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.
all-L -β3-Penta-, hexa-, and heptapeptides with the proteinogenic side chains of valine, leucine, serine, cysteine, and methionine have been prepared by previously described procedures ( 12 , 13 , 14 , 15 ; Schemes 2 – 5). Thioether cleavage with Na/NH3 in β-HMet residues has also provided a β3-hexapeptide with homocysteine (CH2CH2S) side chains ( 13e ). The HS−(CH2)n groups were positioned on the β-peptidic backbone in such a way that, upon disulfide-bridge formation, the corresponding β-peptide was expected to maintain either a 31-helical secondary structure ( 1 , 2 ) (Fig. 1) or to be forced to adopt another conformation ( 3 , 4 ). The 13-, 17-, 19-, and 21-membered-ring macrocyclic disulfide derivatives and their open-chain precursors, as well as all synthetic intermediates, were purified (crystallization, flash or preparative HPL chromatography; Fig. 5) and fully characterized (m.p., [α]D, CD, IR, NMR, FAB or ESI mass spectroscopy, and elemental analysis, whenever possible; Fig. 2 and Exper. Part). The structures in MeOH and H2O of the new β-peptides were studied by CD spectroscopy (Figs. 3 and 4), where the characteristic 215-nm-trough/200-nm-peak pattern was used as an indicator for the presence or absence of (M)-31-helical conformations. A CH2−S2−CH2 and, somewhat less so, a (CH2)2−S2−(CH2)2 bracket between residues i and i+3 ( 1 vs. 12d , and 2 vs. 13e in Fig. 3) give rise to CD spectra which are compatible with the presence of 31-helical structures, while CH2−S2−CH2 brackets between residues i and i+2 ( 3 vs. 14c ) or i and i+4 ( 4 vs. 15c in Fig. 4) do not.  相似文献   

2.
《Chemical physics letters》1985,116(5):371-373
Formation of NH(A 3Πi) has been studied in the photolysis of HN3 at 121.6 nm. Measurements of time-resolved fluorescence intensity show that NH(A 3Πi) is largely formed by secondary reactions involving two energetic intermediates which have lifetimes of 2.7 and 18 μs. These intermediates are tentatively assigned to N2(B 3Πg, υ′) and N2(B′3Σu), respectively.  相似文献   

3.
A series of novel bischelate bridging ligands, CH3NH(CH2)2N(CH3)(CH2) n N(CH3)(CH2)2NHCH3 (n = 9, 10, 11, and 12) were synthesized as hydrochloride salts and characterized by elemental analyses, electrospray mass spectrometry, and 1H and 13C NMR spectroscopy. These ligands form [2]pseudorotaxanes with α-cyclodextrin (α-CD) and the stability constants have been determined from 1H NMR titrations in D2O. The kinetics and mechanism of the assembly and dissociation of a [2]pseudorotaxane in which α-CD has been threaded by the CH3NH2(CH2)2N(CH3)(CH2)12N(CH3)(CH2)2NH2CH 3 2+ ligand were determined in aqueous solution using 1H NMR spectroscopy. A weak inclusion of the dimethylethylenediamine end group precedes the passage of the α-CD onto the hydrophobic dodecamethylene chain.  相似文献   

4.
《Polyhedron》2002,21(25-26):2639-2645
Unsymmetrical diphosphine ligands of the type Ph2P(CH2)nNHPPri 2 [n=2 (1), 3 (2)] have been obtained by reacting the appropriate (diphenylphosphino)alkylamine, Ph2P(CH2)nNH2 with chlorodi-iso-propylphosphine, in the presence of triethylamine. Reaction of Ph2P(CH2)2NHPPri 2 with PdCl2(PhCN)2, PtCl2(PhCN)2, PtMe2(cod) and PtClMe(cod), NiCl2·6H2O and Fe(CO)25-C5H5)I gives the corresponding chelate complexes, PdCl2L, PtX2L, NiCl2L and Fe(CO)(η5-C5H5)L. Reaction of Ph2P(CH2)3NHPPri 2 with PtCl2(PhCN)2, PtMe2(cod) and PdCl2(PhCN)2 yields the chelate complexes and reaction with PtClMe(cod) led to a 50:50 mixture of chelate isomers.  相似文献   

