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 31 毫秒
1.
The structure of the respiratory nitrate reductase (NapAB) from Rhodobacter sphaeroides, the periplasmic heterodimeric enzyme responsible for the first step in the denitrification process, has been determined at a resolution of 3.2 A. The di-heme electron transfer small subunit NapB binds to the large subunit with heme II in close proximity to the [4Fe-4S] cluster of NapA. A total of 57 residues at the N- and C-terminal extremities of NapB adopt an extended conformation, embracing the NapA subunit and largely contributing to the total area of 5,900 A(2) buried in the complex. Complex formation was studied further by measuring the variation of the redox potentials of all the cofactors upon binding. The marked effects observed are interpreted in light of the three-dimensional structure and depict a plasticity that contributes to an efficient electron transfer in the complex from the heme I of NapB to the molybdenum catalytic site of NapA.  相似文献   

2.
Seven genes, napKEFDABC, encoding the periplasmic nitrate reductase system were cloned from the denitrifying phototrophic bacterium Rhodobacter sphaeroides f. sp. denitrificans IL106. Two transmembrane proteins, NapK and NapE, an iron-sulfur protein NapF, a soluble protein NapD, a catalytic subunit of nitrate reductase precursor NapA, a soluble c-type diheme cytochrome precursor NapB, and a membrane-anchored c-type tetraheme cytochrome NapC were deduced as the gene products. Every mutant in which each nap gene was disrupted by omega-cassette insertion lost nitrate reductase activity as well as the ability of cells to grow with nitrate under anaerobic-dark conditions. A transconjugant of the napD-disrupted mutant with a plasmid bearing the napKEFDABC genes recovered both nitrate reductase activity and nitrate-dependent anaerobic-dark growth of cells. Denitrification activity, which was not observed in the napD mutant, was also restored by the conjugation. These results indicate that the periplasmic nitrate reductase encoded by the napKEFDABC genes is the enzyme responsible for denitrification in this phototroph, although the presence of a membrane-bound nitrate reductase has been reported in the same strain.  相似文献   

3.
The periplasmic nitrate reductase, NapA, from Escherichia coli was identified as a 90 kDa molybdoprotein which comigrated during polyacrylamide gel electrophoresis with the di-haem c-type cytochrome, NapB. The DNA sequence of the 5' end of the napA gene and the N-terminal amino acid sequences of both NapA and NapB were determined. The 36 residue leader peptide for NapA includes the double-arginine motif typical of proteins to which complex redox cofactors are attached in the cytoplasm prior to targeting to the periplasm. The pre-NapA leader sequence is both unexpectedly long and, unless two successive proteolysis steps are involved, is cleaved at the unprecedented sequence G-Q-Q-. Nap activity was suppressed during growth in the presence of tungstate and was absent from a mutant unable to synthesise the molybdopterin cofactor.  相似文献   

4.
The periplasmic nitrate reductase (NapAB) from Cupriavidus necator is a heterodimeric protein that belongs to the dimethyl sulfoxide reductase family of mononuclear Mo-containing enzymes and catalyzes the reduction of nitrate to nitrite. The protein comprises a large catalytic subunit (NapA, 91 kDa) containing the molybdenum active site plus one [4Fe-4S] cluster, as well as a small subunit (NapB, 17 kDa), which is a diheme c-type cytochrome involved in electron transfer. Crystals of the oxidized form of the enzyme diffracted beyond 1.5 Å at the European Synchrotron Radiation Facility. This is the highest resolution reported to date for a nitrate reductase, providing true atomic details of the protein active center, and this showed further evidence on the molybdenum coordination sphere, corroborating previous data on the related Desulfovibrio desulfuricans NapA. The molybdenum atom is bound to a total of six sulfur atoms, with no oxygen ligands or water molecules in the vicinity. In the present work, we were also able to prepare partially reduced crystals that revealed two alternate conformations of the Mo-coordinating cysteine. This crystal form was obtained by soaking dithionite into crystals grown in the presence of the ionic liquid [C4mim]Cl. In addition, UV-Vis and EPR spectroscopy studies showed that the periplasmic nitrate reductase from C. necator might work at unexpectedly high redox potentials when compared to all periplasmic nitrate reductases studied to date.  相似文献   

