Poole 2017

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

ACCEPTED MANUSCRIPT • OPEN ACCESS

Laser-driven ion acceleration via TNSA in the relativistic transparency


regime
To cite this article before publication: Patrick Poole et al 2017 New J. Phys. in press https://doi.org/10.1088/1367-2630/aa9d47

Manuscript version: Accepted Manuscript


Accepted Manuscript is “the version of the article accepted for publication including all changes made as a result of the peer review process,
and which may also include the addition to the article by IOP Publishing of a header, an article ID, a cover sheet and/or an ‘Accepted
Manuscript’ watermark, but excluding any other editing, typesetting or other changes made by IOP Publishing and/or its licensors”

This Accepted Manuscript is © 2017 The Author(s). Published by IOP Publishing Ltd on behalf of Deutsche Physikalische
Gesellschaft.

As the Version of Record of this article is going to be / has been published on a gold open access basis under a CC BY 3.0 licence, this Accepted
Manuscript is available for reuse under a CC BY 3.0 licence immediately.

Everyone is permitted to use all or part of the original content in this article, provided that they adhere to all the terms of the licence
https://creativecommons.org/licences/by/3.0

Although reasonable endeavours have been taken to obtain all necessary permissions from third parties to include their copyrighted content
within this article, their full citation and copyright line may not be present in this Accepted Manuscript version. Before using any content from this
article, please refer to the Version of Record on IOPscience once published for full citation and copyright details, as permissions may be required.
All third party content is fully copyright protected and is not published on a gold open access basis under a CC BY licence, unless that is
specifically stated in the figure caption in the Version of Record.

View the article online for updates and enhancements.

This content was downloaded from IP address 80.82.77.83 on 18/12/2017 at 07:36


Page 1 of 19 AUTHOR SUBMITTED MANUSCRIPT - NJP-107183.R1

1
2
3
4

pt
5
6
7
8
9
Laser-driven ion acceleration via TNSA in the

cri
10
11
12 relativistic transparency regime
13
14
15 P. L. Poole1∗
16

us
17
18 L. Obst2,3
19
20
21 G. E. Cochran4
22
23
24
25
26
27
28
J. Metzkes2

H.-P. Schlenvoigt2
an
dM
29
30 I. Prencipe2
31
32
33
34 T. Kluge2
35
36
37 T. Cowan2,3
38
39
U. Schramm2,3
pte

40
41
42
43 D. W. Schumacher4
44
45
46 K. Zeil2
47 1
Lawrence Livermore National Laboratory, 7000 East Ave, Livermore, CA 94551,
48
ce

49 USA
2
50 Helmholtz-Zentrum Dresden-Rossendorf, Dresden 01314, Germany
3
51 Technische Universität Dresden,Dresden 01062, Germany
4
52 The Ohio State University, 191 West Woodruff Ave, Columbus, OH 43210, USA
53
E-mail: poole11@llnl.gov
Ac

54
55
March 2016
56
57
58 Abstract. We present an experimental study investigating laser-driven proton
59 acceleration via Target Normal Sheath Acceleration (TNSA) over a target thickness
60 range spanning the typical TNSA-dominant regime (∼ 1 μm) down to below the
onset of relativistic laser-transparency (< 40 nm). This is done with a single target
AUTHOR SUBMITTED MANUSCRIPT - NJP-107183.R1 Page 2 of 19

1
2
3 Laser-driven ion acceleration via TNSA in the relativistic transparency regime 2
4

pt
5 material in the form of freely adjustable films of liquid crystals along with high
6 contrast (via plasma mirror) laser interaction (∼ 2.65 J, 30 fs, I > 1 × 1021 W/cm2 ).
7 Thickness dependent maximum proton energies scale well with TNSA models down to
8 the thinnest targets, while those under ∼ 40 nm indicate the influence of relativistic
9 transparency on TNSA, observed via differences in light transmission, maximum proton

cri
10
energy, and proton beam spatial profile. Oblique laser incidence (45◦ ) allowed the
11
12 fielding of numerous diagnostics to determine the interaction quality and details: ion
13 energy and spatial distribution was measured along the laser axis and both front
14 and rear target normal directions; these along with reflected and transmitted light
15 measurements on-shot verify TNSA as dominant during high contrast interaction,
16 even for ultra-thin targets. Additionally, 3D particle-in-cell simulations qualitatively

us
17 support the experimental observations of target-normal-directed proton acceleration
18 from ultra-thin films.
19
20
21
22 PACS numbers: 52.38.-r, 64.70.M-, 41.75.Jv
23
24
25
26
27
28
1. Introduction
an
A primary research endeavor of ultra-intense laser plasma interaction is efficient
secondary radiation production. Selection of various laser and target parameters
dM
29
30 can cause transfer of laser energy into electron beams [1], x-rays or neutron beams
31 for radiography or remote detection applications [2, 3], positron-electron plasmas
32
33
for fundamental studies related to astrophysics [4], and ion beams, which relate to
34 radiography [5], generation of warm dense matter [6], and cancer therapy [7, 8, 9, 10].
35 Efforts towards ion beam applications are hastened due to high power, high repetition
36
37 rate laser facilities currently under development [11, 12, 13].
38 The realization of ion beam application goals faces two critical challenges. From a
39 practical perspective, one needs the ability to produce sufficient ion flux through efficient
pte

40
41 procedures–rapid target insertion, laser operation, and optimized on-line diagnostics.
42 More fundamentally, a thorough understanding of the underlying processes that govern
43 ion acceleration in order to better control aspects such as kinetic energy, particle number,
44
45 and beam profile is required. For this, increases in laser energy and power have not
46 yielded ion energies necessary to achieve applications like cancer therapy due in part to
47
48
poor scaling with laser intensity and also to stringent requirements on laser parameters
ce

49 like pulse contrast, which measures the intensity of unwanted but often unavoidable “pre-
50 pulse” light preceding the main beam relative to that of the peak pulse. While several
51
52 mechanisms have been identified for ion acceleration [14, 15, 16, 17, 18] and multiple
53 reviews have been done to accumulate experiment, simulation, and model descriptions of
Ac

54 these [19, 20], the optimization of a given mechanism for an application on any particular
55
56 facility remains difficult.
57 Target Normal Sheath Acceleration [14, 15], or TNSA, involves rear surface layer
58
particles accelerated in the target normal direction up to tens of MeV/nucleon energies
59
60 by a space-charge electric field set up as fast electrons originating in the front-side pre-
plasma are laser-accelerated through the target. TNSA can benefit from pre-pulses–the
Page 3 of 19 AUTHOR SUBMITTED MANUSCRIPT - NJP-107183.R1

1
2
3 Laser-driven ion acceleration via TNSA in the relativistic transparency regime 3
4

pt
5 resulting target pre-plasma is abundant with electrons available for laser acceleration,
6 yielding a larger rear-surface sheath field–but will suffer if the target is sufficiently
7
8 deformed by these early pulses or if the pre-pulse scale length becomes too long [21, 20].
9 Because of this TNSA is typically maximized for targets robust enough to withstand

cri
10 initial interaction with the laser pre-pulse and main pulse rising edge, typically above
11
12 several hundred nm. TNSA is nevertheless observed on both short and long pulse
13 systems: current experimental benchmarks for TNSA include 40 MeV protons for ultra-
14
short pulse interaction [22] and 85 MeV for longer pulse lasers [23], both achieved with
15
16 ∼ 1 μm thick foils.

