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 15 毫秒
1.
Abstract

The synthesis of a series of 2-(3,5-dimethylpyrazol-1-yl)phenyl-based organoselenium compounds, (dmpzC6H4Se)2 (1), dmpzC6H4SeR (dmpz = 3,5-dimethylpyrazol-1-yl; R = (CH2)nY; Y = OH, NH2, and COOH), and dmpzC6H4SeX (X = Cl, Br, or I) is described. The compounds are characterized by IR, NMR (1H, 13C{1H}, 77Se{1H}), and mass spectral (MS) data. The molecular structures of (dmpzC6H4Se)2, dmpzC6H4SeCH2COOH, and dmpzC6H4SeCH2CH2OH have been established by X-ray crystallography. The two latter compounds are associated in the solid state through intermolecular hydrogen bonding between the OH proton and the pyrazolyl nitrogen atom of the adjacent molecule. Glutathione peroxidase (GPx) like catalytic activity of these compounds has been evaluated by using hydrogen peroxide (H2O2) as substrate and dithiothreitol (DTTred) as thiol cofactor in CD3OD, and the progress of the reaction was monitored by 1H NMR spectroscopy. All the compounds exhibited the GPx-like catalytic activity. Among these, the ones containing alkylamino groups showed the best activity.

Supplemental materials are available for this article. Go to the publisher's online edition of Phosphorus, Sulfur, and Silicon and the Related Elements to view the free supplemental file.  相似文献   

2.
The structures and solid-state stereochemistry of the two P-epimeric compounds ( 1 and 2 ) formed in a reaction of butyl phenylvinylphosphinite with (-)-L-menthyl bromoacetate were studied by X-ray crystallography and CP–MAS 13C NMR spectroscopy. The two compounds were also analyzed in solution by means of 2D 1H NMR, 13C NMR, IR, UV, and dipole moment measurement techniques. Compound 1 , C20H29O3P, crystallized in the orthorhombic group P 212121 with Z = 4, a = 20.491(2), b = 16.719(1), and c = 5.910(1). Compound 2 , C20H29O3P, crystallized in the monoclinic space group P 21 with Z = 2, a = 9.266(1), b = 9.852(1), c = 10.954(1), and β = 95.20(1)°. In the crystals both compounds possess their CC PO fragments in an s-cis array, and have their PO and CO dipoles oriented uniformly in a syn, nearly parallel, fashion. In solution, however, an anti arrangement of these two dipoles is slightly preferred.  相似文献   

3.
Some natural and semisynthetic carotenoids were examined by means of FT‐IR spectroscopy. The IR bands of the characteristic functional groups (CH3, CH2, CC, CO, OH, etc.) were assigned when possible. Some special functional groups – without H‐atoms – such as CCC, ‘cross epoxides', etc., which cannot be easily identified by 1H‐NMR methods, were also detected in the FT‐IR spectra.  相似文献   

4.
The 77Se NMR spectra of selenate were studied under various circumstances, such as concentration, pH, temperature, ionic strength, and D2O:H2O ratio, in order to examine its potential as a water-soluble internal chemical shift standard. The performance of selenate as a chemical shift reference and that of other attempted ones from the literature (dimethyl selenide, tetramethylsilane/TMS, and 3-(trimethylsilyl)propane-1-sulfonate/DSS) was also explored. The uncertainty in the resulting chemical shift relative to the effective spectral width is comparable to that of DSS. Compared to the currently prevalent water-soluble external chemical shift reference, selenic acid solution, the properties of internal selenate are much more favorable in terms of ease of use. We have also demonstrated that selenate can be used in reducing media, which is inevitable for the analysis of selenol compounds. Thus, it can be stated that sodium selenate is a robust internal chemical shift reference in aqueous media for 77Se NMR measurements; the chemical shift of this reference in a solution containing 5 V/V% D2O at 25°C and 0.15 mol·dm−3 ionic strength is 1048.65 ppm relative to 60 V/V% dimethyl selenide in CDCl3 and 1046.40 ppm relative to the 1H signal of 0.03 V/V% TMS in CDCl3. In summary, a water-soluble, selenium-containing internal chemical shift reference compound was introduced for 77Se NMR measurements for the first time in the literature, and with the aforementioned results all previous 77Se measurements can be converted to a unified scale defined by the International Union of Pure and Applied Chemistry.  相似文献   

