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 531 毫秒
1.
N-terminal or C-terminal arms that extend from folded protein domains can play a critical role in quaternary structure and other intermolecular associations and/or in controlling biological activity. We have tested the role of an extended N-terminal arm in the structure and function of a periplasmic enzyme glucose-fructose oxidoreductase (GFOR) from Zymomonas mobilis. We have determined the crystal structure of the NAD(+) complex of a truncated form of the enzyme, GFORDelta, in which the first 22 residues of the N-terminal arm of the mature protein have been deleted. The structure, refined at 2.7 A resolution (R(cryst)=24.1%, R(free)=28.4%), shows that the truncated form of the enzyme forms a dimer and implies that the N-terminal arm is essential for tetramer formation by wild-type GFOR. Truncation of the N-terminal arm also greatly increases the solvent exposure of the cofactor; since GFOR activity is dependent on retention of the cofactor during the catalytic cycle we conclude that the absence of GFOR activity in this mutant results from dissociation of the cofactor. The N-terminal arm thus determines the quaternary structure and the retention of the cofactor for GFOR activity and during translocation into the periplasm. The structure of GFORDelta also shows how an additional mutation, Ser64Asp, converts the strict NADP(+) specificity of wild-type GFOR to a dual NADP(+)/NAD(+) specificity.  相似文献   

2.
Recombinant 1,5-anhydro-d-fructose reductase (AFR) from Sinorhizobium morelense S-30.7.5 was crystallized in complex with the cofactor NADP(H) and its structure determined to 2.2 A resolution using selenomethionine SAD (refined R(work) and R(free) factors of 18.9 and 25.0%, respectively). As predicted from the sequence and shown by the structure, AFR can be assigned to the GFO/IDH/MocA protein family. AFR consists of two domains. The N-terminal domain displays a Rossmann fold and contains the cofactor binding site. The intact crystals contain the oxidized cofactor NADP(+), whose attachment to the cofactor binding site is similar to that of NADP(+) in glucose-fructose oxidoreductase (GFOR) from Zymomonas mobilis. Due to variations in length and sequence within loop regions L3 and L5, respectively, the adenine moiety of NADP(+) adopts a different orientation in AFR caused by residue Arg38 forming hydrogen bonds with the 2'-phosphate moiety of NADP(+) and cation-pi stacking interactions with the adenine ring. Amino acid replacements in AFR (S10G, A13G, and S33D) showed that Ala13 is crucial for the discrimination between NADPH and NADH and yielded the A13G variant with dual cosubstrate specificity. The C-terminal domain contains the putative substrate binding site that was occupied by an acetate ion. As determined by analogy to GFOR and by site-directed mutagenesis of K94G, D176A, and H180A, residues Lys94, Asp176, and His180 are most likely involved in substrate binding and catalysis, as substitution of any of these residues resulted in a significant decrease in k(cat) for 1,5-AF. In this context, His180 might serve as a general acid-base catalyst by polarizing the carbonyl function of 1,5-AF to enable the transfer of the hydride from NADPH to the substrate. Here we present the first structure of an AFR enzyme catalyzing the stereoselective reduction of 1,5-AF to 1,5-anhydro-d-mannitol, the final step of a modified anhydrofructose pathway in S. morelense S-30.7.5. We also emphasize the importance of the A13G variant in biocatalysis for the synthesis of 1,5-AM and related derivatives.  相似文献   

3.
Abstract The enzyme glucose-fructose oxidoreductase (GFOR) from the Gram-negative ethanologenic bacterium Zymomonas mobilis was purified to homogeneity and was shown to be a tetrameric protein with a subunit size of M r 42 500. Using immunogold-labelling in combination with electron microscopy, ultrathin sections of Z. mobilis wild type cells showed that the enzyme GFOR is located in the periplasm off the bacterial cells. Z. mobilis strains which carried the cloned gfo gene on plasmid pSUP104, had 5–6-fold increased GFOR enzyme activities. Moreover, these cells accumulated large amounts of a presumable unprocessed pre-GFOR protein ( M r 48 000).  相似文献   