us
17 Other ion acceleration mechanisms can arise for thinner targets (below 100 nm),
18
19 but therefore often require intensity contrasts superior to 10−10 :1 for times at least one
20 ps before the main pulse arrives to preserve target integrity. Sufficiently thin targets can
21 undergo light-sail Radiation Pressure Acceleration [16], where the target receives initial
22
momentum from the laser and then travels along with the pulse to continually gain
23
24
25
26
27
28
an
energy [20]. This process requires irradiation with high intensity, exceptionally spatially
homogeneous pulses of circularly polarized light at normal incidence to minimize early
target heating and expansion. Other mechanisms can be categorized as enhanced-
TNSA[17, 18] where an increase in laser penetration due to a relativistic reduction of the
dM
29
30
plasma frequency (an effect known as relativistic transparency) results a larger volume
31 of electrons being accelerated to generate the sheath field. While these mechanisms
32 may scale more favorably with laser intensity than TNSA [20] they are also difficult to
33
34 produce in experiments due in part to facility limits but also to interference from plasma
35 instabilities and other complex interactions [24, 25, 26].
36 What follows is the first experimental study of ultra-short pulse proton acceleration
37
38 investigated with oblique laser incidence from the TNSA-dominant thickness regime
39 down to that of relativistic transparency with a single target material. This is enabled
pte

40
first by high contrast (via a plasma mirror) interaction, as superior laser pulse quality
41
42 is required for successful ion acceleration from ultra-thin targets, and secondly by an
43 in-situ target formation system using freely suspended liquid crystal films [27]. This
44
45
material can be drawn across an aperture in a rigid frame with changes in wiping
46 speed, temperature, and liquid crystal volume to form films several mm in diameter at
47 repetition rates of 0.1 Hz and with thicknesses from 10 nm to several 10s of μm [28].
48
ce

49 Critically, the in-situ target formation device forms films to within 2 μm of the same
50 position each formation, removing the necessity for between-shot target alignment and
51 thus allowing rapid data collection. While this on-demand thickness manipulation has
52
53 been demonstrated as an avenue toward high formation rate plasma mirrors [29], it is
utilized here to enable a detailed investigation of the influence of target thickness on ion
Ac

54
55
acceleration mechanisms, including the ultra-thin regime (< 40 nm). Furthermore, the
56
57 low density (∼ 1 g/cm−3 ) of liquid crystal used and thin target formation capabilities
58 allowed easy access to the transparency regime.
59
60
Ultra-short (τ = 30 fs) pulses of ∼ 2.65 J were incident on targets at 45◦ in order to
distinguish between laser-axis and target-normal directed acceleration. A single plasma
AUTHOR SUBMITTED MANUSCRIPT - NJP-107183.R1 Page 4 of 19

1
2
3 Laser-driven ion acceleration via TNSA in the relativistic transparency regime 4
4
mirror was optional to ascertain differences between moderate(10−6 :1) and ultra-high

pt
5
6 (10−10 :1) contrast. Oblique target incidence also allowed multiple secondary diagnostics
7
8 to corroborate the target conditions during ultra-thin target proton beam production
9 including ion spatial and energy diagnostics in the front normal, rear normal, and

cri
10 laser axis directions, as well as reflected and transmitted light diagnostics able to
11
12 measure during laser interaction due to the optical quality of the liquid crystal film
13 surface. The data show efficient acceleration yielding high quality proton beams (beam
14
profile, divergence) in the target-normal direction and ion energy scaling with target
15
16 thickness in agreement with TNSA models down to the ultra-thin region. Increased laser

us
17 transparency is observed below the ∼ 40 nm relativistic skin depth, and the thinnest (∼
18
19 10 nm) targets exhibit higher average proton energy but larger shot-to-shot fluctuations
20 along with changes in proton spatial characteristics including less homogeneity at large
21 angles from target normal.
22
23
24
25
26
27
28
2. Experimental setup and apparatus
an
The experiment was performed on the Draco laser facility at HZDR, which is a 3.5 J,
30 fs, 10 Hz Titanium:sapphire based system capable of intensities exceeding 1 × 1021
W/cm2 in a 3 μm diameter FWHM focal spot. Figure 1a shows a simplified schematic
dM
29
30
31 of the primary diagnostic setup, which included an optional single plasma mirror for
32 pulse contrast enhancement. Contrast measurements with and without a plasma mirror
33
are shown in Fig. 1b. The optics design focused light onto and then recollected it from
34
35 the plasma mirror with off axis parabolas which allowed the setup to be bypassed if
36 desired by moving an additional mirror into the beam-path. The entire plasma mirror
37
38
optics path had an energy throughput of 80%, resulting in energies between 2.55 - 2.75
39 J on target. Near-field measurements after plasma mirror reflection were monitored on
pte

40 each shot and did not reveal modulations that would suggest a reduction in final focal
41
42 spot quality. The plasma mirror substrate had dimensions of 50 mm × 95 mm which
43 allowed about 450 few mm spots to be used before it needed to be replaced.
44 The primary ion diagnostics were multi channel plate Thomson parabola
45
46 spectrometers (TPS) situated along the laser axis (0◦ ) and target normal (45◦ )
47 directions, with energy resolution of 0.5 – 1 MeV from 0 to 30 MeV. Additionally, spatial
48
ce

information of accelerated protons was recorded for a number of shots on radiochromic


49
50 film (RCF) packs, which could be moved into place on-demand at a distance of 55 mm
51 from the target.
52
53
Additional diagnostics were implemented to investigate the reflected and
transmitted light upon interaction with the optical quality target surface. This would
Ac

54
55 not only reveal the onset of transparency effects and allow their interaction with ion
56
57 acceleration to be observed, but also would allow target morphology conditions to be
58 determined at the time of laser interaction. A piece of spectralon (Lambertian scatterer)
59 was used to collect the transmitted beam mode, and a ceramic (MACOR) screen served
60
the same purpose for the reflected light. Two cameras filtered for 1 ω and 2 ω observed
Page 5 of 19 AUTHOR SUBMITTED MANUSCRIPT - NJP-107183.R1

1
2
3 Laser-driven ion acceleration via TNSA in the relativistic transparency regime 5
4

pt
5
6
7
8
9

cri
10
11
12
13
14
15
16

us
17
18
19
20
21
22
23
24
25
26
27
28
an
Figure 1. a) Diagram of experimental setup including recollimating geometry
plasma mirror, Linear Slide Target Inserter (LSTI) for in-situ variable thickness
dM
29 target formation, Thomson parabola spectrometers along the laser (0◦ ) and target
30
normal (45◦ ) axes, scatter screens (spectralon) for reflected and transmitted light
31
32 measurement, and a LANEX screen for measurement of specularly reflected electrons.
33 b) Measured ps intensity contrast without plasma mirror (blue line) and with plasma
34 mirror (black dots); dashed line indicates ionization threshold. c) Forming liquid
35 crystal film with the 4 mm diameter aperture LSTI device.
36
37
38 scattered light imaged from the ceramic screen along the same line of sight by using a
39
dichroic mirror. A final diagnostic was a sheet of LANEX, which fluoresces upon impact
pte

40
41 by energetic electrons, to reveal the spatial distribution of target front-surface electrons
42 ejected in the reflection direction. On some shots a rough electron energy spectrum was
43
44 obtained using transmission through a wedged aluminum piece in front of the screen.
45 Liquid crystal films were formed within a Linear Slide Target Inserter device, or
46 LSTI [28], as shown in Fig. 1c. The 4 mm diameter circular aperture was oriented at
47
48 45◦ with respect to the incoming laser axis. 10 μL of the liquid crystal 4-cyano-4’-
ce

49 octylbiphenyl (8CB) was applied to the wiper, which allowed several dozen films to be
50
formed before the chamber needed to be opened for new volume application. A cooled
51
52 water line pumped by a small chiller unit was installed through the chamber wall to
53 maintain the LSTI frame at the desired temperature for film control, typically around
Ac

54
55
28.0 ◦ C. Film thicknesses were measured on-demand by white light interferometry, which
56 was relayed in from outside the chamber. The LSTI design utilizes the same liquid
57 crystal mesophase surface tension that enables freely suspended film formation to draw
58
59 films at the front aperture edge within 2 μm of the same position each time, which
60 eliminates the need for target alignment after it is performed on the first film. This
AUTHOR SUBMITTED MANUSCRIPT - NJP-107183.R1 Page 6 of 19