5.
Functional pairing between cellular glycoconjugates and tissue lectins like galectins has wide (patho)physiological significance. Their study is facilitated by nonhydrolysable derivatives of the natural O-glycans, such as S- and Se-glycosides. The latter enable extensive analyses by specific 77Se NMR spectroscopy, but still remain underexplored. By using the example of selenodigalactoside (SeDG) and the human galectin-1 and -3, we have evaluated diverse 77Se NMR detection methods and propose selective 1H,77Se heteronuclear Hartmann–Hahn transfer for efficient use in competitive NMR screening against a selenoglycoside spy ligand. By fluorescence anisotropy, circular dichroism, and isothermal titration calorimetry (ITC), we show that the affinity and thermodynamics of SeDG binding by galectins are similar to thiodigalactoside (TDG) and N-acetyllactosamine (LacNAc), confirming that Se substitution has no major impact. ITC data in D2O versus H2O are similar for TDG and LacNAc binding by both galectins, but a solvent effect, indicating solvent rearrangement at the binding site, is hinted at for SeDG and clearly observed for LacNAc dimers with extended chain length.  相似文献   

6.
A systematic investigation of thirty-four CF3Se(II, IV) and eight CF3Te(II, IV) compounds by 13C, 19F, 77Se and 125Te NMR spectroscopy resulted in some general features for chemical shifts and coupling constants which agree with the trends of reported 19F and new 13C NMR data of CF3S(II, IV) compounds. Moreover, the NMR spectra of molecules of the type E=CXY (E = chalcogen, X, Y = halogen) and substances containing a CSe double bond have been studied. From the comparison of these NMR data with those of CF3 substituted chalcogen compounds, a partial double bond character of the carbon-fluorine and carbon-chalcogen bond in CF3 substituted chalcogen compounds can be derived:
  相似文献   

7.
Abstract

A series of α-substituted selenenyl acetophenone derivatives of the types, [PhC(OCH2CH2O)CH2Se]2, [PhC(OCH2CH2O)CH2SeR], (PhCOCH2Se)2, and [PhCOCH2SeR] have been prepared. These compounds have been characterized by elemental analyses, IR and NMR (1H, 13C, 77Se) spectroscopy. The compounds, [PhC(OCH2CH2)CH2Se]2 and (PhCOCH2Se)2 have been structurally characterized by single crystal X-ray diffraction analyses. The former shows intra-molecular Se‐?‐?‐O interaction, while the latter exhibits inter-molecular nonbonding Se‐?‐?‐O interaction.

[Supplementary materials are available for this article. Go to the publisher's online edition of Phosphorus, Sulfur, and Silicon and the Related Elements for the following free supplemental files: Additional text and figures.]  相似文献   

8.

Weak nonbonded interaction between a divalent selenium and an oxygen atom (i.e., Se···O interaction) frequently plays important roles in chemical and biological functions of selenium compounds. To establish that 77 Se NMR is an easy experimental probe to diagnose the strength of an Se···O interaction, 3 series of 2-substituted benzeneselenenyl derivatives, which have an intramolecular Se···O interaction in solution, were employed. By comparing the 77 Se NMR chemical shifts (δ Se ) with those observed for other series of selenium compounds, which have an intramolecular Se···Y (Y = N, O, F, Cl, or Br) interaction, approximate linear correlation was found between the δ Se values and the strengths of the nonbonded Se···Y interactions evaluated by natural bond orbital analysis at the B3LYP level. The correlation will be useful for estimating the strength of an Se···O interaction simply from the 77 Se NMR chemical shift. By extending the chemistry of nonbonded Se···O interactions to structural biology, analogous S···O interactions have been discovered in protein architecture. The directional features were, however, different from those of Se···O and S···O interactions of small organic compounds.  相似文献   

