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.
2.
3.
4.
5.
Superoxide dismutase 2 (SOD2) is one of the rare mitochondrial enzymes evolved to use manganese as a cofactor over the more abundant element iron. Although mitochondrial iron does not normally bind SOD2, iron will misincorporate into Saccharomyces cerevisiae Sod2p when cells are starved for manganese or when mitochondrial iron homeostasis is disrupted by mutations in yeast grx5, ssq1, and mtm1. We report here that such changes in mitochondrial manganese and iron similarly affect cofactor selection in a heterologously expressed Escherichia coli Mn-SOD, but not a highly homologous Fe-SOD. By x-ray absorption near edge structure and extended x-ray absorption fine structure analyses of isolated mitochondria, we find that misincorporation of iron into yeast Sod2p does not correlate with significant changes in the average oxidation state or coordination chemistry of bulk mitochondrial iron. Instead, small changes in mitochondrial iron are likely to promote iron-SOD2 interactions. Iron binds Sod2p in yeast mutants blocking late stages of iron-sulfur cluster biogenesis (grx5, ssq1, and atm1), but not in mutants defective in the upstream Isu proteins that serve as scaffolds for iron-sulfur biosynthesis. In fact, we observed a requirement for the Isu proteins in iron inactivation of yeast Sod2p. Sod2p activity was restored in mtm1 and grx5 mutants by depleting cells of Isu proteins or using a dominant negative Isu1p predicted to stabilize iron binding to Isu1p. In all cases where disruptions in iron homeostasis inactivated Sod2p, we observed an increase in mitochondrial Isu proteins. These studies indicate that the Isu proteins and the iron-sulfur pathway can donate iron to Sod2p.Metal-containing enzymes are generally quite specific for their cognate cofactor. Misincorporation of the wrong metal ion can be deleterious and tends to be a rare occurrence in biology. A prime example of metal ion selectivity is illustrated by the family of manganese- and iron-containing superoxide dismutases (SODs)3. This large family of enzymes utilizes either manganese or iron as cofactors to scavenge superoxide anion. The iron- and manganese-containing forms are highly homologous to one another at primary, secondary, and tertiary levels and have virtually identical metal binding and catalytic sites (13). Despite this extensive homology, Mn- and Fe-SODs are only active with their cognate metal. Misincorporation of iron into Mn-SOD or vice versa alters the redox potential of the enzyme''s active site and prohibits superoxide disproportionation (4, 5). Nevertheless, misincorporation of iron into Mn-SOD does occur in vivo (6, 7). The isolated Mn-SOD from Escherichia coli is found as a mixture of manganese- and iron-bound forms (7); binding of manganese is favored under oxidative stress, whereas iron binding is increased under anaerobic conditions (3, 8). It has been proposed that changes in bioavailability of manganese versus iron determine the metal selectivity of Mn-SOD in bacterial cells (3, 8). But is this also true for Fe-SOD? Currently, there is no documentation of manganese misincorporation into Fe-SOD in vivo.Unlike bacteria that co-express Mn- and Fe-SOD molecules in the same cell, eukaryotic mitochondria generally harbor only one member of the Fe/Mn-SOD family, a tetrameric Mn-SOD typically known as SOD2 (9). In some organisms, SOD2 is essential for survival (1012), and mitochondria have therefore evolved to prevent iron-SOD2 interactions despite high levels of mitochondrial iron relative to manganese. Using a yeast model system, we have shown previously that metal ion mis-incorporation can occur with Saccharomyces cerevisiae Sod2p (7). Specifically, iron binds and inactivates yeast Sod2p when cells are either starved for manganese or have certain disruptions in mitochondrial iron homeostasis. These disruptions include mutations in MTM1, a mitochondrial carrier protein that functions in iron metabolism (7, 13), and mutations in GRX5 or SSQ1, involved in iron-sulfur biogenesis (14). We proposed that these disruptions lead to expansion of a mitochondrial pool of so-called SOD2-reactive iron (7). Currently, it is unknown whether SOD2-reactive iron represents a major shift in the chemistry of bulk mitochondrial iron or whether it is just a small pool of the metal emerging from one or more specific sites.The grx5 and ssq1 mutants that promote iron-SOD2 interactions encode just two of many components of a complex pathway for iron-sulfur biogenesis (15, 16). One of the key components is a well conserved iron-sulfur scaffold protein originally described for bacteria as IscU, also known as mammalian ISCU and S. cerevisiae Isu1p and Isu2p, referred collectively herein as “Isu proteins” (1722). The iron-sulfur clusters on Isu proteins are labile and can be transferred to target iron-sulfur proteins through the aid of mitochondrial factors including Grx5p and Ssq1p (15, 16). It is not clear whether disruption of the iron-sulfur pathway per se is sufficient to promote iron interactions with yeast Sod2p or whether this effect is specific to grx5, ssq1, and mtm1 mutants.In the current study, we explore the nature of mitochondrial iron that can interact with Sod2p. We find that the changes in mitochondrial metal homeostasis that shift metal binding in yeast Sod2p likewise alter metal cofactor selection in a heterologously expressed Mn-SOD, but not in a Fe-SOD molecule. Through x-ray absorption near edge structure (XANES) and extended x-ray absorption fine structure (EXAFS) analyses of mitochondrial iron, we detected no major change in bulk mitochondrial iron under conditions that promote iron-SOD2 interactions. SOD2-reactive iron appears to represent a small pool of the metal, and we provide evidence that the iron-sulfur scaffold Isu1p can act as an important source of this reactive iron.  相似文献   

6.
