Thermodynamics of the Am Phi Boles - Anthophyllite-Ferroanthophyllite and the Ortho-clino Phase Loop

download Thermodynamics of the Am Phi Boles - Anthophyllite-Ferroanthophyllite and the Ortho-clino Phase Loop

of 12

Transcript of Thermodynamics of the Am Phi Boles - Anthophyllite-Ferroanthophyllite and the Ortho-clino Phase Loop

  • 7/31/2019 Thermodynamics of the Am Phi Boles - Anthophyllite-Ferroanthophyllite and the Ortho-clino Phase Loop

    1/12

    American Mineralogist , Volume 86, pages 640651, 2001

    0003-004X/01/0506640$05.00 640

    INTRODUCTION

    Accurate representation of calculated phase boundaries for

    mineral polymorphs in P-T-X phase diagrams requires that

    molar Gibbs free-energy differences be known with high pre-

    cision, preferably better than 100 or 200 J/mol. This degree

    of precision is seldom attained in calorimetric measurements

    of heats of solution or experimental bracketing of mineral re-

    actions. Thus, constraints of the usual kinds on the free ener-

    gies of polymorphs from mutually independent sources will

    generally fail to represent their P-T-Xrelations correctly. Di-rect experimental reversal of polymorphic reaction boundaries

    quickly becomes difficult if not totally impossible as tempera-

    ture falls much below 1000 C, and extrapolation from high-

    temperature conditions is not always possible. In these

    circumstances, the occurrence and compositions of polymor-

    phs in well-defined natural situations may provide the best in-

    formation with regard to their relative stabilities. This seems

    to be the case for the phase relations of the orthorhombic and

    monoclinic ferromagnesian amphiboles.

    * E-mail: [email protected]

    Thermodynamics of the amphiboles: Anthophyllite-ferroanthophyllite and the ortho-clino

    phase loop

    BERNARD W. EVANS,*,1 MARK S. GHIORSO,1 HEXIONG YANG,2AND OLAF MEDENBACH3

    1Department of Geological Sciences, Box 351310, University of Washington, Seattle, Washington 98195-1310, U.S. A.2Geophysical Laboratory, Carnegie Institution of Washington, 5251 Broad Branch Road, N.W., Washington, D.C. 20015-1305, U.S.A.

    3Institut fr Mineralogie, Ruhr-Universitt Bochum, D-44780 Bochum, Germany

    ABSTRACT

    Ten new single-crystal X-ray structure refinements of unheated and heat-treated anthophyllite,

    new measurements of the optical indicatrix of anthophyllite, and previously published data from

    Mssbauer spectroscopy of heated anthophyllite, show that temperature-dependent long-range or-

    der of Fe2+ and Mg on the M-sites of cummingtonite-grunerite and anthophyllite may be considered

    identical for the purpose of thermodynamic modeling. The difference in solution properties be-

    tween the monoclinic and orthorhombic series, as expressed in the composition (XFe) dependence of

    lnKD in natural amphibole pairs, is accomodated through adjustment of an enthalpic term that is

    independent of order-disorder.End-member thermodynamic properties of cummingtonite and ferroanthophyllite are derived

    from those already known for anthophyllite and grunerite respectively, using intercrystalline KDdata and a fit of the T-XFephase loop to two critical field constraints: middle amphibolite-facies

    amphibolites and upper amphibolite-facies metaperidotites. Amphibolites suggest a transition tem-

    perature in the system FMSH at 555 C andXFe 0.3, whereas metaperidotites suggest a transition

    temperature of 650 C at XFe 0.1. LnKD for Fe-Mg exchange between cummingtonite and

    anthophyllite passes through zero atXFe 0.7, and as a result the T-XFe phase loop shows a minimum

    at this composition.

    Extrapolated end-member transition temperatures are estimated to be 800 C (Mg) and 450

    C (Fe). At its breakdown to enstatite + quartz + H2O (790 C at 5 kbar), anthophyllite is marginally

    stable with respect to end-member cummingtonite, and the addition of Ca renders the breakdown

    reaction metastable. A stability field is possible for end-member ferroanthophyllite. Cummingtonite-

    anthophyllite phase relations mirror those of the analogous clino- and orthopyroxene.

    Compositionally the simplest of all amphibole groups, the

    ferromagnesian amphiboles [(Mg,Fe)7Si8O22(OH)2] occur in a

    broad range of metamorphic rock types in the form of mono-

    clinic cummingtonite-grunerite (predominantly C2/m, but P21/

    m in low-temperature cummingtonite) and orthorhombic

    anthophyllite (Pnma). In this paper we attempt a quantitative

    assessment of their mutual stabilities, using what we already

    know about the thermodynamics of the cummingtonite-

    grunerite series (Ghiorso et al. 1995; Evans and Ghiorso 1995).

    Significant variables are pressure, temperature, and composi-tion [mole-fraction Fe/(Fe + Mg) or XFe]. Our goal is a set of

    thermodynamic properties for anthophyllite-ferroanthophyllite

    solutions, and an isobaric TXFe phase diagram, the phase loop,

    for the dimorphs. Aluminous orthoamphiboles (Al-anthophyllite

    and gedrite) will be considered in a later contribution.

    Some preliminary reasoning allows us to conclude that both

    ends (Fe and Mg) of the phase loop possibly occur at tempera-

    tures found in the Earths crust. A compilation of Fe/Mg

    intercrystalline partition data from natural parageneses, most

    of which probably equilibrated between 500 and 650 C, shows

    that lnKD for the exchange reaction:

  • 7/31/2019 Thermodynamics of the Am Phi Boles - Anthophyllite-Ferroanthophyllite and the Ortho-clino Phase Loop

    2/12

    EVANS ET AL.: THERMODYNAMICS OF ANTHOPHYLLITE-FERROANTHOPHYLLITE 641

    grunerite + anthophyllite = cummingtonite + ferro-anthophyllite

    Fe7Si8O22(OH)2 Mg7Si8O22(OH)2 Mg7Si8O22(OH)2 Fe7Si8O22(OH)2

    (1)

    falls in the range 0.125 to + 0.025 for Al-free compositions

    (Evans and Ghiorso 1995, Fig. 3). Assuming some cancella-

    tion of non-ideal solution properties between ortho- and clino-amphibole, this range translates into a standard-state Gibbs

    free-energy of exchange somewhere between +1.0 and 0.2

    kJ/M-cation, perhaps near the average value of 0.4 kJ/M-cat-

    ion. Equation 1 may be rearranged:

    anthophyllite cummingtonite = ferro-anthophyllite gruneriteMg7Si8O22(OH)2 Mg7Si8O22(OH)2 Fe7Si8O22(OH)2 Fe7Si8O22(OH)2

    (2)

    to show that the same small standard reaction free-energy is

    also a measure of the P or Tdifference between the end-mem-

    ber ortho-to-clino reactions. For example, when the Fe end-

    member transition (RHS of 2) is at equilibrium, the Mg

    end-member transition will be out of equilibrium by an iso-

    baric temperature increment Tgiven by:

    T= G1/Str (2800)/(4 to 5)

    adopting the average free-energy change and assuming that the

    transition entropies Str (ortho to clino) are temperature inde-

    pendent and both 4 to 5 J/K.mol (Ghiorso et al. 1995, Table 7;

    Holland and Powell 1998, Table 5). This back-of-the-envelope

    calculation suggests that the difference in temperature of the

    ortho-to-clino transitions between the Fe and Mg end-mem-

    bers is possibly on the order of 500700 C. Given the pub-

    lished record of coexisting cummingtonite and anthophyllite

    in nature, we have reason therefore to expect that one of the

    end-member transitions, if not both, may be accessible under

    the P and Tconditions found in the crust. The difference in the

    Mg and Fe end-member transition temperatures for the analo-

    gous, Ca-free pyroxenes (Pbca and C2/c) is just under 600 C

    at 1 bar (Lindsley 1983; Sack and Ghiorso 1994).

    The end-member Gibbs free energies of anthophyllite-Pnma

    and grunerite-C2/m are known from bracketing experiments,

    and have been incorporated into thermodynamic databases (e.g.,

    Berman 1988; Robie and Hemingway 1995; Gottschalk 1997;

    Holland and Powell 1998; Chernosky et al. 1998). The solu-

    tion properties of the cummingtonite-grunerite series have been

    evaluated from data on temperature-dependent, long-range Mg-

    Fe order-disorder and experimental equilibria with

    orthopyroxene, quartz, and H2O (Ghiorso et al. 1995). In this

    paper, we use new single-crystal X-ray structure refinements

    of anthophyllite, measurements of its optical indicatrix, and

    previously published Mssbauer spectra, to compare long-range

    Mg-Fe order in the orthorhombic series with that of the mono-

    clinic cummingtonite-grunerite series. If the temperature-de-

    pendent Mg-Fe order is the same in the two series, it greatly

    simplifies the problem.

    Anthophyllite exhibits long-range, non-convergent order of

    Fe2+ and Mg over the four sites M1, M2, M3, and M4. The

    strong preference of Fe2+ for the M4 site in anthophyllite has

    long been known from X-ray structure refinements and

    Mssbauer and infrared spectra, as reviewed in Deer et al.

    (1997). The temperature dependence and kinetics of the order-

    disorder process were studied by Seifert and Virgo (1975) and

    Seifert (1977, 1978). The temperature and composition depen-

    dence of order-disorder of Fe2+ and Mg on the M-sites of

    cummingtonite-grunerite are also known (Hirschmann et al.1994). We shall show below that systematic differences in the

    degree of Mg-Fe2+ order-disorder between the two series are

    not detectable by the methods currently available. This means

    that the real, small difference in solution properties between

    the two series, evident from the compositional (XFe) dependence

    of lnKD, the intercrystalline partition coefficient of Mg and Fe

    (Evans and Ghiorso 1995), can be accommodated by minor

    adjustment of one energy parameter in our solution model that

    is not a function of ordering state.

