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 37 毫秒
1.
The ratios of the carbon isotopes 13C and 12C of twenty-four graphite samples from Bogala Mine, Sri Lanka have been determined. The graphites are isotopically very similar (mean value ?7.76? °13CPDB; range <2?). Three models for the biogenic or abiogenic origin of the graphite deposit are discussed.  相似文献   

2.
The coprecipitation of Na and K was experimentally investigated in aragonite. The distribution functions were determined at pH 6.8 and 8.8 over aqueous Na and K concentrations of between 5 × 10?4and 2.0 M and temperatures of between 25 and 75°C.The mole fractions of Na and K in aragonite are related to the aqueous ratios of Na and Ca by a function of the form
log XNa2CO3,K2CO3 = C0 + C1loga2Na ? ,K?aCa2+
where C0 and C1 are constants at a given temperature. This equation was derived by a statistical model assuming a heterogeneous energy distribution for the sites of incorporation. The independence of the coprecipitation process from aqueous anion activities suggests that carbonate is the only anionic species in the solid solution.  相似文献   

3.
We have calculated the total individual ion activity coefficients of carbonate and calcium, γTCO32? and γTCa2+, in seawater. Using the ratios of stoichiometric and thermodynamic constants of carbonic acid dissociation and total mean activity coefficient data measured in seawater, we have obtained values which differ significantly from those widely accepted in the literature. In seawater at 25°C and 35%. salinity the (molal) values of γTCO23? and γTCa2+ are 0.038 ± 0.002 and 0.173 ± 0.010, respectively. These values of γTCO32? and γTCa2+ are independent of liquid junction errors and internally consistent with the value γTCl? = 0.651. By defining γTCa2+ and γTCO32? on a common scale (γTCl?), the product γTCa2+γTCO32? is independent of the assigned value of γCl? and may be determined directly from thermodynamic measurements in seawater. Using the value γTCa2+γTCO32? = 0.0067 and new thermodynamic equilibrium constants for calcite and aragonite, we show that the apparent constants of calcite and aragonite are consistent with the thermodynamic equilibrium constants at 25°C and 35%. salinity. The demonstrated consistency between thermodynamic and apparent constants of calcite and aragonite does not support a hypothesis of stable Mg-calcite coatings on calcite or aragonite surfaces in seawater, and suggests that the calcite critical carbonate ion curve of Broecker and Takahashi (1978, Deep-Sea Research25, 65–95) defines the calcite equilibrium boundary in the oceans, within the uncertainty of the data.  相似文献   

4.
The 3He4He ratios measured in 27 Southern Africa diamond stones, four from Premier Mine and the rest of unidentified origin, range from 4.2 × 10?8 to 3.2 × 10?4, with three stones above 1 × 10?4. We conclude that the initial helium isotopic ratio (3He4He)0 in the earth was significantly higher than that of the planetary helium-A (3He4He = 1.42 × 10?4), but close to the solar helium (3He4He ? 4 × 10?4).The apparent K-Ar ages for the twelve diamonds of unidentified origin show enormously old age, indicating excess argon-40. 3He4He evolution in diamonds suggests that the diamonds with the high 3He4He ratio (>2 × 10?4) may be as old as the earth.Noble gas elemental abundance in the diamonds relative to the air noble gas abundance shows monotonie decrease with a decreasing mass number.This paper discusses the implications of these observations on the early solar system and the origin of diamonds.  相似文献   

5.
In a soil developed on the Cretaceous chalk of the Eastern Paris basin, calcite dissolution begins at the surface. The soil water is rapidly saturated in calcite. Calcite dissolution follows two different pathways according to seasonal pedoclimatic conditions.During winter: the soil is only partly saturated in water and the CO2 partial pressure is low (Ca 10?3 atm.). As a consequence total inorganic dissolved carbon (TIDC) is a hundred times the carbon content of the gaseous phase. Equilibrium is usually observed between the two phases. It is a closed system. The measured carbon 14 activity (87,5%) and 13C content (δtidc13C = ?12,2%0) of the drainage water are very close to theoretical values calculated for an ideal mixing system between gaseous and mineral phases (respectively characterized by the following isotopic values: δG13C = ?21,5%0; AG14C = 118%; δM13C = +2,9%0; AM14C = 28%).During spring and summer: the soil moisture decreases, the input of biogenic CO2 induces an increase of the soil CO2 partial pressure (Ca from 3.10?3 atm to 7.10?3 atm). The carbon content of the gaseous phase is higher by an order of magnitude compared to winter conditions. Therefore the aqueous phase is undersaturated in CO2 with respect to the latter. This disequilibrium occurs as a result of unbalanced rates of CO2 dissolution and CO2 effusion toward atmosphère. It is an open system. The carbon isotopic ratio of the aqueous phase is regulated by that of the gaseous phase, as demonstrated by the agreement between measured and calculated isotopic compositions (respectively δL mes = from ?9,4%0 to ?11,5%0, δl calc = from ?9,8%0 to ?13,9%0 AL mes = 119%, AL calc = from 119% to 125%).The solutions originating from both systems (open and closed) move downwards without significant mixing together. It has also been observed that no significant variation of the TIDC isotopic composition occurs during precipitation of secondary calcite.  相似文献   

