C. Michaut et al- Jump conditions in hypersonic shocks: Quantitative effects of ionic excitation and radiation

Embed Size (px)

Citation preview

  • 8/3/2019 C. Michaut et al- Jump conditions in hypersonic shocks: Quantitative effects of ionic excitation and radiation

    1/12

    Eur. Phys. J. D 28, 381392 (2004)DOI: 10.1140/epjd/e2003-00322-3 THE EUROPEAN

    PHYSICAL JOURNAL D

    Jump conditions in hypersonic shocks

    Quantitative effects of ionic excitation and radiation

    C. Michaut1,a, C. Stehle1, S. Leygnac1, T. Lanz1,2,3, and L. Boireau1

    1 LUTH, UMR 8102 du CNRS, Observatoire de Paris, 5 place J. Janssen , 92195 Meudon, France2 Department of Astronomy, University of Maryland, College Park, MD 20742, USA3 NASA Goddard Space Flight Center, Code 681, Greenbelt, MD 20771, USA

    Received 7 April 2003 / Received in final form 3 October 2003Published online 6 January 2004 c EDP Sciences, Societa Italiana di Fisica, Springer-Verlag 2004

    Abstract. We study the quantitative effects of excitation, ionization, radiation energy and pressure, on thejump conditions in hypersonic shocks in a real gas. The ionization structure and excitation energies arecalculated from the local temperature and density, using the Screened Hydrogenic Model. We assume anoptically thick medium and no radiation flux through the shock front. We investigate the jump conditions indifferent gases and propose a phenomenological description of compression for different shock velocities. Wefind that the excitation energy term is the dominant term in ionized gases at low velocities. Consequently,higher shock velocities than the values predicted by standard calculations in a perfect gas must be reachedin order to observe the effects of radiation in the compression ratio. Our results provide constraints for thedesign of future radiative shock experiments on the next generation of powerful nanosecond lasers or onZ-pinches.

    PACS. 52.35.Tc Shock waves and discontinuities 95.30.Dr Atomic processes and interactions 95.30.Lz Hydrodynamics

    1 Introduction

    Radiative shocks are strong hypersonic shocks. Due to gasheating in the shock, atoms and molecules are excited,dissociated, ionized and, as a result, they radiate awaysome of the absorbed energy. The radiation travels faster,ahead through the shock front, and heats the unshockedmedia creating a radiative precursor. The presence of aradiative precursor often is the criterion which refers theshock as radiative. The total optical depth of the mediumin which radiative shocks occur determines the detailed

    structure of the shock (see, e.g. [1]). Several characteristicregions could be distinguished in a radiative shock oc-curring in an optically-thin, or moderately optically-thickplasma. As mentioned above, the main characteristic zoneis the radiative precursor. Just after the precursor, we findseveral zones where different equilibria are achieved: (1)the kinetic equilibrium for each species of particles, (2) azone where the electronic temperature comes to equilib-rium with the ionic temperature, and possibly (3) a zonewhere radiative equilibrium is achieved. Large departuresfrom the Local Thermodynamic Equilibrium (LTE) arethus a typical feature of the shock region in optically thinplasmas. Conversely, the situation of optically-thick me-

    dia is much simpler: the temperature remains roughly con-a e-mail: [email protected]

    stant in the shock region, and tends to decrease slightlyin the radiative precursor. In astrophysical conditions, ra-diative shocks occur in a wide variety of objects, at verydifferent densities, hence in media with very different to-tal optical depths. Examples include shocks in supernovaremnants, in atmospheres and envelopes of pulsating stars[24], in accretion processes during star formation [5], andin exploding supernovae [6], ranging from (very) small to(very) large total optical depths. With the exception ofthe latter case, departures from LTE and radiative cool-ing are major drivers of the shock structure, in particular

    of the compression ratio.

    Shocks and related discontinuities are a difficult nu-merical problem to solve in radiation hydrodynamics. Itis therefore crucial to validate the computer codes whichare applied to model astrophysical objects or laboratoryplasmas. A major goal of recent experimental work isthus to provide benchmarks for radiation hydrodynam-ics codes [7]. Particular attention has been recently de-voted to the experimental study of 2-D and 3-D expandingblast waves [810] and 1-D radiative shocks [1113]. Theyshow strong hydrodynamical discontinuities and radiativeprecursors. These experiments require high-power energyinstallations like high-power lasers or Z-pinches. The hy-drodynamics and geometry of these shocks are quite differ-ent: the blast waves follow a point-like explosion induced

  • 8/3/2019 C. Michaut et al- Jump conditions in hypersonic shocks: Quantitative effects of ionic excitation and radiation

    2/12

    382 The European Physical Journal D

    by intense ultra-short lasers. The front shock decelerateswith time and can never be stationary. On the other hand,1-D radiative shock experiments are conducted with high-energy lasers with longer duration pulses. The front shock

    velocity remains nearly constant. A radiative precursor de-velops ahead of the shock front, at higher velocities. Theduration of the laser pulses generally remains too short toreach the stationary limit in the whole shock structure.

    The eventual goal is to reach a regime where the wholeshock structure is dominated by radiation. This objectiverequires to drive a shock at higher velocities than val-ues reached in present experiments, as well as to reachthe optically-thick limit. These conditions are hard tofulfill experimentally nowadays. The limited total opticaldepths in current experimental settings imply that radia-tive shock structures in the laboratory are more or lessdominated by radiative losses.

