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 943 毫秒
1.
Replacement and Oxidation Reactions of N-Dichlorophosphanyl Triphenylphosphazene, Ph3P?N? PCl2 The title compound ( 1 ) reacts with MeOH, EtOH, PhOH, EtSH, and water forming N-phosphanyl or N-phosphinoyl phosphazenes, resp., Ph3P?N? PX2 (X ? OPh( 8 ), SEt( 9 )) or Ph3P?N? PH(O)X (X ? Cl( 3 ), OH( 4 ), OMe( 5 ), OEt( 7 )). The reaction of 1 with P(NEt2)3 yields Ph3P?N? P(NEt2)2 ( 10 ). Ph3P?N? PF2( 11 ) and Ph3P?N? PH(O)F ( 12 ) are obtained by chlorine-fluorine exchange. The N-phosphanyl compounds 1 , 8 , 9 and 11 are oxidized by NO2 yielding the corresponding N-phosphoryl derivatives, Ph3P?N? P(O)X2 (X ? Cl( 2 ), OPh( 13 ), SEt ( 14 ), F( 15 )). The thiophosphoryl compounds, (Ph3P?N? P(S)X2 (X ? Cl( 16 ), OPh( 17 ), F( 18 )) are obtained by oxidizing 1 , 8 , and 11 with sulfur.  相似文献   

2.
Reactions of the oxorhenium(V) complexes [ReOX3(PPh3)2] (X = Cl, Br) with the N‐heterocyclic carbene (NHC) 1,3,4‐triphenyl‐1,2,4‐triazol‐5‐ylidene (LPh) under mild conditions and in the presence of MeOH or water give [ReOX2(Y)(PPh3)(LPh)] complexes (X = Cl, Br; Y = OMe, OH). Attempted reactions of the carbene precursor 5‐methoxy‐1,3,4‐triphenyl‐4,5‐dihydro‐1H‐1,2,4‐triazole ( 1 ) with [ReOCl3(PPh3)2] or [NBu4][ReOCl4] in boiling xylene resulted in protonation of the intermediately formed carbene and decomposition products such as [HLPh][ReOCl4(OPPh3)], [HLPh][ReOCl4(OH2)] or [HLPh][ReO4] were isolated. The neutral [ReOX2(Y)(PPh3)(HLPh)] complexes are purple, airstable solids. The bulky NHC ligands coordinate monodentate and in cis‐position to PPh3. The relatively long Re–C bond lengths of approximate 2.1Å indicate metal‐carbon single bonds.  相似文献   

3.
The Influence of Phosphoryl Substituents on the Properties of P‐Substituted 2‐Methylimidazolium Ions and 2‐Methyleneimidazolines [1] The imidazolines ImCHP(E)Ph2 [ 6 , E = S ( a ), Se ( b )] are obtained from ImCHPPh2 ( 4 ) and sulfur or selenium. HBF4 reaction yields the corresponding imidazolium salts [ImCH2P(E)Ph2][BF4] [ 5 , E = S ( a ), Se ( b )]. 1, 3, 4, 5‐Tetramethyl‐2‐methylenimidazoline ( 1 , ImCH2) reacts with Ph2P(O)Cl to give the corresponding phosphane salt [ImCH2P(O)Ph2]Cl ( 7 ) from which the vinyl compound ImCHP(O)Ph2 ( 8 ) is formed through deprotonation. 8 reacts with excess HBF4 to give the phosphine oxide BF3 adduct [ImCH2P(O)Ph2 · BF3][BF4] ( 9 ). The crystal structures of 5a , 5b , 6b , 7 · CH2Cl2 and 9 · H2O as well as preliminary data of 8 are reported and discussed on comparison with the phosphanes [ImCH2PPh2][BF4] ( 3b ) and ImCHPPh2 ( 4 ). From structural data, π‐electron delocalisation is concluded for 6b and 8 .  相似文献   

