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 22 毫秒
1.
Variable‐temperature 1H and 77Se NMR data for 3‐phenylselenenyl‐1‐phenyl‐1‐propene (1) in the presence of Rh2(MTPA)4 (Rh*) prove that the equilibria are strongly shifted towards the adduct Rh*···1; free selenide molecules cannot be detected as long as uncomplexed rhodium atoms are available. In the case of excess Rh*, both 1 : 2 and 1 : 1 adducts (Rh* vs 1) are formed, and the latter is slightly favoured. With excess selenide, the system strongly favours the complexation of two selenide molecules (1 : 2 adduct), i.e. one at each rhodium atom. In this situation, intermolecular selenide exchange can be monitored by variable‐temperature 1H NMR spectroscopy and the energy barrier is estimated to be 54–55 kJ mol?1. Copyright © 2001 John Wiley & Sons, Ltd.  相似文献   

2.
Complexation of the oxygen atom in 2‐butylphenylethers and sulfur in 2‐butylphenylthioethers to a rhodium atom in dirhodium tetracarboxylate Rh(II)2[(R)‐(+)‐MTPA]4 is compared. Oxygen atoms complex via electrostatic attraction exclusively leading to an increase in α effects on C‐2 complexation shifts in the sequence OCH3 > F > Br > NO2. However, that trend is opposite in thioethers. This can be rationalized by an additional highest occupied molecular orbital (HOMO)–LUMO interaction and the response of this interaction upon complex formation shifts. Thereby, an experimental evidence was found for the existence of the HOMO–LUMO binding mechanism which has been proposed previously based on theoretical considerations and indirect spectroscopic evidence. Sulfones hardly bind to Rh(II)2[(R)‐(+)‐MTPA]4. Diastereomeric dispersion effects at 13C and 1H signals can be observed for all compounds indicating that enantiodifferentiation is easy in all classes of functionalities. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

3.
Redistribution reactions between diorganodiselenides of type [2‐(R2NCH2)C6H4]2Se2 [R = Et, iPr] and bis(diorganophosphinothioyl disulfanes of type [R′2P(S)S]2 (R = Ph, OiPr) resulted in the hypervalent [2‐(R2NCH2)C6H4]SeSP(S)R′2 [R = Et, R′ = Ph ( 1 ), OiPr ( 2 ); R = iPr, R′ = Ph ( 3 ), OiPr ( 4 )] species. All new compounds were characterized by solution multinuclear NMR spectroscopy (1H, 13C, 31P, 77Se) and the solid compounds 1 , 3 , and 4 also by FT‐IR spectroscopy. The crystal and molecular structures of 3 and 4 were determined by single‐crystal X‐ray diffraction. In both compounds the N(1) atom is intramolecularly coordinated to the selenium atom, resulting in T‐shaped coordination arrangements of type (C,N)SeS. The dithio organophosphorus ligands act monodentate in both complexes, which can be described as essentially monomeric species. Weak intermolecular S ··· H contacts could be considered in the crystal of 3 , thus resulting in polymeric zig‐zag chains of R and S isomers, respectively.  相似文献   

4.
Reactions of the 16e halfsandwich complexes Cp*M[Se2C2(B10H10)] ( 5 M = Rh, 6 M = Ir) with both methyl acetylene monocarboxylate and dimethyl acetylene dicarboxylate were studied in order to obtain information on the influence of the chalcogen (selenium versus sulfur), as well as further evidence for B–H activation, ortho‐metalation and substitution of the carborane. In the case of the rhodium‐selenium complex 5 , the reaction with methyl acetylene monocarboxylate gave products which were all structurally different compared to those of the sulfur analogue of 5 : a polycyclic derivative 12 with a B(6)‐substituted carborane cage was obtained as one of the final products; in addition, both geometrical isomers containing a Rh–B bond ( 10 , 11 ) and isomers without a Rh–B bond ( 8 , 9 ) were isolated, the latter being the result of twofold insertion into one of the Rh–Se bonds. In the case of the iridium‐selenium complex 6 , the reaction with methyl acetylene monocarboxylate led to the geometrical isomers 13 and 14 (similar to 10 and 11 ) with structures possessing an Ir–B bond. Both 5 and 6 reacted with dimethyl acetylene dicarboxylate at room temperature to give the complexes 15 and 16 which are formed by addition of the C≡C unit to the metal center and insertion into one of the metal‐selenium bonds. The proposed structures in solution were deduced from NMR data (1H, 11B, 13C, 77Se, 103Rh NMR), and an X‐ray structural analysis was carried out for the rhodium complex 12 .  相似文献   

