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 15 毫秒
1.
To study structure-activity relationships of growth hormone-releasing hormone (GHRH), a competitive binding assay was developed using cloned porcine adenopituitary GHRH receptors expressed in human kidney 293 cells. Specific binding of [His1,125I-Tyr10,Nle27]hGHRH(1–32)-NH2 increased linearly with protein concentration (10–45 μg protein/tube). Binding reached equilibrium after 90 min at 30°C and remained constant for at least 240 min. Binding was reversible to one class of high-affinity sites (Kd = 1.04 ± 0.19 nM, Bmax = 3.9 ± 0.53 pmol/mg protein). Binding was selective with a rank order of affinity (IC50) for porcine GHRH (2.8 ± 0.51 nM), rat GHRH (3.1 ± 0.69 nM), [N-Ac-Tyr1, -Arg2]hGHRH(3–29)-NH2 (3.9 ± 0.58 nM), and [ -Thr7]GHRH(1–29)-NH2 (189.7 ± 14.3 nM), consistent with their binding to a GHRH receptor. Nonhydrolyzable guanine nucleotides inhibited binding. These data describe a selective and reliable method for a competitive GHRH binding assay that for the first time utilizes rapid filtration to terminate the binding assay.  相似文献   

2.
Erythromycin may stimulate gastrointestinal motor activity via its effect upon motilin receptors. We have studied the ability of the derivative EM-523 [de(N-methyl)-N-ethyl-8,9-anhydroerythromycin A 6,9-hemiacetal] to induce contractions in duodenal smooth muscle strips and to displace labeled motilin bound to antral smooth muscle, in man and in rabbit. In both species EM-523 approached the potency of motilin for inducing contractions. Thus pED50 values were 7.84 +/- 0.11 and 8.69 +/- 0.12 for motilin in, respectively, man and rabbit, against 6.08 +/- 0.13 and 8.19 +/- 0.10 for EM-523. In rabbit the efficacy of both compounds decreased in parallel aborally, the responses to EM-523 could not be blocked by atropine (10(-7) M) or TTX (10(-7) M), and both compounds were unable to further enhance the maximum effect to the other compound. In binding studies the order of potency was the same as in the contraction studies. The pIC50 values were: motilin (8.84 +/- 0.31, 9.17 +/- 0.20) greater than EM-523 (7.89 +/- 0.1, 8.40 +/- 0.10). A Schild plot revealed that EM-523 was a competitive inhibitor of motilin receptor binding in man and in rabbit. We conclude that EM-523 is a potent motilin agonist.  相似文献   

3.
The effect of structural changes in the N-terminal amino acid of AIV, with respect to AT4 receptor binding, was examined by competition with [125I]AIV in bovine adrenal membranes. Analogues with modifications of the first residue -amino group possessed lower affinities than the primary amine-containing parent compound. Peptides with a residue 1 -carbon in the conformation exhibited poor affinity for the AT4 receptor. Modifications of the residue 1 R-group demonstrate that a straight chain aliphatic moiety containing four carbons is optimal for receptor-ligand binding, as evidenced by the extremely high affinity of [Nle1]AIV (Ki = 3.59±0.51 pM). Replacement of the 1–2 peptide bond of AIV with the methylene bond isostere Ψ (CH2-NH), increased the Ki approximately fivefold, indicating that the peptide bond may be replaced wihle maintaining relatively high-affinity receptor binding.  相似文献   

4.
Liu H  Yang Y  Xin R  Liu X  Cao Y  Ni J  Wang R 《Peptides》2008,29(6):1048-1056
Previously, five synthetic peptides derived from endomorphin-1 (Tyr1-Pro2-Trp3-Phe4-NH2, EM-1), including Tyr-d-Ala-Trp-p-Cl-Phe-NH2 (HDAPC), Tyr-d-Ala-Trp-Phe-NH2 (HDADC), N-amidino-Tyr-d-Ala-Trp-p-Cl-Phe-NH2 (GDAPC), N-amidino-Tyr-d-Ala-Trp-Phe-NH2 (GDADC) and N-amidino-Tyr-d-Pro-Gly-Trp-p-Cl-Phe-NH2 (GBDPC), were described to elicit analgesia by subcutaneous administration with enhanced metabolic stabilities. To further our knowledge of the influences of particular modification on the pharmacological activities of EM-1, the present study was undertaken to investigate cardiovascular effects of these peptides in anesthetized rats by intravenous injection. Our results showed that the four d-Ala-containing peptides decreased the systemic arterial pressure (SAP) and heart rate (HR) through a naloxone-sensitive mechanism. Different patterns, potencies and durations of cardiovascular effects were observed among these peptides. When compared to EM-1, the hemodynamic responses to these four tetrapeptides were significantly lower in magnitude but much longer in duration. Surprisingly, intravenous administration of the only pentapeptide GBDPC produced fairly prolonged hypertensive and tachycardiac effects, which was naloxone-insensitive, thus providing evidence that changes in the primary structure of a peptide can profoundly affect its pharmacological activity. Comparisons of the cardiovascular effects between these peptides showed that each modification introduced into EM-1, including N-amidination, chloro-halogenation and unnatural amino acid substitution, played a role in the influence on the cardiovascular regulation of these peptides.  相似文献   

