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.
Amino Derivatives of α‐P4S3, α‐P4Se3, and P3Se4; Data and Analyses of their 31P NMR Spectra in Solution α‐P4S3I2, α‐P4Se3I2, and P3Se4I were reacted with primary and secondary amines in CS2. The reaction yields exo‐exo isomeres of α‐P4S3L2 and α‐P4Se3L2, the N‐bridged compounds α‐P4S3L′ and P3Se4L, with L = NHR1, NPhR2, THC (R1 = tBu, Ad, Ph, Flu, TPMP; R2 = Me, Et, iPr), and L′ = NR1. The 31P NMR data of the compounds in CS2 solution were measured. By the reaction of α‐P4Se3I2 with primary amines NH2tBu and NH2Ad in CS2 an asymmetric isomer α‐P4Se3Iendo(NHR1)exo was observed for the first time in the 31P NMR spectra. The influence of the ligands L on the 31P NMR parameter of α‐P4S3L2, α‐P4Se3L2, and P3Se4L is discussed.  相似文献   

2.
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   

3.
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   

4.
Studies on the aggregation degrees of negatively charged phosphides derived from the nucleophilic P4 functionalization could help understand the pathway of phosphorus atoms degradation or aggregation. In this report, we have isolated and characterized four phosphorus cluster anions (P73–, P144–, P162–, and P264–) from the nucleophilic functionalization of P4 with 1,4‐dilithio‐1,3‐butadienes. These phosphorus clusters could be rationalized as the P‐atom‐containing products besides the main phospholyl lithium. Their structural features and 31P NMR behaviors are discussed based on single crystal X‐ray diffraction analysis and 31P{1H} COSY NMR analysis.  相似文献   

5.
The diamide exo, exoβ‐P4S3(NHCH(Me)Ph)2 has been made in solution using enantiomerically pure or racemic PhCH(Me)NH2, and its three diastereomers characterised by complete analysis of their 31P{1H} NMR spectra.The unsymmetric diastereomer contains phosphorus atoms, made chemically non‐equivalent by the chirality of the substituents, which show a large 2J(P—P—P) coupling to each other (225.2 Hz).  相似文献   

6.
Reaction of [U(TrenTIPS)] [ 1 , TrenTIPS=N(CH2CH2NSiiPr3)3] with 0.25 equivalents of P4 reproducibly affords the unprecedented actinide inverted sandwich cyclo‐P5 complex [{U(TrenTIPS)}2(μ‐η55‐cyclo‐P5)] ( 2 ). All prior examples of cyclo‐P5 are stabilized by d‐block metals, so 2 shows that cyclo‐P5 does not require d‐block ions to be prepared. Although cyclo‐P5 is isolobal to cyclopentadienyl, which usually bonds to metals via σ‐ and π‐interactions with minimal δ‐bonding, theoretical calculations suggest the principal bonding in the U(P5)U unit is polarized δ‐bonding. Surprisingly, the characterization data are overall consistent with charge transfer from uranium to the cyclo‐P5 unit to give a cyclo‐P5 charge state that approximates to a dianionic formulation. This is ascribed to the larger size and superior acceptor character of cyclo‐P5 compared to cyclopentadienyl, the strongly reducing nature of uranium(III), and the availability of uranium δ‐symmetry 5f orbitals.  相似文献   

7.
Reaction of a mixture of bicyclic phosphorus sulfide selenide iodides α‐P4SnSe3−nI2 (n = 0–3) with PriNH2 and Et3N gave corresponding diamides α‐P4SnSe3−n(NHPri)2 (n = 0–3) and imides α‐P4SnSe3−n(μ‐NPri) (n = 2–3), identified in solution by 31P NMR. In one isomer of α‐P4S2Se(μ‐NPri), the C2 symmetry of imides such as α‐P4S3(μ‐NPri) was broken, allowing relative assignment of 2J NMR couplings to the PNP bridge and the PSP bridge opposite to it. The coupling through the sulfur bridge was found to be reduced to ca. zero, in contrast to previous assumptions for this class of compounds. Ab initio models were calculated at the MPW1PW91/svp level for the sulfide selenide imides and for a selection of bond rotamers of the diamides, and at the MPW1PW91/LanL2DZ(d) level for the sulfide selenide diiodides. Different skeletal isomers were prevalent for the mixed chalcogenide diamides than for the diiodides, showing that exchange of chalcogen between skeletal positions took place in the amination reaction even at room temperature. Similar differences to those observed were predicted by the models, suggesting that equilibrium was attained.  相似文献   

