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.
Bovine dopamine beta-hydroxylase was examined spectroscopically for the presence of covalently bound pyrroloquinoline quinone (PQQ). Pure dopamine beta-hydroxylase had a featureless UV-visible spectrum above 300 nm. An equimolar solution of dopamine beta-hydroxylase and exogenously added PQQ (1 PQQ/active site) had a strong absorption maximum at 333 nm. Dialysis removed the added PQQ, indicating that dopamine beta-hydroxylase does not bind PQQ irreversibly. Reaction of dopamine beta-hydroxylase with 6 mM phenylhydrazine in the presence of 15 mM ascorbate caused 96% inactivation within 20 min and did not produce any spectrally detectable amounts of the phenylhydrazone adduct of PQQ, as reported by van der Meer et al. (van der Meer, R.A., Jongejan, J.A., and Duine, J.A. (1988) FEBS Lett. 231, 303-307). The peptide profile of phenylhydrazine inactivated dopamine beta-hydroxylase was monitored at 316 nm and did not reveal any peptides that might contain a PQQ-phenylhydrazone adduct. Thus, the absence of any spectrally detectable PQQ-phenylhydrazone adducts under these conditions demonstrates that the mechanism of phenylhydrazine inactivation does not involve covalent modification of PQQ at the active site of dopamine beta-hydroxylase and provides strong evidence that the native enzyme does not contain PQQ.  相似文献   

2.
Plasma amine oxidases catalyze the oxidative deamination of amines to aldehydes, followed by a 2e- reduction of O2 to H2O2. Pyrroloquinoline quinone (PQQ), previously believed to be restricted to prokaryotes, has recently been proposed to be the cofactor undergoing reduction in the first half-reaction of bovine plasma amine oxidase (Ameyama, M., Hayashi, U., Matsushita, K., Shinagawa, E., and Adachi, O. (1984) Agric. Biol. Chem. 48, 561-565; Lobenstein-Verbeek, C. L., Jongejan, J. A., Frank, J., and Duine, J. A. (1984) FEBS Lett. 170, 305-309). This result is unexpected, since model studies with PQQ implicate Schiff's base formation between a reactive carbonyl and substrates, whereas experiments with bovine plasma amine oxidase have failed to provide evidence for a carbonyl cofactor. We have, therefore, re-examined putative adducts between substrate and enzyme-bound cofactor, employing a combination of [14C]benzylamine and [3H]NaCNBH3. The use of the relatively weak reductant, NaCNBH3, affords Schiff's base specificity and permits the study of enzyme below pH 7.0. As we show, enzyme can only be inactivated by NaCNBH3 in the presence of substrate, leading to the incorporation of 1 mol of [14C]benzylamine/mol of enzyme subunit at complete inactivation. By contrast, we are unable to detect any labeling with [3H]NaCNBH3, analogous to an earlier study with [3H]NaCNBH4 (Suva, R. H., and Abeles, R. H. (1978) Biochemistry 17, 3538-3545). We conclude, first, that our inability to obtain adducts containing both carbon 14 and tritium rules out the reductive trapping either of amine substrate with pyridoxal phosphate or of aldehyde product with a lysyl side chain and, second, that the observed pattern of labeling is fully consistent with the presence of PQQ at the active site of bovine plasma amine oxidase.  相似文献   

3.
We have detected weak absorption bands in the near infra-red region of reduced mammalian cytochrome c oxidase, analogous to those that we have recently reported to be present in the bacterial cytochrome o (Ingledew, W.J., Bacon, M. and Rich, P.R. (1992) FEBS Lett. 305, 167-170). The major band is centred at 784 nm and has an sigma mM-1.cm-1 of around 0.1. It is shifted to 760 nm in the carbon monoxide compound and is absent in the reduced cyanide complex. We attribute it to a charge transfer band of ferrohaem a3, equivalent to the 'band III' or 'conformational band' of haemoglobin.  相似文献   

