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 531 毫秒
1.
Oxidation of the trans-[Cr(cyca)(OH)2]+ complex, where cyca = meso-5,5,7,12,12,14-hexamethyl-1,4,8,11-tetraazacyclotetradecane, by [Fe(CN)6 ]3- ion in strongly alkaline media, leading to [CrV O(cycaox )]3+ ion, has been studied using electronic and e.p.r. spectroscopy. The kinetics of the CrIII → CrIV transformation have been studied using a large excess of the reductant and OH- ion over the oxidant. The reaction is a second order process: first order in [CrIII] and [FeIII] at constant [OH-]. The second order rate constant is higher than linearly dependent on the OH- concentration. The mechanism of the reaction has been discussed. A relatively inert intermediate chromium(V) species was detected based on characteristic bands in the visible region and the e.p.r. signal at giso = 1.987 for the systems where an excess of oxidant was used. The hyperfine structure of the main e.p.r. signal is consistent with the d1 -electron interactions with four equivalent nitrogen nuclei and [CrV = O(cycaox)]3+ formula, where cycaox = oxidized cyca, can be postulated for the intermediate CrV complex.  相似文献   

2.
The stereoselective Rauhut–Currier (RC) reaction catalyzed by a cysteine derivative has been explored computationally with density functional theory (M06‐2X). Both methanethiol and a chiral cysteine derivative were studied as nucleophiles. The complete reaction pathway involves rate‐determining elimination of the thiol catalyst from the Michael addition product. The stereoselective Rauhut–Currier reaction, catalyzed by a cysteine derivative as a nucleophile, has also been studied in detail. This reaction was experimentally found to be extremely sensitive to the reaction conditions, such as the number of water equivalents and the effect of potassium counterion. The E1cB process for catalyst elimination has been explored computationally for the eight possible stereoisomers. The effect of explicit water solvation and the presence of counterion (either K+ or Na+) has been studied for the lowest energy enantiomer pair (1S, 2R, 3S)/(1R, 2S, 3R).  相似文献   

3.
The kinetics and mechanisms of the copper(II)‐catalyzed GSH (glutathione) oxidation are examined in the light of its biological importance and in the use of blood and/or saliva samples for GSH monitoring. The rates of the free thiol consumption were measured spectrophotometrically by reaction with DTNB (5,5′‐dithiobis‐(2‐nitrobenzoic acid)), showing that GSH is not auto‐oxidized by oxygen in the absence of a catalyst. In the presence of Cu2+, reactions with two timescales were observed. The first step (short timescale) involves the fast formation of a copper–glutathione complex by the cysteine thiol. The second step (longer timescale) is the overall oxidation of GSH to GSSG (glutathione disulfide) catalyzed by copper(II). When the initial concentrations of GSH are at least threefold in excess of Cu2+, the rate law is deduced to be ?d[thiol]/dt=k[copper–glutathione complex][O2]0.5[H2O2]?0.5. The 0.5th reaction order with respect to O2 reveals a pre‐equilibrium prior to the rate‐determining step of the GSSG formation. In contrast to [Cu2+] and [O2], the rate of the reactions decreases with increasing concentrations of GSH. This inverse relationship is proposed to be a result of the competing formation of an inactive form of the copper–glutathione complex (binding to glutamic and/or glycine moieties).  相似文献   

4.
Kinetic analyses were carried out for formation of disulfide crosslinkages between thiol groups on linear polymers, poly(acrylamide‐coN‐acrylcysteamine) (P‐SH). Disulfide crosslinkages were formed by auto‐oxidation of pendant thiol groups or through the thiol‐disulfide exchange reaction induced by addition of disulfide compounds gluthathione. In the auto‐oxidation reaction, the rate constant for disulfide formation highly depended on pH values of the reaction mixtures and the P‐SH concentrations. Gelation rate is too slow to enclose living cells into hydrogel under physiological pH 7.4. The hydrogel formation rate can be accelerated by addition of disulfides, such as oxidized glutathione. In the later case, oxygen in the reaction mixture is not consumed. The thiol‐disulfide exchange reaction is much more suitable for the cell encapsulation than the thiol auto‐oxidation reaction. These findings give a basis for enclosure of living cells in a hydrogel. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2011  相似文献   