4.
Glucose-fructose oxidoreductase (GFOR) of the gram-negative bacterium Zymomonas mobilis is a periplasmic enzyme with the tightly bound cofactor NADP. The preprotein carries an unusually long N-terminal signal sequence of 52 amino acid residues. A sorbitol-negative mutant strain (ACM3963) was found to be deficient in GFOR activity and was used for the expression of plasmid-borne copies of the wild-type gfo gene or of alleles encoding alterations in the signal sequence of the pre-GFOR protein. Z. mobilis cells with the wild-type gfo allele translocated pre-GFOR, at least partially, via the Sec pathway since CCCP (carboxylcyanide-m-chlorophenylhydrazone; uncoupler of proton motive force) or sodium azide (inhibitor of SecA) abolished the processing of GFOR. A gfo allele with the hydrophobic region of the signal sequence removed (residues 32–46; Δ32–46) led to a protein that was no longer processed, but showed full enzymatic activity (180 U/mg) and had the cofactor NADP firmly bound. A deletion in the n-region of the signal sequence (residues 2–20; Δ2–20) or exchange of the entire GFOR signal sequence with the signal sequence of gluconolactonase of Z. mobilis led to active and processed GFOR. Strain ACM3963 could not grow in the presence of high sugar concentrations (1 M sucrose) unless sorbitol was added. The presence of the plasmid-borne gfo wild-type allele or of the Δ2–20 deletion led to the restoration of growth on media with 1 M sucrose, whereas the presence of the Δ32–46 deletion led to a growth behavior similar to that of strain ACM3963, with no sorbitol formation from sucrose. Received: 14 December 1995 / Accepted: 4 March 1996  相似文献   

5.
Glucose-fructose oxidoreductase (E.C. 1.1.99.-) from the ethanol-producing Gram-negative bacterium Zymomonas mobilis is a periplasmic, soluble enzyme that forms a homotetramer of 160 kDa with one NADP(H) cofactor per subunit that is tightly, but noncovalently, bound. The enzyme was crystallized by the hanging drop vapor diffusion method using sodium citrate as precipitant. The obtained crystals belong to the space group P2(1)2(1)2, with unit cell constants of 84.6 A, 94.1 A, and 117.0 A, consistent with two monomers in the asymmetric unit. They diffract to a resolution of about 2 A and are suitable for X-ray structure determination.  相似文献   

6.
The quaternary protein structure of two methanol:N,N'-dimethyl-4-nitrosoaniline (NDMA) oxidoreductases purified from Amycolatopsis methanolica and Mycobacterium gastri MB19 was analyzed by electron microscopy and image processing. The enzymes are decameric proteins (displaying fivefold symmetry) with estimated molecular masses of 490 to 500 kDa based on their subunit molecular masses of 49 to 50 kDa. Both methanol:NDMA oxidoreductases possess a tightly but noncovalently bound NADP(H) cofactor at an NADPH-to-subunit molar ratio of 0.7. These cofactors are redox active toward alcohol and aldehyde substrates. Both enzymes contain significant amounts of Zn2+ and Mg2+ ions. The primary amino acid sequences of the A. methanolica and M. gastri MB19 methanol:NDMA oxidoreductases share a high degree of identity, as indicated by N-terminal sequence analysis (63% identity among the first 27 N-terminal amino acids), internal peptide sequence analysis, and overall amino acid composition. The amino acid sequence analysis also revealed significant similarity to a decameric methanol dehydrogenase of Bacillus methanolicus C1.  相似文献   

