The formation of reverse shocks in magnetized high energy ......The formation of reverse shocks in...

11
The formation of reverse shocks in magnetized high energy density supersonic plasma flowsa) S. V. Lebedev, L. Suttle, G. F. Swadling, M. Bennett, S. N. Bland, G. C. Burdiak, D. Burgess, J. P. Chittenden, A. Ciardi, A. Clemens, P. de Grouchy, G. N. Hall, J. D. Hare, N. Kalmoni, N. Niasse, S. Patankar, L. Sheng, R. A. Smith, F. Suzuki-Vidal, J. Yuan, A. Frank, E. G. Blackman, and R. P. Drake Citation: Physics of Plasmas (1994-present) 21, 056305 (2014); doi: 10.1063/1.4874334 View online: http://dx.doi.org/10.1063/1.4874334 View Table of Contents: http://scitation.aip.org/content/aip/journal/pop/21/5?ver=pdfcov Published by the AIP Publishing Articles you may be interested in A comparison of planar, laser-induced fluorescence, and high-sensitivity interferometry techniques for gas-puff nozzle density measurementsa) Rev. Sci. Instrum. 79, 10E717 (2008); 10.1063/1.2979871 Highly resolved measurements of defect evolution under heated-and-shocked conditionsa) Phys. Plasmas 14, 056314 (2007); 10.1063/1.2720799 The evolution of magnetic tower jets in the laboratorya) Phys. Plasmas 14, 056501 (2007); 10.1063/1.2436479 Dynamic hohlraum radiation hydrodynamicsa) Phys. Plasmas 13, 056301 (2006); 10.1063/1.2177640 Steady supersonically rotating plasmas in the Maryland Centrifugal Experimenta) Phys. Plasmas 12, 055704 (2005); 10.1063/1.1896954 This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 155.198.208.245 On: Thu, 01 May 2014 12:29:46

Transcript of The formation of reverse shocks in magnetized high energy ......The formation of reverse shocks in...

  • The formation of reverse shocks in magnetized high energy density supersonicplasma flowsa)S. V. Lebedev, L. Suttle, G. F. Swadling, M. Bennett, S. N. Bland, G. C. Burdiak, D. Burgess, J. P. Chittenden,

    A. Ciardi, A. Clemens, P. de Grouchy, G. N. Hall, J. D. Hare, N. Kalmoni, N. Niasse, S. Patankar, L. Sheng, R.

    A. Smith, F. Suzuki-Vidal, J. Yuan, A. Frank, E. G. Blackman, and R. P. Drake

    Citation: Physics of Plasmas (1994-present) 21, 056305 (2014); doi: 10.1063/1.4874334 View online: http://dx.doi.org/10.1063/1.4874334 View Table of Contents: http://scitation.aip.org/content/aip/journal/pop/21/5?ver=pdfcov Published by the AIP Publishing Articles you may be interested in A comparison of planar, laser-induced fluorescence, and high-sensitivity interferometry techniques for gas-puffnozzle density measurementsa) Rev. Sci. Instrum. 79, 10E717 (2008); 10.1063/1.2979871 Highly resolved measurements of defect evolution under heated-and-shocked conditionsa) Phys. Plasmas 14, 056314 (2007); 10.1063/1.2720799 The evolution of magnetic tower jets in the laboratorya) Phys. Plasmas 14, 056501 (2007); 10.1063/1.2436479 Dynamic hohlraum radiation hydrodynamicsa) Phys. Plasmas 13, 056301 (2006); 10.1063/1.2177640 Steady supersonically rotating plasmas in the Maryland Centrifugal Experimenta) Phys. Plasmas 12, 055704 (2005); 10.1063/1.1896954

    This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

    155.198.208.245 On: Thu, 01 May 2014 12:29:46

    http://scitation.aip.org/content/aip/journal/pop?ver=pdfcovhttp://oasc12039.247realmedia.com/RealMedia/ads/click_lx.ads/www.aip.org/pt/adcenter/pdfcover_test/L-37/1173914813/x01/AIP-PT/PoP_ArticleDL_0514/aipToCAlerts_Large.png/5532386d4f314a53757a6b4144615953?xhttp://scitation.aip.org/search?value1=S.+V.+Lebedev&option1=authorhttp://scitation.aip.org/search?value1=L.+Suttle&option1=authorhttp://scitation.aip.org/search?value1=G.+F.+Swadling&option1=authorhttp://scitation.aip.org/search?value1=M.+Bennett&option1=authorhttp://scitation.aip.org/search?value1=S.+N.+Bland&option1=authorhttp://scitation.aip.org/search?value1=G.+C.+Burdiak&option1=authorhttp://scitation.aip.org/search?value1=D.+Burgess&option1=authorhttp://scitation.aip.org/search?value1=J.+P.+Chittenden&option1=authorhttp://scitation.aip.org/search?value1=A.+Ciardi&option1=authorhttp://scitation.aip.org/search?value1=A.+Clemens&option1=authorhttp://scitation.aip.org/search?value1=P.+de+Grouchy&option1=authorhttp://scitation.aip.org/search?value1=G.+N.+Hall&option1=authorhttp://scitation.aip.org/search?value1=J.+D.+Hare&option1=authorhttp://scitation.aip.org/search?value1=N.+Kalmoni&option1=authorhttp://scitation.aip.org/search?value1=N.+Niasse&option1=authorhttp://scitation.aip.org/search?value1=S.+Patankar&option1=authorhttp://scitation.aip.org/search?value1=L.+Sheng&option1=authorhttp://scitation.aip.org/search?value1=R.+A.+Smith&option1=authorhttp://scitation.aip.org/search?value1=R.+A.+Smith&option1=authorhttp://scitation.aip.org/search?value1=F.+Suzuki-Vidal&option1=authorhttp://scitation.aip.org/search?value1=J.+Yuan&option1=authorhttp://scitation.aip.org/search?value1=A.+Frank&option1=authorhttp://scitation.aip.org/search?value1=E.+G.+Blackman&option1=authorhttp://scitation.aip.org/search?value1=R.+P.+Drake&option1=authorhttp://scitation.aip.org/content/aip/journal/pop?ver=pdfcovhttp://dx.doi.org/10.1063/1.4874334http://scitation.aip.org/content/aip/journal/pop/21/5?ver=pdfcovhttp://scitation.aip.org/content/aip?ver=pdfcovhttp://scitation.aip.org/content/aip/journal/rsi/79/10/10.1063/1.2979871?ver=pdfcovhttp://scitation.aip.org/content/aip/journal/rsi/79/10/10.1063/1.2979871?ver=pdfcovhttp://scitation.aip.org/content/aip/journal/pop/14/5/10.1063/1.2720799?ver=pdfcovhttp://scitation.aip.org/content/aip/journal/pop/14/5/10.1063/1.2436479?ver=pdfcovhttp://scitation.aip.org/content/aip/journal/pop/13/5/10.1063/1.2177640?ver=pdfcovhttp://scitation.aip.org/content/aip/journal/pop/12/5/10.1063/1.1896954?ver=pdfcov

  • The formation of reverse shocks in magnetized highenergy density supersonic plasma flowsa)