5.
An iron(III)–catecholate complex [L1FeIII(DBC)] ( 2 ) and an iron(II)–o‐aminophenolate complex [L1FeII(HAP)] ( 3 ; where L1=tris(2‐pyridylthio)methanido anion, DBC=dianionic 3,5‐di‐tert‐butylcatecholate, and HAP=monoanionic 4,6‐di‐tert‐butyl‐2‐aminophenolate) have been synthesised from an iron(II)–acetonitrile complex [L1FeII(CH3CN)2](ClO4) ( 1 ). Complex 2 reacts with dioxygen to oxidatively cleave the aromatic C? C bond of DBC giving rise to selective extradiol cleavage products. Controlled chemical or electrochemical oxidation of 2 , on the other hand, forms an iron(III)–semiquinone radical complex [L1FeIII(SQ)](PF6) ( 2ox‐PF6 ; SQ=3,5‐di‐tert‐butylsemiquinonate). The iron(II)–o‐aminophenolate complex ( 3 ) reacts with dioxygen to afford an iron(III)–o‐iminosemiquinonato radical complex [L1FeIII(ISQ)](ClO4) ( 3ox‐ClO4 ; ISQ=4,6‐di‐tert‐butyl‐o‐iminobenzosemiquinonato radical) via an iron(III)–o‐amidophenolate intermediate species. Structural characterisations of 1 , 2 , 2ox and 3ox reveal the presence of a strong iron? carbon bonding interaction in all the complexes. The bond parameters of 2ox and 3ox clearly establish the radical nature of catecholate‐ and o‐aminophenolate‐derived ligand, respectively. The effect of iron? carbon bonding interaction on the dioxygen reactivity of biomimetic iron–catecholate and iron–o‐aminophenolate complexes is discussed.  相似文献   

6.
LAC (hydroxylactone (1R,5S)‐1‐hydroxy‐3,6‐dioxabicyclo[3.2.1]octan‐2‐one) is one of the most interesting products of the pyrolysis of cellulose and represents a useful chiral building block in organic synthesis. A computational investigation at the DFT level on the mechanism of formation of LAC shows that this species can be obtained following two reaction paths, path A and path B , starting from a well‐known pyrolysis product (ascopyrone P). A series of internal rearrangements involving in all cases a proton transfer leads directly to LAC ( path B ). An alternative path ( path A ) can be also followed. From this path, via a “gate” connecting the two reaction channels, it is possible to reach path B and form LAC. In both cases, the rate‐determining step of the process is the initial keto‐enol isomerization. We found that water, which is present in the reaction mixture, “catalyzes” the reaction by assisting the proton transfers present in all the steps of the process. In particular, water lowers the barrier of the rate‐determining step that becomes 40.9 kcal mol?1 (79.4 kcal mol?1 in the absence of water). The corresponding computed rate constant is 4.3×10 s?1 at 500 °C, a value which is consistent with the presence of LAC in the absence of metal catalysts. The results of this study on the non‐catalyzed process underpin the important role played by water in the formation of pyrolysis products of cellulose where proton transfer is a key mechanistic step.  相似文献   

7.
We present here a study of the collision induced dissociation (CID) of deprotonated cysteic acid containing peptides produced by MALDI. The effect of cysteic acid (Cox) position is interrogated by considering the positional isomers, CoxLVINVLSQG, LVINVLSQGCox, and LVINVCoxLSQG. Although considerable variation between the CID spectra is observed, the mechanistic picture that emerges involves charge retention at the deprotonated cysteic acid side chain. Fragmentation occurs in the proximity of the cysteic acid group by charge directed mechanisms as well as remote from this group to form ions, which may be rationalized by charge remote mechanisms. Additionally, the formation of the SO3–• ion is observed in all cases. Fragmentation of CoxLVINVLSQCox provides both N- and C-terminal, y and b ions, respectively indicating that the negative charge may be retained at either of the cysteic acids; however, there is some evidence that charge retention at the C-terminal cysteic acid may be preferred. Fragmentation of tryptic type peptides containing a C-terminal arginine or lysine residue is considered through comparison of three peptides CoxLVINKLSQG, CoxLVINVLSQK, and CoxLVINVLSQR. Lastly, we rationalize the formation of b n–1 + H2O and a n–1 ions through a mechanism involving rearrangement of the C-terminal residue to form a mixed anhydride intermediate.  相似文献   

8.
A systematic search for reaction pathways for the vinylogous Mannich‐type reaction was performed by the artificial force induced reaction method. This reaction affords δ‐amino‐γ‐butenolide in one pot by mixing 2‐trimethylsiloxyfuran, imine, and water under solvent‐free conditions. Surprisingly, the search identified as many as five working pathways. Among them, two concertedly produce anti and syn isomers of the product. Another two give an intermediate, which is a regioisomer of the main product. This intermediate can undergo a retro‐Mannich reaction to give a pair of intermediates: an imine and 2‐furanol. The remaining pathway directly generates this intermediate pair. The imine and 2‐furanol easily react with each other to afford the product. Thus, all of these stepwise pathways finally converge to give the main product. The rate‐determining step of all five (two concerted and three stepwise) pathways have a common mechanism: concerted Si? O bond formation through the nucleophilic attack of a water molecule on the silicon atom followed by proton transfer from the water molecule to the imine. Therefore, these five pathways have comparable barriers and compete with each other.  相似文献   

