首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 293 毫秒
1.
2.
To determine if singlet oxygen (O2(1 delta g)) is produced by neutrophils (PMNs) during the process of phagocytosis, glass beads were coated with a specific chemical trap for O2(1 delta g), 9,10-diphenylanthracene (DPA). Singlet oxygen, but not other reactive oxygen species, reacts rapidly with DPA at a rate of kr = 1.3 x 10(6) M-1 s-1 to form a stable product, DPA-endoperoxide (Corey, E. J., and Taylor, W. C. (1964) J. Am. Chem. Soc. 86, 3881-3882; Wasserman, H. H., Scheffer, J. R., and Cooper, J. L. (1972) J. Am. Chem. Soc. 94, 4991-4996; Turro, N. J., Chow, M.-F., and Rigaudy, J. (1981) J. Am. Chem. Soc. 103, 7218-7224). The production of DPA-endoperoxide was determined by ultraviolet spectroscopy as a decrease in DPA absorbance at 355 nm. The absorbance of DPA was normalized to the absorbance of perylene, which was included in the coating on the beads as a nonreactive, internal standard. In the present study, DPA- and perylene-coated beads were initially allowed to adhere to fibronectin-coated coverslips. PMNs were then added to the bead-coated coverslips and allowed to adhere and phagocytose the beads for 1 h at 37 degrees C. In some experiments, 4B-phorbol-12-myristate-13-acetate (PMA) (1 ng/2.5 x 10(7) cells/ml), a known activator of the PMN NADPH-oxidase, was added as a co-stimulant. The amount of O2(1 delta g) produced by phagocytically stimulated PMNs was calculated to be 11.3 +/- 4.9 nmol of O2(1 delta g)/1.25 x 10(6) cells. Low dose PMA co-stimulation increased the production of O2(1 delta g) to 14.1 +/- 4.1 nmol/1.25 x 10(6) cells. Averaged together these amounts represent approximately 19 +/- 5.0% of the total oxygen consumed by PMNs in response to DPA- and perylene-coated beads. The specificity of the DPA reaction with O2(1 delta g) was confirmed by warming to 120 degrees C, which releases O2(1 delta g) from the DPA-endoperoxide, regenerating the parent DPA compound (Wasserman et al., 1972; Turro et al., 1981) and the absorbance at 355 nm. In addition, beta-carotene, an avid quencher of O2(1 delta g), was included in the coating of some bead preparations; assays in which these beads were used showed no change in the absorbance at 355 nm. Singlet oxygen production by myeloperoxidase was also measured using the coated bead assay and the results suggest that this is a major pathway by which singlet oxygen is generated in phagocytically stimulated PMNs.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

3.
B H Oh  E S Mooberry  J L Markley 《Biochemistry》1990,29(16):4004-4011
Multinuclear two-dimensional NMR techniques were used to assign nearly all diamagnetic 13C and 15N resonances of the plant-type 2Fe.2S* ferredoxin from Anabaena sp. strain PCC 7120. Since a 13C spin system directed strategy had been used to identify the 1H spin systems [Oh, B.-H., Westler, W. M., & Markley, J. L. (1989) J. Am. Chem. Soc. 111, 3083-3085], the sequence-specific 1H assignments [Oh, B.-H., & Markley, J. L. (1990) Biochemistry (first paper of three in this issue)] also provided sequence-specific 13C assignments. Several resonances from 1H-13C groups were assigned independently of the 1H assignments by considering the distances between these nuclei and the paramagnetic 2Fe.2S* center. A 13C-15N correlation data set was used to assign additional carbonyl carbons and to analyze overlapping regions of the 13C-13C correlation spectrum. Sequence-specific assignments of backbone and side-chain nitrogens were based on 1H-15N and 13C-15N correlations obtained from various two-dimensional NMR experiments.  相似文献   

