Supplemental material for Bound electron nonlinearity beyond the ionization threshold

Similar documents
Optimizing the time resolution of supercontinuum spectral interferometry

Construction of a 100-TW laser and its applications in EUV laser, wakefield accelerator, and nonlinear optics

Direct measurement of the acoustic waves generated by femtosecond filaments in air

A.J. Verhoef, A.V. Mitrofanov, D. Kartashov, A. Baltuska Photonics Institute, Vienna University of Technology. E.E. Serebryannikov, A.M.

Generation and Applications of High Harmonics

SUPPLEMENTARY INFORMATION

Development of a table top TW laser accelerator for medical imaging isotope production

Measuring the temporal intensity of ultrashort laser pulses by triple correlation

Laser heating of noble gas droplet sprays: EUV source efficiency considerations

Industrial Applications of Ultrafast Lasers: From Photomask Repair to Device Physics

Richard Miles and Arthur Dogariu. Mechanical and Aerospace Engineering Princeton University, Princeton, NJ 08540, USA

Detection of Single Photon Emission by Hanbury-Brown Twiss Interferometry

Molecular quantum wake-induced pulse shaping and extension of femtosecond air filaments

MODELLING PLASMA FLUORESCENCE INDUCED BY FEMTOSECOND PULSE PROPAGATION IN IONIZING GASES

stabilized 10-fs lasers and their application to laser-based electron acceleration

Let us consider a typical Michelson interferometer, where a broadband source is used for illumination (Fig. 1a).

SUPPLEMENTARY INFORMATION

Highly Efficient and Anomalous Charge Transfer in van der Waals Trilayer Semiconductors

B. Miao, L. Feder, Milchberg MD probe pulse. of the concave. ), where. cross

1 Mathematical description of ultrashort laser pulses

Assessment of Threshold for Nonlinear Effects in Ibsen Transmission Gratings

Two-electron systems

Part II. Interaction with Single Atoms. Multiphoton Ionization Tunneling Ionization Ionization- Induced Defocusing High Harmonic Generation in Gases

gives rise to multitude of four-wave-mixing phenomena which are of great

Thomson Scattering from Nonlinear Electron Plasma Waves

Diffuse reflection BBSFG optical layout

Atmospheric applications of laser filamentation from space

Supplementary information

Ultrafast Dynamics and Single Particle Spectroscopy of Au-CdSe Nanorods

Demonstration of long-lived high power optical waveguides in air

Confocal Microscopy Imaging of Single Emitter Fluorescence and Hanbury Brown and Twiss Photon Antibunching Setup

Nonlinear Optics (WiSe 2016/17) Lecture 9: December 16, 2016 Continue 9 Optical Parametric Amplifiers and Oscillators

Generation of tunable and broadband far-infrared laser pulses during two-color filamentation

Nanosecond Broadband Spectroscopy For Laser-Driven Compression Experiments

plasma optics Amplification of light pulses: non-ionised media

Technique of the experiment

Revival Structures of Linear Molecules in a Field-Free Alignment Condition as Probed by High-Order Harmonic Generation

l* = 109 nm Glycerol Clean Water Glycerol l = 108 nm Wavelength (nm)

Carrier dynamics of rubrene single-crystals revealed by transient broadband terahertz

No. 9 Experimental study on the chirped structure of the construct the early time spectra. [14;15] The prevailing account of the chirped struct

EXTREME ULTRAVIOLET AND SOFT X-RAY LASERS

Supplementary Figures

SUPPLEMENTARY INFORMATION

R&D experiments at BNL to address the associated issues in the Cascading HGHG scheme

Introduction. Procedure. In this experiment, you'll use the interferometer to EQUIPMENT NEEDED: Lens 18mm FL. Component holder.

Optics, Light and Lasers

Where are the Fringes? (in a real system) Div. of Amplitude - Wedged Plates. Fringe Localisation Double Slit. Fringe Localisation Grating

Supporting Online Material for

Observation of Electron Trapping in an Intense Laser Beam

A laser-produced plasma extreme ultraviolet (EUV) source by use of liquid microjet target

3.5 Cavities Cavity modes and ABCD-matrix analysis 206 CHAPTER 3. ULTRASHORT SOURCES I - FUNDAMENTALS

Hiromitsu TOMIZAWA XFEL Division /SPring-8

Probing the Propagation Dynamics in a Femtosecond Laser Filament

Towards using molecular ions as qubits: Femtosecond control of molecular fragmentation with multiple knobs

Supplementary Material for In situ frequency gating and beam splitting of vacuum- and extreme-ultraviolet pulses

Direct measurement of spectral phase for ultrashort laser pulses

Nonlinear Optics (NLO)

Suppression of thermal lensing effects in intra-cavity coherent combining of lasers

Some Topics in Optics

Particle-in-cell (PIC) simulation output for the temporal evolution of magnetic fields.

