mcwilliams.sm_revised.pdf

Upload: umberto-genovese

Post on 04-Apr-2018

212 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/30/2019 McWilliams.SM_Revised.pdf

    1/31

    Originally published 22 November 2012; corrected 7 December 2012

    www.sciencemag.org/cgi/content/full/science.1229450/DC1

    Supplementary Materials for

    Phase Transformations and Metallization of Magnesium Oxide atHigh Pressure and Temperature

    R. Stewart McWilliams,* Dylan K. Spaulding, Jon H. Eggert, Peter M. Celliers,Damien G. Hicks, Raymond F. Smith, Gilbert W. Collins, Raymond Jeanloz

    *To whom correspondence should be addressed. E-mail: [email protected]

    Published 22 November 2012 on Science ExpressDOI: 10.1126/science.1229450

    This PDF file includes:

    Materials and MethodsFigs. S1 to S6Tables S1 to S3

    References

    Correction: References were corrected.

    mailto:[email protected]
  • 7/30/2019 McWilliams.SM_Revised.pdf

    2/31

    1

    Materials and Methods

    Materials and Targets

    A laser target is shown in Figure 1A. Multi-layer targets for temperature and

    reflectivity measurements consisted of a 50 m thick Al buffer with a 300 m thick

    single crystal of MgO attached, stacked to produce a uniform-thickness sandwich ~ 1 mm

    in diameter. The synthetic MgO, supplied by Marketech International, was oriented for

    shock propagation in the crystallographic direction. The initial density and index

    of refraction (at 532 nm) were taken to be 3.584 g/cm 3 and 1.743, respectively ( 38 ). A

    broadband anti-reflection coating centered at 532 nm was deposited on the free-surface.

    Samples were kept in dry or vacuum conditions as much as possible to prevent water

    adsorption. The Al buffer was diamond-turned to mirror finish, and coated on the

    ablation surface with ~10 m of CH polymer.

    The P -V measurements were referenced to a quartz standard ( 35, 39 ), so used a 30

    m thick c-cut quartz (

    -SiO 2) crystal inserted between the MgO and the Al buffer ( 34 ),

    taken to have an initial density and index of refraction (532 nm) of 2.65 g/cm 3 and 1.546,

    respectively ( 40 ).

    Parts were bonded using Norland-63 optical adhesive with a glue thickness of less

    than ~5 m for Al-sample bonds and less than ~1 m for SiO 2-MgO bonds. Preheating

    effects are negligible in these types of targets for the drive laser parameters used ( 41 ).

    Drive Lasers

    Experiments were conducted using the Omega (Laboratory for Laser Energetics,

    University of Rochester) and Jupiter (Lawrence Livermore National Laboratory) laser

    facilities ( 12, 13, 42 ). At the Omega Laser Facility, the laser drive consisted of six beams

  • 7/30/2019 McWilliams.SM_Revised.pdf

    3/31

    2

    of the Omega laser, smoothed by SG8 phase plates to provide an 800-m diameter spot

    of uniform irradiation on target to ensure planar loading. Laser energies of 850 J in a 1

    ns pulse were used. At the Jupiter Laser Facility, the drive laser consisted of two beams

    of the Janus laser; CPP phase plates provided uniform irradiation of a 600-m diameter

    spot; laser energies up to 500 J in a 2 ns pulse were used.

    One-dimensional loading is evidenced in our data by lateral uniformity in shock

    features over ~ 500 m on the target (Figs 1B, 1C).

    VISAR

    A line-imaging laser interferometer, or line VISAR (Velocity InterferometerSystem for Any Reflector), measured the velocity and reflectivity of shock fronts in MgO

    and quartz as functions of time and position in the target ( 42 ). A representative dataset is

    shown in Fig. 1B. The shock velocity is

    [ ] ++=0)1(2

    )()()(n

    t bt t U S

    (Eq. S1)

    where t is the time, (t ) is the observed fringe phase (a fraction of the fringe period, such

    that a full period is unity), and b(t ) is the integer number of base fringe shifts. The right

    term is referred to as the VPF (velocity-per-fringe) in which is the wavelength of the

    VISAR probe laser (532 nm), is the optical delay time introduced by the etalon placed

    in one leg of the interferometer, (1+ ) is a correction for optical dispersion in the etalon,

    and n0 is the ambient index of refraction in the material where the reflecting shock is

    propagating. For Fig. 1B, a VPF of 1.5673 km/s is used. Two independent VISAR

    systems having different VPF were used in each shot to constrain b(t ) unambiguously.

    Background fringes in the VISAR from immobile reflecting surfaces in the optical path

  • 7/30/2019 McWilliams.SM_Revised.pdf

    4/31

    3

    (ghost fringes) were subtracted from our signal fringes with Fourier analysis ( 43 ), e.g.

    between e 1 and e 6 in Fig. 1B.

    Shock reflectivity, the coefficient of reflection between the shocked (opaque,

    reflecting) and unshocked (transparent, ambient) states, was determined as

    00

    )()(

    I t I

    Rt R = (Eq. S2)

    where I (t ) is the fringe intensity, corrected for probe beam power variation with time, and

    I 0 is the initial fringe intensity due to reflection from a reference surface of known

    reflectivity R0 (in this case, diamond-turned aluminum against an adhesive layer, with R0

    ~ 88%, which is observed early in the VISAR records). Fringe intensity was determined

    from the fringe amplitude, measured via Fourier transform by filtering intensity

    contributions from background (long wavelength) signals. At the Omega facility, the

    probe power variation was measured in a reference image of the target obtained before

    shocking it. While the intensity vs. time profile of the probe was constant, its timing

    could shift slightly between images, so reference intensity was time-shifted when

    normalizing the data; the accuracy of this shift was tested, for example, by ensuring

    fringe intensity prior to shock breakout (i.e. before event e 1 in Fig. 1) was constant after

    normalization. At the Jupiter facility, a portion of the probe beam was diverted before

    reaching the sample and focused at the edge of the streak camera image to provide a

    time-resolved intensity reference simultaneously with the experiment.

    Hugoniot Compressibility Measurements

    Our P -V - E -U S -U P Hugoniot measurements for MgO at P ~ 0.6 TPa (Figs 2D, S1,

    S6, and Table S1) were accomplished in the manner of Ref. ( 34 ) using the quartz

    Hugoniot ( 35, 39 ) and a thermodynamic description of the quartz reshock Hugoniot ( 34 ).

  • 7/30/2019 McWilliams.SM_Revised.pdf

    5/31

    4

    This technique consists of driving a reflecting shock through a standard material (quartz)

    into the sample material (MgO). Measurement of shock velocity in each material at the

    moment the shock transits the standard-sample interface allows a determination of the

    shock Hugoniot by impedance matching.

