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 relative permittivities (?r20) of 4-dimethylaminostyrene (I) and of 0.5–3.0 M solutions of N,N-dimethyl-4-toluidine (DMT) were measured in nitrobenzene (NB) and in a mixture of NB with diisopropyl ether (NB + E), the DMT solutions modelled those of I. An increase in [DMT] from 0.5 to 3.0 M caused a decrease in ?r20 from 32.3 to 18.3 in NB, while only from 20.8 to 18.3 in (NB + E) at fixed [NB]. The great difference between the ?r20 of 0.5 and 3.0 M DMT solutions in NB explains the previously observed drop in the percentage rate of polymerization (RP) of I, in NB initiated with CCl3COOH, with increasing [I]o. In (NB + E) mixture with fixed [NB], the percentage RP and the limiting viscosity number [ν] of polyl increase with increasing [I]o. The observed dependences of both RP (in Msec?1) and [ν] on [I]0 confirm the validity of the suggested polymerization scheme for I initiated with CCl3COOH, and make possible determination of the characteristics of the elementary polymerization reactions.  相似文献   

2.
The reaction of trans-[RhCl(CO)(DPM)]2 (DPM = Ph2PCH2PPh2) with dimethylacetylenedicarboxylate (DMA) and hexafluoro-2-butyne (HFB) yield the novel species [Rh2Cl2(μ-CO)(μ-Acet)(DPM)2] (Acet = DMA, HFB). The X-ray structure determination of the DMA derivative indicates that the complex has the acetylene molecule coordinated as a cis-dimetallated olefin and also contains a ketonic carbonyl ligand. The long Rh?Rh separation (3.3542(9) Å) suggests no metal—metal bond and the RhC(O)Rh angle (116.0(4)°) suggest sp2 hybridization of the carbonyl carbon atom. Similarly the geometry at the acetylene ligand and the CC distance of the coordinated acetylene moiety (1.32(1) Å) are consistent with the dimetallated olefinic formulation. This represents the first reported characterization of a ketonic carbonyl complex outside the Ni triad. These novel complexes have also been formed by the direct insertion of the acetylene molecules into the formal RhRh bond in [Rh2Cl2(μ-CO)(DPM)2].  相似文献   

3.
The complexation of manganese(II), cobalt(II) and nikel(II) with bromide ions has been studied in N,N-dimethylacetamide(DMA) by calorimetry and spectrophotometry. The formation of [MBr]+, [MBr2] and [MBr3] (M=Mn, Co, Ni) was revealed in all the metal systems. Interestingly, the complexation is significantly enhanced in DMA over N,N-dimethylformamide (DMF). This is unusual because physicochemical properties of DMA and DMF as solvent are similar. Furthermore, extracted electronic spectra of individual complexes of NiII suggested the presence of a geometry equilibrium, [NiBr(DMA)5]+=[NiBr(DMA)4]++ DMA, in DMA. A similar geometry equilibrium is also suggested, [NiBr2(DMA)3]=[NiBr2(DMA)2]+DMA. Such geometry equilibria were not observed in DMF. With regard to cobalt(II), electronic spectra show the presence of the four-coordinated [CoBr(DMA)3]+ complex in DMA, unlike the six-coordinated [CoBr(DMF)5]+ one in DMF. These facts suggest that a specific strong steric interaction operates between coordinating solvent molecules, which plays a key role in the complexation behavior of the divalent transition metal ions in DMA.  相似文献   

4.
Photopolymerization of methyl methacrylate in bulk and in solution at 40° using triethylamine nitrobenzene (TEA-NB) complex as photoinitiator was studied kinetically. Initiator order x, given by the relation Rpα([TEA][NB])x, was 0.28 for [TEA][NB] < 25 × 10 ?8 mol2 · l?2; for higher values of [TEA][NB], x was practically zero. Monomer order was 1.1 in benzene and pyridine but much less than unity (0.65–0.70) in carbon tetrachloride and chloroform. Kinetic analysis indicated that the initiation process was monomer and solvent dependent. The halomethane solvents enhanced the polymerization rate through their active participation in the initiation or radical-generation steps. End-group analysis indicated incorporation of basic (amino) end-groups in the polymers. The kinetic non-ideality was explained on the basis of significant initiator-dependent termination through primary radicals or via degradative transfer to initiator.  相似文献   

