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 26 毫秒
1.
Cool LG 《Phytochemistry》2005,66(2):249-260
Ten sesquiterpenes, many with unusual carbon skeletons, were identified in foliage of Cupressus macrocarpa. These are (-)-10-epi-beta-acoradiene; ent-widdra-2,4(14)-diene; (E)-iso-gamma-bisabolene, i.e., (4E)-4-(1,5-dimethylhex-5-enylidene)-1-methylcyclohexene; (-)-cumacrene, i.e., (4S)-4-[(1R,2S)-2-isopropenyl-1-methylcyclobutyl]-1-methylcyclohexene; (-)-alpha-chamipinene, i.e., (1S,6S,7S)-2,2,6,8-tetramethyltricyclo[5.3.1.01,6]undec-8-ene; and five sesquiterpenes with a 3,3,4'-trimethyl-1,1'-bi(cyclohexyl) skeleton for which the trivial name macrocarpane is proposed. The possible single-enzyme biogenesis of these sesquiterpenes is discussed.  相似文献   

2.
Imai Y  Tajima N  Sato T  Kuroda R 《Chirality》2002,14(7):604-609
Molecular recognition in solid-state crystallization involving derivatives of 1,1'-bi-2-naphthol and benzoquinone was studied. No adduct crystal was formed when crystals of biphenyl were further added as a third component to a grinding mixture of crystals of chiral 1,1'-bi-2-naphthol and benzoquinone, which by itself did not form an adduct. This contrasts with the case in which further addition of naphthalene crystals to the same mixture produced a new red crystal. Adduct formations using chiral 6,6'-dibromo-1,1'-bi-2-naphthol in place of 1,1'-bi-2-naphthol were also studied. In this case, adducts were produced either with or without biphenyl as a third compound, but the colors of the adducts differed significantly in the two cases: red and bluish-black. The same three-component adduct crystals were produced from solid-grinding and solution crystallization and the structure was determined by X-ray diffractometry. Based on the crystal structures, theoretical calculations were carried out to compare the mechanism of colorations in the binary and the ternary complexes.  相似文献   

3.
Kubota T  Yamamoto C  Okamoto Y 《Chirality》2002,14(5):372-376
Cyclopentyl and (+/-)-exo-2-norbornylcarbamates of cellulose and amylose were prepared and their chiral recognition abilities as chiral stationary phases for high-performance liquid chromatography (HPLC) were evaluated. Among these carbamates, cellulose tris(cyclopentylcarbamate) and amylose tris((+/-)-exo-2-norbornylcarbamate) showed particularly high chiral recognition, which is comparable to that of several well-known phenylcarbamate derivatives. The chiral recognition mechanism of cellulose tris(cyclohexylcarbamate), which was previously found to be an effective chiral stationary phase for HPLC, was investigated using NMR spectroscopy. The derivative dissolved in chloroform exhibited the chiral discrimination of several enantiomers in NMR as well as in HPLC. For example, the 1,1'-bi-2-naphthol enantiomers were distinctly discriminated in the (1)H, (13)C, and 2D-NOESY spectra.  相似文献   

4.
The palladium catalysed coupling reactions of 1-iodo-cyclohexene have been investigated with the aim of the synthesis of model compounds towards unsaturated aryl-alkenyl ketones of practical interest in carbonylative Suzuki reaction. In addition to the target compounds, the formation of 1-aryl-cyclohexene derivatives, 1,1'-bi(cyclohex-1-en-1-yl), 1,2-dicyclohex-1-en-1-yl-ethane-1,2-dione, dicyclohex-1-en-1-yl-ketone and 1-aryl-2-cyclohexenyl-cyclohexene derivatives, as well as their isomerization products, have been observed in parallel and consecutive reactions, respectively. The complex reaction mixtures have been analysed with GC-MS providing valuable information to the investigation of the homogeneous catalytic reaction.  相似文献   

