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 906 毫秒
1.
The process of salvaging sugars released from extracellular matrix, during plant cell growth and development, is not well understood, and many molecular components remain to be identified. Here we identify and functionally characterize a unique Arabidopsis gene encoding an α-d-galacturonic acid-1-phosphate kinase (GalAK) and compare it with galactokinase. The GalAK gene appeared to be expressed in all tissues implicating that glycose salvage is a common catabolic pathway. GalAK catalyzes the ATP-dependent conversion of α-d-galacturonic acid (d-GalA) to α-d-galacturonic acid-1-phosphate (GalA-1-P). This sugar phosphate is then converted to UDP-GalA by a UDP-sugar pyrophosphorylase as determined by a real-time 1H NMR-based assay. GalAK is a distinct member of the GHMP kinase family that includes galactokinase (G), homoserine kinase (H), mevalonate kinase (M), and phosphomevalonate kinase (P). Although these kinases have conserved motifs for sugar binding, nucleotide binding, and catalysis, they do have subtle difference. For example, GalAK has an additional domain near the sugar-binding motif. Using site-directed mutagenesis we established that mutation in A368S reduces phosphorylation activity by 40%; A41E mutation completely abolishes GalAK activity; Y250F alters sugar specificity and allows phosphorylation of d-glucuronic acid, the 4-epimer of GalA. Unlike many plant genes that undergo duplication, GalAK occurs as a single copy gene in vascular plants. We suggest that GalAK generates GalA-1-P from the salvaged GalA that is released during growth-dependent cell wall restructuring, or from storage tissue. The GalA-1-P itself is then available for use in the formation of UDP-GalA required for glycan synthesis.d-Galacturonic acid (d-GalA)3 is a quantitatively major glycose present in numerous plant polysaccharides including pectins and arabinogalactan proteins (1, 2). The synthesis of these polysaccharides requires a large number of glycosyltransferases and diverse nucleotide-sugar (NDP-sugar) donors (1, 3). Some of these NDP-sugars are formed by interconversion of pre-existing NDP-sugars and by salvage pathways. For example, the main pathway for UDP-GalA formation is the 4-epimerization of UDP-GlcA, a reaction catalyzed by UDP-GlcA 4-epimerase (46). However, in ripening Fragaria fruit d-GalA is incorporated into pectin (7). It is likely that a sugar kinase converts the d-GalA to GalA-1-P (8), which is then converted to UDP-GalA by a nonspecific UDP-sugar pyrophosphorylase (9). Myo-inositol may also be a source of GalA for polysaccharide biosynthesis (10).Galacturonic acid is likely to be generated by enzyme-catalyzed hydrolysis of pectic polysaccharides in plant tissues. Polysaccharide hydrolase activities are present in germinating seeds (11, 12), in germinating and elongating pollen (1315), and in ripening fruit (14). Thus, monosaccharide salvage pathways may be required for normal plant growth and development.Numerous sugar-1-P kinases, including d-Gal-1-P kinase (16), l-Ara-1-P kinase (17), and l-Fuc-1-P kinase (18), have been described (19), but no d-GalA-1-P kinase has been identified in any species to account for the hydrolysis and recycle of pectic polymers. The subsequent pyrophosphorylation of UDP-sugars could be carried out by UDP-sugar pyrophosphorylases (20).Here, we report the identification and characterization of a functional galacturonic acid kinase (GalAK). We compared the activity of GalAK with a previously uncharacterized Arabidopsis GalK and discussed the evolution of these sugar kinase members of the GHMP kinase.  相似文献   

2.
3.
The peptidoglycan cortex of endospores of Bacillus species is required for maintenance of spore dehydration and dormancy, and the structure of the cortex may also allow it to function in attainment of spore core dehydration. A significant difference between spore and growing cell peptidoglycan structure is the low degree of peptide cross-linking in cortical peptidoglycan; regulation of the degree of this cross-linking is exerted by d,d-carboxypeptidases. We report here the construction of mutant B. subtilis strains lacking all combinations of two and three of the four apparent d,d-carboxypeptidases encoded within the genome and the analysis of spore phenotypic properties and peptidoglycan structure for these strains. The data indicate that while the dacA and dacC products have no significant role in spore peptidoglycan formation, the dacB and dacF products both function in regulating the degree of cross-linking of spore peptidoglycan. The spore peptidoglycan of a dacB dacF double mutant was very highly cross-linked, and this structural modification resulted in a failure to achieve normal spore core dehydration and a decrease in spore heat resistance. A model for the specific roles of DacB and DacF in spore peptidoglycan synthesis is proposed.Peptidoglycan (PG) is the structural element of the bacterial cell wall which determines cell shape and which resists the turgor pressure within the cell. The bacterial endospores produced by species of Bacillus, Clostridium, and several other bacterial genera are modified cells that are able to survive long periods and extreme conditions in a dormant, relatively dehydrated state. The PG wall within the endospore is required for maintenance of the dehydrated state (10, 11), which is the major determinant of spore heat resistance (2, 17, 22). Spore PG appears to be comprised of two distinct though contiguous layers. The thin inner layer, the germ cell wall, appears to have a structure similar to that of the vegetative wall and serves as the initial cell wall of the germinated spore (1, 20, 21, 31). The thicker outer layer, the spore cortex, has a modified structure which may determine its ability to carry out roles specific to the spore, and is rapidly degraded during spore germination (1, 20, 35, 37). The most dramatic of the cortex structural modifications results in partial cleavage or complete removal of ∼75% of the peptide side chains from the glycan strands. Loss of these peptides limits the cross-linking potential of the PG and results in the formation of only one peptide cross-link per 35 disaccharide units in the spore PG, compared to one peptide cross-link per 2.3 to 2.9 disaccharide units in the vegetative PG (1, 20, 36). This low degree of cross-linking has been predicted to give spore PG a flexibility that allows it to have a role in attainment of spore core dehydration (14, 34) in addition to its clear role in maintenance of dehydration. We are studying the structure and mechanism of synthesis of spore PG in an attempt to discern the roles of this structure and its individual components in determining spore properties.A family of proteins called the penicillin-binding proteins (PBPs) polymerizes PG on the external surface of the cell membrane (reviewed in reference 7). The high-molecular-weight (high-MW) members of this family (generally ≥60 kDa) carry the transglycosylase and transpeptidase activities involved in polymerization and cross-linking of the glycan strands. The low-MW PBPs have commonly been found to possess d,d-carboxypeptidase activity. This activity can remove the terminal d-alanine of the peptide side chains and thereby prevent the side chain from serving as a donor in the formation of a peptide cross-link. Analysis of the B. subtilis genome reveals six low-MW PBP-encoding genes: dacA (33), dacB (4), dacC (19), dacF (38), pbpE (23), and pbpX (accession no. Z99112). The four dac gene products exhibit very high sequence similarity to proven d,d-carboxypeptidases, and this activity has been demonstrated in vitro for the dacA and dacB products, PBP5 (12) and PBP5* (32), respectively. The sequences of the pbpE and pbpX products are more distantly related, and no activity has yet been established or ruled out for them.PBP5 is the major penicillin-binding and d,d-carboxypeptidase activity found in vegetative cells (12). Although dacA expression declines significantly during sporulation, a significant amount of PBP5 remains during the time of spore PG synthesis (29). A dacA-null mutation results in no obvious effects on vegetative growth, sporulation, spore characteristics, or spore germination (3, 33). However, loss of PBP5 does result in a reduction of cleavage of peptide side chains from the tetrapeptide to the tripeptide form in the spore PG (20). PBP5* is expressed only during sporulation and only in the mother cell compartment of the sporangium, under the control of the RNA polymerase ςE subunit (4, 5, 28, 29). A dacB-null mutation leading to loss of this d,d-carboxypeptidase results in a fourfold increase in the effective cross-linking of the spore PG (1, 20, 22). This structural change is accompanied by only slight decreases in spore core dehydration and heat resistance (3, 22). The suspected d,d-carboxypeptidase activities of the products of the dacC and dacF genes have not been demonstrated. The latter two genes are expressed only during the postexponential growth phase: dacC is expressed during early stationary phase under the control of ςH (19) and dacF is expressed only within the forespore under the control of ςF (27, 38). Null mutations effecting either gene result in no obvious phenotype and no change in spore PG structure (19, 38).The multiplicity of these proteins in sporulating cells and the lack of effect of loss of some of them suggested redundancy of function among these proteins, a situation observed previously with PBPs of a high-MW class (25, 30, 39). In order to examine this possibility we have constructed mutants lacking multiple low-MW PBPs and have examined their sporulation efficiency, spore PG structure, spore heat resistance and wet density, and spore germination and outgrowth. The present study demonstrates a role for the dacF gene product in synthesis of spore PG, and we also present a model for the roles of the dacB and dacF gene products in spore PG formation.  相似文献   

