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 31 毫秒
1.
The synthesis of the new binary Cs8?xSi46 (shown here) completes the series of binary alkali metal silicides with a clathrate‐I structure M8?xSi46 (M=Na, K, Rb, Cs). In contrast with the lighter homologues, Cs8?xSi46 can be prepared only at elevated pressures. The compound was obtained at 1200 °C between 2–10 GPa and the Cs content rises with applied pressure.

  相似文献   


2.
The crystal structures of Tb(Al0.15Si0.85), (Tb0.70Zr0.30)(Al0.17Si0.83) and Zr(Al0.22Si0.78) have been refined from single‐crystal X‐ray diffraction data. The three compounds crystallize with CrB‐type structures (Pearson symbol oS8, space group Cmcm): Tb(Al0.15Si0.85): a = 4.2715(5), b = 10.5595(15), c = 3.8393(5) Å; (Tb0.70Zr0.30)(Al0.17Si0.83): a = 4.163(2), b = 10.423(5), c = 3.8543(18) Å; Zr(Al0.22Si0.78): a = 3.7824(6), b = 10.0164(16), c = 3.7795(5) Å. The existence of a significant CrB‐type solid solution in the quaternary system Tb‐Zr‐Al‐Si, based on the ternary compound Tb(Al0.15Si0.85) and extending toward the solid solution based on the binary compound ZrSi in the Zr‐Al‐Si system, cannot be excluded.  相似文献   

3.
The1H NMR spectra of the hydrated monocationic forms of clinoptilolite M6[Al6Si30O72]·nH2O (M=Li, Na, K, Cs, NH4; n=12–22) and M 3 [Al6Si30O72]·nH2O (M′=Mg, Ca, Sr, Ba; n=20–24) and heulandite M8[Al8Si28O72]·21H2O (M=Na, K) are divided into three types differing in the symmetry of tensors of magnetic dipole-dipole interactions of protons in zeolite water molecules. On the basis of model calculations it is shown that water molecules in the Cs, K, and Ba forms of clinoptilolite and the K form of heulandite are ordered in structural positions. Institute of Inorganic Chemistry, Siberian Branch, Russian Academy of Sciences, Zeorex Ltd., Sofia. Translated fromZhurnal Strukturnoi Khimii, Vol. 37, No. 5, pp. 891–900, September–October, 1996. Translated by L. Smolina  相似文献   

4.
Colorless and highly air‐ and moisture‐sensitive powders of M[o‐C6H4O(OH)] with M = K, Rb, or Cs have been synthesized from reaction mixtures of the appropriate alkali metal and catechol in thf. All compounds were structurally characterized by means of powder X‐ray diffraction using the Rietveld profile refinement technique including restraints for the C—C/C—O bond distances and the C—C—C angles. The atomic arrangements of M[o‐C6H4O(OH)] (K: monoclinic P21/c; Rb/Cs: orthorhombic Pbcm) are characterized by polymeric chains of [M1[4]O2[2]η6] units connected by hydrogen bonds, thereby making up layered structures similar to the one of catechol. The coordinatively unsaturated alkali metals are forming edge‐sharing MO4 pyramids and exhibit asymmetrical η6‐interactions with the phenylene rings. The symmetry of the unit cells increases with increasing size of the cation, and this results in a decrease of the monoclinic angle from 118.5° (catechol) to 93.7° (K compound), eventually leading to orthorhombic cells for the Rb and Cs compounds.  相似文献   

