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 218 毫秒
1.
Abstract

Reaction of two equivalents of N-mono- or di-substituted 3-amino-4-(n-butoxy)-3-cyclobutene-1,2-diones with a 1,2-diaminoethane gave N-mono- or di-substituted 1,2-bis((2-amino-1-cyclobutene-3,4-dione)amino)-ethane derivatives (bis(squaramides)). Reaction of the bis(squaramides) with excess P4S10 gave the analogous tetrathio derivatives (bis(dithiosquaramides), LH2) of formula (NR1R2)C4S2(NHCH2CH2NH)-C4S2(NR1R2) (R1=n-Bu, R2=H; R1=R2=Et, n-Bu). The new bis(dithiosquaramide) ligands were characterized by elemental analysis, IR, 1H NMR, 13C NMR, electronic, and mass spectroscopic methods. The complexes of these ligands with nickel(II) were prepared, isolated and characterized. The isolated complexes are neutral 2:2 species of formula Ni2L2, as evidenced by results from mass spectrometry, and they exhibit thermochromic behaviour in pyridine solution. Additional spectroscopic data (IR, NMR) are consistent with the ligands being coordinated only through sulfur donor atoms and a structure for the complexes is proposed.  相似文献   

2.
Abstract

A thermotropic ionic lamellar phase from non-stoichiometric pyridinium octyl-phosphates has been investigated by multinuclear N.M.R. and X-ray diffraction. At room temperature and above, this phase is formed for pyridine to octylphosphoric acid molar ratios from 0.2 to 0.8.2H and 13C relaxation experiments show that the pyridinium ion undergoes a very anisotropic motion with Dzz > Dxx ? Dyy, z and x being the perpendicular direction to the ring and the c 2 symmetry axis, respectively. The order parameters given by the 2H quadrupolar splittings and the 13C chemical shift anisotropy (CSA) are Szz = 0.13, Syy = -0.08 and Sxx = -0.05, showing that the pyridinium ring is preferentially oriented parallel to the lamellar plane. The 31P CSA and the C1-P dipolar splitting yield Szz = 0.33 and Sxx ? Syy for the octylphosphate anion. The order parameters of alkyl C-H bonds have been obtained from the J resolved two-dimensional 13C N.M.R. spectra of oriented samples. Two limiting conformational models have been considered to calculate the S CH. One of them is reasonably consistent with the structure derived from X-ray experiments and has been used to calculate the dipolar 31P relaxation. Taking into account the CSA contribution, the relaxation measurements performed at 36, 121 and 202 MHz show that the octylphosphate anion undergoes a quasi-axial reorientation about the long molecular axis x with D∥/D⊥ = 4 and D⊥ ? 107 rad/s at 300 K.  相似文献   

3.
Reaction of bicyclic β‐P4S3I2 with enantiomerically pure (R)‐Hpthiq (1‐phenyl‐1,2,3,4‐tetrahydroisoquinoline) and Et3N gave a solution of a single diastereomer of the unusually stable diamide β‐P4S3(pthiq)2, accounting for 83 % of the phosphorus content. Despite the steric bulk of the substituents, each amide group of this could adopt either of two rotameric positions about their P–N bonds, so that, at 183 K, 31P NMR multiplets for four rotamers could be observed and the spectra of three of them analysed fully. The large 2J(P–P–P) coupling became greater (253, 292, 304 Hz) with decreasing abundance of the individual rotamers. The rotamers were modelled at the ab initio RHF/3–21G* level, and relative NMR chemical shifts predicted by the GIAO method using a locally dense basis set, allowing the observed spectra to be assigned to structures. Calculations at the same level for the model compound α‐P4S3(pthiq)Cl confirmed the assignments of low‐temperature rotamers of α‐P4S3(pthiq)I reported previously. Changes in observed P–P coupling constants and 31P chemical shifts, on rotating a pthiq substituent, could then be compared between β‐P4S3(pthiq)2 and α‐P4S3(pthiq)I, confirming both sets of assignments. The most abundant rotamer of β‐P4S3(pthiq)2 was not the one with the least sterically crowded sides of both pthiq substituents pointing towards the P4S3 cage, because of interaction between the two substituents. Only by using a DFT method could relative abundances of rotamers of β‐P4S3(pthiq)2 be predicted to be in the observed order. Use of racemic Hpthiq gave also the two diastereomers of β‐P4S3(pthiq)2 with Cs symmetry, for which the room temperature 31P{1H} NMR spectra were analysed fully.  相似文献   