Coherent ultra-broadband laser-assisted injection radiation from a laser plasma accelerator

SEMATECH 157nm Technical Review

Time-resolved pump-probe system based on a nonlinear imaging technique with phase object

Plasma Formation and Self-focusing in Continuum Generation

LIST of SUPPLEMENTARY MATERIALS

Spatially and temporally resolved temperature measurements of plasma generated in percussion drilling with a diode-pumped Nd:YAG laser

Control of the filamentation distance and pattern in long-range atmospheric propagation

Continuous-wave biexciton lasing at room temperature using solution-processed quantum wells

The structure of laser pulses

Supporting Online Material for

Supplementary Materials for

Speed of Light in Air

File name: Supplementary Information Description: Supplementary Figures, Supplementary Notes and Supplementary References

Laser pulse propagation in a meter scale rubidium vapor/plasma cell in AWAKE experiment

Ruby crystals and the first laser A spectroscopy experiment

AMO physics with LCLS

Jitter measurement by electro-optical sampling

Ultrafast Lateral Photo-Dember Effect in Graphene. Induced by Nonequilibrium Hot Carrier Dynamics

Multi-GeV electron acceleration using the Texas Petawatt laser

Formation of laser plasma channels in a stationary gas

Experiment 6: Interferometers

SELF-ORGANIZATION OF SPATIAL SOLITONS

Measurements of electric-field strengths in ionization fronts during breakdown Wagenaars, E.; Bowden, M.D.; Kroesen, G.M.W.

Final Report for AOARD grant FA Measurement of the third-order nonlinear susceptibility of graphene and its derivatives

Michelson Interferometer

Multiphoton transitions for delay-zero calibration in attosecond spectroscopy arxiv: v1 [physics.atom-ph] 12 Jun 2014

Lab 3-4 : Confocal Microscope Imaging of Single-Emitter Fluorescence and Hanbury-Brown and Twiss Set Up, Photon Antibunching

37. 3rd order nonlinearities

2.B Laser Ionization of Noble Gases by Coulomb Barrier Suppression

Appendix A Detector Calibration

Physics 476LW Advanced Physics Laboratory Michelson Interferometer

Developments for the FEL user facility

Breakdown threshold and plasma formation in femtosecond laser solid interaction

37. 3rd order nonlinearities

Quantum Control of Molecular Gas Hydrodynamics

Digital Holographic Measurement of Nanometric Optical Excitation on Soft Matter by Optical Pressure and Photothermal Interactions

Ultrashort Phase Locked Laser Pulses for Asymmetric Electric Field Studies of Molecular Dynamics

Measurement of wakefields in hollow plasma channels Carl A. Lindstrøm (University of Oslo)

Measurement of high order Kerr refractive index of major air components

Laser and pinching discharge plasmas spectral characteristics in water window region

Transcription:

Supplemental material for Bound electron nonlinearity beyond the ionization threshold 1. Experimental setup The laser used in the experiments is a λ=800 nm Ti:Sapphire amplifier producing 42 fs, 10 mj pulses at a 1 khz repetition rate. A beam splitter deflects up to 90% of the energy out of the amplifier before it is compressed. This beam is compressed by an external compressor just before entering the vacuum chamber. For pump pulses, up to ~0.5 mj is focused by a 75 cm focal length lens to a full width at half maximum (FWHM) spot size of 65 µm in the gas tube target. The other 10% of the amplifier output is compressed by the internal compressor and focused into a Xe gas cell, generating supercontinuum pulses of energy ~1 μj covering the range 500-700 nm. This is split into probe and reference pulses using a Michelson interferometer and sent into the vacuum chamber, focused by a 50 cm focal length lens to a spot size of ~200 µm, which overfills the hole in the target. The pump and probe beams are linearly polarized with parallel polarization. The probe pulse spectral phase is characterized by time-dependent cross phase modulation [21]. The spectral phase is well fit by φ(ω) = β 2 (ω - ω 0 ) 2 + β 3 (ω - ω 0 ) 3, where β 2 = 1050 fs 2, β 3 = 100 fs 3, ω 0 = 3.15 fs -1. The gas flow target is made from a cylindrical stainless steel tube with 89 µm thick walls, pinched to the desired outer thickness of 200-400 µm and then drilled by the laser, producing 80-200 µm diameter holes in the front and back walls. Gas is slowly fed from one end of the tube and the other end is terminated. The target assembly is mounted on a translation stage in a vacuum chamber, with a flexible tube connecting it to a source of gas. The chamber is pumped by a roots blower. To avoid long-lived, cumulative thermal effects that can alter the gas density [25], the repetition rate of the pump laser is reduced to 20 Hz. Figure S1. (a) Typical interferogram showing light propagating through the hole (left) and around the target (right). (b) Phase shift through the target measured in N 2 as a function of time as the gas input is shut off suddenly at t = 0 s. 1

