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 15 毫秒
1.
AmyL, an extracellular α-amylase from Bacillus licheniformis, is resistant to extracellular proteases secreted by Bacillus subtilis during growth. Nevertheless, when AmyL is produced and secreted by B. subtilis, it is subject to considerable cell-associated proteolysis. Cell-wall-bound proteins CWBP52 and CWBP23 are the processed products of the B. subtilis wprA gene. Although no activity has been ascribed to CWBP23, CWBP52 exhibits serine protease activity. Using a strain encoding an inducible wprA gene, we show that a product of wprA, most likely CWBP52, is involved in the posttranslocational stability of AmyL. A construct in which wprA is not expressed exhibits an increased yield of α-amylase. The potential role of wprA in protein secretion is discussed, together with implications for the use of B. subtilis and related bacteria as hosts for the secretion of heterologous proteins.The cell envelope of the gram-positive bacterium Bacillus subtilis consists of a single (cytoplasmic) membrane surrounded by a relatively thick cell wall consisting of similar proportions of peptidoglycan and covalently attached anionic polymers. The absence of an outer membrane means that there is no equivalent of the membrane-enclosed periplasm found in gram-negative bacteria. However, by virtue of its thickness and high density of negative charge, the cell wall may perform some of the roles of the periplasm in gram-positive bacteria.The absence of an outer membrane in gram-positive bacteria also simplifies the secretion pathway, and, consequently, B. subtilis and its close relatives have the potential to secrete proteins directly into the growth medium, at concentrations in excess of 5 grams per liter (4). Despite its extensive use in the production of commercially important Bacillus enzymes (e.g., α-amylases and alkaline proteases), attempts to exploit B. subtilis for the production of heterologous proteins at high concentrations have proved disappointing (8). One reason for this failure is the production and release into the culture medium of several extracellular proteases (24, 28, 37). Although native Bacillus proteins are generally resistant to these proteases, heterologous proteins are often rapidly degraded in their presence. As a result, strains of B. subtilis that are multiply deficient in extracellular proteases have been developed (11, 37). The more developed of these strains have less than 1% of the proteolytic activity of the wild type (37). To date, efforts have concentrated mainly on the proteases which reside in a truly extracellular location, while those which remain cell associated have been largely overlooked.Although strains deficient in extracellular proteases have improved the productivity of B. subtilis for the production of heterologous proteins, they have only partially overcome problems of unexpectedly low yields. We and others have recently shown (22, 31) that significant amounts of secretory protein are degraded within minutes of being synthesized. This degradation is observed even for Bacillus proteins that are highly resistant to proteases released into the culture medium, suggesting that a component of this degradation is cell associated.Margot and Karamata recently reported the identification of a cell-wall-associated protease encoded by the wprA gene (21). The primary product of this gene is a 96-kDa polypeptide that is processed into two previously identified cell wall proteins, namely, CWBP52 and CWBP23. The processing of the WprA precursor during secretion accompanies the targeting of CWBP52 and CWBP23 to the cell wall and is analagous to the processing of another B. subtilis cell-wall-bound protein, namely, WapA (5). The amino acid sequence of CWBP52 shows a high degree of similarity with serine proteases of the subtilisin family, and phenylmethylsulfonile fluoride (PMSF)-sensitive protease activity was detected in proteins extracted from the cell wall of a wprA+ strain, but not one in which this gene had been insertionally inactivated (21). In the absence of homology to proteins in the databases, the N-terminal CWBP23 moiety was presumed to function as a chaperone-like propeptide that is proteolytically processed on the trans side of the membrane. In this paper, we report on a potential role of products of wprA in the integrity of secretory proteins during late stages in the secretion pathway. We also discuss the potential of wprA mutants to increase the productivity of B. subtilis for secretory proteins.  相似文献   

