Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Dynamic position shifts of X-ray emission from a water film induced by a pair of time-delayed femtosecond laser pulses

Open Access Open Access

Abstract

Femtosecond double-pulsed laser excitation of a water film in air showed enhancements of X-ray intensity as compared with single pulse irradiation. The position of the highest yield of X-rays strongly depends on temporal separation between the pre-pulse and the main-pulse (energy ratios where ∼ 1 : 10). The strongest X-ray emission was observed at 10–15 ns delay of the main-pulse. Nanoscale roughening of water surface can account for the observation.

© 2017 Optical Society of America

1. Introduction

Intense femtosecond (fs-) laser interaction with condensed matter involves linear and nonlinear absorption [1], which results in series of processes including plasma formation [2], ablation [3], and X-ray pulse emission [4]. The characteristics of such X-ray pulses have been considered as an important research topic for applications of time-resolved X-ray diffraction [5] and X-ray absorption fine structure (XAFS) measurements [6]. When the X-ray pulses are applied to the above measurements, the X-ray intensity is one of the most significant parameters. It has become an important subject to enhance X-ray intensity with a constant amount of fs-laser photon flux. The enhancement of X-ray intensity is attributed to an effective increase of the absorption efficiency and an enhanced generation of high-energy electrons from the highly ionized plasma [7].

Solid targets such as metals and transparent glasses [8, 9] with flat surfaces have been used for X-ray generation under intense fs-laser irradiation. Such X-ray generation is considered to be due to surface plasmon resonance effects [10]. The enhancements of X-ray intensity from solid targets by changing the structure of the layered surfaces have been reported [11, 12]. In contrast to the solid targets, liquids/solutions are promising since they can be reused continuously by a circulative pump. Moreover, the X-ray emission spectra of aqueous solution can be easily controlled by changing the concentrations or species of solutes which include electron and metallic suspensions [13,14]. These features are advantageous for time-resolved XAFS measurements since long-time count accumulation is indispensable and the characteristics of solutes can be used as wavelength standards [15].

X-ray intensity enhancements under the double-pulsed excitation to aqueous solution surfaces, with the pre-pulse and the main pulse irradiations, have been reported elsewhere [16]. The pre-pulse irradiation with relatively low intensity results in various time-dependent phenomena such as plasma formation/decay in the picosecond range or ablation in the nanosecond range. Such various phenomena change the solution surface conditions, then we expect that the optimum position for X-ray emission of the solution surface to the laser focus changes as the delay time increases. As practical developments of bench-top X-ray sources, this knowledge is indispensable.

In this study, we perform the axial-scan of the water film exposure for the X-ray intensity measurements using the automatic positioning system [17] under the double-pulsed excitation conditions. Controlled time separation from tens-of-fs to tens-of-ns was investigated for the fixed ratio of pre-pulse and the main pulse.

2. Samples and procedures

The experimental setup is shown in Fig. 1. Transform-limited fs-laser pulses (tp = 40 fs, λ = 800 nm, 1 kHz, horizontally-polarized, Mantis, Legend Elite, HE USP, Coherent, Inc.) were separated and combined collinearly as double pulses with half-wave plates (65–906, Edmund Optics) and polarization beam splitters (47-048, Edmund Optics). Vertically and horizontally polarized pulses are defined as a pre-pulse and a main-pulse, respectively. The vertically-polarized pre-pulse intensity is well low and its irradiation does not induce X-ray emission by itself therefore the obtained X-ray emission under the double pulses irradiation conditions is always a single X-ray pulse. For the delay time of the main pulse relative to the pre-pulse, there are three time zones, I: 0 – 5 ns, II: 5 – 10 ns and III: 10 – 15 ns [Fig. 1], set in the optical path of the pre-pulse. The optical delay for each time zone is obtained with an 80 cm-long variable delay line (SGSP 46–800, Sigma Koki). Change of the time zones from I to II or III can be done with two independent fixed retro-reflectors. The optical path of the main-pulse has the same optical length as the longest optical path of the pre-pulse. After combination, the pulses were tightly focused onto a solution film with an off-axis parabolic mirror (the effective focus length of f = 50.8 mm, the reflection angle of 90 degrees, and the numerical aperture NA = 0.25, 47-097, Edmund Optics) in air. At these focusing conditions, the theoretical diameter of the focal spot is 2w0 = 1.22λ/NA = 3.9 μm and the axial extent along the propagation estimated as two Rayleigh lengths 2zR=2nλNA2=25.5μm where n = 1 is refractive index of air. Noteworthy, the focal region is longer than spatial extent of the pulse ctp = 12 μm; c is the speed of light.

 figure: Fig. 1

Fig. 1 Setup used for X-ray detection using two fs-laser pulses with controlled time separation, Δt; HWP is the half-wave plate, PBS is the polarized beam splitter, RR is the retro-reflector, ODL is the optical delay line, OAPM is the off-axis parabolic mirror, S is the sample solution, GM is the Geiger Müller counter, APS is the automatic positioning system, DS is the displacement sensor.