1
2
3 Laser-driven ion acceleration via TNSA in the relativistic transparency regime 6
4

pt
5 alignment, performed with a scattered light imaging system, was verified by observing
6 the accelerated proton energy with on-shot diagnostics as the target was shifted along
7
8 the incoming laser axis. All data shown here was taken at best focus.
9

cri
10
11
3. Experimental data and analysis
12
13 3.1. TNSA behavior
14
15 The central experimental result is shown in Fig. 2a): maximum proton cutoff energy
16 recorded on the target normal Thomson parabola spectrometer for film thicknesses

us
17
18 ranging from > 1 μm to 10 nm. Open blue circles indicate the maximum energy
19 observed on the target normal Thomson parabola, and the filled circles are averages of
20
these values with bin size indicated by shaded background region. Pink circles are from
21
22 RCF stacks and serve to corroborate the Thomson parabola data. While shots without a
23
24
25
26
27
28
an
plasma mirror (triangles) demonstrated predictably decreasing maximum proton energy
as film thicknesses decreased below 1 μm, the high contrast data shown here exhibits an
overall increase in proton cutoff energy along the target normal direction as thickness
decreases down to 300 nm, at which point the cutoff energies plateau. This behavior is
in agreement with TNSA scaling law predictions [30, 31, 32], where maximum proton
dM
29
30 energy depends on the electron density within the Debye sheath. Assuming that electron
31 density depends on their divergence angle and path length to target rear (i.e. target
32
33 thickness), a saturation for small thicknesses is expected. The fit shown is calculated
34 using 1D TNSA model [30] with electron temperature scaling [32] and a best fit free
35
parameter a0 ∼ 17; this vector potential corresponds to intensity I = 6 × 1020 W/cm2 ,
36
37 reasonably close to the estimated experimental peak intensity of 1021 W/cm2 .
38 Carbon ion (C6+ ) cutoff energies in the target normal direction exhibit the same
39
target thickness dependence as the protons down to the thinnest targets, with a ratio of
pte

40
41 proton/carbon cutoff energies whose average remains near 3, shown in Fig. 2 with error
42 bars. This behavior suggests that both protons and carbon ions gain energy in the same
43
44 sheath field for each target as thickness is reduced, and in general is in agreement with
45 predictions for TNSA [31].
46 The high proton energy achieved for ultra-thin targets suggests a high contrast
47
48 laser interaction, where the target surfaces remain intact before the main pulse arrives.
ce

49 This is supported by observations of the specular-directed optical emission during laser


50
51
interaction recorded with both 1ω and 2ω filters. Here we expand the measurement
52 of previous target emission studies [33, 34] by always using peak laser intensity but
53 with deliberate pulse contrast differences: Figure 3 shows representative images of the
Ac

54
55 moderate (no plasma mirror, left) and high contrast laser reflection (right). High
56 contrast results in significantly enhanced reflected light levels and a better-defined
57 spatial mode. Under this condition of unperturbed target surface at the initiation
58
59 of high intensity laser interaction, 2ω emission becomes visible, with expected spatial
60 modulations due to the strongly intensity-dependent generation process. This improved
Page 7 of 19 AUTHOR SUBMITTED MANUSCRIPT - NJP-107183.R1

1
2
3 Laser-driven ion acceleration via TNSA in the relativistic transparency regime 7
4

pt
5
6
7
8
9

cri
10
11
12
13
14
15
16

us
17
18
19
20
21
22
Figure 2. a) Maximum cutoff energy of target normal accelerated protons increasing
23
24
25
26
27
28
an
with decreasing thickness. Here triangles indicate data taken without the plasma
mirror while circles indicate it was in place for contrast enhancement. Open markers
are single data points (blue = Thomson parabola, pink = Radiochromic film (RCF)),
shaded are average energies from thickness bins indicated by background shading.
Arrows indicate the RCF data is a lower limit. The fit is a 1-D scaling [30, 32] with
free parameter a0 = 16.7 ± 0.1. b) Ratio of maximum proton and carbon 6+ energies
dM
29
30 measured along the target normal direction, where the consistency suggests TNSA
31 occurred for all target thicknesses. Here as well open circles are single data points,
32 shaded circles are averages.
33
34
35
36 reflection quality indicates a minimally expanded critical surface in the high contrast
37 case [35] and was observed even for ultra-thin targets.
38 High contrast interaction is further corroborated by the LANEX diagnostic
39
observing electrons ejected in the laser reflection direction during target interaction.
pte

40
41 Up to 30 nC/sr was measured in electron distributions with energies exceeding 9
42
43
MeV, obtained by recording the calibrated [36] LANEX fluorescence behind different
44 thicknesses of the Al wedge. This is similar to recent results [37] where electrons ejected
45 in this manner required laser interaction with a sharp density profile originating from
46
47 high contrast pulse irradiation.
48 Figure 4 shows target normal ion energy traces for various thickness targets (blue
ce

49 lines) for comparison to an average of those laser axis ion spectra observed (red). While
50
51 all targets have similar spectral shape at low proton energies, the thinnest targets also
52 exhibit increased high energy component suggesting a greater hot electron population.
53 The majority of shots (> 90%) at high contrast showed no laser axis ion signal for
Ac

54
55 any thickness, and when present the laser axis energy cutoff and yield were always
56 significantly lower than that observed along target normal for the same shot.
57
58
The target normal spatial distribution of proton emission, shown for various
59 thicknesses in Fig. 5, was of consistently small divergence (< 10◦ ) at low and high energy
60 for all thicknesses. This was true even for ultra-thin targets near the transparency
AUTHOR SUBMITTED MANUSCRIPT - NJP-107183.R1 Page 8 of 19

1
2
3 Laser-driven ion acceleration via TNSA in the relativistic transparency regime 8
4

pt
5
6
7
8
9

cri
10
11
12
13
14
15
16

us
17
18
19
20
21
22
23
24
25
26
27
28
an
Figure 3. 1ω and 2ω reflected modes from moderate and high contrast pulses, showing
a significant difference in the target morphology at the time of laser interaction for
dM
29
30 these cases. The inset in the left hand images shows the moderate contrast reflection
31 raw images (unscaled dynamic range), which reveal significantly less overall reflection
32 compared to high contrast conditions.
33
34
35
36
37
38
39
pte

40
41
42
43
44
45
46
47
48
ce

49
50
51
52
53
Ac

54
55
56
57
58
59 Figure 4. Proton energy spectra in the target normal (blue) and laser axis (red)
60 directions. The latter is averaged over the few shots for which appreciable signal was
detected, and as such represents an upper limit for ions in this direction.
Page 9 of 19 AUTHOR SUBMITTED MANUSCRIPT - NJP-107183.R1

1
2
3 Laser-driven ion acceleration via TNSA in the relativistic transparency regime 9
4

pt
5
6
7
8
9

cri
10
11
12
13
14
15
16

us
17
18
19
20
21
22
23
24
25
26
27
28
an
Figure 5. Radiochromic film (RCF) stacks demonstrating spatial distribution of
accelerated protons for different target thicknesses (indicated along top) irradiated
with high contrast pulses. Consistency in both beam divergence and overall shape
dM
29 is observed except for ultra-thin films. The rightmost sheet is a representative RCF
30 layer for a moderate contrast shot, demonstrating a higher ion divergence under these
31 conditions.
32
33
34 threshold with the exception of radial streaking in the low energy ion signal (as in the
35
36 4.7 MeV, 6 nm sheet in Fig. 5). For comparison, the rightmost film is one layer from
37 a moderate contrast, 6500 nm target result, demonstrating significantly larger beam
38
divergence.
39
The proton acceleration was also measured with RCF film from both the forward
pte

