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.
4-Monodeuterated NAD(P)H model compounds (1,4,6,7-tetrahydro-1,6,11-trimethyl-5-oxo-5H-benzo[c]pyrido[2,3-e]az epin; 11Me-MMPAH) have been oxidized with a series of p-benzoquinone and its derivatives in the presence of Mg2+. The models have an axial chirality with respect to the orientation of carbonyl dipole, the dihedral angle of which is larger than 55 degrees out of the plane of dihydropyridine ring. Without Mg2+, the anti- (with respect to the carbonyl dipole) hydrogen is 3 to 32 times more reactive than the corresponding syn-hydrogen, whereas, when Mg2+ is present in the system, the selectivity is shifted toward the syn-preferency. Mg2+ plays the role of a Lewis acid catalyst to control the stereochemistry at the same time as it catalyzes the reaction.  相似文献   

2.
The hallmarks of pyridine nucleotide-dependent dehydrogenase reactions are the stereo- and regiospecific hydride transfer between the nicotinamide coenzyme and the corresponding substrate. When the hydride is delivered from NAD(P)H to reduce the keto-substrate, the site of attack is always at the carbonyl carbon. However, the apparent regioselectivity of the hydride transfer is reversed when difluoromethylene is used as a carbonyl mimic in the NADH-dependent enzyme, TDP-l-rhamnose synthase, which catalyzes the conversion of TDP-6-deoxy-l-lyxo-4-hexulose to TDP-l-rhamnose. The observed reversed regioselectivity can be explained by two mechanisms. One involves the formation of a carbene intermediate followed by a rearrangement involving 1,2-H shift. This mechanistic proposal is theoretically sound and would represent a rare example implicating the intermediacy of a carbene species in an enzyme reaction. However, our results are also consistent with a second mechanism in which the hydride addition to the difluoromethylene moiety occurs at the difluorinated end, opposite from the site predicted on the basis of the reduction of a normal keto functional group. Such a regioselectivity is well precedented in chemical models because nucleophilic addition to fluoroalkenes prefers a route in which the number of fluorines beta to the electron-rich carbon in the transition state is maximized. In this mechanism, the difluoromethylene group may be regarded as a carbonyl mimic with reversed polarity in enzyme catalysis. While further experiments are needed to discriminate between these mechanistic possibilities, the results reported here suggest that the apparent regioselectivity of hydride transfer in a pyridine nucleotide-dependent enzyme can be changed by altering the electrochemical properties of the reaction center.  相似文献   

3.
The kinetics of the hydride exchange reaction between NADH model compound 10-methyl-9,10-dihydroacridine (MAH) and 1-benzyl-3-cyanoquinolinium (BQCN+) ion in acetonitrile were studied at temperatures ranging from 291 to 325 K. The extent of reaction-time profiles during the first half-lives are compared with theoretical data for the simple single-step mechanism and a 2-step mechanism involving initial donor/acceptor complex formation followed by unimolecular hydride transfer. The profiles for the reactions of MAH deviate significantly from those expected for the simple single-step mechanism with the deviation increasing with increasing temperature. The deviation from simple mechanism behavior is much less pronounced for the reactions of 10-methyl-9,10-dihydroacridine-10,10-d2 (MAD) which gives rise to extent of reaction dependent apparent kinetic isotope effects (KIEapp). Excellent fits of the experimental extent of reaction-time profiles with theoretical data for the 2-step mechanism, in the pre-steady-state time period, were observed in all cases. Resolution of the kinetics of the hydride exchange reaction into the microscopic rate constants over the entire temperature range resulted in real kinetic isotope effects for the hydride transfer step ranging from 40 (291 K) to 8.2 (325 K). That the reaction involves significant hydride tunnelling was verified by the magnitudes of the Arrhenius parameters; Ea D - EaH = 8.7 kcal mol-1 and AD/AH = 8 x 10(4). An electron donor acceptor complex (lambda max = 526 nm) was observed to be a reaction intermediate. Theoretical extent of reaction-time profile data are discussed for the case where a reaction intermediate is formed in a non-productive side equilibrium as compared to the case where it is a real intermediate on the reaction coordinate between reactants and products. The common assumption that the two cases are kinetically indistinguishable is shown to be incorrect.  相似文献   

