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 93 毫秒
1.
N,N-Dimethylneopentylamine reacts with Pd(MeCO2)2 to give a novel trinuclear cyclopalladated complex [Me2NCH2CMe2CH2Pd(μ-MeCO2)2Pd(μ-MeCO2)2PdCH2CMe2CH2NMe2]?-0.5C6H6 (I). The reaction of I with PPh3 affords both trans-[Pd(MeCO2)2(PPh3)2] (II) and [Pd(CH2CMe2CH2NMe2)(MeCO2)(PPh3)] (III). The reaction of III with LiCl yields a mononuclear cyclopalladated complex, [Pd(CH2CMe2CH2NMe2)Cl(PPh3)] (IV).  相似文献   

2.
Reactions of 2,5‐dibromothiophene, 1 , with [Pd2(dba)3]?dba [Pd(dba)2; dba = dibenzylideneacetone] in the presence of N‐donor ligands such as 2,2′‐bipyridine (bpy) and 4,4′‐di‐tert‐butyl‐2,2′‐bipyridine (dtbbpy) give arylpalladium complexes of cis‐[2‐(5‐BrC4H2S)PdBrL2], 2a, b [L2 = bpy ( 2a ), L2 = dtbbpy ( 2b )], and cis‐cis‐L2PdBr[2,5‐(C4H2S‐)PdBr(L2)], 3a, b [L2 = bpy ( 3a ), L2 = dtbbpy ( 3b )]. Treatment of cis complexes 2a, b and 3a, b with CO causes the insertion of CO into the Pd? C bond to give the aroyl derivatives of palladium complexes of cis‐[2‐(5‐BrC4H2S)COPdBrL2], 4a, b [L2 = bpy ( 4a ), L2 = dtbbpy ( 4b )], and cis‐cis‐[(L2)(CO)BrPdC4H2S‐PdBr(CO)(L2)], 5a, b [L2 = bpy ( 5a ) and L2 = dtbbpy ( 5b )], respectively. Treating complexes 2a, b with 1 mole equivalent of isocyanide XyNC (Xy = 2,6‐dimethylphenyl) gave iminoacyl complexes cis‐[2‐(5‐BrC4H2S)C?NXyPdBrL2], 6a, b [L2 = bpy ( 6a ), L2 = dtbbpy ( 6b )], and a 3‐fold excess of isocyanide XyNC (Xy = 2,6‐dimethylphenyl) gave triiminoacyl complexes [2‐(5‐BrC4H2S)(C?NXy)3 PdBr], 7 . Cyclization reactions of 6a, b with 3 mole equivalents of isocyanide XyNC (Xy = 2,6‐dimethylphenyl) or cyclization reaction of 7 with 1 mole equivalent of isocyanide XyNC (Xy = 2,6‐dimethylphenyl) both gave tetraiminoacyl complexes of [2‐(5‐BrC4H2S)(C?NXy)4PdBr], 8 , which was also obtained by the reaction of 1 or 2a, b with a 4‐fold excess of isocyanide XyNC with or without add Pd(dba)2. Similarly, complexes 3a and b were also reacted with 2 mole equivalents of isocyanide XyNC (Xy = 2,6‐dimethylphenyl) to give iminoacyl complexes cis‐cis‐[(L2)(CNXy)BrPdC4H2S‐PdBr(CNXy)(L2)], 10a, b [L2 = bpy ( 10a ), L2 = dtbbpy ( 10b )] and an 8‐fold excess of isocyanide XyNC (Xy = 2,6‐dimethylphenyl) afforded tetraiminoacyl complexes of [2,5‐(C4H2S)(C?NXy)8Pd2Br2], 11 . Complexes 2a, b and 3a, b reacted with TlOTf [(TfO = CF3SO3)] in CH2Cl2 to give 9a, b and 12a, b , respectively, in a moderate yield. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