    Temperature Measurements

    A Streaked Optical Pyrometer (SOP) (44, 45 ) was used to measure thermal

    emission from the shock front as a function of time and space, integrated over a 350 to

    900 nm spectral window. A representative SOP dataset is shown in Figure 1C. In our

    analysis, we apply a calibration based on the known shock temperatures in quartz ( 12 ).To do this, experiments on quartz alone were conducted in parallel to the MgO

    experiments to correlate emission intensity with a standard temperature ( 14 ). In these

    calibration experiments, as in the MgO experiments, U S (t ) and the spectral radiance L(t )

    are simultaneously measured. Assuming grey-body behavior, L(t ) is given by Plancks

    Law as

    [ ] 1)( / exp12

    )()( 52

    =

    t kT hchc

    t t L

    (Eq. S3)

    where the emissivity (t ) is given by Kirchhoffs Law assuming an optically thick shock

    front

    )(1)( t Rt = (Eq. S4)

    Comparison of quartz measurements with the standard ( 12 ) gives a relationship between

    the observed SOP intensity I SOP (t ) and the temperature, defined as ( 44, 45 )

    +=

    )()(

    1ln)(

    0

    t I At

    t T T

    SOP

    (Eq. S5)

  • 7/30/2019 McWilliams.SM_Revised.pdf

    6/31

    5

    where T 0 and A are fit parameters (Fig. S5) which are independent of material and

    constant for all experiments provided the optical configuration of the pyrometer is

    unchanged.

    Calibration to a standard material ( 14 ), rather than a standard lamp of known

    spectral radiance (44, 45 ), provides several advantages, including the ability to fit Eq. S5

    to a range of temperatures (rather than to a single point) and to temperatures which are

    characteristic of the experiment (as opposed to the low temperature of a lamp), as shown

    in Fig. S5. Straightforward corrections to temperature are possible with future

    improvements to the standard.Equation-of-State Modeling

    A linear U S -U P model of compressibility on the Hugoniot provided a good

    representation of the P -V - E -U S -U P data (Figs 2D and S1). This was defined by a

    weighted linear fit, obtained by orthogonal distance regression, to appropriate MgO

    Hugoniot data, giving U S = (6.743 0.070) + (1.316 0.017) U P . The y-intercept of this

    fit agrees well with the bulk sound speed of MgO, 6.73 0.01 ( 32 ), showing the

    reliability of the measurements and the fit. Data below U S = 9.3 km/s were excluded

    because the shock exhibits two-wave structure in this regime ( 46 ) which can lead to

    nonlinear U S -U P behavior. Also excluded was a datum obtained using the free-surface

    release assumption U f = 2 U P which clearly deviates from other Hugoniot data ( 47 ). It is

    noted that the data of Marsh ( 30 ) used in this fit appears to be identical to the earlier data

    reported by van Thiel ( 48 ) but shifted, possibly from changes in a standard Hugoniot; it is

    presumed the later dataset was accurate, and the earlier dataset was omitted. For our high

    pressure data, the quartz Hugoniot of Ref. ( 35 ) was assumed (Table S1).

  • 7/30/2019 McWilliams.SM_Revised.pdf

    7/31

    6

    Our multi-phase equation-of-state (EOS) for MgO is based on finite-strain and

    Mie-Grneisen-Debye equations-of-state ( 16, 18 ). Basic thermodynamic differentials

    (49 )

    PdV TdS dE = (Eq. S6)

    dT C dV V C T

    PdE V V +

    += (Eq. S7)

    dT V

    C dV

    V K

    dP V T += (Eq. S8)

    dT

    T

    C dV

    V

    C dS V V += (Eq. S9)

    are used, where S is entropy, the Grneisen parameter, and K T the isothermal

    compressibility. For shock states in a given pure phase and at thermodynamic

    equilibrium, Hugoniot pressure P H can be defined as ( 16, 18 )

    2 / ))( / (1))()( / ()(

    )(0 H H

    tr H S H H S H H V V V

    E V E V V PV P

    +=

    (Eq. S10)

    where the subscript H refers to conditions on the Hugoniot of the given phase, centered at

    the initial P , T and V (P 0, T 0, V 0, respectively, where V 0 is the ambient volume of the

    initial phase), the subscript S refers to conditions on the isentrope of the given phase

    (centered at P 0, T 0, and V 0 HP , the ambient volume of the given phase), corresponds to

    the given phase and is assumed to depend only on volume, and the energy of

    transformation E tr is a difference in internal energy between the given phase and the

    initial phase ( E tr 0 if the given phase is the initial phase). The Rankine-Hugoniot

    relations ( 15, 49 ) give

    [ ] 2 / 1000 ) /()( H H S V V PPV U = (Eq. S11)

  • 7/30/2019 McWilliams.SM_Revised.pdf

    8/31

    7

    [ ] 2 / 100 ))(( H H P V V PPU = (Eq. S12)

    2 / ))(( 00 H H H V V PP E += (Eq. S13)

    The isentrope is defined by a third order Birch-Murnaghan equation ( 18 )

    ++= f K f f K f P S S S )4(23

    1)21(3)( / 02 / 5

    0 (Eq. S14)

    where f = (1/2)[( V 0HP / V )2/3-1], and K 0S and K 0S

    / are the ambient bulk modulus and its first

    pressure derivative, respectively, for isentropic conditions. From Eqs. S6-S9

    = dV PV E S S )( (Eq. S15)

    = dV V T

    V T S

    )( (Eq. S16)

    and

    +

    +=)(

    )(

    )()( H H

    tr H S

    V E

    E V E V H S H H C

    dE V T V T (Eq. S17)

    The Debye model is used to determine S and C V (16 )

    )]1ln()()3 / 4[(6 3 x

    gas e x D RS = (Eq. S18)

    )]1 /(3)(4[6 3 =x

    gasV e x x D RC (Eq. S19)

    where

    = x

    zedz z

    x x D

    0

    3

    33 13

    )( (Eq. S20)

    and x = / T , where is the Debye temperature. This model gives C V 6 Rgas in the high-

    temperature regime observed in our experiments, consistent with our specific heat results

    (Fig. 2).

  • 7/30/2019 McWilliams.SM_Revised.pdf

    9/31

    8

    It is assumed that / V = constant, and correspondingly that =V d d ln / ln

    (16 ) in a given phase. We assume K T at given volume is temperature independent and

    estimate it as ( 49 )

    +

    =

    H

    H V H

    H H

    H T V

    T C PP

    V P

    V V V

    K

    22

    )( 00 (Eq. S21)

    (where C V corresponds to conditions on the Hugoniot) or, where the Hugoniot is

    undefined (possible for high-pressure phases at large volume), as ( 50 )

    1

    2

    2

    1

    +

    +=T K V C

    T K V K K

    T V

    S S T

    (Eq. S22)

    (where K S , T , and C V correspond to the isentrope, and the thermal expansivity

    V K C T V / = ). Consistent with the assumption that K T (V ) is temperature independent,

    these two estimates are virtually identical for MgO at V where both can be calculated.

