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.
Photolysis of t-BuHgCl/KI with PhC(R2)C(R1)NO2 forms PhC(R2)C(R1)Bu-t when R1 = R2 = H or in low yield when R1 = H, R2 = Ph. When R1 ≠ H, or when R2 = Ph, reactions with t-BuHgI/KI/hv proceed mainly via PhC(R2)C(R1)NO2·-, PhC(R2)C(R1)N(OBu-t)OHgX+, PhC(R2)C(R1)NO and PhC(R2)C(R1)N(OBu-t)HgX to form a variety of novel products including the dimeric bisnitronic esters ( 6 ) with R1 = Me or Ph and R2 = H; PhCH(R2)C(R1) = NOBu-t with R1 = Me or Ph and R2 = H or R1 = H and R2 = Ph; PhC(R2)(OBu-t)C(R1)NOH with R1 = H or Me and R2 = Ph; and 3-phenyl-2-R1-indoles with R1 = H, Me, Ph, PhS or t-BuS and R2 = Ph. Nitrosoaromatics react with t-BuHgX in the dark to form ArN(OBu-t)(OBu-t)HgX+ which condenses with ArNO to form the azoxy compound. tert-Butyl radicals will add to RNO2 [R = Ph, Ph2CCH, Ph2CC(Ph)] in the presence of t-BuHgI2 to form products derived from RN(OBu-t)OHgI+.  相似文献   

2.
Carbon-11-labeled nimesulide analogs, N-[11C]methyl-N-(2-benzyloxy-4-nitrophenyl)methanesulfonamide ([11C]4a), N-[11C]methyl-N-[2-(4′-methylbenzyloxy)-4-nitrophenyl]methanesulfonamide ([11C]4b), N-[11C]methyl-N-[2-(4′-fluorobenzyloxy)-4-nitrophenyl]methanesulfonamide ([11C]4c), N-[11C]methyl-N-[2-(4′-nitrobenzyloxy)-4-nitrophenyl]methanesulfonamide ([11C]8a), N-[11C]methyl-N-[2-(β-naphthylmethoxy)-4-nitrophenyl]methanesulfonamide ([11C]8b), and N-[11C]methyl-N-[2-(2′-phenylbenzyloxy)-4-nitrophenyl]methanesulfonamide ([11C]8c), have been synthesized as new potential positron emission tomography (PET) selective aromatase expression regulator (SAER) radiotracers for imaging of aromatase expression in breast cancer. The target tracers were prepared by N-[11C]methylation of their corresponding precursors using [11C]CH3OTf under basic conditions (NaH) and isolated by reversed-phase high-pressure liquid chromatography (HPLC) method in 30–50% radiochemical yields decay corrected to end of bombardment (EOB) with 25–30 min overall synthesis time and 111–148 GBq/μmol specific activity at end of synthesis (EOS).  相似文献   

3.
Abstract— –On in situ photolysis (Λ= 250–400 nm) of aqueous oxaloacetic acid solutions, between pH 5 and 10, the radicals -O2CCH2C(O->=C(O+)CH2CO2- and -O2CCH2C(OH)CO2 are identified. With acetone present, CH2CO2, CH3C(OH)CO-2, CH3C(O-)=C(O)CH2CO2 and -O2CCHCOCO2 are also observed. CO2 and CO are identified as reaction products. The experimental results are explained in terms of α-cleavage of the electronically excited keto-isomer dianion of oxaloacetic acid to yield O2CCH2CO and CO2-. -O2CCH2CO adds to the keto-isomer of oxaloacetic acid and to pyruvic acid, which is formed from oxaloacetic acid by thermal decarboxylation, to yield -O2CCH2C(O-)= C(O-)CH2C0- and CH3C(O-)=C(O)CH2CO-2, respectively, via a decarboxylase substitution reaction. CH2CO2 is derived from -O2CCH2CO by decarbonylation. CO2- is scavenged by oxaloacetic acid and pyruvic acid to yield O2CCH2C(OH)CO2 and CH3C(OH)CO2-, respectively.  相似文献   

