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.
This investigation conducted reaction channels of weakly bound complexes CO2…HF, CO2…HF…NH3, CO2…HF…NH2CH3, CO2…HF…NH(CH3)2, and CO2…HF…N(CH3)3 systems, using the Gaussian 98 package at the B3LYP/6‐311++G(3df,2pd) level. The syn‐fluoroformic acid or syn‐fluoroformic acid plus NH3 or amine conformers are more stable than the related anti‐fluoroformic acid or anti‐fluoroformic acid plus NH3 or amine conformers. However, the above‐mentioned weakly bound complexes are more stable than both the related syn‐ and anti‐type fluoroformic acid or acid plus NH3 or amine conformers and their related decomposed into CO2 + HF or CO2 + NHR3F (R?H, CH3) combined molecular systems. Five reaction channels of the weakly bound complexes exist. Each channel includes weakly bound complexes and their related above‐mentioned systems. Moreover, each reaction channel contains two transition states. The transition state between the weakly bound complex and anti‐fluoroformic acid type structure (T13) is significantly higher than that of internal rotation (T23) between syn‐ and anti‐FCO2H (or FCO2H…NR3) structures. However, the above‐mentioned T13 can significantly decrease its energy by adding the third molecule NH3 or NR3 (R?H or CH3). The more CH3 that is substituted in NR3 of the reaction channel, the lower the activation energy of the transition state in the system is affected. © 2005 Wiley Periodicals, Inc. Int J Quantum Chem, 2005  相似文献   

2.
In this investigation, reaction channels of weakly bound complexes CO2HF, CO2HFH2O, CO2HFNH3, CO2HFCH3OH, CO2HFNH2CH3, CO2HFNH(CH3)2 and CO2HFN(CH3)3 systems were studied at the B3LYP/6-311++G(3df,2pd) level. The conformers of syn-fluoroformic acid or syn-fluoroformic acid plus a third molecule (H2O, NH3, CH3OH, NH2CH3, NH(CH3)2 or N(CH3)3) were found to be more stable than the conformers of the related anti-fluoroformic acid or anti-fluoroformic acid plus a third molecule (H2O, NH3, CH3OH, NH2CH3, NH(CH3)2 or N(CH3)3). However, the weakly bound complexes were found to be more stable than either the related syn- and anti- type fluoroformic acid or the acid plus third molecule (H2O, NH3, CH3OH, NH2CH3, NH(CH3)2 or N(CH3)3) conformers. They decomposed into CO2+HF, CO2+H3OF, CO2+NH4F, CO2+(CH3)OH2F, CO2+NH3(CH3)F, CO2+NH2(CH3)2F, or CO2+NH(CH3)3F combined molecular systems. The weakly bound complexes have seven reaction channels, each of which includes weakly bound complexes and their related systems. Moreover, each reaction channel includes two transition state structures. The transition state between the weakly bound complex and anti-fluoroformic acid type structure (T13) is significantly higher than that of internal rotation (T23) between the syn- and anti-FCO2H (or FCO2HH2O, FCO2HNH3, FCO2HCH3OH, FCO2HNH2CH3, FCO2HNH(CH3)2, or FCO2HN(CH3)3) structures. However, adding the third molecule H2O, NH3, CH3OH, NH2CH3, NH(CH3)2 or N(CH3)3 can significantly reduce the activation energy of T13. The catalytic strengths of the third molecules are predicted to follow the order H2O<NH3<CH3OH<.NH2CH3<NH(CH3)2<N(CH3)3.  相似文献   

