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 640 毫秒
1.
A convenient method for the synthesis of polyamides containing hydroxyl and amino substituents on the aromatic rings of the backbones was developed. These polymers were prepared readily by the chemoselective polycondensation of dicarboxylic acids with diamines with hydroxyl and amino functional groups via the activating agent diphenyl(2,3‐dihydro‐2‐thioxo‐3‐bezoxazolyl)phosphonate. The model reactions were studied in detail to demonstrate the feasibility of chemoselective polycondensation. The direct polycondensation of 5‐hydroxy or 5‐aminoisophthalic acid with 4,4′‐diamino‐4″‐hydroxytriphenylmethane proceeded smoothly under mild conditions and produced the desired polyamides with inherent viscosities up to 0.73 dL · g−1. The polymers obtained were characterized by IR, 1H NMR, and 13C NMR spectroscopies. The polymers were readily soluble in aprotic polar solvents such as N‐methyl‐2‐pyrrolidinone, N,N‐dimethyl formamide, and dimethyl sulfoxide. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 3875–3882, 2000  相似文献   

2.
The equimolar mixture of N‐(hydroxyalkyl)pyrazoles and borane formed boric ester complex, in which the remaining borane was stabilized by the adjacent nitrogen of thr pyrazole ring. The borane complex derived from the chiral pyrazoles such as 3‐phenyl‐l‐menthopyrazole reduced p‐methylacetophenone ( 21 ) enantioselectively. When (2′S)‐2‐(2′‐phenyl‐2′‐hydroxyethyl)‐3‐phenyl‐l‐menthopyrazole ((2′S)‐ 10b ) was used, 21 was reduced into (S)‐p‐methylphenyl‐1‐ethanol ( 22 ) in moderate chemical and optical yields. Due to the inconvenience of the preparation and the lower optical yield, the use of N‐(α‐hydroxyalkyl)pyrazoles was unpromising for the enantioselective reduction of ketones by borane.  相似文献   

3.
A series of N‐methyl‐N‐(2‐triorganylsiloxyethyl)‐1,2,3,4‐tetrahydro(iso)quinolinium iodides has been synthesized via dehydrocondensation reaction of N‐(2‐hydroxyethyl)‐1,2,3,4‐tetrahydroisoquinoline, N‐(2‐hydroxyethyl)‐1,2,3,4‐tetrahydroquinoline and 4,4‐dimethyl‐N‐(2‐hydroxyethyl)‐4‐sila‐1,2,3,4‐tetrahydroisoquinoline with trialkyl(aryl)hydrosilanes and subsequent alkylation, and characterized by 1H, 13C and 29Si NMR and mass spectroscopy. The biological activity data exhibited a marked enhancement of inhibitory activity against tumour cell lines and almost all the test bacterial/fungal strains in comparison with their 2‐hydroxyethyl precursors. Cytotoxicity in the microgram range against HT‐1080 (human fibrosarcoma) and MG‐22A (mouse hepatoma) cancer cell lines was observed for most of compounds. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

4.
Herein, the reaction between CO2 and piperidine, as well as commercially available functionalised piperidine derivatives, for example, those with methyl‐, hydroxyl‐ and hydroxyalkyl substituents, has been investigated. The chemical reactions between CO2 and the functionalised piperidines were followed in situ by using attenuated total reflectance (ATR) FTIR spectroscopy. The effect of structural variations on CO2 absorption was assessed in relation to the ionic reaction products identifiable by IR spectroscopy, that is, carbamate versus bicarbonate absorbance, CO2 absorption capacity and the mass‐transfer coefficient at zero loading. On absorption of CO2, the formation of the carbamate derivatives of the 3‐ and 4‐hydroxyl‐, 3‐ and 4‐hydroxymethyl‐, and 4‐hydroxyethyl‐substituted piperidines were found to be kinetically less favourable than the carbamate derivatives of piperidine and the 3‐ and 4‐methyl‐substituted piperidines. As the CO2 loading of piperidine and the 3‐ and 4‐methyl‐ and hydroxyalkyl‐substituted piperidines exceeded 0.5 moles of CO2 per mole of amine, the hydrolysis of the carbamate derivative of these amines was observed in the IR spectra collected. From the subset of amines analysed, the 2‐alkyl‐ and 2‐hydroxyalkyl‐substituted piperidines were found to favour bicarbonate formation in the reaction with CO2. Based on IR spectral data, the ability of these amines to form the carbamate derivatives was also established. Computational calculations at the B3LYP/6‐31+G** and MP2/6‐31+G** levels of theory were also performed to investigate the electronic/steric effects of the substituents on the reactivity (CO2 capture performance) of different amines, as well as their carbamate structures. The theoretical results obtained for the 2‐alkyl‐ and 2‐hydroxyalkyl‐substituted piperidines suggest that a combination of both the electronic effect exerted by the substituent and a reduction in the exposed area of the nitrogen atom play a role in destabilising the carbamate derivative and increasing its susceptibility to hydrolysis. A theoretical investigation into the structure of the carbamate derivatives of these amines revealed shorter N? C bond lengths and a less‐delocalised electron distribution in the carboxylate moiety.  相似文献   