4.
d-Serine is an endogenous coagonist for the N-methyl-d-aspartate receptor and is involved in excitatory neurotransmission in the brain. Mammalian pyridoxal 5′-phosphate-dependent serine racemase, which is localized in the mammalian brain, catalyzes the racemization of l-serine to yield d-serine and vice versa. The enzyme also catalyzes the dehydration of d- and l-serine. Both reactions are enhanced by Mg·ATP in vivo. We have determined the structures of the following three forms of the mammalian enzyme homolog from Schizosaccharomyces pombe: the wild-type enzyme, the wild-type enzyme in the complex with an ATP analog, and the modified enzyme in the complex with serine at 1.7, 1.9, and 2.2 Å resolution, respectively. On binding of the substrate, the small domain rotates toward the large domain to close the active site. The ATP binding site was identified at the domain and the subunit interface. Computer graphics models of the wild-type enzyme complexed with l-serine and d-serine provided an insight into the catalytic mechanisms of both reactions. Lys-57 and Ser-82 located on the protein and solvent sides, respectively, with respect to the cofactor plane, are acid-base catalysts that shuttle protons to the substrate. The modified enzyme, which has a unique “lysino-d-alanyl” residue at the active site, also exhibits catalytic activities. The crystal-soaking experiment showed that the substrate serine was actually trapped in the active site of the modified enzyme, suggesting that the lysino-d-alanyl residue acts as a catalytic base in the same manner as inherent Lys-57 of the wild-type enzyme.d-Serine, which is present at a high level in the mammalian brain, serves as an endogenous coagonist for the N-methyl-d-aspartate (NMDA)5 receptor selectively localized on the postsynaptic membrane of the excitatory synapse (15) and is involved in excitatory neurotransmission and higher brain functions such as learning and memory (3, 6, 7). Stimulation of the NMDA receptor requires the binding of d-serine as well as the agonist l-glutamate. The major enzyme for d-serine synthesis from l-serine in the brain is considered to be pyridoxal 5′-phosphate (PLP)-dependent serine racemase (SR) (810). d-Serine and SR are localized on protoplasmic astrocytes that have the α-amino-3-hydroxy-5-methylisoxazole-4-propionic acid receptor. Glutamate released from presynaptic neurons approaches and activates the α-amino-3-hydroxy-5-methylisoxazole-4-propionic acid receptor, which in turn induces SR to produce d-serine and is followed by d-serine release from astrocytes that act on the NMDA receptor. Recently, it was shown that not only glia but also neurons synthesize and release d-serine involved in signaling (11). SR also catalyzes α,β-elimination of water from d- or l-serine to form pyruvate and ammonia as well as the conversion of l-serine into d-serine and vice versa and is presumed to link d-serine synthesis and energy metabolism of astrocytes (12) and to control the d-serine level (13). Mg·ATP, which is fully bound to SR under physiological conditions, stimulates racemization and the α,β-elimination reaction catalyzed by SR (12, 14).SR was first discovered in pupae of the silkworm Bombyx mori (15), which was followed by purification of the enzyme from a rat brain and cloning of the mouse and human genes (8, 9). The primary structure of mammalian SR is distinct from those of racemases from prokaryotes but is similar to those of fold-type II PLP-dependent enzymes (1618). We have cloned and expressed the Schizosaccharomyces pombe gene homologous to human and mouse SRs, the sequence identities being 35.1 and 37.4%, respectively, in Escherichia coli. The protein product is a bifunctional enzyme that catalyzes racemization and the α,β-elimination reaction of D, l-serine as mammalian SR does. SR from S. pombe (spSR) comprises 322 residues (the N-terminal Met is removed in the purified enzyme) and one PLP per subunit, the subunit molecular weight being 34,917. The mammalian SR homolog, spSR, is an interesting target enzyme for the development of a novel therapeutic compound controlling the d-serine level because d-serine is the product of an SR-catalyzed reaction. In our recent report, the active site of spSR was shown to be modified with its natural substrate serine by mass spectroscopic and x-ray studies (19). Interestingly, the catalytic lysine, which originally forms a Schiff base with PLP, is converted to a lysino-d-alanyl residue through the reaction with the substrate, serine (Fig. 1). The modified enzyme exhibits racemase (54% of the wild-type enzyme) and α,β-elimination (68% of the wild-type enzyme) activities with the amino group of the d-alanyl moiety of the lysinoalanyl residue forming a Schiff base with PLP in place of the lysine (19). In addition, the mammalian SR seems to be possibly modified to have a lysinoalanyl residue at the active site, as observed in spSR (20).Open in a separate windowFIGURE 1.Covalent modification of the active site. The catalytic Lys-57 in spSRw is converted to lysino-d-alanyl residue. The α-amino group (indicated with “α”) of the d-alanyl moiety in the residue acts as a catalytic base in spSRm. The circled P is a phosphate group.Although the structure of modified spSR (spSRm) has been determined (19), the structure-function relationship of essential wild-type spSR (spSRw), the binding mode of activator Mg·ATP, the catalytic base to shuttle protons to the substrate d-serine, and the substrate recognition of the modified enzyme have not yet been uncovered. We now report the three-dimensional structures of unliganded spSRw in the open form, spSRw·AMP-PCP in the open form, and spSRm·serine in the closed form.  相似文献   

5.
The pathway of d-xylose degradation in archaea is unknown. In a previous study we identified in Haloarcula marismortui the first enzyme of xylose degradation, an inducible xylose dehydrogenase (Johnsen, U., and Schönheit, P. (2004) J. Bacteriol. 186, 6198–6207). Here we report a comprehensive study of the complete d-xylose degradation pathway in the halophilic archaeon Haloferax volcanii. The analyses include the following: (i) identification of the degradation pathway in vivo following 13C-labeling patterns of proteinogenic amino acids after growth on [13C]xylose; (ii) identification of xylose-induced genes by DNA microarray experiments; (iii) characterization of enzymes; and (iv) construction of in-frame deletion mutants and their functional analyses in growth experiments. Together, the data indicate that d-xylose is oxidized exclusively to the tricarboxylic acid cycle intermediate α-ketoglutarate, involving d-xylose dehydrogenase (HVO_B0028), a novel xylonate dehydratase (HVO_B0038A), 2-keto-3-deoxyxylonate dehydratase (HVO_B0027), and α-ketoglutarate semialdehyde dehydrogenase (HVO_B0039). The functional involvement of these enzymes in xylose degradation was proven by growth studies of the corresponding in-frame deletion mutants, which all lost the ability to grow on d-xylose, but growth on glucose was not significantly affected. This is the first report of an archaeal d-xylose degradation pathway that differs from the classical d-xylose pathway in most bacteria involving the formation of xylulose 5-phosphate as an intermediate. However, the pathway shows similarities to proposed oxidative pentose degradation pathways to α-ketoglutarate in few bacteria, e.g. Azospirillum brasilense and Caulobacter crescentus, and in the archaeon Sulfolobus solfataricus.d-Xylose, a constituent of the polymer xylan, is the major component of the hemicellulose plant cell wall material and thus one of the most abundant carbohydrates in nature. The utilization of d-xylose by microorganisms has been described in detail in bacteria and fungi, for which two different catabolic pathways have been reported. In many bacteria, such as Escherichia coli, Bacillus, and Lactobacillus species, xylose is converted by the activities of xylose isomerase and xylulose kinase to xylulose 5-phosphate as an intermediate, which is further degraded mainly by the pentose phosphate cycle or phosphoketolase pathway. Most fungi convert xylose to xylulose 5-phosphate via xylose reductase, xylitol dehydrogenase, and xylulose kinase. Xylulose 5-phosphate is also an intermediate of the most common l-arabinose degradation pathway in bacteria, e.g. of E. coli, via activities of isomerase, kinase, and epimerase (1).Recently, by genetic evidence, a third pathway of xylose degradation was proposed for the bacterium Caulobacter crescentus, in analogy to an alternative catabolic pathway of l-arabinose, reported for some bacteria, including species of Azospirillum, Pseudomonas, Rhizobium, Burkholderia, and Herbasprillum (2, 3). In these organisms l-arabinose is oxidatively degraded to α-ketoglutarate, an intermediate of the tricarboxylic acid cycle, via the activities of l-arabinose dehydrogenase, l-arabinolactonase, and two successive dehydration reactions forming 2-keto-3-deoxy-l-arabinoate and α-ketoglutarate semialdehyde; the latter compound is further oxidized to α-ketoglutarate via NADP+-specific α-ketoglutarate semialdehyde dehydrogenase (KGSADH).2 In a few Pseudomonas and Rhizobium species, a variant of this l-arabinose pathway was described involving aldolase cleavage of the intermediate 2-keto-3-deoxy-l-arabinoate to pyruvate and glycolaldehyde, rather than its dehydration and oxidation to α-ketoglutarate (4). Because of the presence of some analogous enzyme activities in xylose-grown cells of Azosprillum and Rhizobium, the oxidative pathway and its variant was also proposed as a catabolic pathway for d-xylose. Recent genetic analysis of Caulobacter crecentus indicates the presence of an oxidative pathway for d-xylose degradation to α-ketoglutarate. All genes encoding xylose dehydrogenase and putative lactonase, xylonate dehydratase, 2-keto-3-deoxylonate dehydratase, and KGSADH were found to be located on a xylose-inducible operon (5). With exception of xylose dehydrogenase, which has been partially characterized, the other postulated enzymes of the pathway have not been biochemically analyzed.The pathway of d-xylose degradation in the domain of archaea has not been studied so far. First analyses with the halophilic archaeon Haloarcula marismortui indicate that the initial step of d-xylose degradation involves a xylose-inducible xylose dehydrogenase (6) suggesting an oxidative pathway of xylose degradation to α-ketoglutarate, or to pyruvate and glycolaldehyde, in analogy to the proposed oxidative bacterial pentose degradation pathways. Recently, a detailed study of d-arabinose catabolism in the thermoacidophilic crenarchaeon Sulfolobus solfataricus was reported. d-Arabinose was found to be oxidized to α-ketoglutarate involving d-arabinose dehydrogenase, d-arabinoate dehydratase, 2-keto-3-deoxy-d-arabinoate dehydratase, and α-ketoglutarate semialdehyde dehydrogenase (3).In this study, we present a comprehensive analysis of the complete d-xylose degradation pathway in the halophilic archaeon Haloferax volcanii. This halophilic archaeon was chosen because it exerts several suitable properties for the analyses. For example, it can be cultivated on synthetic media with sugars, e.g. xylose, an advantage for in vivo labeling studies in growing cultures. Furthermore, a shotgun DNA microarray of H. volcanii is available (7) allowing the identification of xylose-inducible genes, and H. volcanii is one of the few archaea for which an efficient protocol was recently described to generate in-frame deletion mutants.Accordingly, the d-xylose degradation pathway was elucidated following in vivo labeling experiments with [13C]xylose, DNA microarray analyses, and the characterization of enzymes involved and their encoding genes. The functional involvement of genes and enzymes was proven by constructing corresponding in-frame deletion mutants and their analysis by selective growth experiments on xylose versus glucose. The data show that d-xylose was exclusively degraded to α-ketoglutarate involving xylose dehydrogenase, a novel xylonate dehydratase, 2-keto-3-deoxyxylonate dehydratase, and α-ketoglutarate semialdehyde dehydrogenase.  相似文献   

