You are on page 1of 6

High Energy Density Physics 3 (2007) 175e180 www.elsevier.

com/locate/hedp

X-ray absorption of a warm dense aluminum plasma created by an ultra-short laser pulse
L. Lecherbourg a,b, P. Renaudin c,*, S. Bastiani-Ceccotti a, J.-P. Geindre a, C. Blancard c, P. Cosse c, G. Faussurier c, R. Shepherd d, P. Audebert a
a

Laboratoire pour lUtilisation des Lasers Intenses, UMR 7605, CNRS-CEA-Universite Paris VI-Ecole Polytechnique, 91128 Palaiseau, France b Universite du Quebec, INRS energie et materiaux, Varennes, Quebec, Canada c Departement de Physique Theorique et Appliquee, CEA/DAM Ile-de-France, BP12, 91680 Bruyeres-le-Chatel Cedex, France ` d Lawrence Livermore National Laboratory, University of California, Livermore, CA 94550, USA Available online 27 February 2007

Abstract Point-projection K-shell absorption spectroscopy has been used to measure absorption spectra of transient aluminum plasma created by an ultra-short laser pulse. The 1s-2p and 1s-3p absorption lines of weakly ionized aluminum were measured for an extended range of densities in a low-temperature regime. Independent plasma characterization was obtained using frequency domain interferometry diagnostic (FDI) that allows the interpretation of the absorption spectra in terms of spectral opacities. A detailed opacity code using the density and temperature inferred from the FDI reproduce the measured absorption spectra except in the last stage of the recombination phase. 2007 Elsevier B.V. All rights reserved.
PACS: 52.50.Jm; 32.30.Rj; 52.25.Os Keywords: Ultra-short laser; Dense plasma; X-ray absorption

1. Introduction The radiative opacities of dense plasmas are important for stellar interior physics [1] and inertial connement fusion plasmas [2]. Experimental investigation of plasma photoabsorption coefcients has been an active eld of research for many years and to test theoretical models measurements of hot dense plasma absorption at well dened temperature and density were measured using volumetrically radiatively heated foils [3]. This method provides uniform heating to temperatures of 80 eV and for densities from 0.001 to a few 0.01 g/cm3 in conditions near local thermodynamic equilibrium (LTE) [3e7]. Higher density plasmas with minimized gradients can be created by intense laser irradiation of solids. When a

* Corresponding author. E-mail address: patrick.renaudin@cea.fr (P. Renaudin). 1574-1818/$ - see front matter 2007 Elsevier B.V. All rights reserved. doi:10.1016/j.hedp.2007.02.035

sub-picosecond laser pulse irradiates a very thin foil (z10s of nm) with intensities in the range of 1015e1016 W/cm2, impulse heating followed by a rapid heat conduction produces a high-density, moderate-temperature plasma before any hydrodynamic expansion occurs [8e10]. Short-pulse X-ray sources of a few picosecond duration emitting in the sub-5-keV range have been generated by irradiating high-Z materials with a sub-picosecond laser pulse [11]. This offers the possibility to use point-projection time-resolved absorption spectroscopy for the study of spectral opacities of dense plasmas. We present quantitative data of transmission spectra in the warm dense matter regime. The 1s-2p and 1s-3p absorption lines of Al were measured for densities up to 1 g/cm3 and temperatures varying from 5 to 30 eV. Our experiment allows one to investigate a higher density plasma regime than that commonly accessible by using volumetrically radiatively heated foils. In the warm dense regime studied, the Al plasma is weakly ionized and density effects are not negligible. The continuum lowering modies the ionization balance and pressure ionization tends to