5.
Dimethylaniline (DMA) induces chain transfer in the zinc bromide-complexed donor–acceptor polymerizations of styrene–acrylonitrile to form alternating copolymers. The Mayo plots are linear, but the rates decrease with increase in DMA and degradative chain transfer occurs. Although conventional free-radical transfer agents have negligible effect on the rates or molecular weights, a twentyfold reduction of molecular weight is obtained with DMA. Spectroscopic data indicate the formation of an equimolar complex of DMA and ZnBr2, but the lowering of molecular weight is not attributable to the reduction of the effective ZnBr2, concentration. A possible mechanism involving a competition between [ZnBr2(DMA)2] and [ZnBr2,DMA,AN] is suggested.  相似文献   

6.
Kinetic and stereochemical studies show nucleophilic assistance by dimethylformamide (DMF), dimethylacetainide (DMA), hexamethylphosphotriamide (HMPT) and N-methylimidazole (NMI) in racemization and solvolysis of menthylchloro(phenyl)phosphonate, 1a, and O-ethylchloro(phenyl)thiophosphonate, 2. Similar orders of nucleophilic reactivity (Nu = NMI?HMPT>DMF>DMA), and identical rate-laws (vrac=k [M-Cl] [Nu]2 and vH2O = k' [M-Cl] [H2O] [Nu]) are consistent with a common mechanism, governed by entropy (?60 u.e< ΔS<?40u.e). Analogies between reaction mechanisms at silicon and phosphorus are clearly evidenced. A two-step process, involving rate-determining attack on a pentacoordinate complex is discussed.  相似文献   

7.
Heck vinylation of 2-bromo-6-methyl-3-substituted pyridines using η3-allylpalladium chloride dimer/P(o-Tol)3 complex/toluene and dimethylacetamide (DMA) as co-solvent with methyl acrylate is reported. Electronic and steric effects were investigated engaging diversely 2-bromo-3,6-disubstituted pyridines. As application, a new synthesis of the 6-methyl cyclopenta[b]pyridinone building-block connecting Heck vinylation, alkene reduction and Dieckmann condensation is described.  相似文献   

8.
Reaction of 4,4′‐tolanedisulfonic acid, H2TDS, with zinc hydroxide in dimethylacetamide, DMA, under solvothermal conditions led to the coordination polymer Zn(TDS)(DMA)3 ( I ). In the crystal structure [trigonal, P3221, Z=3, a=1175.0(1) pm, c=1949.5(1) pm, R1; wR2 (Io> 2σ(Io))=0.0393; 0.0921] the disulfonate anions linked the Zn2+ ions into helical chains according to 1[Zn(DMA)3/1(TDS)2/2] ( I ) causing the chirality of the compound. By using higher concentrations of H2TDS in the starting mixture the compound [HDMA]2[Zn(TDS)2(DMA)3](DMA)2 ( II ) was formed. The structure [monoclinic, Cc, Z=4, a=1201.5(1) pm, b=1996.0(1) pm, c=2749.2(2) pm, β=101.897(2)°, R1; wR2 (Io> 2σ(Io))=0.0699; 0.2017] displayed the complex anion [Zn(TDS)2(DMA)3]2? which was a perfect cut‐off of the helical chain in I . Charge compensation was achieved by protonated DMA molecules. If N‐methylpyrrolidone, NMP, was chosen as a solvent, the sulfonate Zn(TDS)(NMP)3 ( III ) [monoclinic, I2/a, Z=4, a=1575.7(1) pm, b=1077.3(1) pm, c=1870.0(1) pm, β=101.189(9)°, R1; wR2 (Io> 2σ(Io))=0.0563; 0.1320] was obtained. Similarly to the findings for I , the formation of chains according to 1[Zn(NMP)3/1(TDS)2/2] was observed. However, due to the more bulky NMP molecules these chains were no longer helical but straight instead.  相似文献   