    SINGLE-CRYSTALREFINEMENTSOFNATURALANDHEAT-TREATEDANTHOPHYLLITE

    X-ray single-crystal refinements were conducted on two

    samples of anthophyllite (7.3.71.10, withXFe = 0.11, and W82-

    009, withXFe = 0.22), both unheated and after separate hydro-

    thermal heat-treatment at 600, 700, and 800 C followed by

    rapid quenching. Anthophyllite sample 7.3.71.10 is the same

    material as studied calorimetrically by Krupka et al. (1985).

    An X-ray diffraction (XRD) analysis of a third sample (BM

    93327, with XFe = 0.25) was carried out for us by F.C.

    Hawthorne, before and after heat-treatment at 700 C. The heat-

    treatments were done at 2 kilobars H2O pressure on the C-CH4buffer, for one month (600 C), two days (700 C), and 3 hours

    (800 C). Microprobe analyses of the crystals studied by XRD

    methods are set down in Table 1. The analyses were done on a

    fully automated JEOL 733 Superprobe at the University of

    Washington, Seattle, using our library of natural mineral stan-

    dards, 15 kV accelerating potential, integration times of 10 to

    40 seconds, and the correction factors of Armstrong (1988).

    Formula contents (Table 1) are calculated on the basis of an

    anion charge of 46 per formula unit (pfu) and all Fe as Fe2+.

    This standard procedure tends to propagate error particularly

    in the Si determination to the assignment of all other cations to

    the C, B, and A groups in the amphibole formula. However,

    our average calculated ratio VIAl/Al is 0.38, which we sug-

    gest is very reasonable given that this ratio is 0.43 in the ideal

    gedrite end-member (Robinson et al. 1971), a composition be-

    lieved appropriate for Mg-rich samples (Spear 1980). Of all

    the formula assignments, the least reliable is that of Na on the

    B and A sites.

    The assumption of exclusively Fe2+ in Mg-rich members of

    the anthophyllite series is supported by Mssbauer absorption

    spectra, in which Fe3+ is typically barely detectable (Bancroft

    et al. 1966; Barabanov and Tomilov 1973; Seifert 1978; Stroink

    et al. 1980; Law 1989; Ferrow and Ripa 1990). The spectrum

    of only one of four anthophyllites studied by Law (1989), his

    most Fe-rich sample, with 2.31 atoms pfu Fe, had resolvable

    Fe3+ (Fe3+/Fe = 0.024).

    Single-crystal X-ray refinements were done at the Univer-

    sity of Washington, Seattle, the Geophysical Laboratory (GL),

    Washington, D.C., and the University of Manitoba, Winnepeg

  • 7/31/2019 Thermodynamics of the Am Phi Boles - Anthophyllite-Ferroanthophyllite and the Ortho-clino Phase Loop

    3/12

    EVANS ET AL.: THERMODYNAMICS OF ANTHOPHYLLITE-FERROANTHOPHYLLITE642

    (UM). At the University of Washington, we used a Huber 512

    four-circle diffractometer with graphite monochromatized

    MoK radiation. XRD intensity data from one quadrant of re-

    ciprocal space up to 2 = 65 were collected in the -scan

    mode at a scan speed of 3 per minute. Three standard reflec-

    tions were checked after every 97 reflections; no systematic or

    significant variation in the intensities of the standard reflec-

    tions was observed. All XRD intensity data were corrected for

    Lorentz and polarization effects, and for absorption by the semi-

    empirical method of North et al. (1968). Only reflections hav-

    ing intensities 2I were considered as observed and included

    in the refinements, where I is the standard deviation deter-

    mined from the counting statistics. The laboratories at GL and

    UM both used a Bruker SMART CCD (charge-coupled device)

    X-ray diffractometer with graphite monochromatized MoK

    radiation, but data collection techniques varied slightly: frame

    widths 0.3 at GL and 0.2 at UM, frame time 30 s at GL and

    60 s at UM, a hemisphere of three-dimensional X-ray data col-

    lected up to 60 2 at GL and a sphere of data to 60 2 at UM.

    Unit-cell parameters were refined on the basis of all integrated

    reflections (>10), using a least-squares technique (GL and

    UM). The three-dimensional data were reduced and corrected

    for Lorentz, polarization, and background effects using the

    Bruker program SAINT (GL and UM). An empirical correc-

    tion for X-ray absorption was made (GL and UM) using the

    program SADABS (G. Sheldrick, unpublished computer pro-

    gram). Equivalent reflections were merged for the final refine-

    ments (GL).

    Preparatory to structure refinement, the following assump-

    tions were made for the assignment of atoms among the crys-

    tallographically distinct sites: (1) IVAl = 8.0 Si pfu with allIVAl ordered on T1; (2) VIAl (= Altotal

    IVAl), Cr, and Ti were

    confined to M2; (3) Ca and BNa were confined to M4; (4) ANa

    and K when present were confined to the A-site; and (5) the

    fractionation of Fe + Mn among the four M cation sites was

    determined from the refinements. After structure refinement,

    Mn was proportioned among the M-sites in the same manner

    as Fe.

    The results of our ten single-crystal X-ray structure refine-

    ments are given in Tables 2 and 3. Site-preferences for Fe2+ are

    M4 >> M1, M3 > M2; they are smaller at higher temperatures

    of equilibration. Preferences can be compared with those of

    similarly equilibrated cummingtonites and grunerites

    (Hirschmann et al. 1994) with the aid of plots of RTlnKD, the

    negative ideal ordering energies - Gid, against macroscopic

    XFe (Fig. 1). Because the energy of exchange of Mg and Fe2+

    between the M1 and M3 sites is small, on the order of 1 kJ/M-

    cation (Fig. 1A), we can combine the occupancies of the M1

    and M3 sites for comparison with M2 and M4 according to a

    3-site model (Fig. 1B). Finally, by using a 2-site model that

    TABLE 1. Chemical analyses of refined crystals of anthophyllite with formulae

    Sample no. 7.3.71.10 W82-009 BM 93327Lab. no. unheated C44 C49 C37 unheated C57a C25 C38 unheated C93SiO2 58.9 59.4 58.8 59.1 56.7 56.8 57.1 56.8 56.18 56.30TiO2 0.01 0.00 0.01 0.01 0.03 0.02 0.02 0.03 0.07 0.05Al2O3 0.73 0.51 1.10 0.95 1.09 1.12 1.13 1.09 2.32 2.25Cr2O3 0.15 0.03 0.13 0.04 0.19 0.19 0.17 0.18 n.d. n.d.FeO* 6.26 6.71 6.57 6.74 12.50 12.62 12.20 12.65 14.23 14.03

    MgO 30.63 30.44 30.37 30.56 25.55 25.71 26.02 25.94 23.96 23.94MnO 0.12 0.15 0.15 0.14 0.35 0.35 0.35 0.35 0.42 0.40NiO 0.14 0.11 0.11 0.12 0.10 0.10 0.10 0.10 n.d. n.d.CaO 0.48 0.54 0.54 0.57 0.65 0.61 0.71 0.62 0.62 0.66Na2O 0.08 0.06 0.10 0.10 0.11 0.12 0.11 0.11 0.16 0.17K2O 0.00 0.00 0.00 0.00 0.01 0.00 0.00 0.00 0.00 0.00F 0.00 0.00 0.00 0.00 0.03 0.00 0.00 0.02 0.00 0.00Cl 0.00 0.00 0.00 0.00 0.01 0.00 0.01 0.00 n.d. n.d.H2O 2.22 2.24 2.23 2.24 2.15 2.17 2.18 2.16 2.16 2.15less O=F,Cl 0.00 0.00 0.00 0.00 0.02 0.00 0.00 0.01 0.00 0.00Total 99.72 100.19 100.11 100.57 99.45 99.83 100.10 100.04 100.12 99.95

    Atomic proportions based on anion charge of 46Si 7.923 7.962 7.892 7.900 7.888 7.879 7.880 7.861 7.814 7.834IVAl 0.077 0.038 0.108 0.100 0.112 0.121 0.120 0.139 0.186 0.166 T 8.000 8.000 8.000 8.000 8.000 8.000 8.000 8.000 8.000 8.000VIAl 0.039 0.043 0.065 0.050 0.067 0.062 0.064 0.039 0.194 0.203Ti 0.001 0.000 0.001 0.001 0.003 0.002 0.002 0.003 0.007 0.005

    Cr 0.016 0.003 0.014 0.004 0.021 0.021 0.019 0.020 0.000 0.000Fe 0.704 0.752 0.737 0.753 1.454 1.463 1.408 1.464 1.655 1.633Mg 6.142 6.083 6.076 6.090 5.299 5.311 5.353 5.352 4.968 4.966Mn 0.014 0.017 0.017 0.016 0.041 0.041 0.041 0.041 0.049 0.047Ni 0.015 0.012 0.012 0.013 0.011 0.011 0.011 0.011 0.000 0.000Ca 0.069 0.078 0.078 0.082 0.097 0.091 0.105 0.092 0.092 0.098BNa 0.000 0.012 0.000 0.000 0.007 0.000 0.000 0.000 0.035 0.046 (B + C) 7.000 7.000 7.000 7.009 7.000 7.002 7.003 7.022 7.000 6.998ANa 0.021 0.004 0.026 0.026 0.023 0.031 0.029 0.030 0.008 0.000K 0.000 0.000 0.000 0.000 0.002 0.000 0.000 0.000 0.000 0.000 A 0.021 0.004 0.026 0.026 0.025 0.031 0.029 0.030 0.008 0.000F 0.000 0.000 0.000 0.000 0.013 0.000 0.000 0.009 0.000 0.000Cl 0.000 0.000 0.000 0.000 0.002 0.000 0.002 0.000 0.000 0.000OH 1.992 2.003 1.996 1.997 1.995 2.002 2.007 1.994 2.004 2.000Note: n.d. = not determined.* Total iron. Calculated H2O.

