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.
Mutagenic specificity: reversion of iso-1-cytochrome c mutants of yeast   总被引:19,自引:0,他引:19  
In previous studies the nucleotide sequences of numerous mutant codons in the cy1 gene have been identified from altered iso-1-cytochromes c. These studies not only revealed the mutant codons that caused the deficiencies but also experimentally determined which of the base pair changes allowed the formation of functional iso-1-cytochromes c. In this investigation we have quantitatively measured the reversion frequencies of eleven cy1 mutants which were treated with 12 mutagens. The cy1 mutants comprised nine mutants having single-base changes of the AUG initiation codon (Stewart et al., 1971), an ochre mutant cy1–9 (Stewart et al., 1972), and an amber mutant cy1–179 (Stewart &; Sherman, 1972). In some cases the types of induced base changes could be inferred unambiguously from the pattern of reversion. Selective G.C to A.T transitions were induced by ethyl methanesulfonate, diethyl sulfate, N-methyl-N′-nitro-N-nitrosoguanidine, 1-nitrosoimidazolidone-2, nitrous acid, [5-3H]uridine and β-propiolactone. There was no apparent specificity with methyl methanesulfonate, dimethyl sulfate, nitrogen mustard and γ-rays. Ultraviolet light induced high rates of reversion of the ochre and amber mutants, but in these instances it appears as if the selective action is due to particular nucleotide sequences and not due to simple types of base pair changes.  相似文献   

2.
We present evidence that two non-allelic genes, located on two non-homologous chromosomes in the yeast Saccharomyces cerevisiae, recombine and in this process generate new composite genes containing portions of both genes. The two genes CYC1 and CYC7 encode, respectively, iso-1-cytochrome c and iso-2-cytochrome c; CYC1 is located on the right arm of chromosome X and CYC7 is located on the left arm of chromosome V. The coding regions of CYC1 and CYC7 and the corresponding iso-1-cytochrome c and iso-2-cytochrome c are approximately 80% homologous. Composite genes were uncovered among revertants of certain but not all cyc1 mutants lacking iso-1-cytochrome c; composite genes were observed in most revertants from the low-reverting strains cyc1-11, cyc1-136 and cyc1-158, and in low proportions of the revertants from the typically reverting strains cyc1-94 and cyc1-156. Protein analysis of 14 composite iso-cytochromes c and DNA sequencing of five composite genes indicated that recombinational events produced replacements of central portions of the cyc1 gene with a corresponding segment from the wild-type CYC7+ gene. The replacements varied in length from 13% to 61% of the translated portion of the CYC1 locus. The formation of composite genes occurred spontaneously at very low frequencies and at low but enhanced frequencies after treatments with mutagens including ultraviolet light, ethylmethane sulfonate, methylmethane sulfonate and nitrous acid. Genetic tests indicated that composite genes are formed mitotically by a conversion-like event in which the wild-type CYC1+ allele remains intact. Recombination between non-allelic genes can lead to identical sequences at different loci and to diverse composite genes. These results support the indirect evidence from other eukaryotic systems that non-allelic genes with extensive but not complete homology recombine during evolution.  相似文献   

3.
The structure of the membrane integral rotor ring of the proton translocating F1F0 ATP synthase from spinach chloroplasts was determined to 3.8 Å resolution by x-ray crystallography. The rotor ring consists of 14 identical protomers that are symmetrically arranged around a central pore. Comparisons with the c11 rotor ring of the sodium translocating ATPase from Ilyobacter tartaricus show that the conserved carboxylates involved in proton or sodium transport, respectively, are 10.6–10.8 Å apart in both c ring rotors. This finding suggests that both ATPases have the same gear distance despite their different stoichiometries. The putative proton-binding site at the conserved carboxylate Glu61 in the chloroplast ATP synthase differs from the sodium-binding site in Ilyobacter. Residues adjacent to the conserved carboxylate show increased hydrophobicity and reduced hydrogen bonding. The crystal structure reflects the protonated form of the chloroplast c ring rotor. We propose that upon deprotonation, the conformation of Glu61 is changed to another rotamer and becomes fully exposed to the periphery of the ring. Reprotonation of Glu61 by a conserved arginine in the adjacent a subunit returns the carboxylate to its initial conformation.ATP synthases found in the energy-transducing membranes of bacteria, mitochondria, and chloroplasts catalyze ATP synthesis and ATP hydrolysis coupled with transmembrane proton or sodium ion transport. The enzymes are multi-subunit complexes composed of an extra-membranous catalytic F1 domain and an interconnected integral membrane F0 domain. The hydrophilic F1 domain consists of five different polypeptides with a stoichiometry of α3β3γδϵ. Detailed structural information obtained with the mitochondrial enzyme (13) in combination with biochemical (4), biophysical (5), and single molecule studies (69) revealed that synthesis or hydrolysis of ATP in the F1 domain is accomplished via a rotary catalytic mechanism. In addition to information on the catalytic mechanism, structure analysis and single molecule studies of the mitochondrial or the chloroplast F1 complex have also unraveled the molecular mechanism of several F1-specific inhibitors (1014). Less detailed information is available on the integral membrane F0 domain, which consists of three different polypeptides (a, b, and c) and mediates the transfer of protons or sodium ions across the membrane. Subunits a and b were shown to reside at the periphery of a cylindrical complex formed by multiple copies of the c subunit (1518). The number of c subunits in the cylindrical subcomplex shows substantial variation in different organisms. Ten protomers are found in ATP synthases from yeast, Escherichia coli and Bacillus PS3 (1921), 11 in Ilyobacter tartaricus, Propionigenium modestum, and Clostridium paradoxum (2224), 13 in the thermoalkalophilic Bacillus TA2.TA1 (25), 14 in spinach chloroplasts (26), and 15 in the cyanobacterium Spirulina platensis (27). The structure of isolated subunits a, b, and c from E. coli has been studied by mutagenesis analysis and by NMR spectroscopy in a mixed solvent that was suggested to mimic the membrane environment (2832). These studies showed that subunit a folds with five membrane-spanning helices. The fourth of these helices directly interacts with subunit c and contains a conserved arginine (Arg210), which is thought to be involved in proton transfer (33). Subunit b, which is present in two copies in the intact F0, contains a single transmembrane helix. Cross-linking data support a direct interaction of the two copies of the b subunit (29). Subunit c was studied at two different pH values to obtain the protonated and deprotonated form of a conserved carboxylate (Asp61 in E. coli) that was shown to be essential for proton transport (34). NMR spectroscopy revealed that the isolated c subunit consists of two long hydrophobic membrane spanning segments connected by a short hydrophilic loop (30, 35). This loop is located close to the γ and ϵ subunit on the F1 side of the membrane (36, 37). Low resolution x-ray crystallography, cryo-electron microscopy, and atomic force microscopy showed that the membrane-spanning helices of the multiple copies of subunit c in the intact F0 complex are tightly packed in two concentric rings (19, 22, 26). Atomic resolution of the c ring was recently provided for the Na+-translocating F-type ATPase from I. tartaricus (38) and the related Na+-translocating V-type ATPase from Enterococcus hirae (39). Rotation of the c ring was demonstrated by cross-linking (18), fluorescence studies (40), and single molecule visualization (41, 42). Based on the structural and biochemical information on F1 and F0, different mechanical models have been proposed describing how the rotation of the c ring is coupled to the rotation of the F1 rotor subunits. This rotation in turn drives sequential conformational shifts at the three catalytic β subunits that result in ATP synthesis (4345). Vice versa hydrolysis of ATP in the F1 domain is thought to drive rotation of the γϵc10–15 subcomplex and transports protons or sodium ions across the membrane.Here we describe the crystal structure of the chloroplast c14 rotor, which is the first structure of an isolated c ring rotor from a proton driven ATPase. The structure was solved by molecular replacement using a tetradecameric search model that was generated from a monomer taken from the I. tartaricus c11 structure. The imposition of noncrystallographic symmetry restraints during refinement substantially improved electron density and structure determination.  相似文献   

4.