Download Full Size | PDF

The solution film was formed by a solution flow system [18] and its position was controlled by 3D-automatic (KS701-20LMS, Suruga Seiki) and rotation stages (KS401-60, Suruga Seiki) [17]. Two orthogonal plastic pipette nozzles (inner diameter of 0.46 mm), which are connected to a circulation pump (PMD-211, Sanso), make two colliding solution flows (flow rate ∼180 mL/min). Such flow collisions result in the formation of a thin solution film. The solution film is oriented perpendicularly to the solution flows from the nozzles and moves at ∼ 5.8 m/s or nm/ns rate. The estimated thickness of the solution film is ∼ 20 μm. The solution film was set on the automatic positioning system and the incident angle of the laser pulse was at 60 degrees to the normal of the solution film for the optimization of X-ray emission by change the position of solution film surface to the laser focus. At this close to Brewster angle position, the reflectivity of water surface for the p-polarized main-pulse is minimum Rp → 0 and favors coupling of light energy into pre-excited water film by the preceding pre-pulse. The operation detail of the automatic positioning system with Labview code was reported elsewhere [17].

In this study, for higher precision of the solution film position, a displacement sensor (precision of 0.2 μm, LK-G80, KEYENCE) was set behind the automatic positioning stage and its signal which shows an absolute value of distance from the stage was used as feedback parameter of automatic positioning system. A Geiger counter (SS315, Southern Scientific) was used for X-ray measurements. The major gas component of the Geiger counter is helium (>95%, ∼0.5 bar) with some halogens as quenching agent. The input window material is mica (1.6 mg/cm2) and its thickness is ∼ 6 μm. The distance between the laser focus (the X-ray point source) and the Geiger counter was kept constant at 15 cm in the entire experiment. The solid angle of the detector was 1.39 × 10−4 steradians. All the experiments were carried out in air under atmospheric pressure (1 atm) at room temperature (296K). Therefore, it is certain that the Geiger counter detects only X-ray, not α nor β-rays.

3. Results

3.1. X-rays out of water irradiation

Figure 2 shows X-ray intensity from the double pulse-irradiated water film at different time delays of the main-pulse. At shorter time scale from 0.6 to 1.6 ps the position of the solution surface had to be shifted up-stream (towards incoming fs-pulse) for the most intense X-ray emission [Fig. 2(a)]. The position z = 0 is the Z position at the highest X-ray intensity when only the main pulse was irradiated. On a longer time scale up to several nanoseconds, a tendency was observed of down-stream shift of intense X-ray emission [Fig. 2(b)]. Figure 2(c) represents slices of the X-ray intensity spatial profiles for the time delays from −48.4 ps to 1 ns. Figure 2(d) represents the X-ray intensity profile at Δt = 1.6 ps, which is with two different Gaussian fitting curves.

 figure: Fig. 2

Fig. 2 X-ray intensity of a water film irradiation by the main-pulse of 700 μJ at different time delays Δt after the pre-pulse on a ps- (a) and ns-time (b) scales. The pre-pulse energy was 80 μJ. Automatic positioning system was used to find the maximum X-ray intensity [17]. (c) Slices of the X-ray intensity spatial profiles for the time delays from −48.4 ps to 1 ns. (d) A high z-axis resolution plot of the X-ray intensity at Δt = 1.6 ps at which the local maximum was observed in the picosecond range. The double peak is deconvoluted into two Gaussian contributions. X-ray emission without the pre-pulse was more than twice lower in the peak amplitude.