5.
Y Kloog  V Nadler  M Sokolovsky 《FEBS letters》1988,230(1-2):167-170
Binding of the labeled anticonvulsant drug [3H]dibenzocycloalkenimine (3H]MK-801) to the N-methyl-D-aspartate (NMDA) receptor and its dissociation from the receptor at 25°C are slow processes, both of which follow first order kinetics (t1/270 and 180 min, respectively). Both reactions are markedly accelerated by glutamate and glycine (t1/22-8 and 4 min, respectively), which allow bimolecular association kinetics of the labeled drug with the receptors whereas equilibrium binding of [3H]MK-801 (Kd 2–4 nM) is hardly affected by glutamate and glycine. The data suggest that MK-801 acts as a steric blocker of the NMDA receptor channel. The competitive antagonist D-(−)-2-amino-5-phosphovaleric acid (AP-5) freezes the receptor in a state which precludes either binding of [3H]MK-801 to the receptor channel or its dissociation from it. These findings have therapeutic implications.  相似文献   

6.
The activity of the muscarinic cholinergic system (acetylcholine, ACh; acetylcholinesterase, AChE; choline acetyltransferase, ChAT; muscarinic acetylcholine receptors) was studied in the carp brain. The ACh content (13.9 ± 1.1 nmol/g wet tissue) was estimated by gas chromatography after microwave irradiation focused to the head. The AChE and ChAT activities were 153 ± 13 nmol/min/mg protein and 817 ± 50 pmol/min/mg protein, respectively. The characteristics of [3H](−)quinuclidinyl benzilate ([3H](−)QNB) and [3H]pirenzepine ([3H]PZ) binding were also studied in brain membranes. Their specific binding was linearly dependent on the protein content and they appeared to bind with high affinity to a single, saturable binding site. A dissociation constant (Kd) of 47 ± 6.3 pM and a maximum number of binding sites (Bmax) of 627 ± 65 fmol/mg protein were obtained for [3H](−)QNB, with a Kd value of 3.85 ± 0.67 nM and a Bmax value of 95.3 ± 6.25 fmol/mg protein for [3H]PZ binding. The [3H]PZ binding amounted to only 15% of the [3H](−)QNB-labeled sites, as estimated from the ratio of the Bmax values of [3H](−)QNB and [3H]PZ, suggesting a low density of M1 subtype. Atropine sulfate, atropine methylnitrate and PZ inhibited the binding of both radioligands with Hill slopes (nH) close to unity. The nH value of AF-DX 116 was close to 1 against [3H](−)QNB binding, while it was 0.75 against [3H]PZ binding. The displacement curves of oxotremorine and carbachol were shallow for the binding of both radioligands. The rank order of potency of muscarinic ligands against [3H](−)QNB binding (Ki nM) was atropine sulfate (0.55) > atropine methylnitrate (1.61) > PZ (61.19) > oxotremorine (156.3) > AF-DX 116 (307) > carbachol (1301), while in the case of [3H]PZ binding it was atropine sulfate (0.24) > atropine methylnitrate (0.34) > PZ (10.38) > AF-DX 116 (55.87) > oxotremorine (62.79) > carbachol (1696). The results indicate the presence of a well-developed muscarinic cholinergic system with predominantly M2 receptors in the carp brain.  相似文献   

7.
The binding of [3H]proctolin to oviduct membranes has been analyzed to investigate the nature of proctolin binding sites in the oviduct. Proctolin binding was found to be time dependent, proportional to concentration of membrane protein, saturable, specific and reversible. Two apparent proctolin binding sites were recognized. The first had a Kd of 400 ± 82 nM and a Bmax of 23.7 ± 6.7 pmol/mg protein. The second had a Kd of 2.4 ± 0.2 μM and a Bmax of 96.3 ± 16.7 pmo/mg protein.