    In order to design such fully radiative 1-D shock exper-iments, Bouquet et al. [14] developed an analytical modelof stationary radiative shocks, assuming a perfect gas,one-dimensional geometry, large total optical depth, LTEand the diffusion approximation to describe the radiationtransport. Relevant to the experimental set-up, Bouquetet al. showed that shocks are more radiative when theambient medium is at low pressure and is constituted ofheavy gas, for example xenon. Following the usual ap-proach in radiation hydrodynamics, Bouquet et al. as-sumed a perfect gas, neglecting thus the excitation andionization of the gas. In an experimental set-up to pro-duce radiative shocks, xenon will be ionized several times

    and a significant fraction of the mechanical energy maytherefore be used to excite and ionize the gas. This effectmay actually dominate the radiation effect.

    In this paper, we revisit the assumption of the per-fect gas, quantitatively estimating the effect of excita-tion and ionization, and of radiative energy and pressure,on the post-shock conditions. We model radiative shocksfor different initial conditions. We solve the GeneralizedRankine-Hugoniot (GRH) relations for ionized gases, ac-counting for radiative energy and radiation pressure as-suming LTE. Whereas the basic equations were alreadywritten down by Zeldovich and Raiser [15], only limitedcomputations for special cases were previously reported.

    For example, the effect of radiative energy and pressurewas studied for perfect gases by [16] and later by [14]; theeffect of excitation and ionization was reported on for spe-cific gases, like nitrogen and oxygen gases [17] but withoutradiation effects, or for hydrogen at high shock velocities,including radiative effects [18]. However, no extensive andquantitative calculations were performed to explore therole of radiation in real gases in different conditions. Ourstudy complements the work of Nieuwenhuijzen et al. [19],who proposed a way to resolve the GRH relations using atabulated Equation of State (EOS). Our approach differsin that we calculate relevant atomic quantities on-the-fly,and focus primarily on deriving the conditions for differ-ent gases where radiation plays a role through the shockfront. We adopt a relatively simple atomic description, theScreened Hydrogenic Model [20,21], which is a reasonable

    description for multiple-charged rare gases. We choose tofocus our study on rare gases because they can be eas-ily handled in experimental conditions. In astrophysicalsettings, hydrogen is the most important case. We will

    consider a detailed atomic and molecular description ofhydrogen in a subsequent paper.Our assumptions are as follows: large total optical

    depth, with very small photon mean path and thusnegligible radiative flux; LTE; stationarity; one-dimensiongeometry; pure gases; and no magnetic fields. Our modeltherefore provides the conditions in the post-shock regionrelative to the initial conditions in the pre-shock region,that is essentially the compression ratio. We aim at assess-ing the effect of excitation, ionization, and radiation, onthis compression ratio, but not at describing the detailedstructure of the shock. This model should not be appliedblindly to astrophysical or experimental settings wherethe total optical depth is generally not as large as as-

    sumed here, where the radiative flux might be important,and LTE breaks down in the shock region. Our intentionis to keep the model simple as to assess the assumptionof perfect gas made in most radiation hydrodynamics cal-culations. For real gases, we will indeed show that thecompression ratio is significantly modified and that the ra-diative effects in the compressed gas arise at higher shockvelocities. In the next stage, we will use this model asthe starting solution of a more realistic model of radia-tive shocks that include a detailed treatment of radiationtransport and departures from LTE.

    This paper is organized as follows. In Section 2, wepresent the Generalized Rankine-Hugoniot relations which

    include the radiation energy and pressure terms. Theatomic model adopted for calculating the relevant micro-scopic quantities is described in Section 3. The characteris-tics of shocks in various gases is presented in Section 4, andthe radiative effects are further discussed in Section 4.3.

    2 Jump relations

    In this section, we introduce the Generalized Rankine-Hugoniot equations which include the contributions of ra-diation energy, Erad, and pressure, Prad. These equations,derived for a stationary 1D shock, give the hydrodynam-ical quantities ahead and behind the shock discontinu-ity [15]. We first apply these equations to the case of aperfect gas with polytropic index .

    Hereafter, is the volumic mass density, Pth the ther-mal pressure, T the temperature, and h the massic en-thalpy. We write the three usual continuity equations(mass, momentum and energy) in a coordinate systemthat moves with the shock front, at the velocity u1

    2u2 = 1u1 (1)

    2u22 + Pth2 + Prad2 = 1u

    21 + Pth1 + Prad1 (2)

    2u2

    h2 +

    1

    2u22

    + u2 [Prad2 + Erad2 ] =

    1u1h1 + 1

    2u21

    + u1 [Prad1 + Erad1] (3)

  • 8/3/2019 C. Michaut et al- Jump conditions in hypersonic shocks: Quantitative effects of ionic excitation and radiation

    3/12

    C. Michaut et al.: Jump conditions in hypersonic shocks 383

    where the indices 1 and 2 stand for the unshocked andshocked parts of the gas, respectively. Therefore, u1 is thevelocity of the unshocked matter relative to the shockfront. The contribution of the radiative fluxes are ne-

    glected, because we solve for quantities far from the shockfront in an optically-thick medium. The radiation energyand pressure are taken at radiative equilibrium. This ap-proximation is only correct at high densities when theopacity is large and the diffusion approximation is validfor the radiative flux [1]. We assume this approximation asan initial description for this paper, deferring an improvedtreatment of the radiative transfer from the shock to theprecursor to a future paper. In our description, we have

    Prad =1

    3Erad =

    1

    3arT

    4. (4)

    2.1 Perfect gas case

    In the case of a polytropic perfect gas, the thermal pres-sure and the enthalpy are respectively given by

