首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 78 毫秒
1.
2.
Equations are developed for calculating the density of aluminosilicate liquids as a function of composition and temperature. The mean molar volume at reference temperature Tr, is given by Vr = ∑XiV?oi + XAV?oA, where the summation is taken over all oxide components except A12O3, X stands for mole fraction, V?oi terms are constants derived independently from an analysis of volume-composition relations in alumina-free silicate liquids, and V?oA is the composition-dependent apparent partial molar volume of Al2O3. The thermal expansion coefficient of aluminosilicate liquids is given by α = ∑Xi\?gaio + XA\?gaAo, where \?gaio terms are constants independent of temperature and composition, and \?gaoA is a composition-dependent term representing the effect of Al2O3 on the thermal expansion. Parameters necessary to calculate the volume of silicate liquids at any temperature T according to V(T) = Vrexp[α(T-Tr)], where Tr = 1400°C have been evaluated by least-square analysis of selected density measurements in aluminosilicate melts. Mean molar volumes of aluminosilicate liquids calculated according to the model equation conform to experimentally measured volumes with a root mean square difference of 0.28 ccmole and an average absolute difference of 0.90% for 248 experimental observations. The compositional dependence of V?oA is discussed in terms of several possible interpretations of the structural role of Al3+ in aluminosilicate melts.  相似文献   

3.
Differences in the chemical composition of metamorphic and igneous pyroxene minerals may be attributed to a transfer reaction, which determines the Ca content of the minerals, and an exchange reaction, which determines the relative Mg:Fe2+ ratios. Natural data for associated Ca pyroxene (Cpx) and orthopyroxene (Opx) or pigeonite are combined with experimental data for Fe-free pyroxenes, to produce the following equations for the Cpx slope of the solvus surface: > 1080°C: T = 1000(0.468 + 0.246XCpx ? 0.123 ln (1–2 [Ca]))< 1080°C: T = 1000(0.054 + 0.608XCpx ? 0.304 ln (1–2 [Ca])), and the following equation for the temperature-dependence of the Mg-Fe distribution coefficient: T = 1130(ln Kp + 0.505), where T is absolute temperature, X is Fe2+(Mg + Fe2+)), [Ca] is Ca(Ca + Mg + Fe2+) in Cpx, and KD is the distribution coefficient, defined as XOpx/(1 ? XOpx) ÷ XCpx/(1 ? Cpx).The transfer and exchange equations form useful temperature indicators, and when applied to 9 sets of well-studied rocks, yield pairs of temperatures that are in good agreement. For example, temperatures obtained for the Bushveld Complex are 1020°C (solvus equation) and 980°C (exchange equation), based on 7 specimens. The uncertainty in these numbers, due to precision and accuracy errors, is estimated to be ±60°.  相似文献   

4.
The stability of the amphibole pargasite [NaCa2Mg4Al(Al2Si6))O22(OH)2] in the melting range has been determined at total pressures (P) of 1.2 to 8 kbar. The activity of H2O was controlled independently of P by using mixtures of H2O + CO2 in the fluid phase. The mole fraction of H2O in the fluid (XH2O1fl) ranged from 1.0 to 0.2.At P < 4 kbar the stability temperature (T) of pargasite decreases with decreasing XH2O1fl at constant P. Above P ? 4 kbar stability T increases as XH2O1fl is decreased below one, passes through a T maximum and then decreases with a further decrease in XH2O1fl. This behavior is due to a decrease in the H2O content of the silicate liquid as XH2O1fl decreases. The magnitude of the T maximum increases from about 10°C (relative to the stability T for XH2O1fl= 1) at P = 5 kbar to about 30°C at P = 8 kbar, and the position of the maximum shifts from XH2O1fl ? 0.6 at P = 5 kbar to XH2O1fl? 0.4 at P = 8 kbar.The H2O content of liquid coexisting with pargasite has been estimated as a function of XH2O1fl at 5 and 8 kbar P, and can be used to estimate the H2O content of magmas. Because pargasite is stable at low values of XH2O1fl at high P and T, hornblende can be an important phase in igneous processes even at relatively low H2O fugacities.  相似文献   