5.
Kumar NN  Swamy KC 《Chirality》2008,20(6):781-789
Diastereoselective synthesis and characterization of chiral unsymmetrical tris-spirocyclic cyclotriphosphazenes based on chiral 1,1'-bi-2-naphthol (BINOL) are reported. Specifically, the chiral compounds (-)N(3)P(3)[1,1'-O(2)(C(10)H(6))(2)](O-2,2'C(6)H(4)-C(6)H(4)O)Cl(2) [(-)-4] and (-)N(3)P(3)[1,1'-O(2)(C(10)H(6))(2)](OCH(2)CH(2)NMe)(2) [(-)-5] are prepared by starting with the chiral mono-spiro compound (-)N(3)P(3)[1,1'-O(2)(C(10)H(6))(2)]Cl(4) [(-)-3]. Synthesis of four other chiral spirocyclics, N(3)P(3)[1,1'-O(2)(C(10)H(6))(2)](OCH(2)CH(2) NMe)(O-2,2'C(6)H(4)-C(6)H(4)O)[(-)-6 and (+)-6], N(3)P(3)[1,1'-O(2)(C(10)H(6))(2)](NMe(2))(4) [(-)-7], N(3)P(3)[1,1'-O(2)(C(10)H(6))(2)](O-2,2'C(6)H(4)-C(6)H(4)O)(NMeCH(2)CH(2)OH)(2) [(-)-8 and (+)-8], and N(3)P(3)[1,1'-O(2)(C(10)H(6))(2)](O-2,2'C(6)H(4)-C(6)H(4)O)[NHCH(2)CH(2)CH(2)Si(OEt)(3)](2) (9) is also reported herein. Compounds 4-6 are obtained in the solid state diastereoselectively and their X-ray structures have been determined and discussed. The diastereoselectivity is also shown by structural characterization of two distinct isomers in the case of 6 [(-)-6 and (+)-6, respectively] by starting with precursor of 3 having (R) or (S)-BINOL residue. The (1)H NMR spectra of 7 and 8 exhibit doublets with virtual coupling for the methyl protons, consistent with the chiral nature of the binaphthoxy residue. The potential of 9, which hydrolyzes readily in CDCl(3) solution, as a useful precursor for chiral polymer applications is highlighted.  相似文献   

6.
(R)-(-)-α-Methoxy phenyl acetic acid, (S)-(-)-1,1'-(2-naphthol), and (R)-(+)-α-methoxy-α-trifluoromethyl phenyl acetic acid were evaluated as chiral shift reagents (CSRs) for (1)H NMR spectroscopic resolution and determination of R and S enantiomers of modafinil (MDL) in bulk drugs and formulations. Effects of the nature of CSR and the weight ratio of substrate to shift reagent on enantiomeric discrimination were investigated. Intramolecular and intermolecular hydrogen bonding interactions between the drug and the CSR seem to be the driving force for desired resolution. A mechanism was proposed to explain the interactions between (R, S)-enantiomers of MDL and (R)-(-)-α-methoxy phenyl acetic acid. The method was validated and applied successfully to determine the enantiomeric purity of MDL in tablet formulations.  相似文献   

7.
The enantiomers of various 1-(alpha-aminobenzyl)-2-naphthol and 1-(aminoalkyl)-2-naphthol analogs were separated on cellulose-tris-3,5-dimethylphenyl carbamate-based chiral stationary phases (Chiralcel OD-H and Chiralcel OD-RH), using n-hexane/2-propanol/diethylamine or phosphate buffer/organic modifier mobile phases. The 3,5-dimethylphenyl carbamoylated cellulose columns were effective in both normal and rev ersed-phase modes. The effects of the mobile phase composition, the pH, the buffer concentration, and the structures of the substituents on the 2-naphthol on the enantioseparations were studied. The absolute configuration and elution sequence were determined for 1-(1-amino-2-methylpropyl)-2-naphthol: the elution sequence was S < R.  相似文献   