4.
2D 1H,89Y heteronuclear shift correlation through scalar coupling has been applied to the chemical‐shift determination of a set of yttrium complexes with various nuclearities. This method allowed the determination of 89Y NMR data in a short period of time. Multinuclear NMR spectroscopy as function of temperature, PGSE NMR‐diffusion experiments, heteronuclear NOE measurements, and X‐ray crystallography were applied to determine the structures of [Y5(OH)5(L ‐Val)4(Ph2acac)6] ( 1 ) (Ph2acac=dibenzoylmethanide, L ‐Val=L ‐valine), [Y( 2 )(OTf)3] ( 3 ), and [Y2( 4 )(OTf)5] ( 5 ) ( 2 : [(S)P{N(Me)N?C(H)Py}3], 4 : [B{N(Me)N?C(H)Py}4]?) in solution and in the solid state. The structures found in the solid state are retained in solution, where averaged structures were observed. NMR diffusion measurements helped us to understand the nuclearity of compounds 3 and 5 in solution. 1H,19F HOESY and 19F,19F EXSY data revealed that the anions are specifically located in particular regions of space, which nicely correlated with the geometries found in the X‐ray structures.  相似文献   

5.
The new symmetrical diphosphonium salt [Ph2P(CH2)2PPh2(CH2C(O)C6H4Br)2] Br2 ( S ) was synthesized in the reaction of 1,2‐bis (diphenylphosphino) ethane (dppe) and related ketone. Further treatment with NEt3 gave the symmetrical α‐keto stabilized diphosphine ylide [Ph2P(CH2)2PPh2(CHC(O)C6H4Br)2] ( Y 1 ). The unsymmetrical α‐keto stabilized diphosphine ylide [Ph2P(CH2)2PPh2(CHC(O)C6H4Br)] ( Y 2 ) was synthesized in the reaction of diphosphine in 1:1 ratio with 2.3′‐dibromoacetophenone, then treatment with NEt3. The reaction of dibromo (1,5‐cyclooctadiene)palladium (II), [PdBr2(COD)] with this ligand ( Y 1 ) in equimolar ratio gave the new C,C‐chelated [PdBr2(Ph2P(CH2)2PPh2(C(H)C(O)C6H4Br)2)] ( 1 ) and with unsymmetrical phosphorus ylide [Ph2P(CH2)2PPh2C(H)C(O)C6H4Br] ( Y 2 ) gave the new P, C‐chelated palladacycle complex [PdBr2(Ph2P(CH2)2PPh2C(H)C(O)Br)] ( 2 ). These compounds were characterized successfully by FT‐IR, NMR (1H, 13C and 31P) spectroscopic methods and the crystal structure of Y 1 and 2 were elucidated by single crystal X‐ray diffraction. The results indicated that the complex 1 was C, C‐chelated whereas complex 2 was P, C‐chelated. These air/moisture stable complexes were employed as efficient catalysts for the Mizoroki‐Heck cross‐coupling reaction of several aryl chlorides, and the Taguchi method was used to optimize the yield of Mizoroki‐Heck coupling. The optimum condition was found to be as followed: base; K2CO3, solvent; DMF and loading of catalyst; 0.005 mmol.  相似文献   

6.
α‐Halogenoacetanilides (X=F, Cl, Br) were examined as H‐bonding organocatalysts designed for the double activation of C?O bonds through NH and CH donor groups. Depending on the halide substituents, the double H‐bond involved a nonconventional C?H???O interaction with either a H?CXn (n=1–2, X=Cl, Br) or a H?CAr bond (X=F), as shown in the solid‐state crystal structures and by molecular modeling. In addition, the catalytic properties of α‐halogenoacetanilides were evaluated in the ring‐opening polymerization of lactide, in the presence of a tertiary amine as cocatalyst. The α‐dichloro‐ and α‐dibromoacetanilides containing electron‐deficient aromatic groups afforded the most attractive double H‐bonding properties towards C?O bonds, with a N?H???O???H?CX2 interaction.  相似文献   

