelling with constant vII , parallel to the field,

the particle motion takes the form of a helix of radius rL, referred to as the Larmor

radius or gyroradius, rotating around a guiding centre that moves with velocity vII.

cL

vr

ω⊥= [2.12]

Now, if both electric and magnetic fields are present then the equation of motion is

described by the Lorentz equation:

[ ]BvEv ×+= edtdm [2.13]

Again the components of the motion can be separated in to components perpendicular

and parallel to the direction of the magnetic field:

IIII E

dtdv

em

BvEdt

dvem

=

+= ⊥⊥⊥

[2.14]

29

29

Which can be described again as a helical motion of Larmor radius and gyro-frequency,

but where the guiding centre accelerates along the magnetic field line, and drifts with a

velocity given by vDrift:.

( )2BDriftBEv ×= [2.15]

This can be generalised to any external force, F, acting on the particle, not just the force

due to the electric field. In this case, the drift velocity becomes:

( )2

1BqDrift

BFv ×= [2.16]

Where there is an additional linear acceleration due to the component of the force

parallel to B. In this scenario the direction of the drift velocity depends on the charge of

the particle.

This treatment works well for uniform fields, and can be extended to non-uniform fields

by the use of orbit theory. The condition for this is that the scale length of change in the

field is much larger than the Larmor radius. In this way the motion can be treated as two

components, the unperturbed gyro-motion of the particle around it’s guiding centre and

the drift of the guiding centre.

For the majority of variations in electric and magnetic fields the effect on the guiding

centre can be determined to find an effective force F due to the non-uniformity that

results in a drift velocity of the guiding centre. The full derivations can be found in the

literature,76,77,78 only the results are quoted here in order to illustrate the effects of a non-

uniform B field on the dynamics of a experiment performed in a magnetic field.

30

30

Field Variation Associated drift velocity73

Non-uniformity in E 2

22

411

BrL

BE ×⎟⎠⎞

⎜⎝⎛ ∇+

Non-uniform B: Gradient of B perpendicular to

B 22

1B

rv LBB ∇×± ⊥

Non-uniform B: Curvature of B 22

2

BRemv

c

II BR c ×

Temporal variation of E dtd

Bc

[2.17]

These results do not include two special cases, the gradient of B being parallel to B, and

B varying slowly in time. The first of these two cases can be analysed by considering an

axisymmetric B field, with a gradient in the direction of propagation along the axis, z .

When averaged over one gyration, the force acting on the particle, Fz, is only in the

direction of z:

zB

zB

Bmv

zB

revF

z

z

zLz

∂∂

−=

∂∂

−=

∂∂

±=

µ

2

2121

[2.18]

Where µ is the magnetic moment of the particle, and is an invariant parameter. This can

be verified by conserving the energy of the particle as it moves through a change in

magnetic field strength. This invariance is the principle behind the trapping of plasma

inside a magnetic ‘bottle’.

The second scenario is that the magnetic field is varying with time. As the magnetic

field does not directly impart energy to a charged particle, the problem needs to be

31

31

analysed from the electric field associated with the magnetic field as given by

Maxwell’s equation:

dtdBE −=×∇ [2.19]

If we assume the field changes slowly, we can show that the associated electric field

alters the perpendicular velocity. Hence, an increase in the magnetic field strength is

coupled directly into an increased kinetic energy of the particles orbiting around the

guiding centre, varying the Larmor radius. This fact can be used to heat a plasma by

increasing the external magnetic field. This derivation can be extended to show that the

magnetic moment, µ, is also invariant in time.

From this we can conclude that in order to prevent a drift velocity, or related effect

perturbing an experiment performed in a magnetic field, the field must be both spatially

and temporally uniform over the extent of the experiment.

2.2.3 Plasma Oscillations

In the previous section we have seen that that particle orbits can be characterised by the

Larmor radius and gyro-frequency. Oscillations in plasmas have their own characteristic

plasma frequency, ωp.

The plasma frequency can be derived from a consideration of the effects of a small

perturbation in the electron distribution in homogeneous plasma80 such as would be

screened within the Debye length. If the background particle density distribution, n0, is

homogeneous for both ions and electrons the electron density function, ne, can be

written as

),(),( 10 tnntne rr += [2.20]

The perturbation in electron density, n1, will give rise to an electric field proportional to

the size of the perturbation. This will in turn accelerate all the surrounding electrons

(ignoring the ion motion due to their inertia) to a small velocity u . Considering that the

perturbation is much smaller than the background density, the change in the

perturbation leads to a restoring motion in the flow such that:

32

32

u⋅∇−=∂

∂0

1 nt

n [2.21]

Due to the force exerted on the electrons, of mass me, by the electric field

Eu et

me −=∂∂

[2.22]

And the electric field is given by Poisson’s equation

en100

11ε

ρε

==⋅∇ E [2.23]

If the temporal derivative of [2.21] and the divergence of [2.22] are substituted into

[2.23] we obtain

010

20

21

2

=⎟⎟⎠

⎞⎜⎜⎝

⎛−

∂∂ n

men

tn

eε [2.24]

Implying that any electron density perturbation will oscillate with a natural frequency,

called the plasma frequency, ωp given by

ep m

en

0

20

εω = [2.25]

2.3 Collisionality

In order that electromagnetic forces and not collisions dominate the behaviour of

particles, the natural frequency of plasma oscillations ωp has to be greater than the

collision frequency between particles. If the mean time between collisions is τ this

condition becomes:

1>τω p [2.26]

33

33

The value of τ can be evaluated from the mean free path, λmfp, of the particles, the

average distance travelled before a collision is expected, that make up the plasma and

the thermal velocity, vth, of the particles:

th

mfp

τ = [2.27]

The mean free path of a particle depends upon the species of the particle and the species

with which it collides. In this instance, the discussion will be limited to the ion-ion

mean free path, λii, as they carry the majority of the inertia of the plasma. For ion-ion

collisions within a plasma then mean free path can be written as81:

( )ii

44

20

lnΛi

iBii nZe

Tkελ = [2.28]

Where kB is the Boltzman constant, Ti is the ion temperature, e is the elementary charge,

Z is the average ionisation state of the plasma, ni is the ion density and Ln Λii is the

Coulomb logarithm. The Coulomb logarithm is defined as:

⎟⎟⎠

⎞⎜⎜⎝

⎛=

min

maxlnlnΛrr

[2.29]

Where rmax and rmin are the maximum and minimum impact parameters of the two

species involved in the collision event. Normally rmax is taken to be the Debye length, as

defined in section 2.2.1 as beyond this distance the coulomb field of the ion is

exponentially attenuated. The value for rmin is taken to be either the classical distance of

closest approach (found by evaluating the distance where the kinetic energy of the

particle is converted into potential energy of the Coulomb field) between the two

particles, or the De Broglie wavelength of the particle, whichever is greater.

This treatment shows that low atomic number, high temperature, low density plasmas

will have longer mean free paths, and be less dominated by collisions. However,

equation [2.28] is derived by studying the motion of a particle with its thermal velocity

with respect to other particles. This treatment does not include the possibility of the

particle having a flow velocity in addition to its thermal velocity with respect to the

particle with which it is colliding. In the situation where the flow of one plasma

34

34

interacts with a second plasma the relative flow velocity between the two can be much

greater than the thermal velocity. In this case the mean free path can be re-written81 in

terms of the relative flow velocity for a test ion species in plasma 1 colliding with an

identical ion species in plasma 2:

12242

412

20

LnΛi

iii nZe

vmελ = [2.30]

Where v12 is the relative flow velocity between the two plasmas and ni2 is the ion

density in the second plasma. In an experimental arrangement, it is therefore possible to

have two collisional plasmas interacting in a collisionless manner.

In order to determine the dominance of collisions in this type of scenario it is normal to

compare the magnitude of the mean free path to a suitable scale length, L, such as the

Debye length or the scale of the system, by use of the collisionality parameter for the

system ς:

Lmfpλ

ς = [2.31]

If this parameter is much larger than 1 then relatively few collisions occur. The lack of

collisions in such a system is a problem if hydrodynamic behaviour is assumed. The

evolution of the flow of the two plasmas would be independent and indistinguishable

from the expansion of a single plasma, i.e. the plasmas should interpenetrate. However,

in the presence of a magnetic field we have already seen in section 2.2.2, charged

particles in the plasma have a helical trajectory around the magnetic field lines. The

scale of the gyration is the Larmor radius, and if this value is less than the mean free

path then this is the scale over which the particles may be localised within the plasma.

In this situation, it is possible for a magnetic field to introduce an effective collisionality

to the system, and if the assumptions of ideal MHD hold it will behave like a

hydrodynamic fluid.

The effect the magnetic field has on the particle dynamics of the plasma can be

characterised by the plasma β, the ratio of the thermal pressure to the magnetic pressure

in the plasma.

35

35

202

Bpµβ = [2.32]

Where µ0 is the permeability of vacuum. For the magnetic field to induce an effective

collisionality into the system without affecting the expansion of the two counter-

streaming plasmas would require a judicious choice of magnetic field strength such that

the value of β is much greater than 1, while the Larmor radius is still smaller than the

mean free path and the scale of the system.

2.4 Collisionless shocks

Collisionless shocks are a common astrophysical phenomenon, but not naturally

occurring on earth. Our nearest example of a collisionless shock is the bow shock

around the Earth82 caused by the interaction of the solar wind with the earth’s magnetic

field. In this section, I will explore the formation of a collisionless shock through

developing the standard theory of a hydrodynamic shock and showing how the same

processes can occur in a collisionless system.

2.4.1 Shock formation

If we consider a pressure wave in a perfect gas where the speed of sound, cs, is given by

cs = √(γp/ρ), where p is the pressure, ρ is the density and γ is the adiabatic index. The

variation of pressure along the wave will lead to locally defined regions moving at

different speeds, as the sound speed is a function of pressure. High-pressure elements of

the wave will move faster than the lower-pressure regions causing the waveform to

become steeper. As the velocity of the fluid can only have a single value at any point,

the fluid elements cannot physically overtake each other and eventually a discontinuity

is formed.

If the magnitude of the pressure variation in the wave is small compared to the

background pressure then this can be considered indistinguishable from an acoustic

wave propagating with uniform sound speed. However, for large variations of pressure

a shock is produced.

36

36

What prevents the wave from overturning is the mechanism for momentum transfer and

wave propagation82. If this mechanism is able to balance the processes that steepen the

wave, then a steady shock wave is produced.

2.4.2 Shock conditions

A shock wave is a discontinuity in the pressure and density of a fluid travelling faster

than the local speed of sound. This discontinuity separates the fluid into two regions, the

upstream region into which the shock is propagating and the downstream region

through which the shock has already passed.

As a mathematical discontinuity is infinitely thin it is possible to apply the normal laws

of conservation for mass, momentum and energy across the interface between the

upstream and downstream regions83. If we denote the upstream and downstream regions

with the subscripts 0 and 1 respectively, these conservation equations can be written as:

22

20

0

00

21

1

11

2000

2111

0011

upup

upup

uu

++=++

+=+

=

ρε

ρε

ρρ

ρρ

[2.33]

Where p is the pressure, ρ is the density, u is the flow velocity relative to the shock

discontinuity and ε is the specific internal energy of each region. By replacing the

density by it’s inverse the specific volume, V, and combining the equations in [2.33]

through the flow velocity terms the following relationship is obtained:

( ) ( ) ( )( )0101000111 21 VVppVpVp +−=+−+ εε [2.34]

This equation, known as the Hugoniot relation, relates the upstream and downstream

conditions under shock compression. If the system can be described as a perfect gas

with constant specific heats then the value of ε can be written as:

pV1

1−

ε [2.35]

37

37

Where γ is the adiabatic index. This allows an explicit form of the Hugoniot relation to

be written:

( ) ( )( ) ( ) 01

10

0

1

1111

VVVV

pp

−−+−−+

=γγγγ

[2.36]

From this, it can be seen that for a set of initial upstream conditions there is a fixed set

of possible downstream conditions, a locus of end points known as the Hugoniot curve.

This also introduces a limit to the amount of compression that can occur due to the

shock equal to:

11

1

0

0

1

−+==

γγ

ρρ

VV

[2.37]

For an ideal gas, γ is 5/3 implying a maximum compression of a factor of four. The

treatment of the properties of the shock can now be classified into two types, strong

shocks where p1/p0 → ∞ and weak shocks where p1 ≈ p0. In the case of a strong shock,

the degree of compression is near the limit imposed by [2.37]. If we consider the case of

the strong shock then the upstream and downstream flow velocities can be written as:

( )( )

21

01

2

1

21

010

121

21

⎟⎟⎠

⎞⎜⎜⎝

+−=

⎟⎠⎞

⎜⎝⎛ +=

Vpu

Vpu

γγ

γ

[2.38]

If these velocities are compared with the speed of sound in each region, cs = √(γpV), it

can be shown that:

( ) ( )

( ) ( )γ

γγ

γ

γγ

2

11

2

11

1

02

1

1

0

12

0

0

pp

cu

pp

cu

s

s

++−=⎟⎟

⎞⎜⎜⎝

++−=⎟⎟

⎞⎜⎜⎝

[2.39]

This shows that the downstream flow is subsonic, whereas the upstream flow is

supersonic, as required by the definition of a shock. The ratio of the upstream flow

38

38

velocity to the speed of sound, termed the Mach number, M, serves as a useful indicator

of the shock strength. For a weak shock M is close to 1, whereas for a strong shock

M>>1.

2.4.3 Shock thickness

In the theoretical treatment of the previous section the shock discontinuity is considered

to be infinitely thin, however in a real shock the transition must occur over a finite

layer. As there is a change in entropy between the upstream and downstream region, a

dissipative mechanism must be responsible for the transition. The dissipative

mechanism is responsible for defining the width of the transitional region in terms of

the characteristic scale length of the mechanism involved to heat the fluid experiencing

the compression. For a fluid where the dominant mechanism for heat transfer is through

binary collisions the characteristic scale length would be the mean free path of the

particles.

However, for a shock to form in a collisionless plasma where binary collisions are rare,

a different form of dissipative mechanism must be dominant. The processes leading to

the wave steepening in a collisionless plasma are more complex than the treatment in

section 2.4.1. A collisionless plasma is a dispersive media, which means that the speed

of the wave propagation is a function of the wavelength, as determined by a dispersion

relation. A derivation of a basic plasma dispersion relation is given in Chapter 3,

however for the purposes of this treatment the relevant dispersion relation is such that

for wavelengths above a critical value the wave velocity is constant. Below the critical

value the wave velocity either increases or decreases with wavelength, depending on the

orientation of the wave propagation to the magnetic field.

If we consider again the concept of a steepening wave, but in this instance, decompose

the wave into its Fourier components. As the wave becomes steeper the short

wavelength Fourier components become more dominant in the series. As the different

wavelength components below the critical value now have different velocities due to

dispersion, the dominant components will physically separate themselves from the bulk

of the wave. The cyclic combination of steepening, dispersion and separation converts

the initial wave into a series of pulses of thickness approximately that of this critical

wavelength. The magnetic field defines a unique direction with respect to the shock. In

39

39

general, there are three orientations of the magnetic field with respect to the flow of the

fluid: perpendicular, parallel and the general case of at an angle. In our experiment, the

magnetic field is perpendicular to the flow of the fluid.

When the magnetic field is perpendicular to the propagation of the wave, the dispersion

relation84 decreases the wave velocity for wavelengths below c/ωp, producing a series of

compression pulses. As the weakest (lowest amplitude) pulses are produced first and

subsequently fall behind the bulk of the wave, the shock front will take the form of a

large pulse followed by a series of pulses of decreasing amplitude. In a collisionless

perpendicular shock, the type of wave involved is a magneto-acoustic wave, with a

characteristic Alfvén velocity85 analogous to the speed of sound in an acoustic wave. In

this regime, it is more appropriate to consider the shock conditions in terms of the ratio

of the flow velocity to the Alfvén velocity, termed the Alfvén Mach number, MA .

When the magnetic field is not perpendicular to the propagation direction, the

dispersion relation will increase the wave velocity for wavelengths below the ion

Larmor radius82. This produces a shock front comprised of a wave train of rarefaction

pulses of increasing amplitude.

The thickness of the shock is now governed by the length of the wave train of pulses,

which in turn is governed by the rate at which the energy of these pulses dissipates.

The possible mechanisms for collisionless dissipation of the wave train again depend on

the orientation of the wave propagation to the magnetic field, and in some cases the

strength of the shock itself. Descriptions of proposed collisionless dissipative

mechanisms, from wave-particle interactions to magnetic turbulence can be found in the

literature 68,82,84,86. A complete treatment of all the possible types of dissipation would

be beyond the scope of this chapter, as the important parameter is the scale length of the

dissipative mechanism, which has already been discussed. If whatever mechanism is

employed is sufficient to balance the entropy difference between the upstream and

downstream regions, it will be achieved on a scale comparable with the size of the

pulses. Hence for a perpendicular shock we would expect the shock thickness to be of

the order of c/ωp, and for a non perpendicular shock we would expect the thickness to

be of the order of the ion Larmor radius.

40

40

Our closest example of a collisionless shock is the bow shock formed by the solar wind

around the Earth’s magnetic field. The thickness of this shock is around 103km whereas

the mean free path of the solar wind is approximately 108km. It is worth noting that in

the region of the shock the magnetic field is between 45 to 50 degrees to the shock

front, and the ion Larmor radius is several hundred kilometres82. Furthermore, the

system has a low plasma β~1, whereas for our system of interest, a 100 year old SNR,

β>>1. The observed shock thickness is very close to the scale length of shock thickness

predicted for a non-perpendicular shock.

2.5 Scaling

From the derivations in the previous sections, we can deduce several important plasma

parameters. These parameters, such as the Debye length, plasma frequency and Larmor

radius, yield scales over which we can infer the behaviour of the plasma to be governed

in a certain way. However, if we wish to compare the behaviour of two different

plasmas we need to introduce the concept of scale invariance.

Scaling laws rely on the use of dimensionless constants relevant to the two systems If

the assumptions of the scaling model hold, then the two systems will show similarity.

For example, if we consider the basic example of a rectangle of sides length a and b . If

we wish to compare this rectangle with another rectangle of sides length a’ and b’, only

if the dimensionless ratio a/b was the same as a’/b’ would the two rectangles be similar.

It can also be seen from this treatment that there exists a scaling transformation between

the two systems which satisfies the matching of the dimensionless constant, namely that

a’ = xa and b’ = xb, a linear transformation of magnification x.

This principle was extended by Connor and Taylor87 to show that if the fundamental

equations describing plasma behaviour are invariant under a given transformation, then

a scaling law derived from those equations must also be invariant under the

transformation. To apply an invariant scale there are constraints placed on the model

used to describe the plasma. Our aim is to create a simulation of a SNR and scale such

an experiment using an argument. The SNR phenomenon is a collisionless, high β

system to which ideal MHD can be applied. Connor and Taylor87 identified an ideal

MHD transformation which they termed ‘E2’. This transformation seems the most

appropriate to seek in experiments.

41

41

Ryutov et al88 developed these concepts further, and propose a set of transformations

for the Euler equations describing the evolution of an ideal compressible hydrodynamic

fluid with the thermodynamic properties of a polytropic gas. This work was later

extended89 to the case of ideal MHD. In this case the Euler equations are:

0=•∇+∂∂ vρρ

t [2.40]

BBvvv ××∇+−∇=⎟⎠⎞

⎜⎝⎛ ∇•+

∂∂ )(1

0µρ p

t [2.41]

BvB ××∇=∂∂

t [2.42]

Where v is the velocity, ρ is the density, p is the pressure and B is the magnetic field.

These equations describe conservation of mass, conservation of momentum and

magnetic induction respectively. A fourth equation, for conservation of energy is also

required:

( ) vv •∇+−=∇•+∂∂ εεε p

t [2.43]

Where ε is the internal energy per unit volume. As the system is described as a

polytropic gas, the internal energy is a linear function of the pressure:

Cp=ε [2.44]

The assumption of a polytropic gas is suitable for describing a fully ionised medium, as

there are no changes in the number of degrees of freedom caused by an increase in

temperature. Substituting [2.44] into [2.43] leads to:

vv •∇−=∇•+∂∂ pp

tp γ [2.45]

Where γ is the adiabatic index equal to 1 + 1/C.

These equations are invariant under the transformations between systems 1 and 2

denoted with subscripts:

42

42

21

21

21

21

21

21

BB

vv

ccb

tcbat

cppbarr

=

=

=

===

ρρ

[2.46]

Where a, b, and c are arbitrary positive numbers. For these transformations Ryutov et

al89 demonstrate the invariance of two dimensionless parameters from an initial value

problem for the similarity of two evolving plasmas:

2

2

1

1

2

22

1

11

pp

Eupp

BB

vv

=

==ρρ

[2.47]

The first invariant, Eu, dubbed the Euler number by Ryutov et al, and when calculated

the velocities, pressures and densities do not have to all be taken from the same point.

However, the values must be taken from corresponding points in both systems. The

second invariant is equivalent to the inverse of the square root of the plasma β, which

has already been defined as a key dimensionless parameter for scaling by Connor and

Taylor87.

This derivation makes a number of assumptions about the plasma, the validity of which

needs to be examined. The primary assumption of any fluid description is that the

medium involved can be treated as a fluid. In the case of a collisionless interaction as

described in section 2.3, this may not always be the case. However, assuming an

external magnetic field is applied to the plasma of sufficient strength that the Larmor

radius is smaller than the mean free path and the system size, then this assumption is

valid.

This treatment assumes gravitational effects are negligible as this force is not included

in any form, which is valid for the laboratory experiment, but may not be for an

astrophysical plasma.

43

43

By describing the plasma by the Euler equations the effects of heat conduction,

viscosity and radiative diffusion are all assumed to be negligible as the treatment does

not include the effects of dissipative terms.90 The use of ideal MHD also implicitly

assumes that the fluid has zero resistivity.91

The scaling argument presented by Ryutov et al is designed to show that two systems

with identical initial conditions will evolve in a similar manner if the scaling conditions

are met. However, it is not always possible to duplicate the initial conditions of an

astrophysical system of interest. In this scenario, we present the assumption that if a

laboratory simulation can evolve to a point where the similarity criteria are met, then for

a limited time the behaviour of the two systems may be similar. This should enable a

‘snapshot’ of the system to be created, suitable for testing the physics of the

astrophysical system in the laboratory.

Hence we have two scaling parameters which we need to match in any experiment, Eu

and β, but there are other factors to consider. In the simulation we are trying to achieve,

a collision-free interaction is required, so the collisionality, ς, needs to be much greater

than 1. We also require a high Alfven mach number, MA, shock, hence this parameter

should be much greater than 1.

In considering the design of an experiment to match these scaling criteria, it is useful to

compare the range of plasma parameters that are observed from different sources. The

table below presents a comparison between the conditions typical of naturally occurring

plasmas and the conditions typical of laboratory produced plasmas. Included in the table

are values for magnetic confinement fusion experiments (MFE) and inertially confined

fusion experiments (IFE).

44

44

RELATIVISTIC PLASMAS

QUANTUMPLASMAS

CLASSICALPLASMAS

stronglycoupledplasmas

Pulsar

MFEIFE

SolarCorona

Dis -charge

Iono -sphere

SolarWind

Magneto-sphere

Non-neutral

Thermalprocessing

Lightning

WhiteDwarfs

Electrons inMetals

SolarInterior

kBT=mc2

k BT=EF

EF=e2n1/ 3

nλD3=1

1010

108

106

104

102

1

Tem

pera

ture

(K)

1 1010 10 20 10 30

Density (cm-3 )

Figure 2.1 Diagram of the temperature and density scales of different plasma sources based on Ref 92 with the addition of the parameters of thin foil laser produced plasmas (TFP) and a wire array z-pinch (Z-P)93

The parameters appropriate to a 100 year old SNR are encapsulated in the solar wind

category, the differences between the two being the relative strength of the magnetic

field governing the plasma and the size of the systems. The designed parameters of our

experiment are shown as TFP, thin foil laser produced plasma, with a density ~1018cm-3,

high velocity ~107cm/s and low Z~3.5.

