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.
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   

2.
On the Reaction of P4E3I2 (E = S, Se) with some Carboxylic Acids and Dithiocarbamic Acids By the reaction of α-P4E3I2 (E = S, Se) with carboxylic acids, dithiobenzoic acid or dithiocarbamic acids in the presence of triethylamin or with (C6H5)3SnR, or of β-P4E3I2 with tin-organic compounds α-P4E3(I)R, α(β)-P4E3R2 [R = ? OC(O)C6H5, ? OC(O)CH3, ? SC(S)NC5H10, ? SC(S)N(C2H5)2], α-P4S3(I)SC(S)C6H5, α-P4S3(SC(S)C6H5)2 and β-P4E3(I)R (R = ? OC(O)C6H5, ? OC(O)CH3) were prepared in solution and identified by 31P NMR spectroscopy. In addition α-P4S3(NC5H10)(SC(S)NC5H10) was detected. The β-isomers could be obtained also with lesser yields by the reaction with the dithiocarbamic acids, too. The substitution of the second iodine ligand in β-P4E3I2 resulted mainly in β-P4S3(Rexo)2 and by inversion of the configuration at a phosphorus atom, in β-P4E3RexoRendo. α-P4S3I2 reacted with methanol in CS2 to α-P4S3(OCH3)(SC(S)OCH3) and α-P4S3(SC(S)OCH3)2. The 31P NMR data of the compounds are discussed. The 31P NMR spectra of the α(β)-P4E3 dithiocarbamates indicate dynamic processes in the solution, e. g. α-P4S3(I)(SC(S)NR2) showed an intramolecular conversion, due to the anisobidentate dithiocarbamate ligand. This behaviour had not previously been noticed for compounds with a P4S3-skeleton.  相似文献   

3.
Reactions of Alcohols and Mercaptans with Tetraphosphorus Trichalcogenide Diiodides α-P4E3(I)R and α-P4E3R2 (R = -OCH3, -OC2H5, -OCH(CH3)2, -OC(CH3)3 and -OC6H5) have been detected in the reaction solutions of α-P4E3I2 (E = S, Se) with alcohols in the presence of triethylamine, or with trimethyltin alkoxides in CS2. β-P4S3I2 reacted with methanol at 243 K in CS2 to yield β-P4S3(I)OCH3, β-P4S3(OCH3)exo(OCH3)endo, and β-P4S3[(OCH3)exo]2. The thiolates α-P4Se3R2′ (R′ = -SC2H5, -n-SC5H11, -n-SC7H15, -SC(CH3)3 and -SC6H5) were found in reaction solutions of α-P4Se3I2 with thiols HR′. The 31P NMR data of these compounds are given.  相似文献   

4.
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   

5.
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   

6.
Bicyclic Diiodine Trichalcogenatetetraphosphaheptanes Two isomers of each of β-P4S2SeI2 and β-P4SSe2I2 have been identified by 31P-NMR spectroscopy. They were prepared by the reaction of P4S3–nSen with I2 in CS2. Systematic changes in chemical shifts and in coupling constants with the alterations in molecular geometry, as sulfur was replaced by selenium, are reported.  相似文献   

7.
Reactions of Tetraphosphorus Trichalcogenides with Alkyl Iodides Reactions of alkyl iodides RI (R = CHI2, CH2I or tert-Butyl) with P4E3 (E = S or Se) under the influence of light resulted in cleavage of the basal P3 ring. β-P4E3(I)R was formed initially, then it rearranged to the more stable α-P4E3(I)R structure. 31P NMR data of these products were measured and discussed, along with 77Se data for α- and β-P477SeSe2(I)CHI2. On reaction of P4S3 with tert-butyl iodide in CS2 or with sec-butyl iodide or iso-propyl iodide in dioxane, the new type of compounds P5S2R was observed. In this a sulfur bridge of P4S3 is replaced by a P? R group. 31P-NMR data for these compounds are reported.  相似文献   