7.
The symmetric and unsymmetric phenylchlorohydrodigermanes can be isolated or characterized via partial halogenation of the Ge? H bonds of the symmetrical phenylhydrodigermanes Ph2(H)GeGe(H)2Ph, Ph(H)2GeGe(H)2Ph by chloromethyl methyl ether and carbontetrachloride. Some of these phenylchlorohydrodigermanes are formed by insertion of phenylchlorogermylene (PhGeCl) on the Ge? H or Ge? Cl bonds of the phenylchlorohydrogermanes. The hydrolysis of the monochloro phenylhydrodigermanes Ph2(Cl)GeGe(H)2 and Ph(Cl)(H)GeGe(H)2Ph leads to the phenyl phenylhydrogermyl digermoxanes [Ph2(H)GeGePh2]2O and [Ph(H)2GeGe(H)Ph]2O. Treatment of these oxides with the concentrated aqueous solutions of hydracides leads to the monofluorinated, brominated and iodinated phenylhydrodigermanes Ph2(H)GeGe(X)Ph2 and Ph(H)2GeGe(H)(X)Ph (X) = F, Br, I). Phenylchlorohydrodigermanes decompose thermally by α-elimination on one germanium atom with formation of germylene and phenylchlorohydrogermane. The physico-chemical IR. and NMR. study of these phenylhalogenohydrodigermanes indicates that, if the vGe? H frequency variations are mostly linked to the inductive effects of the substituents on the same germanium, the variations of the chemical shifts of the Ge? H protons seem to be due to many factors and especially to the inductive effect of the substituents on the germanium and the magnetic anisotropy of the Ge? X bonds.  相似文献   

8.
N‐chlorotriphenylphosphaneimine and its Application as an Educt for the Synthesis of Asymmetric PNP Cations. Crystal Structures of Ph3PNCl and [Ph3PNPEt3]Cl Ph3PNCl ( 1 ) originates in good yield as pale yellow crystals from the reaction of Ph3PNSiMe3 with phenyliodine dichloride. According to the crystal structure analysis 1 has a monomeric molecular structure without perceptible intermolecular contacts with distances P–N of 161.0 pm, N–Cl of 175.9 pm, and with a PNCl bond angle of 110.31°. 1 reacts with phosphines PR3 forming asymmetric PNP salts [Ph3PNPR3]Cl. This was tested by reactions with PEt3 and bis‐diphenyl phosphano ferrocene (DPPF). The crystal structure analysis of [Ph3PNPEt3]Cl ( 2 ) shows an almost symmetric PNP bridge with distances PN of 158.6 and 157.0 pm, and with a bond angle of 145.9°.  相似文献   

9.
[Ph3PNSbPh3Cl] ( 1 ) was prepared by oxidative addition of ClNPPh3 to triphenylstibine in dichloromethane solution. The compound is characterized by IR spectroscopy and by an X‐ray structure determination. 1 crystallizes in the monoclinic space group P21/c with four formula units per unit cell. Lattice dimensions at 193 K: a = 925.3(1), b = 1777.2(1), c = 1825.5(1) pm, β = 94.07(1)°, R1 = 0.0228. 1 forms monomeric molecules with tetrahedrally coordinated phosphorus and trigonal‐bipyramidally coordinated antimony atom, the atoms N and Cl being in axial positions. The bond lengths PN and SbN are 155.0(2) and 198.4(2) pm, respectively, the PNSb angle is 138.6(1)°. 1 reacts with iron trichloride to give the known phosphoraneiminato complex [FeCl2(NPPh3)]2.  相似文献   

10.
Polysulfonylamines. CLXXXIV. Crystal Structures of Molecular Triphenylphosphanegold(I) Di(4‐X‐benzenesulfonyl)amides: Isomorphism and Close Packing (X = Me, F, Cl, NO2) vs. Structure‐Determining C–X···Au/O Halogen Bonds (X = Br, I) In order to study the structure‐determining influence that halogen bonding can exert during the course of crystallization, solid‐state structures are compared for two previously reported and four new molecular gold(I) complexes of the type Ph3P–Au–N(SO2–C6H4–4‐X)2, each featuring linear P,N coordination at gold and two phenyl rings with varying p‐substituents X = Me, F, Cl, NO2, Br or I. The compounds were synthesized by reactions of Ph3PAuX (X = Cl or I) with the corresponding silver di(arenesulfonyl)amides, crystallized from dichloromethane, and characterized by low‐temperature X‐ray diffraction. The Me, F, Cl and NO2 congeners are isomorphic and crystallize without solvent inclusion in the chiral orthorhombic space group P212121 (Z′ = 1). These structures are governed by isotropic close packing via three‐dimensional 21 symmetry, incidentally supported by an invariant set of C–H···O=S hydrogen bonds, CH/π interactions and π/π stackings of aromatic rings; in particular, the hard halogen atoms of the fluoro and the chloro homologues are not involved in X···Au, X···O or X···X interactions. The higher homologues, with soft halogen atoms, were obtained as a dichloromethane hemisolvate for X = Br and a corresponding monosolvate for X = I, each triclinic in the centrosymmetric space group (Z′ = 1). Here, the primary structural effect is implemented by infinite chains in which translation‐related molecules are connected for the bromo compound by a bifurcated Au···Br(2)···O=S interaction, for the iodo congener by an equivalent Au···I(2)···O=S interaction and a short halogen bond C–I(1)···O=S. The latter bond is stronger than a similar C–Br···O=S interaction and induces a conformational adjustment of the (CSO2)2N group from the normal twofold symmetry in the bromo compound to an energetically unfavourable asymmetric form in the iodo homologue. In both cases, pairs of antiparallel molecular catemers are associated into strands via sixfold phenyl embraces, the strands are stacked to form layers, the solvent molecules are intercalated between adjacent layers, and the crystal packings are reinforced by a number of C–H···O=S hydrogen bonds and interactions of aromatic rings.  相似文献   