2.6 Laser-plasma interactions

The formation of plasma from a solid target using a high power laser is the technique

used to produce the required plasma parameters to satisfy the scaling arguments

presented in the previous section. The physics of laser-plasma formation is well

documented in the literature94,95. In this section, I will focus on the specific example of

the production of plasma from a 100nm thick C6H8 target by a pulsed laser of 1µm

wavelength. The laser irradiation occurs at 1014 W/cm2 for a pulse duration of 80ps. The

modelling approaches that can be used to simulate the evolution of this type of system

are also discussed.

TFP

Z-P

45

45

Initially the laser is incident on the surface of the solid target, and the high average

electric field (~2.7x1010 V/m in vacuum) removes electrons from the solid surface over

a period of a few laser cycles96. These electrons are further accelerated by the electric

field and will cause ionisation through collisions with the solid. This will produce

mobile ions, which expand into the vacuum. After approximately 10-12s a plasma has

been formed with an roughly exponential density profile decaying away from the solid

surface. The laser is now interacting with a plasma, which has its own natural response

frequency to electric field perturbations, the plasma frequency as defined in section

2.2.1. The effect of the plasma natural frequency is to place an upper limit on the

density of the plasma to which the laser light can penetrate. At this point the frequency

of the laser beam matches the electron plasma frequency. The electron density at which

this occurs is termed the critical density, nc. As will be discussed in the next chapter, the

refractive index of the plasma at this point falls to zero. A useful formula for calculating

the value of the critical density for any wavelength is:

[ ] [ ]2

213

µm101.1cm

λ×=−

cn [2.49]

Where λ[µm] is the wavelength of the laser in µm and nc[cm-3] is the critical density in

cm-3. For a 1µm laser, the critical density surface is at 1.1x1021cm-3, two orders of

magnitude below solid density at around 1023cm-3.

At the intensity of 1014 W/cm2, the laser energy is mostly absorbed in the under-dense

(below critical density) region of the plasma through a process known as inverse

bremsstrahlung. Bremsstrahlung radiation can be considered as the emission of

radiation by an electron undergoing acceleration in the field of an ion97. Hence, inverse

bremsstrahlung can be considered to be the absorption of a photon during the collision

of an electron with an ion. The absorption coefficient for this process, αib, can be

written as96:

22

2

ωω

α+

=ei

peiib vnc

v [2.50]

46

46

Where n is the refractive index, c is the speed of light in vacuum, ω is the angular

frequency of the laser, ωp is the plasma frequency and vei is the electron-ion collision

frequency given by:

gZen

Tkmv e

eBeei 3

12

161

20

423

21

πεππ ⎟⎟

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛= [2.51]

Where Z is the ionisation state, ne the electron density, kB the Boltzman constant, Te is

the electron temperature, me is the mass of an electron and g is the Gaunt factor ~1.

From this, we can see that the rate of absorption is going to increase with plasma

density, and that the absorption is most efficient for cold electrons. Once the electrons

become heated, the collision frequency will decrease and the electrons will remain at a

fixed temperature, but be collisionless with respect to the ions.

Diffusive transport by the heated electrons into the over-dense plasma propagates the

laser energy to the solid surface where the solid is ablated to form more plasma. The

lower energy particles produced at the ablation front drift back to the under-dense

region where they are heated and the process cycles. In thick targets the pressure of the

ablation process can form a shock wave that compresses the target, however in thin foils

the entire target is completely ablated before a shock would be able to develop.

Once the foil has been completely ablated, the plasma is heated for the remainder of the

laser pulse. After the laser has been switched off, the plasma’s behaviour can be

modelled as an adiabatic expansion98.

By treating the plasma as a single fluid, the adiabatic expansion can be modelled from

the ideal hydrodynamic equations in one dimension99:

( )

xxvv

tv

xv

t

∂∂−=⎟

⎠⎞

⎜⎝⎛

∂∂+

∂∂

=∂

∂+∂∂

ρρ

ρρ 0 [2.52]

Where ρ is the density, p is the pressure and v is the velocity of the fluid. If we assume

that the plasma is uniformly heated and that the instantaneous velocity is going to be a

linear function of position based on a time-dependent scale factor, L:

47

47

xLdt

dLv

⎟⎟⎟

⎜⎜⎜

⎛= [2.53]

For each instant, the density function will take the form of a Gaussian distribution:

( ) ⎥⎦

⎤⎢⎣

⎡−= 2

2

)(2exp

)(1,

tLx

tLtxne [2.54]

Where for a given time the value of the scale factor L can be found from the speed of

sound in the plasma, cs, and the original width of the foil L0:

⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛=

0

ln2LL

cLt

s

[2.55]

This analytical model can be extended further99 to include the rate of heating during the

laser pulse and the heating of the plasma after the plasma has expanded to below critical

density. However, as in the case of the 80ps laser pulse we are considering it should

take ~250ps for the plasma to become under-dense.

More detailed modelling of the production of thin foil plasma can be achieved through

the use of numerical simulations. For the conditions already outlined, we utilised a

modified version of the simulation code Medusa100. This is a one dimensional

lagrangian numerical simulation of the hydrodynamic and thermodynamic behaviour of

a plasma irradiated by an intense laser. The modifications made to Medusa for these

simulations allow the plasma to expand to a fixed point in vacuum where it reaches a

barrier. In addition, the code was modified by R.G. Evans to include flux limited

thermal conductivity, required for shock simulation. In this way, the single fluid code

should be able to model the two extreme cases of a counter propagating plasma system.

Firstly, if the system is completely collisionless, yet a fluid description can be applied,

the evolution of one plasma should be independent of the presence of the second

plasma, hence only one plasma needs to be modelled. Secondly, if the plasmas cannot

interpenetrate then the evolution of each plasma should be self-contained up to the

halfway point between the two plasmas.

48

48

A series of numerical investigations into possible experimental configurations were

carried out101 to find a suitable experimental geometry for producing conditions suitable

for matching the scaling criteria. The experimental design is based on the conditions

considered in this section, the production of plasma from two 100nm thick C6H8 targets

by 1µm laser irradiation at 1014 W/cm2 with a gaussian 80ps pulse. The foils are placed

face parallel in a counter-propagating geometry with a separation of 1mm. To ensure the

one-dimensional treatment of the expansion is valid, experimentally a 1mm diameter

focal spot is required. The results of the experimental simulations are presented

below102:

49

49

Figure 2.2 Medusa simulations of the proposed colliding foils scheme102, with the initial foil positions marked at +/- 0.5mm. The plasma is allowed to expand in vacuum up to the centre point around which the data is mirrored. Graph a) shows the electron density, b) shows the thermal ion temperature, c) shows the velocity and d) the pressure at 100ps intervals.

The parameters of the simulated plasma match the scaling criteria as outlined in the

table below, assuming the experiment is performed in a 20T magnetic field. The values

50

50

presented in the middle column are believed to be typical of a SNR 100 years after the

supernova explosion:

Parameter 100 yr SNR Experiment

Collisionality, ς 3.00x104 56

Euler Number, Eu 3.1 2.8

Plasma β 400 400

Mach Number 100 5

Alfven Mach Number 3.00x103 80

Figure 2.3 Scaling parameters are matched from the experimental design to a 100 year old SNR.

Both systems are collisionless, as the collisionality is much greater than unity. The

behaviour of the systems should be similar as the plasma β and Euler number values are

nearly identical, and the shocks in both systems are strong shocks with Mach numbers

much greater than unity. The experimental conditions are designed to match the values

given in Figure 2.3 approximately 500ps after laser irradiation.

2.7 Conclusions

In this chapter, I have introduced the concept of a plasma, a quasi-neutral macroscopic

collection of charged and neutral particles. I have developed the primary concepts of the

response of a plasma to electric field perturbations in terms of the Debye length and

natural plasma frequency, and the behaviour of single particles in the presence of

electric and magnetic fields. The ability of a plasma to interact through long range

Coulomb interactions permits fluid like behaviour without the need for binary

collisions. This is only possible in the presence of a magnetic field to localise the

particles involved within their Larmor radius. I have discussed the possibility of

producing a collision free interaction using counter-propagating plasmas to increase the

ion-ion mean free path to greater than the system size. I have described how a shock can

51

51

form in such a collisionless system, including the characteristic shock thickness that

would be expected to form. I have introduced the scaling laws relevant to a plasma

which can be described by ideal MHD.

I have demonstrated how a plasma can be produced from a thin foil by laser irradiation,

and the parameters of such a plasma can be modelled by analytical and numerical

methods. I have also shown that the parameters of a laser-plasma experiment can be

designed to be similar to a 100 year old SNR.

52

52

3 Chapter 3 - Optical probes.

3.1 Introduction

This chapter is a review of the properties of optical laser probes, and how these probes

can be used to measure parameters such as the refractive index of a plasma. The

interaction of a laser probe with a plasma can be divided in to two categories, a)

scattering processes through the interaction of an electromagnetic wave with a particle,

and b) beam propagation through a plasma. The former processes, which are not of

interest to the work described in this thesis, include Thomson, Mie, Rayleigh, Raman,

and Brillouin scattering103, which represent an extremely useful array of diagnostics.

The latter processes include reflection, transmission, refraction, polarisation, scattering,

and absorption.

The experiments described in Chapters 5 and 6 use a probe in transmission. These

experiments are designed to take advantage of the non-perturbative, accurate, and

sensitive nature of an optical probe to produce space and time resolved data. There are

constraints on the use of a probe set by the probe wavelength, and the spatial resolution

and dynamic range and resolution of the detector. The experiments are designed around

these constraints using computer simulation to determine both the experimental size and

expected electron densities. In these experiments, the probe was used to measure a

number of parameters simultaneously, these are:

1. image of the experiment

2. electron density, ne,

3. electron density gradients, ∇ne, and

4. magnetic field, B

To demonstrate how these measurements can be made, and the underlying assumptions

required I will discuss the propagation of a probe beam through an inhomogeneous

plasma. This description begins with a discussion of the dispersion relation and

refractive index of an ideal plasma to illustrate the conditions for which this treatment is

valid. I will then use the geometrical optics approximation and treat the probe

53

53

propagation in terms of rays and consider a plasma of varying refractive index. This sets

the foundation for describing the design factors necessary in building an optical probe

beam line and imaging system. A number of different probe designs and imaging

systems have been used in experiments, the design specifications and the advantages

and disadvantages are described.

3.2 Theory of propagation of light in an under-dense plasma

In general the passage of an electromagnetic ray through a gas or plasma is affected by

all the scattering, and beam propagation processes mentioned in Section 3.1 and are

illustrated in Figure 3.1. Generally, scattering is observed either at high intensities or

from a large plasma, since scattering cross-sections tend to be very small. The laser

intensities are low, below the damage threshold of delicate optical components

(approximately 0.5 J.cm-2), and the plasma volume and density low, thus scattering is

ignored.

Figure 3.1 Illustration of the processes that occur during the propagation of a laser probe through a plasma. With absorption, these processes tend to attenuate the transmitted beam.

In this Section dispersion and the refractive index of a plasma is described and the

geometrical optics approximation introduced.

Refraction

Scattering

Reflection Transmission Plasma

54

54

3.2.1 The refractive index of a plasma

The effects of a plasma on an electromagnetic wave propagating through it can be

described using the dispersion relation, and its representation as the refractive index, n.

For the derivation of the dispersion relation for an arbitrary plasma with any magnetic

field the reader is directed to the literature where it is well documented 104,105. In this

section I will focus on the simple case of a homogeneous cold plasma and then illustrate

how the more complex the problem of inhomogeneous plasma is treated. The treatment

will be sufficient to develop solutions applicable to the experimental situations

encountered in this research.

If we consider the plasma to be a continuous conducting media then, the propagation of

an electromagnetic wave must satisfy the relevant Maxwell’s equations: -

t

t

∂∂+=×∇

∂∂−=×∇

EjB

BE

000 εµµ [3.1]

Where E is the electric field intensity vector, B is the magnetic induction vector, µ0 is

the permeability of vacuum, ε0 is the permittivity of vacuum, and j is the total current

density. If we eliminate B from the equations, we get

( ) 0000 =⎟⎠⎞

⎜⎝⎛

∂∂+

∂∂+×∇×∇

ttEjE εµµ [3.2]

By using the vector identity:

EEE 2)()( ∇−•∇∇=×∇×∇ [3.3]

By decomposing the electric field into it’s Fourier components, with k as the wave

number and ω is the angular frequency we obtain: -

∫ −•= ωω ω ddt ti kkExE xk 3)(exp),(),( [3.4]

Then we can analyse each Fourier mode E(k,ω) separately and assume the solution to

be valid for all modes. If we re-write [3.2] using [3.4] for a single Fourier mode, and we

obtain

55

55

0)()(

0)()(

0002

000

=++−•

=++××

EjEkEkk

EjEkk

ωµεµωωµεµω

ii

ii [3.5]

As we know the wave we are investigating is transverse we can state that

0=• Ek [3.6]

Also by substituting for the relation

00

2 1µε

=c [3.7]

We get the result

jEk

EjEk

0

222

2

2

0

2

)(εωω

ωεω

ic

ci

−=−

−= [3.8]

If we consider that the current, j, in the plasma is entirely due to electron motion with

electron velocity ve and electron density ne we can write

evj ene−= [3.9]

Where e is the elementary charge. From the equation of motion for an electron

accelerated by an electric field we can also write

ωe

e

ime

tme

Ev

vE

e

e

=

∂∂

= [3.10]

Where me is the mass of an electron. If we substitute this into [3.9] and then [3.8] we get

2222

2

0

2222

2

)(

p

pe

e

e

e

c

menc

imen

ωω

ωε

ω

ω

=−

==−

−=

k

EEEk

Ej

[3.11]

56

56

This is the simple dispersion relation for non-magnetised plasma, which yields the

refractive index n, and introduces the plasma frequency ωp.

2

2

1ωω

ωpcn −== k

[3.12]

However, we are interested in the refractive index as a function of the electron density,

ne. Hence if we define a critical value for the electron density nc such that at nc, ω =ωp

then we can re-write equation 3.12 as

c

e

ec

nnn

emn

−=

=

1

20

2εω

[3.13]

This result is valid for both uniform non-magnetised media but also for a non-

homogeneous magnetised media if a few constraints are satisfied. Firstly the

assumptions required for a non-homogeneous media require that the properties of the

plasma vary slowly enough that the change over one wavelength is negligible, i.e. that

the local conditions surrounding the wave can be considered uniform. This is called the

WKB (Wentzel, Kramers and Brillouin), WKBJ (and Jeffreys), eikonal or geometrical

optics approximation. In all cases the assumption is:

12 <<∇k

k [3.14]

In this situation the homogeneous relation stands locally and can be integrated over the

ray path. This is equivalent to assuming the wavelength of light is negligibly small.

The constraint for the relations to hold to a first order approximation in a magnetised

plasma is that the direction of the ray propagation is parallel with the direction of the

magnetic field vector. The derivation of a more complete dispersion relation will be

treated later in chapter 4 where polarimetry is described Section 4.4.

57

57

3.2.2 Propagation of light in geometrical optics approximation

If the geometrical optics approximation, the assumption that the wavelength of light, λ,

is negligible, produces reliable results for the propagation of a ray of light, then it is

possible to derive the ray path purely from knowledge of the refractive index through

which the ray passes. This is the principal assumption made here, and the approach uses

Maxwell’s equations in the limit that λ→0.106 The resulting equation is commonly

called the eikonal equation, the basic equation of geometrical optics, and is written as

( ) 22 n=∇ζ [3.15]

Where ζ(x,y,z) is the eikonal function or the ‘optical path’, this a scalar function of

position. The eikonal function is the trajectory orthogonal to the geometrical wave

fronts. The route taken by a ray propagating in the medium of refractive index n can be

deduced as a solution such that ζ = constant. If the position vector r of a point on the

ray path can be described as a function of a length of arc s along the ray path then [3.15]

can be re-written as

sr

ddn=∇ζ [3.16]

However this still specifies the ray path in terms of ζ, which is not defined elsewhere.

On differentiating [3.16] with respect to s we can obtain a solution only in terms of the

refractive index n.

( )

( )

( )

( )[ ]2

2

21

21

1

nn

n

n

dddd

ddn

dd

∇=

∇∇=

∇∇•∇=

∇∇•=

∇=⎟⎠⎞

⎜⎝⎛

ζ

ζζ

ζ

ζ

srss

rs

[3.17]

58

58

Which simplifies to:

nddn

dd ∇=⎟

⎠⎞

⎜⎝⎛

sr

s [3.18]

For a homogeneous media, where n is constant (i.e. ∇n= 0) this reduces again to the

solution:

02

2

=sr

dd

[3.19]

This indicates the curvature is zero and a straight-line path obtained as expected. If the

media has a linearly varying refractive index in one direction, say along the x-axis,

defined by a constant α, then the refractive index is:

xx

ˆ0

αα

=∇+=

nnn

[3.20]

From equation 3.18 the ray path is given as

xs

rxs ˆ)ˆ(

2

2

αα =•dd

[3.21]

This implies a parabolic trajectory in the direction of the increase in refractive index.

3.3 The Vulcan Laser System

The probes described in this chapter were all constructed to use beams provided by the

VULCAN laser107 at the Rutherford Appleton Laboratory.

The VULCAN laser is a Master-Oscillator Power Amplifier (MOPA) high power

Nd:Glass laser with 8 separate beamlines. These beams are divided into six 108mm

diameter beams and two 150mm diameter backlighter beams. In these experiments the

beams are all derived from the same oscillator to remove any timing jitter on a shot to

shot basis. The seed pulse is 80ps FWHM formed by a YAG oscillator every 9 seconds.

59

59

3.4 Probe beams

The properties of the beam used to probe the plasma need to be carefully tailored to the

plasma conditions to be diagnosed. In this section I will discuss the main considerations

for the type of laser beam used to probe the plasmas produced in the experiments

described in Chapters 5 and 6. These plasmas are created by direct drive irradiation at

1054 nm in a 1 mm diameter focal spot, typically from 100nm thick plastic (CH) or

aluminium foil targets. The important characteristics of the plasma produced are an

expansion speed exceeding 107 cms-1, and an average electron density of around 1018

cm-3, and dimensions of approximately 1 mm.

Over the course of work reported in this thesis, three separate experiments have been

performed using the high energy Nd:Glass VULCAN laser at the Central Laser Facility

of the Rutherford-Appleton Laboratory. During this period the design of the probe was

developed and modifications made as a result of both scientific and technical needs, this

development is divided into two sections, in this Section the development of the front-

end, i.e. the probe before it reaches the plasma, in the next section, Section 3.5, the

imaging system is discussed. The diagnostic detail is left until the next chapter.

3.4.1 Utilised probe beam designs

Three separate probe designs were used, with each design developing from the previous

over the course of the experimental program. The three probe front end designs are

outlined schematically in Figure3.2a,b and c respectively.

60

60

a)

1054nm pulse

527nm pulse

532nm CW laserDielectric mirror

KDP Crystal

Timing slide

Screen

CCD Camera

To Target Chamber

From VULCAN

Diffraction GratingLens

Figure 3.2a Probe front-end layout for the first experiment. The essential features are the frequency doubling, and the timing slide .

b)

Timing slide

Screen

CCD Camera

To Target Chamber

From VULCAN

Figure 3.2b Probe front-end layout for the second experiment. Here the Vulcan laser beam is temporally compressed using a grating pair to take advantage of the chirped laser pulses, before frequency doubling.

61

61

Timing slide

Screen

CCD Cameras

From VULCAN

Beamsplitter

Variable aperture

c)

Glan-Taylor Polariser

To Target Chamber

750mm EFL

750mm EFL

200 micron pinholeVacuum Chamber

Figure 3.2c Probe front end layout for the third experiment. In addition to pulse compression, frequency doubling, this design uses a vacuum spatial filter (VSF) as the last optical component before entering the vacuum chamber. The VSF is essential to improving the probe near field.

62

62

3.4.2 Temporal resolution requirements

For the probe to produce a snapshot image the plasma must appear to be stationary over

the duration of the measurement, otherwise motional blurring results in loss of data.

This can be quantified, as the distance the plasma will move over the duration of the

probe pulse should be less than the imaging resolution of the diagnostics.

In the first experiment a simple 80ps pulse from the oscillator was used, over which

time the distance plasma would have moved is 80 µm. This degree of blurring was

apparent on the data produced. From the second experiment onwards, the probe pulse

was compressed using a grating pair in air to a pulse length of 15ps. This is a standard

technique on VULCAN, as this form of temporal pulse compression is one of the stages

of the chirped pulse amplification technique (CPA) which allows VULCAN to deliver

intensities above 1020 Wcm-2. The pulse compression utilises the bandwidth of the pulse

to introduce a change in the length of the beampath for the different wavelengths of the

pulse as they pass through a pair of diffraction gratings as shown in Figure 3.3.

Grating

Grating

Figure 3.3 Schematic of a grating pair arranged for pulse compression

If we consider the pulse to be made of two separate component pulses at slightly

different wavelengths with a temporal separation. Then by adjusting the separation

between the two diffraction gratings we can alter the temporal separation between the

pulses by ensuring that one pulse, or wavelength, will take a slightly longer route.

The experimental arrangement of the grating pair for compression was designed to

match a wavelength ‘chirp’ in the seed pulse to introduce a wavelength difference

between the start and end of the pulse. The gratings were initially aligned using a

532nm HeNe laser along the reference axis of the probe beam, and the second order

diffraction of the HeNe along a calculated beam path. Once the initial alignment is fixed

63

63

the matching of the gratings can be achieved by studying the focal spot of the

compressed beam, as shown in Figure 3.4.

Figure 3.4 The effect of a mismatched grating angle (Top) and the addition of a rotation to the second grating (Bottom) on the focal spot of the CPA beam.

If the gratings are perfectly aligned then the beam should form a circular focus. If the

angle between the two gratings is not matched perfectly to the emerging beam the focus

will be elongated in the direction of the discrepancy. If there is any rotational difference

between the rulings on the two gratings, this will rotate the major axis of the elliptical

focus.

The degree of compression is then purely a function of the distance between the two

gratings varying the path length difference between the front and rear of the pulse. The

limit of the degree of pulse compression in air occurs when the refractive index of the

air becomes dominated by an intensity dependant term108. The pulse length produced is

measured using an optical streak camera, as shown in Figure 3.5.

64

64

0

50

100

150

200

250

-100 -80 -60 -40 -20 0 20 40 60 80 100

Time (ps)

Inte

nsity

(a.u

.)

CPA probe pulse80ps Probe pulse

Figure 3.5 Probe beam pulse shapes for the first experiment (red) and the second and third experiments (blue)

3.4.3 Wavelength selection

The first requirement for selecting an appropriate wavelength to use as a probe for

propagation through the plasma is that the critical density for the wavelength selected is

much greater than the expected electron density in the plasma. This is in order to

minimise refractive effects and to allow simplifications of the treatment of the refractive

index to be valid (see section 3.2.1). As the plasma we are studying is of a relatively

low-density of 1018 cm-3, visible wavelength light at 527nm with a critical density of

4x1021 cm-3 satisfies these criteria. For probing higher densities, X-ray lasers have been

used for interferometric studies109.

The VULCAN beams used for the production of plasma are all at the fundamental

wavelength of the laser at 1054nm. As there will be a large amount of light scattered off

the target at this wavelength it is important that a different wavelength is used to probe

the plasma. The method employed is frequency doubling of the beam by use of a

Potassium-Dihydrogen-Phosphate (KDP) crystal. This process is another standard

technique employed on VULCAN for wavelength conversion, and is typically 50%

efficient.

65

65

When the beam is frequency doubled in the crystal, there is a small difference between

the incident and exiting direction of the beam. This has to be accounted for in any

alignment system.

The drawback of utilising frequency-doubled light is that this harmonic is also

generated by the laser plasma interaction. The possibility of Raman shifting110 the probe

to 622nm using an ethanol cell was investigated experimentally to overcome this

problem. Unfortunately, we were unable to produce a significant degree of conversion

through the Raman cell during our experimental set-up time. However, the level of

harmonic emission measured in the plasma is between two and three orders of

magnitude below the probe beam hence the Raman shift was not required.

3.4.4 Intensity selection

The intensity of the probe beam must be sufficient that the level of self-emission and

harmonic generation from the plasma at the probe wavelength is comparatively

insignificant. Working with high intensity lasers poses a significant problem when using

delicate optics. The optical components used to construct beamlines and diagnostics can

be damaged by the laser if it is sufficiently intense. Manufacturers publish guidelines on

the damage threshold of each component, normally in terms of the power of a

continuous beam or the energy of a standard 1ns pulse.