3.
Unit cell dimensions of seven oxofluorometallates of transition metals were investigated by powder X-ray diffraction method. The compounds K3NbO2F4 and K3TaO2F4 were found to be isomorphous with cubic (FCC) structure and having lattice parameters 8.885 and 8.942 Å respectively. Similarly, the compounds K2NbOF5 (a = 8.367 Å, c = 13.038 Å) and K2TaOF5 (a = 8.463 Å, c = 13.139 Å) were also found to be isomorphous with a tetragonal structure. The compound K3Zr2O2F7 (a = 9.367 Å) was found to possess a cubic (FCC) structure. Both K2V2O5F2 (a = 6.739 Å, c = 10.635 Å) and K2VO3F (a = 5.984 Å, c = 10.914 Å) have a hexagonal structure.  相似文献   

4.
The PH bond of dialkylphosphites (dimethylphosphite, 5,5-dimethyl-1,3-dioxa-2-phosphorinane and 4,4,5,5-tetramethyl-1,3-dioxa-2-phospholane) oxidatively adds to irClL2(L = PPh3, AsPh3) and IrCl(PMe2Ph)3 generated in situ to give six-coordinate hydrido(dialkylphosphonato)iridium(III) complexes, e.g. IrHClL2[{(MeO)2-PO}2H] and IrHCl(PMe2Ph)3[PO(OMe)2]. Addition of triphenylphosphine to a solution containing [IrCl(C8H14)2]2 and dimethylphosphite in a 1:2 mol ratio gives a five-coordinate hydrido (dimethylphosphonato)iridium(III) complex IrHCl(PPh3)2{PO(OMe)2}, from which six-coordinate pyridine and acetylacetonato complexes IrHCl(PPh3)2(C5H5N){PO(OMe)2} and IrH(PPh3)2(acac){PO(OMe)2} can be obtained. The ligand arrangements in the various complexes are inferred from IR, 1H and 31P NMR data.  相似文献   

5.
The iodo-bridged sulfur ylide complex [Pd(μ-I)((CH2)2(SO)(CH3))]2 (1) when treated with dithiolates, acetylacetone and various Lewis bases gave [Pd((CH2)2(SO)(CH3))(S ∼ S)] (S ∼ S = S2CN(C2H5)2, S2COC2H5 and S2P(OC2H5)2), [Pd((CH2)2(SO)(CH3))(acac)] (acac = acetylacetonate) and [PdI((CH2)2(SO)(CH3))(base)]a (base = PPh3, (P(OMe)3, P(OPh)3 and C5H5N). In the presence of a phase transfer catalyst (PTC). The reactions rates and yields were greatly increased. Reaction of several related sulfur ylide complexes with I2, HI or aqueous NaOH gave 1. The single crystal structure of [Pd((CH2)2(SO)(CH3))2] was determined (orthorhombic, Pbcn, a 13.379(2), b 8.081(1), c 9.048(2) Å, V 978.2 Å3, Z = 4). The compound has a rather long PdCH2 bond (2.096(1) Å, mean).  相似文献   

6.
Single crystals of (NMe4)(HF2) were obtained during attempted recrystallization of NMe4F from fluoroolefin. X‐ray diffraction data show that (NMe4)(HF2) crystallizes in the orthorhombic space group Pmmn with unit cell dimensions a = 6.535(2), b = 8.688(3), and c = 5.333(2) Å. The symmetric and virtually linear HF2 anions exhibit a short F···F distance of 2.256(2) Å. The both crystal structures of (NMe4)(H2F3) (orthorhombic, Pbca, a = 8.509(1), b = 11.273(2), and c = 14.880(2) Å) and CsH2F3 (orthorhombic, P212121, a = 7.345(3), b = 9.126(4), and c = 11.444(4) Å) contain dihydrogentrifluoride anions, H2F3?, which have a bent shape and F···F distances of 2.30‐2.34Å.  相似文献   

