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 62 毫秒
1.
The sorption of Cu, Pb, and Zn onto natural lake particles, suspended in 0.005 M NaNO3 solution and in a natural lake water (Esthwaite Water, Cumbria, UK), was studied as a function of pH and time in a series of laboratory experiments, under environmentally realistic conditions. The sorption of all three metals increased with increasing pH and reaction time (2 h and 7 days). In 0.005 M NaNO3 solution, the well-defined sorption edges spanned 2–2.5 pH units for Cu and Pb, and ≈ 4 pH units for Zn. In the natural lake water, the Cu sorption edge was broader and both Cu and Zn were less strongly sorbed. The binding stability decreased in the order Pb>Cu>Zn. Competitive adsorption onto surface sites appeared to be the main factor determining the observed sorption behaviour. Application of a macroscopic metal exchange model to the 7 day NaNO3 results enabled the surface site concentration to be estimated as 0.79 ± 0.07 mmol g−1. The modelling exercise suggested that an observed shift in the sorption edge of Zn, in the presence of Pb and Cu, was due to competition for surface sites. The experimental data are in good general agreement with field observations of trace metal behaviour in Cumbrian lakes. The almost total sorption of Pb by lake particles throughout the in-situ pH range is compatible with previous field measurements including trace metal budgets and residence times. Dissolved Zn concentrations in the lake are lower than predicted by the sorption experiments, but the lower lake concentrations are consistent with the previously observed scavenging of Zn by planktonic algae. Both the decreased sorption of Cu in the experiments with natural lake water, compared to that in NaNO3 solution, and the relatively small-scale removal of dissolved Cu by particles in the lake itself can partially be explained by humic complexation.  相似文献   

2.
The cyclization of the RNA model 2-hydroxypropyl p-nitrophenyl phosphate (HPNPP, 1) promoted by Zn2+ alone and the 1,5,9-triazacyclododecane complex of Zn2+ (Zn2+:[12]aneN3) is studied in ethanol in the presence of 0.5 equiv of -OEt/Zn2+ to investigate the effect of a low polarity/dielectric medium on a metal-catalyzed reaction of biological relevance. Ethanol exerts a medium effect that promotes strong binding of HPNPP to Zn2+, followed by a dimerization to form a catalytically active complex (HPNPP:Zn2+)2 in which the phosphate undergoes cyclization with a rate constant of kcat = 2.9 s(-1) at s(s)pH 7.1. In the presence of the triaza ligand:Zn2+ complex, the change from water to methanol and then to ethanol brings about a mechanism where two molecules of the complex, suggested as EtOH:Zn2+:[12]aneN3 and its basic form, EtO-:Zn2+:[12]aneN3, bind to HPNPP and catalyze its decomposition with a rate constant of kcat of 0.13 s(-1) at s(s)pH 7.1. Overall, the acceleration exhibited in these two situations is 4 x 10(14)-fold and 1.7 x 10(12)-fold relative to the background ethoxide-promoted reactions at the respective s(s)pH values. The implications of these findings are discussed within the context of the idea that enzymatic catalysis is enhanced by a reduced effective dielectric constant within the active site.  相似文献   

3.
The collisionally activated mass spectral fragmentations of N-(2,4-dinitrophenyl)alanine and phenylalanine [M - H](-) may be gas-phase analogs of the base-catalyzed cyclization of N-(2,4-dinitrophenyl)amino acids in aqueous dioxane. This latter reaction is one source of the 2-substituted 5-nitro-1H-benzimidazole-3-oxides, which are antibacterial agents. The fragmentation of both compounds, established by tandem mass spectrometric experiments and supported by molecular modeling using DFT methods, indicate that the [M - H](-) ions dissociate via sequential eliminations of CO(2) and H(2)O to produce deprotonated benzimidazole-N-oxide derivatives. The gas-phase cyclization reactions are analogous to the base-catalyzed cyclization in solution, except that in the latter case, the reactant must be a dianion for the reaction to occur on a reasonable time scale. The cyclization of N-(2-nitrophenyl)phenylalanine, which has one less nitro group, requires a stronger base for the cyclization than the compound with a second nitro group at the 4-position. Following losses of CO(2) and H(2)O are expulsions of both neutral molecules and free radicals, the latter being examples of violations of the even-electron ion rule.  相似文献   