8.
The regiospecific oxidation of the polycyclic aromatic hydrocarbons naphthalene and fluorene was examined with Escherichia coli strains expressing wildtype toluene 4-monooxygenase (T4MO) from Pseudomonas mendocina KR1, toluene para-monooxygenase (TpMO) from Ralstonia pickettii PKO1, toluene ortho-monooxygenase (TOM) from Burkholderia cepacia G4, and toluene/ortho-xylene monooxygenase (ToMO) from P. stutzeri OX1. T4MO oxidized toluene (12.1+/-0.8 nmol/min/mg protein at 109 microM), naphthalene (7.7+/-1.5 nmol/min/mg protein at 5 mM), and fluorene (0.68+/-0.04 nmol/min/mg protein at 0.2 mM) faster than the other wildtype enzymes (2-22-fold) and produced a mixture of 1-naphthol (52%) and 2-naphthol (48%) from naphthalene, which was successively transformed to a mixture of 2,3-, 2,7-, 1,7-, and 2,6-dihydroxynaphthalenes (7%, 10%, 20%, and 63%, respectively). TOM and ToMO made 1,7-dihydroxynaphthalene from 1-naphthol, and ToMO made a mixture of 2,3-, 2,6-, 2,7-, and 1,7-dihydroxynaphthalene (26%, 22%, 1%, and 44%, respectively) from 2-naphthol. TOM had no activity on 2-naphthol, and T4MO had no activity on 1-naphthol. To take advantage of the high activity of wildtype T4MO but to increase its regiospecificity on naphthalene, seven engineered enzymes containing mutations in T4MO alpha hydroxylase TmoA were examined; the selectivity for 2-naphthol by T4MO I100A, I100S, and I100G was enhanced to 88-95%, and the selectivity for 1-naphthol was enhanced to 87% and 99% by T4MO I100L and G103S/A107G, respectively, while high oxidation rates were maintained except for G103S/A107G. Therefore, the regiospecificity for naphthalene oxidation was altered to practically pure 1-naphthol or 2-naphthol. All four wildtype monooxygenases were able to oxidize fluorene to different monohydroxylated products; T4MO oxidized fluorene successively to 3-hydroxyfluorene and 3,6-dihydroxyfluorene, which was confirmed by gas chromatography-mass spectrometry and 1H nuclear magnetic resonance analysis. TOM and its variant TomA3 V106A oxidize fluorene to a mixture of 1-, 2-, 3-, and 4-hydroxyfluorene. This is the first report of using enzymes to synthesize 1-, 3-, and 4-hydroxyfluorene, and 3,6-dihydroxyfluorene from fluorene as well as 2-naphthol and 2,6-dihydroxynaphthalene from naphthalene.  相似文献   

9.
Trichloroethylene (TCE) is the most frequently detected groundwater contaminant, and 1-naphthol is an important chemical manufacturing intermediate. Directed evolution was used to increase the activity of toluene ortho-monooxygenase (TOM) of Burkholderia cepacia G4 for both chlorinated ethenes and naphthalene oxidation. When expressed in Escherichia coli, the variant TOM-Green degraded TCE (2.5 +/- 0.3 versus 1.39 +/- 0.05 nmol/min/mg of protein), 1,1-dichloroethylene, and trans-dichloroethylene more rapidly. Whole cells expressing TOM-Green synthesized 1-naphthol at a rate that was six times faster than that mediated by the wild-type enzyme at a concentration of 0.1 mM (0.19 +/- 0.03 versus 0.029 +/- 0.004 nmol/min/mg of protein), whereas at 5 mM, the mutant enzyme was active (0.07 +/- 0.03 nmol/min/mg of protein) in contrast to the wild-type enzyme, which had no detectable activity. The regiospecificity of TOM-Green was unchanged, with greater than 97% 1-naphthol formed. The beneficial mutation of TOM-Green is the substitution of valine to alanine in position 106 of the alpha-subunit of the hydroxylase, which appears to act as a smaller "gate" to the diiron active center. This hypothesis was supported by the ability of E. coli expressing TOM-Green to oxidize the three-ring compounds, phenanthrene, fluorene, and anthracene faster than the wild-type enzyme. These results show clearly that random, in vitro protein engineering can be used to improve a large multisubunit protein for multiple functions, including environmental restoration and green chemistry.  相似文献   