5.
[2‐(Me2NCH2)C6H4]Se? S(S)PR2 [R = Ph (1), OiPr (2)] were prepared by reacting [2‐(Me2NCH2)C6H4]2Se2 with the appropriate disulfanes, [R2P(S)S]2. The compounds were characterized by multinuclear magnetic resonance (1H, 13C, 31P). The molecular structures of 1 and 2 were determined by single‐crystal X‐ray diffraction. Both compounds are monomeric and the nitrogen atom of the pendent CH2NMe2 arm is strongly coordinated to the selenium atom. The organophosphorus ligands are monodentate, thus resulting in a T‐shaped coordination geometry around selenium. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

6.
Supermesityl selenium diimide [Se{N(C6H2tBu3‐2, 4, 6)}2; Se{N(mes*)}2] can be prepared in a good yield from the reaction of SeCl4 and (mes*)NHLi. The molecule adopts an unprecedented anti, anti‐conformation, as deduced by DFT calculations at PBE0/TZVP level of theory and supported by 77Se NMR spectroscopy and a crystal structure determination. An analogous reaction involving (C6H2Me3‐2, 4, 6)NHLi [(mes)NHLi] unexpectedly lead to the reduction of selenium and afforded the selenium diamide Se{NH(mes)}2 that was characterized by X‐ray crystallography and 77Se NMR spectroscopy. The Se‐N bonds of 1.847(3) and 1.852(3) Å show normal single bond lengths. The <NSeN bond angle of 109.9(1)° also indicates a tetrahedral AX2E2 bonding arrangement around selenium. Two N‐H···N hydrogen bonds link the Se{NH(mes)}2 molecule with two discrete (mes)NH2 molecules. In the solid state selenium diamide adopts the anti‐conformation, whereas in solution the presence of both syn‐ and anti‐isomers could be observed. PBE0/TZVP calculations of the shielding tensors of 28 different types of selenium‐containing molecules, for which the 77Se chemical shifts are unambiguously known, were carried out to assist the spectral assignment of Se{N(mes*)}2 and Se{NH(mes)}2.  相似文献   

7.
The trans‐bis(trimethylsilyl)chalcogenolate palladium complexes, trans‐[Pd(ESiMe3)2(PnBu3)2] [E = S ( 1 ) and Se ( 2 )] were synthesized in good yields and high purity by reacting trans‐[PdCl2(PBu3)2] with LiESiMe3 (E = S, Se), respectively. These complexes were characterized by 1H, 13C{1H}, 31P{1H} (and 77Se{1H}) NMR spectroscopy and single‐crystal X‐ray analysis. The reaction of 2 with propionyl chloride led to the formation of trans‐[Pd(SeC(O)CH2CH3)2(PnBu3)2] ( 3 ), a trans‐bis(selenocarboxylato) palladium complex and thus established a new method for the formation of this type of complex. Complex 3 was characterized by 1H, 13C{1H}, 31P{1H} and 77Se{1H} NMR spectroscopy and a single‐crystal X‐ray structure analysis.  相似文献   

8.
Unambiguous resonance assignments of diastereotopic CH2 protons in the anomeric side chain of nine alkyl‐ and aralkylselenoglycosides have been carried out on the basis of experimental CPMG‐HSQMBC measurements and theoretical second order polarization propagator approach (SOPPA) calculations of geminal 77Se‐1H spin‐spin coupling constants involving diastereotopic pro‐R and pro‐S protons. Theoretical conformational analyses have been performed at the MP2/6‐311G** level. The conformational space of each of the selenoglycosides under study could be adequately described as a mixture of six interconverting conformers with the molar fractions depending on the nature of the side chain substituent at the selenium atom. The good agreement observed between measured and the weighted conformational averaged values of the calculated coupling constants provides a basis for reliable diastereotopic assignments in this type of carbohydrate structures. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