    To summarize, within a given phase , and K T are functions of V alone; C V

    depends on V and T . Off-Hugoniot states in a given phase are determined using Eqs. S7-

    S9. A phase is thus defined by 6 parameters: V 0HP (ambient volume, equal to V 0 for initial

    phase), K 0S and K 0S / (ambient elastic constants), 0 and 0, (ambient thermodynamic

    constants), and E tr (energy difference between phases).

    Phase boundaries are located in P -T space where the Gibbs free energies of the

    two relevant phases equalize; comparing the equations-of-state along the boundary yield

    the entropies and volumes of transformation ( S tr and V tr , respectively). To model the

    phase transformation regimes, the Hugoniot was assumed to fall on the phase boundary in

    P -T space (i.e. the transformation was modeled as if it occurred at equilibrium ( 15 )), and

    a linear U S -U P EOS was assumed for these regions. Model parameters (18 total) were

  • 7/30/2019 McWilliams.SM_Revised.pdf

    10/31

    9

    adjusted to match the data, with emphasis on fitting regions of the Hugoniot

    corresponding to pure phases, and on locating phase boundaries at the observed

    conditions of transformation. Phase transition properties (Table I) were established by

    examining a variety of models that fit the data. Two example models are shown in Fig.

    S6 with associated parameters given in Table S3. The ranges given for phase transition

    properties (Table I) include the possibility that our T -U S data in transformation regimes is

    representative of thermodynamic equilibrium (e.g. best-fit model) and the possibility

    that it does not represent thermodynamic equilibrium, either due to nonequilibrium

    transformation effects such as superheating or due to time-resolution limitations (e.g.relaxed model).

    Since transformation parameters and high-pressure phase equations-of-state are

    determined relative to the B1 phase, it was important that the B1 phase model was as

    accurate as possible (Table S3). We found it to be satisfactorily described by parameters

    appropriate for B1 MgO determined near ambient conditions ( 19 ): density 0 = 3.584

    g/cm 3, bulk modulus and its pressure derivative K 0S = 163.2 ( 0.7) GPa and K 0S = 4.13

    ( 0.02), and Grneisen parameter 0 = 1.52 (0.02). We take 0 for the B1 phase to be

    760 K. For our complete multi-phase models, we used somewhat different K 0S and K 0S

    (K 0S = 162, K 0S = 4.3) to enhance model agreement with the highest-pressure gas-gun

    data ( 31, 33 ) (Figs S1, S6), presumed to also be in the B1-phase.

    Details on Compilation of Table I

    The results in Table I were compiled as follows. First, we characterized the

    observed phase transformations using the equation-of-state modeling described above.

    This established experimental ranges for the five quantities in Table I describing the

  • 7/30/2019 McWilliams.SM_Revised.pdf

    11/31

    10

    phase transformations ( T , P , ( ) BT P / , S tr , and V tr ). The ranges of T and P

    encompass the beginning and end of the transformations on the Hugoniot, and include

    some model-dependant uncertainty; the ranges for the three other quantities represent

    model-dependant uncertainties assessed by examining different parameters for the high

    pressure phases. There are tradeoffs in selecting parameters for the high-pressure phases

    that fit the measurements; for example, a large volume decrease at the first transition

    must be accompanied by a relatively large volume increase at the second, to ensure good

    agreement with compressibility data at higher pressure. The quantities S tr , T , and P are

    found to be insensitive to the model parameters selected: that is, they are well constrained

    by our data. The quantities V tr and ( ) BT P / are poorly constrained, as indicated by

    the larger fractional uncertainty of these quantities in Table I; this is due in part to the

    lack of compressibility data in the first high pressure phase, such that the volume of this

    phase is poorly constrained.

    Second, we examined the properties of theoretical phase transformations that

    most closely matched our observed phase transformations, for comparison to our

    measurements. The P , T and ( ) BT P / of the theoretical transformations were taken

    from the regions of the theoretical phase boundaries that were closest to the observed

    transitions. Since predicted V tr and S tr on calculated phase boundaries were not

    available in the literature, we estimated these quantities using various assumptions. For

    the first transition, two limiting assumptions were examined. In one, V tr / V was

    presumed constant on the phase boundary (and defined by the zero-temperature value

    predicted in first-principles calculations), which gives an approximate upper bound for

    V tr and S tr at high temperature. In the other, S tr was presumed constant on the phase

  • 7/30/2019 McWilliams.SM_Revised.pdf

    12/31

    11

    boundary and presumed to be infinitesimal (as it must be at low temperature), which sets

    a lower bound of S tr ~ V tr ~ 0 at high temperature (since the Clapeyron slope, also

    infinitesimal at low temperature, becomes finite at high temperature, requiring V tr

    become small). For the second transition, we assume that volume of transformation on

    melting is roughly constant at high pressure, regardless of local phase.

    Specific Heat Calculation

    The specific heat analysis used here has been described previously (Refs ( 12, 13 )

    and references therein). From Eq. S7

    ( )( ) T V V T PV E C

    H

    H V ) / ( /

    / ++= (Eq. S23)

    For the results in Fig. 2B, was taken from first-principles calculations for pure solid and

    liquid MgO at relevant conditions ( 1.1) (3). In using a corresponding to a pure

    phase in this calculation, the C V measured are accurate where the Hugoniot accesses a

    pure-phase state; this is likely the case where the measurements approach the Dulong-

    Petit limit. In regions of the Hugoniot corresponding to phase transformation, C V should

    deviate from the Dulong-Petit limit (as observed), however here C V is not accurate

    because in a mixed-phase regime can differ from that in a pure phase. Put another way,

    the apparent specific heat in these regimes is a combination of specific heat and latent

    heat.

    Decay PredictionBelow 0.5-0.6 TPa, reflection from the shock front could not be measured (R

    V V V V

    PPPP

    (Eq. S24)

    where subscript 1 indicates conditions at the peak Hugoniot state in the low-pressure

    phase, and subscript 2 indicates a state on the Hugoniot above this (i.e. within the

    subsequent phase transition regime). Geometrically, this criterion can be evaluated by

    examining the Hugoniot in P -V space (Fig. S6). Given the possible large increase in

    density at the first transition, we considered whether this could lead to formation of a

    two-wave structure. However, all EOS models tested are consistent with the occurrence

    of only a single shock. For example, in an extreme model (with large volume decrease at

    the first transition, Fig. S6), the left side of Eq. S24 is ~ 2; and it is larger for other

    models.

    Drude-Semiconductor Conductivity Model:

    Shock front reflectivities beyond a few percent in insulating materials suggest the

    presence of delocalized electrons and enhanced electronic conductivity ( 12-14, 43 ). The

  • 7/30/2019 McWilliams.SM_Revised.pdf

    15/31

    14

    model used here to invert R, T , and V in shock-compressed MgO for electronic

    conductivity is similar to that used in numerous prior studies (Refs ( 12, 43 ) and

    references therein). For wide band-gap insulators, the quantitative results of this model

    have been broadly supported by independent theory, e.g. for SiO 2 (12 ) and LiF ( 43 ).