5.
《Tetrahedron letters》2014,55(50):6915-6918
We prepared β-amino-α-trifluoromethyl-α-amino acids through ring-opening reaction of N-tosyl-2-trifluoromethyl-2-ethoxycarbonylaziridine with aromatic and benzylic amines, and investigated the intramolecular interaction between the trifluoromethyl (CF3) group at the α-position and the NH group at the β-position (NHβ). NMR, UV/vis, and circular dichroism measurements indicated that the conformation of these compounds is fixed by intramolecular interaction of CF3 with NHβ to form a six-membered ring.  相似文献   

6.
The relationships enthalpy of mixing and excess Gibbs energyvs. composition were studied. We report hereH E andG E for 2-CH3-c-C5H4N (α-picoline)+ (1?x) CH3CH(OH)CH3, (1?x) CH3CHCH3CH2OH, (1 ?x) CH3CH2(OH)CHCH3 or (1 ?x) CH3C(CH3) (OH)CH3  相似文献   

7.
Specific ion/molecule reactions are demonstrated that distinguish the structures of the following isomeric organosilylenium ions: Si(CH3) 3 + and SiH(CH3)(C2H5)+; Si(CH3)2(C2H5)+ and SiH(C2H5) 2 + ; and Si(CH3)2(i?C3H7)+, Si(CH3)2(n?C3H7)+, Si(CH3)(C2H5) 2 + , and Si(CH3)3(π?C2H4)+. Both methanol and isotopically labeled ethene yield structure-specific reactions with these ions. Methanol reacts with alkylsilylenium ions by competitive elimination of a corresponding alkane or dehydrogenation and yields a methoxysilylenium ion. Isotopically labeled ethene reacts specifically with alkylsilylenium ions containing a two-carbon or larger alkyl substituent by displacement of the corresponding olefin and yields an ethylsilylenium ion. Methanol reactions were found to be efficient for all systems, whereas isotopically labeled ethene reaction efficiencies were quite variable, with dialkylsilylenium ions reacting rapidly and trialkylsilylenium ions reacting much more slowly. Mechanisms for these reactions and differences in the kinetics are discussed.  相似文献   

8.
The electronic structures of oxalate bridged oligomers of MM quadruply bonded complexes (M?=?Mo or W) have been calculated by density functional theory. The calculations predict the formation of a ?? band width ~1.0?eV. With increasing number of oxalate bridges, the metal-to-ligand (oxalate) charge transfer band moves to lower energy. Two synthetic approaches have been investigated. (1) The direct reaction between oxalic acid and the homoleptic compound M2(T i PB)4, where T i PB?=?2,4,6-triisopropylbenzoate and (2) the metathetic reaction between the herein reported salt Mo2(T i PB)2(CH3CN)4(BF4)2 and ( n Bu4N+)2(oxalate2?). Neither reaction yielded the desired oligomers though the latter resulted in the formation of the molecular triangle [Mo2(T i PB)2(O2CCO2)]3.  相似文献   