2.
The inactivation of Bacillus subtilisα-amylase by acid was shown to be reversible. In the experiment, two different Bac. subtilisα-amylases, saccharifying and liquefying types, were used and the reversibility was investigated deviding into two processes of inactivation and reactivation. Both amylases showed the reversibility in a similar degree and in general the inactivated enzymes by acid were reactivated only by adjusting the pH to slightly alkaline values followed by incubation under certain conditions. However, the reversibility, especially, the reactivation was greatly influenced by several chemicals, the effect of certain chemicals being different according to the type of the bacterial amylase. Contrary to liquefying amylase, saccharifying amylase was insensitive to metal chelators but, nevertheless, the reactivation of the amylase was prevented by metal chelators. Also the reactivation of saccharifying amylase was inhibited by sulfhydryl reagents, although the native enzyme was quite insensitive to the chemicals. In the acid-inactivation and reactivation process, a reversible change in the ultraviolet absorption spectra of the enzymes was observed, and some discussion of the implication was presented.  相似文献   

3.
Inactivation of Bacillus subtilisα-amylase by heat was found to be reversible under a certain condition, and the factors affecting there were investigated, distinguishing into two groups: those influencing on the inactivation process by heat and those on the reactivation at the subsequent incubation after heating. Generally, the amylase heated in borate buffer solution was best in the reactivation degree. For reactivation of the heat-inactivated enzyme there was found an optimum in temperature, pH and concentration of enzyme, respectively. The reactivation was temporarily prevented by urea, but irreversibly inhibited by either calcium salts or calcium binding agents. In the reversible heat-inactivation of the enzyme was also found a reversible change in the absorption spectra as well as in the behavior of the enzyme toward proteinase.  相似文献   

4.
The combined effect of macronutrients in the extraction medium on α-amylase produced by Bacillus subtilis were studied by using response surface methodology in shaken flask cultures. The production of amylase was significantly affected by the interaction between wheat bran and the cotton seed extract in the extraction medium and by the interaction between the cotton seed extract and starch. The optimal combination in the extraction medium for maximum α-amylase production was determined as 10.80 g·L?1 of wheat bran, 9.90 g·L?1 of cotton seed extract, 0.5 g·L?1 of starch, 2.0 g·L?1 of yeast extract, 5.00 g·L?1 of NaCl and 2.00 g·L?1 of CaCl2. A 12.55-fold increase of enzyme activity was recorded in the optimized medium compared to the result acquired in a minimum essential medium. The optimized medium was used to compare different cultivation strategies in fermenters. The pH-stat strategy for reducing cellular stress response and the substrate concentration-stat strategy for reducing substrate inhibition were independently investigated. The temperature-limited strategy has been proposed to solve the proteolytic digestion problem, although the high-pressure strategy resulted in high productivity. A hybrid strategy simultaneously controlling pH, temperature, substrate concentration and pO2 was finally investigated to enhance the efficiency of the process. This hybrid strategy resulted in high activity of α-amylase, increasing the productivity almost three-fold as compared to an ordinary fed-batch culture.  相似文献   

5.
We constructed a reporter system to detect a superoxide-generating methyl viologen using SoxRS of Escherichia coli and GFP of Aequorea victoria. E. coli carrying this plasmid exhibited strong fluorescence when grown in the presence of a superoxide-generating reagent methyl viologen. The fluorescence intensity observed in the stationary phase culture of the transformant increased in response to the methyl viologen concentration in a range of 0.01 μM to 10 μM.  相似文献   

6.
The gene for a novel glucanotransferase, isocyclomaltooligosaccharide glucanotransferase (IgtY), involved in the synthesis of a cyclomaltopentaose cyclized by an α-1,6-linkage [ICG5; cyclo-{→6)-α-D-Glcp-(1→4)-α-D-Glcp-(1→4)-α-D-Glcp-(1→4)-α-D-Glcp-(1→4)-α-D-Glcp-(1→}] from starch, was cloned from the genome of B. circulans AM7. The IgtY gene, designated igtY, consisted of 2,985 bp encoding a signal peptide of 35 amino acids and a mature protein of 960 amino acids with a calculated molecular mass of 102,071 Da. The deduced amino-acid sequence showed similarities to 6-α-maltosyltransferase, α-amylase, and cyclomaltodextrin glucanotransferase. The four conserved regions common in the α-amylase family enzymes were also found in this enzyme, indicating that this enzyme should be assigned to this family. The DNA sequence of 8,325-bp analyzed in this study contained two open reading frames (ORFs) downstream of igtY. The first ORF, designated igtZ, formed a gene cluster, igtYZ. The amino-acid sequence deduced from igtZ exhibited no similarity to any proteins with known or unknown functions. IgtZ was expressed in Escherichia coli, and the enzyme was purified. The enzyme acted on maltooligosaccharides that have a degree of polymerization (DP) of 4 or more, amylose, and soluble starch to produce glucose and maltooligosaccharides up to DP5 by a hydrolysis reaction. The enzyme (IgtZ), which has a novel amino-acid sequence, should be assigned to α-amylase. It is notable that both IgtY and IgtZ have a tandem sequence similar to a carbohydrate-binding module belonging to a family 25. These two enzymes jointly acted on raw starch, and efficiently generated ICG5.  相似文献   

