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 46 毫秒
1.
The reactions of three different tetracoordinated Ir complexes, [Ir(troppph)2]n (n=+1, 0, −1), which differ in the formal oxidation state of the metal from +1 to −1, with proton sources and dihydrogen were investigated (tropp=5‐(diphenylphosphanyl)dibenzo[a,d]cycloheptene). It was found that the cationic 16‐electron complex [Ir(troppph)2]+ ( 2 ) cannot be protonated but reacts with NaBH4 to the very stable 18‐electron IrI hydride [IrH(troppph)2] ( 5 ), which is further protonated with medium strong acids to give the 18‐electron IrIII dihydride [IrH2(troppph)2]+ ( 6 ; pKs in CH2Cl2/THF/H2O 1 : 1 : 2 ca. 2.2). Both, the neutral 17‐electron Ir0 complex [Ir(troppph)2] ( 3 ) and the anionic 18‐electron complex [Ir(troppph)2] ( 4 ) react rapidly with H2O to give the monohydride 5 . In reactions of 3 with H2O, the terminal IrI hydroxide [Ir(OH)(troppph)2] ( 8 ) is formed in equal amounts. All these complexes, apart from 5 , which is inert, do react rapidly with dihydrogen. The complex 2 gives the dihydride 6 in an oxidative addition reaction, while 3 , 4 , and 8 give the monohydride 5 . Interestingly, a salt‐type hydride (i.e., LiH) is formed as further product in the unexpected reaction with [Li(thf)x]+[Ir(troppph)2] ( 4 ). Because 3 undergoes disproportionation into 2 and 4 according to 2 3 ⇄ 2 + 4 (Kdisp=2.7⋅10−5), it is likely that actually the diamagnetic species and not the odd‐electron complex 3 is involved in the reactions studied here, and possible mechanisms for these are discussed.  相似文献   

2.
The two dinuclear IrI complexes [Ir2(μ‐Cl)2 {(R)‐(S)‐PPF‐PPh2}2] ( 1 ; (R)‐(S)‐PPF‐PPh2=(S)‐1‐(diphenylphosphino)‐2‐[(R)‐1‐(diphenylphosphino)ethyl]ferrocene and [Ir2(μ‐Cl)2{(R)‐binap}2] ( 3 ; (R)‐binap=(R)‐[1,1′‐binaphthalene]‐2,2′‐diylbis[diphenylphosphine]) smoothly react with 4 equiv. of the lithium salt of aniline to afford the new bis(anilido)iridate(I) (=bis(benzenaminato)iridate(1‐)) complexes Li[Ir(NHPh)2{(R)‐(S)‐PPF‐PPh2}] ( 4 ) and Li[Ir(NHPh)2{(R)‐binap}] ( 5 ), respectively. The anionic complexes 4 and 5 react upon protonolysis to give the dinuclear aminato‐bridged derivatives [Ir2(μ‐NHPh)2{(R)‐(S)‐PPF‐PPh2}2] ( 6 ) and [Ir2(μ‐NHPh)2{(R)‐binap}2] ( 7 ), which were characterized by X‐ray crystallography. None of the new complexes 4 – 7 shows catalytic activity in the hydroamination of olefins.  相似文献   

3.
Reduction of a variety of extremely bulky amido Group 12 metal halide complexes, [LMX(THF)0,1] (L=amide; M=Zn, Cd, or Hg; X=halide) with a magnesium(I) dimer gave a homologous series of two‐coordinate metal(I) dimers, [L′MML′] (L′=N(Ar?)(SiMe3), Ar?=C6H2{C(H)Ph2}2Pri‐2,6,4); and the formally zinc(0) complex, [L*ZnMg(MesNacnac)] (L*=N(Ar*)(SiPri3); Ar*=C6H2{C(H)Ph2}2Me‐2,6,4; MesNacnac=[(MesNCMe)2CH]?, Mes=mesityl), which contains the first unsupported Zn? Mg bond. Two equivalents of [L*ZnMg(MesNacnac)] react with ZnBr2 or ZnBr2(tmeda) to give the mixed valence, two‐coordinate, linear tri‐zinc complex, [L*ZnIZn0ZnIL*], and the first zinc(I) halide complex, [L*ZnZnBr(tmeda)], respectively. The analogues [L*ZnMZnL*] (M=Cd or Hg), were also prepared, the Cd species contains the first Zn? Cd bond in a molecular compound. Metal–metal bonding was studied by DFT calculations.  相似文献   