5.
Summary: Poly(alkylene hydrogen phosphonate)s with a number‐average molecular weight of about 3 000 Da were obtained by a transesterification of dimethyl hydrogen phosphonate with poly(ethylene glycol) (PEG 400) under microwave irradiation with a very short reaction time (55 min) relative to that of classical thermal heating (9 h). The structure of the resulting polymer was confirmed by 1H, 31P, and 13C NMR spectroscopy. The molecular weight was determined by 1H, 31P{H} NMR spectroscopy, MALDI‐TOF, and GPC.

The transesterification of dimethyl hydrogen phosphonate with poly(ethylene glycol).  相似文献   


6.
IntroductionThepotentialforuseoftwo photonabsorbingmoleculesinapplicationsrangingfromopticallimiting1 3tothreedimensional (3D)fluorescencemicroscopy4 and 3Dmicrofabricationandopticaldatastorage5,6 hasstimulatedresearchonthedesign ,synthesis ,andcharacterizationofnewmoleculeswithlargetwo photonabsorptivities .7,8Thetwo photonabsorption (2PA)processconsideredherein volvesthesimultaneousabsorptionoftwophotons ,eitherdegeneratingornondegenerating ,atwavelengthswellbe yondthelinearabsorptionspectr…  相似文献   

7.
Dedicated to Dr. János Császár on the occasion of his 70th birthday Ring transformation of 2‐cyanoimido‐3‐methyl‐1,3‐oxazolidine ( 10 ) yielded 5‐amino‐3‐[N‐(2‐hydrox‐yethyl)‐N‐methyl]amino‐1H‐1,2,4‐triazole ( 6 ) that was ring closed with different β‐keto esters to 2‐[N‐(2‐hydroxyethyl)‐N‐methyl]amino‐1,2,4‐triazolo[1,5‐a]pyrimidinones ( 4 ). Cyclisation of derivatives 4 led to imidazo[2′,1′:3,4][1,2,4]triazolo[1,5‐a]pyrimidines ( 2 ) and imidazo[1′,2′:2,3][1,2,4]triazolo[1,5‐a]pyrim‐idines ( 3 ) representing 10 novel ring systems. Besides spectroscopical evidence of structure of derivatives 2 and 3 X‐ray diffraction analysis of derivative 2b was also performed.  相似文献   

8.
Polycondensation of 1,10‐decanediol with dimethyl‐H‐phosphonate taken in excess leads to oligomers with methyl‐H‐phosphonate end groups. The polytransesterification of the resulting oligomer as well as the related model reactions were studied. The synthesis of poly(decamethylene‐H‐phosphonate) was analyzed and the final product had n = 1.4–1.9 104 (from end groups, vpo, and n of the derived polymers). The exchange of the ester groups between two homoesters (dimethyl and diethyl phosphonates) used as models, conducted at r.t. and catalyzed by metal alkoxide provides mixed (hetero) ester in a few minutes. If the concentration of the catalyst is not high enough, then the reaction does not go to equilibrium, because the alcoholate anions are converted into the anions of monoesters of the H‐phosphonic acid, catalytically inactive at this temperature. However, these monoesters become catalytically active at higher temperature, i.e., at the conditions used for preparing higher molecular‐weight products by transesterification. The apparent rate constants () of the ester exchange catalyzed by monoester salt (modeling the propagation step in polytransesterification) were determined by two independent methods; at 130°C ∼ 1.0 · 10−2 mol−1 · L · s−1. The detailed study of the model polytransesterification, and particularly of the polymer end groups appearance and disappearance (studied by 1H‐, 13C‐, and 31P‐NMR) allowed postulation of the reaction mechanism and confirmed our previous work, describing formation at these conditions of polymers with n > 104. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 1365–1381, 1999  相似文献   