Unfortunately, the specifications of the probe beamlines utilised in our experiments are

far from the standard pulses used by optics manufacturers. The scaling of damage

threshold with pulse duration has been investigated at VULCAN111 for CPA pulse

lengths similar to the utilised probe designs. It was determined that for a 10 ps pulse the

damage threshold for a coated mirror would be 0.9 J/cm2, and for a 50ps pulse it would

be 1.5 J/cm2. This damage threshold is not an absolute value for damage to occur,

rather it is a value at which damage is significantly likely to occur. Intensities below the

threshold may still cause damage over time. In all three of the experiments performed

the background emission from the plasma is between two and three orders of magnitude

below the probe beam.

66

66

3.4.5 Wavefront quality

Measurements of the near-field intensity pattern of the probe beam during the second

experiment showed that the pulse suffered from severe intensity fluctuations as shown

in Figure 3.6: -

Figure 3.6 Near field intensity pattern of the probe from the second experiment

The overall rectangular shape of the beam is due to apodisation of the beam in the

VULCAN laser hall, which avoids over-filling the diffraction gratings. It is clear that

the near field quality needs to be improved for the any quantitative intensity based

measurements to be feasible. To achieve this a vacuum spatial filter (VSF) was added to

the probe line immediately prior to the target chamber to clean the beam for the third

experiment. This also has the advantage of allowing the shape of the beam to be tailored

to requirement using a variable aperture prior to the VSF, see Figure 3.2c. Any high

spatial frequency distortions due to diffraction around the edge of the aperture will be

removed from the beam.

20mm

67

67

3.5 Imaging Systems

Once the probe has passed through the plasma the beam can then be analysed by an

array of diagnostics. The recording of information from these diagnostics is normally

performed by charge coupled device (CCD) detectors outside the target vacuum

chamber. As the detectors may be several meters from the plasma, and may require the

plasma to be magnified to maximise the resolution at the detector, an imaging system is

normally required.

The design requirements of imaging an optical probe are a compromise between

different factors, most notably the space available and the resolution requirements. Over

the course of the experimental work, three separate experiments were performed with

different imaging systems. In this section, I will describe the systems used and their

capabilities.

In our first experiment, we planned to use a customised triplet lens to form a x4

magnification image. Unfortunately, during vacuum testing of the chamber it was

noticed that the lens was sealed with a non-vacuum compatible sealant. Hence, we had

to make use of the best available alternative at short notice, a 2” achromatic doublet

with a effective focal length (EFL) of 750mm as shown in Figure3.7.

934mm 1166mm

Target

750mm EFL2” Achromat

VacuumChamber

ImagePlane

2584mm

Figure 3.7 First experimental imaging system layout using a simple single lens system. Note that the focus of the laser is contained within the vacuum.

68

68

For our second experiment a high-resolution image was required, but at a lower

magnification of x2.5. A customised f/2.5 spaced doublet lens pair was built based on a

design developed using the Zemax™ optics design package112 in conjunction with the

VULCAN laser staff, this is detailed in Figure 3.8:

284mm 1376mm

Target

108mm diameterf/2.5 doublet

VacuumChamber

ImagePlane

529mm

108mm diameteraspheric f/10

10.4mm

Figure 3.8 Second experimental imaging system using a more complex lens system. Two f/2.5 plano-convex lenses are used inside the chamber with a custom built spacer separating them by 10.4mm The image is then formed by a third lens outside the vacuum chamber.

In our third experiment, the diagnostics required an imaging system that utilised an

expanding beam to form the image, as opposed to the second experimental design

where the image is formed by a slowly converging beam. Diagnostics for this

experiment were to include a self-referencing interferometer. For this design various

catalogue single lens systems were evaluated using the OSLO™ optics package113. The

final design utilised a 3” 750mm EFL NPAC097 lens at x2.5 magnification as shown in

Figure 3.9.

69

69

1047mm 1053mm

Target

750mm EFL3” NPAC097

VacuumChamber

ImagePlane

2599mm

Figure 3.9 Third experimental imaging system again using a simple single lens system and containing the focus within the vacuum chamber.

The different aspects of the performance of these different designs are examined in the

rest of this section.

3.5.1 Resolution requirements

For an imaging system, the resolution limit of the system is the smallest distance

between two point sources that can be distinguished in the image, a rough

approximation is given by Rayleigh’s criterion114. However, this value does not give all

the information required to determine of how well the object is transferred to the image

plane. This information can be derived from the modulation transfer function (MTF) of

the optics, which plots the fractional transmission of a spatial frequency. When the MTF

drops below roughly 0.3 features cannot be resolved. The MTF for the three different

imaging systems used are plotted in Figure 3.10.

1610mm

70

70

a)

b)

c)

Figure 3.10 Modulation Transfer functions for the three imaging systems used. a) 750mm 2” f/10 Achromat at x4 magnification. b) f/2.5 doublet and Aspheric f/10 at x2.5 magnification. c) 750mm 3” f/6 Achromat at x2.5 magnification. Note the change in the horizontal scales.

71

71

The resolution limit, assuming 30% transmission as the cut-off, is calculated from the

magnification and the MTF of the system. This is due to the MTF being calculated in

image space by OSLO™. For the first experimental system this corresponds to a value

of 42 µm, for the second experimental set-up this value is 6.6 µm and for the third it is

44 µm.

This is the resolution limit from the optics, however the resolution of the CCD detector

has to be taken into consideration too. The CCD cells have a finite size, and the smallest

separation that can be resolved as two point sources would be three adjacent pixels.

Three different cell size detectors were used on each experiment, 27.7 µm, 13.3 µm and

7.2 µm cells respectively. These correspond to image space spatial frequency points of

12 cycles/mm, 25 cycles/mm and 46 cycles/mm on the MTF plots in Figure 3.10.

From this we can deduce that in the first experiment the optics are limiting the imaging.

In the second experiment, the detectors are the limiting factor, and in the third, the

optical resolution is approximately the same as the 27.7 µm cameras. Measurements of

the resolution using a test grid for the third experiment produced a resolution of 36 µm

using a 7.2 µm CCD.

3.5.2 Diffraction around target, diffraction limit

Optical resolution is not the only limit on object visibility, as diffraction of the laser

around solid objects will also play a role. To evaluate the effect of diffraction on our

resolution, test images were taken using the second imaging system using a knife edge.

This should produce the diffraction pattern for a semi-infinite screen115: -

21

0

0

2

0

2

220

2

2sin)(

2cos)(

)(21)(

21

2

⎟⎟⎠

⎞⎜⎜⎝

⎛=

⎟⎟⎠

⎞⎜⎜⎝

⎛=

⎟⎟⎠

⎞⎜⎜⎝

⎛=

⎟⎟⎠

⎞⎜⎜⎝

⎛⎥⎦⎤

⎢⎣⎡ −+⎥⎦

⎤⎢⎣⎡ −=

rzv

dwww

dwww

vvI

I

w

w

p

λ

πξ

πς

ξς

[3.1]

72

72

Where I0 is the background intensity, Ip is the diffraction pattern, z is the distance

moved from the edge of the screen in the plane of the screen, λ is the wavelength and r0

is the distance between the object and image planes.

By substituting values for the imaging system the diffraction function can be calculated

and compared with the experimental results. This is illustrated in Figure 3.11

0

10

20

30

40

50

60

100 150 200 250 300

Pixel

Valu

e

Background line-outKnife edge line-outBackground x diffractionDiffraction function x 10

Figure 3.11 The comparison between theoretical and observed diffraction around a semi-infinite screen

The background lineout, in blue, when multiplied by the diffraction function produces

the values of the red line. This red line matches the intensity decay around the knife

edge, in pink. This produces a 20 pixel wide decay instead of a top-hat function.

This implies that when analysing experimental images, data within 20 pixels of a solid

obstacle will be dominated by diffraction off the surface. Note that it is possible to

utilise the fact that at z =0 the value of Ip is 0.25 to determine that the true position of

the knife edge.

The decaying modulations away from the knife edge are reasonably matched in the test

data. Differences may be attributed to the use of a CW HeNe laser to form the image,

73

73

which produces slight time dependent intensity variations in the expanded beam.

Diffraction modulations in experimental data are distinguished by their monotonic

decay parallel to the surface that generated them.

3.5.3 Imaging for Interferometry

If any form of interferometric data is to be taken using the probe it is important that the

imaging system does not distort the wavefront of the beam. The method used to

evaluate this is the optical path difference (OPD). This is a function for each point on

the image plane, which describes the path length difference between all the possible

routes that could be taken by a ray to form the image relative to a reference ray. In

Figure 3.12, the OPD of each of the three imaging systems is plotted for three test

points. The reference ray used by OSLO™ to calculate the OPD is the central ray

through the system.

The OPD plots show that the first system benefits from having a high magnification,

helping to produce a roughly constant OPD over the field with a maximum difference of

0.027 wavelengths. The second and third imaging systems show more variation off axis,

with the second system becoming noticeably distorted at 3mm, the expected maximum

size of the plasma during the experimental timescale.

The maximum OPD of the second design would produce a 0.65 wavelength distortion,

enough to be easily measured by an interferometer. The third imaging system has a

maximum OPD of 0.009 wavelengths, which would require an interferometer with a

resolution of 1/100 of a fringe to resolve.

74

74

a)

b)

c)

Figure 3.12 OPD plots for the three imaging systems, a) 750mm 2” Achromat at x4 magnification. b) f/2.5 doublet and Aspheric f/10 at x2.5 magnification. c) 750mm 3” Achromat at x2.5 magnification. Note the change of colour scale between the three plots.

75

75

3.5.4 Limitations due to refraction

For a measurement using refraction, for example Schlieren imaging, that has occurred in

the plasma, the upper limit of the measurement is governed by the collecting angle of

the imaging system. In all three imaging systems, the limiting aperture in this case is the

first collecting lens. Hence for the first experiment the collection angle is 1.55 degrees,

for the second experiment it is 20.8 degrees and for the third experiment it is 2.08

degrees.

3.6 Conclusions

In this chapter, I have shown propagation of light through a plasma can be treated using

the geometric optics approach and a knowledge of the refractive index of a plasma. This

has been used to develop a beamline for optically probing a plasma suitable for different

experimental requirements. An evaluation of the performance of the implemented

designs has been presented in terms of the resolution of the imaging, the distortion to

the wavefront, the effects of diffraction around the target holder and the limitations on

refraction based measurements.

In our first experiment, we find that the resolution of the system implemented is limited

by the duration of the pulse length. In our second design the resolution is limited by the

diagnostic recording media and not the optical system through the use of a CPA pulse

and custom designed optics. However, this design is not suitable for all types of

diagnostic. The further addition of a VSF to improve the nearfield and a single lens

system was used for the final experiment to allow interferometry and polarimetry

measurements to be made with a resolution comparable with the recording media.

76

76

4 Chapter 4 - Optical probe diagnostics

4.1 Introduction

Following the description of the probe beam and the imaging system in the previous

chapter, the diagnostics used to study how the plasma modifies the probe beam are now

described. These diagnostics include shadowgraphy to produce an ‘image’ of the

plasma, interferometery used to extract quantitative ne measurements, Schlieren imaging

used to extract qualitative measurements of ∇ne, and polarimetry to extract B. The

theory supporting these measurements, the designs, and the techniques used to analyse

the data are described.

4.2 Interferometry

4.2.1 Theory

The basic principle of an interferometer is the measurement of phase differences

between two coherent beams of light. If two beams interfere with each other the sum of

the electric field vectors is observed. For two beams with a phase difference θ which

can be described as

))(exp()exp(

θωω

+titi

2

1

EE

[4.1]

Then on interference the electric fields add to give

)exp())exp(( tii ωθ21t EEE += [4.2]

Which when measured on an intensity based detector leads to a cos(θ) distribution.

When a probe beam propagates for a length l, through medium of refractive index n

(where n≠1), the phase of the beam will be retarded by an amount φ where

∫= ldc

n ωφ [4.3]

77

77

If the probe beam then interferes with a reference beam that has just propagated through

an equal distance in vacuum (where n=1) then the phase difference between the two

beams will be

∫ −= ldc

n ωθ )1( [4.4]

This result is useful as the refractive index of the plasma to a first order approximation

is a function of the electron density of the plasma (see derivation of [3.13])

c

e

nnn −= 1 [4.5]

Which for a value of the electron density much smaller than the critical density can be

approximated to

c

e

nnn

21−= [4.6]

If this result is substituted into the phase difference [4.4] we can see that

∫∫ == ll dncn

dcn

ne

cc

e

22ωωθ [4.7]

Implying that the phase shift is proportional to the path integrated electron density.

Hence by measuring the interference pattern it is possible to evaluate the electron

density of the plasma, ne.

A simple assumption for estimating the interference pattern produced by the plasma is

to assume the path of the ray will be a straight line. This should be approximately valid

for ne << nc, the condition already assumed in this derivation.

For this case the phase difference can be considered as an addition to a background

interference pattern on the interferometer with a given fringe spacing. For a 1mm

diameter cylindrical plasma modelled by Medusa propagating the probe perpendicular

to the cylinder’s axis, the resulting fringe pattern for a x3.6 magnification imaging

interferometer recorded on a 7µm cell CCD camera with a background fringe spacing of

around 60µm should resemble Figure 4.1.

78

78

Target holder

1mm

Figure 4.1 Simulated interference pattern after 500ps showing initial positions of the foil holders, 1mm apart

A more rigorous approach would be to calculate the path length using a ray tracing

procedure to determine the optical path through the plasma. This simulated interference

pattern shows some of the design considerations that need to be considered when

choosing an interferometer for use with an optical probe.

The electron density measurement is based on being able to determine the number of

fringe shifts that occur in a particular part of the image is with respect to the background

interference pattern. This can only be achieved if the recording media can resolve the

fringes. The fringe separation is governed by the type of interferometer used, and is

usually adjustable within a range of values. Ideally, the fringe spacing should be much

larger than the recording unit size of the media being used (grain size of film or pixel

size of CCD). It could also be suggested that the fringe spacing should be larger than the

resolution of the imaging system in the image plane.

As can be seen in Figure 4.1 the distortion of the background interference pattern causes

the fringes to bunch together in regions of high electron density, which leads to a

reduction in fringe visibility. This can be overcome by a suitable choice of

magnification, such that for a given fringe spacing the number of fringe shifts expected

in the interference pattern would be smaller than the image of the object.

Although useful for establishing the geometric parameters of the interferometer design,

the simulated interference pattern assumes that there is perfect fringe contrast over the

whole of the interference pattern. However, when dealing with a short pulse of laser

light the limited coherence of the pulse will reduce the fringe visibility. This is due to

79

79

the polychromatic nature of the source, where if an interference pattern is formed for

one particular wavelength the other wavelength components that form the beam will

gradually slip out of phase and interfere destructively the further away from zero path

difference you go. Hence, the further away from zero path difference a point on the

image lies, the worse the fringe contrast becomes, and in effect limits the number of

fringes that would be visible in the interference pattern.

The coherence length of the probe is governed predominantly by the bandwidth ∆λ of

the pulse for a short pulse probe. The coherence length for a beam of wavelength λ can

be considered to be116: -

λλ∆

≈∆2

tc [4.8]

The values for the system used in our experiments are λ = 527nm and ∆λ = 6nm, this

gives a coherence length of 46µm or a total of 87 visible fringes.

4.2.2 Interferometer Designs

There are several different types of interferometer design that can be used for

measurements in plasmas117. As noted above, the coherence length of the probe is likely

to be very short which requires that the interferometer introduces a zero path length

difference between the reference and data wavefronts. Hence, a form of interferometer

has to be used that is self referencing, to avoid coherence length complications. In our

experiments, we have used Vertical shearing and modified Normarski interferometer

designs. The modified Normarski interferometer is based on the use of a polarisation

splitting prism, in our case a Wollaston prism118. The prism is used to provide a

reference from a section of the beam that is not distorted by the object being studied as

shown in Figure 4.2 below. The main advantage of this design is the simplicity of the

set-up, as there are no delicate alignment issues and the plane in the middle of the

overlapping beam will always have zero path difference.

80

80

Probe pulse45° polarised

Object

WollastonPrism

Polariserset to 45°

0° polarised

90° polarised Interferencepattern wherebeams overlap

Figure 4.2 Principle of a Wollaston prism based interferometer.

The main drawback of this interferometer design is that when used in an imaging mode

the beam should ideally be expanding when forming the image. This is due to the

dependence of the fringe spacing on the relative position of the Wollaston prism with

respect to the Fourier plane of the imaging system and the image plane.

FourierPlane

ImagePlane

Prism

a b

Lens

θ

Figure 4.3 Schematic of an imaging Wollaston prism based interferometer

The fringes produced by such an interferometer can be visualised as being produced by

two virtual point sources as shown in Figure 4.3. Varying the position of the prism with

respect to the image and Fourier planes will alter the separation between the virtual

point sources, and correspondingly alter the fringe spacing observed in the image. The

fringe spacing for a collimated beam is governed purely by the wavelength of the

beam,λ, and the separation angle of the prism, θ. When the ratio of the lengths a and b

in Figure 4.3 is 1:1 then the fringe spacing will be that of the collimated beam.

Otherwise, the fringe spacing, d, is: -

81

81

abd

)sin(θλ= [4.9]

The vertical shearing interferometer works in a similar fashion as shown in Figure 4.4.

Here the reference beam is reflected once more than the test beam, inverting it and

allowing the interference pattern to form. Unfortunately, during experimental testing of

this design we were unable to match the path lengths between the separate arms closely

enough to produce an interference pattern with a CPA pulse from VULCAN.

Probe pulse

Object

Interference patternwhere the beams overlap

Beamsplitter

Mirror

Beamsplitter

Mirror

Mirror

Figure 4.4 Principle of the vertical shearing interferometer

4.2.3 Interferometery Analysis Technique

The basic principle of interferogram analysis is to convert the interferogram into phase

difference information by evaluating the change in the interferogram with respect to the

background interference pattern. Analysing interference patterns can be performed both

manually and with the assistance of a software package.

The changes relative to the reference fringe spacing can be detected by several methods.

These range from performing a Fourier analysis119 on the image to extract changes in

the dominant spatial frequency in areas, to simply counting the changes in intensity

along the path of a reference fringe.

82

82

Manual extraction of phase difference information has the distinct advantage of using

our natural image interpretation abilities to identify the most likely pattern within the

raw data. However, this task is very time consuming especially if two dimensional

phase difference information is required. For this reason, I wrote a software package to

automatically extract the phase difference information from interferograms of the type

produced in our experiments. In this section I will detail the processes used in the

automatic fringe extraction and compare the results with a manual analysis for a sample

interferogram.

The fringe extraction code reads in the interferogram as a bitmap image as produced by

the CCD software. The image may not be perfectly aligned on the camera, but for the

analysis technique used in the code the reference fringes are assumed to run

horizontally. The first stage of the extraction is therefore to rotate the image so that the

reference fringes are horizontal as shown in Figure 4.5.

Raw data

Image is rotatedto make fringeshorizontal1mm

Figure 4.5 Rotating an interferogram to allow fringe extraction. Image a) is the raw data, b) is the image rotated such that the reference fringe is horizontal.

The image rotation is performed using a standard function call to the Graphics32

library120, which uses a bilinear interpolation to construct the rotated image. Once the

reference fringes are horizontal, the background fringe spacing is determined from a

user-defined region of interest where the fringes are unperturbed. In this region, the

83

83

intensity of the pixels in the image is scanned vertically and the average separation

between local maxima and minima is calculated.

When the background fringe spacing is defined the image is then scanned vertically

over the data region for the positions of local maxima and minima using a hill climbing

algorithm. When a maximum or minimum is found the phase of this point is fixed in the

data, and the intervening pixels’ phase is set as a linear interpolation between the

current point’s phase and the previous maximum or minimum’s value. This is chosen to

avoid incorrect assignment of phase between maxima and minima based on intensity

alone.Once the phase of the interferogram is mapped out, the background phase pattern

can then be subtracted to generate the phase difference as shown Figure 4.6

0 1 2 3 4 5 6 7 8 9 10 11

Reference pixel data

Extracted background phase

Background phase

0 1 2 3 4 5 6 7 8 9 10 11

Pixel data

Extracted phase

Phase

Phase comparison

Phase difference

a)

b)

c)

d)

e)

f)

g)

h)

Figure 4.6 The process of fringe extraction to produce a phase difference plot. a) plots the raw pixel values from the reference area. From detection of the positions of the maxima and minima in a) the background phase pattern can be extracted, and plotted in b) and c). A similar process can be applied to the pixel data from the region of interest plotted in d). The extracted phase shown in e) and f) can then be subtracted and compared with the background in g). This leads to a determination of the phase difference between the region of interest and the background which is plotted in h).

84

84

The phase difference can then be converted into an electron density through [4.7] under

the assumption that the plasma is a uniform cylinder. This requires only the length of

the chord that the ray has passed through to be determined. This is determined in the

extraction code by marking where the extremities and centre of the cylinder of plasma

lie on the interferogram and entering the scale of the image. This allows a chord

averaged electron density plot to be produced in two dimensions.

To demonstrate the effectiveness of this technique compared to a manual fringe

extraction the interferogram in Figure 4.7a was analysed using both techniques, the

results of which are plotted in Figure 4.7b.

a)

1mm

b)

0.0E+00

5.0E+18

1.0E+19

1.5E+19

2.0E+19

2.5E+19

3.0E+19

3.5E+19

0 0.05 0.1 0.15 0.2

Distance (cm)

Elec

tron

den

sity

(cm

-3)

Evaluate by handChord average

Figure 4.7 a) Sample interferogram taken from the third experiment of a 100nm CH foil plasma 500ps after irradiation and b) the manual and automatic extraction of the electron density across the centre of the foil.

Both sets of analysis are in agreement within the estimated error for the manual

extraction. It is interesting to note the differences in the performance of the two

85

85

techniques. In the higher density regions the manual extraction produces results that are

more credible as the human eye is able to discern rapid fringe shifts far more accurately

than the analysis code. On the other hand, at low densities the automatic analysis is able

to detect very small shifts in the fringe positions better than the manual extraction. This

is significant as the data we are interested in extracting from the interferograms is in this

low-density expansion region.

4.2.4 Abel inversion

The chord averaged electron density profiles described in the previous section rely on

the assumption that the plasma has a uniform radial density profile. However, this may

not always be the case. If the electron density profile contains a symmetric radial

component this can be extracted from the phase difference data using the technique of

Abel inversion. The Abel transformation converts a chordal integral, such as the phase

difference measurement, into a radial value, the electron density, as described by

Hutchinson121.

As there are often greater than 200 separate measurements of the chordal integral from

the data collected, it was possible to perform Abel inversion on the phase difference

data without having to fit a curve to the data. The test of the Abel inversion method is to

extract an electron density profile where there is known to be a cylindrically symmetric

feature in the plasma. This was achieved by using 100nm thick CH foil targets with a

200µm diameter aluminium dot coated onto its surface. The emission from this plasma

in the direction viewed by the probe beam polarimetry diagnostic in the absence of a

probe pulse and is shown in Figure 4.8.

86

86

1mm

Figure 4.8 Broadband emission from a 200µm diameter Al dot coated onto a 100nm thick CH foil irradiated by two 40J 80ps beams from VULCAN.

The extent of the aluminium portion of the plasma can clearly be seen in the emission as

a tapering spike. Analysis of the interferometery data for this type of target

demonstrates the ability of the Abel inversion technique to extract radial detail from the

data as shown in Figure 4.9

.

1mm

1019cm-3 1018cm-3

b)

1019cm-3 1018cm-3

Figure 4.9 Contour plots of the a) chord averaged data and b) the Abel inverted data. The images plot electron density contours every decade with the peak contour in both images occurring at 1019 cm-3

87

87

Figure 4.9b shows a central tapering density spike in the Abel inverted data of the type

observed in the broadband emission of Figure 4.8, which is not present in the chord

averaged data. Closer examination shows the diameter of this central feature as close to

the initial foil position to be 186 +/- 5µm, very close to the nominal 200µm diameter of

the coated dot.