4.
The fluxionality of [Ir4(CO)82-CO)3L] (L = Br?, I?, SCN?, NO2?, P(4-ClC6H4)3, PPh3, P(4-MeOC6H4)3, P(4-Me2NC6H4)3), as studied by 2D-13C-NMR in solution, is due to two successive scrambling processes: the merry-go-round of six basal CO's and CO bridging to alternative faces of the Ir4 tetrahedron. The basicity of the ligand L has no significant effect on the activation parameters. The scrambling process of lowest activation energy in [Ir4(CO)72-CO)3(PMePh2)2] correspond to the two possible synchronous CO bridging about a unique face of the metal tetrahedron swapping the relative axial and radial positions of the ligands L. The disubstituted clusters [Ir4(CO)102-L? L)] with one edge-bridging ligand have a ground-state geometry with three edge-bridging CO's (L? L = bis(diphenylphosphino)methane, bis(diphenylarsino)methane, bis(diphenylphosphino)propane) or with all terminal CO's (L? L = CH3SCH2SCH3). In all cases, the fluxional process of lowest activation energy in the merry-go-round of six CO's about a unique triangular face. For the P and As donor ligands, this process is followed by the rotation of terminal CO's bonded to two Ir-atoms residing on the mirror plane of the unbridged intermediate.  相似文献   

5.
Here, we report an iridium(III) coordination system with 2‐aminoethanethiolate (aet), which shows the formation of S?H???S hydrogen and S?S disulfide bonds in a controlled manner. Treatment of fac‐[Ir(aet)3] with aqueous HBF4 under aerobic conditions gave dinuclear [Ir2(aet)4(cysta)]2+ ([ 1 ]2+; cysta=cystamine) with a single S?S disulfide bond, while dimeric [Ir2(aet)3(Haet)3](BF4)3 ([ 2 ](BF4)3) with a triple S?H???S hydrogen bond was formed by similar treatment under anaerobic conditions. Upon exposure to air, [ 2 ]3+ was converted to dinuclear [Ir2(aet)2(Haet)2(cysta)]4+ ([ 3 ]4+), in which two IrIII centers are spanned by a double S?H???S hydrogen bond and a single S?S disulfide bond. Complex [ 3 ]4+ was interconvertible with [ 1 ]2+ via the removal/addition of protons on S donors, accompanied by the intermolecular exchange of the fac‐[Ir(aet)3] units. Complexes [ 1 ]2+, [ 2 ]3+, and [ 3 ]4+, isolated as BF4? salts, were fully characterized by single‐crystal X‐ray crystallography.  相似文献   

6.
We have prepared and fully characterized two isomers of [IrIV(dpyp)2] (dpyp=meso‐2,4‐di(2‐pyridinyl)‐2,4‐pentanediolate). These complexes can cleanly oxidize to [IrV(dpyp)2]+, which to our knowledge represent the first mononuclear coordination complexes of IrV in an N,O‐donor environment. One isomer has been fully characterized in the IrV state, including by X‐ray crystallography, XPS, and DFT calculations, all of which confirm metal‐centered oxidation. The unprecedented stability of these IrV complexes is ascribed to the exceptional donor strength of the ligands, their resistance to oxidative degradation, and the presence of four highly donor alkoxide groups in a plane, which breaks the degeneracy of the d‐orbitals and favors oxidation.  相似文献   

7.
The synergistic Ag+/X2 system (X=Cl, Br, I) is a very strong, but ill‐defined oxidant—more powerful than X2 or Ag+ alone. Intermediates for its action may include [Agm(X2)n]m+ complexes. Here, we report on an unexpectedly variable coordination chemistry of diiodine towards this direction: ( A )Ag‐I2‐Ag( A ), [Ag2(I2)4]2+( A ?)2 and [Ag2(I2)6]2+( A ?)2?(I2)x≈0.65 form by reaction of Ag( A ) ( A =Al(ORF)4; RF=C(CF3)3) with diiodine (single crystal/powder XRD, Raman spectra and quantum‐mechanical calculations). The molecular ( A )Ag‐I2‐Ag( A ) is ideally set up to act as a 2 e? oxidant with stoichiometric formation of 2 AgI and 2 A ?. Preliminary reactivity tests proved this ( A )Ag‐I2‐Ag( A ) starting material to oxidize n‐C5H12, C3H8, CH2Cl2, P4 or S8 at room temperature. A rough estimate of its electron affinity places it amongst very strong oxidizers like MF6 (M=4d metals). This suggests that ( A )Ag‐I2‐Ag( A ) will serve as an easily in bulk accessible, well‐defined, and very potent oxidant with multiple applications.  相似文献   