5.
The Escherichia coli NapA (periplasmic nitrate reductase) contains a [4Fe-4S] cluster and a Mo-bis-molybdopterin guanine dinucleotide cofactor. The NapA holoenzyme associates with a di-heme c-type cytochrome redox partner (NapB). These proteins have been purified and studied by spectropotentiometry, and the structure of NapA has been determined. In contrast to the well characterized heterodimeric NapAB systems ofalpha-proteobacteria, such as Rhodobacter sphaeroides and Paracoccus pantotrophus, the gamma-proteobacterial E. coli NapA and NapB proteins purify independently and not as a tight heterodimeric complex. This relatively weak interaction is reflected in dissociation constants of 15 and 32 mum determined for oxidized and reduced NapAB complexes, respectively. The surface electrostatic potential of E. coli NapA in the apparent NapB binding region is markedly less polar and anionic than that of the alpha-proteobacterial NapA, which may underlie the weaker binding of NapB. The molybdenum ion coordination sphere of E. coli NapA includes two molybdopterin guanine dinucleotide dithiolenes, a protein-derived cysteinyl ligand and an oxygen atom. The Mo-O bond length is 2.6 A, which is indicative of a water ligand. The potential range over which the Mo(6+) state is reduced to the Mo(5+) state in either NapA (between +100 and -100 mV) or the NapAB complex (-150 to -350 mV) is much lower than that reported for R. sphaeroides NapA (midpoint potential Mo(6+/5+) > +350 mV), and the form of the Mo(5+) EPR signal is quite distinct. In E. coli NapA or NapAB, the Mo(5+) state could not be further reduced to Mo(4+). We then propose a catalytic cycle for E. coli NapA in which nitrate binds to the Mo(5+) ion and where a stable des-oxo Mo(6+) species may participate.  相似文献   

6.
Many sulphate reducing bacteria can also reduce nitrite, but relatively few isolates are known to reduce nitrate. Although nitrate reductase genes are absent from Desulfovibrio vulgaris strain Hildenborough, for which the complete genome sequence has been reported, a single subunit periplasmic nitrate reductase, NapA, was purified from Desulfovibrio desulfuricans strain 27774, and the structural gene was cloned and sequenced. Chromosome walking methods have now been used to determine the complete sequence of the nap gene cluster from this organism. The data confirm the absence of a napB homologue, but reveal a novel six-gene organisation, napC-napM-napA-napD-napG-napH. The NapC polypeptide is more similar to the NrfH subgroup of tetraheme cytochromes than to NapC from other bacteria. NapM is predicted to be a tetra-heme c-type cytochrome with similarity to the small tetraheme cytochromes from Shewanella oneidensis. The operon is located close to a gene encoding a lysyl-tRNA synthetase that is also found in D. vulgaris. We suggest that electrons might be transferred to NapA either from menaquinol via NapC, or from other electron donors such as formate or hydrogen via the small tetraheme cytochrome, NapM. We also suggest that, despite the absence of a twin-arginine targeting sequence, NapG might be located in the periplasm where it would provide an alternative direct electron donor to NapA.  相似文献   

7.
The complete amino acid sequence of a 26-kDa low redox potential cytochrome c-551 from Rhodocyclus tenuis was determined by a combination of Edman degradation and mass spectrometry. There are 240 residues including two heme binding sites at positions 41, 44, 128, and 132. There is no evidence for gene doubling. The only known homolog of Rc. tenuis cytochrome c-551 is the diheme cytochrome c-552 from Pseudomonas stutzeri which contains 268 residues and heme binding sites at nearly identical positions. There is 44% overall identity between the Rc. tenuis and Ps. stutzeri cytochromes with 10 internal insertions and deletions. The Ps. stutzeri cytochrome is part of a denitrification gene cluster, whereas Rc. tenuis is incapable of denitrification, suggesting different functional roles for the cytochromes. Histidines at positions 45 and 133 are the fifth heme ligands and conserved histidines at positions 29, 209, and 218 and conserved methionines at positions 114 and 139 are potential sixth heme ligands. There is no obvious homology to the low-potential diheme cytochromes characterized from other purple bacterial species such as Rhodobacter sphaeroides. There are therefore at least two classes of low-potential diheme cytochromes c found in phototrophic bacteria. There is no more than 11% helical secondary structure in Rc. tenuis cytochrome c-551 suggesting that there is no relationship to class I or class II c-type cytochromes.  相似文献   