7.
The amino acid esters of hydroxypropyl cellulose (HPC) [R′ = H ( 2a ), CH3 ( 2b ), CH2CH(CH3)2 ( 2c ), CH2CONH2 ( 2d ), CH2CH2CONH2 ( 2e ), CH2CH2CH2CH2 NHOCOC(CH3)3 ( 2f )] were synthesized in good yield by the reaction of t‐butoxycarbonyl (t‐Boc)‐protected amino acids with hydroxy groups of HPC ( 1 ; molar substitution (MS), 4.61). The amino acid functionalities displaying varied chemical nature, shape, and bulk were utilized and the bulk of the substituent on the α‐carbon of amino acids was elucidated to be of vital significance for the observed degree of incorporation (DSEst). The 1H NMR spectra and elemental analysis were employed to determine the degree of incorporation of amino acid moiety (DSEst) and almost complete substitution of the hydroxy protons was revealed for 2a , 2b , and 2f . The presence of the peaks characteristic of the carbonyl group in the FTIR spectra furnished further evidence for the successful esterification of HPC. The starting as well as the resulting polymers ( 1 and 2a – f ) were soluble in polar organic solvents; however, the esterification of 1 with bulky organic moieties resulted in an increased hydrophobicity as all of the amino acid‐functionalized polymers ( 2a – f ) were insoluble in water. The onset temperatures of weight loss of 2a – f were 175–230 °C, indicating fair thermal stability. The amino acid functionalization led to the enhanced polymer chain stiffness, and the glass transition temperatures of the derivatized polymers were 30–40 °C higher than that of 1 (Tg 3.9 °C; cf. Tg of 2a – f , 35.1–43.3 °C). © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 2326–2334, 2008  相似文献   

8.
Various cyclic phosphonium structures are formed in high yield by the deprotection of unstable phosphine-aldehydes in acidic solution. When there is a methylene spacer between the phosphine and the aldehyde, a phosphonium ion [PHR2CH2CH(OEt)2]Br2, R=iPrOH, Et is obtained. Reaction of these phosphonium salts with water produces the dimers [-PR2CH2CH(OH)-]2[Br]2 R = iPr, Et. When there is an ethylene spacer as in PPh2CH2CH2CH(OCH2CH2O), a remarkable tetramer with a 16-membered ring [-PPh2CH2CH2CH (OH)-]4[Cl]4 forms as one diastereomer in hydrochloric acid solution. Reaction of HCl with the protected phosphine-aldehyde with a propylene spacer (PPh2CH2CH2CH2CH(OCH2CH2O)) results in the formation of the monomeric phosphonium salt [-PPh2 CH2CH2CH2CH(OH)-]Cl with a 5-membered ring. Solid state structures of different ring types were determined using X-ray diffraction experiment.  相似文献   

9.
Building upon our earlier results on the synthesis of electron‐precise transition‐metal–boron complexes, we continue to investigate the reactivity of pentaborane(9) and tetraborane(10) analogues of ruthenium and rhodium towards thiazolyl and oxazolyl ligands. Thus, mild thermolysis of nido‐[(Cp*RuH)2B3H7] ( 1 ) with 2‐mercaptobenzothiazole (2‐mbtz) and 2‐mercaptobenzoxazole (2‐mboz) led to the isolation of Cp*‐based (Cp*=η5‐C5Me5) borate complexes 5 a , b [Cp*RuBH3L] ( 5 a : L=C7H4NS2; 5 b : L=C7H4NOS)) and agostic complexes 7 a , b [Cp*RuBH2(L)2], ( 7 a : L=C7H4NS2; 7 b : L=C7H4NOS). In a similar fashion, a rhodium analogue of pentaborane(9), nido‐[(Cp*Rh)2B3H7] ( 2 ) yielded rhodaboratrane [Cp*RhBH(L)2], 10 (L=C7H4NS2). Interestingly, when the reaction was performed with an excess of 2‐mbtz, it led to the formation of the first structurally characterized N,S‐heterocyclic rhodium‐carbene complex [(Cp*Rh)(L2)(1‐benzothiazol‐2‐ylidene)] ( 11 ) (L=C7H4NS2). Furthermore, to evaluate the scope of this new route, we extended this chemistry towards the diruthenium analogue of tetraborane(10), arachno‐[(Cp*RuCO)2B2H6] ( 3 ), in which the metal center possesses different ancillary ligands.  相似文献   