4.
Contrary to a previous report, the sulfurisation of phosphorus(III) derivatives by 3-amino-1,2,4-dithiazole-5-thione (xanthane hydride) does not yield carbon disulfide and cyanamide as the additional reaction products. The reaction of xanthane hydride with triphenyl phosphine or trimethyl phosphite yields triphenyl phosphine sulfide or trimethyl thiophosphate, respectively, and thiocarbamoyl isothiocyanate which has been trapped with nucleophiles. The reaction pathway involves initial nucleophilic attack of the phosphorus at sulfur next to the thiocarbonyl group of xanthane hydride followed by decomposition of the phosphonium intermediate formed to products. The Hammett rho-values for the sulfurisation of substituted triphenyl phosphines and triphenyl phosphites in acetonitrile are approximately -1.0. The entropies of activation are very negative (-114+/-15 J mol-1 K-1) with little dependence on solvent which is consistent with a bimolecular association step leading to the transition state. The negative values of DeltaS(not equal) and rho values indicate that the rate limiting step of the sulfurisation reaction is formation of the phosphonium ion intermediate which has an early transition state with little covalent bond formation. The site of nucleophilic attack has been also confirmed using computational calculations.  相似文献   

5.
The first 1,3-dipolar reaction of azomethine ylides with optically pure vinyl sulfoxide are reported. The presence of the sulfinyl group increase the reactivity of the acrylate moiety as a dipolarophile, and the reactions evolve with complete regio- and endo-selectivities. Nevertheless, mixtures of the two diastereoisomers 4 and 5 (75-88% de) resulting from the anti dipole/s-cis dipolarophile and syn dipole/s-trans dipolarophile approaches, respectively, are obtained. The stereoselectivity can be controlled by using THF or MeCN as solvents or by changing the reaction temperature in MeCN. After separation of the cycloadducts, optically pure 2,5-dihydro-1H-pyrroles are easily obtained by pyrolytic desulfinylation.  相似文献   

6.
Solutions of the zinc hydroxide complex [Tp(Bu(t),Me)]ZnOH in alcohols (ROH; R = Me, Et, Pr(i)) achieve hydride transfer to the NAD(+) model, 10-methylacridinium perchlorate. Deuterium labeling studies, however, demonstrate that the source of the hydride is not the alcohol but, rather, the B [bond] H group of the [Tp(Bu(t),Me)] ligand. A further example in which a [Tp(Bu(t),Me)] ligand acts as a hydride donor is provided by the reaction of the aqua complex [[Tp(Bu(t),Me)]Zn(OH(2))][HOB(C(6)F(5))(3)] with MeOH to generate the zinc hydride complex [Tp(Bu(t),Me)]ZnH. The present study therefore provides a caveat for the often assumed inertness of the B [bond] H group in tris(pyrazolyl)hydroborato ligands, especially in the presence of reactive cationic species.  相似文献   

7.
Cations 2-methyl-5-nitroisoquinolinium (IQ+), tropylium (T+), and xanthylium (Xn+) were treated by an NAD(P)H model 1-(p-substituted phenyl)-1.4-dihydronicotinamide series (1) in buffered aqueous solution to give the corresponding reduced products by accepting hydride. Effects of the 4-substituents of 1 on the reaction rates were investigated. Hammett's linear free energy relationship analysis on the three reactions of 1 provides the reaction constants of -0.48, -2.2, and -1.4 with IQ+, T+, and Xn+ as the hydride acceptors, respectively. Comparison of the present reactions with the reaction examples whose mechanisms are well-known, such as the reaction of 1 with a one-electron oxidant Fe(CN)6(-3), shows that the active site of 1 in the oxidation with IQ+ is at the 4-position on the dihydropyridine ring but that the active site of 1 in the oxidations with T+ and Xn+ is at the 1-position, which is in agreement with the results from the Br?nsted-type linear analysis and the relation studies of the logarithm of the second-order rate constants with the oxidation potentials of the hydride donors. According to the dependence of the reaction mechanism on the active site of 1, a conclusion can be made that the reaction of 1 with IQ+ proceeds by direct one-step hydride transfer mechanism, but the reactions of 1 with T+ and Xn+ would take place via multistep hydride transfer mechanism initiated by one-electron transfer.  相似文献   

