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.
Oligonucleotides containing halogenated `purine' and pyrimidine bases were synthesized. Bromo and iodo substituents were introduced at the 7‐position of 8‐aza‐7‐deazapurine‐2,6‐diamine (see 2b , c ) or at the 5‐position of uracil residues (see 3b , c ). Phosphoramidites were synthesized after protection of 2b with the isobutyryl residue and of 2c with the benzoyl group. Duplexes containing the residues 2b or 2c gave always higher Tm values than those of the nonmodified counterparts containing 2′‐deoxyadenosine, the purine‐2,6‐diamine 2′‐deoxyribonucleoside ( 1 ), or 2a at the same positions. Six 2b residues replacing dA in the duplex 5′‐d(TAGGTCAATACT)‐3′ ( 11 )⋅5′‐d(AGTATTGACCTA)‐3′ ( 12 ) raised the Tm value from 48 to 75° (4.5° per modification (Table 3)). Contrary to this, incorporation of the 5‐halogenated 2′‐deoxyuridines 3b or 3c into oligonucleotide duplexes showed very little influence on the thermal stability, regardless of which `purine' nucleoside was located opposite to them (Tables 4 and 5). The positive effects on the thermal stability of duplexes observed in DNA were also found in DNA⋅RNA hybrids or in DNA with parallel chain orientation (Tables 8 and 9, resp.).  相似文献   

2.
2‐amino‐3, 4, 5, or 6‐nitrobenzoic acids were reacted with PPh3(SCN)2 and alkyl isothiocyanates to give 5, 6, 7, or 8‐nitro‐2‐thioxo‐3‐substituted quinazolin‐4‐ones, respectively. The position of the nitro group was found to have significant influence on the outcome of the reactions. Similarly, the nature of the substituent at position 8 (NO2, NH2, NH(C═O)CH3) in 8‐substituted‐2‐methylthio quinazolin‐4‐ones was also found to significantly influence their reactivity towards morpholine. A selection of the products were also tested for in vitro antibacterial activity but little activity was observed.  相似文献   

3.
《化学:亚洲杂志》2017,12(21):2819-2826
Mono‐, di‐, tri‐, and tetraarylated thieno[3,2‐b ]thiophenes were synthesized by direct site‐selective Pd‐catalyzed C−H activation reactions with various aryl bromides in the presence of a phosphine‐free Pd(OAc)2/KOAc catalyst system in N ,N ‐dimethylacetamide (DMAc). The arylation of 2‐arylthieno[3,2‐b ]thiophene took place at the C3 position if the 2‐aryl substituents possessed electron‐withdrawing groups and at the C5 position if they were bulky and possessed electron‐donating groups.  相似文献   

4.
A series of novel 3‐((4‐(t‐butyl)‐2‐(2‐benzylidenehydrazinyl)thiazol‐5‐yl)methyl)quinolin‐2(1H)‐ones ( 7a – 7z ) were designed, synthesized and evaluated for their ability of inhibiting neuraminidase (NA) of in?uenza H1N1 virus. Some compounds displayed moderate influenza NA inhibitory activity. Compound 7l with the scaffold of 2‐(2‐(2‐methoxybenzylidene)hydrazinyl)thiazole was the best one, exhibiting moderate NA inhibitory activity with IC50 of 44.66 µmol/L. Structure‐activity relationship showed that compounds with methoxy or hydroxy groups at the ortho position, fluorine and nitro groups at the meta position and chlorine and bromine groups at the para position of phenyl ring were more active. Docking study indicated that compound 7l has important interactions with some key residues (including Asp151, Glu119, Arg292, Tyr406, and Asn347) and binds to 430‐cavity adjacent to NA active site.  相似文献   

