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.
Berberine bridge enzyme (BBE) is a member of the recently discovered family of bicovalently flavinylated proteins. In this group of enzymes, the FAD cofactor is linked via its 8α-methyl group and the C-6 atom to conserved histidine and cysteine residues, His-104 and Cys-166 for BBE, respectively. 6-S-Cysteinylation has recently been shown to have a significant influence on the redox potential of the flavin cofactor; however, 8α-histidylation evaded a closer characterization due to extremely low expression levels upon substitution. Co-overexpression of protein disulfide isomerase improved expression levels and allowed isolation and purification of the H104A protein variant. To gain more insight into the functional role of the unusual dual mode of cofactor attachment, we solved the x-ray crystal structures of two mutant proteins, H104A and C166A BBE, each lacking one of the covalent linkages. Information from a structure of wild type enzyme in complex with the product of the catalyzed reaction is combined with the kinetic and structural characterization of the protein variants to demonstrate the importance of the bicovalent linkage for substrate binding and efficient oxidation. In addition, the redox potential of the flavin cofactor is enhanced additively by the dual mode of cofactor attachment. The reduced level of expression for the H104A mutant protein and the difficulty of isolating even small amounts of the protein variant with both linkages removed (H104A-C166A) also points toward a possible role of covalent flavinylation during protein folding.Since the discovery of the first known example of a covalent bond between a flavin cofactor and an amino acid side chain occurring in enzymes in the 1950s (1), a number of different types of linkages have been identified: 8α-histidylation (either to N1 or to N3), 8α-O-tyrosylation, 8α-S-cysteinylation, and 6-S-cysteinylation. For current reviews relating to these modes of flavin attachment, see Refs. 2 and 3. Recently, another way of covalent tethering of FAD to proteins was discovered in x-ray crystallographic studies on glucooligosaccharide oxidase (GOOX)4 from Acremonium strictum (4). The mode of flavin linkage observed in this case employs both 8α-histidylation and 6-S-cysteinylation to form a bicovalently attached cofactor. Representative members of all these groups have been studied in detail, and several explanations for the role of the covalent flavinylation have been put forward. Some of the suggestions tend to be rather specific for the system being studied, e.g. prevention of cofactor inactivation at the C-6 position for trimethylamine dehydrogenase (5) or facilitation of electron transfer from the flavin to the cytochrome subunit for p-cresol methylhydroxylase (6). Other explanations including the increase of the flavin redox potential due to the covalent linkage (79) and the prevention of cofactor dissociation (10, 11) were found for several enzymes also harboring different types of cofactor attachments. Taking into account that protein stability (12) and optimal binding of substrate molecules (11, 13) are also positively influenced by covalent tethering of the flavin, one might speculate that no generally applicable explanation for the covalent attachment of flavins to proteins exists. Therefore, it seems likely that the large variety of systems operating with one of the above mentioned modes of cofactor tethering might have evolved to also adapt to a diversity of enzymatic challenges.Berberine bridge enzyme (BBE) from Eschscholzia californica is a plant enzyme involved in alkaloid biosynthesis, catalyzing the challenging oxidative cyclization of (S)-reticuline to (S)-scoulerine (Scheme 1). This enzyme was recently shown to belong to the group of flavoenzymes with a bicovalently attached FAD (14). After the discovery of this unusual mode of linkage in the crystal structure of GOOX (4), several members of this group, all belonging to the vanillyl-alcohol oxidase family (15), were identified by biochemical methods (1618) and also structural studies (19). Because some of the suggested benefits of a covalent cofactor attachment can easily be brought about by a single linkage, e.g. prevention of cofactor dissociation or stabilization of the tertiary structure, the two amino acids attached to FAD might have different and individual functions as well as an additive effect on physicochemical properties such as redox potentials or substrate binding and oxidation. To elucidate the relative importance for the overall enzymatic functioning of members of this group, more detailed studies have been performed on GOOX (11), chito-oligosaccharide oxidase (ChitO) from Fusarium graminearum (17), and BBE (20). Common results of these analyses show that the bicovalent FAD has a redox potential of about +130 mV, which is among the highest potentials reported for flavoenzymes. Replacement of one of the amino acids involved in anchoring of the cofactor generally reduces the rate of cofactor reduction and the steady-state turnover rate, but whether this can be directly linked to reduced redox potentials of these mutant proteins has been under debate (11).Open in a separate windowSCHEME 1.Overall reaction catalyzed by BBE.To address these issues further, we report the expression of the H104A mutant protein of BBE. A biochemical characterization of this protein variant with respect to the redox potential, transient kinetics, and steady-state analysis is combined with the structural analysis of both the H104A and the C166A mutant proteins. In addition, a structure of wild type (WT) BBE in complex with the product of the enzyme-catalyzed reaction is presented, which provides further insights toward the involvement of active site amino acids during the course of the reaction. Together with the recently reported x-ray crystal structure of WT BBE with and without substrate bound (21) and the biochemical characterization of the C166A mutant protein (20), these results provide interesting insights into the role of bicovalent FAD attachment in enzymes.  相似文献   

2.
