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 562 毫秒
1.
Three new alkali metal transition metal sulfate‐oxalates, RbFe(SO4)(C2O4)0.5 · H2O and CsM(SO4)(C2O4)0.5 · H2O (M = Mn, Fe) were prepared through hydrothermal reactions and characterized by single‐crystal X‐ray diffraction, solid state UV/Vis/NIR diffuse reflectance spectroscopy, infrared spectra, thermogravimetric analysis, and powder X‐ray diffraction. The title compounds all crystallize in the monoclinic space group P21/c (no. 14) with lattice parameters: a = 7.9193(5), b = 9.4907(6), c = 8.8090(6) Å, β = 95.180(2)°, Z = 4 for RbFe(SO4)(C2O4)0.5 · H2O; a = 8.0654(11), b = 9.6103(13), c = 9.2189(13) Å, β = 94.564(4)°, Z = 4 for CsMn(SO4)(C2O4)0.5 · H2O; and a = 7.9377(3), b = 9.5757(4), c = 9.1474(4) Å, β = 96.1040(10)°, Z = 4 for CsFe(SO4)(C2O4)0.5 · H2O. All compounds exhibit three‐dimensional frameworks composed of [MO6] octahedra, [SO4]2– tetrahedra, and [C2O4]2– anions. The alkali cations are located in one‐dimensional tunnels.  相似文献   

2.
Seven new mixed oxochalcogenate compounds in the systems MII/XVI/TeIV/O/(H), (MII = Ca, Cd, Sr; XVI = S, Se) were obtained under hydrothermal conditions (210 °C, one week). Crystal structure determinations based on single‐crystal X‐ray diffraction data revealed the compositions Ca3(SeO4)(TeO3)2, Ca3(SeO4)(Te3O8), Cd3(SeO4)(Te3O8), Cd3(H2O)(SO4)(Te3O8), Cd4(SO4)(TeO3)3, Cd5(SO4)2(TeO3)2(OH)2, and Sr3(H2O)2(SeO4)(TeO3)2 for these phases. Peculiar features of the crystal structures of Ca3(SeO4)(TeO3)2, Ca3(SeO4)(Te3O8), Cd3(SeO4)(Te3O8), Cd3(H2O)(SO4)(Te3O8), and Sr3(H2O)2(SeO4)(TeO3)2 are metal‐oxotellurate(IV) layers connected by bridging XO4 tetrahedra and/or by hydrogen‐bonding interactions involving hydroxyl or water groups, whereas Cd4(SO4)(TeO3)3 and Cd5(SO4)2(TeO3)2(OH)2 crystallize as framework structures. Common to all crystal structures is the stereoactivity of the TeIV electron lone pair for each oxotellurate(IV) unit, pointing either into the inter‐layer space, or into channels and cavities in the crystal structures.  相似文献   

3.
The reduction of cupric ion in ammonia solution by aqueous sulfur dioxide was studied. Each run was carried out at constant initial cupric concentration, stirred rate and total mixed gas flow rate. The effect of temperature, partial pressure of sulfur dioxide in gas phase and cupric ion concentration of the solution was investigated. The reaction of Cu3++SO2(aq.)→Cu++SO42? was carried out by bubbling mixed gas (SO2/N2) through the aqueous ammonia complex of copper (II). The color change for the system was from deep blue, green, yellow to white. The pH values in the system changed from 10 to 4. The product of the reaction was identified by the analyses of IR spectrum and X-ray diffraction, having the formula of 7Cu2SO3· 2CuSO3·3(NH4)2SO3·24H2O. The kinetic model of the reduction was proposed as: –d[Cu2+]/dt = k exp(–E/RT)[Cu2+]α[SO2%] According to the experiments, the parameters were determined as: α=1.64±0.03, þ=1.20, E=13.7 Kcal/mol and k = (1.77±0.20)×1010 (g-equ./?)?0.64min?2.  相似文献   