6.
Xanthates are used in the flotation of sulfide ores although their aqueous solutions are not stable under certain conditions. Their stability in acidic and weakly acidic aqueous solutions was therefore investigated, as these media are required for some processes.The peak absorbances of ethylxanthate ion and carbon disulfide were first determined in aqueous solution. The decomposition of ethylxanthate ion was analyzed by measuring variations in absorbance (at 301 nm) and pH with respect to time. A pH regulation system was then used while measuring variations in absorbance and productions of protons caused by xanthate decomposition.The results concerning xanthate half-lives show good agreement with the literature, but the kinetic results deviate substantially. The following relation was obtained for half-life:
T12=9.67×10?6(pH)11;4?7;T12in seconds
We established that ethylxanthate decomposition at pH 4 is a first order reaction with respect to ethylxanthate concentration, and postulating this order to the other pH values, the following kinetic relation was found:
v= ?(1.22×104[H+]?1.36×10?2)([EtX?]) (4?pH?7)
where v is the rate of decomposition (mol l?1 min?1), and [EtX?] is the ethylxanthate concentration when the decomposition equilibria are reached (mol l?1). The better concentration was found to obey the law:
[EtX?]=3.142×10?5 pH ? 1.255 × 10?4 (4?pH?6)
  相似文献   

7.
High precision mass spectrometric determination of calcium isotope ratios allows the 40K → 40Ca radioactive decay to be used for dating a much broader range of geologic materials than is suggested by previous work. 40Ca42Ca is used to monitor enrichments in 40Ca and can be measured to ±0.01% (2σ) using an exponential mass discrimination correction (Russell et al., 1978) and large ion currents. The earth's mantle has such a low KCa (~0.01) that it has retained “primordial” 40Ca42Ca = 151.016 ± 0.011 (normalized to 42Ca44Ca = 0.31221), as determined by measurements on two meteorites, pyroxene from an ultramafic nodule, metabasalt, and carbonatite. 40Ca42Ca ratios can be conveniently expressed relative to this value as ?Ca in units of 10?4. To test the method for age dating, a mineral isochron has been obtained on a sample of Pikes Peak granite, which has been shown to have concordant KAr, RbSr, and UPb ages. Plagioclase, K-feldspar, biotite, and whole rock yield an age of 1041 ± 32 m.y. (2σ) in agreement with previous age determinations (λK = 0.5543 b.y.?1, λβ?λK = 0.8952, 40K = 0.01167%). The initial 40Ca42Ca of 151.024 ± 0.016 (?Ca = +0.5 ± 1.0), indicates that assimilation of high K/Ca crust was insufficient to affect calcium isotopes. Measurements on two-mica granite from eastern Nevada indicate that the magma sources had K/Ca ≈ 1, similar to intermediate-composition crustal rocks. These results show that the KCa system can be used as a precise geochromometer for common felsic igneous and metamorphic rocks, and may prove applicable to sedimentary rocks containing authigenic K minerals. The relatively short half-life of 40K, the non-volatile daughter, and the fact that potassium and calcium are stoichiometric constituents of many minerals, make the KCa system complementary to other dating methods, and potentially applicable to a variety of geologic problems.  相似文献   

8.
The decay constant 87Rb has been redetermined by measuring the amount of radiogenic 87Sr produced over a period of 19 years, in 20 g samples of purified RbClO4, using isotope dilution techniques. The rubidium sample was spiked with 84Sr and the nanogram quantities of strontium separated by coprecipitation with Ba(NO3)2. Analyses were carried out on a 25cm, 90° sector mass spectrometer equipped with a Spiraltron electron multiplier. Measurement of three independent ratios permitted continuous monitoring of the ion beam fractionation. The average of nine determinations gives a value for the decay constant of 1.419(±0.012) × 10?11 yr?1 (2σ). [τ12 = 4.89(±0.04) × 1010yr.]  相似文献   