8.
Anaerobic degradation of hydrocarbons was discovered a decade ago, and ethylbenzene dehydrogenase was one of the first characterized enzymes involved. The structure of the soluble periplasmic 165 kDa enzyme was established at 1.88 A resolution. It is a heterotrimer. The alpha subunit contains the catalytic center with a molybdenum held by two molybdopterin-guanine dinucleotides, one with an open pyran ring, and an iron-sulfur cluster with a histidine ligand. During catalysis, electrons produced by substrate oxidation are transferred to a heme in the gamma subunit and then presumably to a separate cytochrome involved in nitrate respiration. The beta subunit contains four iron-sulfur clusters and is structurally related to ferredoxins. The gamma subunit is the first known protein with a methionine and a lysine as axial heme ligands. The catalytic product was modeled into the active center, showing the reaction geometry. A mechanism consistent with activity and inhibition data of ethylbenzene-related compounds is proposed.  相似文献   

9.
The periplasmic nitrate reductase (NapAB), a member of the DMSO reductase superfamily, catalyzes the first step of the denitrification process in bacteria. In this heterodimer, a di-heme NapB subunit is associated to the catalytic NapA subunit that binds a [4Fe-4S] cluster and a bis(molybdopterin guanine dinucleotide) cofactor. Here, we report the kinetic characterization of purified mutated heterodimers from Rhodobacter sphaeroides. By combining site-directed mutagenesis, redox potentiometry, EPR spectroscopy, and enzymatic characterization, we investigate the catalytic role of two conserved residues (M153 and R392) located in the vicinity of the molybdenum active site. We demonstrate that M153 and R392 are involved in nitrate binding: the Vm measured on the M153A and R392A mutants are similar to that measured on the wild-type enzyme, whereas the Km for nitrate is increased 10-fold and 200-fold, respectively. The use of an alternative enzymatic assay led us to discover that NapAB is uncompetitively inhibited by Zn2+ ions (Ki' = 1 microM). We used this property to further probe the active site access in the mutant enzymes. It is proposed that R392 acts as a filter by preventing a direct reduction of the Mo atom by small reducing molecules and partially protecting the active site against zinc inhibition. In addition, we show that M153 is a key residue mediating this inhibition likely by coordinating Zn2+ ions via its sulfur atom. This residue is not conserved in the DMSO reductase superfamily while it is conserved in the periplasmic nitrate reductase family. Zinc inhibition is therefore likely to be specific and restricted to periplasmic nitrate reductases.  相似文献   

10.
Protease activity present in aerobically grown cells of Pseudomonas perfectomarina, protease apparently copurified with cytochrome c-552, and trypsin achieved a limited proteolysis of the diheme cytochrome c-552. That partial lysis conferred cytochrome c peroxidase activity upon cytochrome c-552. The removal of a 4000-Da peptide explains the structural changes in the cytochrome c-552 molecule that resulted in the appearance of both cytochrome c peroxidase activity (with optimum activity at pH 8.6) and a high-spin heme iron. The oxidized form of the modified cytochrome c-552 bound cyanide to the high-spin ferric heme with a rate constant of (2.1 +/- 0.1) X 10(3) M-1 s-1. The dissociation constant was 11.2 microM. Whereas the intact cytochrome c-552 molecule can be half-reduced by ascorbate, the cytochrome c peroxidase was not reducible by ascorbate, NADH, ferrocyanide, or reduced azurin. Dithionite reduced the intact protein completely but only half-reduced the modified form. The apparent second-order rate constant for dithionite reduction was (7.1 +/- 0.1) X 10(2) M-1 s-1 for the intact protein and (2.2 +/- 0.1) X 10(3) M-1 s-1 for the modified form. In contrast with other diheme cytochrome c peroxidases, reduction of the low-spin heme was not necessary to permit ligand binding by the high-spin heme iron.  相似文献   

11.
The periplasmic nitrate reductase (NAP) from Paracoccus pantotrophus is a soluble two-subunit enzyme (NapAB) that binds two c-type haems, a [4Fe-4S] cluster and a bis-molybdopterin guanine dinucleotide cofactor that catalyses the reduction of nitrate to nitrite. In the present work the NapAB complex has been studied by magneto-optical spectroscopy to probe co-ordination of both the NapB haems and the NapA active site Mo. The absorption spectrum of the NapAB complex is dominated by features from the NapB c-type cytochromes. Using a combination of electron paramagnetic resonance spectroscopy and magnetic circular dichroism it was demonstrated that both haems are low-spin with bis-histidine axial ligation. In addition, a window between 600 and 800 nm was identified in which weak absorption features that may arise from Mo could be detected. The low-temperature MCD spectrum shows oppositely signed bands in this region (peak 648 nm, trough 714 nm) which have been assigned to S-to-Mo(V) charge transfer transitions.  相似文献   