Download Full Size | PDF

Figure 3 shows the location of the solution surface position along the axis of the laser propagation should be set for the maximum of X-ray emission depending on the time separation between pre-pulse and the main-pulse. The maximum of X-ray emission was achieved for a Δt ≃ 1.6 ps delay of the main-pulse. This enhancement at the up-stream position (towards the incoming laser pulse) was short lived and another maximum was developed after the location of nanosecond time scale. Another location of intense X-rays was located down-stream and evolved on a ns-time frame.

 figure: Fig. 3

Fig. 3 (a) Dependence of the solution surface position of the Δt time-delayed main-pulse for the maximum intensity of X-rays from a 20-μm-thick water film (thickness of film is marked by shaded area). Error bars are plotted as standard deviation for 20 pulse measurement. Pre-pulse energy was 80 μJ, the main-pulse was 700 μJ. Inset schematically shows contributions of the main three peak positions at which X-ray emission is maximized. (b) Schematics of two-pulse irradiation of a solution film for the three typical positions of the solution surface when the maximum of X-rays is generated; focusing position is at the main pulse. The width of film and axial extent of focal region for low intensity focus are comparable. (c) Intensity profiles for NA = 0.25 focusing in air; length of pulse ctp ≃ 12 μm. Intensity is presented on logarithmic scale, I = 0.5 at zR is marked by a contour line (corresponds to the depth of focus). For comparison, the width of water film, h, is marked; in experiments, the irradiation angle was θ = 60° and the effective width of water film was twice larger h/ cos θ.

Download Full Size | PDF

The exact geometrical focusing of a Gaussian beam corresponding to the used NA = 0.25 off-axis parabolic mirror is shown in Fig. 3(c), where the outline of the outer intensity contour is following the waist defining profile w(z)=w01+(z/zr)2 with zR=πw02/λ. The axial extent of the ultra-short laser pulse, ctp, is almost twice shorter as the depth-of-focus, 2zR. This defines favorable conditions to reach higher intensity on the target surface before onset of strong air ionization. The molecular number density of air is ∼ 2.5 × 1019 cm−3 which cause an intensity clamping during filamentation in air breakdown by ultra-short laser pulses [2]; water has 1.34 × 103 larger molecular number density. It was shown that for a higher-NA the intensity of the filament clamping in air is increasing and for NA = 0.12 is I = (2 − 3) × 1014 W/cm3 as determined from side imaging and shadowgraphy of the focal region [2], where the plasma density of ∼ 3 × 1019 cm3 is reached (a full ionization of air). Larger clamping intensity can be predicted for NA = 0.25 used in our study, however, due to a very short pulse duration the localization of filamentation and air breakdown should be close to focus. Estimation of irradiance from geometrical focus [Fig. 3(c)] gives Ip = 6.6 × 1014 W/cm2 for a Ep = 100 μJ pre-pulse (for the NA = 0.12 used in [2], Liu, et al. with clamping at I = (2 − 3) × 1014 W/cm2 the theoretical average irradiance would reach Ip = 1.5 × 1014 W/cm2 and the peak intensity is twice larger). Evidently, in the case of a pre-pulse excitation of water with ∼ 103 larger molecular number, it is expected that larger plasma densities can be created.

3.2. Nanosecond time delays

Figure 4 shows optimization of X-ray intensity by axial positioning of the solution surface at temporal separation between the pulses Δt at 5 and 15 ns. The Gaussian fits show strong localization of X-ray intensity peak at the down-stream with 50 ± 10 μm precision. Up to 75–90% X-ray intensity peak enhancement was observed at 15 ns delay for the focus position on the film (z = 0). Approximately twice more intense X-ray emission was observed during optimized main-pulse placement after 10–15 ns delay as compared with the first spike of X-ray emission during 1–2 ps window [Fig. 5].

 figure: Fig. 4

Fig. 4 X-ray intensity at different positions of the water flow film at 5 ns (a) and 15 ns (b) delays. The fit lines in green were constructed by multi-Gaussian function. Under double-pulsed excitation, the spatial profiles of X-ray intensity are significantly different at various delays. The data set for the main-pulse-only irradiation is shown with 50 counts offset for clarity.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 X-ray intensity and the axial z-position of the water film at short ps (a) and long ns (b) delays. The positions with the maximum X-ray intensity showing in positive z-axis at ps delays and negative z-axis at ns delays represent up-stream shifts and down-stream shifts as pattern 2 and pattern 3 shown in the inset.