4.
Light-induced forward electron transfer in the bacterial photosynthetic reaction center from Rhodobacter sphaeroides was investigated by time-resolved infrared spectroscopy. Using a highly sensitive kinetic photometer based on a tunable IR diode laser source [M?ntele, W., Hienerwadel, R., Lenz, F., Riedel, W. J., Grisar, R., & Tacke, M. (1990a) Spectrosc. Int. 2, 29-35], molecular processes concomitant with electron-transfer reactions were studied in the microsecond-to-second time scale. Infrared (IR) signals in the 1780-1430-cm-1 spectral region, appearing within the instrument time resolution of about 0.5 microseconds, could be assigned to molecular changes of the primary electron donor upon formation of a radical cation and to modes of the primary quinone electron acceptor QA and its environment upon formation of QA-. These IR signals are consistent with steady-state FTIR difference spectra of the P+Q- formation [M?ntele, W., Nabedryk, E., Tavitian, B. A., Kreutz, W., & Breton, J. (1985) FEBS Lett. 187, 227-232; M?ntele, W., Wollenweber, A., Nabedryk, E., & Breton, J. (1988) Proc. Natl. Acad. Sci. U.S.A. 85, 8468-8472; Nabedryk, E., Bagley, K. A., Thibodeau, D. L., Bauscher, M., M?ntele, W., & Breton, J. (1990) FEBS Lett. 266, 59-62] and with time-resolved FTIR studies [Thibodeau, D. L., Nabedryk, E., Hienerwadel, R., Lenz, F., M?ntele, W., & Breton, J. (1990) Biochim. Biophys. Acta 1020, 253-259]. At given wavenumbers, kinetic components with a half-time of approximately 120 microseconds were observed and attributed to QA----QB electron transfer. The time-resolved IR signals, in contrast to steady-state experiments where full protein relaxation after electron transfer can occur, allow us to follow directly the modes of QA and QB and their protein environment under conditions of forward electron transfer. Apart from signals attributed to the primary electron donor, signals are proposed to arise not only from the C = O and C = C vibrational modes of the neutral quinones and from the C-O and C-C vibrations of their semiquinone anion form but also from amino acid groups forming their binding sites. Some of the signals appearing with the instrument rise time as well as the transient 120-microseconds signals are interpreted in terms of binding and interaction of the primary and secondary quinone electron acceptor in the Rb. sphaeroides reaction center and of the conformational changes in their binding site.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

5.
Two human genes encoding precursors for two calcitonin gene-related peptides (CGRP) I (or alpha) and II (or beta) have been identified (Steenbergh, P. H., H?ppener, J. W. M., Zandberg, J., Lips, C. J. M., and Jansz, H. S. (1985) FEBS Lett. 183, 403-407). The amino acid sequence of CGRP-I was obtained in medullary thyroid carcinoma extracts (Morris, H. R., Panico, M., Etienne, T., Tippins, J., Girgis, S. I., and MacIntyre, I. (1984) Nature 308, 746-748), but not in normal human tissues. The human CGRP-II peptide remained to be discovered. Here we have determined in the human spinal cord the amino acid composition and the partial amino acid sequence of the DNA-predicted CGRP-I and -II. The data indicate for the first time the existence of a second CGRP different from the known CGRP-I. CGRP-II has been identified in the central nervous system, pituitary, thyroid, and in medullary thyroid carcinoma as a major CGRP form together with CGRP-I.  相似文献   

6.
Nitrile hydratase has been proved to be a quinoprotein with pyrroloquinoline quinone (PQQ) as a prosthetic group. The broad shoulder from 300 to 500 nm in the absorption spectrum of Brevibacterium nitrile hydratase suggested the presence of PQQ. Since PQQ was attached to the enzyme through a covalent linkage, the chromophores were isolated by acid hydrolysis, protease digestion and successive chromatographic separation. The isolated chromophores showed the similar spectroscopic characteristics to those of obtained from the amine oxidase of Aspergillus niger, in which PQQ is covalently linked. The isolated chromophores potently activated apo-D-glucose dehydrogenase (EC 1.1.99.17), supporting the presence of PQQ or a PQQ-like compound in nitrile hydratase. The finding of PQQ in nitrile hydratase strongly suggests a new function of PQQ, i.e., the activation of H2O in the enzymatic hydration reaction.  相似文献   