    S. V. Lebedev,1,b) L. Suttle,1 G. F. Swadling,1 M. Bennett,1 S. N. Bland,1 G. C. Burdiak,1

    D. Burgess,2 J. P. Chittenden,1 A. Ciardi,3 A. Clemens,2 P. de Grouchy,1 G. N. Hall,1,c)

    J. D. Hare,1 N. Kalmoni,1 N. Niasse,1 S. Patankar,1 L. Sheng,1,4 R. A. Smith,1

    F. Suzuki-Vidal,1 J. Yuan,1,5 A. Frank,6 E. G. Blackman,6 and R. P. Drake71Blackett Laboratory, Imperial College, London SW7 2BW, United Kingdom2Astronomy Unit, School of Physics and Astronomy, Queen Mary University of London, London E1 4NS,United Kingdom3LERMA, Observatoire de Paris and �Ecole Normale Sup�erieure Universit�e Pierre et Marie Curie,UMR 8112 CNRS, 75231 Paris, France4Northwest Institute of Nuclear Technology, Xi’an 710024, China5Institute of Fluid Physics, China Academy of Engineering Physics, Mianyang 621900, China6Department of Physics and Astronomy, University of Rochester, Rochester, New York 14627, USA7University of Michigan, Ann Arbor, Michigan 48109, USA

    (Received 28 November 2013; accepted 25 February 2014; published online 1 May 2014)

    A new experimental platform was developed, based on the use of supersonic plasma flow from the

    ablation stage of an inverse wire array z-pinch, for studies of shocks in magnetized high energy

    density physics plasmas in a well-defined and diagnosable 1-D interaction geometry. The

    mechanism of flow generation ensures that the plasma flow (ReM� 50, MS� 5, MA� 8,Vflow� 100 km/s) has a frozen-in magnetic field at a level sufficient to affect shocks formed by itsinteraction with obstacles. It is found that in addition to the expected accumulation of stagnated

    plasma in a thin layer at the surface of a planar obstacle, the presence of the magnetic field leads to

    the formation of an additional detached density jump in the upstream plasma, at a distance of �c/xpifrom the obstacle. Analysis of the data obtained with Thomson scattering, interferometry, and local

    magnetic probes suggests that the sub-shock develops due to the pile-up of the magnetic flux

    advected by the plasma flow. VC 2014 AIP Publishing LLC. [http://dx.doi.org/10.1063/1.4874334]

    I. INTRODUCTION

    The collision of a supersonic plasma flow with an obsta-

    cle can lead to the formation of a reverse shock in the flow

    with the structure depending on many parameters—including

    the degree of flow collisionality, radiative cooling of the

    post-shock plasma, and the dynamic importance of magnetic

    fields frozen into the plasma. Shocks formed by supersonic

    plasma flows are ubiquitous in astrophysics and include

    accretion shocks,1 internal shocks in astrophysical jets,2 and

    more locally in the interaction of the solar wind with space-

    craft and planets.3–5 Shocks are also of interest in inertial

    confinement fusion (ICF) research, such as those produced

    by ablated material in laser driven hohlraums and in z-

    pinches.6–8 An understanding of the dynamics of reverse

    shock formation and of their properties is far from complete

    despite considerable computational, observational, and ex-

    perimental efforts. This is particularly true in relation to

    shocks in magnetized, high energy density physics (HEDP)

    plasmas, where a pile-up of the magnetic field could affect

    plasma compressibility and the shock structure, and are

    therefore of interest for scalable studies of astrophysical phe-

    nomena9 and ICF.

    In this paper, we describe and present data from a new

    experimental platform designed to study magnetized shocks

    formed at HED using the ablation flow produced by an

    inverse wire array z-pinch. The plasma flow is produced by

    the ablation of fine metallic wires driven by a MA level elec-

    tric current pulse, which introduces a strong (�1–2 T) mag-netic field, embedded into and advected by the flow and

    sustained for the duration of the experiment (�250 ns). Thecollision of the plasma flow with a conducting (metallic) ob-

    stacle leads to the formation of a reverse shock in a configu-

    ration which allows good diagnostic access for detailed

    studies of the formation, evolution, and properties of the

    shock.

    The experiments show that in addition to the expected

    accumulation of stagnated plasma in a thin layer slowly

    expanding from the surface of the obstacle, an additional

    detached density jump is formed in the upstream plasma

    flow ahead of the stagnated plasma. This additional sub-

    shock is first detected at a distance of �c/xpi from the obsta-cle, and remains almost stationary for the duration of the

    experiment. Initially, at the time of formation, both the den-

    sity and flow velocity jumps in the sub-shock are relatively

    small ( 5)of the upstream flow. Later in time, the measurements by the

    Thomson scattering (TS) diagnostic show a larger (factor of

    �3.5) jump in the flow velocity at the sub-shock, while stillno significant heating of the plasma is observed. The analysis

    of the experimental data suggests that the sub-shock

    a)Paper BI3 6, Bull. Am. Phys. Soc. 58, 25 (2013).b)Invited speaker. E-mail addresses: [email protected] and

    [email protected])Present address: Lawrence Livermore National Laboratory, California

    94550, USA.

    1070-664X/2014/21(5)/056305/10/$30.00 VC 2014 AIP Publishing LLC21, 056305-1

    PHYSICS OF PLASMAS 21, 056305 (2014)

    This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

    155.198.208.245 On: Thu, 01 May 2014 12:29:46

    http://dx.doi.org/10.1063/1.4874334http://dx.doi.org/10.1063/1.4874334mailto:[email protected]:[email protected]://crossmark.crossref.org/dialog/?doi=10.1063/1.4874334&domain=pdf&date_stamp=2014-05-01

  • develops as a result of evolution of a magnetic precursor,

    due to the pile-up of the magnetic flux which is frozen into

    the plasma electrons and advected by the plasma flow.

    The paper is organized as follows. Section II describes

    an inverse wire array z-pinch configuration used for the for-

    mation of supersonic plasma flow, and the experimental

    setup used to diagnose this. Section III presents the experi-

    mental results; included here are a description of the parame-

    ters of the undisturbed plasma flow used in these

    experiments, the formation of a sub-shock structure in the

    reverse shock produced by the interaction of the flow with a

    planar obstacle, and the late time evolution of the sub-shock.

    Section IV discusses the results obtained with particular

    focus on the balance between the flow ram pressure and the

    magnetic field pressure. Section V presents conclusions.

    II. EXPERIMENTAL SETUP AND DIAGNOSTICS

    The experimental set-up used to study the interaction of

    the magnetized supersonic plasma flow with a planar con-

    ducting obstacle is shown in Fig. 1. The plasma flow is pro-

    duced by the ablation of thin metallic (Al) wires driven by a

    1.4 MA, 250 ns current pulse at the MAGPIE pulsed power

    facility.10 The wires are arranged to form an inverse wire

    array z-pinch,11 where the current is applied through a cen-

    tral (cathode) electrode and then returns through the sur-

    rounding array of wires, which are positioned coaxially with

    the central electrode. The formation of the ablated flow in

    this configuration is similar to that of a standard imploding

    cylindrical wire array;12 however, in this inverse setup the

    J�B force drives the ablated plasma radially outwards intothe external region which is initially free from magnetic

    field. The ablated plasma is accelerated to velocity of

    �1� 107 cm/s in the close vicinity of the wires (inside thefirst �1–2 mm), and then propagates with almost constantvelocity. The plasma flow is expected to possess a frozen-in

    magnetic field, the advection of the magnetic field having

    been observed previously in standard wire arrays,13–15

    although not measured for inverse arrays prior to these

    experiments. The wire arrays used here consist of 16 Al

    wires, each 40 lm in diameter and 15 mm in length, andpositioned on a diameter of 20 mm coaxially with the central

    (8 mm diameter) electrode (Figs. 1(a) and 1(b)).