11.
Mechanistic aspects of the effect of the X and Y substituents (X = Me, H, CF3, CN, Br, Cl, F, OH, NH2; Y = H, NMe2, NH2, CN, NO2) on the carbonyl bond in 4-YC6H4C(O)X compounds are discussed on the basis of the 13C and 17O NMR data.  相似文献   

12.
As the first diphospha‐urea with P‐bonded protons, [TrtP(H)]2C=O ( 3 ) was found to be of amazing stability, which is thought to be due to the presence of the triphenylmethyl groups. Unlike known cyclic or non‐cyclic analogues, 3 showed next to no tendency to eliminate carbon monoxide. 3 was obtained by reaction of the dimeric phospha‐isocyanate (TrtPCO)2 ( 1 ) with LiAlH4, in which the intermediary phosphaalkene 2 was observed. Caused by its two asymmetric phosphorus atoms, 3 appeared as a mixture of two isomers, meso‐3 and rac‐3 (ratio: 20 : 1). Theoretical considerations, and the analysis of the proton‐coupled 31P NMR spectrum (spin system: AA′XX′), allowed the assignment of the signals to the two isomers. The action of anhydrous hydrogen chloride on 3 led to the cleavage of one P–C(:O)‐bond, and formation of an equimolar mixture of TrtPH2 ( 5 ) and TrtP(H)C(:O)Cl ( 6 ). Cleavage of a P–C(:O)‐bond in 3 was also observed in its reaction with tetramethylguanidine (TMG) or ammonia. As proved by 31P NMR spectroscopy in the case of TMG, the reaction proceeded via the phosphaalkene intermediate 8 . Acting as nucleophiles, TMG and ammonia substituted TrtP(H) in 3 , and the P,N‐ureas 9 and 10 , with TrtPH2 ( 5 ) as a side product, were obtained.  相似文献   

13.
Organometallic Compounds of Copper. XVIII. On the Reaction of the Alkyne Copper(I) Complexes [CuX(S‐Alkyne)] (X = Cl, Br, I; S‐Alkyne = 3,3,6,6‐Tetramethyl‐1‐thiacyclohept‐4‐yne) with the Phosphanes PMe3 and Ph2PCH2CH2PPh2 (dppe) The alkyne copper(I) halide complexes [CuX(S‐Alkyne)]n ( 2 ) ( 2 a : X = Cl, 2 b : X = Br, 2 c : X = I; S‐Alkyne = 3,3,6,6‐tetramethyl‐1‐thiacyclohept‐4‐yne; n = 2, ∞) add the phosphanes PMe3 and Ph2PCH2CH2PPh2 (dppe) to form the mono‐ and dinuclear copper compounds [(S‐Alkyne)CuX(PMe3)] ( 6 ) ( 6 a : X = Cl, 6 b : X = Br) and [(S‐Alkyne)CuX(μ‐dppe)CuX(S‐Alkyne)] ( 7 a : X = Cl, 7 b : X = Br, 7 c : X = I), respectively. By‐product in the reaction of 2 a with dppe is the tetranuclear complex [(S‐Alkyne)Cu(μ‐X)2Cu(μ‐dppe)2Cu(μ‐X)2Cu(S‐Alkyne)] ( 8 ). In case of the compounds 7 prolonged reaction times yield the alkyne‐free dinuclear copper complexes [Cu2X2(dppe)3] ( 9 ) ( 9 a : X = Cl, 9 b : X = Br, 9 c : X = I)). X‐ray diffraction studies were carried out with the new compounds 6 a , 6 b , 7 b , 8 , and 9 c .  相似文献   

