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.
Thiazole synthase catalyzes the formation of the thiazole moiety of thiamin pyrophosphate. The enzyme from Saccharomyces cerevisiae (THI4) copurifies with a set of strongly bound adenylated metabolites. One of them has been characterized as the ADP adduct of 5-(2-hydroxyethyl)-4-methylthiazole-2-carboxylic acid. Attempts toward yielding active wild-type THI4 by releasing protein-bound metabolites have failed so far. Here, we describe the identification and characterization of two partially active mutants (C204A and H200N) of THI4. Both mutants catalyzed the release of the nicotinamide moiety from NAD to produce ADP-ribose, which was further converted to ADP-ribulose. In the presence of glycine, both the mutants catalyzed the formation of an advanced intermediate. The intermediate was trapped with ortho-phenylenediamine, yielding a stable quinoxaline derivative, which was characterized by NMR spectroscopy and ESI-MS. These observations confirm NAD as the substrate for THI4 and elucidate the early steps of this unique biosynthesis of the thiazole moiety of thiamin in eukaryotes.  相似文献   

2.
The reactions of 2-aminothiazole (1a), 4-methyl-2-aminothiazole (1b), and 4,5-dimethyl-2-aminothiazole (1c) with superelectrophilic 4,6-dinitrobenzofuroxan (DNBF) have been studied in acetonitrile and a 70/30 (v/v) H2O/Me2SO mixture. While exhibiting a somewhat higher nitrogen basicity than that of anilines, 1a and 1b do not react as nitrogen nucleophiles, affording exclusively anionic C-bonded sigma-adducts (C-1a and C-1b) through electrophilic S(E)Ar substitution of the thiazole ring by DNBF. Only in the case of the 4,5-dimethyl derivative 1c a N-adduct, N-1c, was obtained. On the basis of 1H-15N correlations, it is demonstrated that this adduct, N-1c;1c,H+, is derived from DNBF addition at the exocyclic amino group and not at the endocyclic nitrogen center of 1c. Rate constants have been determined in the two solvents for the formation of the adducts, revealing a reactivity sequence which accounts well for the finding that 1a and 1b behave preferentially as carbon rather than nitrogen nucleophiles. The enaminic character of these thiazoles is assessed through an estimation of the pKa values for their C-protonation in aqueous solution as well as through a positioning of their reactivity on the nucleophilicity scale recently developed by Mayr et al. (Acc. Chem. Res. 2003, 36, 66). With N values of the order of 6.80 and 5.56, 1b and 1a have a carbon nucleophilicity comparable to that of N-methylindole and indole, respectively.  相似文献   

3.
Several new heterocyclic compounds such as thiophene derivatives have been synthesized by the reactions of the versatile unreported 3‐mercapto‐2‐(4‐methyl‐2‐(tosylamino)thiazole‐5‐carbonyl)‐3‐phenylaminopropenal ( 7 ) with chloroacetonitrile, ethyl 2‐bromoacetate, and 1‐aryl‐2‐bromoethanone derivatives. Reaction of potassium salt intermediate 6 with hydrazonyl halides ( 14a , 14b , 14c ) to afford 1,3,4‐thiadiazole derivatives incorporating a thiazole moiety has been described. All newly synthesized compounds were elucidated by considering the data of both elemental and spectral analysis.  相似文献   

4.
A one-pot, two-step method to prepare 3-halohydrofurans efficiently by TfOH-catalyzed hydroxylation/halocyclization of cyclopropyl methanols with H(2)O and N-halosuccinimide (NXS, X=1, Br, Cl) or Selectfluor is described. The reactions proceed rapidly under mild and operationally straightforward conditions with a catalyst loading as low as 1 mol % and afford the 3-halohydrofuran products in moderate to excellent yields and, in most cases, with preferential cis diastereoselectivity. The method was shown to be applicable to cyclopropyl methanols containing electron-withdrawing, electron-donating, and sterically demanding functional groups and electrophilic halide sources. The mechanism is suggested to involve protonation of the alcohol substrate by the Br?nsted acid catalyst and ionization of the starting material. This results in ring-opening of the cyclopropane moiety and in situ formation of a homoallylic alcohol intermediate, which undergoes subsequent intramolecular halocyclization on treating with the electrophilic halide source to give the halohydrofuran. The observed cis product selectivity is thought to be determined by the reaction proceeding through an in situ generated unsaturated alcohol intermediate that contains a (Z)-alkene moiety under the kinetically controlled conditions.  相似文献   