  • 7/31/2019 Thermodynamics of the Am Phi Boles - Anthophyllite-Ferroanthophyllite and the Ortho-clino Phase Loop

    4/12

    EVANS ET AL.: THERMODYNAMICS OF ANTHOPHYLLITE-FERROANTHOPHYLLITE 643

    averages the occupancies of the M1, M2, and M3 sites, we can

    incorporate into our comparison (Fig. 1C) the Mssbauer data

    for heat-treated anthophyllite obtained by Seifert (1978).

    In cummingtonite (but not in grunerite), there is a small

    calculated temperature dependence of RTlnKD (see Fig. 3 in

    Ghiorso et al. 1995), reflective of excess entropy related to or-

    der-disorder. The precision and accuracy of site occupancies

    measured by XRD diffraction, in both cummingtonite

    (Hirschmann et al. 1994) and anthophyllite (this work), are not

    sufficient to show this dependence over the range 600 to 750

    C, although a decrease in RTlnKD for the 2-site model showsclearly in the Mssbauer data of Seifert (1978). More impor-

    tant in the present context is that, allowing for 2 and all other

    possible uncertainties, Figures 1A, 1B, and 1C show no sys-

    tematic differences in Fe-Mg order between anthophyllite and

    cummingtonite-grunerite, at least in the range ofXFe of the

    anthophyllites studied by XRD and Mssbauer spectroscopy

    (0.11 to 0.25). Except for exchange between the M1 and M3

    sites, cummingtonite and anthophyllite in fact seem to share

    small changes in Gid as the Mg end-members are approached.

    This similarity may reflect energetic consequences of a com-

    positional change that in cummingtonite eventually induces a

    change from C2/m to P21/m symmetry.

    Only rarely is anthophyllite encountered in nature more Fe-

    rich thanXFe 0.30 without it containing significant amounts

    of Al. More Fe-rich natural anthophyllite trends composition-

    ally through Al-anthophyllite toward gedrite, although a broad

    solvus gap below about 600 C separates the two (Robinson et

    al. 1971; Spear 1980). X-ray and Mssbauer data (Papike and

    Ross 1970; Seifert 1978) show that the uptake of Al in

    orthoamphibole is accompanied by increasing Fe2+-Mg disor-

    der. Thus, a direct comparison of order-disorder in the simple

    system MFSH cannot be taken to more Fe-rich compositions,

    unless we resort to the use of synthetic anthophyllite. A rangeof intermediate Fe, Mg-compositions with orthorhombic sym-

    metry have been synthesized in the system FMSH (Cameron

    1975; Popp et al. 1976), but their space group symmetry was

    not determined unambiguously; they are not anthophyllite-

    Pnma. This material might be worthy of study in the future.

    THEOPTICALINDICATRIXOFANTHOPHYLLITEANDCUMMINGTONITE-GRUNERITE

    The optical properties of anthophyllite and cummingtonite

    provide supporting evidence for their indistinguishable intrasite

    exchange energies. Evans and Medenbach (1997) showed that

    the shape of the optical indicatrix (2V) of cummingtonite-

    TABLE 2. Cell dimensions and refinement statistics of anthophyllite crystals

    Sample T(C) a() b() c() V(3) Refls. R Rw Note7.3.71.10. R.T 18.5219(9) 17.9740(8) 5.2725(4) 1755.3 1779 0.036 0.031 GLC44 600 18.5008(30) 17.9695(22) 5.2738(6) 1753.3 1971 0.034 0.044 UWC49 700 18.5004(25) 17.9646(29) 5.2710(7) 1751.8 1343 0.033 0.043 UWC37 800 18.4908(36) 17.9785(24) 5.2743(6) 1753.4 2000 0.037 0.049 UW

    W82-009 R.T. 18.5705(9) 18.0361(8) 5.2876(2) 1771.0 2061 0.030 0.028 GL

    C57 600 18.5654(33) 18.0271(21) 5.2845(6) 1768.6 1941 0.036 0.045 UWC25 700 18.5760(10) 18.0367(10) 5.2847(3) 1770.6 1749 0.036 0.031 GLC38 800 18.5668(9) 18.0306(8) 5.2867(3) 1769.8 1979 0.031 0.028 GL

    BM 93327 R.T 18.5716(9) 18.0307(8) 5.2903(2) 1771.5 1401 0.029 0.023 UMC93 700 18.5711(6) 18.0256(6) 5.2885(2) 1770.4 1794 0.029 0.024 UMNote: GL, UW, UM, X-ray data collected at the Geophysical laboratory, the University of Washington, and the University of Manitoba.

    TABLE 3. Site occupancies of anthophyllite

    7.3.71.10 C44 C49 C37 W82-009 C57 C25 C38 BM 93327 C93T(oC) 600 700 800 600 700 800 700M1 Fe 0.017(2) 0.049(3) 0.063(3) 0.065(3) 0.049(2) 0.114(3) 0.119(3) 0.132(2) 0.087 0.151(3)

    Mg 0.983 0.950 0.936 0.934 0.950 0.883 0.878 0.864 0.910 0.845Mn 0.000 0.001 0.001 0.001 0.001 0.003 0.003 0.004 0.003 0.004

    M2 Fe 0.005(2) 0.024(3) 0.022(3) 0.038(3) 0.014(2) 0.049(4) 0.059(2) 0.071(2) 0.019 0.078(4)Mg 0.967 0.952 0.938 0.934 0.940 0.908 0.897 0.896 0.880 0.815

    Al 0.020 0.022 0.033 0.025 0.034 0.031 0.032 0.020 0.095 0.100Mn 0.000 0.001 0.000 0.001 0.000 0.001 0.002 0.002 0.001 0.002Cr 0.008 0.001 0.007 0.002 0.011 0.010 0.009 0.010Ti 0.001 0.001 0.001 0.001 0.001 0.005 0.005

    M3 Fe 0.022(4) 0.032(5) 0.048(5) 0.046(5) 0.034(3) 0.093(5) 0.097(4) 0.115(3) 0.058 0.126(4)Mg 0.978 0.967 0.951 0.953 0.965 0.904 0.900 0.882 0.940 0.870Mn 0.000 0.001 0.001 0.001 0.001 0.003 0.003 0.003 0.002 0.004

    M4 Fe 0.328 0.291 0.266 0.258 0.672 0.528 0.483 0.474 0.704 0.505Mg 0.631 0.657 0.689 0.696 0.257 0.412 0.451 0.467 0.215 0.405Mn 0.006 0.007 0.006 0.005 0.019 0.015 0.014 0.013 0.021 0.015Ca 0.035 0.039 0.039 0.041 0.049 0.045 0.052 0.046 0.045 0.050Na 0.006 0.003 0.015 0.025

    A Na 0.021 0.004 0.026 0.026 0.023 0.031 0.029 0.030 0.008 0.000K 0.002

    Note: Assignment of Cr to the M3 as well as the M2 site (e.g., Fialips-Gudon et al. 2000) markedly improves the consistency of intrasite Gid valuesin samples C44, C49, and C37, where Cr is variable.

  • 7/31/2019 Thermodynamics of the Am Phi Boles - Anthophyllite-Ferroanthophyllite and the Ortho-clino Phase Loop

    5/12

    EVANS ET AL.: THERMODYNAMICS OF ANTHOPHYLLITE-FERROANTHOPHYLLITE644

    grunerite is a sensitive indicator of Fe-Mg order-disorder in a

    sample of known composition (XFe). Figure 2 incorporates new

    measurements made with a double spindle-stage (Table 4) and

    shows that, in the range XFe = 0.0 to 0.3, the 2Vz of natural

    anthophyllite is indistinguishable from that of natural

    cummingtonite. In this compositional range, the state of order

    for natural samples, as measured by s123, the 2-site order pa-

    rameter where s123 =XFeM4 average (XFeM1M2M3), changes from

    0.0 (end-member cummingtonite) to slightly more than 0.8 at

    XFe = 0.30 (Evans and Medenbach 1997). The coincidence in

    the behavior of 2Vz vs.XFe is interpreted as indicative of identi-

    cal order-disorder in slowly cooled anthophyllite and

    cummingtonite. The 2Vz of more Fe-rich natural anthophyllite

    is smaller than that of cummingtonite (Fig. 2), but the differ-

    ences may reflect the larger amounts of the gedrite component,

    as well perhaps as different cooling rates. Significantly,

    anthophyllite shows a minimum 2Vz at XFe 0.3, just like

    cummingonite, and for the same reason (Evans and Medenbach

    1997).

    We also measured the optical properties of two anthophyllite

    crystals after equilibration in the laboratory at high tempera-

    ture, followed by quenching (Table 4). The 2Vz of sample W82-

    009 increased from 83.7 to 93.1 when equilibrated at 600 C,

    and the 2Vz of sample BM 93327 increased from 80.7 to 92.0

    after treatment at 700 C. If the order-disorder behavior of

    anthophyllite and cummingtonite is assumed to be identical,

    then we can extract values of the two-site order parameter s123based on (1) the measured 2Vz and XFe of anthophyllite using

    the cummingtonite Equation 1 of Evans and Medenbach (1997);

    and (2) the isothermal fits to the cummingtonite ordering data

    shown in Figure 2 of Evans and Ghiorso (1995). For sample

    W82-009 at 600 C these values ofs123 are (1) 0.482 and (2)

    0.488, whereas the measured value of anthophyllite from Table

    3 is 0.477. All three values are in excellent agreement, sug-

    gesting that the assumptions made above are reasonable. Cor-

    responding values for sample BM 93327 equilibrated at 700

    C are (1) 0.481, (2) 0.486, and a measured value of 0.434. In

    this case, givenXFe and temperature, the 2Vz of anthophyllite is

    behaving exactly like cummingtonite. The smaller measured

    s123 of the crystal of BM 93327 is not fully understood; it may

    reflect the greater amount of gedrite substitution (Al = 0.4

    pfu in this sample), or possibly some interlaboratory differ-

    ences. This sample plots off the general trend for cummingtonite

    in Figure 1C.