3.
Using four basis sets, 6‐311G(d,p), 6‐31+G(d,p), 6‐311++G(2d,2p), and 6‐311++G(3df,3pd), the optimized structures with all real frequencies were obtained at the MP2 level for dimers CH2O? HF, CH2O? H2O, CH2O? NH3, and CH2O? CH4. The structures of CH2O? HF, CH2O? H2O, and CH2O? NH3 are cycle‐shaped, which result from the larger bend of σ‐type hydrogen bonds. The bend of σ‐type H‐bond O…H? Y (Y?F, O, N) is illustrated and interpreted by an attractive interaction of a chemically intuitive π‐type hydrogen bond. The π‐type hydrogen bond is the interaction between one of the acidic H atoms of CH2O and lone pair(s) on the F atom in HF, the O atom in H2O, or the N atom in NH3. By contrast with above the three dimers, for CH2O? CH4, because there is not a π‐type hydrogen‐bond to bend its linear hydrogen bond, the structure of CH2O? CH4 is a noncyclic shaped. The interaction energy of hydrogen bonds and the π‐type H‐bond are calculated and discussed at the CCSD(T)/6‐311++G(3df,3pd) level. © 2005 Wiley Periodicals, Inc. Int J Quantum Chem, 2005  相似文献   

4.
Using four basis bets, (6‐311G(d,p), 6‐31+G(d,p), 6‐31++G(2d,2p), and 6‐311++G(3df,3pd), the optimized structures with all real frequencies were obtained at the MP2 level for the dimers CH2O? HF, CH2O? H2O, CH2O? NH3, and CH2O? CH4. The structures of CH2O? HF, CH2O? H2O, and CH2O? NH3 are cycle‐shaped, which result from the larger bend of σ‐type hydrogen bonds. The bend of σ‐type H‐bond O…H? Y (Y?F, O, N) is illustrated and interpreted by an attractive interaction of a chemically intuitive π‐type hydrogen bond. The π‐type hydrogen bond is the interaction between one of the H atoms of CH2O and lone pair(s) on the F atom in HF, the O atom in H2O, or the N atom in NH3. In contrast with the above three dimers, for CH2O? CH4, because there is not a π‐type hydrogen bond to bend its linear hydrogen bond, the structure of CH2O? CH4 is noncyclic shaped. The interaction energy of hydrogen bonds and the π‐type H‐bond are calculated and discussed at the CCSD (T)/6‐311++G(3df,3pd) level. © 2005 Wiley Periodicals, Inc. Int J Quantum Chem, 2005  相似文献   

5.
Hydrogen bonding and crystal packing play major roles in determining the conformations of ethyl methyl 2‐(triphenyl­phospho­ranyl­idene)malonate, Ph3P=C(CO2CH3)CO2CH2CH3 or C24H23O4P, (I), and dimethyl 2‐(triphenyl­phosphor­anyl­idene)malonate, Ph3P=C(CO2CH3)2 or C23H21O4P, (II). In (I), the acyl O atom of the ethyl ester group is anti to the P atom, while the O atom of the methyl ester group is syn. In (II), the dimethyl diester is a 1:1 mixture of antianti and synanti conformers.  相似文献   

6.
In the title compounds, 3-(dihydroxyboryl)anilinium bisulfate monohydrate, C6H9BNO2+·HSO4·H2O ( I ), and 3-(dihydroxyboryl)anilinium methyl sulfate, C6H9BNO2+·CH3SO4 ( II ), the almost planar boronic acid molecules are linked by pairs of O—H…O hydrogen bonds, forming centrosymmetric motifs that can be described by the graph-set R22(8) motif. In both crystals, the B(OH)2 group acquires a synanti conformation (with respect to the H atoms). The presence of the hydrogen-bonding functional groups B(OH)2, NH3+, HSO4, CH3SO4 and H2O generates three-dimensional hydrogen-bonded networks, in which the bisulfate (HSO4) and methyl sulfate (CH3SO4) counter-ions act as the central building blocks within the crystal structures. Furthermore, in both structures, the packing is stabilized by weak boron–π interactions, as shown by noncovalent interactions (NCI) index calculations.  相似文献   

