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.
Resonance Raman spectroscopy has been used to probe the structure of the organic cofactor in copper-containing amine oxidases from bovine plasma, porcine kidney, pea seedlings, and the bacterium Arthrobacter P1. The enzymes were first derivatized with phenylhydrazine or p-nitrophenylhydrazine; resonance Raman spectra were obtained on the intact derivatized enzymes and on a derivatized active-site peptide isolated from bovine plasma amine oxidase. Spectra of the intact amine oxidase phenylhydrazones are practically identical, consistent with the enzymes examined containing a similar cofactor. Only minor frequency shifts and some intensity variations are detected between the resonance Raman spectra of intact bovine plasma amine oxidase and the isolated peptide. These spectral perturbations are attributable to differences in the micro-environment between the intact, folded protein and the isolated small peptide in aqueous solution. This rules out the possibility that a new structure is formed during the isolation of the derivatized active-site peptide. Importantly, the resonance Raman spectra of the phenylhydrazine and p-nitrophenylhydrazine derivatives of the bovine plasma amine oxidase peptide are identical to the spectra of the corresponding derivatives of topa quinone (6-hydroxydopa quinone). Hence these data provide strong, independent support for the recent identification of topa as the organic functional group in bovine plasma amine oxidase (Janes, S. M., Mu, D., Wemmer, D., Smith, A. J., Kaur, S., Maltby, D., Burlingame, A. L., and Klinman, J.P. (1990) Science 248, 981-987).  相似文献   

2.
Resonance Raman spectroscopy has been used to probe the structure of the covalently bound quinone cofactor in methylamine dehydrogenase from the bacterium W3A1. Spectra were obtained on the phenylhydrazine and 2-pyridylhydrazine derivatives of the native enzyme, on the quinone-containing subunit labeled with phenylhydrazine, and on an active-site peptide also labeled with phenylhydrazine. Comparisons of these spectra to the corresponding spectra of copper-containing amine oxidase derivatives indicate that the quinones in these two classes of quinoproteins are not identical. The resonance Raman spectra of the native enzyme and small subunit have also been measured. 16O/18O exchange permitted the carbonyl modes of the quinone to be identified in the resonance Raman spectrum of oxidized methylamine dehydrogenase: a band at 1614 cm-1, together with a shoulder at 1630 cm-1, are assigned as modes containing substantial C = O stretching character. D2O/H2O exchange has pronounced effects on the resonance Raman spectrum of the oxidized enzyme, suggesting that the quinone may have numerous hydrogen bonds to the protein or that it is unusually sensitive to the local environment. Resonance Raman spectra of the isolated small subunit, and its phenylhydrazine derivative, are considerably different from the corresponding spectra of the intact protein. An attractive explanation for these observations is that the quinone cofactor in methylamine dehydrogenase from W3A1 is located at the interface between the large and small subunits, as found for the enzyme from Thiobacillus versutus [Vellieux, F. M. D., Huitema, F., Groendijk, H., Kalk, K. H., Frank, J. Jzn., Jongejan, J. A., & Duine, J. A. (1989) EMBO J. 8, 2171-2178].(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
Quino-cofactors have been found in a wide variety of prokaryotic and eukaryotic organisms. Two variants have, thus far, been demonstrated to derive from tyrosine precursors: these are the 2,4, 5-trihydroxyphenylalanine quinone (topa quinone or TPQ) [Janes, S. M. , et al. (1990) Science 248, 98] and an o-quinone analogue containing the side chain of a lysine residue (lysyltyrosine quinone or LTQ) [Wang, S. Z., et al. (1996) Science 273, 1078]. Additionally, a third variant of the family of tyrosine-derived cofactors has been reported to exist in an Aspergillus niger amine oxidase AO-I. This was described as an o-quinone cross-linked to the side chain of a glutamate residue [Frebort, I. (1996) Biochim. Biophys. Acta 1295, 59]. We have synthesized model compounds related to the proposed structure. Characterization of the redox properties for the model compound and spectral properties of its 4-nitrophenylhydrazine derivative lead us to conclude that the cofactor in A. niger amine oxidase AO-I has been misidentified. A TPQ carboxylate ester is considered an unlikely candidate for a biologically functional quino-cofactor.  相似文献   