176

L. Lecherbourg et al. / High Energy Density Physics 3 (2007) 175e180


1s-2p A14+ A15+ 1s-3p A13+ A14+

delocalize atomic orbitals. In the spectral range studied, many 1s-np absorption lines merge in the K-shell photoionization. Therefore, our experimental data allow us to test the accuracy of approximations that have been proposed to account for density effects in spectral opacity calculations. The experimental setup uses two ultrafast lasers to produce both the plasma and the X-ray probe. The rst ultra-short laser pulse created a thin, high-density plasma slab fairly uniform in temperature. An ultra-short X-ray pulse, produced by the second ultra-short laser, was used to backlight the slab. To interpret the absorption spectrum in terms of spectral opacities, it is necessary to (i) infer the density and temperature of the layer during the absorption measurement, (ii) take into account the temporal and longitudinal gradients in the analysis, (iii) check that non-equilibrium effects due to time-dependent population kinetics are negligible at the time of the X-ray probe. All three of these points will be presented and discussed. 2. Experimental setup The experiment was performed at the 100 TW LULI facility, using two 300-fs laser beams. The experimental setup is described in detail previously [12]. A 0.3-J (1.06-mm) laser heated a 45-nm thick Al layer, deposited on a self-standing 25-nm Si3N4 substrate. The heating beam was focused on the Si3N4 side of the target, limiting the longitudinal gradients in the Al. The measured prepulse intensity contrast ratio was 107 at the fundamental frequency. A 8-J energy, 0.53-mm beam focused on a Sm sample using an off-axis parabola to a spot size of 20-mm, produced the backlighter. The time-delay between the X-ray pulse and the heating laser at the target was measured before each shot with an uncertainty of 1 ps. Diagnostics viewed the rear side of the target, i.e., the thin Al layer side. A single-shot frequency domain interferometry (FDI) diagnostic system [13] monitored the phase of a chirped probe laser beam (1.06-mm) of 35-ps duration, reected at a 47 incidence angle on the rear surface of the plasma. X-ray transmission spectra in the 1470e1610 eV spectral region were recorded with a conically-bent potassiumehydrogene phthalate (KAP) crystal spectrometer coupled to a cooled 1024 1024 16-bits CCD camera. The spatial resolution was limited by the X-ray source dimension, i.e., 20 mm. The spectral resolution was estimated from the width of the various 1s-2p lines measured more than 50 ps after the heating pulse, when instrumental width dominates the other broadening processes. The resolution was determined to be 2 0.2 eV. For the spectral region studied here, the Sm plasma provided a 4-ps duration quasicontinuum backlighter spectrum [14]. 2.1. Time-resolved point-projection absorption spectroscopy Fig. 1a shows an example of a space-resolved absorption spectrum measured 12 ps after the heating laser at a laser intensity of 1.5 1015 W/cm2. The spectra covered the 1s-2p absorption lines of Al4 and Al5, the K-edge position at 1560 eV and the Al3 and Al4 1s-3p absorption lines. The

(a)
Space

Laser pulse

Energy

(b)
Transmission

1.

0.9

0.8

Cold target data (12 g/cm2 of Al) 1540 1560 1580 1600

1520

Energy (eV)
Fig. 1. (a) Space-resolved absorption spectrum obtained 12 ps after the heating laser at a laser intensity of 1.5 1015 W/cm2. The solid vertical line gives the position of the Al K-edge at 1560 eV. (b) Lineout outside the focal spot showing the transmission of different cold foils (markers) as well as the adjusted tabulated cold target transmission (solid line).

cold Al K-edge is visible outside the focal spot. The initial areal density of the Al layer was measured for each experiment by comparing the tabulated cold Al and Si3N4 transmission near the K-edge with the measured transmission (Fig. 1b). We found that the areal density of the Al layer was equal to 12 4 mg/cm2. The uncertainty was inferred from the statistic analysis of the different measured transmissions. 2.2. Frequency domain interferometry The experimental setup of the single-shot FDI diagnostic system used here is shown schematically in Fig. 2. The plasma was imaged on the entrance beam-splitter prism of the Mache Zehnder interferometer that was set with slightly unequal arms. The entrance prism was imaged again on the entrance slit of a 1-m CzernyeTurner spectrometer and the spectrum was recorded on a 12-bit CCD camera. The chirp of the beam associated with spectral dispersion provides the timeresolution for the diagnostic. A calibration performed without the main laser beams was necessary to dene the instrumental null phase shift. Focal spot interferometric images, as the one shown in inset of the Fig. 2, are processed to get the complex response of the fringe perturbation, that is, the phase shift and the reection coefcient introduced by the plasma. 3. Plasma characterization 3.1. Plasma expansion The energy is delivered to the target in a time short compared to any hydrodynamic expansion. The density and

