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.
3‐Fucosyllactose (3‐FL), one of the major oligosaccharides in human breast milk, is produced in engineered Escherichia coli. In order to search for a good α‐1,3‐fucosyltransferase, three bacterial α‐1,3‐fucosyltransferases are expressed in engineered E. coli deficient in β‐galactosidase activity and expressing the essential enzymes for the production of guanosine 5′‐diphosphate‐l ‐fucose, the donor of fucose for 3‐FL biosynthesis. Among the three enzymes tested, the fucT gene from Helicobacter pylori National Collection of Type Cultures 11637 gives the best 3‐FL production in a simple batch fermentation process using glycerol as a carbon source and lactose as an acceptor. In order to use glucose as a carbon source, the chromosomal ptsG gene, considered the main regulator of the glucose repression mechanism, is disrupted. The resulting E. coli strain of ?LP‐YA+FT shows a much lower performance of 3‐FL production (4.50 g L?1) than the ?L‐YA+FT strain grown in a glycerol medium (10.7 g L?1), suggesting that glycerol is a better carbon source than glucose. Finally, the engineered E. coli ?LW‐YA+FT expressing the essential genes for 3‐FL production and blocking the colanic acid biosynthetic pathway (?wcaJ) exhibits the highest concentration (11.5 g L?1), yield (0.39 mol mol?1), and productivity (0.22 g L?1 h) of 3‐FL in glycerol‐limited fed‐batch fermentation.  相似文献   

2.
Galactooligosaccharides (GOS) are prebiotics produced from lactose through an enzymatic reaction. Employing an immobilized enzyme may result in cost reductions; however, the changes in its kinetics due to immobilization has not been studied. This study experimentally determined the optimal reaction conditions for the production of GOS from lactose by β‐galactosidase (EC 3.2.1.23) from Kluyveromyces lactis covalently immobilized to a polysiloxane‐polyvinyl alcohol (POS‐PVA) polymer activated with glutaraldehyde (GA), and to study the transgalactosylation kinetics. Yield immobilization was 99 ± 1.1% with 78.5 ± 2.4% enzyme activity recovery. An experimental design 24 with 1 center point and 2 replicates was used. Factors were lactose [L], enzyme concentration [E], pH and temperature (T). Response variables were glucose and galactose as monosaccharides [G1], residual lactose [Lac]r and GOS as disaccharides [G2] and trisaccharides [G3]. Best conditions were pH 7.1, 40 °C, 270 gL?1 initial lactose concentration and 6 U mL?1 enzyme concentration, obtaining 25.46 ± 0.01 gL?1 yield of trisaccharides. Although below the HPLC‐IR detection limit, tetrasaccharides were also identified after 115 min of reaction. The immobilization protocol was then optimized by diminishing total reactant volumes : support ratio, resulting in improved enzyme activity synthesizing 43.53 ± 0.02 gL?1 of trisaccharides and 13.79 ± 0.21 gL?1 of tetrasaccharides, and after four cycles remaining relative activity was 94%. A reaction mechanism was proposed through which a mathematical model was developed and rate constants were estimated, considering a pseudo steady‐state hypothesis for two concomitant reactions, and from this simplified analysis, the reaction yield could eventually be improved. © 2017 American Institute of Chemical Engineers Biotechnol. Prog., 33:1568–1578, 2017  相似文献   

3.
The enzyme encapsulation is a very well‐known stabilization pathway. However, there are some challenges in order to avoid the enzyme denaturation under encapsulation conditions. The β‐galactosidase from Bacillus circulans was immobilized through sol‐gel encapsulation route assisted by Triton X‐100 surfactant and sugars. The effects of sugar presence in the immobilization process and the gelation time on the biocatalyst activity/stability were explained taking into account the characteristics of the formed silica matrix and the changes of the enzyme environment. The enzyme was effectively immobilized by this strategy, with high immobilization yield in terms of activity (29%) and expressed activity (47 IU/g). The immobilization through silica sol‐gel in the presence of 1×10?3 M Triton X‐100 and fructose conferred 28.4‐fold higher stability to the enzyme compared with the soluble form. This is an advantage for its use in the synthesis of the galacto‐oligosaccharides at 50ºC. The total lactose conversion to galacto‐oligosaccharides was 26%wt, which is comparable with that reported in the literature. The obtained biocatalyst is useful for the synthesis of galacto‐oligosaccharides and its catalytic behavior is rationalized in this work.  相似文献   