    To create a planar flow for the interaction, the wires

    were arranged in pairs so that the distance between the wires

    in each pair was slightly smaller than the average inter-wire

    separation (an angular separation of 20� for the pair insteadof 2p/N¼ 22.5�). This increased proximity acts to magneti-cally focus the flow from the adjacent wires, and the interac-

    tion between the two plasma streams creates two standing

    oblique shocks, between which the flow is essentially planar.

    The resulting flow then interacts with a planar, 5 or 10 mm

    wide, 10 mm tall, and 15 lm thick Al foil, installed perpen-dicular to the flow at a distance of 10 mm from the wires.

    For this foil position, the flow reaches the obstacle �100 nsafter the start of the driving current. The rate of mass injec-

    tion into the flow is determined by the ablation rate of the

    wires, which is described with reasonable accuracy by the

    “rocket” ablation model,12 giving an ablation rate dm/dt/ I2,where I¼ 1.4� 106 sin2(pt/500 ns) is a good approximationof current applied to the wire array. For the rising current

    pulse (first 250 ns experimental time), the mass density in the

    flow increases with time and the spatial distribution of mass

    density for a fixed moment in time can be found using the

    ablation rate and taking into account the time-of-flight deter-

    mined by the flow velocity.11,16

    The collision of the plasma flow with an obstacle leads to

    accumulation of material in front of the foil and the formation

    of a reverse shock in the incoming plasma. The high degree of

    azimuthal symmetry of the system and the formation of eight

    nominally identical plasma flows allows several obstacles to

    be used simultaneously to create additional interaction regions

    to maximize the amount of information that can be collected

    by different diagnostics in each experiment. This also allows

    parameters of undisturbed plasma flow to be measured in the

    same experiment, as illustrated in Fig. 1(b).

    A number of complementary diagnostics were used to

    study the interaction of the flow with the obstacle. Gated

    multi-frame cameras were used to obtain time-resolved

    images of the interaction region and follow the temporal evo-

    lution of the shocks. An optical gated camera (12 frames

    with 5 ns exposure time and 10 or 20 ns inter-frame separa-

    tion) imaged the interaction from the end-on (z) direction,

    while the extreme ultraviolet (XUV) camera17 (photon

    energy >30 eV, 4 frames gated at 2 ns with 10 ns or 30 nsinter-frame separation) recorded side-on images. Laser prob-

    ing (532 and 355 nm, 0.3 ns) of the interaction region was

    performed in side-on and end-on directions using interferom-

    etry and shadow channels. Details on the interferometer

    FIG. 1. (a) Side-on and (b) end-on views of the experimental setup and the

    diagnostics. (c) The ion feature of the TS spectra measured in the upstream

    plasma. kTS is the Doppler-shifted wavelength of the spectrum and k0 is thatof superimposed stray light at the original wavelength.

    056305-2 Lebedev et al. Phys. Plasmas 21, 056305 (2014)

    This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

    155.198.208.245 On: Thu, 01 May 2014 12:29:46

  • set-up and of the analysis technique used to obtain electron

    density distributions from the interferograms can be found in

    Ref. 18. The magnetic field in the plasma flow was measured

    using pairs of identical, overlapped, inductive probes with a

    loop area of 2.7mm2 (1.55 � 1.75 mm) and which werewound in opposite directions. The probes were shielded by a

    thin layer of conductive paint to suppress possible electro-

    static noise from variations of the plasma potential, and sig-

    nals from each probe were recorded separately.

    A TS diagnostic15 was used to measure the plasma flow

    velocity and plasma temperature using the ion feature of the

    scattering spectra. For the parameters of the plasma in these

    experiments (Te� 20 eV, ne� 5� 1017�2� 1018 cm�3), theTS diagnostic operates in the collective scattering regime,19

    with a scattering parameter a¼1.5–4. In most of the experi-ments, the focused probing laser beam (532 nm, 5 ns

    FWHM, 3 J, �200 lm spot size) propagated through thecentre of the array as shown in Fig. 1(b). The beam propa-

    gated through a small hole of �1 mm diameter in the middleof the obstacle foil and through a hole in the central elec-

    trode, or, in some of the experiments, in the opposite direc-

    tion through the electrode towards the obstacle. A different

    probing direction was also used in early experiments, with

    the laser beam passing next to the central electrode in the

    direction towards the foil. In all cases, the scattered signal

    was observed in the r-h plane, in the directions parallel to thefoil surface (Fig. 1(b)). Several spectra corresponding to dif-

    ferent spatial positions along the probing laser beam were

    obtained using fiber optic bundles with 7 or 14 fibers

    (200 lm or 100 lm diameter, respectively), which deliveredscattered light to the input of a 0.5 m ANDOR SHAMROK

    imaging spectrometer, coupled with an ANDOR intensified

    charge-coupled device (ICCD) camera which provided a gat-

    ing time of 4 ns. The spectral resolution was determined by

    the size of the optical fibers used (100 or 200 lm) and for the2400 g/mm grating used in these experiments was equal to

    0.25 A or 0.45 A, respectively.

    III. EXPERIMENTAL RESULTS

    A. Overview of the flow-obstacle interaction

    We begin presentation of the experimental results by

    describing the overall structure developing during the inter-

    action between the plasma flow and the obstacle. Fig. 2

    shows typical laser probing images of the interaction region

    obtained using the perpendicular, side-on probing direction.

    These images were obtained at t � 260 ns after the start ofthe current and correspond to a late phase of the interaction,

    when all the observed features are well defined. The first fea-

    ture observed to form is a dense thin layer of plasma at the

    surface of the foil, which at the time of this image is

    �0.7 mm thick (Figs. 2(a) and 2(c)). Measurements per-formed at different times and with different diagnostics

    (laser probing, optical and XUV imaging) show that this

    layer becomes observable from �90–100 ns after the start ofthe current and expands with a constant velocity of

    �4� 105 cm/s. The delay between the appearance of thestagnated plasma at the obstacle and the start of the driving

    current agrees well with the time-of-flight needed for the

    plasma flow to reach the obstacle. The stagnation layer has a

    sharp density jump and the initial rise of the electron density

    is observable by the interferometry. There is also an increase

    in intensity of XUV emission and an increase in the ioniza-

    tion state of Al ions, as observed by spatially and temporally

    resolved XUV spectroscopy, indicating a temperature

    increase. The details of the measurements of the properties

    of this stagnation shock will be published separately.

    The most prominent feature seen in the images is a

    detached, shock-like structure (which we refer to as a “sub-

    shock”) formed upstream from the obstacle and positioned at

    this time at a distance of �3.5 mm from the obstacle surface.Analysis of the interferogram using the procedure described

    in Ref. 18, yields a map of electron line density (neL) which

    is shown in Fig. 2(b). From this density map it is seen that

    the increase of the density at the sub-shock is relatively

    FIG. 2. Formation of a sub-shock during the collision of the plasma flow with a planar Al foil obstacle: (a) the side-on 532 nm interferometry recorded at

    t¼ 257 ns, (b) the corresponding map of line-integrated electron density (neL), there “0” of the horizontal scale corresponds to obstacle surface, (c) side-onshadow image obtained under the same conditions and a pre-shot image (d). In all images the plasma flow is from left to right, with the foil obstacle at 10 mm

    from the ablating wires.