9.
Enantiopure α‐amino acids were converted to 4‐substituted 2‐aryl‐ and 2‐alkyl‐5(4H)‐oxazolones under partial racemization. These nonracemic mixtures were dissolved in CDCl3, an equimolar amount of the chiral dirhodium complex [(R)? (+)? MTPA]4 (MTPA‐H = Mosher's acid) was added, and the 1H NMR spectra of the resulting samples were recorded (dirhodium method). The relative intensities of 1H signals dispersed by the formation of diastereomeric adducts allow to determine the absolute configuration (AC) of the starting α‐amino acids. Binding atoms in the adducts were identified by comparing the 1H and 13C chemical shifts of the oxazolones in the absence and presence of [(R)? (+)? MTPA]4. Thereby, information about the scope and limits of this method can be extracted. A protocol how to use this method is presented. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

10.
Three different kinds of substituted chiral adamantane molecules—adamantanones, dioxolanoadamantanes and dithiolano—adamantanes—were studied in the dirhodium experiment (NMR measurement with 1:1 molar mixtures with Rh(II)2[(R)‐(+)‐MTPA]4 in CDCl3). Their different behavior in adduct formation is described, and the possibility of determining enantiomeric purities and absolute configurations is explored. Detailed inspection of one‐ and two‐dimensional NMR experiments allowed for an interpretation of steric and electronic intra‐adduct interaction showing that the phenyl groups of Rh* tend to enwrap the bound adamantane ligand so that through‐space effects over a range of 6–7 Å away from the binding rhodium atom can be observed. Even slight differences in the relative orientation of phenyl groups can be monitored when comparing diastereomeric adducts via NMR signal dispersion. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

11.
Synthesis and Spectroscopic Characterization of [Rh(SeCN)6]3– and trans ‐[Rh(CN)2(SeCN)4]3–, Crystal Structure of (Me4N)3[Rh(SeCN)6] Treatment of RhCl3 with KSeCN in acetone yields a mixture of selenocyanato‐rhodates(III), from which [Rh(SeCN)6]3– and trans‐[Rh(CN)2(SeCN)4]3– have been isolated by ion exchange chromatography on diethylaminoethyl cellulose. The X‐ray structure determination on a single crystal of (Me4N)3[Rh(SeCN)6] (trigonal, space group R3, a = 14.997(2), c = 24.437(3) Å, Z = 6) reveals, that the compound crystallizes isotypically to (Me4N)3[Ir(SCN)6]. The exclusively via Se coordinated selenocyanato ligands are bonded with the average Rh–Se distance of 2.490 Å and the Rh–Se–C angle of 104.6°. In the low temperature IR and Raman spectra the metal ligand stretching modes ν(RhSe) of (n‐Bu4N)3[Rh(SeCN)6] ( 1 ) and trans‐(n‐Bu4N)3[Rh(CN)2(SeCN)4] ( 2 ) are in the range of 170–250 cm–1. In 2 νas(CRhC) is observed at 479 cm–1. The vibrational spectra are assigned by normal coordinate analysis based on the molecular parameters of the X‐ray determination. The valence force constants are fd(RhSe) = 1.08 ( 1 ), 1.10 ( 2 ) and fd(RhC) = 3.14 mdyn/Å ( 2 ). fd(RhS) = 1.32 mdyn/Å is determined for [Rh(SCN)6]3–, which has not been calculated so far. The 103Rh NMR resonances are 2287 ( 1 ), 1680 ppm ( 2 ) and the 77Se NMR resonances are –32.7 ( 1 ) and –110.7 ppm ( 2 ). The Rh–C bonding of the cyano ligand in 2 is confirmed by a dublett in the 13C NMR spectrum at 136.3 ppm.  相似文献   