    Reflectivity is analyzed in terms of the Fresnel equation for normal incidence, R =

    |(n-n0)/(n+n0)|2, where n = [ b+ 4 i ( )/ ]1/2 is the complex index of refraction, b is the

    contribution of the bound electrons, is the frequency, and ( ) is the conductivity,

    assumed to follow the Drude formulation ( ) = (2 nee2 / meff )(1- i )-1, where ne is the

    carrier density,

    is the carrier relaxation time, and meff is the carrier effective mass. It is

    assumed that scattering occurs at the Ioffe-Regel limit ( 51 ), such that ev / = where ve is

    the carrier velocity and 3 / 1)4 / 3(2 F A N N MV = is the interatomic distance, where M is

    molar mass, N A is Avogadros number, and N F is the number of atoms in a formula unit.

    Carrier activation is assumed to follow the behavior of an intrinsic semiconductor ( 52 ),

    i.e. )]2 /([)2 / (2 2 / 12 / 32

    kT E f kT mn geff e = , where dye y x f x ym

    m )1 /( / 2][ 0

    +=

    is the Fermi-Dirac integral and E g is the activation gap energy. Carrier velocity is given

    by )])2 /([ / )]2 /([)( / 2( 2 / 12 / 3 kT E f kT E f mkT v ggeff e = . This model has three

    independent variables based on observation ( V , R and T ) and three unknown parameters

    ( b, E g and meff ) which in this study are fit to the measurements.

    The unknown parameters were further parameterized for optimum fitting in the

    presence of distinct phase transitions. We assume a volume-dependent energy gap ( E g)

    between valence and conduction bands ( 12, 43 ), assumed to be sigmoidal in form as

  • 7/30/2019 McWilliams.SM_Revised.pdf

    16/31

    15

    +=

    T T g E E E exp0 (Eq. S25)

    where the density = 1/ V , and E 0, E T , T , and are fit parameters. This allows for rapid

    changes in the gap at phase transitions or for more continuous change across a broader

    pressure range. In the latter limit, Eq. S25 is similar to the linear function of density

    commonly used for materials that lack a sharp phase transition (e.g. refs ( 12, 43 ) and

    references therein). In one present fit, the parameter E 0 was fixed at 8 eV, the ambient-

    pressure gap of the B1 phase, and the predicted gap of the B2 phase ( 1); this choice was

    made considering that the reflectivity at low pressure can be attributed solely to bound

    electron effects (Figs S2 and S3) and therefore cannot meaningfully constrain the initial

    gap. A second fit with an unconstrained initial gap gave broadly similar conclusions

    regarding gap closure at the second transition (Fig. 2C and Fig. S4).

    The quantity b was fit either as a constant, or such that the quantity b (equal

    to the real index of refraction at low conductivity) was a linear function of density.

    Constrained by the low-pressure end of the reflectivity data, where the reflectivity is

    attributable to bound electron effects alone (Fig. S2), b gives an independent estimate for

    the real index of refraction of MgO under pressure (Fig. S3). This model-based

    measurement of the real index agrees reasonably well with extrapolation of shock data

    from lower pressures ( 17 ).

    Optical Depth of Shocked MgO

    A small optical depth in shocked MgO is required in the present experiments for

    accurate interpretation of the temperature data, such that emission originates from near

    shock-front, and hence Hugoniot states, rather than from well behind the shock, where

  • 7/30/2019 McWilliams.SM_Revised.pdf

    17/31

    16

    material undergoing rarefaction may lie off the Hugoniot or where target interfaces can

    contribute anomalously to the emission. That is, we require that shocked MgO have an

    optical thickness on the order of or less than the distance U S t where t is the time

    resolution of the diagnostics (~100 ps), or ~1 m.

    Several lines of evidence support a small optical depth in shocked MgO at the

    conditions of our experiments:

    1) The DC electronic conductivity is expected to be ~ 1000 S/m or larger in these

    experiments, based on the Drude-semiconductor model results (Fig. S4), and also on

    predicted carrier densities for warm solid MgO

    (1). Estimating the optical penetration

    depth d at a conducting surface (here, the shock front) as 8 / ~ cd (strong conductor)

    and k cd 2 / = (arbitrary material) ( 53 ), where k is the imaginary part of the index and

    can be estimated from our model, we estimate a maximum d of ~ 0.3 m and ~ 2 m,

    respectively, in our experiments.

    2) It is observed that above T ~ 8000 K and P ~ 0.3 TPa the data show good consistency

    with prior measurements ( 36 ) and predictions ( 3, 36 ) of shock temperature, and as

    discussed adhere closely to the Dulong-Petit limit. These observations would be highly

    unlikely if the temperature data were in error. Below T ~ 8000 K, however, the apparent

    temperature deviates from previously reported temperatures, suggestive of a large optical

    skin depth at low T and P . This is not unexpected given that nearly-critical d are inferred

    for the solid at conditions near the first transition. These observations evince the

    accuracy of our temperature measurements except at low pressures ( P < 0.3 TPa) where

    the data were discarded.

  • 7/30/2019 McWilliams.SM_Revised.pdf

    18/31

    17

    3) We note that shocked MgO was opaque to VISAR in all experiments accessing the

    present pressure range (including experiments with peak pressures at P < 0.3 TPa); this is

    consistent with our conclusion of strongly absorbing conditions in solid MgO at high

    temperature, though other sources of opacity (i.e. scattering) could contribute.

    Data Quality and Measurement Uncertainties

    Data collected at the Omega laser are considered to be of substantially higher

    quality than those collected at the Jupiter laser (as discussed below), and T -U S - R data

    from Omega have thus been used in our quantitative analyses here; lower-accuracy and

    lower-precision measurements of T -U S - R from the Janus laser have been included todemonstrate that primary observations the existence of a temperature minimum on the

    Hugoniot (Table S2), and the onset of shock reflectivity (Fig. S2) can be reproduced in

    very different experimental configurations.

    Error in the temperature (Fig. 2A error bars) is primarily systematic, such that

    derivatives of temperature, e.g. ( ) H V T / , and quantities calculated from derivatives,

    e.g. C V , are known more precisely. A comparison between Omega and Jupiter results for

    the first transition temperature are presented in Table SII. Random error in temperature

    is small, as is evident in the high reproducibility of measurements at a given facility (Figs

    2A and 2B, and Table SII). Temperatures at the Omega laser are considered more

    accurate due to improved diagnostic quality and improved accuracy in the relative timing

    of VISAR and SOP datasets. Error in reflectivity at Omega (Fig. 2C and Fig. S2 error

    bars) is also mostly systematic; the reproducibility of the measurements confirms that

    random errors are small except at very high pressure. At Jupiter (Fig. S2 error bars)

    reflectivity uncertainty is much larger and principally random in nature.

  • 7/30/2019 McWilliams.SM_Revised.pdf

    19/31

    18

    Compressibility data were collected at the Jupiter laser. Uncertainty in these

    results (Table S1) is considered random, consistent with the observed scatter in the

    datapoints; additional systematic sources of error due to the choice of reference EOS

    have been discussed (Figs S1 and S6), and are not included in Table S1.