4.
The eucaryotic microorganism, Neurospora crassa, is able under specified conditions (Zamir, L.O., Jung, E., and Jensen, R.A. (1982) J. Biol. Chem. 258, 6492-6496) to synthesize a cyclohexadienyl derivative of prephenic acid having the novel structure of a spiro-gamma-lactam. This L-gamma-(spiro-4-hydroxy-2,5-cyclohexadienyl)-pyroglutamate is herein given the trivial name, spiro-arogenate, to indicate its close relationship to the amino acid, L-arogenate. Spiro-arogenate is quantitatively converted to phenylalanine at mildly acidic pH and can be converted to arogenate by boiling at basic pH. The structure of spiro-arogenate was established through the application of spectroscopic techniques (ultraviolet, 1H-NMR, 13C-NMR, and mass spectrometry). The 1H-NMR and 13C-NMR spectra of spiro-arogenate isolated as the natural product conformed to the spectrum of spiro-arogenate prepared by chemical synthesis by S. Danishefsky and co-workers (Danishefsky, S., Morris, J., and Clizbe, L.A. (1981) J. Am. Chem. Soc. 103, 1602-1604). Circular dichroism established the S configuration of the asymmetric carbon at C-8 of spiro-arogenate.  相似文献   

5.
Previously, it was reported that a newly isolated microbial culture, Clavibacter sp. strain ALA2, produced trihydroxy unsaturated fatty acids, diepxoy bicyclic fatty acids, and tetrahydroxyfuranyl fatty acids (THFAs) from linoleic acid (C. T. Hou, J. Am. Oil Chem. Soc. 73:1359-1362, 1996; C. T. Hou and R. J. Forman III, J. Ind. Microbiol. Biotechnol. 24:275-276, 2000; C. T. Hou, H. Gardner, and W. Brown, J. Am. Oil Chem. Soc. 75:1483-1487, 1998; C. T. Hou, H. W. Gardner, and W. Brown, J. Am. Oil Chem. Soc. 78:1167-1169, 2001). In this study, we found that Clavibacter sp. strain ALA2 produced novel THFAs, including 13,16-dihydroxy-12-THFA, 15-epoxy-9(Z)-octadecenoic acid (13,16-dihydroxy-THFA), and 7,13,16-trihydroxy-12, 15-epoxy-9(Z)-octadecenoic acid (7,13,16-trihydroxy-THFA), from alpha-linolenic acid (9,12,15-octadecatrienoic acid). The chemical structures of these products were determined by gas chromatography-mass spectrometry and proton and (13)C nuclear magnetic resonance analyses. The optimum incubation temperature was 30 degrees C for production of both hydroxy-THFAs. 13,16-Dihydroxy-THFA was detected after 2 days of incubation, and the concentration reached 45 mg/50 ml after 7 days of incubation; 7,13,16-trihydroxy-THFA was not detected after 2 days of incubation, but the concentration reached 9 mg/50 ml after 7 days of incubation. The total yield of both 13,16-dihydroxy-THFA and 7,13,16-trihydroxy-THFA was 67% (wt/wt) after 7 days of incubation at 30 degrees C and 200 rpm. In previous studies, it was reported that Clavibacter sp. strain ALA2 oxidized the C-7, C-12, C-13, C-16, and C-17 positions of linoleic acid (n-6) into hydroxy groups. In this case, the bond between the C-16 and C-17 carbon atoms is saturated. In alpha-linolenic acid (n-3), however, the bond between the C-16 and C-17 carbon atoms is unsaturated. It seems that enzymes of strain ALA2 oxidized the C-12-C-13 and C-16-C-17 double bonds into dihydroxy groups first and then converted them to hydroxy-THFAs.  相似文献   

