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 15 毫秒
1.
Hexaaquamagnesium(II) sulfate pentahydrate, [Mg(H2O)6]SO4·5H2O, and hexaaquamagnesium(II) chromate(II) pentahydrate, [Mg(H2O)6][CrO4]·5H2O, are isomorphous, being composed of hexaaquamagnesium(II) octahedra, [Mg(H2O)6]2+, and sulfate (chromate) tetrahedral oxyanions, SO42− (CrO42−), linked by hydrogen bonds. There are two symmetry‐inequivalent centrosymmetric octahedra: M1 at (0, 0, 0) donates hydrogen bonds directly to the tetrahedral oxyanion, T1, at (0.405, 0.320, 0.201), whereas the M2 octahedron at (0, 0, ) is linked to the oxyanion via five interstitial water molecules. Substitution of CrVI for SVI leads to a substantial expansion of T1, since the Cr—O bond is approximately 12% longer than the S—O bond. This expansion is propagated through the hydrogen‐bonded framework to produce a 3.3% increase in unit‐cell volume; the greatest part of this chemically induced strain is manifested along the b* direction. The hydrogen bonds in the chromate compound mitigate ∼20% of the expected strain due to the larger oxyanion, becoming shorter (i.e. stronger) and more linear than in the sulfate analogue. The bifurcated hydrogen bond donated by one of the interstitial water molecules is significantly more symmetrical in the chromate analogue.  相似文献   

2.
A rubidium titanyl arsenate single‐crystal has been studied by neutron diffraction (λ = 1.207 Å). The polished sample used was 5 × 3 × 2 mm and was cut from a crystal made by top‐seeded solution growth. The crystal showed severe extinction. It was, however, possible to obtain a structural model with well defined oxy­gen sites and reasonable anisotropic displacement parameters.  相似文献   

3.
The polar crystal structure of diammonium [octaoxidoditellurato(IV)]tungstate, (NH4)2WTe2O8, was studied at high pressures using single‐crystal X‐ray diffraction in a diamond‐anvil cell at the HASYLAB synchrotron (DESY, Hamburg, Germany). No phase transition was observed up to 7.16 GPa. However, a full structure determination at 5.09 GPa shows that the coordination number of one of the two non‐equivalent Te atoms has decreased from four to three.  相似文献   

4.
The crystal structure of synthetic penkvilksite‐2O, disodium titanium tetrasilicate dihydrate, Na2TiSi4O11·2H2O, a microporous titanosilicate, confirms the major features of a previous model that had been obtained by order–disorder (OD) theory from the known structure of penkvilksite‐1M. An important difference from the previous model involves the hydrogen bonding of the water molecule which, on the basis of a Raman spectrum and the finding of only one of the two H atoms, is proposed to be disordered about a fixed O–H direction. The structure of penkvilksite‐2O is based on (100) silicate layers linked by isolated TiO6 octahedra to form a heteropolyhedral framework. The layer is strongly corrugated, based on interlaced spiral chains, and is crossed by two different channels that have an effective channel width of about 3 Å.  相似文献   

5.
K2TiSi3O9·H2O     
Single crystals of dipotassium titanium trisilicate hydrate were synthesized and the crystal structure was refined using data from single‐crystal X‐ray diffraction. The structure is a three‐dimensional mixed framework and contains channels formed by six‐ and eight‐membered rings. K+ ions and water mol­ecules are located in the channels.  相似文献   

6.
Rubidium chromium(III) dioxalate dihydrate [di­aqua­bis(μ‐oxalato)­chromium(III)­rubidium(I)], [RbCr(C2O4)2(H2O)2], (I), and dicaesium magnesium dioxalate tetrahydrate [tetra­aqua­bis(μ‐oxalato)­magnesium(II)­dicaesium(I)], [Cs2Mg(C2­O4)2(H2O)4], (II), have layered structures which are new among double‐metal oxalates. In (I), the Rb and Cr atoms lie on sites with imposed 2/m symmetry and the unique water molecule lies on a mirror plane; in (II), the Mg atom lies on a twofold axis. The two non‐equivalent Cr and Mg atoms both show octahedral coordination, with a mean Cr—O distance of 1.966 Å and a mean Mg—O distance of 2.066 Å. Dirubid­ium copper(II) dioxalate dihydrate [di­aqua­bis(μ‐oxalato)­copper(II)­dirubidium(I)], [Rb2Cu(C2O4)2(H2O)2], (III), is also layered and is isotypic with the previously described K2‐ and (NH4)2CuII(C2O4)2·2H2O compounds. The two non‐equivalent Cu atoms lie on inversion centres and are both (4+2)‐coordinated. Hydro­gen bonds are medium‐strong to weak in the three compounds. The oxalate groups are slightly non‐planar only in the Cs–Mg compound, (II), and are more distinctly non‐planar in the K–Cu compound, (III).  相似文献   