6.
7.
Bacillus cereus 569 (ATCC 10876) germinates in response to inosine or to l-alanine, but the most rapid germination response is elicited by a combination of these germinants. Mutants defective in their germination response to either inosine or to l-alanine were isolated after Tn917-LTV1 mutagenesis and enrichment procedures; one class of mutant could not germinate in response to inosine as a sole germinant but still germinated in response to l-alanine, although at a reduced rate; another mutant germinated normally in response to inosine but was slowed in its germination response to l-alanine. These mutants demonstrated that at least two signal response pathways are involved in the triggering of germination. Stimulation of germination in l-alanine by limiting concentrations of inosine and stimulation of germination in inosine by low concentrations of l-alanine were still detectable in these mutants, suggesting that such stimulation is not dependent on complete functionality of both these germination loci. Two transposon insertions that affected inosine germination were found to be located 2.2 kb apart on the chromosome. This region was cloned and sequenced, revealing an operon of three open reading frames homologous to those in the gerA and related operons of Bacillus subtilis. The individual genes of this gerI operon have been named gerIA, gerIB, and gerIC. The GerIA protein is predicted to possess an unusually long, charged, N-terminal domain containing nine tandem copies of a 13-amino-acid glutamine- and serine-rich sequence.Bacillus species have the ability, under certain nutrient stresses, to undergo a complex differentiation process resulting in the formation of a highly resistant dormant endospore (6). These spores can then persist in the environment for prolonged periods until a sensitive response mechanism detects specific environmental conditions, initiating the processes of germination and outgrowth (9, 21, 25). Germination can be initiated by a variety of agents (12), including nutrients, enzymes, or physical factors, such as abrasion or hydrostatic pressure.The molecular genetics of spore germination has been most extensively studied in Bacillus subtilis 168 (21). B. subtilis spores can be triggered to germinate in response to either l-alanine or to a combination (29) of asparagine, glucose, fructose, and potassium ions (AGFK). Mutants of B. subtilis which are defective in germination responses to one or to both types of germinant have been isolated previously (20, 27). Analysis of these mutants suggests that the germinants interact with separate germinant-specific complexes within the spore (21). This in some way leads to activation of components of the germination apparatus common to both responses, such as germination-specific cortex lytic enzymes, leading in turn to complete germination of the spore (10, 22). The mutations within the gerA operon of B. subtilis specifically block germination initiated by l-alanine (34). The predicted amino acid sequences of the three GerA proteins encoded in the operon suggest that these proteins could be membrane associated, and they are the most likely candidates to represent the germinant receptor for alanine (21).The amino acid l-alanine has been identified as a common but not universal germinant in a variety of Bacillus species, often requiring the presence of adjuncts such as electrolytes and sugars. Ribosides, such as inosine, represent another type of common germinant, although many species are unable to germinate rapidly in response to these without the addition of l-alanine (9).The food-borne pathogen Bacillus cereus is a major cause of food poisoning of an emetic and diarrheal type (13, 16). The germination and growth of Bacillus cereus spores during food storage can lead to food spoilage and the potential to cause food poisoning (16). B. cereus has been shown to germinate in response to l-alanine and to ribosides (11, 18, 23). Spore germination can be triggered by l-alanine alone, but at high spore densities this response becomes inhibited by d-alanine, generated by the alanine racemase activity associated with the spores (8, 11). This auto-inhibition of l-alanine germination can be reduced by the inclusion of a racemase inhibitor (O-carbamyl-d-serine) with the germinating spores (11).Inosine is the most effective riboside germinant for B. cereus T, while adenosine and guanosine are less potent (28). The rate of riboside-triggered germination has been reported to be enhanced dramatically by the addition of l-alanine (18). It is unclear whether ribosides can act as a sole germinant, or whether there is an absolute requirement for l-alanine (28).An attempt has been made to analyze genetically the molecular components of the germination apparatus in B. cereus in order to dissect the germination responses of this species and to determine whether riboside-induced germination involves components related to those already described for amino acid and sugar germinants in B. subtilis.  相似文献   

8.
The Arabidopsis thaliana locus At5g06580 encodes an ortholog to Saccharomyces cerevisiae d-lactate dehydrogenase (AtD-LDH). The recombinant protein is a homodimer of 59-kDa subunits with one FAD per monomer. A substrate screen indicated that AtD-LDH catalyzes the oxidation of d- and l-lactate, d-2-hydroxybutyrate, glycerate, and glycolate using cytochrome c as an electron acceptor. AtD-LDH shows a clear preference for d-lactate, with a catalytic efficiency 200- and 2000-fold higher than that for l-lactate and glycolate, respectively, and a Km value for d-lactate of ∼160 μm. Knock-out mutants showed impaired growth in the presence of d-lactate or methylglyoxal. Collectively, the data indicated that the protein is a d-LDH that participates in planta in the methylglyoxal pathway. Web-based bioinformatic tools revealed the existence of a paralogous protein encoded by locus At4g36400. The recombinant protein is a homodimer of 61-kDa subunits with one FAD per monomer. A substrate screening revealed highly specific d-2-hydroxyglutarate (d-2HG) conversion in the presence of an organic cofactor with a Km value of ∼580 μm. Thus, the enzyme was characterized as a d-2HG dehydrogenase (AtD-2HGDH). Analysis of knock-out mutants demonstrated that AtD-2HGDH is responsible for the total d-2HGDH activity present in A. thaliana. Gene coexpression analysis indicated that AtD-2HGDH is in the same network as several genes involved in β-oxidation and degradation of branched-chain amino acids and chlorophyll. It is proposed that AtD-2HGDH participates in the catabolism of d-2HG most probably during the mobilization of alternative substrates from proteolysis and/or lipid degradation.l- and d-lactate dehydrogenases belong to evolutionarily unrelated enzyme families (1). l-Lactate is oxidized by l-lactate:NAD oxidoreductase (EC 1.1.1.27), which catalyzes the reaction l-lactate + NAD → pyruvate + NADH, and by l-lactate cytochrome c oxidoreductase (l-lactate cytochrome c oxidoreductase, EC 1.1.2.3), which catalyzes the reaction l-lactate + 2 cytochrome c (oxidized) → pyruvate + 2 cytochrome c (reduced). Both groups are found in eubacteria, archebacteria, and eukaryotes. All known plant sequences belong to the EC 1.1.1.27 group (1). On the other hand, d-lactate is oxidized by d-lactate:NAD oxidoreductase (d-lactate:NAD oxidoreductase, EC 1.1.1.28), which catalyzes the reaction d-lactate + NAD → pyruvate + NADH, and by d-lactate cytochrome c oxidoreductase (d-lactate cytochrome c oxidoreductase, EC 1.1.2.4), which catalyzes the reaction d-lactate + 2 cytochrome c (oxidized) → pyruvate + 2 cytochrome c (reduced).Although l-lactate dehydrogenase belongs to the most intensely studied enzyme families (2, 3), our knowledge about the structure, kinetics, and biological function of d-LDH3 is limited. d-LDHs have mainly been identified in prokaryotes and fungi where they play an important role in anaerobic energy metabolism (410). In Saccharomyces cerevisiae and Kluyveromyces lactis, a mitochondrial flavoprotein d-lactate ferricytochrome c oxidoreductase (d-lactate cytochrome c oxidoreductase), catalyzing the oxidation of d-lactate to pyruvate, is required for the utilization of d-lactate (8, 11). In S. cerevisiae it was suggested that d-LDH is involved in the metabolism of methylglyoxal (MG) (12).In eukaryotic cells, d-lactate results from the glyoxalase system (13, 14). This system is the main MG catabolic pathway, comprising the enzymes glyoxalase I (lactoylglutathione lyase, EC 4.4.1.5) and glyoxalase II (hydroxyacylglutathione hydrolase, EC 3.1.2.6). MG (CH3-CO-CHO; see structure in Fig. 4) is a cytotoxic compound formed primarily as a by-product of glycolysis through nonenzymatic phosphate elimination from dihydroxyacetone phosphate and glyceraldehyde 3-phosphate (15), and its production in various plants is enhanced under stress conditions such as salt, drought, cold, and heavy metal stress (16, 17). Moreover, the overexpression of glyoxalase I or II was shown to confer resistance to salt stress in tobacco and rice (17, 18). It is assumed that the role of the MG pathway, from MG synthase to d-lactate cytochrome c oxidoreductase in the extant metabolism, is to detoxify MG, whereas in the early state of metabolic development it might function as an anaplerotic route for the tricarboxylic acid cycle (15).Open in a separate windowFIGURE 4.Scheme showing the involvement of AtD-LDH in the methylglyoxal pathway and of AtD-2HGDH in the respiration of substrates from proteolysis and/or lipid degradation. d-Lactate resulting from the glyoxalase system is converted to pyruvate by AtD-LDH. The electrons originated may be transferred to the respiratory chain through cytochrome c in the intermembrane space. d-2-HG produced in the peroxisomes (as shown in supplemental Fig. S3) is transported to the mitochondria and converted to 2-ketoglutarate by AtD-2HGDH. Electrons are donated to the electron transport chain through the ETF/ETFQO system. Dotted files represent possible transport processes. 2-KG, 2-ketoglutarate. CIII, complex III. CIV, complex IV. e, electron. ETF, electron transfer protein. ETFQO, ETF-ubiquinone oxidoreductase. GSH, glutathione. Pyr, pyruvate. TCA cycle, tricarboxylic acid cycle; UQ, ubiquinone.Glyoxalase I catalyzes the formation of S-d-lactoylglutathione from the hemithioacetal formed nonenzymatically from MG and glutathione, although glyoxalase II catalyzes the hydrolysis of S-d-lactoylglutathione to regenerate glutathione and liberate d-lactate. Glyoxalase I and II activities are present in all tissues of eukaryotic organisms. Glyoxalase I is found in the cytosol, whereas glyoxalase II localizes to the cytosol and mitochondria (13, 19, 20). Although glyoxalase I and II were extensively characterized, there are only few reports on the characterization of d-LDH. Recently, Atlante et al. (13) showed that externally added d-lactate caused oxygen consumption by mitochondria and that this metabolite was oxidized by a mitochondrial flavoprotein in Helianthus tuberosus.The complete sequence of Arabidopsis thaliana opened the way to search for genes encoding d-LDHs. Based on similarity with the d-LDH from S. cerevisiae (DLD1), an A. thaliana ortholog was identified. In this study, the isolation and structural and biochemical characterization of the recombinant mature d-LDH from A. thaliana (AtD-LDH) and its paralog, which was found to be a d-2-hydroxyglutarate dehydrogenase (AtD-2HGDH), is described. Whereas AtD-LDH has a narrow substrate specificity and the preferred substrates are d-lactate and d-2-hydroxybutyrate, AtD-2HGDH showed activity exclusively with d-2-hydroxyglutarate. Based on gene coexpression analysis and analysis of corresponding knock-out mutants, the participation of these previously unrecognized mitochondrial activities in plant metabolism is discussed.  相似文献   