4.3 Shadowgraphy and Schlieren Imaging

4.3.1 Theory

When light from a probe beam exits inhomogeneous plasma it can be seen from section

3.2.2, that the exit angle of the ray is different to the entrance angle. The angular

deviation is proportional to the integrated gradient of the refractive index of the plasma

over the ray’s path. As already shown in section 3.2.1, the refractive index, to a first

approximation, is a function of the electron density of the plasma. This implies that the

degree of deflection is a function of the electron density gradient. This deflection can be

visualised by dark field Schlieren imaging.

Schlieren imaging is a standard technique for visualising refractive index fluctuations

that would normally be transparent in an image122. The dark field schlieren image is

produced by focussing the probe beam onto a beam stop and imaging the rays that have

been deflected and pass around this beam stop, as shown Figure 4.10. Many varieties of

beam stops can be used, with the most common for light field schlieren imaging being

the knife-edge. This type of stop allows all of the light from one half of the imaged

object to form a background image, on top of which an additional component due to the

deviated rays is added. Dark field imaging uses a circular beam stop so that only the

deviated rays form the resultant image. The degree of deflection that is required for the

ray to arrive at the image plane is equal to the angle subtended by the stop at the focal

length of the lens being used. Hence, the sensitivity of the image is governed by the size

of the stop. The upper limit on the amount of refraction that can be detected is the

limiting aperture of the imaging system, as described in section 3.5.4.

88

88

Plasma

Lens

Stop

Rays refracted sufficiently by plasma do not focus on stop

Image plane

Lens focuses nonrefracted beamonto the stop

Probe beam

Rays passing throughplasma are refracted

Image is formedfrom the refracted

rays only

Non refracted beam

Figure 4.10 Schlieren imaging schematic showing the process whereby an image is formed from rays refracted by the plasma.

Shadowgraphy works on the same principle although there is no stop. For the purpose

of this analysis, shadowgraphs will be treated as schlieren images with an infinitely

small stop.

The schlieren technique is useful for visualising regions of sharp refractive index

change, which relate to areas of high electron density gradient. In an interferometric

measurement this data becomes difficult to recover, as very high electron density

gradients give rise to rapid fringe shifts.

4.3.2 Schlieren Designs

The designs used for our schlieren imaging diagnostics utilise custom manufactured

stops, produced by coating Aluminium through a photo fabricated mask onto an

optically flat glass plate. Multiple stops were coated onto different sections of each plate

allowing rapid alteration of the diagnostic sensitivity. The basic arrangement for each

experiment was to use a 500mm effective focal length (EFL) 2” diameter achromatic

cemented doublet lens to image relay the image plane of the optical probe, and to place

a stop at the focus of the beam. For our second experiment, this technique was applied

in triplicate, as outlined in Figure 4.11, to allow three different stop types to be used on

the same shot. The utilised stop sizes were chosen to give the same diagnostic

sensitivity over two different channels for either frequency doubled or Raman shifted

wavelengths.

89

89

1000mm

500mm

1000mm

500mm

500mm

20%

80% 50%

50%

500mm EFL2” Achromatic Lens

ø1.1mm

ø2.1mm

ø1.6mm

HR @527nm mirror

Beam splitter

Image capture PC

Image plane

Figure 4.11 Schlieren imaging arrangement for second experiment

4.3.3 Ray tracing analysis

When interpreting schlieren images it is not possible to directly infer from an image the

exact magnitude of electron density gradient present in the plasma. It is possible to

deduce the approximate sensitivity of the schlieren imaging system by assuming a

known path length through the plasma and a constant density gradient, although this

will only give an approximate cut-off value. Forming multiple simultaneous images

with different size stops can enable a rough contour map of electron density gradient to

be produced. It is possible to make accurate positional measurements of features in the

images, which can be compared to the scale length of features present in the plasma.

However, a more complete form of analysis would be to form simulated schlieren

images123,124 that can be compared to the experimental data.

If there was an analytical solution to the ray path of a photon through the plasma then a

simulated image could be directly calculated. Simulated schlieren images can also be

formed through iterative ray tracing123 through an assumed plasma profile. Assuming

the deflection of the light passing through the plasma is sufficiently small, then the path

of the ray can be approximated as a straight-line124. For the plasma profile provided by

90

90

our simulations using Medusa I decided that an iterative ray tracing solution would

provide the best solution for the problem.

4.3.4 Code design

The code is designed to iteratively propagate photons through a 1mm diameter cylinder

of plasma with axial profile taken from one-dimensional hydrodynamic simulation data

for the experiment, and a uniform radial profile. The probe beam is simulated by an

array of rays with starting positions evenly distributed in x and y behind the axis of the

cylinder of the plasma. Each ray has an initial velocity of magnitude c in the z direction,

as shown in Figure 4.12.

y x

z

ne

xProbe rays

Rays refract accordingto the electron density

gradient

1mm diameter cylinder ofplasma with axial

electron density profile

Figure 4.12 Geometry of the ray tracing code

The design of the code is based around an iterative loop, where the deflection of the ray

velocity is based on the solution outlined by Decker et al123: -

)(21 2

2

2

engradaλτ

−=∂∂ r

[4.10]

91

91

Where a is a constant of value 8.9 x 10-14cm, r is the position vector, λ is the

wavelength of the ray in cm, and τ is the path length.

For each time step the ray is propagated based on its current position and velocity

vector. The velocity is then deflected according to Decker’s solution and the electron

density gradient of the current position if the ray remains inside the plasma. The ray is

propagated for one period with this new velocity and the loop repeats. The iteration

terminates when the ray has passed through the plasma to a fixed plane in the z direction

or a maximum number of iterations have been reached. The final position, velocity and

angular deflection of the ray are recorded in a data file.

4.3.5 Benchmarking the code.

In order for the results of the ray tracing to be deemed valid, the output should be

checked against a known analytical solution. In this case, an analytical solution derived

by Pert125 was used. This solution is for a spherically symmetric plasma with density

profile given as a power law fall off with distance:

ne /nc = (r/R)-2 [4.11]

Where ne is the electron density at radius r and nc is the critical density for the

wavelength of light which forms a critical density surface at radius R.

92

92

z x

y

1.E-01

1.E+00

1.E+01

1.E+02

1.E+03

0 0.5 1 1.5 2 2.5 3 3.5 4

rne

(nc)

R=1.5

Figure 4.13 Spherical plasma geometry and radial electron density profile with nc at r=1.5

To test the code properly it is necessary to check that it produces a correct result, and

that the iteration process is stable. This is tested using two parameters from the analysis

of Pert, the distance of closest approach, b, and the impact parameter, B. The distance

of closest approach can be evaluated for the traced ray path and compared with the

calculated value given by:

bn(b) = B [4.12]

Where b is the distance of closest approach, n is the refractive index as a function of

radius and B is the impact parameter given by:

n(r) r sin(φ) = B [4.13]

Where φ is the angle between the ray trajectory and the vector position of the ray

relative to the centre of the plasma. The impact parameter B should remain constant

along the ray’s path. A series of 527nm rays with different starting positions were

propagated into the plasma. The value of B was recorded for each iteration, along with

the distance of closest approach to the centre.

93

93

1.0E-08

1.0E-07

1.0E-06

1.0E-05

0.0E+00 1.0E-05 2.0E-05 3.0E-05 4.0E-05 5.0E-05 6.0E-05 7.0E-05

Distance of closest approach(m)

Erro

r (m

)

Figure 4.14 Values of the calculated minimum radius compared with simulation results, using the solution of Decker et al, for a critical density surface at 1.0 x 10-5m

The initial positions of the rays were selected to give a large variation in deflection

angle, ranging from complete reflection at the critical density surface for a normally

incident ray through to a deflection of 10-2 radians for the furthest ray used.

When evaluating the error in the simulation there are two processes at work. Firstly,

during each iteration the velocity of the ray is normalised to the local speed of light and

propagated for a fixed time, and can introduce inaccuracy for large changes in the local

speed of light over one iteration. Secondly, the calculations of the deflection, according

to the iterative solution may be in error. The differences between these two processes

can be seen in the special case of a normal incident ray. In this case, the magnitude of

the error is purely a reflection on the normalisation process. For a normally incident ray,

any error in the degree of deflection will be masked by the normalisation of the velocity

stopping the ray at the critical density surface. For a non-normally incident ray, any

error in the deflection calculation would be expected to fall off in line with the

integrated magnitude of the electron density gradient traversed by the ray. This can be

evaluated for the case of a ray with closest approach at an infinite distance from the

plasma centre, which should experience no error in propagation due to the deflection

calculation. For rays in between these two extremes, the error would be expected to be

approximately a 1/r2 profile; i.e. the error should decrease by an order of magnitude for

a threefold increase in distance of closest approach for a simulation purely limited by

94

94

iterative accuracy. It would also be expected that rays of near normal incidence should

have an error dominated by the normalisation process, producing a value close to that of

the normally incident ray.

The values in Figure 4.14 show that the solution of Decker et al produces a roughly

constant error in the distance of closest approach of 1 x 10-7m. The error for the non-

normal ray is approximately one order of magnitude larger than the normally incident

ray. This is not expected and would suggest an error in the calculation of the amount of

deflection. The accuracy of the calculation can be further investigated by examining the

stability of the impact parameter B throughout the simulation, as shown in Figure 4.15.

1.0E-08

1.0E-07

1.0E-06

1.0E-05

1.0E-04

1.0E-03

1.0E-02

1.0E-01

1.0E+00

0.0E+00 1.0E-05 2.0E-05 3.0E-05 4.0E-05 5.0E-05 6.0E-05 7.0E-05

Distance of closest approach (m)

Var

iatio

n of

B

FinalAverageMaximum

Figure 4.15 Variation of the impact parameter B with respect to the initial value at the start of each ray’s path for the simulation results presented in Figure 4.14

Each ray undergoes thousands of iterations during propagation, only three values are

plotted for each ray in this graph. The ‘Final’ curve in Figure 4.15 indicates the

difference between the value of B in the first and last iterations. The ‘Maximum’ curve

records the peak of the variation of B, normally occurring at the distance of closest

approach. The ‘Average’ curve denotes the mean value of the variation of B over all the

iterations of that ray.

It would be expected that all three curves should have low values, as B is invariant over

the ray path. In addition, for a symmetric ray path the final value should always be

95

95

identical to the initial value, with a certain tolerance for numerical rounding. For

example, given a ray path with 10,000 iterations it would only require a factor of 10-6

numerical error on each iteration to produce a 1% variation in the final value. However,

there will be exceptions due to the iterative nature of the propagation. When the ray is

close to the plasma centre there is a greater uncertainty in determining the angle

between the position vector and the velocity vector when calculating the value of B . As

both vectors are normalised by their magnitude, for small radii positions, the length of

the propagation during the iteration will influence the calculation.

It would be expected that the ‘Average’ curve in Figure 4.15 should follow the same

behaviour as the ‘Final’ curve, if the iteration process is stable and accurate with only a

small number of anomalous values close to the turning point. The ‘Average’ curve

should also be lower than the ‘Final’ curve as the amount of variation is expected to

compound itself with every iteration, if the solution is limited only by numerical

accuracy.

However, Figure 4.15 shows that for Decker’s solution the ‘Average’ curve follows the

behaviour of the ‘Maximum’ curve, indicating that a significant amount of the iterations

differ from the expected behaviour indicated by the ‘Final’ curve. In addition, the

‘Average’ curve is larger than the ‘Final’ curve for distances of closest approach greater

than 5.5 x 10-5m. This suggests that the solution proposed by Decker et al123 is not

correct.

In the treatment derived by Decker et al a constant of 8.9x10-14cm is multiplied by the

square of the wavelength, with no explanation of the source of this constant. The

derivation of the full expression can be derived from Pert’s analysis125. Here we treat

the geometric optics problem as a single particle motion problem where the ratio of the

particles potential energy, V, of the particle to its total energy, E, is given by:

c

e

nn

EV = [4.14]

We can write the change in the particle velocity, v, caused by a variation in electron

density ne over an fixed distance ∆l as:

96

96

llv

⎟⎟⎠

⎞⎜⎜⎝

⎛−−⎟⎟

⎞⎜⎜⎝

⎛ ∆+−

=∆∆

21

21

11c

e

c

ee

nn

cn

nnc

[4.15]

Where c is the speed of light. This can be approximated for small scale length variations

in the electron density, using the geometric optics assumption, to:

llv

⎟⎟⎠

⎞⎜⎜⎝

⎛ ∆

≈∆∆

2c

e

nn

c [4.16]

By converting this velocity gradient into an acceleration we get:

⎟⎠

⎞⎜⎝

⎛∆

∆−=

∆∆

lv e

c

nnnc

t 2

2

[4.17]

As the wavelength dependency of Decker’s solution is included within the calculation

of the critical density, this solution could be re-written in the same format as [4.10].

This would produce a solution with a constant of 8.1705 x10-14cm, rather than 8.9 x

10-14cm. Modifying the ray tracing code to use the solution given in [4.17] produces the

results shown in Figure 4.16: -

1.0E-10

1.0E-09

1.0E-08

1.0E-07

0.0E+00 1.0E-05 2.0E-05 3.0E-05 4.0E-05 5.0E-05 6.0E-05 7.0E-05

Distance of closest approach(m)

Erro

r (m

)

Figure 4.16 Results of the modified ray tracing solution using identical input parameters for comparison with Figure 4.14

97

97

As expected, these results show an identical error value of 10-8m for the normally

incident ray as for Decker’s solution, Figure 4.14. However, the error for rays of near

normal incidence is of the same order of magnitude as that of the normally incident ray,

and the error falls off with an approximately 1/r2 dependency. The variation of B for this

simulation is shown below in Figure 4.17.

1.0E-08

1.0E-07

1.0E-06

1.0E-05

1.0E-04

1.0E-03

1.0E-02

1.0E-01

1.0E+00

0.0E+00 1.0E-05 2.0E-05 3.0E-05 4.0E-05 5.0E-05 6.0E-05 7.0E-05

Distance of closest approach (m)

Var

iatio

n of

B

FinalAverageMaximum

Figure 4.17 Variation of the impact parameter B with respect to the initial value at the start of each ray’s path for comparison with the data presented in Figure 4.15

The magnitude of the ‘Maximum’ curve is reduced by an order of magnitude compared

to the results using Decker’s solution, presented in Figure 4.15. In addition, the

‘Average’ curve follows the behaviour of the ‘Final’ curve and is always lower than the

‘Final’ value.

These results show that the ray tracing solution developed here is both an accurate and

stable method for simulating the passage of light through a plasma of varying electron

density.

4.3.6 Producing simulated schlieren images.

When forming images of an optical probe there is no physical object plane to form an

image from. This is due to the extended size of the plasma. Experimentally, diagnostics

are set to image the central plane of the plasma, defined as the plane z=0 in our

98

98

simulations. To simulate the effect of an extended plasma a separate post processor code

was written to visualise the ray tracing data. The traced rays are read in to the post

processor from the data file and then projected back from their final position to the

plane defined as the object plane along the vector of their final velocity as shown in

Figure 4.18.

y x

z

Rays output from ray tracing code

Initial entry point on CCD

Back propagatedpoint on CCD

CCD cells at centreplane of plasma

Figure 4.18 Back propagation of rays onto a CCD at the assumed object plane

Rays with deflection angles greater than the angle subtended by the schlieren stop but

smaller than the acceptance angle of the imaging system are then allowed to form an

image. The numbers of rays that fall in each cell of a simulated CCD on the object plane

are integrated and a simulated image is formed. This simulated image can then be

compared to the experimental data. It is worth noting that the simulated image will not

be affected by aberrations in the experimental imaging system. The effect of pulse

duration on the image is modelled using multiple ray traces through using different time

steps in the Medusa simulation to vary the plasma profile. The output from these

multiple ray traces can then be integrated on the simulated CCD to form a more

accurate simulated image.

4.3.7 Analysis of the ray tracing

Ray tracing of Medusa simulation data was undertaken as a way to analyse schlieren

image data taken during the experiments. The simulation parameters are outlined in

Figure 4.19

99

99

6040200

18cm

Ne[x10

-3]

-0.1 0.0 0.1Distance [cm]

a)

b)

c)

Figure 4.19 A simulated 1D plasma profile a) derived from the Medusa simulations of two CH foil targets expanding from initial positions at +/- 0.05 cm after 350ps is ray traced using a 527nm 20ps beam and a 1mm diameter cylinder of plasma b). The output of the ray trace is used to construct an ideal image of the centre of the plasma c) for a 1.6mm diameter stop and 500mm lens combination with the imaging system of the second experiment..

The resultant image can be compared with the experimental data with parameters

matching the simulation in Figure 4.20

Figure 4.20 Ray traced image (yellow) overlaid on experimental schlieren data (blue)

100

100

An example of the comparison between the experimental data and the simulated image

shows many similarities, not only in the size and relative positions of the features

observed, but also in the intensity of the features. This intensity correspondence is only

possible through the use of a simulated pulse length.

The comparison with the simulated image allows us to interpret the experimental data in

terms of the model used to provide the plasma profile for the ray tracing. The

assumption that the plasma is expanding in a one-dimensional manner is supported by

the close correspondence between the shape of the simulated and actual intensity

features. If the plasma were not expanding in a one-dimensional manner the assumed

cylindrical nature of the problem would not be valid.

The correspondence between the horizontal intensity variations and position of the

features in the images shows that the evolution of the plasma is being adequately

modelled by the Medusa code. The features in the experimental image that are not

reproduced in the simulated image are outside the capabilities of the Medusa code. As

can be observed there are vertical intensity variations in the features of the experimental

image, implying the existence of fine structure in the plasma. This structure cannot be

modelled by a one-dimensional code and consequently does not appear on the simulated

image. The sources of these vertical intensity variations are discussed in Chapter 5.

4.4 Polarimetry

4.4.1 Theory

A polarimeter is a device used to measure the polarisation state of light. This can be a

useful plasma diagnostic as propagation of light through a plasma in the presence of a

magnetic field alters the polarisation state of the light126.

When there is a magnetic field in a plasma the refractive index of the plasma becomes

related to the magnetic field as well as the electron density. In this situation, we cannot

assume that the only modes of wave propagation are going to be transverse modes as in

section 3.2.1. Hence, the refractive index derivation from the point of equation [3.5],

reproduced here as [4.18], needs to be revised127.

0)()( 000 =++−• EjEkEkk 2 ωµεµω ii [4.18]

101

101

Where E is the electric field intensity vector, j is the total current density, µ0 is the

permeability of vacuum, ε0 is the permittivity of vacuum, k is the wave number and ω is

the angular frequency. If we treat the current as a linear function of the electric field,

from Ohm’s law we can write

),(),(),( ωωω kEkσkj •= [4.19]

Where σ is the conductivity tensor, which when combined with [4.18] leads to

0)(0

2

22 =⎟⎟

⎞⎜⎜⎝

⎛+•+−• EEσEEkk

ωεω ic

k [4.20]

Where c is the speed of light. As can be seen this leads to two types of solution, the

transverse case as derived previously in section 3.2.1 and a longitudinal case. However,

it is also correct to consider the conductivity of the plasma not only as an electric field

effect as in [3.10], but by also considering the force due to the external magnetic field.

Therefore, we should rewrite [3.10] as:

( )t

me e ∂∂=×+ e

evBvE 0 [4.21]

Where B0 is the external magnetic field, me is the mass of an electron, e is the

elementary charge and ve is the velocity of the electron. Assuming B0 is in the z

direction also we can also split the motion into components, in the knowledge that the

Fourier mode of E scales as exp(-iωt) : -

ω

ω

ω

ime

imeeim

ee

e

ze

e

eye

e

exe

z

x

y

y

x

−−

=

−+−

=

−−=

Ev

vBEv

vBEv

0

0

[4.22]

Solving these equations in terms of E gives solutions in terms of the electron gyro-

frequency, ωc, as defined in section 2.2.2

102

102

ec

ze

z

xb

yce

y

yb

xce

x

me

mie

imie

imie

0

2

2

2

2

1

1

1

1

B

Ev

EEv

EEv

=

−=

⎟⎠

⎞⎜⎝

⎛ +−

−=

⎟⎠

⎞⎜⎝

⎛ −−

−=

ω

ω

ωω

ωωω

ωω

ωωω

[4.23]

The conductivity therefore has a matrix representation given by

⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢

−=

•=−=

2

22

2

2

100

01

01

1

1

ωω

ωω

ωω

ωωω

b

b

b

be

e

e

i

i

mein

en

σ

Eσvj e

[4.24]

If we combine this result with [4.20] we get

EEEkk •+−•= εω2

22)(0

ck [4.25]

Where

⎟⎟⎠

⎞⎜⎜⎝

⎛+= Ii σε

0ωε [4.26]

( )

( )

⎟⎟⎟⎟⎟⎟⎟⎟⎟

⎜⎜⎜⎜⎜⎜⎜⎜⎜

⎥⎥⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢⎢⎢

−−

−−−

=

2

2

22

2

22

2

22

2

22

2

100

01

01

ωω

ωωω

ωωωωω

ωωωωω

ωωω

p

c

p

c

cp

c

bp

c

p

i

i

[4.27]

103

103

Where I is the unit identity matrix. This complicated matrix can be simplified in terms

of the dimensionless parameters X and Y defined as

ωωωω

c

p

Y

X

=

= 2

2

[4.28]

⎟⎟⎟⎟⎟⎟⎟

⎜⎜⎜⎜⎜⎜⎜

⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢

−−

−−

−−−

=

XY

XY

iXYY

iXYY

X

100

01

11

011

1

22

22

ε [4.29]

If we then define the axes of k such that

)cos,sin,0( θθ•= kk [4.30]

where θ is the angle between k and B0, and we use the definition of the refractive index

in [3.12] it is possible to solve [4.29] for the refractive index127

( )21

2222

2222

2

cos1sin21sin

211

)1(1

⎥⎥⎦

⎢⎢⎣

⎡−+⎟

⎠⎞

⎜⎝⎛±−−

−−=

θθθ YXYYX

XXn

[4.31]

This expression is commonly known as the Appleton-Hartree formula for the refractive

index. It can be seen from this that if there is no magnetic field, Y goes to zero and the

formula reduces back to [3.12]. If we simplify [4.31] retaining only the first order terms

in Y we get:

θcos12 XYXn ±−= [4.32]

This indicates that with a magnetic field applied the plasma becomes birefringent. If we

consider the case for our experiments were the probe will propagate parallel to the

magnetic field (θ = 0) and that the linearly polarised probe beam is composed of two

104

104

circularly polarised beams which propagate through the plasma with refractive indexes

of n+2=1-X+XY and n-

2=1-X-XY.

It can be seen that due to the difference between the two refractive indexes as the wave

propagates in z there will be a phase difference introduced between the two waves of:

zc

nn ωφ )( −+ −=∆ [4.33]

Hence there will be a change in the angle of polarisation by the process of Faraday

rotation, α, when the waves are superimposed of half the phase difference:

( )z

cX

XY ωφα⎥⎥

⎢⎢

−=∆=

21

121

2 [4.34]

If the plasma is well below critical density X tends to zero so this can be rewritten as

zcm

eB

e

p2

2

21

ωω

α = [4.35]

By substitution for the definition of the refractive index in terms of the electron density

in [3.13] and integrating [4.35] over a path in z we can write:

∫ •= dlBece

ncnm

e2

α [4.36]

Hence, from a measurement of the rotation of the angle of polarisation of the probe

beam and an interferometric measurement of the chord integrated electron density it is

possible to evaluate the magnetic field.

4.4.2 Polarimeter Designs

The diagnostic used to measure the polarisation state of the probe after propagation

through the plasma has to be a single shot device. For this reason a Wollaston prism is

used to separate the probe beam into two divergent, orthogonally polarised beams which

are imaged onto separate areas of a CCD detector.

105

105

Lens

Plasma

WollastonPrism

S-polarised beamvarying intensity

P-polarised beamvarying intensity

Wollaston prism splits beaminto two divergent

orthogonally polarised beams

Probe beamuniform intensity

mixed polarisation

Probe beamuniform intensity

45° Polarised

Image Plane

Figure 4.21 Wollaston prism based polarisation analyser.

This arrangement allows the relative intensity of the two images to be directly compared

and the polarisation angle of each part of the beam to be deduced. This method relies on

the beam having a well characterised polarisation angle prior to entering the plasma, for

this reason a Glan-Taylor polariser with 105 extinction ratio is used in the probe front

end.