7.
Shokes JE  Duin EC  Bauer C  Jaun B  Hedderich R  Koch J  Scott RA 《FEBS letters》2005,579(7):1741-1744
Heterodisulfide reductase (HDR) catalyzes the formation of coenzyme M (CoM-SH) and coenzyme B (CoB-SH) by the reversible reduction of the heterodisulfide, CoM-S-S-CoB. This reaction recycles the two thiol coenzymes involved in the final step of microbial methanogenesis. Electron paramagnetic resonance (EPR) and variable-temperature magnetic circular dichroism spectroscopic experiments on oxidized HDR incubated with CoM-SH revealed a S=1/2 [4Fe-4S]3) cluster, the EPR spectrum of which is broadened in the presence of CoM-33SH [Duin, E.C., Madadi-Kahkesh, S., Hedderich, R., Clay, M.D. and Johnson, M.K. (2002) Heterodisulfide reductase from Methanothermobacter marburgensis contains an active-site [4Fe-4S] cluster that is directly involved in mediating heterodisulfide reduction. FEBS Lett. 512, 263-268; Duin, E.C., Bauer, C., Jaun, B. and Hedderich, R. (2003) Coenzyme M binds to a [4Fe-4S] cluster in the active site of heterodisulfide reductase as deduced from EPR studies with the [33S]coenzyme M-treated enzyme. FEBS Lett. 538, 81-84]. These results provide indirect evidence that the disulfide binds to the iron-sulfur cluster during reduction. We report here direct structural evidence for this interaction from Se X-ray absorption spectroscopic investigation of HDR treated with the selenium analog of coenzyme M (CoM-SeH). Se K edge extended X-ray absorption fine structure confirms a direct interaction of the Se in CoM-SeH-treated HDR with an Fe atom of the Fe-S cluster at an Fe-Se distance of 2.4A.  相似文献   

8.
Pseudomonas aeruginosa ATCC 17933 when grown on ethanol produces high levels of a quinoprotein ethanol dehydrogenase, which amounts to 7% of the soluble protein. The enzyme has been purified to homogeneity and it crystallizes readily in the presence of polyethylene glycol 1550 or 6000. The ethanol dehydrogenase (Km(ethanol) = 14 microM) resembles the dye-dependent quinoprotein methanol dehydrogenases of methylotrophic bacteria, but has a low affinity for methanol (Km (methanol) = 94mM). In addition the enzyme oxidizes secondary alcohols. With its catalytic properties the ethanol dehydrogenase is similar to the enzyme isolated from P. aeruginosa LMD 80.53 (Groen, B., Frank, J. Jzn. & Duine, J.A. (1984) Biochem. J. 223, 921-924). In contrast to this enzyme from P. aeruginosa LMD 80.53, which is a monomer, the ethanol dehydrogenase isolated from P. aeruginosa ATCC 17933 is a dimer of identical subunits of relative molecular mass 60,000. The N-terminal amino acid is lysine. Inactivation with cyclopropanone ethylhemiketal reveals one molecule of pyrroloquinoline quinone per subunit. As shown by active enzyme sedimentation, the dimer is the enzymatically active form.  相似文献   

9.
A new extracellular quinoprotein oxidase named enacyloxin oxidase (ENX oxidase), which is involved in biosynthesis of ENX IIa, a congener of ENX, was found in the culture supernatant of Frateuria sp. W-315 and purified as a homogeneous protein on SDS–PAGE. ENX oxidase was shown to have a molecular mass of 73 kDa by SDS–PAGE and 79 kDa by gel filtration. The enzyme was inhibited by various carbonyl reagents and the activity was stimulated by addition of PQQ. This is the first report on a quinoprotein oxidase that is secreted into the culture medium in the logarithmic growth phase, and acts for biosynthesis of the antibiotic.  相似文献   