5.
The reaction of 4,6‐dinitrobenzofuroxan (DNBF) with 1‐trimethylsilyloxybuta‐1,3‐diene ( 8 ) is shown to afford a mixture of [2+4] diastereomeric cycloadducts ( 10 , 11 ) through stepwise addition–cyclization pathways. Zwitterionic intermediate σ‐adduct 9 , which is involved in the processes, has been successfully characterized by 1H and 13C NMR spectroscopy and UV/visible spectrophotometry in acetonitrile. A kinetic study has been carried out in this solvent that revealed that the rate of formation of 9 nicely fits the three‐parameter equation log k=s(E+N) developed by Mayr to describe the feasibility of nucleophile–electrophile combinations. This significantly adds to the NMR spectroscopic evidence that the overall cycloadditions take place through a stepwise mechanism. The reaction has also been studied in dichloromethane and toluene. In these less polar solvents, the stability of 9 is not sufficient to allow direct characterization by spectroscopic methods, but a kinetic investigation supports the view that stepwise processes are still operating. An informative comparison of our reaction with previous interactions firmly identified as prototype stepwise cycloadditions is made on the basis of the global electrophilicity index, ω, defined by Parr within the density functional theory, and highlighted by Domingo et al. as a powerful tool for understanding Diels–Alder reactions.  相似文献   

6.
A series of new 5-(heteroaryldiazenyl)thiazole incorporating pyrazole moiety have been synthesized through coupling of the thiazole with the appropriate heteroaryldiazonium salts. The newly synthesized compounds were characterized by elemental analysis, spectroscopic (IR, 1H NMR, and Mass) data, and alternative synthesis whenever possible.  相似文献   

7.
[Reaction: see text]. A collection of 13 unsymmetrical ketones, each one featuring a sugar (d-glucosyl, d-galactosyl, d-mannosyl, and l-fucosyl) and an aglycone moiety (phenyl, 2-thiazolyl, TMS-ethynyl, allyl, and 1-propenyl) was prepared by a uniform route based on the use of benzothiazole as a carbonyl group equivalent. Succinctly, C-glycosylbenzothiazoles readily prepared by addition of 2-lithiobenzothiazole to sugar lactones and deoxygenation, were subjected to a one-pot reaction sequence involving N-methylation of the heterocyclic ring by MeOTf, treatment of the N-methylbenzothiazolium salt with a Grignard reagent, and HgCl(2)-promoted hydrolysis of the benzothiazoline thus formed. The resulting ketones were isolated in yields varying from 35 to 80%. Treatment of the sugar ketones with various organometals containing the phenyl, 2-thiazolyl, TMS-ethynyl, or ethynyl group as a substituent afforded chiral tertiary alcohols. These addition reactions were highly stereoselective as observed by crude NMR analysis and isolation of a single epimer in high yield in each case examined. However, because of the complexity of the reagents involved, the stereochemical outcome of these reactions appears to be difficult to rationalize by simple classical steric models, thus, ab initio studies taking into account the role of the sugar fragment are advisable. An interesting synthetic elaboration of a propargylic alcohol containing the thiazole ring into a propargylic alcohol bearing the formyl and carboxylate groups is reported.  相似文献   