5.
Solution calorimetric measurements compared with solubility determinations from the literature for the same samples of gibbsite have provided a direct thermochemical cycle through which the Gibbs free energy of formation of [Al(OH)4 aq?] can be determined. The Gibbs free energy of formation of [Al(OH)4 aq?] at 298.15 K is ?1305 ± 1 kJ/mol. These heat-of-solution results show no significant difference in the thermodynamic properties of gibbsite particles in the range from 50 to 0.05 μm.The Gibbs free energies of formation at 298.15 K and 1 bar pressure of diaspore, boehmite and bayerite are ?9210 ± 5.0, ?918.4 ± 2.1 and ?1153 ± 2 kJ/mol based upon the Gibbs free energy of [A1(OH)4 aq?] calculated in this paper and the acceptance of ?1582.2 ± 1.3 and ?1154.9 ± 1.2 kJ/mol for the Gibbs free energy of formation of corundum and gibbsite, respectively.Values for the Gibbs free energy formation of [Al(OH)2 aq+] and [AlO2 aq?] were also calculated as ?914.2 ± 2.1 and ?830.9 ± 2.1 kJ/mol, respectively. The use of [AlC2 aq?] as a chemical species is discouraged.A revised Gibbs free energy of formation for [H4SiO4aq0] was recalculated from calorimetric data yielding a value of ?1307.5 ± 1.7 kJ/mol which is in good agreement with the results obtained from several solubility studies.Smoothed values for the thermodynamic functions CP0, (HT0 - H2980)T, (GT0 - H2980)T, ST0 - S00, ΔH?,2980 kaolinite are listed at integral temperatures between 298.15 and 800 K. The heat capacity of kaolinite at temperatures between 250 and 800 K may be calculated from the following equation: CP0 = 1430.26 ? 0.78850 T + 3.0340 × 10?4T2 ?1.85158 × 10?4T212 + 8.3341 × 106 T?2.The thermodynamic properties of most of the geologically important Al-bearing phases have been referenced to the same reference state for Al, namely gibbsite.  相似文献   

6.
Light hydrocarbon (C1-C3) concentrations in the water from four Red Sea brine basins (Atlantis II, Suakin, Nereus and Valdivia Deeps) and in sediment pore waters from two of these areas (Atlantis II and Suakin Deeps) are reported. The hydrocarbon gases in the Suakin Deep brine (T = ~ 25°C, Cl? = ~ 85‰, CH4 =~ 711) are apparently of biogenic origin as evidenced by C1(C2 + C3) ratios of ~ 1000. Methane concentrations (6–8 μl/l) in Suakin Deep sediments are nearly equal to those in the brine, suggesting sedimentary interstitial waters may be the source of the brine and associated methane.The Atlantis II Deep has two brine layers with significantly different light hydrocarbon concentrations indicating separate sources. The upper brine (T = ~ 50°C, Cl? = ~ 73‰, CH4 = ~ 155 μl/l) gas seems to be of biogenic origin [C1(C2 + C3) = ~1100], whereas the lower brine (T = ~ 61°C, Cl? = ~ 155‰, CH4 = ~ 120μl/l) gas is apparently of thermogenic origin [C1(C2 + C3) = ~ 50]. The thermogenic gas resulting from thermal cracking of organic matter in the sedimentary column apparently migrates into the basin with the brine, whereas the biogenic gas is produced in situ or at the seawater-brine interface. Methane concentrations in Atlantis II interstitial waters underlying the lower brine are about one half brine concentrations; this difference possibly reflects the known temporal variations of hydrothermal activity in the basin.  相似文献   

7.
The effect of presure on the solubility of minerals in water and seawater can be estimated from In
(KPspK0sp) + (?ΔVP + 0.5ΔKP2)RT
where the volume (ΔV) and compressibility (ΔK) changes at atmospheric pressure (P = 0) are given by
ΔV = V?(M+, X?) ? V?[MX(s)]ΔK = K?(M+, X?) ? K?[MX(s)]
Values of the partial molal volume (V?) and compressibilty (K?) in water and seawater have been tabulated for some ions from 0 to 50°C. The compressibility change is quite large (~10 × 10?3 cm3 bar?1 mol?1) for the solubility of most minerals. This large compressibility change accounts for the large differences observed between values of ΔV obtained from linear plots of In Ksp versus P and molal volume data (Macdonald and North, 1974; North, 1974). Calculated values of KPspKosp for the solubility of CaCO3, SrSO4 and CaF2 in water were found to be in good agreement with direct measurements (Macdonald and North, 1974). Similar calculations for the solubility of minerals in seawater are also in good agreement with direct measurements (Ingle, 1975) providing that the surface of the solid phase is not appreciably altered.  相似文献   

