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 234 毫秒
1.
海水硫酸盐参与许多发生在海洋水中和海相沉积物中的氧化还原作用,并且是多种沉积物和矿床中硫的来源。海水硫酸盐的硫同位素组成与海相环境中各种含硫化合物的硫同位素组成有着直接或间接的成因联系。在很大程度上,海水硫酸盐的硫同位素组成提供了海相环境中硫同位素演变的起点。找到这一起点才能正确阐明同时代海相沉积物中硫同位素之间的关系,进而探寻其演变规律。目前,对确定古海洋硫同位素组成最方便的研究对象是海相石膏。  相似文献   

2.
硫同位素地质研究工作中,经常遇到的研究对象是硫酸盐矿物。如何把这些硫酸盐矿物转化为适于质谱测定硫同位素组成的SO2气体,是我国硫同位素地质研究中急待建立的实验手段之一。 经典的方法中,可溶于水的硫酸盐,通常是先把它沉淀为BaSO4,然后通过一系列的化学反应转化为SO2。 七十年代初期,B.D.Holt等人提出直接加热分解BaSO4制备SO2的方法。  相似文献   

3.
郭建 《铀矿地质》2016,(2):128-128
正铀广泛分布于地球的各种地质体中,由于其化学性质活泼,通常以+3、+4、+5、+6价态的化合物状态存在,其中+4和+6价铀化合物稳定。四价态铀通常以晶质铀矿(或沥青铀矿)(UO2)形成于岩浆、热液、沉积和变质作用产物中,而六价态通常以铀酰离子(UO2+2)化合物溶于水体中或在沉积、蒸发和氧化的条件下形成硫酸盐、碳酸  相似文献   

4.
锂型蒙脱石——锂的快离子导体   总被引:1,自引:0,他引:1       下载免费PDF全文
快离子导体的离子电导率很高(大于1×10-3Ω-1&3183;cm-1),电子电导率很低(小于总电导率的1%)。它的一个重要应用是作固体电池的电解质隔膜。已知的离子导体不下千种,而快离子导体却为数甚少,锂的快离子导体则几乎尚未发现。从研制高能量密度电池的角度考虑,锂的快离子导体是关键材料,这是近几年来对探索锂的快离子导体给予极大重视的原因。在锂型快离子导体中,Li3N是研究得最多的材料,室温下多晶片的离子电导率为6.6×10-4Ω1&3183;cm-1(B.A.Boukamp and R.A.Huggins,1978),但分解电压太低(~0.45V),稳定性差,不可能在电池中应用它。  相似文献   

5.
三叠纪海的硫同位素   总被引:10,自引:0,他引:10       下载免费PDF全文
前言自Ault和Kulp(1959)发表第一批海相硫酸盐岩的硫同位素分析数据以来,海相硫酸盐岩(石膏、硬石膏)的硫同位素研究工作已进行了二十余年。  相似文献   

6.
研究矿物的晶体结构与其光学性质之间的关系,这是结构光性矿物学重要的研究内容。如何把矿物晶体结构中原子的性质及相互之间的关系应用到晶体光学中,有效地定量地解释矿物的光学参数,从而又推动晶体光学进一步发展,这是矿物学家所共同关心的问题。早在1924年布拉格就注意到方解石与文石光性的差异,并用晶体内部电场的理论进行解释。帕伯斯特(A.Pabst,1973,1974)在研究两个碳酸盐矿物的折射率时,根据布拉格的方法,计算了它们的折射度,得到了满意的结果。叶大年(1974)在论述具有氖型结构离子的化合物的折射率时,指出矿物的折射率取决于电子层结构和核电荷,而不是取决于原子量。  相似文献   