4.
We found that the structural gene for monoamine oxidase was located at 30.9 min on the Escherichia coli chromosome. Deletion analysis showed that two amine oxidase genes are located in this region. The nucleotide sequence of one of the two genes was determined. The peptide sequence of the first 40 amino acids from the N terminus of monoamine oxidase purified from E. coli agrees with that deduced from the nucleotide sequence of the gene. The leader peptide extends over 30 amino acids. The nucleotide sequence of the gene and amino acid sequence of the predicted mature enzyme (M.W. 81,295) were highly homologous to those of the maoAK gene and monoamine oxidase from Klebsiella aerogenes, respectively. From these results and analysis of the enzyme activity, we concluded that the gene encodes for monoamine oxidase (maoAE). The tyrosyl residue, which may be converted to topa quinone in the E. coli enzyme, was located by comparison with amino acid sequences at the cofactor sites in other copper/topa quinone-containing amine oxidases.  相似文献   

5.
《Gene》1996,179(2):279-286
A 4040-bp cDNA was cloned from a human placenta library by screening with a polymerase chain reaction-amplified fragment. The fragment was generated from the library using primers corresponding to conserved sequences encompassing the topa quinone (TPQ) cofactor sites of the copper-containing proteins, bovine serum amine oxidase (BSAO) and human kidney diamine oxidase (DAO). The cloned cDNA contains a coding sequence from positions 161 to 2449. Between bases 2901 and 2974, in a very long 1591-bp 3′-untranslated region, there is a G/A-rich region in the minus strand, which contains a (AGG)5 tandem repeat. The human placenta cDNA sequence and its translated amino acid sequence are 84% and 81% identical to the corresponding BSAO sequences, while the identities for the placenta sequences and those for human kidney DAO are 60% and 41%, respectively. The TPQ consensus nucleotide and protein sequences are identical for the placenta enzyme and BSAO, but the corresponding sequences for human kidney DAO are nonidentical. Three His residues that have been identified as Cu(II) ligands in other amine oxidases are conserved in the human placenta amine oxidase protein sequence. It was concluded that the placenta cDNA open-reading frame codes for a copper-containing, TPQ-containing monoamine oxidase. A putative 19-amino acid signal peptide was identified for human placenta amine oxidase. The resulting mature protein would be composed of 744 amino acids, and would have a Mr of 82 525. Comparison of the human placenta amine oxidase with DNA sequences found in GenBank suggests that the gene for this enzyme is located in the q21 region of human chromosome 17, near the BRCA1 gene.  相似文献   

6.
For bovine serum amine oxidase, two different mechanisms of substrate-induced inactivation have been proposed. One consists of a slow oxidation by H2O2 of a conserved residue in the reduced enzyme after the fast turnover phase [Pietrangeli, P., Nocera, S., Fattibene, P., Wang, X.T., Mondovì, B. & Morpurgo, L. (2000) Biochem. Biophys. Res. Commun.267, 174-178] and the other of the oxidation by H2O2 of the dihydrobenzoxazole in equilibrium with the product Schiff base, during the catalytic cycle [Lee, Y., Shepard, E., Smith, J., Dooley, D.M. & Sayre, L.M. (2001) Biochemistry40, 822-829]. To discriminate between the two mechanisms, the inactivation was studied using Lathyrus cicera (red vetchling) amine oxidase. This, in contrast to bovine serum amine oxidase, formed the Cu+-semiquinolamine radical with a characteristic UV-vis spectrum when oxygen was exhausted by an excess of any tested amine in a closed cuvette. The inactivation, lasting about 90 min, was simultaneous with the radical decay and with the formation of a broad band (shoulder) at 350 nm. No inactivation occurred when a thousand-fold excess of amine was rapidly oxidized in an L. cicera amine oxidase solution stirred in open air. Thus, the inactivation is a slow reaction of the reduced enzyme with H2O2, following the turnover phase. Catalase protected L. cicera amine oxidase from inactivation. This effect was substrate-dependent, varying from full protection (benzylamine) to no protection (putrescine). In the absence of H2O2, a specific inactivating reaction, without formation of the 350 nm band, was induced by some aldehydes, notably putrescine. Some mechanisms of inactivation are proposed.  相似文献   