12.
Seleno‐carbohydrates are those in which the oxygen of the glycosidic bond or the hydroxyl group is artificially replaced with selenium. This substitution changes 1H and 13C chemical shifts and produces spin coupling constants involving 77Se. Coupling constants, such as 2‐3J(77Se, 1H), are likely to be useful for conformational analyses of glycans because such couplings are never observed in natural glycans. Several papers have discussed the relationship between 2‐3J(77Se, 1H) and conformation; however, only few reports describe 1‐3J(77Se, 13C), which could also be useful. Here, we obtain 77Se coupling constants of seleno‐carbohydrates from 77Se‐selective HR‐HMBC and 77Se satellites in 1D 13C spectra and examine their conformations using the Newman projection scheme.  相似文献   

13.
The synthesis and characterization of selenium‐containing stannanes, (o‐MeSeC6H4CH2)Sn(Ph)3–nCln [n = 0 ( 1Se ); 1 ( 2Se ); 2 ( 3Se )], is presented. The increasing Lewis acidity at tin in the series 1Se → 2Se → 3Se is reflected in their respective solid state arrangements and supramolecular architecture by interactions of the type Se ··· Se, Sn ··· Se, and Cl ··· H–C. Overall the capacity of the selenium atom to form bidentate interactions creates geometric assemblies distinctly different to those of the oxygen and sulfur analogs.  相似文献   

14.
One‐bond spin–spin coupling constants involving selenium of seven different types, 1 J(Se,X), X = 1H, 13C, 15 N, 19 F, 29Si, 31P, and 77Se, were calculated in the series of 14 representative compounds at the SOPPA(CCSD) level taking into account relativistic corrections evaluated both at the RPA and DFT levels of theory in comparison with experiment. Relativistic corrections were found to play a major role in the calculation of 1 J(Se,X) reaching as much as almost 170% of the total value of 1 J(Se,Se) and up to 60–70% for the rest of 1 J(Se,X). Scalar relativistic effects (Darwin and mass‐velocity corrections) by far dominate over spin–orbit coupling in the total relativistic effects for all 1 J(Se,X). Taking into account relativistic corrections at both random phase approximation and density functional theory levels essentially improves the agreement of theoretical results with experiment. The most ‘relativistic’ 1 J(Se,Se) demonstrates a marked Karplus‐type dihedral angle dependence with respect to the mutual orientation of the selenium lone pairs providing a powerful tool for stereochemical analysis of selenoorganic compounds. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

15.
Cationic R2P5+ cage compounds ( 1 +) have been synthesized by the stoichiometric reaction of R2PCl, GaCl3 and P4. The reaction conditions depend on the substituent R. Alkyl‐substituted derivatives ( 1 a – 1 d [GaCl4]) are best synthesized under solvent‐free conditions, whereas aryl‐substituted derivatives ( 1 e – 1 h [GaCl4]) are formed in C6H5F. All compounds have been prepared on a multi‐gram scale in good to excellent yields and have been fully characterized with an emphasis on 31P NMR spectroscopy in solution and single‐crystal structure determination. Subsequent chalcogenation reactions of cations R2P5+ ( 1 a +, 1 e +) and trication Ph6P73+ ( 3 3+) with elemental sulfur (α‐S8) or grey selenium (Segrey) yielded a series of unique polyphosphorus–chalcogen cations ( 4 a +, 4 e +, 5 a +, 6 2+ and 7 2+), possessing nortricyclane‐type molecular structures. An in‐depth study of the 31P{1H} and 77Se NMR spectroscopic parameters is presented, and correlations between the substitution pattern and the observed structural features have been investigated in detail.  相似文献   