Rotary catalysis in F1F0 ATP synthase is powered by proton translocation through the membrane-embedded F0 sector. Proton binding and release occur in the middle of the membrane at Asp-61 on transmembrane helix (TMH) 2 of subunit c. Previously the reactivity of Cys substituted into TMH2 revealed extensive aqueous access at the cytoplasmic side as probed with Ag+ and other thiolate-directed reagents. The analysis of aqueous accessibility of membrane-embedded regions in subunit c was extended here to TMH1 and the periplasmic side of TMH2. The Ag+ sensitivity of Cys substitutions was more limited on the periplasmic versus cytoplasmic side of TMH2. In TMH1, Ag+ sensitivity was restricted to a pocket of four residues lying directly behind Asp-61. Aqueous accessibility was also probed using Cd2+, a membrane-impermeant soft metal ion with properties similar to Ag+. Cd2+ inhibition was restricted to the I28C substitution in TMH1 and residues surrounding Asp-61 in TMH2. The overall pattern of inhibition, by all of the reagents tested, indicates highest accessibility on the cytoplasmic side of TMH2 and in a pocket of residues around Asp-61, including proximal residues in TMH1. Additionally subunit a was shown to mediate access to this region by the membrane-impermeant probe 2-(trimethylammonium)ethyl methanethiosulfonate. Based upon these results and other information, a pocket of aqueous accessible residues, bordered by the peripheral surface of TMH4 of subunit a, is proposed to extend from the cytoplasmic side of cTMH2 to Asp-61 in the center of the membrane.F1F0 ATP synthase utilizes the energy stored in an H+ or Na+ electrochemical gradient to synthesize ATP in bacteria, mitochondria, and chloroplasts (14). The ATP synthase complex is composed of two sectors, i.e. a water-soluble F1 sector that is bound to a membrane-embedded F0 sector. In bacteria, F1 is composed of five subunits in an α3β3γδϵ ratio and contains three catalytic sites for ATP synthesis and/or hydrolysis centered at the α-β subunit interfaces. F0 is composed of three subunits in an a1b2c10–15 ratio and functions as the ion-conducting pathway (59). Ion translocation through F0 drives rotation of a cylindrical ring of c-subunits that is coupled to rotation of the γ subunit within the (αβ)3 hexamer of F1 to force conformational changes in the three active sites and in turn drive synthesis of ATP by the binding change mechanism (14, 1013).Subunit c of F0 folds in the membrane as a hairpin of two extended α-helices. In Escherichia coli, 10 copies of subunit c pack together to form a decameric ring with TMH12 on the inside and TMH2 on the periphery (6, 14). An atomic resolution structure of the Na+-translocating c11-ring from Ilyobacter tartaricus was recently published by Meier et al. (8). In the c11 structure, the Na+ binding site is formed by two interacting c subunits. The essential Na+-binding Glu residue, which corresponds to Asp-61 in E. coli, is located in TMH2 at the middle of the lipid bilayer. Subunit a consists of five transmembrane helices, four of which likely interact as a four-helix bundle (1518). Subunit a lies on the periphery of the c-ring with TMHs 4 and 5 from subunit a and TMH2 from subunit c forming the a-c interface (1821). During ion translocation through F0, the essential Arg-210 on TMH4 of subunit a is postulated to facilitate the protonation/deprotonation cycle at Asp-61 of subunit c and cause the rotation of the c-ring past the stationary subunit a (3, 4, 19).Chemical modification of cysteine-substituted transmembrane proteins has been widely used as a means of probing the aqueous accessible regions (2224). The reactivity of a substituted cysteine to thiolate-directed probes provides an indication of aqueous accessibility because the reactive thiolate species is preferentially formed in an aqueous environment. The aqueous accessibility of the five TMHs in subunit a of E. coli F0 has been probed using Ag+ and NEM (19, 2527). The results suggest the presence of an aqueous accessible channel in subunit a in the center of TMHs 2–5 extending from the periplasm to the center of the membrane. Protons entering through this periplasmic access channel are postulated to bind to the essential Asp-61 residues of the c-ring and exit to the cytoplasm by a still uncertain pathway at the peripheral face of aTMH4 with protonation/deprotonation of Asp-61 driving c-ring rotation.During H+-driven ATP synthesis, two models for the pathway by which H+ or Na+ exit to the cytoplasm have been proposed. The first model proposes that the ions bound at Asp-61 exit to the cytoplasm via a half-channel composed at least partially by residues in TMH4 of subunit a (2527). Chemical modification studies of Cys-substituted subunit a of E. coli revealed an aqueous accessible surface of TMH4 that includes the essential Arg-210 residue, which extended from the center of the membrane to the cytoplasm, suggesting that the ion exit channel may lie at the a-c interface (19, 25). Alternatively studies of the c-ring from the I. tartaricus enzyme indicate that Na+ can access Glu-65 in the absence of other F0 subunits, suggesting an intrinsic channel in subunit c (28, 29). However, no such channel was apparent in the crystal structure of the c11-ring (8). In a previous study (30), we probed the thiolate reactivity of Cys substitutions in the cytoplasmic half of TMH2 in subunit c. These experiments revealed extensive reactivity to sulfhydryl-directed reagents on the peripheral face of cTMH2, supporting the presence of the cytoplasmic exit channel at the a-c interface. In this study, we extended the survey of aqueous accessibility in transmembrane regions by probing thiolate reactivity of Cys substitutions in TMH1 and in the periplasmic half of TMH2. The reactivity of Cys substituted into these regions proved to be more limited. Only a small region of TMH1, lying directly behind Asp-61, was reactive with Ag+. In addition to Ag+, we used Cd2+ as a complementary, membrane-impermeant probe for aqueous accessibility. The survey of Cd2+ sensitivity confirmed that aqueous accessibility from the cytoplasm is much greater for residues packing at the periphery of the c-ring. The experiments reported here distinguish the aqueous accessible and inaccessible regions of the c-ring and strengthen evidence that the cytoplasmic H+ exit channel is situated at the a-c interface.  相似文献   

5.
6.
7.