    Pth = RT

    Aand h =

    1RT

    A(5)

    where A is the atomic mass of the gas, and the molargas constant R is related to the Boltzmann constant k,and to the number n of particles per unit of volume byR = nkA/. The preceding equations then become

    2u2 = 1u1 (6)

    2u22 + 2

    RT2A

    +1

    3arT

    42 = 1u

    21 + 1

    RT1A

    +1

    3arT

    41 (7)

    2u2

    12RT2A

    +1

    2u22

    + u2

    4

    3arT

    42 =

    1u1

    11RT1A

    +1

    2u21

    + u1

    4

    3arT

    41 . (8)

    The Mach number M is the ratio of the shock speed u1to the sound velocity in the unshocked material, which is

    given for a perfect gas by

    vs,perf =

    RT

    A, (9)

    that is vs,perf = 13 km s1 for kT = 1 eV, = 5/3, and

    A = 1 g/mol (atomic hydrogen gas).As described by Bouquet et al. [14], the variations of

    the compression ratio with the shock velocity u1 show aplateau close to the asymptotic compression ratio with-out radiation, 2/1 = (+ 1)/( 1) (i.e., 2/1 = 4for = 5/3). At higher shock velocities, radiative ef-fects appear and the compression ratio increases towardsthe theoretical asymptotical limit of 7. This is illustratedin Figure 1 for hydrogen, assuming an initial density1 = 5104 g cm3 and different initial temperatures T1.

    Fig. 1. Compression ratio, 2/1, and Prad/Pth for hydrogen,considered as a perfect gas with = 5/3, for different initialtemperatures T1 (kT1 = 0.01 eV (dashed lines, squares), 0.1 eV(diamond), 1 eV (dashed lines, triangles) and 10 eV (circles),versus the shock velocity u1 in kms

    1. The initial density 1is equal to 5 104 g cm3. The curves Prad/Pth for the dif-ferent initial temperatures in bold are indistinguishable on thegraphic scale.

    This figure shows together the compression ratio 2/1

    and the ratio of radiative to thermal pressure.The shock velocity, u1 = urad, for which the radiativeand thermal pressures are equal is independent of andis given by [14]:

    urad = 77/6

    R

    A

    2/3 1ar

    1/6. (10)

    This relation is valid at high Mach number (thus for acompression ratio close to 7). For hydrogen and an initialdensity of1 = 5 10

    4 g cm3, we note indeed that uradis usually independent of the initial temperature T1. Wechoose initial conditions 1 and T1 that correspond tothe experimental conditions in xenon [11]. The variationsof urad with the initial temperature will be discussed inSection 4.3.

    In Figure 2, we report the variations of the tempera-ture T2 in the shocked gas versus the shock velocity. Thistemperature varies, as indicated in [14], as u21 at low veloc-

    ities and as u1/21 at high velocities, whereas it would also

    vary as u21 at high velocities if the radiative terms were ne-glected. Radiative effects reduce the heating of the gas athigh shock velocities. In Figure 3, we show the variationof the compression ratio and the ratio of radiative to ther-mal pressures for hydrogen and xenon versus the shockvelocity for the same initial density (5 104 g cm3)and kT1 = 10 eV. At each velocity u1, we verify thatthe radiative effects are larger in xenon than in hydro-gen: Prad/Pth is larger in xenon by about two orders of

  • 8/3/2019 C. Michaut et al- Jump conditions in hypersonic shocks: Quantitative effects of ionic excitation and radiation

    4/12

    384 The European Physical Journal D

    Fig. 2. Same as Figure 1 for the temperature T2 (in eV) ofthe compressed gas.

    Fig. 3. 2/1 (circles) and Prad/Pth (squares) versus the shockvelocity u1 (kms

    1) for hydrogen (full line, full markers) andxenon (dashed line, empty markers), both considered as perfectgases with = 5/3, for an initial temperatures kT1 = 10 eVand density 1 = 5 10

    4 g cm3. The two compression ratiocurves converge towards 7 at high shock velocities.

    magnitude, which follows from the higher atomic mass ofxenon (A = 131). The kinetic energy is larger at a givenvelocity for the heavier element, resulting in more warm-ing and in larger radiative effects.

    From this simple description of a radiative shock in aperfect gas, we conclude that the radiative effects mod-ify the shock structure at high shock velocities, enhancing

    the compression ratio on one hand, and changing the de-pendency of the shocked gas temperature with the shock

    velocity (in u1/21 instead of in u

    21 without radiative ef-

    fects) on the other hand. In this approach, the radiative

    effects (quantified by the ratio Prad/Pth) are predictedto be larger for heavier gases (at constant initial volumicmass density).

    This description assumes that the mechanical energytaken from the piston is used to compress and heat the gas.It neglects the competitive effect of excitation and ioniza-tion processes which may also take a large fraction of themechanical energy, resulting in a lesser increase of tem-perature and smaller radiative effects (which vary as T4).

    2.2 Generalized Rankine-Hugoniot for ionized gas

    We now consider the case of an atomic gas (atomic num-

    ber Z) which can be partially ionized. We denote by Zthe mean ionization stage which is also the ratio of thenumber of electrons to the number of ions. The excitationenergy per atom is denoted by exc. We adopt hereafterthe energy of the fully ionized isolated element as the ref-erence zero energy. Let and i denote the ionizationstages (0 for neutral, Z for the fully ionized atom) andatomic levels, respectively. Finally, let P,i be the frac-tion of species in ionization stage and atomic state ihaving also energy E,i. One has

    Z =

    =0,Z

    P,i (11)

    exc =,i

    P,iE,i. (12)

    The level populations P,i are derived from Boltzmannstatistics assuming LTE.