8.
Contributions to the Chemistry of Phosphorus. 231. Li3P7S3 and Li2HP7S2 — the First Sulfido Heptaphosphanes(3) The novel sulfido heptaphosphanes(3) Li3P7S3 ( 1 ) and Li2HP7S2 ( 2 ) have been obtained by the reaction of Li3P7 with sulfur in tetrahydrofuran under suitable conditions. The compounds 1 and 2 are also formed from LiP5 and sulfur and are only stable in solution below room temperature. According to a complete analysis of the 31P{1H}-NMR spectra, in each case, the sulfur atoms are bonded as sulfido groups exocyclically to the heptaphosphanortricyclene skeleton. Compound 2 reacts with chloro(trimethyl)silane or acetylacetone at one of the two sulfido groups while compound 1 does not form any product with retention of the P7(3) framework.  相似文献   

9.
An efficient NaBArF4‐catalyzed oxidative cyclization of readily available 1,5‐ and 1,6‐diynes has been developed. Importantly, this transition metal‐free oxidative catalysis proceeds via a presumable Lewis acid‐catalyzed SN2’ pathway, which is distinct from the relevant oxidative rhodium and gold catalysis. This method leads to the facile and practical construction of a diverse range of synthetically useful γ‐ and δ‐lactams in mostly good to excellent yields with broad substrate scope.  相似文献   

10.
Pale blue, lath‐shaped single crystals of K2NdP2S7 (≡ K4Nd2[PS4]2[P2S6]; monoclinic, P21/n, a = 904.76(8), b = 677.38(6), c = 1988.7(2) pm, β = 97.295(5)°, Z = 2) are obtained by the reaction of Nd, S and P2S5 with an excess of KCl as a flux in evacuated silica tubes at 750 °C (7 d) which should produce Nd[PS4] instead. Beside isolated [PS4]3– tetrahedra, the crystal structure contains discrete ethane‐analogous [P2S6]4– (≡ [S3P–PS3]4–) units in staggered conformation with tetravalent phosphorus cations and a P–P distance of 219 pm. The two crystallographically different potassium cations show coordination numbers of nine and ten in the shape of distorted mono‐ and bicapped square antiprisms. Finally, the Nd3+ cations are surrounded by eight sulfur atoms arranged as (uncapped) square antiprisms. The entire structure is dominated by (K1)+ containing {(Nd2[PS4]2[P2S6])4–} layers parallel (101) which are three‐dimensionally interconnected by (K2)+ cations.  相似文献   

11.
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   

12.
The Molecular Composition of Solidified Phosphorus-Sulfur Melts and the Crystal Structure of β-P4S6 Phosphorus sulfur melts were annealed for one week at 673 K and then quenched in ice water. The solids were dissolved in CS2 and the concentrations of phosphorus sulfides were determined by 31P NMR spectroscopy. Samples containing between 44 and 70 mol% sulfur dissolved completely in CS2. Between 0 and 42 mol% remains an insoluble residue of red phosphorus. Above 72 mol% it consisted of sulfur chains linked by phosphorus atoms. The solutions contained mainly the congruently melting compounds P4S3, P4S7, and P4S10 having maximum concentrations at their stoichiometric compositions. Other compounds P4Sn (n = 4–9) which decompose on heating, according to the phase diagram, were also found in surprisingly high concentrations. One of these was β-P4S6 which crystallizes in the monoclinic space group P21/c with the lattice parameters a = 702.4(2), b = 1 205.6(2), c = 1 148.9(6) pm and β = 103.4(2)°. Reaction of white phosphorus with sulfur was also investigated. In contrast to the results of previous authors, who described the system P4–S8 below 373 K as eutectic, we found that the elements reacted below this temperature.  相似文献   