Iron is a crucial cofactor in numerous redox-active proteins operating in bioenergetic pathways including respiration and photosynthesis. Cellular iron management is essential to sustain sufficient energy production and minimize oxidative stress. To produce energy for cell growth, the green alga Chlamydomonas reinhardtii possesses the metabolic flexibility to use light and/or carbon sources such as acetate. To investigate the interplay between the iron-deficiency response and growth requirements under distinct trophic conditions, we took a quantitative proteomics approach coupled to innovative hierarchical clustering using different “distance-linkage combinations” and random noise injection. Protein co-expression analyses of the combined data sets revealed insights into cellular responses governing acclimation to iron deprivation and regulation associated with photosynthesis dependent growth. Photoautotrophic growth requirements as well as the iron deficiency induced specific metabolic enzymes and stress related proteins, and yet differences in the set of induced enzymes, proteases, and redox-related polypeptides were evident, implying the establishment of distinct response networks under the different conditions. Moreover, our data clearly support the notion that the iron deficiency response includes a hierarchy for iron allocation within organelles in C. reinhardtii. Importantly, deletion of a bifunctional alcohol and acetaldehyde dehydrogenase (ADH1), which is induced under low iron based on the proteomic data, attenuates the remodeling of the photosynthetic machinery in response to iron deficiency, and at the same time stimulates expression of stress-related proteins such as NDA2, LHCSR3, and PGRL1. This finding provides evidence that the coordinated regulation of bioenergetics pathways and iron deficiency response is sensitive to the cellular and chloroplast metabolic and/or redox status, consistent with systems approach data.The green alga Chlamydomonas reinhardtii has an enormous metabolic versatility (1) and possesses the flexibility to grow in the presence of different carbon sources. It may use carbon dioxide (CO2) for photoautotrophic, acetate for heterotrophic, and both carbon sources for mixotrophic growth. In this alga CO2 is fixed via the Calvin Benson Bassham cycle (2), while acetate can be taken up, converted to acetyl-CoA, and enter the glyoxylate cycle where it may be incorporated into C4 acids (3). In addition to the use of acetate as a source of energy and carbon backbone for biosynthetic processes, acetate can control respiration and photosynthesis in conjunction with the light intensity and CO2 availability (46). Moreover, acclimation responses to iron- and copper-deficiencies significantly vary in photoautotrophic versus heterotrophic conditions (710), indicating that the metabolic status of the cells influence overall cellular acclimation responses.Transition metals like copper, manganese, and iron possess the ability to donate and accept electrons, making these metals suitable cofactors in enzymes that catalyze redox reactions. In particular, iron is used as a cofactor in numerous biochemical pathways and is therefore an essential nutrient. Cells require relatively high levels of iron because it is present in heme-, iron-sulfur and other proteins that function in respiratory and photosynthetic energy transducing. Correspondingly, in eukaryotic cells, the mitochondrion is a major iron-utilizing compartment. It is well established that iron is transported into mitochondria for heme synthesis and iron-sulfur cluster assembly. This is required for the formation of a functional respiratory electron transport machinery (11). Therefore, mitochondrial metabolism in mammals, fungi and plants is significantly affected under iron deficiency, as demonstrated by a number of studies (1214). In plants, the chloroplasts are a primary target of iron deficiency. Changes in chloroplast structure, photosynthetic capacity and the composition of thylakoid membranes have been described for plants deprived of iron (1521).Plants have devised various strategies for acquiring iron (22). Generally, iron deficiency leads to the activation of the iron uptake systems in photosynthetic organisms. For example, the accumulation of the ferroxidase, a component of the high affinity iron uptake system in C. reinhardtii, is very rapidly enhanced when iron becomes limiting (23). Inactivation of IRT1, the most prevalent Fe2+ transporter in Arabidopsis thaliana leads to a dramatic iron deficiency that is reflected by chlorosis (2426). Despite the evolution of elaborate iron-uptake mechanisms in plants, iron deficiency-induced chlorosis remains a major agricultural problem (27, 28).The global impact of iron deficiency on photosynthetic productivity has been also shown in vast ocean regions, which are severely limited for iron (29, 30). Generally, one can conclude that photosynthesis in the oceans and on land can occur in environment where iron availability is restricted.Photosystem I (PSI) is a prime target of iron deficiency as it