    The thermal pressure can be decomposed into ionicand electronic contributions. The ratio between thesetwo contributions is equal to Z. The enthalpy now in-cludes the contribution due to the excitation energy, exc.One has

    Pth = RT

    A(1 + Z) (13)

    h =5

    2RT

    A(1 + Z) + exc. (14)

    The three-equation system therefore becomes

    2u2 = 1u1 (15)

    2u22 + 2

    RT2A

    (1 + Z2) +1

    3arT

    42 =

    1u21 + 1

    RT1A

    (1 + Z1) +1

    3arT

    41 (16)

    2u2

    5

    22RT2A

    (1 + Z2) + exc, 2 +1

    2u22

    + u2

    4

    3arT

    42 =

    1u1 521RT1

    A(1 + Z1) + exc

    ,1 +

    1

    2u21 + u1 4

    3arT41

    .

    (17)

  • 8/3/2019 C. Michaut et al- Jump conditions in hypersonic shocks: Quantitative effects of ionic excitation and radiation

    5/12

    C. Michaut et al.: Jump conditions in hypersonic shocks 385

    We recall that Z and exc both depend on and T,and are calculated for each necessary couple (, T) usedduring the solution process. Consequently, the equationsare strongly nonlinear, and one needs to build an iterative

    algorithm that uses convergence criteria which automati-cally adapt to the stiffness of the system.The differences between the cases of perfect and ion-

    ized gas come from the ionization degree and excitationprocesses that appear in the Z and exc terms. We shallillustrate the role played by these two contributions. Ouraim is to derive the general trends due to both the radia-tion and the ionization effects in the generalized Rankine-Hugoniot equations. In an initial stage, we use simple ef-ficient procedures to compute the excitation energy andthe equation of state. In particular, the equation of state isgiven by the perfect gas law. This approximation is validfor a large range of density and temperature conditions,but fails at high densities. This issue may concern the

    conditions in the shocked gas at very high shock velocitiesand/or at high initial densities. However, at high shockvelocities, the radiation effects tend to dominate the ther-mal pressure, and the details of the equation of state aretherefore not important. The resulting effects of departurefrom the perfect gas equation of state will be discussed ina future paper. The goal of this work is to highlight theradiative effects and the importance of microscopic param-eters on the shock structure, resulting in large differencesin the compression ratios between a perfect gas and a morerealistic gas where ionization is taken into account.

    3 Atomic physics

    This study concerns pure atomic gases like hydrogen, he-lium, neon, argon, krypton and xenon, which are good can-didates for laboratory experiments. Among these gases,hydrogen is very interesting for two reasons. First, manyradiative shocks occur in astrophysics, and hydrogen isthe dominant cosmic species. As radiation is treated inthe blackbody approximation, contributions from heav-ier, less abundant species to the radiative flux, energy andpressure are not relevant. Moreover, the equation of stateof such mixtures is close to the case of pure hydrogen.Second, the simple atomic structure of hydrogen allows to

    better understand the relevant physics before addressingmore complicated gases.Throughout this paper, we adopt an ion-sphere model:

    the Screened Hydrogenic Model (SHM) which allows us tocompute very quickly all necessary atomic quantities forgases having a very large number of electrons. A moredetailed description provided by standard atomic pack-ages requires impracticable computer time and memoryin our context. In the Screened Hydrogenic Model [20, 21],the bound electrons of the ion with nuclear charge Z aredistributed in shells characterized by the principal quan-tum number n. The effects of the angular orbital mo-mentum l are not included. The effects of exchange andinteractions between the electrons are included throughscreening constants [22]. On average, the action on oneelectron by the other electrons consists in a screening of

    Fig. 4. Ionization energies versus the ionization stage for Ne,Ar, Kr and Xe, calculated from SHM.

    the nuclear charge. Each bound state is characterized byan integer population of electrons in this state and by aneffective charge. Density effects are qualitatively incorpo-rated by a simple model of lowering the ionization energy.

    This model stipulates that the lowering of the ionizationenergy is proportional to the number of ions in the consid-ered ionic stage. For ionization stage (0 for the neutral),the ionization lowering energy is related to the volumic and atomic (A) masses as described in [23]:

    E [eV] 35 [eV cm]

    g cm3/A [g]

    1/3. (18)

    Initially, we have adopted the SHM due to its very ad-vantageous coupling to hydrodynamics calculations. How-ever, the predictions of spectroscopic quantities is thenonly qualitative, and this model is not suitable to pre-dict or interpret spectra, which is not our purpose here.In future stages of our study, we plan to use an improvedatomic description for light elements and molecules thatare most important in astrophysical phenomena. At thisinitial stage, however, detailed physical atomic models aretoo unwieldy for our approach that aims at comparing var-ious gases and the effects of the atomic mass.

    To illustrate the SHM, the ionization energies of var-ious ionization stages are reported in Figure 4 for differ-ent rare gases. The Saha-Boltzmann equations allow usto compute the ionization stage of the plasma. We havetested the accuracy of this model for the computation ofthe ionization stage against results obtained with moreaccurate atomic physics, and we found an overall goodagreement.

  • 8/3/2019 C. Michaut et al- Jump conditions in hypersonic shocks: Quantitative effects of ionic excitation and radiation

    6/12

    386 The European Physical Journal D

    Fig. 5. Compression 2/1 (thin solid line, full circles), ioniza-tion Z2 (diamonds), temperature T2 (thin solid line, invertedtriangles) in eV, Prad/Pth (thick) of the shocked hydrogen gas,versus the shock velocity in km s1, including radiation effects.The quantities 2/1, T2 and Z2 are also reported when ra-diation is neglected in the Rankine-Hugoniot relations (dashedlines, empty markers). Initial conditions are kT1 = 0.1 eV and1 = 5 10

    4 g cm3.