9.
New rhodium(I)- and rhodium(III)-β-diketonato complexes of the type [Rh(FcCOCHCOR)(P(OPh)3)2] and [Rh(FcCOCHCOR)(P(OPh)3)2(CH3)(I)], with Fc = ferrocenyl and R = Fc, CH3 and CF3, have been synthesized. The reactivity of complexes of the type [Rh(β-diketonato)(P(OPh)3)2] increase in the order: β-diketonato = (CF3COCHCOCF3) < (CF3COCHCOPh) < (CF3COCHCOCH3) < (PhCOCHCOPh) < (CF3COCHCOFc) < (CH3COCHCOPh) < (CH3COCHCOCH3) < (CH3COCHCOFc) < (FcCOCHCOFc), giving linear relationships between the kinetic parameter ln k2 and the parameters that are related to the electron density on the rhodium centre; the sum of the group electronegativities of the β-diketonato side groups (χR + χR′) and the pKa of the uncoordinated β-diketone RCOCH2COR′. The large negative values of the volume and entropy of activation indicated a mechanism which occurs via a polar transition state. A density functional theory study, at the PW91/TZP level of theory, indicates that oxidative addition of iodo methane to [Rh(FcCOCHCOCF3)(P(OCH3)3)2] occurs via a two-step mechanism. This mechanism involves a nucleophilic attack by the metal on the methyl carbon to displace iodide to form a metal-carbon bond and the coordination of iodide to the five-coordinated intermediate to give a six-coordinated trans alkyl product.  相似文献   

10.
Silica from leached chrysotile fibers (SILO) was silanized with trialkoxyaminosilanes to yield inorganic–organic hybrids designated SILx (x=1–3). The greatest amounts of the immobilized agents were quantified as 2.14, 1.90, and 2.18 mmol g−1 on SIL1, SIL2, and SIL3 for –(CH2)3NH2,–(CH2)3NH(CH2)2NH2, and –(CH2)3NH(CH2)2NH(CH2)2NH2 groups attached to the inorganic support. The infrared spectra for all modified silicas showed the absence of the Si–OH deformation mode, originally found at 950 cm−1, and the appearance of asymmetric and symmetric C–H stretching bands at 2950 and 2840 cm−1. Other important bands associated with the organic moieties were assigned to νas(NH) at 3478 and νsym(NH) at 3418 cm−1. The NMR spectrum of the solid precursor material suggested two different kinds of silicon atoms: silanol and siloxane groups, between −90 and 110 ppm; however, additional species of silicon that contain the organic moieties bonded to silicon at −58 and −66 ppm appeared after chemical modification. These modified silicas showed a high adsorption capacity for cobalt and copper cations in aqueous solution, in contrast to the original SILO matrix, confirming the unequivocal anchoring of silylating agents on the silica surface.  相似文献   

11.
Synthesis and Characterization of New Intramolecularly Nitrogen‐stabilized Organoaluminium‐ and Organogallium Alkoxides The intramolecularly nitrogen stabilized organoaluminium alkoxides [Me2Al{μ‐O(CH2)3NMe2}]2 ( 1a ), Me2AlOC6H2(CH2NMe2)3‐2,4,6 ( 2a ), [(S)‐Me2Al{μ‐OCH2CH(i‐Pr)NH‐i‐Pr}]2 ( 3a ) and [(S)‐Me2Al{μ‐OCH2CH(i‐Pr)NHCH2Ph}]2 ( 4 ) are formed by reacting equimolar amounts of AlMe3 and Me2N(CH2)3OH, C6H2[(CH2NMe2)3‐2,4,6]OH, (S)‐i‐PrNHCH(i‐Pr)CH2OH, or (S)‐PhCH2NHCH(i‐Pr)CH2OH, respectively. An excess of AlMe3 reacts with Me2N(CH2)2OH, Me2N(CH2)3OH, C6H2[(CH2NMe2)3‐2,4,6]OH, and (S)‐i‐PrNHCH(i‐Pr)CH2OH producing the “pick‐a‐back” complexes [Me2AlO(CH2)2NMe2](AlMe3) ( 5 ), [Me2AlO(CH2)3NMe2](AlMe3) ( 1b ), [Me2AlOC6H2(CH2NMe2)3‐2,4,6](AlMe3)2 ( 2b ), and [(S)‐Me2AlOCH2CH(i‐Pr)NH‐i‐Pr](AlMe3) ( 3b ), respectively. The mixed alkyl‐ or alkenylchloroaluminium alkoxides [Me(Cl)Al{μ‐O(CH2)2NMe2}]2 ( 6 ) and [{CH2=C(CH3)}(Cl)Al{μ‐O(CH2)2NMe2}]2 ( 8 ) are to obtain from Me2AlCl and Me2N(CH2)2OH and from [Cl2Al{μ‐O(CH2)2NMe2}]2 ( 7 ) and CH2=C(CH3)MgBr, respectively. The analogous dimethylgallium alkoxides [Me2Ga{μ‐O(CH2)3NMe2}]2 ( 9 ), [(S)‐Me2Ga{μ‐OCH2CH(i‐Pr)NH‐i‐Pr}]n ( 10 ), [(S)‐Me2Ga{μ‐OCH2CH(i‐Pr)NHCH2Ph}]n ( 11 ), [(S)‐Me2Ga{μ‐OCH2CH(i‐Pr)N(Me)CH2Ph}]n ( 12 ) and [(S)‐Me2Ga{μ‐OCH2(C4H7NHCH2Ph)}]n ( 13 ) result from the equimolar reactions of GaMe3 with the corresponding alcohols. The new compounds were characterized by elemental analyses, 1H‐, 13C‐ and 27Al‐NMR spectroscopy, and mass spectrometry. Additionally, the structures of 1a , 1b , 2a , 2b , 3a , 5 , 6 and 8 were determined by single crystal X‐ray diffraction.  相似文献   