7.
8.
The pyridine nucleotide transhydrogenase of Escherichia coli has an alpha 2 beta 2 structure (alpha: Mr, 54,000; beta: Mr, 48,700). Hydropathy analysis of the amino acid sequences suggested that the 10 kDa C-terminal portion of the alpha subunit and the N-terminal 20-25 kDa region of the beta subunit are composed of transmembranous alpha-helices. The topology of these subunits in the membrane was investigated using proteolytic enzymes. Trypsin digestion of everted cytoplasmic membrane vesicles released a 43 kDa polypeptide from the alpha subunit. The beta subunit was not susceptible to trypsin digestion. However, it was digested by proteinase K in everted vesicles. Both alpha and beta subunits were not attacked by trypsin and proteinase K in right-side out membrane vesicles. The beta subunit in the solubilized enzyme was only susceptible to digestion by trypsin if the substrates NADP(H) were present. NAD(H) did not affect digestion of the beta subunit. Digestion of the beta subunit of the membrane-bound enzyme by trypsin was not induced by NADP(H) unless the membranes had been previously stripped of extrinsic proteins by detergent. It is concluded that binding of NADP(H) induces a conformational change in the transhydrogenase. The location of the trypsin cleavage sites in the sequences of the alpha and beta subunits were determined by N- and C-terminal sequencing. A model is proposed in which the N-terminal 43 kDa region of the alpha subunit and the C-terminal 30 kDa region of the beta subunit are exposed on the cytoplasmic side of the inner membrane of E. coli. Binding sites for pyridine nucleotide coenzymes in these regions were suggested by affinity chromatography on NAD-agarose columns.  相似文献   

9.
During growth of the halophilic archaeon Haloarcula marismortui on D-xylose, a specific D-xylose dehydrogenase was induced. The enzyme was purified to homogeneity. It constitutes a homotetramer of about 175 kDa and catalyzed the oxidation of xylose with both NADP+ and NAD+ as cosubstrates with 10-fold higher affinity for NADP+. In addition to D-xylose, D-ribose was oxidized at similar kinetic constants, whereas D-glucose was used with about 70-fold lower catalytic efficiency (kcat/Km). With the N-terminal amino acid sequence of the subunit, an open reading frame (ORF)-coding for a 39.9-kDA protein-was identified in the partially sequenced genome of H. marismortui. The function of the ORF as the gene designated xdh and coding for xylose dehydrogenase was proven by its functional overexpression in Escherichia coli. The recombinant enzyme was reactivated from inclusion bodies following solubilization in urea and refolding in the presence of salts, reduced and oxidized glutathione, and substrates. Xylose dehydrogenase showed the highest sequence similarity to glucose-fructose oxidoreductase from Zymomonas mobilis and other putative bacterial and archaeal oxidoreductases. Activities of xylose isomerase and xylulose kinase, the initial reactions of xylose catabolism of most bacteria, could not be detected in xylose-grown cells of H. marismortui, and the genes that encode them, xylA and xylB, were not found in the genome of H. marismortui. Thus, we propose that this first characterized archaeal xylose dehydrogenase catalyzes the initial step in xylose degradation by H. marismortui.  相似文献   

10.
Shikimate dehydrogenase (EC 1.1.1.25) catalyses the fourth step of the shikimate pathway which is required for the synthesis of the aromatic amino acids and other aromatic compounds in bacteria, microbial eukaryotes, and plants. The crystal structures of the shikimate dehydrogenase AroE from Thermus thermophilus HB8 in its ligand-free form, binary complexes with cofactor NADP+ or substrate shikimate, and the ternary complex with both NADP(H) and shikimate were determined by X-ray diffraction method at atomic resolutions. The crystals are nearly isomorphous with the asymmetric unit containing a dimer, each subunit of which has a bi-domain structure of compact alpha/beta sandwich folds. The two subunits of the enzyme display asymmetry in the crystals due to different relative orientations between the N- and C-terminal domains resulting in a slightly different closure of the interdomain clefts. NADP(H) is bound to the more closed form only. This closed conformation with apparent higher affinity to the cofactor is also observed in the unliganded crystal form, indicating that the NADP(H) binding to TtAroE may follow the selection mode where the cofactor binds to the subunit that happens to be in the closed conformation in solution. Crystal structures of the closed subunits with and without NADP(H) show no significant structural difference, suggesting that the cofactor binding to the closed subunit corresponds to the lock-and-key model in TtAroE. On the other hand, shikimate binds to both open and closed subunit conformers of both apo and NADP(H)-liganded holo enzyme forms. The ternary complex TtAroE:NADP(H):shikimate allows unambiguous visualization of the SDH permitting elucidation of the roles of conserved residues Lys64 and Asp100 in the hydride ion transfer between NADP(H) and shikimate.  相似文献   