Binding was specific in thatcompetition experiments with a wide range of peptides showed that only Arg-Tyr-Leu-Pro-Ala was an effective competitor at μM concentrations. All other peptides examined weekly reduced proctolin binding at concentrations above 50 μM. Certain peptides were found to potentiate [3]pproctolin binding at low μM concentrations (1–10 μM) and to inhibit proctolin binding at higher concentrations. The significance of these findings is discussed.  相似文献   


8.
The kinetics of substitution reactions of [η-CpFe(CO)3]PF6 with PPh3 in the presence of R-PyOs have been studied. For all the R-PyOs (R = 4-OMe, 4-Me, 3,4-(CH)4, 4-Ph, 3-Me, 2,3-(CH)4, 2,6-Me2, 2-Me), the reactions yeild the same product [η5-CpFe(CO)2PPh3]PF6, according to a second-order rate law that is first order in concentrations of [η5-CpFe(CO)3]PF6 and of R-PyO but zero order in PPh3 concentration. These results, along with the dependence of the reaction rate on the nature of R-PyO, are consistent with an associative mechanism. Activation parameters further support the bimmolecular nature of the reactions: ΔH = 13.4 ± 0.4 kcal mol−1, ΔS = −19.1 ± 1.3 cal k−1 mol−1 for 4-PhPyO; ΔH = 12.3 ± 0.3 kcal mol−1, ΔS = 24.7 ±1.0 cal K−1 mol−1 for 2-MePyO. For the various substituted pyridine N-oxides studied in this paper, the rates of reaction increase with the increasing electron-donating abilities of the substituents on the pyridine ring or N-oxide basicities, but decrease with increasing 17O chemical shifts of the N-oxides. Electronic and steric factors contributing to the reactivity of pyridine N-oxides have been quantitatively assessed.  相似文献   

9.
This study investigates the gastroprokinetic effects of motilin and erythromycin A (EM-A) and its potential mechanism in guinea pigs Cavia porcellus in vitro. Guinea pig stomach strips were mounted under organ baths containing Krebs solution. Motilin,EM-A,Nω-Nitro-L-arginine (L-NNA),L-arginine (L-AA) were added to the bathing solution in a non-cumulative way. Then the effects of motilin and EM-A was studied during electrical field stimulation (EFS) in the absence and presence of L-NNA and L-AA in the gastri...  相似文献   

10.
The effects of brain ischemia on the maximum binding capacity (Bmax) and affinity (Kd) of A1 receptors were studied in the rat cerebral cortex, with an in vitro approach. The results were correlated with changes in 3H-adenosine release, studied under identical experimental conditions. Fifteen minutes of in vitro ‘ischemia’ (hypoxic, glucose-free medium) induced a significant increase in both Bmax (2398±132 fmol/mg protein, 151% of the control, P<0.05) and in Kd (2.43±0.12 nM, 161% of the control, P<0.01). At the same time, an increase in tritium efflux from [3H]-adenosine labeled cerebral cortex slices to 324% of the control was observed. A trend toward normalization was evident 5–15 min after ‘reoxygenation’ (restoring normal medium), but the binding parameters were still altered after 60 min (Bmax 2110±82 fmol/mg protein, Kd 2.26±0.14 nM, P<0.01 vs the corresponding control) as was adenosine release (196% of the control). These findings suggest that the increased availability of adenosine and its receptors may be a defense mechanism against ischemic injury, while the reduced affinity of A1 receptors, possibly due to desensitization, may be a sign of ischemia-induced cellular damage.  相似文献   

11.
Two functional isoforms (1) and + (3) of the Na,K-ATPase catalytic subunit coexist in canine cardiac myocytes [J. Biol. Chem. (1987) 262, 8941-8943]. The in vitro turnover rates of ATP hydrolysis have been determined in sarcolemma preparations by comparing [3H]ouabain-binding and Na,K-ATPase activity at various doses of ouabain (0.3–300 nM). The correlation between the occupancy of the ouabain-binding sites and the degree of Na,K-ATPase inhibition was not linear. The results showed that the form of low-affinity for ouabain (Kd = 300–700 nM) exhibited a lower turnover rate (88 ± 10 vs. 147 ± 15 molecules of ATP hydrolyzed per second per ouabain-binding site) than the high affinity form (Kd = 1–8 nM). Thus our results indicate this specific isoform kinetic difference could contribute to differences in the cardiac cellular function.  相似文献   