5.
A group of 1,2‐diphenyl‐3,5‐dioxopyrazolidines possessing a methylsulfonyl ( 11 ) or sulfonamide ( 15 ) substituent at the para position of the N1‐phenyl ring, in conjunction with a hydrogen, methyl or fluoro sub‐stituent at the para position of the N2‐phenyl ring, and a C‐4 n‐butyl, methyl or spiro‐cyclopropyl substituent were synthesized for evaluation as potential cyclooxygenase‐2 (COX‐2) selective inhibitor antiinflammatory agents. The title compounds 11 and 15 were synthesized using a four‐step and a three‐step reaction sequence, respectively. Thus, the acetic acid promoted condensation of a nitrosobenzene 5 with an aniline derivative ( 6, 12 ) gave the corresponding azobenzene product ( 8, 13 ) which was reduced with zinc dust in the presence of ammonium chloride to yield the corresponding hydrazobenzene ( 9, 14 ). Base‐catalyzed condensation of 9 and 14 with a malonyl dichloride ( 10 ) afforded the target 3,5‐dioxopyrazolidine product ( 11,15 ). 4‐n‐Butyl‐1‐(4‐methylsulfonylphenyl)‐2‐phenyl‐3,5‐dioxopyrazolidine ( 11a ) was a selective COX‐1 inhibitor (COX‐1 IC50 = 8.48 μM). In contrast, 4‐n‐butyl‐1‐(4‐methylsulfonylphenyl)‐2‐(4‐tolyl)‐3,5‐dioxopyrazolidine ( 11b , COX‐2 IC50 = 11.45 μM) and 4‐n‐butyl‐1‐(4‐methylsulfonylphenyl)‐2‐(4‐fluorophenyl)‐3,5‐dioxopyrazoli‐dine ( 11c , COX‐2 IC50 = 9.86 μM) were about 46‐fold and 20‐fold less selective COX‐2 inhibitors respectively, relative to the reference drug celecoxib.  相似文献   

6.
Various (hetero)arenecarboxylic acids were converted to the corresponding Daugulis amides and nitrated selectively in the ortho‐position in the presence of [CuNO3(PPh3)2] and AgNO2 at 50 °C. A microwave‐assisted saponification allows regenerating the carboxylate group within minutes, which may then be removed tracelessly by protodecarboxylation, or substituted by aryl‐ or alkoxy‐groups via decarboxylative cross‐coupling.  相似文献   

7.
HO–adduct radicals have been investigated and confirmed as the common initial intermediates in TiO2 photocatalysis and Fenton degradations of water‐insoluble aromatics. However, the evolution of HO–adduct radicals to phenols has not been completely clarified. When 4‐d‐toluene and p‐xylene were degraded by TiO2 photocatalysis and Fenton reactions, respectively, a portion of the 4‐deuterium or 4‐CH3 group (18–100 %) at the attacked ipso position shifted to the adjacent position of the ring in the formed phenols (NIH shift; NIH is short for the National Institutes of Health, to honor the place where this phenomenon was first discovered). The results, combined with the observation of a key dienyl cationic intermediate by in situ attenuated total reflectance FTIR spectroscopy, indicate that, for the evolution of HO–adduct radicals, a mixed mechanism of both the carbocation intermediate pathway and O2‐capturing pathway occurs in both aqueous TiO2 photocatalysis and aqueous Fenton reactions.  相似文献   

8.
A series of novel neonicotinoids analogs were designed by modifying the pharmacophore of imidacloprid to 1,3,5‐hexahydrotriazine conjugated to nitroimine (?NNO2) and introducing the phenyl or arylmethyl at the 5‐position, and their insecticidal activities were evaluated. Introducing a heterocyclic methyl at 5‐position increased the insecticidal activities, whereas other phenyl, phenylmethyl or phenylethyl substituents were unfavorable to activities. Molecular docking study was also performed to clarify the interactions of the most potent analog 1‐((6‐chloropyridin‐3‐yl)methyl)‐5‐(3‐pyridylmethyl)‐1,3,5‐hexahydrotriazine‐2‐(N‐nitro) imine ( 7s ) with the target nicotinic acetylcholine receptor, which explained the structure‐activity relationships observed in vitro, and revealed further possibilities for insecticide development. J. Heterocyclic Chem., (2011).  相似文献   

9.
7‐(o‐Substituted phenyl)‐2,6‐dimethyl‐1,4‐benzoquinone methides which have an electron‐donating methoxy‐(o‐OMe, 2a ) and methyl‐ (o‐Me, 2b ) substituents or an electron‐withdrawing cyano‐ (o‐CN, 2c ) and trifluoromethyl‐ (o‐CF3, 2d ) substituents at the ortho‐position of the aromatic ring and 7‐(m‐substituted phenyl)‐2,6‐dimethyl‐1,4‐benzoquinone methide with an electron‐withdrawing trifluoromethyl‐ (m‐CF3, 2e ) substituent at the meta‐position of the aromatic ring were synthesized, and their asymmetric anionic polymerizations using the complex of lithium 4‐isopropylphenoxide with (?)‐sparteine were carried out in toluene at 0 °C. The polymers with negative optical activity were obtained for all of five monomers, and their specific rotation values largely changed depending upon the substituents of the monomers. On the basis of the comparison of various substituents effects, it was found that the specific rotation of obtained polymers is significantly affected by the electronic effects such as inductive and resonance effects rather than the steric and electrostatic effects of the substituent. © 2017 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2017 , 55 , 1048–1058  相似文献   

