+ All Categories
Home > Documents > Art Diamante

Art Diamante

Date post: 03-Jun-2018
Category:
Upload: cris1104
View: 237 times
Download: 0 times
Share this document with a friend

of 13

Transcript
  • 8/11/2019 Art Diamante

    1/13

    Pergamon

    OOOS-6223 96)00170-O

    Carbon Vol. 34, No. 2, pp. 141-153,1996

    Copyright 0 1996 Elsevier Science Ltd

    Printed in Great Britain.

    All rights reserved

    000%6223/96 15.00+ 0.00

    REVIEW ARTICLE

    THE PRESSURE-TEMPERATURE PHASE AND

    TRANSFORMATION DIAGRAM FOR CARBON; UPDATED

    THROUGH 1994

    F. P. BUNDY,~ W.

    A.

    BASSETT,~ M. S. WEATHERS,~ R. J. HEMLEY,~ H. K. MAO~

    and

    A. F. GONCHAROV~

    Retired General Electric R & D Center; Home address, 4607 Swallow Court, Lebanon, OH 45036-9541,

    U.S.A.

    bDepartment of Geological Sciences, Snee Hall, Cornell University, Ithaca, NY 14853-1504, U.S.A.

    Geophysical Laboratory and Center for High-Pressure Research, Carnegie Institution of Washington,

    5251 Broad Branch Rd, NW, Washington DC, 20015-1305, U.S.A.

    (Recei ved 5

    June 1995;

    accepted

    in

    revi sed orm 12 Sept ember 1995)

    Abstract-In recent years, important advances in our understanding of the pressure-temperature phase

    and transformation diagram for carbon have occurred as a result of developments in both experimental

    and theoretical techniques. Graphite, diamond, liquid and vapor remain the major thermodynamically

    stable forms of carbon, However, due to the high activation energies for solid-state transformations and

    the specific effects of reaction paths, other metastable forms and a wide spectrum of complex hybrid

    forms may be generated, and possibly quenched-in, to survive metastably. This paper focuses primarily

    on developments since the last review of the carbon phase diagram published in 1989, but also includes

    references to the reliable older work. Some of the newer conclusions include the following: the Clapeyron

    slope of the diamond melting line, dT,/dP, is positive; the liquid is metallic and there appears to be no

    evidence for a transformation between electrically conducting and non-conducting forms; melted droplets

    of carbon less than 0.2 pm in diameter quench to a giant fullerene structure even in the stability field of

    diamond; graphite transforms to a transparent phase on compression at room temperature; this phase

    reverts to graphite on decompression at this temperature from pressures as high as 100 GPa.

    1. INTRODUCTION

    The purpose of this article is to bring up to date

    what is known about the thermodynamically stable

    and metastable phases of elemental carbon and the

    reaction dynamics among them over a very wide

    range of pressure, temperature and reaction condi-

    tions. Earlier reports of this type were published

    [l&6] as new results, knowledge, and interpretations

    of experimental results and theory developed. In

    recent years new experiments and results, some of

    which are unpublished, have been generated.

    Therefore, it is deemed appropriate to present the

    topic as an updated whole. The plan for this article

    is to present the entire phase diagram as we currently

    understand it and then discuss each part giving the

    salient references and brief descriptions of the work

    upon which it is based. For the new parts the

    experimental background and results will be covered

    in more detail.

    The binding energy between atoms of carbon is

    very large; for example, the cohesive energy of dia-

    mond is 717 kJ/mol. This property is also demon-

    strated by the extremely high melting temperatures

    of its solid forms (- 5000 K). In addition, once carbon

    atoms are locked into a given phase configuration,

    typically a large amount of activation energy is

    required to produce a different stable phase; in other

    words, very high temperatures are often required to

    initiate spontaneous transformations from one solid

    phase to another. Recent work, however, has shown

    that this is not always true: new pressure-induced

    transformations have been documented at room tem-

    perature, as discussed below.

    In addition to the well known crystalline forms of

    carbon, i.e. graphite and diamond, there are amor-

    phous forms, such as glassy carbon and carbon

    black, and possibly metastable solid forms referred

    to as carbynes. The latter are believed to be solid

    condensations of linear molecules of carbon but

    remain controversial. Also, there are the more

    recently discovered crystalline forms of pure carbon

    molecules, the fullerenes such as ChO buckyballs and

    C,, buckyfootballs. Because of the high cohesive and

    activation energy, carbon polymorphs typically exist

    metastably well into a T,P region where a different

    solid phase is thermodynamically stable. For example,

    diamond survives indefinitely at room conditions

    where graphite is the thermodynamically stable form.

    Conversely, except at very high temperatures, graph-

    ite stubbornly persists at pressures far into the dia-

    mond stability field. The same is true of fullerene and

    amorphous carbons.

    There are two preferred forms of electronic bonding

    of carbon atoms in the solid state, viz.: (i) the sp

    type in which a given atom is bonded to three

    141

  • 8/11/2019 Art Diamante

    2/13

    1 4 2 F

    P. BUNDY et al

    equidistant nearest neighbors 120 apart in a plane,

    as in graphite; and (ii) the sp3 type in which a given

    carbon atom is bonded to four equidistant nearest

    neighbors arranged in tetrahedral symmetry, as in

    diamond. The amorphous or glassy carbon forms are

    thought to be micro-zonal mixtures exhibiting these

    two types of bonding.

    2 THE PHASE AND REACTION DIAGRAM

    The T,P phase and reaction diagram over a wide

    range of pressures for pure carbon is presented in

    Fig. 1. The topology of stability fields of the thermo-

    dynamically stable phases is quite simple: (i) the

    boundary between the graphite and diamond stable

    regions which runs from 1.7 GPa/O K, to the

    graphite/diamond/liquid triple point at about

    12 GPa/5000 K; (ii) the melting line of graphite

    extending from the graphite/liquid/vapor triple point

    at 0.011 GPa/5000 K to the graphite/diamond/liquid

    triple point at 12 GPa/5000 K; and (iii) the diamond

    melting line that runs to higher P and T above the

    triple point. The graphite/liquid/vapor triple point,

    and the graphite/vapor and the liquid/vapor phase

    boundaries, which occur at pressures too low for the

    scale of Fig. 1, will be shown and discussed later.

    Other features illustrated in the diagram include the

    thermal reaction rate thresholds for the solid-solid

    transformations; these will also be discussed later.

    In discussing the various features of the diagram

    shown in Fig. 1, it is important to clarify that there

    are two principal crystallographic forms of both

    graphite and diamond: hexagonal and rhombohe-

    dral graphite; and hexagonal and cubic diamond.

    The basic chemical bonding is the same in both kinds

    of graphite (i.e. sp) and the same in both kinds of

    diamond (i.e. sp3). The slight crystallographic differ-

    ences are the result of different sequences of layering

    in the crystal. In each of the hexagonal types of

    graphite and diamond the layering sequence is

    ABABAB--; that is the third layer of atoms exactly

    superposes the first layer. In rhombohedral graphite

    and cubic diamond the layering sequence is

    ABCABC-; that is the fourth layer exactly super-

    poses the first layer. The two kinds of diamond and

    hexagonal graphite will be referred to in the following

    discussion of Fig. 1.

    Region A in Fig. 1 is the T,P region utilized for

    the high-pressure commercial synthesis of diamond

    from graphite. Although the process starts with solid

    graphite and ends with solid diamond, it is not a

    truly solid-solid transition because in this process

    carbon from the graphite source dissolves into the

    ambient catalyst-solvent fluid metal layer and precipi-

    I

    J

    I

    I31

    ND

    F

    I

    I

    E

    \

    \

    .

    LIQ.

    GRAPHITE

    I

    0 00

    Temperature K)

    Fig. 1. P,T phase and transition diagram for carbon as understood from experimental observations through 1994. Solid

    lines represent equilibrium phase boundaries. A: commercial synthesis of diamond from graphite by catalysis; B: P/7

    threshold of very fast (less than 1 ms) solid-solid transformation of graphite to diamond; C:

    P/T

    threshold of very fast

    transformation of diamond to graphite; D: single crystal hexagonal graphite transforms to retrievable hexagonal-type

    diamond; E: upper ends of shock compression/quench cycles that convert hex-type graphite particles to hex-type diamond;

    F: upper ends of shock compression/quench cycles that convert hex-type graphite to cubic-type diamond; B,F,G: threshold

    of fast P/T cycles, however generated, that convert either type of graphite or hexagonal diamond into cubic-type diamond;

    H,l,J: path along which a single crystal hex-type graphite compressed in the c-direction at room temperature loses some

    graphite characteristics and acquires properties consistent with a diamond-like polytype, but reverses to graphite upon

    release of pressure.