Complex I (EC 1.6.99.3) of the bacterium Escherichia coli is considered to be the minimal form of the type I NADH dehydrogenase, the first enzyme complex in the respiratory chain. Because of its small size and relative simplicity, the E. coli enzyme has become a model used to identify and characterize the mechanism(s) by which cells regulate the synthesis and assembly of this large respiratory complex. To begin dissecting the processes by which E. coli cells regulate the expression of nuo and the assembly of complex I, we undertook a genetic analysis of the nuo locus, which encodes the 14 Nuo subunits comprising E. coli complex I. Here we present the results of studies, performed on an isogenic collection of nuo mutants, that focus on the physiological, biochemical, and molecular consequences caused by the lack of or defects in several Nuo subunits. In particular, we present evidence that NuoG, a peripheral subunit, is essential for complex I function and that it plays a role in the regulation of nuo expression and/or the assembly of complex I.

Complex I (NADH:ubiquinone oxidoreductase; EC 1.6.99.3), a type I NADH dehydrogenase that couples the oxidation of NADH to the generation of a proton motive force, is the first enzyme complex of the respiratory chain (2, 35, 47). The Escherichia coli enzyme, considered to be the minimal form of complex I, consists of 14 subunits instead of the 40 to 50 subunits associated with the homologous eukaryotic mitochondrial enzyme (17, 29, 30, 4850). E. coli also possesses a second NADH dehydrogenase, NDH-II, which does not generate a proton motive force (31). E. coli complex I resembles eukaryotic complex I in many ways (16, 17, 30, 49): it performs the same enzymatic reaction and is sensitive to a number of the same inhibitors, it consists of subunits homologous to those found in all proton-translocating NADH:ubiquinone oxidoreductases studied thus far, it is comprised of a large number of subunits relative to the number that comprise other respiratory enzymes, and it contains flavin mononucleotide and FeS center prosthetic groups. Additionally, it possesses an L-shaped topology (14, 22) like that of its Neurospora crassa homolog (27), and it consists of distinct fragments or subcomplexes. Whereas eukaryotic complex I can be dissected into a peripheral arm and a membrane arm, the E. coli enzyme consists of three subcomplexes referred to as the peripheral, connecting, and membrane fragments (29) (Fig. (Fig.1A).1A). The subunit composition of these three fragments correlates approximately with the organization of the 14 structural genes (nuoA to nuoN) (49) of the nuo (for NADH:ubiquinone oxidoreductase) locus (Fig. (Fig.1B),1B), an organization that is conserved in several other bacteria, including Salmonella typhimurium (3), Paracoccus denitrificans (53), Rhodobacter capsulatus (12), and Thermus thermophilus (54). The 5′ half of the locus contains a promoter (nuoP), previously identified and located upstream of nuoA (8, 49), and the majority of genes that encode subunits homologous to the nucleus-encoded subunits of eukaryotic complex I and to subunits of the Alcaligenes eutrophus NAD-reducing hydrogenase (17, 29, 30, 49). In contrast, the 3′ half contains the majority of the genes that encode subunits homologous to the mitochondrion-encoded subunits of eukaryotic complex I and to subunits of the E. coli formate-hydrogen lyase complex (17, 29, 30, 49). Whereas the nuclear homologs NuoE, NuoF, and NuoG constitute the peripheral fragment (also referred to as the NADH dehydrogenase fragment [NDF]), the nuclear homologs NuoB, NuoC, NuoD, and NuoI constitute the connecting fragment. The mitochondrial homologs NuoA, NuoH, NuoJ, NuoK, NuoL, NuoM, and NuoN constitute the membrane fragment (29). E. coli complex I likely evolved by fusion of preexisting protein assemblies constituting modules for electron transfer and proton translocation (1719, 30). Open in a separate windowFIG. 1Schematic of E. coli complex I and the nuo locus. Adapted with permission of the publisher (17, 29, 30, 49). (A) E. coli complex I is comprised of three distinct fragments: the peripheral (light gray), connecting (white), and membrane (dark gray) fragments (17, 29). The peripheral fragment (NDF) is comprised of the nuclear homologs NuoE, -F, and -G and exhibits NADH dehydrogenase activity that oxidizes NADH to NAD+; the connecting fragment is comprised of the nuclear homologs NuoB, -C, -D, and -I; and the membrane fragment is comprised of the mitochondrial homologs NuoA, -H, and -J to -N and catalyzes ubiquinone (Q) to its reduced form (QH2). FMN, flavin mononucleotide. (B) The E. coli nuo locus encodes the 14 Nuo subunits that constitute complex I. The 5′ half of the locus contains a previously identified promoter (nuoP) and the majority of genes that encode the peripheral and connecting subunits (light gray and white, respectively). The 3′ half of the locus contains the