9.
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   

10.
The planarity of acetamides 1a–3a , thioacetamides 4a–6a , and selenoacetamides 7a–9a , R1R2NC(=E)CH3 where E = O, S, Se, and R1, R2 = H or CH3, was investigated using theoretical calculations at the density functional theory (DFT) level. The calculations showed that the methyl substitution on nitrogen and the change from the amide moiety (NCO) to NCS or NCSe group increased the double bond character of the N C bond. In other words, the planarity of these compounds ( 1a–9a ) increases in the order NH2 < NHCH3 < N(CH3)2 and O < S < Se. The calculations of bending energy suggest that the planar geometry represents the lowest energy conformation for all compounds investigated in this work. N,N‐Dimethyl‐selenoacetamide ( 9a ), (CH3)2NC(Se) CH3, has the largest bending energy of 10.37 kcal/mol, which suggests that it possesses the greatest planarity among the compounds 1a–9a . However, the solid phase molecular structure of 9a was found to be slightly nonplanar by X‐ray crystallography. The slight nonplanarity observed experimentally is very likely the consequence of intermolecular interactions arising within the crystal packing. © 2002 Wiley Periodicals, Inc. Heteroatom Chem 13:380–386, 2002; Published online in Wiley Interscience (www.interscience.wiley.com). DOI 10.1002/hc.10056  相似文献   

11.
Polymeric Si/C/O/N xerogels, with the idealized polymer network structure comprising [Si O Si(NCN)3]n moieties, were prepared by reactions of hexachlorodisiloxane (Cl3Si O SiCl3) with bis(trimethylsilyl)carbodiimide (Me3Si NCN SiMe3, BTSC). NMR and FTIR spectra indicate the existence of ‐NCN‐ and Si O Si‐ units in the xerogels and also in the ceramic materials obtained upon pyrolysis. The feasibility of this reaction protocol was confirmed on the molecular level by the deliberate synthesis of the macrocyclic compound [SiPh2 O SiPh2(NCN)]2, the crystal structure and spectroscopic data of which are reported. The influence of pyridine as a catalyst for the cross‐linking reaction was studied. The degree of cross‐linking increased within the polymers with the addition of pyridine. It was shown by the reaction of hexachlorodisiloxane with excess pyridine that the latter appears to activate only one out of the two ‐SiCl3 moieties under formation of hexacoordinated silicon compounds. The crystal structure of Cl3Si O SiCl3(pyridine)2 is presented. Quantum chemical calculations are in support of this adduct being a potential intermediate in the pyridine catalyzed sol–gel process. The ceramic yield after pyrolysis of the Si/C/O/N‐xerogels at 1000 °C, which reaches values up to 50%, was found to depend on the aging protocol (time, temperature), whereas no correlation was found with the amount of pyridine added for xerogel synthesis. The Si/C/N/O‐ceramics obtained after pyrolysis at 1000 °C under NH3 are completely amorphous. Chemically they have to be considered as hybrids between an ideal [SiOSi(NCN)3]n network and glass‐like Si2N2O. The products are mesoporous with closed pores and a broad pore size distribution. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

12.
New bis(thio)substituted, S‐,O‐substituted, and S‐,S‐substituted benzoquinone compounds were synthesized from the reaction of p‐chloranil ( 1 ) with S‐,O‐substituted thiols, dithiols, and monothiols. The 13C NMR spectra and the IR spectra of heterocyclic compounds 3 , 4 and 7 , 8 showed different behavior; that of 3 , 7 showed a carbon signal and a >CO group band for the carbonyl group and that of 4 , 8 showed two carbon signals and split bands for the carbonyl group. The structures of the novel compounds were characterized by microanalysis, FT‐IR, 1H NMR, 13C NMR, MS, and UV–vis spectroscopy. The crystal structure of 2,3,5,6‐tetrakis(4‐fluorobenzylthio)cyclohexa‐2,5‐diene‐1,4‐dione ( 15 ) was determined by the X‐ray diffraction method. © 2010 Wiley Periodicals, Inc. Heteroatom Chem 21:446–452, 2010; View this article online at wileyonlinelibrary.com . DOI 10.1002/hc.20634  相似文献   