Oxidoreductases of the cytochrome bc1/b6f family transfer electrons from a liposoluble quinol to a soluble acceptor protein and contribute to the formation of a transmembrane electrochemical potential. The crystal structure of cyt b6f has revealed the presence in the Qi site of an atypical c-type heme, heme ci. Surprisingly, the protein does not provide any axial ligand to the iron of this heme, and its surrounding structure suggests it can be accessed by exogenous ligand. In this work we describe a mutagenesis approach aimed at characterizing the ci heme and its interaction with the Qi site environment. We engineered a mutant of Chlamydomonas reinhardtii in which Phe40 from subunit IV was substituted by a tyrosine. This results in a dramatic slowing down of the reoxidation of the b hemes under single flash excitation, suggesting hindered accessibility of the heme to its quinone substrate. This modified accessibility likely originates from the ligation of the heme iron by the phenol(ate) side chain introduced by the mutation. Indeed, it also results in a marked downshift of the ci heme midpoint potential (from +100 mV to −200 mV at pH 7). Yet the overall turnover rate of the mutant cytochrome b6f complex under continuous illumination was found similar to the wild type one, both in vitro and in vivo. We propose that, in the mutant, a change in the ligation state of the heme upon its reduction could act as a redox switch that would control the accessibility of the substrate to the heme and trigger the catalysis.The cytochrome b6f complex of oxygenic photosynthesis is the integral membrane protein, the quinol:plastocyanin oxido-reductase activity of which allows the linear electron flow between the two photosystems (PSI and PSII).4 The turnover of the cytochrome b6f complex depends on the steady state of its redox partners, the liposoluble plastoquinol (PQH2 reduced and protonated plastoquinone PQ) formed by the PSII, and the hydrosoluble plastocyanin oxidized by the PSI. In the Qo site, exposed to the lumenal side, the quinol is oxidized, and this oxidation is coupled to the release of two protons into the lumen. The two electrons provided by the quinol are transferred along two bifurcated pathways, the high and low potential chains. The high potential chain involves two lumenal redox partners, the membrane-anchored flexible Rieske [2Fe-2S] cluster and the cytochrome f, which ultimately interacts with the soluble plastocyanin. In the low potential chain, electrons are transferred to the stroma via the low and high potential b hemes (bL and bH) of the transmembrane b6 subunit. Two turnovers of the cyt b6f complex lead to the reduction of the low potential chain, thereby allowing the reduction of a quinone molecule in the stromal Qi pocket. This mechanism, which recycles reducing equivalents, is referred to as the “Q cycle,” initially described by Mitchell (1) and modified later by Crofts et al. and others (2, 3).Although this quinol:cytochrome oxidoreductase activity is involved in both the respiratory and photosynthetic electron transfer chains, recent x-ray data (46) have evidenced major structural differences between the b6f complex and its mitochondrial counterpart the bc1 complex (for reviews see Refs. 710). Indeed an additional heme localized in close contact with heme bH stands as another putative electron carrier as proposed earlier by Lavergne (11). Since it was brought to light by the x-ray studies, knowledge of the basic properties of this heme, named ci in reference to the Qi site (5) or cn in reference to its proximity to the negatively charged side of the membrane (4), has significantly improved. The proteins involved in the assembly machinery of the heme have been identified in Chlamydomonas reinhardtii and Arabidopsis thaliana (12, 13). Consistent with the structure, according to which the only axial ligand could be a water molecule interacting with the proponiate chain of the bH heme, the spectroscopic properties of this heme are those of a high spin heme (14, 15). Evidences for a high spin heme covalently bound to the cytochrome b subunit were also found in Heliobacterium modesticaldum and Heliobacillus mobilis (16).In the b6f complex from the oxygenic photosynthetic chain, EPR (15) and structural data (17) have shown that NQNO (2-n-nony l-4-hydroxyquinoline N-oxide), an inhibitor of the Qi pocket (18, 19), can act as an axial ligand to ci. This ligation is accompanied by a significant change in the redox properties of ci, because, in the presence of NQNO, at least two titrations waves were observed (13, 14), one with a midpoint potential (Em) similar to that observed in the absence of NQNO and the other with a midpoint potential downshifted by ∼250 mV. This, together with the widespread range of redox potential found for heme ci (11, 14, 15, 20), points to a structural plasticity of the ci ligand network.This plasticity may arise from the unusual coordination properties of the heme ci. As a matter of fact, the x-ray models of the complex from C. reinhardtii and Mastigocladus laminosus evidenced a water or hydroxyl molecule as a fifth ligand. The sixth position of coordination is directed toward the Qi pocket and appears as free. Nevertheless, the side chain of Phe40 of subunit IV protrudes above the heme plane, leaving little space for any axial ligand to the heme ci. Besides, modeling a quinone analogue in the electron density found in the Qi pocket of structures obtained in presence of Tridecyl- Stigmatellin or NQNO implies a steric clash with the native position of the Phe40 aromatic ring.5 The Phe40 residue of subunit IV therefore stands as a key residue for the plasticity of the site, making it an ideal mutagenesis target when attempting to alter the possible interactions between ci and the quinone or quinol (4, 5) (Fig. 1). Here we present the consequences of the substitution of Phe40 by a tyrosine on the properties and function of the ci heme.Open in a separate windowFIGURE 1.A view of the Qi site from the dimer interface. The coordinates are from the Protein Data Bank 1Q90 model. The Van der Waal''s surface of the peptide chains was drawn with Pymol. Phe40 is in van der Waal''s contact with the plane of the ci, heme.  相似文献   

8.