Download Full Size | PDF

To visualize light intensity distribution on a micro-film of water film under excitations which corresponds to plasma formation at close-to-critical densities several patterns on water film were simulated [Fig. 6]. The modeling results have a qualitative character since only take into account a linear light-matter interaction, however, is informative to reveal possible field enhancement effects due to nanotexture/roughness of surfaces. A thin skin-layer of plasma is formed on the surface of dielectric [19] (e.g., water) which can support surface wave at the plasma-air interface with the period close to the wavelength of light in air, Λ = λ, while on the inner interface plasma-water surface excitation has period Λ = λ/n/2, where n = 1.33 is the refractive index of water. This plasma wave model is consistent with experimental observation of sub-wavelength ripples’ formation on surface and inside dielectrics [19,20]. These two cases qualitatively represented with larger and smaller period ripples on the two different sides of film. The random pattern of water nano-protrusions and droplet formation are also shown (noteworthy, the droplets are formed on much longer time scales of tens-to-hundreds nanoseconds). Surface capillary wave is a candidate for explanation of surface roughening, film, and droplet formation. The size and time scale of those evens are important to estimate. The period of capillary Λ=σh/ρ42πtw, where σ is surface tension, ρ is the mass density, τw is the lifetime of the wave (e.g., lifetime of the molten phase on solid materials [21]), h is the height of the wave (molten phase).

 figure: Fig. 6

Fig. 6 FDTD modeling of 800 nm wavelength light (plane wave) traversing water film with several generic surface ripple patterns expected to develop under pulsed excitation. The patterns are: the period Λ = λ on the front surface and depth profile ∼ λ/2, the back-side period λ/n, random pattern of similar size protrusions caused by light filamentation pattern, and initial stages of departing droplets. Insets show schematically the pattern simulated. Thickness of film is taken 10 μm to reduce simulation time; IntensityE2.

Download Full Size | PDF

The finite-difference time-domain (FDTD) simulations (Lumerical) show light field intensity enhancement up to 3 times for several typical expected surface morphologies of pre-excited water films [Fig. 6]. Even larger light localization was observed in the case of single droplets departing from film, however, such events would be evolving at μs time scale beyond the tested time window. Strong Talbot pattern of refocusing occurs inside the water film and can be instrumental to reach more efficient excitation and heating of water.

4. Discussion

As a development of bench-top X-ray sources, emission spectra, pulse width, and X-ray photon numbers are basic and indispensable characteristics. Hatanaka, et al. reported typical X-ray emission spectra from various aqueous solution [13–15,22,23], including distilled water [13,15] as in the current case, which shows structureless spectra in the energy region of 3 – 20 keV (0.4 – 0.06 nm) indicating bremsstrahlung without characteristic X-ray lines in this energy region because only the light elements such as hydrogen and oxygen are under laser excitation. Even under the double-pulsed excitation conditions with aqueous solutions, X-ray emission spectra have been reported [16], which show the X-ray intensity enhancements and the intensity increase in the higher photon energy region. As for the pulse-width measurements, there are limited numbers of reports on synchrotron radiation, X-ray free-electron laser, and femtosecond laser-based X-ray emission. Tabashi, et al. have reported that X-ray pulses at 14 keV from synchrotron radiation are measured to be 32 ps by intensity interferometry [24]. Helml, et al. have reported that X-ray pulses at 1.7 keV from X-ray free-electron laser are measured to be 4.4 fs by near-infrared streaking spectroscopy [25]. Zamponi, et al. have reported that X-ray pulses at 8 keV from near-IR femtosecond laser-based X-ray emission are measured to be less than 300 fs by time-resolved X-ray diffraction [26]. Recently, Holtz, et al. reported that X-ray pulses at 8 keV from near-IR femtosecond laser-based X-ray emission are measured to be 100 fs by shot-noise limited time-resolved diffraction [27].

As for the conversion efficiency from near IR (800 nm) to hard X-rays has been estimated to be 10−8 under single pulse excitation condition [16] at the comparable conditions as in this study. X-ray emission photon numbers are also estimated, under the assumption that X-ray pulse emits equally into a 4π solid angle. We obtain 3.8 × 1010 photons s−14πsr−1 at 3 – 20 keV under double-pulsed excitation conditions used in this study.