10.
A new method for the synthesis of fluorinated α‐C‐glycosides is described. The reactions between highly electrophilic radicals (fluorinated or unfluorinated) and a 2‐benzyloxyglucal or galactal provide 2‐keto‐D ‐arabino‐ or 2‐keto‐D ‐lyxo‐hexopyranosides through an addition/fragmentation process. Sodium borohydride mediated or Meerwein–Ponndorf–Verley (MPV) reduction of these compounds provides α‐C‐glycosides that feature appropriate anchoring groups for further synthetic elaboration. The presence of CF2CO2iPr or CF2Br groups at the pseudo‐anomeric position allows efficient reduction/olefination or Br/Li‐exchange/nucleophilic‐addition sequences. These transformations open the way for the synthesis of fluorinated C‐glycosidic analogues of glycoconjugates.  相似文献   

11.
Imidazolium salts (NHCewg ? HCl) with electronically variable substituents in the 4,5‐position (H,H or Cl,Cl or H,NO2 or CN,CN) and sterically variable substituents in the 1,3‐position (Me,Me or Et,Et or iPr,iPr or Me,iPr) were synthesized and converted into the respective [AgI(NHC)ewg] complexes. The reactions of [(NHC)RuCl2(CHPh)(py)2] with the [AgI(NHCewg)] complexes provide the respective [(NHC)(NHCewg)RuCl2(CHPh)] complexes in excellent yields. The catalytic activity of such complexes in ring‐closing metathesis (RCM) reactions leading to tetrasubstituted olefins was studied. To obtain quantitative substrate conversion, catalyst loadings of 0.2–0.5 mol % at 80 °C in toluene are sufficient. The complex with the best catalytic activity in such RCM reactions and the fastest initiation rate has an NHCewg group with 1,3‐Me,iPr and 4,5‐Cl,Cl substituents and can be synthesized in 95 % isolated yield from the ruthenium precursor. To learn which one of the two NHC ligands acts as the leaving group in olefin metathesis reactions two complexes, [(FL‐NHC)(NHCewg)RuCl2(CHPh)] and [(FL‐NHCewg)(NHC)RuCl2(CHPh)], with a dansyl fluorophore (FL)‐tagged electron‐rich NHC ligand (FL‐NHC) and an electron‐deficient NHC ligand (FL‐NHCewg) were prepared. The fluorescence of the dansyl fluorophore is quenched as long as it is in close vicinity to ruthenium, but increases strongly upon dissociation of the respective fluorophore‐tagged ligand. In this manner, it was shown for ring‐opening metathesis ploymerization (ROMP) reactions at room temperature that the NHCewg ligand normally acts as the leaving group, whereas the other NHC ligand remains ligated to ruthenium.  相似文献   

12.

An alternative synthetic route for the preparation of key intermediate synthons 7‐α‐bromoacetyl‐2‐diphenylaminofluorene (α‐BrDPAF‐H) and 7‐α‐bromoacetyl‐9,9‐dialkyl‐2‐diphenylaminofluorene (α‐BrDPAF‐Cn) was demonstrated. The latter reactions involved the first step of dialkylation of 2‐bromofluorene at C9 position of the fluorene moiety, the second step of a diphenylamino group attachment at C2 position of the resulting dialkylfluorene, and the third step of Friedel‐Craft acylation of α‐bromoacetyl group at C7 position of dialkylated diphenylaminofluorene. From the intermediates α‐BrDPAF‐H and α‐BrDPAF‐Cn, a series of C60keto‐DPAF nanostructures, such as the fullerene monoadducts C60(>DPAF‐H) and C60(>DPAF‐Cn), where n is 2, 4, or 10, were synthesized in a reasonable yield. Molecular mass ions of the dyads C60(>DPAF‐H), C60(>DPAF‐C2), C60(>DPAF‐C4), and C60(>DPAF‐C10) were clearly detected in positive ion matrix‐assisted laser desorption ionization mass spectrum (MALDI–MS) that confirmed the composition mass of each compound synthesized.  相似文献   