  • 8/11/2019 Art Diamante

    3/13

    P T hase and transformation diagram

    for carbon

    143

    tates out on the growing diamond sink. The activation

    energy of solution is only about 125-165 kJ/mol, so

    this process can proceed at relatively low temper-

    atures and pressures at a rate useful for synthesis.

    The B region on the dashed line B,F,G marks the

    temperature/pressure threshold of very fast (< 1 ms)

    and complete solid-solid transformation of graphite

    to diamond. This transformation always yields cubic-

    type diamond. Experimentally, it is done by pressuriz-

    ing graphite above 12 GPa and heating the sample

    with a pulse of electric current or laser radiation.

    The shaded line, C, marks the temperature/pressure

    threshold of very fast and complete transformation

    of diamond to graphite. It is accomplished experimen-

    tally by compressing diamond crystals embedded in

    a graphite rod and then flash-heating the assembly

    with a pulse of electric current to the threshold

    temperature (or above). No graphitization of the

    diamond specimen occurs unless the critical temper-

    ature is reached, at which point graphitization is

    complete.

    Region D delineates the general area in which

    single-crystal hexagonal graphite transforms grad-

    ually to retrievable hexagonal-type diamond. This

    transformation occurs when the graphite is subjected

    to a deviatoric stress of at least 12 GPa with the

    principal stress parallel to the c crystallographic axis

    and annealing temperatures of 800-2000 K. Region

    E marks the upper ends of shock compression/quench

    cycles that convert hex-type graphite particles,

    embedded in a metal matrix for thermal quenching,

    to hex-type diamond. Note that there is an upper

    limit on the temperature if hex-type diamond is to

    be retrieved.

    Region F designates the upper ends of shock

    compression/quench cycles that convert hex-type

    graphite (embedded in a thermal-quenching metal

    matrix) to cubic-type diamond. The dashed line

    B,F,G, marks the threshold of fast pressure/

    temperature cycles that, however generated, convert

    either type of graphite or hexagonal diamond into

    cubic-type diamond. There have been a number of

    relatively recent experimental results in this area

    which will be described later in the article.

    When single-crystal hex-type graphite is com-

    pressed in the c-direction at room temperature along

    the path H,I,J on the diagram, it maintains its com-

    pressed graphite structure until, in the pressure region

    H-I, it loses some of its graphite characteristics such

    as reflectivity, conductivity and optical opacity. From

    pressures of I (about 23 GPa) up to J and above, the

    Raman spectrum is similar to that of amorphous

    carbon. If such a specimen, under pressure, is heated

    from region J to G, it very quickly organizes itself

    into cubic-type diamond. The experimental evidence

    indicates that the dashed line BFG, as noted above,

    marks the activation energy threshold for the fast

    transformation of compressed graphite, or hex-type

    diamond, into the thermodynamically stable cubic-

    type diamond. According to one theoretical study

    [ 71 perfect single crystal graphite, compressed at zero

    0.1

    0

    0 -

    4000

    5000 6000

    Temperature (K)

    Fig. 2. P,Tphase diagram for the lower pressure region. The

    question marks following carbynes indicate that the exis-

    tence of carbyne solid phases in this P,Tregion is controver-

    sial (see text). The question marks in the liquid phase region

    indicate that the newest more rigorous experimental results

    do not support the earlier experimental interpretations that

    a non-conducting liquid phase of carbon exists (see text).

    K, could be compressed to a pressure of 80 GPa

    before it would spontaneously transform to diamond.

    These results are discussed in more detail in a later

    section.

    The low pressure part of the carbon T,P phase

    diagram (O-O.3 GPa) is shown in Fig. 2. The major

    features are the graphite (and carbyne?)/vapor line,

    the liquid/vapor line and the graphite (carbyne?)

    melting line. Evidence has been reported that the low

    pressure liquid is electrically non-conducting whereas

    the higher pressure liquid is conducting. Indeed, a

    phase boundary between the two kinds of liquids has

    been suggested, as illustrated in the figure. The

    liquid/vapor line should end at a critical point which

    has been estimated to be roughly in the vicinity of

    0.2 GPa/6800 K. At high temperatures below melting

    there is some evidence for transformations from

    graphite to other solid phases which are proposed to

    be carbyne (linear molecules) or chaoite-like [ S,9], -

    but this remains controversial and unresolved [ 3, lo].

    2.1. Diamond/graphite equilibrium line

    The position of the diamond/graphite equilibrium

    line has been quite accurately established by

    thermodynamic calculations based upon the mea-

    sured physical properties of graphite and diamond in

    the temperature range from 300 to about 1200 K

    [ 1 I-131 and by experiments on growing or graphit-

    ization of diamond [ 14,151. The position of the

    upper part of the line, where it approaches the

    graphite/diamond/liquid triple point, is established

    by the confluence of the lines B and C, Fig. 1, which

    correspond to very fast transition of graphite to

    diamond and the reverse, respectively [ 16-191.

  • 8/11/2019 Art Diamante

    4/13

    144 F. P. BUNDY t u/.

    2.2. Gruphite melting line and properties o iquid

    carbon

    The graphite melting line at low pressures is now

    well established [20-261. However, there are differ-

    ences of hundreds of degrees in the temperatures

    reported in the region of the graphite/liquid/vapor

    triple point. There are many difficulties involved in

    obtaining reliable temperature measurements under

    such high temperature conditions. With improve-

    ments in pyrometric equipment and procedures over

    the decades, the quality of the measurements has

    become considerably better. The temperature

    assigned to the triple point has been moved up from

    about 4000 K in the 1940s to about 5000 K in recent

    years. In Fig. 1 we show 5000 K as being the most

    probable value for the temperature at the triple

    point [ 271.

    The graphite melting line has been investigated

    experimentally by Bundy [ 161, Fateeva and

    Vereshchagin [ 281 and Togaya er al. [ 191 using flash

    electrical resistance heating of graphite specimens in

    static high pressure apparatus. All studies found the

    melting temperature to increase with pressure to a

    maximum at 5-6GPa, then to decrease toward the

    graphite/diamond/liquid triple point. The temper-

    ature of the latter is about the same as that of the

    graphite/liquid/vapor triple point. These studies dis-

    agree, however, on the magnitude of the temperature

    at the maximum. Fateeva and Vereshchagin [28]

    reported that temperature at the maximum was

    1200 K above the triple point, whereas Bundy [ 161

    reported a temperature difference of 600 K and

    Togaya et ~2. found 200 K. The study of Togaya et al.

    [19] utilized the most sophisticated apparatus and

    instrumentation, so it may be considered the most

    reliable. In this work, the temperatures were based

    on his measured values of enthalpy to melting, which

    were then converted to temperatures by applying the

    data of Scheindlin and Senchenko 1251 on enthalpy

    versus temperature at conditions near the melting

    point.

    Experimental measurements of the heat of fusion

    of graphite to liquid were reported by Bundy 1161,

    Heremans et al. [26] and Baitin

    et cd.

    27]. The data

    of Bundy [16] and of Heremans et al. [26] both

    indicate a value of about 105 kJ/mol and the observa-

    tions of Baitin et ul. [27] show it to be greater than

    80 kJ/mol. These investigations, together with those

    of Gathers er LI/.1231 and Togaya et ul. [ 191, indicate

    that the enthalpy increase of graphite from room

    temperature to melting. lies in the range of

    92-t 15 kJ/mol. The value 115 kJ/mol is probably the

    most reliable.

    Observations on the electrical conductivity of

    liquid carbon are varied and contradictory.