X-ray intensity is driven by the absorbed energy and resonant absorption at the interface of water/air and nanoparticle/air in the case of a metal nanoparticles. It can also have a black body contribution which can be significant at longer times after the pulse. The absorbed energy and its spatial localization can be estimated. The skin depth in optically excited material is determined by the imaginary part of the refractive index nε=n+iκ as labs = c/(ωκ) = λ/(2πκ); this is for E-field and labs/2 for intensity. The absorbed energy density [J/cm3] at the end of the laser pulse Wabs = 2AFp/labs, where Fp=0tpI(t)dt is the integral fluence per pulse, A is the absorbance, I(t) is the temporal envelope of intensity [28]. At the pre-breakdown conditions in dielectrics (e.g., air or water), the change of the imaginary part of permittivity is [28]:

(Δεd)imωpe2ω2νeω=nencrνeω,
where ωpe=4πe2ne/me is the cyclic electron plasma frequency, e, me are the electron charge and mass, respectively, ne is the electron density, νe is the electron relaxation frequency (electron-phonon in solid state material), ncr = ω2me/(4πe2) is the critical plasma density which was ∼ 1.7 × 1021 cm−3 at the used λ = 800 nm wavelength. The permittivity at the pre-breakdown in dielectric medium is εdn02+i×(Δεd)im=n02+i2n0κ, where n0 is the real part of an unperturbed refractive index [28]. Hence, the skin depth with use of Eqn. 1 can be expressed as [28]:
labs=λ2πκ=n0λπ(Δεd)im=2cn0νencrne
and is defining localization of the energy deposition. The absorbed energy density is then:
Wabs=A0νecn0nencrFpnencrFp,
where the unperturbed absorbance A0 ≡ 4n0/[(n0 + 1)2 + κ2] ≃ 4n0/(n0 + 1)2 is neglecting the imaginary part of the refractive index, κ. Equation 3 shows that as the electron density, ne, is approaching critical, the most efficient energy deposition occurs. When a nonlinear absorption via n–photon process generates electrons neIpn~Fpn, it begins to saturate as the critical density is approached. Then, the breakdown region behaves as metal where the absorbed energy is proportional to the fluence, WabsFp.

From the scaling shown above (Eqn. 3), the role of pre-pulse in creation of initial plasma density facilitates the larger energy deposition of the main-pulse, however, the nonlinearity of X-ray generation is decreasing as nencr. The tendency observed in Fig. 7, which follows a nonlinear absorption intensity (fluence) window with slope n = 2. X-ray emission is proportional to the absorbed energy per volume and according to Eqn. 3 can be interpreted as being driven by linear process of electron generation Wabs~ne×FpFp2.

 figure: Fig. 7

Fig. 7 X-ray emission at different pre-pulse energies at Δt = 1.6 ps when there is most intense peak developed up-stream. The line represents a γ = 2 power law. Sample: water film.

Download Full Size | PDF

The used wavelength of 800 nm is in the window of a solvated electron absorption [29] which are produced by water ionization resulting in generation of electrons and H2+ via a H2O2+ intermediate which is specific for the photo-ionizaiton of water by ultra-short pulses at high irradiance [30]. Absorption band of solvated electron at 1.7–1.8 eV is coinciding with the thermodynamical water splitting potential EH2O2/2H2O = 1.77 V at standard hydrogen electrode (SHE) with two electrons produced [31, 32]. These photo-generated electrons, ne, constitutes species which undergo laser driven heating by direct absorption Wabs~ne×FpFp2 in the spectral absorbance window of solvated electrons.

5. Conclusion

By using a highly accurate ∼ 1 μm resolution axial beam steering method, a two pulse breakdown of water micro-films was investigated. Two distinct regions at short and long delay times, Δt, were observed when the up-stream and down-stream (in respect to the incoming fs-laser pulse) positions caused the most intense X-ray emission. The most intense X-rays were generated when the main-pulse was 10–15 ns delated. This time is consistent to establishment of nanorough-ness/ripples on water film surface which facilitates stronger light localization inside film and local light enhancement.

Funding

Australian Research Council (ARC) DP170100131; Melbourne Centre for Nanofabrication (MCN); Australian National Fabrication Facility (ANFF).

Acknowledgment

FDTD simulations were performed on the swinSTAR supercomputer at Swinburne University of Technology.