7.
The cDNA coding for copper amine oxidase has been cloned from etiolated pea seedlings (Pisum sativum). The deduced amino acid sequence, consisting of 674 residues including the signal peptide, agreed well with those reported for the enzymes from a different cultivar of P. sativum and other plant sources, except for several evolutionary replacements located mostly on the molecular surface. A heterologous expression system for the cloned pea enzyme was constructed with the yeast Pichia pastoris, using the AOX1 promoter and the yeast alpha-factor secretion signal. Adding copper to the culture medium increased the secretion of an active, quinone-containing enzyme. Furthermore, the inactive enzyme produced in a copper-deficient medium was activated considerably by subsequent incubation with excess cupric ions. These results strongly suggest that the Tyr-derived redox cofactor, 2,4,5-trihydroxyphenylalanylquinone (topa quinone, TPQ), is produced in the plant enzyme by post-translational modification that proceeds through the copper-dependent, self-processing mechanism, as in the enzymes from bacteria and yeast.  相似文献   

8.
The topa quinone (TPQ) cofactor of copper amine oxidase is generated by copper-assisted self-processing of the precursor protein. Metal ion specificity for TPQ biogenesis has been reinvestigated with the recombinant phenylethylamine oxidase from Arthrobacter globiformis. Besides Cu2+ ion, some divalent metal ions such as Co2+, Ni2+, and Zn2+ were also bound to the metal site of the apoenzyme so tightly that they were not replaced by excess Cu2+ ions added subsequently. Although these noncupric metal ions could not initiate TPQ formation under the atmospheric conditions, we observed slow spectral changes in the enzyme bound with Co2+ or Ni2+ ion under the dioxygen-saturating conditions. Resonance Raman spectroscopy and titration with phenylhydrazine provided unambiguous evidence for TPQ formation by Co2+ and Ni2+ ions. Steady-state kinetic analysis showed that the enzymes activated by Co2+ and Ni2+ ions were indistinguishable from the corresponding metal-substituted enzymes prepared from the native copper enzyme (Kishishita, S., Okajima, T., Kim, M., Yamaguchi, H., Hirota, S., Suzuki, S., Kuroda, S., Tanizawa, K., and Mure, M. (2003) J. Am. Chem. Soc. 125, 1041-1055). X-ray crystallographic analysis has also revealed structural identity of the active sites of Co- and Ni-activated enzymes with Cu-enzyme. Thus Cu2+ ion is not the sole metal ion assisting TPQ formation. Co2+ and Ni2+ ions are also capable of forming TPQ, though much less efficiently than Cu2+.  相似文献   

9.
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   

10.
The presence of pyrroloquinoline quinone (PQQ) as the organic cofactor of Dactylium dendroides galactose oxidase and lentil (Lens culinaris) seedling amine oxidase, purported PQQ-containing oxidoreductases (Van der Meer, R. A., Jongejan, J. A., and Duine, J. A. (1989) J. Biol. Chem. 264, 7792-7794; Citro, G., Verdina, A., Galati, R., Floris, G., Sabatini, S., and Finazzi-Argo', A. (1989) FEBS Lett. 247, 201-204), was reinvestigated using the nitro blue tetrazolium redoxcycling method (Paz, M. A., Gallop, P. M., Torrelio, B. M., and Flückiger, R. (1988) Biochem. Biophys. Res. Commun. 154, 1330-1337; Paz, M. A., Flückiger, R., Boak, A., Kagan, H. M., and Gallop, P. M. (1991) J. Biol. Chem. 266, 689-692) and the enzyme-linked immunosorbent assay with polyclonal antibodies against PQQ. The possible quinoprotein nature of the laccases from Polyporus versicolor and Rhus vernicifera was also investigated because of the similarities in spectroscopic and kinetic features of these enzymes and the laccase from Phlebia radiata, reported to be a PQQ protein (Karhunen, E., Niku-Paavola, M.-L., Viikari, L., Haltia, T., Van der Meer, R. A., and Duine, J. A. (1990) FEBS Lett. 267, 6-8). The presence of a quinonoid cofactor in lentil seedling amine oxidase is confirmed, whereas galactose oxidase and both laccases do not display any quinoprotein nature.  相似文献   