5.
4‐Nitrobenzoic acid (PNBA) has proved to be a useful ligand for the preparation of metal complexes but the known structures of the alkali metal salts of PNBA do not include the rubidium salt. The structures of the isomorphous potassium and rubidium polymeric coordination complexes with PNBA, namely poly[μ2‐aqua‐aqua‐μ3‐(4‐nitrobenzoato)‐potassium], [K(C7H4N2O2)(H2O)2]n, (I), and poly[μ3‐aqua‐aqua‐μ5‐(4‐nitrobenzoato)‐rubidium], [Rb(C7H4N2O2)(H2O)2]n, (II), have been determined. In (I), the very distorted KO6 coordination sphere about the K+ centres in the repeat unit comprise two bridging nitro O‐atom donors, a single bridging carboxylate O‐atom donor and two water molecules, one of which is bridging. In Rb complex (II), the same basic MO6 coordination is found in the repeat unit, but it is expanded to RbO9 through a slight increase in the accepted Rb—O bond‐length range and includes an additional Rb—Ocarboxylate bond, completing a bidentate O,O′‐chelate interaction, and additional bridging Rb—Onitro and Rb—Owater bonds. The comparative K—O and Rb—O bond‐length ranges are 2.7352 (14)–3.0051 (14) and 2.884 (2)–3.182 (2) Å, respectively. The structure of (II) is also isomorphous, as well as isostructural, with the known structure of the nine‐coordinate caesium 4‐nitrobenzoate analogue, (III), in which the Cs—O bond‐length range is 3.047 (4)–3.338 (4) Å. In all three complexes, common basic polymeric extensions are found, including two different centrosymmetric bridging interactions through both water and nitro groups, as well as extensions along c through the para‐related carboxylate group, giving a two‐dimensional structure in (I). In (II) and (III), three‐dimensional structures are generated through additional bridges involving the nitro and water O atoms. In all three structures, the two water molecules are involved in similar intra‐polymer O—H...O hydrogen‐bonding interactions to both carboxylate and water O‐atom acceptors. A comparison of the varied coordination behaviour of the full set of Li–Cs salts with 4‐nitrobenzoic acid is also made.  相似文献   

6.
The researches of all‐metal aromatic clusters have been a thermic theme in inorganic aromaticity domain both experimentally and theoretically since the Al4L? (L = Li, Na, Cu) clusters were created by laser vaporization. In systemic determination of the lowest structures of 20 gaseous all‐metal aromatic clusters M4L2 (M = Al, Ga, In, Tl; L = Li, Na, K, Rb, Cs), the isomer energy differences of four low‐lying structures of each cluster were evaluated at high‐quality quantum chemistry levels. Single point calculations at the coupled cluster level were performed at geometries optimized at the MP2, B3LYP, and B3PW91 levels, and harmonic frequency calculations and zero point energy corrections were implemented following optimizations at the B3LYP and B3PW91 levels. In addition to Li‐ and Na‐containing species, theoretical investigations came down to those new clusters including K, Rb, and Cs. For many clusters, the most convincing theoretical evidences indicate that the lowest structures are a square bipyramidal isomer rather than an edge‐caped square pyramidal species. A few discrepancies were addressed at the MP2, B3LYP, and B3PW91 levels in comparison with the coupled cluster results. These findings are significant because some clusters were generated by laser vaporization and served as theoretical prototypes to test the new means for assessing aromaticity. © 2009 Wiley Periodicals, Inc. Int J Quantum Chem, 2010  相似文献   

7.
Volumetric properties of poly(acrylic acid) alkali-metal salts (Li, Na, K, Rb, and Cs) with different degrees of neutralization and water contents were studied in the range from pure solid to highly concentrated solutions. The apparent partial molar volume ?2 of the polymer and the partial molar volume V1 of water were calculated from density data. The value of ?2 decreased with decreasing polymer concentration and eventually leveled off. Values of V1, which at low water contents were much smaller than that of free water, increased with increasing water content and eventually reached a constant value equivalent to that of free water, thus indicating the appearance of free water. Water contents corresponding to the appearance of free water increased in the order of Li < Na < K < Rb < Cs, differing from the usual trend of hydration numbers observed in dilute solutions. The change of the slope of the plots of V1 versus composition suggested a change in the hydration mechanism. For Li, Na, and K salts, the limiting values of V1 at very low water content is considerably smaller than the 18 cm3/mol of free water. In contrast, for Rb and Cs salts, these values were relatively large, indicating the relatively weak electrostriction effects of these ions.  相似文献   