8.
A multistep synthesis of novel pyrene-based thiazole moiety been has been realized following some synthetic challenges and complications. The chemical structure of the synthesized compound has been established on the basis of both spectroscopic and analytical tools. Its nucleophilic reactivity with 4,6-dinitrobenzofuroxan (DNBF) has been successfully studied in solution. A kinetic study of the covalent electrophile/nucleophile combination of dinitrobenzofuroxan (DNBF, electrophile) and 4-(pyren-1-yl)thiazol-2-amine (nucleophile) resulting in the formation of the corresponding σ-adduct in solution is reported. The rate constant (k1) of the second-order relating to the CC bond forming step of this complexation process has been found to fit into the linear correlation log k = sN (N + E), thereby permitting the evaluation of the nucleophilicity parameter (N) of the 4-(pyren-1-yl)thiazol-2-amine. 4-(Pyren-1-yl)thiazol-2-amine has been subsequently ranked according to its reactivity profile on the general nucleophilicity scale developed recently by Mayr et al., leading to an interesting and direct comparison over a large domain of π-, σ-, and n-nucleophiles.  相似文献   

9.
This work clarifies the structural characterization and antioxidant activity between aromatic and 3-arylsydnonyl substituted hydrazino-thiazoles by further synthesizing a series of aromatic ring-substituted hydrazino-thiazole derivatives 8a-h and 9a-h. Hydrazino-thiazole derivatives 8a-h and 9a-h were obtained by reacting aromatic or heterocyclic aromatic aldehyde thiosemicarbazones 7a-h with cyclization reagents ethyl 2-chloroacetoacetate (2a) and 2-bromoacetophenone (2b), respectively. The ORTEP drawings of compounds 8g, 8h and 9f provide strong evidence of the structure of aromatic thiazole derivatives 8a-h and 9a-h. Undoubtedly, the structure of compounds 3e-h and 4e-h synthesized by the reaction of 3-aryl-4-formylsydnone thiosemicarbazones 1e-h with cyclization reagents 2a and 2b in the previous work should have the thiazole moiety, and not the thiazoline moiety. Both the new thiazole derivatives 8a-h and 9a-h and the 3-arylsydnonyl-substituted derivatives 3e-h and 4e-h were investigated to determine their antioxidant activity by two tests that have been highly documented-the direct scavenging effect on a stable free 1,1-diphenyl-2-picrylhydrazyl (DPPH) radical and the inhibition of the 2,2-azino-bis(3-ethylbenzthiazoline-6-sulfonic acid) (ABTS) radical. Results of this study demonstrate that not only the thiazole ring and the aryl ring has the contribution to the antioxidant activities, the sydnone ring of 3-arylsydnonyl moiety also has its considerable contribution.  相似文献   

10.
以2-氨基噻唑和溴乙酰基吡啶为原料合成了二个咪唑[2,1-b]噻唑的甲酰基化合物3,4,然后与N1-氨基丁基-N1,N4-二叔丁氧羰基-1,4-丁二胺经缩合后用NaBH4还原,产物提纯后脱保护得目标产物7、8,并通过IR,^1H NMR,^13C NMR,ESI-MS对目标化合物的结构进行了表征.  相似文献   

11.
A kinetic study of the reactions of potentially bioactive 2-amino-4-arylthiazoles with highly reactive 4,6-dinitrobenzofuroxan (DNBF) is reported herein in acetonitrile solution. The complexation reaction was followed by recording the UV–vis spectra with time at λmax = 482 nm. Electronic effects of substituents influencing the rate of reaction have been studied using structure-reactivity relationships. It is shown that the Hammett plot relative to the reaction of DNBF with 2-amino-4-(4-chlorophenyl)thiazole exhibit positive deviation from the log k1 versus σ correlation, while it showed excellent linear correlation in terms of Yukawa–Tsuno equation. It has be noticed that the nonlinear Hammett plot observed for 2-amino-4-(4-chlorophenyl) thiazole is not attributed to a change in rate-determining step but is due to nature of electronic effect of substituent caused by the resonance of stabilization of substrates. The second-order rate constant (k1) relating to the bond C–C and C-N forming step of the complexation processes of DNBF with 4-substituted-aminothiazoles and 2-amino-5-methyl-4-phenylthiazole, respectively, is fit into the linear relationship log k = sN (N + E), thereby permitting the assessment of the nucleophilicity parameter (N) of the 2-amino-4-arylthiazoles of the range (4.90 < N < 6.85). 2-amino-4-arylthiazoles is subsequently ranked by positioning its reactivity on the general nucleophilicity scale developed recently by Mayr and coworkers (2003) leading an interesting and a direct comparison over a large domain of π-, σ -, and n-nucleophiles. The global electrophilicity/nucleophilicity reactivity indexes of the 2-amino-4-arylthiazoles have been investigated by means of a density functional theory (DFT) method. .  相似文献   

