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 46 毫秒
1.
2.
3.
All of the 2,6-dideoxy sugars contained within the structure of chromomycin A3 are derived from d-glucose. Enzyme assays were used to confirm the presence of hexokinase, phosphoglucomutase, UDPG pyrophosphorylase (UDPGP), and UDPG oxidoreductase (UDPGO), all of which are involved in the pathway of glucose activation and conversion into 2,6-dideoxyhexoses during chromomycin biosynthesis. Levels of the four enzymes in Streptomyces spp. cell extracts were correlated with the production of chromomycins. The pathway of sugar activation in Streptomyces spp. involves glucose 6-phosphorylation by hexokinase, isomerization to G-1-P catalyzed by phosphoglucomutase, synthesis of UDPG catalyzed by UDPGP, and formation of UDP-4-keto-6-deoxyglucose by UDPGO.Dideoxy sugars occur commonly in the structures of cardiac glycosides from plants, in antibiotics like chromomycin A3 (Fig. (Fig.1),1), and in macrolides produced by microorganisms. On the basis of stable isotope-labeling experiments, biosynthetic studies conducted in Rosazza’s laboratory have indicated that all the deoxy sugars of chromomycin A3 are derived from d-glucose (21). While the assembly of the polyketide aglycone is reasonably well understood, relatively little is known of the details of 2,6-dideoxy sugar biogenesis in streptomycetes. Earlier studies with Streptomyces rimosus indicated that TDP-mycarose is synthesized from TDP-d-glucose (TDPG) and S-adenosyl-l-methionine (10, 23). The reaction requires NADPH as a cofactor, and TDP-4-keto-6-deoxy-d-glucose is an intermediate. Formation of TDP-4-keto-6-deoxy-d-glucose was catalyzed by the enzyme TDPG oxidoreductase (TDPG-4,6-dehydratase; EC 4.2.1.46). Similar 4-keto sugar nucleotides are intermediates for the biosynthesis of polyene macrolide antibiotic amino sugars (18). Similar pathways have been elaborated for the formation of 2,6-dideoxy-d-threo-4-hexulose of granaticin in Escherichia coli (6, 25) and 2,6-dideoxy-d-arabino-hexose of chlorothricin (12). The initial 6-deoxygenation of glucose during 3,6-dideoxy sugar formation involves a similar mechanism (32). In all of these processes, glucose is first activated by conversion into a sugar nucleotide such as UDPG followed by NAD+ oxidation of the 4 position to the corresponding 4-oxo derivative. Position 6 deoxygenation involves a general tautomerization, dehydration, and NADH,H+-catalyzed reduction process (6, 12, 25). A similar tautomerization and dehydration followed by reduction may produce C-3-deoxygenated products, such as CDP-3,6-dideoxyglucose (27). The pathway for formation of 3,6-dideoxyhexoses from CDPG in Yersinia pseudotuberculosis was clearly elucidated by Liu and Thorson (14). However, none of this elegant work was focused on the earlier steps of hexose nucleotide formation. Open in a separate windowFIG. 1Structures of chromomycins A2 and A3.On the basis of previous work (7), it is reasonable to postulate that the biosynthesis of 2,6-dideoxyglucose in Streptomyces griseus involves phosphorylation to glucose-6-phosphate by hexokinase (HK; E.C.2.7.7.1), as in glycolysis; conversion to glucose-1-phosphate by phosphoglucomutase (PGM; EC 2.7.5.1); reaction with UTP to form UDPG in a reaction catalyzed by UDPG pyrophosphorylase (UDPGP) (glucose-1-phosphate uridylyltransferase; EC 2.7.7.9), and C-6 deoxygenation catalyzed by UDP-d-glucose-4,6-dehydratase with NAD+ as a cofactor (Fig. (Fig.2).2). UDPG and GDPG have been detected in cell extracts of S. griseus and Streptomyces sp. strain MRS202, suggesting that these compounds are active sugar nucleotides involved in the formation of dideoxyhexoses (15). UDPGP genes from several bacteria have been cloned and sequenced (1, 3, 4, 11, 29, 30). Although nucleotidyl diphosphohexose-4,6-dehydratases (NDP-hexose-4,6-dehydratases) have been purified and characterized from several sources (5, 8, 9, 13, 19, 25, 26, 31, 33), the occurrence of the glucose-activating enzymes HK, PGM, UDPGP, and UDPG oxidoreductase (UDPGO) involved in 2,6-dideoxyhexose formation has not been established in streptomycetes. This work provides evidence for the presence of these enzymes involved in the biosynthetic activation of glucose to the 2,6-dideoxyhexoses in chromomycin A3.Open in a separate windowFIG. 2Proposed pathway for the formation of 2,6-dideoxy sugars in streptomycetes involving HK, PGM, UDPGP, and UDPGO.  相似文献   

4.
Hyperhomocysteinemia has long been associated with atherosclerosis and thrombosis and is an independent risk factor for cardiovascular disease. Its causes include both genetic and environmental factors. Although homocysteine is produced in every cell as an intermediate of the methionine cycle, the liver contributes the major portion found in circulation, and fatty liver is a common finding in homocystinuric patients. To understand the spectrum of proteins and associated pathways affected by hyperhomocysteinemia, we analyzed the mouse liver proteome of gene-induced (cystathionine β-synthase (CBS)) and diet-induced (high methionine) hyperhomocysteinemic mice using two-dimensional difference gel electrophoresis and Ingenuity Pathway Analysis. Nine proteins were identified whose expression was significantly changed by 2-fold (p ≤ 0.05) as a result of genotype, 27 proteins were changed as a result of diet, and 14 proteins were changed in response to genotype and diet. Importantly, three enzymes of the methionine cycle were up-regulated. S-Adenosylhomocysteine hydrolase increased in response to genotype and/or diet, whereas glycine N-methyltransferase and betaine-homocysteine methyltransferase only increased in response to diet. The antioxidant proteins peroxiredoxins 1 and 2 increased in wild-type mice fed the high methionine diet but not in the CBS mutants, suggesting a dysregulation in the antioxidant capacity of those animals. Furthermore, thioredoxin 1 decreased in both wild-type and CBS mutants on the diet but not in the mutants fed a control diet. Several urea cycle proteins increased in both diet groups; however, arginase 1 decreased in the CBS+/− mice fed the control diet. Pathway analysis identified the retinoid X receptor signaling pathway as the top ranked network associated with the CBS+/− genotype, whereas xenobiotic metabolism and the NRF2-mediated oxidative stress response were associated with the high methionine diet. Our results show that hyperhomocysteinemia, whether caused by a genetic mutation or diet, alters the abundance of several liver proteins involved in homocysteine/methionine metabolism, the urea cycle, and antioxidant defense.Homocysteine (Hcy)1 is a thiol-containing amino acid that is produced in every cell of the body as an intermediate of the methionine cycle (Fig. 1, Reactions 1–5) (1). Once formed, the catabolism of homocysteine occurs via three enzymatic pathways. 1) Hcy is remethylated back to methionine using vitamin B12-dependent methionine synthase (Fig. 1, Reaction 4) and/or 2) betaine-homocysteine methyltransferase (BHMT) (Fig. 1, Reaction 5), and 3) Hcy is converted to cysteine via the transsulfuration pathway using CBS and γ-cystathionase (Fig. 1, Reactions 6 and 7). Under normal conditions ∼40–50% of the Hcy that is produced in the liver is remethylated, ∼40–50% is converted to cysteine, and a small amount is exported (13). However, when Hcy production is increased (i.e. increased dietary methionine/protein intake) or when Hcy catabolism is decreased (i.e. CBS deficiency or B vitamin deficiencies), excess Hcy is exported into the extracellular space, resulting in hyperhomocysteinemia (15).Open in a separate windowFig. 