References and links

1. T. A. Labutin, V. N. Lednev, A. A. Ilyin, and A. M. Popov, “Femtosecond laser-induced breakdown spectroscopy,” J. Anal. At. Spectrom. 31, 90–118 (2016). [CrossRef]  

2. X. L. Liu, X. Lu, X. Liu, T. T. Xi, F. Liu, J. L. Ma, and J. Zhang, “Tightly focused femtosecond laser pulse in air: from filamentation to breakdown,” Opt. Express 18(25), 26007–26017 (2010). [CrossRef]   [PubMed]  

3. E. G. Gamaly, “The physics of ultra-short laser interaction with solids at non-relativistic intensities,” Phys. Rep. 508, 91–243 (2011). [CrossRef]  

4. S. Mondal, I. Chakraborty, S. Ahmad, D. Carvalho, P. Singh, A. D. Lad, V. Narayanan, P. Ayyub, and R. Kumar, “Highly enhanced X-ray emission from oriented metal nanorod arrays excited by intense femtosecond laser pulses,” Phys. Rev. B. 83, 035408 (2011). [CrossRef]  

5. M. Bargheer, N. Zhavoronkov, Y. Gritsai, J. C. Woo, D. S. Kim, M. Woerner, and T. Elsaesser, “Coherent atomic motions in a nanostructure studied by femtosecond X-ray diffraction,” Science 306, 1771–1773 (2004). [CrossRef]   [PubMed]  

6. T. Lee, Y. Jiang, C. G. Rose-Petruck, and F. Benesch, “Ultrafast tabletop laser-pump-X-ray probe measurement of solvated Fe(CN)64,” J. Chem. Phys. C 122, 084506 (2005). [CrossRef]  

7. S. Bulanov, L. Kovrizhnykh, and A. Sakharov, “Regular mechanisms of electron and ion acceleration in the interaction of strong electromagnetic waves with a plasma,” Physics Reports 186(1), 1–51 (1990). [CrossRef]  

8. K. Hatanaka, K.-I. Yomogihata, H. Ono, K. Nagafuchi, H. Fukumura, M. Fukushima, T. Hashimoto, S. Juodkazis, and H. Misawa, “Hard X-ray generation using femtosecond irradiation of PbO glass,” J. Non-Crystall. Solids 354, 5485–5490 (2008). [CrossRef]  

9. K. Hatanaka and H. Fukumura, “Femtosecond laser-induced X-ray pulse emission from transparent materials including glasses,” New Glass 19, 61–66 (2008).

10. L. M. Chen, M. Kando, M. H. Xu, Y. T. Li, J. Koga, M. Chen, H. Xu, X. H. Yuan, Q. L. Dong, Z. M. Sheng, S. V. Bulanov, Y. Kato, J. Zhang, and T. Tajima, “Study of X-ray emission enhancement via a high-contrast femtosecond laser interacting with a solid foil,” Phys. Rev. Lett. 100, 045004 (2008). [CrossRef]   [PubMed]  

11. S. Kahaly, S. K. Yadav, W. M. Wang, S. Sengupta, Z. M. Sheng, A. Das, P. K. Kaw, and G. R. Kumar, “Near-complete absorption of intense, ultrashort laser light by sub-gratings,” Phys. Rev. Lett. 101, 145001 (2008). [CrossRef]  

12. S. Juodkazis, V. Mizeikis, S. Matsuo, K. Ueno, and H. Misawa, “Three-dimensional micro- and nano-structuring of materials by tightly focused laser radiation,” Bull. Chem. Soc. Jpn. 81(4), 411–448 (2008). [CrossRef]  

13. K. Hatanaka, T. Miura, and H. Fukumura, “Ultrafast X-ray Pulse Generation by Focusing Femtosecond Infrared Laser Pulses onto Aqueous Solutions of Alkali Metal Chloride,” Appl. Phys. Lett. 80(21), 3925–3927 (2002). [CrossRef]  

14. K. Hatanaka, T. Miura, and H. Fukumura, “White X-ray pulse emission of alkali halide aqueous solutions irradiated by focused femtosecond laser pulses: a spectroscopic study on electron temperatures as function of laser intensity, solute concentration, and solute atomic number,” Chem. Phys. 299, 265–270 (2004). [CrossRef]  

15. K. Hatanaka and H. Fukumura, “X-ray emission from CsCl aqueous solutions when irradiated by intense femtosecond laser pulses and its application to time-resolved XAFS measurement of I in aqueous solution,” X-ray Spectrom. 41, 195–200 (2012). [CrossRef]  

16. K. Hatanaka, H. Ono, and H. Fukumura, “X-ray pulse emission from cesium chloride aqueous solutions when irradiated by double-pulsed femtosecond laser pulses,” App. Phys. Lett. 93, 064103 (2008). [CrossRef]  