7.
Because of their versatile coordination modes and strong coordination ability for metals, triazole ligands can provide a wide range of possibilities for the construction of metal–organic frameworks. Three transition‐metal complexes, namely bis(μ‐1,2,4‐triazol‐4‐ide‐3‐carboxylato)‐κ3N 2,O :N 13N 1:N 2,O‐bis[triamminenickel(II)] tetrahydrate, [Ni2(C3HN3O2)2(NH3)6]·4H2O, (I), catena‐poly[[[diamminediaquacopper(II)]‐μ‐1,2,4‐triazol‐4‐ide‐3‐carboxylato‐κ3N 1:N 4,O‐[diamminecopper(II)]‐μ‐1,2,4‐triazol‐4‐ide‐3‐carboxylato‐κ3N 4,O :N 1] dihydrate], {[Cu2(C3HN3O2)2(NH3)4(H2O)2]·2H2O}n , (II), (μ‐5‐amino‐1,2,4‐triazol‐1‐ide‐3‐carboxylato‐κ2N 1:N 2)di‐μ‐hydroxido‐κ4O :O‐bis[triamminecobalt(III)] nitrate hydroxide trihydrate, [Co2(C3H2N4O2)(OH)2(NH3)6](NO3)(OH)·3H2O, (III), with different structural forms have been prepared by the reaction of transition metal salts, i.e. NiCl2, CuCl2 and Co(NO3)2, with 1,2,4‐triazole‐3‐carboxylic acid or 3‐amino‐1,2,4‐triazole‐5‐carboxylic acid hemihydrate in aqueous ammonia at room temperature. Compound (I) is a dinuclear complex. Extensive O—H…O, O—H…N and N—H…O hydrogen bonds and π–π stacking interactions between the centroids of the triazole rings contribute to the formation of the three‐dimensional supramolecular structure. Compound (II) exhibits a one‐dimensional chain structure, with O—H…O hydrogen bonds and weak O—H…N, N—H…O and C—H…O hydrogen bonds linking anions and lattice water molecules into the three‐dimensional supramolecular structure. Compared with compound (I), compound (III) is a structurally different dinuclear complex. Extensive N—H…O, N—H…N, O—H…N and O—H…O hydrogen bonding occurs in the structure, leading to the formation of the three‐dimensional supramolecular structure.  相似文献   

8.
应用B3LYP方法,结合6-31G**、cc-pVDZ、aug-cc-pVDZ和cc-pVTZ基组对硫代乙酸的两种异构体CH3C(O)SH和CH3C(S)OH在基态势能面上的9个单分子反应进行了研究。本文计算预测硫代乙酸主要以CH3C(O)SH的形式存在,两种异构体均以顺式构象为优势构象。通过对比CH3C(O)SH、CH3C(S)OH和 CH3C(O)OH的反应性差异,我们可以得出结论:CH3C(O)OH中-OH基团的O被S取代后,只有当-SH作为一个整体参加反应时才对分子解离过程有较大影响;而C=O或C=S对反应性影响较小。  相似文献   

9.
The crystal structures of 2‐hydroxy‐5‐[(E)‐(4‐nitrophenyl)diazenyl]benzoic acid, C13H9N3O5, (I), ammonium 2‐hydroxy‐5‐[(E)‐phenyldiazenyl]benzoate, NH4+·C13H9N2O3, (II), and sodium 2‐hydroxy‐5‐[(E)‐(4‐nitrophenyl)diazenyl]benzoate trihydrate, Na+·C13H8N3O5·3H2O, (III), have been determined using single‐crystal X‐ray diffraction. In (I) and (III), the phenyldiazenyl and carboxylic acid/carboxylate groups are in an anti orientation with respect to each other, which is in accord with the results of density functional theory (DFT) calculations, whereas in (II), the anion adopts a syn conformation. In (I), molecules form slanted stacks along the [100] direction. In (II), anions form bilayers parallel to (010), the inner part of the bilayers being formed by the benzene rings, with the –OH and –COO substituents on the bilayer surface. The NH4+ cations in (II) are located between the bilayers and are engaged in numerous N—H...O hydrogen bonds. In (III), anions form layers parallel to (001). Both Na+ cations have a distorted octahedral environment, with four octahedra edge‐shared by bridging water O atoms, forming [Na4(H2O)12]4+ units.  相似文献   