10.
Abstract

Some bisacylphosphonates are biologically active in calcium related disorders, such as bone resorption and ectopic calcification.1 In the course of our studies directed towards the preparation of stable, pharmaceutically acceptable salts of bisacylphosphonates, we found that in the presence of N,N'-dibenzylethylenediamine (benzathine), adipoylbisphosphonate (AdBP) underwent cyclization to 2-hydroxy-2-phosphonocyclopentanecarbonylphosphonic acid. Similarly, pimeloylbisphosphonate (PiBP) cyclized to 2-hydroxy-2-phosphonocyclohexanecarbonylphosphonic acid, although at a rate about 30 times slower than AdBP. Study of the catalytic effect of amines on the cyclization of PiBP revealed a striking dependence on the pH, the chain length of the diamine, and the amine used. Thus the following relative efficacy was observed for the different amines at pH 5: Me2N(CH2)2NH2 (120), H2N(CH2)2NH2 (100), H2N(CH2)3NH2 (4), H2N(CH2)4NH2 (1), PhCH2NHCH2CH2NHCH2Ph (0.1), Me2NCH2CH2NMe2 (0.02), MeNH2 (0.02). These data show that depending on the chain length, 1,2-diamines are far more efficient catalysts than longer chain diamines and monoamines in these cyclizations. The mechanism which accommodates these results involves attack of a primary amine group on one of the keto groups, followed be removal of an alpha proton by the second amine group and the formation of an enamine. The latter then cyclizes by attacking the second keto group.  相似文献   

11.
Previous studies of different solvates of 2-methylpyridyllithium (2-picolyllithium) have uncovered electronic structures corresponding to aza-allyl and enamido resonance forms of the metallated pyridine-based compounds. Here, we report the synthesis and characterization of [2-CH2Li(THF)2C5H4N], a new THF solvate. X-ray crystallographic studies reveal a dimeric arrangement featuring a non-planar eight-membered [NCCLi]2 ring, in which the primary cation-anion interaction is between the central Li atom and the C atom of the deprotonated methyl group [length, 2.285(2) Å], suggesting a new carbanionic resonance structure for this 2-picolyllithium series. The significant carbanionic character of [2-CH2Li(THF)2C5H4N] was confirmed by gas-phase DFT calculations [B3LYP/6-311+G(d)] with the calculated electron density interrogated by means of quantum theory of atoms in molecules (QTAIM) and natural bond orbital (NBO) analyses. For comparison these computational analyses were also performed on the literature structures of [2-CH2Li(2-Picoline)C5H4N] and [2-CH2Li(PMDETA)C5H4N]. In a reactivity study, [2-CH2Li(THF)2C5H4N] was found to undergo nucleophilic addition to pyridine to generate dipyridylmethane in a good yield.  相似文献   