11.
D Nurizzo  D Halbig  G A Sprenger  E N Baker 《Biochemistry》2001,40(46):13857-13867
The NADP(H)-dependent enzyme glucose-fructose oxidoreductase (GFOR) is a classic example of a redox protein that is translocated across a membrane in fully folded form. GFOR is synthesized in the cytoplasm with a 52-residue signal peptide, giving a precursor form, preGFOR, that is fully active and has its cofactor tightly bound. A twin-arginine motif in the signal peptide directs it to a Sec-independent pathway by which it is translocated, in fully folded form, into the periplasm where it functions to produce sorbitol for osmoprotection. We have determined the crystal structures of four different forms of preGFOR, (i) oxidized preGFOR, with succinate bound in the active site, (ii) oxidized preGFOR with glycerol bound, (iii) reduced preGFOR in 0.3 M glucose, and (iv) reduced preGFOR in 1.5 M sorbitol, at resolutions of 2.2, 2.05, 2.5, and 2.6 A, respectively. In all four crystal structures, the signal peptide is disordered, implying a flexibility that may be important for its interaction with the translocation apparatus; a factor contributing to this disorder may be the high positive charge of the protein surface in the region where the signal peptide emerges. This may disfavor a stable association between the signal peptide and the rest of the protein. The crystal structures show that the mature enzyme portion of preGFOR is identical to native GFOR, in structure and cofactor binding, explaining the enzymatic activity of the precursor form. In the glycerol complex, preGFOR(gll), a bound glycerol molecule models the binding of the glucose substrate, with its O1 atom hydrogen bonded to the essential acid/base catalyst, Tyr269, and C1 only 3 A from C4 of the nicotinamide. In the glucose-soaked structure, preGFOR(glu), we identify a conformational change of the nearby Lys181 that probably results from the oxidation of glucose to gluconolactone, and functions to prevent rebinding of glucose prior to the binding of fructose. In this conformational change, the Lys181 side chain moves closer to the nicotinamide ring, stabilized by its increased negative charge.  相似文献   

12.
Azospirillum brasilense glutamate synthase is a complex iron-sulfur flavoprotein that catalyses the NADPH-dependent reductive transfer of glutamine amide group to the C(2) carbon of 2-oxoglutarate to yield L-glutamate. Its catalytically active alphabeta protomer is composed of two dissimilar subunits (alpha subunit, 164.2 kDa; beta subunit, 52.3 kDa) and contains one FAD (at Site 1, the pyridine nucleotide site within the beta subunit), one FMN (at Site 2, the 2-oxoglutarate/L-glutamate site in the alpha subunit) and three different iron-sulfur clusters (one 3Fe-4S center on the alpha subunit and two 4Fe-4S clusters of unknown location). A plasmid harboring the gltD and gltB genes, the genes encoding the glutamate synthase beta and alpha subunits, respectively, each one under the control of the T7/lac promoter of pET11a was found to be suitable for the overproduction of glutamate synthase holoenzyme in Escherichia coli BL21(DE3) cells. Recombinant A. brasilense glutamate synthase could be purified to homogeneity from overproducing E. coli cells by ion exchange chromatography, gel filtration and affinity chromatography on a 2',5' ADP-Sepharose 4B column. The purified enzyme was indistinguishable from that prepared from Azospirillum cells with respect to cofactor content, N-terminal sequence of the subunits, aggregation state, kinetic and spectroscopic properties. The study of the recombinant holoenzyme allowed us to establish that the tendency of glutamate synthase to form a stable (alphabeta)4 tetramer at high protein concentrations is a property unique to the holoenzyme, as the isolated beta subunit does not oligomerize, while the isolated glutamate synthase alpha subunit only forms dimers at high protein concentrations. Furthermore, the steady-state kinetic analysis of the glutamate synthase reaction was extended to the study of the effect of adenosine-containing nucleotides. Compounds such as cAMP, AMP, ADP and ATP have no effect on the enzyme activity, while the 2'-phosphorylated analogs of AMP and NADP(H) analogs act as inhibitors of the reaction, competitive with NADPH. Thus, it can be ruled out that glutamate synthase reaction is subjected to allosteric modulation by adenosine containing (di)nucleotides, which may bind to the putative ADP-binding site at the C-terminus of the alpha subunit. At the same time, the strict requirement of a 2'-phosphate group in the pyridine nucleotide for binding to glutamate synthase (GltS) was established. Finally, by comparing the inhibition constants exhibited by a series of NADP+ analogs, the contribution to the binding energy of the various parts of the pyridine nucleotide has been determined along with the effect of substituents on the 3 position of the pyridine ring. With the exception of thio-NADP+, which binds the tightest to GltS, it appears that the size of the substituent is the factor that affects the most the interaction between the NADP(H) analog and the enzyme.  相似文献   