9.
A low-specificity l-threonine aldolase (l-TA) gene from Pseudomonas sp. strain NCIMB 10558 was cloned and sequenced. The gene contains an open reading frame consisting of 1,041 nucleotides corresponding to 346 amino acid residues. The gene was overexpressed in Escherichia coli cells, and the recombinant enzyme was purified and characterized. The enzyme, requiring pyridoxal 5′-phosphate as a coenzyme, is strictly l specific at the α position, whereas it cannot distinguish between threo and erythro forms at the β position. In addition to threonine, the enzyme also acts on various other l-β-hydroxy-α-amino acids, including l-β-3,4-dihydroxyphenylserine, l-β-3,4-methylenedioxyphenylserine, and l-β-phenylserine. The predicted amino acid sequence displayed less than 20% identity with those of low-specificity l-TA from Saccharomyces cerevisiae, l-allo-threonine aldolase from Aeromonas jandaei, and four relevant hypothetical proteins from other microorganisms. However, lysine 207 of low-specificity l-TA from Pseudomonas sp. strain NCIMB 10558 was found to be completely conserved in these proteins. Site-directed mutagenesis experiments showed that substitution of Lys207 with Ala or Arg resulted in a significant loss of enzyme activity, with the corresponding disappearance of the absorption maximum at 420 nm. Thus, Lys207 of the l-TA probably functions as an essential catalytic residue, forming an internal Schiff base with the pyridoxal 5′-phosphate of the enzyme to catalyze the reversible aldol reaction.β-Hydroxy-α-amino acids constitute an important class of compounds. They are natural products in their own right and are components of a range of antibiotics, for example, cyclosporin A, lysobactin, and vancomycin (30) and bouvardin and deoxybouvardin (6). 4-Hydroxy-l-threonine is a precursor of rizobitoxine, a potent inhibitor of pyridoxal 5′-phosphate (PLP)-dependent enzymes (32). 3,4,5-Trihydroxyl-l-aminopentanoic acid is a key component of polyoxins (32). l-threo-3,4-Dihydroxyphenylserine is a new drug for Parkinson’s disease therapy (13). However, the industrial production of β-hydroxy-α-amino acids has been limited to chemical synthesis processes, which need multiple steps to isolate the four isomers (l-threo form, d-threo form, l-erythro form, and d-erythro form). Threonine aldolase (EC 4.1.2.5), which stereospecifically catalyzes the retro-aldol cleavage of threonine, is a potentially useful catalyst for the synthesis of substituted amino acids from aldehyde and glycine (27, 31, 32).Two different types of threonine aldolases are known so far. l-allo-Threonine aldolase (l-allo-TA), isolated and purified from Aeromonas jandaei DK-39 (8), stereospecifically catalyzes the reversible interconversion of l-allo-threonine and glycine. Low-specificity l-threonine aldolase (l-TA) catalyzes the cleavage of both l-threonine and l-allo-threonine to glycine and acetaldehyde, as well as the reverse reaction, aldol condensation. The enzymes have been purified and characterized from Candida humicola (9, 34) and Saccharomyces cerevisiae (12). Low-specificity l-TA activity has also been shown to exist in mammals (7, 23, 26) and a variety of other microbial species (2, 4, 17, 35). The enzyme is physiologically important for the synthesis of cellular glycine in yeast (12, 15, 16). Threonine aldolases with distinct stereospecificities are ideal targets for enzymology studies on structural and functional relationships. However, information on the primary structures of threonine aldolases was limited to our recent studies (11, 12). The construction of an overproduction system for threonine aldolase will be indispensable for the industrial biosyntheses of β-hydroxy-α-amino acids.The present work focuses on the cloning, sequencing, and overexpression in Escherichia coli cells of the low-specificity l-TA gene from Pseudomonas sp. strain NCIMB 10558, the purification and characterization of the recombinant enzyme, and the identification of the active-site lysine residue of the enzyme by site-directed mutagenesis. Evidence is presented that Lys207 of low-specificity l-TA probably functions as a catalytic residue, forming an internal Schiff base with the PLP of the enzyme to catalyze the reversible aldol reaction. This is the first report showing a purified enzyme with l-β-3,4-dihydroxyphenylserine aldolase and l-β-3,4-methylenedioxyphenylserine aldolase activities, providing a new route for the industrial production of these important unnatural amino acids.  相似文献   

10.
Burkholderia cenocepacia is an opportunistic pathogen that displays a remarkably high resistance to antimicrobial peptides. We hypothesize that high resistance to antimicrobial peptides in these bacteria is because of the barrier properties of the outer membrane. Here we report the identification of genes for the biosynthesis of the core oligosaccharide (OS) moiety of the B. cenocepacia lipopolysaccharide. We constructed a panel of isogenic mutants with truncated core OS that facilitated functional gene assignments and the elucidation of the core OS structure in the prototypic strain K56-2. The core OS structure consists of three heptoses in the inner core region, 3-deoxy-d-manno-octulosonic acid, d-glycero-d-talo-octulosonic acid, and 4-amino-4-deoxy-l-arabinose linked to d-glycero-d-talo-octulosonic acid. Also, glucose is linked to heptose I, whereas heptose II carries a second glucose and a terminal heptose, which is the site of attachment of the O antigen. We established that the level of core truncation in the mutants was proportional to their increased in vitro sensitivity to polymyxin B (PmB). Binding assays using fluorescent 5-dimethylaminonaphthalene-1-sulfonyl-labeled PmB demonstrated a correlation between sensitivity and increased binding of PmB to intact cells. Also, the mutant producing a heptoseless core OS did not survive in macrophages as compared with the parental K56-2 strain. Together, our results demonstrate that a complete core OS is required for full PmB resistance in B. cenocepacia and that resistance is due, at least in part, to the ability of B. cenocepacia to prevent binding of the peptide to the bacterial cell envelope.Burkholderia cenocepacia is a Gram-negative opportunistic pathogen ubiquitously found in the environment (1, 2). Although generally harmless to healthy individuals, B. cenocepacia affects immunocompromised patients (1) such as those with cystic fibrosis and chronic granulomatous disease. Infected cystic fibrosis patients commonly develop chronic lung infections that are very difficult to treat because these bacteria are intrinsically resistant to virtually all clinically useful antibiotics as well as antimicrobial peptides (APs)5 (1, 3).Lipopolysaccharide (LPS) is the major surface component of Gram-negative bacteria and consists of lipid A, core oligosaccharide (OS), and in some bacteria O-specific polysaccharide or O antigen (4, 5). The O antigen acts as a protective barrier against desiccation, phagocytosis, and serum complement-mediated killing, whereas the core OS and the lipid A contribute to maintain the integrity of the outer membrane (4, 5). The lipid A also anchors the LPS molecule to the outer leaflet of the outer membrane and accounts for the endotoxic activity of LPS (4, 6). Lipid A is a bisphosphorylated β-1,6-linked glucosamine disaccharide substituted with fatty acids ester-linked at positions 3 and 3′ and amide-linked at positions 2 and 2′ (4). The core OS can be subdivided into the inner core and outer core. The inner core OS typically consists of one or two 3-deoxy-d-manno-octulosonic acid (Kdo) residues linked to the lipid A and three l-glycero-d-manno-heptose residues linked to the first Kdo (4). The outer core OS in enteric bacteria typically consists of 8–12 branched sugars linked to heptose II of the inner core. As a result of phosphate groups on the lipid A and core OS, the bacterial surface has a net negative charge. This plays an important role in the interaction of the bacterial surface with positively charged compounds such as cationic APs, which are cationic amphipathic molecules that kill bacteria by membrane permeabilization. In response to a series of environmental conditions such as low magnesium or high iron, bacteria can express modified LPS molecules that result in a less negative surface. This reduces the binding of APs and promotes resistance to these compounds. Previous studies have shown that Burkholderia LPS molecules possess unique properties. For example, Kdo cannot be detected by classic colorimetric methods in LPS from Burkholderia pseudomallei and Burkholderia cepacia, and the covalent linkage between Kdo and lipid A is more resistant to acid hydrolysis than in conventional LPS molecules (7). In B. cepacia, 4-amino-4-deoxy-l-arabinose (l-Ara4N) is bound to the lipid A by a phosphodiester linkage at position 4 of the nonreducing glucosamine (GlcN II) (8) and is also present as a component of the core OS. Also, instead of two Kdo molecules, the B. cepacia core OS has only one Kdo and the unusual Kdo analog, d-glycero-d-talo-octulosonic acid (Ko), which is nonstoichiometrically substituted with l-Ara4N forming a 1→8 linkage with α-Ko (7, 9). Although this is also the case for the inner core OS of B. cenocepacia J2315 (10), it is not a common feature for the core OS in all Burkholderia. For example, the inner core of Burkholderia caryophylli consists of two Kdo residues and does not possess l-Ara4N (11).Burkholderia species, including B. cenocepacia, are intrinsically resistant to human and non-human APs such as these produced by airway epithelial cells (12, 13), human β-defensin 3 (14), human neutrophil peptides (15), and polymyxin B (PmB) (15, 16). The minimum inhibitory concentration determined for some of these peptides in several Burkholderia species is greater than 500 μg/ml, which could aid these microorganisms during colonization of the respiratory epithelia (13). It has been proposed that the resistance of B. cepacia to cationic APs stems from ineffective binding to the outer membrane, as a consequence of the low number of phosphate and carboxylate groups in the lipopolysaccharide (17), but a systematic analysis of the molecular basis of AP resistance in B. cenocepacia and other Burkholderia is lacking. We have previously reported that a heptoseless B. cenocepacia mutant (SAL1) is significantly more sensitive than the parental clinical strain K56-2 to APs (15). This mutant has a truncated inner core and lacks the outer core, suggesting that a complete core OS is required for resistance of B. cenocepacia to APs.Apart from heptoses, the role of other sugar moieties of the B. cenocepacia core OS in AP resistance is not known. In this study, we report the structure of the core OS for B. cenocepacia strain K56-2 and its isogenic mutants XOA3, XOA7, and XOA8, which carry various core OS truncations. The structural analysis, combined with mutagenesis, allowed us to assign function to the majority of the genes involved in core OS biosynthesis and ligation of the O antigen and to establish that the degree of truncation of the core OS correlates with increased binding and bacterial sensitivity to PmB in vitro and reduced bacterial intracellular survival in macrophages.  相似文献   