12.
The photochemical reaction of piperazine with C70 produces a mono‐adduct (N(CH2CH2)2NC70) in high yield (67 %) along with three bis‐adducts. These piperazine adducts can combine with various Lewis acids to form crystalline supramolecular aggregates suitable for X‐ray diffraction. The structure of the mono‐adduct was determined from examination of the adduct I2N(CH2CH2)2NI2C70 that was formed by reaction of N(CH2CH2)2NC70 with I2. Crystals of polymeric {Rh2(O2CCF3)4N(CH2CH2)2NC70}n?nC6H6 that formed from reaction of the mono‐adduct with Rh2(O2CCF3)4 contain a sinusoidal strand of alternating molecules of N(CH2CH2)2NC70 and Rh2(O2CCF3)4 connected through Rh?N bonds. Silver nitrate reacts with N(CH2CH2)2NC70 to form black crystals of {(Ag(NO3))4(N(CH2CH2)2NC70)4}n?7nCH2Cl2 that contain parallel, nearly linear chains of alternating (N(CH2CH2)2NC70 molecules and silver ions. Four of these {Ag(NO3)N(CH2CH2)2NC70}n chains adopt a structure that resembles a columnar micelle with the ionic silver nitrate portion in the center and the nearly non‐polar C70 cages encircling that core. Of the three bis‐adducts, one was definitively identified through crystallization in the presence of I2 as 12{N(CH2CH2)2N}2C70 with addends on opposite poles of the C70 cage and a structure with C2v symmetry. In 12{I2N(CH2CH2)2N}2C70, individual 12{I2N(CH2CH2)2N}2C70 units are further connected by secondary I2???N2 interactions to form chains that occur in layers within the crystal. Halogen bond formation between a Lewis base such as a tertiary amine and I2 is suggested as a method to produce ordered crystals with complex supramolecular structures from substances that are otherwise difficult to crystallize.  相似文献   

13.
The Stoichiometry of thermal decomposition was studied for the following compounds: Ni(NCS)2(2-Mepy)2 (I), (Me=methyl, py=pyridine), Ni(NCS)2(2-Etpy)2 (II) (Et=ethyl), Ni(NCS)2(2-Clpy)2 (III), Ni(NCS)2(2-Brpy)2 (IV), Ni(NCS)2(2-NH2py)2 (V), Ni((NCS)2(2-NH2py)2·3/4 (C2H5)2O (VI). The release of volatile ligands 2-Rpy is a one-step process for complexes I, II, III and IV, while for V and VI it is a two-step process, Ni(NCS)2(2-NH2py)1 (VII) being formed as an intermediate complex. It was found that complexes I and II are square-planar; the others exhibited pseudo-octahedral geometry. The differences in stereochemistry of the above complexes are explained by the different electronic properties of 2-Rpy.  相似文献   

14.
Treatment of bis(cyanamide) [M(N≡CNEt2)2L4](BPh4)2 and bis(cyanoguanidine) [M{N≡CN(H)C(NH2)=NH}2L4](BPh4)2 complexes [M = Fe, Ru, Os; L = P(OEt)3] with an excess of amine RNH2 (R = nPr, iPr) affords mixed‐ligand complexes with cyanamide and amine [M(NH2R)(N≡CNEt2)L4](BPh4)2 ( 1a – 5a ) and [M(NH2R){N≡CN(H)C(NH2)=NH}L4](BPh4)2 ( 1b , 2b ). The complexes were characterized by spectroscopy and X‐ray crystal structure determination of [M(NH2iPr)(N≡CNEt2){P(OEt)3}4](BPh4)2 [M = Ru ( 3a ), Os ( 5a )].  相似文献   

15.
Dimeric chlorobridge complex [Rh(CO)2Cl]2 reacts with two equivalents of a series of unsymmetrical phosphine–phosphine monoselenide ligands, Ph2P(CH2)nP(Se)Ph2 {n = 1( a ), 2( b ), 3( c ), 4( d )}to form chelate complex [Rh(CO)Cl(P∩Se)] ( 1a ) {P∩Se = η2‐(P,Se) coordinated} and non‐chelate complexes [Rh(CO)2Cl(P~Se)] ( 1b–d ) {P~Se = η1‐(P) coordinated}. The complexes 1 undergo oxidative addition reactions with different electrophiles such as CH3I, C2H5I, C6H5CH2Cl and I2 to produce Rh(III) complexes of the type [Rh(COR)ClX(P∩Se)] {where R = ? C2H5 ( 2a ), X = I; R = ? CH2C6H5 ( 3a ), X = Cl}, [Rh(CO)ClI2(P∩Se)] ( 4a ), [Rh(CO)(COCH3)ClI(P~Se)] ( 5b–d ), [Rh(CO)(COH5)ClI‐(P~Se)] ( 6b–d ), [Rh(CO)(COCH2C6H5)Cl2(P~Se)] ( 7b–d ) and [Rh(CO)ClI2(P~Se)] ( 8b–d ). The kinetic study of the oxidative addition (OA) reactions of the complexes 1 with CH3I and C2H5I reveals a single stage kinetics. The rate of OA of the complexes varies with the length of the ligand backbone and follows the order 1a > 1b > 1c > 1d . The CH3I reacts with the different complexes at a rate 10–100 times faster than the C2H5I. The catalytic activity of complexes 1b–d for carbonylation of methanol is evaluated and a higher turnover number (TON) is obtained compared with that of the well‐known commercial species [Rh(CO)2I2]?. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