    Experiments have been performed in which the speci-

    men is confined by solid refractory walls and flash-

    heated with a short pulse of electrical current through

    the specimen itself [ 16,191. All such experiments at

    higher pressures indicate that the resistivity of the

    liquid phase is about 60-70s of that of the solid

    graphite at the melting temperature. In Bundys [ 161

    experiments at 4.8 GPa the resistivity of the liquid

    carbon was observed to be about 350 @cm, whereas

    the data of Togaya

    et (II. [

    191 at 5.6 GPa show about

    460 @cm.

    Experiments have also been performed in which a

    rod or bar-shaped specimen is mounted in a chamber

    with a high pressure atmosphere of an inert gas (such

    as argon) to prevent vaporization and heated with

    an electric current, usually in milliseconds or micro-

    seconds [20,22,23,26]. These experiments yield

    widely different values for the resistivity of the liquid

    phase. Ludwig [ZO] and Jones [22] reported the

    resistivity of the liquid formed from melting graphite

    to be very high, practically an insulator. Gathers

    et d. 1231 measured values of about 1000 @cm.

    Heremans vt (II. [26] melted graphite fibers under

    similar conditions and report values of 30-70 @2cm.

    Vaporization and melting of graphite have also

    been investigated by intensive surface-heating with

    focused lasers. Both continuous wave and pulsed

    experiments as short as 90 fs have been performed

    [27.29%321. Among the objectives of these investiga-

    tors were the determination of the optical, electrical

    and thermal properties of liquid carbon. The results

    tend to be contradictory, depending upon the energy

    insertion time, the method of monitoring the sample

    spot, etc. Experiments reported before 1988 are dis-

    cussed by Bundy 143 and will not be treated further

    here. Baitin et

    al.

    [27] found the emissivity of liquid

    carbon at i=O.65 jtrn to be about 0.60; this is close

    to values for liquid metals and is therefore consistent

    with liquid carbon being metallic. Reitze

    et al.

    [32]

    used 90 fs laser pulses which. due to inertial effects,

    result in heating at constant volume. The sample

    surface was probed with approximately 100 fs reso-

    lution without interference from vaporized carbon.

    According to the argument of Bundy [4] regarding

    thermal pressure due to heating at nearly constant

    volume (see Figs 3 and 4 in Ref. [4]), the pressure in

    the zone of the sample melted by the laser would be

    in the range of 5-10 GPa and the carbon melt would

    be electrically conducting with a resistivity of about

    400 &?cm. Optically derived d.c. resistivities from the

    experiments of Reitze et ul. [32] are about 600 2cm.

    which is in quite good agreement with the static

    pressure results of Bundy [ 161 and Togaya

    clt LI/.

    [ 191. All of the results listed above, except for those

    of Malvezzi et a/. [31], are consistent with liquid

    carbon having semimetallic properties.

    Consideration of all the reported experimental

    observations brings up the question of whether there

    is in fact an electrically insulating phase of liquid

    carbon. In exploding wire types of experiments, a

    wire or rod specimen is heated into the liquid state

    very rapidly in a gas or vacuum environment. As

    soon as the sample begins to melt, its physical shape

    becomes unstable due to surface tension and electro-

    magnetic forces, as demonstrated by stroboscopic

    and streak camera photographic measurements. Even

    in the most recent carbon melting experiments, in

    which the best circuit control and diagnostic tech-

    niques are used. it has not generally been possible to

  • 8/11/2019 Art Diamante

    5/13

    P,T phase and transformation diagram for carbon

    145

    reach complete melting without disruptive instabilit-

    ies occurring. In order not to lose the specimen by

    disintegration, the power to the specimen may be cut

    off at the instability point by crow bar shunt cir-

    cuitry. There appears to be no firm evidence from

    the most recent experiments that supports the exis-

    tence of an electrically insulating form of liquid

    carbon.

    Recent first-principles calculations by Galli et al.

    [33] indicate that liquid carbon at low pressure and

    5000 K should have mainly two-fold and three-fold

    coordination. Similar calculations for liquid carbon

    at 9000 K and pressures above 100 GPa [ 341 predict

    the coordination to be mainly four-fold. A simple

    theoretical approach to modelling liquid carbon has

    been to consider it as a pseudo-binary mixture of

    graphite-like and diamond-like states in which the

    composition changes smoothly with pressure [ 351.

    This approach leads to a melting curve maximum, as

    is found experimentally for graphite. On the other

    hand, the first-principles molecular-dynamics calcula-

    tions performed along high-density isochores indicate

    that the melting curve of diamond should have a

    positive slope, i.e. (dP/dT, >O) to above 100 GPa

    [34]. In the section on melting of diamond, it will

    be shown that there is some experimental evidence

    that liquid carbon at pressures above the

    graphite/diamond/liquid triple point does contain

    graphite-like and diamond-like linkages [ 36,371.

    2.3. The diamond melting line

    Prior to the mid-1980s it was thought that the

    behavior of cubic-type diamond would be analogous

    to that of the diamond-structure forms of the heavier

    Group IV elements Si and Ge. Experiments demon-

    strated that diamond-structure Si and Ge melt to

    form metallic liquids of higher density than the solid,

    which requires dT,/dP to be negative. It was also

    known experimentally that at higher pressures, dia-

    mond-structure Si and Ge undergo transitions at

    room temperature to denser, metallic phases [6,38].

    By analogy, the earlier

    T,P

    phase diagrams for carbon

    showed the melting line of diamond with a negative

    slope, extending to a diamond/metallic solid/liquid

    triple point at higher pressure and lower temperature.

    The melting line of the (supposed) metallic solid was

    believed to have a positive slope, as found in Si and

    Ge [ 1,2].

    This was the case in the first experimental observa-

    tions of melting of diamond at pressures above the

    graphite/diamond/liquid triple point reported by

    Bundy [ 16,171. The data indicated that at the triple

    point, the slope of the diamond melting line,

    dT,JdP, is about zero. However, by analogy with Si

    and Ge the melting line was thought to curve back

    to give a negative slope at higher pressures.

    Shaner, er al. [39] shock compressed pyrolytic

    graphite and measured the sound velocity in carbon

    at shock pressures from 80 to 140GPa and corre-

    sponding shock temperatures of 1500 to 5500 K,

    respectively. The sound derived velocities are close

    to the elastic longitudinal wave in solid diamond,

    which is much higher than that for a bulk wave in a

    carbon melt. They concluded that the diamond

    carbon had not melted even though the shock states

    reached far into the supposed region of liquid carbon

    and that the diamond melting line must therefore

    have a positive slope. At this same time diamond-

    anvil cell experiments indicated the stability of cubic

    diamond to extend to at least 275 GPa [40]. These

    experimental results are consistent with theoretical

    calculations of the phase stability of diamond relative

    to the possible metallic forms [41-431 which indi-

    cated that the cubic diamond form of carbon would

    remain stable relative to all the likely metallic forms

    up to pressures in the range of 1300-2300 GPa and

    that the diamond melting line must have a positive

    slope. This difference in behavior between carbon

    and Si and Ge is attributed to the absence of

    p-electron states in the 1s atomic core which allows

    the p-character sp3 bonding electrons in diamond to

    be held close to the nucleus [41].

    Gold et al. [44] melted diamond at high pressure

    in a diamond anvil cell. Later, Weathers and Bassett

    [36] reported detailed results of diamond-anvil cell

    experiments of melting small particles of diamond or

    graphite in a NaCl matrix at pressures of 5530 GPa

    by a pulsed Nd:YAG laser. They found that liquid

    carbon is immiscible with molten NaCl and that the

    carbon specimens quenched to spherical particles

    ranging in size from - 1 pm down to less than a few

    nanometers (Fig. 3). For the samples melted at

    30 GPa the larger spherules

    (>

    0.2 pm) were polycrys-

    talline diamond with either a granular or radial

    texture. The smaller spherules (< 0.2 pm) yielded

    electron diffraction patterns with four diffuse rings

    that correspond to the 002, 100, 004 and 110 d-

    spacings of graphite - a pattern that is typical of

    disordered graphite randomly oriented around the

    c-axis. Dark-field electron microscope images pro-

    duced by aperturing the 002 diffraction ring showed

    that in the tiny spherules the c-axis had a radial

    orientation. Also, it was found that the 002 and 004

    (c-axis) spacings in quenched unloaded samples were

    smaller in those samples that had been pulse heated

    at higher pressures, while the 110 and 100 (a,b-axis)

    spacings were insensitive to the pressure at the time

    of pulse heating. This effect was interpreted to indicate

    a graphitic structure with some sp3 diamond bond

    stitching to hold the layers closer together on the

    average and also that the diamond melt contains a

    significant amount of sp2 as well as sp3 coordination

    at pressures of 30 GPa. In 1993, after the fullerene

    structures of carbon became well known and estab-

    lished, Bassett and Weathers [ 371 re-interpreted their

    1987 [ 361 results for the smaller

    (

    0.2 mm) solidified as polycrystal-

    line diamond, it was interesting to note that the

    spherules of borderline size commonly had a polycrys-

    talline diamond core surrounded by an onion-struc-

    tured fullerene-like mantle. This indicates that under

    special circumstances the hybrid fullerene structure

    can be energetically stable relative to diamond even

    at conditions well within the diamond stability field.