    All uncertainties reported here are equivalent to one standard deviation.

  • 7/30/2019 McWilliams.SM_Revised.pdf

    20/31

    19

    Fig. S1:

    Linear U S -U P fit to available shock data, as discussed in Supplementary Materials and

    Methods, plotted with a theoretical U S -U P EOS ( 3) and two EOS models from this study

    (see Fig. S6). Present measurements are shown using the standard quartz Hugoniots from

    Ref. ( 39 ) (HEA) and Ref. ( 35 ) (KD); the latter was used in the linear fit.

  • 7/30/2019 McWilliams.SM_Revised.pdf

    21/31

    20

    Fig. S2:

    Shock reflectivity R vs. shock velocity U S . Grey and black curves are separate shots from

    Omega; blue curve is a compilation of reflectivities from four Jupiter shots. At both

    facilities, reflectivity became observable above U S =17.3 km/s, though the data from

    Omega are more precise. The green curve shows the predicted reflectivity based on

    extrapolation of the real index of refraction from shock data to P =0.023 TPa ( U S =7.8

    km/s) ( 17 ), which agrees well with the lowest reflectivities observed here. This suggests,

    independent of our Drude-semiconductor model analysis, that the reflectivity observed

    below the second transition is principally due to bound electron contributions, and that

    the rise beginning at ~18 km/s can be attributed to an increase in conductivity associated

    with the second transformation.

  • 7/30/2019 McWilliams.SM_Revised.pdf

    22/31

    21

    Fig. S3:

    Real component of the index of refraction of MgO at low conductivity. Green and blue

    solid lines are given by b in the present Drude-semiconductor model fits (for assumed

    constant b or b varying linearly in density, respectively) with dashed portions

    representing extrapolations beyond the reflectivity data. Solid black line is from Ref.

    (17 ), with dashed line indicating extrapolation to high density.

  • 7/30/2019 McWilliams.SM_Revised.pdf

    23/31

    22

    Fig. S4:

    DC conductivities ( =0) and carrier densities inferred from the Drude-semiconductor

    model, showing results for a constrained ( E 0=8 eV) and unconstrained ( E 0 fit) initial gap.

    The predicted carrier density on a 10,000 K isotherm in solid (B2) MgO (1) is consistent

    with our model results; this isotherm overlaps the Hugoniot at 0.49 TPa (here, predicted

    carrier densities are ~10 19 cm -3 while observed carrier densities are 10 19 to 10 20 cm -3

    depending on model assumptions).

  • 7/30/2019 McWilliams.SM_Revised.pdf

    24/31

    23

    Fig. S5:

    Pyrometer temperature-intensity relationship determined from measurements on a

    decaying shock in quartz (red). The data is fit (dashed black line) to obtain T 0 and A in

    Eq. S5. Note that the pyrometer is calibrated at temperatures corresponding to the phase

    transitions in MgO.

  • 7/30/2019 McWilliams.SM_Revised.pdf

    25/31

    24

    Fig. S6:

    Mie-Grneisen-Debye, finite-strain EOS models for MgO; two example models are

    presented with prior gas gun data ( 30-33, 36 ) and present measurements of T vs. U S (2

    shots, colored as in Fig. 2), and P vs. V (2 shots). The best fit model required that

    Hugoniot temperature data is precisely fit in mixed-phase regimes, whereas the relaxed

    model examined the possibility of small density changes at transformation and deviations

    from the data in mixed-phase regimes. Systematic uncertainty due to the reference EOS

    of quartz (HEA ( 39 ), KD ( 35 )) is of the same order of magnitude as random error, and

    was accounted for. Also shown is the Hugoniot model of Ref. ( 3) which overlaps our

    model exactly in the B1 phase, and is similar to the relaxed model for the first phase

    transformation.

  • 7/30/2019 McWilliams.SM_Revised.pdf

    26/31

    25

    Table S1.Hugoniot compressibility measurements from two experiments. Quartz Hugoniot of Ref.(35 ) used.

    Shot Laserfacility

    MgO U S (km/s)

    MgO U P (km/s)

    MgO P (TPa)

    MgO (g/cm 3)

    Quartz U S (km/s)

    MgO6 Jupiter 18.82 (0.49) 9.48 (0.42) 0.639 (0.030) 7.22 (0.40) 18.41 (0.46)MgO7 Jupiter 18.88 (0.49) 8.96 (0.40) 0.606 (0.028) 6.82 (0.34) 17.98 (0.45)

  • 7/30/2019 McWilliams.SM_Revised.pdf

    27/31

    26

    Table S2.Temperature of the first transition as measured at the Jupiter and Omega laser facilities.The temperature minimum at the high-pressure end of the coexistence regime ( U S 16.5

    km/s) was selected as a reference point.

    Shot Laserfacility

    Temperature minimum (K)(at P ~ 0.44 TPa)

    MgO16 Jupiter 9540 (+850,-620)MgO17 Jupiter 9522 (+850,-620)MgO19 Jupiter 9068 (+850,-620)MgO29 Jupiter 9930 (+850,-620)MgO31 Jupiter 9881 (+850,-620)57513 Omega 8610 (+790, -752)57514 Omega 8432 (+790, -752)

    Jupiter Average 9588 (+850,-620)Omega Average 8521 (+790, -752)

  • 7/30/2019 McWilliams.SM_Revised.pdf

    28/31

    27

    Table S3.

    Equation-of-state parameters for the two example models shown in Figs S1 and S6.

    Best-fit model Relaxed model

    B1 Phase

    V 0 (cc/g) 0.2790

    K 0S (GPa) 162

    K 0S / (unitless) 4.3

    0 (unitless) 1.5

    0 (K) 760

    E tr (J) 0

    First High Pressure Phase

    V 0HP / V 0-B1 0.862 0.927

    K 0S / K 0S -B1 1.3 1.2

    K 0S / / K 0S -B1 / 1.1 1.1

    0 / 0-B1 1 1

    0 / 0-B1 0.52 0.58

    E tr / (V 0-B1 K 0S -B1) 0.3 0.175

    Second High Pressure Phase

    V 0HP / V 0-B1 0.92 0.972

    K 0S / K 0S -B1 1.2 1.2

    K 0S / / K 0S -B1 / 1.18 1

    0 / 0-B1 1 1

    0 / 0-B1 0.415 0.44

    E tr / (V 0-B1 K 0S -B1) 0.25 0.225

  • 7/30/2019 McWilliams.SM_Revised.pdf

    29/31

    28

    References

    1. K. Umemoto, R. M. Wentzcovitch, P. B. Allen, Dissociation of MgSiO 3 in the cores of gas giantsand terrestrial exoplanets. Science 311 , 983 (2006). doi:10.1126/science.1120865 Medline