12.
We purified the nitrate reductase from the soluble fraction of Magnetospirillum magnetotacticum MS-1. The enzyme was composed of 86- and 17-kDa subunits and contained molybdenum, non-heme iron, and heme c. These properties are very similar to those of the periplasmic nitrate reductase found in Paracoccus pantotrophus. The M. magnetotacticum nap locus was clustered in seven open reading frames, napFDAGHBC. The phylogenetic analyses of NapA, NapB, and NapC suggested a close relationship between M. magnetotacticum nap genes and Escherichia coli nap genes, which is not consistent with the 16S rDNA data. This is the first finding that the alpha subclass of Proteobacteria possesses a napFDAGHBC-type nap gene cluster. The nap gene cluster had putative fumarate and nitrate reduction regulatory protein (Fnr) and NarL protein binding sites. Furthermore, we investigated the effect of molybdate deficiency in medium on the total iron content of the magnetosome fraction and discussed the physiological function of nitrate reductase in relation to the magnetite synthesis in M. magnetotacticum.  相似文献   

13.
The aerobic respiratory chain of the Gram-positive Corynebacterium glutamicum involves a bc(1) complex with a diheme cytochrome c(1) and a cytochrome aa(3) oxidase but no additional c-type cytochromes. Here we show that the two enzymes form a supercomplex, because affinity chromatography of either strep-tagged cytochrome b (QcrB) or strep-tagged subunit I (CtaD) of cytochrome aa(3) always resulted in the copurification of the subunits of the bc(1) complex (QcrA, QcrB, QcrC) and the aa(3) complex (CtaD, CtaC, CtaE). The isolated bc(1)-aa(3) supercomplexes had quinol oxidase activity, indicating functional electron transfer between cytochrome c(1) and the Cu(A) center of cytochrome aa(3). Besides the known bc(1) and aa(3) subunits, few additional proteins were copurified, one of which (CtaF) was identified as a fourth subunit of cytochrome aa(3). If either of the two CXXCH motifs for covalent heme attachment in cytochrome c(1) was changed to SXXSH, the resulting mutants showed severe growth defects, had no detectable c-type cytochrome, and their cytochrome b level was strongly reduced. This indicates that the attachment of both heme groups to apo-cytochrome c(1) is not only required for the activity but also for the assembly and/or stability of the bc(1) complex.  相似文献   

14.
The complete sequence of the 21-kDa cytochrome subunit of the flavocytochrome c (FC) from the purple phototrophic bacterium Chromatium vinosum has been determined to be as follows: EPTAEMLTNNCAGCHG THGNSVGPASPSIAQMDPMVFVEVMEGFKSGEIAS TIMGRIAKGYSTADFEKMAGYFKQQTYQPAKQSF DTALADTGAKLHDKYCEKCHVEGGKPLADEEDY HILAGQWTPYLQYAMSDFREERRPMEKKMASKL RELLKAEGDAGLDALFAFYASQQ. The sequence is the first example of a diheme cytochrome in a flavocytochrome complex. Although the locations of the heme binding sites and the heme ligands suggest that the cytochrome subunit is the result of gene doubling of a type I cytochrome c, as found with Azotobacter cytochrome c4, the extremely low similarity of only 7% between the two halves of the Chromatium FC heme subunit rather suggests that gene fusion is at the evolutionary origin of this cytochrome. The two halves also require a single residue internal deletion for alignment. The first half of the Chromatium FC heme subunit is 39% similar to the monoheme subunit of the FC from the green phototrophic bacterium Chlorobium thiosulfatophilum, but the second half is only 9% similar to the Chlorobium subunit. The N-terminal sequence of the Chromatium FC flavin subunit was determined up to residue 41 as AGRKVVVVGGGTGGATAAKYIKLADPSIEVTLIEP NTKYYT. It shows more similarity to the Chlorobium FC flavin subunit (60%) than do the two heme subunits. The N terminus of the flavin subunit is homologous to a number of flavoproteins, including succinate dehydrogenase, glutathione reductase, and monamine oxidase. There is no obvious homology to the Pseudomonas putida FC flavin subunit, which suggests that the two types of flavocytochrome c arose by convergent evolution. This is consistent with the dissimilar enzyme activities of FC as sulfide dehydrogenase in the phototrophic bacteria and as p-cresol methylhydroxylase in Pseudomonas. We also present a sequence "fingerprint" pattern for the recognition of FAD-binding proteins which is an extended version of the consensus sequence previously presented (Wierenga, R. K., Terpstra, P., and Hol, W. G. J. (1986) J. Mol. Biol. 187, 101-107) for nucleotide binding sites.  相似文献   