    4 Shock characteristics for ionized gases

    4.1 Hydrogen

    In Figures 5 and 6, we analyze the influence of the radia-tion on the compression ratio, 2/1, on the shocked tem-perature T2, and on the ionization stage Z2 for hydrogen.We have assumed an initial density, 1 = 510

    4 g cm3,and unshocked temperatures kT1 equal to 0.1 and 10 eV.

    At kT1 = 0.1 eV, the effects of atomic structure arevery important, resulting in a peak of the compression ra-tio of 10.5 for a shock velocity of 30 km s1. The compres-sion ratio curve is very different from the case of a perfectgas (see Fig. 1). At higher temperature, we see that the

    strong compression peak found at kT1 = 0.1 eV disappearsbecause the initial hydrogen gas is almost fully ionized. Forboth initial temperatures, thermal and radiative pressuresare equal for a shock velocity, urad 1100 km s1. Whenthe radiation pressure becomes larger than the thermalpressure, the effects of the atomic structure on T2 disap-pear, but they are still visible (less than 10%) in the com-pression ratio. Generally, the shocked temperature, T2, isless affected by atomic physics effects than the compres-sion ratio. At high velocities, the compression ratio tendstowards a value of 7 instead of 4 when radiation is ne-glected. The temperature of the shocked gas is loweredwhen radiation is accounted for.

    In Figure 7, we show that the atomic structure stronglyaffects the compression ratio at 0.1 eV. With a compres-sion peak of 10.5 at shock velocity of 30 km s1, this curve

    Fig. 6. Same as Figure 5 for kT1 = 10 eV.

    Fig. 7. 2/1 (thin line, full circles), exc/u2

    1 (thin line, nomarker) in hydrogen as ionized gas, compared to 2/1 (dashedline, empty circles) in hydrogen as perfect gas. Initial condi-tions: kT1 = 0.1 eV and 1 = 5 10

    4 g cm3.

    differs significantly from the perfect gas curve. The com-pression peak is attributed to the strong increase of theexcitation energy due to heating. This increase is alsopresent when radiation is neglected. At these shock veloci-ties, the excitation energy dominates the enthalpy term inequation (14). The radiation effects are negligible at thesevelocities, and the dominant terms in the left member ofequation (17) are the excitation energy and the mechan-ical term in u22. The increasing in exc = exc,2 exc,1is taken from the kinetic energy (lowering thus u22, see

  • 8/3/2019 C. Michaut et al- Jump conditions in hypersonic shocks: Quantitative effects of ionic excitation and radiation

    7/12

    C. Michaut et al.: Jump conditions in hypersonic shocks 387

    Fig. 8. 2/1, Z2, kT2 (eV) of the shocked gas, versus theshock velocity in kms1 for hydrogen at kT1 = 0.1 eV and1 = 5 10

    4 g cm3. These quantities are reported for twoatomic models of pure atomic hydrogen gas: detailed model(full lines, full circles for 2/1 and kT2, full diamonds for Z2)and SHM (dashed lines, no marker, empty diamonds for Z2).

    Eq. (17)); due to mass conservation (Eq. (15)), this re-

    sults in increasing 2. The difference in excitation energy(exc 96.4, in eV), normalized to u

    21 (u1 in km s

    1) isplotted together with the compression peak in Figure 7.We note that the variations of the compression ratio andexcitation energy are qualitatively very similar.

    In Figure 8, we illustrate the effects of the adoptedatomic model on the shock description for the low ini-tial temperature (kT1 = 0.1 eV). We consider the case ofa pure atomic hydrogen gas, and we use both a detailedmodel and the SHM. The detailed atomic model [24,25]includes the effect of pressure ionization by the formalismof level dissolution, and uses exact energies. The varia-

    tions of different quantities, like 2/1, T2, Z2, occur inthe same range of velocities. The amplitude of the com-pression peak differs between the two models with a valueof 6.5 at u1 = 60 kms1 for the detailed atomic model,while the compression ratio reaches a factor of 10.5 atu1 = 30 km s1 in the SHM. As expected, the differencesoccur in the regime where hydrogen is partially ionized.They are due to the qualitative description of the atomicphysics for hydrogen with an ionization energy of 9.7 eVinstead of 13.6 eV and to a qualitative inclusion of pressureionization effects in the SHM. The differences disappearfor larger initial temperatures, because the atomic physicseffects are much reduced. We have shown that the SHMyields results in good qualitative agreement with a betteratomic description, and we postpone further discussion toa future paper.

    Fig. 9. 2/1 (full circles, thin line), kT2 in eV (invertedfull triangles, thin line) in helium for initial conditions kT1 =0.1 eV, 1 = 5 10

    4 g cm3, versus shock velocity in km s1,with (full markers, full lines) and without (empty markers,dashed lines) radiation. The ionization stage Z2 is plottedwith diamonds. Prad/Pth is plotted in thick continuous lineand full squares.