contains 12 atoms of iron per core complex. In algae, the degradation of PSI is also linked to remodeling of PSI-associated light-harvesting antenna (LHCI) (3133). Cyanobacteria respond to iron deficiency by degradation of light harvesting phycobilisomes (34) and induction of the “iron-stress-induced” gene isiA. The ISIA protein, which has significant sequence similarity with CP43, a chlorophyll a-binding protein of photosystem II (PSII; (35, 36), forms a ring of 18 molecules around a PSI trimeric reaction center, as shown by electron microscopy (37, 38). The overall reorganization of the PSI complex from 900 kDa into 1.7 MDa complex highlights the large adaptive nature of the cellular response to iron deficiency, which helps to optimize the architecture of the photosynthetic apparatus to conditions in which iron is a limiting factor.The marine diatom Thalassiosira oceanica shows a remarkable retrenchment of cellular metabolism and remodeling of bioenergetic pathways in response to iron availability (39). Low iron triggers a reduction in the level of iron-rich photosynthetic proteins while iron-rich mitochondrial proteins are preserved. Furthermore, iron deprivation causes a remodeling of the photosynthetic machinery resulting in the adjustment of light energy use to an overall decline in the level of photosynthetic electron transport complexes (39). These responses, reported for green algae such as C. reinhardtii (31, 40, 41), are important for minimizing photo-oxidative stress and optimizing photosynthetic function. As observed for T. oceanica, under conditions of low iron availability (in the presence of organic carbon) a hierarchy of iron allocation responses in C. reinhardtii result in the down-regulation of iron-rich photosynthetic complexes while iron-rich mitochondrial complexes remain stable (41). Notably, under photoautotrophic and mixotrophic conditions C. reinhardtii displays distinct iron deprivation responses, suggesting that the cell''s response to iron deficiency is also dependent on trophic conditions (79). Thus bioenergetics pathways are remodeled in response to iron availability as well as to the type of carbon source available. Moreover, recent data has indicated that the regulation of iron-induced remodeling of the photosynthetic apparatus is linked to energy metabolism. Depletion of Proton Gradient Regulation Like1 protein (PGRL1) in C. reinhardtii has revealed a decreased efficiency of cyclic electron transfer under low iron conditions resulting in higher vulnerability toward iron deprivation (42).It was our aim to generate a more comprehensive picture of how the proteome of C. reinhardtii varies in response to low iron under distinct trophic conditions and how these changes compare with differences observed for cells grown under photoautotrophic and photoheterotrophic iron replete conditions. Quantitative proteomics in conjunction with a novel hierarchical clustering approach revealed information about the responses of C. reinhardtii to low iron conditions and the iron requirements of photoautotrophic growth. These analyses provide novel insights into the relationships between protein networks required for photosynthesis and iron deprivation-elicited stress responses; these studies are providing the knowledge required for modulating the level of available iron to improve the photosynthetic performance of plants (43, 44).  相似文献   

7.
8.
Two methods were used to compare the biodegradation of six polychlorinated biphenyl (PCB) congeners by 12 white rot fungi. Four fungi were found to be more active than Phanerochaete chrysosporium ATCC 24725. Biodegradation of the following congeners was monitored by gas chromatography: 2,3-dichlorobiphenyl, 4,4′-dichlorobiphenyl, 2,4′,5-trichlorobiphenyl (2,4′,5-TCB), 2,2′,4,4′-tetrachlorobiphenyl, 2,2′,5,5′-tetrachlorobiphenyl, and 2,2′,4,4′,5,5′-hexachlorobiphenyl. The congener tested for mineralization was 2,4′,5-[U-14C]TCB. Culture supernatants were also assayed for lignin peroxidase and manganese peroxidase activities. Of the fungi tested, two strains of Bjerkandera adusta (UAMH 8258 and UAMH 7308), one strain of Pleurotus ostreatus (UAMH 7964), and Trametes versicolor UAMH 8272 gave the highest biodegradation and mineralization. P. chrysosporium ATCC 24725, a strain frequently used in studies of PCB degradation, gave the lowest mineralization and biodegradation activities of the 12 fungi reported here. Low but detectable levels of lignin peroxidase and manganese peroxidase activity were present in culture supernatants, but no correlation was observed among any combination of PCB congener biodegradation, mineralization, and lignin peroxidase or manganese peroxidase activity. With the exception of P. chrysosporium, congener loss ranged from 40 to 96%; however, these values varied due to nonspecific congener binding to fungal biomass and glassware. Mineralization was much lower, ≤11%, because it measures a complete oxidation of at least part of the congener molecule but the results were more consistent and therefore more reliable in assessment of PCB biodegradation.