    056305-3 Lebedev et al. Phys. Plasmas 21, 056305 (2014)

    This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

    155.198.208.245 On: Thu, 01 May 2014 12:29:46

  • small: not more than a factor of �2, while the shape of theinterference fringes in Fig. 2(a) gives an apparent width of

    the density jump of �0.4 mm. The actual width of the den-sity jump is smaller as due to the presence of a small curva-

    ture in the sub-shock (in the r-h plane), which is seen in theend-on images at late times, and is responsible for the over-

    estimation of the width of the transition as measured by the

    interferometry. The shadow laser probing images (Figs. 2(c)

    and 2(d)), which are sensitive to the electron density gra-

    dients, suggest that the width of the sub-shock is indeed

    ��0.1 mm.The rest of Sec. III describes the formation, properties,

    and the evolution of the sub-shock, starting from the descrip-

    tion of the parameters of the undisturbed plasma flow collid-

    ing with the obstacle.

    B. Parameters of the upstream plasma flow

    The parameters of the plasma in the upstream flow were

    determined by comparing the measurements obtained with

    and without the obstacle present. Such measurements were

    performed either in the same experiment using flows pro-

    duced by different pairs of wires of the inverse wire array (as

    illustrated in Fig. 1(b)) or in different experiments using the

    observed high symmetry and reproducibility of the plasma

    flow generation. These measurements show that the plasma

    flow parameters in the upstream region are not affected by

    the presence of the obstacle, as is expected for a supersonic

    flow. The density distribution was measured using side-on

    interferometry images similar to that shown in Fig. 2(a), and

    also using end-on interferometry images discussed later in

    the paper. These measurements show that for the times rele-

    vant to the experiments described in this paper the electron

    density at positions just upstream of the sub-shock is in the

    range of 4� 1017 cm�3 (for t� 180 ns with the foil at 10 mmfrom the wires) to 2� 1018 cm�3 (for t� 250 ns with the foilat 7 mm from the wires).

    The velocity of the plasma flow and the product ZTewere measured using the Thomson scattering diagnostic and

    Fig. 1(c) shows a typical TS spectra. The Doppler shift of the

    ion feature provides measurements of the plasma flow

    velocity, which for the upstream flow is equal to

    Vfl¼ 1.1� 107 cm/s at early times (�180 ns), decreasing to0.9� 107 cm/s later (at �250 ns). The shape of the TS spec-tra shows the well-separated ion-acoustic peaks, which indi-

    cates that ZTe is much greater than the ion temperature Ti,19

    with the fitting of the theoretical scattering form-factor yield-

    ing ZTe¼ 60 eV for the upstream plasma flow. The ionsound speed calculated using this value of ZTe is equal to

    Cs¼ (cZkBTe/mi)1/2� 2� 106 cm/s, giving the sonic Machnumber of the upstream plasma Ms¼ 5. Simultaneous meas-urements of velocity in the interaction region and in an

    unperturbed flow on the opposite side of the array (pre-

    formed as shown in Fig. 1) show that the presence of the ob-

    stacle only changes the flow parameters at positions close to

    the foil surface, but does not change Vfl and ZTe upstream of

    the sub-shock. In some of the experiments, the TS spectra

    were recorded simultaneously from two opposite (690�)directions to the probing beam, which allowed the full

    velocity vector to be determined in the r-h plane (similar tothat discussed in Ref. 20). These measurements show that in

    the region of the plasma flow between the two standing

    oblique shocks (Fig. 1), the plasma velocity direction is nor-

    mal to the obstacle, and that the velocity component in the

    direction parallel to the obstacle surface is negligible

    (

  • the probe signals (dB/dt) in the absence of the obstacle

    repeats the shape of the derivative of the driving current

    (dI/dt), which is consistent with the advection of the mag-

    netic field by the plasma flow, as would be expected for

    plasma flow with a magnetic Reynolds number (ReM) larger

    than unity. Indeed, for the measured parameters of the

    upstream plasma (Vflow¼ 107 cm/s, ZTe¼ 60 eV) and a char-acteristic spatial scale of the distance between the wires and

    the obstacle (10 mm), the magnetic Reynolds number is suf-

    ficiently large, ReM� 50–100.Comparison of the probe signals obtained both with and

    without the obstacle (Fig. 3) show that they behave identi-

    cally until �175 ns. After that, which agrees well with thetime when the sub-shock is first detected, a larger magnetic

    field is observed with the obstacle present. We should how-

    ever emphasise that the magnetic probe measurements do

    not provide quantitative information on the field strength

    behind the sub-shock due to the relatively large size of the

    probes which is comparable to the distance between the sub-

    shock and the obstacle.

    Integration of the dB/dt signals yields the magnetic field

    evolution (Fig. 3(c)), showing that at the time of the sub-

    shock formation at t� 170 ns the field magnitude reaches�1 T. (We note here that while the probes are designed tominimize possible perturbations to the flow and to the meas-

    urements of the upstream magnetic field, it is difficult to

    fully exclude possibility of this from the probe measure-

    ments alone. However, recently performed measurements24

    with Faraday rotation technique show the same upstream

    magnetic field as measured by the probes, thus suggesting

    that these perturbations are relatively small.) Using this

    value of the magnetic field, and estimating the mass density

    of the flow from the measured upstream electron density and

    the measured ZTe¼ 60 eV (corresponding for Al to an aver-age ionisation of Z� 4 and T� 15 eV), we can estimate theAlfv�en velocity as VA� 1� 106 cm/s. Thus, we can con-clude that the upstream plasma flow is both supersonic

    (Ms� 5) and super-Alfv�enic (MA¼Vfl/VA� 8), and the flowis supersonic with respect to the fast magneto-sonic speed

    (VMS¼ (Cs2þVA2)1/2), with MMS¼Vfl/VMS� 4.5.The presented characterisation of the upstream flow pa-

    rameters shows that the sub-structure in the reverse shock

    observed in these experiments is formed as a result of the

    interaction of a super-magnetosonic flow with the obstacle.

    The level of magnetic field advected by the flow is relatively

    small, in the sense that the ram pressure of the flow is signifi-

    cantly larger that the magnetic field pressure

    (qV2/(B2/8p)� 100), though the thermal plasma beta of theflow is not large (bth� 3).

    C. Formation of the sub-shock

    Fig. 4 shows typical images of the flow-obstacle interac-

    tion observed by end-on diagnostics. The optical image in

    Fig. 4(a) obtained at t¼ 210 ns after the current start showsthe plasma streams from a pair of ablating wires and the two

    standing oblique shocks indicated in Fig. 1, which are analo-

    gous to the oblique shocks observed in standard Al wire

    arrays.18 Strong optical emission is also seen from a thin

    layer positioned immediately at the foil surface. The thick-

    ness of this layer increases in time with a constant velocity

    of only 4� 105 cm/s, starting from zero at t¼ 90 ns when thefirst plasma arrives at the obstacle, and reaching �0.6 mm att¼ 230 ns after the current start. Laser probing images (Fig.4(b)) also show a narrow opaque region of the same thick-

    ness in front of the foil surface, and the formation of this

    layer is consistent with the accumulation of stagnated plasma

    at the foil.