    4.2 Rare gases

    We now turn to radiative shocks in rare gases, from he-lium to xenon. In this section, we choose the same ini-tial temperatures (low, 0.1 eV, and high, 10 eV) thanin hydrogen. We have also used the same initial density(1 = 5 104 g cm3). We have adopted the SHM sincethis atomic model is sufficient to describe qualitative be-haviors of the shocked gas quantities, especially when con-sidering radiative effects occurring at high temperatures,thus at large shock velocities. We have tested the perti-nence of this approximation of the atomic physics againstmore detailed atomic models. We proceeded similarly thanfor hydrogen. We found that differences are negligible athigh temperatures, whereas they are more important at

    low shock velocities and low initial temperature. In thiscase, the differences in compression ratio between SHMand detailed atomic models remain for helium of the sameorder than for hydrogen, and the differences decrease withincreasing atomic number.

    We have solved the GRH relations and derived shockcharacteristics for He, Ne, Ar, Kr and Xe. The correspond-ing compression ratios and temperatures are displayed inFigures 9, 11, 13, 15, 17 for the low temperature and 10,12, 14, 16, 18 for the high initial temperature (10 eV), forthe ionized gas case, with and without radiation.

    The helium case is very similar to hydrogen with apeak in the compression at low temperature which van-ishes at high temperature. This bump results from theexcitation of neutral helium at shock velocities below100 kms1 and from the excitation of He+ at shock

  • 8/3/2019 C. Michaut et al- Jump conditions in hypersonic shocks: Quantitative effects of ionic excitation and radiation

    8/12

    388 The European Physical Journal D

    Fig. 10. Same as Figure 9 for kT1 = 10 eV, 1 = 5 104 g cm3.

    Fig. 11. Same as Figure 9 for Ne (kT1 = 0.1 eV, 1 = 5 104 g cm3).

    velocities between 100 and 200 kms1. At kT1 = 10 eV,the mean ionization degree is about 1.5 at low velocities,reaching the upper limit of 2 around 300 km s1. There-fore, the effect of excitation is strongly reduced for shockvelocities under 200 km s1. It results that the compres-sion ratio goes up almost monotonically towards the ratioof 7 contrary to the case at kT1 = 0.1 eV where an emerg-ing peak is visible in the compression rate, at velocitiesunder 200 km s1.

    To understand the complicate structures of the com-pression ratios for other rare gases, it is necessary to con-

    Fig. 12. Same as Figure 11 for kT1 = 10 eV, 1 = 5 104 g cm3. Prad/Pth curve, not displayed, is very similar tothe same curve in Figure 11.

    Fig. 13. Same as Figure 9 for Ar (kT1 = 0.1 eV, 1 = 5 104 g cm3).

    sider the details of the ionization energies (Fig. 4). Theionization energies show jumps related to the ionic shellstructure (see an example in [26] for xenon). We mightthus expect that these jumps show up in the predictedcompression ratios.

    As pointed out previously, the order of magnitude ofthe excitation energy for the shocked gas is such that itdominates the enthalpy in equation (17) and its variationsare directly reflected in the term, 0.5 u22 of the sameequation. The ratio u22/u21 = 21/22 follows the variationsofexc/u21. This can been seen for example in Figure 19,

  • 8/3/2019 C. Michaut et al- Jump conditions in hypersonic shocks: Quantitative effects of ionic excitation and radiation

    9/12

    C. Michaut et al.: Jump conditions in hypersonic shocks 389

    Fig. 14. Same as in Figure 13 for T1 = 10 eV, 1 = 5 104 g cm3. Prad/Pth curve, not displayed, is very similar tothe same curve in Figure 13.

    Fig. 15. Same as Figure 9 for Kr (kT1 = 0.1 eV, 1 = 5 104 g cm3).

    where we have plotted the variations of the compressionratio, exc/u

    21 in arbitrary units and Z2 for krypton

    at 0.1 eV. The three compression bumps at 7, 100, and380 kms1, are visible on the normalized excitation en-ergy variation. They follow from the variations of excand u21, though the jumps in the variations ofexc ver-sus u1 are less pronounced. They are slightly visible in achange of the slope of the corresponding curve.

    In the minima, the excitation energy varies less rapidlythan u21, and the compression ratio therefore decreases.This occurs when the gas becomes more difficult to ion-

    Fig. 16. Same as Figure 15 for kT1 = 10 eV, 1 = 5 104 g cm3. Prad/Pth curve, not displayed, is very similar tothe same curve in Figure 15.

    Fig. 17. Same as Figure 9 for Xe (kT1 = 0.1 eV, 1 = 5 104 g cm3).

    ize/excite and is connected to the jumps in ionization stepsbetween atomic shells (see Fig. 4). For example, the firstpeak for Kr in Figure 19 (kT1 = 10 eV) is associated tothe progressive ionization of the outer shell, n = 4, andthe second peak to shells, n = 3, and n = 2 (the configura-tion of the ground state of Kr i is (1s2, 2s2, 2p6, 3s2, 3p6,3d10, 4s2, 4p6). We can see in Figure 19 that the meanionization degree of Kr varies between 0 and 7 for shockvelocities from 1 to 35 kms1, which corresponds to anexcitation/ionization ofn = 4 shell. Let us note that it isthe mean ionization degree divided by four (Z2/4) which

  • 8/3/2019 C. Michaut et al- Jump conditions in hypersonic shocks: Quantitative effects of ionic excitation and radiation

    10/12

    390 The European Physical Journal D

    Fig. 18. Same as Figure 17 for kT1 = 10 eV, 1 = 5 104 g cm3. Prad/Pth curve, not displayed, is very similar tothe same curve in Figure 17.