8.
Optical and analytical studies were performed on 400 N2 + CO2 gas bearing inclusions in dolomites and quartz from Triassic outcrops in northern Tunisia. Other fluids present include brines (NaCl and KCl bearing inclusions) and rare liquid hydrocarbons. At the time of trapping, such fluids were heterogeneous gas + brine mixtures. In hydrocarbon free inclusions the N2(N2+ CO2) mole ratio was determined using two different non-destructive and punctual techniques: Raman microprobe analysis, and optical estimation of the volume ratios of the different phases selected at low temperatures. In the observed range of compositions, the two methods agree reasonably well.The N2 + CO2 inclusions are divided into three classes of composition: (a) N2(N2 + CO2) > 0,57: Liquid nitrogen is always visible at very low temperature and homogenisation occurs in the range ?151°C to ? 147°C (nitrogen critical temperature) dry ice (solid CO2) sublimates between ?75°C and ?60°C; (b) 0,20 < N2(N2 + CO2) ? 0,57: liquid nitrogen is visible at very low temperature but dry ice melts on heating; liquid and gas CO2 homogenise to liquid phase between ?51°C to ?22°C; (c) N2(N2 + CO2) ? 0,20: liquid nitrogen is not visible even at very low temperature (?195°C) and liquid and gas CO2 homogenise to liquid phase between ?22°C and ?15°C. The observed phases changes are used to propose a preliminary phase diagram for the system CO2-N2 at low temperatures.Assuming additivity of partial pressures, isochores for the CO2-N2 inclusions have been computed. The intersection of these isochores with those for brine inclusions in the same samples may give the P and T of trapping of the fluids.  相似文献   

9.
Calculations based on approximately 350 new measurements (CaT-PCO2) of the solubilities of calcite, aragonite and vaterite in CO2-H2O solutions between 0 and 90°C indicate the following values for the log of the equilibrium constants KC, KA, and KV respectively, for the reaction CaCO3(s) = Ca2+ + CO2?3: Log KC = ?171.9065 ? 0.077993T + 2839.319T + 71.595 log TLog KA = ?171.9773 ? 0.077993T + 2903.293T +71.595 log TLog KV = ?172.1295 ? 0.077993T + 3074.688T + 71.595 log T where T is in oK. At 25°C the logarithms of the equilibrium constants are ?8.480 ± 0.020, ?8.336 ± 0.020 and ?7.913 ± 0.020 for calcite, aragonite and vaterite, respectively.The equilibrium constants are internally consistent with an aqueous model that includes the CaHCO+3 and CaCO03 ion pairs, revised analytical expressions for CO2-H2O equilibria, and extended Debye-Hückel individual ion activity coefficients. Using this aqueous model, the equilibrium constant of aragonite shows no PCO2-dependence if the CaHCO+3 association constant is Log KCahco+3 = 1209.120 + 0.31294T — 34765.05T ? 478.782 log T between 0 and 90°C, corresponding to the value logKCahco+3 = 1.11 ± 0.07 at 25°C. The CaCO03 association constant was measured potentiometrically to be log KCaCO03 = ?1228.732 ? 0.299444T + 35512.75T + 485.818 log T between 5 and 80°C, yielding logKCaCO03 = 3.22 ± 0.14 at 25°C.The CO2-H2O equilibria have been critically evaluated and new empirical expressions for the temperature dependence of KH, K1 and K2 are log KH = 108.3865 + 0.01985076T ? 6919.53T ? 40.45154 log T + 669365.T2, log K1 = ?356.3094 ? 0.06091964T + 21834.37T + 126.8339 log T — 1684915.T2 and logK2 = ?107.8871 ? 0.03252849T + 5151.79/T + 38.92561 logT ? 563713.9/T2 which may be used to at least 250°C. These expressions hold for 1 atm. total pressure between 0 and 100°C and follow the vapor pressure curve of water at higher temperatures.Extensive measurements of the pH of Ca-HCO3 solutions at 25°C and 0.956 atm PCO2 using different compositions of the reference electrode filling solution show that measured differences in pH are closely approximated by differences in liquid-junction potential as calculated by the Henderson equation. Liquid-junction corrected pH measurements agree with the calculated pH within 0.003-0.011 pH.Earlier arguments suggesting that the CaHCO+3 ion pair should not be included in the CaCO3-CO2-H2O aqueous model were based on less accurate calcite solubility data. The CaHCO+3 ion pair must be included in the aqueous model to account for the observed PCO2-dependence of aragonite solubility between 317 ppm CO2 and 100% CO2.Previous literature on the solubility of CaCO3 polymorphs have been critically evaluated using the aqueous model and the results are compared.  相似文献   