40
41 and backward target normals for a single target thickness of 36 nm, selected as near
42
43 the relativistic transparency transition. Figure 6 shows that protons from these two
44 directions are similar in maximum energy, yield, and beam divergence, indicating
45 similarly strong accelerating fields on both surfaces. This is a hallmark of the TNSA
46
47 mechanism and has been observed previously [38, 39] in the case where the target
48 front is sufficiently unperturbed by laser pre-pulse. A slight offset is observed in
ce

49 the proton emission centroid from exactly the target normal direction: toward the
50
51 reflected laser in the backward direction and toward the transmitted laser in the forward
52 direction (as indicated by the inset). This angular shift is attributed to an intra-
53
pulse TNSA effect whereby the electron distribution accelerated in the direction of the
Ac

54
55 reflected/transmitted beam will cause a sheath field asymmetry which likewise impacts
56 the proton beam direction [40].
57
58
59
60
AUTHOR SUBMITTED MANUSCRIPT - NJP-107183.R1 Page 10 of 19

1
2
3 Laser-driven ion acceleration via TNSA in the relativistic transparency regime 10
4

pt
5
6
7
8
9

cri
10
11
12
13
14
15
16 Figure 6. Representative RCF layers placed in the forward (within incoming laser,

us
17 top images) and backward (bottom images) target normal axis. The maximum energies
18
for these are 12 MeV and 10 MeV, respectively, for this 36 nm thick target. The inset
19
20 shows the orientation of the target and RCF stacks, along with the direction protons
21 are shifted with respect to the target normal axis (dark central hole).
22
23
24
25
26
27
28
an
dM
29
30
31
32
33
34
35
36
37
38
39
pte

40
Figure 7. Transmitted light measurements with respect to target thickness collected
41
42 by imaging scatter from a spectralon sheet, absolutely calibrated to shots with no
43 target present. Dark circles are transmission averages over thickness bins indicated by
44 the shaded regions, and the curve is present to guide the eye. Transparency onset is
45 near 40 nm.
46
47
48 3.2. Transition to transparency
ce

49
50 While the energy spectra, spatial distribution, and directionality of the acceleration
51 ions suggests TNSA as the dominant mechanism for low and high contrast shots for
52
53 thicknesses from 40 nm and above, additional characteristics were observed for ultra-
thin targets in the relativistic transparency regime. This state was verified in several
Ac

54
55
ways, but most directly by measuring the laser pulse transmitted through the target.
56
57 These values are shown in Fig. 7 as a function of film thickness. Thicknesses above 40
58 nm exhibit a nearly constant average of  5 % transmission, but those below show both
59
60 increased average transmission as well as greater shot-to-shot fluctuation.
There is a similar increase in maximum proton energy fluctuations for these ultra-
Page 11 of 19 AUTHOR SUBMITTED MANUSCRIPT - NJP-107183.R1

1
2
3 Laser-driven ion acceleration via TNSA in the relativistic transparency regime 11
4

pt
5 thin targets. Near 10 nm target thickness proton energies ranged from 16 MeV to
6 26 MeV (see Fig. 2a))–despite this the average maximum energy was higher for these
7
8 thinnest targets. A third observation from the transparency regime was a radial streak
9 pattern outside the primary proton spatial structure visible in the lowest energy RCF

cri
10 layers, as in the top left stack of Fig. 5. These thinnest targets still showed a high
11
12 quality reflection during laser interaction, so this burst proton pattern is suspected to
13 originate not from target expansion via pre-pulse but rather from late-time TNSA fields
14
still accelerating particles as the target volume expands after laser interaction.
15
16 Although fluctuations in the transmitted light, maximum proton energy, and low-

us
17 energy spatial distribution were observed for ultra-thin target interaction, none of these
18
19 effects are seen to correlate strongly with each other, nor with other measured values such
20 as the quality of the reflected or transmitted modes. Additionally, the proton energy
21 only weakly correlates with the ± 2 % laser energy fluctuations on target. Critically, the
22
ratio of proton to carbon 6+ does not fluctuate largely at the thinnest targets (Fig. 2b)),
23
24
25
26
27
28
an
suggesting that rear surface acceleration and hence the target integrity is not affected.
As such the source of these fluctuations is believed to be related to underlying laser
plasma interaction in this relativistic regime, the details of which will be further studied
with particle-in-cell simulations.
dM
29
30
31 4. Simulations
32
33
Simulation efforts for comparison to experimental observations first focused on ion
34
35 acceleration behavior with thickness: Fig. 8 shows snapshots in time of the Poynting
36 vector magnitude S, accelerated proton trajectory, and energy spectrum. These fully 3D
37
38
particle-in-cell simulations were performed using the code Large Scale Plasma (LSP) [41]
39 with 8CB targets using an implicit field solver and particle advance as well as particle
tracking. A p-polarized laser pulse (λ = 800 nm, τ = 30 fs FWHM sin2 intensity
pte

40
41
42 temporal envelope) with a peak intensity of 1 × 1021 W/cm2 is incident at 45◦ on
43 target, with x as the target normal direction. The target is composed of fully ionized
44 carbon and hydrogen atoms in their stoichiometric ratio in the liquid crystal 8CB, with
45
46 an electron density of 184 ncr and an initial electron temperature Te = 10 keV. There
47 were 125 particles per species per cell in the 300 nm target simulation, and 8 particles
48
ce

per species per cell in the 30 nm target simulation.


49
50 For the 300 nm target, the target normal cell size Δx = 15 nm and Δy = Δz =
51 45 nm. In the case of the 30 nm target, Δx = 1.5 nm in the 150 nm directly around
52
53
the target, expanding out to 12 nm; Δy = Δz = 12 nm throughout the grid. In order
to maximize vacuum acceleration length while limiting computational cost, the spot
Ac

54
55 size used (1.2 μm, Gaussian FWHM) was smaller than in the experiment to decrease
56
57 transverse target extent. Additionally, the cells used do not resolve the Debye length
58 (roughly 1 nm); as such the energy-conserving force interpolation of LSP was employed
59 to mitigate numerical heating. These simulations were conducted with a 0.75 Courant
60
timestep.
AUTHOR SUBMITTED MANUSCRIPT - NJP-107183.R1 Page 12 of 19

1
2
3 Laser-driven ion acceleration via TNSA in the relativistic transparency regime 12
4

pt
5 Figures 8a and 8b show the strong dependence of laser penetration on target
6 thickness. In Fig. 8a the magnitude of the Poynting vector is plotted at t0 + 20 fs, where
7
8 t0 is the time at which the peak of the laser pulse reaches the focus. The laser pulse
9 is largely reflected from the still opaque 300 nm target. Plotting the same quantity

cri
10 at the same point in time in Fig. 8b shows the laser has penetrated through the 30
11
12 nm target. The total flux passing through simulation planes at -1.8 μ m and 2.5 μ
13 m was recorded for both simulations, allowing measurement of total reflectivity and
14
transmission. The 300 nm target showed 73% reflectivity and 2% transmission, while
15
16 the 30 nm target showed a marked decrease in reflectivity (61%) and a slight increase

us
17 in transmission (5%). This is consistent with the trend seen in experiment, where the
18
19 amount of transmitted light increased substantially around a 30 nm target thickness
20 (see Fig. 7).
21 Figure 8c shows the late-time angular distribution of accelerated protons for t0 +300
22
f s for two film thicknesses:the 300 nm target (blue) displays protons directed along
23
24
25
26
27
28
an
target normal in a tight grouping, while the 30 nm target (orange) also shows
general target normal acceleration although now with a somewhat broader distribution
angle. Dominant proton acceleration in the target normal direction for both thick and
relativistically transparent targets are observed, consistent with experimental results.
dM
29
30
This deviation from true target normal is in a consistent direction with that shown in
31 Fig. 6, coming from a similar sheath field asymmetry [40].
32 Finally, Fig. 8d shows the experimental average proton cutoff energies from Fig. 2a
33
34 in blue with the cutoff energy obtained from the two simulation thicknesses overlaid
35 in red. The simulation cutoff energies fall reasonably close to experimental values and
36 importantly reflect the general experimental trend of highest energies from thinnest
37
38 targets. The discrepancy seen is most likely due to the simulation laser pulse not
39 properly accounting for the small amount of laser pre-pulse remaining after plasma
pte