10.
We report a new polymorph of (1E,4E)‐1,5‐bis(4‐fluorophenyl)penta‐1,4‐dien‐3‐one, C17H12F2O. Contrary to the precedent literature polymorph with Z′ = 3, our polymorph has one half molecule in the asymmetric unit disordered over two 50% occupancy sites. Each site corresponds to one conformation around the single bond vicinal to the carbonyl group (so‐called anti or syn). The other half of the bischalcone is generated by twofold rotation symmetry, giving rise to two half‐occupied and overlapping molecules presenting both anti and syn conformations in their open chain. Such a disorder allows for distinct patterns of intermolecular C—H…O contacts involving the carbonyl and anti‐oriented β‐C—H groups, which is reflected in three 13C NMR chemical shifts for the carbonyl C atom. Here, we have also assessed the cytotoxicity of three symmetric bischalcones through their in vitro antitumour potential against three cancer cell lines. Cytotoxicity assays revealed that this biological property increases as halogen electronegativity increases.  相似文献   

11.
Although it has been generally assumed that electron attachment to disulfide derivatives leads to a systematic and significant activation of the S? S bond, we show, by using [CH3SSX] (X=CH3, NH2, OH, F) derivatives as model compounds, that this is the case only when the X substituents have low electronegativity. Through the use of MP2, QCI and CASPT2 molecular orbital (MO) methods, we elucidate, for the first time, the mechanisms that lead to unimolecular fragmentation of disulfide derivatives after electron attachment. Our theoretical scrutiny indicates that these mechanisms are more intricate than assumed in previous studies. The most stable products, from a thermodynamic viewpoint, correspond to the release of neutral molecules; CH4, NH3, H2O, and HF. However, the barriers to reach these products depend strongly on the electronegativity of the X substituents. Only for very electronegative substituents, such as OH or F, the loss of H2O or HF is the most favorable process, and likely the only one observed. This is possible because of two concomitant factors, 1) the extra electron for [CH3SSX]? (X=OH, F) occupies a σ*(S? X) MO, which favors the cleavage of the S? X bond, and 2) the activation barriers associated with the hydrogen transfer process to produce H2O and HF are rather low. Only when the substituents are less electronegative (X=H, CH3, NH2) the extra electron is located in a σ*(S? S) orbital and the cleavage of the disulfide bridge becomes the most favorable process. The intimate mechanism associated with the S? S bond dissociation process also depends strongly on the nature of the substituent. For X=H or CH3 the process is strictly adiabatic, while for X=NH2 it proceeds through a conical intersection ( CI ) associated with the charge reorganization necessary to obtain, from a molecular anion with the extra electron delocalized in a σ*(S? S) antibonding orbital, two fragments with the proper charge localization.  相似文献   

12.
《Analytical letters》2012,45(6):1567-1580
Abstract

The effects of β-cyclodextrin were investigated on the fluorescence emission and excitation as well as on the UV absorption and solubility of certain bimanes in aqueous solution. Three syn-bimanes with differing water solubilities were examined, namely, syn-(CH2OCOCH3, CH3) bimane, syn-(CH3, CH3) bimane, and syn-(C6H3, Cl) bimane. the anti-(CH3, CH3) bimane was also examined. In dilute solutions, the syn-(CH3, CH3) bimane and syn-(C6H3, Cl) bimane showed enhancement of their relative fluorescence intensities upon the addition of β-cyclodextrin as did anti-(CH3, CH3) bimane. Only the anti-(CH3, CH3) bimane showed significant changes in its UV absorption upon the addition of β-cyclodextrin. Both syn-(CH2OCOCH3, CH3) bimane and syn-(CH3, CH3) bimane solubilities were increased in the presence of β-CD. the formation of β-cyclodextrin inclusion complexes is proposed as a possible interpretation of these observations.  相似文献   