8.
Stereochemistry of reductions of a pyridinium and quinoliniums with an asymmetric sulfinyl group has been studied. Dithionite and borohydride prefer the face characterized by a lone pair on sulfur atom. Dihydropyridine prefers the face characterized by a S-O bond.  相似文献   

9.
[reaction: see text] Data on the apparent dipole moment of thianthrene-5-oxide (1) and (1)H NMR spectra in different solvents support the conformational mobility of 1, which flaps between two limit boat conformations with the sulfinyl group in pseudoequatorial and pseudoaxial positions, respectively. The conformational equilibrium of 1 occurs too fast for the (1)H NMR (500 MHz) time-scale even at -130 degrees C, and the equilibrium constant has not been determined. The apparent dipole moments of 1 in n-hexane and 1,4-dioxane and the (1)H NMR spectra of 1 and the model compounds cis- and trans-thianthrene-5,10-dioxides (2) and thianthrene (5) in different solvents and at various temperatures confirm that the relative position of the conformational equilibrium of 1 is solvent-dependent, and more polar solvents favor the conformation with the sulfoxide group in the pseudoaxial position (1(')(ax)). Variable-temperature (1)H NMR spectra have established the interconversion barrier of trans-2 and confirmed that the conformational equilibrium of cis-2 is strongly displaced toward the conformation with both sulfinyl groups in the pseudoequatorial position. The (1)H NMR data support the transannular interaction of the functional groups in 1 and trans-2.  相似文献   

10.
报道了5种N-芳基芴亚胺在酸性条件下被烟酰胺辅酶模型(Hantzsch酯,BNAH)还原的反应。结果表明:亚胺的结构、酸的强度以及溶剂的不同均会影响亚胺的还原效率,本文结合反应的结构效应、溶剂效应和同位素效应,对其可能的酸催化氢负离子转移机理进行了讨论。  相似文献   

11.
对Hantzsch酯在Mg2+存在和不存在的情况下还原N-芳基芴亚胺的反应进行了研究, 并与BNAH的类似还原做了系统的比较.研究结果表明:Mg2+在该还原反应中起亲电催化剂的作用;还原能力较BNAH弱的Hantzsch酯在反应中所呈现的强的反应性是由于其3, 5-位两个极性谈基氧通过静电作用降低过渡态的能量的缘故;本文反应届H-一步转移机理.  相似文献   

12.
A new and easily regenerable NAD(P)H model 9,10-dihydrophenanthridine (DHPD) has been designed for biomimetic asymmetric hydrogenation of imines and aromatic compounds. This reaction features the use of hydrogen gas as terminal reductant for the regeneration of the DHPD under the mild condition. Therefore, the substrate scope is not limited in benzoxazinones; the biomimetic asymmetric hydrogenation of benzoxazines, quinoxalines, and quinolines also gives excellent activities and enantioselectivities. Meanwhile, an unexpected reversal of enantioselectivity was observed between the reactions promoted by the different NAD(P)H models, which is ascribed to the different hydride transfer pathway.  相似文献   

13.
Abstract

The Thio-ARBUZOV-Reactions has proved to be a facile synthetic route for the conversion of alkoxysulfanes into sulfinyl derivatives. Scope and limitations of the reaction are discussed. BEcause of their donor properties the alkoxysulfanes are suitable ligands which has been confirmed by the preparatoin of palladium(II) complexes.  相似文献   