12.
2‐Thiazolylhydrazonomalononitrile ( 1 ) was used as key intermediate for the synthesis of polyfunctionally substituted heterocycles (e.g. pyridazine, tetrazole, pyrimidines, 1,5‐benzo diazepine, benzoimidazo[1,2‐a]pyrimidine, triazolo[4,3‐a]pyrimidine, pyrazole, pyrazolo[1,5‐a]triazines, pyrazolo[1,5‐a]pyrimidine and 1,5‐benzoxazocine) incorporating thiazole moiety via its reactions with some carbon and nitrogen nucleophiles. Structures of the newly synthesized compounds were established by elemental analysis and spectral data. The mechanistic aspects for the formation of the new compounds were also discussed.  相似文献   

13.
The reactions 4-methylphenyl 4-nitrophenyl carbonate (MPNPC), 4-chlorophenyl 4-nitrophenyl carbonate (CIPNPC), 4-methylphenyl 2,4-dinitrophenyl carbonate (MPDNPC), and 4-chlorophenyl 2,4-dinitrophenyl carbonate (CIPDNPC) with a homogeneous series of phenoxide anions are subjected to a kinetic investigation in aqueous solution (25.0 degrees C, ionic strength 0.2 M (KCI)). Under an excess of phenoxide with respect to the substrate, all of these reactions obey pseudo-first-order kinetics and are first order in phenoxide. The Br?nsted-type plots for the nucleophilic rate constants (k(N)) are linear, with slopes beta = 0.48 (MPNPC), 0.67 (ClPNPC), 0.41 (MPDNPC), and 0.32 (ClPDNPC). The magnitude of these slopes and the absence of a curvature in the Br?nsted plot at pK(a) = 7.1 for the CIPNPC reactions are consistent with concerted mechanisms (one step). The carbonates MPDNPC and ClPDNPC are more reactive than MPNPC and CIPNPC, respectively, toward phenoxide nucleophiles. This can be explained by the presence of a second nitro group in the nucleofuge of the dinitro derivatives, which (i) leaves their carbonyl carbon more positively charged, making them better electrophiles, and (ii) makes 2,4-dinitrophenoxide a better leaving group than 4-nitrophenoxide. The 4-chloro derivatives are more reactive than the corresponding 4-methyl derivatives. This should be due to the greater electron withdrawal of 4-chloro than 4-methyl, which makes the former carbonyl more electrophilic. Comparison of the concerted phenolysis of MPNPC with the stepwise reactions of secondary alicyclic amines with the same substrate indicates that substitution of a secondary alicyclic amine group in a zwitterionic tetrahedral intermediate by a phenoxy group greatly destabilizes the intermediate. An equation is deduced for log k(N) in terms of the basicity of the nucleophile, the nonleaving moiety, and the leaving group. This equation shows that for these reactions, the sensitivity of log k(N) to the basicity of the nonleaving moiety (beta(nlg) = -0.27) is very similar to that of the nucleofuge (beta(lg) = -0.25).  相似文献   

14.
Readily available 6-(methylsulfinyl)hexanoic acid (1) is employed as a substitute for DMSO in Swern oxidation reactions using oxalyl chloride to smoothly convert primary or secondary alcohols to corresponding aldehydes or ketones in high yield. The resulting 6-(methylthio)hexanoic acid (2) is easily separable by aqueous extraction or by filtration through silica gel and can be reoxidized to 1 with sodium metaperiodate in 97% yield. Low temperature (-60 degrees C) (13)C NMR spectrometry is used to examine the intermediates of this Swern process. The results indicate that any residual unoxidized alcohol is generated during Pummerer elimination of the alkoxysulfonium intermediate and can be minimized by extended exposure to triethylamine at -40 degrees C. Reaction of the potassium salt of 1 with cross-linked chloromethyl polystyrene affords a polymer-bound reagent 12 that quantitatively oxidizes borneol to camphor when used in two-fold excess.  相似文献   