6.
K J Gruys  C J Halkides  P A Frey 《Biochemistry》1987,26(24):7575-7585
The synthesis of 2-acetylthiamin pyrophosphate (acetyl-TPP) is described. The synthesis of this compound is accomplished at 23 degrees C by the oxidation of 2-(1-hydroxyethyl)thiamin pyrophosphate using aqueous chromic acid as the oxidizing agent under conditions where Cr(III) coordination to the pyrophosphoryl moiety and hydrolysis of both the pyrophosphate and acetyl moieties were prevented. Although the chemical properties exhibited by acetyl-TPP are similar to those of the 2-acetyl-3,4-dimethylthiazolium ion examined by Lienhard [Lienhard, G.E. (1966) J. Am. Chem. Soc. 88, 5642-5649], significant differences exist because of the pyrimidine ring in acetyl-TPP. Characterization of acetyl-TPP by ultraviolet spectroscopy, 1H NMR, 13C NMR, and 31P NMR provided evidence that the compound in aqueous solution exists as an equilibrium mixture of keto, hydrate, and intramolecular carbinolamine forms. The equilibria for the hydration and carbinolamine formation reactions at pD 1.3 as determined by 1H NMR are strongly dependent on the temperature, showing an increase in the hydrate and carbinolamine forms at the expense of the keto form with decreasing temperature. The concentration of keto form also decreases with increasing pH. Acetyl-TPP is stable in aqueous acid but rapidly deacetylates at higher pH to form acetate and thiamin pyrophosphate. Trapping of the acetyl moiety in aqueous solution occurs efficiently with 1.0 M hydroxylamine at pH 5.5-6.5 to form acetohydroxamic acid and to a much smaller extent with 1.0 M 2-mercaptoethanol at pH 4.0 and 5.0 to form thio ester. Transfer of the acetyl group to 0.5 M dihydrolipoic acid at pH 5.0 and 1.0 M phosphate dianion at pH 7.0 is not observed to any significant extent in water. The kinetic and thermodynamic reactivity of acetyl-TPP with water and other nucleophiles is compatible with a hypothetical role for acyl-TPPs as enzymatic acyl-transfer intermediates.  相似文献   

7.
Previously, it was reported that a newly isolated microbial culture, Clavibacter sp. strain ALA2, produced trihydroxy unsaturated fatty acids, diepxoy bicyclic fatty acids, and tetrahydroxyfuranyl fatty acids (THFAs) from linoleic acid (C. T. Hou, J. Am. Oil Chem. Soc. 73:1359-1362, 1996; C. T. Hou and R. J. Forman III, J. Ind. Microbiol. Biotechnol. 24:275-276, 2000; C. T. Hou, H. Gardner, and W. Brown, J. Am. Oil Chem. Soc. 75:1483-1487, 1998; C. T. Hou, H. W. Gardner, and W. Brown, J. Am. Oil Chem. Soc. 78:1167-1169, 2001). In this study, we found that Clavibacter sp. strain ALA2 produced novel THFAs, including 13,16-dihydroxy-12-THFA, 15-epoxy-9(Z)-octadecenoic acid (13,16-dihydroxy-THFA), and 7,13,16-trihydroxy-12, 15-epoxy-9(Z)-octadecenoic acid (7,13,16-trihydroxy-THFA), from α-linolenic acid (9,12,15-octadecatrienoic acid). The chemical structures of these products were determined by gas chromatography-mass spectrometry and proton and 13C nuclear magnetic resonance analyses. The optimum incubation temperature was 30°C for production of both hydroxy-THFAs. 13,16-Dihydroxy-THFA was detected after 2 days of incubation, and the concentration reached 45 mg/50 ml after 7 days of incubation; 7,13,16-trihydroxy-THFA was not detected after 2 days of incubation, but the concentration reached 9 mg/50 ml after 7 days of incubation. The total yield of both 13,16-dihydroxy-THFA and 7,13,16-trihydroxy-THFA was 67% (wt/wt) after 7 days of incubation at 30°C and 200 rpm. In previous studies, it was reported that Clavibacter sp. strain ALA2 oxidized the C-7, C-12, C-13, C-16, and C-17 positions of linoleic acid (n-6) into hydroxy groups. In this case, the bond between the C-16 and C-17 carbon atoms is saturated. In α-linolenic acid (n-3), however, the bond between the C-16 and C-17 carbon atoms is unsaturated. It seems that enzymes of strain ALA2 oxidized the C-12-C-13 and C-16-C-17 double bonds into dihydroxy groups first and then converted them to hydroxy-THFAs.  相似文献   