    The additional shock-like feature, the sub-shock, is seen

    upstream from the stagnation layer at a distance �2 mm fromthe foil. The optical camera images show that this feature

    becomes detectable at t¼ 170 ns as a region of enhancedemission, which expands with a speed of �1�1.5� 106 cm/s(�10%�15% of upstream flow velocity), reaching a position�3 mm from the foil at t¼ 250 ns. Fig. 4(b) shows an inter-ferogram obtained at t¼ 225 ns using probing in the end-ondirection. Tracing the positions of interference fringes and

    comparing them with the fringe positions recorded before the

    experiment allows the distribution of the electron density to

    be reconstructed following the procedure described in Ref. 18.

    The resulting electron density map (Fig. 4(c)) shows that the

    density inside the two oblique shocks is fairly uniform in the

    direction parallel to the foil, which is consistent with the 1-D

    character of the flow. The density map shows an increase of

    the electron density at �2 mm from the foil, at the same posi-tion where the sub-shock is seen in other diagnostics. The

    FIG. 4. Experimental images showing the interaction of the plasma flow with the obstacle, including (a) end-on fast-frame optical emission, (b) end-on inter-

    ferometry, and (c) the areal electron density (neL) obtained from the interferogram. In all images plasma flow is from the left, with the foil surface at the right

    (“0” of the horizontal scale corresponds to obstacle surface).

    056305-5 Lebedev et al. Phys. Plasmas 21, 056305 (2014)

    This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

    155.198.208.245 On: Thu, 01 May 2014 12:29:46

  • electron density measured in the upstream region immediately

    before the density jump is equal to 5� 1017 cm�3, and themaximum density in the downstream region directly before

    the dense stagnated layer is�1.3� 1018 cm�3.Fig. 5 shows spatial profiles of the flow velocity meas-

    ured with the TS diagnostic at two different times, 180 ns

    and 225 ns, together with electron density profiles measured

    using end-on interferometry in the same experiments at the

    same times. Velocity profiles show that only a relatively

    small drop in the flow velocity occurs at the sub-shock, to

    about 60% of the upstream value. Thereafter, the velocity

    continues to decrease reaching �30% of the upstream veloc-ity at the boundary with the dense plasma layer accumulated

    at the foil surface. The electron density profiles show that

    there is only a small electron density increase at the sub-

    shock, of not more than a factor of �2 in. comparison to theupstream density, and the data show a steepening of the elec-

    tron density profiles with time (Fig. 5).

    We should note here that the interferometry provides

    density averaged along the probing path, and it is possible

    that due to the presence of a small curvature of the sub-

    shock in the r-z plane the density immediately after the sub-

    shock could be slightly higher, with the thickness of the

    transition being slightly smaller. In contrast, the Thomson

    scattering provides local measurements of the flow velocity,

    and is thus a better indicator of the flow behaviour at the

    sub-shock.

    It is instructive to compare the measured electron den-

    sity profiles ne(x) with mass density distributions which can

    be estimated using the measured velocity profiles. For a

    steady flow with time-independent density in the upstream

    flow, the mass density should be inversely proportional to

    the flow velocity, such that q(x)� 1/Vfl(x). Density profilescalculated in this way are shown in Figs. 5(b) and 5(d). For

    the temporally increasing density of the flow, this estimate

    represents the upper limit for the mass density profile.

    Comparison of the electron and mass density profiles indi-

    cates there could be some small increase in the average

    plasma ionization at the sub-shock (due to the divergence of

    their profiles in the flow immediately downstream). This is

    consistent with the results of the TS measurements, which

    show only a small increase in ZTe, from �60 eV in theupstream flow to �110 eV immediately after the sub-shock,decreasing to �90 eV at the boundary with the dense stag-nated layer. The TS data also do not show a measurable

    increase of the ion temperature, the ion-acoustic peaks

    remain well separated meaning that Ti remains much smaller

    than ZTe. The absence of measurable heating of the ions will

    be discussed in more detail later in Sec. III D.

    These results, the small level of velocity and density

    jumps at the sub-shock together with only a small tempera-

    ture increase, suggest that this feature cannot be explained as

    a “standard” shock formed by a high Mach number flow

    (MMS� 5) which should produce a much larger jump of �4(or higher if the effective adiabatic index c is smaller than5/3). Instead, we interpret the observed global structure of

    the interaction region as the development of a sub-shock at

    the front edge of a magnetic precursor created by the pile-up

    of the magnetic flux, advected by the flow towards the obsta-

    cle. Stagnation of the plasma flow containing a frozen-in

    magnetic flux creates (in the reference frame of the obstacle)

    a motional electric field E¼�V2�B2 which drives an elec-tric current on the surface of the obstacle (Fig. 6). The corre-

    sponding perturbation of the magnetic field in the plasma

    (field pile-up) acts differently on the electrons and ions of

    the flow. At a spatial scale of the order of the ion inertial

    length c/xpi the ion and electron responses decouple and theperturbation of the electron pressure could propagate ahead

    of the stagnation shock, e.g., as a whistler mode,21 creating a

    magnetic precursor and ultimately leading to the formation

    FIG. 5. Plasma flow velocity profiles measured from Thomson scattering at (a) t¼ 180 ns and (c) 225 ns, as well as profiles of electron density (interferometry)and relative mass density calculated from the velocity profiles at these times (b) and (d). The sub-shock is at xs� 2 mm from the obstacle (at x¼ 0).

    056305-6 Lebedev et al. Phys. Plasmas 21, 056305 (2014)

    This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

    155.198.208.245 On: Thu, 01 May 2014 12:29:46

  • of a cross-shock electric field which would decelerate the

    ions. The distance between the sub-shock and the stagnated

    layer (�1.5 mm) at the time when it becomes detectable isindeed very close to the c/xpi� 1 mm. The observed reduc-tion in the flow velocity (V/Vfl� 1/2 at the sub-shock) corre-sponds to the presence of the decelerating potential of

    Du� 3/4Ei/(eZ)� 300 V. The presence of an electric poten-tial at the sub-shock is consistent with detection of reflected

    ions, which are observed by Thomson scattering (Fig. 5(a)),

    showing the reflected ions as an additional oppositely shifted

    feature in the scattering spectra. The reflected ions are only

    seen in the upstream flow immediately in front of the

    sub-shock, and not between the sub-shock and the stagnation

    plasma layer. The average velocity of these reflected ions is

    the same as that in the incoming flow (Fig. 5(a)), which is

    consistent with reflection by the electric potential. The

    reflected ions are only observed at the time just after the

    sub-shock formation, and are not seen at later times, suggest-

    ing that the reflection is a transient phenomenon accompany-

    ing formation of the potential, but further investigation is

    needed here.

    D. Parameters of the sub-shock at late time

    The experiments show that the sub-shock formed ahead

    of the obstacle retains its overall structure for the duration of

    the experiment, which is at least �150 ns after it is firstformed. Side-on XUV images in Fig. 7 obtained in two dif-

    ferent experiments at t� 170 ns and t� 270 ns after the cur-rent start show that the sub-shock remains uniform and keeps

    the same shape, but the distance from the obstacle gradually

    increases with time at a speed of 1.5� 106 cm/s (�15% ofthe flow velocity). It is seen that at later time the intensity of

    emission becomes more concentrated at the sub-shock front,

    with a more pronounced decrease towards the obstacle sur-

    face. The apparent steepening of the sub-shock is seen by

    several diagnostics, but the most accurate and detailed meas-

    urements of the sub-shock parameters are provided by the

    Thomson scattering diagnostic.