16.
The portfolio of acyclic diaminocarbenes (ADACs) has been substantially expanded, owing to the synthesis of eleven new formamidinium salts, mostly of the type [(iPr2N)CH(NRR′)][PF6], for use as immediate carbene precursors. The corresponding ADACs (iPr2N)C(NRR′) were sufficiently stable for isolation in the case of NRR′=2‐methylpiperidino ( 13 ), 3‐methylpiperidino ( 14 ), 4‐methylpiperidino ( 15 ), morpholino ( 17 ) and NiPrPh ( 20 ), but had to be trapped in situ in the case of NRR′=2,2,6,6‐tetramethylpiperidino ( 12 ) and NiPrMe ( 19 ). The tetraaryl‐substituted ADACs (Ph2N)2C ( 22 ) and (Ph2N)C[N(C6F5)2] ( 24 ) also could only be generated and trapped in situ. Trapping with elemental selenium was particularly efficient, affording the corresponding selenourea derivative in all cases, whereas trapping with [{Rh(μ‐Cl)(cod)}2] did not work for 12 and 24 . The 77Se NMR chemical shifts, δ(77Se), of the selenourea compounds derived from the new ADACs lie in the range 450–760 ppm, which indicates a much higher electrophilicity and π‐accepting capability of ADACs in comparison with NHCs, which typically exhibit δ(77Se)<200 ppm. The extreme low‐field shift of 758 ppm observed for 12 Se can be rationalised by the results of DFT calculations, which revealed that ADAC 12 has a minimum energy conformation with the 2,2,6,6‐tetramethylpiperidino unit perpendicular to the N2C plane, which suppresses the π donation of this amino group and causes an unusually low LUMO energy and high electrophilicity.  相似文献   

17.
Zinc(II), cadmium(II) and mercury(II) complexes of thiourea (TU) and selenourea (SeU) of general formula M(TU)2Cl2 or M(SeU)2Cl2 have been prepared. The complexes were characterized by elemental analysis and NMR (1H, 13C, 15N, 77Se and 113Cd) spectroscopy. A low-frequency shift of the C=S resonance of thiones in 13C NMR and high-frequency shifts of N–H resonances in 1H and 15N NMR are consistent with sulfur or selenium coordination to the metal ions. The Se nucleus in Cd(SeU)2Cl2 in 77Se NMR is deshielded by 87?ppm on coordination, relative to the free ligand. In comparison, the analogous Zn(II) and Hg(II) complexes show deshielding by 33 and 50?ppm, respectively, indicating that the orbital overlap of Se with Cd is better. Principal components of 77Se and 113Cd shielding tensors were determined from solid-state NMR data.  相似文献   

18.
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   

19.
The irreversible inhibition of δ‐chymotrypsin with the enantiomerically pure, P(3)‐axially and P(3)‐equatorially X‐substituted cis‐ and trans‐configurated 2,4‐dioxa‐3‐phospha(1,5,5‐2H3)bicyclo[4.4.0]decane 3‐oxides (X=F, 2,4‐dinitrophenoxy) was monitored by 31P‐NMR spectroscopy. 1H‐Correlated 31P{2H}‐NMR spectra enabled the direct observation of the vicinal coupling (3J) between the P‐atom of the inhibitor and the CH2O moiety of Ser195 (=‘Ser195’(CH2O)), thus establishing the covalent nature of the ‘Ser195’(CH2O? P) bond in the inhibited enzyme. The stereochemical course of the phosphorylation is dependent on the structure of the inhibitor, and neat inversion, both inversion and retention, as well as neat retention of the configuration at the P‐atom was found.  相似文献   

20.
Stereochemical structure of nine Z‐2‐(vinylsulfanyl)ethenylselanyl organyl sulfides has been investigated by means of experimental measurements and second‐order polarization propagator approach calculations of their 1H–1H, 13C–1H, and 77Se–1H spin–spin coupling constants together with a theoretical conformational analysis performed at the MP2/6‐311G** level. All nine compounds were shown to adopt the preferable skewed s‐cis conformation of their terminal vinylsulfanyl group, whereas the favorable rotational conformations with respect to the internal rotations around the C–S and C–Se bonds of the internal ethenyl group are both skewed s‐trans. Stereochemical trends of 77Se–1H spin–spin coupling constants originating in the geometry of their coupling pathways and the selenium lone pair effect were rationalized in terms of the natural J‐coupling analysis within the framework of the natural bond orbital approach. Copyright © 2012 John Wiley & Sons, Ltd.  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