10.
Peptide-analogues of the A and B families of silkmoth chorion proteins form amyloid fibrils under a variety of conditions [Iconomidou, V.A., Vriend, G. Hamodrakas, S.J. 2000. Amyloids protect the silkmoth oocyte and embryo. FEBS Lett. 479, 141-145; Iconomidou,V.A., Chryssikos, G.D.,Gionis, V., Vriend, G., Hoenger, A., Hamodrakas, S.J., 2001. Amyloid-like fibrils from an 18-residue peptide-analogue of a part of the central domain of the B-family of silkmoth chorion protein. FEBS Lett. 499, 268-273; Hamodrakas, S.J. Hoenger, A., Iconomidou, V. A., 2004 . Amyloid fibrillogenesis of silkmoth chorion protein peptide-analogues via a liquid crystalline intermediate phase. J. Struct. Biol. 145, 226-235.], which led us to propose that silkmoth chorion is a natural protective amyloid. In this study, we designed and synthesized two mutant peptide-analogues of the central conservative domain of the A family: (a) one, cA_m1, with a length half of that of the central domain of the A family, which folds and self-assembles, in various conditions, into amyloid fibrils very similar in properties and structure to the fibrils formed by the cA peptide, which corresponds to the entire length of the A family central domain [Iconomidou, V.A., Vriend, G. Hamodrakas, S.J. 2000. Amyloids protect the silkmoth oocyte and embryo. FEBS Lett. 479, 141-145.], in full support of our previous proposal, (b) the second, cA_m2, differing from cA_m1 at three positions, where three glutamates have replaced two valines and one alanine residues, does not form amyloid fibrils in any conditions. It appears that (a) the amyloidogenic properties of silkmoth chorion peptides are encoded into the tandemly repeating hexapeptides comprising the central domain of silkmoth chorion proteins, and, that (b) suitable mutations, properly and carefully designed, greatly affect the strong amyloidogenic properties inherent in certain aminoacid sequences and may inhibit amyloid formation.  相似文献   

11.
Caldesmon has two calmodulin-binding domains   总被引:1,自引:0,他引:1  
Chicken gizzard caldesmon was cleaved with chymotrypsin or CNBr, and the calmodulin-binding fragments were isolated using an affinity column. Limited chymotryptic digestion gives rise to a 38 kDa calmodulin-binding fragment (CT40) as described previously (Szpacenko, A. & Dabrowska, R., FEBS Lett. 202, 182-186, 1986; Fujii, T., Imai, M., Rosenfeld, G. C. & Bryan, J., J. Biol. Chem. 261, 16155-16160, 1987; Yazawa, M., Yagi, K. & Sobue, K., J. Biochem. 102, 1065-1073, 1987). In the case of CNBr cleavage a 37 kDa calmodulin-binding fragment (CB40) was obtained. Both CT40 and CB40 contain a reactive thiol group, but these thiols are apparently in different environments as judged by the responses of attached fluorescent labels to calmodulin-binding. A comparison of the N-terminal sequences of CB40 and CT40 with the complete sequence of caldesmon shows that the two calmodulin-binding fragments in fact originate from different parts of the parent molecule. Thus there exist two calmodulin-binding sites in caldesmon, one in the N-terminal half and the other in the C-terminal half of the molecule. This is consistent with the recent finding that up to two calmodulin molecules can be crosslinked to each caldesmon molecule (Wang, C.-L.A., Biochem. Biophys. Res. Commun., 156, 1033-1038, 1988).  相似文献   

12.
The precursor of mitochondrial aspartate aminotransferase accumulates in the cytosol of cultured chicken embryo fibroblasts if its import into mitochondria is inhibited by an uncoupling agent. However, its accumulation is limited by degradation with a half-life of only approximately 5 min (Jaussi, R., Sonderegger, P., Flückiger, J., and Christen, P. (1982) J. Biol. Chem. 257, 13334-13340). The aim of the present study was the characterization of the proteolytic system(s) responsible for this very rapid intracellular degradation. On depleting chicken embryo fibroblasts of ATP, the rate of degradation of the precursor was lowered by approximately 70%. Chicken embryo fibroblasts depleted of divalent metal ions showed a degradative activity of 10% of the initial value. Reconstitution of these cells with Mg2+ and Ca2+ increased the degradative activity from 10 to 107 and 24%, respectively. Thiol reagents almost completely prevented the degradation, whereas specific peptide inhibitors of cysteine proteases or inhibitors of intralysosomal proteolysis decreased the rate of degradation by only approximately 30%. Inhibitors of serine proteases had little effect. No rapid degradation of the precursor was observed in crude extracts of chicken embryo fibroblasts. The data indicate that the bulk of the precursor accumulated under conditions of import block is degraded by one or several cytosolic proteases dependent on ATP, Mg2+, and thiol groups of unknown localization, conceivably by proteolytic enzymes identical with or similar to one of the high molecular weight cytosolic proteases (Waxman, L., Fagan, J.M., Tanaka, K., and Goldberg, A. L. (1985) J. Biol. Chem. 260, 11994-12000). The rest of the precursor appears to be degraded by lysosomes.  相似文献   