majority of the genes encoding the membrane subunits (dark gray). The 3′ end of nuoG encodes a C-Terminal region (CTR) of the NuoG subunit (hatched).Because of its smaller size and relative simplicity, researchers recently have begun to utilize complex I of E. coli, and that of its close relative S. typhimurium, to identify and characterize the mechanism(s) by which cells regulate the synthesis and assembly of this large respiratory complex (3, 8, 46) and to investigate the diverse physiological consequences caused by defects in this enzyme (4, 6, 10, 40, 59). Such defects affect the ability of cells to perform chemotaxis (40), to grow on certain carbon sources (4, 6, 10, 40, 57), to survive stationary phase (59), to perform energy-dependent proteolysis (4), to regulate the expression of at least one gene (32), and to maintain virulence (5).To begin dissecting the processes by which E. coli cells regulate the expression of nuo and the assembly of complex I, we undertook a genetic analysis of the nuo locus. Here, we present the results of studies, performed on an isogenic collection of nuo mutants, that focus on the physiological, biochemical, and molecular consequences caused by the lack of or defects in several Nuo subunits. In particular, we present evidence that NuoG, a peripheral subunit, is essential for complex I function and that it plays a role in the regulation of nuo expression and/or the assembly of complex I.  相似文献   

3.
Cryptochromes and DNA photolyases are related flavoproteins with flavin adenine dinucleotide as the common cofactor. Whereas photolyases repair DNA lesions caused by UV radiation, cryptochromes generally lack repair activity but act as UV-A/blue light photoreceptors. Two distinct electron transfer (ET) pathways have been identified in DNA photolyases. One pathway uses within its catalytic cycle, light-driven electron transfer from FADH* to the DNA lesion and electron back-transfer to semireduced FADHo after photoproduct cleavage. This cyclic ET pathway seems to be unique for the photolyase subfamily. The second ET pathway mediates photoreduction of semireduced or fully oxidized FAD via a triad of aromatic residues that is conserved in photolyases and cryptochromes. The 5,10-methenyltetrahydrofolate (5,10-methenylTHF) antenna cofactor in members of the photolyase family is bleached upon light excitation. This process has been described as photodecomposition of 5,10-methenylTHF. We show that photobleaching of 5,10-methenylTHF in Arabidopsis cry3, a member of the cryptochrome DASH family, with repair activity for cyclobutane pyrimidine dimer lesions in single-stranded DNA and in Escherichia coli photolyase results from reduction of 5,10-methenylTHF to 5,10-methyleneTHF that requires the intact tryptophan triad. Thus, a third ET pathway exists in members of the photolyase family that remained undiscovered so far.DNA photolyases and cryptochromes (cry)2 form a large family of related flavoproteins with DNA repair activity and photoreceptor function, respectively. Members of this protein family were identified in all kingdoms of life and can be grouped in at least nine subclades (1). DNA photolyases repair cytotoxic and mutagenic DNA lesions that are formed during exposure of DNA to UV-B. These DNA lesions are cyclobutane pyrimidine dimers (CPDs) or pyrimidine-pyrimidone (6-4) photoproducts. According to their substrate specificity, DNA photolyases are designated as CPD photolyases or (6-4) photolyases (2). The repair of both types of DNA lesions by photolyase requires the catalytic fully reduced and anionic flavin cofactor FADH that, when photoexcited, injects an electron directly into the DNA lesion (1) as shown in Fig. 1A (electron transfer pathway 1). During extraction from the cell and purification under aerobic conditions the flavin cofactor is usually oxidized to the semireduced and eventually to the fully oxidized form. Reduction of these flavin species to FADH in vitro can be achieved by illumination of the enzyme in the presence of reducing agents such as dithiothreitol or β-mercaptoethanol. This process is named photoactivation (1). Photoactivation in vitro requires photoexcitation of the flavin and a triad of redox-active residues in the protein moiety that is highly conserved in DNA photolyases (3, 4) as shown in Fig. 1A (electron transfer pathway 2). These residues are generally tryptophans that allow transport of an electron from the protein surface to the U-shaped flavin cofactor, which is buried within the C-terminal α-helical domain (59). Whether the same mechanism is used by photolyase to photoreduce FAD in vivo is a matter of debate (10). Photoreduction of the flavin cofactor was also observed in cryptochrome blue/UV-A photoreceptors. However, instead of fully reduced flavin, semireduced flavin species (either anionic flavin semiquinone radical or neutral semiquinone radical) accumulate. This form of the photoreceptor is considered