4.
Contributions to the Chemistry of Phosphorus. 212. Tetraisopropyldodecaphosphane(4), P12i-Pr4 – Preparation, Properties, and Molecular Dynamics According to an earlier crystal structure analysis, tetraisopropyldodecaphosphane(4) ( 1 ) exhibits the symmetry C2, and the substituents are arranged in all-trans position [3]. We have now found by NMR spectroscopic studies that in solution a second configurational isomer of the symmetry CS ( 1b ) exists in addition to the molecule present in the crystal ( 1a ). The transformation of 1a into 1b , which can only occur through a quasi synchronous inversion at the atoms P3 and P4 or P9 and P10, takes place at a noticeable rate already below room temperature.  相似文献   

5.
A New Phosphorus Sulfide with Adamantane Structure: δ‐P4S7 By sulfur abstraction from α‐P4S9/P4S10 with triphenylphosphine a new phosphorus sulfide δ‐P4S7 with adamantane skeleton and an additional sulfur in exo‐position was identified in CS2‐solution by 31P‐NMR spectroscopy. Product distribution and 31P‐NMR parameter are given.  相似文献   

6.
In an earlier publication (J. Am. Chem. Soc. 2002 , 124, 7111) we showed that polymeric cationic [Ag(P4S3)n]+ complexes (n=1, 2) are accessible if partnered with a suitable weakly coordinating counterion of the type [Al(ORF)4]? (ORF: poly‐ or perfluorinated alkoxide). The present work addresses the following questions that could not be answered in the initial report: How many P4S3 cages can be bound to a Ag+ ion? Why are these complexes completely dynamic in solution in the 31P NMR experiments? Can these dynamics be frozen out in a low‐temperature 31P MAS NMR experiment? What are the principal binding sites of the P4S3 cage towards the Ag+ ion? What are likely other isomers on the [Ag(P4S3)n]+ potential energy surface? Counterion influence: Reactions of P4S3 with Ag[Al{OC(CH3)(CF3)2}4] (Ag[hftb]) and Ag[{(CF3)3CO}3Al‐F‐Al{OC(CF3)3)}3] (Ag[al‐f‐al]) gave [(P4S3)Ag[hftb]] ( 7 ) as a molecular species, whereas [Ag2(P4S3)6]2+[al‐f‐al]?2 ( 8 ) is an isolated 2:1 salt. We suggest that a maximum of three P4S3 cages may be bound on average to an Ag+ ion. Only isolated dimeric dications are formed with the largest cation, but polymeric species are obtained with all other smaller aluminates. Thermodynamic Born–Haber cycles, DFT calculations, as well as solution NMR and ESI mass spectrometry indicate that 8 exhibits an equilibrium between the dication [Ag2(P4S3)6]2+ (in the solid state) and two [Ag(P4S3)3]+ monocations (in the gas phase and in solution). Dynamics: 31P MAS NMR spectroscopy showed these solid adducts to be highly dynamic, to an extent that the 2JP,P coupling within the cages could be resolved (J‐res experiment). This is supported by DFT calculations, which show that the extended PES of [Ag(P4S3)n]+ (n=1–3) and [Ag2(P4S3)2]+ is very flat. The structures of α‐ and γ‐P4S3 were redetermined. Their variable‐temperature 31P MAS NMR spectra are discussed jointly with those of all four currently known [Ag(P4S3)n]+ adducts with n=1, 2, and 3.  相似文献   

7.
Abstract

The dissociative ionization by electron impact (70 eV) of tetraphosphorus trisulfide, P4S3, yields [P3S]+ and [PZS]+0 ions whose structures have been investigated by means of Coliisionai Activation (CA) mass spectra. Using the technique of Neutralization-Reionization mass spectrometry (NRMS) it is shown that both ions can be reduced to the corresponding neutral molecules. Thus, triphosphorus sulfide and diphosphorus sulfide are viable molecules in the rarefied gas phase. The results of ab initio MO calculations were used to interpret the experimental findings. Tetrahedral (C3v) and triangular (C2v) structures are proposed for the [P,S]+ and [PZS]+0 ions respectively.  相似文献   

