共查询到18条相似文献,搜索用时 15 毫秒
1.
Isabella M. Acquistapace Monika A. Zietek Arthur W. H. Li Melissa Salmon Imke Kühn Mike R. Bedford Charles A. Brearley Andrew M. Hemmings 《The Journal of biological chemistry》2020,295(51):17724
Highly engineered phytases, which sequentially hydrolyze the hexakisphosphate ester of inositol known as phytic acid, are routinely added to the feeds of monogastric animals to improve phosphate bioavailability. New phytases are sought as starting points to further optimize the rate and extent of dephosphorylation of phytate in the animal digestive tract. Multiple inositol polyphosphate phosphatases (MINPPs) are clade 2 histidine phosphatases (HP2P) able to carry out the stepwise hydrolysis of phytate. MINPPs are not restricted by a strong positional specificity making them attractive targets for development as feed enzymes. Here, we describe the characterization of a MINPP from the Gram-positive bacterium Bifidobacterium longum (BlMINPP). BlMINPP has a typical HP2P-fold but, unusually, possesses a large α-domain polypeptide insertion relative to other MINPPs. This insertion, termed the U-loop, spans the active site and contributes to substrate specificity pockets underpopulated in other HP2Ps. Mutagenesis of U-loop residues reveals its contribution to enzyme kinetics and thermostability. Moreover, four crystal structures of the protein along the catalytic cycle capture, for the first time in an HP2P, a large ligand-driven α-domain motion essential to allow substrate access to the active site. This motion recruits residues both downstream of a molecular hinge and on the U-loop to participate in specificity subsites, and mutagenesis identified a mobile lysine residue as a key determinant of positional specificity of the enzyme. Taken together, these data provide important new insights to the factors determining stability, substrate recognition, and the structural mechanism of hydrolysis in this industrially important group of enzymes. 相似文献
2.
The complex N-glycan structures on glycoproteins play important roles in cell adhesion and recognition events in metazoan organisms. A critical step in the biosynthetic pathway leading from high mannose to these complex structures includes the transfer of N-acetylglucosamine (GlcNAc) to a mannose residue by the inverting N-acetylglucosaminyltransferase I (GnT-I). The catalytic mechanism of this enzymatic reaction is explored herein using DFT quantum chemical methods. The computational model used to follow the reaction is based on the X-ray crystallographic structure of GnT-I and contains 127 atoms that represent fragments of residues critical for the substrate binding and catalysis. The mechanism of the catalytic reaction was monitored by means of a 2D potential energy map calculated as a function of predefined reaction coordinates at the B3LYP/6-31G** level. This potential energy surface revealed one transition state associated with a reaction pathway following a concerted mechanism. The reaction barrier was estimated, and the structure of the transition state was characterized at the B3LYP/6-311++G**// B3LYP/6-31G** level. 相似文献
3.
The structure of trypanosomal triosephosphate isomerase (TIM)has been solved at a resolution of 2.1Å in a new crystal form grown at pH 8.8 from PEG6000. In this new crystal form (space group C2, cell dimensions 94.8 Å, 48.3 Å, 131.0 Å, 90.0°, 100.3°, 90.0°), TIM is present in a ligand-free state. The asymmetric unit consists of two TIM subunits. Each of these subunits is part of a dimer which is sitting on a crystallographic twofold axis, such that the crystal packing is formed from two TIM dimers in two distinct environments. The two constituent monomers of a given dimer are, therefore, crystallographically equivalent. In the ligand-free state of TIM in this crystal form, the two types of dimer are very similar in structure, with the flexible loops in the “Open” conformation. For one dimer (termed molecule-1), the flexible loop (loop-6) is involved in crystal contacts. Crystals of this type have been used in soaking experiments with 0.4 M ammonium sulphate (studied at 2.4 Å resolution), and with 40 μM phosphoglycolohydroxamate (studied at 2.5 Å resolution). It is found that transfer to 0.4 M ammonuum sulphate (equal to 80 times the Ki of sulphate for TIM), gives rise to significant sulphate binding at the active site of one dimer (termed molecule-2), and less significant binding at the active site of the other. In neither dimer does sulphate induce a “closed” conformation. In a mother liquor containing 40 μM phosphoglycolohydroxamate (equal to 10 times the Ki of phosphoglycolohydroxamate for TIM), an inhibitor molecule binds at the active site of only that dimer of which the flexible loop is free from crystal contacts (molecule-2). In this dimer, it induces a closed conformation. These three structures are compared and discussed with respect to the mode of binding of ligand in the active site as well as with respect to the conformational changes resulting from ligand binding. © 1993 Wiley-Liss, Inc. 相似文献
4.