11.
Crude extract of Aspergillus niger AKU 3302 mycelia incubated with methylamine showed a single amine oxidase activity band in a developed polyacrylamide gel that weakly cross-reacted with the antibody against a copper/topa quinone-containing amine oxidase (AO-II) from the same strain induced by n-butylamine. Since the organism cannot grow on methylamine and the already known quinoprotein amine oxidases of the organism cannot catalyze oxidation of methylamine, the organism was forced to produce another enzyme that could oxidize methylamine when the mycelia were incubated with methylamine. The enzyme was separated and purified from the already known two quinoprotein amine oxidases formed in the same mycelia. The purified enzyme showed a sharp symmetric sedimentation peak in analytical ultracentrifugation showing S20,w0 of 6.5s. The molecular mass of 133 kDa estimated by gel chromatography and 66.6 kDa found by SDS-PAGE confirmed the dimeric structure of the enzyme. The purified enzyme was pink in color with an absorption maximum at 494 nm. The enzyme readily oxidized methylamine, n-hexylamine, and n-butylamine, but not benzylamine, histamine, or tyramine, favorite substrates for the already known two quinoprotein amine oxidases. Inactivation by carbonyl reagents and copper chelators suggested the presence of a copper/topa quinone cofactor. Spectrophotometric titration by p-nitrophenylhydrazine showed one reactive carbonyl group per subunit and redox-cyclic quinone staining confirmed the presence of a quinone cofactor. pH-dependent shift of the absorption spectrum of the enzyme-p-nitrophenylhydrazone (469 nm at neutral to 577 nm at alkaline pH) supported the identity of the cofactor with topaquinone. Nothern blot analysis indicated that the methylamine oxidase encoding gene is largely different from the already known amine oxidase in the organism.  相似文献   

12.
Ruggiero CE  Dooley DM 《Biochemistry》1999,38(10):2892-2898
The stoichiometry of the topa quinone biogenesis reaction in phenylethylamine oxidase from Arthrobacter globiformis (AGAO) has been determined. We have shown that the 6e- oxidation of tyrosine to topa quinone (TPQ) consumes 2 mol of O2 and produces 1 mol of H2O2/mol of TPQ formed. The rate of H2O2 production is first-order (kobs = 1.0 +/- 0.2 min-1), a rate only slightly lower than the rate of TPQ formation directly determined previously (kobs = 1.5 +/- 0.2 min-1). This gives the following net reaction stoichiometry for TPQ biogenesis: E-Tyr + 2O2 --> E-TPQ + H2O2. This stoichiometry is in agreement with recently proposed mechanisms for TPQ biogenesis, and rules out several possible alternatives.  相似文献   

13.
A peptide derived from p34cdc2, cdc2(6-20)NH2 with the amino acid sequence of KVEKIGEGTYGVVYK-amide, was found to be a specific and efficient substrate for a pp60c-src-related protein tyrosine kinase from bovine spleen (STK). Glu-12 and Thr-14 were identified to be substrate specificity determinants in this peptide (Cheng, H.-C., Litwin, C. M. E., Hwang, D. M., and Wang, J. H. (1991) J. Biol. Chem. 266, 17919-17925). In this study, we demonstrated the presence of cdc2(6-20)NH2 peptide tyrosine kinase activity in the membrane fractions of bovine brain, spleen, thymus, lung, liver, and kidney. Hydroxylapatite column chromatography of thymus membrane extract revealed four protein tyrosine kinases, TK-I, TK-II, TK-III, and TK-IV, with different relative activities toward cdc2(6-20)NH2 and a general tyrosine kinase substrate, poly(Glu/Tyr). Only TK-I and TK-II showed significant activity toward cdc2(6-20)NH2, they were suggested as belonging to the src-family by virtue of their cross-reactivity with an antibody against a synthetic peptide corresponding to a conserved sequence of src-family kinases. Further immunological characterization using antibodies specific to individual src-related protein tyrosine kinases suggested that TK-I, TK-II, and STK are bovine homologs of p56lck, p55fyn, and p56lyn, respectively. Substrate specificity and kinetic characterization of src-family tyrosine kinases including human platelet pp60c-src, bovine p56lyn, p56lck, and p55fyn, as well as several non-src-related tyrosine kinases including epidermal growth factor receptor, p43v-abl, TK-III, and TK-IV showed that all the src-family tyrosine kinases but none of the other kinases displayed efficient cdc2(6-20)NH2 phosphorylation. In all cases, the high efficiency of cdc2(6-20)NH2 peptide phosphorylation could be markedly attenuated when Glu-12 and Thr-14 of the peptide were substituted, respectively, by valine and serine.  相似文献   