Programmed cell death is an event displayed by many different organisms along the evolutionary scale. In plants, programmed cell death is necessary for development and the hypersensitive response to stress or pathogenic infection. A common feature in programmed cell death across organisms is the translocation of cytochrome c from mitochondria to the cytosol. To better understand the role of cytochrome c in the onset of programmed cell death in plants, a proteomic approach was developed based on affinity chromatography and using Arabidopsis thaliana cytochrome c as bait. Using this approach, ten putative new cytochrome c partners were identified. Of these putative partners and as indicated by bimolecular fluorescence complementation, nine of them bind the heme protein in plant protoplasts and human cells as a heterologous system. The in vitro interaction between cytochrome c and such soluble cytochrome c-targets was further corroborated using surface plasmon resonance. Taken together, the results obtained in the study indicate that Arabidopsis thaliana cytochrome c interacts with several distinct proteins involved in protein folding, translational regulation, cell death, oxidative stress, DNA damage, energetic metabolism, and mRNA metabolism. Interestingly, some of these novel Arabidopsis thaliana cytochrome c-targets are closely related to those for Homo sapiens cytochrome c (Martínez-Fábregas et al., unpublished). These results indicate that the evolutionarily well-conserved cytosolic cytochrome c, appearing in organisms from plants to mammals, interacts with a wide range of targets on programmed cell death. The data have been deposited to the ProteomeXchange with identifier PXD000280.Programmed cell death (PCD)1 is a fundamental event for the development of multicellular organisms and the homeostasis of their tissues. It is an evolutionarily conserved mechanism present in organisms ranging from yeast to mammals (13).In mammals, cytochrome c (Cc) and dATP bind to apoptosis protease-activating factor-1 (Apaf-1) in the cytoplasm, a process leading to the formation of the Apaf-1/caspase-9 complex known as apoptosome. This apoptosome subsequently activates caspases-3 and -7 (4, 5). In other organisms, such as Caenorhabditis elegans or Drosophila melanogaster, however, Cc is not essential for the assembly and activation of the apoptosome (6) despite the presence of proteins homologous to Apaf-1—cell death abnormality-4 (CED-4) in C. elegans and Drosophila Apaf-1-related killer (Dark) in D. melanogaster—which have been found to be essential for caspase cascade activation. Furthermore, other organisms such as Arabidopsis thaliana lack Apaf-1 (7). In fact, only highly distant caspase homologues (metacaspases) (8, 9), serine proteases (saspases) (10), phytaspases (11) and VEIDases (1214) with caspase-like activity have been detected in plants; however, their targets remain veiled and whether they are activated by Cc remains unclear.Intriguingly, the release of Cc from mitochondria into the cytoplasm during the onset of PCD is an evolutionarily conserved event found in organisms ranging from yeast (15) and plants (16) to flies (17), and mammals (18). However, understanding of the roles of this phenomenon in different species can be said to be uneven at best. In fact, the release of Cc from mitochondria has thus far been considered a random event in all organisms, save mammals. Thus, the participation of Cc in the onset and progression of PCD needs to be further elucidated.Even in the case of mammals, the role(s) of Cc in the cytoplasm during PCD remain(s) controversial. Recently, new putative functions of Cc, going beyond the already-established apoptosome assembly process, have been proposed in the nucleus (19, 20) and the endoplasmic reticulum (2123). Neither these newly proposed functions nor other arising functions, such as oxidative stress (24), are as yet fully understood. This current state of affairs demands deeper exploration of the additional roles played by Cc in nonmammalian species.In this study, putative novel Cc-partners involved in plant PCD were identified. For this identification, a proteomic approach was employed based on affinity chromatography and using Cc as bait. The Cc-interacting proteins were identified using nano-liquid chromatography tandem mass spectrometry (NanoLC-MS/MS). These Cc-partners were then further confirmed in vivo through bimolecular fluorescence complementation (BiFC) in A. thaliana protoplasts and human HEK293T cells, as a heterologous system. Finally, the Cc-GLY2, Cc-NRP1 and Cc-TCL interactions were corroborated in vitro using surface plasmon resonance (SPR).These results indicate that Cc is able to interact with targets in the plant cell cytoplasm during PCD. Moreover, they provide new ways of understanding why Cc release is an evolutionarily well-conserved event, and allow us to propose Cc as a signaling messenger, which somehow controls different essential events during PCD.  相似文献   

9.
The facultative phototrophic bacterium Rhodobacter capsulatus contains only one form of cytochrome (cyt) c oxidase, which has recently been identified as a cbb3-type cyt c oxidase. This is unlike other related species, such as Rhodobacter sphaeroides and Paracoccus denitrificans, which contain an additional mitochondrial-like aa3-type cyt c oxidase. An extensive search for mutants affected in cyt c oxidase activity in R. capsulatus led to the isolation of at least five classes of mutants. Plasmids complementing them to a wild-type phenotype were obtained for all but one of these classes from a chromosomal DNA library. The first class of mutants contained mutations within the structural genes (ccoNOQP) of the cyt cbb3 oxidase. Sequence analysis of these mutants and of the plasmids complementing them revealed that ccoNOQP in R. capsulatus is not flanked by the oxygen response regulator fnr, which is located upstream of these genes in other species. Genetic and biochemical characterizations of mutants belonging to this group indicated that the subunits CcoN, CcoO, and CcoP are required for the presence of an active cyt cbb3 oxidase, and unlike in Bradyrhizobium japonicum, no active CcoN-CcoO subcomplex was found in R. capsulatus. In addition, mutagenesis experiments indicated that the highly conserved open reading frame 277 located adjacent to ccoNOQP is required neither for cyt cbb3 oxidase activity or assembly nor for respiratory or photosynthetic energy transduction in R. capsulatus. The remaining cyt c oxidase-minus mutants mapped outside of ccoNOQP and formed four additional groups. In one of these groups, a fully assembled but inactive cyt cbb3 oxidase was found, while another group had only extremely small amounts of it. The next group was characterized by a pleiotropic effect on all membrane-bound c-type cytochromes, and the remaining mutants not complemented by the plasmids complementing the first four groups formed at least one additional group affecting the biogenesis of the cyt cbb3 oxidase of R. capsulatus.The gram-negative facultative photosynthetic bacterium Rhodobacter capsulatus has a highly branched electron transport chain, resulting in its ability to grow under a wide variety of conditions (52). Its light-driven photosynthetic electron transfer pathway is a cyclic process between the photochemical reaction center and the ubihydroquinone cytochrome (cyt) c oxidoreductase (cyt bc1 complex) (30). On the other hand, the respiratory electron transfer pathways of R. capsulatus are branched after the quinone pool and contain two different terminal oxidases, previously called cyt b410 (cyt c oxidase) and cyt b260 (quinol oxidase) (3, 27, 29, 53). The branch involving cyt c oxidase is similar to the mitochondrial electron transfer chain in that it depends on the cyt bc1 complex and a c-type cyt acting as an electron carrier. The quinol oxidase branch circumvents the cyt bc1 complex and the cyt c oxidase by taking electrons directly from the quinone pool to reduce O2 to H2O. The pronounced metabolic versatility, including the ability to grow under dark, anaerobic conditions (50, 52), makes these purple non-sulfur bacteria excellent model organisms for studying microbial energy transduction.Marrs and Gest (29) have reported the first R. capsulatus mutants which were defective in the respiratory electron transport chain. Of these mutants, M5 was incapable of catalyzing the α-naphthol plus N′,N′-dimethyl-p-phenylenediamine (DMPD) plus O2→indophenol blue plus H2O reaction (NADI reaction) and unable to grow by respiration (Res), and hence was deficient in both terminal oxidases. Another mutant, M4, was also NADI but Res+ due to the presence of an active quinol oxidase. Marrs and Gest have also described two different spontaneous revertants of M5, called M6 and M7, which regained the ability to grow by respiration (29). M6 regained cyt c oxidase activity and became concurrently NADI+ and sensitive to low concentrations of cyanide and the cyt bc1 inhibitor myxothiazol, but remained quinol oxidase. On the other hand, M7 regained the quinol oxidase activity but remained cyt c oxidase (thus, NADI and resistant to myxothiazol, a phenotype identical to that of M4). All of these mutants remained proficient for phototrophic (Ps) growth.The cyt c oxidase of R. capsulatus has been purified previously and characterized as being a novel cbb3-type cyt c oxidase without a CuA center (15). It is composed of at least a membrane-integral b-type cyt (subunit I [CcoN]) with a low-spin heme b and a high-spin heme b3-CuB binuclear center, and two membrane-anchored c-type cyts (CcoO and CcoP). It has a unique active site that possibly confers a very high affinity for its substrate oxygen (49). The structural genes of this enzyme (ccoNOQP) have been sequenced recently from R. capsulatus 37b4 (45) and aligned to the partial amino acid sequence of the purified enzyme from R. capsulatus MT1131 (15). Although a ccoN mutant of strain 37b4 was reported to lack cyt c oxidase activity (45), the observed discrepancies between the amino acid sequence and the nucleotide sequence do not entirely exclude the possible presence of two similar cb-type cyt c oxidases in this species. The presence of a similar cyt c oxidase has also been demonstrated in several other bacteria, including P. denitrificans (9), R. sphaeroides (13), and Rhizobium spp. In the latter species, the homologs of ccoNOQP have been named fixNOQP (23, 34) and are required to support respiration under oxygen-limited growth during symbiotic nitrogen fixation (36).The biogenesis of a multisubunit protein complex containing several prosthetic groups, such as cyt cbb3 oxidase, is likely to require many accessory proteins involved in various posttranslational events, including protein translocation, assembly, cofactor insertion, and maturation (46). Thus, insights into this important biological process, about which currently little is known, may be gained by searching for mutants defective in cyt c oxidase activity. In this work, we describe the isolation of such mutants and their molecular genetic characterization, including those already available, such as M4, M5, and M7G. These studies indicate that in R. capsulatus, gene products of at least five different loci are involved in the formation of an active cyt cbb3 oxidase.  相似文献   

10.