14.
The Schiff base enaminones (3Z)‐4‐(5‐ethylsulfonyl‐2‐hydroxyanilino)pent‐3‐en‐2‐one, C13H17NO4S, (I), and (3Z)‐4‐(5‐tert‐butyl‐2‐hydroxyanilino)pent‐3‐en‐2‐one, C15H21NO2, (II), were studied by X‐ray crystallography and density functional theory (DFT). Although the keto tautomer of these compounds is dominant, the O=C—C=C—N bond lengths are consistent with some electron delocalization and partial enol character. Both (I) and (II) are nonplanar, with the amino–phenol group canted relative to the rest of the molecule; the twist about the N(enamine)—C(aryl) bond leads to dihedral angles of 40.5 (2) and −116.7 (1)° for (I) and (II), respectively. Compound (I) has a bifurcated intramolecular hydrogen bond between the N—H group and the flanking carbonyl and hydroxy O atoms, as well as an intermolecular hydrogen bond, leading to an infinite one‐dimensional hydrogen‐bonded chain. Compound (II) has one intramolecular hydrogen bond and one intermolecular C=O...H—O hydrogen bond, and consequently also forms a one‐dimensional hydrogen‐bonded chain. The DFT‐calculated structures [in vacuo, B3LYP/6‐311G(d,p) level] for the keto tautomers compare favourably with the X‐ray crystal structures of (I) and (II), confirming the dominance of the keto tautomer. The simulations indicate that the keto tautomers are 20.55 and 18.86 kJ mol−1 lower in energy than the enol tautomers for (I) and (II), respectively.  相似文献   

15.
The title compounds, (9‐fluoro‐4H‐chromeno[4,3‐c]isoxazol‐3‐yl)methanol, C11H8FNO3, (I), and (9‐chloro‐4H‐chromeno[4,3‐c]isoxazol‐3‐yl)methanol, C11H8ClNO3, (II), crystallize in the orthorhombic space group Pbca with Z′ = 1 and the triclinic space group P with Z′ = 6, respectively. The simple replacement of F by Cl in the main molecular scaffold of (I) and (II) results in significant differences in the intermolecular interaction patterns and a corresponding change in the point‐group symmetry from D2h to Ci = S2. These striking differences are manifested through the presence of C—H...F and the absence of O—H...O and C—H...O interactions in (I), and the absence of C—H...Cl and the presence of O—H...O and C—H...O interactions in (II). However, the geometry of the synthons formed by the O—H...N and O—H...X (X = F or Cl) interactions observed in the constitution of the supramolecular networks of both (I) and (II) remains similar. Also, C—H...O interactions are not preferred in the presence of F in (I), while they are much preferred in the presence of Cl in (II).  相似文献   

16.
Structures of Ionic Di(arenesulfonyl)amides. 2. Silver(I) Di(arenesulfonyl)amides and a Silver(I) (Arenesulfonyl)(alkanesulfonyl)amide: From Ribbons to Lamellar Layers Exhibiting Short C–H…Hal–C or C–Br…Br–C Interlayer Contacts Low‐temperature X‐ray crystal structures are reported for AgN(SO2C6H4‐4‐X)2 · H2O, where X is Cl ( 4 ) or Br ( 5 ), and for AgN(SO2Ph)(SO2Me) ( 6 ). Compounds 4 and 5 and the previously described F analogue ( 3 ) are isotypic, though not strictly isostructural (monoclinic, space group P21/c, Z = 4, but egregiously large discrepancies of x and z coordinates for corresponding atoms). Throughout this triad, glide‐plane related formula units are linked along the z axis to form infinite ribbons [(ArSO2)2N–Ag(μ‐H2O)], in which Ag extends its coordination number to five by accepting one Ag–O bond from each of the (ArSO2)2N ligands in the adjacent units. By means of O–H…O(S) hydrogen bonds, the ribbons are associated into lamellar layers parallel to the xz plane. Owing to the folded conformation of the anions, the layers display an inner polar region of Ag atoms, H2O molecules and N(SO2)2 groups, outer apolar regions of stacked pairs of aryl rings, and interlayer regions hosting the halogen atoms. Inspection of the latter areas provides sound evidence that the distinct juxtapositions of adjacent layers arise from specific interlamellar attractions and repulsions ( 3 : two C–H…F, all F…F beyond the van der Waals limit dW; 4 : one C–H…Cl, close packing of Cl atoms at Cl…Cl ≈ dW; 5 : one C–H…Br, one short Br…Br contact < dW, all other Br…Br > dW). Structure 6 (monoclinic, P21/n, Z = 4) consists of a lamellar coordination polymer, in which the cation accepts one Ag–N and three Ag–O bonds drawn from four different anions. On account of crystal symmetry, the extended ligand has its Ph and Me groups distributed on both sides of the sheet, the phenyl rings forming the apolar regions of the lamella, whereas the smaller methyl groups are integrated into the corrugated inorganic region by means of weak C–H…O hydrogen bonds.  相似文献   