7.
硫酸盐热化学还原作用的启动机制研究   总被引:1,自引:0,他引:1  
硫酸盐热化学还原作用(TSR)是导致高含硫化氢天然气生成和聚集、碳酸岩盐储层酸化和溶蚀的重要因素,是地质盆地内烃-水-岩三者之间的复杂反应。本文利用黄金管热模拟实验,对TSR反应的可能启动机制及控制因素进行了研究。通过不同盐溶液与原油的热解实验,证实了硫酸盐的存在是启动TSR反应的必要因素,MgSO4比CaSO4和Na2SO4更容易启动TSR反应,体系中盐度的增加会加速H2S的生成。实验结果表明,不同水介质条件下,TSR反应的程度与溶液的离子强度呈正相关,弱酸性环境并不足以启动TSR反应;原油中不稳定含硫化合物的含量越高越有利于TSR反应的发生,饱和链烷烃比原油中其它组分更容易引发TSR反应,且大分子烷烃比小分子烷烃更容易被硫酸盐氧化。  相似文献   

8.
三塘湖盆地上二叠统芦草沟组烃源岩地球化学特征   总被引:3,自引:0,他引:3  
三塘湖盆地上二叠统芦草沟组分布面积广、厚度大、有机碳含量高,具好的生烃潜力,是盆地内主力烃源岩。烃源岩样品可溶有机质中含有丰富的脂肪酸、脂肪酸酯等含羰基化合物,随着埋深增加,饱和烃和芳烃含量递增,非烃含量和有机碳则呈递减趋势,表明含羰基化合物对源岩生烃有较大的贡献。饱和烃的各项指标显示,烃源岩的沉积环境为咸化湖相还原环境,母质输入以水生生物为主,陆源物质输入较少。诸多证据表明,在沉积初期和成岩早期细菌输入较多,细菌输入有机质构成了原始有机质的一部分,细菌改造原始有机质并产生前述含羰基化合物,在热作用和微生物作用下均可成烃。  相似文献   

9.
一、络合物在地球化学上的意义络合物化学的研究,自本世纪20-30年代以来由于新兴技术的需要而得到迅速的发展。在实验室中为了促使某些物理化学性质相类似的成对元素得以分离,往往可以利用络合物的特性从而取得成效。因此当制备和提纯无机化合物时,应用络合物化学原理发挥了一定的作用。例如锆同铪,铌同钽等元素都由于其各形成络合物以后,显然从根本上改变了它们的性质,因此得到分离。事实上在地球化学作用中,几乎周期系内许多元素都具有络合物的形式,反映了络合物在成矿成岩过程中的变化,同样是十分重要的。有关这一方面的探索,将有助于深入了解地质学上的某些重要问题。从络合物的形成理论上认识,其构成的内配位层和外配位层的离子,对于中心离子间的关系各有不同,从而使它们相互之间的联系与结合情况也各有差异。外配位层的离子  相似文献   

10.
长江口盐沼硫酸盐还原菌的分布特征与环境机制   总被引:3,自引:0,他引:3  
陈庆强  唐媛  杨艳  谢冰  吕宝一 《沉积学报》2012,30(6):1088-1098
选择长江口崇明东滩两类植被区(互花米草、土著植被)纵向剖面,根据不同高程部位柱样土壤和植被根际硫酸盐还原菌(SRB)与异养菌的数量,颗粒有机碳(POC)含量与δ13C值,孔隙水Cl-/SO2-   4摩尔比等,研究盐沼SRB的空间分布特征与机制。自高潮滩向光滩,柱样氧含量降低,SO2-   4含量增加,导致SRB含量增加,SRB在有机质矿化中的作用增强。高潮滩柱样不同深度层位的异养菌数量明显大于中潮滩和光滩柱样的相应层位,这与由陆向海柱样氧含量及POC含量降低有关。SRB对植被根际环境的变化较敏感,芦苇根际最有利于SRB生长,藨草根际次之;互花米草根际不利于SRB生长,根系分泌物可能对SRB有抑制作用。土壤有机质含量是导致不同纵向剖面相同高程部位柱样之间微生物数量差异的主要因素。同一纵向剖面不同高程部位柱样之间在土壤氧含量与SO2-   4含量方面的不同,导致这些柱样微生物数量存在差异。盐沼柱样孔隙水Cl-/SO2-4摩尔比总体偏低,多小于19.33,指示有非海源SO2- 4加入。非海源SO2-4的加入掩盖了硫酸盐还原作用对SO2-4的消耗,增加了利用孔隙水Cl-/SO2-4摩尔比定量研究盐沼硫酸盐还原作用强度的难度。柱样中SRB越多,Cl-/SO2-4摩尔比就越大,硫酸盐还原作用就越明显。  相似文献   