    2. A. B. Belonoshko, S. Arapan, R. Martonak, A. Rosengren, MgO phase diagram from firstprinciples in a wide pressure-temperature range. Phys. Rev. B 81 , 054110 (2010).doi:10.1103/PhysRevB.81.054110

    3. N. de Koker, L. Stixrude, Self-consistent thermodynamic description of silicate liquids, withapplication to shock melting of MgO periclase and MgSiO 3 perovskite. Geophys. J. Int. 178 ,162 (2009). doi:10.1111/j.1365-246X.2009.04142.x

    4. D. Alf, Melting curve of MgO from first-principles simulations. Phys. Rev. Lett. 94 , 235701(2005). doi:10.1103/PhysRevLett.94.235701 Medline

    5. A. Aguado, P. A. Madden, New insights into the melting behavior of MgO from moleculardynamics simulations: The importance of premelting effects. Phys. Rev. Lett. 94 , 068501(2005). doi:10.1103/PhysRevLett.94.068501 Medline

    6. A. Zerr, R. Boehler, Constraints on the melting temperature of the lower mantle from high-pressure experiments on MgO and magnesiowstite. Nature 371 , 506 (1994).doi:10.1038/371506a0

    7. L. Zhang, Y. W. Fei, Melting behavior of (Mg,Fe)O solid solutions at high pressure. Geophys. Res. Lett. 35 , L13302 (2008). doi:10.1029/2008GL034585

    8. B. Karki, D. Bhattarai, L. Stixrude, First-principles calculations of the structural, dynamical, andelectronic properties of liquid MgO. Phys. Rev. B 73 , 174208 (2006).doi:10.1103/PhysRevB.73.174208

    9. A. R. Oganov, M. J. Gillan, G. D. Price, Ab initio lattice dynamics and structural stability of MgO. J. Chem. Phys. 118 , 10174 (2003). doi:10.1063/1.1570394

    10. D. Alfe et al ., Quantum Monte Carlo calculations of the structural properties and the B1-B2phase transition of MgO. Phys. Rev. B 72 , 7 (2005). doi:10.1103/PhysRevB.72.014114

    11. Materials and methods are available as supplementary materials on Science Online.

    12. D. G. Hicks et al ., Dissociation of liquid silica at high pressures and temperatures. Phys. Rev. Lett. 97 , 025502 (2006). doi:10.1103/PhysRevLett.97.025502 Medline

    13. J. H. Eggert et al ., Melting temperature of diamond at ultrahigh pressure. Nat. Phys. 6 , 40(2010). doi:10.1038/nphys1438

    14. D. K. Spaulding et al ., Evidence for a phase transition in silicate melt at extreme pressure andtemperature conditions. Phys. Rev. Lett. 108 , 065701 (2012).doi:10.1103/PhysRevLett.108.065701 Medline

    15. Y. B. Zel'dovich, Y. P. Raizer, Physics of Shock Waves and High-Temperature HydrodynamicPhenomena , W. D. Hayes, R. F. Probstein, Eds. (Dover Publications, Mineola, NY, 2002).

    16. R. G. McQueen, S. P. Marsh, J. W. Taylor, J. N. Fritz, W. J. Carter, in High-Velocity Impact Phenomena , R. Kinslow, Ed. (Academic Press, New York, 1970), pp. 293417.

    17. G. D. Stevens, L. R. Veeser, P. A. Rigg, R. S. Hixson, Suitability of magnesium oxide as aVISAR window. AIP Conf. Proc. 845 , 1353 (2006). doi:10.1063/1.2263575

    http://dx.doi.org/10.1126/science.1120865http://dx.doi.org/10.1126/science.1120865http://dx.doi.org/10.1126/science.1120865http://dx.doi.org/10.1103/PhysRevB.81.054110http://dx.doi.org/10.1103/PhysRevB.81.054110http://dx.doi.org/10.1111/j.1365-246X.2009.04142.xhttp://dx.doi.org/10.1111/j.1365-246X.2009.04142.xhttp://dx.doi.org/10.1111/j.1365-246X.2009.04142.xhttp://dx.doi.org/10.1103/PhysRevLett.94.235701http://dx.doi.org/10.1103/PhysRevLett.94.235701http://dx.doi.org/10.1103/PhysRevLett.94.235701http://dx.doi.org/10.1103/PhysRevLett.94.068501http://dx.doi.org/10.1103/PhysRevLett.94.068501http://dx.doi.org/10.1103/PhysRevLett.94.068501http://dx.doi.org/10.1038/371506a0http://dx.doi.org/10.1038/371506a0http://dx.doi.org/10.1029/2008GL034585http://dx.doi.org/10.1029/2008GL034585http://dx.doi.org/10.1029/2008GL034585http://dx.doi.org/10.1103/PhysRevB.73.174208http://dx.doi.org/10.1103/PhysRevB.73.174208http://dx.doi.org/10.1063/1.1570394http://dx.doi.org/10.1063/1.1570394http://dx.doi.org/10.1063/1.1570394http://dx.doi.org/10.1103/PhysRevB.72.014114http://dx.doi.org/10.1103/PhysRevB.72.014114http://dx.doi.org/10.1103/PhysRevB.72.014114http://dx.doi.org/10.1103/PhysRevLett.97.025502http://dx.doi.org/10.1103/PhysRevLett.97.025502http://dx.doi.org/10.1103/PhysRevLett.97.025502http://dx.doi.org/10.1038/nphys1438http://dx.doi.org/10.1038/nphys1438http://dx.doi.org/10.1038/nphys1438http://dx.doi.org/10.1103/PhysRevLett.108.065701http://dx.doi.org/10.1103/PhysRevLett.108.065701http://dx.doi.org/10.1063/1.2263575http://dx.doi.org/10.1063/1.2263575http://dx.doi.org/10.1063/1.2263575http://dx.doi.org/10.1063/1.2263575http://dx.doi.org/10.1103/PhysRevLett.108.065701http://dx.doi.org/10.1103/PhysRevLett.108.065701http://dx.doi.org/10.1038/nphys1438http://dx.doi.org/10.1103/PhysRevLett.97.025502http://dx.doi.org/10.1103/PhysRevLett.97.025502http://dx.doi.org/10.1103/PhysRevB.72.014114http://dx.doi.org/10.1063/1.1570394http://dx.doi.org/10.1103/PhysRevB.73.174208http://dx.doi.org/10.1029/2008GL034585http://dx.doi.org/10.1038/371506a0http://dx.doi.org/10.1103/PhysRevLett.94.068501http://dx.doi.org/10.1103/PhysRevLett.94.068501http://dx.doi.org/10.1103/PhysRevLett.94.235701http://dx.doi.org/10.1103/PhysRevLett.94.235701http://dx.doi.org/10.1111/j.1365-246X.2009.04142.xhttp://dx.doi.org/10.1103/PhysRevB.81.054110http://dx.doi.org/10.1126/science.1120865http://dx.doi.org/10.1126/science.1120865
  • 7/30/2019 McWilliams.SM_Revised.pdf