Polychlorinated biphenyls (PCBs) are produced by chlorination of biphenyl, resulting in up to 209 different congeners. Commercial mixtures range from light oily fluids to waxes, and their physical properties make them useful as heat transfer fluids, hydraulic fluids, solvent extenders, plasticizers, flame retardants, organic diluents, and dielectric fluids (1, 21). Approximately 24 million lb are in the North American environment (19). The stability and hydrophobic nature of these compounds make them a persistent environmental hazard.To date, bacterial transformations have been the main focus of PCB degradation research. Aerobic bacteria use a biphenyl-induced dioxygenase enzyme system to attack less-chlorinated congeners (mono- to hexachlorobiphenyls) (1, 5, 7, 8, 22). Although more-chlorinated congeners are recalcitrant to aerobic bacterial degradation, microorganisms in anaerobic river sediments reductively dechlorinate these compounds, mainly removing the meta and para chlorines (1, 6, 10, 33, 34).The degradation of PCBs by white rot fungi has been known since 1985 (11, 18). Many fungi have been tested for their ability to degrade PCBs, including the white rot fungi Coriolus versicolor (18), Coriolopsis polysona (41), Funalia gallica (18), Hirneola nigricans (35), Lentinus edodes (35), Phanerochaete chrysosporium (3, 11, 14, 17, 18, 35, 39, 4143), Phlebia brevispora (18), Pleurotus ostreatus (35, 43), Poria cinerescens (18), Px strain (possibly Lentinus tigrinus) (35), and Trametes versicolor (41, 43). There have also been studies of PCB metabolism by ectomycorrhizal fungi (17) and other fungi such as Aspergillus flavus (32), Aspergillus niger (15), Aureobasidium pullulans (18), Candida boidinii (35), Candida lipolytica (35), Cunninghamella elegans (16), and Saccharomyces cerevisiae (18, 38). The mechanism of PCB biodegradation has not been definitively determined for any fungi. White rot fungi produce several nonspecific extracellular enzymes which have been the subject of extensive research. These nonspecific peroxidases are normally involved in lignin degradation but can oxidize a wide range of aromatic compounds including polycyclic aromatic hydrocarbons (37). Two peroxidases, lignin peroxidase (LiP) and Mn peroxidase (MnP), are secreted into the environment of the fungus under conditions of nitrogen limitation in P. chrysosporium (23, 25, 27, 29) but are not stress related in fungi such as Bjerkandera adusta or T. versicolor (12, 30).Two approaches have been used to determine the biodegradability of PCBs by fungi: (i) loss of the parent congener analyzed by gas chromatography (GC) (17, 32, 35, 42, 43) and (ii) mineralization experiments in which the 14C of the universally labeled 14C parent congener is recovered as 14CO2 (11, 14, 18, 39, 41). In the first method, the loss of a peak on a chromatogram makes it difficult to decide whether the PCB is being partly degraded, mineralized, adsorbed to the fungal biomass, or bound to glassware, soil particles, or wood chips. Even when experiments with killed-cell and abiotic controls are performed, the extraction efficiency and standard error can make data difficult to interpret. For example, recoveries can range anywhere from 40 to 100% depending on the congener used and the fungus being investigated (17). On the other hand, recovery of significant amounts of 14CO2 from the cultures incubated with a 14C substrate provides definitive proof of fungal metabolism. There appears to be only one report relating data from these two techniques (18), and in that study, [U-14C]Aroclor 1254, rather than an individual congener, was used.In this study, we examined the ability of 12 white rot fungal strains to metabolize selected PCB congeners to determine which strains were the most active degraders. Included in this group was P. chrysosporium ATCC 24725, a strain used extensively in PCB studies (3, 14, 18, 35, 39, 4143). Six PCB congeners were selected to give a range of chlorine substitutions and therefore a range of potential biodegradability which was monitored by GC. One of the chosen congeners was 14C labeled and used in studies to compare the results from a mineralization method with those from the GC method.  相似文献   

9.