    SOURCEOFSOLUTIONPROPERTYDIFFERENCES

    We are not able to compare the ordering behavior of more

    Fe-rich anthophyllite with that of compositionally similar mem-

    bers of the cummingtonite-grunerite series because low-Al

    orthoamphibole with XFe > 0.30 is extremely rare in nature.

    Such anthophyllite or ferroanthophyllite is largely unstable with

    respect to cummingtonite-grunerite gedrite (see below) in

    amphibolite-facies parageneses where it might occur; and, at

    lower temperatures that would favor anthophyllite rather than

    cummingtonite, both ferromagnesian amphiboles are unstable

    with respect to more hydrated and carbonated equivalents, such

    as minnesotaite, siderite or ankerite + quartz iron oxides.

    Accordingly, for modeling solution properties, we shall make

    the crystal-chemically reasonable assumption that the long-

    range Fe-Mg ordering behavior of more Fe-rich anthophyllite

    continues to be indistinguishable from its compositionally

    FIGURE 1. Comparison of RTlnKD vs. macroscopic XFe for site

    partitioning in heat-treated cummingtonite-grunerite and anthophyllite.

    A = four-site model, B = three-site model, C = two-site model. Filled

    circles, cummingtonite-grunerite (Hirschmann et al. 1994); open

    circles, anthophyllite (this work); triangles, anthophyllite by Mssbauer

    spectrometry (Seifert 1978). Error bars are 2 for instrumental

    uncertainty (in most cases smaller than the symbol).

    70

    80

    90

    100

    110

    120

    130

    0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

    (Fe+Mn)/(Fe+Mn+Mg)

    2Vz

    FIGURE 2. 2Vz of anthophyllite-ferroanthophyllite (Table 4)

    compared to metamorphic cummingtonite-grunerite (Evans and

    Medenbach 1997). Symbols as in Figure 1. Included are data for two

    samples of proto-ferroanthophyllite from Sueno et al. (1998).

  • 7/31/2019 Thermodynamics of the Am Phi Boles - Anthophyllite-Ferroanthophyllite and the Ortho-clino Phase Loop

    6/12

    EVANS ET AL.: THERMODYNAMICS OF ANTHOPHYLLITE-FERROANTHOPHYLLITE 645

    equivalent monoclinic analog.

    Except for M1 vs. M3, intersite (ideal) exchange energies

    in both the monoclinic and orthorhombic ferromagnesian am-

    phiboles are large in comparison to those related to the accu-

    racy and precision of measured site occupancies, and to those

    related to possible differences in order-disorder between

    anthophyllite and cummingtonite-grunerite. The energetic con-

    sequence of temperature dependent order-disorder plays a ma-

    jo r ro le in de te rm in in g th e so lu ti on pr ope rt ie s of

    cummingtonite-grunerite (see in Fig. 2 Ghiorso et al. 1995),

    and the same must also be true for the anthophyllite series.

    The two series are nevertheless not identical in their solu-

    tion properties. They differ macroscopically because it can be

    shown that intercrystalline Fe-Mg exchange between them is

    not independent of their Fe contents. A compilation of data for

    natural clino-ortho ferromagnesian amphibole pairs was fit to

    the expression: lnKD = 0.125 + 0.150XFe + 0.046XFe wt%

    Al2O3 (see in Fig. 3 Evans and Ghiorso 1995), where XFe and

    wt%Al2O3 refer to the orthoamphibole. This expression implies

    for Al-free pairs a difference of 0.15 in lnKD in going fromXFe= 0 toXFe = 1, and zero lnKD atXFe = 0.75. Others have shown

    that the zero, or extremum, composition is close toXFe = 0.4

    (e.g., Elliot-Meadows et al. 1999), but this value applies to

    cummingtonite coexisting with Al-anthophyllite.

    Although we cannot rule out small differences in order-dis-

    order behavior between the two series, the macroscopic differ-

    ence in solution properties is likely to be due, in our opinion,

    to one or more other crystal-structure factors. The dimorphs

    differ fundamentally in the manner of stacking parallel to c

    between adjacent tetrahedral layers (Gibbs 1969; Hawthorne 1983);

    the sequence in Pnma anthophyllite is (..++ ++ ..) whereas

    in C2/m and P21/m cummingtonite it is (..++++..). There are

    also important differences in bond lengths at the M4 site, and

    kinking of the tetrahedral chains (Boffa Ballaran et al. 2001).

    These considerations allow us to derive the solution prop-

    erties of anthophyllite-ferroanthophyllite from those of

    cummingtonite-grunerite in a straightforward and simple man-

    ner. We shall assume that all the solution parameters that are

    determined by order-disorder in cummingtonite-grunerite, as

    calibrated by single-crystal X-ray refinements of heat-treated

    crystals (Table 9 in Ghiorso et al. 1995), also apply to

    anthophyllite-ferroanthophyllite.Ghiorso et al. (1995, Eq. 5)

    expressed the vibrational molar Gibbs free energy of

    cummingtonite-grunerite solution in terms of a truncated sec-

    ond-order Taylor expansion in composition and ordering vari-

    ables. Only three coefficients in this expansion are independent

    of order-disorder. Two of them are determined from standard-

    state properties. The third coefficient, G

    *r,r, was determined

    for cummingtonite-grunerite from heterogeneous phase equi-

    librium constraints. G

    *r,r differs between the two series by an

    amount that is determined by the difference (0.15) in the lnKDfor intercrystalline exchange at the Mg and Fe extremes (Fig.

    3). The required adjustment in G

    *r,ris from 11.2 kJ/mol in

    cummingtonite-grunerite to 12.2 kJ/mol in anthophyllite-

    ferroanthophyllite. G

    *r,r is equivalent to W/4 where Wis an

    effective macroscopic regular solution parameter. The abso-

    lute values of lnKD (the y-axis of Fig. 3) depend on end-mem-

    ber standard-state properties.

    END-MEMBERTHERMODYNAMICPROPERTIES

    Our recommended thermodynamic end-member properties

    are listed in Table 5. For end-member anthophyllite, we use

    thermodynamic properties given in Ghiorso et al. (1995, Table

    7), which are virtually identical to those of Chernosky et al.

    (1998, Table 6), and consistent with other minerals in the

    Berman (1988) database. The heat capacity, expansivity, and

    compressibility of end-member ferro-anthophyllite are assumed

    to be identical to those of grunerite. The reaction entropy for

    the ortho clino transition of Fe7Si8O22(OH)2 is taken to be 5

    J/Kmol, as was done for Mg7Si8O22(OH)2 (Ghiorso et al. 1995),

    making the monoclinic structure the high-temperature form.

    Hirschmann et al. (1994) showed that, corrected Ca- and Mn-

    free, the cell volume of end-member anthophyllite, if at all, is

    smaller than that of Mg-cummingtonite by no more than 0.05%.

    The mean refractive indices (Table 2 and Evans and Medenbach

    1997) for the two series betweenXFe = 0.1 and 0.3 differ by a

    little more than 0.001, which, from the Gladstone and Dale

    relation (Jaffe 1988), translates into a density difference of

    0.15%. If we split the difference between these two estimates,

    we arrive at molar volumes that differ by 0.03 J/barmol for

    anthophyllite and cummingtonite respectively (Table 5). It is

    not known whether the same volume change holds up to the Fe

    TABLE 4. Optical properties of anthophyllite

    Cations per formula unitSample no. IVAl VIAl Fe Mn Mg Ca Na XFe nx ny nz 2VzUnheated31 0.000 0.038 0.025 0.062 6.730 0.052 0.056 0.013 1.6040 127.07.31.71.10 0.049 0.041 0.779 0.015 6.081 0.077 0.020 0.115 1.6120 1.6248 1.6336 103.896-34 0.033 0.024 0.899 0.036 5.959 0.077 0.017 0.136 1.6147 1.6266 1.6359 99.0121-73 0.000 0.007 1.156 0.029 5.731 0.054 0.004 0.171 1.6290 91.3

    W82-009 0.154 0.077 1.460 0.043 5.325 0.090 0.030 0.220 1.6263 1.6359 1.6475 83.7BM 93327 0.150 0.153 1.614 0.049 5.081 0.088 0.024 0.246 1.6292 1.6379 1.6507 80.76A9 0.503 0.439 2.282 0.110 3.986 0.119 0.194 0.375 1.6450 1.6510 1.6610 74.02H-348 0.571 0.480 3.277 0.081 3.076 0.048 0.167 0.522 1.6580 1.6670 1.6810 76.6

    HeatedW82-009(600) 0.093 0.091 1.457 0.038 5.301 0.097 0.033 0.220 1.6241 1.6359 1.6482 93.1BM 93327(700)0.228 0.182 1.657 0.048 4.948 0.135 0.059 0.256 1.6289 1.6401 1.6515 92.0Notes: XFe = (Fe + Mn)/(Fe + Mn + Mg).Sources of samples. 3-1: Klein (1968), 7.3.71.10: Krupka et al. (1985), 96-34 and 121-73: Fabris and Perseil (1971) (alsoSeifert 1978, samples AG1 and AG2), W82-009: Hirschmann et al. (1994), BM 93327: British Museum of Natural History,6A9: Robinson and Jaffe (1969), 2H-348: Guiraud et al. (1996).In parentheses: temperature of heat-treatment.