9.
The diffusivity of oxygen was determined in melts of Jadeite (NaAlSi2O6) and diopside (CaMgSi2O6) compositions using diffusion couples with 18O as a tracer. In the Jadeite melt, the diffusivity of oxygen increases from 6.87?0.25+0.28 × 10?10cm2/sec at 5 Kb to 1.32 ± 0.08 × 10?9cm2/sec at 20 Kb at constant temperature (1400°C), whereas in the diopside melt at 1650°C, the diffusivity decreases from 7.30?0.180.29 × 10?7cm2/sec at 10 Kb to 5.28?0.55+0.60 × 10?7cm2/sec at 17 Kb. These results demonstrate that the diffusivity is inversely correlated with the viscosity of the melt. For the jadeite melt, in particular, the inverse correlation is very well approximated by the Eyring equation using the diameter of oxygen ions as a unit distance of translation, suggesting that the viscous flow is rate-limited by the diffusion of individual oxygen ions. In the diopside melt, the activation volume is slightly greater than the molar volume of oxygen ion, indicating that the individual oxygen ion is the diffusion unit. The negative activation volume obtained for the jadeite melt is interpreted as the volume decrease associated with a diffusive jump of an oxygen ion due to local collapse of the network structure.  相似文献   

10.
Measurement of solubility as a function of pressure allows calculation of 3V?1. Using this experimental approach, the best estimate of 3V?1 for the dissolution of aged amorphous silica in salt water or seawater at 0–2°C is ?9.9 cm3 mol?1 (standard error = 0.4 cm3 mol?1). This gives V?Si(OH)4(aq)= 55 ± 5 cm3mol?1, which compares well with other published values of V?Si(OH)4(aq).  相似文献   

11.
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   

12.
Hydrogen which is highly enriched in deuterium is present in organic matter in a variety of meteorites including non-carbonaceous chondrites. The concentrations of this hydrogen are quite large. For example Renazzo contains 140 μmoles/g of the 10,000‰ δD hydrogen. The DH ratios of hydrogen in the organic matter vary from 8 × 10?5 to 170 × 10?5 (δD ranges from ? 500‰ to 10,000‰) as compared to 16 × 10?5 for terrestrial hydrogen and 2 × 10?5 for cosmic hydrogen. The majority of the unequilibrated primitive meteorites contain hydrogen whose DH ratios are greater than 30 × 10?5. If the DH ratios in these compounds were due to enrichment relative to cosmic hydrogen by isotope exchange reactions, it would require that these reactions take place below 150 K. In addition the organic compounds having DH ratios above 50 × 10?5 would require temperatures of formation of < 120 K. These types of deuterium enrichments must take place by ion-molecule reactions in interstellar clouds where both ionization and low temperatures exist. Astronomically observed DH ratios in organic compounds in interstellar clouds are typically 180 × 10?5 and range between about 40 × 10?5 and 5000 × 10?5. The DH values we have determined are the lower limits for the organic compounds derived from interstellar molecules because all processes subsequent to their formation, including terrestrial contamination, decrease their DH ratios.In contrast, the DH ratios of hydrogen associated with hydrated silicates are relatively uniform for the meteorites we have analyzed with an average value of 14 × 10?5; very similar to the terrestrial value. These phyllosilicates values suggest equilibration of H2O with H2 in the solar nebula at temperatures of about 200 K and higher.The 13C12C ratios of organic matter, irrespective its DH ratio, lie well within those observed for the earth. If organic matter originated in the interstellar medium, our data would indicate that the 13C12C ratio of interstellar carbon five billion years ago was similar to the present terrestrial value.Our findings suggest that other interstellar material, representing various inputs from various stars, in addition to the organic matter is preserved and is present in the meteorites which contain the high DH ratios. We feel that some elements existing in trace quantities which possess isotopic anomalies in the meteorites may very well be such materials.  相似文献   

13.
The relative reactivities of pulverized samples (100–200 mesh) of 3 marcasite and 7 pyrite specimens from various sources were determined at 25°C and pH 2.0 in ferric chloride solutions with initial ferric iron concentrations of 10?3 molal. The rate of the reaction:
FeS2 + 14Fe3+ + 8H2O = 15Fe2+ + 2SO2?4 + 16H+
was determined by calculating the rate of reduction of aqueous ferric ion from measured oxidation-reduction potentials. The reaction follows the rate law:
?dmFe3+dt = k(AM)mFe3+
where mFe3+ is the molal concentration of uncomplexed ferric iron, k is the rate constant and AM is the surface area of reacting solid to mass of solution ratio. The measured rate constants, k, range from 1.0 × 10?4 to 2.7 × 10?4 sec?1 ± 5%, with lower-temperature/early diagenetic pyrite having the smallest rate constants, marcasite intermediate, and pyrite of higher-temperature hydrothermal and metamorphic origin having the greatest rate constants. Geologically, these small relative differences between the rate constants are not significant, so the fundamental reactivities of marcasite and pyrite are not appreciably different.The activation energy of the reaction for a hydrothermal pyrite in the temperature interval of 25 to 50°C is 92 kJ mol?1. This relatively high activation energy indicates that a surface reaction controls the rate over this temperature range. The BET-measured specific surface area for lower-temperature/early diagenetic pyrite is an order of magnitude greater than that for pyrite of higher-temperature origin. Consequently, since the lower-temperature types have a much greater AM ratio, they appear to be more reactive per unit mass than the higher temperature types.  相似文献   