8.
The synthesis and structures of the two CuI halide complexes [Cu5(dppm)(dppm?)2(OtBu)Cl2] and [Cu3(dppm)3Br2][CuBr2] (dppm = Ph2PCH2PPh2, dppm? = [Ph2PCHPPh2]?) are reported. The compounds were obtained by treating reaction mixtures of [CuOtBu] and dppm with dichloromethane or dibromomethane.  相似文献   

9.
Several new donor–acceptor adducts of niobium and tantalum pentaazide with N‐donor ligands have been prepared from the pentafluorides by fluoride–azide exchange with Me3SiN3 in the presence of the corresponding donor ligand. With 2,2′‐bipyridine and 1,10‐phenanthroline, the self‐ionization products [MF4(2,2′‐bipy)2]+[M(N3)6]?, [M(N3)4(2,2′‐bipy)2]+[M(N3)6]? and [M(N3)4(1,10‐phen)2]+[M(N3)6]? were obtained. With the donor ligands 3,3′‐bipyridine and 4,4′‐bipyridine the neutral pentaazide adducts (M(N3)5)2?L (M=Nb, Ta; L=3,3′‐bipy, 4,4′‐bipy) were formed.  相似文献   

10.
Complexes [NiI3(mpta)2]I ( 1 ) and [NiI3(ppta)2]I ( 2 ) have been synthesized by reaction of nickel(II) halide salts with ‐1‐methyl‐1‐azonia‐3,5‐diaza‐7‐phosphatricyclo[3.3.1.13,7]decane iodide (mpta+I?) and 1‐(n‐propyl)‐1‐azonia‐3,5‐diaza‐7‐phosphatricyclo[3.3.1.13,7]decane bromide (ppta+Br?) respectively. The crystal structures of compounds 1 and 2 are described and are similar, with both compounds crystallizing in monoclinic space groups. The geometry about both nickel atoms is that of a trigonal bipyramid with the cationic phosphine ligands found in the axial positions and the iodide ligands arranged in the equatorial plane.  相似文献   

11.
Reaction of the iridium tetracarbonylate [PPN][Ir(CO)4] (1a) with triphenylcyclopropenyl tetrafluoroborate [C3Ph3][BF4] afforded two dinuclear species Ir2(CO)4(μ,η12-C3Ph3)(μ,η23-C3Ph3) (2) and Ir2(CO)4(μ,η44-C6Ph6) (3a) resulting from the ring opening and in the latter case, coupling of the resulting acyclic, propenyl ligands. The analogous reaction with [PPN][Rh(CO)4] (1b) afforded only the rhodium analogue for 3a.  相似文献   

12.
The water‐soluble phosphine ligands, 1,3,5‐triaza‐7‐phosphatricyclo[3.3.1.13,7]decane (tpa) and 1‐alkyl‐1‐azonia‐3,5‐diaza‐7‐phosphatricyclo[3.3.1.13,7]decane iodides (Rtpa+I), with alkyl=methyl(mtpa+I), ethyl (etpa+I) and n‐propyl, (ptpa+I), and mtpa+Cl react with [Rh2Cl2(CO)4] giving the rhodium(I) complexes [RhCl(CO)(tpa)2], [RhI(CO)(Rtpa+I)2], [RhCl‐­(CO)(mtpa+Cl)3] and [RhI(CO)(Rtpa+I)3]. The properties and reactivities of the complexes have been investigated using 1H and 31PNMR and IR spectroscopies. The five‐coordinate complexes in solutions show dynamic properties. The complexes are catalysts of the water‐gas shift reaction, the hydrogenation of CC and CO bonds, the hydroformylation of alkenes and the isomerization of unsaturated compounds. Copyright © 1999 John Wiley & Sons, Ltd.  相似文献   