8.
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   

9.
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   

10.
Novel A4B3 Molecules in the System P4Se3–As4Se3 By means of 31P-NMR and masspectroscopic measurements in the system P4Se3–As4Se3 was shown that in the melt and vapour phase at all compositions molecules of the type P4 ? nAsnSe3 are formed. A separation was possible by liquid chromatography (RP 18-column). The concentration distribution of the different species is nearly statistical. In the solid state at ambient temperature regions of solid solubility with α-P4Se3, α+-phase, α-P4S3 and α-As4Se3 structure were observed. P3AsSe3 could be transformed into a plastically-crystalline phase with β-P4S3 structure. At higher temperatures the phase decomposes slowly. The thermal behaviour of PAs3Se3 is strongly influenced by the heating rate. Using low heating rates it decomposes into an amorphous phase, by fast heating a transformation into a metastable plastically-crystalline modification was achieved. During long extraction with CS2 molecules P4 ? nAsnS3 ? mSem are formed by an exchange reaction. They can also be prepared by melting the proper amounts of the elements.  相似文献   

11.
Neues vom P4Se4     
New Results on P4Se4 Preparation of P4Se4 from the elements yields always the β-modification of P4Se4. α-P4Se4 is obtained only with selenium deficient samples. However, it is also observed, when P4Se3 is annealed and then extracted with CS2. The insoluble part has the X-ray pattern of α-P4Se4. A reversible α-β transition is not observed. MAS-31P-NMR investigations on solid P4Se4 by Eckert et al. [2] reveal P2Se4/2 building units, which are, in view of our results, not dimer but linked to a polymeric network. Well-crystallized samples of β-P4Se4 are obtained only at measuring temperatures above 573 K. The structure is of monoclinic symmetry with the space group P21/n (a = 114.9, b = 729.0, c = 1211.0 pm, β = 120.80°). The reaction of α-P4Se3I2 with bis-(trimethyltin)selenide in CS2 at low temperature yields molecular α-P4Se4, which can be detected in solution by 31P-NMR spectroscopy. α-P4Se4 has D2d-symmetry like α-P4S4. It polymerizes at higher temperature. α-P4Se3I(SeSnMe3) and α-P4Se3(SeSnMe3)2 were observed in the course of this reaction, too. The analogous reaction of α-P4Se3I2 with bis-(trimethyltin)sulfide leads to comparable results.  相似文献   

12.
Reactions of P4S3I2 with Bifunctional Ligands Nitrogen, oxygen, selenium or carbon atoms were introduced as bridges into the P4S3 skeleton by the reaction of α- or β-P4S3I2 with bifunctional ligands. Among compounds in the series α-P4S3-α-E, the oxide μ-P4SO was made for the first time. High concentrations of α-P4S4Se, made by a new route, allowed observation of 77Se satellites in its 31P NMR spectrum and hence the assignment of 31P chemical shifts. Polymeric species were more stable than these monomers, leading to low yields in both reactions. α- and β-isomers of P4S3I2 reacted with diethyl malonate. While β-P4S3I2 gave traces of β-P4S3(CR2)(R = C2H5CO2) and of β-P4S3(CHR2)exo(CHR2)endo, along with insoluble products, α-P4S3I2 yielded α-P4S3(CR2), which could be isolated. P4S2(CR2), a new skeleton similar to that of P4S3, was formed on storage of CS2 solutions of a-P4S3(CR2) for two days. The 31P NMR data of the molecules are given.  相似文献   