Assimilatory nitrite reductase (aNiR) reduces nitrite to an ammonium ion and has siroheme and a [Fe(4)S(4)] cluster as prosthetic groups. A reaction mechanism for Nii3, an aNiR from tobacco, is proposed based on high resolution X-ray structures and UV-Vis (ultraviolet-visible) microspectroscopy of Nii3-ligand complexes. Analysis of UV-Vis spectral changes in Nii3 crystals with increasing X-ray exposure showed prosthetic group reductions. In Nii3-NO2(-) structures, X-ray irradiation enhanced the progress of the reduction reaction, and cleavage of the N-O bond was observed when X-ray doses were increased. Crystal structures of Nii3 with other bound ligands, such as Nii3-NO and Nii3-NH(2)OH, were also determined. Further, by combining information from these Nii3 ligand-bound structures, including that of Nii3-NO2(-), with UV-Vis microspectral data obtained using different X-ray doses, a reaction mechanism for aNiR was suggested. Cleavage of the two N-O bonds of nitrite was envisaged as a two-step process: first, the N-O bond close to Lys224 was cleaved, followed by cleavage of the N-O bond close to Arg109. X-ray structures also indicated that aNiR-catalyzed nitrite reduction proceeded without the need for conformation changes in active site residues. Geometrical changes in the ligand molecules and the placement of neighboring water molecules appeared to be important to the stability of the active site residue interactions (Arg109, Arg179, and Lys224) and the ligand molecule. These interactions may contribute to the efficiency of aNiR reduction reactions. 相似文献
5.
The NADP-dependent glycerol dehydrogenase (EC 1.1.1.72) from Gluconobacter oxydans is a member of family 11 of the aldo-keto reductase (AKR) enzyme superfamily; according to the systematic nomenclature within the AKR superfamily, the term AKR11B4 has been assigned to the enzyme. AKR11B4 is a biotechnologically attractive enzyme because of its broad substrate spectrum, combined with its distinctive regioselectivity and stereoselectivity. These features can be partially rationalized based on a 2-Å crystal structure of apo-AKR11B4, which we describe and interpret here against the functional complex structures of other members of family 11 of the AKR superfamily. The structure of AKR11B4 shows the AKR-typical (β/α)8 TIM-barrel fold, with three loops and the C-terminal tail determining the particular enzymatic properties. In comparison to AKR11B1 (its closest AKR relative), AKR11B4 has a relatively broad binding cleft for the cosubstrate NADP/NADPH. In the crystalline environment, it is completely blocked by the C-terminal segment of a neighboring protomer. The structure reveals a conspicuous tryptophan residue (Trp23) that has to adopt an unconventional and strained side-chain conformation to permit cosubstrate binding. We predict and confirm by site-directed mutagenesis that Trp23 is an accelerator of (co)substrate turnover. Furthermore, we show that, simultaneously, this tryptophan residue is a critical determinant for substrate binding by the enzyme, while enantioselectivity is probably governed by a methionine residue within the C-terminal tail. We present structural reasons for these notions based on ternary complex models of AKR11B4, NADP, and either octanal, d-glyceraldehyde, or l-glyceraldehyde. 相似文献
6.
7.