10.
The direct separation of the enantiomers of 1-(α-aminoarylmethyl)-2-naphthol, 1-(α-aminoalkyl)-2-naphthol, 2-(α-aminoarylmethyl)-1-naphthol analogs, and 2-(1-amino-2-methylpropyl)-1-naphthol) was performed on a newly developed chiral stationary phase containing isopropyl carbamate-cyclofructan6 as chiral selector, with n-heptane/alcohol/trifluoroacetic acid as mobile phase. The effects of the mobile-phase composition, the nature and concentration of the alcoholic and acidic modifiers, and the structures of the analytes on the retention and resolution were investigated. In some cases, separations were carried out at constant mobile-phase compositions in the temperature range 5-40°C. Thermodynamic parameters and T(iso) values were calculated from plots of ln k' or ln α versus 1/T. -Δ(ΔH°) ranged from 2.8 to 3.2 kJ mol(-1) , -Δ(ΔS°) from 7.7 to 10.1 J mol(-1) K(-1) , and -Δ(ΔG°) from 0.2 to 0.5 kJ mol(-1) . It was found that the enantioseparations were enthalpy driven. The sequence of elution of the stereoisomers determined in some cases was (R) < (S).  相似文献   

11.
We developed a new 3-D HPLC method for on-line clean-up and simultaneous quantification of two important naphthalene metabolites, 1-naphthol and 2-naphthol, in human urine. Except an enzymatic hydrolysis no further sample pre-treatment is necessary. The metabolites are stripped from urinary matrix by on-line extraction on a restricted access material pre-column (RAM RP-8), transferred in backflush mode onto a silica-based CN-(cyano)phase column for further purification from interfering substances. By another successive column switching step both analytes are transferred with a minimum of overlapping interferences onto a C12 bonded reversed phase column with trimethylsilyl endcapping where the final separation is carried out. The entire arrangement is software controlled. Eluting analytes are quantified by fluorescence detection (227/430 nm) after an external calibration. Within a total run time of 40 min we can selectively quantify both naphthols with detection limits in the lower ppb range (1.5 and 0.5 microg/l for 1- and 2-naphthol, respectively) with excellent reliability (ensured by precision, accuracy, matrix-independency and FIOH quality assurance program participation). First results on a collective of 53 occupationally non exposed subjects showed mean levels of 11.0 microg/l (1-naphthol) and 12.9 microg/l (2-naphthol). Among smokers (n=21) a significantly elevated mean level of urinary naphthols was determined (1-naphthol: 19.2 microg/l and 2-naphthol: 23.7 microg/l) in comparison to non smokers (n=32; 1-naphthol: 5.6 microg/l, 2-naphthol: 5.6 microg/l).  相似文献   

12.
Experiments with mini-alphaA-crystallin (KFVIFLDVKHFSPEDLTVK) showed that Phe(71) in alphaA-crystallin could be essential for the chaperone-like action of the protein (Sharma, K. K., Kumar, R. S., Kumar, G. S., and Quinn, P. T. (2000) J. Biol. Chem. 275, 3767-3771). In the present study we replaced Phe(71) in rat alphaA-crystallin with Gly by site-directed mutagenesis and then compared the structural and functional properties of the mutant protein with the wild-type protein. There were no differences in molecular size or intrinsic tryptophan fluorescence between the proteins. However, 1,1'-bi(4-anilino)naphthalene-5,5'-disulfonic acid interaction indicated a higher hydrophobicity for the mutant protein. Both wild-type and mutant proteins displayed similar secondary structure during far UV CD experiments. Near UV CD signal showed a slight difference in the tertiary structure around the 285-295 region for the two proteins. The mutant protein was totally inactive in suppressing the aggregation of reduced insulin, heat-denatured citrate synthase, and alcohol dehydrogenase. However, a marginal suppression of beta(L)-crystallin aggregation was observed when mutant alphaA-crystallin was included. These results suggest that Phe(71) contributes to the chaperone-like action of alphaA-crystallin. Therefore we conclude that the 70-88-region in alphaA-crystallin, identified by us earlier, is the functional chaperone site in alphaA-crystallin.  相似文献   