    Fig. 8 shows results from TS measurements obtained at

    t¼ 216 ns after the current start for an obstacle positioned at7 mm from the wires. The smaller distance between the wires

    and the obstacle did not affect the sub-shock formation and

    evolution, but in comparing the measurements obtained for

    the two different positions of the obstacle one needs to take

    into account that the closer position of the obstacle effec-

    tively corresponds to the later stages of the sub-shock evolu-

    tion, approximately equivalent to a time-of-flight delay of

    �0.3 cm/107cm/s¼ 30 ns.The image in Fig. 8(a) represents the raw TS data from

    14 spatial positions along the probing laser beam, which

    propagates in the flow direction, perpendicular to the obsta-

    cle surface (as illustrated in Fig. 1(b)). Accompanying the

    raw data are the corresponding spectral fits for the scattered

    FIG. 6. A cartoon illustrating the formation of the sub-shock. The motional

    electric field drives the current (��V2 � B2) along the conducting obstaclesurface and some fraction of it closes at the edges of the plasma flow, creat-

    ing a region of enhanced magnetic field (magnetic precursor) between the

    obstacle and the sub-shock (which is formed at a distance of �c/xpi).

    FIG. 7. XUV images obtained in two different experiments showing the

    temporal evolution of the sub-shock. The edge wire of the array is seen at

    the left boundary of the images, with the obstacle foil positioned at a dis-

    tance of 10 mm from this and visible on the right. The first image (177 ns)

    corresponds to the time close to the sub-shock formation time. The two other

    images show the propagation of the sub-shock.

    FIG. 8. Data from the Thomson scattering diagnostic. (a) The scattered spec-

    trum obtained for 14 spatial positions covering a length of �4 mm along thedirection perpendicular to the obstacle surface, from the foil (top) towards

    the wires. The four spectra at the bottom correspond to the upstream flow.

    (b) The ion feature of the Thomson scattering spectra measured in the

    upstream flow and (c) after the sub-shock. The Doppler shift of the spectra

    provides measurement of the flow velocity, while the fit to the scattering

    form-factor provides measurement of ZTe.

    056305-7 Lebedev et al. Phys. Plasmas 21, 056305 (2014)

    This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

    155.198.208.245 On: Thu, 01 May 2014 12:29:46

  • signals recorded at positions upstream and downstream of

    the sub-shock (Fig. 8(b)).

    The spatial profile of the flow velocity obtained from the

    Doppler shift of the scattering spectra is shown in Fig. 9(a).

    It is seen that at this late time the velocity reduction across

    the sub-shock has increased to a factor of �3.4, which is sig-nificantly larger than the jump of only �2 observed for theearlier times. Such larger velocity jumps of �3.5�4 werereproducibly measured in all experiments corresponding to

    the late stages of the sub-shock evolution, and could, on a

    first glance, be interpreted as an evidence of the development

    of a strong shock. The distribution of the average electron

    density measured with the end-on interferometry in the same

    experiment (Fig. 9(b)), with ne determined by dividing the

    measured neL by the height of the shock, does not show a

    large density jump. This is, however, primarily due to the

    presence of a curvature of the sub-shock front in the r-z

    plane, which leads to a significant underestimation of the

    electron density immediately (inside the first �1 mm) afterthe sub-shock. Finding the local electron density distribution

    from the neL measurements could be potentially possible by

    combining the side-on and end-on interferometry measure-

    ments, but this has not yet been attempted here.

    Another way to obtain information about the local den-

    sity distribution is by using the measured relative total

    intensities of the TS signals (integrated over wavelength)

    and Fig. 9(b) shows the corresponding spatial profile

    obtained from the scattering spectra shown in Fig. 8. It is

    seen that the TS intensity distribution has an abrupt increase

    by a factor of �3.5 at the position of the sub-shock. At thesame position the velocity decreases by approximately

    the same factor. To discuss a possible relation between the

    measured TS intensity and the density distribution, it is in-

    structive to compare the TS intensity profile with the

    expected mass density distribution, which can be calculated

    using the measured velocity profile. For a shock formed by a

    steady (time-independent) flow, the mass density distribution

    would be inversely proportional to the velocity distribution

    due to the mass flow continuity at the shock. In this experi-

    ment the flow density is not constant in time, but is propor-

    tional to the ablation rate of the wires.

    Our previous experiments with both the standard and

    inverse wire arrays11,12,16 show that the density distributions

    are well approximated by the rocket ablation model,12 and

    we will use it to find the expected mass density profiles at

    the sub-shock.

    The mass density q1(x,t0) in the undisturbed (upstream)flow at the time of observation t0 is proportional to

    q1ðx; t0Þ /dm

    dtt0 �

    x

    V1

    � �/ I2 t0 �

    x

    V1

    � �; (1)

    where x is the distance from the ablating wire and V1 is the

    flow velocity in the upstream flow, and x/V1 accounts for the

    time-of-flight delay. After the sub-shock, the density changes

    first due to the reduction of the flow velocity (which

    increases by a factor of V1/V2), and second due to the change

    in the time-of-flight delay in the post-shock flow, and is

    given by

    q2ðx; t0Þ /V1V2� dm

    dtt0 �

    xSV1� ðx� xSÞ

    V2

    � �; (2)

    where xs is the position of the sub-shock and V2 is the flow

    velocity in the downstream region, which is assumed con-

    stant following the measurements shown in Fig. 9(a).

    The expected relative mass density distribution, deter-

    mined using Eqs. (1) and (2) with the measured V1 and V2(from TS) and the shape of the driving current I(t), is shown

    in Fig. 9(b), together with the relative profile of the TS inten-

    sity. It is seen that the two profiles show a remarkably good,

    though very unexpected, agreement and possible implica-

    tions of this are discussed next.

    The total intensity of the ion feature of the Thomson scat-

    tering signal (ITS) is given by the well-known19 expression

    ITS / neZa4

    1þ a2ð Þ 1þ a2 þ a2 ZTeTi� � : (3)

    For the plasma parameters relevant to the measurements

    shown in Figs. 8 and 9 the scattering parameter a is greaterthan unity (a� 3), so that a2 1. In addition, the shape ofthe TS signal, i.e., the presence of the well-defined

    FIG. 9. Spatial profile of the flow velocity (top), determined from the analy-

    sis of TS data shown in Fig. 8. Plots at the bottom show: the electron density

    profile measured by end-on interferometry (green line), the mass density

    profile calculated from Eqs. (1) and (2) (blue dashes), and the profile of total

    intensity of the TS signal (red dots). The sub-shock is at xs� 2.7 mm fromthe obstacle positioned at x¼ 0 (7 mm from the wires in this experiment).

    056305-8 Lebedev et al. Phys. Plasmas 21, 056305 (2014)

    This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

    155.198.208.245 On: Thu, 01 May 2014 12:29:46

  • ion-acoustic peaks seen in both the upstream and down-

    stream regions, means that the ratio ZTe/Ti 1. This allowsEq. (3) to be simplified by retaining only a2 in the firstbracket and only the last term in the second bracket, so that

    the TS intensity becomes approximately equal to

    ITS � neTiTe� qTi

    Z

    Te; (4)

    where q is the mass density and Z is the average charge ofthe ions.

    The observed good agreement between the measured TS

    intensity profile and the calculated mass density profile thus

    could be interpreted as that the ratio TiZ/Te does not change

    between the upstream and the downstream flow regions.