4.4.3 Polarisation analysis code

The determination of the polarisation state of the probe beam is performed by an

analysis code which reads in the data from the experimental CCD cameras.

Firstly, the image is calibrated against the measured flat-field profile of the camera.

Then the transposition of one image onto the other image on the CCD is determined

either from user input or from extraction of the autocorrelation of the whole CCD

image. This autocorrelation produces a bright peak at the origin as the self-correlation is

always perfect, but there will also be a secondary maximum at the offset that maps one

image onto the other.

The angle of polarisation of the images produced by the Wollaston prism analyser is

then determined using the method outlined by Lochte and Holtgraven128. As the angle

between the polarisation of the two images recorded on the CCD is 90° then the rotation

angle, α, can be determined as:

106

106

⎟⎟⎠

⎞⎜⎜⎝

⎛+−

= −

21

211sin21

IIIIα [4.37]

Where I1 and I2 are the intensities for corresponding elements of each image. It is

assumed that the sum of these two values is equal to the original intensity of the beam.

This assumption holds provided there is no significant amount of emission from the

plasma, and that the transmission of any optical components used in the diagnostic is

the same for both s and p polarised light. If there is emission from the plasma it should

be of random polarisation, and therefore should not effect the value of I1 – I2, but a

separate measurement of the original intensity of the beam is required.

If there is a difference in transmission of any optical components, calibration of the

diagnostic should enable this effect to be compensated for.

This analysis process is performed for each corresponding element of the two images

that can be mapped out. To increase the accuracy of the result at the expense of spatial

resolution, the intensities used can be taken from individual pixels, or the sum of a 3x3,

5x5, 7x7 or 9x9 rectangle centred around the pixel on a rolling average. For each

calculation that is performed, the statistical error in the calculation due to the random

counting error129 is evaluated. The accuracy of this technique is highest for comparable

values of I1 and I2, hence the initial input angle is set experimentally at 45 degrees.

The accuracy of the polarisation angle calculation improves with the number of events

counted, which directly relates to the dynamic resolution of the CCD cameras used.

This can be seen from the degree of numerical error in the calculation of [4.37] based on

the value of I1 + I2 being constant at 256 counts for an 8-bit CCD and 10,000 counts for

a 16-bit CCD as shown in Figure 4.22.

107

107

0

1

2

3

4

5

6

7

0 15 30 45 60 75 90

Polarisation angle (degrees)

Stat

istic

al e

rror

(deg

rees

)

8 - bit CCD 16 - bit CCD

Figure 4.22 Statistical error in the calculation of equation [4.37] for different CCD dynamic resolutions.

The effectiveness of an imaging polarimeter is shown in Figure 4.23

10 20 26 28

30 32 34 36

38 40 42 44

46 48 50 60

K ey : B lack= 90 , B lue= 55 , G reen= 45 , Y ellow = 35 , R ed= 25, M agent a= 15 , W hit e= 0

Figure 4.23 Polarimeter testing showing the output of the analysis code for various input polarisation angles. The polarograms are colour coded with the determined polarisation angle in the image, as shown in the key.

Input angle

Polarogram

Input angle

Polarogram

Input angle

Polarogram

Input angle

Polarogram

108

108

The input polarisation of the beam was set at the angle noted above each polarogram

using a sheet polariser with an extinction ratio of 103. The polarograms were generated

using an imaging analyser as described in Figure 4.21 and recorded on 8-bit CCD

cameras. As can be seen in Figure 4.23 the polarogram for an reference angle of 10

degrees shows significant amounts of magenta (equivalent to an angle of 15 degrees)

implying an error of approximately 5 degrees, whereas closer to 45 degrees the apparent

error is less an 1 degree. This would appear to validate the error curve for this

arrangement in Figure 4.22.

In the experiment, 16-bit CCD cameras were used to improve the sensitivity of the

diagnostic. From the calculation of the angle of rotation, with the experimental

conditions of a density of around 1019 cm-3 and a 10T magnetic field and a propagation

distance of 1mm a rotation of 0.041 degrees would be predicted. The statistical error in

such a measurement would be 0.0003 degrees. However, the level of background self-

emission from the plasma at the probe intensity used was sufficient to limit the effective

accuracy of the diagnostic to +/- 0.1 degrees. In light of this, the polarimetry diagnostic

was to be used to measure the amount of Faraday rotation introduced by the

experimental magnetic field when the probe propagated through a 1mm thick piece of

SF57 glass.

4.5 Conclusions

In this chapter, I have presented diagnostic methods for measuring the electron density

and magnetic field structure in a plasma and for observing the electron density gradient

in a plasma. This has been achieved through a discussion of the theory behind, and

experimental designs used for Interferometery, Schlieren Imaging, Shadowgraphy and

Polarimetry diagnostics. I have also discussed the analysis techniques used to interpret

the experimental data produced by these diagnostics. This has included the development

of computational tools to assist the analysis. For the analysis of interferometric data, a

code has been developed which enables the extraction of small fringe shifts in the data

more accurately than by eye. The extracted data can also be used in conjunction with an

Abel inversion routine to study the internal structure of cylindrically symmetric plasma.

To assist in the interpretation of Schlieren images an iterative ray tracing code was

developed, which is in agreement with published analytical ray tracing solutions125. This

109

109

code can be used to compare the experimental data with numerical simulations of the

plasma by producing simulated schlieren images.

The analysis of polarimetry data where both s and p polarised images are formed on a

single CCD can be performed automatically using a third code. The transposition

between the s and p polarised images can be determined through autocorrelation or by

user defined input. The polarisation angle is calculated for each point in the image along

with an estimation of the statistical error in the calculation.

110

110

5 Chapter 5 - Producing plasma from a single thin foil

5.1 Introduction

In this chapter the production and characterisation of laser-exploded thin films is

described. The production of a plasma from a thin film by direct laser irradiation is well

established technique and used in many types laser plasma experiments, including X-

ray laser production130, hydrodynamic instability studies131,132 and colliding plasma

inter-penetration experiments133. The definition of thin foil thickness changes depending

on the experiment, in some cases these foils are around a few microns thick, in our case

the foil thickness is typically 100 nm.

These thin foils are necessary as high expansion velocities (>107 cms-1) and low

densities (~1018 cm-3) are required to ensure the interaction between two similar foils in

a colliding geometry is collisionless. This collisionless interaction is discussed in the

next chapter.

Plasma formation from exploding thin foils is sensitive to laser beam intensity non-

uniformity across the focal spot, and the presence of target imperfections in terms of

thickness and density, and surface structures, which density uniformity in the expanding

plasma. To use exploding plasmas for scaled astrophysical simulations it is important to

ensure the plasma is uniform. This is to ensure that the physical processes of interest are

not masked by a highly structured plasma. Furthermore, a clear understanding of the

parameters of the exploding plasma is required before studying the more complex

situation of colliding plasma experiments.

In this chapter I will begin by describing the experimental technique used to produce an

exploding plasma, including how the plasma is diagnosed, and then present data

showing spatial and time resolved measurements of the subsequent expansion. This

experimental data is compared with both an analytical model and numerical simulation

of the plasma.

Density non-uniformities are present in the exploding plasmas and are observed to

persist through the expansion. Sources of plasma non-uniformity are described,

including a discussion of target quality, and then I shall demonstrate that by varying the

111

111

laser focal spot intensity profile that at least part of the observed density non-

uniformities result from the intensity structure in the focal spot. I will introduce the

concept of pre-pulse smoothing and present experimental evidence for the success of

this innovative technique before drawing conclusions.

5.2 Experimental technique

5.2.1 Introduction

These experiments were conducted at the Central Laser Facility using the VULCAN134

laser and Target Area East. The experiments and results described in this chapter were

conducted as part of three colliding plasma experiments conducted during 2000, 2001

and 2002. In the following sub-sections target design, manufacture and the experimental

target area configuration are described.

5.2.2 Target Design and Manufacture

Target design was dictated by the need to produce uniform, large area plasmas suitable

for studying collisionless plasma physics on a kilo-Joule scale laser such as Vulcan.

Laser energy limitations, and Medusa simulations, suggested 1 mm diameter 100 nm

thick foils would produce a rapidly expanding one-dimensional plasma with densities

and dimensions suitable for optical probing. In addition, the target and the holders

needed to be sufficiently small so that they would easily fit in to a mm-scale Helmholtz

coil. These Helmholtz coils are described in Chapter 6.

The target holders were manufactured either by punching them from 250µm thick mylar

sheets or from 50µm thick photo-etched copper. In both cases the holder geometry was

the same, an 8mm by 2mm rectangle with a 1.2mm diameter hole centred 1mm from

one end as shown in Figure 5.1. The thin foils are mounted across the 1.2mm diameter

hole.

112

112

2mm square foil

8mm

φ1.2 mm hole

2mm

Figure 5.1 Design of the target holder used in all experiments

The targets used in our experiments were manufactured by the CLF Target Preparation

group and are constructed from 100nm thick sheets of either C6H8 e-type parylene

(referred to as CH or plastic) or aluminium (Al), cut to size then mounted on a holder.

The CH plastic foils are grown by polymerisation, from bulk material, on to an optical

flat (flat to λ/10 at 637nm, 10 cm2) glass slide coated with detergent that acts as a

releasing agent. The Al foils were manufactured by thermal evaporation, also onto an

optical flat.

After deposition onto the glass, the film is cut into individual 2 mm squares and then

floated off the glass plate in a water bath. The individual foil sections are then mounted

on target holders by hand, being attached by surface tension.

The CH foils were then gently warmed using a hot air blower and allowed to cool. This

additional process results in a smoother foil surface.

5.2.3 Experimental Setup

The foil target is irradiated simultaneously by two overlaid 40 J, 80 ps pulses from the

YLF oscillator of the VULCAN Nd:Glass laser at 1054 nm with a 2 nm bandwidth. The

beams formed a 1mm diameter focus by use of phase zone plates (PZP) or random

113

113

phase plates (RPP) and 108mm diameter f/10 lenses. This results in a nominal intensity

on target of 1014 W/cm2. However, after consideration of the energy losses due to back

reflections and the efficiency of the diffractive optics reduces this to of 3 x 1013 W/cm2.

The main beams were timed to arrive on target within ± 5ps by measuring the scattered

light from a reference target using a streak camera.

In each of the three experiments, an optical probe beam propagating perpendicular to

the plasma expansion was used as shown in Figure 5.2. The design of the probe beam

and imaging system used in each experiment is detailed in Chapter 3.

Probe beam

∅108mm f/10 lensPZP,RPP or defocused∅1mm Spot

Imagingsystem

80ps 1054nm~1014 W/cm2

Pinhole cameras

Optical probediagnostics

100nm thick foil target∅1.2mm

a) b)

VacuumChamber

Figure 5.2 Generic plan view of the target chamber layout for single foil experiments.

The optical probe has been primarily used to study refractive index changes in the

plasma, caused by electron density variations, through a combination of dark field

schlieren imaging and interferometery.

114

114

An interferometer was used during the third experiment, which was a modified

Normarski interferometer as described in section 4.2, imaged onto a 16 bit Andor CCD

with 13 µm square cells. The interferometer was adjusted to produce a reference fringe

pattern with 1 fringe approximately every 10 pixels.

The schlieren images were formed using 500mm focal length f/10 achromatic lenses

and 1.1mm or 1.6mm diameter stops, as described in section 4.3. The resolution of the

schlieren images is governed by the imaging optics and camera combination, which is

discussed in Chapter 3.

The X-ray emission from the foils was monitored using two Pulnix 8-bit CCD X-ray

pinhole cameras mounted inside the target chamber in the second experiment. These

cameras were set to image X-ray emission from the foil with camera a) perpendicular to

the foil, at x3 magnification, and camera b) at 37.5° to the foil surface normal at x5

magnification.

A second optical imaging system was employed to study the scattered focal spot

structure on the foil surface. The scattered light was imaged onto a 16 bit CCD camera,

as shown in Figure 5.3. This imaging system is identical to the one described for the

second experiment in Chapter 3.

115

115

f/10 lensPZP or RPP∅1mm Spot

80ps 1054nm1014 W/cm2

f/2.5f/10

16bit CCD

Focal spot imaging

100nm thick foil target∅1.2mm

Figure 5.3 Focal spot imaging system to study the west foil spot profile during the second experiment.

These focal spot measurements are time integrated, and resolution limited by the 27µm

CCD cell size rather than the optics.

5.3 Results

In this section, I will describe the measurement of the expansion speed, the electron

density profile and the non-uniform structure of the plasma. I will then report the

measurements of the possible sources of these non-uniform structures through

measurements of the laser focal spot profile and the characterisation of the targets.

5.3.1 Expansion

The expansion velocity of the plasma is inferred from the interferograms by locating the

point of furthest detectable fringe shift attributable to the plasma. This process was

performed using the automatically extracted phase shift map, which as described in

section 4.2.3 has the advantage of accurately detecting small fringe shifts. However,

there is still an element of noise in the extracted data introducing an error in locating

this position of +/- 10 pixels. The distances from the foil’s initial position to the leading

edge are plotted against the elapsed time in Figure 5.4.

116

116

0

100

200

300

400

500

600

700

800

900

1000

0 100 200 300 400 500 600 700 800

Time (ps)

Dis

tanc

e (m

icro

ns)

1.55 microns/ps1.3 microns/ps1.13 microns/ps1.51 microns/ps

Figure 5.4 Graph of plasma expansion, showing expansion up to 500 ps at 1.55 µm/ps and up to 750 ps at 1.3 µm/ps

The velocity of the leading edge is assumed to be constant over the expansion. There

may be a residual error in the assumed position of the initial foil surface, hence the

velocities are calculated both assuming the line passes through zero (solid lines) and

linear best fit to the data (dashed lines).

At 750 ps as the plasma density is decreasing during the expansion the leading edge cut

off may have dropped below the sensitivity of the interferometer at this time, making

the later measurements less accurate. Also once the plasma has expanded beyond

500µm the expansion is can longer be considered one-dimensional. For this reason, the

velocity is deduced with this data point (blue lines) and without (purple lines). The

velocities for the expansion up to 500ps provide a good match to a constant velocity,

within the error of the measurement, with only a small correction for the potential

misplacement of the initial foil position. The 750ps data point appears to be erroneous,

as the linear fit falls outside the error of the measurements, and does not fit well with

the zero point. This supports the concept of a one-dimensional expansion up to 500µm

from the foil surface.

117

117

5.3.2 Density Profile

The determination of the electron density profile by provides details of the evolution of

the plasma without perturbing its structure. It also allows us to study the internal radial

structure of the plasma through Abel inversion of interferometric data, as described in

section 4.2.4. This type of measurement is important as it directly provides quantitative

plasma parameter data.

Electron density profiles have been measured for various points in the evolution of a

thin foil plasma by varying the relative delay of the optical probe beam on a timing

slide. Interference patterns were recorded between 250 ps and 1500 ps after the main

drive beams are presented in Figure 5.5.

118

118

Time (ps) Interferogram Extracted Fringe Pattern

350

500

750

Figure 5.5 Interferometric images of 100nm thick CH foils mounted on Cu holders at various time delays after simultaneous irradiation by two ~40J 80ps Vulcan beams focussed to a 1mm spot using PZPs and a 5mm spacer. Note how the fringes close to the initial foil position and are not extracted by the code.

0 5mm

0 5mm

0 5mm

0 5mm

0 5mm

0 5mm

119

119

Samples of the raw data and extracted interference fringes are presented at times

relevant to the colliding system described in Chapter 6.

These results were analysed using the techniques described in section 4.2.4 to provide

Abel inverted electron density profiles at the core of the plasma. The resulting profiles

are shown in Figure 5.6 with the initial foil position at distance 0.

Figure 5.6 Abel inverted electron density profiles for 100nm thick CH foils after 350ps (black), 500ps (blue) and 750ps (red) from the interferograms in Figure 5.5

Referring back to Figure 5.5 it is noted that there exists a region surrounding the initial

foil position where fringes are not discernible. This is believed to be due to refraction of

the probe beam outside of the collecting angle of the imaging system. This lack of data

is reflected in the extracted profiles in Figure 5.6.

The dominant term for error in the analytical process comes from the determination of

the fringe pattern. Assuming the accuracy of fringe extraction to be ± 1 pixel on the

interferogram this leads to a sliding accuracy scale. At low densities, this accuracy

corresponds to an error of ± 0.1 fringe shifts, which would imply an error of ± 4x1017

cm-3. However as the density increases the fringes narrow, until they become

indistinguishable when every other pixel is a different fringe. At this point the error in

measurement can rise to ± 0.5 fringe shifts, or 2 x 1018 cm-3. Taking these values into

consideration, the 4x1017 cm-3 extraction where the measured fringe shift is equal to the

120

120

potential for error the error itself is ±100%. However at the highest densities measured

at ~3x1019 cm-3 the maximum measurement error is only ±6%.

5.3.3 Plasma non-uniformity

An interesting feature observed in the interferometer data is the onset of non-

uniformities in the expanding foil. These can be seen to develop as filaments, which are

first noticeable after 350ps in Figure 5.5 and have been observed in other

experiments135. After 750ps, the distortion caused by the filamentation of the plasma is

quite distinct in the fringe pattern.

Evidence of plasma non-uniformity has been observed in all three experiments. Results

supporting the presence of non-uniformity come mainly from the dark field schlieren

imaging diagnostic as described in section 4.3.

The schlieren images provide a measurement of the scale lengths of the structures that

can be seen, although it is not possible to directly infer the magnitude of the electron

density gradient structure. In this case the measurement is limited only by the resolution

of the optical probe system, which is discussed for each experiment in Chapter 3. In a

uniform thin foil plasma each plasma should produce a schlieren image with two

vertical features, one either side of the target holder. No horizontal structures are

expected from uniform plasma, as is shown by the ray tracing simulations in section

4.3.4.

The results from the first experiment136 showed significant amounts of horizontal

structure in the schlieren images, as shown in Figure 5.7. These structures are observed

to change as the focussing conditions of the laser are altered.

121

121

a) b)

0.5mm

Laser Laser

Figure 5.7 Schlieren images from the first experiment of a) the plasma expanding from a CH foil 750 ps after irradiation by two defocused beams and b) a CH foil 350 ps after irradiation using PZP and a 5mm spacer.

In our second experiment higher resolution schlieren images were taken different target

compositions and focussing conditions. Samples of the schlieren data from this

experiment are presented in Figure 5.8 and in Figure 5.9.

a) b)1 mm

Laser Laser

Figure 5.8 Schlieren images taken during the second experiment 500 ps after the main pulse. Image a) shows a CH foil irradiated using Phase Zone Plates and a 5mm spacer. Image b) shows an identical foil irradiated using Random Phase Plates and a 5mm spacer.

Altering the focussing conditions between PZP and RPP focussing has a dominant

effect on the shape of the expanding plasma envelope, but produces similar scale

structures in both cases.

122

122

a) b)

Laser Laser

1mm

Figure 5.9 Detail of high resolution schlieren images taken during the second experiment, 500ps after irradiation of a) a CH foil and b) an Al foil mounted on a Cu holder using a PZP with 5mm spacer

Changing the target composition introduces different scale structures, as is shown in

Figure 5.9.

1 mm

a)

b)

c)

d)

1 mm

Laser

Laser

Figure 5.10 X-ray pinhole camera images taken by the pinhole camera ‘a)’ in Figure 5.2 perpendicular to the plasma expansion in the second experiment. Image a) is CH plasma and b) is Aluminium plasma. The foils were mounted on CH holders and irradiated using a PZP with 5mm spacer. Images c) and d) are the corresponding schlieren images to a) and b) respectively.

123

123

In this experiment we were also able to show a correspondence between the X-ray

emissions from the plasma and the features observed in the schlieren images, as shown

in Figure 5.10.

5.3.4 Laser focussing conditions

Laser focal spots were formed using RPP and PZP diffractive optics, however flaws in

the PZP and RPP manufacture137 tend to result in a large intensity spike at best focus.

This is undesirable for uniform plasma production. Moving the lens out of best focus by

a known amount defocuses this central intensity spike as shown in Figure 5.11

0 0.5 1 1.5mm5mm SPACER PZP

BEST FOCUS PZP

0 0.5 1 1.5mm

0 0.5 1 1.5mm5mm SPACER RPP

BEST FOCUS RPP

0 0.5 1 1.5mm

a) b)

c) d)

e) f)

g) h)

1mm

Figure 5.11 Normalised focal spot images a), c), e) and f) and the corresponding intensity profiles b), d), f) and g) taken over the central horizontal region for different focussing conditions. The intensity profiles are averaged over a 10 pixel wide strip.

124

124

To determine if there is any dominant feature size within the focal spot envelope both

the RPP and PZP foci were processed through a two dimensional Fourier transform

using the Scion Image138 package. The resultant power spectra were then extracted and

their central horizontal and vertical cross sections plotted in Figure 5.12. As the power

spectra should be symmetric the data is reflected around the central point to highlight

any asymmetry.

a)

PZP Power Spectrum

100

150

200

250

0 10 20 30 40 50

Distance (microns)

Pow

er (a

.u.)

Horizontal LHSHorizontal RHSVertical UpperVertical Lower

b)

RPP Power Spectrum

100

150

200

250

0 10 20 30 40 50

Distance (microns)

Pow

er (a

.u.)

Horizontal LHSHorizontal RHSVertical UpperVertical Lower

Figure 5.12 2D FFT Power spectra for a) PZP focussing and b) RPP focussing

As can be seen, the power spectra for both RPP and PZP are horizontally symmetric but

have vertical asymmetries. This asymmetry may be caused by the focussing geometry,

125

125

where both beams are incident at 35° to the target plane vertically. The PZP focal spot

has a secondary maximum at 30µm, whereas the RPP secondary maxima occurs lower

at 24µm. Given that the resolution of the images is limited by the CCD cells size to a

measurement accuracy of ± 5µm, this would imply a speckle size for the PZP focus of

30 ± 5µm and 24 ± 5µm for the RPP focus. Using f/10 focussing optics the speckle size

for a 1µm wavelength beam using a RPP139 to smooth the beam should be 24.4µm.

5.3.5 Target characterisation

One of our major concerns in the use of the thin foils is the presence of structures in the

targets that can produce density perturbations and directly seed hydrodynamic

instabilities in the plasma. These structures could take the form of both surface and

thickness modulations. To investigate the amount of surface structure several samples

of mounted targets were analysed using a scanning electron microscope. CH and Al

foils were analysed when mounted on 50µm thick Cu and 250µm thick Mylar holders as

described in section 5.2.2. In order for the electron microscope to detect feaures on the

Mylar holders these targets were given a flash coating of Au to make them conductive.

The images of foils mounted on Mylar holders are presented in Figure 5.13, and those

for foils mounted on Cu holders in Figure 5.14

Figure 5.13 Electron micrograph of a 100nm thick CH foil (left) and a 100nm thick Al foil (right) mounted on a 250µm thick Mylar holder

Visual comparison between the two images reveals that the Al foil contains large

perturbations to its surface which appear to be randomly oriented but are of roughly

100µm or greater scale. No such perturbations are visible on the CH foil, except around

the interface between the Mylar holder and the suspended foil, where imperfections

126

126

around the hole edge affect the foil. These holes are punched into the holder leaving

burrs around the edge.

Figure 5.14 Electron micrograph of a 100nm thick CH foil (left) ) and a 100nm thick Al foil (right) mounted on a 50µm thick Cu holder

A similar comparison is made with the Cu mounts in Figure 5.14. These holders are

manufactured by photoetching and are free of the edge defects observed on the mylar

holders. The Al foil surface without a flash coating of Au reveals a multitude of hairline

fractures over the foil surface with an approximate separation of 100 - 200µm between

fractures. At low magnification no structure is observed on the CH foil. Hence Figure

5.15 shows a high magnification image of the boundary where the foil lies over the

holder, as this is the point where most distortion in the CH foils are observed.

Figure 5.15 High magnification electron micrograph of the boundary where a 100nm thick CH foil rests on a 50µm thick Cu holder and a particle of dust has been trapped between the foil and the holder.

127

127

Here it is possible to see the very fine structures in the CH foil visible as several sub

micron width horizontal lines in groups roughly 7µm apart. It is also possible to see the

effect of a surface deformation on the holder, and that the surface tension of the CH foil

draws the surface flat within 20µm.