13.
The influence of the potentially chelating imino group of imine‐functionalized Ir and Rh imidazole complexes on the formation of functionalized protic N‐heterocyclic carbene (pNHC) complexes by tautomerization/metallotropism sequences was investigated. Chloride abstraction in [Ir(cod)Cl{C3H3N2(DippN=CMe)‐κN3}] ( 1 a ) (cod=1,5‐cyclooctadiene, Dipp=2,6‐diisopropylphenyl) with TlPF6 gave [Ir(cod){C3H3N2(DippN=CMe)‐κ2(C2,Nimine)}]+[PF6]? ( 3 a +[PF6]?). Plausible mechanisms for the tautomerization of complex 1 a to 3 a +[PF6]? involving C2?H bond activation either in 1 a or in [Ir(cod){C3H3N2(DippN=CMe)‐κN3}2]+[PF6]? ( 6 a +[PF6]?) were postulated. Addition of PR3 to complex 3 a +[PF6]? afforded the eighteen‐valence‐electron complexes [Ir(cod)(PR3){C3H3N2(DippN=CMe)‐κ2(C2,Nimine)}]+[PF6]? ( 7 a +[PF6]? (R=Ph) and 7 b +[PF6]? (R=Me)). In contrast to Ir, chloride abstraction from [Rh(cod)Cl{C3H3N2(DippN=CMe)‐κN3}] ( 1 b ) at room temperature afforded [Rh(cod){C3H3N2(DippN=CMe)‐κN3}2]+[PF6]? ( 6 b +[PF6]?) and [Rh(cod){C3H3N2(DippN=CMe)‐κ2(C2,Nimine)}]+[PF6]? ( 3 b +[PF6]?) (minor); the reaction yielded exclusively the latter product in toluene at 110 °C. Double metallation of the azole ring (at both the C2 and the N3 atom) was also achieved: [Ir2(cod)2Cl{μ‐C3H2N2(DippN=CMe)‐κ2(C2,Nimine),κN3}] ( 10 ) and the heterodinuclear complex [IrRh(cod)2Cl{μ‐C3H2N2(DippN=CMe)‐κ2(C2,Nimine),κN3}] ( 12 ) were fully characterized. The structures of complexes 1 b , 3 b +[PF6]?, 6 a +[PF6]?, 7 a +[PF6]?, [Ir(cod){C3HN2(DippN=CMe)(DippN=CH)(Me)‐κ2(N3,Nimine)}]+[PF6]? ( 9 +[PF6]?), 10? Et2O ? toluene, [Ir2(CO)4Cl{μ‐C3H2N2(DippN=CMe)‐κ2(C2,Nimine),κN3}] ( 11 ), and 12? 2 THF were determined by X‐ray diffraction.  相似文献   

14.
The synthesis and 119Sn NMR characteristics of new five-coordinate tris(trichlorostannato) complexes of RhI, IrI and PtII are reported. The RhI and IrI complexes are complex dianions of the form (PPN)2[M(SnCl3)3L2] where L can be CO, CN (cyclohexyl) or L2, a diolefin such as 1,5-COD or NBD (norbornadiene). The anionic platinum complexes (PPN)[Pt(SnCl3)3L2] contain similar L ligands. A number of neutral monotrichlorostannato complexes of type [M(SnCl3)L4] including [Ir(SnCl3)(NBD)(1,5-COD)] have been prepared and characterized. Their δ(119Sn), δ(13C), δ(195Pt) as well as 1J(103Rh, 119Sn), 1J(195Pt, 119Sn), 2J(119Sn, 117Sn) and 2J(119Sn, 13C) data are given. A trans influence series, based on 1J(195Pt, 119Sn), reveals the following sequence: H? > PR3 > AsR3 > SnCl3? > olefin > Cl?.  相似文献   