7.
An α-amylase which produces maltohexaose as the main product from strach was found in the culture filtrate of Bacillus circulans G-6 which was isolated from soil and identified by the author.

The enzyme was purified by means of ammonium sulfate fractionation, DEAE-Sepharose column chromatography and Sephadex G-200 column chromatography. The purified enzyme was homogeneous on disc electrophoresis. The optimum pH and temperature of the enzyme were around pH 8.0 and around 60°C, respectively. The enzyme was stable in the range of pH 5–10. Metal ions such as Hg2+, Cu2+, Zn2+, Fe2+ and Co2+ inhibited the enzyme activity. The molecular weight was about 76,000. The yield of maltohexaose from soluble starch of DE (dextrose equivalent*) 1.8-12.6 was about 30%, and the combined action of the enzyme and pullulanase or isoamylase increased the yield of maltohexaose.  相似文献   

8.
《Process Biochemistry》2004,39(11):1745-1749
A moderately thermophilic Bacillus subtilis strain, isolated from fresh sheep’s milk, produced extracellular thermostable α-amylase. Maximum amylase production was obtained at 40 °C in a medium containing low starch concentrations. The enzyme displayed maximal activity at 135 °C and pH 6.5 and its thermostability was enhanced in the presence of either calcium or starch. This thermostable α-amylase was used for the hydrolysis of various starches. An ammonium sulphate crude enzyme preparation as well as the cell-free supernatant efficiently degraded the starches tested. The use of the clear supernatant as enzyme source is highly advantageous mainly because it decreases the cost of the hydrolysis. Upon increase of reaction temperature to 70 °C, all substrates exhibited higher hydrolysis rates. Potato starch hydrolysis resulted in a higher yield of reducing sugars in comparison to the other starches at all temperatures tested. Soluble and rice starch took, respectively, the second and third position regarding reducing sugars liberation, while the α-amylase studied showed slightly lower affinity for corn starch and oat starch.  相似文献   

9.
Using a method consisting of two repetitions of “prophage transformation,” the thermostable α-amylase gene in Bacillus subtilis has been cloned in temperate phage ρ11.  相似文献   

10.
An artificially inserted extra peptide (21 amino acid peptide) between the B. subtilis α-amylase signal peptide and the mature thermostable α-amylase was completely cleaved by B. subtilis alkaline protease in vitro. The cleavage to form a mature enzyme was observed between pH 7.5 and 10, but not between pH 6.0 and 6.5, although a similar protease activity toward Azocall was observed between pH 6.0 and 7.5. To analyze the effects of pH on the cleavage, CD spectra at pH 6, 8, and 11 of the NH2-terminally extended thermostable α-amylase were analyzed and the results were compared with those of the mature form of the α-amylase. It is suggesteded that the cleavage of the NH2-terminally extended peptide is controlled by the secondary and tertiary structure of the precursor enzyme. Similar cleavage of different NH2-terminally extended peptides by the alkaline protease was also found in other hybrid thermostable α-amylases obtained.  相似文献   

11.
Aspergillus oryzae RIB40 has three α-amylase genes (amyA, amyB, and amyC), and secretes α-amylase abundantly. However, large amounts of endogenous secretory proteins such as α-amylase can compete with heterologous protein in the secretory pathway and decrease its production yields. In this study, we examined the effects of suppression of α-amylase on heterologous protein production in A. oryzae, using the bovine chymosin (CHY) as a reporter heterologous protein. The three α-amylase genes in A. oryzae have nearly identical DNA sequences from those promoters to the coding regions. Hence we performed silencing of α-amylase genes by RNA interference (RNAi) in the A. oryzae CHY producing strain. The silenced strains exhibited a reduction in α-amylase activity and an increase in CHY production in the culture medium. This result suggests that suppression of α-amylase is effective in heterologous protein production in A. oryzae.  相似文献   