    Fig. 19. 2/1 (full circles), Z2 (full diamonds), normalizedexcitation energy (exc/u

    2

    1) in krypton (no marker) for initialconditions kT1 = 0.1 eV, 1 = 5 10

    4 g cm3, versus shockvelocity in km s1.

    is reported in the figure in order to fit the axis scale. Thesecond peak, between 35 and 200 km s1 corresponds tothe variation of the ionization stage between 8 and 25 andis associated to the n = 3 shell. The last small bump isrelated to n = 2 and n = 1 shells, and is more pronouncedwhen radiation is accounted for.

    In the energy range considered, the initial tempera-tures are far below the necessary energy to fully ionize thestudied rare gases (i.e., 73 eV, 2121 eV, 9739 eV, 56482 eV

    Table 1. Shock velocity urad (in kms1) for which Pth =

    Prad. Top, perfect gas (PG); bottom, ionized gas (IG). Initialconditions are: kT1 = 0.1 eV, 1 = 5 10

    4 g cm3.

    H He Ne Ar Kr Xe

    urad (PG) 720 286 97 63 37 28

    urad (IG) 1100 580 515 450 300 410

    Table 2. u and uT at the same temperature and density fordifferent gases (kT1 = 0.1 eV, 1 = 5 10

    4 g cm3).

    H He Ne Ar Kr Xe

    u 450 300 300 250 250 200

    uT 400 300 250 220 220 200

    and 154090 eV, for He, Ne, Ar, Kr and Xe, respectively),

    explaining therefore the strong influence of the excitationenergy.

    4.3 Discussion

    Different criteria may be defined to quantify radiation ef-fects on macroscopic quantities in the shocked gas.

    A first possibility is the equality of thermal and radia-tive pressures, which defines the velocity urad. Valuesof this velocity are reported in Table 1 for the real andperfect gas cases of the different species. In the case ofthe real rare gas, we find that urad stays in the range

    of 500 kms1

    with increasing atomic number. Whenthe criteria is lowered to Prad/Pth = 0.1, the veloci-ties are equal, in the case of the ionized gas, to 490,260, 270, 250, 210 and 180 for H, He, Ne, Ar, Kr andXe, respectively. They are approximatively smaller bya factor of 2.

    Alternatively, we may determine the shock velocity ufor which the compression ratios 2/1 differ by a givenfactor (say 10%), if radiation is included or not in thecalculation.

    A final choice is to use the mimimum velocity, uT,for which the shocked temperatures T2 differ by 10%when radiation is included or not in the calculation.

    The corresponding velocities uT and u are reportedin Table 2 for an initial temperature of 0.1 eV and agas density of 5104 g cm3. These two quantities arequite similar. They are smaller than urad and decreasewith increasing gas mass.

    We now examine the variations of urad with the initialtemperature and density.

    From equation (10) derived for the case of perfect gas,we expect that urad does not vary with T1. However,the radiative effects in the unshocked gas should be-come more important as temperature increases. Whenthe radiative pressure equals the thermal pressure inthe unshocked gas, urad will tend to zero because thecondition Prad/Pth = 1 is already satisfied at initialconditions. The radiative pressure therefore dominates

  • 8/3/2019 C. Michaut et al- Jump conditions in hypersonic shocks: Quantitative effects of ionic excitation and radiation

    11/12

    C. Michaut et al.: Jump conditions in hypersonic shocks 391

    Fig. 20. Shock velocity urad in km s1 for which Prad = Pth

    for hydrogen (full lines) as a perfect gas (full circles) and asa ionized gas (no marker) at 5 104 g cm3, compared withxenon (dot-dashed lines) as a perfect gas (full diamonds) andas a ionized gas (no marker) at 5 104 g cm3. Molecularequilibrium of H is taken into account in the case of the ionizedgas.

    the hydrodynamics before and after the shock. Thislimit occurs for the perfect gas at a temperature, Tlim,defined by

    Tlim =

    3R

    aA

    1/3, (19)

    which corresponds to 220 eV for hydrogen and 43 eVfor xenon at 1 = 5 104 g cm3. This limit is largerin the case of ionized gas. The corresponding variationsfor hydrogen and xenon are reported in Figure 20.

    From equation (10), we note that urad varies as 1/6. InFigure 21, we have studied numerically the variationsofurad with the initial density in the case of xenon at

    kT1 = 0.1 eV. The two curves indicate a linear varia-tion with 1/6, extending the relation obtained for theperfect gas to ionized gases.

    5 Conclusions

    Radiation hydrodynamics calculations are generally per-formed assuming a perfect gas. In this paper, we have de-veloped a simple radiative shock model where we exploreextensively the differences between perfect and real gases,namely the effects of excitation, ionization, radiation en-ergy and pressure, on the jump conditions in hypersonicshocks. We have assumed a large total optical depth, withvery small photon mean path and thus negligible radiative

    Fig. 21. Variations with 1 of the shock velocity uradin km s1 in xenon as a perfect gas (thin line) and as a ionizedgas (thick line). Initial temperature is kT1 = 0.1 eV.

    flux, LTE, stationarity, 1-D geometry, pure gases, and nomagnetic field. Our model thus provides the macroscopic

    conditions in the post-shock region, far from the shock,relative to the initial conditions. Compared to the perfectgas case, we found that the excitation energy becomes thedominant term in the ionized gas case, and directly af-fects the compression ratio. Lower initial (pre-shock) den-sities or higher initial temperatures result in larger ra-diative effects that are best evidenced in the compressionratio and in the shocked gas temperature. We have pre-sented quantitative estimates for several rare gases whichare most often used in experimental settings. For exam-ple, for xenon at an initial density of 5 104 g cm3 andkT1 = 0.1 eV, we found that radiation effects dominate inshocks with velocities larger than 120 km s1. This limitis significantly higher than the value predicted from theBouquet et al. model [14], assuming a perfect gas. Thisconclusion is crucial in designing future radiative shocksexperiments, at energies which are large enough to tran-scend the excitation and ionization of the gas. At thesehigh energies, shocks will be fully structured by radiation.Such high energies and shock velocities will be reachablewith future high-power lasers (LIL, LMJ in France or NIFin the USA), which will therefore provide the necessaryexperimental setup to study radiative shocks.