A Synechococcus sp. strain PCC 7002 psaAB::cat mutant has been constructed by deletional interposon mutagenesis of the psaA and psaB genes through selection and segregation under low-light conditions. This strain can grow photoheterotrophically with glycerol as carbon source with a doubling time of 25 h at low light intensity (10 E m–2 s–1). No Photosystem I (PS I)-associated chlorophyll fluorescence emission peak was detected in the psaAB::cat mutant. The chlorophyll content of the psaAB::cat mutant was approximately 20% that of the wild-type strain on a per cell basis. In the absence of the PsaA and PsaB proteins, several other PS I proteins do not accumulate to normal levels. Assembly of the peripheral PS I proteins PsaC,PsaD, PsaE, and PsaL is dependent on the presence of the PsaA and PsaB heterodimer core. The precursor form of PsaF may be inserted into the thylakoid membrane but is not processed to its mature form in the absence of PsaA and PsaB. The absence of PS I reaction centers has no apparent effect on Photosystem II (PS II) assembly and activity. Although the mutant exhibited somewhat greater fluorescence emission from phycocyanin, most of the light energy absorbed by phycobilisomes was efficiently transferred to the PS II reaction centers in the absence of the PS I. No light state transition could be detected in the psaAB::cat strain; in the absence of PS I, cells remain in state 1. Development of this relatively light-tolerant strain lacking PS I provides an important new tool for the genetic manipulation of PS I and further demonstrates the utility of Synechococcus sp. PCC 7002 for structural and functional analyses of the PS I reaction center.Abbreviations ATCC
American type culture collection
- Chl
chlorophyll
- DCMU
3-(3,4-dichlorophyl)-1,1-dimethylurea
- DBMIB
2,5-dibromo-3-methyl-6-isopropyl-p-benzoquinone
- HEPES
N-[2-hydroxyethyl]piperazine-N-[2-ethanesulfonic acid]
- PCC
Pasteur culture collection
- PS I
Photosystem I
- PS II
Photosystem II
- SDS
sodium dodecyl sulfate 相似文献
8.
The structure of the recombinant (-) gamma-lactamase from an Aureobacterium species has been solved at 1.73A resolution in the cubic space group F23 with unit cell parameters a=b=c=240.6A. The trimeric enzyme has an alpha/beta hydrolase fold and closely resembles the cofactor free haloperoxidases. The structure has been solved in complex with a covalently bound ligand originating from the host cell and also in the unligated form. The associated density in the former structure has been interpreted as the two-ring ligand (3aR,7aS)-3a,4,7,7a-tetrahydro-benzo [1,3] dioxol-2-one which forms a tetrahedral complex with OG of the catalytic Ser98. Soaks of these crystals with the industrial substrate gamma-lactam or its structural analogue, norcamphor, result in the displacement of the ligand from the enzyme active site, thereby allowing determination of the unligated structure. The presence of the ligand in the active site protects the enzyme from serine hydrolase inhibitors. Cyclic ethylene carbonate, the first ring of the ligand, was shown to be a substrate of the enzyme. 相似文献
9.
The kinetic mechanism and evaluation of several potential inhibitors of purified arginine kinase from the cockroach (Periplanta americana) were investigated. This monomeric phosphagen kinase is important in maintaining ATP levels during the rapid energy demands of muscle required for contraction and motility. Analysis reveals the following dissociation constants (mM) for the binary complex: E.Arg P-->E+Arg P, K=1.0; E.Arg-->E+Arg, K=0.45; E.MgATP-->E+MgATP, K=0.17; E.MgADP-->E+MgADP, K=0.12; and the ternary complex: Arg P.E.MgADP-->E.MgADP+Arg P, K=0.94; Arg.E.MgATP-->E.MgATP+Arg, K=0.49; MgATP.Enz.Arg-->E.Arg+MgATP, K=0.14; MgADP.E.Arg P-->E.Arg P+MgADP, K=0.09. For a particular substrate, the ratio of the dissociation constants for the binary to ternary complex is close to one, indicating little, if any, cooperativity in substrate binding for the rapid equilibrium, random addition mechanism. The time course of the arginine kinase reaction exhibits a pronounced curvature, which, as described for enzyme from other sources, is attributed to formation of an inhibitory catalytic dead-end complex, MgADP.E.Arg. The curvature is accentuated by the addition of monovalent anions, including borate, thiocyanate, and, most notably, nitrite and nitrate. This effect is attributed to stabilization of the dead-end complex through formation of a transition state analog. However, the substantial decrease in initial velocity (92%) caused by nitrate is due to an additional inhibitory effect, further characterized as non-competitive inhibition (Ki=8.0 mM) with the substrate L-arginine. On the other hand, borate inhibition of the initial velocity is only 30% with significant subsequent curvature, suggesting that this anion functions as an inhibitor mainly by formation of a transition state analog. However, some component of the borate inhibition appears to be mediated by an apparent partial competitive inhibition with L-arginine. D-arginine is not a substrate for arginine kinase from the cockroach, but is an effective competitive inhibitor with a Ki=0.31 mM. L-Canavanine is a weak substrate for arginine kinase (Km=6.7 mM) with a Vmax for the pure enzyme that is approximately one-third that of L-arginine. However, initial velocity experiments of substrate mixtures suggest that competition between L-canavanine and L-arginine may not be a simple summation effect and may involve a structural modification. Sensitivity of arginine kinase activity to D-arginine as well as nitrate and borate anions, coupled with the fact that L-arginine is an essential amino acid for the cockroach, suggest that arginine kinase could be a useful chemotherapeutic target for the control of cockroach proliferation. 相似文献
10.