13.
Complementary strategies to 2,5‐disubstituted pyrazolo[4,3‐c ]cinnolin‐3‐ones are reported herein, providing late stage substituent introduction at either the 2‐ or the 5‐position. Treating a readily prepared 4‐thiocinnoline ester with substituted hydrazines afforded late stage access to the 2‐position, while late stage substituent introduction at the 5‐position was achieved via two different strategies: alkylation of 4‐hydrazonopyrazol‐3‐ones, followed by a ring‐closing intramolecular SNAr tactic and direct reaction of 5‐(2‐fluorophenyl)‐2,4‐dihydro‐3H‐pyrazol‐3‐ones with aryl diazonium salts, followed by cyclization. The strategies described herein provide practical and general methods to prepare 2,5‐disubstituted pyrazolo[4,3‐c ]cinnolin‐3‐ones.  相似文献   

14.
In the negative‐ion collision‐induced dissociation mass spectra of most organic sulfonates, the base peak is observed at m/z 80 for the sulfur trioxide radical anion (SO3–·). In contrast, the product‐ion spectra of a few sulfonates, such as cysteic acid, aminomethanesulfonate, and 2‐phenylethanesulfonate, show the base peak at m/z 81 for the bisulfite anion (HSO3). An investigation with an extensive variety of sulfonates revealed that the presence of a hydrogen atom at the β‐position relative to the sulfur atom is a prerequisite for the formation of the bisulfite anion. The formation of HSO3 is highly favored when the atom at the β‐position is nitrogen, or the leaving neutral species is a highly conjugated molecule such as styrene or acrylic acid. Deuterium‐exchange experiments with aminomethanesulfonate demonstrated that the hydrogen for HSO3 formation is transferred from the β‐position. The presence of a peak at m/z 80 in the spectrum of 2‐sulfoacetic acid, in contrast to a peak at m/z 81 in that of 3‐sulfopropanoic acid, corroborated the proposed hydrogen transfer mechanism. For diacidic compounds, such as 4‐sulfobutanoic acid and cysteic acid, the m/z 81 ion can be formed by an alternative mechanism, in which the negative charge of the carboxylate moiety attacks the α‐carbon relative to the sulfur atom. Experiments conducted with deuterium‐exchanged and deuterium‐labeled analogs of sulfocarboxylic acids demonstrated that the formation of the bisulfite anion resulted either from a hydrogen transfer from the β‐carbon, or from a direct attack by the carboxylate moiety on the α‐carbon. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

15.
Double Horner–Wadsworth–Emmons reaction of (E)‐2,3‐diaryl‐1,4‐bis(diethylphosphonyl)but‐2‐ene with (p‐substituted) benzaldehydes gave (1E,3E,5E)‐1,3,4,6‐tetraarylhexa‐1,3,5‐trienes in moderate to good yields. Substitution of electron‐withdrawing or ‐donating groups at the para position of the 1,6‐diphenyl groups induced a slight bathochromic shift of UV spectra measured in CHCl3 compared with that of the parent 1,3,4,6‐tetraphenylhexa‐1,3,5‐triene. Although fluorescence was not observed with all the trienes in CHCl3, they markedly emitted visible light in powder forms with quantum yields of 0.15–0.44. Introduction of amino groups at the para position of the 3,4‐diphenyl groups induced a bathochromic shift of emission maxima with good solid‐state quantum yields. Thus, the tetraarylated triene framework is found to serve as a new class of fluorophores that exhibit aggregation‐induced emission.  相似文献   

16.
Nitrogen‐centered urazole radicals exist in equilibrium with tetrazane dimers in solution. The equilibrium established typically favors the free‐radical form. However, 1‐arylurazole radicals bearing substituents at the ortho position favor the dimeric form. We were able to determine the structure of one of the dimers (substituted at both ortho positions with methyl groups), namely 1,2‐(2,4‐dimethylphenyl)‐2‐[2‐(2,4‐dimethylphenyl)‐4‐methyl‐3,5‐dioxo‐1,2,4‐triazolidin‐1‐yl]‐4‐methyl‐1,2,4‐triazolidine‐3,5‐dione, C24H28N6O4, via X‐ray crystallography. The experimentally determined structure agreed well with the computationally obtained geometry at the B3LYP/6‐311G(d,p) level of theory. The preferred syn conformation of these 1‐arylurazole dimers results in the two aromatic rings being proximate and nearly parallel, which leads to some interesting shielding effects of certain signals in the 1H NMR spectrum. Armed with this information, we were able to decipher the more complicated 1H NMR spectrum obtained from a dimer that was monosubstituted at the ortho position with a methyl group.  相似文献   