4.
《中国化学会会志》2017,64(1):43-54
White microcrystalline diamagnetic oxoperoxotungstate(VI) complexes K[WO(O2)2F]·H2O, K2[WO(O2)2(CO3)]·H2O, [WO(O2)(SO4)(H2O)2] have been synthesized from reaction of Na2WO4·2H2O with aqueous HF, solid KHCO3, aqueous H2SO4 (W:F 1:3; W: CO3 2 1:1; and W: SO4 2 1:3), and an excess of 30% H2O2 at pH 7.5–8. Precipitation was completed by the addition of precooled acetone. The occurrence of terminal WO and triangular bidentate O2 2 (C 2 v ) in the synthesized compounds was ascertained from IR spectra. The IR spectra also suggested that the F and SO4 2 ions in K[WO(O2)2F]·H2O and [WO(O2)(SO4)(H2O)2] were bonded to the WO +4 center in monodentate manner, while CO3 2 ion in K2[WO(O2)2(CO3)]·H2O binds the metal center in bidentate chelating fashion. The complex [WO(O2)(SO4)(H2O)2] is stable upto 110°C. The water molecule in [WO(O2)(SO4)(H2O)2] is coordinated to the WO +4 center, whereas it occurs as water of crystallization in the corresponding peroxo(fluoro) and peroxo(carbonato) compounds. Mass spectra of the compounds are in good agreement with the molecular formulae of the complexes. K2[WO(O2)2(CO3)]·H2O acts as an oxidant for bromide in the aqueous‐phase bromination of organic substrates to the corresponding bromo‐organics, and the complex also oxidizes Hantzsch‐1,4‐dihydropyridine to the corresponding pyridine derivative in excellent yield at room temperature. Density functional theory computation was carried out to compute the frequencies of relevant vibrational modes and electronic properties, and the results are in agreement with the experimentally obtained data.  相似文献   

5.
A series of manganese(II) complexes of general formula MnLn(H2O)m (where H2Ln are substituted N,N′‐bis(salicylidene)‐1,2‐diimino‐2,2‐dimethylethane) have been prepared by electrochemical synthesis and characterized by analytical and spectroscopic techniques, magnetism and by studying their redox reversibility character by cyclic and normal pulse voltammetry. The reactivity of these complexes with sulphur dioxide has been investigated in the solid state and in toluene slurries at room temperature. The studies of the reversibility of the reaction (desorption studies) by thermogravimetrical analysis (TGD) have shown a different behaviour among the SO2‐adducts (from irreversible to totally reversible fixing), pointing to different SO2 binding modes. Thus, adducts 10 , 12 and 14 , kept the SO2 after TGD, signifying S–bridged SO2 binding mode, while TGD for 8 , 9 and 13 revealed the lability of their SO2, attributable to ligand bound SO2 coordination. The manganese(II) precursor 4 is that one which has the ability of reversily fixing a major quantity of SO2 and undergoes the sulphato reaction to form 11 also.  相似文献   

6.
The mixed oxochalcogenate compounds Mg2(SO4)(TeO3)(H2O), Mg3(SO4)(TeO3)(OH)2(H2O)2, Zn2(SeO4)(TeO3), and Zn4(SO4)(TeO3)3 were obtained under hydrothermal conditions (210 °C, autogenous pressure). Structure analyses using single‐crystal X‐ray data revealed tellurium in all four compounds to be present in oxidation state +IV, whereas sulfur or selenium atoms exhibit an oxidation state of +VI. In the crystal structures of the two magnesium compounds, [MgO5(H2O)] octahedra [Mg2(SO4)(TeO3)(H2O) structure, isotypic with the Co and Mn analogues] or [MgO4(OH)2] and [MgO4(OH)2(H2O)2] octahedra [Mg3(SO4)(TeO3)(OH)2(H2O)2 structure, novel structure type] as well as trigonal‐pyramidal TeO32– anions make up metal oxotellurate sheets, which are bridged by SO42– anions. The polar crystal structure of Zn2(SeO4)(TeO3) is isotypic with Zn2(MoO4)(TeO3) and consists of [ZnO4] tetrahedra, [ZnO6] octahedra, SeO42– and TeO32– anions as principal building units that are connected into a framework structure. Such a structural arrangement, with basically the same coordination polyhedra as in Zn2(SeO4)(TeO3) but with SO42– instead of SeO42– anions, is also found in the tellurium‐rich compound Zn4(SO4)(TeO3)3 that crystallizes in a novel structure type.  相似文献   