【目的】对我国西藏地区来源的不同酵母菌株进行有机酸发酵性能测试,此外,对具有良好产酸性能的分离自松萝内部的酿酒酵母菌株Saccharomy cescerevisiae 2-2进行耐酸性能分析,并探究其耐酸较强的分子机制。【方法】比较不同糖浓度培养基液体发酵培养过程中pH的变化,并比较低pH胁迫条件下菌株的生长,检测酿酒酵母菌株的产酸潜力和耐酸特性;对菌株2-2和模式酵母菌株S288C进行比较基因组分析,并利用实时荧光定量聚合酶链式反应(real-time fluorescence quantitative polymerase chain reaction,RT-qPCR)分析关键基因的转录,探究菌株2-2耐酸分子机制。【结果】松萝内生酿酒酵母2-2在所有检测的菌株中产酸潜力较大,耐酸性能较好。在菌株2-2中与胁迫耐受性相关的基因PDR15、PDR12和SUR1在低pH胁迫条件下存在显著的上调或下调,但这些基因转录变化趋势与菌株S288C相反。【结论】松萝内生酿酒酵母2-2是一株产酸耐酸性能较好的菌株,对其独特的调节机制进行深入分析,有希望选育性能更好的产酸酵母菌株。 相似文献
11.
The crystal structure of maleylacetate reductase from Rhizobium sp. strain MTP‐10005 provides insights into the reaction mechanism of enzymes in its original family 下载免费PDF全文
Tomomi Fujii Ai Sato Yuko Okamoto Takae Yamauchi Shiro Kato Masahiro Yoshida Tadao Oikawa Yasuo Hata 《Proteins》2016,84(8):1029-1042
Maleylacetate reductase plays a crucial role in catabolism of resorcinol by catalyzing the NAD(P)H‐dependent reduction of maleylacetate, at a carbon–carbon double bond, to 3‐oxoadipate. The crystal structure of maleylacetate reductase from Rhizobium sp. strain MTP‐10005, GraC, has been elucidated by the X‐ray diffraction method at 1.5 Å resolution. GraC is a homodimer, and each subunit consists of two domains: an N‐terminal NADH‐binding domain adopting an α/β structure and a C‐terminal functional domain adopting an α‐helical structure. Such structural features show similarity to those of the two existing families of enzymes in dehydroquinate synthase‐like superfamily. However, GraC is distinct in dimer formation and activity expression mechanism from the families of enzymes. Two subunits in GraC have different structures from each other in the present crystal. One subunit has several ligands mimicking NADH and the substrate in the cleft and adopts a closed domain arrangement. In contrast, the other subunit does not contain any ligand causing structural changes and adopts an open domain arrangement. The structure of GraC reveals those of maleylacetate reductase both in the coenzyme, substrate‐binding state and in the ligand‐free state. The comparison of both subunit structures reveals a conformational change of the Tyr326 loop for interaction with His243 on ligand binding. Structures of related enzymes suggest that His243 is likely a catalytic residue of GraC. Mutational analyses of His243 and Tyr326 support the catalytic roles proposed from structural information. The crystal structure of GraC characterizes the maleylacetate reductase family as a third family in the dehydroquinate synthase‐like superfamily. Proteins 2016; 84:1029–1042. © 2016 Wiley Periodicals, Inc. 相似文献
12.
Previously we reported an antifungal protein specific to Pythium porphyrae, a causative agent of red rot disease afflicting seaweed Porphyra spp. This study was carried out to identify the antifungal mechanism of the antifungal protein to P. porphyrae. When we first examined the effect of an anti- Pythium protein (SAP) on the P. porphyrae cell walls, SAP did not decompose the six structural polysaccharides in Pythium cell walls. However, hyphal growth was significantly inhibited in Pythium cells treated with 50 microg/ml of SAP by MTT assay. Protoplasmic leakage was observed in P. porphyrae hyphae treated with SAP for 1 h, followed by hyphal swelling and disintegration, using SYTOX Green, and SAP permeabilized the membrane of P. porphyrae in a dose-dependent manner. Treating P. porphyrae cells with SAP in the presence of carbonyl cyanide m-chlorophenylhydrazone (CCCP), a membrane-depolarizing agent, significantly reduced the membrane permeability to SYTOX Green. Moreover, a similar effect was observed when the P. porphyrae cells were treated with SAP in the presence of MgCl2. In contrast, identical treatment in the presence of KCl significantly increased the membrane permeability to SYTOX Green. These results suggested that anti- Pythium mechanism of SAP was related to alteration of the membrane permeability in P. porphyrae. 相似文献
13.