13.
Three sterically crowded peri‐substituted naphthalene phosphines, Nap[PPh2][ER] (Nap=naphthalene‐1,8‐diyl; ER=SEt, SPh, SePh) 1–3 , which contain phosphorus and chalcogen functional groups at the peri positions have been prepared. Each phosphine reacts to form a complete series of PV chalcogenides Nap[P(E′)(Ph2)(ER)] (E′=O, S, Se). The novel compounds were fully characterised by using X‐ray crystallography and multinuclear NMR spectroscopy, IR spectroscopy and MS. X‐ray data for 1 , 2 , n O , n S , n Se (n=1–3) are compared. Eleven molecular structures have been analysed by naphthalene ring torsions, peri‐atom displacement, splay angle magnitude, X???E interactions, aromatic ring orientations and quasi‐linear arrangements. An increase in the congestion of the peri region following the introduction of heavy chalcogen atoms is accompanied by a general increase in naphthalene distortion. P???E distances increase for molecules that contain bulkier atoms at the peri positions and also when larger chalcogen atoms are bound to phosphorus. The chalcogenides adopt similar conformations that contain a quasi‐linear E???P? C fragment, except for 3 O , which displays a twist‐axial‐twist conformation resulting in the formation of a linear O???Se? C alignment. Ab initio MO calculations performed on 2 O , 3 O , 3 S and 3 Se reveal Wiberg bond index values of 0.02 to 0.04, which indicates only minor non‐bonded interactions; however, calculations on radical cations of 3 O , 3 S and 3 Se reveal increased values (0.14–0.19).  相似文献   

14.
Organocalcium compounds have been reported as efficient catalysts for various alkene transformations. In contrast to transition metal catalysis, the alkenes are not activated by metal–alkene orbital interactions. Instead it is proposed that alkene activation proceeds through an electrostatic interaction with a Lewis acidic Ca2+. The role of the metal was evaluated by a study using the metal‐free catalysts: [Ph2N][Me4N+] and [Ph3C][Me4N+]. These “naked” amides and carbanions can act as catalysts in the conversion of activated double bonds (CO and CN) in the hydroamination of Ar NCO and R NCN R (R=alkyl) by Ph2NH. For the intramolecular hydroamination of unactivated CC bonds in H2CCHCH2CPh2CH2NH2 the presence of a metal cation is crucial. A new type of hybrid catalyst consisting of a strong organic Schwesinger base and a simple metal salt can act as catalyst for the intramolecular alkene hydroamination. The influence of the cation in catalysis is further evaluated by a DFT study.  相似文献   

15.
We report a method for the screening of interactions between proteins and selenium‐labeled carbohydrate ligands. SEAL by NMR is demonstrated with selenoglycosides binding to lectins where the selenium nucleus serves as an NMR‐active handle and reports on binding through 77Se NMR spectroscopy. In terms of overall sensitivity, this nucleus is comparable to 13C NMR, while the NMR spectral width is ten times larger, yielding little overlap in 77Se NMR spectroscopy, even for similar compounds. The studied ligands are singly selenated bioisosteres of methyl glycosides for which straightforward preparation methods are at hand and libraries can readily be generated. The strength of the approach lies in its simplicity, sensitivity to binding events, the tolerance to additives and the possibility of having several ligands in the assay. This study extends the increasing potential of selenium in structure biology and medicinal chemistry. We anticipate that SEAL by NMR will be a beneficial tool for the development of selenium‐based bioactive compounds, such as glycomimetic drug candidates.  相似文献   