    Moreover, for the Al plasma at the conditions of our experi-

    ment (ne� 1018 cm�3, Te in the range 10�50 eV), the ratioZ/Te only weakly depends on the electron temperature and

    varies by less than �20%, according to calculations withPRISM-SPECT software22 for the non-LTE ionization equi-

    librium. This in turn suggests that the ion temperature does

    not change between the upstream and the downstream

    regions. This is contrary to the expected strong heating of

    the ions in a “standard” strong shock transition. For the con-

    ditions of our experiment, with the Al ions moving at a ve-

    locity of �107 cm/s, an ion temperature of �100 eV could beexpected immediately after a strong shock. Thomson scatter-

    ing spectra shown in Fig. 8(b), however, do not show high

    ion temperatures in the post sub-shock region. Indeed, a fit-

    ting of the theoretical TS form-factors to the measured spec-

    tra show that ZTe increases at the sub-shock from

    ZTe� 60 eV in the upstream region to ZTe� 110 eV imme-diately after the sub-shock. At the same time these fittings

    suggest that the ion temperature in the downstream region

    remains sufficiently small (best fit at Ti� 25 eV). Anincrease of the Ti to 50 eV or 75 eV leads to a significant and

    clearly observable disagreement between the measured and

    theoretical TS spectra, as can be seen in Fig. 8, and we esti-

    mate from the fittings that the Ti in the downstream region

    does not exceed �40 eV.The observed absence of the measurable heating of the

    ions in the sub-shock suggests that even at the late stages of

    the evolution the main effect on the ions is produced by the

    cross-shock electric potential formed by the pile-up of mag-

    netic field, which decelerates the ions with very little, if any,

    increase of the ion temperature. To produce the observed

    jump in the flow velocity of a factor of 3.5, a larger electro-

    static potential of Du� 0.9Ei/(eZ) is needed for these laterstages of the sub-shock evolution. To explain the observed

    unexpectedly low ion heating, measurements of the variation

    of plasma parameters across the sub-shock front with better

    spatial resolution will be attempted in future experiments.

    IV. DISCUSSION

    The presented experimental data show the formation of

    a sub-structure developing in a reverse shock produced by

    the collision of a magnetized HED plasma with an obstacle,

    due to the pile-up of the magnetic field advected by the flow.

    This magnetic field initially acts on the electrons of the flow,

    leading to the formation of a cross-shock electric potential

    affecting the ions. An interesting and important issue in this

    picture is the balance of the ram pressure of the incoming

    flow at the sub-shock. The observed steady motion of the

    sub-shock in the direction away from the obstacle means that

    the post sub-shock pressure exceeds the ram pressure (qV2)of the flow. The magnetic pressure should play a key role in

    the balance since the thermal pressure, calculated using the

    measured post-shock plasma parameters (ZTe, ne), is an

    order of magnitude too small to achieve this balance alone.

    A contribution to the thermal pressure from a non-thermal

    population of ions appears to be unlikely being inconsistent

    with the measurements of TS spectra in the downstream

    plasma. The average post-shock magnetic field Bav, arising

    from the accumulation of the magnetic flux brought by the

    plasma flow, can be estimated using the magnetic field B(t)

    measured by probes, such that

    Bav ¼Vf l

    ðBðtÞdt

    xs: (5)

    Using the magnetic field measurements presented in

    Fig. 3, Eq. (5) yields Bav� 3 T at t¼ 225 ns. An estimate ofthe ram pressure from the plasma parameters measured at

    the same time (Fig. 5) suggests that in order to balance the

    jump in the ram pressure, a magnetic field larger than Bav by

    a factor of �2.5 is needed.The origin for this discrepancy is not understood at pres-

    ent and requires further investigation; in particular additional

    measurements of the magnetic field in the flow by a different

    diagnostic are required. The first attempts to measure the

    magnetic field distribution using a newly fielded Faraday

    rotation diagnostic23,24 indeed show the presence of a 7–8 T

    magnetic field in the region between the sub-shock and the

    obstacle, in the immediate vicinity to the stagnated plasma at

    the obstacle surface. The field in all other regions after the

    sub-shock is however smaller at �4 T, which is reasonablyclose to the Bav field estimated above, while the magnetic

    field in the upstream plasma is also in close agreement with

    that measured by the magnetic probes. More detailed meas-

    urements of the magnetic field distribution and of other

    plasma parameters will be attempted in future experiments

    to address this issue. In particular it will be important to

    investigate the possibility of the presence of other compo-

    nents of the magnetic field in the post sub-shock region, as

    the present diagnostics have only been sensitive to the

    B-field directed parallel to the obstacle in r-h plane (whichshould be the only component present in the laminar flow ge-

    ometry of this experiment). An additional understanding

    could be gained by attempting to reproduce the formation of

    the sub-shock observed in these experiments using numerical

    simulations. This would require at least a two-fluid MHD

    model, treating the electrons and ions separately, or possibly

    full Particle In Cell (PIC) simulations of the problem, per-

    formed in at least a 2-D setup to model the possible current

    loop (Fig. 6) in the plasma responsible for the pile-up of the

    advected magnetic field.

    056305-9 Lebedev et al. Phys. Plasmas 21, 056305 (2014)

    This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

    155.198.208.245 On: Thu, 01 May 2014 12:29:46

  • V. CONCLUSIONS

    The HEDP z-pinch driven experiment described in this

    paper presents a novel and interesting platform for creating

    supersonic plasma flow with an embedded and advected

    magnetic field. The generated flow is both supersonic and

    super-Alfv�enic, with a fast magneto-sonic Mach number of�5, and has Reynolds and magnetic Reynolds numbersmuch greater than unity. The present paper describes the use

    of this platform for the study of a reverse shock formed at

    the collision of the flow with an obstacle in the geometry of

    a perpendicular magnetised shock. We have observed that

    the flow-obstacle interaction leads to the formation of a sub-

    shock positioned at a distance of �c/xpi from the obstacle,presumably at the edge of the magnetic precursor. This

    sub-shock arises due to the accumulation of the magnetic

    flux advected by the flow, which is believed to act directly

    on electrons of the incoming flow, resulting in the develop-

    ment of a cross-shock potential decelerating the ions.

    Measurements with Thomson scattering and other diagnos-

    tics show that the ions decelerate at the sub-shock without

    being heated, which is not understood at present.

    The experimental platform described in this paper can

    be modified to investigate the interaction of two counter-

    streaming magnetized supersonic flows. In particular it is

    possible to create the colliding flows with oppositely directed

    magnetic fields and use this configuration to study the mag-

    netic reconnection in high beta, radiatively cooled plasma

    flows.

    ACKNOWLEDGMENTS

    This work was supported in part by EPSRC Grant No.

    EP/G001324/1 and by DOE Cooperative Agreement Nos.

    DE-F03-02NA00057 and DE-SC-0001063.

    1M. Camenzind, Rev. Mod. Astron. 3, 234 (1990).2P. Hartigan, J. M. Foster, B. H. Wilde, R. F. Coker, P. A. Rosen, J. F.

    Hansen, B. E. Blue, R. J. R. Williams, R. Carver, and A. Frank, ApJ 705,1073 (2009).

    3F. D. Drell, H. M. Foley, and A. M. Ruderman, J. Geophys. Res. 70, 3131,doi:10.1029/JZ070i013p03131 (1965).

    4P. Thompson, D. J. Southwood, and S. Goodman, J. Geophys. Res. 101,19849, doi:10.1029/96JA01283 (1996).