40
mirror reflection in the experiment, instead using an idealized pulse shape and beginning
41
42 with a pre-ionized, initially hot target plasma. A full study of these pre-pulse related
43 effects on the ultra-thin targets is planned to further investigate the ion acceleration
44
45
fluctuations observed in this regime.
46
47
5. Comparison to literature
48
ce

49
50 Figure 9 shows maximum proton energy thickness scans taken from numerous literature
51 references, where those with > 3 distinct thicknesses and linearly polarized pulses were
52
53
selected. Here normal incidence data are indicated as triangles, non-normal as circles,
and the data presented in this work as stars (also non-normal). Filled points are short
Ac

54
55 pulses (< 50 fs) and open are long (> 500 fs). In general higher laser energies result
56
57 in higher maximum proton energies, but the efficiency can vary strongly depending on
58 other laser conditions.
59 Foremost, laser contrast plays an important role in the nature of proton acceleration
60
as thickness is varied: maximized TNSA cutoff energy will occur at that thickness where
Page 13 of 19 AUTHOR SUBMITTED MANUSCRIPT - NJP-107183.R1

1
2
3 Laser-driven ion acceleration via TNSA in the relativistic transparency regime 13
4

pt
5
6
7
8
9

cri
10
11
12
13
14
15
16

us
17
18
19
20
21
22
23
24
25
26
27
28
an
Figure 8. a) Magnitude of the Poynting vector taken at t0 + 20 fs for a 300 nm target
(logarithmic scale). The laser has not yet penetrated the target thickness, and is
dM
29
predominantly reflected along the specular direction. Target initial position marked in
30
31 dashed cyan. b) Magnitude of the Poynting vector taken at t0 +20 fs for a 30 nm target
32 (logarithmic scale). The laser has penetrated the target and moved several microns
33 beyond its initial position at this point in time, and laser reflection along the specular
34 direction is minimal. c) Angular distribution of protons from the 300 nm (blue)
35 and 30 nm (orange) simulations at t0 + 300 fs, both displaying predominantly target
36 normal directed distributions. d) Cutoff energies obtained from the two simulation
37 thicknesses (red) plotted with the averaged experimental proton energies from Fig.2a.,
38 demonstrating the trend of higher energies from thinner targets
39
pte

40
41
42 electron divergence through the target and pre-pulse expansion from imperfect contrast
43 are both minimized [31, 52]. Several previous thickness scans [42, 45, 50, 51] therefore
44 exhibit a peak of maximum proton cutoff energy at a certain thickness. If the laser
45
46 pre-pulse can be sufficiently suppressed, proton cutoff energies tend to rise slightly or
47 plateau as thickness decreases to 100 nm and below, as seen in [47, 43, 44, 48, 49, 26, 23].
48
ce

Those datasets with thicknesses below the transparency regime (∼ 40 nm) can exhibit a
49
50 slight increase in proton energy here possibly due to a related enhancement effect. This
51 is true as well of the averaged data presented in this work (black stars), which was also
52
53
high contrast.
The uniqueness of the dataset presented here is then twofold: first, no other
Ac

54
55 data spans the range from few nm to few μm with a single target material, enabling
56
57 measurement of TNSA over the entire range of thicknesses previously seen as optimum
58 with other laser conditions. The absence of a strong peak confirms the pre-pulse/electron
59 divergence trade-off described previously. Secondly, this data is the only one to examine
60
non-normal incidence below 100 nm, and the continued increase of cutoff energy in
AUTHOR SUBMITTED MANUSCRIPT - NJP-107183.R1 Page 14 of 19

1
2
3 Laser-driven ion acceleration via TNSA in the relativistic transparency regime 14
4

pt
5
6
7
8
9

cri
10
11
12
13
14
15
16

us
17
18
19
20
21
22
23
24
25
26
27
28
an
Figure 9. Proton energy target thickness scans reported in the literature, separated
dM
29
30 by incident laser energy. Data represented by circles are near 45◦ incidence ([23]
31 (averaged data), [42], [38], [43], [44], [45], [46] (22◦ )) and triangles indicate normal or
32 near-normal (< 10◦ ) incidence interactions ([26], [47], [48], [49], [50] (averaged data),
33 [51]). Filled data points are from pulses of 50 fs duration or less, while open points
34 indicate greater than 500 fs. The averaged results from this experiment (Fig. 2) are
35 shown as black stars; this is the first dataset to span the full range from few nm to few
36 μm with a single target material.
37
38
39
the target normal direction here suggests TNSA as a relevant acceleration mechanism
pte

40
41 for thicknesses down to the onset of laser transparency. Full characterization of the
42 enhancement regime for targets near 10 nm requires further study with exceptional
43
44 laser contrast and in particular with a quantity of shots capable of fully mapping the
45 fluctuation regime observed in this experiment, which is only now possible due to laser
46 and solid target technology advancements.
47
48
ce

49 6. Conclusion
50
51
52 Presented here is a high granularity thickness scan of ultrashort pulse ion acceleration,
53 the first to study oblique incidence interaction from the TNSA regime down below
Ac

54
the onset of relativistic transparency, where proton energies are observed beyond 25
55
56 MeV with ∼ 2.65 J, 30 fs pulses. Oblique laser incidence was used to discriminate
57 between target normal and laser axis directed ion acceleration, which was verified with
58
59
angularly separated ion and optical diagnostics that simultaneously ascertained high
60 contrast laser-target interaction. In particular, the data set presented shows consistent
TNSA energies for thicknesses down to ∼ 100 nm if sufficient contrast can be achieved,
Page 15 of 19 AUTHOR SUBMITTED MANUSCRIPT - NJP-107183.R1

1
2
3 Laser-driven ion acceleration via TNSA in the relativistic transparency regime 15
4

pt
5 providing an avenue toward repeatable, predictable ion acceleration from such targets.
6 Additionally, target normal acceleration in the transparency regime–below 40 nm
7
8 for the conditions used here–reveal inconsistencies in laser transmission, maximum
9 proton energy, and proton beam spatial profile, the origins of which warrant further

cri
10 simulation and experiment study. In particular, additional experimental control in the
11
12 form of deliberate pre-pulses that tailor the critical surface and pre-plasma scale length
13 could shed light on further acceleration dynamics.
14
The large data set allowing a credible statistical evaluation of the acceleration
15
16 dynamics was able to be collected due to a combination of rapid shot-on-demand, ultra-

us
17 high contrast laser, diagnostic, and target operation, and serves as an important step
18
19 towards applications that will require optimized ion acceleration from high repetition
20 rate laser plasma interaction. In particular, the observed proton energy stability in the
21 regime below 100 nm but above the relativistic transparency threshold from consistent
22
laser conditions in combination with the rapid target formation of the LSTI device
23
24
25
26
27
28
an
has achieved the robustness necessary for progress toward applications. Additionally,
the low mass of these targets and minimal, optimally directed plasma ejecta of liquid
crystal compared to traditional metallic foils both present excellent prospects for low-
debris, high rate laser-target interaction. In this setup, the last remaining cost-intensive
dM
29
30
consumable is the plasma mirror substrate, which can in principle be replaced with a
31 liquid-crystal based setup [29]. This combination of liquid crystal plasma mirror and
32 target film formation in the 40 – 100 nm range is hence a credible path toward future
33
34 application-relevant laser-driven ion sources.
35
36
37
Acknowledgment
38
39 This work was supported by the DARPA PULSE program through AMRDEC, by the
pte