9.
The reaction of 5‐chloro‐3‐methyl‐1‐phenyl‐1H‐pyrazole‐4‐carbaldehyde and N‐benzylmethylamine under microwave irradiation gives 5‐[benzyl(methyl)amino]‐3‐methyl‐1‐phenyl‐1H‐pyrazole‐4‐carbaldehyde, C19H19N3O, (I). Subsequent reactions under basic conditions, between (I) and a range of acetophenones, yield the corresponding chalcones. These undergo cyclocondensation reactions with hydrazine to produce reduced bipyrazoles which can be N‐formylated with formic acid or N‐acetylated with acetic anhydride. The structures of (I) and of representative examples from this reaction sequence are reported, namely the chalcone (E )‐3‐{5‐[benzyl(methyl)amino]‐3‐methyl‐1‐phenyl‐1H‐pyrazol‐4‐yl}‐1‐(4‐bromophenyl)prop‐2‐en‐1‐one, C27H24BrN3O, (II), the N‐formyl derivative (3RS )‐5′‐[benzyl(methyl)amino]‐3′‐methyl‐1′,5‐diphenyl‐3,4‐dihydro‐1′H ,2H‐[3,4′‐bipyrazole]‐2‐carbaldehyde, C28H27N5O, (III), and the N‐acetyl derivative (3RS )‐2‐acetyl‐5′‐[benzyl(methyl)amino]‐5‐(4‐methoxyphenyl)‐3′‐methyl‐1′‐phenyl‐3,4‐dihydro‐1′H ,2H‐[3,4′‐bipyrazole], which crystallizes as the ethanol 0.945‐solvate, C30H31N5O2·0.945C2H6O, (IV). There is significant delocalization of charge from the benzyl(methyl)amino substituent onto the carbonyl group in (I), but not in (II). In each of (III) and (IV), the reduced pyrazole ring is modestly puckered into an envelope conformation. The molecules of (I) are linked by a combination of C—H…N and C—H…π(arene) hydrogen bonds to form a simple chain of rings; those of (III) are linked by a combination of C—H…O and C—H…N hydrogen bonds to form sheets of R 22(8) and R 66(42) rings, and those of (IV) are linked by a combination of O—H…N and C—H…O hydrogen bonds to form a ribbon of edge‐fused R 24(16) and R 44(24) rings.  相似文献   

10.
A new synthetic route to 6‐substituted‐imidazo[4,5‐c]pyridin‐2‐ons from 4‐aminopyridine has been investigated. 4‐Aminopyridine protected as alkyl carbamates were nitrated with dinitrogen pentoxide to the corresponding methyl, i‐propyl and t‐butyl 3‐nitropyridin‐4‐yl carbamates ( 5a‐c ) in 51‐63 % yields. Attempts to substitute these in the 6‐position by the ONSH and the VNS techniques succeeded with butyl‐amine and the t‐butyl carbamate 9 . From the methyl or t‐butyl 3‐nitropyridin‐4‐yl carbamates 5a, 5c 1,3‐dihydro‐2H‐imidazo[4,5‐c]pyridin‐2‐one ( 1 ) was formed in 73 and 39 % yields, respectively. t‐Butyl 6‐N‐butylamin‐3‐aminopyridin‐4‐yl carbamate ( 6 ) gave 6‐butylamino‐1,3‐dihydro‐2H‐imidazo[4,5‐c]‐pyridin‐2‐one (7) in 53 % yield.  相似文献   