8.
The Li, Rb and Cs complexes with the herbicide (2,4‐dichlorophenoxy)acetic acid (2,4‐D), namely poly[[aqua[μ3‐(2,4‐dichlorophenoxy)acetato‐κ3O1:O1:O1′]lithium(I)] dihydrate], {[Li(C8H5Cl2O3)(H2O)]·2H2O}n, (I), poly[μ‐aqua‐bis[μ3‐(2,4‐dichlorophenoxy)acetato‐κ4O1:O1′:O1′,Cl2]dirubidium(I)], [Rb2(C8H5Cl2O3)2(H2O)]n, (II), and poly[μ‐aqua‐bis[μ3‐(2,4‐dichlorophenoxy)acetato‐κ5O1:O1′:O1′,O2,Cl2]dicaesium(I)], [Cs2(C8H5Cl2O3)2(H2O)]n, (III), respectively, have been determined and their two‐dimensional polymeric structures are described. In (I), the slightly distorted tetrahedral LiO4 coordination involves three carboxylate O‐atom donors, of which two are bridging, and a monodentate aqua ligand, together with two water molecules of solvation. Conjoined six‐membered ring systems generate a one‐dimensional coordination polymeric chain which extends along b and interspecies water O—H...O hydrogen‐bonding interactions give the overall two‐dimensional layers which lie parallel to (001). In hemihydrate complex (II), the irregular octahedral RbO5Cl coordination about Rb+ comprises a single bridging water molecule which lies on a twofold rotation axis, a bidentate Ocarboxy,Cl‐chelate interaction and three bridging carboxylate O‐atom bonding interactions from the 2,4‐D ligand. A two‐dimensional coordination polymeric layer structure lying parallel to (100) is formed through a number of conjoined cyclic bridges, including a centrosymmetric four‐membered Rb2O2 ring system with an Rb...Rb separation of 4.3312 (5) Å. The coordinated water molecule forms intralayer aqua–carboxylate O—H...O hydrogen bonds. Complex (III) comprises two crystallographically independent (Z′ = 2) irregular CsO6Cl coordination centres, each comprising two O‐atom donors (carboxylate and phenoxy) and a ring‐substituted Cl‐atom donor from the 2,4‐D ligand species in a tridentate chelate mode, two O‐atom donors from bridging carboxylate groups and one from a bridging water molecule. However, the two 2,4‐D ligands are conformationally very dissimilar, with one phenoxyacetate side chain being synclinal and the other being antiperiplanar. The minimum Cs...Cs separation is 4.4463 (5) Å. Structure extension gives coordination polymeric layers which lie parallel to (001) and are stabilized by intralayer water–carboxylate O—H...O hydrogen bonds.  相似文献   

9.
Seven 1,4‐phenylenebisphosphonates of monovalent ions, A(HO3PC6H4PO3H2) (A = Li, K, Rb, Cs, Tl, Ag and NH4), were synthesized and characterized by single‐crystal X‐ray diffraction, spectroscopic and thermal methods. These compounds and the reported sodium analogue have four structure types. The sodium compound, one‐dimensional lithium compound and pillared‐layered cesium compounds have different structure types, whereas the potassium, rubidium, thallium, ammonium and silver compounds have a pillared ladder‐like structure. They undergo initial thermal decomposition in the range of 120–270 °C. Moreover, the single crystal X‐ray structure of 1,4‐phenylenebisphosphonic acid was determined.  相似文献   

10.
This survey analyzes the results of studies into Li(Na,K,Rb,Cs)/W/Mn/SiO2 composites, which are used as catalysts for oxidative coupling of methane (OCM). The focus is on phase states. Our analysis shows that the SiO2 matrix is an active constituent of the composites and not an inert carrier of additives and the OCM heterogeneous process involves alkali-metal tungstate melts, along with polycrystalline manganese oxides. The effects of the cation ratio and synthetic routes on the phase composition of Li(Na,K,Rb,Cs)/W/Mn/SiO2 are assessed.  相似文献   

11.
M[m‐C6H4O(OH)] (M = Li—Cs) have been obtained as highly air‐ and moisture‐sensitive powders from reaction mixtures of the appropriate alkali metals and resorcinol in thf. Both the potassium and rubidium compounds were structurally characterized by means of powder X‐ray diffraction using the Simulated Annealing method and the Rietveld profile refinement technique including C—C/C—O bond distance and C—C—C angle restraints. K[m‐C6H4O(OH)] (orthorhombic P212121) forms infinite alternating chains of meta‐hydroxyphenolate anions connected by K—O bonds and short charge‐assisted hydrogen bonds, thereby generating a three‐dimensional network of corrugated layers similar to the structure of pure resorcinol. The potassium cations are surrounded by a triangle of oxygen and, moreover, coordinated by six adjacent phenylene rings to form a distorted octahedron. The complex crystal structure of Rb[m‐C6H4O(OH)] (monoclinic Pa) is characterized by layers of hydrogen‐bonded meta‐hydroxyphenolate triple units separated by corrugated rubidium layers. The three crystallographically different Rb atoms are coordinated by three, four, and five oxygens with irregular polyhedra, and the rubidiums are also involved in further electrostatic interactions by up to eight phenylene rings.  相似文献   