  • 7/31/2019 Thermodynamics of the Am Phi Boles - Anthophyllite-Ferroanthophyllite and the Ortho-clino Phase Loop

    7/12

    EVANS ET AL.: THERMODYNAMICS OF ANTHOPHYLLITE-FERROANTHOPHYLLITE646

    end-members, but, for simplicity, we shall assume that it does.

    These small volume uncertainties affect the pressure depen-

    dence of lnKD by a minor amount on the order of 0.001/kbar.

    Their effect on the temperature of the transition is small but

    not trivial, however, although a meaningful estimate of the dP/dT

    of the transition is not possible at the present time.

    Still to be evaluated are the enthalpies of the end-members

    ferro-anthophyllite and cummingtonite. The Gibbs free ener-

    gies of these end-members are extremely close to those of

    grunerite and anthophyllite, respectively, and, in the absence

    of any experimental constraints, the only way to evaluate the

    small differences involved is to analyze data from the field.

    Two kinds of field data can be used to constrain the free ener-

    gies of the solutions: (1) field information that serves to bracket

    the ortho = clino inversion temperature for ideally Al-free

    (quadrilateral) ferromagnesian amphiboles as a function ofXFe;

    and (2) compositional data that give robust estimates of lnKDfor equilibrium pairs of cummingtonite + low-Al anthophyllite

    for specific values ofXFe. These data in combination, together

    with the MgFe solution properties of the two phases derived

    above, and the remaining end-member properties (Table 5), will

    constrain the end-member enthalpies.

    FIELDCONSTRAINTS

    It has long been known that low-Al anthophyllite is a me-dium-grade metamorphic mineral typical of Mg-rich bulk com-

    positions (metamorphosed ultrabasic rocks), whereas

    cummingtonite and grunerite are characteristic of intermediate

    FeMg and Fe-rich compositions (low-Ca amphibolites, low-K

    pelites, uralitized gabbros, and metamorphosed iron-formation

    or BIF) equilibrated at metamorphic grades reaching into the

    granulite facies. Cummingtonite (XFe from 0.29 to 0.5) also

    occurs as phenocrysts in H2O-rich silicic volcanic rocks and as

    a possible magmatic mineral in a small percentage of plutonic

    rocks. With slow cooling, cummingtonite inverts to

    anthophyllite and liberates a small quantity of Ca-amphibole

    (Ross et al. 1969; Evans et al. 1974; Carpenter 1982). These

    observations, and the sign of the partition coefficient lnKD,

    imply that the transition temperature, at least at first, declines

    with increasing XFe. The most useful constraints on the T-XFelocation of the transition in the FMSH system are supplied by

    amphibolites and metaperidotites.

    In Figure 4 we have plotted total Al (apfu) against theXFefor microprobe-analyzed anthophyllite and cummingtonite from

    five well-documented examples of middle amphibolite-facies

    metamorphism of low-Ca amphibolites and low K-pelites (Stout

    1972; Spear 1980, 1982; Clark 1978; Early and Stout 1991;

    Schneidermann and Tracy 1991). Cummingtonite and

    anthophyllite commonly coexist in these samples and in some

    cases gedrite and/or Ca-amphibole occur as well. Metamor-

    phic temperatures estimated by these authors fall in the range

    535 to 575 C, and pressures from 3 to 6 kbar. Possible field

    boundaries for these conditions have been inserted on Figure

    4. Except for the base of the plot and the three-phase assem-

    blage anthophyllite-gedrite-cummingtonite (which is virtually

    coplanar in composition space), phase boundaries are some-

    what dependent on the compositions of coexisting minerals, as

    FIGURE 3. Computed correlation of lnKD andXFe of cummingtonite-

    grunerite for Mg/Fe partition between cummingtonite-grunerite and

    anthophyllite-ferroanthophyllite. KD = (Fe/Mg)Ath/(Fe/Mg)Cum

    TABLE 5. Internally consistent thermodynamic properties of end-members

    anthophyllite cummingtonite ferro-anthophyllite gruneriteMg7Si8O22(OH)2 Mg7Si8O22(OH)2 Fe7Si8O22(OH)2 Fe7Si8O22(OH)2

    H0

    f,Tr,Pr (kJ) 12073.132 12067.920 9627.015 9623.550S

    0Tr,Pr(J/K) 535.259 540.259 720.000 725.000

    V0

    Tr,Pr(J/bar) 26.31 26.34 27.81 27.84k0 1233.8 1347.83

    k1 102 71.3398 93.5691k2 105 221.638 202.285k3 107 233.394 303.919v1 106 1.1394 1.6703v2 1012 8.68919v3 106 28.105 28.400v4 1012 62.894

    Notes:C kk k k

    T,P

    o

    r= + + +

    0

    1 2

    2

    3

    3T T T

    (J/K)

    V

    Vv P P v P P v T T v T T

    T,P

    o

    T ,P

    o r r r r

    r r

    = + ( ) + ( ) + ( ) + ( )1 1 22

    3 4

    2

    (J/bar)

    H0

    f,Tr,Pr and S0

    Tr,Pr of anthophyllite and cummingtonite are given with greater precision than individually known in order to maintain internal consistency.

  • 7/31/2019 Thermodynamics of the Am Phi Boles - Anthophyllite-Ferroanthophyllite and the Ortho-clino Phase Loop

    8/12

    EVANS ET AL.: THERMODYNAMICS OF ANTHOPHYLLITE-FERROANTHOPHYLLITE 647

    illustrated, for example, by the reaction: Cum + Plag = Al-

    anthophyllite + Ca-amphibole (Robinson and Jaffe 1969; see

    also Spear 1982). Extrapolating to zero Al, the data points sug-

    gest a narrow gap between anthophyllite and cummingtonite,

    centered onXFe 0.30. From KD data we know the gap here is

    about 2 mol% wide. The field of (subsolvus) Al-anthophyllite

    terminates at these temperatures atXFe 0.40 (Fig. 4), beyond

    which it reacts to cummingtonite + gedrite (Stout 1972). The

    many examples of anthophyllite that plot in Figure 4 in the

    field of cummingtonite + gedrite (and the overlap of

    anthophyllite and cummingtonite compositions) are interpreted

    to represent late-stage growth of anthophyllite at lower than

    maximum metamorphic temperatures, by one (or more) of the

    following mechanisms; (1) grain-scale exsolution from gedrite;

    (2) from the reaction gedrite + cummingtonite anthophyllite;

    or (3) simple inversion from cummingtonite to anthophyllite.

    Spear (1982) noted in some of his samples that anthophyllite

    rims gedrite and was the last amphibole to grow. Also,

    cummingtonite surrounded by a rim of anthophyllite has been

    observed in many descriptions of amphibolites (Eskola 1914;

    Tilley 1937; Tella and Eade 1978; Elliott-Meadows et al. 1999).

    There are, in general, far too many reported examples of two

    or three coexisting ferromagnesian amphiboles for them all to

    be considered to represent equilibrium frozen at one set of

    metamorphic conditions; collectively they represent a very

    small portion of potential amphibole composition space

    (Robinson et al. 1982). Because anthophyllite can form from

    cummingtonite on cooling, Figure 4 should probably be used

    to extract a maximum XFe for the phase boundary under the

    average ofP-T conditions estimated for the five field areas.

    Roughly half the samples plotted in Figure 4 were saturated in

    Ca-amphibole. In these, the anthophyllites and cummingtonites

    contain 23.5 and 36% Ca-amphibole (actinolite or horn-

    blende) in solid solution, respectively.

    In various kinds of metaperidotites, low-Ca amphibolites,

    and metasomatic rocks, numerous analyses in the literature (e.g.,

    Rabbitt 1948; Deer et al. 1997) show that low-Al anthophyllite

    (

  • 7/31/2019 Thermodynamics of the Am Phi Boles - Anthophyllite-Ferroanthophyllite and the Ortho-clino Phase Loop

    9/12

    EVANS ET AL.: THERMODYNAMICS OF ANTHOPHYLLITE-FERROANTHOPHYLLITE648

    H2O and Tlc + Fo = 5 En + H2O. Although these results are not

    definitive with respect to the relative stability of cummingtonite

    and anthophyllite in the system MSH, they are not in conflict

    with our field-based boundary. One must bear in mind, how-

    ever, that the apparently orthorhombic Fe, Mg-amphiboles syn-

    thesized by Cameron (1975) and Popp et al. (1976) have

    compositions that place them (from the field evidence) clearly

    in the T-XFe stability field of the monoclinic form. An occur-

    rence of virtually end-member anthophyllite at the Wight mine,

    Balmat, New York (sample 31, Klein 1968, with 0.15% F thiswork), believed to have formed at a maximum of 675 25 C

    (E.J. Essene, personal communication), provides a low-tem-

    perature field bracket on the end-member transition tempera-

    ture.

    The constraint on the phase loop supplied by amphibolites

    (555 C, XFe 0.30) seems to agree better with the MFSH

    loop than the CMFSH loop, even though hornblende is present

    in many of the samples. An MSH transition set at 850 rather

    than 800 C would fit the amphibolite data better, but the

    metaperidotites less well. An overall flatter dT/dXFe slope for

    the phase-loop might reasonably fit both the metaperidotites

    and the amphibolites, but then the standard-state Gibbs free

    energy of the exchange reaction, as we found by trial and error,would be too small to satisfy the natural KD data, because it

    would shift the isotherms (Fig. 3) to more positive values of

    lnKD.