8.
Near-UV circular dichroic (CD) spectra of three colchicine analogues that differ at the C-10 position have been obtained in the presence and absence of tubulin. All three colchicine analogues show dramatic alterations in the low-energy near-UV CD band upon tubulin binding that cannot be mimicked by solvent, but in no event does the rotational strength of the CD band decrease to nearly zero as in the case of colchicine [Detrich, H. W., III, Williams, R. C., Jr., Macdonald, T. L., & Puett, D. (1981) Biochemistry 20, 5999-6005]. The effect of self-association of colchicine and one of the C-10 analogues, thiocolchicine, on the near-UV CD band was also investigated. A qualitative similarity was seen between the near-UV CD spectra of colchicine and thiocolchicine dimers and the spectra of these molecules bound to tubulin. These observations support the previous suggestion that ligands bound to the colchicine site on tubulin may be interacting with an aromatic amino acid in the colchicine binding site [Hastie, S. B., & Rava, R. P. (1989) J. Am. Chem. Soc. 110, 6993-7001].  相似文献   

9.
Spectra of intermediates in oxidation and reduction of cytochrome c oxidase   总被引:1,自引:0,他引:1  
Two kinetic components with distinct difference spectra occur during reduction of cytochrome c oxidase by ruthenium hexamine. They are attributed to reduction of heme a (fast phase) and heme a3 (slow phase) (Scott, R. A., and Gray, H. B. (1980) J. Am. Chem. Soc. 102, 3219-3774). Two spectra seen during oxidation of cytochrome c oxidase by molecular oxygen have also been attributed to oxidation of hemes a3 and a (Greenwood, C., and Gibson, Q. H. (1967) J. Biol. Chem. 242, 1782-1787). We now report that spectra for the reductive and oxidative reactions obtained with the same preparations and the same apparatus under similar conditions are significantly different. The reactions appear to populate different reaction intermediates. Reconstitution into phospholipid vesicles does not affect these two spectra significantly. During turnover, the chief intermediates are those of the reductive pathway (Scott and Gray type intermediates). Reduction of heme a3 occurs approximately 70 times faster after turnover than the reduction of the resting enzyme. This is probably a dramatic "pulsing" effect (Wilson, M. T., Peterson, J., Antonini, E., Brunori, M., Colosimo, A., and Wyman, J. (1981) Proc. Natl. Acad. Sci. U.S.A. 7115-7118).  相似文献   

10.
The type I Cu site in the Cys457Ser mutant of Myrothecium verrucaria bilirubin oxidase was vacant, but the trinuclear center composed of a type II Cu and a pair of type III Cu's was fully occupied by three Cu ions. Cys457Ser could react with dioxygen, affording reaction intermediate I with absorption maxima at 340, 470, and 675 nm. This intermediate corresponds to that obtained from laccase, whose type I Cu is cupric and type II and III Cu's are cuprous [Zoppellaro, G., Sakurai, T., and Huang, H. (2001) J. Biochem. 129, 949-953] or whose type I Cu is substituted with Hg [Palmer, A. E., Lee, S. K., and Solomon, E. I. (2001) J. Am. Chem. Soc. 123, 6591-6599]. Another type I Cu mutant, Met467Gln, with modified spectroscopic properties and redox potential, afforded reaction intermediate II with absorption maxima at 355 and 450 nm. This intermediate corresponds to that obtained during the reaction of laccase [Sundaram, U. M., Zhang, H. H., Hedman, B., Hodgson, K. O., and Solomon, E. I. (1997) J. Am. Chem. Soc. 119, 12525-12540; Huang, H., Zoppellaro, G., and Sakurai, T. (1999) J. Biol. Chem. 274, 32718-32724]. According to a three-dimensional model of bilirubin oxidase, Asp105 is positioned near the trinuclear center. Asp105Glu and Asp105Ala exhibited 46 and 7.5% bilirubin oxidase activity compared to the wild-type enzyme, respectively, indicating that Asp105 conserved in all multi-copper oxidases donates a proton to reaction intermediates I and II. In addition, this amino acid might be involved in the formation of the trinuclear center and in the binding of dioxygen based on the difficulties in incorporating four Cu ions in Asp105Ala and Asp105Asn and their reactions with dioxygen.  相似文献   