10.
Thermal analysis of the halite liquidus in the system NaCl-H2O has been conducted for NaCl mole fractions (XNaCl) greater than 0.25 (i.e., > 50 wt. % NaCl) at pressures between 0.3 and 4.1 kb and temperatures greater than 450°C. The position of the liquidus was located by differential thermal analysis (DTA) of cooling scans only, as heating scans did not produce definitive DTA peaks. The dP/dT slope of the liquidus is positive and steep at high pressures, but at high XNaCl, and pressures below 0.5 kb it appears to reverse slope and intersects the three-phase curve (liquid-halite-vapour) at a shallow angle. However, due to the complex nature of the DTA signal when P <- 0.5 kb, there is considerable doubt about exactly what event has been recorded in the experiments conducted at these low pressures.The solubility of halite can be expressed as a function of the mole-fractional-based activity of NaCl in the liquid phase (L) in temperature (T, °K) and pressure (P, bars) In αNaCl(L.T.P) = ?19.884 ? 0.001275T ? 1388T + 3.2305 In (T) ? 0.07574PT Our liquidus data (based on 10 compositions) above 500 bars for these brines were combined with this equation to generate activity coefficients of NaCl which were fit within their experimental uncertainties to the following one parameter Margules equation In γNaCl(L.T.P) = (0.7268 ? 695.7T ? 0.1217PT)(1 ? XNaCl)2. Concentrated solutions of NaCl show negative deviations from ideality which rapidly increase in magnitude with decreasing XNaCl.  相似文献   

11.
For a phase at equilibrium in which two cation species are partitioned ideally between two sub-lattice sites, the excess functions of mixing (free energy, enthalpy and entropy) are directly related to the bulk composition of the phase and ΔGE°(T, P), the standard-state intra- crystalline exchange free energy. If the phase is not at equilibrium internally, an additional ordering parameter is necessary to fix the excess free energy of mixing, GmixEX, unambiguously. Conversely, for any fixed GmixEX there exists an infinity of possible intracrystalline cation dis- tributions, only one of which is the equilibrium distribution for the specified temperature and pressure. As ideal intraphase cation ordering becomes more pronounced, GmixEX decreases. In response, the total free energy of mixing for the phase decreases progressively for non-end member compositions, approaching, at the limits of ordering, values appropriate for stabilizing compounds of intermediate composition.The model-dependent activity coefficient for component A in the phase, γAT, can be calculated for any bulk composition, XAT, either from GmixEX directly or from more basic equations involving the interrelation of chemical potentials at equilibrium. A general form for γAT is ln γAT= 1n[2(XAαXAβ)12/(XAα+XAβ)]+Y, where Xjκ denotes the mole fraction of species j in site κ. The first term on the right-hand side of this equation is the contribution to γAT from ideal intracrystalline partitioning, and is common to the several theories lately presented to model intraphase cation partitioning. It can be shown rigorously that this term contributes to a negative deviation from ideality for the bulk phase. The second term is the contribution to the macroscopic activity coefficient from non-ideal intraphase partitioning, and is related to an enthalpy of mixing, HmixN in excess of that resulting from ideal inter-site cation ordering. While the expression represented by Y can take several functional forms, the additional enthalpy can be evaluated explicitly for specific non-ideal partitioning models from the relation HmixN = 2RT(1? XAT) ∝ Y(1 ? XAT)2dXAT.In those cases, GmixEX can also be determined exactly.  相似文献   

12.
A differential rate equation for silica-water reactions from 0–300°C has been derived based on stoichiometry and activities of the reactants in the reaction SiO2(s) + 2H2O(l) = H4SiO4(aq)
(?aH4SiO4?t)P.T.M. = (AM)(γH4SiO4)(k+aSiO2a2H2O ? k_aH4SiO4)
where (AM) = (the relative interfacial area between the solid and aqueous phases/the relative mass of water in the system), and k+ and k? are the rate constants for, respectively, dissolution and precipitation. The rate constant for precipitation of all silica phases is log k? = ? 0.707 ? 2598T(T, K) and Eact for this reaction is 49.8 kJ mol?1. Corresponding equilibrium constants for this reaction with quartz, cristobalite, or amorphous silica were expressed as log K = a + bT + cT. Using K =k+k?, k was expressed as log k + = a + bT + cT and a corresponding activation energy calculated:
  相似文献   