14.
While the sonochemical grafting of molecules on silicon hydride surface to form stable Si–C bond via hydrosilylation has been previously described, the susceptibility towards nucleophilic functional groups during the sonochemical reaction process remains unclear. In this work, a competitive study between a well-established thermal reaction and sonochemical reaction of nucleophilic molecules (cyclopropylamine and 3-Butyn-1-ol) was performed on p-type silicon hydride (111) surfaces. The nature of surface grafting from these reactions was examined through contact angle measurements, X-ray photoelectron spectroscopy (XPS) and atomic force microscopy (AFM). Cyclopropylamine, being a sensitive radical clock, did not experience any ring-opening events. This suggested that either the Si–H may not have undergone homolysis as reported previously under sonochemical reaction or that the interaction to the surface hydride via a lone-pair electron coordination bond was reversible during the process. On the other hand, silicon back-bond breakage and subsequent surface roughening were observed for 3-Butyn-1-ol at high-temperature grafting (≈150 °C). Interestingly, the sonochemical reaction did not produce appreciable topographical changes to surfaces at the nano scale and the further XPS analysis may suggest Si–C formation. This indicated that while a sonochemical reaction may be indifferent towards nucleophilic groups, the surface was more reactive towards unsaturated carbons. To the best of the author’s knowledge, this is the first attempt at elucidating the underlying reactivity mechanisms of nucleophilic groups and unsaturated carbon bonds during sonochemical reaction of silicon hydride surfaces.  相似文献   

15.
Density functional theory calculations (B3LYP) have been carried out to understand the mechanism and stereochemistry of an asymmetric reductive aldol reaction of benzaldehyde and tert-butyl acrylate with hydrosilanes catalyzed by Rh(Phebox-ip)(OAc)(2)(OH(2)). According to the calculations, the reaction proceeds via five steps: (1) oxidative addition of hydrosilane, (2) hydride migration to carbon-carbon double bond of tert-butyl acrylate, which determines the chirality at C2, (3) tautomerization from rhodium bound C-enolate to rhodium bound O-enolate, (4) intramolecular aldol reaction, which determines the chirality at C3 and consequently the anti/syn-selectivity, and (5) reductive elimination to release aldol product. The hydride migration is the rate-determining step with a calculated activation energy of 23.3 kcal mol(-1). In good agreement with experimental results, the formation of anti-aldolates is found to be the most favorable pathway. The observed Si-facial selectivity in both hydride migration and aldol reaction are well-rationalized by analyzing crucial transition structures. The Re-facial attack transition state is disfavored because of steric hindrance between the isopropyl group of the catalyst and the tert-butyl acrylate.  相似文献   

16.
Abstract

Some β-keto sulfoxides containing different groups at the sulfinyl moiety were submitted to sulfenylation reaction employing two procedures: in homogeneous media and by PTC to give the corresponding monosulfenylated products. The Z configuration for the enolates of the β-keto sulfoxides independently of the group at the sulfur moiety is proved by 1H NMR NOE difference experiments. The diastereoselectivity observed in the case of a t-butyl sulfinyl derivative is rationalized by preferred attack of the sulfur electrophile to the unhindered face of enolate not shielded by t-butyl group, as shown by X-ray analysis of the corresponding enol ether.  相似文献   

17.
Structures of the iridoid synthase nepetalactol synthase in the presence of NAD+, NADPH or NAD+/10‐oxogeranial were solved. The 10‐oxogeranial substrate binds in a transoid‐O1‐C3 conformation and can be reduced by hydride addition to form the byproduct S‐10‐oxo‐citronellal. Tyr178 Oζ is positioned 2.5 Å from the substrate O1 and provides the second proton required for reaction. Nepetalactol product formation requires rotation about C1–C2 to form the cisoid isomer, leading to formation of the cis‐enolate, together with rotation about C4–C5, which enables cyclization and lactol production. The structure is similar to that of progesterone‐5β‐reductase, with almost identical positioning of NADP, Lys146(147), Tyr178(179), and F342(343), but only Tyr178 and Phe342 appear to be essential for activity. The transoid 10‐oxogeranial structure also serves as a model for β‐face hydride attack in progesterone 5β‐reductases and is of general interest in the context of asymmetric synthesis.  相似文献   