15.
The periplasmic nitrate reductase (Nap) is wide-spread in proteobacteria. NapA, the nitrate reductase catalytic subunit, contains a Mo-bisMGD cofactor and one [4Fe-4S] cluster. The nap gene clusters in many bacteria, including Rhodobacter sphaeroides DSM158, contain an napF gene, disruption of which drastically decreases both in vitro and in vivo nitrate reductase activities. In spite its importance in the Nap system, NapF has never been characterized biochemically, and its role remains unknown. The NapF protein has four polycysteine clusters that suggest that it is an iron-sulfur-containing protein. In the present study, a His(6)-tagged NapF protein was overproduced in Escherichia coli and purified anaerobically. The purified NapF protein was used to obtain polyclonal antibodies raised in rabbit, and cellular fractionation of R. sphaeroides followed by immunoprobing with anti-NapF antibodies revealed that the native NapF protein is located in the cytoplasm. This contrasts with the periplasmic location of the mature NapA. However, NapA could not be detected in an isogenic napF(-) strain of R. sphaeroides. The His(6)-tagged NapF protein displayed spectral properties indicative of Fe-S clusters, but these features were rapidly lost, suggesting cluster lability. However, reconstitution of the Fe-S centers into the apo-NapF protein was achieved in the presence of Azotobacter vinelandii cysteine desulfurase (NifS), and this allowed the recovery of nitrate reductase activity in NapA protein that had previously been treated with 2,2'-dipyridyl to remove the [4Fe-4S] cluster. This activity was not recovered in the absence of NapF. Taking into account the cytoplasmic localization of NapF, the presence of labile Fe-S clusters in the protein, the napF(-) strain phenotype, and the NapF-dependent reactivation of the 2,2'-dipyridyl-treated NapA, we propose a role for NapF in assembling the [4Fe-4S] center of the catalytic subunit NapA.  相似文献   

16.
Cytochromes from the SoxAX family have a major role in thiosulfate oxidation via the thiosulfate-oxidizing multi-enzyme system (TOMES). Previously characterized SoxAX proteins from Rhodovulum sulfidophilum and Paracoccus pantotrophus contain three heme c groups, two of which are located on the SoxA subunit. In contrast, the SoxAX protein purified from Starkeya novella was found to contain only two heme groups. Mass spectrometry showed that a disulfide bond replaced the second heme group found in the diheme SoxA subunits. Apparent molecular masses of 27,229 +/- 10.3 Da and 20,258.6 +/- 1 Da were determined for SoxA and SoxX with an overall mass of 49.7 kDa, indicating a heterodimeric structure. Optical redox potentiometry found that the two heme cofactors are reduced at similar potentials (versus NHE) that are as follows: +133 mV (pH 6.0); +104 mV (pH 7.0); +49 (pH 7.9) and +10 mV (pH 8.7). EPR spectroscopy revealed that both ferric heme groups are in the low spin state, and the spectra were consistent with one heme having a His/Cys axial ligation and the other having a His/Met axial ligation. The His/Cys ligated heme is present in different conformational states and gives rise to three distinct signals. Amino acid sequencing was used to unambiguously assign the protein to the encoding genes, soxAX, which are part of a complete sox gene cluster found in S. novella. Phylogenetic analysis of soxA- and soxX-related gene sequences indicates a parallel development of SoxA and SoxX, with the diheme and monoheme SoxA sequences located on clearly separated branches of a phylogenetic tree.  相似文献   

17.
Azotobacter vinelandii is an obligately aerobic bacterium in which aerotolerant dinitrogen fixation requires cytochrome bd. This oxidase comprises two polypeptide subunits and three hemes, but no copper, and has been studied extensively. However, there remain apparently conflicting reports on the reactivity of the high spin heme b(595) with ligands. Using purified cytochrome bd, we show that absorption changes induced by CO photodissociation from the fully reduced cytochrome bd at low temperatures demonstrate binding of the ligand with heme b(595). However, the magnitude of these changes corresponds to the reaction with CO of only about 5% of the heme. CO binding with a minor fraction of heme b(595) is also revealed at room temperature by time-resolved studies of CO recombination. The data resolve the apparent discrepancies between conclusions drawn from room and low temperature spectroscopic studies of the CO reaction with cytochrome bd. The results are consistent with the proposal that hemes b(595) and d form a diheme oxygen-reducing center with a binding capacity for a single exogenous ligand molecule that partitions between the hemes d and b(595) in accordance with their intrinsic affinities for the ligand. In this model, the affinity of heme b(595) for CO is about 20-fold lower than that of heme d.  相似文献   