13.
In order to develop orally active CCR5 antagonists, 1-propyl- or 1-isobutyl-1-benzazepine derivatives containing a sulfoxide moiety have been designed, synthesized, and evaluated for their biological activities. Sulfoxide compounds containing a 2-pyridyl group were first investigated, which led to discovering that the presence of a methylene group between the sulfoxide moiety and 2-pyridyl group was necessary for increased inhibitory activity in a binding assay. After further chemical modification, it was found that replacement of the pyridyl group with an imidazolyl or 1,2,4-triazolyl group enhanced activity in the binding assay and that S-sulfoxide compounds were more active than R-isomers. Particularly, compounds (S)-4r, (S)-4s, and (S)-4w exhibited highly potent CCR5 antagonistic activities (IC50=1.9, 1.7, 1.6 nM, respectively) and inhibitory effects (IC50=1.0, 2.8, 7.7 nM, respectively) in the HIV-1 envelope mediated membrane fusion assay, together with good pharmacokinetic properties in rats. In addition, we established the synthesis of (S)-4r and (S)-4w by asymmetric oxidation with titanium-(S)-(-)-1,1'-bi-2-naphthol complex.  相似文献   

14.
Kinetics of interaction of C1 inhibitor with complement C1s   总被引:3,自引:0,他引:3  
M Lennick  S A Brew  K C Ingham 《Biochemistry》1986,25(13):3890-3898
The kinetics of inhibition of the complement serine protease, C1s, by its only known inhibitor, C1 inhibitor, have been measured by a variety of methods. One method continuously monitors the loss of esterolytic activity with a synthetic substrate coupled to a chromogen while another monitors the formation of a stable (covalent) complex by high-pressure size-exclusion chromatography under dissociating conditions. Additional methods employ fluorescence probes to follow the formation of bimolecular complexes but are not expected to distinguish between covalent product and noncovalent (reversible) intermediates. There was good agreement between rate constants obtained by the various methods over a broad range of inhibitor concentrations, suggesting that noncovalent intermediates do not accumulate to a significant extent. The reaction appears to be pure second order with a bimolecular rate constant of 6.0 X 10(4) M-1 s-1 at 30 degrees C, independent of Ca2+, and an activation energy of 11.0 kcal/mol. The rate increases up to 35-fold in the presence of heparin which was shown to bind to all three components (enzyme, inhibitor, and complex) with similar affinity (Kd = 2.0-3.3 microM). The fluorescent probe 1,1'-bis(anilino)-4-,4'-bi(naphthalene)-8,8'-disulfonate [bis(ANS)] bound to the complex with Kd = 0.26 microM under conditions where the individual components had little affinity for the dye, consistent with the generation of one or more hydrophobic binding sites on the protein surface during complex formation.  相似文献   

15.
The enantiomers of 1,1′-bi-2-naphthol, 1,1′-binaphthyl diyl hydrogen phosphate, and 1,1′-binaphthyldicarboxylic acid are separated using capillary electrophoresis with cyclodextrins added to the running buffer. It is demonstrated that the type and concentration of cyclodextrin employed are critical for maximum enantiomeric resolution. A modified version of a previously described model of enantiomeric separations in capillary electrophoresis is shown to support the observed separation behavior. Molecular modeling is employed to calculate interaction energies between the various enantiomers and cyclodextrins. A reasonable correlation between these computationally derived interaction energies and separation behavior resulted from a statistical mechanical treatment of the molecular modeling data. The importance of hydrogen bonding in inclusion complex formation was probed and the effects of minimization and solvation in molecular modeling calculations are also discussed. © 1995 Wiley-Liss, Inc.  相似文献   