9.
Protein disulfide isomerase (PDI) and glutathione peroxidase 7 (GPx7) cooperatively promote the oxidative folding of disulfide (SS)‐containing proteins in endoplasmic reticulum by recognizing the nascent proteins to convert them into the native folds by means of SS formation and SS isomerization and by catalyzing reoxidation of reduced PDI with H2O2, respectively. In this study, new amphiphilic selenides with a long‐chain alkyl group were designed as hybrid mimics of PDI and GPx7 and were applied to the refolding of reduced hen egg‐white lysozyme (HEL‐R). Competitive SS formation at pH 4 using HEL‐R and glutathione (GSH) in the presence of the selenide catalyst and H2O2 showed that the amphiphilic selenides can preferentially catalyze SS formation of HEL‐R, probably on account of hydrophobic interactions between the protein and the catalyst. In contrast, simple water‐soluble selenides did not exhibit such behavior. In addition, when the pH of the solution was adjusted to 8.5 after the SS formation, surviving GSH promoted the SS isomerization of misfolded HEL to recover the native SS linkages. Thus, the amphiphilic selenides designed here could mimic the function of the PDI‐GPx7 system. The combination of a water‐soluble selenide and a long‐chain alkyl group would be a useful motif in designing medicines for both protein misfolding diseases and antioxidant therapy.  相似文献   

10.
The peroxyoxalate reaction is a highly efficient chemiluminescence system, its chemiexcitation process involving the intermolecular interaction between an activator (ACT) and the high‐energy intermediate (HEI) of the reaction. Typically, the HEI is generated through the reaction of an oxalate ester with H2O2, in the presence of a basic/nucleophilic catalyst, such as imidazole (IMI‐H). IMI‐H, besides catalyzing the formation of the HEI, is also known to decompose this peroxidic intermediate. Despite that, up to now, no rate constant value has been determined for such significant interaction. Through kinetic measurements, we have observed that IMI‐H is roughly four times more efficient than 9,10‐diphenylanthracene (DPA), a classic ACT, in catalyzing the decomposition of the HEI by a bimolecular electron transfer reaction through a Chemically Initiated Electron Exchange Luminescence‐like process. For instance, when IMI‐H and DPA are at the same concentration, 78% of the generated HEI is actually consumed by the nonemissive bimolecular interaction with IMI‐H. We have obtained an average singlet excited state formation quantum yield, at infinite ACT concentration, of (5.5 ± 0.5) × 10?2 E mol?1, determined at five different IMI‐H concentrations. This ultimately suggests that the yield of formation of HEI actually does not depend on the IMI‐H concentration.  相似文献   

11.
The hairy poly(methacrylic acid‐co‐divinylbenzene)‐g‐poly(N‐isopropylacrylamide) (P(MAA‐co‐DVB)‐g‐PNIPAm) nanocapsules with pH‐responsive P(MAA‐co‐DVB) inner shell and temperature‐responsive PNIPAm brushes were prepared by combined distillation–precipitation copolymerization and surface thiol‐ene click grafting reaction using 3‐(trimethoxysilyl)propyl methacrylate‐modified silica (SiO2‐MPS) nanospheres as a sacrificial core material. The well‐defined PNIPAm was synthesized by a reversible addition fragmentation chain transfer (RAFT) polymerization. The chain end was converted to a thiol by chemical reduction. The PNIPAm was integrated into the nanocapsules via thiol‐ene click reaction. The surface thiol‐ene click reaction conduced to tunable grafting density of PNIPAm brushes. The grafting densities decreased from 0.70 chains nm?2 to 0.15 chains nm?2 with increasing the molecular weight of grafted PNIPAm chains. Using water soluble doxorubicin hydrochloride (DOX·HCl) as a model molecular, the tunable shell permeability of the nanocapsule was investigated in detail. The permeability constant can be tuned by controlling the thickness of the P(MAA‐co‐DVB) inner shell, the grafting density of PNIPAm brushes, and the environmental pH and temperature. The tunable shell permeability of these nanocapsules results in the release of the loaded guest molecules with manipulable releasing kinetics. © 2014 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2014 , 52, 2202–2216  相似文献   