18.
The 16-kDa diheme cytochrome c from the bacterium Shewanella baltica OS155 (Sb-DHC) was cloned and expressed in Escherichia coli and investigated through UV–vis, magnetic circular dichroism, and 1H NMR spectroscopies and protein voltammetry. The model structure was obtained by means of comparative modeling using the X-ray structure of Rhodobacter sphaeroides diheme cytochrome c (Rs-DHC) (with a 37% pairwise sequence identity) as a template. Sb-DHC folds into two distinct domains, each containing one heme center with a bis-His axial ligation. Both secondary and tertiary structures of the N-terminal domain resemble those of class I cytochrome c, displaying three α-helices and a compact overall folding. The C-terminal domain is less helical than the corresponding domain of Rs-DHC. The two heme groups are bridged by Tyr26 in correspondence with the shortest edge-to-edge distance, a feature which would facilitate fast internal electron transfer. The electronic properties of the two prosthetic centers are equivalent and sensitive to two acid–base equilibria with pK a values of approximately 2.4 and 5, likely corresponding to protonation and detachment of the axial His ligands from the heme iron and a pH-linked conformational change of the protein, respectively. Reduction potentials of −0.144 and −0.257 V (vs. the standard hydrogen electrode), were determined for the C- and N-terminal heme groups, respectively. An approach based on the extended Debye–Hückel equation was applied for the first time to a two-centered metalloprotein and was found to reproduce successfully the ionic strength dependence of E°′.  相似文献   

19.
Behr J  Michel H  Mäntele W  Hellwig P 《Biochemistry》2000,39(6):1356-1363
By specific (13)C labeling of the heme propionates, four bands in the reduced-minus-oxidized FTIR difference spectrum of cytochrome c oxidase from Paracoccus denitrificans have been assigned to the heme propionates [Behr, J., Hellwig, P., M?ntele, W., and Michel, H. (1998) Biochemistry 37, 7400-7406]. To attribute these signals to the individual propionates, we have constructed seven cytochrome coxidase variants using site-directed mutagenesis of subunit I. The mutant enzymes W87Y, W87F, W164F, H403A, Y406F, R473K, and R474K were characterized by measurement of enzymatic turnover, proton pumping activity, and Vis and FTIR spectroscopy. Whereas the mutant enzymes W164F and Y406F were found to be structurally altered, the other cytochrome c oxidase variants were suitable for band assignment in the infrared. Reduced-minus-oxidized FTIR difference spectra of the mutant enzymes were used to identify the ring D propionate of heme a as a likely proton acceptor upon reduction of cytochromic oxidase. The ring D propionate of heme a(3) might undergo conformational changes or, less likely, act as a proton donor.  相似文献   

20.
The sequential flow of electrons in the respiratory chain, from a low reduction potential substrate to O(2), is mediated by protein-bound redox cofactors. In mitochondria, hemes-together with flavin, iron-sulfur, and copper cofactors-mediate this multi-electron transfer. Hemes, in three different forms, are used as a protein-bound prosthetic group in succinate dehydrogenase (complex II), in bc(1) complex (complex III) and in cytochrome c oxidase (complex IV). The exact function of heme b in complex II is still unclear, and lags behind in operational detail that is available for the hemes of complex III and IV. The two b hemes of complex III participate in the unique bifurcation of electron flow from the oxidation of ubiquinol, while heme c of the cytochrome c subunit, Cyt1, transfers these electrons to the peripheral cytochrome c. The unique heme a(3), with Cu(B), form a catalytic site in complex IV that binds and reduces molecular oxygen. In addition to providing catalytic and electron transfer operations, hemes also serve a critical role in the assembly of these respiratory complexes, which is just beginning to be understood. In the absence of heme, the assembly of complex II is impaired, especially in mammalian cells. In complex III, a covalent attachment of the heme to apo-Cyt1 is a prerequisite for the complete assembly of bc(1), whereas in complex IV, heme a is required for the proper folding of the Cox 1 subunit and subsequent assembly. In this review, we provide further details of the aforementioned processes with respect to the hemes of the mitochondrial respiratory complexes. This article is part of a Special Issue entitled: Cell Biology of Metals.  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