11.
Several N1‐(2‐furanidyl)‐5‐fluorouracil derivatives of α‐hydroxythiophosphonates were synthesized via oxidation by Moffatt's method of N1‐(2‐furanidyl)‐N3‐(hydroxyalkyl)‐5‐fluorouracil, followed by the addition of diethyl thiophosphite. The phosphonate products were obtained by the oxidation of the corresponding thiophosphonates with m‐chloroperoxybenzoic acid. The crystal structure of compound 6a was determined by X‐ray diffraction. © 2002 Wiley Periodicals, Inc. Heteroatom Chem 13:211–215, 2002; Published online in Wiley Interscience (www.interscience.wiley.com). DOI 10.1002/hc.10021  相似文献   

12.
The title compounds, rac‐(1′R,2R)‐tert‐butyl 2‐(1′‐hydroxyethyl)‐3‐(2‐nitrophenyl)‐5‐oxo‐2,5‐dihydro‐1H‐pyrrole‐1‐carboxylate, C17H20N2O6, (I), rac‐(1′S,2R)‐tert‐butyl 2‐[1′‐hydroxy‐3′‐(methoxycarbonyl)propyl]‐3‐(2‐nitrophenyl)‐5‐oxo‐2,5‐dihydro‐1H‐pyrrole‐1‐carboxylate, C20H24N2O8, (II), and rac‐(1′S,2R)‐tert‐butyl 2‐(4′‐bromo‐1′‐hydroxybutyl)‐5‐oxo‐2,5‐dihydro‐1H‐pyrrole‐1‐carboxylate, C13H20BrNO4, (III), are 5‐hydroxyalkyl derivatives of tert‐butyl 2‐oxo‐2,5‐dihydropyrrole‐1‐carboxylate. In all three compounds, the tert‐butoxycarbonyl (Boc) unit is orientated in the same manner with respect to the mean plane through the 2‐oxo‐2,5‐dihydro‐1H‐pyrrole ring. The hydroxyl substituent at one of the newly created chiral centres, which have relative R,R stereochemistry, is trans with respect to the oxo group of the pyrrole ring in (I), synthesized using acetaldehyde. When a larger aldehyde was used, as in compounds (II) and (III), the hydroxyl substituent was found to be cis with respect to the oxo group of the pyrrole ring. Here, the relative stereochemistry of the newly created chiral centres is R,S. In compound (I), O—H...O hydrogen bonding leads to an interesting hexagonal arrangement of symmetry‐related molecules. In (II) and (III), the hydroxyl groups are involved in bifurcated O—H...O hydrogen bonds, and centrosymmetric hydrogen‐bonded dimers are formed. The Mukaiyama crossed‐aldol‐type reaction was successful when using the 2‐nitrophenyl‐substituted hydroxypyrrole, or the unsubstituted hydroxypyrrole, and boron trifluoride diethyl ether as catalyst. The synthetic procedure leads to a syn configuration of the two newly created chiral centres in all three compounds.  相似文献   