4.
Utilization of sodium [1-14C]-, [2–14C]-, and [1,2-13C]-acetates, [1-14C]-, [1-13C]-, or [2-14C]-propionates, [1-14C]-or [2-14C]-malonates, of [1-14C]- or of [1-14C]-myristic acid, or of [1-14C]- and [1-14C]-palmitic acid in the biosynthesis of cytochalasin D ( 1 ) by Zygosporium masonii was determined by degradation studies or by carbon magnetic resonance spectroscopy. The precursors were incorporated primarily via the acetate-malonate pathway to generate 1 from nine intact acetate units, eight of which are coupled in a head to tail fashion to form the C16-polyketide moiety.  相似文献   

5.
By feeding the ant Lasius fuliginosus LATR . with [14C]-1-acetate, [14C]-2-acetate, [14C]-2-mevalonate, [14C]-2-mevalonate, [14C]-1-glucose and [14C]-U-glucose, incorporation ratios of 10?4 – 0,15% were obtained in the sesquiterpenoid dendrolasin. It was shown by analysis of the labelling pattern in dendrolasin that the insertions were spread over the whole molecule in exactly the manner that would be expected from terpene biosynthesis.  相似文献   

6.
Digallane [L1Ga−GaL1] ( 1 , L1=dpp-bian=1,2-[(2,6-iPr2C6H3)NC]2C12H6) reacts with RN=C=O (R=Ph or Tos) by [2+4] cycloaddition of the isocyanate C=N bonds across both of its C=C−N−Ga fragments to afford [L1(O=C−NR)Ga−Ga(RN−C=O)L1] (R=Ph, 3 ; R=Tos, 4 ). The reactions with both isocyanates result in new C−C and N−Ga single bonds. In the case of allyl isocyanate, the [2+4] cycloaddition across one C=C−N−Ga fragment of 1 is accompanied by insertion of a second allyl isocyanate molecule into the Ga−N bond of the same fragment to afford compound [L1Ga−Ga(AllN− C=O)2L1] ( 5 ) (All=allyl). In the presence of Na metal, the related digallane [L2Ga−GaL2] ( 2 ; L2=dpp-dad=[(2,6-iPr2C6H3)NC(CH3)]2) is converted into the gallium(I) carbene analogue [L2Ga:] ( 2 A ), which undergoes a variety of reactions with isocyanate substrates. These include the cycloaddition of ethyl isocyanate to 2 A affording [Na2(THF)5]{L2Ga[EtN−C(O)]2GaL2} ( 6 ), cleavage of the N=C bond with release of 1 equiv. of CO to give [Na(THF)2]2[L2Ga(p-MeC6H4)(N−C(O))2−N(p-MeC6H4)]2 ( 7 ), cleavage of the C=O bond to yield the di-O-bridged digallium compound [Na(THF)3]2[L2Ga-(μ-O)2-GaL2] ( 8 ), and generation of the further addition product [Na2(THF)5][L2Ga(CyNCO2)]2 ( 9 ). Complexes 3 – 9 have been characterized by NMR (1H, 13C), IR spectroscopy, elemental analysis, and X-ray diffraction analysis. Their electronic structures have been examined by DFT calculations.  相似文献   

7.
The stability constants of the Sm(NO3)2+ complex were determined at three temperatures, using the solvent extraction method. It was found that:K 1 0 =63.6 at 17°C, 30.3 at 35°C, 20.1 at 50°C. This corresponds with the formation of a Sm(H2O)(NO3)2+ complex at 17°C and a Sm(H2O)2(NO3)2+ complex at 50°C.
Der Einfluß der Temperatur auf die Bildung von Samarium Nitrato Komplexen
Zusammenfassung Die Stabilitätskonstanten von Sm(NO3)2+ Komplexen wurden mittels der Lösungsmittelextraktionsmethode bei drei Temperaturen bestimmt. Dabei ergab sichK 1 0 =63.6 bei 17°C, 30.3 bei 35°C und 20.1 bei 50°C. Das entspricht der Bildung eines Sm(H2O)(NO3)2+ Komplexes bei 17°C und eines Sm(H2O)2(NO3)2+ Komplexes bei 50°C.
  相似文献   