17.
Herein the sodium alkylmagnesium amide [Na4Mg2(TMP)6(nBu)2] (TMP=2,2,6,6‐tetramethylpiperidide), a template base as its deprotonating action is dictated primarily by its 12 atom ring structure, is studied with the common N‐heterocyclic carbene (NHC) IPr [1,3‐bis(2,6‐diisopropylphenyl)imidazol‐2‐ylidene]. Remarkably, magnesiation of IPr occurs at the para‐position of an aryl substituent, sodiation occurs at the abnormal C4 position, and a dative bond occurs between normal C2 and sodium, all within a 20 atom ring structure accommodating two IPr2?. Studies with different K/Mg and Na/Mg bimetallic bases led to two other magnesiated NHC structures containing two or three IPr? monoanions bound to Mg through abnormal C4 sites. Synergistic in that magnesiation can only work through alkali‐metal mediation, these reactions add magnesium to the small cartel of metals capable of directly metalating a NHC.  相似文献   

18.
The phenylazo moiety and its donor‐ and acceptor‐substituted derivatives are studied as effective auxochromes to improve their sensitivity and resolution for distinguishing between the spiro (SP; OFF) and mero (ON) forms in molecular photoswitching applications. Thus, 13 azospiropyran derivatives were synthesized and their spectroscopic and photokinetic behaviors were studied. The quality of photochromic reactions of the synthesized photochromic compounds were compared using a dose–response model. Interestingly, by replacing the nitro group in 6‐nitrospiropyran (ε = 0.42 × 104 M?1 cm?1) with a simple phenylazo moiety, the SP form is still colorless and the color intensity of the merocyanine (MC) form is improved desirably by extending the conjugation length ( 1a , ε = 1.35 × 104 M?1 cm?1). The presence of a hydrophilic OH group or a CH3 group at the para position of phenylazo moiety revealed more or less the same photochromic properties as 1a . The OCH3 group substituted at position 6 of the phenylazo moiety at the para position of the azobenzene moiety effectively increased the photochromic properties with the maximum k‐value for SP to MC switching. Meanwhile, Cl, Br, COOH, and NO2 groups at the para position of the azobenzene moiety revealed the reduction in photochromic properties compared to 1a .  相似文献   

19.
Substituted 7‐aryl‐2,6‐dimethyl‐1,4‐benzoquinone methides which have an electron‐donating methoxy substituent at the para‐position (p‐OMe, 2a ) or an electron‐withdrawing chloro one at the para‐ (p‐Cl, 2b ), meta‐ (m‐Cl, 2c ) , and ortho‐positions (o‐Cl, 2d ) of the benzene ring were synthesized, and their asymmetric anionic polymerizations using the complex of lithium 4‐isopropylphenoxide with (?)‐sparteine were carried out in toluene at 0 °C. The polymers with negative specific rotation were obtained for all of four monomers, and the polymer obtained from 2a showed smaller specific rotation value than that of polymer having no substituent (p‐H, 1 ) on the phenyl group and the polymers obtained from 2b–d showed larger ones. It was found that the kind of a substituent and its substitution position on the phenyl group affect significantly the optical activity of polymers. The largest specific rotation value of [α]435= ?153.2° was obtained in the polymerization of 2d with an ortho‐chloro substituent. © 2014 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2015 , 53, 437–444  相似文献   

20.
Two representatives of a new type of β‐amino acids, carrying two functionalized side chains, one in the 2‐ and one in the 3‐position, have been prepared stereoselectively: a β‐Ser derivative with an additional CH2OH group in the 2‐position (for β‐peptides with better water solubility; Scheme 2) and a β‐HCys derivative with an additional CH2SBn group in the 2‐position (for disulfide formation and metal complexation with the derived β‐peptides; Scheme 3). Also, a simple method for the preparation of α‐methylidene‐β‐amino acids is presented (see Boc‐2‐methylidene‐β‐HLeu‐OH, 8 in Scheme 3). The two amino acids with two serine or two cysteine side chains are incorporated into a β‐hexa‐ and two β‐heptapeptides ( 18 and 23/24 , resp.), which carry up to four CH2OH groups. Disulfide formation with the β‐peptides carrying two CH2SH groups generates very stable 1,2‐dithiane rings in the centre of the β‐heptapeptides, and a cyclohexane analog was also prepared (cf. 27 in Scheme 6). The CD spectra in H2O clearly indicate the presence of 314‐helical structures of those β‐peptides ( 18 , 23 , 24 , 27b ) having the `right' configurations at all stereogenic centers (Fig. 2). NMR Measurements (Tables 1 and 2, and Fig. 4) in aqueous solution of one of the new β‐peptides ( 24 ) are interpreted on the assumption that the predominant secondary structure is the 314‐helix, a conformation that has been found to be typical for β‐peptides in MeOH or pyridine solution, according to our previous NMR investigations.  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