7.
Two organic–inorganic hybrid layered materials, namely poly[(μ‐1,4‐diaminobenzene‐κ2N:N′)[μ3‐sulfato(VI)‐κ4O:O′:O′′,O′′′]manganese], [Mn(SO4)(C6H8N2)]n, 1 , and poly[(μ‐1,4‐diaminobenzene‐κ2N:N′)[μ3‐sulfato(VI)‐κ4O:O′:O′′,O′′′]copper], [Cu(SO4)(C6H8N2)]n, 2 , have been synthesized using 1,4‐phenylenediamine (PPD) as an organic template and component (linker). Both materials form three‐dimensional frameworks. The crystal structures were determined using data from powder X‐ray diffraction measurements. The purity and morphology of the compounds were studied by elemental analyses and SEM investigations, and their thermal stabilities were determined by thermogravimetric and nonambient powder X‐ray diffraction measurements, which indicated that 1 is stable up to 537 K and 2 is stable up to 437 K.  相似文献   

8.
Tungsten‐183 NMR data are reported for the complexes cis‐[W(CO)4(PPh3)(4‐RC5H4N)] (R = H, Me, Ph, COMe, COPh, OMe, NMe2, Cl, NO2). The 183W chemical shift (obtained by indirect detection using 31P) is found to correlate with the Hammett σ function for the group R, with 183W shielding increasing approximately linearly with the donor strength of the pyridine over a range of 93 ppm. The X‐ray structures of cis‐[W(CO)4(PPh3)(4‐MeOC5H4N)] and cis‐[W(CO)4(PPh3)(4‐PhCOC5H4N)] are also reported. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

9.
A series of tributyltin(IV) complexes of 2‐[(E)‐2‐(3‐formyl‐4‐hydroxyphenyl)‐1‐diazenyl]benzoic acid and 4‐[((E)‐1‐{2‐hydroxy‐5‐[(E)‐2‐(2‐carboxyphenyl)‐1‐diazenyl]phenyl}methylidene)amino]aryls have been investigated by electrospray mass spectrometry (ESI‐MS) and tandem mass spectrometry (MSn) techniques. The assignments are facilitated by agreement between observed and calculated isotopic patterns and MSn studies. Single‐crystal X‐ray crystallography of (Bu3Sn[O2CC6H4{N?N(C6H3‐4‐OH(C(H)?NC6H4OCH3‐4))}‐o])n reveals a polymeric structure. Toxicity studies of the tributyltin(IV) complexes of the 4‐[((E)‐1‐{2‐hydroxy‐5‐[(E)‐2‐(2‐carboxyphenyl)‐1‐diazenyl]phenyl}methylidene)amino]aryls on the second larval instar of the Aedes aegypti and Anopheles stephensi mosquito larvae are also reported. The LC50 values indicate that the complexes are effective larvicides, which range from a low of 0.36 ppm to a high of 0.69 ppm against the Ae. aegypti larvae and between 0.82 and 1.17 ppm against the An. stephensi larvae. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

10.
A series of 4‐X‐1‐methylpyridinium cationic nonlinear optical (NLO) chromophores (X=(E)‐CH?CHC6H5; (E)‐CH?CHC6H4‐4′‐C(CH3)3; (E)‐CH?CHC6H4‐4′‐N(CH3)2; (E)‐CH?CHC6H4‐4′‐N(C4H9)2; (E,E)‐(CH?CH)2C6H4‐4′‐N(CH3)2) with various organic (CF3SO3?, p‐CH3C6H4SO3?), inorganic (I?, ClO4?, SCN?, [Hg2I6]2?) and organometallic (cis‐[Ir(CO)2I2]?) counter anions are studied with the aim of investigating the role of ion pairing and of ionic dissociation or aggregation of ion pairs in controlling their second‐order NLO response in anhydrous chloroform solution. The combined use of electronic absorption spectra, conductimetric measurements and pulsed field gradient spin echo (PGSE) NMR experiments show that the second‐order NLO response, investigated by the electric‐field‐induced second harmonic generation (EFISH) technique, of the salts of the cationic NLO chromophores strongly depends upon the nature of the counter anion and concentration. The ion pairs are the major species at concentration around 10?3 M , and their dipole moments were determined. Generally, below 5×10?4 M , ion pairs start to dissociate into ions with parallel increase of the second‐order NLO response, due to the increased concentration of purely cationic NLO chromophores with improved NLO response. At concentration higher than 10?3 M , some multipolar aggregates, probably of H type, are formed, with parallel slight decrease of the second‐order NLO response. Ion pairing is dependent upon the nature of the counter anion and on the electronic structure of the cationic NLO chromophore. It is very strong for the thiocyanate anion in particular and, albeit to a lesser extent, for the sulfonated anions. The latter show increased tendency to self‐aggregate.  相似文献   