7.
We have identified a new compound in the glycine–MgSO4–water ternary system, namely glycine magnesium sulfate trihydrate (or Gly·MgSO4·3H2O) {systematic name: catena‐poly[[tetraaquamagnesium(II)]‐μ‐glycine‐κ2O:O′‐[diaquabis(sulfato‐κO)magnesium(II)]‐μ‐glycine‐κ2O:O′]; [Mg(SO4)(C2D5NO2)(D2O)3]n}, which can be grown from a supersaturated solution at ∼350 K and which may also be formed by heating the previously known glycine magnesium sulfate pentahydrate (or Gly·MgSO4·5H2O) {systematic name: hexaaquamagnesium(II) tetraaquadiglycinemagnesium(II) disulfate; [Mg(D2O)6][Mg(C2D5NO2)2(D2O)4](SO4)2} above ∼330 K in air. X‐ray powder diffraction analysis reveals that the trihydrate phase is monoclinic (space group P21/n), with a unit‐cell metric very similar to that of recently identified Gly·CoSO4·3H2O [Tepavitcharova et al. (2012). J. Mol. Struct. 1018 , 113–121]. In order to obtain an accurate determination of all structural parameters, including the locations of H atoms, and to better understand the relationship between the pentahydrate and the trihydrate, neutron powder diffraction measurements of both (fully deuterated) phases were carried out at 10 K at the ISIS neutron spallation source, these being complemented with X‐ray powder diffraction measurements and Raman spectroscopy. At 10 K, glycine magnesium sulfate pentahydrate, structurally described by the `double' formula [Gly(d5)·MgSO4·5D2O]2, is triclinic (space group P, Z = 1), and glycine magnesium sulfate trihydrate, which may be described by the formula Gly(d5)·MgSO4·3D2O, is monoclinic (space group P21/n, Z = 4). In the pentahydrate, there are two symmetry‐inequivalent MgO6 octahedra on sites of symmetry and two SO4 tetrahedra with site symmetry 1. The octahedra comprise one [tetraaquadiglcyinemagnesium]2+ ion (centred on Mg1) and one [hexaaquamagnesium]2+ ion (centred on Mg2), and the glycine zwitterion, NH3+CH2COO, adopts a monodentate coordination to Mg2. In the trihydrate, there are two pairs of symmetry‐inequivalent MgO6 octahedra on sites of symmetry and two pairs of SO4 tetrahedra with site symmetry 1; the glycine zwitterion adopts a binuclear–bidentate bridging function between Mg1 and Mg2, whilst the Mg2 octahedra form a corner‐sharing arrangement with the sulfate tetrahedra. These bridged polyhedra thus constitute infinite polymeric chains extending along the b axis of the crystal. A range of O—H…O, N—H…O and C—H…O hydrogen bonds, including some three‐centred interactions, complete the three‐dimensional framework of each crystal.  相似文献   

8.
Tetra­ammonium disodium decavanadate decahydrate crystallizes in the triclinic system in space group P. The structure contains typical centrosymmetric OV6 double octahedra and centrosymmetric pairs of edge‐shared NaO6 double octahedra forming a layered structure. In contrast to other monovalent cationic decavanadates, the NaO6 double octahedra are integrated in the layer.  相似文献   