So-called sulfur-turf microbial mats, which are macroscopic white filaments or bundles consisting of large sausage-shaped bacteria and elemental sulfur particles, occur in sulfide-containing hot springs in Japan. However, no thermophiles from sulfur-turf mats have yet been isolated as cultivable strains. This study was undertaken to determine the phylogenetic positions of the sausage-shaped bacteria in sulfur-turf mats by direct cloning and sequencing of 16S rRNA genes amplified from the bulk DNAs of the mats. Common clones with 16S rDNA sequences with similarity levels of 94.8 to 99% were isolated from sulfur-turf mat samples from two geographically remote hot springs. Phylogenetic analysis showed that the phylotypes of the common clones formed a major cluster with members of the Aquifex-Hydrogenobacter complex, which represents the most deeply branching lineage of the domain bacteria. Furthermore, the bacteria of the sulfur-turf mat phylotypes formed a clade distinguishable from that of other members of the Aquifex-Hydrogenobacter complex at the order or subclass level. In situ hybridization with clone-specific probes for 16S rRNA revealed that the common phylotype of sulfur-turf mat bacteria is that of the predominant sausage-shaped bacteria.Microbial mats develop in a wide variety of aquatic environments, including geothermal hot springs and hydrothermal vents. There are several types of thermophilic microbial mats, e.g., those of cyanobacteria, anoxygenic phototrophic bacteria, and chemotrophic sulfur bacteria, which differ according to the physical and chemical conditions they favor and other environmental factors (10, 38). These microbial mats in thermal habitats have been studied extensively as a peculiar microbial community of the ecosystem, in relation to the phylogeny and evolution of thermophilic prokaryotes, or as a source of new functional enzymes.So-called sulfur-turf microbial mats are macroscopic bundles of white filaments consisting of colorless sulfur bacteria and elemental sulfur particles that form in shallow streams of sulfide-containing high-temperature hot springs. Since first reported by Miyoshi in 1897 (33), this kind of microbial mat has been recorded for several geographically remote hot springs in Japan, although there have been only scattered reports of sulfur-turf microbial mats or chemotrophic sulfur streamers in geothermal springs in other countries (9, 13, 14). The sulfur-turf mats generally develop within a temperature range of 45 to 73°C, within a pH range of 6 to 9, and at discrete sulfide-oxygen interfaces in geothermal springs. These characteristics suggest that the major constituents of the sulfur-turf prokaryotic community are (hyper)thermophilic, neutrophilic, microaerophilic, and chemolithotrophic bacteria. Early studies of these sulfur-turf mats distinguished microscopically three morphotypes of bacteria, two of which were tentatively named Thiovibrio miyoshi and Thiothrix miyoshi (15). Moreover, in situ ecophysiological and microscopic studies have shown that one of these bacteria, the large sausage-shaped “Thiovibrio miyoshi,” predominates in sulfur-turf mats and oxidizes environmental sulfide to elemental sulfur and then to sulfate via thiosulfate (2731). So far, however, it has not been possible to isolate and cultivate any thermophilic prokaryotes from the sulfur-turf mats predominated by these sausage-shaped bacteria with artificial media, and no attempt has been made to clarify their taxonomic and phylogenetic positions.Determination of 16S rRNA genes is a useful research strategy for identifying uncultivated prokaryotes and is now commonly performed in ecological studies. This technique, involving PCR amplification of 16S rRNA genes or synthesis of cDNAs from bulk 16S rRNAs of natural mixed microbial populations, has been used successfully for the phylogenetic characterization of prokaryotes in hydrothermal environments (6, 7, 34, 40, 41, 47, 48). In the present study, this approach was applied to characterize the sausage-shaped bacteria in sulfur-turf mats without isolating and cultivating them. Here we report that sulfur-turf mats contain novel thermophilic bacteria belonging to the earliest-branching lineage of the domain bacteria.  相似文献   

10.
Mathematical tools developed in the context of Shannon information theory were used to analyze the meaning of the BLOSUM score, which was split into three components termed as the BLOSUM spectrum (or BLOSpectrum). These relate respectively to the sequence convergence (the stochastic similarity of the two protein sequences), to the background frequency divergence (typicality of the amino acid probability distribution in each sequence), and to the target frequency divergence (compliance of the amino acid variations between the two sequences to the protein model implicit in the BLOCKS database). This treatment sharpens the protein sequence comparison, providing a rationale for the biological significance of the obtained score, and helps to identify weakly related sequences. Moreover, the BLOSpectrum can guide the choice of the most appropriate scoring matrix, tailoring it to the evolutionary divergence associated with the two sequences, or indicate if a compositionally adjusted matrix could perform better.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29]  相似文献   

11.
12.