The gas density distribution in the gas tube hole and outside the tube is characterized by folded wavefront interferometry, using a 635 nm diode laser. The measured phase shift is proportional to the axially integrated gas density sampled by the 635 nm probe laser as it propagates through the gas. To determine the integrated gas density through the target, the probe beam propagates through the hole (along the path of the SSSI probe light) with part of the beam outside the tube. The interference between these two beam segments provides the phase shift and refractive index change from which the mean target density N 0 is determined [8]. Typical longitudinal interferometry results are shown in Fig. S1. Figure S2 shows the phase shift imposed on the probe beam directed perpendicular to the hole axis and skimming one side of the tube, showing that the gas density drops to ~0.5N 0 near the plane of the hole and falls roughly exponentially, reaching 0.2N 0 about 100 µm beyond the hole. Figure S2. Phase shift image of density profile of gas flowing out of the tube hole, showing that the N 2 density falls off quickly outside the tube (the tube edge is at z = 0, and the hole center is at x = 0). The vertical ripples are from diffraction from the tube edge. The flow gas pressure (density) in the target is adjusted to maximize the probe phase shift while limiting ionization-induced refraction of the pump and probe. Typically this demanded running at lower gas pressure at high laser intensities and at higher pressure for lower laser intensities. A roots pump maintains a background chamber pressure of <1 mbar (~100 Pa). The effective interaction length L eff is much smaller than the pump and probe confocal parameters of 30 mm and 200 mm respectively, helping ensure that the pump intensity is independent of the axial coordinate within the gas, improving the time resolution [S1]. To record and monitor the pump beam shape during the experiment, we image residual pump (~0.1%) light transmitted through the dichroic mirror DM1 in Fig. 1 of the main text using a CCD camera (not shown in the figure) placed at the equivalent focal plane of the pump focusing lens. We refer to this as the pre-target camera. Approximately 0.1% of the 800 nm pump beam is transmitted through dichroic mirror that separates the pump and probe light (DM2 in Fig. 1 of the main text). This residual pump light is reflected by another dichroic mirror (DM3) and imaged using another CCD camera (the post-target camera) placed at the equivalent plane of the entrance slit of the imaging spectrometer. We find that for 2

the peak intensities and gases used in the experiment, the images of the pump beam pre- and post-target agree within error. At higher laser pulse energy, effects from plasma defocusing are observed. For each 2D+1 data set, a total of ~10 4 shots is used to construct ΔΦ(x,y,t), as discussed in the main text. For each laser shot, an image of the pump spot is also captured, enabling measurement of y i as the probe beam is swept back and forth, and allowing post-collection compensation in the data analysis of pump beam intensity and pointing fluctuations and drift. The interferograms are sorted into ~30 bins by their y i value, as determined from the corresponding pump spot image for each laser shot. The interferograms in each bin are summed before phase extraction. The number of interferograms in each y i bin varies from tens to hundreds. Recording the full 2D+1 phase shift and amplitude change of the probe beam has multiple advantages. Among them are that the spatial peak of the pump spot is easily located, even in the presence of pointing drift, and that a single 2D phase shift profile contains many intensity points. 3