8.
The crystal structures of six members of the homologous series with general formula [BiQX]2[AgxBi1?xQ2?2xX2x?1]N+1 (Q = S, Se; X = Cl, Br; 1/2 ≤ x ≤ 1) and N = 4, 5, or 7 were determined by single‐crystal X‐ray diffraction. The series are characterized by the parameters N and x and are denoted (N, x)P. Ag3Bi4S6Cl3 (x = 0.60) (I) , Ag3.5Bi3.5S5Br4 (x = 0.70) (II) and Ag3.65Bi3.35Se4.70Br4.30 (x = 0.73) (III) belong to (4, x)P series Ag5xBi7?5xQ12?10xX10x?3 and adopt the AgBi6S9 structure type. The (5, x)P compound Ag3.66Bi4.34S6.68Br3.32 (IV) , which corresponds to x = 0.61 in Ag6xBi8?6xS14?12xBr12x?4, crystallizes isostructurally to AgBi3S5. The compounds Ag4.56Bi5.44Se8.88Br3.12 (x = 0.57) (V) and Ag5.14Bi4.86S7.76Br4.24 (x = 0.64) (VI) , which are members of (7, x)P series Ag8xBi10?8xQ18?16xBr16x?6, adopt the Ag3Bi7S12 structure type. In the monoclinic crystal structures (space group C2/m) two kinds of layered modules alternate along [001]. Modules of type A uniformly consist of paired rods of face‐sharing monocapped trigonal prisms around Bi atoms with octahedra around mixed occupied metal positions (M = Ag/Bi) between them. Modules of type B are composed of [MZ6] octahedra, which are arranged in NaCl‐type fragments of thickness N. All structures exhibit Ag/Bi disorder in octahedrally coordinated metal positions as well as Q/X mixed occupation of some anion positions. Corresponding to their black color, all compounds are narrow‐gap semiconductors (Eg = 0.35 eV for (II) ). General characteristics of the entire class of (N, x)P compounds are gathered in a catalogue.  相似文献   

9.
Reactions of α‐P4S3I2 with (S)‐ or racemic‐RNH2 (R = CH(Me)Ph) have given solutions containing exo, exo‐ or endo, exo‐α‐P4S3(NHR)2, α‐ or β‐P4S3(μ‐NR), or P4S2(μ‐NR), as the first compounds with polycyclic phosphorus sulfide skeletons to contain chiral substituents. The expected diastereomers have been characterised by complete analysis of their 31P{1H} NMR spectra, and relative configuration has been assigned in most cases. Considerable diastereomeric differences in coupling constants and chemical shifts were found.  相似文献   

10.
Reactions of bicyclic α‐P4S3I2 with Hpthiq gave solutions containing α‐P4S3(pthiq)I and α‐P4S3(pthiq)2, where Hpthiq is the conformationally constrained chiral secondary amine 1‐phenyl‐1,2,3,4‐tetrahydroisoquinoline. The expected diastereomers have been characterised by complete analysis of their 31P{1H} NMR spectra. Hindered P–N bond rotation in the amide iodide α‐P4S3(pthiq)I caused greater broadening of peaks in the room‐temperature spectrum of one diastereomer than in that of the other. At 183 K, spectra of two P–N bond rotamers for each diastereomer were observed and analysed. The minor rotamers showed strong evidence for steric crowding, having large diastereomeric differences in 1J(P–P) and 2J(P–S–P) couplings (49 Hz, 16 % of value, and 4.4 Hz, 19 % of value, respectively).  相似文献   

11.
The reaction of (S)-1,1,2-triphenylethanediol (3) with phosphorus trichloride leads to the diastereoselective formation of (S C,R P)-2-chloro-1,3,2-dioxaphospholane (2). Its configuration has been determined by single crystal X-ray diffraction. When reacted with racemic secondary alcohols, diastereomeric phosphites are obtained, which display substantial shift differences in the 31P NMR spectra. Thus, chlorodioxaphospholane 2 can serve as derivatizing reagent for chiral secondary alcohols permitting to determine their enantiomeric excess.  相似文献   