13.
The enthalpies of solution of a suite of 19 high-structural state synthetic plagioclases were measured in a Pb2B2O5 melt at 970 K. The samples were crystallized from analyzed glasses at 1200°C and 20 kbar pressure in a piston-cylinder apparatus. A number of runs were also made on Amelia albite and Amelia albite synthetically disordered at 1050–1080°C and one bar for one month and at 1200°C and 20 kbar for 10 hr. The component oxides of anorthite, CaO, Al2O3 and SiO2, were remeasured.The ΔH of disorder of albite inferred in the present study from albite crystallized from glass is 3.23 kcal, which agrees with the 3.4 found by Holm and Kleppa (1968). It is not certain whether this value includes the ΔH of a reversible displacive transition to monoclinic symmetry, as suggested by Helgesonet al. (1978) for the Holm-Kleppa results. The enthalpy of solution value for albite accepted for the solid solution series is based on the heat-treated Amelia albite and is 2.86 kcal less than for untreated Amelia albite.The enthalpy of formation from the oxides at 970 K of synthetic anorthite is ?24.06 ± 0.31 kcal, significantly higher than the ?23.16 kcal found by Charluet al. (1978), and in good agreement with the value of ?23.89 ± 0.82 given by Robieet al. (1979), based on acid calorimetry.The excess enthalpy of mixing in high plagioclase can be represented by the expression, valid at 970 K: ΔHex(±0.16 kcal) = 6.7461 XabX2An + 2.0247 XAnX2Ab where XAb and XAn are, respectively, the mole fractions of NaAlSi3O8 and CaAl2Si2O8. This ΔHex, together with the mixing entropy of Kerrick and Darken's (1975) Al-avoidance model, reproduces almost perfectly the free energy of mixing found by Orville (1972) in aqueous cation-exchange experiments at 700°C. It is likely that Al-avoidance is the significant stabilizing factor in the high plagioclase series, at least for XAn≥ 0.3. At high temperatures the plagioclases have nearly the free energies of ideal one-site solid solutions. The Al-avoidance model leads to the following Gibbs energy of mixing for the high plagioclase series: ΔGmix = ΔHex + RT XAbln[X2Ab(2 ? XAb)]+ XAnln[XAn(1+XAn)2]4. The entropy and enthalpy of mixing should be very nearly independent of temperature because of the unlikelihood of excess heat capacity in the albite-anorthite join.  相似文献   

14.
Calcium-45 was used as a radiotracer to measure self-diffusion coefficients for Ca in a sodium-calcium-aluminosilicate melt (29% Na2O, 5% CaO, 10% Al2O3, 56% SiO2) at temperatures in the range 1100–1400°C and pressures to 30 kbar. Calcium diffusivity (DCa) was found to depend upon both temperature and pressure in a complex but systematic manner: (?DCa?P)T is always negative and has a larger absolute value at lower temperatures; (?DCa?T)P is positive and increases with increasing pressure. The overall dependence of DCa upon T and P is given approximately by DcaT.P = [0.0025 exp(-23,107RT)] exp [P(0.7297T ? 1261.32)RT]. When expressed in terms of volume (Va) and energy (E) of activation, the results are as follows: Va ranges from 2.2 cm3/mole at 1400°C to 11.9 cm3/mole at 1100°C. and E ranges from 25.4 kcal/mole (1 kban to 49.8 kcal/mole (20 kbar).From the systematic dependence of DCa upon T and P, it is concluded that diffusion of Ca2+ in silicate melts does not take place by means of a vacant site mechanism, but is controlled instead by the amount and distribution of free volume in the melt structure.If it is assumed that the viscosity of the melt used in this study decreases with increasing pressure (Kushiro, 1976, J. Geophys. Res.81, 6351–6356) as DCa does, then the Stokes-Einstein inverse relation between viscosity and diffusivity is clearly violated, and its validity for silicate melts must be questioned. Thus, it appears that in silicate melts, unlike many liquids, viscous flow and diffusion are fundamentally different transport processes, involving different structural units.The effect of pressure on calcium diffusion is too small to invalidate kinetic models of upper mantle processes that have been based upon diffusivity values measured at 1 atm. Pressure may, however, induce significant reductions in the diffusion rates of large ions such as Rb+ or SiO4?4 in silicate melts.  相似文献   