13.
Previous crystallographic studies of the antibacterial trimethoprim in complexes with bacterial and avian dihydrofolate reductases have shown substantial differences in the mode of binding, providing plausible explanations for the origin of the remarkable species selectivity of this inhibitor (Matthews, D. A., Bolin, J. T., Burridge, J. M., Filman, D. J., Volz, K. W., Kaufman, B. T., Beddell, C. R., Champness, J. N., Stammers, D. K., and Kraut, J. (1985) J. Biol. Chem. 260, 381-391; Matthews, D. A., Bolin, J. T., Burridge, J. M., Filman, D. J., Volz, K. W., and Kraut, J. (1985) J. Biol. Chem. 260, 392-399). A major species difference between the active sites is that the only carboxylate present is always Glu in vertebrates and Asp in bacteria. Crystallographic studies of the wild-type and E30D mutant of the enzyme from mouse now reveal that in both cases trimethoprim is bound in an identical fashion to that observed with the bacterial enzyme, and there is no obvious single explanation for the origin of the 10(5)-fold selectivity of trimethoprim binding. In an earlier study of a mouse wild-type enzyme using more limited data it was proposed that trimethoprim bound in the avian mode (Stammers, D. K., Champness, J. N., Beddell, C. R., Dann, J. G., Eliopoulos, E. E., Geddes, A. J., Ogg, D., and North, A. C. T. (1987) FEBS Lett. 218, 178-184), but a re-examination indicates that the occupancy of the active site by trimethoprim is less than had been thought, and we are currently unable to make an unambiguous interpretation of the electron density maps and cannot confirm the avian mode of binding in those crystals.  相似文献   

14.
Several mutants of quinoprotein glucose dehydrogenase (GDH) in Escherichia coli, located around its cofactor pyrroloquinoline quinone (PQQ), were constructed by site-specific mutagenesis and characterized by enzymatic and kinetic analyses. Of these, critical mutants were further characterized after purification or by different amino acid substitutions. H262A mutant showed reduced affinities both for glucose and PQQ without significant effect on glucose oxidase activity, indicating that His-262 occurs very close to PQQ and glucose, but is not the electron acceptor from PQQH(2). W404A and W404F showed pronounced reductions of affinity for PQQ, and the latter rather than the former had equivalent glucose oxidase activity to the wild type, suggesting that Trp-404 may be a support for PQQ and important for the positioning of PQQ. D466N, D466E, and K493A showed very low glucose oxidase activities without influence on the affinity for PQQ. Judging from the enzyme activities of D466E and K493A, as well as their absorption spectra of PQQ during glucose oxidation, we conclude that Asp-466 initiates glucose oxidation reaction by abstraction of a proton from glucose and Lys-493 is involved in electron transfer from PQQH(2).  相似文献   

15.
Milk lactose is hydrolysed to galactose and glucose in the small intestine of mammals by the lactase/phlorizin hydrolase complex (LPH; EC 3.2.1.108/62). The two enzymatic activities, lactase and phlorizin hydrolase, are located in the same polypeptide chain. According to sequence homology, mature LPH contains two different regions (III and IV), each of them homologous to family 1 glycosidases and each with a putative active site. There has been some discrepancy with regard to the assignment of enzymatic activity to the two active sites. Here we show differential reactivity of the two active sites with mechanism-based glycosidase inhibitors. When LPH is treated with 2',4'-dinitrophenyl 2-deoxy-2-fluoro-beta-D-glucopyranoside (1) and 2', 4'-dinitrophenyl-2-deoxy-2-fluoro-beta-D-galactopyranoside (2), known mechanism-based inhibitors of glycosidases, it is observed that compound 1 preferentially inactivates the phlorizin hydrolase activity whereas compound 2 is selective for the lactase active site. On the other hand, glycals (D-glucal and D-galactal) competitively inhibit lactase activity but not phlorizin hydrolase activity. This allows labeling of the phlorizin site with compound 1 by protection with a glycal. By differential labeling of each active site using 1 and 2 followed by proteolysis and MS analysis of the labeled fragments, we confirm that the phlorizin hydrolysis occurs mainly at the active site located at region III of LPH and that the active site located at region IV is responsible for the lactase activity. This assignment is coincident with that proposed from the results of recent active-site mutagenesis studies [Zecca, L., Mesonero, J.E., Stutz, A., Poiree, J.C., Giudicelli, J., Cursio, R., Gloor, S.M. & Semenza, G. (1998) FEBS Lett. 435, 225-228] and opposite to that based on data from early affinity labeling with conduritol B epoxide [Wacker, W., Keller, P., Falchetto, R., Legler, G. & Semenza, G. (1992) J. Biol. Chem. 267, 18744-18752].  相似文献   