    Togaya [ 181 reported experiments in which speci-

    mens of boron-doped semiconducting diamond were

    melted at ambient pressures of 6-18 GPa by flash-

    heating them with controlled current from a capaci-

    tor. Although the behavior of the resistance was

    complicated somewhat by changes associated with

    the doping element, and at the lower pressures by

    the transformation to solid graphite, there were clear

    indications that the T, of diamond increases with

    pressure. This result is consistent with Togayas obser-

    vation, and that of Bundy [ 171, that at confining cell

    pressures of 14-18 GPa (well into the diamond-stable

    region) the central pocket of liquid carbon, enclosed

    by a rigid shell of solid diamond always freezes to a

    mix of diamond and graphite. The explanation is that

    the strong container of solid diamond with its low

    coefficient of thermal expansion serves as a nearly

    constant volume space for the freezing process. The

    first liquid freezes as diamond having a smaller

    specific volume than the liquid. The chamber being

    constrained to constant volume loses ambient pres-

    sure until the equilibrium pressure for graphite is

    reached after which the remaining liquid freezes as

    graphite. According to the Clapeyron equation, a

    solid of smaller volume than the melt requires

    dT,/dP to be positive since the entropy of the melt

    is always larger than that of the solid.

    The positive slope of diamond melting is further

    supported by the early observation of Bundy [ 161

    that the threshold T P line for the fast transforma-

    tion of diamond to graphite at pressures below the

    graphite/diamond/liquid triple point (i.e. metastable

    melting of diamond) has a positive slope (C in Fig. 1).

    Additional experimental evidence that the slope,

    dT,/dP, of the diamond melting line is positive was

    obtained from an unpublished experimental study by

    Bassett and Weathers. A mixture of Pt grains and

    NaCl loaded in a diamond-anvil cell was brought up

    to pressure and the individual Pt grains were heated

    with a focused Nd-YAG laser. The temperature of

    each grain was determined using spectroradiometric

    techniques. After unloading, the Pt/NaCl specimen

    was removed with the position of each Pt grain

    indexed; the diamond anvil faces were then searched

    for craters that would indicate melting of the diamond

    face. The melt craters were observed and found to

    correlate with adjacent Pt grains that were heated to

    high temperatures.

    In this manner, the P-T plot shown in Fig. 4 was

    generated. The boundary line marks the minimum

    temperature at which melt craters were observed.

    This line is quite well defined and has a positive

    slope. Note, however, that there is a large discrepancy

    between the temperatures measured here and those

    found for melting at the triple point graphite-dia-

    mond-melt in quite different types of experiments;

    this is shown by the dashed lines which are based on

    the data presented in Fig. 1. Although the measure-

    ment of absolute temperatures by this spectroradio-

  • 8/11/2019 Art Diamante

    7/13

    P,T phase and transformation diagram for carbon

    147

    Temperature, K

    Fig. 4. Plot of relative temperatures of melting of carbon as a function of pressure. Solid circles represent P,Tconditions

    at which a melt crater was formed in the diamond anvil surface when a millisecond laser pulse heated a platinum grain

    embedded in NaCl a few micometers from the anvil surface. Temperatures were based on black-body spectra of incandescent

    light and pressures were measured by the ruby method. The dashed line corresponds to the melt lines shown for graphite

    and diamond in Fig. 1 (see text regarding the displacement).

    metric method is subject to many technical difficulties,

    the relative temperature measurements are probably

    quite reliable. The significant point in this case is that

    all the data points shown in Fig. 4 were taken with

    the same physical and optical set-up and so the slope

    of the indicated melting line for diamond should be

    meaningful, especially the indication that the sign of

    the slope is positive.

    Theoretical studies of the phase stability of carbon

    at extremely high

    T,P

    by Young and Grover [46]

    indicate the melting line of diamond and the

    solid/solid phase boundary between diamond and the

    (supposed) BC8 metallic phase are as shown in Fig. 5.

    The maximum pressures and temperatures in the

    interiors of the outer planets Uranus and Neptune

    are in the range of 600 GPa/7000 K; under these

    conditions, calculations by Ross [47] indicate that

    the elemental carbon there would be in the diamond

    phase. Available data indicate that free carbon in the

    Earths mantle (to 135 GPa, 250%3500 K) would

    also be solid diamond phase as well.

    1500

    G

    &

    w 1000

    2

    9

    8

    PI

    500

    0

    DIAMOND

    IQUID

    I

    -I

    5000

    10,000

    Temperature K)

    I I

    Fig. 5. P,Tphase diagram for carbon up to extremely high

    pressures based on theoretical calculations [46].

  • 8/11/2019 Art Diamante

    8/13

    148

    F. P. BUNDY et al.

    2.4. The graphite-diamond direct conversion line

    The dashed line, B-F-G, in Fig.

    1

    marks the thresh-

    old of very fast (i.e. ms-ii) transition of highly

    compressed graphite, or its low temperature deriva-

    tives, to cubic-type diamond. The lower pressure

    portion of the line near the graphite/diamond/liquid

    triple point was first explored by Bundy [ 171 by

    millisecond resistance heating of specimens in a high

    compression belt apparatus. Recently, this threshold

    was investigated by Bassett and Weathers [unpub-

    lished] up to 35 GPa using the laser-heated diamond-

    cell technique like that described above. In these

    experiments. a shutter was installed to chop the

    continuous infrared laser beam (i= 1.06 mm) into

    pulses of approximately 200 ms. The spectrum was

    recorded as a function of time and fit to black-body

    curves by a least squares method. The wavelength of

    the fluorescence emission from a chip of ruby [48,49]

    in the sample was used to measure the pressure before

    heating. Measurement of the temperature and pres-

    sure of the graphite-diamond transition is particu-

    larly well suited for investigation by this technique.

    When graphite is converted to diamond, the sample

    becomes transparent (Fig. 6) and the absorption of

    the infrared light ceases. Thus, the temperature rises

    to the transition temperature and remains there while

    the conversion is taking place.

    Samples for this study were prepared by placing

    graphite fibers c 1 kern in diameter in a sodium chlo-

    ride matrix and compressing the mixture in a stainless

    steel gasket between diamond anvils ranging from

    0.3 to 0.5 mm in diameter. The fibers were provided

    by Dr Gary Tibbitts of General Motors. Temperature

    measurements on the graphite-diamond conversion

    can be made quite accurately and repeatably with

    the technique for two reasons: (1) the graphite

    behaves as an excellent black body; and (2) the

    temperature at the conversion ceases to rise when the

    carbon is converted to the transparent diamond form

    and no longer absorbs the infrared radiation from

    the laser.

    The direct conversion graphite-diamond boundary

    extends from the graphite/diamond/melt triple point

    along a nearly straight line to 2500 K/l5 GPa with

    a negative slope. It then makes a sharp turn and

    follows a nearly straight line with a gentle negative

    slope to the highest attainable pressure point at

    2000 K/35 GPa (Fig. 7). This line is in fairly good

    agreement with that determined by Bundy [ 171.

    The large number of new data points collected in

    the vicinity of 20 GPa/2500 K indicate a rather

    abrupt change in slope. Figure 7 shows this abrupt

    change in slope along with the metastable room

    temperature transition (discussed below) reported by

    Hanfland rr al. [SO], Goncharov [ 5

    I]

    and Yagi rt trl.