11.
The inhibition of the zinc enzyme carbonic anhydrase (CA, EC 4.2.1.1) with (thio)coumarins has been recently reported (Maresca et al., J. Am. Chem. Soc. 2009, 131, 3057). Here we demonstrate that a series of γ- and δ-(thio)lactones also act as mechanism based, prodrug type CA inhibitors, similar to the (thio)coumarins. Through the esterase activity of CA, these compounds are hydrolyzed in situ to the corresponding hydroxy/keto/mercapto acids which thereafter act as inhibitors. CA isoforms I and IX were efficiently inhibited by simple such compounds, with K(I)s in the range of 0.92-19.1μM, whereas CA II was not inhibited at all. Isoform-selective CA inhibitors which spare the ubiquitous off-target CA II may have interesting applications for example for selectively inhibiting the tumor-associated CA IX, a validated anticancer target.  相似文献   

12.
Clapp CH  Strulson M  Rodriguez PC  Lo R  Novak MJ 《Biochemistry》2006,45(51):15884-15892
Soybean lipoxygenase-1 (SBLO-1) catalyzes the oxygenation of polyunsaturated fatty acids to produce conjugated diene hydroperoxides. Previous work from our laboratories has demonstrated that SBLO-1 will also catalyze the oxygenation of monounsaturated acids (Clapp, C. H., Senchak, S. E., Stover, T. J., Potter, T. C., Findeis, P. M., and Novak, M. J. (2001) Soybean Lipoxygenase-Mediated Oxygenation of Monounsaturated Fatty Acids to Enones, J. Am. Chem. Soc. 123, 747-748). Interestingly, the products are alpha,beta-unsaturated ketones rather than the expected allylic hydroperoxides. In the present work, we provide evidence that the monoolefin substrates are initially converted to allylic hydroperoxides, which are subsequently converted to the enone products. The hydroperoxide intermediates can be trapped by reduction to the corresponding allylic alcohols with glutathione peroxidase plus glutathione or with SnCl2. Under some conditions, the hydroperoxide intermediates accumulate and can be detected by HPLC and peroxide assays. Kinetics measurements at low concentrations of [1-14C]-9(Z)-octadecenoic acid indicate that oxygenation of this substrate at 25 degrees C, pH 9.0 occurs with kcat/Km = 1.6 (+/-0.1) x 10(2) M-1 s-1, which is about 105 lower than kcat/Km for oxygenation of 9(Z),12(Z)-octadecadienoic acid (linoleic acid). Comparison of the activities of 9(Z)-octadecenoic acid and 12(Z)-octadecenoic acid implies that the two double bonds of linoleic acid contribute almost equally to the C-H bond-breaking step in the normal lipoxygenase reaction. The results are consistent with the notion that SBLO-1 functionalizes substrates by a radical mechanism.  相似文献   