12.
The neuropeptides vasoactive intestinal peptide (VIP), neuropeptide Y (NPY), and substance P (SP) as well as insulin and insulin-like growth factor 1 (IGF-1) were labeled with biotin, fluorescent dyes, and radioactivity to characterize the expression of peptide receptor of a novel cancer cell line, established from a human glioblastoma multiforme. Thus, not only binding sites could be detected but advantages and disadvantages of the different labels could be compared, too. With all three markers, the presence or absence of the receptors could be answered rapidly and sensitively. The glioblastoma cells express receptors for VIP (IC50 = 9 nM ± 30%), insulin (Kd = 0.66 nM ± 14%, Bmax = 0.028 nM ± 13%), and IGF-1 (Kd = 21 nM ± 25%, Bmax = 1.65 nM ± 24%), but there are no binding sites for NPY and SP. As especially VIP and IGF-1 receptors are expressed in huge amounts, these receptors might be an interesting target for tumor diagnostics and therapy.  相似文献   

13.
High affinity Ins(1,4,5)P3-binding sites of permeabilized hepatocytes are probably the ligand recognition sites of the receptors that mediate the effects of Ins91,4,5)P3 on intracellular Ca2+ mobilization. We have now solubilized these sites from rat liver membranes in the zwitterionic detergent, CHAPS, and shown that the solubilized bind Ins(1,4,5)P3 with an affinity (Kd = 7.26 ± 0.52 nM, Hill coefficient H = 1.05 ± 0.06) similar to that of the sites in native membranes (Kd = 6.02 ± 0.02). ATP and a range of inositol phosphates (Ins(2,4,5)P3 Ins(4,5)P2, and inositol 1,4,5-trisphosphorothioate) also bound with similar affinities to the native and solubilized sites. Solubilization of the liver InsP3 receptor will allow its further characterization, purification, and comparison of its properties with those of InsP3 receptors already purified from cerebellum and smooth muscle.  相似文献   

14.
The binding kinetics of the specific dopamine D2 antagonist [3H]raclopride to dopamine D2 receptors in rat neostriatum were studied. The pseudo-first-order rate constants of [3H]raclopride binding with these membranes revealed a hyperbolic dependence upon the antagonist concentration, indicating that the reaction had at least two consecutive and kinetically distinguished steps. The first step was fast binding equilibrium, characterized by the dissociation constant KA = 12 ± 2 nM. The following step corresponded to a slow isomerization of the receptor-antagonist complex, characterized by the isomerization equilibrium constant Ki = 0.11. The dissociation constant Kd = 1.3 nM, calculated from these kinetic data, was similar to Kd = 2.4 nM, determined from equilibrium binding isotherm for the radioligand. Implications of the complex reaction mechanism on dopamine D2 receptor assay by [3H]raclopride were discussed.  相似文献   

15.
The enthalpies of reaction of HMo(CO)3C5R5 (R = H, CH3) with diphenyldisulfide producing PhSMo(CO)3C5R5 and PhSH have been measured in toluene and THF solution (R = H, ΔH= −8.5 ± 0.5 kcal mol−1 (tol), −10.8 ± 0.7 kcal mol−1 (THF); R = CH3, ΔH = −11.3±0.3 kcal mol−1 (tol), −13.2±0.7 kcal mol−1 (THF)). These data are used to estimate the Mo---SPh bond strength to be on the order of 38–41 kcal mol−1 for these complexes. The increased exothermicity of oxidative addition of disulfide in THF versus toluene is attributed to hydrogen bonding between thiophenol produced in the reaction and THF. This was confirmed by measurement of the heat of solution of thiophenol in toluene and THF. Differential scanning calorimetry as well as high temperature calorimetry have been performed on the dimerization and subsequent decarbonylation reactions of PhSMo(CO)3Cp yielding [PhSMo(CO)2Cp]2 and [PhSMo(CO)Cp]2. The enthalpies of reaction of PhSMo(CO)3Cp and [PhSMo(CO)2Cp]2 with PPh3, PPh2Me and P(OMe)3 have also been measured. The disproportionation reaction: 2[PhSMo(CO)2Cp]2 → 2PhSMo(CO)3Cp + [PhSMP(CO)Cp]2 is reported and its enthalpy has also been measured. These data allow determination of the enthalpy of formation of the metal-sulfur clusters [PhSMo(CO)nC5H5]2, N = 1,2.  相似文献   