myo1c is a single-headed myosin that dynamically links membranes to the actin cytoskeleton. A putative pleckstrin homology domain has been identified in the myo1c tail that binds phosphoinositides and soluble inositol phosphates with high affinity. However, the kinetics of association and dissociation and the influence of phospholipid composition on the kinetics have not been determined. Stopped-flow spectroscopy was used to measure the binding and dissociation of a recombinant myo1c construct containing the tail and regulatory domains (myo1cIQ-tail) to and from 100-nm diameter large unilamellar vesicles (LUVs). We found the time course of association of myo1cIQ-tail with LUVs containing 2% phosphatidylinositol 4,5-bisphosphate (PtdIns(4,5)P2) followed a two-exponential time course, and the rate of the predominant fast phase depended linearly upon the total lipid concentration. The apparent second-order rate constant was approximately diffusion-limited. Increasing the molar ratio of anionic phospholipid by adding phosphatidylserine, additional PtdIns(4,5)P2, or by situating PtdIns(4,5)P2 in a more physiologically relevant lipid background increased the apparent association rate constant less than 2-fold. myo1cIQ-tail dissociated from PtdIns(4,5)P2 at a slower rate (2.0 s−1) than the pleckstrin homology domain of phospholipase C-δ (13 s−1). The presence of additional anionic phospholipid reduced the myo1cIQ-tail dissociation rate constant >50-fold but marginally changed the dissociation rate of phospholipase C-δ, suggesting that additional electrostatic interactions in myo1cIQ-tail help to stabilize binding. Remarkably, high concentrations of soluble inositol phosphates induce dissociation of myo1cIQ-tail from LUVs, suggesting that phosphoinositides are able to bind to and dissociate from myo1cIQ-tail as it remains bound to the membrane.Myosin-I isoforms are low molecular weight members of the myosin superfamily that link cell membranes with the actin cytoskeleton and play crucial roles driving a diverse array of dynamic membrane processes (15). Cell biological studies have shown that myosin-I isoforms localize and fractionate with cell membranes (2, 6), and biochemical experiments have shown myosin-I isoforms bind directly to lipid membranes (710). Thus, a key property of some myosin-I isoforms is their ability to bind membranes.myo1c is a widely expressed vertebrate myosin-I isoform that has roles in a variety of important membrane events, including insulin-stimulated fusion of vesicles containing glucose transporter-4 with the plasma membrane (2, 11), compensatory endocytosis following regulated exocytosis (12), and tensioning of mechano-sensitive ion channels (3). The mechanisms of myo1c targeting and anchoring to specific regions on the membrane to support these functions are not well understood. However, evidence is building that myo1c targeting requires direct binding of myo1c to phosphoinositides in cell membranes (1316).We have shown that binding of myo1c to membranes is mediated by a putative pleckstrin homology (PH)3 domain in its tail that binds phosphatidylinositol 4,5-bisphosphate (PtdIns(4,5)P2) and other phosphoinositides with high affinity (Kd <0.5 μm in terms of accessible phosphoinositide concentration) (13). myo1c also binds soluble inositol phosphates (e.g. inositol 1,4,5-trisphosphate (InsP3)) with similar affinity. Point mutations of amino acids known to be essential for phosphoinositide binding in other PH domains inhibit myo1c binding to PtdIns(4,5)P2 in vitro, and these mutations disrupt membrane localization in vivo (13). The affinity of myo1c for PtdIns(4,5)P2-containing membranes is increased by the presence of additional anionic phospholipids in the membrane. This increased affinity may be due to nonspecific electrostatic interactions between the anionic phospholipids and positively charged regions within the myo1c tail or regulatory domain (13, 17), which is similar to what has been found for the guanine nucleotide exchange factor, ARNO (18). However, high affinity membrane binding via these nonspecific electrostatic interactions (i.e. binding in the absence of PtdIns(4,5)P2) requires the membrane composition to contain a nonphysiological mole fraction (e.g. >40% phosphatidylserine) of anionic phospholipids (13, 14).Because phosphoinositide binding is important for the cellular localization and function of myo1c (13), it is important to determine the physical constants that define this interaction. Determining the kinetics of membrane attachment will provide insight into the relationship between membrane attachment and actin attachment lifetimes and will also provide details about the role of anionic lipids in regulating membrane attachment. Therefore, we used stopped-flow kinetics to measure the in vitro association and dissociation kinetics of myo1c from LUVs as a function of phosphoinositide composition and anionic charge.  相似文献   

11.