12.
A new solid substrate-room temperature phosphorescence (SS-RTP) quenching method for the determination of trace As(V) has been developed, based on the facts that 9-hydroxy-2,3,4,9-tetrahydro-1,10-anthraquinone (R) can emit intense and stable SS-RTP on solid substrate, and α,α'-dipyridyl can activate As(V) catalysis of the reaction of H2O2 oxidizing R to non-phosphorescence compound R', which can cause the sharp quenching of SS-RTP. Under the optimum condition, the relationship between the ΔIp of the emitting intensity and 1.60-160 fg·spot^-1 As(V) (corresponding concentration: 0.0040-0.40 ng·mL^-1, sample volume: 0.4 μL·spot^-1) conformed to Beer' law. The regression equation of working curve can be expressed as ΔIp= 20.46+0.5492CAs(v) fig·spot^-1) (r= 0.9995, n = 6). The limit detection (LD) is 0.27 fg·spot^-1 [As(V) corresponding concentration: 6.8 × 10^-13 g·mL^-1, n=11]. The samples containing 0.0040 and 0.40 ng·mL^-1 As(V) were repeatedly determined for 11 times. RSD are 3.0% and 2.7% respectively. The SS-RTP mechanism was also discussed. R was synthesized in this paper. Meanwhile, the structure was determined by NMR, IR, mass spectra and elemental analysis.  相似文献   

13.
In this study, a novel polymer electrolyte membrane, poly(vinylbenzyl sulfonic acid)‐grafted poly(tetrafluoroethylene‐co‐hexafluoropropylene) (FEP‐g‐PVBSA), has been successfully prepared by simultaneous irradiation grafting of vinylbenzyl chloride (VBC) monomer onto a FEP film and taking subsequent chemical modification steps to modify the benzyl chloride moiety to the benzyl sulfonic acid moiety. The chemical reactions for the sulfonation were carried out via the formation of thiouronium salt with thiourea, base‐catalyzed hydrolysis for the formation of thiol, and oxidation with hydrogen peroxide. Each chemical conversion process was confirmed by FTIR, elemental analysis, and SEM‐EDX. A chemical stability study performed with Fenton's reagent (3% H2O2 solution containing 4 ppm of Fe2+) at 70 °C revealed that FEP‐g‐PVBSA has a higher chemical stability than the poly(styrene sulfonic acid)‐grafted membranes (FEP‐g‐PSSA). © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 48: 563–569, 2010  相似文献   

14.
With the aid of density functional theory (DFT) calculations, we have investigated the mechanisms and stereoselectivities of the tandem cross Rauhut–Currier/cyclization reaction of methyl acrylate R1 with (E)‐2‐benzoyl‐3‐phenyl‐acrylonitrile R2 catalyzed by a tertiary amine DABCO. The results of the DFT calculations indicate that the favorable mechanism (mechanism A) includes three steps: the first step is the nucleophilic attack of DABCO on R1 to form intermediates Int1 and Int1‐1, the second step is the reaction of Int1 and Int1‐1 with R2 to generate intermediate Int2(SS,RR,SR&RS), and the last step is an intramolecular SN2 process to give the final product P(SS,RR,SR&RS) and release catalyst DABCO. The SN2 substitution is computed to be the rate‐determining step, whereas the second step is the stereoselectivity‐determining step. The present study may be helpful for understanding the reaction mechanism of similar tandem reactions.  相似文献   

15.
α‐Ferrocenylalkylation of azoles or S‐nucleophiles with FcCH(R)OH (Fc = ferrocenyl) can be accomplished under acid‐free conditions as one‐pot process via an intermediate formation of the α‐ferrocenylalkyl carbonates FcCH(R)OC(O)OEt. The reaction allows the alkylation of acid sensitive substrates like imidazole derivatives or sodium N,N‐diethyldithiocarbamate. The reaction with ambident azoles proceeds as the N‐alkylation. Some α‐ferrocenylalkyl azoles were found to exhibit plant growth stimulating or herbicidal effects on corn seeds or act as the herbicide safeners against sulfonylurea herbicides.  相似文献   

16.
17.
The catalytic reactivity of the high‐spin MnII pyridinophane complexes [(Py2NR2)Mn(H2O)2]2+ (R=H, Me, tBu) toward O2 formation is reported. With small macrocycle N‐substituents (R=H, Me), the complexes catalytically disproportionate H2O2 in aqueous solution; with a bulky substituent (R=tBu), this catalytic reaction is shut down, but the complex becomes active for aqueous electrocatalytic H2O oxidation. Control experiments are in support of a homogeneous molecular catalyst and preliminary mechanistic studies suggest that the catalyst is mononuclear. This ligand‐controlled switch in catalytic reactivity has implications for the design of new manganese‐based water oxidation catalysts.  相似文献   