12.
Bacillus subtillis ATCC 21770 was entrapped in a carrageenan gel, especially formulated for immobilization. Bacterial growth and α-amylase (1,4-α-d-glucan glucanohydrolase EC 3.2.1.1) production were tested. The bead suspensions were submitted to two aeration modes, one consisting of bubbling air into a round flask, the other involving sparging of air into an airlift fermenter. The latter system, which produces microbubbles, gave 40–70% increase in enzyme production over the former and a doubling of bacterial density within the beads was measured. The use of CaCl2instead of KCl as polymerization agent led to a better yield of α-amylase.  相似文献   

13.
α-Amylase from the antarctic psychrophile Alteromonas haloplanktis is synthesized at 0 ± 2°C by the wild strain. This heat-labile α-amylase folds correctly when overexpressed in Escherichia coli, providing the culture temperature is sufficiently low to avoid irreversible denaturation. In the described expression system, a compromise between enzyme stability and E. coli growth rate is reached at 18°C.Psychrophilic enzymes possess specific properties, such as high activity at low temperatures and weak thermal stability, which promise to allow the use of these enzymes as industrial biocatalysts, as biotechnological tools, or for fundamental research (6, 8, 11). For instance, substantial energy savings can be obtained if heating is not required during large-scale processes which take advantage of the efficient catalytic capacity of cold-adapted enzymes in the range 0 to 20°C. The pronounced heat lability of psychrophilic enzymes also allows their selective inactivation in a complex mixture, as illustrated by an antarctic bacterial alkaline phosphatase which is available for molecular biology research (7). Finally, psychrophilic enzymes represent the lower natural limit of protein stability (3) and are useful tools for studies in the field of protein folding.Large-scale fermentation of psychrophilic microorganisms suffers from two main drawbacks, however: the low production levels of wild strains and the prohibitive cost of growing wild strains at low temperatures. A possible alternative is to overexpress the gene coding for a psychrophilic protein in a mesophilic host for which efficient expression systems have been designed. In this context, two crucial questions remain to be solved: (i) what is the folding state of an enzyme normally synthesized at 0°C when it is expressed by the mesophilic genetic machinery at higher temperatures, and (ii) is there a temperature at which a compromise can be reached between the stability of the psychrophilic enzyme and the mesophilic growth rate? To address these questions, the heat-labile α-amylase from the antarctic psychrophile Alteromonas haloplanktis (2, 4) was expressed in Escherichia coli at various temperatures.

Construction of the expression vector and α-amylase production.

The α-amylase gene (2) was cloned downstream from the lacZ promoter in pUC12 by ligating the SmaI site of the polylinker to the HpaI site located 60 nucleotides upstream from the formylmethionine codon. This construction is devoid of the C-terminal peptide cleaved by the wild strain following α-amylase secretion. The recombinant enzyme was expressed in E. coli RR1 with the constitutive assistance of lacZ (without IPTG [isopropyl-β-d-thiogalactopyranoside] induction) in a medium containing 16 g of bactotryptone, 16 g of yeast extract, 5 g of NaCl, 2.5 g of K2HPO4, 0.1 μM CaCl2, and 100 mg of ampicillin per liter. The effect of the culture temperature on α-amylase production by E. coli is illustrated in Fig. Fig.1.1. Within the range of temperatures used, maximal enzyme production was reached below 18°C, whereas higher temperatures induced a gradual decrease of α-amylase activity in cultures. Three independent cultures were pooled for the purification of the recombinant enzymes produced at 18 and 25°C. Open in a separate windowFIG. 1Temperature dependence of α-amylase production by E. coli. Results are expressed as percent mean maximal activity recorded at 18°C.

α-Amylase purification.