Damage of presynaptic mitochondria could result in release of proapoptotic factors that threaten the integrity of the entire neuron. We discovered that α-synuclein (Syn) forms a triple complex with anionic lipids (such as cardiolipin) and cytochrome c, which exerts a peroxidase activity. The latter catalyzes covalent hetero-oligomerization of Syn with cytochrome c into high molecular weight aggregates. Syn is a preferred substrate of this reaction and is oxidized more readily than cardiolipin, dopamine, and other phenolic substrates. Co-localization of Syn with cytochrome c was detected in aggregates formed upon proapoptotic stimulation of SH-SY5Y and HeLa cells and in dopaminergic substantia nigra neurons of rotenone-treated rats. Syn-cardiolipin exerted protection against cytochrome c-induced caspase-3 activation in a cell-free system, particularly in the presence of H2O2. Direct delivery of Syn into mouse embryonic cells conferred resistance to proapoptotic caspase-3 activation. Conversely, small interfering RNA depletion of Syn in HeLa cells made them more sensitive to dopamine-induced apoptosis. In human Parkinson disease substantia nigra neurons, two-thirds of co-localized Syn-cytochrome c complexes occurred in Lewy neurites. Taken together, these results indicate that Syn may prevent execution of apoptosis in neurons through covalent hetero-oligomerization of cytochrome c. This immediate protective function of Syn is associated with the formation of the peroxidase complex representing a source of oxidative stress and postponed damage.Lewy bodies (LBs),3 mitochondrial impairment, and oxidative stress are cardinal features of Parkinson disease (PD) and several related neurodegenerative disorders (1, 2). Aggregation of α-synuclein (Syn), an abundant protein in synaptic terminals, plays a major role in the formation of LBs (3, 4). Neither the mechanisms of LB production nor their pathogenic or protective roles in neurodegeneration are well understood.In nigrostriatal dopaminergic synaptic terminals, mitochondria, harboring a host of death-initiating factors, are in peril of damage by reactive oxygen species generated by disrupted electron transport and/or oxidative metabolism of dopamine (DA). Because cytochrome c (cyt c)-dependent formation of apoptosomes and activation of caspases designates a point of no return in the apoptotic program, release of proapoptotic factors from synaptic mitochondria could threaten the integrity of the entire neuron. How neurons protect themselves against inadvertent release of death signals from damaged synaptic mitochondria is not known.The N-terminal fragment of Syn contains six variants of an 11-amino acid consensus motif that include an apolipoprotein-like class A2 helix participating in binding of different lipids, particularly anionic phospholipids (5). This domain is believed to be important for Syn functions in regulation of neuronal lipid metabolism, particularly turnover of a mitochondria-specific phospholipid, cardiolipin (CL) (6). However, the relevance of the Syn lipid binding capacity in regulating neuronal injury (apoptotic) responses has not been established.It is believed that oxidative stress participates in the accumulation of LB and Lewy neurites (LN) through yet to be identified pathways (7). Reportedly, Syn is co-localized with cyt c in LBs (8), indicating a potential interaction between the two proteins. Because cyt c is a redox-active hemeprotein (9, 10), its presence in the LBs in conjunction with Syn may also provide a mechanistic link of LBs with oxidative stress. We have recently reported that cyt c interacts with CL in mitochondria early in apoptosis and with phosphatidylserine (PS) in the plasma membrane after its release into the cytosol (11, 12). In both cases, this results in redox activation of cyt c and the production of complexes with high peroxidase activity that effectively catalyze peroxidation of the respective phospholipids (13).Based on these facts, we hypothesize and provide experimental evidence that Syn acts as a sacrificial scavenger of cytosolic cyt c inadvertently released from synaptic mitochondria to prevent its migration into the soma, i.e. spread of the proapoptotic signal and cell death. This vital function is realized through the emergence of a peroxidase activity of the cyt c-Syn-phospholipid complex that cross-links its components and yields covalently conjugated protein-lipid hetero-oligomers. The latter maintain lingering peroxidase activity. Thus protection against acute apoptotic cell death comes with a penalty of Syn-cyt c aggregation into a peroxidase complex capable of inducing protracted oxidative stress. Our results present a novel biochemical mechanism likely involved in Lewy body formation and explain a known paradox of a dual protective and deleterious role that Syn plays in neuronal cells.  相似文献   

12.
13.
14.
The temperature-dependent rotation of F1-ATPase γ subunit was observed in Vmax conditions at low viscous drag using a 60-nm gold bead (Nakanishi-Matsui, M., Kashiwagi, S., Hosokawa, H., Cipriano, D. J., Dunn, S. D., Wada, Y., and Futai, M. (2006) J. Biol. Chem. 281, 4126–4131). The Arrhenius slopes of the speed of the individual 120° steps and reciprocal of the pause length between rotation steps were very similar, indicating a flat energy pathway followed by the rotationally coupled catalytic cycle. In contrast, the Arrhenius slope of the reciprocal pause length of the γM23K mutant F1 was significantly increased, whereas that of the rotation rate was similar to wild type. The effects of the rotor γM23K substitution and the counteracting effects of βE381D mutation in the interacting stator subunits demonstrate that the rotor-stator interactions play critical roles in the utilization of stored elastic energy. The γM23K enzyme must overcome an abrupt activation energy barrier, forcing it onto a less favored pathway that results in uncoupling catalysis from rotation.F-ATPase (FoF1), consisting of the catalytic sector F13β3γδϵ) and the transmembrane proton transport sector Fo (ab2c10), synthesizes or hydrolyzes ATP coupled with proton transport (for reviews, see Ref. 16). As Abrahams et al. (7) discovered in the first high resolution x-ray structure, a critical feature of the F1-ATPase is the inherent asymmetry of the three β subunits in different conformations, βTP, βDP, and βE, referring to the nucleotide bound in each catalytic site, ATP, ADP, or empty, respectively. A rotational mechanism has been firmly established mostly based on direct observation in single molecule experiments of the behavior of the rotor complex ϵγc10, relative to the stator complex α3β3δab2 (reviewed in Ref. 1). ATP hydrolysis-dependent rotation of the γ and ϵ subunits in purified bacterial F1 (8, 9), the ϵγc10 complex in detergent solubilized FoF1 (1013), and the ϵγc10 complexin FoF1 in lipid bilayers (14) were shown experimentally by single molecule observations using fluorescent actin filament as a probe. Relative rotation of the single copy Fo a subunit was also shown in F0F1, which was immobilized through the ring of ∼10 c subunits, suggesting that the rotor and stator are interchangeable mechanical units (14). ATP synthesis by F-ATPase is believed to follow the reverse mechanism of ATP hydrolysis because mechanically induced rotation of the γ subunit in immobilized F1 in the presence of ADP and Pi results in net ATP synthesis (15, 16). There remain many questions about the mechanism of coupling between catalysis and transport via mechanical rotation. In particular, the mechanism of coupling H+ transport to rotation of the subunit c10 ring is still not well understood (4).In contrast, there is considerably more information on the mechanism of coupling catalysis to γ and ϵ subunit rotation. Observations of γ subunit rotation in the catalytic F1 sector are consistent with Boyer''s binding change model (17); thus coupling between the chemistry and rotation can be assessed by studies of the soluble F1, and these findings relate to the mechanism of the entire ATP synthase complex. The γ subunit rotates relative to the α3β3 hexamer in distinct 120° steps. A 120° rotation step consisting of pause and rotation substeps appears to correspond to the hydrolysis of one ATP, assuming that three ATP molecules are hydrolyzed per 360° revolution (18). Additional pauses observed at low ATP concentrations are attributed to the “ATP waiting” dwell (19). Yasuda et al. (19) and Shimabukuro et al. (20) further resolved that each 120° step occurred in two substeps: an 80° substep whose onset was dependent upon the Mg·ATP concentration, and a 40° substep, which was not affected by substrate concentration (19). The pause before the 80° substep, the ATP waiting dwell became shorter with increasing [Mg·ATP]. In contrast, the pause duration before the 40° rotation step was modulated by the slow hydrolysis rate of ATPγS2 or by the catalytic site mutant βE190D (in the Bacillus PS3 F1), which was found to significantly increase the length of the catalytic dwell (20). These data together indicate that the dwell before the 40° step is the “catalytic dwell” (20) and defines the order of the substeps during the 120° rotation step observed in high Mg·ATP concentrations (21).In this paper, we address the question of when the rate-limiting step of steady state catalysis occurs, with respect to the rotational behavior. Pre-steady state analysis of the burst kinetics of ATP hydrolysis at nearly Vmax conditions demonstrated that the rate-limiting transition state occurs after the reversible hydrolysis/synthesis step and before release of phosphate (Pi) (22, 23). The rate-limiting step is likely associated with a rotation step because a γ-β cross-linked enzyme is still able to undergo the initial ATP hydrolysis, but the rotation-impeded enzyme is unable to release Pi (23). Significantly, the kinetics of steady state hydrolysis can only be assessed when the Mg·ATP concentration is high enough to fill all three catalytic sites. The only model consistent with these data is one that involves all three catalytic sites. During each 120° catalytic cycle, one site binds ATP, a different site carries out reversible hydrolysis/synthesis, and the third site releases product Pi and ADP (22, 23).Steady state analyses, which take advantage of a particular γ subunit mutation γM23K (24), are consistent with this model. Replacement of the conserved γMet-23 with lysine causes an uncoupling between catalysis and γ subunit rotation caused by altered interactions between γ and β subunits (25). Importantly, Al-Shawi and Nakamoto (26) and Al-Shawi et al. (25, 27) found that the γM23K mutation strongly affected the rate-limiting transition state of steady state ATP hydrolysis and ATP synthesis. The slope of the Arrhenius plots and thus the energy of activation were significantly increased in the mutant enzyme. Several second site suppressor mutations, mostly in the γ subunit (28, 29) but also in the β subunits (30, 31), were genetically identified because they restored coupled ATP synthesis. Significantly, all were in the γ-β interface. Thermodynamic analyses found that the second site suppressors generally compensated for the primary γM23K mutations by reducing the increased activation energy (25, 27, 31). Although most of the second site mutations were found distant from the γM23K site, the x-ray crystal structures (7) suggested that γM23K may directly interact with conserved βGlu-381. As expected, replacement of βGlu-381 with aspartate also suppressed the uncoupling effects of γM23K (31).To identify the rate-limiting transition state step in the rotational behavior, we analyzed the temperature dependence of the γM23K mutant in Vmax conditions observed in single molecule experiments. Interestingly, direct observation of this mutant using the micron-length actin filaments did not detect differences in the rotation behavior at room temperature (9). In contrast, we find in the data presented here that there is dramatic effect of the mutation on the temperature dependence of the length of the catalytic dwell or pause between the 120° rotation steps. This is likely because of two factors: first, we used a bead small enough not to invoke a drag on the rotation (32), and second, the temperature dependence of the rate of the rotation steps is critical for the analyses of the mechanism.  相似文献   

15.
The C-terminal domain of subunit ε of the bacterial FoF1 ATP synthase is reported to be an intrinsic inhibitor of ATP synthesis/hydrolysis activity in vitro, preventing wasteful hydrolysis of ATP under low-energy conditions. Mutants defective in this regulatory domain exhibited no significant difference in growth rate, molar growth yield, membrane potential, or intracellular ATP concentration under a wide range of growth conditions and stressors compared to wild-type cells, suggesting this inhibitory domain is dispensable for growth and survival of Escherichia coli.FoF1 ATP synthases are ubiquitous enzymes that synthesize ATP using a transmembrane electrochemical potential of protons or proton motive force (PMF) generated by the respiratory chain across the cytoplasmic membrane of bacteria, the thylakoid membrane of chloroplasts, or the mitochondrial inner membrane (4, 5, 37). The enzyme consists of two parts: membrane-embedded Fo subcomplex (a complex of subunits a, b, and c in bacteria) and hydrophilic F1 subcomplex (composed of subunits α, β, γ, δ, and ε). The enzyme is also known as a molecular motor, which is composed of the stator subcomplex (α, β, δ, a, and b) and the rotor subcomplex (γ, ε, and c), and its rotation is coupled to ATP synthesis and proton flow across the membrane (20, 31, 52). The reaction of the enzyme is reversible; ATP is hydrolyzed into ADP and inorganic phosphate, the rotor subcomplex rotates in reverse, and protons are extruded to the periplasmic side, resulting in the generation of PMF. Although some bacteria utilize the reverse reaction under particular conditions, the primary function of FoF1 ATP synthase is generation of ATP from the PMF. Therefore, the direction of the activity of FoF1 ATP synthase is regulated to avoid wasteful ATP hydrolysis.Subunit ε in bacterial FoF1 has been known to be an intrinsic inhibitor of F1 and FoF1 complex (18, 21, 23) and is proposed to have a regulatory function (10, 11, 42). Although the inhibitory effects of subunit ε vary among species, in general, ε inhibits ATP hydrolysis activity while repressing ATP synthesis activity to a lesser degree (14, 27). This regulatory function of the ε subunit is mediated almost exclusively by the C-terminal region of ε, which is comprised of two antiparallel α-helices (18, 49, 50). Biochemical and crystallographic studies have revealed that the C-terminal helices can adopt two different conformations (34, 46, 47, 48). In the retracted conformation, the α-helices form a hairpin-like structure and sit on the N-terminal β-sandwich domain of the ε subunit. When the ε subunit exhibits an inhibitory effect, it adopts a more extended conformation in which the C-terminal α-helices extend along the γ subunit, which composes the central stalk. It has also been shown that basic, positively charged residues on the second α-helix of the ε subunit interact with negatively charged residues in the DELSEED segment of subunit β to exert the inhibitory effect (12).Escherichia coli mutants deleted in the entire ε subunit exhibit a reduced growth rate and growth yield, and this effect is proposed to be a result of a deficiency in assembly of the Fo and F1 complexes (21). The N-terminal β-sandwich domain of the ε subunit is responsible for the assembly of Fo and F1 and is therefore important for efficient coupling between proton translocation through Fo and ATP synthesis/hydrolysis in F1 (15, 39). Deletion of the ε subunit leads to dissociation of the FoF1 complex and wasteful ATP hydrolysis by free (cytoplasmic) F1 and dissipation of PMF through free Fo (21, 22, 51).While the importance of the entire ε subunit in the whole-cell physiology of E. coli is fairly well established, the role of the regulatory C-terminal region of ε has received little attention and warrants investigation to determine if the regulatory functions (e.g., inhibition of ATP hydrolysis) observed in vitro are manifested in the physiology of E. coli under various growth conditions. To address this question, we constructed isogenic E. coli mutants that were deleted in the C-terminal region of ε subunit (εDC) and used these strains to compare physiological properties of wild-type versus εDC cells under a wide range of environmental conditions and stressors.  相似文献   

16.
17.