L. Lecherbourg et al. / High Energy Density Physics 3 (2007) 175e180


30
31015 W/cm2

177

25

1.51015 W/cm2 51014 W/cm2

20

Phase (rad)
Fig. 2. Diagram of the different component of the FDI diagnostic with the probe pulse reected at the rear side of the target, the MacheZehnder interferometer, and the spectrograph.

15

10

0 0 5 10 15 20

Time

(10-12

s)

temperature during the plasma expansion depend mainly on the maximum temperature achieved at the end of the laser pulse. To infer the plasma parameters at the time of the absorption measurements we have used the one-dimensional hydrodynamic MULTI-FS code [15]. The uncertainty in the laser interaction parameter measurements leads to a non-negligible uncertainty in the laser intensity and, therefore, on the density and temperature calculated with the hydrodynamic code. Nevertheless a precise determination of the plasma parameters is necessary to interpret the absorption spectra in terms of spectral opacities. To overcome this problem, we used the FDI diagnostic system to infer the density and temperature of the expanding plasma during the absorption measurements. Indeed, the phase shift of the reected light provided by the FDI diagnostic depends on the velocity of the critical surface, which is related to the laser intensity [16]. The Helmholtz equation is solved to calculate the propagation of the probe beam in the subcritical plasma and its reection near the critical surface [17]. These calculations provide the phase variation of the probe beam as a function of the laser intensity and were used as a postprocessor for the MULTI-FS code. Fig. 3 shows the temporal evolution of the phase of the chirped probe laser beam reected at the rear critical surface of the plasma. The best agreement between theoretical calculations and experimental data is obtained by adjusting the laser intensity in MULTI-FS. An input laser intensity of IL 10% is determined from the best t of the measured phase shift, whereas IL and the measured laser intensities are equal within 30%. Thus, the accuracy of the FDI analysis allows us to reduce the uncertainty in the plasma parameters. An input laser intensity of IL 10% leads to a 1 eV-temperature variation at the time of the absorption measurements, i.e., a lower variation than the one due to the temporal integration during the backlighter. For each experimental absorption spectrum, FDI analysis provides an independent characterization of the plasma expansion.

Fig. 3. Temporal evolution of the phase of the probe laser beam reected at the rear critical surface of the plasma, showing the best agreement between hydrodynamic simulations performed with MULTI-FS (solid lines) and experimental data (markers).

3.2. Plasma parameters Fig. 4 shows the evolution of the thermodynamic path at different times after the heating pulse (9 1014 W/cm2) for the Al plasma along the absorption axis. MULTI-FS simulations were performed with 45-nm-thick Al layer deposited on 25-nm thick Si3N4 substrate irradiated with laser intensity corresponding to the FDI measurement. The Si3N4 substrate does not contribute to the spectral transmission and allows one to reduce the gradients in the Al layer. The simulations indicate the plasma parameters during the 4-ps of the X-ray probe. The layer temperature is mainly driven by the timedelay between the heating pulse and the backlighter beam whereas the density is mainly dependent on the laser intensity. The temperature gradients are reduced at late times. An error of 10% in the FDI measurement leads to an error of 10% of the densities and the temperatures. 3.3. Stationary effect As our opacity calculations assume steady-state ions population, we have calculated the inuence of non-equilibrium effects on the hydrodynamic evolution of the plasma. The temporal evolution of the average charge and detailed populations of the plasma were calculated with the collisional-radiative code TRANSPEC using the superconguration code AVERROES [18] as a postprocessor of the MULTI-FS code. Fig. 5 shows the temporal evolution of the relative percentage of ground level population of Al ions and mean ionization, Z, calculated with AVERROESeTRANSPEC under stationary or non-stationary conditions. Because of the spatial gradients in density and temperature in Al, the calculated data are weighted by the areal mass of the cells. Non-stationary effects become