Thermus thermophilus HB27 is an extremely thermophilic eubacteria with a high frequency of natural competence. This organism is therefore often used as a thermophilic model to investigate the molecular basis of type IV pili–mediated functions, such as the uptake of free DNA, adhesion, twitching motility, and biofilm formation, in hot environments. In this study, the phosphoproteome of T. thermophilus HB27 was analyzed via a shotgun approach and high-accuracy mass spectrometry. Ninety-three unique phosphopeptides, including 67 in vivo phosphorylated sites on 53 phosphoproteins, were identified. The distribution of Ser/Thr/Tyr phosphorylation sites was 57%/36%/7%. The phosphoproteins were mostly involved in central metabolic pathways and protein/cell envelope biosynthesis. According to this analysis, the ATPase motor PilF, a type IV pili–related component, was first found to be phosphorylated on Thr-368 and Ser-372. Through the point mutation of PilF, mimic phosphorylated mutants T368D and S372E resulted in nonpiliated and nontwitching phenotypes, whereas nonphosphorylated mutants T368V and S372A displayed piliation and twitching motility. In addition, mimic phosphorylated mutants showed elevated biofilm-forming abilities with a higher initial attachment rate, caused by increasing exopolysaccharide production. In summary, the phosphorylation of PilF might regulate the pili and biofilm formation associated with exopolysaccharide production.Thermus thermophilus HB27 is a Gram-negative, rod-shaped, and extremely thermophilic eubacterium isolated from a geothermal area (1). This organism grows at temperatures up to 85 °C and has an optimal growth temperature of 70 °C. The thermostable enzymes obtained from members of the genus Thermus are of considerable interest because of their potential in research, biotechnological, and industrial applications (2, 3). In addition, T. thermophilus HB27 is a suitable laboratory model for genetic manipulation, as it is easily cultured under laboratory conditions and has a natural transformation system that is much more efficient than those of other Thermus spp. (4). Intriguingly, thermophiles are also found in biofilms, enclosed within a matrix consisting of extracellular polymeric substances, in various natural and artificial thermal environments (5, 6). Bacteria form biofilms in order to adapt and survive in harsh environments (7, 8). Over the past few decades, biofilm formation has been a major focus of microbial research and, as such, has been studied in relationship to bacterial pathogenesis, immunology, biofouling, microbial technology, and industrial applications (7, 912).Members of the genus Thermus, like many other thermophiles, have evolved two main mechanisms for thermoadaption. One is biofilm formation, which confers protection against environmental stresses such as high temperature and the presence of antibiotics (8). In previous studies, a novel exopolysaccharide, TA-1, was isolated from a T. aquaticus YT-1 biofilm, and both its primary structure and its immunological activity were determined (13). In addition, we showed that the overexpression of uridine diphosphate (UDP)-galactose-4′-epimerase (GalE), which catalyzes the reversible interconversion of UDP-galactose and UDP-glucose, in T. thermophilus HB27 increases biofilm production because of the enzyme''s involvement in an important step of exopolysaccharide (EPS)1 biosynthesis (14). The other mechanism that enables Thermus to thrive in extreme habitats is natural transformation (i.e. the ability to take up free DNA). In hot environments, natural transformation allows the horizontal exchange of genetic information between extremophiles, including of genes that promote thermoadaptation (1517). Recent studies showed that the type IV pili (T4P) on the cell surface of T. thermophilus HB27 not only are required for natural transformation (18, 19), but also mediate adhesion and twitching motility (20). Also, together with the degree of EPS production, the presence of T4P on the bacterial cell surface contributes to the regulation of biofilm formation (21). However, despite extensive research on the physiological, biochemical, and genetic traits of thermophiles, the mechanisms underlying these functions and their role in thermal adaptation have not been fully elucidated (16, 2224).Advances in the field of phosphoproteomics have come from high-resolution mass spectrometry and prokaryotic genome sequencing, which have confirmed the phosphorylation of many bacterial proteins on serine/threonine and tyrosine residues (25, 26). In surveys of phosphorylation-related functions, bacterial serine, threonine, and tyrosine phosphoproteins have been shown to regulate many physiological and adaptation processes, such as central carbon catabolism, the heat shock response, osmolarity, starvation, EPS synthesis, virulence, and sporulation (2527). These observations have been followed by more detailed, species-specific phosphoproteomics investigations, including in Bacillus subtilis (28), Escherichia coli (29), Lactococcus lactis (30), Halobacterium salinarum (31), Klebsiella pneumonia (32), Pseudomonas spp. (33), Rhodopseudomonas palustris (34), and T. thermophilus HB8 (35). In this study, the role played by the global phosphorylation network of the thermophile T. thermophilus HB27 in the physiological processes that mediate the stress responses and thermotolerance of this bacterium was examined. Specifically, we used strong cation exchange (SCX) chromatography and titanium dioxide (TiO2) (2830) enrichment to characterize the phosphoproteomic map of T. thermophilus HB27. Genetic manipulation of this strain indicated that phosphorylation of the PilF protein, which contains an ATP-binding motif (TTC1622/pilF) and drives T4P formation, is involved in both EPS production and piliation, thereby influencing the biofilm formation during thermophilic adaptation.  相似文献   

13.