13.
The new cyclotriphosphazene derivative N3P3(OC6H3OCH3COH)6 ( 1 ) was synthesized from hexachlorocyclotriphosphazene, N3P3Cl6, and 4‐hydroxy‐3‐methoxybenzaldehyde in acetonitrile in the presence of K2CO3. The structure of 1 was verified by means of elemental analysis, IR, 1H NMR, 13C NMR, 31P NMR spectra, thermal analysis and X‐ray diffraction.  相似文献   

14.
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   

15.
The organic acid–base complex 1,1,3,3‐tetramethylguanidinium 4‐methylbenzenesulfonate, C5H14N3+·C7H7O3S, was obtained from the corresponding 1,1,3,3‐tetramethylguanidinium 4‐methylbenzenesulfinate complex, C5H14N3+·C7H7O2S, by solid‐state oxidation in air. Comparison of the two crystal structures reveals similar packing arrangements in the monoclinic space group P21/c, with centrosymmetric 2:2 tetramers being connected by four strong N—H...O=S hydrogen bonds between the imine N atoms of two 1,1,3,3‐tetramethylguanidinium bases and the O atoms of two acid molecules.  相似文献   

16.
Treatment of 1,8‐bis(diphenylphosphino)naphthalene (dppn, 1 ) with stoichiometric amounts of sulfur or selenium in toluene at 80 °C selectively afforded the diphosphine monochalcogenides 1‐Ph2P(C10H6)‐8‐P(:S)Ph2 (dppnS, 2 a ) and 1‐Ph2P(C10H6)‐8‐P(:Se)Ph2 (dppnSe, 2 b ). The 31P{1H} NMR spectrum of 2 b showed an unusually large 5J(P–Se) value, which indicates a significant through‐space coupling component. The monosulfide acted as a bidentate P,S‐ligand towards platinum(II) ( 3 a ), whereas the corresponding monoselenide complex ( 3 b ′) lost elemental selenium with formation of the previously reported complex [PtCl2(dppn)‐P,P′] ( 3 ). Treatment of dppnSe with [(nor)Mo(CO)4] (nor = norbornadiene) led to formation of [(dppnSe)Mo(CO)4P,Se] ( 3 b ). Solutions of the latter slowly deposited Se with formation of [(dppn)Mo(CO)4P,P′] ( 4 ) which was also obtained by independent synthesis from 1 and [(nor)Mo(CO)4]. All isolated new compounds were characterised by a combination of 31P, 1H, 13C and 77Se ( 2 b ) NMR spectroscopy, IR spectroscopy, mass spectrometry and elemental analysis. Single‐crystal X‐ray structure determinations were performed for dppnSe ( 2 b ), [PtCl2(dppnS)‐P,S] ( 3 a ), [(dppnSe)Mo(CO)4P,Se] ( 3 b ) and [(dppn)Mo(CO)4P,P′] ( 4 ). In 2 b steric effects cause the naphthalene ring to be distorted and force the phosphorus atoms by 65 and 59 pm to opposite sides of the best naphthalene plane. In the metal complexes 3 a , 3 b and 4 the phosphino‐phosphinochalcogenyl systems act as bidentate ligands through the P and the chalcogen atoms. The naphthalene systems are again distorted. The two independent molecules of 4 differ in their conformations.  相似文献   