The gram-negative A. haloplanktis was cultivated at 4°C, and α-amylase was purified from the culture supernatants by ion-exchange chromatography on DEAE-agarose followed by gel filtration on Sephadex G-100 and Ultrogel AcA54 as previously described (2, 4). The recombinant α-amylases were purified by the protocol developed for the wild-type enzyme except that concentration by ammonium sulfate precipitation at 70% saturation was required before the first chromatographic step. Recombinant enzyme production at 18 and 25°C ranged between 60 and 100 mg/liter of culture, which corresponds to a 10-fold improvement over production by the wild strain.

Characterization of the recombinant α-amylases.

N- and C-terminal amino acid sequences (determined on an Applied Biosystems Procise analyzer and by carboxypeptidase Y digestion, respectively) of α-amylase produced at 18 and 25°C indicated that the signal peptide is correctly cleaved in E. coli and that no additional posttranslational cleavage occurred. The isoelectric point (5.5) and the molecular mass (49,340 Da as determined from the sequence and 49,342 ± 8 Da as determined from electrospray mass spectroscopy measurements) were identical to the values recorded for the wild-type enzyme. Dynamic light scattering (DynaPro-801; DLS Instruments) also showed that the purified recombinant enzymes are homogeneous, without any evidence of aggregated forms.

Comparison of the wild-type and recombinant α-amylases.

Several properties of the wild-type enzyme produced at 4°C and the recombinant α-amylase expressed in E. coli at 18°C were compared (Table (Table1).1).

TABLE 1

Kinetic parameters, dissociation constants, and free thiol groups for the wild-type and recombinant α-amylases
α-Amylasekcat (s−1)Km (μM)kcat/Km (s−1 · μM−1)Kd
Cysteinesa (mol−1)Free thiol (mol−1)
Cl (mM)Ca (M)
Wild-type (produced at 4°C)780 ± 25174 ± 84.65.9 ± 0.22.10−880.03
Recombinant (produced at 18°C)792 ± 34168 ± 144.76.1 ± 0.22.10−880.05
Recombinant (produced at 25°C)609 ± 29186 ± 223.36.0 ± 0.32.10−880.05
Open in a separate windowaFrom the amino acid sequence. 

(i) Kinetic and ion binding parameters.

4-Nitrophenyl-α-d-maltoheptaoside-4,6-O-ethylidene (EPS) was used as the substrate in a coupled assay with α-glucosidase at 25°C. The absorption coefficient for 4-nitrophenol was 8,990 M−1 · cm−1 at 405 nm, and a stoichiometric factor of 1.25 was applied for kcat (turnover number) calculation. Dissociation constants were determined by activation kinetics following Cl or Ca2+ titration of the apoenzyme obtained by dialysis against 25 mM HEPES-NaOH (pH 7.2) and 25 mM HEPES-NaOH–5 mM EGTA (pH 8.0), respectively. The saturation curves were computer fitted by a nonlinear regression analysis of the Hill equation in the form v = kcat [I]h/Kd + [I]h where [I] is the ion concentration and h is the Hill coefficient. The free calcium concentrations were set by calcium titration in the presence of 5 mM EGTA at pH 8.0. Kinetic parameters (kcat, Km and kcat/Km) for the hydrolysis of EPS as well as dissociation constants (Kd) for calcium and chloride ions were found to be identical in the wild-type and recombinant enzymes produced at 18°C (Table (Table1).1). Owing to the stringent structural requirements for functional active site and ion binding site conformation, it can be safely concluded that the recombinant enzyme is properly folded at 18°C.

(ii) Disulfide bond integrity.

Free thiol content was determined by DTNB (5,5′-dithiobis-2-nitrobenzoic acid) titration of the unfolded enzyme in 8 M urea in order to promote −SH group accessibility. The eight cysteine residues of A. haloplanktis α-amylase are engaged in disulfide linkages (4). Thus, the lack of free sulfhydryl groups, as detected by DTNB titration of both the native and the unfolded enzymes (Table (Table1),1), indicates that the four disulfide bonds are formed in the recombinant α-amylase samples.

(iii) Conformational stability.