as the signaling state (1114).Open in a separate windowFIGURE 1.Electron transfer pathways in cry3 and structures of folates. A, indicated are the distances of the tryptophans in the tryptophan triad (Trp-356, -409, -432) of Trp-432 to FADH and of FADH to the 5,10-methenylTHF (MTHF) cofactor in cry3. Shown are also the two established routes of electrons from FADH to the DNA lesion (Route 1) and within the tryptophan triad to FAD (Route 2). The third electron transfer pathway from FADH to 5,10-methenylTHF (Route 3) is the subject of this study. B, chemical structures of folates and their molecular masses. Folypolyglutamate molecules have a pteridin and a p-aminobenzoate moiety linked with a glutamate chain with a variable number of glutamic acids. The various THF species differ in their oxidation state of the C1 unit that is attached at the N-5 or N-10 position or form a bridge between both.A recently discovered subclade of the DNA photolyase/cryptochrome family are DASH cryptochromes, which have members in plants, bacteria, and aquatic animals (6, 1517). Because DASH cryptochromes were found to lack repair activity for CPDs in double-stranded DNA, they were considered as cryptochrome-type photoreceptors (6, 16). However, it was recently shown that DASH cryptochromes repair CPDs in single-stranded DNA (18) and loop structures of double-stranded DNA (19) and, thus, belong to the CPD photolyase group. In contrast to conventional CPD photolyases, DASH cryptochromes are unable to flip the CPD lesion out of the DNA duplex (7).Besides the flavin cofactor that is essential for enzymatic activity, DNA photolyases and most likely all cryptochromes contain a second chromophore (1). Like the catalytic flavin, the second chromophore is non-covalently attached to the protein moiety. The majority of DNA photolyases and, as far as studied, the cryptochromes including the DASH-type like cry3 from Arabidopsis thaliana contain polyglutamated 5,10-methenyltetrahydrofolate (5,10-methenylTHF) as the second chromophore (1, 12, 17, 20, 21) (see Fig. 1B for folate structures). Several organisms like the cyanobacterium Anacystis nidulans (Synechococcus elongatus) produce deazariboflavins (7,8-didemethyl-8-hydroxy-5-deazariboflavin) and utilize them as second cofactor (22). In photolyases of thermophilic bacteria and Archaea of the genus Sulfolobus, FMN and FAD, respectively, were found as second cofactors (23, 24). The sole function of the second cofactors demonstrated at present is transfer of excitation energy to the catalytic flavin cofactor via a Förster-type mechanism. The crystal structures of DNA photolyases and DASH cryptochromes revealed that the second chromophores are located in a cleft between the N-terminal α/β domain and the C-terminal α domain (79). The centroid distances between the catalytic FAD and the second chomophore are in the range of 15–18 Å. The close distances and the angles between the transition dipole moments of the two cofactors are favorable for efficient energy transfer. Indeed, energy transfer efficiencies are about 70% for Escherichia coli photolyase (25), close to 100% for A. nidulans photolyase (26), and between 78% (dark-adapted) and 87% (light-adapted) for Arabidopsis cry3 (27). Although the second cofactors are not essential for catalysis (28, 29), they increase the efficiency of repair and possibly of photoactivation by having higher extinction coefficients than FADH in the near UV and blue region (30). The spectral overlap between 5,10-methenylTHF emission and the absorption of the different flavin redox states is on the order FADHo > FADox > FADH (31).Illumination in vitro of photolyase that contains fully oxidized or semireduced flavin results in light-induced absorbance changes. The decrease in absorption in the 450–470-nm region reflects a decrease in the amount of fully oxidized FAD concomitant with transient increase in absorption above 500 nm, which indicates the formation of a neutral semiquinone radical. Excitation of the 5,10-methenylTHF antenna chromophore at its absorption peak at 380 nm causes a likewise photoreduction of the catalytic FAD (1, 27, 28, 30, 31). However, irreversible bleaching of the 380-nm peak is observed under high irradiance UV-A or camera flash illumination (28, 30). This irreversible bleaching goes along with release of the folate cofactor from the protein moiety (30) and was named photodecomposition of 5,10-methenylTHF (28). However, the identity of the formed folate species remained unknown (30). In our previous spectroscopic characterization of Arabidopsis cry3, a similar bleaching of the 380-nm peak was observed (27).Here we show that a third electron transfer pathway exists in photolyase and DASH cryptochome, where the 5,10-methenylTHF cofactor is photoreduced to 5,10-methyleneTHF. Thus, bleaching at 380 nm does not simply reflect destruction but is a specific chemical conversion of the second chromophore.  相似文献   

4.