15.
Trifluoromethylation of AuCl3 by using the Me3SiCF3/CsF system in THF and in the presence of [PPh4]Br proceeds with partial reduction, yielding a mixture of [PPh4][AuI(CF3)2] ( 1′ ) and [PPh4][AuIII(CF3)4] ( 2′ ) that can be adequately separated. An efficient method for the high‐yield synthesis of 1′ is also described. The molecular geometries of the homoleptic anions [AuI(CF3)2]? and [AuIII(CF3)4]? in their salts 1′ and [NBu4][AuIII(CF3)4] ( 2 ) have been established by X‐ray diffraction methods. Compound 1′ oxidatively adds halogens, X2, furnishing [PPh4][AuIII(CF3)2X2] (X=Cl ( 3 ), Br ( 4 ), I ( 5 )), which are assigned a trans stereochemistry. Attempts to activate C? F bonds in the gold(III) derivative 2′ by reaction with Lewis acids under different conditions either failed or only gave complex mixtures. On the other hand, treatment of the gold(I) derivative 1′ with BF3?OEt2 under mild conditions cleanly afforded the carbonyl derivative [AuI(CF3)(CO)] ( 6 ), which can be isolated as an extremely moisture‐sensitive light yellow crystalline solid. In the solid state, each linear F3C‐Au‐CO molecule weakly interacts with three symmetry‐related neighbors yielding an extended 3D network of aurophilic interactions (Au???Au=345.9(1) pm). The high $\tilde \nu $ CO value (2194 cm?1 in the solid state and 2180 cm?1 in CH2Cl2 solution) denotes that CO is acting as a mainly σ‐donor ligand and confirms the role of the CF3 group as an electron‐withdrawing ligand in organometallic chemistry. Compound 6 can be considered as a convenient synthon of the “AuI(CF3)” fragment, as it reacts with a number of neutral ligands L, giving rise to the corresponding [AuI(CF3)(L)] compounds (L=CNtBu ( 7 ), NCMe ( 8 ), py ( 9 ), tht ( 10 )).  相似文献   

16.
Exceptional water oxidation (WO) turnover frequencies (TOF=17 000 h?1), and turnover numbers (TONs) close to 400 000, the largest ever reported for a metal‐catalyzed WO reaction, have been found by using [Cp*IrIII(NHC)Cl2] (in which NHC=3‐methyl‐1‐(1‐phenylethyl)‐imidazoline‐2‐ylidene) as the pre‐catalyst and NaIO4 as oxidant in water at 40 °C. The apparent TOF for [Cp*IrIII(NHC)X2] ( 1 X , in which X stands for I ( 1 I ), Cl ( 1 Cl ), or triflate anion ( 1 OTf )) and [(Cp*‐NHCMe)IrIIII2] ( 2 ) complexes, is kept constant during almost all of the O2 evolution reaction when using NaIO4 as oxidant. The TOF was found to be dependent on the ligand and on the anion (TOF ranging from ≈600 to ≈1100 h?1 at 25 °C). Degradation of the complexes by oxidation of the organic ligands upon reaction with NaIO4 has been investigated. 1H NMR, ESI‐MS, and dynamic light‐scattering measurements (DLS) of the reaction medium indicated that the complex undergoes rapid degradation, even at low equivalents of oxidant, but this process takes place without formation of nanoparticles. Remarkably, three‐month‐old solution samples of oxidized pre‐catalysts remain equally as active as freshly prepared solutions. A UV/Vis feature band at λmax=405 nm is observed in catalytic reaction solutions only when O2 evolves, which may be attributed to a resting state iridium speciation, most probably Ir–oxo species with an oxidation state higher than IV.  相似文献   

17.
Reactions of bis(phosphinimino)amines LH and L′H with Me2S ? BH2Cl afforded chloroborane complexes LBHCl ( 1 ) and L′BHCl ( 2 ), and the reaction of L′H with BH3 ? Me2S gave a dihydridoborane complex L′BH2 ( 3 ) (LH=[{(2,4,6‐Me3C6H2N)P(Ph2)}2N]H and L′H=[{(2,6‐iPr2C6H3N)P(Ph2)}2N]H). Furthermore, abstraction of a hydride ion from L′BH2 ( 3 ) and LBH2 ( 4 ) mediated by Lewis acid B(C6F5)3 or the weakly coordinating ion pair [Ph3C][B(C6F5)4] smoothly yielded a series of borenium hydride cations: [L′BH]+[HB(C6F5)3]? ( 5 ), [L′BH]+[B(C6F5)4]? ( 6 ), [LBH]+[HB(C6F5)3]? ( 7 ), and [LBH]+[B(C6F5)4]? ( 8 ). Synthesis of a chloroborenium species [LBCl]+[BCl4]? ( 9 ) without involvement of a weakly coordinating anion was also demonstrated from a reaction of LBH2 ( 4 ) with three equivalents of BCl3. It is clear from this study that the sterically bulky strong donor bis(phosphinimino)amide ligand plays a crucial role in facilitating the synthesis and stabilization of these three‐coordinated cationic species of boron. Therefore, the present synthetic approach is not dependent on the requirement of weakly coordinating anions; even simple BCl4? can act as a counteranion with borenium cations. The high Lewis acidity of the boron atom in complex 8 enables the formation of an adduct with 4‐dimethylaminopyridine (DMAP), [LBH ? (DMAP)]+[B(C6F5)4]? ( 10 ). The solid‐state structures of complexes 1 , 5 , and 9 were investigated by means of single‐crystal X‐ray structural analysis.  相似文献   