12.
The chemical kinetics, studied by UV/Vis, IR and NMR, of the oxidative addition of iodomethane to [Rh((C6H5)COCHCOR)(CO)(PPh3)], with R = (CH2)nCH3, n = 1-3, consists of three consecutive reaction steps that involves isomers of two distinctly different classes of RhIII-alkyl and two distinctly different classes of RhIII-acyl species. Kinetic studies on the first oxidative addition step of [Rh((C6H5)COCHCOR)(CO)(PPh3)] + CH3I to form [Rh((C6H5)COCHCOR)(CH3)(CO)(PPh3)(I)] revealed a second order oxidative addition rate constant approximately 500-600 times faster than that observed for the Monsanto catalyst [Rh(CO)2I2]. The reaction rate of the first oxidative addition step in chloroform was not influenced by the increasing alkyl chain length of the R group on the β-diketonato ligand: k1 = 0.0333 ([Rh((C6H5)COCHCO(CH2CH3))(CO)(PPh3)]), 0.0437 ([Rh((C6H5)COCHCO(CH2CH2CH3))(CO)(PPh3)]) and 0.0354 dmmol−1 s−1 ([Rh((C6H5)COCHCO(CH2CH2CH2CH3))(CO)(PPh3)]). The pKa and keto-enol equilibrium constant, Kc, of the β-diketones (C6H5)COCH2COR, along with apparent group electronegativities, χR of the R group of the β-diketones (C6H5)COCH2COR, give a measurement of the electron donating character of the coordinating β-diketonato ligand: (R, pKa, Kc, χR) = (CH3, 8.70, 12.1, 2.34), (CH2CH3, 9.33, 8.2, 2.31), (CH2CH2CH3, 9.23, 11.5, 2.41) and (CH2CH2CH2CH3, 9.33, 11.6, 2.22).  相似文献   

13.
Herein we report on the Si grafting of two Fe4 derivatives, [Fe4(Li)2(tmhd)6], in which tmhd is 2,2,6,6-tetramethylheptane-3,5-dionate and H3Li = R–C(CH2OH)3 is a tripodal ligand with R = CH2CH–CH2–O–CH2 (H3L1) and CH2CH–(CH2)9–O–CH2 (H3L2). These complexes were specifically designed to be directly anchored on the H-terminated silicon surface via the hydrosilylation reaction. The complexes were grafted by a one pot route based on the photoinduced hydrosilylation followed by a ligand exchange step in the same reaction solution. The resulting decorated surfaces were characterized using X-ray photoelectron spectroscopy (XPS), attenuated total reflection infrared spectroscopy (ATR-IR) and atomic force microscopy (AFM).  相似文献   