13.
Bioassay‐directed isolation and purification of the methanol extract of Moringa oleifera fruits yielded bioactive N‐benzyl thiocarbamates, N‐benzyl carbamates, benzyl nitriles, and a benzyl ester. Among these, methyl 2‐[4‐(α‐L ‐rhamnopyranosyl)phenyl]acetate ( 2 ), N‐[4‐(β‐L ‐rhamnopyranosyl)benzyl]‐1‐Oα‐D ‐glucopyranosylthiocarboxamide ( 3 ), 1‐O‐phenyl‐α‐L ‐rhamnopyranoside ( 5 ), and 4‐[(β‐D ‐glucopyranosyl)‐(1→3)‐(α‐L ‐rhamnopyranosyl)]phenylacetonitrile ( 6 ) are novel, and their structures were determined by spectroscopic methods. The known compounds isolated and characterized from the MeOH extract were niazirin (=4‐(α‐L ‐rhamnopyranosyl)phenylacetonitrile; 1 ), niazicin A (=methyl N‐{4‐[(4′‐O‐acetyl‐α‐L ‐rhamnopyranosyl)benzyl]}thiocarbamate; 4 ), methyl N‐{4‐[(α‐L ‐rhamnopyranosyl)benzyl]}carbamate ( 7 ), and methyl N‐{4‐[(4′‐O‐acetyl‐α‐L ‐rhamnopyranosyl)benzyl]}carbamate ( 8 ). The combined yield of these compounds from dried M. oleifera fruits was 1.63%. In rodent pancreatic β‐cells (INS‐1), compounds 4, 5, 6, 7 , and 8 at 100 ppm significantly stimulated insulin release. Cyclooxygenase‐1 (COX‐1) and cyclooxygenase‐2 (COX‐2) enzyme inhibition assays revealed that 5 and 6 were most active at 83 ppm. Compound 6 , however, demonstrated greater specificity for inhibition of COX‐2 enzyme (46%) than COX‐1 enzyme. Lipid peroxidation assays revealed that 4 and 6 at 50 ppm inhibited peroxidation reactions by 80 and 95%, respectively, while 3 and 8 inhibited lipid peroxidation by 35%. These compounds did not inhibit the cell growth when tested with human breast (MCF‐7), central nervous system (CNS, SF‐268), lung (NCI‐H460), or colon (HCT‐116) cancer cell lines. Moreover, these compounds were not cytotoxic at the concentrations tested.  相似文献   

14.
Four new pyrrole alkaloids, methyl 2‐[2‐formyl‐5‐(methoxymethyl)‐1H‐pyrrol‐1‐yl]propanoate ( 1 ), methyl 2‐[2‐formyl‐5‐(methoxymethyl)‐1H‐pyrrol‐1‐yl]‐3‐(4‐hydroxyphenyl)propanoate ( 2 ), dimethyl 2‐[2‐formyl‐5‐(methoxymethyl)‐1H‐pyrrol‐1‐yl]butanedioate ( 3 ), and dimethyl 2‐[2‐formyl‐5‐(methoxymethyl)‐1H‐pyrrol‐1‐yl]pentanedioate ( 4 ), were isolated from the AcOEt extract of the fruits of Lycium chinense Miller (Solanaceae). The stereogenic center C(2) in the bulky N‐alkyl side chain in each of 1 – 4 seems to hold the H‐atoms of nearby CH2 groups, CH2(7′) and CH2(3) (if R≠H), leading to two different chemical shifts in the 1H‐NMR spectrum due to their diastereotopic characteristics. In the 1H‐NMR data of each of 2 – 4 , the enhancement of H? C(2) signal was inhibited by the R group, probably due to steric hindrance, and its chemical shift was influenced by the anisotropy effect. The structures of 1 – 4 were elucidated by analysis of various spectroscopic data, including 1D‐ and 2D‐NMR.  相似文献   

15.
The one-pot synthesis of N-substituted (3-oxobutanyl)carbamates via the tandem condensation of primary amines with methyl chloroformate, followed by the conjugate addition of the resulting carbamate with methyl vinyl ketone in the presence of Sn4+ modified zeolite Hβ (Hβ-SnA) at room temperature is described.  相似文献   

16.
17.
N‐acetyl‐4‐nitrotryptophan methyl ester (2), N‐acetyl‐5‐nitrotryptophan methyl ester (3), N‐acetyl‐6‐nitrotryptophan methyl ester (4) and N‐acetyl‐7‐nitrotryptophan methyl ester (5) were synthesized through a modified malonic ester reaction of the appropriate nitrogramine analogs followed by methylation with BF3‐methanol. Assignments of the 1H and 13C NMR chemical shifts were made using a combination of 1H–1H COSY, 1H–13C HETCOR and 1H–13C selective INEPT experiments. Copyright © 2008 Crown in the right of Canada. Published by John Wiley & Sons, Ltd  相似文献   