40 NNSA (DE-NA0001976), by EC Horizon 2020 LASERLAB-EUROPE/LEPP (654148),


41
42 by the German Federal Ministry of Education and Research (BMBF, 03Z1O511), and
43 by an allocation of computing time from the Ohio Supercomputer Center.
44
45
46 Bibliography
47
48 [1] W.P. Leemans, A.J. Gonsalves, H.-S. Mao, K. Nakamura, C. Benedetti, C.B. Schroeder, Cs. Tóth,
ce

49 J. Daniels, D.E. Mittelberger, S.S. Bulanov, J.-L. Vay, C.G.R. Geddes, and E. Esarey. Multi-
50 gev electron beams from capillary-discharge-guided subpetawatt laser pulses in the self-trapping
51 regime. Phys. Rev. Lett., 113:245002, 2014.
52
[2] Margaret M. Murnane, Henry C. Kapteyn, Mordecai D. Rosen, and Roger W. Falcone. Ultrafast
53
x-ray pulses from laser-produced plasmas. Science, 251(4993):531–536, 1991.
Ac

54
55 [3] K. W. D. Ledingham, P. McKenna, and R. P. Singhal. Applications for nuclear phenomena
56 generated by ultra-intense lasers. Science, 300(5622):1107–1111, 2003.
57 [4] Hui Chen, S. C. Wilks, J. D. Bonlie, S. N. Chen, K. V. Cone, L. N. Elberson, G. Gregori, D. D.
58 Meyerhofer, J. Myatt, D. F. Price, M. B. Schneider, R. Shepherd, D. C. Stafford, R. Tommasini,
59 R. Van Maren, and P. Beiersdorfer. Making relativistic positrons using ultraintense short pulse
60 lasers. Physics of Plasmas, 16(12), 2009.
AUTHOR SUBMITTED MANUSCRIPT - NJP-107183.R1 Page 16 of 19

1
2
3 Laser-driven ion acceleration via TNSA in the relativistic transparency regime 16
4

pt
5 [5] Teresa Bartal, Mark E. Foord, Claudio Bellei, Michael J. Key, Kirk A. Flippo, Sandrine A. Gaillard,
6 Dustin T. Offerman, Pravesh K. Patel, Leonard C. Jarrott, Drew P. Higginson, Markus Roth,
7 Anke Otten, Dominik Kraus, Richard B. Stephens, Harry S. McLean, Emilio M. Giraldez,
8 Mighsheng S. Wei, Donald C. Gautier, and Farhat N. Beg. Focusing of short-pulse high-intensity
9 laser-accelerated proton beams. Nat Phys, 8(2):139–142, 2012.

cri
10
[6] M Roth, I Alber, V Bagnoud, C R D Brown, R Clarke, H Daido, J Fernandez, K Flippo, S Gaillard,
11
12 C Gauthier, M Geissel, S Glenzer, G Gregori, M Gnther, K Harres, R Heathcote, A Kritcher,
13 N Kugland, S LePape, B Li, M Makita, J Mithen, C Niemann, F Nrnberg, D Offermann,
14 A Otten, A Pelka, D Riley, G Schaumann, M Schollmeier, J Schtrumpf, M Tampo, A Tauschwitz,
15 and An Tauschwitz. Proton acceleration experiments and warm dense matter research using
16 high power lasers. Plasma Physics and Controlled Fusion, 51(12):124039, 2009.

us
17 [7] S. V. Bulanov, H. Daido, T. Zh. Esirkepov, V. S. Khoroshkov, J. Koga, K. Nishihara, F. Pegoraro,
18 T. Tajima, and M. Yamagiwa. Feasibility of using laser ion accelerators in proton therapy. AIP
19 Conference Proceedings, 740(1):414–429, 2004.
20
[8] S D Kraft, C Richter, K Zeil, M Baumann, E Beyreuther, S Bock, M Bussmann, T E Cowan,
21
Y Dammene, W Enghardt, U Helbig, L Karsch, T Kluge, L Laschinsky, E Lessmann, J Metzkes,
22
D Naumburger, R Sauerbrey, M. Schrer, M Sobiella, J Woithe, U Schramm, and J Pawelke.
23
24
25
26
27
28
Journal of Physics, 12(8):085003, 2010.
an
Dose-dependent biological damage of tumour cells by laser-accelerated proton beams. New

[9] A. Yogo, T. Maeda, T. Hori, H. Sakaki, K. Ogura, M. Nishiuchi, A. Sagisaka, H. Kiriyama,


H. Okada, S. Kanazawa, T. Shimomura, Y. Nakai, M. Tanoue, F. Sasao, P. R. Bolton,
M. Murakami, T. Nomura, S. Kawanishi, and K. Kondo. Measurement of relative biological
dM
29 effectiveness of protons in human cancer cells using a laser-driven quasimonoenergetic proton
30
beamline. Applied Physics Letters, 98(5):053701, 2011.
31
[10] K. Zeil, M. Baumann, E. Beyreuther, T. Burris-Mog, T. E. Cowan, W. Enghardt, L. Karsch,
32
33 S. D. Kraft, L. Laschinsky, J. Metzkes, D. Naumburger, M. Oppelt, C. Richter, R. Sauerbrey,
34 M. Schürer, U. Schramm, and J. Pawelke. Dose-controlled irradiation of cancer cells with laser-
35 accelerated proton pulses. Applied Physics B, 110(4):437–444, 2013.
36 [11] S. Gales. Laser driven nuclear science and applications: The need of high efficiency, high power
37 and high repetition rate laser beams. Eur. Phys. J. Special Topics, 224:2631–2637, 2015.
38 [12] J. Schreiber, P. R. Bolton, and K. Parodi. Invited Review Article: ”hands-on” laser-driven ion
39 acceleration: A primer for laser-driven source development and potential applications. Review
pte

40
of Scientific Instruments, 87(7), 2016.
41
[13] A. Irman M. Siebold K. Zeil D. Albach C. Bernert S. Bock F. Brack J. Branco J.P. Couperus
42
43 T. Cowan A. Debus C. Eisenmann M. Garten R. Gebhardt S. Grams U. Helbig A. Huebl T.
44 Kluge A. Kohler J.M. Kramer S. Kraft F. Kroll M. Kuntzsch U. Lehnert M. Loeser J. Metzkes
45 P. Michel L. Obst R. Pausch M. Rehwald R. Sauerbrey H.-P. Schlenvoigt K. Steiniger O. Zarini
46 U. Schramm, M. Bussmann. First results with the novel peta-watt laser acceleration facility in
47 dresden. IOP Conf. Proc. accepted, 2017.
48
ce

[14] R. A. Snavely, M. H. Key, S. P. Hatchett, T. E. Cowan, M. Roth, T. W. Phillips, M. A. Stoyer,


49 E. A. Henry, T. C. Sangster, M. S. Singh, S. C. Wilks, A. MacKinnon, A. Offenberger, D. M.
50
Pennington, K. Yasuike, A. B. Langdon, B. F. Lasinski, J. Johnson, M. D. Perry, and E. M.
51
Campbell. Intense high-energy proton beams from petawatt-laser irradiation of solids. Phys.
52
53 Rev. Lett., 85:2945–2948, 2000.
[15] Stephen P. Hatchett, Curtis G. Brown, Thomas E. Cowan, Eugene A. Henry, Joy S. Johnson,
Ac

54
55 Michael H. Key, Jeffrey A. Koch, A. Bruce Langdon, Barbara F. Lasinski, Richard W. Lee,
56 Andrew J. Mackinnon, Deanna M. Pennington, Michael D. Perry, Thomas W. Phillips, Markus
57 Roth, T. Craig Sangster, Mike S. Singh, Richard A. Snavely, Mark A. Stoyer, Scott C. Wilks,
58 and Kazuhito Yasuike. Electron, photon, and ion beams from the relativistic interaction of
59 petawatt laser pulses with solid targets. Physics of Plasmas (1994-present), 7(5), 2000.
60 [16] T. Esirkepov, M. Borghesi, S. V. Bulanov, G. Mourou, and T. Tajima. Highly efficient relativistic-
Page 17 of 19 AUTHOR SUBMITTED MANUSCRIPT - NJP-107183.R1