11.
 绝大部分的亚硝酸盐和氰化物属于钙钛矿型结构,但有部分化合物为“反”钙钛矿型结构,充填空洞A位的阳离子的拓扑体积为0。阴离子团NO2-和CN-的拓扑体积都近乎常数,分别为41.73Å3和40.03Å3,标准方差σn-1=1.516和σn-1=1.25。由此,进一步证明了原来的一套离子拓扑体积数值的普适性。“反”钙钛矿型结构中A位阳离子的平均半径rA>1.00Å时,三价阳离子会产生扩容效应,其扩容系数大约是其本身拓扑体积的2倍;正钙钛矿型结构中,当A位是由半径较大的Cs+充填时,Cs+也会产生扩容效应,其扩容系数约为24Å3。氰化物中结晶水分子的拓扑体积近乎常数,为24Å3或0。  相似文献   

12.
Hydrogen incorporation is critical for explaining defect energies, structure parameters and other physical characteristics of minerals and understanding mantle dynamics. This work analyzed the hydrogen complex defects in jadeite by the plane-wave pseudo-potential method based on density functional theory, and optimized the atomic positions and lattice constants in all configurations(different defective systems). Incorporation mechanisms considered for hydrogen(H) in jadeite include:(1) hydrogen incorporating with the O2 site oxygen and coexisting with M2 vacancy;(2) one H atom combined with an Al atom replacing Si in tetrahedron;(3) 4H atoms directly replacing Si in tetrahedron and(4) 3H atoms replacing Al on the M1 site. The four incorporation mechanisms mentioned above form the corresponding V_(Na)-H_i, Al_(Si)-H_i, V_(Si)-4H_i and V_(Al)-3H_i point defects. The molecular dynamics simulation to the ideal, V_(Na)-H_i, Al_(Si)-H_i, V_(Si)-4H_i and V_(Al)-3H_i point defects under the P-T conditions of 900 K, 2 GPa, the V_(Na)-H_i and Al_(Si)-H_i point defects under different pressures at T = 900 K, and Al_(Si)-H_i point defects under different temperatures at P = 3 GPa was performed to examine the preferential mode of hydrogen incorporation in jadeite by means of first-principles calculations. The calculations show that the averaged O–H bond-length in the hydrogen point defects system decreased in the order of Al_(Si)-H_i, V_(Na)-H_i, V_(Si)-4H_i and V_(Al)-3H_i. V_(Na)-H_i complex defects result in a contraction of the jadeite volume and the presence of Al_(Si)-H_i, V_(Si)-4H_i and V_(Al)-3H_i defects could increase the supercell volume, which is the most obvious in the V_(Al)-3H_i defects. The energy of formation of Al_(Si)-H_i and V_(Al)-3H_i complex defects was much lower than that of other defect systems. The V_(Al)-3H_i defects system has the lowest energy and the shortest O–H bond-length, suggesting that this system is the most favorable. The analytical results of vacancy formation energy, O–H bondlength, and the stability of the hydrogen defects in jadeite have suggested that the preferred hydration incorporation mode in jadeite is V_(Al)-3H_i complex defect.  相似文献   