17. W. H. Hsu, F. C. P. Masim, M. Porta, M. T. Nguyen, T. Yonezawa, A. Balčytis, X. Wang, L. Rosa, S. Juodkazis, and K. Hatanaka, “Femtosecond laser-induced hard X-ray generation in air from a solution flow of Au nano-sphere suspension using an automatic positioning system,” Opt. Express 24(18), 19994–20001 (2016). [CrossRef]   [PubMed]  

18. G. I. Taylor, “Formation of thin flat sheets of water,” Proc. R. Soc. Lond. A 259, 1–17 (1960). [CrossRef]  

19. L. Wang, B. Xu, X. Cao, Q. Li, W. Tian, Q. Chen, S. Juodkazis, and H. Sun, “Competition between subwavelength and deep-subwavelength structures ablated by ultrashort laser pulses,” Optica 4(6), 637–642 (2017). [CrossRef]  

20. R. Buividas, M. Mikutis, and S. Juodkazis, “Surface and bulk structuring of materials by ripples with long and short laser pulses: recent advances,” Prog. Quant. Electronics 38, 119–156 (2014). [CrossRef]  

21. M. Trtica, B. Gakovic, B. Radak, D. Batani, T. Desai, and M. Bussoli, “Periodic surface structures on crystalline silicon created by 532 nm picosecond Nd:YAG laser pulses,” Appl. Surf. Sci. 254, 1377–1381 (2007). [CrossRef]  

22. F. C. P. Masim, M. Porta, W. H. Hsu, M. T. Nguyen, T. Yonezawa, A. Balčytis, S. Juodkazis, and K. Hatanaka, “Au nanoplasma as efficient hard X-ray emission source,” ACS Photonics 3(11), 2184–2190 (2016). [CrossRef]  

23. K. Hatanaka, M. Porta, F. C. P. Masim, W.-H. Hsu, M. T. Nguyen, T. Yonezawa, A. Balčytis, and S. Juodkazis, “Efficient X-ray generation from gold colloidal solutions,” arXiv:1604.07541 (2016).

24. M. Yabashi, K. Tamasaku, and T. Ishikawa, “Measurement of X-Ray Pulse Widths by Intensity Interferometry,” Phys. Rev. Lett. 88, 244801 (2002). [CrossRef]   [PubMed]  

25. W. Helml, A. R. Maier, W. Schweinberger, I. Grgurál, P. Radcliffe, G. Doumy, C. Roedig, J. Gagnon, M. Messerschmidt, S. Schorb, C. Bostedt, F. Grüner, L. F. DiMauro, D. Cubaynes, J. D. Bozek, Th. Tschentscher, J. T. Costello, M. Meyer, R. Coffee, S. Düsterer, A. L. Cavalieri, and R. Kienberger, “Measuring the temporal structure of few-femtosecond free-electron laser X-ray pulses directly in the time domain,” Nat. Photon 8(12), 950–957 (2014). [CrossRef]  

26. F. Zamponi, Z. Ansari, C. v K. Schmising, P. Rothhardt, N. Zhavoronkov, M. Woerner, T. Elsaesser, M. Bargheer, T. Trobitzsch-Ryll, and M. Haschke, “Femtosecond hard X-ray plasma sources with a kilohertz repetition rate,” Appl. Phys. A 96, 51–58 (2009). [CrossRef]  

27. M. Holtz, C. Hauf, J. Weisshaupt, A.-A. H. Salvador, M. Woerner, and T. Elsaesser, “Towards shot-noise limited diffraction experiments with table-top femtosecond hard x-ray sources,” Struct. Dyn. 4(5), 054304 (2017). [CrossRef]   [PubMed]  

28. E. G. Gamaly, Femtosecond Laser-Matter Interactions: Theory, Experiments, and Applications (Pan Stanford, 2011).

29. W. Marbach, A. N. Asaad, and P. Krebs, “Optical absorption of solvated electrons in water and tetrahydrofuran/water mixtures,” J. Phys. Chem. A 103, 28–32 (1999). [CrossRef]  

30. F. A. Rajgara, A. K. Dharmadhikari, D. Mathur, and C. P. Safvan, “Strong fields induce ultrafast rearrangement of H atoms in H2O molecule,” J. Chem. Phys. 130, 231104 (2009). [CrossRef]  