16.
Differential scanning calorimetry, circular dichroism, and visible absorption spectrophotometry were employed to elucidate the structural stability of thermophilic phycocyanin derived from Cyanidium caldarium, a eucaryotic organism which contains a nucleus, grown in acidic conditions (pH 3.4) at 54°C. The obtained results were compared with those previously reported for thermophilic phycocyanin derived from Synechococcus lividus, a procaryote containing no organized nucleus, grown in alkaline conditions (pH 8.5) at 52°C. The temperature of thermal unfolding (td) was found to be comparable between C. caldarium (73°C) and S. lividus (74°C) phycocyanins. The apparent free energy of unfolding (ΔG[urea]=0) at zero denaturant (urea) concentration was also comparable: 9.1 and 8.7 kcal/mole for unfolding the chromophore part of the protein, and 5.0 and 4.3 kcal/mole for unfolding the apoprotein part of the protein, respectively. These values of td and ΔG[urea]=0 were significantly higher than those previously reported for mesophilic Phormidium luridum phycocyanin (grown at 25°C). These findings revealed that relatively higher values of td and ΔG[urea]=0 were characteristics of thermophilic proteins. In contrast, the enthalpies of completed unfolding (ΔHd) and the half-completed unfolding (ΔHd)1/2 for C. caldarium phycocyanin were much lower than those for S. lividus protein (89 versus 180 kcal/mole and 62 versus 115 kcal/mole, respectively). Factors contributing to a lower ΔHd in C. caldarium protein and the role of charged groups in enhancing the stability of thermophilic proteins were discusse.  相似文献   

17.
A series of dihydroxamic acid ligands of the formula [RN(OH)C(O)]2(CH2)n, (n = 2, 4, 6, 7, 8; R = CH3, H) has been studied in 2.0 M aqueous sodium perchlorate at 25.0 °C. These ligands may be considered as synthetic analogs to the siderophore rhodotorulic acid. Acid dissociation constants (pKa) have been determined for the ligands and for N-methylacetohydroxamic acid (NMHA). The pKa1 and pKa2 values are: n = 2, R = CH3 (8.72, 9.37); N = 4, R = CH3 (8.79, 9.37); N = 6, R = CH3; N = 7, R = CH3 (8.95, 9.47); N = 8, R = CH3 (8.93, 9.45); N = 8, R = H (9.05, 9.58). Equilibrium constants for the hydrolysis of coordinated water (log K) have been estimated for the 1:1 feeric complexes of the ligands n = 2, 4, 8; R = CH3. The N = 8 ligand forms a monomeric complex with Fe(III) while the n = 2 and 4 ligands form dimeric complexes. For hydrolysis of the n = 8 monomeric complex, log K1 = −6.36 and log K2 = −9.84. Analysis of the spectrophotometric data for the dimeric complexes indicates deprotonation of all four coordinated waters. The successive hydrolysis constants, log K1–4, for the dimeric complexes are as follows: n = 2 (−6.37, −5.77, −10.73, −11.8); n = 4 (−5.54, −5.07, −11.57, −10.17). The log K2 values for the dimers are unexpectedly high, higher in fact than log K1, inconsistent with the formation of simple ternary hydroxo complexes. A scheme is proposed for the hydrolysis of the ferric dihydroxamate dimers, which includes the possible formation of μ-hydroxo and μ-oxo bridges.  相似文献   