18.
Semiempirical calculations were used to conduct a comprehensive study of the thiolysis of the fundamental core of penicillins and cephalosporins. The significance of the intramolecular protonation of the β‐lactam nitrogen in the formation and cleavage of the tetrahedral intermediate ( T in Scheme 1 ) was examined in two thiols bearing substituents of different basicity in β with respect to the thiol group in the attacking nucleophile, namely 2‐mercaptoethanol ( 6 ) and 2‐mercaptoethylamine ( 7 ). Based on the results, the rate‐determining step in the reaction of penicillins is the cleavage of the tetrahedral intermediate, consistent with an intramolecular acid catalysis process in their thiolysis by 2‐mercaptoethylamine. On the other hand, the rate‐determining step in the reaction of cephalosporins, which possess an appropriate leaving group at position 3', is the formation of the tetrahedral intermediate, so the desolvation energy of the nucleophile is a major contributor to the overall energy of the process. This differential behavior between the two types of β‐lactam bicycles arises from the presence of the acetate group at 3' and the delocalization of π electrons over the N5–C4–C3 system in cephalosporins; this favors the formation of a thiolate with the 5‐ethoxymethylene‐1,3‐thiazine group in the cleavage of the tetrahedral intermediate, which is stabilized by an intramolecular hydrogen bond between N5 and the alcohol or amine group in β of the attacking thiol. The theoretical results are consistent with previous experimental data showing that, unlike penicillins, cephalosporins undergo no intramolecular acid catalysis in their thiolysis. © 2005 Wiley Periodicals, Inc. Int J Chem Kinet 37: 434–443, 2005  相似文献   

19.
Exciplexes of 9,10‐dicyanoanthracene (DCA) with alkylbenzene donors in cyclohexane show structureless emission spectra, typical of exciplexes with predominantly charge‐transfer (CT) character, when the donor has a relatively low oxidation potential (Eox), e.g. hexamethylbenzene (HMB). With increasing Eox and stronger mixing with a locally excited (LE) state, vibrational structure begins to appear with 1,2,3,5‐tetramethylbenzene and becomes prominent with p‐xylene (p‐Xy). A simple theoretical model reproduces the spectra and the radiative rate constants, and it reveals several surprises: Even in this nonpolar solvent, the fractional CT character of a highly mixed exciplex varies widely in response to fluctuations in the microscopic environment. Environments that favor the LE (or CT) state contribute more to the blue (or red) side of the overall spectrum. It is known that sparsely substituted benzene radical cations, e.g., p‐Xy?+, are stabilized more in acetonitrile than the heavily substituted HMB?+. Remarkably, ion pairing with DCA?– in cyclohexane leads to even larger differences in the stabilization of these radical cations. The spectra of the low‐Eox donors are almost identical except for displacements that approximately equal the differences in Eox, even though the exciplexes have varying degrees of CT character. These similarities result from compensation among several nonobvious, but quantified factors.  相似文献   

20.
Polysiloxanes are commonly used in a myriad of applications, and the “click” nature of the thiol‐ene reaction is well suited for introducing alternative functionalities or for crosslinking these ubiquitous polymers. As such, understanding of the thiol‐ene reaction in the presence of silicones is valuable and would lead to enhanced methodologies for modification and crosslinking. Here, the thiol‐ene reaction kinetics were investigated in functionalized oligosiloxanes having varying degrees of thiol functionalization (SH), π–π interactions (from diphenyls, DP), and ene types (C?C). In the ene‐functionalized oligomers, π–π interactions were controlled through the use of dioctyl repeats (DO). The polymerization rate and rate‐limiting steps were determined for all systems containing an allyl‐functionalized oligomer, and rates ranging from 0.10 to 0.54 mol L?1 min?1 were seen. The rate‐limiting step varied with the oligomer composition; examples of rate‐limited propagation (5:3:2 C?C:DP:DO/1:1 SH:DP) or chain transfer (5:3:2 C?C:DP:DO/3:1 SH:DP) were found in addition to cases with similar reaction rate constants (5:2:3 C?C:DP:DO/1:1 SH:DP). None of the siloxanes were found to exhibit autoacceleration despite their relatively high viscosities. Instead, the allyl‐, vinyl‐, and acrylate‐functionalized siloxanes were all found to undergo unimolecular termination based on their high α scaling values (0.98, 0.95, and 0.82, respectively) in the relation RpRiα. © 2013 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2013  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