Biogenesis of mitochondrial cytochrome c oxidase (COX) relies on a large number of assembly factors, among them the transmembrane protein Surf1. The loss of human Surf1 function is associated with Leigh syndrome, a fatal neurodegenerative disorder caused by severe COX deficiency. In the bacterium Paracoccus denitrificans, two homologous proteins, Surf1c and Surf1q, were identified, which we characterize in the present study. When coexpressed in Escherichia coli together with enzymes for heme a synthesis, the bacterial Surf1 proteins bind heme a in vivo. Using redox difference spectroscopy and isothermal titration calorimetry, the binding of the heme cofactor to purified apo-Surf1c and apo-Surf1q is quantified: Each of the Paracoccus proteins binds heme a in a 1:1 stoichiometry and with Kd values in the submicromolar range. In addition, we identify a conserved histidine as a residue crucial for heme binding. Contrary to most earlier concepts, these data support a direct role of Surf1 in heme a cofactor insertion into COX subunit I by providing a protein-bound heme a pool.Leigh syndrome (LS)3 is an autosomal recessive inherited neurodegenerative disorder characterized by focal, bilateral lesions in one or more areas of the central nervous system (1). Symptoms start in early childhood, and the disease usually progresses rapidly. Although mutations in various mitochondrial enzymes can result in LS, its most frequent trigger is deficiency of cytochrome c oxidase (COX) caused by mutations in the SURF1 gene, as identified in LS patients (2, 3). Human SURF1, the first gene of the SURFEIT gene locus on chromosome 9, encodes a 30-kDa protein related to COX assembly (2, 3).Mitochondrial COX consists of up to 13 subunits (SU). The three core SU encoded by the mitochondrial genome carry all of the redox-active cofactors, two heme a moieties, and three copper ions. These three SU are highly conserved among different organisms and represent the main components of bacterial oxidase complexes as well (4, 5). The assembly process of mitochondrial COX is only marginally understood, involving the interplay of a large number of auxiliary proteins (69).Despite intensive efforts over more than a decade to unravel Surf1 function, its exact role in COX assembly still remains unclear. Surf1 is not strictly essential for COX assembly because patients with LS have residuals of assembled oxidase with remaining activity of approximately 10–20% in all tissues (2, 3). Located in the inner mitochondrial membrane, Surf1 is predicted to form two transmembrane helices connected by a long loop facing the intermembrane space (10, 11). Sequence alignments confirm the presence of Surf1 homologs in many eukaryotes and prokaryotes (12).One of the best studied Surf1 proteins is the yeast homolog Shy1p, which has been discovered and characterized in the context of pet mutants (10). Deletion of the gene leads to a strongly decreased COX level, although the residual enzyme appears fully functional. This points to a role of Shy1p in assembly or stabilization of COX (13), most likely during the formation of an early assembly intermediate consisting of the highly conserved core SU I and II (14).So far, only three bacterial homologs have been inspected in closer detail (15, 16). In Paracoccus denitrificans, two Surf1 homologs were identified and named Surf1c and Surf1q for their specific role in serving a heme aa3-type COX and a related heme ba3-type quinol oxidase, respectively (15). With the function of Surf1 in COX assembly still being speculative, a role in heme a insertion into COX SU I seemed conceivable (15, 16).Here we show that P. denitrificans Surf1c and Surf1q are able to bind heme a both in vivo and in vitro. This novel finding suggests that Surf1 proteins promote heme a insertion into SU I of either cytochrome c oxidase or quinol oxidase. In addition, Surf1 may modulate heme a synthase activity and provide a heme a cofactor pool in a safe, chelated form for COX SU I biogenesis.  相似文献   

18.
19.
Cytochrome c (cyt c) is a heme-containing protein that participates in electron transport in the respiratory chain and as a signaling molecule in the apoptotic cascade. Here we addressed the effect of removing mammalian cyt c on the integrity of the respiratory complexes in mammalian cells. Mitochondria from cyt c knockout mouse cells lacked fully assembled complexes I and IV and had reduced levels of complex III. A redox-deficient mutant of cyt c was unable to rescue the levels of complexes I and IV. We found that cyt c is associated with both complex IV and respiratory supercomplexes, providing a potential mechanism for the requirement for cyt c in the assembly/stability of complex IV.The mitochondrial electron transport chain consists of four multisubunit complexes, namely, NADH-ubiquinone oxidoreductase (complex I),2 succinate-ubiquinone oxidoreductase (complex II), ubiquinone-cytochrome c oxidoreductase (complex III), and cytochrome c oxidase (complex IV, COX). Cytochrome c (cyt c) shuttles electrons from oxidative phosphorylation complex III to complex IV. Electrons are transferred from reduced cyt c sequentially to the CuA site, heme a, heme a3, and CuB binuclear center in the complex IV before being finally transferred to molecular oxygen to generate water (1). Respiratory complexes are assembled into supercomplexes (also called respirasomes). These contain complex I bound to dimeric complex III and a variable copy number of complex IV (2).In Saccharomyces cerevisiae, cyt c is encoded by two genes: CYC1 and CYC7. Mutagenesis studies in yeast have shown that cyt c is required for the assembly of COX (3, 4). In yeast lacking both the cyt c genes (CYC1 and CYC7), COX assembly was absent. It was also shown that cyt c is only structurally required for COX assembly, because a catalytic mutant of cyt c (W65S) was sufficient to bring about near normal levels of COX. However, because yeast lacks complex I, they could not analyze the role of cyt c in the assembly/stability of complex I. Mammals possess two different isoforms of cyt c encoded on different chromosomes: the somatic (cyt cS)- and testis (cyt cT)-specific isoforms. In mouse, the cDNAs bear 74% homology, whereas the proteins possess 86% identity with most dissimilarity in the C terminus.Cardiolipin (CL) is an anionic phospholipid present almost exclusively in the mitochondrial membranes and constitutes 25% of its total phospholipids (5). Work from several laboratories showed that CL is essential for the membrane anchorage of the respiratory supercomplexes. CL has two main roles in the mitochondrial structure and function, namely, stabilization of mitochondrial membranes and specific interactions with proteins. CL deficiency results in inefficient energy transformation by oxidative phosphorylation, swelling of mitochondria, decreased ATP/oxygen ratio, and reduced membrane potential (6, 7). In accordance, in S. cerevisiae lacking CL synthase, the supercomplex comprising complexes III and IV is unstable (8). Assembly mutants of COX had significantly reduced CL synthase activity, whereas assembly mutants of respiratory complex III and complex V showed less inhibition (9). Subsequently, the proton gradient across the inner mitochondrial membrane was found to be important for CL formation and that CL synthase was stimulated by alkaline pH at the matrix side (10). In this study, we investigated the role of cyt c depletion on CL levels by examining its content and composition in cyt c null cells.Here we aimed to answer the following questions: What is the role of cyt c in the assembly and maintenance of the different respiratory complexes in mammals? Are there changes in the content/composition of lipids in the cyt c-ablated cells? Analysis of mouse fibroblasts revealed that cyt c is essential for the assembly/stability of COX, and a catalytically mutant form of cyt c cannot rescue the COX defect in the cyt c null cells. CL and triacylglycerols showed significant differences in the cyt c null cells, both in content and composition.  相似文献   

20.
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