13.
By combining results from a variety of mass spectrometric techniques (metastable ion, collisional activation, collision-induced dissociative ionization, neutralization-reionization spectrometry, 2H, 13C and 18O isotopic labelling and appearance energy measurements) and high-level ab initio molecular orbital calculations, the potential energy surface of the [CH5NO]+ ˙ system has been explored. The calculations show that at least nine stable isomers exist. These include the conventional species [CH3ONH2]+ ˙ and [HO? CH2? NH2]+ ˙, the distonic ions [O? CH2? NH3]+ ˙, [O? NH2? CH3]+ ˙, [CH2? O(H)? NH2]+ ˙, [HO? NH2? CH2]+ ˙, and the ion-dipole complex CH2?NH2+ …? OH˙. Surprisingly the distonic ion [CH2? O? NH3]+ ˙ was found not to be a stable species but to dissociate spontaneously to CH2?O + NH3+ ˙. The most stable isomer is the hydrogen-bridged radical cation [H? C?O …? H …? NH3]+ ˙ which is best viewed as an immonium cation interacting with the formyl dipole. The related species [CH2?O …? H …? NH2]+ ˙, in which an ammonium radical cation interacts with the formaldehyde dipole is also a very stable ion. It is generated by loss of CO from ionized methyl carbamate, H2N? C(?O)? OCH3 and the proposed mechanism involves a 1,4-H shift followed by intramolecular ‘dictation’ and CO extrusion. The [CH2?O …? H …? NH2]+ ˙ product ions fragment exothermically, but via a barrier, to NH4+ ˙ HCO…? and to H3N? C(H)?O+ ˙ H˙. Metastable ions [CH3ONH2]+…? dissociate, via a large barrier, to CH2?O + NH3+ + and to [CH2NH2]+ + OH˙ but not to CH2?O+ ˙ + NH3. The former reaction proceeds via a 1,3-H shift after which dissociation takes place immediately. Loss of OH˙ proceeds formally via a 1,2-CH3 shift to produce excited [O? NH2? CH3]+ ˙, which rearranges to excited [HO? NH2? CH2]+ ˙ via a 1,3-H shift after which dissociation follows.  相似文献   

14.
The crystal structures of rare‐earth diaryl‐ or dialkylphosphate derivatives are poorly explored. Crystals of bis[bis(2,6‐diisopropylphenyl)phosphato‐κO ]chloridotetrakis(methanol‐κO )neodymium methanol disolvate, [Nd(C24H34O4P)Cl(CH4O)4]·2CH3OH, (1), and of the lutetium, [Lu(C24H34O4P)Cl(CH4O)4]·2CH3OH, (2), and yttrium, [Y(C24H34O4P)Cl(CH4O)4]·2CH3OH, (3), analogues have been obtained by reactions between lithium bis(2,6‐diisopropylphenyl)phosphate and LnCl3(H2O)6 (in a 2:1 ratio) in methanol. Compounds (1)–(3) crystallize in the C 2/c space group. Their crystal structures are isomorphous. The molecule possesses C 2 symmetry with a twofold crystallographic axis passing through the Ln and Cl atoms. The bis(2,6‐diisopropylphenyl)phosphate ligands all display a κ1O‐monodentate coordination mode. The coordination polyhedron for the metal atom [coordination number (CN) = 7] is a distorted pentagonal bipyramid. Each [Ln{O2P(O‐2,6‐iPr2C6H3)2}2Cl(CH3OH)4] molecular unit exhibits two intramolecular O—H…O hydrogen bonds, forming six‐membered rings, and two intramolecular O—H…Cl interactions, forming four‐membered rings. Intermolecular O—H…O hydrogen bonds connect each unit via four noncoordinating methanol molecules with four other units, forming a two‐dimensional hydrogen‐bond network. Crystals of bis[bis(2,6‐diisopropylphenyl)phosphato‐κO ]tetrakis(methanol‐κO )(nitrato‐κ2O ,O ′)neodymium methanol disolvate, [Nd(C24H34O4P)(NO3)(CH4O)4]·2CH3OH, (4), have been obtained in an analogous manner from NdCl3(H2O)6. Compound (4) also crystalizes in the C 2/c space group. Its crystal structure is similar to those of (1)–(3). The κ2O ,O ′‐bidentate nitrate anion is disordered over a twofold axis, being located nearly on it. Half of the molecule is crystallographically unique (CNNd = 8). Unlike (1)–(3), complex (4) exhibits disorder of all three methanol molecules, one isopropyl group of the phosphate ligand and the NO3 ligand. The structure of (4) displays intra‐ and intermolecular O—H…O hydrogen bonds similar to those in (1)–(3). Compounds (1)–(4) represent the first reported mononuclear bis[bis(diaryl/dialkyl)phosphate] rare‐earth complexes.  相似文献   