9.
The reaction of acetone-4-(2-methoxy-phenyl)thiosemicarbazone with triosmium cluster Os3(CO)11(NCMe) results in the formation of the cluster with the μ2 chelate-bridging ligand coordinated by S and N1 atoms, which was studied by X-ray diffraction analysis. Reaction of (3aR, 3bR, 4aR, 5aS)-5a-hydroxy-3,4,4-trimethyl-3a,3b,4,4a,5,5a-hexahydrocyclopropa[3, 4]cyclopenta[1,2-c]pyrazole-1-carbothioic acid amide with Os3(CO)11(NCMe) gives rise to the complex with the bridging ligand coordinated by sulphur atom. Further transformation of the complex in hot benzene results in tautomeric rearrangement of the organic ligand and the cleavage of the pyrazolinol cycle to form an open chain tautomer. Unusual silica gel induced oxidative cleavage of the cyclopropane ring in the open chain derivative and epoxidation of cycloalcane C-C bond are observed on the air.  相似文献   

10.
By an example of cysteamine iron nitrosyl complex {Fe2[S(CH2)2NH3]2(NO)4}SO4 ? 2.5H2O (CAC) it was shown for the first time that the NO donor hydrolysis in the presence of ferricytochrome c (cyt c3+) affords the iron nitrosyl complex NO—cyt c3+. It was found that cyt c3+ can serve as a depot for NO evolved during the hydrolysis of CAC. In the presence of CAC, the rate of NO—cyt c3+ complex decomposition to NO and cyt c3+ depends on the molar ratio [cyt c3+]: [CAC] and at [cyt c3+]: [CAC] = 0.3 it was found to be lower than that in decomposition of CAC in the absence of cyt c3+. As a result, the total NO evolving process becomes 5.6 times more prolonged. The number of NO groups evolved from CAC can be determined by the reaction of CAC with cyt c3+ in the presence of ferricyanide: at most one NO group is evolved to a solution in the spontaneous hydrolysis of CAC (pH 7.0), and no less than three of them are evolved from oxidized CAC.  相似文献   

11.
Synthesis and Characterization of 2‐O‐Functionalized Ethylrhodoximes and ‐cobaloximes 2‐Hydroxyethylrhodoxime and ‐cobaloxime complexes L—[M]—CH2CH2OH (M = Rh, L = PPh3, 1 ; M = Co, L = py, 2 ; abbr.: L—[M] = [M(dmgH)2L] (dmgH2 = dimethylglyoxime, L = axial base) were obtained by reaction of L—[M] (prepared by reduction of L—[M]—Cl with NaBH4 in methanolic KOH) with BrCH2CH2OH. H2O—[Rh], prepared by reduction of H[RhCl2(dmgH)2] with NaBH4 in methanolic KOH, reacted with BrCH2CH2OH followed by addition of pyridine yielding py—[Rh]—CH2CH2OH ( 3 ). Complexes 1 and 3 were found to react with (Me3Si)2NH forming 2‐(trimethylsilyloxy)ethylrhodoximes L—[Rh]—CH2CH2OSiMe3 (L = PPh3, 4 ; L = py, 5 ). Treatment of complex 1 with acetic anhydride resulted in formation of the 2‐(acet oxy)ethyl complex Ph3P—[Rh]—CH2CH2OAc ( 6 ). All complexes 1 — 6 were isolated in good yields (55—71 %). Their identities were confirmed by NMR spectroscopic investigations ( 1 — 6 : 1H, 13C; 1 , 4 , 6 : 31P) and for [Rh(CH2CH2OH)(dmgH)2(PPh3)]·CHCl3·1/2H2O ( 1 ·CHCl3·1/2H2O) and py—[Rh]—CH2CH2OSiMe3 ( 5 ) by X‐ray diffraction analyses, too. In both molecules the rhodium atoms are distorted octahedrally coordinated with triphenylphosphine and the organo ligands (CH2CH2OH and CH2CH2OSiMe3, respectively) in mutual trans position. Solutions of 1 in dmf decomposed within several weeks yielding a hydroxyrhodoxime complex “Ph3P—[Rh]—OH”. X‐ray diffraction analysis exhibited that crystals of this complex have the composition [{Rh(dmg)(dmgH) (H2O)(PPh3)}2]·4dmf ( 7 ) consisting of centrosymmetrical dimers. The rhodium atom is distorted octahedrally coordinated. Axial ligands are PPh3 and H2O. One of the two dimethylglyoximato ligands is doubly deprotonated. Thus, only one intramolecular O—H···O hydrogen bridge (O···O 2.447(9)Å) is formed in the equatorial plane. The other two oxygen atoms of dmgH and dmg2—, respectively, act as hydrogen acceptors each forming a strong (intermolecular) O···H′—O′ hydrogen bridge to the H′2O′ ligand of the other molecule (O···O′ 2.58(2)/2.57(2)Å).  相似文献   