13.
A gene encoding glutamate dehydrogenase (GDH) was found in the genome sequence of a commensal thermophile, Symbiobacterium toebii. The amino acid sequence deduced from the gdh I of S. toebii was well conserved with other thermostable GDHs. The gdh I which encodes GDH consisting of 409 amino acids was cloned and expressed in E. coli DH5 under the control of a highly constitutive expression (HCE) promoter in a pHCE system. The recombinant GDH was expressed without addition of any inducers in a soluble form. The molecular mass of the GDH was estimated to be 263 kDa by Superose 6 HR gel filtration chromatography and 44 kDa by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) indicating that the GDH was composed of hexameric form. The optimal temperature and pH of the purified enzyme were 60 °C and 9.0, respectively, and the purified GDH retained more than 75% of its original activity after an incubation at 70 °C for 30 min. Although NADP(H) was the preferred cofactor, S. toebii GDH was able to utilize either NADP(H) or NAD(H) as coenzyme.  相似文献   

14.
Ribitol dehydrogenase from Zymomonas mobilis (ZmRDH) catalyzes the conversion of ribitol to d-ribulose and concomitantly reduces NAD(P)(+) to NAD(P)H. A systematic approach involving an initial sequence alignment-based residue screening, followed by a homology model-based screening and site-directed mutagenesis of the screened residues, was used to study the molecular determinants of the cofactor specificity of ZmRDH. A homologous conserved amino acid, Ser156, in the substrate-binding pocket of the wild-type ZmRDH was identified as an important residue affecting the cofactor specificity of ZmRDH. Further insights into the function of the Ser156 residue were obtained by substituting it with other hydrophobic nonpolar or polar amino acids. Substituting Ser156 with the negatively charged amino acids (Asp and Glu) altered the cofactor specificity of ZmRDH toward NAD(+) (S156D, [k(cat)/K(m)(,NAD)]/[k(cat)/K(m)(,NADP)] = 10.9, where K(m)(,NAD) is the K(m) for NAD(+) and K(m)(,NADP) is the K(m) for NADP(+)). In contrast, the mutants containing positively charged amino acids (His, Lys, or Arg) at position 156 showed a higher efficiency with NADP(+) as the cofactor (S156H, [k(cat)/K(m)(,NAD)]/[k(cat)/K(m)(,NADP)] = 0.11). These data, in addition to those of molecular dynamics and isothermal titration calorimetry studies, suggest that the cofactor specificity of ZmRDH can be modulated by manipulating the amino acid residue at position 156.  相似文献   

15.
An easy procedure for cell free biotransformation of pineapple juice sugars into dietetic derivatives was accomplished using a commercial invertase and an oxidoreductase from Zymomonas mobilis. First, pineapple juice sucrose was quantitatively converted into glucose and fructose by invertase, thus increasing the concentration of each monosaccharide in the original juice to almost twice. In a second step, glucose-fructose oxidoreductase (GFOR) transformed glucose into gluconolactone, and fructose into the low calorie sweetener sorbitol. The advantage of using GFOR is simultaneous reduction of fructose and oxidation of glucose, allowing the continuous regeneration of the essential coenzyme NADP(H), that is tightly bound to the enzyme. The yield of GFOR catalyzed sugar conversion depends on initial pH and control of pH during the reaction. At optimal conditions (pH control at 6.2) a maximum of 80% (w/v) sugar conversion was obtained. Without pH control, GFOR is inactivated rapidly due to gluconic acid formation. Therefore, conversion yields are relatively low at the natural pH of pineapple juice. The application of this process might be more advantageous on juices of other tropical fruits (papaya, jackfruit, mango) due to their naturally given higher pH.  相似文献   