1.Homocysteine metabolism in liver and kidney. In classical homocystinuria, the initial enzyme of the transsulfuration pathway, CBS (Reaction 6), is deficient. MTHF, methylenetetrahydrofolate; THF, tetrahydrofolate; DHF, dihydrofolate; MeCbl, methylcobalamin; DMG, dimethylglycine; PLP, pyridoxal 5′-phosphate.Homocystinuria was first described in the 1960s by Carson et al. (6): they observed 10 pediatric patients with severely elevated levels of Hcy in the urine and hypermethioninemia. Normal concentrations of plasma total homocysteine (tHcy) range from 5 to 12 μm (7); however, in homocystinuria, tHcy levels can exceed 100 μm. Homocystinuric patients present with mental retardation, abnormal bone growth, fine hair, malar flush, and dislocation of the lens of the eye, and most die from premature cardiovascular disease (6, 8). Autopsy findings indicate widespread thromboembolism, arteriosclerosis, and fatty livers (6, 8). Mudd et al. (9, 10) identified the cause of homocystinuria as a defect in the enzyme cystathionine β-synthase. A recent study of newborn infants in Denmark estimated the birth prevalence for CBS heterozygosity to be about 1:20,000 (11).Plasma tHcy concentrations are also directly correlated with dietary methionine/protein intake (12, 13). Guttormsen et al. (13) demonstrated that a protein-rich meal affected tHcy for at least 8–24 h. When normal subjects were fed a low protein-containing breakfast (12–15 g), plasma methionine levels increased slightly after 2 h (22.5–27.5 μm), but tHcy levels did not change significantly. However, when these same subjects were fed a high protein meal (52 g), plasma methionine levels peaked after 4 h (38 μm), and tHcy rose steadily until a maximum level was reached 8 h postmeal (7.6 versus 8.5 μm) (13). Thus, the following questions can be raised. How does the hepatic proteome respond to a hyperhomocysteinemic diet, and are the changes that accompany such a diet the same as or different from those that may be observed in gene-induced hyperhomocysteinemia?Because hyperhomocysteinemia is a strong independent risk factor for cardiovascular, cerebrovascular, and peripheral vascular disease, most of the current research has focused on the mechanisms involved in Hcy-induced endothelial dysfunction (1424). The results of those studies have concluded that Hcy induces intracellular oxidative stress by generating ROS, which in turn lead to decreased bioavailable nitric oxide (NO), altered gene expression, increased endoplasmic reticulum stress, and activation of cholesterol biosynthesis. Also, several studies have examined the association between hyperhomocysteinemia and alcoholic liver disease, but few have looked at the effect of Hcy on the non-alcoholic liver even though fatty liver is a constant finding in homocystinuria (6, 8), and the liver is the major source of circulating Hcy (4, 5, 10). We hypothesize that 1) the liver proteome will respond to hyperhomocysteinemia by altering the expression of proteins involved in methionine/homocysteine metabolism and antioxidant defense and that 2) the set of proteins that are expressed when hyperhomocysteinemia is induced by CBS deficiency will differ from those expressed as a result of a high methionine diet. In the present study, we use a well established mouse model of CBS deficiency to study the early changes in the liver proteome that accompany hyperhomocysteinemia (25).  相似文献   

5.
6.
7.
Folding and stability are parameters that control protein behavior. The possibility of conferring additional stability on proteins has implications for their use in vivo and for their structural analysis in the laboratory. Cyclic polypeptides ranging in size from 14 to 78 amino acids occur naturally and often show enhanced resistance toward denaturation and proteolysis when compared with their linear counterparts. Native chemical ligation and intein-based methods allow production of circular derivatives of larger proteins, resulting in improved stability and refolding properties. Here we show that circular proteins can be made reversibly with excellent efficiency by means of a sortase-catalyzed cyclization reaction, requiring only minimal modification of the protein to be circularized.Sortases are bacterial enzymes that predominantly catalyze the attachment of surface proteins to the bacterial cell wall (1, 2). Other sortases polymerize pilin subunits for the construction of the covalently stabilized and covalently anchored pilus of the Gram-positive bacterium (35). The reaction catalyzed by sortase involves the recognition of short 5-residue sequence motifs, which are cleaved by the enzyme with the concomitant formation of an acyl enzyme intermediate between the active site cysteine of sortase and the carboxylate at the newly generated C terminus of the substrate (1, 68). In many bacteria, this covalent intermediate can be resolved by nucleophilic attack from the pentaglycine side chain in a peptidoglycan precursor, resulting in the formation of an amide bond between the pentaglycine side chain and the carboxylate at the cleavage site in the substrate (9, 10). In pilus construction, alternative nucleophiles such as lysine residues or diaminopimelic acid participate in the transpeptidation reaction (3, 4).When appended near the C terminus of proteins that are not natural sortase substrates, the recognition sequence of Staphylococcus aureus sortase A (LPXTG) can be used to effectuate a sortase-catalyzed transpeptidation reaction using a diverse array of artificial glycine-based nucleophiles (Fig. 1). The result is efficient installation of a diverse set of moieties, including lipids (11), carbohydrates (12), peptide nucleic acids (13), biotin (14), fluorophores (14, 15), polymers (16), solid supports (1618), or peptides (15, 19) at the C terminus of the protein substrate. During the course of our studies to further expand sortase-based protein engineering, we were struck by the frequency and relative ease with which intramolecular transpeptidation reactions were occurring. Specifically, proteins equipped with not only the LPXTG motif but also N-terminal glycine residues yielded covalently closed circular polypeptides (Fig. 1). Similar reactivity using sortase has been described in two previous cases; however, rigorous characterization of the circular polypeptides was absent (16, 20). The circular proteins in these reports were observed as minor components of more complex reaction mixtures, and the cyclization reaction itself was not optimized.Open in a separate windowFIGURE 1.Protein substrates equipped with a sortase A recognition sequence (LPXTG) can participate in intermolecular transpeptidation with synthetic oligoglycine nucleophiles (left) or intramolecular transpeptidation if an N-terminal glycine residue is present (right).Here we describe our efforts toward applying sortase-catalyzed transpeptidation to the synthesis of circular and oligomeric proteins. This method has general applicability, as illustrated by successful intramolecular reactions with three structurally unrelated proteins. In addition to circularization of individual protein units, the multiprotein complex AAA-ATPase p97/VCP/CDC48, with six identical subunits containing the LPXTG motif and an N-terminal glycine, was found to preferentially react in daisy chain fashion to yield linear protein fusions. The reaction exploited here shows remarkable similarities to the mechanisms proposed for circularization of cyclotides, small circular proteins that have been isolated from plants (2123).  相似文献   

8.