Fe2+ is now shown to weaken binding between ferritin and mitochondrial aconitase messenger RNA noncoding regulatory structures ((iron-responsive element) (IRE)-RNAs) and the regulatory proteins (IRPs), which adds a direct role of iron to regulation that can complement the well known regulatory protein modification and degradative pathways related to iron-induced mRNA translation. We observe that the Kd value increases 17-fold in 5′-untranslated region IRE-RNA·repressor complexes; Fe2+, is studied in the absence of O2. Other metal ions, Mn2+ and Mg2+ have similar effects to Fe2+ but the required Mg2+ concentration is 100 times greater than for Fe2+ or Mn2+. Metal ions also weaken ethidium bromide binding to IRE-RNA with no effect on IRP fluorescence, using Mn2+ as an O2-resistant surrogate for Fe2+, indicating that metal ions bound IRE-RNA but not IRP. Fe2+ decreases IRP repressor complex stability of ferritin IRE-RNA 5–10 times compared with 2–5 times for mitochondrial aconitase IRE-RNA, over the same concentration range, suggesting that differences among IRE-RNA structures contribute to the differences in the iron responses observed in vivo. The results show the IRE-RNA·repressor complex literally responds to Fe2+, selectively for each IRE-mRNA.Iron (e.g. ferrous sulfate, ferric citrate, and hemin) added to animal cells changes translation rates of messenger RNAs encoding proteins of iron traffic and oxidative metabolism (14). To cross cell membranes, iron ions are transported by membrane proteins such as DMT1 or carried on proteins such as transferrin. Inside the cells, iron is mainly in heme, FeS clusters, non-heme iron cofactors of proteins, and iron oxide minerals coated by protein nanocages (ferritins). Iron in transit is thought to be Fe2+ in labile “pools” accessible to small molecular weight chelators, and/or bound loosely by chaperones.When iron concentrations in the cells increase, a group of mRNAs with three-dimensional, noncoding structures in the 5′-untranslated region (UTR)3 are derepressed (Fig. 1A), i.e. the fraction of the mRNAs in mRNA·repressor protein complexes, which inhibit ribosome binding, decreases and the fraction of the mRNAs in polyribosomes increases (57). The three-dimensional, noncoding mRNA structure, representing a family of related structures, is called the iron-responsive element, or IRE, and the repressors are called iron regulatory proteins (IRPs). Together they are one of the most extensively studied eukaryotic messenger RNA regulatory systems (14). In addition to large numbers of cell studies, structures of IRE-RNAs are known from solution NMR (812), and the RNA·protein complex from x-ray crystallography (13). Recent data indicate that demetallation of IRP1 and disruption of the [4Fe-4S] cluster that inhibits IRP1 binding to RNA will be enhanced by phosphorylation and low iron concentrations (1, 2, 1416). Such results can explain the increased IRP1 binding to IRE-mRNAs and increased translational repression when iron concentrations are abnormally low. However, the mechanism to explain dissociation of IRE-RNA·IRP complexes, thereby allowing ribosome assembly and increased proteosomal degradation of IRPs (1, 2, 14, 15) (Fig. 1A), when high iron concentrations are abnormally high, is currently unknown.Open in a separate windowFIGURE 1.IRE-RNA·IRP complexes and a model for depression by excess iron. A, a representative model of iron-induced translation of 5′-UTR IRE-RNAs. This figure is modified from Ref. 7. B, IRE-RNA sites influenced by metal binding related to the crystal structure of the ferritin-IRE-RNA·IRP complex from Ref. 13. The figure was created by T. Tosha using Discovery Studio 1.6 and Protein Data Bank file 2IPY. ■, hydrated Mg2+, determined by solution NMR; ▴, Cu1+-1.10-phenanthroline, determined by RNA cleavage in O2.Metal ion binding changes conformation and function of most RNA classes, e.g. rRNA (17), tRNA (18, 19), ribozymes (2023), riboswitches (24, 25), possibly hammerhead mRNAs in mammals (26), and proteins. Although the effects of metal ion binding on eukaryotic mRNAs have not been extensively studied, Mg2+ is known to cause changes in conformation, shown by changes in radical cleavage sites of IRE-RNA with 1,10-phenanthrolene-iron and proton shifts in the one-dimensional NMR spectrum (12, 27). The Mg2+ effects are observed at low magnesium concentrations (0.1–0.5 mm) and low molar stoichiometries (1:1 and 2:1 = Mg:RNA).We hypothesized that Fe2+ could directly change the binding of the IRE-mRNA to the iron regulatory protein for several reasons. First, other metal ions influence the IRE-RNA structure (12, 27). Second, in IRE-RNA/IRP cocrystals there are exposed RNA sites in the IRE-RNA/IRP complex that are accessible for interactions (13) (Fig. 1B). Third, regions in the IRE-RNA are hypersensitive to Fe2+-EDTA/ascorbate/H2O2, suggesting selective interactions with metals and/or solvent (28). We now report that Fe2+ weakens IRE-RNA/IRP binding, whereas Mg2+ requires 100 times the concentration and Mn2+ is comparable with Fe2+; the Fe2+ effect was diminished in mutant IRE-RNA and IRE-RNA selective in wild type sequences: ferritin IRE-RNA > mt-aconitase IRE-RNA.  相似文献   

14.