18.
Oxygenation of [CuII(fla)(idpa)]ClO4 (fla=flavonolate; IDPA=3,3′-iminobis(N,N-dimethylpropylamine)) in dimethylformamide gives [CuII(idpa)(O-bs)]ClO4 (O-bs=O-benzoylsalicylate) and CO. The oxygenolysis of [CuII(fla)(idpa)]ClO4 in DMF was followed by electronic spectroscopy and the rate law −d[{CuII(fla)(idpa)}ClO4]/dt=kobs[{CuII(fla)(idpa)}ClO4][O2] was obtained. The rate constant, activation enthalpy and entropy at 373 K are kobs=6.13±0.16×10−3 M−1 s−1, ΔH=64±5 kJ mol−1, ΔS=−120±13 J mol−1 K−1, respectively. The reaction fits a Hammett linear free energy relationship and a higher electron density on copper gives faster oxygenation rates. The complex [CuII(fla)(idpa)]ClO4 has also been found to be a selective catalyst for the oxygenation of flavonol to the corresponding O-benzoylsalicylic acid and CO. The kinetics of the oxygenolysis in DMF was followed by electronic spectroscopy and the following rate law was obtained: −d[flaH]/dt=kobs[{CuII(fla)(idpa)}ClO4][O2]. The rate constant, activation enthalpy and entropy at 403 K are kobs=4.22±0.15×10−2 M−1 s−1, ΔH=71±6 kJ mol−1, ΔS=−97±15 J mol−1 K−1, respectively.  相似文献   

19.
1,10-Phenanthroline-5,6-dione (C12H6N2O2 (1)) reacts with V(η6-mesitylene)2 and Ti(η6-toluene)2 affording coordination compounds of general formula M(O,O′---C12H6N2O2)3 (M=Ti (2); M=V (3)) which further react with TiCl4 or TiCp2(CO)2 yielding the tetrametallic species M(O,O′---C12H6N2O2---N,N′)3(M′Ln)3 (M=V, M′Ln=TiCl4 (4); M=Ti, M′Ln=TiCp2 (5); M=V, M′Ln=TiCp2 (6)). The complex salt [Fe(N,N′---C12H6N2O2)3][PF6]2 (7) has been obtained from iron(II) chloride tetrahydrate and 1 in the presence of NH4PF6. The reaction of 7 with TiCp2(CO)2 affords the tetrametallic derivative [Fe(N,N′---C12H6N2O2---O,O′)3(TiCp2)3][PF6]2 (8). TiCl2(THF)2 reacts with MCp2(O,O′---C12H6N2O2) to give MCp2(O,O′---C12H6N2O2---N,N′)TiCl2 (M=Ti (9); M=V (10)). By reaction of TiCp2(O,O′---C12H6N2O2---N,N′)TiCl2 (9) with C12H6N2O2, the bimetallic derivative TiCp2(O,O′---C12H6N2O2---N,N′)TiCl2(O,O′---C12H6N2O2) (11) has been prepared, which readily adds to TiCl4, to give the trimetallic titanium derivative TiCp2(O,O′---C12H6N2O2---N,N′)TiCl2(O,O′---C12H6N2O2---N,N′)TiCl4 (12). VCp2(O,O′---C12H6N2O2---N,N′)TiCl2 (10) reacts with the tris-chelate iron(II) cation 7 affording the heptametallic cationic complex [Fe(N,N′---C12H6N2O2---O,O′)TiCl2(N,N′---C12H6N2O2---O,O′)VCp2]3 +2 isolated as the hexafluorophosphate 13.  相似文献   

20.
The ester cleavage of R- and S-isomers N-CBZ-leucine p-nitrophenyl ester intermolecularly catalyzed by R- (a) and S-stereoisomers (b) of the Pd(II) metallacycle [Pd(C6H4C*HMeNMe2)Cl(py)] (3) follows the rate expression kobs = ko + kcat [3], where the rate constants kcat equal 25.8 ± 0.4 and 7.6 ± 0.5 dm3 mol−1 s−1 for the S- and R-ester, respectively, in the case of 3a, but are 5.7 ± 0.6 and 26.7 ± 0.5 dm3 mol−1 s−1 for the S- and R-ester, respectively, in the case of 3b (pH 6.23 and 25°C). Thus, the best catalysis occurs when the asymmetric carbons of the leucine ester and Pd(II) complex are R and S, or S and R configured, respectively. Molecular modeling suggests that the stereoselection results from the spatial interaction between the CH2CHMe2 radical of the ester and the -methyl group of 3. A hydrophobic/stacking contact between the leaving 4-nitrophenolate and the coordinated pyridine also seems to play a role. Less efficient intramolecular enantioselection was observed for the hydrolysis of N-t-BOC-S-metthionine p-nitrophenyl ester with R- and S-enantiomers of [Pd(C6H4C*HMeNMe2)Cl] coordinated to sulfur.  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