14.
Infrared spectra of creatinine (H3CNC(NH)NHCOCH2) (creat), cis-Pt(creat)2(NO2)2 and Pt(creat)4(CIO4)2 have been recorded in the range 50–4000 cm−1. The fundamental vibrations for the creatinine molecule were assigned by normal coordinate analysis in the generalized valence force field approximation. The spectrum of cis-Pt(creat)2(NO2)2 was interpreted by comparison with the creatinine vibrational modes. Additionally the Pt(creat)4(ClO4)2 infrared spectrum has been involved to help the assignment.  相似文献   

15.
《Tetrahedron》1986,42(22):6253-6262
Thermal electron attachment to EtO2CCHN2 produced the parent anion radical EtO2CCHN2-(m/z 114). The ion-molecule reactions of m/z 114 with 30 neutral substrates were examined. Using the bracketing method, PA(EtO2CCHN2XXX) = 355±4 kcal mol-1 was determined. The reaction of m/z 4 with CH3SH produced H2CSXXX, the product of β-elimination of H+2 from the thiol. From a series of bracketing studies and determination of the equilibrium constant for the electron transfer (ET) process CH3C(O)C(O)CH3-+EtO2CCHN2 = CH3C(O)C(O)CH3+EtO2CCHN2XXX EA(EtO2CCHN2) = 19.7 kcal mol-1 was derived. Both associative and dissociative ET were observed in the reactions of m/z 114 with certain perhalomethanes (depending on their EA) as well as halogen-atom abstraction from BrCCl3. While no reaction was observed between m/z 114 with CH3CHO or (CH3)2CO, m/z 114 reacted with certain other ketones and esters (CF3CO2R) mainly to yield enolate anions of the β-keto esters, EtO2CCHC(O-)R. These enolate anions are believed to be formed by nucleophilic addition of Cα of m/z 114 to the carbonyl group of the neutral substrate followed by loss of N2 and radical β-fragmentation.  相似文献   

16.
The reaction of [IrCl(dmso)3] with trisphosphinomethylborato ligand Li(THF){PhB(CH2PiPr2)3} at room temperature results in intramolecular C-H activation of one of the iPr substituents affording two diastereomers of cyclometalated iridium(III) complex [Ir(H)(dmso){PhB(CH2PiPr2)2(CH2PiPrCHMeCH2)}] (1) in high yield in approximately equimolar ratio. NMR spectroscopic characterization indicates that only the diastereomers with the hydride ligand in cis position with respect to the metalacyclic phosphorous atom are formed as confirmed by single crystal X-ray diffraction. Facile ring opening with H2 at room temperature gives dihydride [Ir(H)2(dmso){PhB(CH2PiPr2)3}] (2). However, C-H activation of benzene was not observed.  相似文献   

17.
The 13C-NMR spectra of the synthetic membrane modifying nonadecapeptide Boc-(Aib-l-Ala)5-Gly-Ala-Aib-Pro-Ala- Aib-Aib-Glu(OBz)-Gln-OMe (Aib = α-aminoisobutyric acid), and of synthetic intermediates were used for conformational analysis in solution. The assignments of the 13C-NMR signals of Aib are based on the magnetic nonequivalence (MNE) of the geminal Cβ-signals in asymmetric environment resulting in a shift difference of 0.2–0.5 ppm due to neighbouring chiral residues. More than 4 ppm MNE are observed due to α-helical conformation and about 2.5 ppm for Aib situated in the corners of a rigid β-turn. The Ala-Cα signal is also sensitive to different secondary structures. The Cα signal for C-terminal alanine is found at 49–50 ppm, and for alanine within unordered oligopeptides it absorbs at 50–51 ppm. α-Helical environment shifts the Ala-Cα signal to lower field down to 54 ppm. In methanolic solution the nonadecapeptide shows a α-helical N-terminal region. For the C-terminus beginning with proline-14 no periodically ordered conformation is observed, and we suggest a sequence of β-turns. Furthermore the typical E/Z isomerism of the prolyl-peptide bond can be observed on proline itself and on its neighbour alanine.  相似文献   