178

L. Lecherbourg et al. / High Energy Density Physics 3 (2007) 175e180

1 ps 2 Al 5 3 7 10 0.1 24 30 40 50 0.01 14

4. LTE opacity code The LTE opacity code OPAS is used to interpret the transmission spectra [7]. For given density and temperature, we use an average-atom model [19] to estimate the Z and the subshell statistical occupations. A large set of non-relativistic congurations is then selected. From subshell statistical occupations the optimized potential method is used to compute self-consistent one-electron wave-functions by ionic stage. Conguration occupations are then deduced from SahaeBoltzmann equilibrium including continuum lowering. The chemical potential is adjusted to obtain the Z of the average-atom calculation. The analysis of the relative variation of average subshell occupation numbers between SahaeBoltzmann equilibrium and average-atom calculations allows to optimize the conguration set. The chemical potential adjustment is generally a few percents at most. The one-electron wave-functions are also used to evaluate the one-electron boundebound and boundefree oscillator strength, where the length form of the electric dipole operator is used. Boundefree opacity is evaluated between pair of congurations using the one-electron oscillator strengths. The energy thresholds are computed including a relativistic shifts and orbital relaxation effects. The edges are convolved by a Gaussian prole with a variance equal to the sum of the initial and nal conguration energy variances. Boundebound opacity is calculated performing detailed line accounting in pure jj coupling using a Multi-Congurational DiraceFock code [20]. All lines are calculated using length gauge and the resulting total oscillator strength is reset to the one-electron oscillator strength to preserve the one-electron oscillator strength sum-rule. A Voigt prole is used for each line shape, where the Gaussian width component is given by Doppler broadening and the Lorentzian width includes the radiative decay, the autoionization rate and the electron impact broadening.

Density (g/cm3)

Si3N4 0.001

7 8 9

7 8 9

10

100

Temperature (eV)
Fig. 4. Thermodynamic path of the Al plasma at different times after the heating pulse (black lines). The laser intensity is equal to 9 1014 W/cm2. MULTI-FS simulations were performed with 45-nm-thick Al layer deposited on 25-nm thick Si3N4 substrate irradiated with laser intensity corresponding to the FDI measurement. The Si3N4 substrate does not contribute to the spectral transmission (red lines). (For interpretation of the references to colour in gure legends, the reader is referred to the web version of this article).

non-negligible when the plasma density decreases, i.e., after 15 ps. At these times, the stationary assumption leads to an underestimate of the ionic fraction. In a previous paper [12], we have shown that despite the rapid temporal evolution of the temperature, the ionization state given by AVERROESe TRANSPEC is in good agreement with the measured Z, suggesting that the plasma is close to LTE.
50
<Z>

5.5

40 Al4+

Al3+

5.0

5. Results The OPAS code is used with the density and temperature proles given at the time of the absorption measurements derived from the simulations that reproduces the FDI measurements. The areal densities used to calculate the transmission spectra are equal to 8.5 mg/cm2, in the lower limit of the measured areal density. The calculated transmission spectra are convolved with the spectral resolution, i.e., 2 eV and the spectral proles in the center of the focal spot were extracted from all the space-resolved spectra, taking into account a 20-mm spatial width compatible with the Sm source diameter. Fig. 6 shows lineouts measured for different laser intensities and different probe-times as well as the theoretical calculations. Each absorption spectrum is dened with the intensity of the heating pulse, and delay time between the pump and the probe. Fig. 7 shows the corresponding thermodynamic paths inferred from the FDI analysis. The density and temperature varies from 0.001 to 1 g/cm3 and from 12 to 30 eV, respectively.