13.
K-substituted gonnardite, K2.18Na0.04Ca0.02[Al2.26Si2.74O10]·2.2H2O, was studied by X-ray powder diffraction method. The structure was refined with the Rietveld technique in the tetragonal space group $I\overline{4} 2d$ with a = 13.65409(16), c = 6.56928(11) Å, V = 1224.74(2) Å3, Z = 4. Most of K+ cations (1.94 apfu) statistically occupy three nearest positions to be considered as the split one. “Excess” cations are located in the position non-typical for K+. Statistics in the cation distribution is defined by the occupation of the additional position. Based on a crystal chemical positional model (C2R2A2) [T5O10], the separation of the local water–cation assemblages from an average statistical pattern has been suggested.  相似文献   

14.
Using single-crystal X-ray diffraction at 293, 200 and 100 K, and neutron diffraction at 50 K, we have refined the positions of all atoms, including hydrogen atoms (previously undetermined), in the structure of coquimbite ( $ P {\bar 3}1c $ , a?=?10.924(2)/10.882(2) Å, c?=?17.086(3) / 17.154(3) Å, V?=?1765.8(3)/1759.2(5) Å3, at 293 / 50 K, respectively). The use of neutron diffraction allowed us to determine precise and accurate hydrogen positions. The O–H distances in coquimbite at 50 K vary between 0.98 and 1.01 Å. In addition to H2O molecules coordinated to the Al3+ and Fe3+ ions, there are rings of six “free” H2O molecules in the coquimbite structure. These rings can be visualized as flattened octahedra with the distance between oxygen and the geometric center of the polyhedron of 2.46 Å. The hydrogen-bonding scheme undergoes no changes with decreasing temperature and the unit cell shrinks linearly from 293 to 100 K. A review of the available data on coquimbite and its “dimorph” paracoquimbite indicates that paracoquimbite may form in phases closer to the nominal composition of Fe2(SO4)3·9H2O. Coquimbite, on the other hand, has a composition approximating Fe1.5Al0.5(SO4)3·9H2O. Hence, even a “simple” sulfate Fe2-x Al x (SO4)3·9H2O may be structurally rather complex.  相似文献   

15.
Molecular orbital calculations have been completed on sulfate monomers and a dimer in a determination of minimum-energy geometries and electron density distributions. SO bond lengths calculated for the monomer and dimer correlate linearly with the fractional s-characters of the bonds, as observed for sulfate groups in crystals. With increasing coordination number of S, the bonded radii of S and O, as determined from electron density maps, increase at the same rate. This is at variance with the assumption that the radius of the oxide ion is nearly constant and that bond length variations arise primarily from changes in cation radii. The dimer shows a relatively large change in energy as its SOS angle is deformed from its minimum-energy value (125.6°) to 180°, in conformity with the small variation among observed angles. This is in contrast to the wide variation of bridging angles observed for silicate and phosphate dimers in crystals and molecules, and may imply that significant differences should be expected in the behavior of sulfates with respect to polymorphism and glass formation. The reaction energy of SO3 + H2O → H2SO4, calculated with second-order Møller-Plesset perturbation theory, agrees with the experimental value. Other properties of H2SO4 are also calculated and compared with experimental observations and previous calculations.  相似文献   