12.
Disorder in Smectites in Dependence of the Interlayer Cation Fluorosmectites, [M0.5]inter[Mg2.5Li0.5]oct[Si4]tetO10F2 (M = Na, K, Rb, Cs), have been synthesised from the melt in gastight Mo crucibles. At the same layer charge of x = 0.5, which lies within the range of smectites, both, the crystallite size and the stacking order increases with the size of the interlayer cation. For microcrystalline Na‐hectorite both rotational (n120° and n60°) and translational (±0.145b) planar defects were identified, whereas for K‐hectorite only ±b/3 translational defects were found. Finally, Rb‐ and Cs‐hectorite show a normal Bragg‐type diffraction pattern. For Cs‐hectorite even single crystals may be found that display no diffuse scattering and allow a structure refinement (monoclinic, 1M‐polytype, C2/m, a = 5.2401(10)Å, b = 9.0942(10) Å, c = 10.7971(10)Å, β = 99.21 (2)°, V = 507.90(12) Å3, Z = 2). These 3D ordered smectites still show satisfactory intracrystalline reactivity and the interlayer cations may readily be exchanged for organocations.  相似文献   

13.
The behavior of Li-exchanged natrolite Li1.92Na0.10[Al2.02Si2.98O10]?2H2O at compression in penetrating (water-containing) medium was studied by in situ synchrotron powder diffraction in diamond anvil cell up to 2.5 GPa. Within 0-1.3 GPa the compression is almost isotropic, and upon the further pressure increase the sample undergoes additional hydration, leading to abrupt volume expansion by 22%, a record value for natrolite. In the proposed model for the high-pressure phase Li2[Al2Si3O10]?6H2O the Li+ cations have no contact with the framework O-atoms and are surrounded by “water-jacket” in the form of semi-octahedron (tetragonal pyramid) composed of five H2O molecules. Such polyhedra, lining up along the channel axis, are joined through their edges and create a “water” column expanding the structure.  相似文献   

14.
A series of M? Pd? Me10CB[5] (M=Li, Na, K, Rb, and Cs; Me10CB[5]=decamethylcucurbit[5]uril) hybrid solid materials have been successfully synthesized for the first time through a simple diffusion method. These as‐prepared hybrid solids have been applied as phosphine‐free precatalysts for Heck cross‐coupling reactions with excellent catalytic performance and good recyclability. In the processes of the catalytic reactions, the activated PdII species were released from the crystalline hybrid precatalysts and transformed into catalytically active Pd nanoparticles, which have been demonstrated as key to carry on the catalytic reactions for the recoverable precatalysts M? Pd? Me10CB[5] (M=K, Rb, and Cs). It has also been rationalized that the introduction of different alkali metals afforded crystalline hybrid precatalysts with different crystal structures, which are responsible for their diversified stability and reusability presented in Heck reactions.  相似文献   

15.
In (1,4,7,10,13,16‐hexaoxacyclooctadecane)rubidium hexachloridoantimonate(V), [Rb(C12H24O6)][SbCl6], (1), and its isomorphous caesium {(1,4,7,10,13,16‐hexaoxacyclooctadecane)caesium hexachloridoantimonate(V), [Cs(C12H24O6)][SbCl6]}, (2), and ammonium {ammonium hexachloridoantimonate(V)–1,4,7,10,13,16‐hexaoxacyclooctadecane (1/1), (NH4)[SbCl6]·C12H24O6}, (3), analogues, the hexachloridoantimonate(V) anions and 18‐crown‐6 molecules reside across axes passing through the Sb atoms and the centroids of the 18‐crown‐6 groups, both of which coincide with centres of inversion. The Rb+ [in (1)], Cs+ [in (2)] and NH4+ [in (3)] cations are situated inside the cavity of the 18‐crown‐6 ring; they are situated on axes and are equally disordered about centres of inversion, deviating from the centroid of the 18‐crown‐6 molecule by 0.4808 (13), 0.9344 (7) and 0.515 (8) Å, respectively. Interaction of the ammonium cation and the 18‐crown‐6 group is supported by three equivalent hydrogen bonds [N...O = 2.928 (3) Å and N—H...O = 162°]. The centrosymmetric structure of [Cs(18‐crown‐6)]+, with the large Cs+ cation approaching the centre of the ligand cavity, is unprecedented and accompanied by unusually short Cs—O bonds [2.939 (2) and 3.091 (2) Å]. For all three compounds, the [M(18‐crown‐6)]+ cations and [SbCl6] anions afford linear stacks along the c axis, with the cationic complexes embedded between pairs of inversion‐related anions.  相似文献   