12.
The complexes Ag(L)n[WCA] (L=P4S3, P4Se3, As4S3, and As4S4; [WCA]=[Al(ORF)4] and [F{Al(ORF)3}2]; RF=C(CF3)3; WCA=weakly coordinating anion) were tested for their performance as ligand-transfer reagents to transfer the poorly soluble nortricyclane cages P4S3, P4Se3, and As4S3 as well as realgar As4S4 to different transition-metal fragments. As4S4 and As4S3 with the poorest solubility did not yield complexes. However, the more soluble silver-coordinated P4S3 and P4Se3 cages were transferred to the electron-poor Fp+ moiety ([CpFe(CO)2]+). Thus, reaction of the silver salt in the presence of the ligand with Fp−Br yielded [Fp−P4S3][Al(ORF)4] ( 1 a ), [Fp−P4S3][F(Al(ORF)3)2] ( 1 b ), and [Fp−P4Se3][Al(ORF)4] ( 2 ). Reactions with P4S3 also yielded [FpPPh3−P4S3][Al(ORF)4] ( 3 ), a complex with the more electron-rich monophosphine-substituted Fp+ analogue [FpPPh3]+ ([CpFe(PPh3)(CO)]+). All complex salts were characterized by single-crystal XRD, NMR, Raman, and IR spectroscopy. Interestingly, they show characteristic blueshifts of the vibrational modes of the cage, as well as structural contractions of the cages upon coordination to the Fp/FpPPh3 moieties, which oppose the typically observed cage expansions that lead to redshifts in the spectra. Structure, bonding, and thermodynamics were investigated by DFT calculations, which support the observed cage contractions. Its reason is assigned to σ and π donation from the slightly P−P and P−E antibonding P4E3-cage HOMO (e symmetry) to the metal acceptor fragment.  相似文献   

13.
Abstract

Aluminum(III) derivatives of O-alkyl or O-aryl trithiophosphate of the type Al[S2P(SH)OR]3 (R = Me, Et, Pri, Bui, Ph, CH2Ph) have been synthesized. The products were obtained as white powdery solids. The monomers are soluble in common organic solvents and were characterized by elemental analyses, molecular weight determinations, and IR and (1H, 27Al, and 31P) NMR spectroscopic studies, which are consistent with six coordinated aluminum and bidentate behavior of the trithiophosphate moiety. The products also exhibit antifungal effectiveness against powdery mildew disease.  相似文献   

14.
A simple and versatile method for general synthesis of uniform one‐dimensional (1D) MxCo3?xS4 (M=Ni, Mn, Zn) hollow tubular structures (HTSs), using soft polymeric nanofibers as a template, is described. Fibrous core–shell polymer@M‐Co acetate hydroxide precursors with a controllable molar ratio of M/Co are first prepared, followed by a sulfidation process to obtain core–shell polymer@MxCo3?xS4 composite nanofibers. The as‐made MxCo3?xS4 HTSs have a high surface area and exhibit exceptional electrochemical performance as electrode materials for hybrid supercapacitors. For example, the MnCo2S4 HTS electrode can deliver specific capacitance of 1094 F g?1 at 10 A g?1, and the cycling stability is remarkable, with only about 6 % loss over 20 000 cycles.  相似文献   

15.
Abstract

The preparation of P4P9 from P4S3 and sulfur in decalin is described. The 31P nmr spectra of P4S9 and P4S10 are presented and discussed.

Die Darstellung von P4S3 aus P4S3 und Schwefel in Dekalin wird beschrieben. Die 31P-NMR-Spektren von P4S9 und P4S10 sind angegeben und interpretiert.  相似文献   

16.
Abstract

Reactions of the monofunctional platinum(II) complex, [PtCl(dien)]+, with different thiols and thioethers, including biologically important molecules, have been studied as a function of temperature (288.2–308.2K) using conventional electronic spectrophotometry in 0.10 M aqueous hydrochloric acid and by 1H NMR spectroscopy. The second-order rate constants, k2, are similar, varying between 1.43 × 10?3 and 46.1 × 10?3 M?1s?1 at 25°C. The reactivity follows the sequences: D-penicillamine ≤ L-cysteine ≤ glutathione ≤ thiodiglycolic acid ≤ thioglycolic acid ≤ L-methionine ≤ S-methylthioglycolic acid ≤ glycyl-D,L-methionine. However, variation in size, bulkiness and solvation of the entering ligands reflect in their properties as nucleophiles. Large negative values of the entropy of activation (ΔS≠), between ?140 and ?190 J K?1 mol?1, indicate that all thiols and thioethers react via the same associative mechanism. Results have been analyzed in relation to the antitumor activity and toxicity of platinum(II) complexes.  相似文献   

17.
    