15.
Reaction of [M(NO)Cl3(NCMe)2] (M=Mo, W) with (iPr2PCH2CH2)2PPh (etpip) at room temperature afforded the syn/anti‐[M(NO)Cl3(mer‐etpip)] complexes (M=Mo, a ; W, b ; 3 a,b (syn,anti); syn and anti refer to the relative position of Ph(etpip) and NO). Reduction of 3 a,b (syn,anti) produced [M(NO)Cl2(mer‐etpip)] ( 4 a,b (syn)), [M(NO)Cl(NCMe)(mer‐etpip)] ( 5 a,b (syn,anti)), and [M(NO)Cl(η2‐ethylene)(mer‐etpip)] ( 6 a,b (syn,anti)) complexes. The hydrides [M(NO)H(η2‐ethylene)(mer‐etpip)] ( 7 a,b (syn,anti)) were obtained from 6 a,b (syn,anti) using NaHBEt3 (75 °C, THF) or LiBH4 (80 °C, Et3N), respectively. 7 a,b (syn,anti) were probed in olefin hydrogenations in the absence or presence of a hydrosilane/B(C6F5)3 mixture. The 7 a,b (syn,anti)/Et3SiH/B(C6F5)3 co‐catalytic systems were highly active in various olefin hydrogenations (60 bar H2, 140 °C), with maximum TOFs of 5250 h?1 ( 7 a (syn,anti)) and 8200 h?1 ( 7 b (syn,anti)) for 1‐hexene hydrogenation. The Et3SiH/(B(C6F5)3 co‐catalyst is anticipated to generate a [Et3Si]+ cation attaching to the ONO atom. This facilitates NO bending and accelerates catalysis by providing a vacant site. Inverse DKIE effects were observed for the 7 a (syn,anti)/Et3SiH/(B(C6F5)3 (kH/kD=0.55) and the 7 b (syn,anti)/Et3SiH/(B(C6F5)3 (kH/kD=0.65) co‐catalytic mixtures (20 bar H2/D2, 140 °C).  相似文献   

16.
An effective method was developed for the synthesis of three cluster‐based frameworks with multifarious secondary building units (SBUs) and various structures, which were formulated as [Me2NH2]2[Zn10(BTC)63‐O)(μ4‐O)(H2O)5] · 3DMA · 9H2O ( FJI ‐ 3 ), [Me2NH2]2[Zn93‐OH)2(BTC)6(H2O)3] · 5DMA · 6H2O ( FJI ‐ 4 ) and [Me2NH2][Zn33‐OH)(BTC)2DMF] · H2O ( FJI ‐ 5 ) (H3BTC = 1,3,5‐benzenetricarboxylic acid, DMA = N,N′‐dimethyl acetamide and DMF = N,N′‐dimethyl formamide), respectively. X‐ray structural analysis reveals that FJI ‐ 3 displays 3D highly porous metal‐organic framework with four kinds of microporous cages constructed by two paddle‐wheel Zn2(CO2)4, trimeric Zn3O(CO2)6, and tetrameric Zn4O(CO2)6 SBUs. FJI ‐ 4 exhibits 3D microporous MOFs with a dodecahedral cavities built by paddle‐wheel Zn2(CO2)4 and trimeric Zn3O(CO2)6. FJI ‐ 5 shows 3D microporous MOFs with an 1D channel assembled by the Zn3O(CO2)6 SBUs. In addition, the fluorescence and sorption properties in these cluster‐based frameworks were also investigated. Furthermore, the method employed in this work may provide an useful approach to the design and synthesis of novel cluster‐based frameworks.  相似文献   