9.
The structure of the neutral heterometal oxide cluster dodecaaqua‐di‐μ3‐hydroxido‐deca‐μ2‐hydroxido‐octacosaoxidotetracobalt(II)dodecamolybdenum(V) dodecahydrate, [Mo12O282‐OH)103‐OH)2{Co(H2O)3}4], is virtually identical to the previously reported NiII analogue [Mo12O282‐OH)103‐OH)2{NiII(H2O)3}4] [Müller, Beugholt, Kögerler, Bögge, Budko & Luban (2000). Inorg. Chem. 39 , 5176–5177], the first molecular magnet to exhibit signs of magnetostriction. The formation kinetics of the neutral cluster species, which is insoluble in water, can be significantly slowed by the use of deuterated reactants in order to grow single crystals of sufficient size for single‐crystal X‐ray diffraction studies using standard diffractometers. One half of the main cluster and six solvent water molecules constitute the asymmetric unit. The main cluster is located on a mirror plane.  相似文献   

10.
NaHC2O4 · H2O crystallizes in space group P 1 with a0 = 6,51, b0 = 6,66, c0 = 5,70 Å, α0 = 95,0°, β0 = 109,8°, γ0 = 74,9° and Z = 2. The structure was solved by direct methods. The refinement was carried out with 679 reflections to R1 = 7,7%. The angle between the O? C? O planes is 12,6°.  相似文献   

11.
Te(OH)6 · 2Na3P3O9 · 6H2O, is hexagonal (P63/m) with a = 11,67(1), c = 12,12(1) Å, Z = 2 and Dx = 2,225 g/cm3. Te(OH)6 · K3P3O9 · 2H2O, is monoklin (P21/c) with a = 19,61(5), b = 7,456(1), c = 14,84(6) Å, = 108,01(4), Z = 4 and Dx = 2,506 g/cm3. Both compounds are the first examples of phosphate tellurates in which the anion phosphate is condensed to the ring anion P3O9. As in phosphate tellurates already described the phosphate groups are independent of the TeO6 octahedra.  相似文献   

12.
Iron(III) Complexes with Ethylene Diamine Tetraacetic Acid or Nitrilotriacetic Acid and Phenol Quarternary complexes of iron(III) with phenol (HR) and ethylene diamine tetraacetic acid (H4Y) as well as nitrilotriacetic acid (H3X) in aqueous solution could be detected. The equilibria have been controlled by spectrophotometric and electrophoretic methods. The measured optical properties and the equilibrium constants are discussed. The following particles are present : (see ?Inhaltsübersicht”?)  相似文献   

13.
The crystal structure of cobalt vanadophosphate dihydrate {systematic name: poly[diaqua‐μ‐oxido‐μ‐phosphato‐hemicobalt(II)vanadium(II)]}, Co0.50VOPO4·2H2O, shows a three‐dimensional framework assembled from VO5 square pyramids, PO4 tetrahedra and Co[O2(H2O)4] octahedra. The CoII ions have local 4/m symmetry, with the equatorial water molecules in the mirror plane, while the V and apical O atom of the vanadyl group are located on the fourfold rotation axis and the P atoms reside on sites. The PO4 tetrahedra connect the VO5 polyhedra to form a planar P–V–O layer. The [Co(H2O)4]2+ cations link adjacent P–V–O layers via vanadyl O atoms to generate an unprecedented three‐dimensional open framework. Powder diffraction measurements reveal that the framework collapses on removal of the water molecules.  相似文献   

14.
15.
Single crystals of calcium bromide enneahydrate, CaBr2·9H2O, calcium iodide octahydrate, CaI2·8H2O, calcium iodide heptahydrate, CaI2·7H2O, and calcium iodide 6.5‐hydrate, CaI2·6.5H2O, were grown from their aqueous solutions at and below room temperature according to the solid–liquid phase diagram. The crystal structure of CaI2·6.5H2O was redetermined. All four structures are built up from distorted Ca(H2O)8 antiprisms. The antiprisms of the iodide hydrate structures are connected either via trigonal‐plane‐sharing or edge‐sharing, forming dimeric units. The antiprisms in calcium bromide enneahydrate are monomeric.  相似文献   