16.
Meridianiite, MgSO4·11H2O, is the most highly hydrated phase in the binary MgSO4–H2O system. Lower hydrates in the MgSO4–H2O system have end-member analogues containing alternative divalent metal cations (Ni2+, Zn2+, Mn2+, Cu2+, Fe2+, and Co2+) and exhibit extensive solid solution with MgSO4 and with one another, but no other undecahydrate is known. We have prepared aqueous MgSO4 solutions doped with these other cations in proportions up to and including the pure end-members. These liquids have been solidified into fine-grained polycrystalline blocks of metal sulfate hydrate + ice by rapid quenching in liquid nitrogen. The solid products have been characterised by X-ray powder diffraction, and the onset of partial melting has been quantified using a thermal probe. We have established that of the seven end-member metal sulfates studied, only MgSO4 forms an undecahydrate; ZnSO4 forms an orthorhombic heptahydrate (synthetic goslarite), MnSO4, FeSO4, and CoSO4 form monoclinic heptahydrates (syn. mallardite, melanterite, bieberite, respectively), and CuSO4 crystallises as the well-known triclinic pentahydrate (syn. chalcanthite). NiSO4 forms a new hydrate which has been indexed with a triclinic unit cell of dimensions a = 6.1275(1) Å, b = 6.8628(1) Å, c = 12.6318(2) Å, α = 92.904(2)°, β = 97.678(2)°, and γ = 96.618(2)°. The unit-cell volume of this crystal, V = 521.74(1) Å3, is consistent with it being an octahydrate, NiSO4·8H2O. Further analysis of doped specimens has shown that synthetic meridianiite is able to accommodate significant quantities of foreign cations in its structure; of the order 50 mol. % Co2+ or Mn2+, 20–30 mol. % Ni2+ or Zn2+, but less than 10 mol. % of Cu2+ or Fe2+. In three of the systems we examined, an ‘intermediate’ phase occurred that differed in hydration state both from the Mg-bearing meridianiite end-member and the pure dopant end-member hydrate. In the case of CuSO4, we observed a melanterite-structured heptahydrate at Cu/(Cu + Mg) = 0.5, which we identify as synthetic alpersite [(Mg0.5Cu0.5)SO4·7H2O)]. In the NiSO4- and ZnSO4-doped systems we characterised an entirely new hydrate which could also be identified to a lesser degree in the CuSO4- and the FeSO4-doped systems. The Ni-doped substance has been indexed with a monoclinic unit-cell of dimensions a = 6.7488(2) Å, b = 11.9613(4) Å, c = 14.6321(5) Å, and β = 95.047(3)°, systematic absences being indicative of space-group P21/c with Z = 4. The unit-cell volume, V = 1,176.59(5) Å3, is consistent with it being an enneahydrate [i.e. (Mg0.5Ni0.5)SO4·9H2O)]. Similarly, the new Zn-bearing enneahydrate has refined unit cell dimensions of a = 6.7555(3) Å, b = 11.9834(5) Å, c = 14.6666(8) Å, β = 95.020(4)°, V = 1,182.77(7) Å3, and the new Fe-bearing enneahydrate has refined unit cell dimensions of a = 6.7726(3) Å, b = 12.0077(3) Å, c = 14.6920(5) Å, β = 95.037(3)°, and V = 1,190.20(6) Å3. The observation that synthetic meridianiite can form in the presence of, and accommodate significant quantities of other ions increases the likelihood that this mineral will occur naturally on Mars—and elsewhere in the outer solar system—in metalliferous brines.  相似文献   

17.
The crystal structure of α-CaSi2O5 synthesized at conditions of 1500°C and 10 GPa, has been solved and refined in centrosymmetric space group P , using single crystal X-ray diffraction data. The composition (Z=4) and unit cell are Ca1.02Si1.99O5 by EPMA analysis and a=7.243(2) Å, b=7.546(4) Å, c=6.501(4) Å, α=81.43(5)°, β=84.82(4)°, γ=69.60(3)°, V=329.5(3) Å3, yielding the density value, 3.55 g/cm3. The structure is closely related to that of titanite, CaTiSiO5 and features the square-pyramid five-fold coordination of silicon by oxygen. The ionic radius for five-coordinated Si calculated from the bond distances is 0.33 Å. The substantial deviation of valence sum for Ca indicates the existence of local strain and the instability of α-CaSi2O5 at room pressure.  相似文献   