13.
Carbon 13 NMR spectra have been obtained for aqueous solutions of DL-2-(alpha-hydroxyethyl)thiamin, DL-2-(alpha-hydroxybenzyl)thiamin, DL-2-(alpha-hydroxybenzyl)oxythiamin, and related N-3 methyl and N-3 benzyl analogs. The unusually large downfield shift of the 13C resonance of C-2 of hydroxyethylthiamin suggests that this carbon bears a partial positive charge. This result stands in contrast to results of x-ray crystallographic studies of hydroxyethylthiamin, which place a partial negative charge on C-2 (Pletcher, J., and Sax, M. (1974) J. Am. Chem. Soc. 96, 155-165). A partial positive charge on C-2 helps to explain the facility of carbanion formation at the alpha carbon both enzymatically and in model systems. The rates of proton-deuteron exchange of (C-alpha)-H with solvent deuterium, and of release of aldehyde to regenerate thiamin have been measured for hydroxyethylthiamin and analogs. The differences in kinetic acidity of (C-alpha)-H and of rates of aldehyde release are rationalized in terms of differing electron-withdrawing abilities of the substituents attached to N-3, and appear not to be related to intramolecular basic catalysis of these processes by the C-4' amino group.  相似文献   

14.
Previously we showed that 24(S),25-epoxycholesterol is formed from acetate, via squalene 2,3(S),22(S),23-dioxide and 24(S),25-oxidolanosterol, during the normal course of cholesterol biosynthesis in S10 rat liver homogenate (Nelson, J. A., Steckbeck, S. R., and Spencer, T. A. (1981) J. Biol. Chem. 256, 1067-1068; Nelson, J. A., Steckbeck, S. R., and Spencer, T. A. (1981) J. Am. Chem. Soc. 103, 6974-6975). Herein we demonstrate that the nonsaponifiable extract from human liver tissue contains 24(S),25-epoxycholesterol in an amount approximately 10(-3) relative to cholesterol. We show that 24(S),25-epoxycholesterol, like many other oxygenated sterols, represses hydroxymethylglutaryl-CoA reductase activity in cultured cells and binds to the cytosolic oxysterol-binding protein. Furthermore, we show that this epoxide is not rapidly metabolized in cultured cells. These results suggest that 24(S),25-epoxycholesterol may participate in the regulation of hepatic cholesterol metabolism in vivo.  相似文献   

15.
Time-resolved resonance Raman spectra have been recorded during the reaction of mixed valence (a3+ a2+(3)) cytochrome oxidase with dioxygen at room temperature. In the spectrum recorded at 10 microseconds subsequent to carbon monoxide photolysis, a mode is observed at 572 cm-1 that shifts to 548 cm-1 when the experiment is repeated with 18O2. The appearance of this mode is dependent upon the laser intensity used and disappears at higher incident energies. The high frequency data in conjunction with the mid-frequency data allow us to assign the 572 cm-1 mode to the Fe-O stretching vibration of the low-spin O2 adduct that forms in the mixed valence cytochrome oxidase/dioxygen reaction. The 572 cm-1 v(Fe2(+)-O2) frequency in the mixed valence enzyme/O2 adduct is essentially identical to the 571 cm-1 frequency we measured for this mode during the reduction of O2 by the fully reduced enzyme (Varotsis, C., Woodruff, W. H., and Babcock, G. T. (1989) J. Am. Chem. Soc. 111, 6439-6440; Varotsis, C., Woodruff, W. H., and Babcock, G. T. (1990) J. Am. Chem. Soc. 112, 1297), which indicates that the O2-bound cytochrome a3 site is independent of the redox state of the cytochrome a/CuA pair. The photolabile oxy intermediate is replaced by photostable low- or intermediate-spin cytochrome a3+(3), with t1/2 congruent to 200 microseconds.  相似文献   