Apoptotic caspases, such as caspase-7, are stored as inactive protease zymogens, and when activated, lead to a fate-determining switch to induce cell death. We previously discovered small molecule thiol-containing inhibitors that when tethered revealed an allosteric site and trapped a conformation similar to the zymogen form of the enzyme. We noted three structural transitions that the compounds induced: (i) breaking of an interaction between Tyr-223 and Arg-187 in the allosteric site, which prevents proper ordering of the catalytic cysteine; (ii) pinning the L2′ loop over the allosteric site, which blocks critical interactions for proper ordering of the substrate-binding groove; and (iii) a hinge-like rotation at Gly-188 positioned after the catalytic Cys-186 and Arg-187. Here we report a systematic mutational analysis of these regions to dissect their functional importance to mediate the allosteric transition induced by these compounds. Mutating the hinge Gly-188 to the restrictive proline causes a massive ∼6000-fold reduction in catalytic efficiency. Mutations in the Arg-187–Tyr-223 couple have a far less dramatic effect (3–20-fold reductions). Interestingly, although the allosteric couple mutants still allow binding and allosteric inhibition, they partially relieve the mutual exclusivity of binding between inhibitors at the active and allosteric sites. These data highlight a small set of residues critical for mediating the transition from active to inactive zymogen-like states.Caspases are a family of dimeric cysteine proteases whose members control the ultimate steps for apoptosis (programmed cell death) or innate inflammation among others (for reviews, see Refs. 1 and 2). During apoptosis, the upstream initiator caspases (caspase-8 and -9) activate the downstream executioner caspases (caspase-3, -6, and-7) via zymogen maturation (3). The activated executioner caspases then cleave upwards of 500 key proteins (46) and DNA, leading to cell death. Due to their pivotal role in apoptosis, the caspases are involved both in embryonic development and in dysfunction in diseases including cancer and stroke (7). The 11 human caspases share a common active site cysteine-histidine dyad (8), and derive their name, cysteine aspartate proteases, from their exquisite specificity for cleaving substrate proteins after specific aspartate residues (913). Thus, it has been difficult to develop active site-directed inhibitors with significant specificity for one caspase over the others (14). Despite difficulties in obtaining specificity, there has been a long-standing correlation between efficacy of caspase inhibitors in vitro and their ability to inhibit caspases and apoptosis in vivo (for review, see Ref. 31). Thus, a clear understanding of in vitro inhibitor function is central to the ability control caspase function in vivo.Caspase-7 has been a paradigm for understanding the structure and dynamics of the executioner caspases (1521). The substrate-binding site is composed of four loops; L2, L3, and L4 are contributed from one-half of the caspase dimer, and L2′ is contributed from the other half of the caspase dimer (Fig. 1). These loops appear highly dynamic as they are only observed in x-ray structures when bound to substrate or substrate analogs in the catalytically competent conformation (1719, 22) (Fig. 1B).Open in a separate windowFIGURE 1.Allosteric site and dimeric structure in caspase-7. A, the surface of active site-bound caspase-7 shows a large open allosteric (yellow) site at the dimer interface. This cavity is distinct from the active sites, which are bound with the active site inhibitor DEVD (green sticks). B, large subunits of caspase-7 dimers (dark green and dark purple) contain the active site cysteine-histidine dyad. The small subunits (light green and light purple) contain the allosteric site cysteine 290. The conformation of the substrate-binding loops (L2, L2′, L3, and L4) in active caspase-7 (Protein Data Bank (PDB) number 1f1j) is depicted. The L2′ loop (spheres) from one-half of the dimer interacts with the L2 loop from the other half of the dimer. C, binding of allosteric inhibitors influences the conformation of the L2′ loop (spheres), which folds over the allosteric cavity (PDB number 1shj). Subunit rendering is as in panel A. Panels A, B, and C are in the same orientation.A potential alternative to active site inhibitors are allosteric inhibitors that have been seeded by the discovery of selective cysteine-tethered allosteric inhibitors for either apoptotic executioner caspase-3 or apoptotic executioner caspase-7 (23) as well as the inflammatory caspase-1 (24). These thiol-containing compounds bind to a putative allosteric site through disulfide bond formation with a thiol in the cavity at the dimer interface (Fig. 1A) (23, 24). X-ray structures of caspase-7 bound to allosteric inhibitors FICA3 and DICA (Fig. 2) show that these compounds trigger conformational rearrangements that stabilize the inactive zymogen-like conformation over the substrate-bound, active conformation. The ability of small molecules to hold mature caspase-7 in a conformation that mimics the naturally occurring, inactive zymogen state underscores the utility and biological relevance of the allosteric mechanism of inhibition. Several structural changes are evident between these allosterically inhibited and active states. (i) The allosteric inhibitors directly disrupt an interaction between Arg-187 (next to the catalytic Cys-186) and Tyr-223 that springs the Arg-187 into the active site (Fig. 3), (ii) this conformational change appears to be facilitated by a hinge-like movement about Gly-188, and (iii) the L2′ loop folds down to cover the allosteric inhibitor and assumes a zymogen-like conformation (Fig. 1C) (23).Open in a separate windowFIGURE 2.Structure of allosteric inhibitors DICA and FICA. DICA and FICA are hydrophobic small molecules that bind to an allosteric site at the dimer interface of caspase-7. Binding of DICA/FICA is mediated by a disulfide between the compound thiol and Cys-290 in caspase-7.Open in a separate windowFIGURE 3.Movement of L2′ blocking arm. The region of caspase-7 encompassing the allosteric couple Arg-187 and Tyr-223 is boxed. The inset shows the down orientation of Arg-187 and Tyr-223 in the active conformation with DEVD substrate mimic (orange spheres) in the active site. In the allosteric/zymogen conformation, Arg-187 and Tyr-223 are pushed up by DICA (blue spheres).Here, using mutational analysis and small molecule inhibitors, we assess the importance of these three structural units to modulate both the inhibition of the enzyme and the coupling between allosteric and active site labeling. Our data suggest that the hinge movement and pinning of the L2-L2′ are most critical for transitioning between the active and inactive forms of the enzyme.  相似文献   

9.
10.
11.
Mammalian glutamate dehydrogenase (GDH) is a homohexameric enzyme that catalyzes the reversible oxidative deamination of l-glutamate to 2-oxoglutarate using NAD(P)+ as coenzyme. Unlike its counterparts from other animal kingdoms, mammalian GDH is regulated by a host of ligands. The recently discovered hyperinsulinism/hyperammonemia disorder showed that the loss of allosteric inhibition of GDH by GTP causes excessive secretion of insulin. Subsequent studies demonstrated that wild-type and hyperinsulinemia/hyperammonemia forms of GDH are inhibited by the green tea polyphenols, epigallocatechin gallate and epicatechin gallate. This was followed by high throughput studies that identified more stable inhibitors, including hexachlorophene, GW5074, and bithionol. Shown here are the structures of GDH complexed with these three compounds. Hexachlorophene forms a ring around the internal cavity in GDH through aromatic stacking interactions between the drug and GDH as well as between the drug molecules themselves. In contrast, GW5074 and bithionol both bind as pairs of stacked compounds at hexameric 2-fold axes between the dimers of subunits. The internal core of GDH contracts when the catalytic cleft closes during enzymatic turnover. None of the drugs cause conformational changes in the contact residues, but all bind to key interfaces involved in this contraction process. Therefore, it seems likely that the drugs inhibit enzymatic turnover by inhibiting this transition. Indeed, this expansion/contraction process may play a major role in the inter-subunit communication and allosteric regulation observed in GDH.Glutamate dehydrogenase (GDH)2 is found in all living organisms and catalyzes the reversible oxidative deamination of l-glutamate to 2-oxoglutarate using NAD(P)+ as coenzyme (1). In eukaryotic organisms, GDH resides within the inner mitochondrial matrix where it catabolizes