14.
J Rucker  Y Cha  T Jonsson  K L Grant  J P Klinman 《Biochemistry》1992,31(46):11489-11499
Previous investigations have indicated a role for hydrogen tunneling in the yeast alcohol dehydrogenase catalyzed oxidation of benzyl alcohol [Cha, Y., Murray, C. J., & Klinman, J. P. (1989) Science 243, 1325] and the bovine plasma amine oxidase catalyzed oxidation of benzylamine [Grant, K.L., & Klinman, J. P. (1989) Biochemistry 28,6597]. In the present studies, values of protium to tritium and deuterium to tritium isotope effects and their temperature dependencies have been measured using ring-substituted substrates for yeast alcohol dehydrogenase and bovine plasma amine oxidase, revealing tunneling in each case. The results of these studies indicate that hydrogen tunneling is a general phenomenon and is not limited to enzyme reactions with degenerate energy levels for bound substrates and products. An analysis of internal thermodynamics in the yeast alcohol dehydrogenase reaction shows that tunneling occurs when delta H degrees is endothermic and that the degree of tunneling appears to increase as delta H degrees decreases toward zero.  相似文献   

15.
From the amino acid sequence of yeast cytochrome c oxidase subunit VIII published previously (Power, S. D., Lochrie, M.A., Patterson, T.E., and Poyton, R.C. (1984) J. Biol. Chem. 259, 6571-6574), we have synthesized a pair of oligonucleotide probes and used them to identify COX8, its structural gene. By genomic Southern blot analysis and disruption of the COX8 chromosomal locus, we have shown that this gene is present in one copy per haploid genome and that its product, subunit VIII, is essential for maximal levels of cellular respiration and cytochrome c oxidase activity. Alignment of the amino acid sequence predicted from the DNA sequence of COX8 with the determined amino acid sequence of subunit VIII indicates that mature subunit VIII is derived from a larger precursor that extends from both the NH2 and COOH termini of the mature polypeptide. Thus, like many other nuclear coded mitochondrial proteins, subunit VIII is derived from a precursor which carries a leader peptide. In addition, this precursor, like that for yeast cytochrome c oxidase subunit VIIa, appears to carry a four-amino acid "trailer peptide" at its COOH terminus.  相似文献   

16.
We have cloned the AMO gene, encoding the microbody matrix enzyme amine oxidase (EC 1.4.3.6) from the yeast Hansenula polymorpha. The gene was isolated by differential screening of a cDNA library, immunoselection, and subsequent screening of a H. polymorpha genomic library. The nucleotide sequence of a 3.6 kilobase stretch of DNA containing the amine oxidase (AMO) gene was determined. The AMO gene contains an open reading frame of 692 amino acids, with a relative molecular mass of 77,435. The 5' and 3' ends of the gene were mapped and show that the transcribed region measures 2134 nucleotides. The derived amino-acid sequence was confirmed by sequencing an internal proteolytic fragment of the purified protein. Amine oxidase contains the tripeptide sequence Ser-Arg-Leu, located 9 residues from the carboxy terminus, which may represent the topogenic signal for protein import into microbodies.  相似文献   

17.
The catalytic function of bovine lysyl oxidase in the absence of copper   总被引:6,自引:0,他引:6  
Bovine lysyl oxidase (BLO) contains two different cofactors, copper (Kagan, H. M. (1986) in Biology of Extracellular Matrix (Mecham, R. P., ed) Vol. 1, pp. 321-398, Academic Press, Orlando, FL) and lysine tyrosyl quinone (LTQ) (Wang, S. X., Mure, M., Medzihradszky, K. F., Burlingame, A. L., Brown, D. E., Dooley, D. M., Smith, A. J., Kagan, H. M., and Klinman, J. P. (1996) Science 273, 1078-1084). By a combination of UV-visible spectroscopy, metal content analysis, and activity measurements, we find that copper-depleted BLO reacts in an irreversible manner with phenylhydrazine, an amine substrate analog, and catalyzes multiple turnovers of the substrate benzylamine. After removal of the majority of enzyme-bound copper, apoBLO exhibits a decrease in the LTQ content, as evidenced by the drop of the 510-520-nm absorbance, suggesting that the copper may play a structural role in stabilizing the LTQ. The remaining intact LTQ in the apoBLO reacted with phenylhydrazine, both in the presence and absence of the chelator, 10 mm 2,2'-dipyridyl. When benzylamine was used as the substrate, the apoBLO turned over at a rate of 50-60% of the native BLO (after correction for the residual copper and the change of LTQ content). Copper contamination from the assay buffer was ruled out by comparison of enzyme activity using different apoBLO concentrations. These studies demonstrate that the mature form of lysyl oxidase retains many of its functions in the absence of copper.  相似文献   