1
2
3 Laser-driven ion acceleration via TNSA in the relativistic transparency regime 17
4

pt
5 ion generation in the laser-piston regime. Phys. Rev. Lett., 92:175003, 2004.
6 [17] Emmanuel dHumires, Erik Lefebvre, Laurent Gremillet, and Victor Malka. Proton acceleration
7 mechanisms in high-intensity laser interaction with thin foils. Physics of Plasmas, 12(6):062704,
8 2005.
9 [18] L. Yin, B. J. Albright, B. M. Hegelich, and J. C. Fernandez. Gev laser ion acceleration from

cri
10
ultrathin targets: The laser break-out afterburner. Laser and Particle Beams, 24:291–298, 6
11
12 2006.
13 [19] Hiroyuki Daido, Mamiko Nishiuchi, and Alexander S Pirozhkov. Review of laser-driven ion sources
14 and their applications. Reports on Progress in Physics, 75(5):056401, 2012.
15 [20] Andrea Macchi, Marco Borghesi, and Matteo Passoni. Ion acceleration by superintense laser-
16 plasma interaction. Rev. Mod. Phys., 85:751–793, 2013.

us
17 [21] J. Badziak, S. Jaboski, P. Parys, M. Rosiski, J. Woowski, A. Szydowski, P. Antici, J. Fuchs, and
18 A. Mancic. Ultraintense proton beams from laser-induced skin-layer ponderomotive acceleration.
19 Journal of Applied Physics, 104(6), 2008.
20
[22] K Ogura, M Nishiuchi, A S Pirozhkov, T Tanimoto, A Sagisaka, T Zh. Esirkepov, M Kando,
21
T Shizuma, T Hayakawa, and H Kiriyama. Proton acceleration to 40\,MeV using a high
22
intensity, high contrast optical parametric chirped-pulse amplification/Ti:sapphire hybrid laser
23
24
25
26
27
28
[23]
an
system. Opt. Lett., 37(14):2868–2870, 2012.
F. Wagner, O. Deppert, C. Brabetz, P. Fiala, A. Kleinschmidt, P. Poth, V.A. Schanz, A. Tebartz,
B. Zielbauer, M. Roth, T. Stöhlker, and V. Bagnoud. Maximum Proton Energy above 85 MeV
from the Relativistic Interaction of Laser Pulses with Micrometer Thick CH2 Targets. Physical
Review Letters, 116(20):205002, 2016.
dM
29 [24] C A J Palmer, J Schreiber, S R Nagel, N P Dover, C Bellei, F N Beg, S Bott, R J Clarke, A E
30
Dangor, S M Hassan, P Hilz, D Jung, S Kneip, S P D Mangles, K L Lancaster, A Rehman,
31
A P L Robinson, C Spindloe, J Szerypo, M Tatarakis, M Yeung, M Zepf, and Z Najmudin.
32
33 Rayleigh-Taylor Instability of an Ultrathin Foil Accelerated by the Radiation Pressure of an
34 Intense Laser. 225002:1–5, 2012.
35 [25] H. W. Powell, M. King, R. J. Gray, D. A. MacLellan, B. Gonzalez-Izquierdo, L. C. Stockhausen,
36 G. Hicks, N. P. Dover, D. R. Rusby, D. C. Carroll, H. Padda, R. Torres, S. Kar, R. J. Clarke,
37 I. O. Musgrave, Z. Najmudin, M. Borghesi, D. Neely, and P. McKenna. Proton acceleration
38 enhanced by a plasma jet in expanding foils undergoing relativistic transparency. New Journal
39 of Physics, 17(10):103033, 2015.
pte

40
[26] N P Dover, C A J Palmer, M J V Streeter, H Ahmed, B Albertazzi, M Borghesi, D C Carroll,
41
J Fuchs, R Heathcote, P Hilz, K F Kakolee, S Kar, R Kodama, A Kon, D A MacLellan,
42
43 P McKenna, S R Nagel, D Neely, M M Notley, M Nakatsutsumi, R Prasad, G Scott, M Tampo,
44 M Zepf, J Schreiber, and Z Najmudin. Buffered high charge spectrally-peaked proton beams in
45 the relativistic-transparency regime. New Journal of Physics, 18(1):013038, 2016.
46 [27] P. L. Poole, C. D. Andereck, D. W. Schumacher, R. L. Daskalova, S. Feister, K. M. George,
47 C. Willis, K. U. Akli, and E. A. Chowdhury. Liquid crystal films as on-demand, variable
48
ce

thickness (50 - 5000nm) targets for intense lasers. Physics of Plasmas (1994-present), 21(6),
49 2014.
50
[28] P. L. Poole, C. Willis, G. E. Cochran, R. T. Hanna, C. D. Andereck, and D. W. Schumacher.
51
Moderate repetition rate ultra-intense laser targets and optics using variable thickness liquid
52
53 crystal films. Applied Physics Letters, 109(15), 2016.
[29] P. L. Poole, A. Krygier, G. E. Cochran, P. S. Foster, G. G. Scott, L. A. Wilson, J. Bailey,
Ac

54
55 N. Bourgeois, C. Hernandez-Gomez, D. Neely, P. P. Rajeev, and D. W. Schumacher. Experiment
56 and simulation of novel liquid crystal plasma mirrors for high contrast, intense laser pulses.
57 Scientific Reports, 6:32041, 2016.
58 [30] P. Mora. Plasma expansion into a vacuum. Phys. Rev. Lett., 90:185002, 2003.
59 [31] J. Schreiber, F. Bell, F. Grüner, U. Schramm, M. Geissler, M. Schnürer, S. Ter-Avetisyan, B. M.
60 Hegelich, J. Cobble, E. Brambrink, J. Fuchs, P. Audebert, and D. Habs. Analytical model for
AUTHOR SUBMITTED MANUSCRIPT - NJP-107183.R1 Page 18 of 19

1
2
3 Laser-driven ion acceleration via TNSA in the relativistic transparency regime 18
4

pt
5 ion acceleration by high-intensity laser pulses. Physical Review Letters, 97:1–4, 2006.
6 [32] T. Kluge, T. Cowan, A. Debus, U. Schramm, K. Zeil, and M. Bussmann. Electron temperature
7 scaling in laser interaction with solids. Phys. Rev. Lett., 107:205003, 2011.
8 [33] A. S. Pirozhkov, I. W. Choi, J. H. Sung, S. K. Lee, T. J. Yu, T. M. Jeong, I J. Kim, N. Hafz,
9 C. M. Kim, K. H. Pae, Y.-C. Noh, D.-K. Ko, J. Lee, A. P. L. Robinson, P. Foster, S. Hawkes,

cri
10
M. Streeter, C. Spindloe, P. McKenna, D. C. Carroll, C.-G. Wahlstrm, M. Zepf, D. Adams,
11
12 B. Dromey, K. Markey, S. Kar, Y. T. Li, M. H. Xu, H. Nagatomo, M. Mori, A. Yogo, H. Kiriyama,
13 K. Ogura, A. Sagisaka, S. Orimo, M. Nishiuchi, H. Sugiyama, T. Zh. Esirkepov, H. Okada,
14 S. Kondo, S. Kanazawa, Y. Nakai, A. Akutsu, T. Motomura, M. Tanoue, T. Shimomura,
15 M. Ikegami, I. Daito, M. Kando, T. Kameshima, P. Bolton, S. V. Bulanov, H. Daido, and
16 D. Neely. Diagnostic of laser contrast using target reflectivity. Applied Physics Letters,

us
17 94(24):241102, 2009.
18 [34] M J V Streeter, P S Foster, F H Cameron, M Borghesi, C Brenner, D C Carroll, E Divall, N P
19 Dover, B Dromey, P Gallegos, J S Green, S Hawkes, C J Hooker, S Kar, P McKenna, S R Nagel,
20
Z Najmudin, C A J Palmer, R Prasad, K E Quinn, P P Rajeev, A P L Robinson, L Romagnani,
21
J Schreiber, C Spindloe, S Ter-Avetisyan, O Tresca, M Zepf, and D Neely. Relativistic plasma
22
surfaces as an efficient second harmonic generator. New Journal of Physics, 13(2):023041, 2011.
23
24
25
26
27
28
[35]