glutamate to feed 2-oxoglutarate to the Krebs cycle. Although there is some debate as to the directionality of the reaction, the high Km for ammonium in the reductive amination reaction seems to prohibit the reverse reaction under normal conditions in most organisms (2). GDH from animals, but not other kingdoms (3), is allosterically regulated by a wide array of ligands (39). GTP (911), and with ∼100-fold lower affinity, ATP (3), is a potent inhibitor of the reaction and acts by increasing the binding affinity for the product, thereby slowing down enzymatic turnover (11). Hydrophobic compounds such as palmitoyl-CoA (12), steroid hormones (13), and steroid hormone analogs such as diethylstilbestrol (5) are also potent inhibitors. ADP is an activator of GDH (3, 6, 10, 11, 14) that acts in an opposite manner to GTP by facilitating product release. Leucine is a poor substrate for GDH but is also an allosteric activator for the enzyme (8). Its activation is akin to ADP but acts at site distinct from ADP (15).The crystal structures of the bacterial (1618) and animal forms (19, 20) of GDH have shown that the general architecture and the locations of the catalytically important residues have remained unchanged throughout evolution. The structure of GDH (Fig. 1) is essentially two trimers of subunits stacked directly on top of each other with each subunit being composed of at least three domains (1922). The bottom domain makes extensive contacts with a subunit from the other trimer. Resting on top of this domain is the “NAD binding domain” that has the conserved nucleotide binding motif. Animal GDH has a long protrusion, “antenna,” rising above the NAD binding domain that is not found in bacteria, plants, fungi, and the vast majority of protists. The antenna from each subunit lies immediately behind the adjacent, counterclockwise neighbor within the trimer. Because these intertwined antennae are only found in the forms of GDH that are allosterically regulated by numerous ligands, it is reasonable to speculate that it plays a role in regulation.Open in a separate windowFIGURE 1.Conformational transitions and locations of ligand binding sites in bovine glutamate dehydrogenase. A, a ribbon diagram of apo-bovine glutamate dehydrogenase with each of the identical subunits represented by different colors. The subunit arrangement is that of a trimer of dimers where anti-parallel β-strands form extensive interactions between the subunits stacked on top of each other. This pairing is represented by different shades of the same color. The conformational changes that during substrate binding are shown by the numbered arrows. As substrate binds, the NAD+ binding domain closes (1). The ascending helix of the antenna moves toward the pivot helix of the adjacent subunit (2). The short helix of the descending strand of the antenna becomes extended and distorted at the carboxyl end (3). Finally, the internal cavity of the helix compresses, bringing the three pairs closer together (4). B shows the structure of ADP (green spheres) bound to the apo-form of GDH and the location of Arg-463 (mauve spheres) that is involved in ADP activation (22). C shows the location of the inhibitor, GTP (mauve spheres), bound to the NADH (gray spheres), and glutamate (orange spheres) abortive complex. The green arrow notes the approximate location of one of the two sites (Lys-420) modified by 5′-FSBA (48). Comparing B and C, the closing of the catalytic cleft and the movement of the pivot helix is evident.From the structures GDH alone and complexed with various ligands, it is clear that GDH undergoes large conformational changes during each catalytic cycle (1922) (the locations of these changes are summarized in Fig. 1). Substrate binds to the deep recesses of the cleft between the coenzyme binding domain and the lower domain. Coenzyme binds along the coenzyme binding domain surface of the cleft. Upon binding, the coenzyme binding domain rotates by ∼18° to firmly close down upon the substrate and coenzyme (Fig. 1, arrow 1). As the catalytic cleft closes, the base of each of the long ascending helices in the antenna appears to rotate out in a counterclockwise manner to push against the “pivot” helix of the adjacent subunit (Fig. 1, arrow 2). There is a short helix in the descending loop of the antenna that becomes distended and shorter as the mouth closes in a manner akin to an extending spring (Fig. 1, arrow 3). The pivot helix rotates in a counterclockwise manner along the helical axes as well as rotating counterclockwise around the trimer 3-fold axis. Finally, the entire hexamer seems to compress as the mouth closes (Fig. 1, arrow 4). The three pairs of subunits that sit on top of each other move as a rigid units toward each other, compressing the cavity at the core of the hexamer. This last conformational change will be further examined in this work. Allosteric regulation is likely exacted by controlling some or all of these conformational changes.The reason for complex animal regulation came from studies that linked GDH regulation with insulin and ammonia homeostasis. The connection between GDH and insulin regulation was initially established using a non-metabolizable analog of leucine (7, 23), BCH (β-2-aminobicycle[2.2.1]-heptane-2-carboxylic acid). These studies demonstrated that activation of GDH was tightly correlated with increased glutaminolysis and release of insulin. In addition, it has also been noted that factors that regulate GDH also affect insulin secretion (24). The in vivo importance of GDH in glucose homeostasis was demonstrated by the discovery that a genetic hypoglycemic disorder, the hyperinsulinemia/hyperammonemia (HHS) syndrome, is caused by loss of GTP regulation of GDH (2527). Children with HHS have increased β-cell responsiveness to leucine and susceptibility to hypoglycemia following high protein meals (28). This is likely due to uncontrolled catabolism of amino acids yielding high ATP levels that stimulate insulin secretion as well as high serum ammonium levels. During glucose-stimulated insulin secretion in normal individuals, it has been proposed that the generation of high energy phosphates inhibits GDH and promotes conversion of glutamate to glutamine, which, alone or combined, might amplify the release of insulin (29, 30). Further support for this contention came from studies on the inhibitory effects of the polyphenolic compounds from green tea on BCH-stimulated insulin secretion (31). This not only lent support for the contention that GDH plays a significant role in insulin homeostasis, but also suggests that the HHS disorder might be directly controlled pharmaceutically. The role of GDH in insulin homeostasis is summarized in Fig. 2.Open in a separate windowFIGURE 2.Link between GDH and insulin homeostasis. This figure shows the role of GDH in BCH stimulated insulin secretion and how GDH inhibitors affect this process (29, 30). In energy-depleted β-cells, a BCH ramp stimulates insulin secretion. Here, the major energy source is glutaminolysis via phosphate-dependent glutaminase and GDH, because the concentration of GDH inhibitors (ATP/GTP) have been depleted and the phosphate-dependent glutaminase activator Pi (inorganic phosphate) has been increased. BCH stimulates glutamine utilization via GDH activation, thus providing the ATP signal necessary for insulin secretion. GDH inhibitors block this process by inhibiting GDH activity.To both find a more stable pharmaceutical agent to control HHS and to better understand the allosteric regulation of GDH, a high throughput screen was performed to identify new GDH inhibitors (32). Of the ∼30,000 compounds tested, ∼20 demonstrated significant activity. Three of the most active compounds, hexachlorophene, GW5074, and bithionol, were chosen for further analysis in this study. As shown here, all three compounds exhibit essentially non-competitive inhibition of the reaction and therefore do not compete with either substrate or coenzyme. Structural studies are presented here that demonstrate that six hexachlorophene (HCP) molecules bind to the inner core of the GDH hexamer, forming an internal ring via aromatic interactions. In contrast, bithionol and GW5074 bind as pairs between dimers of GDH subunits further away from the core of the enzyme. None of these compounds induce significant conformational changes in their immediate vicinity, and the mechanism of action is not clear from the location of their binding sites. However, detailed analysis of the various GDH complexes shows all of the drugs are binding to contact areas in the core of the hexamer that appear to be expanding and contracting during catalytic turnover. Therefore, inhibition is likely due interference with this “breathing” process.  相似文献   

12.