    [ 521. A metastable phase boundary may extend from

    this room temperature metastable transition to the

    point of abrupt slope change in our curve. resulting

    in a metastable triple point there. The me&table

    transition at room temperature is discussed in more

    detail below.

    The construction of a hypothetical phase line across

    H,D,B (Fig.

    1)

    as a metastable extension of the graph-

    ite melting curve would imply structural similarity, if

    not thermodynamic continuity, between the liquid

    and the diamond-like phase produced in the low-

    temperature compression experiments. In other

    words, such topology would suggest that the material

    is glassy. The transition could then be analogous to

    the pressure-induced amorphization of ice-1 [ 533 also

    observed for other materials in which the melting

    curve has a negative slope combined with a frustrated

    10pm

    Fig. 6. Optical micrograph showing conversion by laser heating of graphite fibers to diamond (circled) at pressures greater

    than 12.5 GPa. The preservation of the morphology of the fibers indicates that the graphite (opaque) does not melt. but

    undergoes direct conversion to diamond (transparent).

  • 8/11/2019 Art Diamante

    9/13

    P T

    phase and transformation diagram for carbon

    149

    PHASES OF CARBON

    I

    CJ

    iamond

    I

    I

    I

    e:

    transparent

    a

    w

    metastable &

    1

    :

    20

    graphite

    b.0 0

    0 I

    a

    \

    01.

    I I. I. I,

    t. I

    0

    1000 2000 3000 4000 5000 6000

    Temperature K)

    Fig. 7. P,Tplot of direct graphite-diamond conversion. The

    dashed line separates

    P T

    oints of graphite fibers that

    remained graphite (solid circles) from those at higher P,T

    conditions that were converted to diamond by millisecond

    laser heating pulses [Bassett and Weathers, unpublished].

    solid-solid transition as in SiO, [54]. This inter-

    pretation remains conjectural as the structural state

    and thermodynamic properties of the carbon material

    at high P-T(e.g. above the proposed metastable triple

    point) has not been fully characterized. Alternatively,

    the high temperature transition near 15 GPa (Fig. 7)

    may be unrelated to the metastable extension of the

    graphite melting curve, as indicated in Fig. 1.

    Recent calculations of Zhang

    et al

    [55] on the

    reaction rate of the direct transition of graphite to

    diamond as a function of P and Tyield a fast reaction

    line on a P,Tchart that is remarkably similar to the

    BFG threshold of Fig. 1 and Fig. 7. This similarity

    implies that the shape of the BFG threshold may be

    determined mainly by the reaction kinetics.

    We also point out that this behavior of carbon is

    analogous to that of the iso-electronic case of boron

    nitride which was investigated by Corrigan and

    Bundy [ 561. These studies of boron nitride indicated

    a reaction rate controlled threshold line between

    wurtzite BN (like hexagonal diamond) and zinc-

    blende BN (like cubic-type diamond) which is very

    similar to that of BFG for carbon in Fig. 1 and Fig. 7.

    2.5. Cool compression ofgraphite h,i,j, Fig. 1)

    Aust and Drickamer [57] reported the first meas-

    urements of the electrical resistivity behavior of single

    crystal graphite compressed at room temperature

    well into the lo-20 GPa pressure region. Compressed

    in the c-direction and monitored in the a,b direction,

    the graphite that they observed slowly underwent a

    decrease of resistivity with increase of pressure to

    - 12 GPa,

    whereupon

    the

    resistivity rapidly

    increased. In some of the better specimens the resistiv-

    ity rise was more than a factor of 10, suggesting a

    change of phase to a non-conducting form.

    Bundy and Kasper [ 581 reported results of similar

    experiments with additional arrangements for heating

    the specimen while under pressure to moderate tem-

    peratures up to about 1500 K. For the room temper-

    ature experiments the same fast rise of resistivity as

    reported by Aust and Drickamer [57] started at

    about 12 GPa, but upon return to room pressure the

    resistivity returned to its initial value and the reco-

    vered specimen proved to be graphite. By contrast,

    when the graphite specimen was compressed into the

    higher resistivity state and then heated to around

    1200 K the resistivity increased significantly and when

    it was cooled and decompressed the resistivity

    increased even more. The recovered specimen was no

    longer shiny like graphite, but a dull gun-metal gray.

    It would scratch sapphire the way diamond does and

    its X-ray diffraction pattern showed it to be mostly

    hexagonal diamond. In general, the static pressure

    experiments indicated that in order to retrieve hexag-

    onal diamond, it was necessary to subject the speci-

    men to P,Tconditions in area D of Fig. 1.

    When diamond anvil cells became available to

    generate pressures of up to many 10s of giga-Pascals,

    the behavior of very small specimens of graphite and

    graphitic carbons could be monitored by optical,

    Raman, and X-ray diffraction techniques in

    situ

    at

    high pressure, in contrast to the previous experiments

    which were limited to electrical conductivity measure-

    ments using opaque tungsten carbide anvils. Of these

    types of measurements, X-ray diffraction and Raman

    are able to add significant structural and crystal

    chemical information to our knowledge of the carbon

    phases. The observed behavior of a given specimen

    during compression depends upon its initial degree

    of crystallinity, the direction and homogeneity of the

    pressure field, the rapidity of compression and the

    type of measurement performed. These effects are

    illustrated by the experimental results of a number

    of different workers, some of which follow.

    Goncharov et al. [ 593, using a diamond anvil cell,

    found that single crystal graphite transforms to an

    optically transparent state at about 35 GPa. Although

    this result suggested the formation of sp3 bonding,

    the Raman spectra showed no hint of the diamond

    Raman band and the spectral absorption edge was

    too broad for that of diamond. Further work by

    Goncharov [Sl] showed that the transition in single-

    crystal graphite under hydrostatic conditions (He

    medium) occurs at about 23 GPa, as indicated by an

    abrupt change of the Raman spectrum and the open-

    ing of the band gap. Specimens with less regular

    stacking between hexagonal layers required higher

    pressures (to 45 GPa) to achieve the same transition.

  • 8/11/2019 Art Diamante

    10/13

    150

    F. P. BUNDY

    et al.

    Goncharov [Sl] also found that the transition pres-

    sure for single-crystal graphite can be lowered to

    about 14 GPa under uniaxial stress directed along

    the c-axis of the crystal. These Raman and optical

    observations were corroborated by Hanfland et ul.

    [60] and Utsumi et ul. [61].

    Hanfland et u[. [SO] used X-ray diffraction and

    Raman spectroscopy to explore the behavior of

    graphite in diamond cells at pressures in the range

    of 14 GPa where the previously reported increase in

    electrical resistance and decrease of reflectivity occurs.

    From the behavior of the widths and shifts of the

    critical Raman bands and the X-ray diffraction pat-

    terns, these authors concluded that the phase trans-

    ition started at about 14 GPa. Yagi et ul. [ 521 used

    a cubic anvil apparatus with sintered diamond piston

    tips and X-rays from a synchrotron source to study

    the compression of graphite. They reported that for

    room temperature compression the behavior is sensi-

    tive to the crystal texture of the specimen, but that

    in general, graphite begins to transform at about

    18 GPa.

    Kertesz and Hoffmann [62] and Fahy et ul. [ 71

    performed theoretical calculations of the structure

    and energy of graphite during cool compression as it

    approached the density of diamond. These studies

    indicated that a very high activation energy is associ-

    ated with transforming the strained graphite to the

    diamond. In particular, the study by Fahy et al. [7]

    indicated that a pressure of over 80 GPa (T=O K)

    would be required to force a spontaneous transition.

    From these experiments and calculations, it may

    be concluded that when single-crystal graphite is

    compressed along the c-axis at room temperature

    (H,I,J of Fig.

    1),

    the material remains a good electrical

    conductor in the a,b plane directions until about

    12-14 GPa. At higher pressures decreasing electrical

    conductivity and optical reflectivity and increasing

    optical transmittance are observed. The fact that the

    material remains transparent to visible light suggests

    that it remains

    electrically non-conducting.