We are also concerned about variations in the thickness and/or density of the foil

surface once mounted. To investigate this, the CH foils were measured

interferometrically using an imaging Wollaston prism interferometer similar to the one

outlined in section 4.2.2. Sample results are shown in Figure 5.16 below.

Figure 5.16 Laser interferogram taken at 627 nm of a nominally 100nm thick CH foil mounted on a 50µm Cu holder. Data fringes are diagonal, low contrast vertical fringes are an artifact of the optical setup.

Taking the refractive index, n, of the e-type parylene to be approximately 1.639, the

published value for c-type and d-type parylene, then the thickness of the foil, l, can be

deduced from the fringe shift, θ, by the formula: -

( )λ

θ ln 1−= [ 5.1]

The displacement of the central fringe across the foil relative to the background fringe

was determined for a set of 10 sample foils. The fringe spacing was set at 100 pixels per

fringe, with an error in measurement of +/- 2 pixels for the fringe shift. The samples

appeared to be split into two groups of thickness 158 +/- 19nm and 196 +/- 19nm. This

may be due to the foils being manufactured from two different batches of coating. Over

the diameter of the foil itself, no noticeable thickness variation was observed.

128

128

5.4 Discussion

In this section I will discuss the measurements of the evolution of the plasma with the

predictions of the analytical and numerical models described in section 2.6. I will then

investigate the observed plasma non-uniformity in terms of the features that can be

attributed to structure in the target and in the focal spot. I will then discuss the

mechanism by which non-uniformities become imprinted into the plasma and the

requirements of a possible solution to smooth the plasma.

5.4.1 Comparison of Expansion with models

Assuming value for the sound speed of 2 x 105 m/s for the analytical model described in

section 2.6, an analytical density profile can be derived for any time during the

expansion of the plasma. As we are interested in the interaction after 500ps the density

profile, n(x), in arbitrary units for this time was numerically solved using Maple140 to

be:

n(x) = 2.5x104 exp(-1.2x104x2) [5.2]

Where x is the distance in meters. By scaling the density profile by a conversion factor,

a reasonable fit with the expansion of the plasma can be seen in Figure 5.17

Figure 5.17 Graph of the analytical model after 500ps (blue) compared with the Abel inverted interferometry measurements (red) from Figure 5.6

Laser

129

129

The analytical model appears to correctly match the expansion of the foil away from the

laser, however agreement is less good when expanding into the laser beam. The

asymmetry in the plasma due to the laser target interaction is not modelled with this

method, as it is assumed that the target is uniformly heated and then adiabatically

expands. Hence this discrepancy is to be expected, but the near match for the expansion

away from the laser would tend to validate the assumption of a hydrodynamic

expansion.

One-dimensional simulations of the laser target interaction were carried out using the

modified version of the Medusa code, as described in section 2.6. The target geometry

modelled was a 150nm thick CH foil, to be in approximate agreement with the target

thickness measured by interferometery in section 5.3.5. The laser pulse was an 80ps

FWHM gaussian pulse with peak irradiance of 3x1013 W/cm2. The simulation used an

ideal gas equation of state for ions and electrons, and flux limited heat transport. The

simulation output is compared with the Abel inverted experimental electron density

profile in Figure 5.18

Figure 5.18 Graph of Medusa simulation compared with experimental measurements

Laser

130

130

The simulation matches closely the evolution of the electron density profile up to the

point where the one-dimensional assumption becomes tenuous (where the expansion is

greater than half the target width).

The measured values of an expansion speed between 1.3 - 1.55x108 cm/s compare well

with Medusa simulation predictions of an expansion speed away from the laser of

1.46x108 cm/s.

5.4.2 Plasma non-uniformity

In the results gathered from the schlieren data we can separate the possible sources of

the non-uniformity into laser focus intensity profile and target structure. From an

analysis of the scale lengths of the structures observed in the schlieren images it is

possibly to deduce different forms of imprinting observed from both possible sources.

For this reason this section is split into a discussion of the target related features and the

laser related features.

5.4.2.1 Target structure

Evidence of plasma structure resulting from target non-uniformity can be most directly

be produced by comparing the scale of the features observed in the plasma with the

surface scans of the sample mounted targets. The first source of experimental evidence

for comparison is the schlieren image data. In interpreting schlieren images, it is

important to recall from section 4.3 that the dark field method is sensitive to electron

density gradients in any direction. Hence, a single density feature in the plasma would

present itself as two visible features on the schlieren. A comparison of the scale of

structure observed in CH and Al plasmas mounted on Cu holders and irradiated using

PZP focussing is shown in Figure 5.19.

131

131

0

0

0

0

0

0

0

0 0.1 0.2 0.3 0.4 0.5 0.6

Distance (mm)

CH FoilAl Foil

Figure 5.19 Schlieren Image lineouts taken over a 70µm wide vertical section of the data presented in Figure 5.9

With the CH plasma in Figure 5.19, a series of regular features is visible. Thirty sets of

peaks can be identified, with an average separation of 13±2µm. This would imply a

typical scale length for periodic density features of 27.2µm. In section 5.3.5 we have

observed that the scale length of features measured on the CH foil surface when

mounted on Cu holders is 1 to 7µm. This value is below the resolution of the imaging

optics, which implies that these features cannot be observed, unless the width of the

features grows with the expansion. However, as the flow is one dimensional and highly

supersonic this is not expected.

The Aluminium plasma in Figure 5.19 shows several clear density features of 130 up to

200µm. This is consistent with the scale and random variations of the structures

observed in the Aluminium foil electron micrograph in Figure 5.13.

Closer examination of the Aluminium profile in Figure 5.19, shows that this plasma also

exhibits a secondary overlaid pattern of small scale structures, most noticeable between

0.45 and 0.55 mm. The observed secondary pattern is indicative of 28.3µm scale density

features, within the standard deviation of the scale of structure identified in the CH

plasma. The presence of this scale structure in both Aluminium and CH plasma would

imply that this structure is either a systematic feature of the diagnostic or that it is

caused by the laser conditions.

Regions of electron density non-uniformity in the plasma would produce spatial

variations in the amount of electron-ion interactions resulting in the emission of

132

132

bremsstrahlung radiation. As the peak bremsstrahlung emission occurs in the soft x-ray

region for hot plasma, any variation on the x-ray emission could be linked to variations

in the plasma density.

As can be seen from Figure 5.10 there is a significant variation in the x-ray emission

from the Al plasma that correlates with the structures observed in the schlieren images

for the same shot. The central feature in both the x-ray and schlieren data is 400µm

wide, of the same order of magnitude as other Aluminium foil surface structures that

have been observed. The CH plasma produces fewer x-rays than the Al plasma, but the

distribution appears spatially uniform at the resolution of the pinhole camera.

5.4.2.2 Laser focal spot intensity structure

The evidence for laser induced structure comes from the variation in imprinted

structures that occurs when the focal spot changes. Although limited by the 30µm

resolution of the imaging optics the comparison of the results in Figure 5.20 from the

first experiment shows a marked contrast in the scale of the structures in the plasma.

0 20 40 60 80 100 120 140Distance (microns) PZP Defocussed

Figure 5.20 50µm wide vertical lineouts across the structures in the schlieren images of CH foils irradiated by either PZP or defocused beams as presented in Figure 5.7

With the measurements of the structure in the focal spot for RPP and PZP, it is possible

to associate a specific scale length with a type of focussing. The extracted image

lineouts from the second experiment in Figure 5.21 demonstrate two different scale

length patterns.

133

133

Figure 5.21 50µm vertical Lineouts across the structure in the schlieren images presented in Figure 5.8

These results both contain significant amounts of structure, the PZP data indicates an

average separation of 13 ± 2µm between features, implying 26 ± 4µm scale density

features. The RPP structure is on average separated by 10 ± 3µm, implying 20 ± 6µm

scale density structures. As both types of focussing are at the same f-number and focal

length and producing the same spot size similar scale structures would be expected, at

24.4µm.

These values agree within error with the focal spot speckle sizes measured in section

5.3.4 for the PZP focus of 30 ± 5 µm and 24 ± 5 µm for the RPP focus. The structures

common to both CH and Aluminium targets irradiated with PZP at 27 ± 4µm and 28 ±

4µm can therefore also be attributed to the laser focus.

As this structure is not observed when the drive beams are defocused, and is observed

with different target surface conditions, we can attribute this structure with a

characteristic scale of 25 ± 5µm to the imprint by the laser.

5.4.3 Thermal smoothing

In direct drive high-power laser experiments any non-uniformity in the drive laser pulse

produces energy density variations in the absorption region. These modulations are

conducted to the ablation surface where they couple into the ablation pressure. Non

134

134

uniform ablation pressure across the target surface at the onset of irradiation launches a

perturbed shock into the target. The rear surface of the target then becomes deformed by

the shock’s passing, producing “feed-in” perturbations141. In our experiments the targets

are so thin that the shock thickness is comparable with the target thickness, and no

“feed-in” perturbations are produced. This is supported by the accuracy of the analytical

model and hydrodynamic simulations in predicting the expansion of the plasma through

hydrodynamic expansion.

As material is ablated the non-uniform ablation pressure profile will be imprinted on the

plasma as it forms, generating “feed-through” perturbations141. The presence of a

plasma ‘atmosphere’ between the absorption region and the ablation surface allows

thermal diffusion to reduce these feed-through perturbations before they reach the

ablation surface. Both of these types of imprint perturbations in the plasma can seed

hydrodynamic instabilities and are a well-known problem for ICF experiments142.

The degree of feed-through perturbation that reaches the rear surface of the target can be

estimated by an evaluation of the amount of thermal smoothing that occurs. The

effectiveness of thermal smoothing has been shown to depend on the wavelength of the

laser143, the spatial scale of the imprinted modulations and the distance over which

smoothing occurs144. Experimental results for direct drive irradiation at 1054 nm and

1014 W/cm2 show that the degree of imprinting by the laser can be interpreted in terms

of [5.3]. Here a fractional change in ablation pressure p, due to a fractional change in

intensity I is smoothed by a factor n. This factor n is based on the spatial irregularities

in the focus of scale length L and the distance between the critical density surface and

the ablation front, D.

⎥⎦⎤

⎢⎣⎡−=→∆=∆

LDn

IIn

pp π4exp

[5.3]

As both the absorption region and the ablation surface are initially in contact at the front

surface of the target, the plasma that is produced from the front surface of the target is

always imprinted. Over time the distance D between the absorption region and the

ablation surface increases as the target ablates and non-uniform plasma is formed, until

this distance is sufficient for the diffusive thermal transport to homogenise the ablation

pressure. Optical smoothing techniques such as the use of Random Phase Plates145

135

135

(RPP) or Phase Zone Plates146 (PZP) improve the effectiveness of such thermal

smoothing by reducing the spatial scale of the intensity modulations in the focal spot.

Thereby also reducing the separation required between the absorption region and

ablation front for thermal smoothing to be effective. To illustrate the problem of

imprinting on very thin targets even using random phase plates to smooth the laser

beam we can consider the following example: -

For a 100 nm thick plastic foil target irradiated at 1012 W/cm2, Medusa simulations

show that the laser pulse would burn through the target in 13 ps. The distance between

the critical density surface and the ablation front is 0.37µm at this time. If these values

are substituted into [5.3] along with the theoretical speckle size of 24.4µm, then we get

a value for n of 0.82, meaning that virtually no thermal smoothing is occurring before

the target is completely ablated.

Hence, in the production of plasma from thin foils in our experiments, there is

insufficient target material to form a plasma atmosphere capable of effective smoothing

before the target has been completely ablated. This problem has been investigated using

radiatively heated low-density foam buffers before the target147. However, as our target

geometry is designed to match scaling criteria148 this experimental solution is not

suitable.

An experimental solution is required which allows the production of plasma from this

existing geometry without imprinting structures into the plasma produced from the rear

surface of the foil.

5.5 Plasma Smoothing experiment

Producing uniform plasma from thin foils would require a uniform laser focus in the

initial phase of irradiation. This has been achieved in 2002 using incoherent X-rays149,

however in our 2001 experiment we utilised a low energy pre-pulse, which is spatially

filtered immediately before the target.

The intention is to form an ablation surface with a uniform ablation pressure and,

providing there are no target surface structures, then pre-form a homogeneous

plasma.After the critical density surface has been separated from the ablation front, the

target can then be irradiated by non-uniform high energy drive beams, heating the

136

136

plasma and driving a supersonic hydrodynamic expansion. Without an ablation surface

to directly couple the laser non-uniformity into the plasma, imprinting on the plasma

formed at the rear surface of the foil should be limited.

Pre-pulse smoothing investigations were performed during the second experiment. In

the following sub-sections the experimental technique and results are presented, with a

discussion of the findings.

5.5.1 Experimental set up

100nm thick CH foil targets mounted on Cu holders were irradiated by a spatially

filtered pre pulse prior to irradiation by the main drive beams at 1014 W/cm2. Precise

measurements of the energy of the pre-pulse on target after the spatial filter were not

possible, however from calorimeter measurements of the pre-pulse beam prior to

entering the target chamber and estimated pinhole transmission, an on target intensity of

1013 W/cm2 is assumed.

f/10 lensPZP or RPP∅1mm Spot

50µm Pinholef/20 lens

17ps 1054nm

f/2.5

f/10

80ps 1054nm1014 W/cm2

Pinhole cameras15ps 527nm

Optical probe

100nm thick CH foil target∅1.2mm

1013 W/cm 2

Figure 5.22 Experimental chamber layout for pre-pulse experiments

137

137

The experimental arrangement was identical to that shown in Figure 5.2, with the

addition of a 50µm diameter pinhole mounted along the central axis of the foil, 20 mm

behind the target, as shown in Figure 5.22, and a CPA pre-pulse beam.

The pre-pulse for smoothing was derived from the same oscillator as the drive beams to

remove timing jitter. Measurements of the wave front quality of the pre-pulse beam150

show that it is at least 3 times the diffraction limit, therefore with a f/20 focus a 50µm

diameter pinhole is used for the spatial filter. Irradiating the pinhole with a high power

laser pulse will form plasma from the edge of the pinhole. We estimate that the

expansion of this plasma will prevent propagation of light through the pinhole after

10ps.

In order to deliver the maximum amount of energy through the pinhole a CPA pre-pulse

beam was compressed in air using a grating pair to a FWHM of 17±5 ps, in an identical

manner as described in chapter 3 for the optical probe. The combination of pulse

evolution and pinhole closure is designed provide optimum focal spot quality to

coincide with the peak intensity. The pre-pulse was focused onto the pinhole 20mm

from the target through a f/20 lens to generate a smooth 1mm focal spot on the foil. The

pre-pulse was designed to arrive up to 500 ps before the main drive beams, and

synchronised to within 5ps with the arrival of the main beams.

5.5.2 Results

High-resolution dark field schlieren images were taken of the expanding plasma at

350ps after irradiation by the main beams. The schlieren images were formed using

500mm focal length f/10 achromatic lenses and a 1.1mm diameter stop, as described in

Chapter 3, for greatest sensitivity to electron density structures in the plasma. The pre-

pulse delay was varied between 0ps and –100ps with respect to the main beams.

Samples of the schlieren images for various pre-pulse delays are presented in Figure

5.23 along with the temporal evolution of the pulses involved, taken from optical streak

camera measurements.

138

138

a)

b)

c)

d)

0

100

200

300

400

500

600

700

-100 -50 0 50 100

Time (ps)

Inte

nsity

(1011

W/c

m2 )

Main Pulse

Pre-pulse 60psearlyPre-pulse 80psearly

Figure 5.23 Images a) to c) show data from the 1.1mm Schlieren channel taken 350 ps after the main beams irradiated a CH foil target using Phase Zone Plates and a 5mm spacer. Image a) is taken with no pre-pulse. Image b) is taken with the pre-pulse 60 ps early, and image c) is taken with the pre-pulse 100ps early. The graph d) represents the evolution of the pre-pulses in time with respect to the main pulse.

139

139

5.5.3 Discussion

The success of this technique depends critically on the timing of the pre-pulse relative

to the main pulse as can be clearly seen in Figure 5.23. When the pre-pulse arrives

inside the envelope of the main pulse, or when the pre-pulse is not spatially filtered, the

identified PZP imprinted structure persists in the plasma. When the pre-pulse arrives

prior to the main pulse but has not ablated the target fully before the main pulse, then a

reduction of imprinted structure is observed. This would be consistent with a degree of

thermal smoothing, where the pressure modulations are significantly reduced during

transport from the absorption region to the ablation front. However, if the pre-pulse

arrives sufficiently early and completely ablates the foil prior to the main pulse, no

imprinting is observed at the sensitivity of our diagnostics. The upper limit on the

timing window for the pre-pulse is set by the plasma expanding to below critical

density.

5.6 Conclusions

An investigation into the production of plasma from 100nm scale thickness targets

using direct drive laser irradiation has been presented. The effects of target surface

structure and laser focus non-uniformity on the uniformity of the plasma have been

studied and characteristic scale length electron density features in the plasma have been

identified. Electron density profiles for various times in the evolution of the plasma

have been measured using an interferometer. Abel inversion of this data has shown

good agreement with one dimensional simulations and analytical models of the plasma

expansion. Measured leading edge expansion velocities are also in agreement with the

simulations. The ability of a spatially filtered pre-pulse to reduce the imprinted laser

focal spot structure has been demonstrated.

140

140

6 Chapter 6 - Colliding magnetised foils

6.1 Introduction

In this chapter, I will describe the experiments and results of a study aimed at producing

a laboratory collisionless shock. As outlined in Chapter 2, the formation of a

collisionless shock depends upon producing a suitable magnetised plasma and

interacting this plasma in a collisionless system. The approach is to take two similar

plasmas in an opposing geometry and allow them to flow together. The flow velocities,

density and ionisation state are chosen such that there are few collisions between the

counter-propagating plasmas. The role of the magnetic field is to introduce new scale

lengths to the system, which should be smaller than the ion-ion binary collision mean

free path. In addition, if the new scale lengths, or Larmor radii, are smaller than the size

of the system the dynamics of the counter-flowing plasmas are expected to be altered.

Scaling the system to a SNR as discussed in Chapter 2 requires a β of ~400. On

achieving such scaling, physics similar to that of a SNR is expected to be reproduced in

the laboratory.

Of considerable uncertainty, and interest, is the effect of an external magnetic field on

the experimentally created plasma. Experimental results presented here imply a

magnetic field of 7.5T, β of ~3000, does not effect the dynamics of an exploding foil.

Yet in a counter-propagating geometry the interaction between two plasmas is affected

by the presence of the external magnetic field. Details of the interaction between the

plasma and magnetic field are unknown, for example whether the magnetic field

penetrates the plasma.

6.2 Production of a magnetic field

The use of an external magnetic field in our experiments introduces several technical

challenges. From the scaling arguments for the experiment to be relevant to a 100 year

old SNR with the anticipated plasma conditions, the field strength needs to be

approximately 20T. To prevent effects of magnetic field variation on the plasma, as

documented in section 2.2.2, the generated field needs to be spatially and temporally

uniform over the extent of the plasma (1mm3) and duration of the experiment (~1ns). In

141

141

addition, the direction of the field should be perpendicular to the direction of the plasma

expansion.

The method used to generate the field must be compatible with the experimental

geometry required to produce two thin foil plasmas as described in Chapter 5. The

magnitude of the magnetic field should also be controllable to enable investigations of

the effect of magnetic field strength on the interaction. Two methods for magnetic field

generation have been investigated, 1) the use of a laser driven Helmholtz coil target and

2) a traditional pulsed power electromagnet.

6.2.1 Helmholtz coil laser target

Producing a multi-tesla, uniform magnetic field, synchronised with a laser-plasma

experiment can be achieved using a laser beam to generate a source of hot electrons,

which is used to drive an intense current. If the current flows in a loop, a magnetic field

is created. The methodology described here is based around the pioneering work of

Daido et al151,152, which was performed using a 10µm wavelength CO2 laser. The

essence of the method is outlined in Figure 6.1.

142

142

e-

I

B

LaserIλ2 > 1016 W/cm2 µm2

PinholeTypically φ500µm

Mylar SpacerCu Sheet50 µm thick

Plasma

Hot Electrons

Figure 6.1 Schematic of the production of a magnetic field through use of a laser target.

The target shown in Figure 6.1 consists of two metal sheets, separated by an insulator

and connected by a wire loop. The sheet closest to the laser has a 500µm pinhole. A

high irradiance laser pulse (Iλ2 > 1016 W/cm2 µm2) is focussed through this pinhole. This

produces plasma on the second sheet with a portion (up to 30%) of the laser energy

being converted into super-thermal electrons by resonant absorption153,154. These ‘hot’

electrons, with approximate temperature of 15keV, preferentially drift down the density

gradient of the expanding plasma. These electrons are subsequently deposited on the

front stripline, to build up a charge imbalance, somewhat like a parallel plate capacitor.

A return current then flows through the wire loop connecting the striplines, generating a

magnetic field. Once the plasma has expanded across the gap the circuit shorts out and

the flow of the return current decays with a characteristic time of (L/R), where L is the

impedance and R is the resistance.

143

143

In Daido’s work, experimental observations of 60T magnetic fields were reported using

a single wire loop. For our work, we require a uniform magnetic field geometry, which

allowed laser and diagnostic access to the experiment. This resulted in the development

of a Helmholtz coil variation of Daido’s design, suggested by M. M. Notley.

Figure 6.2 Photograph155 of the helmholtz coil target geometry showing the location of two thin foils within the central portion of the magnetic field, and the geometry of the lasers used to drive the experiment (red) and probe the plasma (green)

The VULCAN156 laser operating wavelength is approximately a factor of 10 shorter

than the 10µm wavelength CO2 laser used by Daido et al. It is necessary to maintain the

same irradiance (Iλ2) of 1016 W/cm2 µm2 in order to generate a hot electron population

similar to that reported by Daido et al. To achieve this, a 300J 1ns pulse is focussed

using a f/2.5 lens passing through best focus between the two sheets. The plasma

produced by this method will have a much higher expansion speed than that reported by

Daido et al, which will cause the system to short-circuit during the laser pulse.

The magnitude of the field depends on the amount of current driven around the coils,

which in turn is governed by the amount of hot electrons produced. From the results

published by Daido et al we estimate a current of 50kA per loop. Applying this analysis

to the Helmholtz geometry used in the experiment a field strength of 40T is

estimated157.

144

144

Once the plasma has crossed the separation between the plates, the field should decay

with the characteristic L/R decay time, which is estimated to be 30ns.

The primary measurement of the magnetic field comes from the use of single turn, 1mm

diameter, search coils 158. The L/R response time of the search coil is estimated to be

1ns. These coils measure the rate of change of the field strength at coil position. A value

of the magnetic field strength can be extrapolated from the assumed shape of the field

produced by the target.

-4

-2

0

Vol

ts

403020100-10Time [ns]

-600

-400

-200

0

Volts-ns

Search coil dataFit, FWHM 1 nsIntegrated data

Figure 6.3 Single turn search coil results from our first experiment157 showing the raw search coil data, the integrated valuus of the search coil data, and a fit of a 1ns FWHM pulse (the response time of the search coil) to the search doil data.

These results suggest a 40T field is generated at the centre of the Helmholtz coil, and

that the measured decay time of the field of 30ns is in agreement with the theoretical

L/R value.

However, it is also possible that this magnitude of field will be caused by the laser-

target interaction itself, not current flowing through the coils. During the laser target

interaction, a toroidal magnetic field around the laser axis can be produced159 if there

exists a temperature gradient in the plasma perpendicular to the density gradient. If this

is the case then the field shape used in this calculation is not valid.

The use of this technique for a scaled SNR simulation has to assume that the plasma

that is used to drive the field does not directly affect the thin foils. However,

experimental evidence shows that the presence of the drive plasma significantly alters

the evolution of the thin foil plasmas:

145

145

TargetHolder

HelmholtzCoil

1mm

Fringedistortion

Figure 6.4 Interferogram of two foils in a Helmholtz coil taken 500ps after foil irradiation, and 2.5ns after the Helmholtz coil drive plasma is created.

As can be seen, there is a distortion to the interference pattern, coming from below the

target holder where the drive plasma is generated. Further evidence of this can be seen

in the schlieren data:

0514

0516

With magnetic field

No magnetic field

MagneticFieldDrivePlas ma1mm

Figure 6.5 Schlieren data taken 500ps after foil irradiation. Initial foil positions 1mm apart are shown in white. With a magnetic field additional features are observed propagating from the magnetic field plasma as highlighted.