Four class IIa bacteriocins (pediocin PA-1, enterocin A, sakacin P, and curvacin A) were purified to homogeneity and tested for activity toward a variety of indicator strains. Pediocin PA-1 and enterocin A inhibited more strains and had generally lower MICs than sakacin P and curvacin A. The antagonistic activity of pediocin-PA1 and enterocin A was much more sensitive to reduction of disulfide bonds than the antagonistic activity of sakacin P and curvacin A, suggesting that an extra disulfide bond that is present in the former two may contribute to their high levels of activity. The food pathogen Listeria monocytogenes was among the most sensitive indicator strains for all four bacteriocins. Enterocin A was most effective in inhibiting Listeria, having MICs in the range of 0.1 to 1 ng/ml. Sakacin P had the interesting property of being very active toward Listeria but not having concomitant high levels of activity toward lactic acid bacteria. Strains producing class IIa bacteriocins displayed various degrees of resistance toward noncognate class IIa bacteriocins; for the sakacin P producer, it was shown that this resistance is correlated with the expression of immunity genes. It is hypothesized that variation in the presence and/or expression of such immunity genes accounts in part for the remarkably large variation in bacteriocin sensitivity displayed by lactic acid bacteria.Many lactic acid bacteria (LAB), including members of the genera Lactococcus, Lactobacillus, Carnobacterium, Enterococcus, and Pediococcus, are known to secrete small, ribosomally synthesized antimicrobial peptides called bacteriocins (26, 29, 34). Some of these peptides undergo posttranslational modifications (class I bacteriocins), whereas others are not modified (class II bacteriocins) (29, 34). Class II bacteriocins contain between 30 and 60 residues and are usually positively charged at a neutral pH. Studies of a large number of class II bacteriocins have led to subgrouping of these compounds (29, 34). One of the subgroups, class IIa, contains bacteriocins that are characterized by the presence of YGNG and CXXXXCXV sequence motifs in their N-terminal halves as well as by their strong inhibitory effect on Listeria (e.g., 3, 4, 22, 23, 27, 28, 31, 38, 45) (Fig. (Fig.1).1). Because of their effectiveness against the food pathogen Listeria, class IIa bacteriocins have potential as antimicrobial agents in food and feed. Open in a separate windowFIG. 1Sequence alignment of class IIa bacteriocins. Residue numbering is according to the sequence of pediocin PA-1. Cysteine residues are printed in boldface; the two known class IIa bacteriocins with four cysteine residues are in the upper group. No attempt was made to optimize the alignment in the C-terminal halves of the peptides. Piscicolin 126 is identical to piscicocin V1a (4). Carnobacteriocin BM1 most probably is identical to piscicocin V1b (4). Sakacin P most probably is identical to bavaricin A (30). Curvacin A is identical to sakacin A (2). The consensus sequence includes residues conserved in at least 8 of the 12 sequences shown; 100% conserved residues are underlined.Class IIa bacteriocins act by permeabilizing the membrane of their target cells (1, 5, 6, 9, 10, 26, 28). The most recent studies on the mode of action of these bacteriocins indicate that antimicrobial activity does not require a specific receptor and is enhanced by (but not fully dependent on) a membrane potential (9, 28). Little is known about bacteriocin structure, and unravelling the relationships between structure and function is one of the great challenges in current bacteriocin research. A logical starting point for structure-function studies is a thorough study of the differences in activity and target cell specificity between naturally occurring homologous bacteriocins. A few such studies have been described, but these suffer from either a very limited number of tested indicator strains or the use of culture supernatants instead of purified bacteriocins (3, 4, 17, 45). The use of purified bacteriocins for comparative analyses is absolutely essential, since it is becoming increasingly evident that bacteriocin producers produce more than one bacteriocin (4, 8, 38, 48; this study).In the present study, the activities of four pure class IIa bacteriocins (pediocin PA-1, enterocin A, curvacin A, and sakacin P) (Fig. (Fig.1)1) were tested against a large number of LAB as well as several strains of the food pathogen Listeria monocytogenes. The bacteriocins were purified from their respective producer strains by use of an optimized purification protocol yielding highly pure samples. The contribution of disulfide formation was assessed and found to be important for activity. The effects of the purified bacteriocins on (noncognate) class IIa bacteriocin-producing strains are described, and the implications of our findings for immunity and resistance are discussed.  相似文献   

13.
14.
15.
The study of metabolically labeled or probe-modified proteins is an important area in chemical proteomics. Isolation and purification of the protein targets is a necessary step before MS identification. The biotin-streptavidin system is widely used in this process, but the harsh denaturing conditions also release natively biotinylated proteins and non-selectively bound proteins. A cleavable linker strategy is a promising approach for solving this problem. Though several cleavable linkers have been developed and tested, an efficient, easily synthesized, and inexpensive cleavable linker is a desirable addition to the proteomics toolbox. Here, we describe the chemical proteomics application of a vicinal diol cleavable linker. Through easy-to-handle chemistry we incorporate this linker into an activity-based probe and a biotin alkyne tag amenable for bioorthogonal ligation. With these reagents, background protein identifications are significantly reduced relative to standard on-bead digestion.The covalent modification of proteins by small molecules within a complex proteome is a major theme in chemical biology and proteomics. An effective method for the detection of posttranslational modifications of proteins is the metabolic incorporation of modified biomolecules such as tagged carbohydrates or lipids (1). Reversible interactions of enzyme inhibitors, natural products, or drugs can be detected when one appends photocrosslinking agents, thereby facilitating target discovery (2, 3). A particularly interesting example of protein labeling is activity-based protein profiling (ABPP)1 (4, 5), which utilizes the intrinsic catalytic activity of a target enzyme for the covalent attachment of an affinity or visualization tag. ABPP makes use of small molecules (activity-based probes (ABPs)) that react with the active form of a specific enzyme or enzyme class by means of a “warhead,” which is often derived from a mechanism-based enzyme inhibitor (Fig. 1A). DCG-04, for example, is based on the naturally occurring inhibitor E-64 and targets the papain family of cysteine proteases via covalent attachment of the epoxysuccinate group to the active site cysteine (Fig. 1B) (6).Open in a separate windowFig. 1.The cleavable linker strategy in ABPP. A, the elements of an ABP. B, the example ABP DCG-04, an epoxysuccinate-containing probe for clan CA cysteine proteases. DCG-04 is based on the naturally occurring protease inhibitor E-64. C, schematic strategy of cleavable linker-mediated target identification. D, the cleavage mechanism of a vicinal diol.Bulky fluorophore or biotin tags on chemical probes might interfere with efficient protein binding. Moreover, they can negatively influence the cell permeability of probes, which therefore limits their applicability in in vitro experiments. Bioorthogonal chemistries, such as the Bertozzi-Staudinger ligation (7) and the 1,3-bipolar cycloaddition of an azide and an alkyne (click chemistry) (8), allow tandem labeling strategies in which a biotin or a fluorophore is attached to an enzyme probe complex in a separate step. Consequently, the probes themselves only carry azide or alkyne groups as “mini-tags.” Tandem labeling using bioorthogonal chemistry has now become a widely used strategy to label biomolecules in lysates and in live cells (911).An essential step in ABPP, as well as in other chemical proteomics approaches, is the elucidation of the tagged proteins. This usually involves a biotin-mediated enrichment step followed by mass-spectrometry-based identification. Although the streptavidin-biotin interaction allows efficient enrichment as a result of the strong binding affinity (Kd ∼ 10−15 m), it also has limitations. The quantitative elution of biotinylated proteins requires harsh conditions (12), which lead to contamination of the sample by endogenous biotinylated and non-specifically bound proteins. These other proteins will be identified together with the real protein targets. Given that subsequent target validation with secondary assays can be a costly and time-consuming process, a reduction in false positive identifications is highly desirable. For cleaner protein identification, cleavable linker strategies (13) that allow the selective release of target proteins have been developed (Fig. 1C). The commercially available disulfide linker can be cleaved under mild conditions, but it suffers from premature cleavage in reducing media such as the intracellular environment and reducing buffers used for click chemistry and in vitro reactions of cysteine proteases. Therefore, a variety of alternative linkers for proteomics applications have been reported, including a sterically hindered disulfide (14), diazobenzenes (1519), hydrazones (20, 21), silanes (22), light sensitive linkers (2325), tobacco etch virus protease sensitive linkers (26, 27), and a levulinoyl-based linker (28). The synthesis of some of these linkers is lengthy or difficult to scale up, which limits their general application in chemical proteomics.Ideally, a cleavable linker is stable under a wide variety of conditions, is efficiently and selectively cleaved, and can be synthesized in a low number of easy chemical transformations. We aimed to meet these requirements by using a vicinal diol as a cleavable linker system. When vicinal diols are treated with sodium periodate (NaIO4), the carbon–carbon bond is cleaved (Fig. 1D). Periodate treatment of proteins can result in side-reactions, such as the cleavage of linked carbohydrates or the oxidation of N-terminal serine and threonine residues. However, these N-termini rarely occur in proteins and are therefore of minor concern. In general, the mild, neutral conditions of periodate cleavage are compatible with proteins. This has been illustrated in the past, for example, by its application in the detection of protein–protein interactions (29) and the creation of unliganded MHC class I molecules (30). In this article, we report the chemical proteomics application of diol cleavable linker probes. We show that the synthesis of the linker and its probe derivatives is straightforward, that the linker is compatible with tandem click labeling, that enrichment and release of probe targets is efficient, and that the identification of targets takes place with significantly lower background than in on-bead digestion protocols.  相似文献   

16.
17.