16.
Three potassium edta (edta is ethylenediaminetetraacetic acid, H4Y) salts which have different degrees of ionization of the edta anion, namely dipotassium 2‐({2‐[bis(carboxylatomethyl)azaniumyl]ethyl}(carboxylatomethyl)azaniumyl)acetate dihydrate, 2K+·C10H14N2O82−·2H2O, (I), tripotassium 2,2′‐({2‐[bis(carboxylatomethyl)amino]ethyl}ammonio)diacetate dihydrate, 3K+·C10H13N2O83−·2H2O, (II), and tetrapotassium 2,2′,2′′,2′′′‐(ethane‐1,2‐diyldinitrilo)tetraacetate 3.92‐hydrate, 4K+·C10H12N2O84−·3.92H2O, (III), were obtained in crystalline form from water solutions after mixing edta with potassium hydroxide in different molar ratios. In (II), a new mode of coordination of the edta anion to the metal is observed. The HY3− anion contains one deprotonated N atom coordinated to K+ and the second N atom is involved in intramolecular bifurcated N—H...O and N—H...N hydrogen bonds. The overall conformation of the HY3− anions is very similar to that of the Y4− anions in (III), although a slightly different spatial arrangement of the –CH2COO groups in relation to (III) is observed, whereas the H2Y2− anions in (I) adopt a distinctly different geometry. The preferred synclinal conformation of the –NCH2CH2N– moiety was found for all edta anions. In all three crystals, the anions and water molecules are arranged in three‐dimensional networks linked via O—H...O and C—H...O [and N—H...O in (I) and (II)] hydrogen bonds. K...O interactions also contribute to the three‐dimensional polymeric architecture of the salts.  相似文献   

17.
The preparation of copper-tetrasodium trimetaphosphate tetrahydrate: Na4Cu(P3O9)2 · 4H2O is described. This salt is the first example of the occurence of CuII in a trimetaphosphate. The triclinic unit cell has the dimensions a = 7.907(5), b = 8.364(5), c = 7.122(5) Å, α = 102.46(5), β = 97.89(5), γ = 84.04(5)°. Crystal structure has been solved by using 2019 independent reflexions with a final R value 0.017. Both copper and sodium atoms are in octahedral coordination. NaO6 octahedra form ribbons running in (011) planes. These ribbons are interconnected by CuO6 octahedra as to form a three dimensional network. Hydrogen atoms have been located and refined. The hydrogen bond scheme is described.  相似文献   

18.
Ammonium cyclooctaphosphate trihydrate, (NH4)8P8O24 · 3H2O, is monoclinic, Cc, with Z = 4 and the following unit-cell dimensions: a = 24.268(14), b = 6.700(3), c = 20.586(13) Å, β = 112,06(6)° Chemical preparation and crystal structure are reported. Using 4 668 unique reflections the atomic arrangement was determined with a final R value of 0.037. The organization of ring-anions and ammonium groups is almost centrosymmetrical but the location of the water molecules is not. The ring anion itself has a strong pseudo centrosymmetric conformation.  相似文献   

19.
The title compound, tri­ammonium cis‐di­aqua‐cis‐dioxo‐trans‐disulfatovanadate 1.5‐hydrate, was obtained by oxidizing VIV to VV in a 2 M sulfuric acid solution of vanadyl­ sulfate and adding ammonium sulfate. Here, the V atom is sandwiched by two sulfate groups by corner‐sharing to form a discrete [VO2(SO4)2(OH2)2]3? anion. The water mol­ecules occupy cis positions in the equatorial plane of the vanadium octahedron.  相似文献   

20.
杨颙  张为俊  高晓明 《中国化学》2006,24(7):887-893
A theoretical study on the blue-shifted H-bond N-H…O and red-shifted H-bond O-H…O in the complexHNO…H_2O_2 was conducted by employment of both standard and counterpoise-corrected methods to calculate thegeometric structures and vibrational frequencies at the MP2/6-31G(d),MP2/6-31 G(d,p),MP2/6-311 q G(d,p),B3LYP/6-31G(d),B3LYP/6-31 G(d,p) and B3LYP/6-311 G(d,p) levels.In the H-bond N-H…O,the calcu-lated blue shift of N-H stretching frequency is in the vicinity of 120 cm~(-1) and this is indeed the largest theoreticalestimate of a blue shift in the X-H…Y H-bond ever reported in the literature.From the natural bond orbital analy-sis,the red-shifted H-bond O-H…O can be explained on the basis of the dominant role of the hyperconjugation.For the blue-shifted H-bond N-H…O,the hyperconjugation was inhibited due to the existence of significant elec-tron density redistribution effect,and the large blue shift of the N-H stretching frequency was prominently due tothe rehybridization of sp~n N-H hybrid orbital.  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