146

146

For this reason, although it appears that the required magnetic field is attainable, this

method is unsuitable for scaled collisionless shock simulations.

6.2.2 Pulsed power electromagnet

Electromagnets are a traditional method for magnetic field production where variable

magnitude of field is required. However, producing the required 20T fields for a scaled

SNR simulation requires a specialised electromagnet. A pulsed power electromagnet

and a 2kV power supply were used, on loan from Dr Karl Krushelnick of Imperial

College. Instead of a wire wound solenoid, a series of Beryllium-Copper alloy discs

separated by mylar sheets form the coil. In order to permit a laser-plasma experiment to

be performed inside the magnetic field the electromagnet is split into two solenoid

sections with a small insulated spacer providing limited access to the centre of the

magnet as shown in Figure 6.6:

Stainless steel frameBe-Cu alloyelectrodes

10mm tube

Tufnol insulator

Be-Cu alloy discs0.3mm thick

50mm

10mm

Figure 6.6 The construction of the pulsed electromagnet

The limited access allows only one laser beam per foil to be used. The experiment can

be probed using the standard probe beam propagating down the 10mm tube. To test the

uniformity of the field produced by the electromagnet a 1mm thick piece of high Verdet

number SF57 glass was mounted in the centre of the magnet, at the interaction point for

the experiment. The amount of polarisation rotation of the probe beam, α, passing

through the glass is directly proportional to the magnetic field strength, B, the Verdet

147

147

number and thickness of the glass. For SF57 the Verdet number is 1900°/T/m at 530nm.

The rotation was recorded by the polarimeter described in Chapter 4.

5mm

0 20 30 40 50 60 90

Figure 6.7 Polarogram of a glass slide inside the pulsed electromagnet, with the polarisation angle (in degrees) of each pixel represented by it’s colour as shown in the scale.

As can be seen from the data, the variation of the magnetic field over the surface of the

glass is negligible, whereas there is a 9° rotation with respect to the background. This

implies a 7.5T field would be generated from a 1.6kV pulse, the safe continuous

operating level of the magnet.

The temporal variation of the magnetic field was measured using a search coil mounted

in the centre of the magnet, and a Rogowski coil160 mounted around the power supply

cables. The integrated search coil signal provides a measurement of the magnetic field

strength, whereas the Rogowski coil diagnostic measures the amount of current flowing

into the magnet. As it is not possible to place a search coil in the magnet and explode

thin foils at the same time, calibrating the Rogowski coil to the search coil allows a

repetitive measurement of the magnetic field to be made.

148

148

-10

0

10

20

30

40

0.0E+00 5.0E-04 1.0E-03 1.5E-03Time(s)

B(T)

search coil(x100 V)Rogowskicoil (x10 V)

Figure 6.8 Evolution of the magnetic field in the electromagnet as inferred by a 1mm diameter search coil and a Rogowski coil

The time-scale of the magnetic field, of the order of 10-4s, is sufficiently large compared

to the experimental time-scale of 10-9s that the field can be considered uniform.

The measurement using the search coil is in agreement with the polarimetry result, and

can be used to calibrate the Rogowski coil data. However, the maximum field generated

by this magnet is predicted to be 10T when driven by a 2kV pulse, which is half the 20T

field required for the scaling of the experiment. In addition, the access for laser beams

and diagnostics into the magnet is not designed for use with the VULCAN laser,

limiting the experiments to one drive beam per foil. The advantages of the

electromagnet in terms of reproducibility of field strengths and ease of field strength

alteration outweigh the major disadvantages of low field strength and accessibility,

which can be improved with future magnet designs.

6.3 Single foil expansion in a magnetic field

Measurements of the evolution of a plasma produced by laser irradiation of a thin CH

foil immersed in a strong magnetic field were made using the pulsed electromagnet

technique. Although a 7.5T field can be generated, the resulting plasma β of 3000 is

much greater than unity, and theory would therefore predict that the expansion of the

plasma should not be influenced by the presence of the field.

149

149

The experimental configuration is dictated by the use of a pulsed electromagnet to

generate the required magnetic field. Due to the limited access to the target position

only one laser beam can be used to irradiate the foil. To avoid possible issues with

surface currents being induced on the target holders the foils are mounted on Mylar

mounts and suspended from above the magnet on a Mylar rod.

The experimental arrangement is described in Figure 6.9:

Probe beam

∅108mm f/10 lensPZP,RPP or defocused∅1mm Spot

Imaging system

80ps 1054nm~3x1013 W/cm2

Optical probediagnostics

100nm thick foil target∅1.2mm

VacuumChamber

Electromagnet

~7.5T Magnetic field

Figure 6.9 Experimental chamber arrangement for magnetised single foil experiments. Note that only one beam per foil can be used due to the use of the electromagnet.

The electromagnet was used with a 1.6kV pulse triggered 300ms earlier than the

irradiating lasers to generate a 7.5T field around the target position to coincide with the

target irradiation.

The following measurements were made during the third experiment, with the

associated diagnostics and probe capabilities as described in Chapters 3 and 4

respectively. To ensure that the evolution of the plasma is not affected by the presence

of the field interferograms were taken after 750ps as shown in Figure 6.10:

150

150

a)

1.5mm

b)

1.5 mm

Figure 6.10 Interferograms taken 750ps after target irradiation with a) a 7.5T field and b) no magnetic field applied

The corresponding laser energies of 39J with magnetic field in a) and 42J without field

in b) are sufficiently close that the expansion should be nearly identical for both CH

foils. The electron density data is extracted from the interferograms using the

methodology described in section 4.2.4 to produce the Abel inverted electron densities

shown in Figure 6.11. Unfortunately, the thickness of the Mylar holder masks most of

the expansion of the foil towards the laser.

1.00E+17

1.00E+18

1.00E+19

-1.00E-01 -5.00E-02 0.00E+00 5.00E-02 1.00E-01

Distance (cm)

Elec

tron

Den

sity

(cm

-3)

0T Data7.5T Data

Figure 6.11 Abel inverted thin CH foil electron density profiles taken after 750ps

Laser Laser

Laser

151

151

The expansion of the plasmas away from the laser for both the magnetised and non

magnetised data shows a similar expansion profile, implying that the magnetic field

does not appear to affect the hydrodynamics of the expansion.

In interpreting the experimental results presented here, there lies one essential

unanswered issue. The fundamental problem is the penetration of the magnetic field into

the plasma. Here I will address this by examining the consequences of the field

penetrating the plasma and the field being excluded from the plasma on the

experimental results I have presented.

For the case of a single foil immersed in the magnetic field, if we assume that the

magnetic field penetrates the plasma, we have already stated that the plasma β is

sufficiently high that the expansion should not be affected by the presence of the field.

If we now assume the magnetic field is excluded from the plasma, the expansion of a

cylindrical plasma of radius r, would be halted161 after a distance LB, when the pressure

of the excluded magnetic field is equal to the initial pressure of the plasma:

220

rBE

LB = [6.1]

Where E0 is the energy of the plasma and B is the magnetic field. If the value of E0 is

approximately 1J, calculated from the kinetic energy of the expanding plasma for our

experiment, the value of LB is 5cm, much larger than the 0.5mm expansion. From this, it

would seem that both treatments show that the expansion up to the times observed

should be independent of the magnetic field. This simple analysis is supported by our

results.

6.4 Counter propagating exploding foil plasmas

The evolution of two counter propagating thin foil plasmas has been studied during all

three of our experiments, using both CH and Al foils. The foils are mounted faces

parallel, separated by 1mm and irradiated by one or two beams from VULCAN as

shown in Figure 6.12:

152

152

Probe beam

∅108mm f/10 lensPZP,RPP or defocused∅1mm Spot

Imaging system

80ps 1054nm~1014 W/cm2

Faraday Cup

Optical probediagnostics

2x100nm thick foil targets∅1.2mm 1mm separation

VacuumChamber

Figure 6.12 Experimental layout for counter-propagating thin foil experiments

In the case of CH foils, the ion-ion mean free path162 of 3.1cm is greater than the scale

of the experiment, and therefore the two plasmas are expected to interpenetrate.

However, Al has an atomic mass of 27, density of 2.7g/cc and an expected ionisation

state of Al11+, which are higher than the values for CH of 6.5, 1.1g/cc and 3.5

respectively. In comparison to CH foils this reduces the ion-ion mean free path for Al

foils to ~1mm.

The results presented here are taken from the third experiment, where the expansion of

the plasma is diagnosed using optical probing techniques and the ion temperature is

measured using Faraday cups. The Faraday cups act as a time of flight detector for

charged ions.

6.4.1 Counter propagating CH foils

The electron density profile of the CH foil expansion is determined by interferometery.

The evolution of the system between the target holders after 500ps is presented in

Figure 6.13.

153

153

1mm

Figure 6.13 Interferogram of the interaction of two counter propagating foils taken 500ps after irradiation

The electron density profile is extracted using the analysis technique outlined in section

4.2.3. If the two plasmas have interpenetrated then the extracted profile should be very

similar to the summation of the electron density profiles of two single foils, as measured

in Chapter 5. As we can see in Figure 6.14, these profiles are in reasonable agreement.

0.00E+00

5.00E+18

1.00E+19

1.50E+19

2.00E+19

-0.05 -0.04 -0.03 -0.02 -0.01 0 0.01 0.02 0.03 0.04 0.05

Distance (cm)

Elec

tron

Den

sity

(cm

-3)

Colliding foils Single foil Summation of single foil densities

Figure 6.14 Expansion of two counter propagating CH foils situated at –0.5 and +0.5 mm after 500ps compared to the linear addition of the density profiles of two single foils

Laser Laser

154

154

The rate of expansion matches the summation curve, except in the central 200µm region

where the density is slightly increased above the prediction.

The results of the schlieren imaging diagnostic also compare well with the results of ray

tracing a simulated electron density profile reflected around a point 0.5mm from the

initial target, as described in Chapter 4.

1mm

Figure 6.15 Schlieren data taken after 500ps (blue) with the ray-traced simulation overlaid (red)

Some of the features caused by plasma non-uniformity are not reproduced by the ray

tracing as discussed in Chapter 4, but the size, shape and position of the main features

are in good agreement. The comparison also highlights a small rotational miss-

alignment issue with the foil targets, caused by mounting the targets inside the

electromagnet on a rod. The Faraday cup measurements indicate an ion temperature of

approximately 100eV with a small (<10%) fraction heated to above 1keV.

6.4.2 Counter propagating Al foils

The Aluminium foils’ expansion profile is again studied by interferometery. Here we

present an interferogram of two counter propagating Al foils 750ps after irradiation:

155

155

2mm

Figure 6.16 Interferogram of two Al foils taken 750ps after irradiation with the initial foil positions shown in white..

As the mean free path is of the order of the scale length, the plasmas are expected to

behave differently to the CH foils. Comparison of the extracted electron density profiles

with the summation of single foil profiles in Figure 6.17 shows this to be the case.

0.00E+00

5.00E+18

1.00E+19

1.50E+19

2.00E+19

-0.05 -0.04 -0.03 -0.02 -0.01 0 0.01 0.02 0.03 0.04 0.05

Distance (cm)

Elec

tron

Den

sity

(cm

-3)

Colliding foils Single foil Summation of single foil densities

Figure 6.17 Expansion of two counter propagating Al foils situated at –0.5 and +0.5 mm after 750ps compared to the linear addition of the density profiles of two single foils. Note the significantly lower density central plateau in the central 200µm region.

The rate of plasma expansion is lower with two plasmas compared to the summation of

two single plasma curves, and a low-density plateau observed over the central 200µm

Laser Laser

156

156

region. Note that the central plateau is at a density below that of even a single foil

expansion, whereas in the interaction of two CH foils in Figure 6.17 the central density

is higher than the summation of two single foil profiles.

6.5 Magnetised counter propagating plasmas

Investigations into the effect of a strong magnetic field on the evolution of counter-

propagating thin foil CH plasmas were undertaken during the third experiment using the

pulsed power electromagnet technique. As noted in the previous sections, without the

presence of a magnetic field the profile of the evolution of two counter propagating CH

foils matches closely the expansion of two single foils, implying that the plasmas are

interpenetrating.

The experiment was performed with the geometry shown in Figure 6.18 below.

Probe beam

∅108mm f/10 lensPZP,RPP or defocused∅1mm Spot

Imaging system

80ps 1054nm~3x1013 W/cm2

Optical probediagnostics

2x100nm thick foil targets∅1.2mm 1mm separation

VacuumChamber

Electromagnet

~7.5T Magnetic field

Figure 6.18 The experimental arrangement for counter propagating exploding foil experiments in a magnetic field

157

157

Again, due to the constraints of using the electromagnet only one beam per foil is

available for target irradiation. In addition, due to the presence of the electromagnet in

the chamber it was not possible to use the Faraday cups diagnostic to measure the ion

temperature. The main measurement of the plasma came from the interferometery

diagnostic, and is presented in Figure 6.19:

1mm

Figure 6.19 Interferogram of two 100nm thick CH foils immersed in a 7.5T magnetic field 500ps after target irradiation.

The electron density profile is extracted from the interferogram as described in Chapter

4. The resultant profile is plotted in Figure 6.20 and compared with the data for two

foils counter-propagating without a magnetic field.

Laser Laser

158

158

0.00E+00

5.00E+18

1.00E+19

1.50E+19

2.00E+19

-0.05 -0.04 -0.03 -0.02 -0.01 0 0.01 0.02 0.03 0.04 0.05

Distance (cm)

Elec

tron

Den

sity

(cm

-3)

0T Colliding foils 7.5T Colliding foils

Figure 6.20 Comparison of magnetised and non-magnetised colliding foil data after 500ps with initial foil positions situated at –0.5 and +0.5 mm. Note the similarity between this graph and Figure 6.17.

As can be seen from this graph, the expansion of the plasma appears to be retarded

compared to the non-magnetised case. Also there is a central 300µm wide low-density

plateau, reminiscent of the feature observed in the Al foil expansion in Figure 6.18. The

data also contains an unexpected central feature, as shown in Figure 6.21:

1mm

Figure 6.21 Interferometry reference channel image, showing the presence of a central feature between the two foils.

This feature is also observed, but with less clarity on other diagnostics. Similar type

structures were observed at 450ps and 550ps, although with less clarity than this shot. It

is also very interesting to note that a similar feature is observed in the data from a

Helmholtz coil driven experiment, as shown in Figure 6.22:

159

159

1mm

Figure 6.22 Shadowgraph from a Helmholtz coil driven magnetised counter-propagating CH foil experiment taken 500ps after foil target irradiation.

The origin of this feature is uncertain, however it has only been observed on shots with

a magnetic field. In the Helmholtz coil driven data, the feature appears near the foil that

is not perturbed by the magnetic field driving plasma.

6.6 Discussion

Differences are identified between the experimental results for counter-propagating CH

foils with and without the presence of a 7.5T magnetic field. The reasons for these

differences are not certain. However, there are two possibilities that I will explore in this

section. Firstly, the magnetic field is unable to penetrate the expanding plasmas, which

are slowed as the magnetic field is compressed as the two plasmas approach. Secondly

the magnetic field penetrates the leading edge of the plasmas and reduces the interaction

scale length such that the plasmas have increased collisionality.

The first hypothesis is tested by developing a simple model of a retarded expansion. The

second hypothesis is tested by comparison of the magnetised CH experiment with the

non-magnetised Al results, where the ion-ion mean free path is shorter and the plasmas

are known to be more collisional.

If the magnetic field does not penetrate either plasma then the field either will move out

of the gap between the plasmas, or will be compressed between them. The field moving

out of the gap would lead to an identical case to the non-magnetised interaction. This

does not agree with the experimental observations.

Assuming the field is completely compressed from an initial size of a cubic millimetre

between the two plasmas then the effect of the compression can be interpreted as an

160

160

increase of the magnetic field pressure. To gain a feeling for how this may affect the

evolution of the plasma we can use the basic analytical model163 with a small

modification.

The analytical model described in Chapter 2 shows the development of the plasma with

a gaussian profile based on a scale length L evolving at a rate based on the speed of

sound in the plasma, cs. The equation governing the development of L in time is given

as164:

⎟⎟⎠

⎞⎜⎜⎝

⎛⎥⎦

⎤⎢⎣

⎡=⇒

=

0

2

2

ln2LLc

Lt

pdt

LdL

s

ρ [6.2]

Where L0 is the initial foil width and p and ρ are the pressure and density respectively.

For a sound speed cs of 2x105m/s this model has been shown to be in agreement with

the expansion of the plasma away from the laser in Chapter 5. For an expansion after

500ps the value for L is 2mm, roughly twice the observable extent of the plasma.

If we now include the effect of the excluded magnetic field, this introduces an

additional term, representing the effect of the magnetic field being compressed by the

expanding plasmas. We know that when the plasma has expanded to the midpoint of the

system the pinched magnetic field should prevent expansion of the plasma, therefore the

expansion speed is zero. Moreover, we know from the results of a single foil expansion

in a magnetic field that the initial expansion speed is not effected by the magnetic field,

and can therefore assume the initial contribution to be negligible. A simple model is to

assume a linear interpolation between the initial and final points, then we can introduce

a correction to the sound speed based on the scale length L in relation to the maximum

possible expansion distance of the plasma and re-write [6.2] as:

⎟⎟⎠

⎞⎜⎜⎝

⎛⎥⎦

⎤⎢⎣

⎡⎟⎟⎠

⎞⎜⎜⎝

⎛−

=

0max

ln22

1LLc

LL

Lt

s

[6.3]

161

161

Where Lmax is the maximum possible expansion distance of the plasma, 1mm. For

expansion after 500ps, this now predicts L to be 0.95mm.

This is the value predicted by the original model [6.2] for the expansion of the plasma,

without magnetic field, after 260ps. If we now compare the simulated expansion of two

single foils after 250ps with the experimental results we do see some agreement.

0.00E+00

5.00E+18

1.00E+19

1.50E+19

2.00E+19

-0.05 -0.04 -0.03 -0.02 -0.01 0 0.01 0.02 0.03 0.04 0.05

Distance (cm)

Elec

tron

Den

sity

(cm

-3)

Simulated single foil expansion after 250ps 7.5T Colliding foils

Figure 6.23 Comparison of Medusa simulation data after 250ps from 150nm thick CH foils at -0.05cm and +0.054cm with the experimental profile of two CH foils in a 7.5T magnetic field after 500ps.

The numerical simulations used here are the ones shown to be in good agreement with

the single foil expansion in Chapter 5. The simulation appears to correctly predict the

early points of the plasma expansion, but does not explain the central density plateau at

1018 cm-3.

If we now consider the second hypothesis, where the magnetic field has penetrated the

plasma, the problem advances beyond the treatment already given. As determined in

Chapter 2, the ion-ion mean free path of the counter-propagating CH plasmas should be

much greater than the scale of the experiment. However, the introduction of a magnetic

field can localise the particles, reducing the scale length of fluid like behaviour to the

Larmor radius. For an estimated electron temperature of 1keV, the electron Larmor

162

162

radius in a 7.5T magnetic field is less than 25µm . For the ions the problem becomes

more complicated. Multi-fluid simulations165 have shown that it is likely that the C6+

ions and the H+ ions will separate, hence the Larmor radius should be calculated for

each species. Assuming an ion temperature of around 100eV as measured by the

Faraday cups for the non-magnetised experiment the Larmor radius in a 7.5T magnetic

field for a proton is ~280µm and for a carbon ion is ~45µm.

Both of these values are much smaller than the system scale length of 1mm and the ion-

ion mean free path of ~3cm, which would imply that the magnetised plasma should

behave similarly to a collisional fluid.

The easiest way to interpret this is to compare the magnetised plasma interaction with

that of a collision dominated interaction where the mean free path of the counter-

propagating fluids is shorter than the scale length, such as a non-magnetised Al foil

experiment. Due to the slower expansion of the Al plasma the foils have not met after

500ps, hence the best data for comparison is the interaction after 750ps. Although the

results are taken at different times in the evolution of the two systems, a correspondence

of the features in the magnetised CH foils can be seen in the non-magnetised Al foils:

0.00E+00

5.00E+18

1.00E+19

1.50E+19

2.00E+19

-0.05 -0.04 -0.03 -0.02 -0.01 0 0.01 0.02 0.03 0.04 0.05

Distance (cm)

Elec

tron

Den

sity

(cm

-3)

7.5T Colliding CH foils after 500ps Colliding Al foils after 750ps

Figure 6.24 Comparison of the expansion of two magnetised CH foils and two Al foils.

163

163

There are similarities in the profiles of both plasmas. These similarities are that both

curves follow the pattern of a transition from high density >1019cm-3 to a turning point

at around 5x1018cm-3 followed by a reduced density gradient from this point to a flat

plateau at around 1x1018cm-3.

If this similarity is due to the effect of the magnetic field, the question of how the field

can penetrate the plasma arises. The theoretical objection166 to the penetration of the

magnetic field is the high conductivity of the plasma, such that the plasma will exclude

the field. The field may diffuse into the plasma, although not on the timescale of Spitzer

resistivity167 as this would rely on collisions. Instead, it is more appropriate to treat the

problem in terms of Bohm diffusion168:

BohmDL

Bohm

eBekT

BohmD

2

216

=

=

τ [6.4]

For our experiment, with a electron temperature, Te, of ~1keV and magnetic field of

7.5T the value of DBohm is 8.3m2/s. For the worst case for diffusion into our plasma of

scale L = 0.5mm the characteristic time, τBohm, is 15ns. For a τBohm of 500ps, the value of

L is 90µm.

The presence of the large magnetic field can cause electrons within the skin depth

(c/ωP) to drift faster than the speed of sound, leading to ion-acoustic turbulence.

Anomalous resistivity169, where the electrons collide with electric field perturbations in

the ion-acoustic instability, can permit the field to penetrate up to the skin depth, around

130µm.

Another possible mechanism may be the penetration of the field as a ‘shock’170. If the

plasma is magnetised, as ours is due to the initial field being frozen into the plasma as it

expands, then a radial density variation in the expanding plasma would permit fast

magnetic field penetration. The estimated velocity of the field penetration in the axial

direction, Vp is given by:

⎟⎠⎞

⎜⎝⎛∂⎟

⎠⎞

⎜⎝⎛=

necBV yp

18π

[6.5]

164

164

Where ∂y(1/n) is the inverse of the radial density gradient. Although the expansion of

the foil is approximately one-dimensional, there is a radial component to the observed

electron density profiles, as shown in Figure 6.25.

1E+18

1E+19

1E+20

0 0.1 0.2 0.3 0.4 0.5 0.6

Radial distance (mm)

Mag

nitu

de o

f rad

ial e

lect

ron

dens

ity g

radi

ent (

cm-3

mm

-1)

Figure 6.25 Abel inverted radial density gradient profile at the midpoint of two counter propagating CH foils immersed in a 7.5T magnetic field.

For our experiment, taking a radial density gradient of around 1019cm-3mm-1 and a 7.5T

field, this gives a value for Vp of 5.5x107 cm/s. This rate of penetration is slower than

the observed plasma expansion speed of 1.55x108 cm/s. Hence, the field should

penetrate ~166µm the plasma. It is interesting to note that the penetration speed is faster

for lower variations in radial density, and therefore any significant radial non-uniformity

may prevent magnetic field penetration. It is also interesting to note that for the

designed experimental magnetic field strength of 20T the predicted velocity of field

penetration would be 1.46 x108 cm/s, approximately the expansion velocity.

All of these treatments show that the plasma regions involved in the initial phase of the

interaction between two counter-propagating CH foils may be magnetised.

6.7 Conclusions

The experiments described in this chapter were designed to demonstrate the

fundamental processes that underpin the formation of a collisionless shock.

165

165

This has been achieved by demonstrating that we are able to produce high β plasma

from thin foils, as the hydrodynamics of the foil expansion are not affected by the

magnetic field. We have then gone on to demonstrate that plasmas can be produced in a

counter propagating collision free system where the two plasmas interpenetrate.

I have also demonstrated the behaviour of a highly collisional interaction in the same

geometry, and proceeded show that this form of behaviour is akin to that exhibited by

high β plasmas in a counter propagating collision free system with the presence of an

external magnetic field.