16.
Optical, resonance Raman, and electron paramagnetic resonance spectroscopies have been used to characterize the ligands and spin state of the chloroplast cytochrome b-559. The protein was isolated from both maize and spinach in a low-potential form. The spectroscopic data indicate that the heme iron in both ferric and ferrous cytochrome b-559 is in its low-spin state and ligated in its fifth and sixth coordination positions by histidine nitrogens. Electron paramagnetic resonance data for the purified spinach cytochrome are in good agreement with those determined by Bergstr?m and V?nng?rd [Bergstr?m, J., & V?nng?rd, T. (1982) Biochim. Biophys. Acta 682, 452-456] for a low-potential membrane-bound form of cytochrome b-559. The g values of high-potential cytochrome b-559 are shifted from those of its low-potential forms; this shift is interpreted as arising from a deviation of the planes of the two axial histidine imidazole rings from a parallel orientation. The model is consistent with the physical data and may also account for the facility with which cytochrome b-559 can be converted between low- and high-potential forms. Recent biochemical and molecular biological data [Widger, W. R., Cramer, W. A., Hermodson, M., Meyer, D., & Gullifor, M. (1984) J. Biol. Chem. 259, 3870-3876; Herrmann, R. G., Alt, J., Schiller, D., Cramer, W. A., & Widger, W. R. (1984) FEBS Lett. 179, 239-244] have shown that two polypeptides, one with 83 residues and a second with 39 residues, most likely constitute the protein of the cytochrome.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
The extent of glycans heterogeneity in a pathological human immunoglobulin M ZAJ has been studied on oligosaccharides released by hydrazinolysis from the purified glycoprotein. After reduction with NaB3H4, asparagine-linked carbohydrate chains were separated by affinity chromatography on concanavalin A-Sepharose into oligomannosidic and N-acetyllactosaminic types. Glycans of the oligomannosidic type were further fractionated by HPLC and those of the N-acetyllactosamine type by preparative high-voltage electrophoresis. The primary structure of the main oligosaccharides was investigated on the basis of micro-methylation analysis, mass spectrometry and sequential exo-glycosidase digestion. Glycans of the oligomannosidic type varied in size from Man5GlcNAc2 to Man9GlcNAc2. N-Acetyllactosaminic glycans were found of the biantennary, bisected-biantennary and triantennary types. They presented a higher degree of heterogeneity due to the presence of a variable number of NeuAc and fucose residues. The new structures we report here were in addition to the major biantennary one we previously described on the basis of methylation analysis and 500 MHz 1H-NMR spectroscopy (Cahour, A., Debeire, P., Hartmann, L., Montreuil, J., Van Halbeek, H. and Vliegenthart, J.F.G. (1984) FEBS Lett. 170, 343-349): NeuAc(alpha 2-6)Gal(beta 1-4)GlcNAc(beta 1-2)Man(alpha 1-3)[Gal(beta 1-4)Glc-NAc(beta 1-2)Man(alpha 1-6)]Man(beta 1-4)]Glc-NAc(beta 1-4) [Fuc(alpha 1-6)]GlcNAc.  相似文献   