14.
The isotopic composition of strontium of pore water and of authigenic minerals leached from the sediment of core 119K with hot aqua regia is similar to that of the brine in the Discovery deep and differs from that of normal seawater. The average 87Sr86Sr ratio of strontium removed by acid leaching is 0.7077 ± 0.0007 (1σ) compared to a value of 0.70904 for the Red Sea. The detrital silicate fraction exhibits an approximate inverse correlation between 87Sr86Sr ratios and strontium concentrations which provides tentative support for a model in which the detrital silicate fraction of deep-sea sediment is considered to be a mixture of terrigenous dust of sialic composition enriched in radiogenic 87Sr and of volcanogenic material of basaltic composition and low 87Sr abundance. The 87Sr86Sr ratios of the shells of foraminifers and pteropods, expressed as δ 87Sr‰ relative to 0.70904 for seawater, decrease from ?0.23 ± 0.17‰ at 90 cm to ?0.82 ± 0.17‰ at 273 cm and remain constant at this value to a depth of 450 cm. The lowering of the δ 87Sr values is attributed both to the presence of aragonite overgrowths on pteropod shells and to possible isotope exchange with strontium in the connate fluid.  相似文献   

15.
Stability constants of hydroxocomplexes of Al(III):Al(OH)2+ and A1(OH)4? have been measured in the 20–70°C temperature range by reactions involving only dissolved species. The stability constant 1K1 of the first complex ion is studied by measuring pH of solutions of aluminium salts at several concentrations. 1β4 of aluminate ion is deduced from equilibrium constants of the reaction between the trioxalato aluminium (III) complex ion and Al3+ in acid medium, and between the same complex ion and A1(OH)4? in alkaline medium. The K values and the associated ΔH are 1K1 = 10?5.00 and ΔH1 = 11.8 Kcal; 1β4 = 10?22.20 and ΔH4 = 42.45 Kcal. These last results are not in agreement with the values of recent tables for ΔG0? and ΔH0? of Al3+ and Al(OH)4?. We suggest a consistent set of data for dissolved and solid Al species and for some aluminosilicates.  相似文献   

16.
Significant amounts of SO42?, Na+, and OH? are incorporated in marine biogenic calcites. Biogenic high Mg-calcites average about 1 mole percent SO42?. Aragonites and most biogenic low Mg-calcites contain significant amounts of Na+, but very low concentrations of SO42?. The SO42? content of non-biogenic calcites and aragonites investigated was below 100 ppm. The presence of Na+ and SO42? increases the unit cell size of calcites. The solid-solutions show a solubility minimum at about 0.5 mole percent SO42? beyond which the solubility rapidly increases. The solubility product of calcites containing 3 mole percent SO42? is the same as that of aragonite. Na+ appears to have very little effect on the solubility product of calcites. The amounts of Na+ and SO42? incorporated in calcites vary as a function of the rate of crystal growth. The variation of the distribution coefficient (D) of SO42? in calcite at 25.0°C and 0.50 molal NaCl is described by the equation D = k0 + k1R where k0 and k1 are constants equal to 6.16 × 10?6 and 3.941 × 10?6, respectively, and R is the rate of crystal growth of calcite in mg·min?1·g?1 of seed. The data on Na+ are consistent with the hypothesis that a significant amount of Na+ occupies interstitial positions in the calcite structure. The distribution of Na+ follows a Freundlich isotherm and not the Berthelot-Nernst distribution law. The numerical value of the Na+ distribution coefficient in calcite is probably dependent on the number of defects in the calcite structure. The Na+ contents of calcites are not very accurate indicators of environmental salinities.  相似文献   