17.
Multianvil Synthesis, X‐ray Powder Diffraction Analysis, 31P‐MAS‐NMR, and FTIR Spektroscopy as well as Material Properties of γ‐P3N5, a High‐Pressure Polymorph of Binary Phosphorus(V) Nitride, Built up from Distorted PN5 Square Pyramids and PN4 Tetrahedra The high‐pressure phase γ‐P3N5 was synthesized at a pressure of 11 GPa and a temperature of 1500 °C in a multianvil apparatus. Partially crystalline P3N5 has been used as a starting material. The crystal structure was solved by direct methods on the basis of X‐ray powder diffraction data and it was refined by the Rietveld method (Imm2, a = 1287.21(4), b = 261.312(6), c = 440.03(2) pm, Z = 2, Rp = 0.073, wRp = 0.094, RF = 0.048). γ‐phosphorus nitride crystallizes in a three‐dimensional network structure built up from corner sharing PN4 tetrahedra and trans‐edge sharing distorted PN5 square pyramids. In the 31P‐MAS‐NMR spectrum two sharp isotropic resonances with an intensity ratio of 1 : 2.02(5) are observed at —11.95(3) and —101.72(7) ppm, respectively. The IR‐spectroscopic and thermal properties of γ‐P3N5 are described. Measurement of the Vickers hardness resulted in a value of 9.7(21) GPa for sintered polycrystalline γ‐P3N5, which is significantly higher than that for the partially crystalline normal pressure modification of P3N5 (5.1(7) GPa).  相似文献   

18.
The structure of an iridolactone isolated from Valeriana laxiflora was established as (4R,4aR,6S,7S,7aS)‐6‐hydroxy‐7‐hydroxy­methyl‐4‐methyl­per­hydro­cyclo­penta­[c]­pyran‐1‐one chloro­form solvate, C10H16O4·CHCl3. The two rings are cis‐fused. The δ‐lactone ring adopts a slightly twisted half‐chair conformation with approximate planarity of the lactone group and the cyclo­pentane ring adopts an envelope conformation. The hydroxy group, the hydroxymethyl group and the methyl group all have β orientations. The absolute configuration was determined using anomalous dispersion data enhanced by the adventitious inclusion of a chloro­form solvent mol­ecule. Hydro­gen bonding, crystal packing and ring conformations are discussed in detail.  相似文献   

19.
The title molecule, 3‐{[4‐(3‐methyl‐3‐phenyl‐cyclobutyl)‐thiazol‐2‐yl]‐hydrazono}‐1,3‐dihydro‐indol‐2‐one (C22H20N4O1S1), was prepared and characterized by 1H NMR, 13C NMR, IR, UV–visible, and single‐crystal X‐ray diffraction. The compound crystallizes in the monoclinic space group P21 with a = 8.3401(5), b = 5.6976(3), c = 20.8155(14) Å, and β = 95.144(5)°. Molecular geometry from X‐ray experiment and vibrational frequencies of the title compound in the ground state has been calculated using the Hartree–Fock with 6‐31G(d, p) and density functional method (B3LYP) with 6‐31G(d, p) and 6‐311G(d, p) basis sets, and compared with the experimental data. The calculated results show that optimized geometries can well reproduce the crystal structural parameters, and the theoretical vibrational frequencies values show good agreement with experimental data. Density functional theory calculations of the title compound and thermodynamic properties were performed at B3LYP/6‐31G(d, p) level of theory. © 2011 Wiley Periodicals, Inc. Int J Quantum Chem, 2012  相似文献   

20.
Heteropentapeptides containing the α‐ethylated α,α‐disubstituted amino acid (S)‐butylethylglycine and four dimethylglycine residues, i.e., CF3CO‐[(S)‐Beg]‐(Aib)4‐OEt ( 4 ) and CF3CO‐(Aib)2‐[(S)‐Beg]‐(Aib)2‐OEt ( 7 ), were synthesized by conventional solution methods. In the solid state, the preferred conformation of 4 was shown to be both a right‐handed (P) and a left‐handed (M) 310‐helical structure, and that of 7 was a right‐handed (P) 310‐helical structure. IR, CD, and 1H‐NMR spectra revealed that the dominant conformation of both 4 and 7 in solution was the 310‐helical structure. These conformations were also supported by molecular‐mechanics calculations.  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