11.
Mithramycin is an antitumor polyketide drug produced by Streptomyces argillaceus that contains two deoxysugar chains, a disaccharide consisting of two d-olivoses and a trisaccharide consisting of a d-olivose, a d-oliose, and a d-mycarose. From a cosmid clone (cosAR3) which confers resistance to mithramycin in streptomycetes, a 3-kb PstI-XhoI fragment was sequenced, and two divergent genes (mtmGI and mtmGII) were identified. Comparison of the deduced products of both genes with proteins in databases showed similarities with glycosyltransferases and glucuronosyltransferases from different sources, including several glycosyltransferases involved in sugar transfer during antibiotic biosynthesis. Both genes were independently inactivated by gene replacement, and the mutants generated (M3G1 and M3G2) did not produce mithramycin. High-performance liquid chromatography analysis of ethyl acetate extracts of culture supernatants of both mutants showed the presence of several peaks with the characteristic spectra of mithramycin biosynthetic intermediates. Four compounds were isolated from both mutants by preparative high-performance liquid chromatography, and their structures were elucidated by physicochemical methods. The structures of these compounds were identical in both mutants, and the compounds are suggested to be glycosylated intermediates of mithramycin biosynthesis with different numbers of sugar moieties attached to C-12a-O of a tetracyclic mithramycin precursor and to C-2-O of mithramycinone: three tetracyclic intermediates containing one sugar (premithramycin A1), two sugars (premithramycin A2), or three sugars (premithramycin A3) and one tricyclic intermediate containing a trisaccharide chain (premithramycin A4). It is proposed that the glycosyltransferases encoded by mtmGI and mtmGII are responsible for forming and transferring the disaccharide during mithramycin biosynthesis. From the structures of the new metabolites, a new biosynthetic sequence regarding late steps of mithramycin biosynthesis can be suggested, a sequence which includes glycosyl transfer steps prior to the final shaping of the aglycone moiety of mithramycin.

Many bioactive drugs contain sugars attached to their aglycones which are usually important or, in some cases, essential for bioactivity. Most of these sugars belong to the family of the 6-deoxyhexoses (6-DOH) (18, 20, 27) and are transferred to the different aglycones as late steps in biosynthesis. Genes involved in the biosynthesis of different 6-DOH have been reported elsewhere and participate in the biosynthesis of erythromycin (9, 12, 31, 38, 39), daunorubicin (13, 26, 36), mithramycin (22), granaticin (2), streptomycin (10, 28), and tylosin (14, 23). However, information about the glycosyltransferases (GTFs) responsible for the transfer of the sugars to the respective aglycones is quite scarce. So far, only two GTFs from antibiotic producers have been biochemically characterized in detail, and they are involved in macrolide inactivation: Mgt, from Streptomyces lividans, a nonmacrolide producer (7, 17); and OleD, from the oleandomycin producer Streptomyces antibioticus (15, 29), which inactivates oleandomycin by addition of glucose to the 2′-OH group of the desosamine attached to the macrolactone ring (40). In the last several years, a few genes have been proposed to encode GTFs involved in the transfer of sugars to various aglycones during biosynthesis: dnrS and dnrH, from Streptomyces peucetius, involved in daunorubicin (26) and baumycin (36) biosynthesis, respectively; gra-orf5, involved in granaticin biosynthesis (2); eryCIII and eryBV, involved in the transfer of desosamine and mycarose, respectively, in erythromycin biosynthesis (12, 32, 38); and tylM2, from Streptomyces fradiae, involved in sugar transfer during tylosin biosynthesis (14).Mithramycin (Fig. (Fig.1)1) is an aromatic polyketide which shows antibacterial activity against gram-positive bacteria and also antitumor activity (30, 37). Together with the chromomycins and the olivomycins, mithramycin constitutes the so-called aureolic acid group of antitumor drugs. The polyketide moiety of mithramycin is derived from the condensation of 10 acetate building blocks in a series of reactions catalyzed by a type II polyketide synthase (5, 21). The mithramycin aglycone is glycosylated at positions 6 and 2 with disaccharide (d-olivose- d-olivose) and trisaccharide (d-olivose-d-oliose-d-mycarose) moieties, respectively. All of these sugars belong to the 6-DOH family. In the mithramycin pathway, two genes (mtmD and mtmE) encoding two enzymes (glucose-1-phosphate:TTP thymidylyl transferase and dTDP-4,6-dehydratase, respectively) involved in the biosynthesis of the mithramycin 6-DOH have been cloned, and their participation in mithramycin biosynthesis has been demonstrated by insertional inactivation (22). Here we report the characterization of two Streptomyces argillaceus genes (mtmGI and mtmGII) that encode two putative GTFs responsible for the formation and transfer of the disaccharide chain. Inactivation of these genes by gene replacement showed identical accumulated compounds and allowed the isolation of four glycosylated compounds which are likely to be intermediates in mithramycin biosynthesis. Open in a separate windowFIG. 1Structures of mithramycin, premithramycinone, and the new premithramycins.  相似文献   

12.
Bacteria which were β-d-galactosidase and β-d-glucuronidase positive or expressed only one of these enzymes were isolated from environmental water samples. The enzymatic activity of these bacteria was measured in 25-min assays by using the fluorogenic substrates 4-methylumbelliferyl-β-d-galactoside and 4-methylumbelliferyl-β-d-glucuronide. The enzyme activity, enzyme induction, and enzyme temperature characteristics of target and nontarget bacteria in assays aimed at detecting coliform bacteria and Escherichia coli were investigated. The potential interference of false-positive bacteria was evaluated. Several of the β-d-galactosidase-positive nontarget bacteria but none of the β-d-glucuronidase-positive nontarget bacteria contained unstable enzyme at 44.5°C. The activity of target bacteria was highly inducible. Nontarget bacteria were induced much less or were not induced by the inducers used. The results revealed large variations in the enzyme levels of different β-d-galactosidase- and β-d-glucuronidase-positive bacteria. The induced and noninduced β-d-glucuronidase activities of Bacillus spp. and Aerococcus viridans were approximately the same as the activities of induced E. coli. Except for some isolates identified as Aeromonas spp., all of the induced and noninduced β-d-galactosidase-positive, noncoliform isolates exhibited at least 2 log units less mean β-d-galactosidase activity than induced E. coli. The noncoliform bacteria must be present in correspondingly higher concentrations than those of target bacteria to interfere in the rapid assay for detection of coliform bacteria.Indicators of pollution (e.g., coliforms, fecal coliforms, and Escherichia coli) are traditionally used for monitoring the microbiological safety of water supplies and recreational water. Several techniques for detection of coliforms and E. coli are based on enzymatic hydrolysis of fluorogenic or chromogenic substrates for β-d-galactosidase and β-d-glucuronidase (9, 20). Current methods of recovery are usually culture based, and the analysis time is 18 to 24 h. In addition to enzymatic activity, these techniques use growth at appropriate temperatures in the presence of inhibitors, combined with demonstration of enzymatic activity, to selectively detect target bacteria.Rapid methods which require less than 6 h and are based on chromogenic, fluorogenic, or chemiluminogenic substrates for detection of coliforms, fecal coliforms, or E. coli have been described (13, 10, 27, 28). These rapid assays are based on the assumption that β-d-galactosidase and β-d-glucuronidase are markers for coliforms and E. coli, respectively. However, when the incubation time is 1 h or less, growth is not a selective step, and all β-d-galactosidase-positive or β-d-glucuronidase-positive microorganisms in a water sample contribute to the activity measured. At low initial concentrations of target bacteria (i.e., E. coli and total coliforms), increasing the preincubation time to 5 to 6 h did not result in a predominance of target bacteria compared to nontarget bacteria (28).The β-d-galactosidase or β-d-glucuronidase activity calculated per cultivable coliform or fecal coliform bacterium in environmental samples can be 1 to 2 log units higher than the activity per induced E. coli cell in pure culture (11, 26). The presence of active, noncultivable bacteria can be one reason for this. Studies of survival (7, 24, 25) and disinfection (26) of E. coli have shown that loss of cultivability does not necessarily result in a loss of β-d-galactosidase activity. The presence of false-positive bacteria can be another reason.β-d-Galactosidase has been found in numerous microorganisms, including gram-negative bacteria (e.g., strains belonging to the Enterobacteriaceae, Vibrionaceae, Pseudomonadaceae, and Neisseriaceae), several gram-positive bacteria, yeasts, protozoa, and fungi (17, 29). β-d-Glucuronidase is produced by most E. coli strains and also by other members of the Enterobacteriaceae, including some Shigella and Salmonella strains and a few Yersinia, Citrobacter, Edwardia, and Hafnia strains. Production of β-d-glucuronidase by Flavobacterium spp., Bacteroides spp., Staphylococcus spp., Streptococcus spp., anaerobic corynebacteria, and Clostridium has also been reported (12).High numbers of false-positive bacteria in sewage and contaminated water have been revealed by enumeration of β-d-galactosidase- and β-d-glucuronidase-positive CFU on nonselective agar supplemented with fluorogenic or chromogenic substrates (11, 28). Whether the activity from nontarget organisms can be neglected in a rapid assay depends on the number of nontarget organisms compared with the number of target bacteria and also on the level of their enzyme activity. Plant and algal biomass must be present at high concentrations to interfere in rapid bacterial β-d-galactosidase and β-d-glucuronidase assays (8).The main objective of this study was to investigate the enzyme characteristics of β-d-galactosidase- and β-d-glucuronidase-positive bacteria isolated from environmental water samples and to evaluate the potential influence of false-positive bacteria in rapid assays for coliform bacteria or E. coli in water. The effect of temperature on enzyme activity and on the interference of nontarget bacteria in the rapid assays was investigated as an important factor.(Some of the results were presented at the 97th General Meeting of the American Society for Microbiology 1997, Miami Beach, Fla., 4 to 8 May 1997.)  相似文献   