4.
The rhodium(II)-catalyzed reaction of -diazo ketones bearing tethered alkyne units represents a new and useful method for the construction of a variety of substituted cyclopentenones. The process proceeds by addition of the rhodium-stabilized carbenoid onto the acetylenic π-bond to give a vinyl carbenoid intermediate. The resulting rhodium complex undergoes a wide assortment of reactions including cyclopropanation, 1,2-hydrogen migration, CH-insertion, addition to tethered alkynes and ylide formation. The exact pathway followed is dependent on the specific metal/ligand employed and is also influenced by the nature of the solvent. Sulfonium ylide formation occurred both intra and intermolecularly when the reaction was carried out in the presence of a sulfide. In the case where an ether oxygen was present on the backbone of the vinyl carbenoid, cyclization afforded an oxonium ylide which underwent a [1,2] or [2,3]-sigmatropic shift to give a rearranged product. These cyclic metallocarbenoids were also found to interact with a neighboring carbonyl π-bond to produce carbonyl ylide dipoles that could be trapped with added dipolarophiles. The domino transformation was also performed intramolecularly by attaching an alkene directly to the carbonyl group. When 2-alkynyl-2-diazo-3-oxobutanoates were treated with a Rh(II)-catalyst, furo[3,4-c]furans were formed in excellent yield. The 1,5-electrocyclization process involved in furan formation has also been utilized to produce indeno[1,2-c]furans. Rotamer population was found to play a significant role in the cyclization of -diazo amide systems containing tethered alkynes. In this account, an overview of our work in this area is presented.  相似文献   

5.
Reaction mechanisms of the imide formation in an N-(o-carboxybenzoyl)-l-amino acid have been studied using density functional theory. Our results suggest that the reaction route initiated by protonation at the oxygen of the carboxyl group of the amino acid is favored, while those initiated by deprotonation at the oxygen of the carboxyl group of phthalic acid and at the amidic nitrogen are minor pathways. During the dehydration process, water functions as a catalyst. These conclusions are in good agreement with the experimental facts that at highly acidic conditions (hydrogen ion concentration H(0) < -1), imide formation is the most favorable pathway, whereas in the pH range 0-5, cyclization to the imide is not the dominant reaction. Our calculations also show that the carboxyl group of the amino acid is involved in the catalytic reaction in both the favored and minor pathways and that solvent effects have little influence on the reaction barriers.  相似文献   

6.
For dephosphorylation of bis(2,4-dinitrophenyl) phosphate (BDNPP) by hydroxylamine in water, pH region 4-12, the observed first-order rate constant, k(obs), initially increases as a function of pH, but is pH-independent between pH 7.2 and pH 10. The initial BDNPP cleavage by nonionic NH(2)OH (<0.2 M) involves attack by the OH group and follows first-order kinetics, but the overall initial reaction of BDNPP liberates ca. 1.7 mol of 2,4-dinitrophenoxide ion (DNP). This initial reaction generates a short-lived O-phosphorylated hydroxylamine, 2, followed by three possible reactions: (1) reaction of 2 with hydroxylamine, generating 2,4-dinitrophenyl phosphate (DNPP, 3), which subsequently forms DNP; (2) intramolecular displacement of the second DNP group and rapid decomposition of the cyclic intermediate to form phosphonohydroxylamine and eventually inorganic phosphate; (3) a novel rearrangement with intramolecular aromatic nucleophilic substitution involving a cyclic intermediate and migration of the 2,4-dinitrophenyl group from O to N. Values of k(obs) increase modestly with pH > 10, the reaction is biphasic, and the yield of DNP increases. An increase in [NH(2)OH] also increases the yield of DNP, due largely to accelerated hydrolysis of DNPP.  相似文献   

7.
The kinetics of formation of amide, 4, from the corresponding carboxylic acid by reaction with the isopropyl ester of methionine (MIPE), mediated by carbodiimide EDCI, 1, and HOBt, 2, have been studied in 1-methyl-2-pyrrolidinone (NMP) using reaction calorimetry. The reaction rates have been found to be independent of the concentration of HOBt, showing that the rate-determining step is the reaction between the carboxylic acid and EDCI to give the corresponding O-acylisourea. The pH dependence of the observed rate constants for O-acylisourea formation is consistent with a second-order reaction between doubly protonated EDCI (EDCIH2(2+), 6) and the carboxylate group. The observed rate constants fall sharply at high pH, as the fraction of EDCI as EDCIH2(2+) continues to fall strongly, whereas the carboxylic acid group is already fully ionized. The rate constant, kP, for reaction between the carboxylate group of acid, 3, and EDCIH2(2+) has a value of kP = 4.1 x 10(4) M(-1) s(-1) at 20 degrees C, some 10(5) times higher than similar rate constants measured in water. The subsequent catalytic cycle, involving reaction of O-acylisourea with HOBt to give HOBt ester, which then reacts with the amine to give the amide with regeneration of HOBt, determines the product distribution. In the case of the amino acid, 3, reaction of the O-acylisourea with MIPE to give amide, 4, is increasingly favored at higher pH values over that with the less basic internal aromatic amine of 3 to give the diamide 5.  相似文献   