16.
Microsomes from maize seedlings are capable of catalyzing the C-24 alkylation of 4,4,14 alpha-trimethyl-9 beta,19-cyclo-5 alpha-cholest-24-en-3 beta-ol (cycloartenol) by (S)-adenosyl-L-methionine (AdoMet) leading to 24-methylene cycloartanol. Derivatives of cycloartenol bearing a nitrogen atom at C-25 have been previously shown to be potent inhibitors of the AdoMet-cycloartenol-C-24-methyltransferase (Narula, A. S., Rahier, A., Benveniste, P., and Schuber, F. (1981) J. Am. Chem. Soc. 103, 2408-2409). In order to determine the molecular parameters of the inhibition and to gain information about its mechanism, various azasteroids and analogues have been synthesized and assayed. The following results have been obtained. i) The presence of a positive charge at position 25 was found to be the major cause of the inhibition since electrostatically neutral isosteric compounds possessing a carbon in place of the nitrogen atom were not inhibitory. The positive charge leading to inhibition may be conferred by a protonated amine, a quaternary ammonium group, as well as by a sulfonium or an arsonium group. ii) A steroid-like structure of the inhibitor was also important. And iii) the presence of a free 3 beta-hydroxy group and the bent conformation of cycloartenol, which are essential molecular features of the substrate for the methylation reaction, were no longer required to observe inhibition. The data obtained strongly support the idea that C-25 heteroatoms (N, As, and S), substituted triterpenoid derivatives possessing a positive charge at position 25, are analogues of a carbocationic high-energy intermediate involved during the reaction catalyzed by the AdoMet-cycloartenol-C-24-methyltransferase.  相似文献   

17.
A C Anusiem  M Kelleher 《Biopolymers》1984,23(7):1147-1167
Interest in the thermodynamics of the iron-binding site in hemoproteins has increased in recent years due to refinements in x-ray crystallographic studies of hemoproteins [see Deathage, J. F., Lee, R. S., Anderson, C. M. & Moffat, K. (1976) J. Mol. Biol. 104 , 687–706; Heidner, E. J., Ladner, R. C. & Perutz, M. F. (1976) J. Mol. Biol. 104 , 707–722; Deathage, J. F., Lee, R. S. & Moffat, K. (1976) J. Mol. Biol. 104 , 723–728; Ladner, R. C., Heidner, E. J. & Perutz, M. F. (1976) J. Mol. Biol. 114 , 385–414; Fermi, G. & Perutz, M. F. (1977) J. Mol. Biol. 114 , 421–431; Takano, T. (1977) J. Mol. Biol. 110 , 537–568 and 569–589], the synthesis and x-ray analysis of model heme compounds [see Scheidt, W. R. (1977) Acc. Chem. Res. 10 , 339–345; Kastner, M. E., Scheidt, W. R., Mashino, T. & Reed, C. A. (1978) J. Am. Chem. Soc. 100 , 666–667; Mashiko, T., Kastner, M. E., Spartalian, K., Scheidt, W. R. & Reed, C. A. (1978) J. Am. Chem. Soc. 100 , 6354–6362; Hill, H. A. O., Skite, P. P., Buchler, J. W., Luchr, H., Tonn, M., Gregson, A. K. & Pellizer, G. (1979) Chem. Commun. 4 , 151–152; and Scheidt, W. R., Cohen, I. A. & Kastner, M. E. (1979) Biochemistry 18 , 3546–3556], and the numerous data on heme–protein interactions that account for the differences observed in ligand binding between the various species of animals. Numerous probes have been used and provide information about the structure and thermodynamics of the binding site, but no single probe can provide the complete picture [see Iizuka, T. & Yonetani, T. (1970) Adv. Biophys. 1 , 157–182; Smith, D. W. & Williams, R. J. P. (1970) Struct. Bond. 7 , 1–45; and Spiro, T. G. (1975) Biochim. Biophys. Acta 416 , 169–189].  相似文献   