16.
Summary: In this paper, films were prepared from soy protein and corn starch in different proportions and thermal stability and kinetic parameters were determined through degradation reactions in an inert atmosphere. Solid residues and decomposition products were identified by infrared spectroscopy. Films from corn starch were less thermally stable than soy protein films. The films containing both components had lower thermal stabilities when compared to those of the pure biopolymers. The mechanism of starch thermal degradation seems to occur in a single step, which can be confirmed by the constant E-values during the thermal degradation reaction. For the pure protein and its mixtures an increase in the activation energy was observed during the reaction. Solid residues for protein at different temperatures showed mainly bands related to CO stretching, angular deformation of N H and C H groups. For starch, absorptions related to free and bound O H, CO stretching of CO2, CO and carbonyl compounds were observed. For the 50/50 mixture bands related to soy protein and corn starch were observed. The gaseous products for soy protein showed absorptions related to CO2, CO, CO, NH3 and C H stretching. For pure starch absorptions related to O H stretching from alcohol, CO from CO2, CO and carbonyl compounds. The 50/50 mixture had the same characteristics as pure soy protein and corn starch.  相似文献   

17.
A new approach to determination of the stereochemical structure of bis-selenium-substituted alkenes using experimental 77Se NMR studies and B3LYP/6-311G(d) quantum-chemical calculations is developed. Joint analysis of experimental and calculated data allows assignment of signals in the 77Se NMR spectrum. The method was evaluated taking the model compounds (PhSe)HC=C(SePh)R (R = COOMe, CH2NMe2, CH2OH, Ph) as examples.  相似文献   

18.
The reactions of Cl3PN P(O)Cl2 ( 1 ) with primary and secondary amines have been studied. The following monophosphazenes, (RRN)3PN P(O)(NRR)2, and bis(phosphinoyl)amines, [(RRN)2P(O)]2NH were isolated: NRR = NHCH2Ph, Net2, NH(CH2)2CH3 groups for monophosphazenes, and Net2, NH(CH2)2CH3 for phosphinoyl amines. The unexpected geminal phosphazene, Cl(RRN)2PN P(O)Cl2, {RRN = N[CH(CH3)2]2}, was also obtained in moderate yield. The spectral data (IR, 1H, 13C, and 31P NMR, and MS) are presented. The structure of 1-(dichlorophosphinyl)-2-chloro-2,2-bis(diisopropylamino)phosphazene ( 5 ) was determined by X-ray crystallography. The basicities of these and related compounds in nonaqueous nitrobenzene solution were obtained by potentiometric titration.  相似文献   

19.
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   

20.
The diorganodiselenides (pzCH2CH2)2Se2 ( 1 ) and (PhtzCH2)2Se2 ( 2 ) were prepared by reacting Na2Se2 with 1‐(2‐bromoethyl)‐1H‐pyrazole and 4‐(chloromethyl)‐2‐phenylthiazole, respectively, while the reactions between 1‐(2‐bromoethyl)‐1H‐pyrazole or 4‐(chloromethyl)‐2‐phenylthiazole and the lithium organoselenolates [2‐(Et2NCH2)C6H4]SeLi and [2‐{O(CH2CH2)2NCH2}C6H4]SeLi in a 1:1 molar ratio resulted in the heteroleptic diorganoselenium(II) compounds [2‐(Et2NCH2)C6H4](R)Se (R = pzCH2CH2 ( 3 ) or PhtzCH2 ( 5 )) and [2‐{O(CH2CH2)2NCH2}C6H4](R)Se (R = pzCH2CH2 ( 4 ) or PhtzCH2 ( 6 )). The diorganotin(IV) bis(organoselenolato) derivatives of type R2Sn(SeCH2CH2pz)2 (R = 2‐(Me2NCH2)C6H4 ( 7 ) or Me ( 8 )) were obtained by reacting (pzCH2CH2)SeNa with the appropriate diorganotin(IV)dichloride in a 2:1 molar ratio. All compounds were investigated using NMR spectroscopy (1H, 13C, 77Se, 119Sn as appropriate) and ESI+ mass spectrometry. The molecular structures of 2 and 6 were determined using single‐crystal X‐ray diffraction. The formation of a 10–Se–3 hypercoordinated species was evidenced for 6 in the solid state, as a consequence of the C,N coordination behaviour of the 2‐{O(CH2CH2)2NCH2}C6H4 group. Compounds 1 , 7 and 8 were investigated for their antiproliferative activity towards the mouse colon carcinoma C26 cell line with the preliminary results showing a better activity than 5‐fluorouracil.  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