16.
Salicylaldehyde thiosemicarbazone (H2saltsc) reacts with [M(PPh3)3X2] (M = Ru, Os; X = Cl, Br) to afford complexes of type [M(PPh3)2(Hsaltsc)2], in which the salicylaldehyde thiosemicarbazone ligand is coordinated to the metal as a bidentate N,S-donor forming a four-membered chelate ring. Reaction of benzaldehyde thiosemicarbazones (Hbztsc-R) with [M(PPh3)3X2] also affords complexes of similar type, viz. [M(PPh3)2(bztsc-R)2], in which the benzaldehyde thiosemicarbazones have also been found to coordinate the metal as a bidentate N,S-donor forming a four-membered chelate ring as before. Reaction of the Hbztsc-R ligands has also been carried out with [M(bpy)2X2] (M = Ru, Os; X = Cl, Br), which has afforded complexes of type [M(bpy)2(bztsc-R)]+, which have been isolated as perchlorate salts. Coordination mode of bztsc-R has been found to be the same as before. Structure of the Hbztsc-OMe ligand has been determined and some molecular modelling studies have been carried out determine the reason for the observed mode of coordination. Reaction of acetone thiosemicarbazone (Hactsc) has then been carried out with [M(bpy)2X2] to afford the [M(bpy)2(actsc)]ClO4 complexes, in which the actsc ligand coordinates the metal as a bidentate N,S-donorformingafive-membered chelate ring. Reaction of H2saltsc has been carried out with [Ru(bpy)2Cl2] to prepare the [Ru(bpy)2(Hsaltsc)]ClO4 complex, which has then been reacted with one equivalent of nickel perchlorate to afford an octanuclear complex of type [Ru(bpy)2(saltsc-H)4Ni4](ClO4)4.  相似文献   

17.
The reaction of a dichloromethane solution of a mixture of cis,trans-[PtCl2(SMe2)2] with a tetrahydrofuran solution of SnBr2 resulted in oxidation of platinum(II) with halogen exchange producing cis,trans-[PtBr4(SMe2)2]. Reaction of a mixture of cis,trans-[PtCl2(SEt2)2], potassium tetrachloroplatinate(II) or potassium hexachloroplatinate(IV) with SnBr2 in hydrochloric acid solution resulted in formation of predominantly anionic five-coordinate trichlorostannyl platinum(II) complexes. Reaction of potassium tetrabromoplatinate(II) with SnCl2 in hydrobromic acid in the presence of tetraphenylphosphonium bromide affords cis-[PPh4]2[PtBr2(SnBr3)2]. The insertion of SnCl2 into Pt–Cl bond of platinum(II) complexes cis-[PtCl2(L2)] {L2 = (PPh3)2; (PMe3)2; {P(OMe)3}2; dppm (bis(diphenylphosphino)methane); dppa (bis(diphenylphosphino)amine); and dppe (1,2-bis(diphenylphosphino)ethane)} is described.  相似文献   