11.
IntroductionThereisatremendousactivityforthepastyearsintheareaofinorganic organichybridmaterialsinviewoftheirdi versifiedstructuresandinterestingproperties.1,2 Theeffortshavebeenmadeinsynthesizingandcharacterizingtheclassofmaterials.3Thecontrolofinorganicstructurebyanorganiccomponentrevealsaninteractivestructuralrelationinthema terials .4 Assemblyofaninorganic organichybridmaterialcanbeachievedbyselectingorganiccomponents (multidentateligands)andinorganiccomponents (transitionmetalions ,metalo…  相似文献   

12.
Substitution of A‐site and/or X‐site ions of ABX3‐type perovskites with organic groups can give rise to hybrid perovskites, many of which display intriguing properties beyond their parent compounds. However, this method cannot be extended effectively to hybrid antiperovskites. Now, the design of hybrid antiperovskites under the guidance of the concept of Goldschmidt's tolerance factor is presented. Spherical anions were chosen for the A and B sites and spherical organic cations for the X site, and seven hybrid antiperovskites were obtained, including (F3(H2O)x)(AlF6)(H2dabco)3, ((Co(CN)6)(H2O)5)(MF6)(H2dabco)3 (M=Al3+, Cr3+, or In3+), (Co(CN)6)(MF6)(H2pip)3 (M=Al3+ or Cr3+), and (SbI6)(AlF6)(H2dabco)3. These new structures reveal that all ions at A, B, and X sites of inorganic antiperovskites can be replaced by molecular ions to form hybrid antiperovskites. This work will lead to the synthesis of a large family of hybrid antiperovskites.  相似文献   

13.
Reaction of copper(I) thiocyanate and triphenylphosphane with the bidentate Schiff base N,N′‐bis(trans‐2‐nitrocinnamaldehyde)ethylenediamine {Nca2en, (1); systematic name (1E,1′E,2E,2′E)‐N,N′‐(ethane‐1,2‐diyl)bis[3‐(2‐nitrophenyl)prop‐2‐en‐1‐imine]}, C20H18N4O4, in a 1:1:1 molar ratio in acetonitrile resulted in the formation of the complex {(1E,1′E,2E,2′E)‐N,N′‐(ethane‐1,2‐diyl)bis[3‐(2‐nitrophenyl)prop‐2‐en‐1‐imine]‐κ2N,N′}(thiocyanato‐κN)(triphenylphosphane‐κP)copper(I)], [Cu(NCS)(C20H18N4O4)(C18H15P)] or [Cu(NCS)(Nca2en)(PPh3)], (2). The Schiff base and copper(I) complex have been characterized by elemental analyses, IR, electronic and 1H NMR spectroscopy, and X‐ray crystallography [from synchrotron data for (1)]. The molecule of (1) lies on a crystallographic inversion centre, with a trans conformation for the ethylenediamine unit, and displays significant twists from coplanarity of its nitro group, aromatic ring, conjugated chain and especially ethylenediamine segments. It acts as a bidentate ligand coordinating via the imine N atoms to the CuI atom in complex (2), in which the ethylenediamine unit necessarily adopts a somewhat flattened gauche conformation, resulting in a rather bowed shape overall for the ligand. The NCS ligand is coordinated through its N atom. The geometry around the CuI atom is distorted tetrahedral, with a small N—Cu—N bite angle of 81.56 (12)° and an enlarged opposite angle of 117.29 (9)° for SCN—Cu—P. Comparisons are made with the analogous Schiff base having no nitro substituents and with metal complexes of both ligands.  相似文献   