15.
16.
17.
18.
In the pigmented dopaminergic neurons of the human substantia nigra pars compacta the system relevant in iron storage is the polymer neuromelanin (NM). Although in most cells this function is mainly accomplished by ferritin, this protein complex appears not to be expressed in NM-containing neurons. Nevertheless the conceivable presence of iron-storing proteins as part of the NM granules has recently been discussed on the basis of Mössbauer spectroscopy and synchrotron x-ray microspectroscopy. Intriguingly by combining subcellular fractionation of NM granules, peptide sequencing via tandem mass spectrometry, and the additional confirmation by multiple reaction monitoring and immunogold labeling for electron microscopy, L-ferritin could now be unambiguously identified and localized in NM granules for the first time. This finding not only supports direct evidence for a regulatory role of L-ferritin in neuroectodermal cell pigmentation but also integrates a new player within a complicated network governing iron homeostasis in the dopamine neurons of the human substantia nigra. Thus our finding entails far reaching implications especially when considering etiopathogenetic aspects of Parkinson disease.Neuromelanin (NM)1 is a dark colored polymeric pigment produced in specific populations of catecholaminergic neurons in the brain (1). Unlike peripheral melanins, which are produced in specialized cells called melanocytes and may be transferred to other cell types, NM granules are believed to be stored in the neurons in which they are produced. NM granules display a unique, more heterogeneous appearance compared with peripheral melanins. Further unlike melanin, NM is traditionally thought to result from a non-enzymatic synthesis pathway with no known pathway for NM catabolism. More recent data, however, are indicative of some regulation of NM synthesis and turnover (1).NM appears in greatest quantities in the human brain and in lesser amounts in some other non-human primates but is absent from the brain of many lower species. Interest in this pigment has seen a resurgence in recent years because of a hypothesized link between NM and the especial vulnerability of NM-containing neurons of the substantia nigra pars compacta (SN) for cell death in Parkinson disease (PD) (2, 3). In particular, the interaction between iron and NM has been a focus of research (48) because a marked accumulation of iron related to disease severity is reported in the parkinsonian SN (911). The cellular location of this apparent increase in iron is unclear, but a variety of changes in iron regulatory systems occur in PD (1215).A potential candidate for intraneuronal iron homeostasis in the SN, however, is NM. NM is able to bind a variety of metals; 7% (w/w) of isolated NM is reported to consist of iron, copper, zinc, manganese, and chromium (16, 17). Iron binding studies using NM isolated from the human SN demonstrated that NM contains high (KD = 7.18 ± 1.08 nm) and low affinity binding sites (KD = 94.31 ± 6.55 nm) for Fe(III) (18). Our recent data showed that a pure Fe(III) signal can be measured from intact frozen SN tissue using Mössbauer spectroscopy (18). These data indicated that iron is directly bound to NM granules in the SN (4, 16, 19) and that this signal is increased in PD (20). In addition, Mössbauer spectroscopy showed that iron binding sites in NM isolated from the human SN are similar to those of human ferritin and hemosiderin (21). Similar results were also reported recently in whole neurons from formalin-fixed and paraffin-embedded human SN sections using synchrotron chemical x-ray microscopy (22). Because ferritin, the main iron storage protein, is primarily located in glia rather than in neurons (23), it seems unlikely that it could regulate neuronal iron levels, and until today the exact iron storing mechanism in the NM-containing neurons of the SN was unknown.The aim of the present study was thus to find direct evidence for the presence of L-ferritin in NM granules isolated from human post-mortem tissue of subjects with no history of neurological, neurodegenerative, or psychiatric diseases by using a targeted MS-based approach. Recently our group reported a method for the isolation of intact NM granules from the human SN to carry out the first protein profile of these organelles (24). The major findings were the identifications of numerous proteins closely associated with lysosome-related organelles originating from the endosomal system (24, 25). In our present study, we report for the first time the identification of L-ferritin as a component of NM granules, pointing to a ferritin-based iron storage mechanism in the NM-containing neurons of the SN, by using an approach combining one-dimensional (1-D) SDS-PAGE, reversed-phase nano-HPLC electrospray ionization tandem mass spectrometry (nano-LC-ESI-MS/MS and nano-LC-ESI-multiple reaction monitoring (MRM)-MS/MS), Western blot analysis, and immunotransmission electron microscopy.  相似文献   

19.
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