4.5

Mean ionization

Pourcentage

30 4.0 20 3.5

10 Al 0 0 5 10 15 20 25
5+

3.0

2.5

Time (10-12 s)
Fig. 5. Temporal evolution of the percentage of ground level population of Al ions and mean ionization Z calculated with AVERROESeTRANSPEC under stationary (solid lines) or non-stationary (dashed lines) conditions (MULTI-FS simulations for a laser intensity of 9 1014 W/cm2).

L. Lecherbourg et al. / High Energy Density Physics 3 (2007) 175e180

179

1.04 1.00 0.96 0.92 0.88 0.84 0.80 0.76 0.72 1480

1s-2p 4+ 5+ 3+

1s-3p 4+

1.04 1.00 0.96 0.92 0.88 0.84 0.80

1s-2p 4+ 5+ 3+

1s-3p 4+

Transmission

Transmission
IL=1.51015 W/cm2 t=t0+7ps

0.76 0.72

IL=1.51015 W/cm2 t=t0+12ps

1520

1560

1600

1480

1520

1560

1600

Energy (eV)
1.04 1.00 0.96 0.92 0.88 0.84 0.80 0.76 0.72 1480 1520 1560 1600 IL=1.51015 W/cm2 t=t0+22ps 0.72 1480 1520 1.04 4+ 4+ 1.00 0.96 0.92 0.88 0.84 0.80 0.76 5+

Energy (eV)

1s-2p 4+ 5+ 3+

1s-3p

3+

4+

Transmission

Transmission

IL=31015 W/cm2 t=t0+51ps

1560

1600

Energy (eV)

Energy (eV)

Fig. 6. Transmission spectra measured at different times after the heating pulse (black lines) compared to the calculated longitudinal-integrated transmission (red lines) with the plasma parameters inferred from FDI analysis (see Fig. 7). (For interpretation of the references to colour in gure legends, the reader is referred to the web version of this article).

The spectra measured at the rst three probe-times, see Fig. 6aec, exhibit extended structures in the Al3 and Al4 1s-3p absorption. They are primarily due to electron impact broadening which is known to be important in high-density and moderate-temperature plasmas [21]. At 7 ps and 12 ps, good agreement is obtained between the experimental spectra and the theoretical transmission. However, the wing absorption of Al4 1s-2p is overestimated, which may be due to the large autoionization width that is too large in the calculations. Satellite lines from excited Al3 congurations and negative shifts of photoionization thresholds due to continuum lowering contribute to enhance the absorption of the red wing of the Al4 1s-3p absorption structure. For an LTE Al

plasma at temperature 16 eV and density 0.09 g/cm3, close to the conditions obtained at 12 ps, the theoretical photoionization threshold shifts are between 21.9 eV and 8.8 eV for an ionic stage between Al6 and Al. Such effects are also present at 22 ps but are negligible at 51 ps. A red wing asymmetry due to unresolved broad resonance and satellite lines of Al can be observed for Al3 1s-3p absorption structure at 12 ps. The experimentally unresolved broad resonance and satellite lines of Al2 are responsible of this asymmetry. Calculations indicate that the blue wing of the Al3 1s-3p absorption structure is unaffected by neighbouring lines. We can test our electron impact broadening calculation based on a semi-empirical method [22], which yield

180
1

L. Lecherbourg et al. / High Energy Density Physics 3 (2007) 175e180

7 ps 12 ps 0.1 22 ps

Density (g/cm3)

varying from 12 to 25 eV. Independent characterization of the plasma parameters allowed an interpretation of the absorption spectra in terms of spectral opacities. Our experimental data are in good agreement with results of our LTE opacity code using a detailed line accounting approach coupled with hydrodynamic simulations. This agreement assists in validating the treatment of the density effects occurring in the experimentally accessed thermodynamic regime.