    Mechanically, the material is extremely strong and

    was observed to fracture the faces of the single-crystal

    diamond anvils that push against it [Hemley and

    Mao, unpublished]. Heating this material causes the

    sample to transform into coherent hexagonal dia-

    mond crystallites as shown by the experiments of

    Bundy and Kasper [SS] and by Utsumi and Yagi

    C631

    The changes in structure and other physical proper-

    ties accompanying cool compression of graphitic

    carbon are still not fully understood. It is likely that

    sp3 bonding begins to develop at about 12-14 GPa,

    thereby removing conduction electrons from the

    system and decreasing the electrical conductivity and

    optical reflectivity and increasing the optical transmit-

    tance. The fact that decompression from this state

    yields graphite implies either a martensitic trans-

    formation or a small activation energy for the back

    transformation. The latter is supported by very recent

    observations of pressure quenching the transparent

    form at low temperatures [Badding, private com-

    munication]. It is also likely that small domains of

    the high-pressure phase are formed initially on room-

    temperature compression. Heat-annealing the mate-

    rial may then cause the small domains to coalesce

    and grow into coherent hexagonal diamond crystal-

    lites that are large enough and stable enough to

    survive decompression 158,631. Another explanation

    is that the transparent high-pressure, room-temper-

    ature phase is a new allotrope of carbon and perhaps

    has the properties described here for reasons other

    than sp3 bonding. Clearly further work is needed in

    order to resolve this question.

    2.6. Shock compression of graphites e,f; Fig. 1)

    Both temperature and pressure rise during shock

    compression. The rise of temperature with pressure

    takes place with the compressional work done on the

    specimen, the PdVwork. The greater the compress-

    ibility and porosity of the specimen material the

    greater the temperature rise for a given increase in

    pressure. On Fig. 1, the T,P paths for shock compres-

    sion of graphite would pass from room T,P toward

    D,E,F. The T,P path is not reversible and unless fast

    thermal quenching is provided, the temperature drops

    much less rapidly than the pressure during the decom-

    pression. In the commercial production of diamond

    powder from graphite powder, the body to be shock

    compressed consists of small particles of graphite

    dispersed in a matrix of metal, such as iron or copper,

    which is a good thermal conductor and less compress-

    ible than graphite. During shock compression, the

    graphite particles heat up more than the surrounding

    metal matrix, transform to diamond, and are ther-

    mally quenched during the decompression by the

    adjacent cooler metal. The type and size of diamond

    particles that can be produced by this method are

    established and limited by the shortness of the reac-

    tion time and thermal/geometry requirements

    [64468].

    Shock compression experiments of this type on

    hexagonal graphite particles yield mostly hexagonal

    diamond if the top of the cycle goes to the E region

    (Fig. 1) and mostly cubic-type diamond if the top of

    the cycle goes past the BFG line. To be retrieved as

    diamond the specimen particles have to be at high

    temperature long enough to allow coalescing of the

    diamond domains into crystallites large enough that,

    with thermal quenching, they can survive decompres-

    sion without graphitizing. Only small particles of

    diamond can be produced by this method because of

    the very short time at high T,P and the necessity

    for the individual carbon particles to be small in

    order to satisfy the fast thermal quench requirements.

    3. DIAMOND SYNTHESIS UNDER METASTABLE

    CONDITIONS

    As early as 1910 Tammann [69] proposed a T,P

    phase diagram for carbon which included a large T,P

    area at lower pressures and temperatures in which

    there would be pseudo-equilibrium between dia-

    mond and graphite. During this time, it also was

  • 8/11/2019 Art Diamante

    11/13

    P,T phase and transformation diagram for carbon

    151

    recognized that the free energy of carbon vapor was

    far higher than in either the graphite or diamond

    solid forms. The quantitative determination by

    Rossini and Jessup [ 111 of the free energies of

    graphite and diamond showed that their difference

    at one atmosphere, 0 K, is only about 2.5 kJ/mol.

    This difference is insignificant compared to

    -712 kJ/mol of the vapor. Hence, with respect to

    free energy, the condensation of carbon vapor to

    graphite or diamond would be of nearly equal prob-

    ability. By proper control of the reaction path and

    the nucleation conditions it should be possible to

    condense carbon vapor as diamond under low pres-

    sure, metastable conditions.

    The long history of attempts to accomplish this

    include von Bolten [70], Ruff [71], Spitsyn and

    Derjaguin [72], Oriani and Rocco [73], Eversole

    [74], Angus

    et al.

    [75], Matsumoto et al. [76],

    Kamo et al. [77] and Spitsyn

    et al.

    [78]. The method

    was generally unsuccessful because of poor control

    of graphite nucleation until the early 1980s when it

    was found that carbon vapor deposition from a

    plasma, which contained atomic hydrogen as well as

    carbon vapor and hydrocarbon molecules, nearly

    eliminated the competition of graphite nuclei [77].

    This allowed only the diamond to exist and grow.

    Under proper conditions it has become possible to

    deposit films, and sheets, of polycrystalline diamond

    which have commercial utility. Recent studies have

    revealed that organic molecules in the vapor play an

    important role by configuring the carbon to precipi-

    tate as diamond [79].

    4. FULLERENE CARBONS

    Rohlfing

    et al.

    [SO] reported some curious results

    from experiments with a cluster molecular beam

    apparatus. They used a laser to produce puffs of

    vapor from a carbon rod which was entrained and

    cooled by expansion through a supersonic nozzle to

    form small nanometer-size clusters of carbon atoms.

    The relative abundance of clusters of different size

    was examined on line by mass spectroscopy. In

    addition to the usual 2-20-atom clusters studied

    previously by numerous workers and different tech-

    niques, carbon clusters in the 40-300-atom mass

    range were observed. Interestingly, only even-num-

    bered atom clusters formed.

    In the following year, 1985, a similar apparatus

    was set up by Kroto et al. [Sl] to synthesize and

    study the chemical reactivity of the small 2-30-atom

    clusters; in particular, to determine whether these

    clusters had the same form as the long carbon linear

    chains known to be abundant in interstellar space.

    The study of the small clusters was successful, but

    the most significant finding was that the C,, cluster

    was outstandingly stable, as well as C,, and other

    species. Thinking that the great stability might be

    related to some kind of closed shell structure the idea

    soon developed that C,, might be a truncated icosa-

    hedral cage made up of hexagons and 12 pentagons

    (like the markings on a soccer ball) with a carbon

    atom at each intersection [ 811. Geometrically, it was

    soon realized, 12 pentagonal defects convert a planar

    hexagonal array of any size into a quasi-icosahedral

    cage [ 823 as in C6,,, C240 and C,,,. C,, was visualized

    as having 12 pentagons and 25 hexagons by Curl

    and Smalley [83]. Kratchmer is credited with the

    discovery that CeO and C,, are soluble in benzene.

    That, together with Huffmans productive soot-

    making apparatus [84,85] made it possible to pro-

    duce macro amounts of C& and C,,, crystallize it

    and determine the crystal structure. The lattice spac-

    ings in the crystal were consistent with the presumed

    size of the closed-cage molecules. The electronic and

    vibrational spectra of the molecules had already been

    probed by Curl and Smalley [S6] and, in these,

    theory and observation agreed, indicating that the

    closed cage structure of the molecules was correct.

    The discovery and development of methods of

    synthesizing and separating out bulk quantities of

    fullerenes have opened up new fields in chemistry

    and physics. No attempt is made to review these

    large and rapid developments in this article. Earlier

    in this article it was pointed out that Bassett and

    Weathers now interpret the graphite-like spherules of

    their quenched nanometer-sized carbon droplets as

    being onion-layered giant fullerenes which formed in

    the stability field of diamond. We note there have

    been a number of studies of high-pressure trans-

    formations of fullerenes. Some of the more interesting

    ones are as follows: Duclos

    et al.

    [87] on the effects

    of pressure and stress on C& fullerite to 20 GPa;

    Samara et al. [ 881 on the pressure dependence of the

    orientational ordering in solid C6,,; Regueiro et al.

    [S9] on crushing of C,, to diamond at room temper-

    ature; Chandrabhas

    et al.

    [90] on reversible pressure-

    induced amorphization in solid (&; di Brozolo

    et al.

    [91] on fullerenes in an impact crater on the LDEF

    space craft; Nunez-Regueiro et al. [92] on polymer-

    ized fullerite structures; Wolk

    et al.