A variety of tellurium ligands has been designed and studied for their complexation reactions in the last decade. Of these hybrid telluroethers, halotellurium ligands and polytellurides are the most notable ones. RTe-andpolytelluride ions have also been used to design clusters. Ligation of ditelluroethers and several hybrid telluroethers is extensively studied in our laboratories. The ditelluroether ligand RTeCH2TeR (where R = 4-MeOC6H4) (1), similar to dppm [1,2-bis(diphenylphosphino)methane], has been synthesized in good yield (∼80%) by reacting CHCl3 with RTe- (generatedin situ by borohydride reduction of R2Te2). Iodine reacts with1 to give tetra-iodo derivative, which has intermolecular Te.I interactions resulting in a macro structure containing rectangular Te-I.Te bridges.1 readily forms four membered rings with Pd(II) and Ru(II). On the formation of this chelate ring, the signal in125Te NMR spectra shifts significantly upfield (50-60 ppm). The bridging mode of1 has been shown in [Ru(p-cymene)Cl2]](μ-l)[Ru(p-cymene)Cl2]. The hybrid telluroether ligands explored are of the types (Tex, Sy ), (Tex, Ny) and ( Tex,Oy ). The tellurium donor site has strongtrans influence, which is manifested more strongly in square planar complexes of palladium(II). The morpholine N-donor site has been found to have weaker donor characteristics in (Tex, Ny) ligands than pyridine and alkylamine donor sites of analogous ligands. The singlet oxygen readily oxidises the coordinated Te. This oxidation follows first order kinetics. The complexation reaction of RuCl3].xH2O with N-[2-(4-methoxyphenyltelluro)ethyl]phthalimide (2) results in a novel (Te, N, O)-heterocycle, Te-chloro,Te-anisyl-1a-aza-4-oxa-3-tellura-1H, 2H, 4aH-9 fluorenone. The (Te, O) ligands can be used as hemilabile ligands, the oxygen atom temporarily protects the vacant coordination site before the arrival of the substrate. The chelate shifts observed in125Te NMR spectra of metal complexes of Te-ligands have a close parallel to those of31P NMR. For the formation of five-membered rings, the value is positive and of the order of 130 ppm whereas for six-membered rings it is negative and ∼30 ppm only.  相似文献   

18.

Crystal growth and characterization by X-ray diffraction and NMR spectroscopy of a new p-phenylenediamonium diphosphate [p-NH3 C 6 H 4 NH 3]H 2 P 2 O 7 are reported. This compound crystallizes in a triclinic unit cell P1 with the parameters a = 7.130(3), b = 9.047(3), c = 9.350(2) Å, α = 133.44(2)°, β = 95.02(2)°, γ = 107.11(4)°, Z = 2, V = 514.3(15) Å3, and D x = 1.848 g.cm? 3. The crystal structure has been solved and refined to R = 0.0273, using 3678 independent reflections. The atomic arrangement is build up by infinite ribbons of [H2 P 2 O 7] 2? anions, extending along the a-direction at y = 1/2. Between these ribbons are located the p-phenylenediammonium entities, which form hydrogen bonds N─H…O with some external oxygen atoms of phosphoric groups. Crystallographic results are correlated with that of the solid state 13C and 31P MAS NMR spectroscopy.  相似文献   

19.
A simple and versatile method for general synthesis of uniform one‐dimensional (1D) MxCo3−xS4 (M=Ni, Mn, Zn) hollow tubular structures (HTSs), using soft polymeric nanofibers as a template, is described. Fibrous core–shell polymer@M‐Co acetate hydroxide precursors with a controllable molar ratio of M/Co are first prepared, followed by a sulfidation process to obtain core–shell polymer@MxCo3−xS4 composite nanofibers. The as‐made MxCo3−xS4 HTSs have a high surface area and exhibit exceptional electrochemical performance as electrode materials for hybrid supercapacitors. For example, the MnCo2S4 HTS electrode can deliver specific capacitance of 1094 F g−1 at 10 A g−1, and the cycling stability is remarkable, with only about 6 % loss over 20 000 cycles.  相似文献   

20.
Abstract

A series of 1,3-dihydro-2λ4-benzotellurole-2,2-diyl di-thiocarbamates C8H8TeR2 and C8H8TeIR ( RS 2CNC5H10, S2CNHC6H5, S2CNC4H8O) have been synthesised by the reactions of C8H8TeI2 with the corresponding ammonium salts of piperidine-, aniline- and morpholine- dithiocarbamates in 1:1 and 1:2 molar ratio, respectively. They have been characterized by FT-IR and (1H, 13C) NMR spectroscopy. The reaction of C8H8TeI2 with (NH4S2CNC5H10) in 1:2 molar ratio gives C8H8Te(S2CNC5H10)2 [IR, (1H, 13C)NMR evidence] and X-ray quality crystals of Te(S2CNC5H10)2 in very low yield, demonstrating the formation of the first Te–C bond-cleaved product.The monomers of Te(S2CNC5H10)2 are connected through intermolecular Te…S secondary bonds and it exists as a dimer in the solid state. These dimers are interconnected through intermolecular S…S secondary bonds to yield 3D-supramolecular 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