    5A. V. Gurevich, A. L. Krylov, and E. N. Fedorov, Sov. Phys. JETP 48,1074 (1978).

    6J. D. Lindl, Inertial Confinement Fusion (Springer, New York, 1999).7C. K. Li, F. H. S�eguin, J. A. Frenje, M. J. Rosenberg, H. G. Rinderknecht,A. B. Zylstra, R. D. Petrasso, P. A. Amendt, O. L. Landen, A. J.

    Mackinnon, R. P. J. Town, S. C. Wilks, R. Betti, D. D. Meyerhofer, J. M.

    Soures, J. Hund, J. D. Kilkenny, and A. Nikroo, Phys. Rev. Lett. 108,025001 (2012).

    8M. G. Haines, Plasma Phys. Controlled Fusion 53, 093001 (2011).9B. A. Remington, R. P. Drake, and D. D. Ryutov, Rev. Mod. Phys. 78, 755(2006).

    10I. H. Mitchell, J. M. Bayley, J. P. Chittenden, J. F. Worley, A. E. Dangor,

    M. G. Haines, and P. Choi, Rev. Sci. Instrum. 67, 1533 (1996).11A. Harvey-Thompson, S. V. Lebedev, S. N. Bland, J. P. Chittenden, G. N.

    Hall, A. Marocchino, F. Suzuki-Vidal, S. C. Bott, J. B. A. Palmer, and C.

    Ning, Phys. Plasmas 16, 022701 (2009).12S. V. Lebedev, F. N. Beg, S. N. Bland, J. P. Chittenden, A. E. Dangor, M.

    G. Haines, K. H. Kwek, S. A. Pikuz, and T. A. Shelkovenko, Phys.

    Plasmas 8, 3734 (2001).13V. V. Aleksandrov, G. S. Volkov, E. V. Grabovski, V. I. Zaitsev, G. G.

    Zukakishvili, S. F. Medovschikov, K. N. Mitrofanov, S. L. Nedoseev, G.

    M. Oleinik, I. Yu. Porofeev, A. A. Samokhin, P. V. Sasorov, V. P.

    Smirnov, I. N. Frolov, and M. V. Fedulov, AIP Conf. Proc. 808, 3 (2006).14J. Greenly, M. Martin, I. Blesener, D. Chalenski, P. Knapp, and R.

    McBride, AIP Conf. Proc. 1088, 53 (2009).15A. Harvey-Thompson, S. V. Lebedev, S. Patankar, S. N. Bland, G.

    Burdiak, J. P. Chittenden, A. Colaitis, P. de Grouchy, H. W. Doyle, G. N.

    Hall, E. Khoory, M. Hohenberger, L. Pickworth, F. Suzuki-Vidal, R. A.

    Smith, J. Skidmore, L. Suttle, and G. F. Swadling, Phys. Rev. Lett. 108,145002 (2012).

    16G. N. Hall, S. V. Lebedev, F. Suzuki-Vidal, G. Swadling, J. P. Chittenden,

    S. N. Bland, A. Harvey-Thompson, P. F. Knapp, I. C. Blesener, R. D.

    McBride, D. A. Chalenski, K. S. Blesener, J. B. Greenly, S. A. Pikuz, T.

    A. Shelkovenko, D. A. Hammer, and B. R. Kusse, Phys. Plasmas 20,022703 (2013).

    17S. N. Bland, S. C. Bott, G. N. Hall, S. V. Lebedev, F. Suzuki, D. J.

    Ampleford, J. B. A. Palmer, S. A. Pikuz, and T. A. Shelkovenko, Rev. Sci.

    Instrum. 77, 10F326 (2006).18G. F. Swadling, S. V. Lebedev, N. Niasse, J. P. Chittenden, G. N. Hall, F.

    Suzuki-Vidal, G. Burdiak, A. J. Harvey-Thompson, S. N. Bland, P. De

    Grouchy, E. Khoory, L. Pickworth, J. Skidmore, and L. Suttle, Phys.

    Plasmas 20, 022705 (2013).19J. Sheffield, D. Froula, S. H. Glenzer, and N. C. Luhmann, Plasma

    Scattering of Electromagnetic Radiation (Academic Press, New York,2010).

    20A. Harvey-Thompson, S. V. Lebedev, S. Patankar, S. N. Bland, G.

    Burdiak, J. P. Chittenden, A. Colaitis, P. de Grouchy, G. N. Hall, E.

    Khoory, M. Hohenberger, L. Pickworth, F. Suzuki-Vidal, R. A. Smith, J.

    Skidmore, L. Suttle, and G. F. Swadling, Phys. Plasmas 19, 056303(2012).

    21V. V. Krasnoselskikh, B. Lembège, P. Savoini, and V. V. Lobzin, Phys.

    Plasmas 9, 1192 (2002).22J. J. MacFarlane, I. E. Golovkin, P. R. Woodruff, D. R. Welch, B. V.

    Oliver, T. A. Mehlhorn, and R. B. Campbell, “Simulation of the ionization

    dynamics of aluminium irradiated by intense short-pulse lasers,” in

    Proceedings of the Third International Conference on Inertial FusionSciences and Applications 2003 (2003).

    23S. N. Bland, D. J. Ampleford, S. C. Bott, A. Guite, G. N. Hall, S. M.

    Hardy, S. V. Lebedev, P. Shardlow, A. Harvey-Thompson, F. Suzuki, and

    K. H. Kwek, Rev. Sci. Instrum. 77, 10E315 (2006).24G. F. Swadling, S. V. Lebedev, G. N. Hall, P. De Grouchy, F. Suzuki-

    Vidal, G. Burdiak, A. J. Harvey-Thompson, S. N. Bland, L. Suttle, and

    R. A. Smith, S. Patankar, “Diagnosing collisions of magnetized, high

    energy density plasma flows using a combination of collective Thomson

    scattering, Faraday rotation and interferometry,” Rev. Sci. Instrum.

    (submitted).

    056305-10 Lebedev et al. Phys. Plasmas 21, 056305 (2014)

    This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

    155.198.208.245 On: Thu, 01 May 2014 12:29:46

    http://dx.doi.org/10.1007/978-3-642-76238-3_17http://dx.doi.org/10.1088/0004-637X/705/1/1073http://dx.doi.org/10.1029/JZ070i013p03131http://dx.doi.org/10.1029/96JA01283http://dx.doi.org/10.1103/PhysRevLett.108.025001http://dx.doi.org/10.1088/0741-3335/53/9/093001http://dx.doi.org/10.1103/RevModPhys.78.755http://dx.doi.org/10.1063/1.1146884http://dx.doi.org/10.1063/1.3077305http://dx.doi.org/10.1063/1.1385373http://dx.doi.org/10.1063/1.1385373http://dx.doi.org/10.1063/1.2159308http://dx.doi.org/10.1063/1.3079752http://dx.doi.org/10.1103/PhysRevLett.108.145002http://dx.doi.org/10.1063/1.4789851http://dx.doi.org/10.1063/1.2336470http://dx.doi.org/10.1063/1.2336470http://dx.doi.org/10.1063/1.4790520http://dx.doi.org/10.1063/1.4790520http://dx.doi.org/10.1063/1.3694671http://dx.doi.org/10.1063/1.1457465http://dx.doi.org/10.1063/1.1457465http://dx.doi.org/10.1063/1.2227443