    Accepting our estimate of the transition temperature for the

    Ca-free, Mg end-members (800 C, 5 kbar), the remaining un-

    known, the Gibbs free energy of ferro-anthophyllite, can be

    obtained from the standard-state Gibbs free energy of the ex-

    change reaction (equivalent to the enthalpy of reaction), which

    is tightly constrained by the natural KD data, in particular, those

    from metaperidotites. This provides leverage on the dT/dXFe of

    the phase loop and the transition temperature at the Fe-end. A

    value of lnKD = 0.1 for coexisting anthophyllite and

    cummingtonite in metaperidotites (XFe = 0.10.2) may be read

    from the fit expression in Figure 3 of Evans and Ghiorso (1995),

    Figure 3 here. This value of lnKD is independently supported

    by a comparison of Fe-Mg partition in a large number of ferro-

    magnesian amphibole + olivine pairs in metaperidotites (Fig.

    6). Despite considerable scatter resulting mostly from a lack of

    perfect exchange equilibrium, Figure 6 shows unmistakablythat, for a given olivine composition, cummingtonite is slightly

    richer in Fe than anthophylliteby an amount corresponding

    to an average lnKD of 0.1. It should be noted that our enthalpy

    for end-member cummingtonite is 250 J/mol more negative

    than adopted before (Table 7 in Ghiorso et al. 1995).

    Our calculated binary two-phase loop for MFSH (Fig. 5)

    has a temperature minimum close toXFe = 0.7 and an Fe end-

    member transition temperature of 450 C, or about 35 C lower

    in the presence of ferroactinolite. This result is consistent with

    the common occurrence of cummingtonite-grunerite (ranging

    inXFe from 0.4 to 0.98) in metamorphosed iron-formation (Klein

    1982). The first occurrence of grunerite in the prograde meta-

    morphism of iron-formation is close to the biotite isograd (Klein

    1978; Haase 1982), that is, not much higher than 400 C. The

    only anthophyllite reported in metamorphosed iron-formation

    occurs in hematite-bearing rocks with high Mg/Fe2+ ratios

    (Klein 1972). A flatter dT/dXFe slope would be in poorer agree-

    ment with the widespread grunerite in BIF as well as the lnKDconstraints. In the Introduction, we guessed at a standard-state

    exchange reaction free-energy of 0.4 kJ/M-cation. Our recom-

    mended dataset (Table 5) has G1 = H1 = 1.75 kJ/mol or 0.25

    kJ/M-cation, equivalent to a 350C difference in transition tem-

    peratures from the Mg to the Fe end of the diagram.

    CONCLUDINGREMARKS

    We believe that our phase diagram (Fig. 5) successfully

    accounts for the occurrence in various rock types of low-Al

    anthophyllite, cummingtonite, and grunerite. In addition, the

    diagram opens up the possibility of a stability field at low tem-

    perature for pure ferro-anthophyllite. End-member

    Fe7Si8O22(OH)2, usually somewhat manganoan, forms in na-

    ture by cooling and hydration of fayalite, and has been described

    FIGURE 5. Calculated TXFe phase loop at 5 kbar for coexisting

    Al-free orthorhombic and monoclinic Fe, Mg-amphiboles (2 continuous

    lines), and a phase loop (3 dashed lines) for three coexisting Al-freeamphiboles in CMFSH. Arrows denote constraints discussed in the

    text.

    FIGURE 6. Mg/Fe partition between natural cummingtonite,

    anthophyllite, and olivine as a function of Fe/(Fe + Mg) in olivine.

    Symbols as in Figure 1. KD = (Fe/Mg)0l/(Fe/Mg)Amph.

  • 7/31/2019 Thermodynamics of the Am Phi Boles - Anthophyllite-Ferroanthophyllite and the Ortho-clino Phase Loop

    10/12

    EVANS ET AL.: THERMODYNAMICS OF ANTHOPHYLLITE-FERROANTHOPHYLLITE 649

    as grunerite (e.g., Bowen and Schairer 1935; Bonnichsen 1969;

    Floran and Papike 1978; Vaniman et al. 1980; Janeczek 1989).

    Recently, it has been shown that proto-ferroanthophyllite-Pnmn

    (Sueno et al. 1998) and ferro-anthophyllite-Pnma (Bozhilov

    and Evans 1999) are alternative possibilities. These observa-

    tions are not fully understood at present (and MnO may play a

    role in the occurrence of protoamphibole), but they certainlyhint at the possibility that an orthorhombic end-member is more

    stable than grunerite at low temperature, but not so low a tem-

    perature as to preclude growth of any Fe-amphibole. This lower

    limit is about 340 C, because below this temperature we can

    expect reaction of end-member Fe 7Si8O22(OH)2, whether

    grunerite or ferroanthophyllite, to minnesotaite in the presence

    of quartz and H2O (Miyano 1978; Rasmussen et al. 1998). Thus,

    Fe end-member, Al-free orthorhombic amphibole may have a

    stability field above 340 C and below 450 C. Probably the

    best place to look for ferro-anthophyllite is in high Fe/Mg hy-

    drothermal ore deposits; in low-grade metamorphosed BIF, the

    required temperature may be too low and, furthermore, grunerite

    will be favored in the presence of Ca.

    Our reference-state Gibbs free-energies of formation for

    monoclinic vs. orthorhombic FeMg-amphibole, not surpris-

    ingly, are quite similar at 298 K and 1 bar, with GAthCum =

    3.875 kJ/mol and GFathGru = 2.125 kJ/mol. When free ener-

    gies are extracted from different sources along independent

    paths, the differences, and the relative uncertainties, tend to be

    considerably larger in the case of polymorphs. For example, the

    corresponding reference-state transition free energies from Hol-

    land and Powell (1998, Table 5) are 10.61 and 12.63 kJ/mol re-

    spectively, and there is no stability field for either of the

    monoclinic end-members.

    None of the published papers on the subsolidus phase dia-

    gram for MSH have considered the possibility of a stability

    field for cummingtonite. Our data suggest that, close to its

    breakdown to enstatite, quartz, and H2O (790 C at 5 kbar), the

    stable amphibole may indeed be anthophyllite, but the uncer-

    tainties are such that it could be cummingtonite. Additional

    study of experimental run-products in the 800 C region in this

    system, by HRTEM for example, would certainly be helpful.

    In the presence of tremolite, the field of anthophyllite is suffi-

    ciently reduced (Fig. 5) that it is unlikely to reach that of

    enstatite, quartz, and H2O. In an attempt to check of our pro-

    posed phase loop, D. M. Jenkins (personal communication,

    October 1998) hydrothermally treated anthophyllite sample BM

    93327 (XFe = 0.25, 12.5 wt% Al2O3) at 5 kbar. He found no

    indication of conversion to cummingtonite at 700 C for 20

    days, at 800 C for 12 days, and at 850 C for 5 days (at which

    temperature partial breakdown to orthopyroxene took place).

    The free-energy drive for the inversion reaction is so small that

    direct experimental bracketing of the inversion curve will

    clearly be extremely difficult. On the other hand, crystallization

    experiments at high oxygen fugacity on hydrous, high Mg/Fe rhyo-

    lite and dacite melts might be able to overcome the kinetic prob-

    lem, in that ferromagnesian amphiboles grow readily from such

    melts (e.g., Nicholls et al. 1992; Rutherford and Devine 1996;

    Scaillet and Evans 1999).

    Figure 4 shows the stabilizing effect of Al on anthophyllite,

    but otherwise the diagram should only be regarded as a tenta-

    tive illustration of the mutual relations of cummingtonite,

    anthophyllite, and gedrite in the system NFMASH. For ex-

    ample, it would be interesting to know if field evidence can be

    brought to bear on the T-XFe behavior of the reaction terminal

    to anthophyllite: Al-Ath = Ged + Cum.

    The analogous inversion in Fe, Mg-pyroxene from

    orthopyroxene to clinopyroxene shows the same behavior asthe ferromagnesian amphiboles in the T-XFe section, with a

    minimum close to the Fe-end (Fig. 3 in Davidson et al. 1988,

    Fig. 7 in Sack and Ghiorso 1994). The 350 C difference in the

    transition temperature from the Mg to Fe end for Fe, Mg-am-

    phibole may be normalized to the average temperature (in K):

    350/898 = 0.390. This normalization yields a number that is

    essentially identical to the corresponding value for pyroxene:

    590/1478 = 0.399 (Fig. 7 in Sack and Ghiorso 1994). The cor-

    respondence may be understood by multiplying both numera-

    tor and denominator of this ratio by the appropriate averaged

    heat capacity of the phase, i.e. (T) (CP,ave) / (Tave ) (CP,ave), which

    may be written (H)/(Have). This demonstrates that the Tnor-

    malized to the average temperature of the transition is a proxy

    for the enthalpy difference between the Mg and Fe end-mem-

    bers of the series, rendered as an intensive quantity through

    division by the average enthalpy. The fact that structures and

    transition mechanisms are similar between amphibole and py-

    roxene suggests that normalized enthalpy differences between

    Mg and Fe end-members of both series should compare favor-

    ably. In fact, this calculation could have been used to predict

    the transition temperatures in the amphibole series from knowl-

    edge of those in the pyroxenes, and it could be argued that the

    favorable correspondence is an independent test of the obser-

    vations that led us to select the transition temperatures adopted

    for the Fe, Mg-amphiboles.

    ACKNOWLEDGMENTSWe thank J. Fabris, M. Guiraud, C. Klein, the Natural History Museum

    London, J.S. Schneidermann, and R.J. Tracy for the loan of samples, M. Schindlerand F.C. Hawthorne for the refinements of sample BM 93327, and D.M. Jenkinsfor high PTwork on this sample. Critical reviews by G. Droop, J.C . Schumacher,and V. Trommsdorff are greatly appreciated. This study was supported by theNational Science Foundation, grant EAR 97-06326.