Fluorescence intensity of α-amylases (50 μg/ml) was recorded in 30 mM MOPS (morpholinepropanesulfonic acid)–50 mM NaCl–1 mM CaCl2 (pH 7.2) at a scanning rate of 1°C/min and at an excitation wavelength of 280 nm and an emission wavelength of 347 nm with a Perkin-Elmer LS 50 spectrofluorimeter. Raw data were corrected for thermal dependence of the fluorescence by using the slopes of the pre- and posttransition regions as described elsewhere (10). The conformational stability (ΔGN⇔U) was determined by reversible, thermally induced unfolding recorded by fluorescence. Both the wild-type and the recombinant α-amylases have melting point (Tm) values of 45 ± 0.2°C and display the same cooperative transition (Fig. (Fig.2).2). Consequently, plots of ΔG as a function of T (constructed by using the relation ΔG = −RTlnK, where K = fraction unfolded/fraction folded) are similar (Fig. (Fig.2,2, inset). These results indicate that the weak interactions stabilizing the folded state of the wild-type and recombinant α-amylases are quantitatively identical. Open in a separate windowFIG. 2Heat-induced unfolding transitions of the wild-type α-amylase (•) and the recombinant enzyme produced at 18°C (○). The fraction of protein in the unfolded state (fU) was calculated as follows: fU = (yF − y)/(yF − yU), where yF and yU are the fluorescence intensities of the native and the fully unfolded states, respectively, and y is the fluorescence intensity at a given temperature. The inset shows a plot of ΔG as a function of the temperature around the melting point (Tm), where ΔG = 0.

Expression at 25 and 37°C.

When cultures of the recombinant E. coli are carried out at 25°C, all parameters determined by activation kinetics and independent of the enzyme concentration, such as Km and Kd, remain constant, as does the free sulfhydryl content (Table (Table1).1). This indicates that the native enzyme fraction is correctly folded. By contrast, the kcat of the recombinant α-amylase is reduced by about 20%, suggesting the occurrence of a corresponding inactive fraction. When expressed at 37°C, no α-amylase activity is recorded; the recombinant heat-labile enzyme could fail to fold at this high temperature, or its denaturation rate could exceed its synthesis rate. Furthermore, Western blotting with rabbit polyclonal antibodies to α-amylase detects only trace amounts of the recombinant gene product, suggesting that the denatured enzyme is quickly degraded by the E. coli cell.

Conclusions.

We have previously shown that cloning of a psychrophilic gene in E. coli and detection of the gene product can be achieved by careful control of the culture conditions: overnight incubation at 25°C of transformed cells followed by 1 to 2 days of incubation at 4°C produced halos of substrate hydrolysis on agar plates (5). The folding state of the recombinant psychrophilic enzymes (e.g., fully or partly active, native or marginal stability, etc.), however, was unknown. The results presented here demonstrate that the genuine properties of a psychrophilic enzyme are preserved when it is expressed in a mesophilic host, providing the culture temperature is sufficiently low to allow correct folding and to avoid irreversible denaturation. In our expression system, a compromise is reached between the stability of the psychrophilic enzyme and the growth rate of the mesophilic host by cultivating the recombinant E. coli at 18°C. It should be noted that commonly used E. coli strains have different growing capacities at that temperature. We found E. coli RR1, HB101, or XL1-Blue (Stratagene) suitable for these culture conditions (the generation times are about 3 h, and stationary phase is reached after approximately 30 h), whereas E. coli DH5α grows twice as slowly at 18°C.The lack of α-amylase expression at 37°C is not an isolated case: under the same conditions, lipases and proteases (1, 5, 9) from antarctic psychrophiles were not expressed in an active form. This illustrates the general heat lability of psychrophilic enzymes, which is thought to arise from their flexible conformation, allowing high catalytic activity at temperatures close to 0°C (3).  相似文献   