2. Experimental results for Kr, Xe, N 2, and O 2 Experimental results, in the same format as the Ar results shown in Fig. 2 of the paper, are shown in Figs. S3-S6 for the other gases studied. The bound electronic response is relatively larger in Kr and Xe than in Ar, but otherwise the results are qualitatively similar. The molecular gases N 2 and O 2 have an additional delayed rotational response that peaks ~80 fs after the peak of the pump pulse [8, 23]. Movies showing the evolution of ΔΦ(x,y,t) for each of these data sets are provided in the supplementary materials. 1 Figure S3. Results for Kr at a pump peak intensity of 84 TW/cm 2. (a) Measured spatiotemporal phase shift ΔΦ(x,y 0 =0,t). (b) Image of the pump spot at the gas target, showing that the Kerr response follows the pump intensity profile. (c) Phase shift ΔΦ(x,y,t 0 14 fs), showing the Kerr response dominating. (d) Phase shift ΔΦ(x,y,t 0 +21 fs), showing the Kerr response on the wings and the growing plasma phase shift in the center of the beam. (e) Phase shift ΔΦ(x,y,t 0 +100 fs), showing the negative plasma phase shift long after the pump pulse. 1 These can be found at http://lasermatter.umd.edu/publications.html#bound_electron_supplementary, or at the APS-PRL website. 4

Figure S4. Results for Xe at a pump peak intensity of 50 TW/cm 2. (a) Measured spatiotemporal phase shift ΔΦ(x,y 0 =0,t). (b) Image of the pump spot at the gas target, showing that the Kerr response follows the pump intensity profile. (c) Phase shift ΔΦ(x,y,t 0 14 fs), showing the Kerr response dominating. (d) Phase shift ΔΦ(x,y,t 0 +14 fs), showing the Kerr response on the wings and the growing plasma phase shift in the center of the beam. (e) Phase shift ΔΦ(x,y,t 0 +100 fs), showing the negative plasma phase shift long after the pump pulse. 5

Figure S5. Results for N 2 at a pump peak intensity of 114 TW/cm 2. (a) Measured spatiotemporal phase shift ΔΦ(x,y 0 =0,t). (b) Image of the pump spot at the gas target, showing that the Kerr response follows the pump intensity profile. (c) Phase shift ΔΦ(x,y,t 0 14 fs), showing the Kerr response dominating. (d) Phase shift ΔΦ(x,y,t 0 +14 fs), showing the Kerr response on the wings and the growing plasma phase shift in the center of the beam. (e) Phase shift ΔΦ(x,y,t 0 +280 fs), showing the negative plasma phase shift long after the pump pulse and the bound electronic and rotational response. 6

Figure S6. Results for O 2 at a pump peak intensity of 88 TW/cm 2. (a) Measured spatiotemporal phase shift ΔΦ(x,y 0 =0,t). (b) Image of the pump spot at the gas target, showing that the Kerr response follows the pump intensity profile. (c) Phase shift ΔΦ(x,y,t 0 14 fs), showing the Kerr response dominating. (d) Phase shift ΔΦ(x,y,t 0 +14 fs), showing the Kerr response on the wings and the growing plasma phase shift in the center of the beam. (e) Phase shift ΔΦ(x,y,t 0 +350 fs), showing the negative plasma phase shift long after the pump pulse and the bound electronic and rotational response. 7

3. Fits of ionization yield data A full intensity scan for a gas species consists of multiple 2D+1 data sets, each at a different nominal peak space-time laser intensity. A total of 9 data sets were used for Ar, 4 for Kr, 6 for Xe, 9 for N2, and 8 for O2 to generate Figs. 3 and 4. The measured ionization yield Y over the intensity range measured here is very well-fit by the expression Y = c I (S1) m 1 0 as shown by the dashed curves in Fig. 3 of the paper. Here I 0 is the temporal peak intensity that occurred at a particular x, y location, so that each 2D+1 SSSI trace ΔΦ(x,y,t1) contains many I0, Y pairs, where t1 > 50 fs for the noble gases and t1 > 250 fs for N2 and O2 The fit for each gas to Eq. (S1) yields parameters c 1 and m given in Table S1. The point of the expression used for Y is not to advance a multiphotonionization (MPI)-like model for ionization; it is to provide an analytic model fit to the ionization yield data to enable separation of the bound and free electron contributions. In fact, the best fit values for m are notably smaller than their corresponding multiphoton ionization values ( ) for each species (also shown in Table S1), indicating the significant contribution of tunneling ionization. Examples of the time-dependent phase shift ΔΦ(x0=0,y0=0,t) in Kr, Xe, N2, and O2, along with fits to the bound response + ionization model described in the main text, are shown in Figs. S7, S8, S9, and S10. These are equivalent to Fig. 4ab in the main text. Panel (a) in each figure shows the response at an intensity below the ionization threshold. Figures S7a and S8a for Kr and Xe show the Kerr response, which follows the intensity pulse envelope. For N2 and O2 in Figs. S9a and S10a, the initial peak, which is the prompt Kerr response, is followed by the delayed rotational response. Panel (b) in each figure shows the response as the peak laser intensity is increased to produce growing levels of ionization. For the atomic gases Ar, Kr and Xe, the time-dependent refractive index shift data is fit to 2 2 t / τ 1/2 Δ nt ( ) ΔnKe Δ np(1 erf[ m t/ τ])/2 while for N2 and O2, it is fit to = + +, (S2) 2 2 t / τ 1/2 Δ nt ( ) ΔnKe Δ nrotrt ( ) Δ np(1 erf[ m t/ τ]) / 2 = + + +, (S3) where in Eq. S3 both Δ K n and Δ rot n are fitting parameters and R(t) is the rotational response function [8, 23] convolved with a Gaussian with a FWHM of 42 fs. Note that the fit was optimum assuming a gas temperature of 200 K (the shape of the rotational response function is sensitive to temperature because the initial population of rotational states is affected). This indicates cooling of the gas as it expands into the vacuum chamber through the holes in the target. 8