16.
A method was developed for simultaneous quantification of urinary 1- and 2-naphthols, 3- and 9-phenanthrols and 1-pyrenol using gas chromatography with mass spectrometry (GC-MS). This method was applied to urine samples from coke oven workers (n =28) and controls (n =22) from Northern China. Geometric mean levels of urinary 1-naphthol (58.8 μg l-1), 2-naphthol (34.1 μg l-1), 3-phenanthrol (7.35 μg l-1), 9-phenanthrol (1.28 μg l-1) and 1-pyrenol (25.4 μg l-1) were significantly higher among coke oven workers than controls. All the substances tested were highest among top-of-oven workers, who had 15-fold higher 1-naphthol, eight-fold higher 2-naphthol and 20-fold higher 1-pyrenol levels compared with controls. Using multiple linear regression models, 72.5% of the variation in 1- and 2-naphthol and 82.8% of the variation in 1-pyrenol were explained by the concentration of naphthalene or pyrene in the urine, the work category and the smoking intensity. Cigarette consumption significantly contributed to levels of urinary 1-pyrenol and naphthols, particularly 2-naphthol. A negative relationship between work category and the ratio of naphthols/1-pyrenol was observed among smokers. Our results suggest that urinary naphthols and phenanthrols reflect polycyclic aromatic hydrocarbon (PAH) exposure as well as the widely used 1-pyrenol, and that interactions between cigarette smoking and PAH exposure result in different patterns of metabolism for individual PAHs.  相似文献   

17.
The kinetic properties of the rat intestinal microsomal 1-naphthol:UDPglucuronosyltransferase (EC 2.4.1.17) were investigated in fully activated microsomes prepared from isolated mucosal cells. The enzyme appeared to follow an ordered sequential bireactant mechanism in which 1-naphthol and UDP-glucuronic acid (UDPGlcUA) are the first and second binding substrates and UDP and 1-naphthol glucuronide the first and second products, respectively. Bisubstrate kinetic analysis yielded the following kinetic constants: Vmax = 102 +/- 6 nmol/min per mg microsomal protein, Km (UDPGlcUA) = 1.26 +/- 0.10 mM, Km (1-naphthol) = 96 +/- 10 microM and Ki (1-naphthol) = 25 +/- 7 microM. The rapid equilibrium random or ordered bireactant mechanisms, as well as the iso-Theorell-Chance mechanism, could be excluded by endproduct inhibition studies with UDP.UDP-N-acetylglucosamine (UDPGlcNAc), usually found to be an activator of UDP glucuronosyltransferase in liver microsomes, acted as a full competitive inhibitor towards UDPGlcUA in rat intestinal microsomes. With regard to 1-naphthol UDPGlcNAc exhibited a dual effect: both inhibition and activation was observed. The effect of activation by MgCl2 and Triton X-100 on the kinetic constants and the inhibition patterns of UDP and UDPGlcNAc were investigated. The results obtained suggest that latency in rat intestinal microsomes may be due to endproduct inhibition by UDP. This endproduct inhibition could be abolished by in vitro treatment with MgCl2 and Triton X-100.  相似文献   