18.
Compound trans-PtBr2(C2H4)(NHEt2) (1) has been synthesized by Et2NH addition to K[PtBr3(C2H4)] and structurally characterized. Its isomer cis-PtBr2(C2H4)(NHEt2) (3) has been obtained from 1 by photolytic dissociation of ethylene, generating the dinuclear trans-[PtBr2(NHEt2)]2 intermediate (2), followed by thermal re-addition of C2H4, but only in low yields. The addition of further Et2NH to 1 in either dichloromethane or acetone yields the zwitterionic complex trans-Pt(−)Br2(NHEt2)(CH2CH2N(+)HEt2) (4) within the time of mixing in an equilibrated process, which shifts toward the product at lower temperatures (ΔH° = −6.8 ± 0.5 kcal/mol, ΔS° = 14.0 ± 2.0 e.u., from a variable temperature IR study). 1H NMR shows that free Et2NH exchanges rapidly with H-bonded amine in a 4·NHEt2 adduct, slowly with the coordinated Et2NH in 1, and not at all (on the NMR time scale) with Pt-NHEt2 or -CH2CH2N(+)HEt2 in 4. No evidence was obtained for deprotonation of 4 to yield an aminoethyl derivative trans-[PtBr2(NHEt2)(CH2CH2NEt2)] (5), except as an intermediate in the averaging of the diasteretopic methylene protons of the CH2CH2N(+)HEt2 ligand of 4 in the higher polarity acetone solvent. Computational work by DFT attributes this phenomenon to more facile ion pair dissociation of 5·Et2NH2+, obtained from 4·Et2NH, facilitating inversion at the N atom. Complex 4 is the sole observable product initially but slow decomposition occurs in both solvents, though in different ways, without observable generation of NEt3. Addition of TfOH to equilibrated solutions of 4, 1 and excess Et2NH leads to partial protonolysis to yield NEt3 but also regenerates 1 through a shift of the equilibrium via protonation of free Et2NH. The DFT calculations reveal also a more favourable coordination (stronger Pt-N bond) of Et2NH relative to PhNH2 to the PtII center, but the barriers of the nucleophilic additions of Et2NH to the C2H4 ligand in 1 and of PhNH2 to trans-PtBr2(C2H4)(PhNH2) (1a) are predicted to be essentially identical for the two systems.  相似文献   

19.
G. Meyer  P. Viout 《Tetrahedron》1977,33(15):1959-1961
The alkaline hydrolysis of p-nitrophenyl acetate and of CH3CO2(CH2)2N+(CH3)2C16H33, Br? was studied in the presence of micelles C16H33N+(CH3)2CH2CH2OH, Br? and CTAB, C16H33N+(CH3)3,Br?. A pathway involving an intermediate is suggested for the hydrolysis of the ester. Hydrolysis rate of the intermediate in the presence of micelles is the same as hydrolysis rate for the ester in the absence of micelles. Consequently, hydrolysis of p-nitrophenyl acetate is not catalysed by one type of micelle, while it is enhanced by another type of micelle.  相似文献   

20.
《Polyhedron》1988,7(6):449-462
The complexes [ML*(NO)Cl(OR)] {L* = HB(3,5-Me2C3HN2)3; M= Mo, R = CH2CH2X, X = Cl, OMe or OEt; (CH2)nOH, n = 2, 5, 6; M = W, R = CH2CH2X, X = Cl, OMe or OEt; (CH2)nOH, n = 2–6; CH2(CF2)3CH2OH; CHMeCH2CMe2OH} and [ML*(NO)(OR)2] {M = Mo, R = CH2CH2X, X = Cl, OMe or OEt; (CH2)nOH, n = 2–6; M = W,R = CH2CH2X, X= Cl, OMe or OEt; (CH2)nOH, n = 2,4–6; CH2(CF2)3CH2OH} have been prepared from [ML*(NO)Cl2] and the appropriate alcohol in the presence of NEt3 or NaCO3, and have been characterized by IR, 1H NMR and mass spectroscopy.  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