14.
Structures of Polar Magnesium Organyls: Synthesis and Structure of Base Adducts of Bis(cyclopentadienyl)magnesium Eight donor‐acceptor complexes of bis(cyclopentadienyl)magnesium ( 1 ) with N‐ and O‐donor Lewis bases have been synthesized and characterized by X‐ray structure analysis. With acetonitrile, dimethoxyethane, diethyleneglycoldimethylether, dioxane, and tetramethylethylenediamine simple 1:1 adducts are formed ( 2 – 6 ). In some cases a change of the hapticity of one cyclopentadienylring from η5 to η2 or η1 is observed ( 4 – 6 ). In the adduct with pentamethyldiethylenetriamine ( 7 ) one C5H5‐ring is removed from the magnesium atom forming the cation [Mg(C5H5)(PMDTA)]+ and an uncoordinated five‐ring anion. With the crown ether 15‐crown‐5 the two ionic Mg compounds 8 and 9 are formed which have a [Mg(15‐crown‐5)L2]2+‐cation [L = pyridine, THF] and two uncoordinated cyclopentadienyl anions. Cyclopentadienyl‐methyl‐magnesium reacts with 15‐crown‐5 to the salt [Mg(CH3)(15‐crown‐5)]+ C5H5? ( 10 ) which has also a free cyclopentadienyl anion.  相似文献   

15.
Acidic Sulfates of Neodymium: Synthesis and Crystal Structure of (H5O2)(H3O)2Nd(SO4)3 and (H3O)2Nd(HSO4)3SO4 Light violett single crystals of (H5O2)(H3O)2 · Nd(SO4)3 are obtained by cooling of a solution prepared by dissolving neodymium oxalate in sulfuric acid (80%). According to X‐ray single crystal investigations there are H3O+ ions and H5O2+ ions present in the monoclinic structure (P21/n, Z = 4, a = 1159.9(4), b = 710.9(3), c = 1594.7(6) pm, β = 96.75(4)°, Rall = 0.0260). Nd3+ is nine‐coordinate by oxygen atoms. The same coordination number is found for Nd3+ in the crystal structure of (H3O)2Nd(HSO4)3SO4 (triclinic, P1, Z = 2, a = 910.0(1), b = 940.3(1), c = 952.6(1) pm, α = 100.14(1)°, β = 112.35(1)°, γ = 105.01(1)°, Rall = 0.0283). The compound has been prepared by the reaction of Nd2O3 with chlorosulfonic acid in the presence of air. In the crystal structure both sulfate and hydrogensulfate groups occur. In both compounds pronounced hydrogen bonding is observed.  相似文献   

16.
A comparative study of the sonoluminescence spectra of water and argon-saturated aqueous H2SO4 solutions was carried out. At an H2SO4 concentration of 18 mol L−1, the sulfuric acid sonoluminescence is fifty times more intense than water sonoluminescence. The sulfuric acid luminescence spectrum differs from the water sonoluminescence spectrum caused by the emission of excited water molecules and OH radicals from the gas phase of cavitation bubbles. The sulfuric acid sonoluminescence spectrum exhibits maxima at 330, 420, 500, and 630 nm. Emitters of sonoluminescence of sulfuric acid are the singlet (330–340 nm) and triplet (∼420 nm) excited SO2 molecules formed by sonolysis of H2SO4 molecules. Another product of sonolysis of H2SO4, atomic oxygen, is assumed to be responsible for the luminescence at λ = 630 nm. __________ Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 8, pp. 1742–1745, August, 2005.  相似文献   

17.
Depending on the reaction partner, the organic ditopic molecule isonicotinic acid (Hina) can act either as a Brønsted acid or base. With sulfuric acid, the pyridine ring is protonated to become a pyridinium cation. Crystallization from ethanol affords the title compound tris(4‐carboxypyridinium) hydrogensulfate sulfate monohydrate, 3C6H6NO2+·HSO4·SO42−·H2O or [(H2ina)3(HSO4)(SO4)(H2O)]. This solid contains 11 classical hydrogen bonds of very different flavour and nonclassical C—H…O contacts. All N—H and O—H donors find at least one acceptor within a suitable distance range, with one of the three pyridinium H atoms engaged in bifurcated N—H…O hydrogen bonds. The shortest hydrogen‐bonding O…O distance is subtended by hydrogensulfate and sulfate anions, viz. 2.4752 (19) Å, and represents one of the shortest hydrogen bonds ever reported between these residues.  相似文献   