18.
A method was developed for simultaneous quantification of urinary 1- and 2-naphthols, 3- and 9-phenanthrols and 1-pyrenol using gas chromatography with mass spectrometry (GC-MS). This method was applied to urine samples from coke oven workers (n =28) and controls (n =22) from Northern China. Geometric mean levels of urinary 1-naphthol (58.8 μg l?1), 2-naphthol (34.1 μg l?1), 3-phenanthrol (7.35 μg l?1), 9-phenanthrol (1.28 μg l?1) and 1-pyrenol (25.4 μg l?1) were significantly higher among coke oven workers than controls. All the substances tested were highest among top-of-oven workers, who had 15-fold higher 1-naphthol, eight-fold higher 2-naphthol and 20-fold higher 1-pyrenol levels compared with controls. Using multiple linear regression models, 72.5% of the variation in 1- and 2-naphthol and 82.8% of the variation in 1-pyrenol were explained by the concentration of naphthalene or pyrene in the urine, the work category and the smoking intensity. Cigarette consumption significantly contributed to levels of urinary 1-pyrenol and naphthols, particularly 2-naphthol. A negative relationship between work category and the ratio of naphthols/1-pyrenol was observed among smokers. Our results suggest that urinary naphthols and phenanthrols reflect polycyclic aromatic hydrocarbon (PAH) exposure as well as the widely used 1-pyrenol, and that interactions between cigarette smoking and PAH exposure result in different patterns of metabolism for individual PAHs.  相似文献   

19.
Alpha-crystallin is a member of the family of small heat-shock proteins (sHSP) and is composed of two subunits, alphaA-crystallin and alphaB-crystallin, which exhibit molecular chaperone-like properties. In a previous study, we found that residues 70-88 in alphaA-crystallin can function like a molecular chaperone by preventing the aggregation and precipitation of denaturing substrate proteins [Sharma, K. K., et al. (2000) J. Biol. Chem. 275, 3767-3771]. In this study, we show that the complementary sequence in alphaB-crystallin, residues 73-92 (DRFSVNLDVKHFSPEELKVK), is the functional chaperone site of alphaB-crystallin. Like the mini-alphaA-crystallin chaperone, the mini-alphaB-crystallin chaperone interacts with 1,1'-bi(4-anilino) naphthalene-5,5'-disulphonic acid (bis-ANS) and also possesses significant beta-sheet and random coil structure. Deletion of four residues (DRFS) from the N-terminus or deletion of C-terminus LKVK residues from the 73-92 peptide abolishes the chaperone-like activity against denaturing alcohol dehydrogenase. However, removal of DRFS or HFSPEELKVK is necessary to completely abolish the antiaggregation property of the peptide in insulin reduction assay. Substitution of Asp at a site corresponding to D80 in alphaB-crystallin with d-Asp or beta-Asp results in a significant loss of chaperone-like activity. Kynurenine modification of His in the peptide abolishes the antiaggregation property of the mini-chaperone. These data suggest that the 73-92 region in alphaB-crystallin is one of the substrate binding sites during chaperone activity.  相似文献   

20.
Conventional methods to determine esterase activity from insects are composed of a three-step process where the enzyme is allowed to hydrolyze a 1-naphthyl acetate substrate, that reaction is quenched by a SDS detergent, and then a Fast Blue B dye complex is formed with 1-naphthol, the product of 1-naphthyl acetate hydrolysis. These methods measure dye-product complex rather than the product, 1-naphthol. A new assay is presented that continuously monitors the formation of 1-naphthol with the hydrolysis of an esterase substrate. The esterase activity was determined as the slope of the linear regression change in absorbance over time at 320 nm. The continuous assay provides a simple, rapid, and sensitive method for measuring esterases extracted from a single diamondback moth in 1-10 min. The detection limit of the assay is approximately 0.6 microM 1-naphthol. The 1-naphthol product from the esterase reaction was confirmed by HPLC analysis. According to the assay, the K(m) and V(max) values of the esterase were 28 +/- 2 microM and 6.0 +/- 0.1 microM/min, respectively, at 37 degrees C for 1-naphthyl acetate. The K(i) value was 9 +/- 2 microM using azadirachtin, an insecticide from neem tree, Azadirachta indica (A.Juss). Azadirachtin was a reversible competitive inhibitor of the esterase activity.  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