    30/31

    29

    18. R. Jeanloz, Shock wave equation of state and finite strain theory. J. Geophys. Res. 94 , (B5),5873 (1989). doi:10.1029/JB094iB05p05873

    19. T. J. Ahrens, Ed., AGU Handbook of Physical Constants , vol. II (1995), pp. 4597;http://www.agu.org/books/rf/v002 .

    20. D. A. Young, Phase Diagrams of the Elements (University of California Press, Berkeley, 1991).

    21. S. Tateno, K. Hirose, Y. Ohishi, Y. Tatsumi, The structure of iron in Earths inner core. Science 330 , 359 (2010). doi:10.1126/science.1194662 Medline

    22. D. Valencia, R. J. O'Connell, D. Sasselov, Internal structure of massive terrestrial planets. Icarus 181 , 545 (2006). doi:10.1016/j.icarus.2005.11.021

    23. T. Guillot, The interiors of giant planets: Models and outstanding questions. Annu. Rev. EarthPlanet. Sci. 33 , 493 (2005). doi:10.1146/annurev.earth.32.101802.120325

    24. B. J. Wood, Phase transformations and partitioning relations in peridotite under lower mantleconditions. Earth Planet. Sci. Lett. 174 , 341 (2000). doi:10.1016/S0012-821X(99)00273-3

    25. S.-i. Karato, Rheological structure of the mantle of a super-Earth: Some insights from mineralphysics. Icarus 212 , 14 (2011). doi:10.1016/j.icarus.2010.12.005

    26. F. Gaillard, G. I. Marziano, Electrical conductivity of magma in the course of crystallizationcontrolled by their residual liquid composition. J. Geophys. Res. Solid Earth 110 , (B6),B06204 (2005). doi:10.1029/2004JB003282

    27. E. Knittle, R. Jeanloz, High pressure metallization of FeO and implications for the Earths core.Geophys. Res. Lett. 13 , 1541 (1986). doi:10.1029/GL013i013p01541

    28. D. J. Stevenson, Planetary Magnetic Fields: Achievements and prospects. Space Sci. Rev. 152 ,651 (2010). doi:10.1007/s11214-009-9572-z

    29. V. Solomatov, in Origin of the Earth and Moon , R. M. Canup, K. Righter, Eds. (University of

    Arizona Press, Tucson, 2000), pp. 323338.30. S. P. Marsh, LASL Shock Hugoniot Data (University of California Press, Berkeley, 1980).

    31. M. S. Vassiliou, T. J. Ahrens, Hugoniot equation of state of periclase to 200 GPa. Geophys. Res. Lett. 8 , 729 (1981). doi:10.1029/GL008i007p00729

    32. T. S. Duffy, T. J. Ahrens, Compressional sound velocity, equation of state, and constitutiveresponse of shock-compressed magnesium oxide. J. Geophys. Res. Solid Earth 100 , (B1),529 (1995). doi:10.1029/94JB02065

    33. L. Zhang, Z. Z. Gong, Y. W. Fei, Shock-induced phase transitions in the MgOFeO system to200 GPa. J. Phys. Chem. Solids 69 , 2344 (2008). doi:10.1016/j.jpcs.2008.04.006

    34. D. G. Hicks et al ., High-precision measurements of the diamond Hugoniot in and above the meltregion. Phys. Rev. B 78 , 174102 (2008). doi:10.1103/PhysRevB.78.174102

    35. M. D. Knudson, M. P. Desjarlais, Shock compression of quartz to 1.6 TPa: redefining a pressurestandard. Phys. Rev. Lett. 103 , 225501 (2009). doi:10.1103/PhysRevLett.103.225501Medline

    36. B. Svendsen, T. J. Ahrens, Shock-induced temperatures of MgO. Geophys. J. R. Astron. Soc. 91 ,667 (1987). doi:10.1111/j.1365-246X.1987.tb01664.x

    http://dx.doi.org/10.1029/JB094iB05p05873http://dx.doi.org/10.1029/JB094iB05p05873http://dx.doi.org/10.1029/JB094iB05p05873http://www.agu.org/books/rf/v002/http://www.agu.org/books/rf/v002/http://dx.doi.org/10.1126/science.1194662http://dx.doi.org/10.1126/science.1194662http://dx.doi.org/10.1126/science.1194662http://dx.doi.org/10.1016/j.icarus.2005.11.021http://dx.doi.org/10.1016/j.icarus.2005.11.021http://dx.doi.org/10.1016/j.icarus.2005.11.021http://dx.doi.org/10.1146/annurev.earth.32.101802.120325http://dx.doi.org/10.1146/annurev.earth.32.101802.120325http://dx.doi.org/10.1146/annurev.earth.32.101802.120325http://dx.doi.org/10.1016/S0012-821X(99)00273-3http://dx.doi.org/10.1016/S0012-821X(99)00273-3http://dx.doi.org/10.1016/S0012-821X(99)00273-3http://dx.doi.org/10.1016/j.icarus.2010.12.005http://dx.doi.org/10.1016/j.icarus.2010.12.005http://dx.doi.org/10.1016/j.icarus.2010.12.005http://dx.doi.org/10.1029/2004JB003282http://dx.doi.org/10.1029/2004JB003282http://dx.doi.org/10.1029/2004JB003282http://dx.doi.org/10.1029/GL013i013p01541http://dx.doi.org/10.1029/GL013i013p01541http://dx.doi.org/10.1029/GL013i013p01541http://dx.doi.org/10.1007/s11214-009-9572-zhttp://dx.doi.org/10.1007/s11214-009-9572-zhttp://dx.doi.org/10.1007/s11214-009-9572-zhttp://dx.doi.org/10.1029/GL008i007p00729http://dx.doi.org/10.1029/GL008i007p00729http://dx.doi.org/10.1029/GL008i007p00729http://dx.doi.org/10.1029/94JB02065http://dx.doi.org/10.1029/94JB02065http://dx.doi.org/10.1029/94JB02065http://dx.doi.org/10.1016/j.jpcs.2008.04.006http://dx.doi.org/10.1016/j.jpcs.2008.04.006http://dx.doi.org/10.1016/j.jpcs.2008.04.006http://dx.doi.org/10.1103/PhysRevB.78.174102http://dx.doi.org/10.1103/PhysRevB.78.174102http://dx.doi.org/10.1103/PhysRevB.78.174102http://dx.doi.org/10.1103/PhysRevLett.103.225501http://dx.doi.org/10.1103/PhysRevLett.103.225501http://dx.doi.org/10.1111/j.1365-246X.1987.tb01664.xhttp://dx.doi.org/10.1111/j.1365-246X.1987.tb01664.xhttp://dx.doi.org/10.1111/j.1365-246X.1987.tb01664.xhttp://dx.doi.org/10.1111/j.1365-246X.1987.tb01664.xhttp://dx.doi.org/10.1103/PhysRevLett.103.225501http://dx.doi.org/10.1103/PhysRevLett.103.225501http://dx.doi.org/10.1103/PhysRevB.78.174102http://dx.doi.org/10.1016/j.jpcs.2008.04.006http://dx.doi.org/10.1029/94JB02065http://dx.doi.org/10.1029/GL008i007p00729http://dx.doi.org/10.1007/s11214-009-9572-zhttp://dx.doi.org/10.1029/GL013i013p01541http://dx.doi.org/10.1029/2004JB003282http://dx.doi.org/10.1016/j.icarus.2010.12.005http://dx.doi.org/10.1016/S0012-821X(99)00273-3http://dx.doi.org/10.1146/annurev.earth.32.101802.120325http://dx.doi.org/10.1016/j.icarus.2005.11.021http://dx.doi.org/10.1126/science.1194662http://dx.doi.org/10.1126/science.1194662http://www.agu.org/books/rf/v002/http://dx.doi.org/10.1029/JB094iB05p05873
  • 7/30/2019 McWilliams.SM_Revised.pdf