Malic enzyme has a dimer of dimers quaternary structure in which the dimer interface associates more tightly than the tetramer interface. In addition, the enzyme has distinct active sites within each subunit. The mitochondrial NAD(P)+-dependent malic enzyme (m-NAD(P)-ME) isoform behaves cooperatively and allosterically and exhibits a quaternary structure in dimer-tetramer equilibrium. The cytosolic NADP+-dependent malic enzyme (c-NADP-ME) isoform is noncooperative and nonallosteric and exists as a stable tetramer. In this study, we analyze the essential factors governing the quaternary structure stability for human c-NADP-ME and m-NAD(P)-ME. Site-directed mutagenesis at the dimer and tetramer interfaces was employed to generate a series of dimers of c-NADP-ME and m-NAD(P)-ME. Size distribution analysis demonstrated that human c-NADP-ME exists mainly as a tetramer, whereas human m-NAD(P)-ME exists as a mixture of dimers and tetramers. Kinetic data indicated that the enzyme activity of c-NADP-ME is not affected by disruption of the interface. There are no significant differences in the kinetic properties between AB and AD dimers, and the dimeric form of c-NADP-ME is as active as tetramers. In contrast, disrupting the interface of m-NAD(P)-ME causes the enzyme to be less active than wild type and to become less cooperative for malate binding; the kcat values of mutants decreased with increasing Kd,24 values, indicating that the dissociation of subunits at the dimer or tetramer interfaces significantly affects the enzyme activity. The above results suggest that the tetramer is required for a fully functional m-NAD(P)-ME. Taken together, the analytical ultracentrifugation data and the kinetic analysis of these interface mutants demonstrate the differential role of tetramer organization for the c-NADP-ME and m-NAD(P)-ME isoforms. The regulatory mechanism of m-NAD(P)-ME is closely related to the tetramer formation of this isoform.Malic enzymes catalyze a reversible oxidative decarboxylation of l-malate to yield pyruvate and CO2 with reduction of NAD(P)+ to NAD(P)H. This reaction requires a divalent metal ion (Mg2+ or Mn2+) for catalysis (13). Malic enzymes are found in a broad spectrum of living organisms that share conserved amino acid sequences and structural topology; such shared characteristics reveal a crucial role for the biological functions of these enzymes (4, 5). In mammals, malic enzymes have been divided into three isoforms according to their cofactor specificity and subcellular localization as follows: cytosolic NADP+-dependent (c-NADP-ME),2 mitochondrial NADP+-dependent (m-NADP-ME), and mitochondrial NAD(P)+-dependent (m-NAD(P)-ME). The m-NAD(P)-ME isoform displays dual cofactor specificity; it can use both NAD+ and NADP+ as the coenzyme, but NAD+ is more favored in a physiological environment (68). Dissimilar to the other two isoforms, m-NAD(P)-ME binds malate cooperatively, and it can be allosterically activated by fumarate; the sigmoidal kinetics observed with cooperativity is abolished by fumarate (912). Mutagenesis and kinetic studies demonstrated that ATP is an active-site inhibitor, although it also binds to the exo sites in the tetramer interface (1315). Structural studies also revealed an allosteric binding site for fumarate residing at the dimer interface. Mutation in the binding site significantly affects the activating effects of fumarate (11, 16, 17).The c-NADP-ME and m-NADP-ME isoforms play an important role in lipogenesis by providing NADPH for the biosynthesis of long-chain fatty acids and steroids. Thus, c-NADP-ME together with acetyl-CoA carboxylase, fatty-acid synthase, and glucose-6-phosphate dehydrogenase are classified as lipogenic enzymes (2, 1821). The m-NAD(P)-ME isoform has attracted much attention because it is involved in glutaminolysis, which is an energy-producing pathway of tumor cells that utilizes glutamine and glutamate. Thus, m-NAD(P)-ME is considered to be a potential target in cancer therapy (2227).Various crystal structures of malic enzymes in complex with substrate, metal ion, coenzyme, regulator, and inhibitor are available in the Protein Data Bank (4, 11, 2832). The overall tertiary structures of these malic enzymes are similar, but there are still some differences that may be significant for catalysis and regulation of the enzyme. Malic enzyme exists as a dimer of dimers with a stronger association of the dimer interface than that of the tetramer interface (Fig. 1A). The dimer interface is formed by subunits A and B or C and D (Fig. 1B), whereas the tetramer interaction consists of contacts between subunits A and D or B and C (Fig. 1C). A hydrophobic interaction is the major driving force for subunit assembly, but hydrogen bonding and ionic interactions also contribute markedly. The homotetramer of the enzyme is composed of four identical monomers each with its own active site. In the structure of human m-NAD(P)-ME, aside from its well defined active site, there are two regulatory sites on the enzyme (Fig. 1A). One of these sites is located at the dimer interface and is occupied by fumarate (Fig. 1B), whereas the other site, which is referred to as the exo site, is located at the tetramer interface and is occupied by either an NAD or an ATP molecule (Fig. 1A). In the ME family, Ascaris suum and human m-NAD(P)-ME were found to be activated by fumarate (11, 1517, 31). However, the relationship between enzyme regulation and subunit-subunit interaction is still unclear.Open in a separate windowFIGURE 1.Dimer and tetramer interfaces of human m-NAD(P)-ME. A, dimer of dimers quaternary structure of human m-NAD(P)-ME (Protein Data Bank code 1PJ3). The active site, fumarate site, and exo site in each subunit are indicated. B, dimer interface between A and B subunits of m-NAD(P)-ME. C, tetramer interface between A and D subunits of m-NAD(P)-ME. The amino acid residues at the dimer interface, Gln-51, Glu-90, Asp-139, His-142, and Asp-568 and C terminus in the tetramer interface, are represented by ball-and-stick modeling. The amino acid residues 51 and 90 in human c-NADP-ME are His and Asp, respectively. This figure was generated with PyMOL (DeLano Scientific LLC, San Carlos, CA).Previous studies have shown that the quaternary structure stability of malic enzyme isoforms is diverse. At neutral pH, pigeon c-NADP-ME exists as a unique tetramer with a sedimentation coefficient of ∼10 S (3335), whereas human m-NAD(P)-ME exists as a mixture of tetramer and dimer of 9.5 S and 6.5 S, respectively (13, 35). Some mutations at the interface will affect the quaternary structure (3437). Although the crystal structure of human c-NADP-ME has not been resolved, it is believed that it is similar to pigeon c-NADP-ME.Here we analyze the essential factors governing quaternary structure stability for human c-NADP-ME and m-NAD(P)-ME. Site-directed mutagenesis at the dimer and tetramer interfaces was used to disrupt the tetramer organization to create a series of c-NADP-ME and m-NAD(P)-ME dimers. Combined with the analytical ultracentrifugation data and kinetic analysis of these interface mutants, we demonstrate the differential role of tetramer organization for the c-NADP-ME and m-NAD(P)-ME isoforms. The regulatory mechanism of m-NAD(P)-ME is highly associated with the tetramer formation of this isoform.  相似文献   

18.
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   

19.