14.
IntroductionIt is widely accepted that ion-pair increases rigidity and thermostability. There are numerous studies on ion-pairs and thermostability, but none are available about the effect of ion-pair on the activity of enzymes. This paper studies whether an ion-pair allows flexible movement in an enzyme molecule and affects its activity.Materials and methodsIon-pairs are designed at the α-helix region of a Bacillus circulans xylanase, and they are far from the active-sites (23.85–25.15 Å). Two ion-pairing mutations are situated at the C-terminus (D151/E151-K154 ion-pairs) of the helix. One mutation is double-site (F48R-N151D), which introduces both the tertiary (R48-D151) and intra-helical (D151-K154) ion-pairs.Results and discussionAll of the mutants enhanced the catalytic efficiency against xylan (1.66–3.58 times). The double-site mutation showed a synergistic effect on the activity. Overall, the ion-pairs decreased the flexibility (increased rigidity) of the α-helix region and increased the active-site flexibility. The ion-pairs were destabilizing and surface-located; this means that the weaker destabilizing ion-pair still allows flexible movement in the active-site. There is higher mobility of the strand B4 where the active site residue E172 is located. Moreover, the residues lining the active-site cleft (strand B8) showed increased flexibility upon substrate binding.ConclusionIncrease in the activity was due to the increase in active-site flexibility and increased mobility of the residues lining the active-site cleft (strand B8).  相似文献   

15.
A Bacillus strain was isolated from soil samples from the campus area of Dicle University. Based on 16S ribosomal RNA sequencing, the microorganism was closely related to Bacillus subtilis. Effects of different culture medium, incubation time, carbon and nitrogen sources, and various starches, flours, and chemicals on α-amylase production were examined. Maximum enzyme production (7516 U/mL) was obtained in a basal medium A containing 0.05% Tween 40 in 24 h. Partially purified enzyme showed maximum activity at 60 °C with an optimum pH of 6.0. The effects of 0.2% detergents (sodium dodecyl sulfate [SDS], CHAPS [3-[(3-cholamidopropyl)dimethylammonio]-1-propanesulfonate], and commercial detergent Omo Matic) on partially purified enzyme activity over a period of time (15-150 min) were examined and the order of inhibition effect from the most to the least was found as SDS > Omo Matic > CHAPS. Different metal ions inhibited α-amylase activity at low concentrations (1.5 mM). Co2? was a mild inhibitor and Hg2? and Cd2? were potent inhibitors, whereas Ca2? and Mg2? increased the enzyme activity. At 20 mM, Ca2? enhanced enzyme activity, and different Ca2? concentrations (10-300 mM) were studied.  相似文献   

16.
We have demonstrated that a mixture of wheat bran (35 g l-1), as a main substrate, and palm seed powder (10 g l-1), as a co-substrate, is appropriate for -mannanase production by Bacillus subtilis. A 2n factorial experimental design was employed as a primary step for medium optimization. The enzyme activity titters obtained at the optimized growth condition were equivalent to about 319% of the -mannanse activity and 114% of the specific activity levels reached by a galactomannan-based culture.  相似文献   