8.
Oxidation of a thiophene‐hexapyrrane hybrid S‐P6 afforded a stable conjugated open‐chain thiaheptapyrrolic helix 1 with the terminal thiophene and confused pyrrole units lying at a long distance that is adverse for further cyclization. Chelation of 1 with copper(II) ion afforded 1‐Cu , which exhibits more distant terminal units. Interestingly, further oxidation of 1 triggered an intramolecular C?N fusion reaction to afford a unique 5,5,5‐tricyclic fused linear thiaheptapyrrin 2 , with the two terminals positioned in proximity, which favors the oxidative ring‐closure reaction to give a unique 5,5,5‐tricyclic fused thiaheptaphyrin (1.1.1.1.1.1.0) 3 under air. The inner‐fusion strategy for positioning the reactive sites in proximity to promote oxidative cyclization offers a new approach for constructing large porphyrinoids through conjugated oligopyrrins without the assistance of metal ions.  相似文献   

9.
The reaction mechanism of (E)‐methyl 3‐(2‐aminophenyl)acrylate ( A ) with phenylisothiocyanate ( B ) as well as the vital roles of substrate A and solvent water were investigated under unassisted, water‐assisted, substrate A ‐assisted, and water‐ A ‐assisted conditions. The reaction proceeds with four processes via nucleophilic addition, deprotonation and protonation, intramolecular cyclization with hydrogen transfer, and keto–enol tautomerization. According to the different H‐shift mode, two possible types of H‐shift P1 and P2 are carefully investigated to identify the most preferred pathway, differing in the ? NH2 group deprotonation and ? CH group of A protonation processes. It is found that substrate A and water not only act as reactant and solvent, but also as catalyst, proton shuttle, and stabilizer in effectively lowering the energy barrier. Therefore, the results demonstrate that the strong donating and accepting ability of ? NH2 group on A and the presence of bulk water are the keys to the title reaction proceed. © 2016 Wiley Periodicals, Inc.  相似文献   

10.
Plots of log k(0) vs pH for the cyclization of trifluoroethyl and phenyl 2-aminomethylbenzoate to phthalimidine at 30 degrees C in H(2)O are linear with slopes of 1.0 at pH >3. The values of the second-order rate constants k(OH) for apparent OH(-) catalysis in the cyclization reactions are 1.7 x 10(5) and 5.7 x 10(7) M(-)(1) s(-)(1), respectively. These rate constants are 10(5)- and 10(7)-fold greater than for alkaline hydrolysis of trifluoroethyl and phenyl benzoate. The k(OH) for cyclization of the methyl ester is 7.2 x 10(3) M(-)(1) s(-)(1). Bimolecular general base catalysis occurs in the intramolecular nucleophilic reactions of the neutral species. The value of the Bronsted coefficient beta for the trifluoroethyl ester is 0.7. The rate-limiting step in the general base catalyzed reaction involves proton transfer in concert with leaving group departure. The mechanism involving rate-determining proton transfer exemplified by the methyl ester in this series (beta = 1.0) can then be considered a limiting case of the concerted mechanism. General acid catalysis of the neutral species reaction or a kinetic equivalent also occurs when the leaving group is good (pK(a) 相似文献   