18.
Preparation, Spectroscopic Characterization and Normal Coordinate Analysis of Nonabromoditechnetate(IV), [Tc2Br9]? . On heating the tetraethylammonium salt (TEA)2[Tc2Br6] with trifluoroacetic acid the face-sharing bioctahedral anion [Tc2Br9]? is formed, which vibronic spectrum is assigned according to point group D3h. Normal coordinate analysis, based on a general valence force field has been performed, resulting in a good agreement of calculated frequencies with the observed IR and Raman bands. Due to stronger bonding of the terminal as compared to the bridging ligands, the valence force constant fd(TcBrt) = 1.045 is significantly higher than fd(TcBrb) = 0.80 mdyn/Å.  相似文献   

19.
We describe the synthesis of base‐free bisborole [Cym?(BC4Ph4)2]—Cym?=(OC)3Mn(η5‐C5H3)—and its transformation into two fully characterized Lewis acid–base adducts with pyridine bases of the type 4‐R? NC5H4 (R=tBu, NMe2). The results of electrochemical, as well as NMR and UV/Vis spectroscopic studies on [Cym?(BC4Ph4)2] and the related monoborole derivative [Cym(BC4Ph4)]—Cym=(OC)3Mn(η5‐C5H4)—provided conclusive evidence for 1) the enhanced Lewis acidity of the two boron centers that result from conjugation of two borole fragments, and 2) the fact that Mn? B bonding interactions between the Lewis acidic borole moieties and the Mn center are considerably less pronounced for bisborole [Cym?(BC4Ph4)2]. In addition, the reduction chemistry of [Cym?(BC4Ph4)2] has been studied in detail, both electrochemically and chemically. Accordingly, chemical reduction of [Cym?(BC4Ph4)2] with magnesium anthracene afforded the corresponding tetraanion, which features a rare Mg? OC bonding mode in the solid state.  相似文献   

20.
[ReNCl2(PPh3)2] and [ReNCl2(PMe2Ph)3] react with the N‐heterocyclic carbene (NHC) 1,3,4‐triphenyl‐1,2,4‐triazol‐5‐ylidene (HLPh) under formation of the stable rhenium(V) nitrido complex [ReNCl(HLPh)(LPh)], which contains one of the two NHC ligands with an additional orthometallation. The rhenium atom in the product is five‐coordinate with a distorted square‐pyramidal coordination sphere. The position trans to the nitrido ligand is blocked by one phenyl ring of the monodentate HLPh ligand. The Re–C(carbene) bond lengths of 2.072(6) and 2.074(6) Å are comparably long and indicate mainly σ‐bonding between the NHC ligand and the electron deficient d2 metal atom. The chloro ligand in [ReNCl(HLPh)(LPh)] is labile and can be replaced by ligands such as pseudohalides or monoanionic thiolates such as diphenyldithiophosphinate (Ph2PS2?) or pyridine‐2‐thiolate (pyS?). X‐ray structure analyses of [ReN(CN)(HLPh)(LPh)] and [ReN(pyS)(HLPh)(LPh)] show that the bonding situation of the NHC ligands (Re–C(carbene) distances between 2.086(3) and 2.130(3) Å) in the product is not significantly influenced by the ligand exchange. The potentially bidentate pyS? ligand is solely coordinated via its thiolato functionality. Hydrogen atoms of each one of the phenyl rings come close to the unoccupied sixth coordination positions of the rhenium atoms in the solid state structures of all complexes. Re–H distances between 2.620 and 2.712Å do not allow to discuss bonding, but with respect to the strong trans labilising influence of “N3?”, weak interactions are indicated.  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