    REFERENCESCITED

    Andrew, A.S. (1984) P-T-X(CO2) conditions in mafic and calc-silicate hornfelsesfrom Oberon, New South Wales, Australia. Journal of Metamorphic Geology,2, 143164.

    Armstrong, J.T. (1988) Bence-Albee after 20 years: review of the accuracy of -factor correction procedures for oxide and silicate minerals. Microbeam Analy-sis, Applications in Geology, 469476.

    Bancroft, G.M., Maddock, A.G., Burns, R.G., and Strens, R.G.J. (1966) Cation dis-

    tribution in anthophyllite from Mssbauer and infra-red spectroscopy. Nature,212, 913915.Barabanov, A.V. and Tomilov, S.B. (1973) Mssbauer study of the isomorphous

    series anthophyllite-gedrite and cummingtonite-grunerite. Geochemistry Inter-national, 12401267.

    Berman, R.G. (1988) Internally-consistent thermodynamic data for stoichiometricminerals in the system Na2O-K2O-CaO-MgO-FeO-Al 2O3-SiO2-TiO2-H2O-CO2.Journal of Petrology, 29, 445522.

    Boffa Ballaran, T., Carpenter, M.A., and Domeneghetti, M.C. (2001) Phase transi-tions and thermodynamic mixing behaviour of the cummingtonite-gruneritesolid solution. Physics and Chemistry of Minerals, in press.

    Bonnichsen, B. (1969) Metamorphic pyroxenes and amphiboles in the Biwabik IronFormation, Dunka River area, Minnesota. Mineralogical Society of AmericaSpecial Paper 2, 217239.

    Bowen, N.L. and Schairer, J.F. (1935) Grunerite from Rockport, Massachusetts,and a series of synthetic fluor-amphiboles. American Mineralogist, 20, 543551.

  • 7/31/2019 Thermodynamics of the Am Phi Boles - Anthophyllite-Ferroanthophyllite and the Ortho-clino Phase Loop

    11/12

    EVANS ET AL.: THERMODYNAMICS OF ANTHOPHYLLITE-FERROANTHOPHYLLITE650

    Bozhilov, K.N. and Evans, B.W. (1999) A HRTEM study of grunerite andferroanthophyllite from Rockport, Massachusetts. Transactions of the Ameri-can Geophysical Union, 80, no. 46, F1107 (abstract).

    Cameron, K.L. (1975) An experimental study of actinolite-cummingtonite phaserelations with notes on the synthesis of Fe-rich anthophyllite. American Miner-alogist, 60, 375390.

    Carpenter, M.A. (1982) Amphibole microstructures: some analogies with phase trans-formations in pyroxenes. Mineralogical Magazine, 46, 395397.

    Chernosky, J.V. and Autio, L.K. (1979) The stability of anthophyllite in the pres-

    ence of quartz. American mineralogist, 64, 294303.Chernosky, J.V., Day, H.W., and Caruso, L.J. (1985) Equilibria in the system MgO-

    SiO2-H2O: experimental determination of the stability of Mg-anthophyllite.American Mineralogist, 70, 223236.

    Chernosky, J.V., Berman, R.B., and Jenkins, D.M. (1998) The stability of tremolite:New experimental data and a thermodynamic assessment. American Mineralo-gist, 83, 726738.

    Clark, M.D. (1978) Amphibolitic rocks from the Precambrian of Grand Canyon:mineral chemistry and phase petrology. Mineralogical Magazine, 42, 199207.

    Davidson, P.M., Lindsley, D.H., and Carlson, W.D. (1988) Thermochemistry ofpyroxenes on the join Mg2Si2O6-CaMgSi2O6: A revision of the model for pres-sures up to 30 kbar. American Mineralogist, 73, 12641266.

    Deer, W.A., Howie, R.A, and Zussman, J. (1997) Rock-forming minerals, Volume2B, Double-chain silicates. The Geological Society, London, 764p.

    Droop, G.T.R. (1994) Triple-chain pyriboles in Lewisian ultramafic rocks. Mineral-ogical Magazine, 390, 120.

    Dymek, R.F., Brothers, S.C., and Schiffries, C.M. (1988) Petrogenesis of ultramaficmetamorphic rocks from the 3800 Ma Isua supracrustal belt, West Greenland.

    Journal of Petrology, 29, 13531397.Early, D. III and Stout, J.H. (1991) Cordierite-cummingtonite facies rocks from the

    Gold Brick District, Colorado. Journal of Petrology, 32, 11691202.Elliott-Meadows, S.R., Froese, E., and Appleyard, E.C. (1999) Cordierite-

    anthophyllite-cummingtonite rocks from the Lar deposit, Laurie Lake, Manitoba.Canadian Mineralogist, 37, 375380.

    Eskola, P. (1914) On the petrology of the Orijrvi region in southwestern Finland.Bulletin de la Commission Gologique de Finlande 40, 277.

    Evans, B.W. and Ghiorso, M.S. (1995) Thermodynamics and petrology ofcummingtonite. American Mineralogist, 80, 649-663.

    Evans, B.W. and Medenbach, O. (1997) The optical properties of cummingtoniteand their dependence on Fe-Mg order-disorder. European Journal of Mineral-ogy, 9, 9931003.

    Evans, B.W. and Trommsdorff, V. (1983) Fluorine hydroxyl titanian clinohumite inAlpine recrystallized garnet peridotite: compositional controls and petrologicsignificance. American Journal of Science, 283-A, 355369.

    Evans, B.W., Ghose, S., Rice, J.M., and Trommsdorff, V. (1974) Cummingtonite-anthophyllite phase transformation in metamorphosed ultramafic rocks, Ticino,

    Switzerland. Transactions of the American Geophysical Union, 55, 469.Fabris, J. and Persil, E.A. (1971) Nouvelles observations sur les amphiboles

    orthorhombiques. Bulletin Societ Franaise Mineralogie et Cristallographie,94, 385395.

    Ferrow, E. A. and Ripa, M. (1990) Al-poor and Al-rich orthoamphiboles: a Mssbauerspectroscopy and TEM study. Mineralogical Magazine, 54, 547552.

    Fialips-Gudon, C.-I., Robert, J.-L., and Delbove, F. (2000) Experimental study ofCr incorporation in pargasite. American Mineralogist, 85, 687693.

    Floran, R.J. and Papike, J.J. (1978) Mineralogy and petrology of the Gunflint ironformation, Minnesota-Ontario: Correlation of compositional and assemblagevariations at low to moderate grade. Journal of Petrology, 19, 215288.

    Frost, B.R. (1975) Contact metamorphism of serpentinite, chloritic blackwall, androdingite at Paddy-Go-Easy Pass, Central Cascades, Washington. Journal ofPetrology, 16, 272313.

    Ghiorso, M.S., Evans, B.W., Hirschmann, M., and Yang, H. (1995) Thermodynam-ics of the amphiboles: Fe-Mg cummingtonite solid solutions. American Miner-alogist, 80, 502519.

    Gibbs, G.V. (1969) Crystal structure of protoamphibole. Mineralogical Society of

    America Special Paper 2, 101109.Gole, M.J., Barnes, S.J., and Hill, R.E.T. (1987) The role of fluids in the metamor-

    phism of komatiites, Agnew nickel deposit, Western Australia. Contributions toMineralogy and Petrology, 96, 151162.

    Gottschalk, M. (1997) Internally consistent thermodynamic data for rock-formingminerals. European Journal of Mineralogy, 9, 175223.

    Greenwood, H.J. (1963) The synthesis and stability of anthophyllite. Journal ofPetrology, 4, 317351.

    Grond, R., Wahl, F., and Pfiffner, M. (1995) Mehrphasige alpine Deformation undMetamorphose in der nrdlichen Cima-Lunga Einheit, Zentralalpen (Schweiz).Schweizerische Mineralogische und Petrographische Mitteilungen, 75, 371386.

    Guiraud, M., Powell, R., and Cottin, J.-Y. (1996) Hydration of orthopyroxene-cordi-erite-bearing assemblages at Laouni, Central Hoggar, Algeria. Journal of Meta-morphic Geology, 14, 467476.

    Haase, C.S. (1982) Metamorphic petrology of the Negaunee Iron Formation,Marquette District, Michigan: mineralogy, metamorphic reactions, and phase

    equilibria. Economic Geology, 77, 6081.Hawthorne, F.C. (1983) The crystal chemistry of the amphiboles. The Canadian

    Mineralogist, 21, 173480.Hirschmann, M., Evans, B.W., and Yang, H. (1994) Composition and temperature

    dependence of Fe-Mg ordering in cummingtonite-grunerite as determined byX-ray diffraction. American Mineralogist, 79, 862877.

    Holland T.J.B. and Powell, R. (1998) An internally consistent thermodynamic dataset for phases of petrological interest. Journal of Metamorphic Geology, 16,309343.

    Jaffe, H.W. (1988) Introduction to crystal chemistry. Cambridge University Press,New York, 161 p.

    Janeczek, J. (1989) Manganoan fayalite and products of its alteration from theStrzegom pegmatites, Poland. Mineralogical Magazine, 53, 315325.

    Kamineni, D.C., Jackson, G.D., and Bonardi, M. (1979) Coexisting magnesian andcalcic amphiboles in meta-ultramafites from Baffin Island (Arctic Canada).Neues Jahrbuch fr Mineralogie, Monatsheft 12, 542555.

    Kisch, H.J. (1969) Magnesiocummingtonite-P21/m : A Ca- and Mn-poorclinoamphibole from New South Wales. Contributions to Mineralogy and Pe-trology, 21, 319331.