5.
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   

6.
A mutant (JY2190) of Streptococcus pneumoniae Rx1 which had acquired the ability to grow in the absence of choline and analogs was isolated. Lipoteichoic acid (LTA) and wall teichoic acid (TA) isolated from the mutant were free of phosphocholine and other phosphorylated amino alcohols. Both polymers showed an unaltered chain structure and, in the case of LTA, an unchanged glycolipid anchor. The cell wall composition was also not altered except that, due to the lack of phosphocholine, the phosphate content of cell walls was half that of the parent strain. Isolated cell walls of the mutant were resistant to hydrolysis by pneumococcal autolysin (N-acetylmuramyl-l-alanine amidase) but were cleaved by the muramidases CPL and cellosyl. The lack of active autolysin in the mutant cells became apparent by impaired cell separation at the end of cell division and by resistance against stationary-phase and penicillin-induced lysis. As a result of the absence of choline in the LTA, pneumococcal surface protein A (PspA) was no longer retained on the cytoplasmic membrane. During growth in the presence of choline, which was incorporated as phosphocholine into LTA and TA, the mutant cells separated normally, did not release PspA, and became penicillin sensitive. However, even under these conditions, they did not lyse in the stationary phase, and they showed poor reactivity with antibody to phosphocholine and an increased release of C-polysaccharide from the cell. In contrast to ethanolamine-grown parent cells (A. Tomasz, Proc. Natl. Acad. Sci. USA 59:86–93, 1968), the choline-free mutant cells retained the capability to undergo genetic transformation but, compared to Rx1, with lower frequency and at an earlier stage of growth. The properties of the mutant could be transferred to the parent strain by DNA of the mutant.Pneumococci differ from other gram-positive bacteria in that their lipoteichoic acid (LTA) and wall teichoic acid (TA) have the same chain structure which is, moreover, unusually complex (Fig. (Fig.1):1): glycerophosphate is replaced by ribitol phosphate (7), and between the ribitol phosphate residues a tetrasaccharide is intercalated (23). It contains d-glucose, 2-acetamido-4-amino-2,4,6-trideoxy-d-galactose (AATGal), and two N-acetyl-d-galactosaminyl residues, one or both of which carry a phosphocholine residue at O-6 (references 3 and 12 and this report). Open in a separate windowFIG. 1Pneumococcal TA and LTA. As shown, in strain R6 most of the repeats carry two phosphocholine residues each, at O-6 of the N-acetyl-d-galactosaminyl residues (3, 12). In strain Rx1 and Rx1/AL, most repeats contain one phosphocholine residue (this report) attached to O-6 of the non-ribitol-linked galactosaminyl residue (14).Pneumococci are not able to synthesize the choline required for the synthesis of these substituents. Moreover, choline is an essential growth factor (2, 30) but can be substituted in this function by nutritional ethanolamine (EA) (38). Phosphoethanolamine is incorporated into LTA and TA in place of phosphocholine (14), but it cannot replace phosphocholine functionally. Phosphocholine-substituted LTA serves to anchor pneumococcal surface protein A (PspA) to the outer layer of the cytoplasmic membrane, with choline-mediated interaction between membrane-associated LTA and the C-terminal repeat region of PspA. In EA-grown bacteria, PspA is no longer retained and is released into the surrounding medium (45). Phosphocholine substituents also play an essential role for the activity of the major pneumococcal autolysin, an N-acetylmuramyl-l-alanine amidase (38). This protein possesses a choline-binding C-terminal domain that is essential for activity but, unlike PspA, is not essential for retention on the pneumococcal cell surface (16, 32). Binding of phosphocholine-substituted LTA to this domain results in potent inhibition of the amidase (21). The inhibitory property is dependent on the micellar structure of LTA (13) and lost by deacylation (5). Phosphocholine-substituted LTA may also participate in the transport of the amidase through the cytoplasmic membrane from the cytosol (5), the location of its synthesis (15). It additionally effects the conversion of the inactive E form of the enzyme into the active C form (5). This conversion is likewise effected by the choline residues of cell wall-linked TA (33, 39). Furthermore, binding of the amidase to the choline residues of TA is prerequisite for the hydrolysis of cell walls by the enzyme (18, 22). It should be noted that the amidase is not essential for growth. Though the enzyme is completely inactive in EA grown cells, the growth rate is not affected. However, cell separation is impaired, and there is a loss of stationary-phase and penicillin-induced cell lysis (38, 40), as well as a loss of genetic transformation (38). After insertional inactivation of the autolysin gene (lytA), the autolysin-deficient mutants (Lyt) grew normally (31) and did not even show impeded cell separation (41).In this report, we describe a mutant which acquired the ability to grow in the absence of choline and analogs. Except for the observation that [3H]choline-substituted LTA is not a precursor of [3H]choline-substituted TA (6), nothing is known about the biosyntheses of pneumococcal LTA and TA and the stage of biosynthesis at which phosphocholine is incorporated. Since the absence of choline incorporation might affect the structure of LTA and TA as well as the composition of cell walls, we included relevant analyses in our study.(A preliminary report of this work was presented in an overview on pneumococcal LTA and TA at the International Meeting on the Molecular Biology of Streptococcus pneumoniae and Its Diseases, Oeiras, Portugal, September 24 to 29, 1996 [10].)  相似文献   

7.