18.
Structures of the iridoid synthase nepetalactol synthase in the presence of NAD+, NADPH or NAD+/10‐oxogeranial were solved. The 10‐oxogeranial substrate binds in a transoid‐O1‐C3 conformation and can be reduced by hydride addition to form the byproduct S‐10‐oxo‐citronellal. Tyr178 Oζ is positioned 2.5 Å from the substrate O1 and provides the second proton required for reaction. Nepetalactol product formation requires rotation about C1–C2 to form the cisoid isomer, leading to formation of the cis‐enolate, together with rotation about C4–C5, which enables cyclization and lactol production. The structure is similar to that of progesterone‐5β‐reductase, with almost identical positioning of NADP, Lys146(147), Tyr178(179), and F342(343), but only Tyr178 and Phe342 appear to be essential for activity. The transoid 10‐oxogeranial structure also serves as a model for β‐face hydride attack in progesterone 5β‐reductases and is of general interest in the context of asymmetric synthesis.  相似文献   

19.
Gold(I)‐chloride‐catalyzed synthesis of α‐sulfenylated carbonyl compounds from propargylic alcohols and aryl thiols showed a wide substrate scope with respect to both propargylic alcohols and aryl thiols. Primary and secondary aromatic propargylic alcohols generated α‐sulfenylated aldehydes and ketones in 60–97 % yield. Secondary aliphatic propargylic alcohols generated α‐sulfenylated ketones in yields of 47–71 %. Different gold sources and ligand effects were studied, and it was shown that gold(I) chloride gave the highest product yields. Experimental and theoretical studies demonstrated that the reaction proceeds in two separate steps. A sulfenylated allylic alcohol, generated by initial regioselective attack of the aryl thiol on the triple bond of the propargylic alcohol, was isolated, evaluated, and found to be an intermediate in the reaction. Deuterium labeling experiments showed that the protons from the propargylic alcohol and aryl thiol were transferred to the 3‐position, and that the hydride from the alcohol was transferred to the 2‐position of the product. Density functional theory (DFT) calculations showed that the observed regioselectivity of the aryl thiol attack towards the 2‐position of propargylic alcohol was determined by a low‐energy, five‐membered cyclic protodeauration transition state instead of the strained, four‐membered cyclic transition state found for attack at the 3‐position. Experimental data and DFT calculations supported that the second step of the reaction is initiated by protonation of the double bond of the sulfenylated allylic alcohol with a proton donor coordinated to gold(I) chloride. This in turn allows for a 1,2‐hydride shift, generating the final product of the reaction.  相似文献   

20.
Organometallic complexes of the general formula [(η(6)-arene)Ru(N?N)Cl](+) and [(η(5)-Cp*)Rh(N?N)Cl](+) where N?N is a 2,2'-dipyridylamine (DPA) derivative carrying a thiol-targeted maleimide group, 2,2'-bispyridyl (bpy), 1,10-phenanthroline (phen) or ethylenediamine (en) and arene is benzene, 2-chloro-N-[2-(phenyl)ethyl]acetamide or p-cymene were identified as catalysts for the stereoselective reduction of the enzyme cofactors NAD(P)(+) into NAD(P)H with formate as a hydride donor. A thorough comparison of their effectiveness towards NAD(+) (expressed as TOF) revealed that the Rh(III) complexes were much more potent catalysts than the Ru(II) complexes. Within the Ru(II) complex series, both the N?N and arene ligands forming the coordination sphere had a noticeable influence on the activity of the complexes. Covalent anchoring of the maleimide-functionalized Ru(II) and Rh(III) complexes to the cysteine endoproteinase papain yielded hybrid metalloproteins, some of them displaying formate dehydrogenase activity with potentially interesting kinetic parameters.  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