16.
Al‐ and Ga‐containing open‐Dawson polyoxometalates (POMs), K10[{Al4(μ‐OH)6}{α,α‐Si2W18O66}] · 28.5H2O ( Al4 ‐ open ) and K10[{Ga4(μ‐OH)6}(α,α‐Si2W18O66)] · 25H2O ( Ga4 ‐ open ) were synthesized by the reaction of trilacunary Keggin POM, [A‐α‐SiW9O34]10–, with Al(NO3)3 · 9H2O or Ga(NO3)3 · nH2O, and unequivocally characterized by single‐crystal X‐ray analysis, 29Si and 183W NMR, and FT‐IR spectroscopy as well as elemental analysis and TG/DTA. Single‐crystal X‐ray analysis revealed that the {M4(μ‐OH)6}6+ (M = Al, Ga) clusters were included in an open pocket of the open‐Dawson polyanion, [α,α‐Si2W18O66]16–, which was constituted by the fusion of two trilacunary Keggin POMs via two W–O–W bonds. These two open‐Dawson structural POMs showed clear difference of the bite angles depending on the size of ionic radii. In cases of both compounds, the solution 29Si and 183W NMR spectra in D2O showed only one signal and five signals, respectively. These spectra were consistent with the molecular structures of Al4 ‐ and Ga4 ‐ open , suggesting that these polyoxoanions were obtained as single species and maintained their molecular structures in solution.  相似文献   

17.
The two‐dimensional polymeric structures of the caesium complexes with the phenoxyacetic acid analogues (4‐fluorophenoxy)acetic acid, (3‐chloro‐2‐methylphenoxy)acetic acid and the herbicidally active (2,4‐dichlorophenoxy)acetic acid (2,4‐D), namely poly[[μ5‐(4‐fluorophenoxy)acetato][μ4‐(4‐fluorophenoxy)acetato]dicaesium], [Cs2(C8H6FO3)2]n, (I), poly[aqua[μ5‐(3‐chloro‐2‐methylphenoxy)acetato]caesium], [Cs(C9H8ClO3)(H2O)]n, (II), and poly[[μ7‐(2,4‐dichlorophenoxy)acetato][(2,4‐dichlorphenoxy)acetic acid]caesium], [Cs(C8H5Cl2O3)(C8H6Cl2O3)]n, (III), are described. In (I), the Cs+ cations of the two individual irregular coordination polyhedra in the asymmetric unit (one CsO7 and the other CsO8) are linked by bridging carboxylate O‐atom donors from the two ligand molecules, both of which are involved in bidentate chelate Ocarboxy,Ophenoxy interactions, while only one has a bidentate carboxylate O,O′‐chelate interaction. Polymeric extension is achieved through a number of carboxylate O‐atom bridges, with a minimum Cs...Cs separation of 4.3231 (9) Å, giving layers which lie parallel to (001). In hydrated complex (II), the irregular nine‐coordination about the Cs+ cation comprises a single monodentate water molecule, a bidentate Ocarboxy,Ophenoxy chelate interaction and six bridging carboxylate O‐atom bonding interactions, giving a Cs...Cs separation of 4.2473 (3) Å. The water molecule forms intralayer hydrogen bonds within the two‐dimensional layers, which lie parallel to (100). In complex (III), the irregular centrosymmetric CsO6Cl2 coordination environment comprises two O‐atom donors and two ring‐substituted Cl‐atom donors from two hydrogen bis[(2,4‐dichlorophenoxy)acetate] ligand species in a bidentate chelate mode, and four O‐atom donors from bridging carboxyl groups. The duplex ligand species lie across crystallographic inversion centres, linked through a short O—H...O hydrogen bond involving the single acid H atom. Structure extension gives layers which lie parallel to (001). The present set of structures of Cs salts of phenoxyacetic acids show previously demonstrated trends among the alkali metal salts of simple benzoic acids with no stereochemically favourable interactive substituent groups for formation of two‐dimensional coordination polymers.  相似文献   