8.
9.
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   

10.
The mechanisms by which mutant variants of Cu/Zn-superoxide dismutase (SOD1) cause familial amyotrophic lateral sclerosis are not clearly understood. Evidence to date suggests that altered conformations of amyotrophic lateral sclerosis mutant SOD1s trigger perturbations of cellular homeostasis that ultimately cause motor neuron degeneration. In this study we correlated the metal contents and disulfide bond status of purified wild-type (WT) and mutant SOD1 proteins to changes in electrophoretic mobility and surface hydrophobicity as detected by 1-anilinonaphthalene-8-sulfonic acid (ANS) fluorescence. As-isolated WT and mutant SOD1s were copper-deficient and exhibited mobilities that correlated with their expected negative charge. However, upon disulfide reduction and demetallation at physiological pH, both WT and mutant SOD1s underwent a conformational change that produced a slower mobility indicative of partial unfolding. Furthermore, although ANS did not bind appreciably to the WT holoenzyme, incubation of metal-deficient WT or mutant SOD1s with ANS increased the ANS fluorescence and shifted its peak toward shorter wavelengths. This increased interaction with ANS was greater for the mutant SOD1s and could be reversed by the addition of metal ions, especially Cu2+, even for SOD1 variants incapable of forming the disulfide bond. Overall, our findings support the notion that misfolding associated with metal deficiency may facilitate aberrant interactions of SOD1 with itself or with other cellular constituents and may thereby contribute to neuronal toxicity.The sequence of events by which more than 100 mutations in the gene encoding Cu/Zn-superoxide dismutase (SOD1)3 cause familial forms of amyotrophic lateral sclerosis (ALS) is unknown. Studies of purified SOD1 proteins and cellular or rodent models of SOD1-linked ALS suggest that impaired metal ion binding or misfolding of mutant SOD1 proteins in the cellular environment may be related to their toxicity (110). Available evidence suggests that partially unfolded mutant SOD1 species could contribute to motor neuron death by promoting abnormal interactions that produce cellular dysfunction (1116).In previous studies we characterized physicochemical properties of 14 different biologically metallated ALS SOD1 mutants (17) and demonstrated altered thermal stabilities of these mutants compared with wild-type (WT) SOD1 (18). These “as-isolated” SOD1 proteins, which contain variable amounts of copper and zinc, were broadly grouped into two classes based on their ability to incorporate and retain metal ions with high affinity. WT-like SOD1 mutants retain the ability to bind copper and zinc ions and exhibit dismutase activity similar to the normal enzyme, whereas metal binding region (MBR) mutants are significantly deficient in copper and/or zinc (17, 19). We also observed that ALS-associated SOD1 mutants were more susceptible than the WT enzyme to reduction of the intrasubunit disulfide bond between Cys-57 and Cys-146 (20). The significance of these results is that even WT-like mutants, which exhibit a nearly normal backbone structure (2123), may be vulnerable to destabilizing influences in vivo. Our group and others subsequently showed that the mutant SOD1 proteins share a susceptibility to increased hydrophobicity under conditions that reduce disulfide bonds and/or chelate metal ions (5) and that similar hydrophobic species exist in tissue lysates from mutant SOD1 transgenic mice (46). One consequence of such hydrophobic exposure could be the facilitation of abnormal interactions between the mutant enzymes and other cellular constituents (e.g. chaperones, mitochondrial components, or other targets), which might influence pathways leading to motor neuron death (15, 16, 2427).Accumulating evidence suggests that metal deficiency of SOD1 is an important factor that can influence SOD1 aggregation or neurotoxicity (4, 2833), but the metal-deficient states of SOD1 that are most relevant to ALS remain unclear. Zinc-deficient, copper-replete SOD1 species, which can be produced in vitro by adding copper to SOD1 that has been stripped of its metal ions at acidic pH, were shown to be toxic to motor neurons in culture (28). However, it has not been shown that