11.
A methoxide-bridged dinuclear Zn(II) complex of 1,3-[N,N'-bis(1,5,9-triazacyclododecane)]propane (1-Zn(II)2:(-OCH3)) was prepared, and its catalysis of the cyclization of a series of 2-hydroxypropyl aryl phosphates (4a-g) was investigated in methanol at pH 9.8, T = 25degreesC by stopped-flow spectrophotometry. An X-ray diffraction structure of the hydroxide analogue of 1-Zn(II)2:(-OCH3), namely 1-Zn(II)2:(-OH), reveals that each of the Zn(II) ions is coordinated by the three N's of the triazacyclododecane units and a bridging hydroxide. The cyclizations of substrates 4a-g reveal a progressive change in the observed kinetics from Michaelis-Menten saturation kinetics for the poorer substrates (4-OCH3 (4g); 4-H (4f); 3-OCH3 (4e); 4-Cl (4d); 3-NO2, (4c)) to second-order kinetics (linear in 1-Zn(II)2:(-OCH3)) for the better substrates (4-NO2,3-CH3 (4b); 4-NO2, (4a)). The data are analyzed in terms of a multistep process whereby a first formed complex rearranges to a reactive complex with a doubly activated phosphate coordinated to both metal ions. The kinetic behavior of the series is analyzed in terms of change in rate-limiting step for the catalyzed reaction whereby the rate-limiting step for the poorer substrates (4g-c) is the chemical step of cyclization of the substrate, while for the better substrates (4b,a) the rate-limiting step is binding. The catalysis of the cyclization of these substrates is extremely efficient. The kcat/KM values for the catalyzed reactions range from 2.75 x 10(5) to 2.3 x 10(4) M-1 s-1, providing an acceleration of 1 x 10(8) to 4 x 10(9) relative to the methoxide reaction (k2OCH3, which ranges from 2.6 x 10(-3) to 5.9 x 10(-6) M-1 s-1 for 4a-g). At a pH of 9.8 where the catalyst is maximally active, the acceleration for the substrates ranges from (1 - 4) x 10(12) relative to the background reaction at the same pH. Detailed energetics calculations show that the transition state for the catalyzed reaction comprising 1-Zn(II)2, methoxide, and 4 is stabilized by about -21 to -23 kcal/mol relative to the transition state for the methoxide reaction. The pronounced catalytic activity is attributed to a synergism between a positively charged catalyst that has high affinity for the substrate and for the transition state for cyclization, and a medium effect involving a reduced polarity/dielectric constant that complements a reaction where an oppositely charged reactant and catalyst experience charge dispersal in the transition state.  相似文献   

12.
Cyclization reactions with polyphosphoric acid of 3-aryl-3-hydroxypropionanilides carrying a p-nitro- 7a or p-amino substituent 10 on the C-3 phenyl group were investigated. In the case of p-nitro substitution the preferred reaction is, instead of cyclization, the elimination of water to give the corresponding cinnamic acid derivative. On the other hand, reaction of the p-amino-substituted analogue gave the new compound 4-(4-aminophenyl)-3,4-dihydroquinolin-2(1H)-one ( 2b ) in a good yield. The intermediate product of the cyclization was isolated and its structure established.  相似文献   

13.
Oligomers containing thianthrene units in the chain are synthesized by AlCl3-catalyzed reaction of diphenylsulfide, thianthrene, poly(p-phenylene sulfide), and poly(m-phenylene disulfide) with sulfur at 80°C. The products are compared with that obtained by the reaction of diphenylsulfide with AlCl3 at 225°C. IR spectra and elemental analyses are consistent with cyclic chain structures and show a higher cyclization for CHCl3-insoluble fractions. X-ray diffraction analysis indicates the same structure for the samples obtained with the different methods of synthesis. A possible structural model is tentatively proposed.  相似文献   

14.
The reaction of a 1,6-enyne with a hydrosilane catalyzed by Rh(acac)(CO)(2), Rh(4)(CO)(12), or Rh(2)Co(2)(CO)(12) under ambient CO atmosphere or N(2) gives 2-methyl-1-silylmethylidene-2-cyclopentane or its heteroatom congener in excellent yield through silylcarbocycization (SiCaC) process. The same reaction, but in the presence of a phosphite such as P(OEt)(3) and P(OPh)(3) under 20 atm of CO, affords the corresponding 2-formylmethyl-1-silylmethylidene-2-cyclopentane or its heteroatom congener with excellent selectivity through carbonylative silylcarbocycization (CO-SiCaC) process. The SiCaC reaction has also been applied to a 1,6-enyne bearing a cyclohexenyl group as the alkene moiety and a 1,7-enyne system. The functionalized five- and six-membered ring systems obtained by these novel cyclization reactions serve as useful and versatile intermediates for the syntheses of natural and unnatural heterocyclic and carbocyclic compounds. Possible mechanisms for the SiCaC and CO-SiCaC reactions as well as unique features of these processes are discussed.  相似文献   