15.
Cationic alkoxycarbene complexes of platinum(II) have been isolated in the reactions of trans-[(PR3)2PtX(R′OH)]PF6 (X  H or Me; R′  Me or Et) with Me3SiCCR′′ (R′′  H, Me or SiMe3). In these reactions cleavage of the carbon-silicon bond by the nucleophilic attack of alcohol has been observed. These carbene complexes have been characterized by elemental analyses and by IR, 1H and 13C NMR spectral data. 13C NMR chemical shift data for carbene carbon atoms suggest that the carbene carbon may be very positively charged.  相似文献   

16.
The structure of the dodecamer d(CGCGAATXCGCG)(2), in which X = Z3dU, 5-(3-aminopropyl)-2'-deoxyuridine, was determined. At neutral pH, Z3dU introduced a positive charge into the major groove. NMR spectroscopy revealed that the Z3dU omega-aminopropyl moiety oriented in the 3'-direction from the site of modification. Watson-Crick base pairing remained intact throughout the dodecamer. The presence of the charged amino group in the major groove resulted in a 0.24 ppm upfield shift of one (31)P NMR resonance in the 3'-direction at the phosphodiester linkage between nucleotides C(9) and G(10). Molecular dynamics calculations restrained by distances obtained from (1)H NOE data and torsion angles obtained from (1)H NMR (3)J coupling data, and in which the omega-amino group was constrained to be proximate to G(10)O(6), predicted from the (31)P NMR data and molecular modeling (Dande, P.; Liang, G.; Chen, F.-X.; Roberts, C.; Nelson, M. G.; Hashimoto, H.; Switzer, C.; Gold, B. Biochemistry 1997, 36, 6024-6032), were consistent with experimental NOEs. These refined structures exhibited bending. The distance from the amino group to the 5'-phosphate oxygen of Z3dU was >5 A, which indicated that in this dodecamer the Z3dU amino group did not participate in a salt bridge to its 5'-phosphate.  相似文献   

17.
Diynes 1a-c [X(CH(2)Ctbd1;CCO(2)Me)(2): X = (CH(2))(2), 1a, X = CH(2), 1b and X = O, 1c] undergo [2 + 2 + 2] ene-diyne cycloaddition reactions with a variety of allenes (n-butylallene 2a, phenylallene 2b, (4-chlorophenyl)allene 2c, (4-bromophenyl)allene 2d, (3-methoxyphenyl)allene 2e, 1-naphthylallene 2f, cyclohexylallene 2g and cyclopentylallene 2h) in the presence of Ni(dppe)Br(2) and Zn powder in CH(3)CN at 80 degrees C for 8 h to give the corresponding polysubstituted benzene derivatives 4a-l in good to excellent yields. Under similar reaction conditions, unsymmetrical diynes 5a-c (HCtbd1;CCH(2)XCH(2)Ctbd1;CCO(2)Me) react with allenes 2 to afford exclusively the corresponding meta-isomers 6a-g in 73-86% yields. The catalytic reaction is highly regioselective and completely chemoselective. This synthetic method is compatible with many functional groups such as Cl, Br, and OMe on the phenyl group of the allene moiety and an ether linkage in a diyne moiety. In this catalytic reaction, allenes are synthetically equivalent to terminal alkynes. Interestingly, unsymmetrical diyne 7 (MeCtbd1;C(CH(2))(4)Ctbd1;CCO(2)Me) undergoes 2:1 cocyclotrimerization with allenes 2a and 2g to afford the corresponding polysubstituted benzene derivatives 9a,b in 87% and 82% yields, respectively. A plausible mechanism involving a nickelacycloheptadiene intermediate is proposed to account for this nickel-catalyzed reaction.  相似文献   