13.
The recombinant industrial Saccharomyces cerevisiae strain MA-R5 was engineered to express NADP+-dependent xylitol dehydrogenase using the flocculent yeast strain IR-2, which has high xylulose-fermenting ability, and both xylose consumption and ethanol production remarkably increased. Furthermore, the MA-R5 strain produced the highest ethanol yield (0.48 g/g) from nonsulfuric acid hydrolysate of wood chips.Successful fermentation of lignocellulosic biomass to ethanol is dependent on efficient utilization of d-xylose, which is the second most common fermentable sugar in the hydrolysate. Although the well-known fermentative yeast Saccharomyces cerevisiae is one of the most effective ethanol-producing organisms for hexose sugars, it is not able to ferment d-xylose. However, S. cerevisiae can metabolize an isomerization product of d-xylose, d-xylulose, which is phosphorylated to d-xylulose 5-phosphate, channeled through the pentose phosphate pathway and glycolysis.S. cerevisiae transformed with the XYL1 and XYL2 genes encoding xylose reductase (XR) and xylitol dehydrogenase (XDH) from Pichia stipitis (referred to as PsXR and PsXDH, respectively) acquires the ability to ferment d-xylose to ethanol (2, 5, 6, 9, 10, 12, 22). Furthermore, overexpression of the XKS1 gene encoding xylulokinase (XK) from S. cerevisiae (ScXK) has been shown to aid d-xylose utilization (4, 7, 11, 16, 23), with xylitol still a major by-product. Kuyper et al. (14) also demonstrated the successful fermentation of d-xylose to ethanol using recombinant S. cerevisiae strains expressing fungal xylose isomerase. However, these approaches are insufficient for industrial bioprocesses, mainly due to the low rate of d-xylose fermentation.Xylitol excretion has been ascribed mainly to the difference in coenzyme specificities between PsXR (with NADPH) and PsXDH (with NAD+), which creates an intracellular redox imbalance (1). Therefore, modifying the coenzyme specificity of XR and/or XDH by protein engineering is an attractive approach for achieving efficient fermentation of ethanol from d-xylose using recombinant S. cerevisiae. Watanabe et al. (24) previously succeeded in generating several PsXDH mutants (e.g., quadruple ARSdR mutant) with a complete reversal of coenzyme specificity toward NADP+ by multiple site-directed mutagenesis on amino acids from the coenzyme-binding domain. The ARSdR mutant (D207A/I208R/F209S/N211R) has more that 4,500-fold-higher catalytic efficiency (kcat/Km) with NADP+ than the wild-type PsXDH. In addition, we recently found that several laboratory recombinant S. cerevisiae strains, in which the ARSdR mutant, along with PsXR and ScXK, is expressed through a strong constitutive promoter, increased ethanol production from d-xylose at the expense of xylitol excretion (17, 18), probably by maintaining the intracellular redox balance. However, commercialization of lignocellulosic hydrolysate fermentation requires industrial strains that are more robust than laboratory strains (5, 19, 21).A potential host for developing d-xylose-fermenting strains requires an active and efficient pentose phosphate pathway linking the d-xylose-to-d-xylulose pathway to glycolysis. In the case of S. cerevisiae, strains have different d-xylulose fermentation abilities (3, 25), indicating inherent differences in the capacities of these strains to ferment pentose sugars. Furthermore, anaerobic d-xylulose fermentation was investigated to identify genetic backgrounds potentially beneficial to anaerobic d-xylose fermentation in strains not exhibiting product formation related to the redox imbalance generated by the first two steps of the eukaryotic d-xylose metabolism (3), although the physiological purpose of the different d-xylulose fermentation abilities of S. cerevisiae is not yet clear. Therefore, we selected an efficient industrial d-xylulose-fermenting S. cerevisiae strain as a host for constructing a recombinant strain through chromosomal integration of the NADP+-dependent XDH gene and the XR and endogenous XK genes. Using this recombinant strain, we characterized the enzyme activity and ability to ferment both d-xylose and lignocellulosic hydrolysate.  相似文献   

14.
α-l-Arabinofuranosidases I and II were purified from the culture filtrate of Aspergillus awamori IFO 4033 and had molecular weights of 81,000 and 62,000 and pIs of 3.3 and 3.6, respectively. Both enzymes had an optimum pH of 4.0 and an optimum temperature of 60°C and exhibited stability at pH values from 3 to 7 and at temperatures up to 60°C. The enzymes released arabinose from p-nitrophenyl-α-l-arabinofuranoside, O-α-l-arabinofuranosyl-(1→3)-O-β-d-xylopyranosyl-(1→4)-d-xylopyranose, and arabinose-containing polysaccharides but not from O-β-d-xylopyranosyl-(1→2)-O-α-l-arabinofuranosyl-(1→3)-O-β-d-xylopyranosyl-(1→4)-O-β-d-xylopyranosyl-(1→4)-d-xylopyranose. α-l-Arabinofuranosidase I also released arabinose from O-β-d-xylopy-ranosyl-(1→4)-[O-α-l-arabinofuranosyl-(1→3)]-O-β-d-xylopyranosyl-(1→4)-d-xylopyranose. However, α-l-arabinofuranosidase II did not readily catalyze this hydrolysis reaction. α-l-Arabinofuranosidase I hydrolyzed all linkages that can occur between two α-l-arabinofuranosyl residues in the following order: (1→5) linkage > (1→3) linkage > (1→2) linkage. α-l-Arabinofuranosidase II hydrolyzed the linkages in the following order: (1→5) linkage > (1→2) linkage > (1→3) linkage. α-l-Arabinofuranosidase I preferentially hydrolyzed the (1→5) linkage of branched arabinotrisaccharide. On the other hand, α-l-arabinofuranosidase II preferentially hydrolyzed the (1→3) linkage in the same substrate. α-l-Arabinofuranosidase I released arabinose from the nonreducing terminus of arabinan, whereas α-l-arabinofuranosidase II preferentially hydrolyzed the arabinosyl side chain linkage of arabinan.Recently, it has been proven that l-arabinose selectively inhibits intestinal sucrase in a noncompetitive manner and reduces the glycemic response after sucrose ingestion in animals (33). Based on this observation, l-arabinose can be used as a physiologically functional sugar that inhibits sucrose digestion. Effective l-arabinose production is therefore important in the food industry. l-Arabinosyl residues are widely distributed in hemicelluloses, such as arabinan, arabinoxylan, gum arabic, and arabinogalactan, and the α-l-arabinofuranosidases (α-l-AFases) (EC 3.2.1.55) have proven to be essential tools for enzymatic degradation of hemicelluloses and structural studies of these compounds.α-l-AFases have been classified into two families of glycanases (families 51 and 54) on the basis of amino acid sequence similarities (11). The two families of α-l-AFases also differ in substrate specificity for arabinose-containing polysaccharides. Beldman et al. summarized the α-l-AFase classification based on substrate specificities (3). One group contains the Arafur A (family 51) enzymes, which exhibit very little or no activity with arabinose-containing polysaccharides. The other group contains the Arafur B (family 54) enzymes, which cleave arabinosyl side chains from polymers. However, this classification is too broad to define the substrate specificities of α-l-AFases. There have been many studies of the α-l-AFases (3, 12), especially the α-l-AFases of Aspergillus species (28, 1215, 17, 22, 23, 2832, 3639, 4143, 46). However, there have been only a few studies of the precise specificities of these α-l-AFases. In previous work, we elucidated the substrate specificities of α-l-AFases from Aspergillus niger 5-16 (17) and Bacillus subtilis 3-6 (16, 18), which should be classified in the Arafur A group and exhibit activity with arabinoxylooligosaccharides, synthetic methyl 2-O-, 3-O-, and 5-O-arabinofuranosyl-α-l-arabinofuranosides (arabinofuranobiosides) (20), and methyl 3,5-di-O-α-l-arabinofuranosyl-α-l-arabinofuranoside (arabinofuranotrioside) (19).In the present work, we purified two α-l-AFases from a culture filtrate of Aspergillus awamori IFO 4033 and determined the substrate specificities of these α-l-AFases by using arabinose-containing polysaccharides and the core oligosaccharides of arabinoxylan and arabinan.  相似文献   