18.
Residues 89-100 of troponin C (C89-100) and 96-116 of troponin I (I96-116) interact with each other in the troponin complex (Dalgarno, D.C., Grand, R.J.A., Levine, B.A. Moir, A., J.G., Scott, G.M.M., and Perry, S.V. (1982) FEBS Lett. 150, 54-58) and are necessary for the Ca2+ sensitivity of actomyosin ATPase (Syska, H., Wilkinson, J.M., Grand, R.J.A., and Perry, S.V. (1976) Biochem. J. 153, 375-387 and Grabarek, Z., Drabikowski, W., Leavis, P.C., Rosenfeld, S.S., and Gergely, J. (1981) J. Biol. Chem. 256, 13121-13127). We have studied Ca2+-induced changes in the region C89-100 by monitoring the fluorescence of troponin C (TnC) labeled at Cys-98 with 5-(iodoacetamidoethyl)aminonaphthalene-1-sulfonic acid. Equilibrium titration of the labeled TnC with Ca2+ indicates that the probe is sensitive to binding to both classes of sites in free TnC as well as in its complex with TnI. When Mg2 X TnC is mixed with Ca2+ in a stopped flow apparatus, there is a rapid fluorescence increase related to Ca2+ binding to the unoccupied sites I and II followed by a slower increase (k = 9.9 s-1) that represents Mg2+-Ca2+ exchange at sites III and IV. In the TnC X TnI complex, the fast phase is much larger and the Mg2+-Ca2+ exchange at sites III and IV results in a small decrease rather than an increase in the fluorescence of the probe. The possibility is discussed that the fast change in the environment of Cys-98 upon Ca2+ binding to sites I and II may be instrumental in triggering activation of the thin filament by facilitating a contact between C89-100 and I96-116.  相似文献   

19.
Factors relevant in bacterial pyrroloquinoline quinone production   总被引:2,自引:0,他引:2  
Quinoprotein content and levels of external pyrroloquinoline quinone (PQQ) were determined for several bacteria under a variety of growth conditions. From these data and those from the literature, a number of factors can be indicated which are relevant for PQQ production. Synthesis of PQQ is only started if synthesis of a quinoprotein occurs, but quinoprotein synthesis does not depend on PQQ synthesis. The presence of quinoprotein substrates is not necessary for quinoprotein and PQQ syntheses. Although the extent of PQQ production was determined by the type of organism and quinoprotein produced, coordination between quinoprotein and PQQ syntheses is loose, since underproduction and overproduction of PQQ with respect to quinoprotein were observed. The results can be interpreted to indicate that quinoprotein synthesis depends on the growth rate whereas PQQ synthesis does not. In that view, the highest PQQ production can be achieved under limiting growth conditions, as was shown indeed by the much higher levels of PQQ produced in fed-batch cultures compared with those produced in batch experiments. The presence of nucleophiles, especially amino acids, in culture media may cause losses of PQQ due to transformation into biologically inactive compounds. Some organisms continued to synthesize PQQ de novo when this cofactor was administered exogenously. Most probably PQQ cannot be taken up by either passive diffusion or active transport mechanisms and is therefore not able to exert feedback regulation on its biosynthesis in these organisms.  相似文献   

20.
An insulin-like growth factor II variant (IGF-II variant) was purified from Cohn fraction IV1 of human plasma by ion exchange, gel filtration, and reversed-phase high pressure liquid chromatography. The amino-terminal sequence of the first 35 amino acid residues showed a replacement of Ser-29 of IGF-II with the tetrapeptide Arg-Leu-Pro-Gly of IGF-II variant. Peptides isolated and sequenced after digestion with endoproteinase Asp-N and endoproteinase Glu-C disclosed no differences with the sequence predicted from an IGF-II variant cDNA clone isolated by Jansen, M., van Shaik, F. M. A., van Tol, H., Van den Brande, J. L., and Sussenbach, J. S. (1985) FEBS Lett., 179, 243-246. The molecular ion of intact IGF-II variant was 7809.4 mass units, as measured by plasma desorption mass spectrometry. This is in close agreement with the molecular ion of 7812.8 mass units calculated from the determined sequence and indicates the entire amino acid sequence had been accounted for. Binding of IGF-II variant to purified insulin-like growth factor I (IGF-I) receptors demonstrated a 2-3-fold lower affinity for this receptor compared with IGF-I or IGF-II. The dissociation constants for IGF-I, IGF-II, and IGF-II variant are 0.23, 0.38, and 0.80 nM, respectively. In a growth assay, the concentration of IGF-II and IGF-II variant required to stimulate the half-maximal growth of MCF-7 cells was 4 and 13 nM, respectively. Finally, the amount of IGF-II variant that can be purified by this method constitutes approximately 25% of the total IGF-II isolated from Cohn fraction IV1 of human plasma.  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