    [93] on pressure-

    induced structural metastability in crystalline C,,. A

    curious reaction difference is indicated for Cc0 versus

    CT0 in that CT0 compresses at room temperature to

    amorphous form, much like graphite and decom-

    presses reversibly back to the C,, crystalline structure,

    whereas the amorphous &,-derived material

    decompresses to a more dense diamond-like form.

    Such behavior illustrates once again how, for carbon,

    the initial state and the reaction paths are important

    factors in what the end product will be.

    5. THEORETICAL STUDIES

    In addition to the new results coming from

    improved experimental methods, the large number of

    increasingly accurate theoretical studies are providing

    important new insights into the stability and physical

    properties of known and predicted carbon poly-

    morphs [94- 1041.

  • 8/11/2019 Art Diamante

    12/13

    152

    F. P.

    BUNDY et (11.

    6. CONCLUSIONS

    Carbon, the lightest of the Group IV elements, is

    certainly one of the most versatile, interesting and

    useful of the chemical elements in respect of the

    materials properties of its various phases and forms.

    Although many new things have been learned during

    the past few decades about its physical forms and

    properties over a wide range of pressures and temper-

    atures there is still much to be established accurately

    and discrepancies to be resolved - possibly new

    forms and properties to be discovered that may be

    both technically useful and fundamentally important.

    A deeper understanding of this important system will

    surely emerge with the continued development of

    experimental methods to probe higher pressures and

    temperatures, with increasing accuracy, precision and

    sensitivity, and with improving theoretical methods.

    Acknowledgemmts~We thank John Badding for useful

    comments and discussion. Work performed at the

    Geophysical Laboratory and Cornell University was sup-

    ported by the National Science Foundation.

    1.

    3

    5:

    4.

    5.

    6.

    7.

    8.

    9.

    10.

    11.

    12.

    13.

    14.

    15.

    16.

    17.

    18.

    19.

    20.

    21.

    22.

    23.

    24.

    25.

    REFERENCES

    F. P. Bundy,

    Koninkl, Nederl. Aknd. wn Wetenschap-

    pen-Amsterdam-Proc. Series B 72, No. 5 1969).

    F. P. Bundy. J. Geophys. Res. 85, 6930 (1980).

    F. P. Bundy, In Solid Srute

    Physics under Pressure

    (Edited by S. Minomura), KTK Sci. Pub]., Tokyo

    (1985).

    F. P. Bundy, Phlsicn A156, 169

    (

    1989).

    P. Gustafson, Carbon 24, 169 (1986).

    D. A. Young, Phase Diagrams

    of the Elements

    niv. of

    Calif. Press (1991).

    S. Fahy, S. Ci. Louie and M. L. Cohen, Phys. Rev. B34,

    1191 (1986).

    B. V. Derjaguin. D. V. Fedoseev, V. P. Varnin and S. P.

    Vnukov,

    Nature

    269, 398 (1977).

    A. G. Whittaker, Science 200, 763

    (

    1978).

    P. P. K. Smith and P. R. Buseck.

    Science

    216. 984

    (1982).

    F. D. Rossini and R. S. Jessup, J. Nut.

    Bur. Stds.

    C21,

    491 (193s).

    0. 1. Leipunskii, usp.

    Khim 8,

    1519 (1939).

    R. Berman and F. Simon, Zeirs. Elektrochm. 59, 333

    1955).

    F. P. Bundy, H. P. Bovenkerk, H. M. Strong and R.

    H. Wentorf, Jr, J. Chem. Phys. 35, 383 (1961).

    C. S. Kennedy and G. C. Kennedy,

    J. Geophys. Rex

    81, 2467 (1976).

    F. P. Bundy, J.

    Chem.

    Phys. 38, 618 (1963).

    F. P. Bundy, J. Chem. Phys. 38, 631 (1963).

    M. Togaya, In Science

    and Technology ofNew

    Diamond,

    p.369. Tokyo, KTKiTSPC (1990).

    M. Togaya, S. Sugiyama and E. Minuhara. In High

    Pressure Sciewe und Technology, Proc. AIRAPT und

    APS Cortf., Colorado Springs, CO (1993). p. 255.

    A. Ludwig, Zuit. Elrktrochrm. 8, 273 (1902).

    M. J. Basset, J. Phys. Radium 10, 217 (1939).

    M. T. Jones, Nat. Carbon Res Lab. Report PRG

    (1958).

    G. R. Gathers, I. W. Shaner and D. A. Young, UCRL

    Rpt., 51644 (1974); Phys.

    Rev.

    L t. 33, 70 (1974).

    A. V. Kirillin, M. D. Kovalenko and M. A. Scheindlin,

    Pisma Zh. Eksp. Tear. Fiz. 40, 452 (1980).

    M. A. Scheindlin and V. N. Senchenko, Ser.

    Phys.

    Dokl.

    33, 142 (1988).

    26.

    27.

    28.

    29.

    30.

    31.

    32.

    33.

    34.

    35.

    36.

    37.

    38.

    39.

    40.

    41.

    42.

    43.

    44.

    45.

    46.

    47.

    48.

    49.

    50.

    51.

    52.

    53.

    54.

    55.

    J. Heremans, C. H. Olk, G. L. Eesly, J. Steinbeck and

    G. Dresselhaus,

    Phys. Rev.

    L&t. 60, 452 (1988).

    A. V. Baitin, A. A. Lebedev, S. V. Romanenko. V. N.

    Senchenko and M. A. Scheindlin,

    High Temp.

    High

    Pres.

    22, 157 (1990).

    N. S. Fateeva and L. F. Vereshchagin.

    JETP Letr. 13,

    157 1971).

    T. Venkatesan, D. C. Jacobson, J. M. Gibson, B. S.

    Elman, G. Braunstein, M. S. Dresselhaus and G. Dres-

    selhaus, Phys. Rec. Lat. 53, 360 (1984).

    J. Steinbeck. G. Braunstein, M. S. Dresselhaus, T.

    Venkatesan and D. C. Jacobson.

    J. Appl.

    Phys. 58,

    4374 (1985).

    A. M. Malvezzi, N. Bloembergen and C. Y. Huang.

    Phys.

    Rrw.

    Lert. 57, 146 (1986).

    D. H. Reitze, H. Ahn and M. C. Downer, Phys. Rev.

    B45, 2677 (1992).

    G. Galli. R. Martin, R. Car and M. Parrinello. Phvs.

    Rev. Lat. 63, 988 1989).

    C.

    Galli, R. M. Martin, R. Car and M. Parrinello.

    Science 250, 1547

    (

    1990).

    M. van Thiel and F. H. Ree,

    Int.

    J. Thermophy\. 10,

    227

    (

    1989).

    M. S. Weathers and W. A. Bassett, Phys. Chem. Minrr-

    a/s

    15, 105 (1987).

    W. A. Bassett and M. S. Weathers. High Pressurr S ,i-

    ewe and 7echnology. Proc. AIRAPT and APS Conf:,

    Colorado Springs, CO ( 1993) p. 65 1.

    F. P. Bundy. J. Charm. Phys. 41, 3809 (1964).

    J. W. Shaner, J. M. Brown, C. A. Swenson and R. G.

    McQueen, J. Phys. 45, 235 (1984).

    K. A. Goettel, H. K. Mao and P. M. Bell. Rev. Sci.

    Imtrum.56, 1420 (1985).

    M. T. Yin and M. L. Cohen, Phys. Rer. Lctr. 50,

    2006 (1983).

    R. Biswas. R. M. Martin, R. J. Needs and 0. H. Nielsen.

    Ph.r.s. Rw. B30, 3210 (1984).

    R. Biswas, R. M. Martin. R. J. Needs and 0. H. Nielsen.

    Phys. Rev. B35, 9559

    (

    1987).

    J. S. Gold, W. A. Bassett, M. S. Weathers and J. M.

    Bird, Science 225, 921

    (

    1984).

    D. Ugarte, Nature 359, 707 (1992).

    D. A. Young and R. Grover, In

    Shock Waves in Con-

    densed Mutter (Edited by Schmidt and Holmes), p. 13

    1.

    North-Holland, Amsterdam (1988).

    M. Ross,

    Noirturr

    92, 435 (1981).