13.
The reaction behaviour of 1, 3, 5‐triaza‐2σ3λ3‐phosphorin‐4, 6‐dionyloxy‐substituted calix[4]arenes towards mono‐ and binuclear rhodium and platinum complexes was investigated. Special attention was directed to structure and dynamic behaviour of the products in solution and in the solid state. Depending on the molar ratio of the reactands, the reaction of the tetrakis(triazaphosphorindionyloxy)‐substituted calix[4]arene ( 4 ) and its tert‐butyl‐derivative ( 1 ) with [(cod)RhCl]2 yielded the mono‐ and disubstituted binuclear rhodium complexes 2 , 3 , and 5 . In all cases, a C2‐symmetrical structure was proved in solution, apparently caused by a fast intramolecular exchange process between cone conformation and 1, 3‐alternating conformation. The X‐ray crystal structure determination of 5 confirmed [(calixarene)RhCl]2‐coordination through two opposite phosphorus atoms with a P ⃜P separation of 345 pm. The complex displays crystallographic inversion symmetry, and the Rh2Cl2 core is thus exactly planar. Reaction of 1 and of the bis(triazaphosphorindionyloxy)‐bis(methoxy)‐substituted tert‐butyl‐calix‐[4]arene ( 7 ) with (cod)Rh(acac) in equimolar ratio and subsequent reaction with HBF4 led to the expected cationic monorhodium complexes 5 and 8 , involving 1, 3‐alternating P‐Rh‐P‐coordination. The cone conformation in solution was proved by NMR spectroscopy and characteristic values of the 1J(PRh) coupling constants in the 31P‐NMR‐spectra. Reaction of equimolar amounts of 4 with (cod)Rh(acac) or (nbd)Rh(acac) led, by substitution of the labile coordinated acetylacetonato and after addition of HBF4, to the corresponding mononuclear cationic complexes 9 and 10 . Only two of the four phosphorus atoms in 9 and 10 are coordinated to the central metal atom. Displacement of either cycloocta‐1, 5‐diene or norbornadiene was not observed. For both compounds, the cone conformation was proved by NMR spectroscopy. Reaction of 4 with (cod)PtCl2 led to the PtCl2‐complex ( 11 ). As for all compounds mentioned above, only two phosphorus atoms of the ligand coordinate to platinum, while two phosphorus atoms remain uncoordinated (proved by δ31P and characteristic values of 1J(PPt)). NMR‐spectroscopic evidence was found for the existence of the cone conformation in the cis‐configuration of 11 .  相似文献   

14.
A series of novel 5‐alkyl(aryl)‐3‐[(3‐trifluoromethyl)anilino]‐4,5‐dihydro‐1H‐pyrazolo[4,3‐d]pyrimidin‐4‐imines were designed and synthesized by the multistep reactions. 1 reacted with 3‐(trifluoromethyl)aniline to obtain N,S‐acetals 2 in the presence of 10 mol% DBU. 2 reacted with hydrazine to form 5‐amino‐3‐[(3‐trifluoromethyl)anilino]‐1H‐pyrazol‐4‐ nitrile 3 , the target compounds 5a , 5b , 5c , 5d , 5e , 5f , 5g , 5h , 5i , 5j , 5k , 5l , 5m were obtained by the reaction of 3 with triethyl orthoformate followed by the cyclization of 4 with various amines. Their structures were confirmed by IR, 1H NMR, EI‐MS, and elemental analyses. The preliminary bioassay indicated that some of them displayed moderate herbicidal activity against dicotyledonous weed Brassica campestris L at the concentration of 100 mg/L. For example, compounds 5f , 5g , 5h , and 5j possessed 60.1%, 63.4%, 67.1%, and 61.3% inhibition against Brassica campestris L at the concentration of 100 mg/L.  相似文献   

15.
Diselenadiphosphetane Diselenides and Triselenadiphospholane Diselenides – Synthesis and Characterization by 31P and 77Se Solid‐State NMR Spectroscopy 1,3‐Diselena‐2,4‐diphosphetane‐2,4‐diselenides (RPSe2)2 with R = Me, Et, t‐Bu, Ph, 4‐Me2NC6H4, 4‐MeOC6H4 have been synthesized by different methods. The insoluble compounds were investigated by 31P and 77Se solid‐state NMR and the purity of the compounds has been checked by their CP MAS sideband NMR spectra. The structure of the investigated compounds has been confirmed by the isotropic and anisotropic values of the chemical shifts and the 1JP–Se coupling constants. In addition, two new 1,2,4‐triselena‐3,5‐diphospholane‐3,5‐diselenides, (RPSe2)2Se (R = Me, Et), formed under similar synthesis conditions, were investigated. Their structure was derived from the 77Se satellites of 31P solution spectra and from solid‐state spectra. For (t‐BuPSe2)2 the experimentally obtained principal values of phosphorus and selenium shielding tensors are compared with values from IGLO calculations (HF und SOS DFPT). The calculated orientations of the principal axes are discussed.  相似文献   