18.
The thermal behaviour of alkali metal hexachlorostannates of general formula A2SnCl6 (A=Li, Na, K, Cs, Rb) has been studied by thermoanalytical methods (dynamic TG, DTG and DTA, and Q-TG). The compounds decompose upon heating with total (A=Li, Na) or partial (A=K, Rb, Cs) release of SnCl4 molecules to the gaseous phase. The enthalpies of the decomposition process have been evaluated on the basis of the Van't Hoff equation and using dynamic TG curves. The thermochemistry of the compounds was thoroughly examined particularly, with regard to the enthalpies of formation and crystal lattice energies, as well as enthalpy changes characterizing the dissociation process. These data revealed that stability of the compounds markedly increase with an increase in the size of alkali metal cation.
Zusammenfassung Mittels thermoanalytischer Methoden (dynamische TG, DTG, DTA, Q-TG) wurde das thermische Verhalten von Alkalimetall-hexachlorostannaten der allgemeinen Formel A2SnCl6 (A=Li, Na, K, Cs, Rb) untersucht. Die Verbindungen zersetzen sich beim Erhitzen unter totaler (A=Li, Na) oder partieller (A=K, Rb, Cs) Freisetzung von SnC4-Molekülen, die in die Gasphase übergehen. Die Enthalpien des Zersetzungsprozesses wurden auf der Grundlage der Van't Hoffschen Gleichung unter Zuhilfenahme dynamischer TG-Kurven ermittelt. Die Thermochemie dieser Verbindungen wurde gründlich untersucht, in besonderem Hinblick auf Bildungsenthalpie und Kristallgitterenergie als auch auf Enthalpieänderungen, die den Dissoziationsprozess näher beschreiben. Die Untersuchungen zeigen, daß die Stabilität der Verbindungen mit anwachsender Alkalimetall-Kationengröße zunimmt.
  相似文献   

19.
In the title monohydrated cocrystal, namely 1,3‐diamino‐5‐azaniumyl‐1,3,5‐trideoxy‐cis‐inositol iodide–1,3,5‐triamino‐1,3,5‐trideoxy‐cis‐inositol–water (1/1/1), C6H16N3O3+·I·C6H15N3O3·H2O, the neutral 1,3,5‐triamino‐1,3,5‐trideoxy‐cis‐inositol (taci) molecule and the monoprotonated 1,3‐diamino‐5‐azaniumyl‐1,3,5‐trideoxy‐cis‐inositol cation (Htaci+) both adopt a chair conformation, with the three O atoms in axial and the three N atoms in equatorial positions. The cation, but not the neutral taci unit, exhibits intramolecular O—H...O hydrogen bonding. The entire structure is stabilized by a complex three‐dimensional network of intermolecular hydrogen bonds. The neutral taci entities and the Htaci+ cations are each aligned into chains along [001]. In these chains, two O—H...N interactions generate a ten‐membered ring as the predominant structural motif. The rings consist of vicinal 2‐amino‐1‐hydroxyethylene units of neighbouring molecules, which are paired via centres of inversion. The chains are interconnected into undulating layers parallel to the ac plane, and the layers are further held together by O—H...N hydrogen bonds and additional interactions with the iodide counter‐anions and solvent water molecules.  相似文献   

20.
The synthesis and crystal structure determination (at 293 K) of the title complex, Cs[Fe(C8H6BrN3OS)2], are reported. The compound is composed of two dianionic O,N,S‐tridentate 5‐bromosalicylaldehyde thiosemicarbazonate(2−) ligands coordinated to an FeIII cation, displaying a distorted octahedral geometry. The ligands are orientated in two perpendicular planes, with the O‐ and S‐donor atoms in cis positions and the N‐donor atoms in trans positions. The complex displays intermolecular N—H...O and N—H...Br hydrogen bonds, creating R44(18) rings, which link the FeIII units in the a and b directions. The FeIII cation is in the low‐spin state at 293 K.  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