18.
Tetranuclear manganese(II) phosphates [Mn(dipp)(bpy)]4?4 H2O ( 1 ) and [Mn4(dmpp)2(dmppH)4(bpy)4(H2O)2]?H2O ( 2 ) have been prepared from Mn(OAc)2?4 H2O and 2,6‐diisopropylphenyl phosphate (dippH2) or 2,6‐dimethylphenyl phosphate (dmppH2) in the presence of 2,2′‐bipyridine (bpy). In contrast, the reaction between [Mn(bpy)2(OAc)(ClO4)]?H2O and dippH2 affords [Mn(bpy)2(dippH)]2?2 ClO4?2 CH3OH ( 3 ). The reactions of Mn(OAc)2?4 H2O, dippH2, and pyridine (py) or 3,5‐dimethylpyrazole (dmpz) in CH3CN under reflux afford hexanuclear complexes [Mn6(dipp)6(py)8]?2CH3CN ( 4 ) and [Mn6(dipp)6(dmpz)6(AcOH)2]?2 H2O ( 5 ), respectively. Although compounds 1 and 2 are tetrameric, the former is a closed cubane‐like structure resembling the D4R secondary building unit of zeolites, whereas the latter exists in a staircase structure with fused Mn2O4P2 rings. The core structure of 3 contains a Mn2O4P2 eight‐membered ring that resembles the S4R building block of zeolites. Single‐crystal X‐ray diffraction studies reveal that compounds 4 and 5 have a similar core structure and differ from each other by the neutral ligands coordinated to manganese ions. All six phosphate ligands exist in a doubly deprotonated [(RO)PO32?] form and exhibit two types of binding modes [5.222] and [3.111]. An interesting feature of compounds 1 – 5 is that although they are oligonuclear complexes, there is an absence of oxido bridges. The magnetic properties of compounds 1 – 5 have been investigated in the temperature range 5–298 K, and it was found that all the compounds obey the Curie law.  相似文献   

19.
The three secondary phosphine oxides [CH2=CH(CH2)4]2HPO ( 1 ), [CH2=CH(CH2)5]2HPO ( 2 ), and [CH2=CH(CH2)6]2HPO ( 3 ), and two diphosphine dioxides, {[CH2=CH(CH2)6]2PO(CH2)7}2 ( 4 ) and {[CH2=CH(CH2)6]2PO(CH2)4}2 ( 5 ), incorporating long methylene chains, are described. The single crystal X‐ray structures of 1 , 2 , and 5 have been determined. The phosphine oxides 3 , 4 , and 5 have been adsorbed on silica in submonolayer quantities to give 3 a – 5 a . The 1H, 13C, and 31P solid‐state NMR spectra of polycrystalline 3 – 5 have been analyzed and compared with those of 3 a – 5 a . The changes of the solid‐state NMR characteristics upon adsorption and the surface mobilities of the phosphine oxides are discussed.  相似文献   

20.
Gas phase IR spectra have been investigated in a series of methylsilyl compounds, including MeSiHDX (X = F, Cl), MeSiHX2 (X = F, Cl), Me2SiHCl, (Me2SiH)2NH, (MeSiH2)3N, (MeSiHD)3N, (MeSiH2)2NMe,* (Me2SiH)2NMe,* MeSiH2NMe2,* Me2SiHNMe2, (MeSiHD)2O and (Me2SiH)2O. In the asterisked molecules, pairs of νSiH bands are seen which confirm the existence of fixed conformations. Amongst the other compounds, the single band seen is attributed on the basis of the α-Me substitutent effect to a fixed conformation in the cases of Me2SiHNMe2 and (Me2SiH)2O. In (Me2SiH)2NH and (MeSiH2)2O, free internal rotation about the SiN and SiO bonds seems likely. Apart from the possible case of (MeSiH2)3N, the spectra appear to be compatible with the electron diffraction evidence, and for the most part in broad agreement with the structures found. 2νSiH bands are in general harder to see in these amines and ethers than in Me2SiHCl. A likely transition to a (1, 1) local mode state is observed in (MeSiH2)3N.Determination of barriers to internal rotation in these molecules should be a high priority.  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