17.
A novel liquefying α-amylase (LAMY) was found in cultures of an alkaliphilic Bacillus isolate, KSM-1378. The specific activity of purified LAMY was approximately 5,000 U mg of protein−1, a value two- to fivefold greater between pH 5 and 10 than that of an industrial, thermostable Bacillus licheniformis enzyme. The enzyme had a pH optimum of 8.0 to 8.5 and displayed maximum activity at 55°C. The molecular mass deduced from sodium dodecyl sulfate-polyacrylamide gel electrophoresis was approximately 53 kDa, and the apparent isoelectric point was around pH 9. This enzyme efficiently hydrolyzed various carbohydrates to yield maltotriose, maltopentaose, maltohexaose, and maltose as major end products after completion of the reaction. Maltooligosaccharides in the maltose-to-maltopentaose range were unhydrolyzable by the enzyme. The structural gene for LAMY contained a single open reading frame 1,548 bp in length, corresponding to 516 amino acids that included a signal peptide of 31 amino acids. The calculated molecular mass of the extracellular mature enzyme was 55,391 Da. LAMY exhibited relatively low amino acid identity to other liquefying amylases, such as the enzymes from B. licheniformis (68.9%), Bacillus amyloliquefaciens (66.7%), and Bacillus stearothermophilus (68.6%). The four conserved regions, designated I, II, III, and IV, and the putative catalytic triad were found in the deduced amino acid sequence of LAMY. Essentially, the sequence of LAMY was consistent with the tertiary structures of reported amylolytic enzymes, which are composed of domains A, B, and C and which include the well-known (α/β)8 barrel motif in domain A.α-Amylase (1,4-α-d-glucan glucanohydrolase [EC 3.2.1.1]) and pullulanase (pullulan 6-glucanohydrolase [EC 3.2.1.41]) are amylolytic enzymes of industrial importance, particularly in the food and detergent industries. We have found and characterized some unique debranching enzymes, such as a high-alkaline pullulanase (2), an alkali-resistant neopullulanase (16), and an alkaline isoamylase (3), from cultures of alkaliphilic Bacillus strains, and these enzymes can be used as effective additives in dishwashing and laundry detergents under alkaline conditions, especially when used in combination with α-amylase. We have also found the first known alkaline amylopullulanase from alkaliphilic Bacillus sp. strain KSM-1378 (4), which is very unique in that it efficiently hydrolyzes the α-1,6 linkages of pullulan, as well as the α-1,4 linkages of various carbohydrates at different active sites (1, 13).Liquefying α-amylases, particularly the Bacillus licheniformis enzyme (BLA) (35), are used widely in technical application fields, such as in bread making, production of glucose and fructose syrup and fuel ethanol from starch materials, and textile treatment. The demand for α-amylase for use in laundry and automatic dishwashing detergents has also been growing for several years (42). However, most of the Bacillus liquefying amylases, such as the enzymes from Bacillus amyloliquefaciens (BAA) and Bacillus stearothermophilus (BSA) (28), including BLA (35), have pH optima of between 5 and 7.5 (44). These neutrophilic enzymes are essentially not good for use in detergents, because the working pH range between 8 and 11 is relevant to washing in detergents (17). Since Horikoshi (15) first reported an alkaline amylase from alkaliphilic Bacillus sp. strain A-40-2, many alkaline amylases have been found in cultures of, for example, Bacillus sp. strain NRRL B-3881 (31), Bacillus sp. strain H-167 (14), Bacillus alcalothermophilus A3-8 (7), and Bacillus sp. strain GM8901 (21). The alkaline amylases from these alkaliphilic Bacillus strains reported to date are all of the saccharifying type, except for the enzymes from Bacillus sp. strain 707 (22, 41) and B. licheniformis TCRDC-B13 (5). However, very limited or no information about enzymatic properties of these two liquefying amylases is available. In this paper, we report the isolation of a novel liquefying α-amylase (LAMY) from cultures of the alkaline amylopullulanase producer Bacillus sp. strain KSM-1378 (13). This enzyme is highly active at alkaline pH compared with those of other liquefying α-amylases reported to date. Furthermore, analysis of the gene for this α-amylase (amyK) indicates that LAMY exhibits low amino acid identity to the reported liquefying α-amylases.  相似文献   

18.
In this study, the Taguchi experimental design was applied to optimize the conditions for α-amylase production by Bacillus subtilis RSKK96, which was purchased from Refik Saydam Hifzissihha Industry (RSHM). Four factors, namely, carbon source, nitrogen source, amino acid, and fermentation time, each at four levels, were selected, and an orthogonal array layout of L(16) (4(5)) was performed. The model equation obtained was validated experimentally at maximum casein (1%), corn meal (1%), and glutamic acid (0.01%) concentrations with incubation time to 72 h in the presence of 1% inoculum density. Point prediction of the design showed that maximum α-amylase production of 503.26 U/mg was achieved under optimal experimental conditions.  相似文献   

19.
An extracellular amylase (AmyKS) produced by a newly isolated Bacillus subtilis strain US572 was purified and characterized. AmyKS showed maximal activity at pH 6 and 60°C with a half-life of 10 min at 70°C. It is a Ca2+ independent enzyme and able to hydrolyze soluble starch into oligosaccharides consisting mainly of maltose and maltotriose. When compared to the studied α-amylases, AmyKS presents a high affinity toward soluble starch with a Km value of 0.252 mg ml−1. Coupled with the size-exclusion chromatography data, MALDI–TOF/MS analysis indicated that the purified amylase is a dimer with a molecular mass of 136,938.18 Da. It is an unusual feature of a non-maltogenic α-amylase. A 3D model and a dimeric model of AmyKS were generated showing the presence of an additional domain suspected to be involved in the dimerization process. This dimer arrangement could explain the high substrate affinity and catalytic efficiency of this enzyme.  相似文献   

20.
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