18.
The topa quinone (TPQ) cofactor of copper amine oxidase is produced by posttranslational modification of a specific tyrosine residue through the copper-dependent, self-catalytic process. We have site-specifically mutated three histidine residues (His431, His433, and His592) involved in binding of the copper ion in the recombinant phenylethylamine oxidase from Arthrobacter globiformis. The mutant enzymes, in which each histidine was replaced by alanine, were purified in the Cu/TPQ-free precursor form and analyzed for their Cu-binding and TPQ-generating activities by UV-visible absorption, resonance Raman, and electron paramagnetic resonance spectroscopies. Among the three histidine-to-alanine mutants, only H592A was found to show a weak activity to form TPQ upon aerobic incubation with Cu(2+) ions. Also for H592A, exogenous imidazole rescued binding of copper and markedly promoted the TPQ formation. Accommodation of a free imidazole molecule within the cavity created in the active site of H592A was suggested by X-ray crystallography. Although the TPQ cofactor in H592A mutant was readily reduced with substrate, its catalytic activity was very low even in the presence of imidazole. Combined with the crystal structures of the mutant enzymes, these results demonstrate the importance of the three copper-binding histidine residues for both TPQ biogenesis and catalytic activity, fine-tuning the position of the essential metal.  相似文献   

19.
6-Hydroxybenzofuran and phenylhydrazine are mechanism-based inhibitors of dopamine beta-hydroxylase (D beta H; EC 1.14.17.1). We report here the isolation and characterization of radiolabeled peptides obtained after inactivation of D beta H with [3H]6-hydroxybenzofuran and [14C]phenylhydrazine followed by digestion with Staphylococcus aureus V8 protease. Inactivation of D beta H with [3H]6-hydroxybenzofuran gave only one labeled peptide, whereas inactivation with [14C]phenylhydrazine gave several labeled peptides. Each inhibitor labeled a unique tyrosine in the enzyme corresponding to Tyr477 in the primary sequence of the bovine enzyme (Robertson, J. G., Desai, P. R., Kumar, A., Farrington, G. K., Fitzpatrick, P. F., and Villafranca, J. J. (1990) J. Biol. Chem. 265, 1029-1035). In addition, [14C]phenylhydrazine also labeled a unique histidine (His249) as well as several other peptides. Examination of the complete peptide profile obtained by high pressure liquid chromatography analysis also revealed the presence of a modified but nonradioactive peptide. This peptide was isolated and sequenced and was identical whether the enzyme was inactivated by 6-hydroxybenzofuran or phenylhydrazine. An arginine at position 503 was missing from the sequence cycle performed by Edman degradation of the modified peptide, but arginine was present in the identical peptide isolated from native dopamine beta-hydroxylase. These data are analyzed based on an inactivation mechanism involving formation of enzyme bound radicals (Fitzpatrick, P. F., and Villafranca, J. J. (1986) J. Biol. Chem. 261, 4510-4518) interacting with active site amino acids that may have a role in substrate binding and binding of the copper ions at the active site.  相似文献   

20.
Copper, a mediator of redox chemistries in biology, is often found in enzymes that bind and reduce dioxygen. Among these, the copper amine oxidases catalyze the oxidative deamination of primary amines utilizing a type(II) copper center and 2,4,5-trihydroxyphenylalanine quinone (TPQ), a covalent cofactor derived from the post-translational modification of an active site tyrosine. Previous studies established the dependence of TPQ biogenesis on Cu(II); however, the dependence of cofactor formation on the biologically relevant Cu(I) ion has remained untested. In this study, we demonstrate that the apoform of the Hansenula polymorpha amine oxidase readily binds Cu(I) under anaerobic conditions and produces the quinone cofactor at a rate of 0.28 h(-1) upon subsequent aeration to yield a mature enzyme with kinetic properties identical to the protein product of the Cu(II)-dependent reaction. Because of the change in magnetic properties associated with the oxidation of copper, electron paramagnetic resonance spectroscopy was employed to investigate the nature of the rate-limiting step of Cu(I)-dependent cofactor biogenesis. Upon aeration of the unprocessed enzyme prebound with Cu(I), an axial Cu(II) electron paramagnetic resonance signal was found to appear at a rate equivalent to that for the cofactor. These data provide strong evidence for a rate-limiting release of superoxide from a Cu(II)(O(2)(.)) complex as a prerequisite for the activation of the precursor tyrosine and its transformation for TPQ. As copper is trafficked to intracellular protein targets in the reduced, Cu(I) state, these studies offer possible clues as to the physiological significance of the acquisition of Cu(I) by nascent H. polymorpha amine oxidase.  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