I have also discussed whether this observed form of behaviour may also be explained by

the effects of magnetic field compression retarding the flow of the plasma. I conclude

that there is stronger experimental evidence to support the view of magnetic field

penetration, and that there are potential theoretical mechanisms to enable this to occur.

However, this awaits conclusive proof by measurement of the strength of the magnetic

field inside the plasmas, possibly through Faraday rotation or Zeeman splitting.

:

166

166

7 Chapter 7 - Conclusions

In drawing together the body of work presented in this thesis, this chapter provides a

summary of the work and it’s conclusions. In this context, I will discuss the possible

avenues for continuation of this research.

7.1 Conclusions

This thesis presents the beginning of an investigation into the interaction of two rapidly

expanding laser-produced plasmas immersed in a strong magnetic field. From the

outset, our aim has been to produce an experimental system that could be used for

scaled simulation studies appropriate to a 100 year old supernova remnant. The

simulation is a snapshot in time, which does not attempt to model the evolution of the

supernova remnant. This approach is different to that used elsewhere in simulating

young supernovae. Considerable progress towards this goal has been made through a

series of experiments performed using the VULCAN laser at the Rutherford – Appleton

Laboratory.

In Chapter 2 I have shown how a collisionless shock may be formed in a plasma. The

formation of a collisionless system, i.e. where the mean free path of the particles in the

plasma is much greater than the scale of the system, can be achieved in a low density,

low Z counter propagating system where the relative velocities of the plasmas involved

is high. If a magnetic field is present within the system it can introduce fluid like

behaviour by localising the particles in the plasma. The corresponding scale length for

this type of interaction is the Larmor radius of the particles. The further constraint is

that the Larmor radius is less than the system size.

I have also introduced the concept of scaling and shown that by numerical simulation an

experiment can be designed to be similar to a 100 year old supernova remnant by

matching a series of four key dimensionless constants between the two systems. Firstly,

the Euler number of the two systems must be similar, which requires that the system can

be described by the ideal fluid equations. Secondly, the plasma β must also be matched

to relate the effect of the magnetic field in both systems. Thirdly the system must

produce a strong shock, with a Mach number greater than 1. Finally, the system must be

167

167

collisionless, with the mean free path greater than the scale length of the system.

Experimental work has been undertaken to attempt to produce such an experiment.

The principle diagnostic used in this experimental work has been optical probing. An

optical probe providing simultaneous shadowgraphy, interferometery, Schlieren

imaging and polarimetry measurements was used. The development of optical probe

beam design for use in our experiments, along with a suitable imaging system tailored

to the requirements of particular experiments have been presented in chapter 3. The

employment of the utilised optical probe diagnostics has been investigated in chapter 4,

where I have demonstrated the theory behind the diagnostic measurements, and also the

development of techniques for analysing experimental data for interferometery,

Schlieren and polarimetry diagnostics.

The experimental scheme described in Chapter 5 is based upon the laser irradiation of

~100nm thick targets. Optical probe data is presented and shows the formation of a

rapidly expanding, low-density plasma using this technique. The plasma evolution was

found to be in good agreement with both analytical and numerical modelling of the

plasma expansion.

However, measurement illustrated that the plasma produced from such thin foil targets

is susceptible to non-uniformity. The source of the structure was found to be both target

surface structure and imprinting of the laser focal spot profile. To eliminate this, two

approaches were taken. Firstly, improvements were made to the targets through more

uniform film production and target holder developments. Secondly, the laser beam used

to explode the foils was smoothed. Experimentally observed imprinting into the plasma

of structure in the focus of the laser has been found to be limited by the use of a

spatially filtered pre-pulse. From this investigation, and implementing these techniques

it is possible to conclude that a smooth plasma can be created.

The experimental production of uniform thin foil plasma can be accurately designed to

meet scaling criteria through the numerical simulation methods used. In addition, the

behaviour of thin foil plasma has been characterised to an extent that enables

interpretation of thin foil interaction experiments.

In Chapter 6, the experimental use of thin foil plasma in a counter propagating geometry

was used to simulate a collisionless interraction. Measurements with and without a

168

168

magnetic field are compared. Generating the required magnetic field is difficult and two

separate methods of magnetic field production have been experimentally investigated,

the use of a laser-plasma driven Helmholtz coil target and a pulsed power

electromagnet. It was determined that the pulsed power electromagnet was the

preferential experimental solution, despite having a lower field strength generated than

required by the scaling criteria, as the mechanism for generating the field did not

interfere with performing an experiment within the field.

It was observed that the magnetic field did not effect the hydrodynamics of a single

foil’s expansion. In addition, the use of counter propagating CH foils produced a

collisionless system as the two plasmas interpenetrated in agreement with expectation

and confirmed by comparison with the summation of the density profiles of two

separate foils. The affect of a magnetic field on the interaction of two counter

propagating CH foil plasmas was to modify the system such that the observed behaviour

is similar to the interaction of two Al foils without a magnetic field.

The plasma parameters we have been able to produce are not in agreement with the

scaling conditions, and we have not produced a collisionless shock in the experiments

presented here. However, this is mainly due to technological constraints on the

production of a sufficiently strong magnetic field, which can be improved upon in

future experiments.

From this thesis, I can conclude that the presence of a suitable magnetic field could be

interpreted as introducing an effective collisionality to the collisionless system we have

created, producing fluid like behaviour. This is the cornerstone of collisionless shock

formation, and hopefully will lead to the lead to the production of a collisionless shock

with continued developments to the experimental technique.

7.2 Further Work

The experimental program of which the first three experiments are presented in this

thesis is ongoing. At the time of writing, another experiment in this series is underway

utilising an electromagnet designed and constructed within our research group. This

development should bring the parameters of the experiment closer to the designed

criteria for a scaled simulation, and possibly the formation of a collisionless shock.

169

169

There are still outstanding issues that need to be addressed within the current

experimental system. Probably the most fundamental of these is the evidence for

magnetic field penetration into the expanding plasma, which may be investigated

experimentally. A measurement of the strength of the magnetic field inside the plasmas

is required to conclusively determine the role of the magnetic field in the interaction.

This may be possible through Faraday rotation if the magnetic field strength is increased

and a more accurate polarimeter design can be constructed. Alternatively, it may be

possible to look for spectroscopic evidence of field penetration from a Zeeman splitting

diagnostic.

The degree of field penetration may also be investigated indirectly by studying the

proposed mechanisms by which field penetration may occur. For example, if the

proposed method of magnetic field penetration by a shock is to be tested, varying the

radial uniformity of the plasma through varying the focussing techniques used should

affect the rate of field penetration.

With sufficient diagnostic access through the new electromagnet the ion temperature

within the plasma may be measured, possibly by Thomson scattering, and possibly

temporally resolved. If this shows an increase in the ion temperature coinciding with the

interaction of the two plasmas, this would provide strong evidence for the start of shock

formation.

As the initial experimental design was constrained by the laser output of VULCAN it

may be possible in the future to take advantages of proposed upgrades to VULCAN or

use other facilities to produce larger scale length plasmas within a similar one-

dimensional geometry. This would decrease the required magnetic field strength to

localise the particles sufficiently for scaling, which is the current experimental

limitation.

If this experimental series enables us to form a collisionless shock, there is the very

interesting possibility of being able to study the ‘injection problem’ of particles being

accelerated to cosmic rays by collisionless shocks.

170

170

List of references

1 B. Remington et al, Science, 284,1488 (1999)

2 G. Collins et al, Physics of Plasmas, 5, 5, 1864, (1998)

3 J. Wark et al, Physics of Plasmas, 4, 5, 2004, (1997)

4 D. Ryutov et al, Ast. J, 518, 821-832, (2000)

5 D. D. Ryutov and B. A. Remington, Plasma Phys. Contr. Fusion, 44, B407, (2002)

6 S. Lebedev et al, Ast. J, 564,113-119, (2002)

7 S. Rose, Physics World, 54, Apr (1994)

8 K. Shigemori et al, Physical Review E, 62, 6, 8838, (2000)

9 D. R. Farley et al, Phys. Rev. Lett. , 83, 10, 1982, (1999)

10 R. Klein et al, Ast J. Supp 127, 379-383 (2000)

11 R. Dendy and J. Kirk, Plasma Phys. Contr. Fusion, 41, A427-A435, (1999)

12 T. J. M. Boyd and J. J. Sanderson, ‘The Physics of Plasmas’, Cambridge University

Press, (2003), Chapter 5.2

13 M. Brown et al. Ast. J. Lett, 577, L63-L66, (2002)

14 M. A. Malkov and P. H. Diamond, Physics of Plasmas, 8, 4, 2401, (2001)

15 A. R. Bell, Mon. Not. R. astr. Soc. 182, 147, (1978)

16 L. Drury, Adv. Space Res. 15, 8/9, 481-487, (1995)

17 Koyama et al, Nature, 378, 255, (1995)

18 G. P. Zank et al, Physics of Plasmas, 8, 10, 4560, (2001)

171

171

19 M. Scholer, H. Kucharek and K. J. Trattner, Adv. Space Res., 21, 4, 533, (1998)

20 R. Sagdeev and C.Kennel, Scientific Americam, 40, Apr (1991)

21 R. P. Drake, Physics of Plasmas, 7, 11, 4690, (2000)

22 B. H. Ripin et al, Laser Part Beams, 8, 183, (1990)

23 B. A. Remington, Physics of Plasmas, 7, 5, 1641, (2000)

24 H. Takabe et al, Plasma Phys. Contr. Fusion, 41, A75, (1999)

25 Y. P. Zakharov, ‘Collisionless Laboratory astrophysics with Lasers’.

26 R. P. Drake, J. Geophys Research, 104, A7, 14505 (1999)

27 R. P. Drake et al, Physics of Plasmas, 7, 5, 2142, (2000)

28 R. P. Drake et al, Astrophysical Journal Supplement Series, 127, 305, (2000)

29 J. Kane et al, Astrophysical Journal, 478, L75, (1997)

30 K. Shigemori et al, Astrophysical Journal, 533, L159, (2000)

31 C. F. McKee and B. T. Draine, Science, 252, 397, (1991)

32 R. I. Klein et al, Astrophysical Journal, 583, 245, (2003)

33 R. J. Taylor, D. R. Baker and H. Ikezi, Phys. Rev. Lett., 24, 5, 206, (1970)

34 D. W. Forslund and C. R. Shonk, Phys. Rev. Lett., 25, 25, 1699, (1970)

35 D. W. Koopman and D. A. Tidman, Phys. Rev. Lett., 18, 14, 533, (1967)

36 Dean et al, Phys. Rev. Lett., 27, 8, 487, (1971)

37 T. J. M. Boyd and J. J. Sanderson, ‘The Physics of Plasmas’, Cambridge University

Press, (2003), Chapter 6.5.1

172

172

38 T. P. Wright, Phys. Rev. Lett., 28, 5, 268, (1972)

39 Dean et al, Phys. Rev. Lett., 29, 9, 569, (1972)

40 A. Y. Cheung, R. R. Goforth and D. W. Koopman, Phys. Rev. Lett., 31, 7, 429,

(1973)

41 A. R. Bell et al, Physical Review A, 38, 3, 1363, (1988)

42 R. Decoste et al, Phys. Rev. Lett., 42, 25, 1673, (1979)

43 R. A. London and M. D. Rosen, Phys. Fluids, 29, 11, 3813, (1986)

44 P Helander, et al. CLF Annual Report 2000-2001, 58 (2001).

45 M. D. Rosen et al, Phys Rev Lett, 54, 2, 106 (1985)

46 E. A. McLean et al, Phys. Rev. Lett., 45, 15, 1246, (1980)

47 J. Grun et al, Phys. Rev. Lett., 66, 21, 2738, (1991)

48 A. Raven et al, Phys. Rev. Lett., 47,15, 1049, (1981)

49 R. R. Whitlock et al, Phys. Rev. Lett., 52, 10, 819, (1984)

50 J. Grun et al, Phys. Rev. Lett., 53, 14, 1352, (1984)

51 T. R. Boehly et al, J. of Applied Physics, 92, 3, 1212, (2002)

52 S. P. Obenschain et al, Phys. Rev. Lett., 46, 21, 1402, (1981)

53 J. H. Gardner and S. E. Bodner, Phys. Rev. Lett., 47, 16, 1137, (1981)

54 A. J. Cole et al, J. Phys. D., 15, 1689, (1982)

55 S. G. Glendinning et al, Phys. Rev. Lett., 80, 9, 1904, (1998)

56 J. Grun et al. Phys. Rev. Lett., 58, 25, 2672, (1987)

173

173

57 D. S. Montgomery et al, Physics of Plasmas, 3, 5, 1728, (1996)

58 P. T. Rumsby, J. W. M. Paul and M. M. Masoud, Plasma Physics, 16, 969, (1974)

59 Berger et al, Phys. Fluids. B., 3, 1, 3, (1991)

60 Bosch et al, Physics of Fluids, 4, 4, 979, (1992)

61 P.W. Rambo and J. Denavit, Physics of Plasmas, 1, 12, 4050, (1994)

62 O. Larroche, Phys. Fluids B., 5, 8, 2816, (1993)

63 P. W. Rambo and R. J. Procassini, Physics of Plasmas, 2, 8, 3130, (1995)

64 T. S. Perry et al, Astrophysical Journal Supplement Series, 127, 437, (2000)

65 O. Rancu et al. Phys. Rev. Lett., 75, 21, 3854, (1995)

66 C. Chenais-Popovics et al, Physics of Plasmas, 4, 1, 190, (1997)

67 F.F. Chen, ‘Introduction to Plasma Physics’, Plenum Press, (1974) p.3

68 T. M. O’Neil and P. V. Coroniti, Rev. Modern Phys., 71, 2, S404, (1999)

69 R.O. Dendy, ‘Plasma Physics: An Introductory Course’, Cambridge University Press,

(1993)

70 F.F. Chen, ‘Introduction to Plasma Physics’, Plenum Press, (1974)

71 T. J. M. Boyd and J. J. Sanderson, ‘The Physics of Plasmas’, Cambridge University

Press, (2003)

72 A.R. Choudhuri, ‘The Physics of Fluids and Plasmas: An introduction for

Astrophysicists’, Cambridge University Press, (1998)

73 R.O. Dendy, ‘Plasma Physics: An Introductory Course’, Cambridge University Press,

(1993) p.27

74 F.F. Chen, ‘Introduction to Plasma Physics’, Plenum Press, (1974), p.4

174

174

75 A.R. Choudhuri, ‘The Physics of Fluids and Plasmas: An introduction for

Astrophysicists’, Cambridge University Press, (1998) p221

76 R.O. Dendy, ‘Plasma Physics: An Introductory Course’, Cambridge University Press,

(1993), Chapter 1

77 F.F. Chen, ‘Introduction to Plasma Physics’, Plenum Press, (1974) Chapter 2

78 T. J. M. Boyd and J. J. Sanderson, ‘Plasma Dynamics’, Nelson, (1969), Chapter 2

79 A.R. Choudhuri, ‘The Physics of Fluids and Plasmas: An introduction for

Astrophysicists’, Cambridge University Press, (1998), Chapter 10

80 T. J. M. Boyd and J. J. Sanderson, ‘Plasma Dynamics’, Nelson, (1969) p.6

81 C. Chenais-Popovics et al, Physics of Plasmas, 4, 1, 190, (1997)

82 R.Z. Sagdeev and C. F. Kennel, Scientific American, 40, April (1991)

83 Y. B. Zel’dovich and Y. P. Raizer, ‘Physics of Shock Waves and High-Temperature

Hydrodynamic Phenomena’, Academic Press, (1966) Chapter I-2 §13

84 T. J. M. Boyd and J. J. Sanderson, ‘The Physics of Plasmas’, Cambridge University

Press, (2003), Chapter 10.5.2

85 T. J. M. Boyd and J. J. Sanderson, ‘The Physics of Plasmas’, Cambridge University

Press, (2003), Chapter 4.8

86 C. F. McKee and B. T. Draine, Science, 397, April (1991)

87 J.W. Connor and J.B. Taylor, Nuclear Fusion, 17, 5, 1047, (1977)

88 D. D. Ryutov et al, Astrophysical Journal, 518, 821-832, June (1999)

89 D. D. Ryutov et al, Physics of Plasmas, 8, 5, 1804, (2001)

90 T. J. M. Boyd and J. J. Sanderson, ‘The Physics of Plasmas’, Cambridge University

Press, (2003), Chapter 4.2

175

175

91 T. J. M. Boyd and J. J. Sanderson, ‘The Physics of Plasmas’, Cambridge University

Press, (2003), Chapter 3.3.2

92 Plasma Science: From Fundamental Research to Technological Applications,

National Academy Press (1995)

93 S. Lebedev et al, Ast. J, 564,113-119, (2002)

94 R. A. Cairns, ‘Plasma Physics’, Blackie, (1985)

95 S. Eliezier, ‘Interactions of high power lasers with plasma’, Institute of Physics

Publishing, (2002), Chapter 1.

96 G. Tallents, J. Phys. D, 36, R1-R18, (2003)

97 T. J. M. Boyd and J. J. Sanderson, ‘The Physics of Plasmas’, Cambridge University

Press, (2003), Chapter 9.4.4

98 P Helander et al, CLF Annual Report 2000-2001, 58 (2001).

99 RA London and MD Rosen, Physics of Fluids, 29, 11, 3813, (1986)

100 J. P. Christiansen, Computer Physics Communications, 7,271, (1974)

101 N. C. Woolsey et al, CLF Annual report RAL-TR-2000-034, 59, (2000)

102 N. C. Woolsey et al, Physics of Plasmas, 9, 2, 729, (2002)

103 I.H. Hutchinson, ‘Principles of Plasma Diagnostics’ , Cambridge University Press,

(1987) Chapter 7

104 F.F. Chen, ‘Introduction to Plasma Physics’ (1974) p100

105 I.H. Hutchinson, ‘Principles of Plasma Diagnostics’ , Cambridge University Press,

(1987) p87

106 M. Born and E. Wolf ‘Principles of Optics’, Pergamon, (1975) 5th Edition p110

176

176

107 C. N. Danson et al, Opt. Commun., 103, 392, (1993)

108 S. Eliezer, ‘The interaction of high power lasers with plasmas’, Institute of Physics

(2002) Chapter 1

109 L. B. Da Silva et al, Phys Rev Lett, 74, 20, 3991 (1995)

110 R.S. Adrain, E. G. Arthurs and W. Sibbett, Optics Communications, 15, 2, 290,

(1975)

111 D. Neely, Private communication.

112 Zemax, http://zemax.com

113 OSLO Light, Lambda Research Corporation.

114 E. Hecht, ‘Optics’, Addison-Wesley, (1987) Chapter 10.2.6

115 E. Hecht, ‘Optics’, Addison-Wesley, (1987) Chapter 10.3.9

116 M. Born and E. Wolf ‘Principles of Optics’ (1975) 5th Edition p264

117 I. H. Hutchinson, ‘Principles of Plasma Diagnostics’ , Cambridge University Press,

(1987) p96

118 R. Benattar et al, Rev. Sci. Instrum. 50,12,1583 (1979)

119 FRAN, http://www.eng.warwick.ac.uk/~espbc/previous/fran.htm

120 Graphics32, http://g32.org

121 I. H. Hutchinson, ‘Principles of Plasma Diagnostics’ , Cambridge University Press,

(1987) p124

122 T. P. Davies, Optics and Laser technology, February (1981) p37

123 G. W. Decker et al, Applied Optics, 24, 6, 823-828, (1985)

124 R. Noll et al, Applied Optics, 25, 5, 769-774, (1986)

177

177

125 G. J. Pert, Plasma Physics, 25, 4, 387-419, (1983)

126 S. E. Serge, Plasma Phys. Contr. Fusion, 41, R57-R100, (1999)

127 I. H. Hutchinson, ‘Principles of Plasma Diagnostics’ , Cambridge University Press,

(1987) Chapter 4.1.2 p90

128 W. Lochte-Holtgraven, ‘Plasma Diagnostics’, North-Holland Publishing Company,

(1968) p601

129 I. H. Hutchinson, ‘Principles of Plasma Diagnostics’ , Cambridge University Press,

(1987) Appendix 2

130 M. D. Rosen et al, Phys Rev Lett, 54, 2, 106 (1985)

131 R. R. Whithook et al, Phys Rev Lett, 55, 10, 819, (1984)

132 A. Raven et al, Phys Rev Lett, 47, 15, 1049, (1981)

133 C. Chenais-Popovics et al, Phys Plasmas, 4, 1, 190, (1997)

134 C. N. Danson et al, Opt. Commun., 103, 392, (1993)

135 O. Willi et al, IEEE Journal of Quantum Electronics, 17, 9 , 1909, (1981)

136 N. C. Woolsey et al, Physics of Plasmas, 8, 5, 2439, (2001)

137 T. Borland et al, Central Laser Facility annual report 2001/2002, 185, (2002)

138 Scion Image, http://www.scioncorp.com

139 S. G. Glendinning et al, Phys Rev Lett, 80, 9, 1904, (1998)

140 Maple 7, Waterloo Maple Inc., www.maplesoft.com

141 K. Shigemori et al, Phys rev E, 65, 045401-1 (2002)

142 T. R. Boehly et al, Physics of Plasmas, 8, 5, 2331 (2001)

178

178

143 J. H. Gardner and S. E. Booker, Phys Rev Lett, 47, 16, 1137 (1981)

144 W.M. Manheimer, D.G. Colombant and J.H. Gardner, Phys Fluids, 25, 1644, (1982)

145 Y. Kato, et al., Phys Rev Lett, 53, 11, 1057, (1984)

146 R. H. Lehmberg, A. J. Schmidt and S. E. Bodner, J Appl Phys, 62, 7, 2680, (1987)

147 D. Hoarty, A. Iwase, C. Meyer, J. Edwards and O. Willi, Phys Rev Lett, 78, 3322,

(1997)

148 D. Ryutov, R. P. Drake, J. Kane and B. A. Remington, Ap. J, 518, 821, (1999);

D. Ryutov, R. P. Drake and B. A. Remington, Ap. J. Supp, 127, 465, (2000)

149 M. Nishikino et al, Physics of Plasmas, 9, 4, 1381, (2002)

150 I. Ross, Central Laser Facility annual report 2001/2002, 181, (2002)

151 H. Daido et al, Japanese Journal of Applied Physics, 26, 8, 1290-1295 (1987)

152 H. Daido et al, Phys. Rev. Lett., 56, 8, 846, (1986)

153 D. W. Forslund, J. M. Kindel and K. Lee, Phys. Rev. Lett., 39, 5, 284, (1977)

154 K. Estabrook and W. L. Kruer, Phys. Rev. Lett., 40, 1, 42, (1978)

155 N. C. Woolsey, R. A. D. Grundy et al, CLF Annual Report RAL-TR-2000-034, 42,

(2000)

156 C. N. Danson et al, Opt. Commun., 103, 392, (1993)

157 N. C. Woolsey R. A. D Grundy et al, Physics of Plasmas, 8, 5, 2439, (2001)

158 G. Decker and D. L. Honca, Journal of Physics E, 5, 484 (1972)

159 S. Eliezier, ‘Interactions of high power lasers with plasma’, Institute of Physics

Publishing, (2002), Chapter 8

179

179

160 I.H. Hutchinson, ‘Principles of Plasma Diagnostics’ , Cambridge University Press,

(1987) p.14

161 C. Courtois, R. A. D Grundy et al, to be published.

162 N. C. Woolsey et al, CLF Annual report RAL-TR-2000-034, 59, (2000)

163 P. Helander, et al. CLF Annual Report 2000-2001, 58 (2001).

164 R. A. London and M. D. Rosen, Phys. Fluids, 29, 11, 3813, (1986)

165 P.W. Rambo and J. Denavit, Physics of Plasmas, 1, 12, 4050, (1994)

166 R. P. Drake, Physics of Plasmas, 9, 2, 727, (2002)

167 F.F. Chen, ‘Introduction to Plasma Physics’, Plenum Press, (1974) Chapter 5.6.2

168 F.F. Chen, ‘Introduction to Plasma Physics’, Plenum Press, (1974) Chapter 5.10

169 J. Aparicio et al, Physics of Plasmas, 5, 9, 3180, (1998)

170 K. Gomberoff, Physics of Plasmas, 10, 6, 2313, (2003)