8.
Isotope-selective IR multiphoton dissociation of CF2C12 molecules in a mixture with HI was experimentally investigated. It was shown that irradiation of the CF2C12-HI mixture leads to the successive buildup of the intermediates CF2HC1 and CF2HI, which also isotopically selectively dissociate simultaneously with the substrate to yield the final product CF2H2 enriched in13C up to 97% at an initial13C concentration of 1.1% in CF2C12. When CF2C12 preliminarily enriched in13C up to 12.3% was used, the attained13C concentration in CF2H2 was as high as ≥99%. Isotopic selectivity and dissociation yields of13C- and12C-containing components of the substrate CF2C12 and both intermediate dissociation products, CF2HC1 and CF2HI, were measured, depending on experimental conditions. The13C distribution over the intermediate and final dissociation products was studied. The side products C2F4C12 and CF2IC1 were detected.  相似文献   

9.
Partial molar volumes, V 2 o and partial molar heat capacities C p,2 o have been determined in aqueous solution at 25°C for the dipeptides glycyl-L-asparagine, glycyl-DL-threonine, glycyl-DL-serine and glycyl-DL-phenylalanine. These results, along with those for some other dipeptides of sequence Gly-X, were used to estimate side chain contributions to V 2 o and C p,2 o . For these dipeptides both V 2 o and C p,2 o were found to be a linear function of the respective thermodynamic property for the amino acid X. The contributions of the glycyl units to V 2 o and C p,2 o of the dipeptide are discussed.  相似文献   

10.
The specific conductivity of dodecyldimethylbenzylammonium bromide (C12BBr) in aqueous solutions, in the temperature range of 15 to 40 °C, has been measured as a function of molality. The two breaks which were found on the conductivity against molality plots were attributed to the critical micelle concentration, cmc, and second critical micelle concentration, 2nd cmc, respectively. The ratio of the slopes, S, of the three linear fragments on the plots, S2/S1 and S3/S1, was attributed to the degree of ionization of the micelles at cmc and 2nd cmc respectively. It was shown that the values of the 2nd cmc estimated above 27 °C are only apparent due to thermal disintegration of the micelles. In the temperature range of 15 to 27 °C, the values of the 2nd cmc increase gradually and the plot of the 2nd cmc against temperature is concave. The ratio of 2nd cmc/cmc for C12BBr at 25 °C amounts to 15 and appears to be high compared to the literature values for other surfactants. For comparative purposes the cmc and 2nd cmc values were also estimated conductometrically for decyldimethylbenzylammonium bromide (C10BBr) at 25 °C. The 2nd cmc value for this surfactant is higher compared to the value for the C12 homologue by a factor of 2.6.The standard Gibbs free energies of micellization at cmc and at the 2nd cmc were estimated from the experimental data for both surfactants at 25 °C.  相似文献   

11.
The reactions of carbethoxycarbene (:CH2-CO2Et, 2) with several acyclic enaminones (RCOCH=CR1NHR2, 3) lead to the unexpected formation of 2-Me, 3-CO2Et, 4-H, 5-R1-pyrroles 4 . Structural variations of the enaminones show that the structural fragments C(3)-CO2Et and C(2)-Me are provided by 2 and that the fragment C(5)-R1NHR2 originates from the enaminones 3 , while the RCO group from 3 is eliminated during the course of reaction. Reactions with cyclic and nitrogen-hindered enaminones do not lead to pyrrole formation but occur by simple insertion of 2 to the Cα-H bond.  相似文献   