    31/31

    37. I. Baraffe, G. Chabrier, T. Barman, Structure and evolution of super-Earth to super-Jupiterexoplanets. Astron. Astrophys. 482 , 315 (2008). doi:10.1051/0004-6361:20079321

    38. R. E. Stephens, I. H. Malitson, Index of refraction of magnesium oxide. J. Res. Natl. Bur. Stand. 49 , 249 (1952). doi:10.6028/jres.049.025

    39. D. G. Hicks et al ., Shock compression of quartz in the high-pressure fluid regime. Phys.

    Plasmas 12 , 082702 (2005). doi:10.1063/1.2009528 40. K. Vedam, T. A. Davis, Nonlinear variation of the refractive indices of -quartz with pressure.

    J. Opt. Soc. Am. 57 , 1140 (1967). doi:10.1364/JOSA.57.001140

    41. T. R. Boehly et al ., Properties of fluid deuterium under double-shock compression to severalMbar. Phys. Plasmas 11 , L49 (2004). doi:10.1063/1.1778164

    42. P. M. Celliers et al ., Line-imaging velocimeter for shock diagnostics at the OMEGA laserfacility. Rev. Sci. Instrum. 75 , 4916 (2004). doi:10.1063/1.1807008

    43. P. M. Celliers et al ., Insulator-to-conducting transition in dense fluid helium. Phys. Rev. Lett. 104 , 184503 (2010). doi:10.1103/PhysRevLett.104.184503 Medline

    44. J. E. Miller et al ., Streaked optical pyrometer system for laser-driven shock-wave experimentson OMEGA. Rev. Sci. Instrum. 78 , 034903 (2007). doi:10.1063/1.2712189 Medline

    45. D. K. Spaulding et al ., New Optical Diagnostics for Equation of State Experiments on the JanusLaser, in AIP Conference Proceedings (AIP, 2007), vol. 955, pp. 10711074;http://proceedings.aip.org/resource/2/apcpcs/955/1/1071_1 .

    46. D. E. Grady, in High Pressure Research: Applications in Geophysics M. H. Manghnani, S.Akimoto, Eds. (Academic Press, Inc., New York, 1977), pp. 389438.

    47. T. J. Ahrens, High-pressure electrical behavior and equation of state of magnesium oxide fromshock wave measurements. J. Appl. Phys. 37 , 2532 (1966). doi:10.1063/1.1782080

    48. M. van Thiel, Compendium of shock wave data Tech. Report No. UCRL-50108 (LawrenceLivermore National Laboratory, 1977).

    49. G. E. Duvall, R. A. Graham, Phase transitions under shock-wave loading. Rev. Mod. Phys. 49 ,523 (1977). doi:10.1103/RevModPhys.49.523

    50. W. C. Overton, Relation between ultrasonically measured properties and the coefficients in thesolid equation of state. J. Chem. Phys. 37 , 116 (1962). doi:10.1063/1.1732931

    51. A. F. Ioffe, A. R. Regel, in Progress in Semiconductors (1960), vol. 4, pp. 237.

    52. C. Kittel, H. Kroemer, Thermal Physics (Macmillan, San Francisco, 1980).

    53. M. Born, E. Wolf, Principles of Optics (Cambridge University Press, Cambridge, ed. 7th, 1999).

    http://dx.doi.org/10.1051/0004-6361:20079321http://dx.doi.org/10.1051/0004-6361:20079321http://dx.doi.org/10.1051/0004-6361:20079321http://dx.doi.org/10.6028/jres.049.025http://dx.doi.org/10.6028/jres.049.025http://dx.doi.org/10.6028/jres.049.025http://dx.doi.org/10.1063/1.2009528http://dx.doi.org/10.1063/1.2009528http://dx.doi.org/10.1063/1.2009528http://dx.doi.org/10.1364/JOSA.57.001140http://dx.doi.org/10.1364/JOSA.57.001140http://dx.doi.org/10.1364/JOSA.57.001140http://dx.doi.org/10.1063/1.1778164http://dx.doi.org/10.1063/1.1778164http://dx.doi.org/10.1063/1.1778164http://dx.doi.org/10.1063/1.1807008http://dx.doi.org/10.1063/1.1807008http://dx.doi.org/10.1063/1.1807008http://dx.doi.org/10.1103/PhysRevLett.104.184503http://dx.doi.org/10.1103/PhysRevLett.104.184503http://dx.doi.org/10.1103/PhysRevLett.104.184503http://dx.doi.org/10.1063/1.2712189http://dx.doi.org/10.1063/1.2712189http://dx.doi.org/10.1063/1.2712189http://proceedings.aip.org/resource/2/apcpcs/955/1/1071_1http://proceedings.aip.org/resource/2/apcpcs/955/1/1071_1http://dx.doi.org/10.1063/1.1782080http://dx.doi.org/10.1063/1.1782080http://dx.doi.org/10.1063/1.1782080http://dx.doi.org/10.1103/RevModPhys.49.523http://dx.doi.org/10.1103/RevModPhys.49.523http://dx.doi.org/10.1103/RevModPhys.49.523http://dx.doi.org/10.1063/1.1732931http://dx.doi.org/10.1063/1.1732931http://dx.doi.org/10.1063/1.1732931http://dx.doi.org/10.1063/1.1732931http://dx.doi.org/10.1103/RevModPhys.49.523http://dx.doi.org/10.1063/1.1782080http://proceedings.aip.org/resource/2/apcpcs/955/1/1071_1http://dx.doi.org/10.1063/1.2712189http://dx.doi.org/10.1063/1.2712189http://dx.doi.org/10.1103/PhysRevLett.104.184503http://dx.doi.org/10.1103/PhysRevLett.104.184503http://dx.doi.org/10.1063/1.1807008http://dx.doi.org/10.1063/1.1778164http://dx.doi.org/10.1364/JOSA.57.001140http://dx.doi.org/10.1063/1.2009528http://dx.doi.org/10.6028/jres.049.025http://dx.doi.org/10.1051/0004-6361:20079321