zinc-deficient, copper-replete SOD1 is produced in vivo as a consequence of ALS mutations, and loading of copper into SOD1 by the copper chaperone for SOD1 (CCS) is not required for toxicity (34, 35). Furthermore, the MBR mutants have a disrupted copper site and have been found to be severely deficient in both zinc and copper (17, 30), yet expression of these SOD1s still produces motor neuron disease (1, 2, 30, 34, 36, 37).When recombinant human SOD1 was overexpressed in insect cells, we instead observed zinc-replete but copper-deficient species for most WT-like mutants, probably because the capacity of the copper-loading mechanism was exceeded (17). These preparations indicate that zinc can be efficiently incorporated into many WT-like mutants in vivo, and much of it is retained after purification. Furthermore, these copper-deficient biologically metallated proteins may be useful reagents to assess the influence of copper binding upon other properties of SOD1 mutants that may be relevant to their neurotoxicity.We previously observed that reduction of the Cys-57—Cys-146 disulfide bond facilitates the ability of metal chelators to alter the electrophoretic mobility and to increase the hydrophobicity of SOD1 mutants (5). This is consistent with the known properties of this linkage to stabilize the dimeric interface, to orient Arg-143 via a hydrogen bond from the carbonyl oxygen of Cys-57 to Arg-143-NH2, and to prevent metal ion loss (3840). However, it remains unclear whether the Cys-57—Cys-146 bond is required to prevent abnormal SOD1 hydrophobic exposure or whether the aberrant conformational change primarily results from metal ion loss. Ablation of the disulfide bond by the experimental (non-ALS) mutants C57S and C146S provides useful reagents to test the relative influence of the disulfide bond and copper binding upon SOD1 properties.In this study we sought to correlate the consequences of copper deficiency, copper and zinc deficiency, and disulfide reduction upon the hydrodynamic behavior and surface hydrophobicity of WT and representative mutant SOD1 enzymes (Fig. 1A). We quantitated the metal contents of as-isolated SOD1 proteins, detected changes in conformation or metal occupancy using native PAGE to assess their electrophoretic mobility, a measure of global conformational change, and correlated these changes to hydrophobic exposure using 1-anilinonaphthalene-8-sulfonic acid (ANS), which is very sensitive to local conformational changes. ANS is a small amphipathic dye (Fig. 1B) that has been used as a sensitive probe to detect hydrophobic pockets on protein surfaces (4144). Free ANS exhibits only weak fluorescence that is maximal near 520 nm, but when ANS binds to a hydrophobic site in a partially or fully folded protein, the fluorescence peak increases in amplitude and shifts to a shorter wavelength (42). ANS also has an anionic sulfonate group that can interact with cationic groups (e.g. Arg or Lys residues) through ion-pair formation which may be further strengthened by hydrophobic interactions (4346).Open in a separate windowFIGURE 1.A, WT SOD1 structure showing the position of the C57-C146 intrasubunit disulfide bond (S–S, yellow), bound copper and zinc ions, and ALS mutant residues. The residues altered in A4V, G85R, G93A, D124V, and S134N SOD1s are indicated as green spheres. The backbone of the β-barrel core and the loops is shown in a rainbow color, from blue at the amino terminus to red at the carboxyl terminus. The figure was generated using PyMOL (84) and PDB entry 1HL5 (22). B, chemical structure of ANS fluorophore.To evaluate further the importance of metal ion binding, we measured spectral changes related to the binding of cobalt and copper to the same SOD1 proteins. We observed that as-isolated WT-like mutants containing zinc could interact with copper ions to produce an electrophoretic mobility and decreased hydrophobicity resembling that of the fully metalated holo-WT SOD1. In contrast, we saw no evidence for copper binding to MBR mutants in a manner that alters their hydrodynamic properties or their hydrophobicity. Our data suggest that binding of both copper and zinc are important determinants of SOD1 conformation and that perturbation of such binding may be relevant to the ALS disease process.  相似文献   

11.
12.
13.
14.
15.
16.
17.
18.
19.