15.
16.
A lysine racemase (lyr) gene was isolated from a soil metagenome by functional complementation for the first time by using Escherichia coli BCRC 51734 cells as the host and d-lysine as the selection agent. The lyr gene consisted of a 1,182-bp nucleotide sequence encoding a protein of 393 amino acids with a molecular mass of about 42.7 kDa. The enzyme exhibited higher specific activity toward lysine in the l-lysine-to-d-lysine direction than in the reverse reaction.Amino acids are the building blocks of proteins and play an important role in the regulation of the metabolism of living organisms. Among two enantiomers of naturally occurring amino acids, l-amino acids are predominant in living organisms, while d-amino acids are found in both free and bound states in various organisms like bacteria (36), yeasts (35), plants (47), insects (11), mammals (17), bivalves (39), and fish (28). The d-amino acids are mostly endogenous and produced by racemization from their counterparts by the action of a racemase. Thus, the amino acid racemases are involved in d-amino acid metabolism (29, 46). Since the discovery of alanine racemase in 1951 (42), several racemases toward amino acids, such as those for glutamate, threonine, serine, aspartate, methionine, proline, arginine, and phenylalanine, have been reported in bacteria, archaea, and eukaryotes, including mammals (1, 2, 15, 30, 31, 44). They are classified into two groups: pyridoxal 5′-phosphate (PLP)-dependent and PLP-independent enzymes (9, 36).Lysine racemase (Lyr, EC 5.1.1.5) was first reported in Proteus vulgaris ATCC 4669 (19) and proposed to be involved in the lysine degradation of bacterial cells (5, 19). Catabolism of lysine occurs via two parallel pathways. In one of the pathways, δ-aminovalerate is the key metabolite, whereas in the other l-lysine is racemized to d-lysine, and l-pipecolate and α-aminoadipate (AMA) are the key metabolites (5). d-Lysine catabolism proceeds through a series of cyclized intermediates which are necessary to regenerate an α-amino acid and comprise the following metabolites (AMA pathway): d-lysine→α-keto-ɛ-amino caproate→Δ1-piperideine-2-carboxylate→pipecolate→Δ1-piperideine-6-carboxylate→α-amino-δ-formylcaproate→α-AMA→α-ketoadipate (6, 7, 12, 27). The final product is converted to α-ketoglutarate via a series of coenzyme A derivatives and subsequently participates as an intermediate in the Krebs cycle. This pathway suggests that the biological function of d-lysine in the bacteria is that of a carbon or nitrogen source. Racemization of added l-lysine to d-lysine by whole cells of Proteus spp. and Escherichia spp. (19) and by the cell extract of Pseudomonas putida ATCC 15070 (5, 20) has been found. However, the enzyme has not been purified to homogeneity, and thus, its molecular and catalytic characteristics, including its gene structure, have not been elucidated. In this study, we explored a metagenomic library constructed from a garden soil to isolate a novel Lyr enzyme. After expression in Escherichia coli, the purified enzyme was characterized in terms of optimal pH and temperature, thermal stability, and racemization activity.  相似文献   

17.
Thermotoga maritima is a Gram-negative, hyperthermophilic bacterium whose peptidoglycan contains comparable amounts of l- and d-lysine. We have determined the fine structure of this cell-wall polymer. The muropeptides resulting from the digestion of peptidoglycan by mutanolysin were separated by high-performance liquid chromatography and identified by amino acid analysis after acid hydrolysis, dinitrophenylation, enzymatic determination of the configuration of the chiral amino acids, and mass spectrometry. The high-performance liquid chromatography profile contained four main peaks, two monomers, and two dimers, plus a few minor peaks corresponding to anhydro forms. The first monomer was the d-lysine-containing disaccharide-tripeptide in which the d-Glu-d-Lys bond had the unusual γ→ϵ arrangement (GlcNAc-MurNAc-l-Ala-γ-d-Glu-ϵ-d-Lys). The second monomer was the conventional disaccharide-tetrapeptide (GlcNAc-MurNAc-l-Ala-γ-d-Glu-l-Lys-d-Ala). The first dimer contained a disaccharide-l-Ala as the acyl donor cross-linked to the α-amine of d-Lys in a tripeptide acceptor stem with the sequence of the first monomer. In the second dimer, donor and acceptor stems with the sequences of the second and first monomers, respectively, were connected by a d-Ala4-α-d-Lys3 cross-link. The cross-linking index was 10 with an average chain length of 30 disaccharide units. The structure of the peptidoglycan of T. maritima revealed for the first time the key role of d-Lys in peptidoglycan synthesis, both as a surrogate of l-Lys or meso-diaminopimelic acid at the third position of peptide stems and in the formation of novel cross-links of the l-Ala1(α→α)d-Lys3 and d-Ala4(α→α)d-Lys3 types.Peptidoglycan (or murein) is a giant macromolecule whose main function is the protection of the cytoplasmic membrane against the internal osmotic pressure. It is composed of alternating residues of N-acetylglucosamine (GlcNAc) and N-acetylmuramic acid (MurNAc)2 cross-linked by short peptides (1). The composition of the peptide stem in nascent peptidoglycan is l-Ala1-γ-d-Glu2-X3-d-Ala4-d-Ala5, where X is most often meso-diaminopimelic acid (meso-A2pm) or l-lysine in Gram-negative and Gram-positive species, respectively (2, 3). In the mature macromolecule, the last d-Ala residue is removed. Cross-linking of the glycan chains generally occurs between the carboxyl group of d-Ala at position 4 of a donor peptide stem and the side-chain amino group of the diamino acid at position 3 of an acceptor peptide stem (4→3 cross-links). Cross-linking is either direct or through a short peptide bridge such as pentaglycine in Staphylococcus aureus (2, 3). The enzymes for the formation of the 4→3 cross-links are active-site serine dd- transpeptidases that belong to the penicillin-binding protein (PBP) family and are the essential targets of β-lactam antibiotics in pathogenic bacteria (4). Catalysis involves the cleavage of the d-Ala4-d-Ala5 bond of a donor peptide stem and the formation of an amide bond between the carboxyl of d-Ala4 and the side chain amine at the third position of an acceptor stem. Transpeptidases of the ld specificity are active-site cysteine enzymes that were shown to act as surrogates of the PBPs in mutants of Enterococcus faecium resistant to β-lactam antibiotics (5). They cleave the X3-d-Ala4 bond of a donor stem peptide to form 3→3 cross-links. This alternate mode of cross-linking is usually marginal, although it has recently been shown to predominate in non-replicative “dormant” forms of Mycobacterium tuberculosis (6).Thermotoga maritima is a Gram-negative, extremely thermophilic bacterium isolated from geothermally heated sea floors by Huber et al. (7). A morphological characteristic is the presence of an outer sheath-like envelope called “toga.” Although the organism has received considerable attention for its biotechnological potential, studies about its peptidoglycan are scarce (811), and in particular the fine structure of the macromolecule is still unknown. In their initial work, Huber et al. (7) showed that the composition of its peptidoglycan was unusual for a Gram-negative species, because it contained both isomers of lysine and no A2pm. Recently, we purified and studied the properties of T. maritima MurE (12); this enzyme is responsible for the addition of the amino acid residue at position 3 of the peptide stem (13, 14). We demonstrated that T. maritima MurE added in vitro l- and d-Lys to UDP-MurNAc-l-Ala-d-Glu. Although l-Lys was added in the usual way, yielding the conventional nucleotide UDP-MurNAc-l-Ala-γ-d-Glu-l-Lys containing a d-Glu(γ→α)l-Lys amide bond, the d-isomer was added in an “upside-down” manner, yielding the novel nucleotide UDP-MurNAc-l-Ala-d-Glu(γ→ϵ)d-Lys. We also showed that the d-Lys-containing nucleotide was not a substrate for T. maritima MurF, the subsequent enzyme in the biosynthetic pathway, whereas this ligase catalyzed the addition of dipeptide d-Ala-d-Ala to the l-Lys-containing tripeptide, yielding the conventional UDP-MurNAc-pentapeptide (12).However, both the l-Lys-containing UDP-MurNAc-pentapeptide and d-Lys-containing UDP-MurNAc-tripeptide were used as substrates by T. maritima MraY with comparable efficiencies in vitro (12). This observation implies that the unusual d-Lys-containing peptide stems are likely to be translocated to the periplasmic face of the cytoplasmic membrane and to participate in peptidoglycan polymerization. Therefore, we have determined here the fine structure of T. maritima peptidoglycan and we have shown that l-Lys- and d-Lys-containing peptide stems are both present in the polymer, the latter being involved in the formation of two novel types of peptidoglycan cross-link.  相似文献   