Linear genome stability requires specialized telomere replication and protection mechanisms. A common solution to this problem in non-eukaryotes is the formation of hairpin telomeres by telomere resolvases (also known as protelomerases). These enzymes perform a two-step transesterification on replication intermediates to generate hairpin telomeres using an active site similar to that of tyrosine recombinases and type IB topoisomerases. Unlike phage telomere resolvases, the telomere resolvase from the Lyme disease pathogen Borrelia burgdorferi (ResT) is a permissive enzyme that resolves several types of telomere in vitro. However, the ResT region and residues mediating permissive substrate usage have not been identified. The relapsing fever Borrelia hermsii ResT exhibits a more restricted substrate usage pattern than B. burgdorferi ResT and cannot efficiently resolve a Type 2 telomere. In this study, we determined that all relapsing fever ResTs process Type 2 telomeres inefficiently. Using a library of chimeric and mutant B. hermsii/B. burgdorferi ResTs, we mapped the determinants in B. burgdorferi ResT conferring the ability to resolve multiple Type 2 telomeres. Type 2 telomere resolution was dependent on a single proline in the ResT catalytic region that was conserved in all Lyme disease but not relapsing fever ResTs and that is part of a 2-amino acid insertion absent from phage telomere resolvase sequences. The identification of a permissive substrate usage determinant explains the ability of B. burgdorferi ResT to process the 19 unique telomeres found in its segmented genome and will aid further studies on the structure and function of this essential enzyme.Replication and protection of telomeric DNA are required to ensure the genomic stability of all organisms with linear replicons. Until quite recently, it was assumed that linearity is a property confined to the replicons of eukaryotes and certain primarily eukaryotic viruses. However, a growing body of evidence indicates that linear DNA is also found in a broad range of bacteriophages (16) and in bacteria themselves (710), including the Borrelia species that cause Lyme disease and relapsing fever (11, 12). A common solution to the end replication and protection problem in non-eukaryotes is the covalent sealing of DNA ends in the form of hairpins (2, 46, 10, 11, 1316). Hairpin DNA is not recognized as a double-strand break, and continuous synthesis of DNA around the hairpin loop abolishes the end replication problem. However, mother and daughter replicons are covalently linked at the junction of their telomeres following DNA replication; separation of the two replicons and formation of new hairpin telomeres require a DNA breakage and reunion process referred to as telomere resolution (17, 18).Resolution of the linear chromosome and plasmids in Borrelia species and of the linear plasmid prophages from Escherichia coli, Yersinia enterocolitica, and Klebsiella oxytoca is performed by telomere resolvases (also referred to as protelomerases) (5, 1921). A growing number of candidate telomere resolvases have been identified in the genomes of eukaryotic viruses, phages, and bacteria (22, 23). Telomere resolvases are DNA cleavage and rejoining enzymes related to tyrosine recombinases and type 1B topoisomerases (19, 21, 22, 24, 25). Telomere resolvase catalyzes a two-step transesterification reaction in which staggered cuts are introduced 6 bp apart on either side of the axis of symmetry in the replicated telomere substrate (5, 19, 21, 24). Cleavage is accompanied by the formation of a 3′-phosphotyrosyl protein-DNA linkage. Subsequent nucleophilic attack on opposing strands by the free 5′-OH groups in the nicked substrate creates covalently closed hairpin telomeres. A recent crystal structure of the Klebsiella phage telomere resolvase (TelK) in complex with its substrate identified the residues involved in catalysis (25); all but one of these residues are conserved in all telomere resolvases (22), implying that the basic catalytic mechanism underlying telomere resolution is conserved. However, telomere resolvase sequences vary substantially outside of the central catalytic region (25, 26), and the enzymes characterized to date demonstrate important differences in substrate usage that likely reflect functionally distinct mechanisms of substrate interaction.The Borrelia burgdorferi telomere resolvase, ResT, appears to be particularly divergent. It is substantially smaller than phage telomere resolvases, and unlike its phage counterparts (5, 20, 21), it cannot efficiently resolve negatively supercoiled DNA (19, 27), presumably reflecting differences in the substrates resolved by phage and Borrelia telomere resolvases in vivo. On the other hand, B. burgdorferi ResT can fuse hairpin telomeres in a reversal of the resolution reaction (28), a function that is not shared with the phage telomere resolvase TelK (25). It can also synapse replicated telomeres and catalyze the formation of Holliday junctions (29). The ability of ResT to promote hairpin fusion has been proposed as the mechanism underlying the ongoing genetic rearrangements that are a prominent feature of the B. burgdorferi genome (18, 28). Finally, B. burgdorferi ResT can tolerate a surprising amount of variation in its substrate (30, 31), a feature that is not shared by phage telomere resolvases (21). Although B. burgdorferi ResT appears to be more permissive with a greater scope of activities than other telomere resolvases, the sequences mediating most of its unique properties have not yet been identified.The B. burgdorferi genome contains a total of 19 distinct hairpin sequences, all of which must be resolved by ResT (31). These sequences can be classified into three groups based on the presence and positioning of the box 1 motif, which is a critical determinant of activity in phage and Borrelia telomere resolvases (see Fig. 1A) (21, 24, 30). A box 1-like motif is also found in many of the hairpin telomeres sequenced to date (6, 14, 3235), although its function in telomere resolution is unknown. The box 1 consensus sequence (TAT(a/t)AT) closely resembles the −10/Pribnow box and TATA box consensus sequences of prokaryotic and eukaryotic promoters (TATAAT and TATA(a/t)A(a/t), respectively), which undergo transient deformations that predispose them to melting (36) and are intrinsically bent and anisotropically flexible (37). Therefore, box 1 may facilitate nucleation of hairpin folding and/or may confer an intrinsic bend or flexibility to substrates that is important for the resolution reaction.Open in a separate windowFIGURE 1.Species-specific resolution of Type I and 2 telomeres. A, a schematic showing the three types of hairpin telomere found on the linear replicons of the B. burgdorferi genome (see Ref. 31). The box 1 sequence in Type 1 and 2 telomeres is situated 1 and 4 nucleotides away from the axis of symmetry, respectively, whereas Type 3 telomeres contain no clear box 1. B, a schematic illustrating the telomere resolution reaction substrate and products is shown along with two ethidium bromide-stained agarose gels showing telomere resolution assays. The gels show resolution kinetics for B. burgdorferi and B. hermsii ResT on Type 1 and 2 telomeres (plasmid substrates pYT1/lp17L and pYT92/chromL, respectively).B. burgdorferi ResT can resolve telomeres in which box 1 is located at positions 1 and 4 nucleotides away from the axis of symmetry (Type 1 and 2 telomeres, respectively), as well as AT-rich telomeres without a box 1 sequence (Type 3 telomeres) (see Fig. 1A) (30, 31). B. burgdorferi ResT cleaves telomeres at a fixed position relative to the axis of symmetry, independent of the location of box 1 (30). Positioning of the enzyme for cleavage in all telomere types is most likely driven by sequence-specific interactions between ResT domains 2 (catalytic) and/or 3 (C-terminal) and a fixed element upstream of box 1 that is positioned 14 nucleotides from the axis of symmetry in all Borrelia telomeres (box 3 and adjacent nucleotides) (see Figs. 1A and and2)2) (26, 30, 31). In contrast, box 1 and axis-flanking nucleotides are not involved in high affinity and/or sequence-specific interactions with ResT and require the ResT N-terminal domain for full protection in DNase footprinting assays (26, 27). The most likely candidate for interactions with box 1 and axis-flanking nucleotides is a Borrelia-specific hairpin-binding region in the N terminus, which is thought to promote a pre-hairpinning step involving strand opening at the axis (38).Open in a separate windowFIGURE 2.Alignment of 11 Borrelia ResT sequences. Shown is ClustalW2 alignment of ResT amino acid sequences from five Lyme disease Borrelia species (B. afzelii, B. spielmanii, B. valaisiana, B. garinii, and B. burgdorferi), five relapsing fever Borrelia species (B. turicatae, B. parkeri, B. hermsii, B. recurrentis, and B. duttonii), and one avian Borrelia species (B. anserina) (generated using ClustalW2 from the EBI web site) (19, 3942, 48, 49). The sequences for B. anserina, B. parkeri, and B. turicatae ResTs are reported for the first time in this study (respective GenBank accession numbers are FJ882620, FJ882621, and FJ882623). Sequences are arranged in order of similarity to neighboring sequences and are colored in JalView using the Zappo coloring scheme for identifying amino acids with similar physicochemical properties (50). Only residues that are identical in 100% of ResTs are indicated by colored shading. Arrows above the alignment indicate ResT domain boundaries identified by chymotrypsin digest, sequence comparison with other proteins, and HHsenser predictions (26, 51). The hairpin-binding motif found in cut-and-paste transposases is indicated beneath the alignment by white text on a black background (38). The positions corresponding to the active site residues in tyrosine recombinases, type IB topoisomerases, and TelK are indicated by blue asterisks below the sequence, with the active site tyrosine nucleophile at position 335 marked by a red asterisk (22, 25). The ringed black dot below position 326 indicates an amino acid in the active site region that differs in Lyme disease and relapsing fever ResTs. Sequences above the black line drawn between B. burgdorferi and B. turicatae are from Lyme disease Borrelia species; sequences below the black line are from relapsing fever Borrelia species. The ResT sequence from the avian Borrelia species B. anserina is shown at bottom.ResT from the relapsing fever Borrelia species Borrelia hermsii exhibits a more restricted substrate usage pattern in vitro when compared with ResT from the Lyme disease pathogen B. burgdorferi (39). Specifically, B. hermsii ResT is unable to efficiently resolve a Type 2 telomere. Therefore, B. burgdorferi ResT appears to be a more permissive enzyme than its relapsing fever counterpart. In this study, we investigated the basis for permissive substrate usage by B. burgdorferi ResT. Using a library of chimeric B. hermsii/B. burgdorferi ResTs, we mapped the sequence determinants in B. burgdorferi ResT that confer the ability to resolve multiple Type 2 telomeres. Surprisingly, this approach indicated that Type 2 telomere resolution was crucially regulated by a single proline residue located in a small Borrelia-specific insertion in the central catalytic region of ResT. The proline at this position was conserved in the ResTs from all Lyme disease Borrelia species but in none of the ResTs from relapsing fever Borrelia species, which were unable to efficiently resolve Type 2 telomeres in vitro. This study has identified a specific residue in ResT responsible for permissive substrate usage patterns.  相似文献   

20.