17.
The mono (bistrifluoromethylamino-oxy)alkanes (CF3)2NOCXYZ (X = Y = F, Z = Cl; X = H, Y = F or Cl, Z = CH3; X = Y = F, Z = CH3; X = H, Y = Cl or Br, Z = CF3; X = Cl, Y = Br, Z = CF3) have been synthesised by treatment of appropriate halogenoalkanes, CHXYZ, with bistrifluoromethyl nitroxide. The 1,2-bis(bistrifluoromethylamino-oxy)alkanes (CF3)2NOCH2CXYON(CF3)2 were obtained as by-products in the reactions involving the ethanes CH3CHXY (X = H, Y = F or Cl; X = Y = F); these products, like their analogues (CF3)2NOCHFCF2ON(CF3)2 and (CF3)2NOCH2CCl2ON(CF3)2, were also prepared via attack of bistrifluoromethyl nitroxide on the corresponding ethenes.  相似文献   

18.
Fluoropropionic acids of the general formula CF3CXYCO2H ( X = F, Cl, Br ; Y = F, Cl, Br, H ) were obtained by the sonochemically promoted reaction of fluorohalogenoethanes CF3CXYZ ( Z = Cl, Br ) with zinc and carbon dioxide. Penta- and tetrafluoroethanes ( X = Y = F and X = F, respectively ) gave good yields ( 35 – 47 % ) of the acids; with trifluoro derivatives the yields were substantially lower. Hydrogenolysis of the CCl and CBr bonds in CF3CFClCO2H and CF3CFBrCO2H afforded 2,3,3,3-tetrafluoropropionic acid.  相似文献   

19.
The positive electrostatic potentials (ESP) outside the σ‐hole along the extension of O? P bond in O?PH3 and the negative ESP outside the nitrogen atom along the extension of the C? N bond in NCX could form the Group V σ‐hole interaction O?PH3?NCX. In this work, the complexes NCY?O?PH3?NCX and O?PH3?NCX?NCY (X, Y?F, Cl, Br) were designed to investigate the enhancing effects of Y?O and X?N halogen bonds on the P?N Group V σ‐hole interaction. With the addition of Y?O halogen bond, the V S, max values outside the σ‐hole region of O?PH3 becomes increasingly positive resulting in a stronger and more polarizable P?N interaction. With the addition of X?N halogen bond, the V S, min values outside the nitrogen atom of NCX becomes increasingly negative, also resulting in a stronger and more polarizable P?N interaction. The Y?O halogen bonds affect the σ‐hole region (decreased density region) outside the phosphorus atom more than the P?N internuclear region (increased density region outside the nitrogen atom), while it is contrary for the X?N halogen bonds. © 2015 Wiley Periodicals, Inc.  相似文献   

20.
The title compound, [CrSn(C6H5)3(C7H6NO2)3Cl][Sn(C6H5)3Cl(CH4O)], was obtained from the reaction of Ph3SnCl with the complex [Cr(C7H6NO2)3] in methanol. The structure contains [Ph3SnCl(MeOH)] (A) and [Ph3SnClCr(C7H6NO2)3] (B) mol­ecules. In mol­ecule A, the Sn atom of Ph3SnCl is coordinated by one methanol mol­ecule. In mol­ecule B, the Sn atom of Ph3SnCl is coordinated by one carboxyl­ate O atom of [Cr(C7H6NO2)3]. Mol­ecules A and B are connected through an O—H⋯O hydrogen bond between a carboxyl­ate O atom and the methanol OH group. Weak C—H⋯Cl inter­actions and O—H⋯O hydrogen bonds extend the components of (I) into a two‐dimensional network.  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