12.
《中国化学》2018,36(10):904-908
The peculiar electronic structure of scandium phosphinoalkylidene complex [LSc{C(SiMe3)PPh2}THF] (L=[MeC(NDIPP)CHC(NDIPP)Me]), DIPP=2,6‐(iPr)2C6H3) leads to an interesting versatile reactivity, which is demonstrated both experimentally and computationally. The complex undergoes [2+2] cycloaddition reactions with alkynes, and easily activates various X—O bonds such as C—O of propylene oxide, N—O of 3,5‐dimethylisoxazole, B—O of pinacolborane and Si—O of triethoxysilane. These reactions occur on the Sc—C bond of the phosphinoalkylidene complex. Interestingly, the Sc—P bond can also be activated as the presence of a Sc—C—P three center π interaction in the complex allows performing C—F activation of 2,6‐difluoropyridine and 1,2 addition with imine or ketone. The complex also reacts with metal complexes, [(COD)RhCl]2 and (Ph3P)AuCl, to form structural intriguing heterobimetallic complexes.  相似文献   

13.
A novel supramolecular graft copolymer(SGP) composed of viologen-containing copolymer(P(DMA-co-diEV)) as the main chain and Np ended PNIPAM(Np-PN1PAm) as the grafts is prepared(DMA:N,N- dimethylacryamide,diEV:ethylviologen dimer,Np:naphthalene,PNIPAM:poly(N-isopropylacrylamide)). The grafting is based on the triple complexation among a host of cucurbit[8]uril(CB[8])and two guests of diEV and Np,which is characterized by UV-vis spectra and ITC.Temperature sensitive property of PNIPAm moiety allows SGP to self-assemble into non-covalently connected micelle(NCCM) at high temperature. The micelles are sensitive to reducing agents,for example Na2S2O3,which breaks the current inclusion complex pair and induces aggregation.  相似文献   

14.
The title antimony(III) complex, [Sb(C32H16N8)]Cl or [SbPc]Cl (where Pc = C32H16N82−), has been obtained from the reaction of pure powdered antimony with 1,2-di­cyano­benzene under a stream of ICl vapour. The asymmetric unit of this complex consists of an [SbPc]+ cation and a Cl anion. The phthalocyaninate residue [SbPc]+ is not planar. The Sb atom lies 1.057 (3) Å from the plane defined by the four iso­indole N atoms. A combination of ionic and donor–acceptor interactions links the [SbPc]Cl mol­ecules to form centrosymmetric [(SbPc)Cl]2 pseudo-dimers in the crystal. The Sb—Cl distances in the pseudo-dimer are not equivalent [3.043 (2) and 3.201 (2) Å]. The pseudo-dimers are weakly linked through Cl⃛H—Cbenzo interactions to form a three-dimensional network. As a result of these interactions, the four Sn—Nisoindole bond lengths in the [SbPc]+ residue are not equivalent and the symmetry of the Sb—N core is only close to Cs.  相似文献   

15.
Five solvates, [CdBr2(DMF)] n , [CdBr2(DMA)] n , [CdI2(DMF)] n , [Cd(DMF)6][Cd2I6], and {[Cd(DMA)6][Cd5I12] n } m , were isolated from the ternary systems CdX2–L–H2O (X = Br, I; L = N,N-dimethylacetamide, N,N-dimethylformamide) and characterized by the X-ray single crystal analysis. The structures of the first three solvates is similar to each other in structures and represent a one-dimensional polymer chain, the fourth solvate has the discrete structure containing [Cd(DMF)6]2+ and [Cd2I6]2– ions, and the fifth solvate contains discrete [Cd(DMA)6]2+ cations and the polymer anionic fragment [Cd5I12] n 2n.  相似文献   