In this study, we report that the purified wild-type FANCI (Fanconi anemia complementation group I) protein directly binds to a variety of DNA substrates. The DNA binding domain roughly encompasses residues 200–1000, as suggested by the truncation study. When co-expressed in insect cells, a small fraction of FANCI forms a stable complex with FANCD2 (Fanconi anemia complementation group D2). Intriguingly, the purified FANCI-FANCD2 complex preferentially binds to the branched DNA structures when compared with either FANCI or FANCD2 alone. Co-immunoprecipitation with purified proteins indicates that FANCI interacts with FANCD2 through its C-terminal amino acid 1001–1328 fragment. Although the C terminus of FANCI is dispensable for direct DNA binding, it seems to be involved in the regulation of DNA binding activity. This notion is further enhanced by two C-terminal point mutations, R1285Q and D1301A, which showed differentiated DNA binding activity. We also demonstrate that FANCI forms discrete nuclear foci in HeLa cells in the absence or presence of exogenous DNA damage. The FANCI foci are colocalized perfectly with FANCD2 and partially with proliferating cell nuclear antigen irrespective of mitomycin C treatment. An increased number of FANCI foci form and become resistant to Triton X extraction in response to mitomycin C treatment. Our data suggest that the FANCI-FANCD2 complex may participate in repair of damaged replication forks through its preferential recognition of branched structures.Fanconi anemia (FA)3 is a genetic disorder characterized by chromosome instability, predisposition to cancer, hypersensitivity to DNA cross-linking agents, developmental abnormalities, and bone marrow failure (19). There are at least 13 distinct FA complementation groups, each of which is associated with an identified gene (2, 9, 10). Eight of them are components of the FA core complex (FANC A, B, C, E, F, G, L, and M) that is epistatic to the monoubiquitination of both FANCI and FANCD2, a key event to initiate interstrand cross-link (ICL) repair (2, 9, 11). Downstream of or parallel to the FANCI and FANCD2 monoubiquitination are the proteins involved in double strand break repair and breast cancer susceptibility (i.e. FANCD1/BRCA2, FANCJ/BRIP1, and FANCN/PALB2) (2, 9).FANCI is the most recently identified FA gene (1113). FANCI protein is believed to form a FANCI-FANCD2 (ID) complex with FANCD2, because they co-immunoprecipitate with each other from cell lysates and their stabilities are interdependent of each other (9, 11, 13). FANCI and FANCD2 are paralogs to each other, since they share sequence homology and co-evolve in the same species (11). Both FANCI and FANCD2 can be phosphorylated by ATR/ATM (ataxia telangiectasia and Rad3-related/ataxia telangiectasia-mutated) kinases under genotoxic stress (11, 14, 15). The phosphorylation of FANCI seems to function as a molecular switch to turn on the FA repair pathway (16). The monoubiquitination of FANCD2 at lysine 561 plays a critical role in cellular resistance to DNA cross-linking agents and is required for FANCD2 to form damage-induced foci with BRCA1, BRCA2, RAD51, FANCJ, FANCN, and γ-H2AX on chromatin during S phase of the cell cycle (1725). In response to DNA damage or replication stress, FANCI is also monoubiquitinated at lysine 523 and recruited to the DNA repair nuclear foci (11, 13). The monoubiquitination of both FANCI and FANCD2 depends on the FA core complex (11, 13, 26), and the ubiquitination of FANCI relies on the FANCD2 monoubiquitination (2, 11). In an in vitro minimally reconstituted system, FANCI enhances FANCD2 monoubiquitination and increases its specificity toward the in vivo ubiquitination site (27).FANCI is a leucine-rich peptide (14.8% of leucine residues) with limited sequence information to indicate which processes it might be involved in. Besides the monoubiquitination site Lys523 and the putative nuclear localization signals (Fig. 1A), FANCI contains both ARM (armadillo) repeats and a conserved C-terminal EDGE motif as FANCD2 does (11, 28). The EDGE sequence in FANCD2 is not required for monoubiquitination but is required for mitomycin C (MMC) sensitivity (28). The ARM repeats form α-α superhelix folds and are involved in mediating protein-protein interactions (11, 29). In addition, FANCI, at its N terminus, contains a leucine zipper domain (aa 130–151) that could be involved in mediating protein-protein or protein-DNA interactions (Fig. 1A) (3033). FANCD2, the paralog of FANCI, was reported to bind to double strand DNA ends and Holliday junctions (34).Open in a separate windowFIGURE 1.Purified human FANCI binds to DNA promiscuously. A, schematic diagram of predicted FANCI motifs and mutagenesis strategy to define the DNA binding domain. The ranges of numbers indicate how FANCI was truncated (e.g. 801–1328 represents FANCI-(801–1328)). NLS, predicted nuclear localization signal (aa 779–795 and 1323–1328); K523, lysine 523, the monoubiquitination site. The leucine zipper (orange bars, aa 130–151), ARM repeats (green bars), and EDGE motif (blue bars) are indicated. Red bars with a slash indicate the point mutations shown on the left. B, SDS-PAGE of the purified proteins stained with Coomassie Brilliant Blue R-250. R1285Q and D1301A are two point mutants of FANCI. All FANCI variants are tagged by hexahistidine. FANCD2 is in its native form. Protein markers in kilodaltons are indicated. C, titration of WT-FANCI for the DNA binding activity. Diagrams of the DNA substrates are shown at the top of each set of reactions. *, 32P-labeled 5′-end. HJ, Holliday junction. Concentrations of FANCI were 0, 20, 40, 60, and 80 nm (ascending triangles). The substrate concentration was 1 nm. Protein-DNA complex is indicated by an arrow. D, supershift assay. 1 nm of ssDNA was incubated with PBS (lane 1), 80 nm FANCI alone (lane 2), and 80 nm FANCI preincubated with a specific FANCI antibody (lane 3) in the condition described under “Experimental Procedures.”In order to delineate the function of FANCI protein, we purified the recombinant FANCI from the baculovirus expression system. In this study, we report the DNA binding activity of FANCI. Unlike FANCD2, FANCI binds to different DNA structures, including single-stranded DNA (ssDNA), double-stranded DNA (dsDNA), 5′-tailed, 3′-tailed, splayed arm, 5′-flap, 3′-flap, static fork, and Holliday junction with preference toward branched structures in the presence of FANCD2. Our data suggest that the dynamic DNA binding activity of FANCI and the preferential recognition of branched structures by the ID complex are likely to be the mechanisms to initiate downstream repair events.  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