18.
Two mononuclear copper complexes, {bis[(3,5‐dimethyl‐1H‐pyrazol‐1‐yl‐κN2)methyl]amine‐κN}(3,5‐dimethyl‐1H‐pyrazole‐κN2)(perchlorato‐κO)copper(II) perchlorate, [Cu(ClO4)(C5H8N2)(C12H19N5)]ClO4, (I), and {bis[(3,5‐dimethyl‐1H‐pyrazol‐1‐yl‐κN2)methyl]amine‐κN}bis(3,5‐dimethyl‐1H‐pyrazole‐κN2)copper(II) bis(hexafluoridophosphate), [Cu(C5H8N2)2(C12H19N5)](PF6)2, (II), have been synthesized by the reactions of different copper salts with the tripodal ligand tris[(3,5‐dimethyl‐1H‐pyrazol‐1‐yl)methyl]amine (TDPA) in acetone–water solutions at room temperature. Single‐crystal X‐ray diffraction analysis revealed that they contain the new tridentate ligand bis[(3,5‐dimethyl‐1H‐pyrazol‐1‐yl)methyl]amine (BDPA), which cannot be obtained by normal organic reactions and has thus been captured in the solid state by in situ synthesis. The coordination of the CuII ion is distorted square pyramidal in (I) and distorted trigonal bipyramidal in (II). The new in situ generated tridentate BDPA ligand can act as a meridional or facial ligand during the process of coordination. The crystal structures of these two compounds are stabilized by classical hydrogen bonding as well as intricate nonclassical hydrogen‐bond interactions.  相似文献   

19.
Poly[[μ4‐4,4′‐bipyridazine‐μ5‐sulfato‐disilver(I)] monohydrate], {[Ag2(SO4)(C8H6N4)]·H2O}n, (I), and poly[[aqua‐μ4‐pyridazino[4,5‐d]pyridazine‐μ3‐sulfato‐disilver(I)] monohydrate], {[Ag2(SO4)(C6H4N4)(H2O)]·H2O}n, (II), possess three‐ and two‐dimensional polymeric structures, respectively, supported by N‐tetradentate coordination of the organic ligands [Ag—N = 2.208 (3)–2.384 (3) Å] and O‐pentadentate coordination of the sulfate anions [Ag—O = 2.284 (3)–2.700 (2) Å]. Compound (I) is the first structurally examined complex of the new ligand 4,4′‐bipyridazine; it is based upon unprecedented centrosymmetric silver–pyridazine tetramers with tetrahedral AgN2O2 and trigonal–bipyramidal AgN2O3 coordination of two independent AgI ions. Compound (II) adopts a typical dimeric silver–pyridazine motif incorporating two kinds of square‐pyramidal AgN2O3 AgI ions. The structure exhibits short anion–π interactions involving noncoordinated sulfate O atoms [O...π = 3.041 (3) Å].  相似文献   

20.
A new copper(II) phosphonatobenzenesulfonate incorporating 4,4′‐bipyridine (4,4′‐bipy) as auxiliary ligand has been discovered through systematic high‐throughput (HT) screening of the system Cu(NO3)2·3H2O/H2O3PC6H4SO3H/4,4′‐bipy using different solvents. The hydrothermal synthesis of [Cu(HO3PC6H4SO3)(C10H8N2)]·H2O ( 1 ) was further optimized by screening various copper(II) salts. The crystal structure of 1 was determined by single‐crystal X‐ray diffraction and unveiled the presence of isolated sixfold coordinated Jahn–Teller‐distorted Cu2+ ions. The isolated CuN2O4 octahedra are interconnected by phosphonate and sulfonate groups to form chains along the c‐axis. The organic groups, namely phenyl rings and 4,4′‐bipy molecules cross‐link the chains into a three‐dimensional framework. Water molecules are found in the narrow voids in the structure which are held by weak hydrogen bonds. Upon dehydration, the structure of 1 undergoes a phase transition, which was confirmed by TG measurements and temperature dependent X‐ray powder diffraction. The new structure of 1‐h was refined with Rietveld methods. Detailed inspection of the structure revealed the directional switching of the Jahn–Teller distortion upon de/rehydration. Weak ferro‐/ferrimagnetic interactions were observed by magnetic investigations of 1 , which switch to antiferromagnetic below 3.5 K. Compounds 1 and 1‐h are further characterized by thermogravimetric and elemental analysis as well as IR spectroscopy.  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