16.
The MJ0490 gene, one of the only two genes of Methanococcus jannaschii showing sequence similarity to the lactate/malate family of dehydrogenases, was classified initially as coding for a putative l-lactate dehydrogenase (LDH). It has been re-classified as a malate dehydrogenase (MDH) gene, because it shows significant sequence similarity to MT0188, MDH II from Methanobacterium thermoautotrophicum strain DeltaH. The three-dimensional structure of its gene product has been determined in two crystal forms: a "dimeric" structure in the orthorhombic crystal at 1.9 A resolution and a "tetrameric" structure in the tetragonal crystal at 2.8 A. These structures share a similar subunit fold with other LDHs and MDHs. The tetrameric structure resembles typical tetrameric LDHs. The dimeric structure is equivalent to the P-dimer of tetrameric LDHs, unlike dimeric MDHs, which correspond to the Q-dimer. The structure reveals that the cofactor NADP(H) is bound at the active site, despite the fact that it was not intentionally added during protein purification and crystallization. The preference of NADP(H) over NAD(H) has been supported by activity assays. The cofactor preference is explained by the presence of a glycine residue in the cofactor binding pocket (Gly33), which replaces a conserved aspartate (or glutamate) residue in other NAD-dependent LDHs or MDHs. Preference for NADP(H) is contributed by hydrogen bonds between the oxygen atoms of the monophosphate group and the ribose sugar of adenosine in NADP(H) and the side-chains of Ser9, Arg34, His36, and Ser37. The MDH activity of MJ0490 is made possible by Arg86, which is conserved in MDHs but not in LDHs. The enzymatic assay showed that the MJ0490 protein possesses the fructose-1,6-bisphosphate-activated LDH activity (reduction). Thus the MJ0490 gene product appears to be a novel member of the lactate/malate dehydrogenase family, displaying an LDH scaffold and exhibiting a relaxed substrate and cofactor specificities in NADP(H) and NAD(H)-dependent malate and lactate dehydrogenase reactions.  相似文献   

17.
The enzymes responsible for sorbitol formation in Zymomonas mobilis were investigated. A previously undescribed enzyme catalyzes the intermolecular oxidation-reduction of glucose and fructose to form gluconolactone and sorbitol. This enzyme has been purified; it had a subunit size of 40,000 daltons and is probably tetrameric at low pH. It contained tightly bound NADP as the hydrogen carrier and did not require any added cofactor for activity. In addition, a gluconolactonase has been isolated, although not completely purified. Together these two enzymes were capable of completely converting a 54% (wt/vol) equimolar mixture of glucose and fructose to sorbitol and sodium gluconate at the optimum pH of close to 6.2. The oxidoreductase had low affinities for its substrates, but natural environmental conditions would expose it to high concentrations of sugars. The amount of the enzyme in Z. mobilis cells was sufficient to account for the rate of sorbitol formation in vivo. However, the enzyme was present in the highest amounts when the cells were grown on glucose alone, and it was repressed by the presence of fructose; this was not the case with the gluconolactonase.  相似文献   

18.
The bacterial twin arginine translocation (Tat) pathway translocates across the cytoplasmic membrane folded proteins which, in most cases, contain a tightly bound cofactor. Specific amino-terminal signal peptides that exhibit a conserved amino acid consensus motif, S/T-R-R-X-F-L-K, direct these proteins to the Tat translocon. The glucose-fructose oxidoreductase (GFOR) of Zymomonas mobilis is a periplasmic enzyme with tightly bound NADP as a cofactor. It is synthesized as a cytoplasmic precursor with an amino-terminal signal peptide that shows all of the characteristics of a typical twin arginine signal peptide. However, GFOR is not exported to the periplasm when expressed in the heterologous host Escherichia coli, and enzymatically active pre-GFOR is found in the cytoplasm. A precise replacement of the pre-GFOR signal peptide by an authentic E. coli Tat signal peptide, which is derived from pre-trimethylamine N-oxide (TMAO) reductase (TorA), allowed export of GFOR, together with its bound cofactor, to the E. coli periplasm. This export was inhibited by carbonyl cyanide m-chlorophenylhydrazone, but not by sodium azide, and was blocked in E. coli tatC and tatAE mutant strains, showing that membrane translocation of the TorA-GFOR fusion protein occurred via the Tat pathway and not via the Sec pathway. Furthermore, tight cofactor binding (and therefore correct folding) was found to be a prerequisite for proper translocation of the fusion protein. These results strongly suggest that Tat signal peptides are not universally recognized by different Tat translocases, implying that the signal peptides of Tat-dependent precursor proteins are optimally adapted only to their cognate export apparatus. Such a situation is in marked contrast to the situation that is known to exist for Sec-dependent protein translocation.  相似文献   