18.
19.
Bacillus anthracis elaborates a poly-γ-d-glutamic acid capsule that protects bacilli from phagocytic killing during infection. The enzyme CapD generates amide bonds with peptidoglycan cross-bridges to anchor capsular material within the cell wall envelope of B. anthracis. The capsular biosynthetic pathway is essential for virulence during anthrax infections and can be targeted for anti-infective inhibition with small molecules. Here, we present the crystal structures of the γ-glutamyltranspeptidase CapD with and without α-l-Glu-l-Glu dipeptide, a non-hydrolyzable analog of poly-γ-d-glutamic acid, in the active site. Purified CapD displays transpeptidation activity in vitro, and its structure reveals an active site broadly accessible for poly-γ-glutamate binding and processing. Using structural and biochemical information, we derive a mechanistic model for CapD catalysis whereby Pro427, Gly428, and Gly429 activate the catalytic residue of the enzyme, Thr352, and stabilize an oxyanion hole via main chain amide hydrogen bonds.Spores of Bacillus anthracis are the causative agents of anthrax disease (1). Upon entry into their hosts, spores germinate and replicate as vegetative bacilli (1). The formation of a thick capsule encasing vegetative forms enables bacilli to escape granulocyte0 and macrophage-mediated phagocytosis, and the pathogen eventually disseminates throughout all tissues of an infected host (2, 3). Bacilli secrete lethal and edema toxins, which cause macrophage necrosis and precipitate anthrax death (47). The genes providing for toxin and capsule formation are carried on two large virulence plasmids, pXO1 and pXO2, respectively (8, 9). Loss of any one plasmid leads to virulence attenuation, a feature that has been exploited for the generation of vaccine-type strains (1014).Unlike polysaccharide-based capsules that are commonly found in bacterial pathogens, the capsular material of B. anthracis is composed of poly-γ-d-glutamic acid (PDGA)3 (3). All the genes necessary for capsule biogenesis are located in the capBCADE gene cluster on plasmid pXO2 (1519). CapD is the only protein of this cluster that is located on the bacterial surface (16). CapD shares sequence similarity with bacterial and mammalian γ-glutamyl transpeptidases (GGTs; EC 2.3.2.2) (17). GGTs belong to the N-terminal nucleophile hydrolases (Ntn) family (Protein Structure Classification (Class (C), Architecture (A), Topology (T) and Homologous superfamily (H)) (CATH) id 3.60.60.10) (20). These enzymes assemble as a single polypeptide chain and acquire activity by undergoing autocatalytic processing to heterodimer.Bacterial GGTs catalyze the first step in glutathione degradation. For example, Helicobacter pylori GGT removes glutamate from glutathione tripeptide via the formation of a γ-glutamyl acyl enzyme. This intermediate is resolved by the nucleophilic attack of a water molecule, causing the release of γ-glutamate (21, 22). Mammalian enzymes transfer the γ-glutamyl intermediate to the amino group of a peptide, thereby completing a transpeptidation reaction (23). The B. anthracis CapD precursor is also programmed for autocatalytic cleavage (17). Similar to mammalian GGTs, CapD also catalyzes a transpeptidation reaction; however, this reaction promotes the covalent linkage of PDGA to the bacterial envelope (16, 24). We have recently demonstrated the cell wall anchor structure of capsule filaments in the envelope of B. anthracis, identifying an amide bond between the terminal carboxyl group of PDGA and the side amino group of m-diaminopimelic acid cross-bridges within muropeptides (24). The CapD-catalyzed transpeptidation reaction could be recapitulated in vitro using purified recombinant CapD, γ-d-Glun peptide, and muropeptide substrates (24). In the absence of the physiological nucleophile (muropeptides), CapD acyl intermediates can be resolved by the nucleophilic attack of water to generate hydrolysis products.Here, we report the high resolution crystal structure of CapD in the absence and presence of a glutamate dipeptide and compare it with the known structures of H. pylori and Escherichia coli GGTs. By combining structural, genetic, and biochemical approaches, we identify the unique features of CapD that distinguish the protein from GGTs and detect several residues that are important for CapD autocatalytic cleavage and PDGA processing. This structural information will further the development of small molecule inhibitors that disrupt CapD activity and that may be useful as anti-infective therapies for anthrax.  相似文献   

20.
A major question about cytokinesis concerns the role of the septin proteins, which localize to the division site in all animal and fungal cells but are essential for cytokinesis only in some cell types. For example, in Schizosaccharomyces pombe, four septins localize to the division site, but deletion of the four genes produces only a modest delay in cell separation. To ask if the S. pombe septins function redundantly in cytokinesis, we conducted a synthetic-lethal screen in a septin-deficient strain and identified seven mutations. One mutation affects Cdc4, a myosin light chain that is an essential component of the cytokinetic actomyosin ring. Five others cause frequent cell lysis during cell separation and map to two loci. These mutations and their dosage suppressors define a signaling pathway (including Rho1 and a novel arrestin) for repairing cell-wall damage. The seventh mutation affects the poorly understood RNA-binding protein Scw1 and severely delays cell separation when combined either with a septin mutation or with a mutation affecting the septin-interacting, anillin-like protein Mid2, suggesting that Scw1 functions in a pathway parallel to that of the septins. Taken together, our results suggest that the S. pombe septins participate redundantly in one or more pathways that cooperate with the actomyosin ring during cytokinesis and that a septin defect causes septum defects that can be repaired effectively only when the cell-integrity pathway is intact.THE fission yeast Schizosaccharomyces pombe provides an outstanding model system for studies of cytokinesis (McCollum and Gould 2001; Balasubramanian et al. 2004; Pollard and Wu 2010). As in most animal cells, successful cytokinesis in S. pombe requires an actomyosin ring (AMR). The AMR begins to assemble at the G2/M transition and involves the type II myosin heavy chains Myo2 and Myp2 and the light chains Cdc4 and Rlc1 (Wu et al. 2003). Myo2 and Cdc4 are essential for cytokinesis under all known conditions, Rlc1 is important at all temperatures but essential only at low temperatures, and Myp2 is essential only under stress conditions. As the AMR constricts, a septum of cell wall is formed between the daughter cells. The primary septum is sandwiched by secondary septa and subsequently digested to allow cell separation (Humbel et al. 2001; Sipiczki 2007). Because of the internal turgor pressure of the cells, the proper assembly and structural integrity of the septal layers are essential for cell survival.Septum formation involves the β-glucan synthases Bgs1/Cps1/Drc1, Bgs3, and Bgs4 (Ishiguro et al. 1997; Le Goff et al. 1999; Liu et al. 1999, 2002; Martín et al. 2003; Cortés et al. 2005) and the α-glucan synthase Ags1/Mok1 (Hochstenbach et al. 1998; Katayama et al. 1999). These synthases are regulated by the Rho GTPases Rho1 and Rho2 and the protein kinase C isoforms Pck1 and Pck2 (Arellano et al. 1996, 1997, 1999; Nakano et al. 1997; Hirata et al. 1998; Calonge et al. 2000; Sayers et al. 2000; Ma et al. 2006; Barba et al. 2008; García et al. 2009b). The Rho GTPases themselves appear to be regulated by both GTPase-activating proteins (GAPs) and guanine-nucleotide-exchange factors (GEFs) (Nakano et al. 2001; Calonge et al. 2003; Iwaki et al. 2003; Tajadura et al. 2004; Morrell-Falvey et al. 2005; Mutoh et al. 2005; García et al. 2006, 2009a,b). In addition, septum formation and AMR function appear to be interdependent. In the absence of a normal AMR, cells form aberrant septa and/or deposit septal materials at random locations, whereas a mutant defective in septum formation (bgs1) is also defective in AMR constriction (Gould and Simanis 1997; Le Goff et al. 1999; Liu et al. 1999, 2000). Both AMR constriction and septum formation also depend on the septation initiation network involving the small GTPase Spg1 (McCollum and Gould 2001; Krapp and Simanis 2008). Despite this considerable progress, many questions remain about the mechanisms and regulation of septum formation and its relationships to the function of the AMR.One major question concerns the role(s) of the septins. Proteins of this family are ubiquitous in fungal and animal cells and typically localize to the cell cortex, where they appear to serve as scaffolds and diffusion barriers for other proteins that participate in a wide variety of cellular processes (Longtine et al. 1996; Gladfelter et al. 2001; Hall et al. 2008; Caudron and Barral 2009). Despite the recent progress in elucidating the mechanisms of septin assembly (John et al. 2007; Sirajuddin et al. 2007; Bertin et al. 2008; McMurray and Thorner 2008), the details of septin function remain obscure. However, one prominent role of the septins and associated proteins is in cytokinesis. Septins concentrate at the division site in every cell type that has been examined, and in Saccharomyces cerevisiae (Hartwell 1971; Longtine et al. 1996; Lippincott et al. 2001; Dobbelaere and Barral 2004) and at least some Drosophila (Neufeld and Rubin 1994; Adam et al. 2000) and mammalian (Kinoshita et al. 1997; Surka et al. 2002) cell types, the septins are essential for cytokinesis. In S. cerevisiae, the septins are required for formation of the AMR (Bi et al. 1998; Lippincott and Li 1998). However, this cannot be their only role, because the AMR itself is not essential for cytokinesis in this organism (Bi et al. 1998; Korinek et al. 2000; Schmidt et al. 2002). Moreover, there is no evidence that the septins are necessary for AMR formation or function in any other organism. A further complication is that in some cell types, including most Caenorhabditis elegans cells (Nguyen et al. 2000; Maddox et al. 2007) and some Drosophila cells (Adam et al. 2000; Field et al. 2008), the septins do not appear to be essential for cytokinesis even though they localize to the division site.S. pombe has seven septins, four of which (Spn1, Spn2, Spn3, and Spn4) are expressed in vegetative cells and localize to the division site shortly before AMR constriction and septum formation (Longtine et al. 1996; Berlin et al. 2003; Tasto et al. 2003; Wu et al. 2003; An et al. 2004; Petit et al. 2005; Pan et al. 2007; Onishi et al. 2010). Spn1 and Spn4 appear to be the core members of the septin complex (An et al. 2004; McMurray and Thorner 2008), and mutants lacking either of these proteins do not assemble the others at the division site. Assembly of a normal septin ring also depends on the anillin-like protein Mid2, which colocalizes with the septins (Berlin et al. 2003; Tasto et al. 2003). Surprisingly, mutants lacking the septins are viable and form seemingly complete septa with approximately normal timing. These mutants do, however, display a variable delay in separation of the daughter cells, suggesting that the septins play some role(s) in the proper completion of the septum or in subsequent processes necessary for cell separation (Longtine et al. 1996; An et al. 2004; Martín-Cuadrado et al. 2005).It is possible that the septins localize to the division site and yet are nonessential for division in some cell types because their role is redundant with that of some other protein(s) or pathway(s). To explore this possibility in S. pombe, we screened for mutations that were lethal in combination with a lack of septins. The results suggest that the septins cooperate with the AMR during cytokinesis and that, in the absence of septin function, the septum is not formed properly, so that an intact system for recognizing and repairing cell-wall damage becomes critical for cell survival.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号