18.
For substituted phenyl‐N‐butyl carbamates (1) and 4‐nitrophenyl‐N‐substituted carbamates (2), linear relationships between values of NH proton chemical shift (δNH), pKa, and logk[OH] and Hammett substituent constant (σ) or Taft substituent constant (σ*) are observed. Carbamates 1 and 2 are pseudo‐substrate inhibitors of porcine pancreatic cholesterol esterase. Thus, the mechanism of the reaction necessitates that the inhibitor molecule and the enzyme form the enzyme‐inhibitor tetrahedral species at the Ki step of the reaction and then form the carbamyl enzyme at the kc step of the reaction. Linear relationships between the logarithms of Ki and kc for cholesterol esterase by carbamates 1 and σ are observed, and the reaction constants (ρs) are ?3.4 and ?0.13, respectively. Therefore, the above reaction forms the negative‐charge tetrahedral species and follows the formation of the relatively neutral carbamyl enzymes. For the inhibition of cholesterol esterase by carbamates 2 except 4‐nitrophenyl‐N‐phenyl carbamate and 4‐nitrophenyl‐N‐t‐butyl carbamate, linear relationships of ‐logKi and logkc with σ* are observed and the ρ* values are ?0.50 and 1.03, respectively. Since the above reaction also forms the negative‐charge tetrahedral intermediate, it is possible that the Ki step of this reaction is further divided into two steps. The first Ki step is the development of the positive‐charge at the carbamate nitrogen from the protonation of the carbamate nitrogen. The second Ki step is the formation of the tetrahedral intermediate with the negative‐charge at the carbonyl oxygen. From Arrhenius plots of a series of inhibition reactions by carbamates 1 and 2, the isokinetic and isoequilibrium temperatures are different from the reaction temperature (25°C). Therefore, the observed ρ and ρ* values only depend upon the electronic effects of the substituents. Taken together, the cholesterol esterase inhibition mechanism by carbamates 1 and 2 is proposed.  相似文献   

19.
Jing Sun  Hui Gong  Chaoguo Yan 《中国化学》2015,33(9):1049-1056
The novel spirooxindoline fused [1,3]oxazines were efficiently synthesized from Diels‐Alder reaction of N‐arylmaleimides with 1,2‐dihydro‐2‐oxospiro[3H‐indole‐3,2′‐[2H,9aH‐pyrido[2,1‐b][1,3]oxazines], which were generated in situ from three‐component reactions of substituted pyridines and isatins with methyl propiolate, or dimethyl acetylenedicarboxylate. The stereochemistry of the products was clearly clarified by the analysis of 1H NMR data and single crystal structures of the obtained polycyclic compounds.  相似文献   

20.
The automated on‐line synthesis of DNA‐3′‐PNA chimeras 1 – 4 and (2′‐O‐methyl‐RNA)‐3′‐PNA chimeras 5 – 8 is described, in which the 3′‐terminal part of the oligonucleotide is linked to the N‐terminal part of the PNA via N‐(ω‐hydroxyalkyl)‐N‐[(thymin‐1‐yl)acetyl]glycine units (alkyl=Et, Ph, Bu, and pentyl). By means of UV thermal denaturation, the binding affinities of all chimeras were directly compared by determining their Tm values in the duplex with complementary DNA and RNA. All investigated DNA‐3′‐PNA chimeras and (2′‐O‐methyl‐RNA)‐3′‐PNA chimeras form more‐stable duplexes with complementary DNA and RNA than the corresponding unmodified DNA. Interestingly, a N‐(3‐hydroxypropyl)glycine linker resulted in the highest binding affinity for DNA‐3′‐PNA chimeras, whereas the (2′‐O‐methyl‐RNA)‐3′‐PNA chimeras showed optimal binding with the homologous N‐(4‐hydroxybutyl)glycine linker. The duplexes of (2′‐O‐methyl‐RNA)‐3′‐PNA chimeras and RNA were significantly more stable than those containing the corresponding DNA‐3′‐PNA chimeras. Surprisingly, we found that the charged (2′‐O‐methyl‐RNA)‐3′‐PNA chimera with a N‐(4‐hydroxybutyl)glycine‐based unit at the junction to the PNA part shows the same binding affinity to RNA as uncharged PNA. Potential applications of (2′‐O‐methyl‐RNA)‐3′‐PNA chimeras include their use as antisense agents acting by a RNase‐independent mechanism of action, a prerequisite for antisense‐oligonucleotide‐mediated correction of aberrant splicing of pre‐mRNA.  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