12.
The neurosteroid trans‐dehydroandrosterone (DHEA) and its analogs with slightly different modifications in the side chain attached to C17, that is, (3S)‐acetoxypregn‐5‐en‐20‐one ( 1 ) and (3S,20R)‐acetoxypregn‐5‐en‐20‐ol ( 2 ), have been synthesized to investigate DHEA–cation interactions. In this study, we applied solid‐state 1H/13C cross‐polarization/magic‐angle spinning (CP/MAS) nuclear magnetic resonance (NMR) spectroscopy to a series of DHEA analog/Mg2+ mixtures at different Mg2+ concentrations. The high‐resolution 13C NMR spectra of 1 /Mg2+ mixtures exhibit two distinct 13C spectral patterns, one attributable to 1 free from Mg2+, and the other attributable to 1 with bound Mg2+. For 2 , the 13C NMR spectra exhibit three distinct spectral patterns; besides that of the free form, the other two can be assigned to Mg2+‐bound forms. Based on the analysis of the chemical shift deviations (CSDs), we conclude that both 1 and 2 might be subject to a cation–π interaction via the C5–C6 double bond, in contrast to that observed previously for DHEA. As demonstrated, DHEA possesses two Mg2+ binding sites, that is, C17–O and C5–C6 double bond, in which the binding affinity of the former is at least three times stronger than that of the latter. The solid‐state 13C NMR investigation allows better understanding of the underlying cation binding effects of neurosteroid molecules in vitro.  相似文献   

13.
A simple, high‐yielding synthesis of dibutyl[14C]formamide ([14C]DBF; 1 ) from 14CO2 was developed (Scheme 1): reaction of LiBEt3H and 14CO2 followed by aqueous workup gave H14CO2H in high yield. Conversion of the [14C]formic acid to 1 was effected by a standard carbodiimide coupling procedure. The utility of 1 as an alternative to dimethyl[14C]formamide ([14C]DMF) in alkylation reactions and in the [14C]Vilsmeier–Haack reaction was demonstrated for several substrates (Table 2). A 14C‐labeled phosphodiesterase‐4 (PDE‐4) inhibitor, [14C]‐ 2 , was synthesized by application of this technology (Scheme 2).  相似文献   

14.
Functionally conjugated enynes, H2CC(R1)CCCR2R3OS(O)Me, undergo 1,5-substitution with alkylsilver(I) reagents, RAGG ☆ 3 LiBr. The purity of the produced alkylated butatrienes, RCH2C(R1)CCCR2R3 depends on the nature of R in RAg ☆ 3 LiBr and on the substituents R1, R2 and R3 in the substrate.  相似文献   

15.
Incorporation experiments using sodium [2-14C]-, [2-3H]-, (3R)-[5-14C]- and [2-3H, 2-14C]-mevalonates and with mevalonates stereospecifically tritiated at C(2) demonstrate the transformation of mevalonic acid ( 8 ) into verrucarinic acid ( 5 ). Degradation experiments showed that this transformation occurs with a hydrogen 1, 2-shift of the ‘pro-2R’ hydrogen atom of mevalonate to C(3) of verrucarinate. A possible mechanistic pathway is discussed.  相似文献   

16.
Stability constants in methanol at 25.0°C were evaluated for the complexes of the divalent cations Ca2+, Ni2+, Zn2+, Pb2+, Mg2+, Co2+ and Cu2+ with the macrocyclic polyethers 15-crown-5 (15C5), 18-crown-6 (18C6), dicyclohexyl-18-crown-6 (DC18C6) and dibenzo-24-crown-8 (DB24C8). The log K values of the 1:1 complexes were generally in the range 2.1–4.2, which is low in comparison to the values of the corresponding crown ether/alkali metal ion complexes. M2L complexes were observed for the systems Pb2+/18C6, Pb2+/DC18C6, Ca2+/DC18C6 and Cu2+/D18C6, whereas ML2 complexes were found for Ca2+/18C6 and Cu2+/18C6. Within the series of complexes studied, there was no clear relationship between cation diameter and hole size.  相似文献   

17.
Microwave studies (26.5–40 GHz) of further isotopic species of selenoketene formed by pyrolysis of 1,2,3-selenodiazole (12CH212C76,77,82Se, 12CH213C80Se and 13CH212C80Se) and by pyrolysis of 5-deuterio-1,2,3-selenodiazole (12CHD12C78,80Se) are reported. In conjunction with earlier results for 12CH12C78,80Se an rs structure has been derived with distances SeC (1.706 Å), CC (1.303 Å), CH (1.0908 A) and a HCH bond angle of 119.7°. The geometry of the CH2C moiety of selenoketene is closer to allene, CH2CCH2, than to ketene, CH2CO.  相似文献   