15.
Xu J  Wang Y  Burton DJ 《Organic letters》2006,8(12):2555-2558
A novel cyclization reaction for the preparation of 3-fluoro-1-substituted-naphthalenes is reported. (E)-Monofluoroenynes, which are prepared by Sonogashira coupling reaction from (Z)-1-bromo-1-fluoroalkenes, undergo cyclization to afford 3-fluoro-1-substituted-naphthalenes in good to excellent yields when treated with DABCO or DBU in refluxing N-methyl-2-pyrrolidinone (NMP). [reaction: see text]  相似文献   

16.
The binding ability of a chiral L-cysteinato cobalt(III) complex, [Co(L-cys-N,S)(en)2]+ (l-H2cys = L-cysteine, en = ethylenediamine), toward a cadmium(II) center, together with the construction of S-bridged CoIIICdII structures that are controlled by anions and pH, is reported. The reaction of Lambda(L)-[Co(L-Hcys-N,S)(en)2](ClO4)2 having a pendent COOH group with CdCl2 in a 1:1 ratio in water, followed by the addition of NaCl, gave an S-bridged CoIIICdII dinuclear complex, Lambda(L)-[CdCl4{Co(L-Hcys-N,S)(en)2}] (1Cl), in which a cadmium(II) ion is weakly coordinated by a thiolato group from a Lambda(L)-[Co(L-Hcys-N,S)(en)2]2+ unit, besides four Cl- anions. The corresponding 1:1 reaction with CdBr2 and NaBr yielded an S-bridged CoIIICdIICoIII trinuclear complex composed of an S-bridged CoIIICdIICoIII trinuclear cation and a [CdBr4]2- anion, (Lambda(L))2-[CdBr3{Co(L-Hcys-N,S)(en)2}{Co(L-cys-N,S)(en)2} ][CdBr4] (2), while a CoIIICdII dinuclear complex analogous to 1Cl, Lambda(L)-[CdBr4{Co(L-Hcys-N,S)(en)2}] (1Br), was obtained by the addition of HBr instead of NaBr. In the CoIIICdIICoIII cation of 2, a CdII center is very weakly coordinated by two thiolato groups from Lambda(L)-[Co(L-Hcys-N,S)(en)2]2+ and Lambda(L)-[Co(L-cys-N,S)(en)2]+ units, besides three Br- anions, with the trinuclear structure being sustained by an intramolecular COOH...OOC hydrogen bond. On the other hand, no S-bridged structure was obtained by the corresponding 1:1 reaction with CdI2 and NaI, giving only a mononuclear CoIII species with a [CdI4]2- counteranion, Lambda(L)-[Co(L-Hcys-N,S)(en)2][CdI4] (3). When Lambda(L)-[Co(L-cys-N,S)(en)2]ClO4 having a deprotonated pendent COO- group was reacted with CdCl2 in a 1:1 ratio in water, followed by the addition of NaCl, a one-dimensional (CoIIICdII)n polymeric complex, (Lambda(L))n-[CdCl3{Co(L-cys-N,S)(en)2}]n (4Cl), in which Lambda(L)-[Co(L-cys-N,S)(en)2]+ units are alternately linked by [CdCl3]- moieties through thiolato and carboxylate groups, was constructed. An analogous (CoIIICdII)n polymeric structure having [Cd(NCS-N)3]- moieties, (Lambda(L))n-[Cd(NCS-N)3{Co(L-cys-N,S)(en)2}]n (4NCS), was also produced by the use of Cd(ClO4)2 and NaSCN.  相似文献   

17.
A series of 2-alkynyl 2-diazo-3-oxobutanoates, when treated with a catalytic quantity of rhodium(II) acetate, afforded furo[3,4-c]furans in good yield. The reaction proceeds by addition of a rhodium-stabilized carbenoid onto the acetylenic pi-bond to give a vinyl carbenoid that subsequently cyclizes onto the neighboring carbonyl group to produce the furan ring. These furo[3,4-c]furans react with various dienophiles, furnishing anisole derivatives derived by loss of water from the initially formed Diels-Alder cycloadducts. The Rh(II)-catalyzed cyclization reaction was quite versatile with regard to the nature of the interacting carbonyl group. The methodology was applied to the synthesis of several oxa-polyheterocyclic systems by first generating a 2-alkoxy-substituted furan and then allowing it to undergo a subsequent intramolecular Diels-Alder cycloaddition. Ring opening of the resulting cycloadduct is followed by deprotonation to furnish a rearranged keto lactone. The potential use of this method for the synthesis of the alkaloid strychnine was probed using suitable model diazo compounds. To establish the viability of this approach, the Rh(II)-catalyzed cyclization/cycloaddition sequence of alpha-diazo amides 64 and 68 were studied. Both compounds underwent the sequential process in good overall yield, leading to novel pentacyclic products. The structural features of the resultant products present numerous opportunities for postcycloaddition manipulations that could be exploited to synthetic advantage.  相似文献   