15.
The carbonato and hydrogencarbonato complexes of Mg2+ were investigated at 25 and 50° in solutions of the constant ClO4? molality (3 M) consisting preponderantly of NaClO4. The experimental data could be explained assuming the following equilibria: Mg2+ + CO2B + H2O ag MgHCO+3 + H+, log 1β1 = ?7.644 ± 0.017 (25°), ?7.462 ± 0.01 1 (50°), Mg2+ + 2 CO2g + 2 H2Oag Mg(HCO3)02 ± 2 H+, log 1β2 = ?15.00 ± 0.14 (25°), ?15.37 ± 0.39 (50°), Mg2+ + CO2g + H2Oag MgCO03 + 2 H+, log 1k1 = ?15.64 ± 0.06 (25°),?15.23 ± 0.02 (50°), with the assumption γMgCO30 = γMg(HCO3)02, ΔG0(I = 0) for the reaction MgCO03 + CO2g + H2O = Mg(HCO3)02 was estimated to be ?3.91 ± 0.86 and 0.6 ± 2.4 kJ/mol at 25 and 50°C, respectively. The abundance of carbonate linked Mg(II) species in fresh water systems is discussed.  相似文献   

16.
The partial molal volume of silicic acid (V?(Si(OH)4)) in 0.725 M NaCl at 1°C was calculated from the measured volume change (ΔV?n) due to the neutralization of anhydrous sodium metasilicate with HCl and the V?(HCl) and V?(NaCl) obtained from the literature. V?(Si(OH)4) = 59.0 cm3mol ? 1, determined under experimental conditions of pH = 2.2, compares favorably with V?(Si(OH)4) = 58.9 cm3 mol?1 calculated from the measured volume change due to the hydrolysis of the meta-silicate salt at pH = 11 and from the partial molal volume due to electrostriction (V?elect) of water by charged Si species present in the solution at the high pH. This agreement lends support to a semiempirical model for calculating V?elect in developed by Millero (1969). V?(NaOH) = ? 5.45 cm3 mol?1 in 0.725 M NaCl needed for this calculation was also determined in this work. The rate of polymerization of Si(OH)4 at 1°C was monitored to insure that the monomer Si(OH)4 was the main Si species present during the determination of V?(Si(OH)4) by neutralization of the alkali silicate. V?(Si(OH)4) determined in this study compares favorably with the value calculated from high pressure solubility measurements.  相似文献   

17.
The distribution coefficients of Eu and Sr for plagioclase-liquid and clinopyroxene-liquid pairs as a function of temperature and oxygen fugacity were experimentally investigated using an oceanic ridge basalt enriched with Eu and Sr as the starting material. Experiments were conducted between 1190° and 1140°C over a range of oxygen fugacities between 10?8 and 10?14 atm.The molar distribution coefficients are given by the equations: log KEuPL = 3320/T?0.15 log?o2?4.22log KCPXEu = 6580/T + 0.04 log?o2?4.37logPLSr = 7320/T ? 4.62logKCPXSr = 18020/T ? 13.10. Similarly, the weight fraction distribution coefficients are given by the equations: log DPLEu =2460/T ? 0.15 log?o2 ? 3.87log DCPXEu = 6350/T + 0.04 log?o2 ? 4.49logDPLSr = 6570/T ? 4.30logDCPXSr = 18434/T ? 13.62.Although the mole fraction distribution coefficients have a smaller dependence on bulk composition than do the weight fraction distribution coefficients, they are not independent of bulk composition, thereby restricting the application of these experimental results to rocks similar to oceanic ridge basalts in bulk composition.Because the Sr distribution coefficients are independent of oxygen fugacity, they may be used as geothermometers. If the temperature can be determined independently — for example, with the Sr distribution coefficients, the Eu distribution coefficients may be used as oxygen geobarometers. Throughout the range of oxygen fugacities ascribed to terrestrial and lunar basalts, plagioclase concentrates Eu but clinopyroxene rejects Eu.  相似文献   