18.
Chlorophyll derivatives that possessed a phenylsulfanyl group at the C31- or C32-position were synthesized and their optical properties were investigated. Methyl 31-phenylsulfanyl-mesopyropheophorbide-a was prepared by substitutions of the corresponding C31-hydroxy-chlorin, methyl bacteriopheophorbide-d, with thiophenol in the presence of zinc iodide or of the corresponding C31-bromo-chlorin with thiophenol. The regioisomeric C32-phenylsulfanyl-chlorin was obtained by addition of thiophenol to the C3-vinyl group of methyl pyropheophorbide-a in the presence of AIBN. Both the synthetic compounds gave similar electronic absorption and emission spectra in chloroform, but fluorescence quantum yield of the C31-sulfanyl-chlorin (0.18) was ca. 30% smaller than those of the C32-sulfanyl-chlorin (0.25) and the C3-ethyl-chlorin (0.24). These observations were consistent with their fluorescence lifetime data. It is suggested that the heavy atom effect of a sulfur atom at the C31-position can tune photophysical properties of the chlorophyll derivatives.  相似文献   

19.
Reactions of alkyl 4-aryl(or 4,4-diaryl)-4-hydroxybut-2-ynoates [Ar(H or Ar')(OH)C4–C3≡C2–CO2Alk] with arenes under the action of triflic acid TfOH or HUSY zeolite result in the formation of two main compounds, aryl substituted furan-2-ones or products of propargylation of electron rich arenes. Key reactive intermediates in these transformations are the corresponding O,O-diprotonated forms of starting butynoates, Ar(H or Ar')(+OH2)C4–C3≡C2– C(=O+H)(OAlk), dehydration of which gives rise to mesomeric propargyl-allenyl cations Ar(H or Ar')(OH)4C+–C3≡C2–C(=O+H)(OAlk) ? Ar(H or Ar')(OH)4C = C3 = 2C+–C(=O+H)(OAlk), having two electrophilic centers on the carbons C4 and C2 respectively. Reactions of these species with arenes at C4 lead to products of arene propargylation, alternatively, reactions at C2 result in allenylation of arenes, followed by further transformation into furan-2-ones. Using quantum chemical calculations by the DFT method, it has been shown that the reactivity of such propargyl-allenyl cations is mainly explained by orbital factors. Plausible reaction mechanism is discussed.  相似文献   

20.
The stability constants (Kf) for the complexation reactions of Cr3+, Mn2+ and Zn2+ metal cations with macrocyclic ligand, 15-crown-5 (15C5), in acetonitrile (AN), ethanol (EtOH) and also in their binary solutions (AN–EtOH) were determined at different temperatures, using conductometric method. 15C5 forms 1:1 complexes with Cr3+, Mn2+ and Zn2+ cations in solutions. A non-linear behaviour was observed for changes of logKf of the metal ion complexes versus the composition of the mixed solvent. The order of stability of the metal–ion complexes in pure AN and in a binary solution of AN–EtOH (mol% AN?=?52) at 25?°C was found to be: (15C5Zn)2+?>?(15C5·Mn)2+?>?(15C5·Cr)3+, but in the case of pure EtOH at the same temperature, it changes to: (15C5·Zn)2+?>?(15C5·Cr)3+?>?(15C5·Mn)2+. The results also show that the stability sequence of the complexes in the other binary solutions of AN–EtOH (mol% AN?=?26 and mol% AN?=?76) varies in order: (15C5·Cr)3+?~?(15C5·Zn)2+?>?(15C5·Mn)2+. The values of the standard thermodynamic quantities (ΔHC°, ΔSC°) for formation of (15C15-Cr3+), (15C5-Mn2+) and (15C5-Zn2+) complexes were obtained from the temperature dependence of the stability constants and the results show that the thermodynamics of complexation reactions is affected by nature and composition of the solvent systems and in most solution systems, the complexes are enthalpy stabilized but entropy destabilized.  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