18.
The crystal structure of a new compound, [(H5O2)(H3O)(H2O)][(UO2)(SeO4)2] (monoclinic, P21/n a = 8.3105(15), b = 11.0799(14), c = 13.227(2) Å, β = 103.880(13)°, V = 1182.4(3) Å3), has been solved by direct methods and refined to R 1 = 0.036. The structure is based on [(UO2)(SeO4)2]2? sheet complexes formed by corner-shared UO7 pentagonal bipyramids and SeO4 tetrahedrons. The sheets are parallel to the ( $ \bar 1 The crystal structure of a new compound, [(H5O2)(H3O)(H2O)][(UO2)(SeO4)2] (monoclinic, P21/n a = 8.3105(15), b = 11.0799(14), c = 13.227(2) ?, β = 103.880(13)°, V = 1182.4(3) ?3), has been solved by direct methods and refined to R 1 = 0.036. The structure is based on [(UO2)(SeO4)2]2− sheet complexes formed by corner-shared UO7 pentagonal bipyramids and SeO4 tetrahedrons. The sheets are parallel to the (01) plane. Oxonium ions and water molecules forming [(H3O)·(H2O)·(H5O2)]2+ complexes are interlayer. Among minerals, the existence of (H5O2)+ has been unambiguously confirmed only in rhomboclase, (H5O2)+[Fe2(SO4)2(H2O)2]. Original Russian Text ? S.V. Krivovichev, 2008, published in Zapiski Rossiiskogo Mineralogicheskogo Obshchestva, 2008, No. 2, pp. 123–130.  相似文献   

19.
Thaumasite, Ca3Si(OH)6(CO3)(SO4)12H2O, occurs as a low-temperature secondary alteration phase in mafic igneous and metamorphic rocks, and is recognized as a product and indicator of sulfate attack in Portland cement. It is also the only mineral known to contain silicon in six-coordination with hydroxyl (OH)? that is stable at ambient PT conditions. Thermal expansion of the various components of this unusual structure has been determined from single-crystal X-ray structure refinements of natural thaumasite at 130 and 298 K. No phase transitions were observed over this temperature range. Cell parameters at room temperature are: a= 11.0538(6) Å, c=10.4111(8) Å and V=1101.67(10) Å3, and were measured at intervals of about 50 K between 130 and 298 K, resulting in mean axial and volumetric coefficients of thermal expansion (×10?5K?1); α a =1.7(1), α c =2.1(2), and α V =5.6(2). Although the unit cell and VIIICaO8 polyhedra show significant positive thermal expansion over this temperature range, the silicate octahedron, sulfate tetrahedron, and carbonate group show zero or negative thermal expansion, with α V (VISiO6) = ?0.6 ± 1.1, α V (IVSO4)=?5.8 ± 1.4, and α R (C–O)= 0.0 ± 1.8 (×10?5 K?1). Most of the thermal expansion is accommodated by lengthening of the R(O...O) hydrogen bond distances by on average 5σ, although the hydrogen bonds involving hydroxyl sites on VISi expand twice as much as those on molecular water, causing the [Ca3Si(OH)6(H2O)12]4+ columns to expand in diameter more than they move apart over this temperature range. The average Si–OH bond length of the six-coordinated Si atom 〈R(VISi–OH)〉 in thaumasite is 1.783(1) Å, being about 0.02 Å (?20σ) shorter than VISi–OH in the dense hydrous magnesium silicate, phase D, MgSi2H2O6.  相似文献   

20.
程敏清 《地质科学》1981,(3):286-290
关于褐钇铌矿族的研究,国内外积累了不少资料。其主要产在黑云母花岗岩、花岗伟晶岩、微斜长石岩、交代变质岩、蚀变花岗岩、白岗岩及花岗岩的残坡积和冲积砂中。1979年我们在进行内蒙白云鄂博铌、稀土、铁矿床物质成分的研究时,在白云石型铌、稀土矿石中发现含铈褐钇铌矿,与该区原已发现的褐铈铌矿等矿物组成了褐钇铌矿-褐铈铌矿系列。  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