    Klein, C. (1968) Coexisting amphiboles. Journal of Petrology, 9, 281330.(1978) Regional metamorphism of Proterozoic iron-formation, Labrador

    Trough, Canada. American Mineralogist, 63, 898912.(1982) Amphiboles in iron-formations. In D.R. Veblen and P.H. Ribbe, Eds.,

    Amphiboles: petrology and experimental phase relations, vol. 9B, p. 8898.Reviews in Mineralogy Mineralogical Society of America, Washington, D.C.

    Kleinschmidt, G., Schubert, W., Olesch, M., and Rettmann, E.S. (1987) Ultramaficrocks of the Lanterman Range in North Victoria Land, Antarctica. Petrology,

    geochemistry, and geodynamic implications. Geologisches Jahrbuch, Hannover,B 66, 231273.

    Krupka, K.M., Hemingway, B.S., Robie, R.A., and Kerrick, D.M. (1985) High tem-perature heat capacities and derived thermodynamic properties of anthophyllite,diopside, enstatite, bronzite, and wollastonite. The American Mineralogist, 70,249260.

    Law, A.D. (1989) Studies of the orthoamphiboles. IV. Mssbauer spectra ofanthophyllites and gedrites. Mineralogical Magazine, 53, 181191.

    Lindsley, D.H. (1983) Pyroxene thermometry. American Mineralogist, 68, 477493.

    Matthes, S. (1986) Die prograde Kontaktmetamorphose von Serpentiniten desOberpflzer Waldes. Geologica Bavarica, 89, 720.

    Matthes, S. and Knauer, E. (1981) The phase petrology of the contact metamorphicserpentinites near Erbendorf, Oberpfalz, Bavaria. Neues Jahrbuch frMineralogie Abhandlungen, 141, 5989.

    Miyano, T. (1978) Phase relations in the system Fe-Mg-Si-O-H and environmentsduring low-grade metamorphism of some Precambrian iron formations. Jour-nal of the Geological Society of Japan, 84, 679690.

    Nicholls, I.A., Oba, T., and Conrad, W.K. (1992) The nature of primary rhyoliticmagmas involved in crustal evolution: evidence from an experimental study ofcummingtonite-bearing rhyolites. Geochimica et Cosmochimica Acta, 56, 955962.

    North, A.C.T, Phillips, D.C., and Matthews, F.S. (1968) A semi-empirical methodof absorption correction. Acta Crystallographica, A24, 351359.

    Papike, J.J. and Ross, M. (1970) Gedrites: crystal structures and intracrystallinecation distributions. American Mineralogist, 55, 19451972.

    Pfeifer, H.R. (1978) Hydrothermal Alpine metamorphism in metaperidotite rocksof the Cima Lunga zone, Valle Verzasca, Switzerland. SchweizerischeMineralogische und Petrographische Mitteilung, 58, 400405.

    (1987) A model for fluids in metamorphosed ultramafic rocks: IV. Metaso-matic veins in metaharzburgites of Cima di Gagnone, Valle Verzasca, Switzer-land. In H.C. Helgeson, Ed., Chemical Transport in Metasomatic Processes,NATO Advanced Study Institute Series, Series C, p. 591632. D Reidel Pub-lishing Company, Dordrecht.

    Popp, R.K., Gilbert, M.C., and Craig, J.R. (1976) Synthesis and X-ray properties ofFe-Mg orthoamphiboles. American Mineralogist, 61, 12671279.

    Rabbitt, J.C. (1948) A new study of the anthophyllite series. American Mineralo-gist, 33, 263323.

    Rasmussen, M.G., Evans, B.W., and Kuehner, S.M. (1998) Low-temperature fayalite,greenalite, and minnesotaite from the Overlook gold deposit, Washington: Phaserelations in the system FeO-SiO2-H2O. Canadian Mineralogist, 36, 147162.

    Rice, J.M., Evans, B.W., and Trommsdorff, V. (1974) Widespread occurrence ofmagnesiocummingtonite in ultramafic schists, Cima di Gagnone, Ticino, Swit-zerland. Contributions to Mineralogy and Petrology, 43, 245251.

    Robie, R.A. and Hemingway, B.S. (1995) Thermodynamic properties of mineralsand related substances at 298.15 K and 1 bar (105 pascals) pressure and at highertemperatures. U.S. Geological Survey Bulletin 2131, 461p.

    Robinson, P. and Jaffe, H.W. (1969) Chemographic exploration of amphibole as-semblages from central Massachusetts and southwestern New Hampshire. Min-eralogical Society of America, Special Paper 2, 251300.

    Robinson, P., Ross, M., and Jaffe, H. (1971) Composition of the anthophyllite-gedriteseries, comparisons of gedrite and hornblende, and the anthophyllite-gedritesolvus. American Mineralogist, 56, 10051041.

  • 7/31/2019 Thermodynamics of the Am Phi Boles - Anthophyllite-Ferroanthophyllite and the Ortho-clino Phase Loop

    12/12

    EVANS ET AL.: THERMODYNAMICS OF ANTHOPHYLLITE-FERROANTHOPHYLLITE 651

    Robinson, P., Spear, F.S., Schumacher, J.C., Laird, J., Klein, C., Evans, B.W., andDoolan, B.L. (1982) Phase relations of metamorphic amphiboles: natural oc-currence and theory. In D.R. Veblen and P.H. Ribbe, Eds., Amphiboles: petrol-ogy and experimental phase relations, vol. 9B, p. 1227. Reviews in Mineral-ogy Mineralogical Society of America, Washington, D.C.

    Ross, M., Papike, J.J., and Shaw, K.W. (1969) Exsolution textures in amphiboles asindicators of subsolidus thermal histories. Mineralogical Society of America,Special Paper, 2, 275299.

    Rutherford, M.J. and Devine, J.D. (1996) Pre-eruption pressure-temperature condi-

    tions and volatiles in the 1991 eruption of Mount Pinatubo magma. In: Newhall,C.G. and Punongbayan, R.S., eds. Fire and Mud. Eruptions and Lahars of MountPinatubo, Philippines. Seattle: University of Washington Press, pp. 751766.

    Sack, R.O. and Ghiorso, M.S. (1994) Thermodynamics of multicomponent py-roxenes: II. Applications to phase relations in the quadrilateral. Contributionsto Mineralogy and Petrology, 116, 287300.

    Scaillet, B. and Evans, B.W. (1999) The 15 June 1991 eruption of Mount Pinatubo.I. Phase equilibria and pre-eruptionP-T-fO2fH2O conditions of the dacite magma.Journal of Petrology, 40, 381411.

    Schneidermann, J.S. and Tracy, R.J. (1991) Petrology of orthoamphibole-cordieritegneisses from the Orijrvi area, southwest Finland. American Mineralogist, 76,942955.

    Seifert, F.A. (1977) Reconstruction of rock cooling paths from kinetic data on theFe2+-Mg exchange reaction in anthophyllite. Philosophical Transactions of theRoyal Society London A, 286, 303311.

    (1978) Equilibrium Mg-Fe2+ cation distribution in anthophyllite. AmericanJournal of Science, 278, 13231333.

    Seifert, F.A. and Virgo, D. (1975) Kinetics of the Fe2+-Mg order-disorder reaction in

    anthophyllites: quantitative cooling rates. Science, 188, 11071109.Spear, F.S. (1980) The gedrite-anthophyllite solvus and the composition limits of

    orthoamphibole from the Post Pond Volcanics, Vermont. American Mineralo-gist, 65, 11031118.

    (1982) Phase equilibria of amphibolites from the Post Pond Volcanics, Mt.Cube quadrangle, Vermont. Journal of Petrology, 23, 383426.

    Srikantappa, C., Raith, M., and Ackermand, D. (1985) High-grade regional meta-morphism of ultramafic and mafic rocks from the Archaean Sargur Terrane,Karnataka, South India. Precambrian Research, 30, 189219.

    Stout, J.H. (1972) Phase petrology and mineral chemistry of coexisting amphibolesfrom Telemark, Norway. Journal of Petrology, 13, 99146.

    Stroink, G., Blaauw, C., White, C.G., and Leiper, W. (1980) Mssbauer characteris-tics of UICE standard reference asbestos samples. Canadian Mineralogist, 18,285290.

    Sueno, S., Matsuura, S., Gibbs, G.V., and Boisen, M.B.Jr. (1998) A crystal chemical

    study of protoanthophyllite: orthoamphiboles with the protoamphibole struc-ture. Physics and Chemistry of Minerals, 25, 366377.

    Tella, S. and Eade, K.E. (1978) Coexisting cordierite-gedrite-cummingtonite fromEdehon Lake map area, Churchill structural province, District of Keewatin.Geological Survey of Canada, Paper 78-1C, 712.

    Tilley, C.E. (1937) Anthophyllite-cordierite granulites of the Lizard. GeologicalMagazine, 74, 300309.

    Todd, C.S. and Engi, M. (1997) Metamorphic field gradients in the Central Alps.Journal of Metamorphic Geology, 15, 513530.

    Trommsdorff, V. and Connolly, J.A. (1996) The ultramafic contact aureole aboutthe Bregaglia (Bergell) tonalite: isograds and a thermal model. SchweizerischeMineralogische und Petrographische Mitteilungen, 76, 537547.

    Vaniman, D.T., Papike, J.J., and Labotka, T. (1980) Contact metamorphic effects ofthe Stillwater Complex, Montana: the concordant iron formation. AmericanMineralogist, 65, 10871102.

    Yakovleva, A.K. and Kolesnikova, V.V. (1967) Characteristics of high-magnesiumcummingtonite from ultrabasic rocks. Mineralogical Abstracts, 1969, Vol. 20,No. 4, p. 314, abstract 69-3245.

    MANUSCRIPTRECEIVED APRIL 3, 2000MANUSCRIPTACCEPTED JANUARY 31, 2001MANUSCRIPTHANDLEDBY JOHN C. SCHUMACHER