19.
The action of six different enzymes on the function and structure of Factor H was investigated by use of sodium dodecyl sulphate/polyacrylamide-gel electrophoresis, haemagglutination, two enzyme-linked immunosorbent assay systems and an assay for Factor I cofactor activity. Six monoclonal antibodies directed against the 38 kDa tryptic fragment of Factor H [which contains the binding site for C3b (a 180 kDa fragment of the third component of complement) and the cofactor activity] were also used to detect cleavage products derived from the same fragment. Elastase, chymotrypsin A4 or trypsin first cleaved Factor H to 36-38 kDa fragments carrying all six monoclonal anti-(Factor H)-binding sites. In parallel, the interaction of Factor H with surface-bound C3b was lost, whereas the cofactor function was preserved. Further cleavage of the 36-38 kDa fragments into two 13-19 kDa fragments (one carrying the MAH4 and MRC OX 24 epitopes, the other the MAH1, MAH2, MAH3 and MRC OX 23 epitopes) destroyed cofactor activity. Pepsin, bromelain or papain rapidly split off a 13-15 kDa fragment of Factor H carrying the MAH1, MAH2, MAH3 and MRC OX 23 epitopes and destroyed all tested functions of Factor H. Ficin cleaved Factor H into disulphide-linked fragments smaller than 25 kDa, but did not affect the functions of the Factor H molecule. The 38 kDa tryptic fragment of Factor H is the N-terminal end of the Factor H molecule, as determined by N-terminal sequence analysis. A model is presented of the substructure of Factor H.  相似文献   

20.
When grown on xanthan as a carbon source, the bacterium Bacillus sp. strain GL1 produces extracellular xanthan lyase (75 kDa), catalyzing the first step of xanthan depolymerization (H. Nankai, W. Hashimoto, H. Miki, S. Kawai, and K. Murata, Appl. Environ. Microbiol. 65:2520-2526, 1999). A gene for the lyase was cloned, and its nucleotide sequence was determined. The gene contained an open reading frame consisting of 2,793 bp coding for a polypeptide with a molecular weight of 99,308. The polypeptide had a signal peptide (2 kDa) consisting of 25 amino acid residues preceding the N-terminal amino acid sequence of the enzyme and exhibited significant homology with hyaluronidase of Streptomyces griseus (identity score, 37.7%). Escherichia coli transformed with the gene without the signal peptide sequence showed a xanthan lyase activity and produced intracellularly a large amount of the enzyme (400 mg/liter of culture) with a molecular mass of 97 kDa. During storage at 4 degrees C, the purified enzyme (97 kDa) from E. coli was converted to a low-molecular-mass (75-kDa) enzyme with properties closely similar to those of the enzyme (75 kDa) from Bacillus sp. strain GL1, specifically in optimum pH and temperature for activity, substrate specificity, and mode of action. Logarithmically growing cells of Bacillus sp. strain GL1 on the medium with xanthan were also found to secrete not only xanthan lyase (75 kDa) but also a 97-kDa protein with the same N-terminal amino acid sequence as that of xanthan lyase (75 kDa). These results suggest that, in Bacillus sp. strain GL1, xanthan lyase is first synthesized as a preproform (99 kDa), secreted as a precursor (97 kDa) by a signal peptide-dependent mechanism, and then processed into a mature form (75 kDa) through excision of a C-terminal protein fragment with a molecular mass of 22 kDa.  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