17.
The synthetic chelating agent ethylenediaminetetraacetic acid (EDTA) has been used to evaluate the stoichiometric solubility product of galena (PbS) at 298°K: Ks2 = aPb2+aHS?aH+ This method circumvents the possible uncertainties in the stoichiometry and stability of lead sulfide complexes. At infinite dilution, Log Ks2 = ?12.25 ±0.17, and at an ionic strength corresponding to seawater (I = 0.7 M), Log Ks2 = ?11.73 ± 0.05. Using the value of Ks2 at infinite dilution, and the free energies of formation of HS? and Pb2+ at 298°K (literature values), the free energy of formation of PbS at 298°K is computed to be ?79.1 ± 0.8 KJ/mol (?18.9 Kcal/mol). Galena is shown to be more than two orders of magnitude more soluble than indicated by calculations based on previous thermodynamic data.  相似文献   

18.
Radioactivity measurements have shown evidence for long-lived 92Nb and 2.03 × 104yr 94Nb in natural niobium. The specific activity of 94Nb was observed to be 0.32 ± 0.03 dis/min. kg Nb and that of 92Nb to be 0.058 ± 0.035 dis/min. kg Nb. With t12 taken as ≈ 1.7 × 108yr, the isotopic abundance of 98Nb is 1.2 × 10?10 per cent.  相似文献   

19.
The South Mountain batholith of southwestern Nova Scotia is a large, peraluminous, granodiorite-granite complex which intrudes mainly greenschist facies metasediments of the Cambro-Ordovician Meguma Group. Using Rb-Sr isochrons constructed from whole rocks and mineral separates, the present study shows a variation in age and initial ratios of the intrusive phases of the batholith as follows: biotite granodiorite (371.8 ± 2.2 Ma, (87Sr86Sr)i ranges from 0.7076 ± 0.0003 to 0.7090 ± 0.0003, with the average = 0.7081); adamellite (364.3 ± 1.3 Ma, (87Sr86Sr)i = 0.70942 ± 35); porphyry (361.2 ± 1.4 Ma, (87Sr86Sr)i = 0.71021 ± 119); using λ87Rb = 1.42 × 10?11yr?1.A suite of Meguma country rock samples showed a variation of 87Sr86Sr = 0.7113?0.7177 at the time of intrusion of the batholith. A number of xenoliths of this material occurring in the marginal granodiorite had partially equilibrated isotopically with the granodiorite at a higher 87Sr86Sr ratio than elsewhere in the granodiorites. This evidence demonstrates that isotopic (and probably some accompanying bulk chemical) contamination by the Meguma rocks has been an important factor in determining the ultimate chemical composition and mineralogy of the South Mountain batholith.The (87Sr86Sr)372 = 0.7081 of the early granodiorites indicates that the parent magma of the South Mountain batholith was derived from a source unlike the Meguma Group. The precise nature of the source region cannot be determined by Rb-Sr work unless the degree of contamination with Megumalike material is known.  相似文献   

20.
Metal and silicate portions from 13 mesosiderites, one pallasite, Bencubbin (“unique”) and Udei Station (‘iron with silicate inclusions’) have been analysed for their content of He, Ne and Ar; in most cases 36Cl could be determined as well. 36Cl-36Ar cosmic ray exposure ages fall between 10 and 160 Myr. Half of the metal samples show a deficit of spallogenic 3He (up to 30%) which we ascribe to a loss of tritium. The observed depletion of 3He in the silicates is correlated with their mineralogical composition: feldspar has lost its 3He in all cases, pyroxene definitely in one and possibly in five others, while olivine has been affected in only two meteorites. The thermal histories during their exposure to the cosmic radiation have been different for different meteoroids. Nevertheless, with the exception of Veramin, the data are compatible with the assumption of a continuous diffusion loss during a considerable fraction of the exposure era. For Veramin, however, an episodic event late in the exposure history is required. The exceptionally high 39Ar36Cl ratio in the metal, which is due to a high 39Ar activity, indicates that the event occurred during the last 500,000 years or so and resulted in an extremely excentric orbit (large aphelion).Production rates of 38,39Ar from Ca and 21,22Ne from Mg are given. The ratio P38CaP21Mg is close to unity. The ratios P38CaP38Fe vary between 20 and 50, and are not correlated with the absolute production rate of 38Ar from metal. The 22Ne21Ne production ratio from Mg is found to be close to but below unity.Of the mesosiderites only Veramin shows unambiguous evidence for primordial rare gases with larger amounts and a higher 20Ne36Ar ratio in the olivine, suggesting in situ fractionation to have at least been partly responsible for the abundance pattern found. Bencubbin contains large amounts of strongly fractionated primordial gases, but again part of the fractionation may have occurred in situ. Udei Station shows an excess of (3.5 ± 0.6) × 10?10 cm3 STP 129Xe/g in the non-magnetic portion.  相似文献   

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