16.
A C32 aetio porphyrin isolated from Gilsonite bitumen (ca. 60 × 106yr) was assigned unambiguously as aetioporphyrin—III by 1H NMR analysis of its bis[porphyrinato-mercury(II) acetato]mercury(II) complex and by comparison with an authentic sample. The occurrence of this compound provides the first direct evidence that the petroporphyrins of Gilsonite are the product of reductive degradation of naturally-occurring chlorophylls, and conversion of the chlorin to the porphyrin system.  相似文献   

17.
The trans-[Fe(cyclam)(NO)Cl]Cl2 complex was synthesized by the reaction of cis-[Fe(cyclam)Cl2]Cl with NO gas. The X-ray structure of the complex showed that the [Fe–NO] moiety is linear, consistent with the NO+ character of the nitric oxide ligand. This suggestion was reinforced by the IR data, which showed the νNO at 1888 cm−1. The cyclic voltammogram of the trans-[Fe(cyclam)(NO)Cl]2+ complex presented three electrochemical processes at −0.70, 0.08 and 0.40 V versus Ag/AgCl. The first and last redox processes are centered at the NO ligand, whereas the second is characteristic of the generated aqua species, trans-[Fe(cyclam)Cl(H2O)]2+. Upon irradiation at 330 nm, pH 3.4, the title complex releases the NO moiety with the concomitant generation of the trans-[Fe(cyclam)(H2O)Cl]+ complex as suggested by electronic and IR spectroscopy as well as by cyclic voltammetry technique.  相似文献   

18.
The copper(II)—l-carnosine (L) system has been re-investigated in aqueous solution, at I = 0.1 mol dm−1, different temperatures (5⩽t⩽45°C) and with metal to ligand ratios ranging from 3:1 to 1:3. Both potentiometry and visible spectrophotometry were employed. From an overall consideration of all experiments, [CuLH]2+, [CuL]+, [CuLH−1]°, [Cu2L2H−2]° and [Cu2LH−1]2+ were recognized as the species which provide the best interpretation of experimental data. The complex formation constants, determined at different temperatures, allowed us to obtain reliable values of ΔH° and good estimates of ΔC°p. From visible spectrophotometric measurements, carried out at different pH and metal to ligand ratios, it was possible to calculate the electronic spectrum of each complex formed in solution. A structure is also proposed for each species, on the basis of thermodynamic and spectral results.  相似文献   

19.
The compound [RuCl2(CO)(DMA)(PPh3)2] [DMA = dimethylacetamide] was obtained from [RuCl3(PPh3)2-(DMA)] · DMA and CO in DMA. Orange crystals of [RuCl2(CO)(DMA)(PPh3)2] · 1/2CH2Cl2 were isolated by slow evaporation of a CH2Cl2/DMA solution and its structure was determined by single crystal X-ray diffraction. The analogous compounds containing DMF and DMSO were obtained from the precursor ttt-[RuCl2(CO)2(PPh3)2]. Characterization of the other complexes is based on i.r. and n.m.r. spectroscopy, including 31P{1H} data.  相似文献   

20.
In the title salt, (C6H8N4)[Mn(C14H8O4)2(C6H6N4)2]·6H2O, the MnII atom lies on an inversion centre and is coordinated by four N atoms from two 2,2′‐biimidazole (biim) ligands and two O atoms from two biphenyl‐2,4′‐dicarboxylate (bpdc) anions to give a slightly distorted octahedral coordination, while the cation lies about another inversion centre. Adjacent [Mn(bpdc)2(biim)2]2− anions are linked via two pairs of N—H...O hydrogen bonds, leading to an infinite chain along the [100] direction. The protonated [H2biim]2+ moiety acts as a charge‐compensating cation and space‐filling structural subunit. It bridges two [Mn(bpdc)2(biim)2]2− anions through two pairs of N—H...O hydrogen bonds, constructing two R22(9) rings, leading to a zigzag chain in the [2] direction, which gives rise to a ruffled set of [H2biim]2+[Mn(bpdc)2(biim)2]2− moieties in the [01] plane. The water molecules give rise to a chain structure in which O—H...O hydrogen bonds generate a chain of alternating four‐ and six‐membered water–oxygen R42(8) and R66(12) rings, each lying about independent inversion centres giving rise to a chain along the [100] direction. Within the water chain, the (H2O)6 water rings are hydrogen bonded to two O atoms from two [Mn(bpdc)2(biim)2]2− anions, giving rise to a three‐dimensional framework.  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