31. J. Juodkazytė, G. Seniutinas, B. Šebeka, I. Savickaja, T. Malinauskas, K. Badokas, K. Juodkazis, and S. Juodkazis, “Solar water splitting: efficiency discussion,” Int. J. Hydrogen Energy 41(28), 11941–11948 (2016). [CrossRef]  

32. K. Juodkazis, J. Juodkazyte, B. Šebeka, and S. Juodkazis, “Reversible hydrogen evolution and oxidation on Pt electrode mediated by molecular ion,” Appl. Surf. Sci. 290, 13–17 (2014). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (7)

Fig. 1
Fig. 1 Setup used for X-ray detection using two fs-laser pulses with controlled time separation, Δt; HWP is the half-wave plate, PBS is the polarized beam splitter, RR is the retro-reflector, ODL is the optical delay line, OAPM is the off-axis parabolic mirror, S is the sample solution, GM is the Geiger Müller counter, APS is the automatic positioning system, DS is the displacement sensor.
Fig. 2
Fig. 2 X-ray intensity of a water film irradiation by the main-pulse of 700 μJ at different time delays Δt after the pre-pulse on a ps- (a) and ns-time (b) scales. The pre-pulse energy was 80 μJ. Automatic positioning system was used to find the maximum X-ray intensity [17]. (c) Slices of the X-ray intensity spatial profiles for the time delays from −48.4 ps to 1 ns. (d) A high z-axis resolution plot of the X-ray intensity at Δt = 1.6 ps at which the local maximum was observed in the picosecond range. The double peak is deconvoluted into two Gaussian contributions. X-ray emission without the pre-pulse was more than twice lower in the peak amplitude.
Fig. 3
Fig. 3 (a) Dependence of the solution surface position of the Δt time-delayed main-pulse for the maximum intensity of X-rays from a 20-μm-thick water film (thickness of film is marked by shaded area). Error bars are plotted as standard deviation for 20 pulse measurement. Pre-pulse energy was 80 μJ, the main-pulse was 700 μJ. Inset schematically shows contributions of the main three peak positions at which X-ray emission is maximized. (b) Schematics of two-pulse irradiation of a solution film for the three typical positions of the solution surface when the maximum of X-rays is generated; focusing position is at the main pulse. The width of film and axial extent of focal region for low intensity focus are comparable. (c) Intensity profiles for NA = 0.25 focusing in air; length of pulse ctp ≃ 12 μm. Intensity is presented on logarithmic scale, I = 0.5 at zR is marked by a contour line (corresponds to the depth of focus). For comparison, the width of water film, h, is marked; in experiments, the irradiation angle was θ = 60° and the effective width of water film was twice larger h/ cos θ.
Fig. 4
Fig. 4 X-ray intensity at different positions of the water flow film at 5 ns (a) and 15 ns (b) delays. The fit lines in green were constructed by multi-Gaussian function. Under double-pulsed excitation, the spatial profiles of X-ray intensity are significantly different at various delays. The data set for the main-pulse-only irradiation is shown with 50 counts offset for clarity.
Fig. 5
Fig. 5 X-ray intensity and the axial z-position of the water film at short ps (a) and long ns (b) delays. The positions with the maximum X-ray intensity showing in positive z-axis at ps delays and negative z-axis at ns delays represent up-stream shifts and down-stream shifts as pattern 2 and pattern 3 shown in the inset.
Fig. 6
Fig. 6 FDTD modeling of 800 nm wavelength light (plane wave) traversing water film with several generic surface ripple patterns expected to develop under pulsed excitation. The patterns are: the period Λ = λ on the front surface and depth profile ∼ λ/2, the back-side period λ/n, random pattern of similar size protrusions caused by light filamentation pattern, and initial stages of departing droplets. Insets show schematically the pattern simulated. Thickness of film is taken 10 μm to reduce simulation time; IntensityE2.
Fig. 7
Fig. 7 X-ray emission at different pre-pulse energies at Δt = 1.6 ps when there is most intense peak developed up-stream. The line represents a γ = 2 power law. Sample: water film.

Equations (3)

Equations on this page are rendered with MathJax. Learn more.

( Δ ε d ) im ω pe 2 ω 2 ν e ω = n e n cr ν e ω ,
l abs = λ 2 π κ = n 0 λ π ( Δ ε d ) im = 2 c n 0 ν e n cr n e
W abs = A 0 ν e c n 0 n e n cr F p n e n cr F p ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.