18.
Potentiometric measurements in dilute sodium borate solutions with added alkali earth chlordie salts yield the following expressions for the dissociation constants of alkali earth borate ion pairs from 10 to 50°C:
pK(MgH2BO3+=1.266+0.001204 T
pK(CaH2BO3+=1.154+0.002170 T
pK(SrH2BO3+=1.033+0.001738 T
pK(BaH2BO3+=1.942+0.001850 T
where T is in °K. Enthalpies for the dissociation reactions at 25°C are less than 1 kcal./mole for all the alkali earth borate ion pairs.Values for pK(NaH2BO3°) from 5 to 55°C computed from the experimental data of Owen and King are in good agreement with those determined potentiometrically. The average value from both methods is 0.22 ± 0.1 at 25°C.Application to seawater of computed pK's for MgH2BO3+, CaH2BO3+ and NaH2BO30 yields an apparent dissociation constant for boric acid of 8.73 vs. 8.70 measured by Lyman, 8.68 by Buch and 8.73 by Byrne and Kester.  相似文献   

19.
A linear correlation exists between the standard Gibbs free energies of formation of calcite-type carbonates (MCO3) and the corresponding conventional standard Gibbs free energies of formation of the aqueous divalent cations (M2+) at 25 °C and 1 bar ΔGMCO30 = m(ΔGf,M2+0) ? 141,200 cal · mole?1 where m is equal to 0.9715. This relationship enables prediction of the standard free energies of formation of numerous hypothetical carbonates with the calcite structure. Associated uncertainties typically range from about ± 250 to 600 cal · mole?1. An important consequence of the above correlation is that the thermodynamic equilibrium constant for the distribution of two trace elements M and N between carbonate mineral and aqueous solution at 25 °C and 1 bar is proportional to the free energy difference between the corresponding two aqueous ions: In KM-N = m ? 1298.15RG?f,M2+0 ? ΔG?f,N2+0)Combination of predicted standard free energies, entropies and volumes of carbonate minerals at 25°C and 1 bar with standard free energies of aqueous ions and the equation of state in Helgesonet al. (1981) enables prediction of the thermodynamic equilibrium constant for trace element distribution between carbonates and aqueous solutions at elevated temperatures and pressures. Interpretation of the thermodynamic equilibrium constant in terms of concentration ratios in the aqueous phase is considerably simplified if pairs of divalent trace elements are considered that have very similar ionic radii (e.g., Sr2+Pb2+, Mg2+Zn2+). In combination with data for the stabilities of complex ions in aqueous solutions, the above calculations enable useful limits to be placed on the concentrations of trace elements in hydrothermal solutions.  相似文献   

20.
The distribution of trace amounts of Na, Rb and Cs, between muscovite, phlogopite, sanidine and hydrothermal solution have been studied by ion exchange in a temperature range from 400 to 800°C.These distributions have been expressed with a partition ratio Paq?mx = (XK)aq(XK)m (where X is Na, Rb or Cs).In the case of Na and Cs in muscovite, even for the dilute solutions, the ratio Paq?mx is not the equilibrium constant kx of exchange reactions. In other cases, Paq?mx does not depend on the trace alkali ion concentration in silicates (X) and is equal to kx. Variations of Px or kx with T are greater for Na and Cs than for Rb. Generally, kx decreases with increase in T. The function log Px = f(1T) is not linear for Na or Cs, but in the case of Rb, f(1T) is linear and the standard enthalpy and entropy of exchange reactions have been estimated by applying the Arrhenius relation.The distribution relations obtained between silicate and vapour phase permit the determination of distributions of Na, Rb and Cs between two minerals mI and mII, relative to K. These have been expressed with the partition ratio Qx =(XK)mI(XK)mII. Variations of Qx with T are not remarkable, and even for Rb between phlogopite and feldspar are negligible. Nevertheless, one may use the distributions of Rb and Cs between muscovite and feldspar for geothermometry. Experimental results have been applied to some rocks by effecting corrections from the major element composition of the natural minerals. Estimated temperatures are near to 400°C in the granites and pegmatite studied here.  相似文献   

abcEact(kJ mol -1)
Quarts1.174-2.028 x 103-415867.4–76.6
α-Cristobalite-0.7390-358668.7
β-Cristobalite-0.9360-339265.0
Amorphous silica-0.369-7.890 x 10-4343860.9–64.9
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号