18.
The solution behavior of (1,3-DimeU-C5)Hg(CH(3)COO) (1a) (1,3-DimeU = 1,3-dimethyluracil) with regard to acetate replacement by anions X (Cl(-), Br(-), I(-), NO(3)(-), SCN(-), CN(-)) and by other model nucleobases (1-methylcytosine, 1-MeC, 1-methyluracil, 1-MeUH, 1-methylthymine, 1-MeTH, 9-ethylguanine, 9-EtGH, and 2-thiouracil, 2-ThioUH) has been studied, primarily by means of (1)H and (199)Hg NMR spectroscopy. Moreover, the bis(1,3-DimeU-C5) complex of Hg has been crystallized and studied by X-ray crystallography. 7a: orthorhombic system, space group Fdd2, a = 14.185(4) ?, b = 25.275(7) ?, c = 7.924(2) ?, V = 2840(2) ?(3), Z = 8. The acetato ligand of 1a is readily displaced by anions X, frequently followed by disproportionation reactions leading to HgX(2) and 7a. The donor atom X trans to C(5) has an effect on (3)J coupling between (199)Hg and H(6) of the 1,3-DimeU ligand according to NO(3)(-) > OAc(-) > Cl(-) approximately Br(-) > I(-) > SCN(-) > CN(-) > 1,3-DimeU-C5 with extremes being 222 (X = NO(3)(-)) and 107 Hz (7a). In the presence of excess metal ions (Ag(+), Hg(2+)), 1a forms hetero- and homonuclear derivatives with the second metal ion probably sitting at O(4). The mixed nucleobase complexes have the second base bound to Hg via N(3) (1-MeU (2a), 1-MeT (3a)), N(4) (1-MeC(-) (4a), 1-MeC (4b)), N(1) (9-EtG (5a)), N(7) (9-EtGH (5b)), and N(1), N(7) (9-EtG (5c)), as well as S(2) (2-ThioU (6a)). With the exception of the 9-ethylguanine complexes 5b and 5c, all the other complexes are inert on the (1)H time scale. In several cases, e.g. 2a, 3a, 4a, and 5a, formation of dinuclear Hg or heteronuclear Ag and Pt derivatives has been established by multinuclear NMR spectroscopy.  相似文献   

19.
Photoenolization as a means to release alcohols   总被引:2,自引:0,他引:2  
We have designed molecules which release alcohols upon exposure to UV light independent of the reaction media, making it possible to liberate alcohols in a controlled manner in applications. Photolysis of 2-(2-isopropylbenzoyl)benzoate ester derivatives 4 in various solvents and in thin films results in the liberation of the alcohol moiety from the ester. The reaction mechanism for the release of the alcohol has been elucidated by time-resolved laser flash photolysis. Upon irradiation the triplet excited state of ketone, 4 is formed, and its lifetime can be estimated to be between 0.08 and 0.8 ns. The triplet excited state decays by efficient intramolecular H-atom abstraction to form a 1,4-biradical, 8, that has a lifetime of less than 17 ns and is trapped by molecular oxygen. In the absence of oxygen, biradical 8 intersystem crosses to form photoenols (Z)-9 and (E)-10 in a ratio of 5:2, respectively. Photoenol (Z)-9 has a lifetime of approximately 3000 ns in protic solvents and returns to the starting material through 1,5 intramolecular hydrogen transfer. The other isomer, (E)-10, is much longer lived (>1 ms) and releases the alcohol moiety through an intramolecular lactonization.  相似文献   

20.
In addition to pyridines, α‐chloroformylarylhydrazine hydrochloride 1 can also react with some N‐heterocyclic compounds. The cycloaddition of 1 with isoquinoline was achieved to obtain 3 . The production of 4, 5, 6 given by cycloaddition of 1 with pyridazine was de pendent on the reaction condition. Some heterocyclic compounds bearing an X‐C=N (X:S, N) group on the ring can react with 1 to gain the derivatives of 2,4‐dihydro‐1,2,4‐triazol‐3‐one. 7, 8, 9 and 10 were given by reaction of 1 with 1,3,5‐triazine, 1,4,5,6‐tetrahydropyrimidine, 1,3‐thiazole and 2‐amino‐1,3‐thiazole, respectively. The reactions for 2‐amino‐1,3,4‐thiadiazole and 3‐amino‐1,2,4‐triazole had the same product 11 .  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