17.
The self‐assembly of ditopic bis(1H‐imidazol‐1‐yl)benzene ligands ( L H) and the complex (2,2′‐bipyridyl‐κ2N,N′)bis(nitrato‐κO)palladium(II) affords the supramolecular coordination complex tris[μ‐bis(1H‐imidazol‐1‐yl)benzene‐κ2N3:N3′]‐triangulo‐tris[(2,2′‐bipyridyl‐κ2N,N′)palladium(II)] hexakis(hexafluoridophosphate) acetonitrile heptasolvate, [Pd3(C10H8N2)3(C12H10N4)3](PF6)6·7CH3CN, 2 . The structure of 2 was characterized in acetonitrile‐d3 by 1H/13C NMR spectroscopy and a DOSY experiment. The trimeric nature of supramolecular coordination complex 2 in solution was ascertained by cold spray ionization mass spectrometry (CSI–MS) and confirmed in the solid state by X‐ray structure analysis. The asymmetric unit of 2 comprises the trimetallic Pd complex, six PF6? counter‐ions and seven acetonitrile solvent molecules. Moreover, there is one cavity within the unit cell which could contain diethyl ether solvent molecules, as suggested by the crystallization process. The packing is stabilized by weak inter‐ and intramolecular C—H…N and C—H…F interactions. Interestingly, the crystal structure displays two distinct conformations for the L H ligand (i.e. syn and anti), with an all‐syn‐[Pd] coordination mode. This result is in contrast to the solution behaviour, where multiple structures with syn/anti‐ L H and syn/anti‐[Pd] are a priori possible and expected to be in rapid equilibrium.  相似文献   

18.
A second polymorph of phenylselenium trichloride, PhSeCl3 or C6H5Cl3Se, is disclosed, which is comprised of asymmetric chlorine‐bridged noncovalent dimer units rather than polymeric chains. These dimers are each weakly bound to an adjacent dimer through noncovalent Se…Cl bonding interactions. Phenyl rings within each dimer are oriented in a syn fashion. Density functional theory (DFT) calculations reveal that the putative anti isomer is within 5 kJ mol?1 of the experimentally observed form. This structure represents the first additional polymorph discovered for an organoselenium trihalide compound.  相似文献   

19.
A two dimensional coordination polymer with pseudo‐Kagomé net [Cu3(btc)2(NH3)8(H2O)] was prepared from Cu(NO3)2 · 3H2O and 1, 3, 5‐benzenetricarboxylic acid (btc) in ammonia aqua solution and was structurally characterized by X‐ray diffraction. The magnetic susceptibility measurements, measured from 2 to 300 K, revealed a weak anti‐ferromagnetic interaction between the CuII ions via the btc ligands.  相似文献   

20.
The FOGO method is used to calculate absolute proton affinities of the molecules H2, HF, NH3, H2O, CH3OH, C2H5OH, H2O2, CH2O, CO, and CH2CO. Comparison with experimental values demonstrates that the geometrical and energetical data resulting from this type of ab initio calculation are of chemical accuracy. Predictive data for higher energy isomers, such as hydroxymethylene and ethynol are given as possible aid for the identification of these species.  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