    The authors thank K. Eidman for his precious help in thisstudy, J.P.C. Chieze, S. Bouquet, M. Cornille and D. Gilles forhelpful discussions. This work was supported by the FrenchNational CNRS program of stellar physics (PNPS). T. Lanzacknowledges the support of Observatoire de Paris.

  • 8/3/2019 C. Michaut et al- Jump conditions in hypersonic shocks: Quantitative effects of ionic excitation and radiation

    12/12

    392 The European Physical Journal D

    References

    1. D. Mihalas, B. Weibel-Mihalas, Foundations of RadiationHydrodynamics (Dover Pub. Inc., Mineola, New York,1999)

    2. A.B. Fokin, D. Gillet, Astron. Astrophys. 325, 1013 (1997)3. P. Mathias, D. Gillet, A. Lebre, Astron. Astrophys. 341,

    853 (1999)4. M. Scholz, P.R. Wood, Astron. Astrophys. 362, 1065

    (2000)5. C. Stehle, J.P. Chieze, Scientific Highlights 2002, edited

    by D. Barret, F. Combes (EDP-Sciences, Les Ulis, 2002),p. 493

    6. H.A. Bethe, Astrophys. J. 490, 765 (1997)7. A. Calder, B. Fryxell, T. Plewa, R. Rosner, L.J. Dursi,

    V.G. Weirs, T. Dupont, H.F. Robey, J.O. Kane, B.A.Remington, R.P. Drake, G. Dimonte, M. Zingale, F.X.Timmes, K. Olson, P. Ricker, P. MacNeice, H.M. Tufo,Astrophys. J. Suppl. 143, 201 (2002)

    8. K. Shigemori, R. Kodama, D.R. Farley, T. Koase, K.G.Estabrook, B.A. Remington, D.D. Ryutov, Y. Ochi, H.Azechi, J. Stone, N. Turner, Phys. Rev. E 62, 8838 (2000)

    9. J.M. Laming, J. Grun, Phys. Rev. Lett. 89, 125002 (2002)10. K.S. Budil, M. Gold, K.G. Estabrook, B.A. Remington,

    J. Kane, P.M. Bell, D.M. Pennington, C. Brown, S.P.Hatchett, J.A. Koch, M.H. Key, M.D. Perry, Astrophys.J. Suppl. 127, 261 (2000)

    11. X. Fleury, S. Bouquet, C. Stehle, M. Koenig, D. Batani, A.Benuzzi-Mounaix, J.-P. Chieze, N. Grandjouan, J. Grenier,T. Hall, E. Henry, J.-P.J. Lafon, S. Leygnac, B. Marchet,H. Merdji, C. Michaut, F. Thais, Las. Part. Beams 20, 263(2002)

    12. P.A. Keiter, R.P. Drake, T.S. Perry, H.F. Robey, B.A.Remington, C.A. Iglesias, R.J. Wallace, J. Knauer, Phys.Rev. Lett. 89, 165003 (2002)

    13. J.E. Bailey, G.A. Chandler, S.A. Slutz, G.R. Bennett, G.Cooper, J.S. Lash, S. Lazier, R. Lemke, T.J. Nash, D.S.Nielsen, T.C. Moore, C.L. Ruiz, D.G. Scroen, R. Smelser,J. Torres, R.A. Vesey, Phys. Rev. Lett. 89, 095004 (2002)

    14. S. Bouquet, R. Teyssier, J.-P. Chieze, Astrophys. J. Suppl.127, 245 (2000)

    15. Ya.B. Zeldovich, Yu.P. Raizer, Physics of Shock Wavesand high-Temperature Hydrodynamic Phenomena, editedby W.D. Hayes, R.F. Probstein (Dover Pub. Inc., Mineola,New York, 2001)

    16. S.I. Pai, A.I. Speth, Phys. Fluid. 4, 1232 (1961)17. R.A. Alpher, H.D. Greyber, Phys. Fluid. 1, 160 (1958)18. R.G. Sachs, Phys. Rev. 69, 514 (1946)19. H. Nieuwenhuijzen, C. de Jager, M. Cuntz, A. Lobel, L.

    Achmad, Astron. Astrophys. 280, 195 (1993)20. K. Eidmann, Las. Part. Beams 12, 223 (1994)21. B. Karcher, Atomphysikalische Beschreibung Ionenstrahl-

    erzeugter Plasmen (MPQ report, 1991), p. 15822. R.M. More, Adv. At. Mol. Phys. 21, 305 (1985)23. G.C. Pomraning, The equations of Radiation Hydrodyna-

    mics (Pergamon Press, Oxford, 1973)24. C. Stehle, S. Jacquemot, Astron. Astrophys. 271, 348

    (1993)25. D.G. Hummer, D. Mihalas, Astrophys. J. 331, 794 (1988)26. C. Michaut, L. Boireau, M. Cornille, S. Leygnac, C. Stehle,

    Scientific Highlights 2002, edited by D. Barret, F. Combes(EDP-Sciences, Les Ulis, 2002), p. 543