16.
Two novel tridentate ligands of 2,6‐bis‐[l‐(2,6‐dibromophenylimino) ethyl] pyridine (L1) and2‐acetyl‐6‐[1‐(2,6‐dibromophenylimino) ethyl] pyridine (L2) have been synthesized. The iron(II) complex of L1 and L2 has been characterized with the crystal structure of [Fe(L1)(L2)]2+ [FeCl4]2 CH2Cl2 [monoclinic, P21 (#11), a = 1.0562(4), b = 2.0928(4), c = 1.2914(2) nm, β = 100.12°, V = 2.810(1) nm3 Dc = 1.879 g/cm3 and Z = 2].  相似文献   

17.
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   

18.
The reactions of P4S3 with As4S3 and of P4Se3 with As4Se3 in the molten state yields molecules of the type P m As4–m S3 and P m As4–m Se3, respectively. A method was developed to separate the different components by the HPLC technique, and to determine their concentrations. The identification of the isomers in the HPLC pattern was achieved with the aid of the LC-MS method. In the selenium system, the distribution of the different species is statistical. In the system P4S3-As4S3, the formation of PAs3S3 with one phosphorus atom in the apical position is favoured.
  相似文献   

19.
While the gold(I)‐catalyzed glycosylation reaction with 4,6‐O‐benzylidene tethered mannosyl ortho‐alkynylbenzoates as donors falls squarely into the category of the Crich‐type β‐selective mannosylation when Ph3PAuOTf is used as the catalyst, in that the mannosyl α‐triflates are invoked, replacement of the ?OTf in the gold(I) complex with less nucleophilic counter anions (i.e., ?NTf2, ?SbF6, ?BF4, and ?BAr4F) leads to complete loss of β‐selectivity with the mannosyl ortho‐alkynylbenzoate β‐donors. Nevertheless, with the α‐donors, the mannosylation reactions under the catalysis of Ph3PAuBAr4F (BAr4F=tetrakis[3,5‐bis(trifluoromethyl)phenyl]borate) are especially highly β‐selective and accommodate a broad scope of substrates; these include glycosylation with mannosyl donors installed with a bulky TBS group at O3, donors bearing 4,6‐di‐O‐benzoyl groups, and acceptors known as sterically unmatched or hindered. For the ortho‐alkynylbenzoate β‐donors, an anomerization and glycosylation sequence can also ensure the highly β‐selective mannosylation. The 1‐α‐mannosyloxy‐isochromenylium‐4‐gold(I) complex ( Cα ), readily generated upon activation of the α‐mannosyl ortho‐alkynylbenzoate ( 1 α ) with Ph3PAuBAr4F at ?35 °C, was well characterized by NMR spectroscopy; the occurrence of this species accounts for the high β‐selectivity in the present mannosylation.  相似文献   

20.
Contributions to the Chemistry of Phosphorus. 221. Stannyl-Substituted Bicyclo[1.1.0]tetraphosphanes: Formation and properties of R3Sn(H)P4 (R ? CH3, C6H5, c-C6H11, o-C7H7) The unsymmetrically substituted bicyclo[1.1.0]tetraphosphanes Me3Sn(H)P4 ( 1 ), Ph3Sn(H)P4 ( 2 ), (c-Hex)3Sn(H)P4 ( 3 ) and (o-Tol)3Sn(H)P4 ( 4 ) have been obtained by reaction of a solution of (Na/K) HP4 with R3 SnCl (R ? Me, Ph, c-Hex, o-Tol) under proper conditions. The structure of the compounds 1 – 4 , which are only stable in solution, has been elucidated by means of 31P-NMR-spectroscopy. Whereas 3 exists at ?60°C as the exo,endo isomer, 1, 2 and 4 are fluctuating molecules at room temperature and probably invert between the three possible configurational isomers (exo,exo-, exo,endo- and endo,endo-form).  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