18.
The first positive evidence for the utilization of a direct C-6' ' oxidation/reduction mechanism by ADP-l-glycero-d-manno-heptose 6-epimerase is reported here. The epimerase (HldD or AGME, formerly RfaD) operates in the biosynthetic pathway of l-glycero-d-manno-heptose, which is a conserved sugar in the core region of lipopolysaccharide (LPS) of Gram-negative bacteria. The stereochemical inversion catalyzed by the epimerase is interesting as it occurs at an "unactivated" stereocenter that lacks an acidic C-H bond, and therefore, a direct deprotonation/reprotonation mechanism cannot be employed. Instead, the epimerase employs a transient oxidation strategy involving a tightly bound NADP(+) cofactor. A recent study ruled out mechanisms involving transient oxidation at C-4' ' and C-7' ' and supported a mechanism that involves an initial oxidation directly at the C-6' ' position to generate a 6' '-keto intermediate (Read, J. A., Ahmed, R. A., Morrison, J. P., Coleman, W. G., Jr., Tanner, M. E. (2004) J. Am. Chem. Soc. 126, 8878-8879). A subsequent nonstereospecific reduction of the ketone intermediate can generate either epimer of the ADP-heptose. In this work, an intermediate analogue containing an aldehyde functionality at C-6' ', ADP-beta-d-manno-hexodialdose, is prepared in order to probe the ability of the enzyme to catalyze redox chemistry at this position. It is found that incubation of the aldehyde with a catalytic amount of the epimerase leads to a dismutation process in which one-half of the material is oxidized to ADP-beta-d-mannuronic acid and the other half is reduced to ADP-beta-d-mannose. Transient reduction of the enzyme-bound NADP(+) was monitored by UV spectroscopy and implicates the cofactor's involvement during catalysis.  相似文献   

19.
E John  F Jhnig 《Biophysical journal》1992,63(6):1536-1543
An analogue of melittin synthesized in the group of E. T. Kaiser (DeGrado, W. F., F. J. Keźdy, and E. T. Kaiser. 1981. J. Am. Chem. Soc. 103:679-681) was investigated by Raman spectroscopy and fluorescence anisotropy decay. In water, the analogue is completely alpha-helical and aggregates in large oligomers of about 50 monomers. In vesicle membranes, it undergoes orientational fluctuations similar to melittin. The most significant difference from melittin, therefore, is the formation of straight helixes and their aggregation in large oligomers in water. We interpret this as a consequence of the lacking proline residue in the analogue. We, furthermore, hypothesize that the increased tendency for aggregation causes the increased hemolytic activity of the analogue.  相似文献   

20.
B H Oh  J L Markley 《Biochemistry》1990,29(16):3993-4004
Complete sequence-specific assignments were determined for the diamagnetic 1H resonances from Anabaena 7120 ferredoxin (Mr = 11,000). A novel assignment procedure was followed whose first step was the identification of the 13C spin systems of the amino acids by a 13C(13C) double quantum correlation experiment [Oh, B.-H., Westler, M. W., Darba, P., & Markley, J. L. (1988) Science 240, 908-911]. Then, the 1H spin systems of the amino acids were identified from the 13C spin systems by means of direct and relayed 1H(13C) single-bond correlations [Oh, B.-H., Westler, W. M., & Markley, J. L. (1989) J. Am. Chem. Soc. 111, 3083-3085]. The sequential resonance assignments were based mainly on conventional interresidue 1H alpha i-1HNi + 1 NOE connectivities. Resonances from 18 residues were not resolved in two-dimensional 1H NMR spectra. When these residues were mapped onto the X-ray crystal structure of the homologous ferredoxin from Spirulina platensis [Fukuyama, K., Hase, T., Matsumoto, S., Tsukihara, T., Katsube, Y., Tanaka, N., Kakudo, M., Wada, K., & Matsubara, H. (1980) Nature 286, 522-524], it was found that they correspond to amino acids close to the paramagnetic 2Fe.2S* cluster. Cross peaks in two-dimensional homonuclear 1H NMR spectra were not observed for any protons closer than about 7.8 A to both iron atoms. Secondary structural features identified in solution include two antiparallel beta-sheets, one parallel beta-sheet, and one alpha-helix.  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-23

京公网安备 11010802026262号