Grayna E. Sroga Ulf Landegren Birgitta Bergman Maria Lagerström-Fermér 《FEMS microbiology letters》1996,136(2):137-145
Abstract A modified capture polymerase chain reaction (CPCR) technique was used to isolate the entire sequence of the nifH gene and its flanking regions from a natural population of Trichodesmium sp. A set of specific CPCR primers derived from a known 72-bp DNA segment of the nifH sequence permitted isolation of both the upstream and the downstream region of Trichodesmium sp. nifH . The 882-bp nifH gene presented here is the first full-length gene isolated from Trichodesmium sp. A sequence similar to a nif -like promoter was found in front of nifH . The nifH open reading frame of Trichodesmium sp. encoded 294 amino acids. Comparative analysis of the Trichodesmium sp. NifH sequence revealed strong similarity with 23 known NifH proteins. Amino acids postulated to be involved in binding of the 4Fe:4S cluster and those subjected to ADP-ribosylation were present. An open reading frame for the nifD gene was identified 189 bp downstream of nifH . A sequence similar to the consensus of the nif -like promoter was also found in front of nifD . 相似文献
14.
Srinivasan Rengachari Gustavo A. Bezerra Lina Riegler-Berket Christian C. Gruber Christian Sturm Ulrike Taschler Andras Boeszoermenyi Ingrid Dreveny Robert Zimmermann Karl Gruber Monika Oberer 《Biochimica et Biophysica Acta (BBA)/Molecular and Cell Biology of Lipids》2012,1821(7):1012-1021
Monoacylglycerol lipases (MGLs) catalyse the hydrolysis of monoacylglycerol into free fatty acid and glycerol. MGLs have been identified throughout all genera of life and have adopted different substrate specificities depending on their physiological role. In humans, MGL plays an integral part in lipid metabolism affecting energy homeostasis, signalling processes and cancer cell progression. In bacteria, MGLs degrade short-chain monoacylglycerols which are otherwise toxic to the organism. We report the crystal structures of MGL from the bacterium Bacillus sp. H257 (bMGL) in its free form at 1.2 Å and in complex with phenylmethylsulfonyl fluoride at 1.8 Å resolution. In both structures, bMGL adopts an α/β hydrolase fold with a cap in an open conformation. Access to the active site residues, which were unambiguously identified from the protein structure, is facilitated by two different channels. The larger channel constitutes the highly hydrophobic substrate binding pocket with enough room to accommodate monoacylglycerol. The other channel is rather small and resembles the proposed glycerol exit hole in human MGL. Molecular dynamics simulation of bMGL yielded open and closed states of the entrance channel and the glycerol exit hole. Despite differences in the number of residues, secondary structure elements, and low sequence identity in the cap region, this first structure of a bacterial MGL reveals striking structural conservation of the overall cap architecture in comparison with human MGL. Thus it provides insight into the structural conservation of the cap amongst MGLs throughout evolution and provides a framework for rationalising substrate specificities in each organism. 相似文献
15.
Indoleglycerol phosphate synthase catalyzes the ring closure of an N-alkylated anthranilate to a 3-alkyl indole derivative, a reaction requiring Lewis acid catalysis in vitro. Here, we investigated the enzymatic reaction mechanism through X-ray crystallography of complexes of the hyperthermostable enzyme from Sulfolobus solfataricus with the substrate 1-(o-carboxyphenylamino) 1-deoxyribulose 5-phosphate, a substrate analogue and the product indole-3-glycerol phosphate. The substrate and the substrate analogue are bound to the active site in a similar, extended conformation between the previously identified phosphate binding site and a hydrophobic pocket for the anthranilate moiety. This binding mode is unproductive, because the carbon atoms that are to be joined are too far apart. The indole ring of the bound product resides in a second hydrophobic pocket adjacent to that of the anthranilate moiety of the substrate. Although the hydrophobic moiety of the substrate moves during catalysis from one hydrophobic pocket to the other, the triosephosphate moiety remains rigidly bound to the same set of hydrogen-bonding residues. Simultaneously, the catalytically important residues Lys53, Lys110 and Glu159 maintain favourable distances to the atoms of the ligand undergoing covalent changes. On the basis of these data, the structures of two putative catalytic intermediates were modelled into the active site. This new structural information and the modelling studies provide further insight into the mechanism of enzyme-catalyzed indole synthesis. The charged epsilon-amino group of Lys110 is the general acid, and the carboxylate group of Glu159 is the general base. Lys53 guides the substrate undergoing conformational transitions during catalysis, by forming a salt-bridge to the carboxylate group of its anthranilate moiety. 相似文献
16.