[36]
2011. an
D. W. Schumacher, G. E. Kemp, A. Link, R. R. Freeman, and L. D. Van Woerkom. The shaped
critical surface in high intensity laser plasma interactions. Physics of Plasmas, 18(1):013102,

A. Buck, K. Zeil, A. Popp, K. Schmid, A. Jochmann, S. D. Kraft, B. Hidding, T. Kudyakov,


C. M. S. Sears, L. Veisz, S. Karsch, J. Pawelke, R. Sauerbrey, T. Cowan, F. Krausz, and
dM
29 U. Schramm. Absolute charge calibration of scintillating screens for relativistic electron
30
detection. Review of Scientific Instruments, 81(3):033301, 2010.
31
[37] M. Thevenet, A. Leblanc, S. Kahaly, H. Vincenti, A. Vernier, F. Quere, and J. Faure. Vacuum
32
33 laser acceleration of relativistic electrons using plasma mirror injectors. Nat Phys, 12(4):355–
34 360, 2016.
35 [38] T. Ceccotti, A. Lévy, H. Popescu, F. Réau, P. D’Oliveira, P. Monot, J. P. Geindre, E. Lefebvre, and
36 Ph Martin. Proton acceleration with high-intensity ultrahigh-contrast laser pulses. Physical
37 Review Letters, 99(18):1–4, 2007.
38 [39] S. Steinke, M. Schnrer, T. Sokollik, A.A. Andreev, P.V. Nickles, A. Henig, R. Hrlein, D. Kiefer,
39 D. Jung, J. Schreiber, T. Tajima, M. Hegelich, D. Habs, and W. Sandner. Optimization of
pte

40
laser-generated ion beams. Contributions to Plasma Physics, 51(5):444–450, 2011.
41
[40] K. Zeil, J. Metzkes, M. Bussmann, T. E. Cowan, S. D. Kraft, R. Sauerbrey, and U. Schramm.
42
43 Direct observation of prompt pre-thermal laser ion sheath acceleration. Nat. Commun., 3(874),
44 2012.
45 [41] D. R. Welch, D. V. Rose, M. E. Cuneo, R. B. Campbell, and T. A. Mehlhorn. Integrated simulation
46 of the generation and transport of proton beams from laser-target interaction. Physics of
47 Plasmas, 13(6), 2006.
48
ce

[42] D. Neely, P. Foster, A. Robinson, F. Lindau, O. Lundh, A. Persson, C.-G. Wahlstrm, and
49 P. McKenna. Enhanced proton beams from ultrathin targets driven by high contrast laser
50
pulses. Applied Physics Letters, 89(2):021502, 2006.
51
[43] D C Carroll, O Tresca, R Prasad, L Romagnani, P S Foster, P Gallegos, S Ter-Avetisyan, J S
52
53 Green, M J V Streeter, N Dover, C A J Palmer, C M Brenner, F H Cameron, K E Quinn,
J Schreiber, A P L Robinson, T Baeva, M N Quinn, X H Yuan, Z Najmudin, M Zepf, D Neely,
Ac

54
55 M Borghesi, and P McKenna. Carbon ion acceleration from thin foil targets irradiated by
56 ultrahigh-contrast, ultraintense laser pulses. New Journal of Physics, 12(4):045020, 2010.
57 [44] K Zeil, S D Kraft, S Bock, M Bussmann, T E Cowan, T Kluge, J Metzkes, T Richter, R Sauerbrey,
58 and U Schramm. The scaling of proton energies in ultrashort pulse laser plasma acceleration.
59 New Journal of Physics, 12(4):045015, 2010.
60 [45] J. S. Green, A. P. L. Robinson, N. Booth, D. C. Carroll, R. J. Dance, R. J. Gray, D. A. MacLellan,
Page 19 of 19 AUTHOR SUBMITTED MANUSCRIPT - NJP-107183.R1

1
2
3 Laser-driven ion acceleration via TNSA in the relativistic transparency regime 19
4

pt
5 P. McKenna, C. D. Murphy, D. Rusby, and L. Wilson. High efficiency proton beam generation
6 through target thickness control in femtosecond laser-plasma interactions. Applied Physics
7 Letters, 104(21):214101, 2014.
8 [46] A. J. Mackinnon, Y. Sentoku, P. K. Patel, D. W. Price, S. Hatchett, M. H. Key, C. Andersen,
9 R. Snavely, and R. R. Freeman. Enhancement of proton acceleration by hot-electron

cri
10
recirculation in thin foils irradiated by ultraintense laser pulses. Phys. Rev. Lett., 88:215006,
11
12 May 2002.
13 [47] A. Henig, S. Steinke, M. Schnürer, T. Sokollik, R. Hörlein, D. Kiefer, D. Jung, J. Schreiber, B. M.
14 Hegelich, X. Q. Yan, J. Meyer-ter Vehn, T. Tajima, P. V. Nickles, W. Sandner, and D. Habs.
15 Radiation-pressure acceleration of ion beams driven by circularly polarized laser pulses. Phys.
16 Rev. Lett., 103:245003, 2009.

us
17 [48] I Jong Kim, Ki Hong Pae, Chul Min Kim, Hyung Taek Kim, Jae Hee Sung, Seong Ku Lee, Tae Jun
18 Yu, Il Woo Choi, Chang-Lyoul Lee, Kee Hwan Nam, Peter V. Nickles, Tae Moon Jeong, and
19 Jongmin Lee. Transition of proton energy scaling using an ultrathin target irradiated by linearly
20
polarized femtosecond laser pulses. Phys. Rev. Lett., 111:165003, 2013.
21
[49] F. Dollar, C. Zulick, T. Matsuoka, C. McGuffey, S. S. Bulanov, V. Chvykov, J. Davis,
22
G. Kalinchenko, G. M. Petrov, L. Willingale, V. Yanovsky, A. Maksimchuk, A. G. R. Thomas,
23
24
25
26
27
28
[50]
of Plasmas, 20(5):056703, 2013.
an
and K. Krushelnick. High contrast ion acceleration at intensities exceeding 1021wcm2. Physics

D Jung, B J Albright, L Yin, D C Gautier, R Shah, S Palaniyappan, S Letzring, B Dromey,


H-C Wu, T Shimada, R P Johnson, M Roth, J C Fernandez, D Habs, and B M Hegelich.
Beam profiles of proton and carbon ions in the relativistic transparency regime. New Journal
dM
29 of Physics, 15(12):123035, 2013.
30
[51] I. Jong Kim, Ki Hong Pae, Il Woo Choi, Chang-Lyoul Lee, Hyung Taek Kim, Himanshu Singhal,
31
Jae Hee Sung, Seong Ku Lee, Hwang Woon Lee, Peter V. Nickles, Tae Moon Jeong, Chul Min
32
33 Kim, and Chang Hee Nam. Radiation pressure acceleration of protons to 93mev with circularly
34 polarized petawatt laser pulses. Physics of Plasmas, 23(7):070701, 2016.
35 [52] M. Kaluza, J. Schreiber, M. I. K. Santala, G. D. Tsakiris, K. Eidmann, J. Meyer-ter Vehn, and
36 K. J. Witte. Influence of the laser prepulse on proton acceleration in thin-foil experiments.
37 Phys. Rev. Lett., 93:045003, 2004.
38
39
pte

40
41
42
43
44
45
46
47
48
ce

49
50
51
52
53
Ac

54
55
56
57
58
59
60

You might also like