Fibronectin (FN) is a large extracellular matrix glycoprotein important for development and wound healing in vertebrates. Recent work has focused on the ability of FN fragments and embryonic or tumorigenic splicing variants to stimulate fibroblast migration into collagen gels. This activity has been localized to specific sites and is not exhibited by full-length FN. Here we show that an N-terminal FN fragment, spanning the migration stimulation sites and including the first three type III FN domains, also lacks this activity. A screen for interdomain interactions by solution-state NMR spectroscopy revealed specific contacts between the Fn N terminus and two of the type III domains. A single amino acid substitution, R222A, disrupts the strongest interaction, between domains 4–5FnI and 3FnIII, and restores motogenic activity to the FN N-terminal fragment. Anastellin, which promotes fibril formation, destabilizes 3FnIII and disrupts the observed 4–5FnI-3FnIII interaction. We discuss these findings in the context of the control of cellular activity through exposure of masked sites.Fibronectin (FN),4 a large multidomain glycoprotein found in all vertebrates, plays a vital role in cell adhesion, tissue development, and wound healing (1). It exists in soluble form in plasma and tissue fluids but is also present in fibrillar networks as part of the extracellular matrix. The structures of many FN domains of all three types, FnI, FnII, and FnIII, are known, for example (24). Although interactions between domains that are close in primary sequence have been demonstrated (3, 5), studies of multidomain fragments generally assume a beads-on-string model (2). There is, however, much evidence for the presence of long range order in soluble FN as a number of functional sites, termed cryptic, are not active in the native molecule, until exposed through conformational change. These include self-association sites (58), sites of protein interactions (9), and sites that control cellular activity (10, 11). Low resolution studies of the FN dimer suggest a compact conformation under physiological conditions (1214); however, attempts to define large scale structure in FN by small angle scattering or electric birefringence (1517) have yielded contradictory results. Interpretation of domain stability changes in terms of interaction sites (18) has also not been straightforward (2), possibly because of domain stabilization through nearest-neighbor effects (19, 20).A FN splicing variant produced in fetal and cancer patient fibroblasts, termed migration stimulation factor (MSF), stimulates migration of adult skin fibroblasts into type I collagen gels (10, 21) and breast carcinoma cells using the Boyden chamber (22). MSF comprises FN domains 1FnI to 9FnI, a truncated 1FnIII, and a small C-terminal extension; a recombinant FN fragment corresponding to 1FnI-9FnI (Fn70kDa) displays the same activity (10). An overview of FN domain structure and nomenclature is presented in Fig. 1a. Further experiments sub-localized full motogenic activity to the gelatin binding domain of FN (GBD, domains 6FnI-9FnI) (23) and partial activity to a shorter fragment spanning domains 7–9FnI (24). Two IGD tripeptides of domains 7FnI and 9FnI were shown to be essential through residue substitutions and reconstitution of partial motogenic activity in synthetic peptides (10, 24, 25); however, similar IGD tripeptides outside the GBD, on domains 3FnI and 5FnI, appear to have little effect (10, 23). Full-length adult FN does not affect cell migration in similar assays (10, 23); thus motogenic activity sites are presumed to be masked in the conformation adopted by soluble FN, although they could be exposed by molecular rearrangement.Open in a separate windowFIGURE 1.Motogenic activity of FN fragments. a, schematic representation of the FN domain structure (top) and enlargement of the FN N terminus (bottom). Type I domains are shown as pentagons; type II domains as hexagons; and type III domains as ovals. b, comparison of motogenic activity versus protein concentration of wild-type Fn70kDa and Fn100kDa fragments. Error bars are derived from duplicate experiments, and a gray band denotes migration activity of media without additives. c, similar comparisons for mutant Fn100kDa fragments. d, analytical size exclusion chromatography of large FN fragments. The trace of UV absorbance at 280 nm versus elution volume shown here indicates a larger hydrodynamic radius for Fn100kDa R222A compared with the wild type, consistent with our model (Fig. 6a).Here we show that a recombinant fragment, closely matching a truncated form of FN identified in zebrafish (26), as well as amphibians, birds, and mammals (27), does not stimulate cell migration. This fragment is similar to MSF but includes the first three FnIII domains (1–3FnIII), suggesting that these domains are responsible for a conformational transition that masks the activity sites in this construct and probably in full-length FN. To identify the mechanism behind this transition, we performed structural studies by solution NMR spectroscopy and identified a specific long range interaction between domains 4–5FnI and 3FnIII as essential for this masking effect. Interestingly, this interaction does not involve direct contacts with the GBD but possibly represses motogenic activity through chain compaction, evident in analytical size exclusion assays. Intramolecular interactions thus provide a mechanism by which conformational rearrangement induced, for example, by tension or splicing variation can result in cellular activity differences.  相似文献   

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

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