A carbamoylase enzyme was purified from a cell-free extract of Agrobacterium sp. with an overall yield of 81%. It was judged to be homogenous on sodium dodecyl sulfate-polyacrylamide gel electrophoresis, with a subunit molecular weight of 38,000 daltons. Further studies on the native enzyme suggested that the active enzyme was present as a dimer, with a pI of 5.5. It was able to cleave a variety of N-carbamoyl substrates, but was strictly D(−) specific. It was found to have a Km of 0.82 m
and a Vmax of 31 U mg−1 for D(−) N-carbamoyl hydroxyphenylglycine in the presence of 10 m
dithiothreitol. It showed no metal ion requirements but was inhibited by iodoacetic acid and iodoacetamide, both thiol reagents. The N-terminal amino acid sequence of the enzyme was elucidated. 相似文献
17.
The first naturally occurring split intein was found in the dnaE gene of Synechocystis sp. PCC6803 and belongs to a subclass of inteins without a penultimate histidine residue. We describe two high-resolution crystal structures, one derived from an excised Ssp DnaE intein and the second from a splicing-deficient precursor protein. The X-ray structures indicate that His147 in the conserved block F activates the side-chain N(delta) atom of the intein C-terminal Asn159, leading to a nucleophilic attack on the peptide bond carbonyl carbon atom at the C-terminal splice site. In this process, Arg73 appears to stabilize the transition state by interacting with the carbonyl oxygen atom of the scissile bond. Arg73 also seems to substitute for the conserved penultimate histidine residue in the formation of an oxyanion hole, as previously identified in other inteins. The finding that the precursor structure contains a zinc ion chelating the highly conserved Cys160 and Asp140 reveals the structural basis of Zn2+-mediated inhibition of protein splicing. Furthermore, it is of interest to observe that the carbonyl carbon atom of Asn159 and N(eta) of Arg73 are 2.6 angstroms apart in the free intein structure and 10.6 angstroms apart in the precursor structure. The orientation change of the aromatic ring of Tyr-1 following the initial acyl shift may be a key switching event contributing to the alignment of Arg73 and the C-terminal scissile bond, and may explain the sequential reaction property of the Ssp DnaE intein. 相似文献
18.
Hatada Y Takeda N Hirasawa K Ohta Y Usami R Yoshida Y Grant WD Ito S Horikoshi K 《Extremophiles : life under extreme conditions》2005,9(6):497-500
A novel alkaline mannanase Man26A has been found in the culture of an alkaliphilic Bacillus sp. strain JAMB-750 and the optimal pH for the mannanase activity of the enzyme was around pH 10 (J Biol Macromol 4: 67–74, 2004). This optimal pH is the highest among those of the mannanases reported to date. The gene man26A coding the enzyme was cloned from the genomic DNA of strain JAMB-750 and sequenced. It encodes a protein of 997 amino acids including a signal peptide. The N-terminal half (Glu27–Val486) of the enzyme exhibited moderate similarities to other mannanases belonging to glycoside hydrolase family 26, such as the enzymes from Cellvibrio japonicus (37% identity), Cellulomonas fimi (33% identity), and Bacillus sp. strain AM-001 (28% identity). The C-terminal half was found to contain four domains. The first, second, third, and fourth domains exhibited similarities to the carbohydrate-binding module, the mannan-binding module, the Homo sapiens collagen type IX alpha I chain, and the membrane anchor region of Gram-positive surface proteins, respectively. Its recombinant mannanase was produced extracellularly using Bacillus subtilis as the host. The optimal pH for the mannanase activity of the recombinant enzyme was around pH 10. The enzyme was very resistant to surfactants, for example, SDS up to 2.0% (w/v). 相似文献