Figure S7. Kr response during pump pulse. (a) Nonlinear index shift in Kr vs. time for peak intensity 32 TW/cm 2 (dots), below the ionization threshold, and a fit to a Gaussian pulse with τ FWHM =42 fs (solid line). (b) Nonlinear index shift vs. time (dots) and fits to Eq. S2 (solid lines), where Δn k is a fitting parameter. The curves have been offset vertically for clarity. Figure S8. Xe response during pump pulse. (a) Nonlinear index shift in Xe vs. time for peak intensity 16 TW/cm 2 (dots), below the threshold for ionization, and a fit to a Gaussian pulse with τ FWHM =42 fs (solid line). (b) Nonlinear index shift vs. time (dots) and fits to Eq. S2 (solid lines), where Δn k is a fitting parameter. The curves have been offset vertically for clarity. 9

Figure S9. N 2 response during pump pulse. (a) Nonlinear index shift in N 2 vs. time for peak intensity 52 TW/cm 2 (dots), below the threshold for ionization, and a fit to Eq. S3 for Δn p =0 and a Gaussian pulse with τ FWHM =42 fs. (b) Nonlinear index shift vs. time (dots) and fits to Eq. S2 (solid lines), where Δn k and Δn rot are fitting parameters. The curves have been offset vertically for clarity. Figure S10. Response during pump pulse. (a) The nonlinear index shift in O 2 as a function of time for peak intensity of 39 TW/cm 2, below the threshold for ionization, and a fit to a Gaussian with FWHM 42 fs. (b) Nonlinear index shift vs. time (dots) and fits to Eq. S2 (solid lines), where Δn k and Δn rot are fitting parameters. The curves have been offset vertically for clarity. 10

Also shown in Table S1 are values of n2 extracted from fits of the bound response to a linear intensity dependence for datasets spanning the full intensity range from no ionization to significant ionization; the fits are to the bound index shift versus intensity data points shown in Fig. 4cd in the main text. Notably, these effective values of for Ar, Kr, and Xe are higher than the values measured for intensities below the onset of ionization [9], owing to an apparent greater than linear intensity dependence of the bound electron response above the ionization threshold, the detailed scaling of which requires more precise measurements. However, for the molecules, the effective value for N2 over the full intensity range is the same as at subionization intensities, whereas the full range value in O2 is lower than at sub-ionization intensities [8]. Table 1. Summary of results Gas c1 [(cm 2 /TW) -m ] mfit n2 (cm 2 /W) (full intensity range from below to above ionization threshold) n2 (cm 2 /W) [8,9] (Below ionization threshold) Ar 2.05 10-13 5.14 11 (12.2±1.6) 10-20 (9.7±1.2) 10-20 Kr 5.17 10-11 4.43 9 (2.7±0.4) 10-19 (2.2±0.4) 10-19 Xe 5.47 10-10 4.36 7 (7.1±0.9) 10-19 (5.8±1.1) 10-19 N2 5.7 10-16 6.28 11 (7.4±1.0) 10-20 (7.4±0.9) 10-20 O2 9.6 10-11 4.09 7 (5.8±0.8) 10-20 (9.5±1.2) 10-20 Supplemental references [S1] K. Kim, I. Alexeev, and H. M. Milchberg, Single-shot measurement of laser-induced double step ionization of helium, Opt. Express 10, 1563 (2002). 11