18.
采用密度泛函理论的B3LYP泛函对AuCl3催化的2-(1-炔基)-2-烯基酮与亲核试剂反应的机理进行了研究, 得到了反应的最优路径. 结果表明, 整个反应的决速步骤是羟基H转移到AuCl3的配体Cl上, 其活化能为49.3 kJ·mol-1. 通过计算发现, 催化剂AuCl3的配体Cl原子在反应中有重要的作用, 它不仅稳定配合物, 而且直接参与反应, 协助质子的转移, 显著降低质子转移的活化能(由71.5 kJ·mol-1降低到49.3 kJ·mol-1). 另外还讨论了HBF4不能催化此反应的可能原因, 计算结果与实验结果一致.  相似文献   

19.
The values of pseudo first‐order rate constants (kobs) for the cleavage of N‐(2‐hydroxyphenyl)phthalamic acid ( 7 ), obtained at 4.9 × 10?2 M HCl, 35°C, and within CH3CN content range 2–80% (v/v) in mixed aqueous solvent are smaller than kobs for the cleavage of N‐(2‐methoxyphenyl)phthalamic acid ( 8 ), obtained under almost similar experimental conditions, by nearly 1.5‐ to 2‐fold. These observations show the absence of expected intramolecular general acid catalysis due to 2‐OH group in 7 . The values of kobs for the cleavage of 7 and 8 decrease by more than 20‐fold with the increase in the content of CH3CN from 2 to 80–82% (v/v) in mixed aqueous solvent. The kinetic data reveal that in acidic aqueous cleavage of 7 , N‐cyclization (leading to the formation of imide) and O‐cyclization (leading to the formation of phthalic anhydride) vary from ~10 to 15% and ~90 to 85%, respectively, with the increase in CH3CN content from 2 to 80% (v/v). Similar increase in CH3CN content causes increase in N‐cyclization from ~0 to 5% and decrease in O‐cyclization from ~100 to 95% in the acidic aqueous cleavage of 8 . Some speculative, yet conceivable, reasons for nearly 10 and 0% N‐cyclization in the cleavage of respective 7 and 8 at low content of CH3CN have been described. © 2006 Wiley Periodicals, Inc. Int J Chem Kinet 38: 746–758, 2006  相似文献   

20.
Many dipeptide boronic acids of the type H(2)N-X-Y-B(OH)(2) are potent protease inhibitors. Interest in these compounds as drugs for cancer, diabetes, and other diseases is growing. Because of the great mutual B-N affinity, cyclization through the N- and B-termini, forming six-membered rings, is a common occurrence at neutral pH and higher where the terminal amino group is unprotonated. Here we report the discovery that when X, the N-terminal amino acid, contains a side chain having a functional group with boron affinity and suitable geometry, additional cyclization in the form of bidentate intramolecular chelation or "autochelation" may occur, predominantly at mid pH. NMR studies of two compounds, l-Aspartyl-l-boroProline (Asp-boroPro) and l-Histidyl-l-boroProline (His-boroPro), are reported here from pH 0.5 to pH 12 by (1)H, (15)N, (13)C, and (11)B NMR. Both of these previously unreported autochelates contain two fused six-membered rings, cis-proline, chiral boron, and -NH(2)(+) protons in slow exchange with water, even at 25 degrees C and pH as high as 4. Using microscopic acid-base equilibrium constants, we show that at high pH (>8 for Asp-boroPro and >10 for His-boroPro) hydroxide competes with the side chains for boron, reducing the chelates from bidentate to monodentate. At low pH (<0.5), proton competition for N-terminal nitrogens causes both compounds to become noncyclic. High chelate stability causes a reduction of the apparent acidic dissociation constant of the protonated N-terminal amino group greater than eight units. In the His-boroPro autochelate, imidazolate anion is produced at the extraordinarily low pH value of approximately 9.  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