0.01

Acknowledgement The authors would like to thank O. Peyrusse for his help in the use of the TRANSPEC/AVERROES code and M. Millerioux (CEA/DAM-Ile-de-France) for making the deposited layer of aluminum. We acknowledge the invaluable support of the LULI laser operations staff. LLNL (Livermore, California) provided the various targets. Partial support for Ludovic Lecherbourg was provided by NSERC.

0.001 51 ps

10

12

14

16

18

20

22

24

Temperature (eV)
Fig. 7. Thermodynamic path at different times after the heating pulse (t0 Dt) for the Al plasma (thick lines). MULTI-FS simulations were performed with 45-nm-thick Al layer deposited on 25-nm thick Si3N4 substrate irradiated at a laser intensity (Ilaser). The Si3N4 substrate does not contribute to the spectral transmission and is not shown. The thin lines indicate the area of density and temperature probed during the 4-ps duration of the backlighter.

References
[1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] ` S. Turck-Chieze, et al., Phys. Rev. Lett. 93 (2004) 211102. J.D. Lindl, Phys. Plasma. 11 (2004) 339. T.S. Perry, et al., Phys. Rev. E 54 (1996) 5617. G. Winhart, K. Eidmann, C.A. Iglesias, A. Bar-Shalom, Phys. Rev. E 53 (1996) R1332. C.A. Back, et al., J. Quant. Spectrosc. Radiat. Transfer 58 (1997) 415. H. Merdji, et al., Phys. Rev. E 57 (1998) 1042. P. Renaudin, et al., J. Quant. Spectrosc. Radiat. Transfer 99 (2006) 511. K. Nazir, et al., App. Phys. Lett. 69 (1996) 24. A. Forsman, A. Ng, G. Chiu, R.M. More, Phys. Rev. E 58 (1998) R1248. S.J. Davidson, et al., J. Quant. Spectrosc. Radiat. Transfer 65 (2000) 151. M. Fajardo, et al., Phys. Rev. Lett. 86 (2001) 1231. P. Audebert, et al., Phys. Rev. Lett 94 (2005) 025004; S. Tzortzakis, et al., J. Quant. Spectrosc. Radiat. Transfer 99 (2005) 614. J.-P. Geindre, et al., Optics Lett. 26 (2001) 1612. C. Chenais-Popovics, et al., SPIE Proceedings, vol. 5196, SPIE, Bellingham, WA, 2004, pp. 205. K. Eidmann, J. Meyer-ter-Vehn, T. Schlegel, S. Huller, Phys. Rev. E 62 (2000) 1202. R. Shepherd, et al., J. Quant. Spectrosc. Radiat. Transfer 71 (2001) 711. P. Celliers, A. Ng, Phys. Rev. E 47 (1993) 3547. O. Peyrusse, J. Phys. B: At. Mol. Opt. Phys. 32 (1999) 683; J. Phys. B: At. Mol. Opt. Phys. 33 (2000) 4303. C. Blancard, G. Faussurier, Phys. Rev. E 69 (2004) 16409. J. Bruneau, J. Phys. B 17 (1984) 3009. D. Hoarty, et al., J. Quant. Spectrosc. Radiat. Transfer 99 (2006) 283. M.S. Dimitrijevic, N. Konjevic, Astron. Astrophys. 172 (1987) 345.

half-width at half-maximum of the Al3 1s-3p line is equal to 1.5 eV. When the spectral resolution is taken into account, this becomes 1.8 eV, which is close to experimental value of 1.86 eV. A higher temporal resolution with a shorter-duration backlighter would allow us to test the line broadening calculations in the higher density plasma regime created at the beginning of the laser interaction. At later times, the 1s-3p absorption structures are in good agreement with the experimental absorption structures. As shown in Fig. 5, the non-stationary effects that tend to slow the plasma recombination may be responsible for the underestimate of the Al5 1s-2p absorption. 6. Conclusion In conclusion, spectral opacity measurements of dense plasmas using high-intensity, ultra-short laser have been demonstrated. Absorption spectra of transient Al plasmas have been measured for densities up to 0.3 g/cm3 and temperatures

You might also like