    G. J. Piermarini, S. Block, J. F. Barnett and R. A.

    Forman,

    J. Appl. Phys.

    46, 2774 (1975).

    P. M. Bell, J. Xu and H. K. Mao, In

    Proc. 4th Amer.

    PI?xs. SM. 7~opical Conf. on

    hock

    Waws

    in Condensed

    Matter

    (Edited by Gupta), p. 125. Plenum Pub]., NY

    [

    1986).

    M. Hanfland. J. Z. Hu, J. F. Shu, R. J. Hemley, H. K.

    Mao and Y. Wu, Bull. Amer. Phys. Sot. 35, 465

    (

    1990).

    A. F. Goncharov, Zh.

    Eksp. Tear. Fiz. 98,

    1824 (1990).

    T. Yagi, W. Utsumi, M. Yamakata, T. Kikegawa and

    0. Shimomura.

    Phys. Rru.

    B46, 6031 (1992).

    0. Mlshima. L. D. Calvert and E. Whalley,

    Nature

    310,

    393 (1984).

    R. J. Hemley, C. T. Prewitt and K. J. Kingma. In

    Reriews irl Mineralogy (Edited by P. J. Heaney, C. T.

    Prewitt and G. V. Gibbs), Vol. 29, p. 41. Mineralogical

    Society of America, Washington, DC

    (

    1994).

    Y. Zhang, F. Zhang and G. Chen,

    Curhon

    32, 1415

    .^.

    lYY4).

    56. F. R. Corrigan and F. P. Bundy,

    J. Chem. Phvs.

    63,

    3x12 (1975).

    57. R. B. Aust and H. G. Drickamer, Scipnc,r 140, Xl7

    (1963).

    5X. F. P. Bundy and J. S. Kasper,

    J. Chem.

    Phy.s. 46,

    2437 (1967).

    59. A. F. Goncharov, I. N. Makarenko and S. M. Stishov,

    Ser. Pkys. JETP 69, 380 1989).

  • 8/11/2019 Art Diamante

    13/13

    P,T phase and transformation diagram for carbon 153

    60. M. Hanfland, H. Beister and K. Syassen, Phys. Rev.

    B39, 2598 (1989)

    61. W. Utsumi, M. Yamakata, T. Yagi and 0. Shimomura,

    In Hi gh Pr essure Science and Technology, Proc.

    AIRAPT and APS Co Colorado Springs, CO

    (1993), p. 535.

    62. M. Kertesz and R. Hoffmann, J. Sol. State Chem. 54,

    313 (1984).

    63. W. Utsumi and T. Yagi, Science 252, 1542 (1991)

    64. P. S. DeCarli and J. C. Jamieson, Science 133, 182

    (1961).

    65. Netherlands Patent Release No. 6506395, Nov. 22,

    1965, E. I. duPont de Nemours and Company.

    66. L. F. Trueb, J.

    Appl.

    Phys. 39, 4707 (1968).

    67. V. I. Trefilov, G. I. Davvakin, V. V. Skorokhod and

    A. F. Khrienko, Sou. Phys. Dokl. 23, 269 (1978).

    68. N. Setaka and Y. Sekihawa, J. bZateria[s Sci. 16, 1728

    (1981).

    69. G. Tammann, Z. Phys. Chem. 69, 569 (1910).

    70. W. von Bolton, Z. Elektrochenz. 17, 971 (1911).

    71. 0. RulT, Z. Anorg. Allgem. Chem. 99, 73 (1917).

    72. B. V. Spitsyn and B. V. Derjaguin, USSR Patent, issued

    S/S/SO (filed 1956, issued 1980).

    73. R. A. Oriani and W. A. Rocco, Gen. Elec. Memo No.

    MA-36, Class IV (Aug. 1957).

    74. W. G. Eversole, U.S. Patent, issued 4/17/62 (filed 1958,

    issued 1962)

    75. J. C. Angus, H. A. Will and W. S. Stanko, J. App2.

    Phys. 39, 29 15 (1968).

    76. S. Matsumoto, Y. Sato, M. Tsutsumi and N. Setaka,

    J. Mater. Sci. 17, 3106 (1982).

    77. M. Kamo, Y. Sato, S. Matsumoto and N. Setaka,

    J. Cryst. Grow th 62, 642 (1983).

    78. B. V. Spitsyn, L. L. Bouilov and B. V. Derjaguin, Prog.

    Crysf. Growth and Charac. 17, 79 (1988).

    79. W. A. Yarbough, In Meto

    Di amond Science und Technof-

    ogy (Edited by R. Messier, J. T. Glass, J. E. Butler and

    R. Roy), p. 291. Materials Research Society, Pittsburgh,

    PA (1990)

    80. E. A. Rohlfing, D. M. Cox and A. Kaldor, J. Chem.

    Phys. 81, 3322 (1984).

    81. H. W. Kroto, J. R. Heath, S. C. OBrien, R. F. Curl

    and R. E. Smalley, Nature 318, 162 (1985)

    82.

    83.

    84.

    85.

    86.

    87.

    88.

    89.

    90.

    91.

    92.

    93.

    94.

    95.

    96.

    97.

    98.

    99.

    100.

    101.

    102.

    103.

    104.

    H. W. Kroto, Science 242, 1139 (1988).

    R. F. Curl and R. E. Smalley,

    Sci. Amer

    265,54 (1991).

    W. Kratchmer, D. D. Lamb, K. Fostiropoulos and D.

    R. Huffman, Nature 347. 354 (1990).

    D. R. Huffman, Phys. Today 44, 22(1991).

    R. F. Curl and R. E. Smalley, Science 242, 1017 (1988).

    S. J. Duclos, K. Brister, R. C. Haddon, A. R. Kortan

    and F. A. Thiel,

    Nature 351, 380

    (1991)

    G. A. Samara, J. E. Schirber, B. Morosin, L. V. Hansen,

    D. Loy and A. P. Sylwester, Phys. Rev. Lett. 67,

    3136, (1991)

    M. N. Regueiro, P. Monceau and J.-L. Hodeau, Nature

    355, 237 ( 1992).

    N. Chandrabhas, A. K. Sood, D. V. S. Muthu. C. S.

    Sundar, A. Bharathi, Y. Hariharan and C. N. R. Rao,

    Phvs.

    Rev.

    Lett. 73. 3411 (1994)

    F. -R. di Brozolo, T. E. Bunch, R. H. Fleming and

    J. Macklin,

    Nature 369, 37 (1994)

    M. Nunez-Regueiro, L. Marques, J. L. Hodeau, 0.

    Bethoux and M. Perroux, Phys. Rev. Lett. 74, 278

    (1995)

    J. A. Wolk, P. J. Horoyski and M. L. W. Thewalt, Phys.

    Rev. Let t. 74, 3483 (1995)

    A. K. McMahan, Phys. Rev. B30, 5835 (1984).

    S. Fahy and S. G. Louie, Phys. Rev. B36, 3373 (1987).

    R. L. Johnston and R. Hoffmann, J. Am. Chem. Sot .

    111, 810 (1989).

    A. Y. Liu, M. L. Cohen, K. C. Hass and M. A. Tamor,

    Phys. Rev. B43, 6742 (1991).

    C. Mailhiot and A. K. McMahan, Phys.

    Rev.

    B44,

    11578 (1991).

    A. Y. Liu and M. L. Cohen, Phys. Rev. B45, 4579

    (1992).

    Y. Bar-Yam, T. Lei, T. D. Moustakas, D. C. Allan and

    M. P. Teter, In MRS Symp. Proc. 242, W ide Band Gap

    Semi conductors (Edited by T. D. Moustakas, J. I. Pan-

    kove and Y. Hamakawajp. 335, (1992).

    M. Teter. Phvs. Rev. B48. 5031 (1993).

    K. Vogel and R. Nesper, J. Mol. Cat aL 82,467 (1993).

    J. Crain, S. J. Clark, G. J. Ackland, M. C. Payne, V.

    Milman, P. D. Hatton and B. J. Reid, Phys. Rev. B49,

    5329 (1994).

    J. Furthmuller, J. Hafner and G. Kresse, Phys. Rev.

    B50, 15606 (1994).


Recommended