4.
A pseudo steady‐state model for the kinetically controlled synthesis of galacto‐oligosaccharides (GOS) with Aspergillus oryzae β‐galactosidase is presented. The model accounts for the dynamics of lactose consumption and production of galactose, glucose, di, tri, tetra, and penta‐oligosaccharides during the synthesis, being able to describe the total GOS content in the reaction medium at the experimental conditions evaluated. Experimental results show that the formation of GOS containing only galactose residues is significant at high conversions of substrate, which was taken into account in the model. The formation of enzyme transition complexes was considered and reasonable assumptions were made to reduce the number of parameters to be determined. The model developed has 8 parameters; 2 of them were experimentally determined and the other 6 were estimated by fitting to the experimental data using multiresponse regression. Temperature effect on kinetic and affinity constants was determined in the range from 40 to 55°C, and the data were fitted to Arrhenius type equation. Parameters of the proposed model are independent from the enzyme load in the reaction medium and, differently from previously reported models, they have a clear biochemical meaning. The magnitude of the kinetic and affinity constants of the enzyme suggests that the liberation of galactose from the galactosyl–enzyme complex is a very slow reaction and such complex is driven into GOS formation. It also suggests that the affinity for sugars of the galactosyl–enzyme complex is higher than that of the free enzyme. Biotechnol. Bioeng. 2011;108: 2270–2279. © 2011 Wiley Periodicals, Inc.  相似文献   

5.
The herbaceous perennial energy crops miscanthus, giant reed, and switchgrass, along with the annual crop residue corn stover, were evaluated for their bioconversion potential. A co‐hydrolysis process, which applied dilute acid pretreatment, directly followed by enzymatic saccharification without detoxification and liquid–solid separation between these two steps was implemented to convert lignocellulose into monomeric sugars (glucose and xylose). A factorial experiment in a randomized block design was employed to optimize the co‐hydrolysis process. Under the optimal reaction conditions, corn stover exhibited the greatest total sugar yield (glucose + xylose) at 0.545 g g?1 dry biomass at 83.3% of the theoretical yield, followed by switch grass (0.44 g g?1 dry biomass, 65.8% of theoretical yield), giant reed (0.355 g g?1 dry biomass, 64.7% of theoretical yield), and miscanthus (0.349 g g?1 dry biomass, 58.1% of theoretical yield). The influence of combined severity factor on the susceptibility of pretreated substrates to enzymatic hydrolysis was clearly discernible, showing that co‐hydrolysis is a technically feasible approach to release sugars from lignocellulosic biomass. The oleaginous fungus Mortierella isabellina was selected and applied to the co‐hydrolysate mediums to accumulate fungal lipids due to its capability of utilizing both C5 and C6 sugars. Fungal cultivations grown on the co‐hydrolysates exhibited comparable cell mass and lipid production to the synthetic medium with pure glucose and xylose. These results elucidated that combining fungal fermentation and co‐hydrolysis to accumulate lipids could have the potential to enhance the utilization efficiency of lignocellulosic biomass for advanced biofuels production. Biotechnol. Bioeng. 2013; 110: 1039–1049. © 2012 Wiley Periodicals, Inc.  相似文献   

6.
Pichia pastoris has become one of the major microorganisms for the production of proteins in recent years. This development was mainly driven by the readily available genetic tools and the ease of high‐cell density cultivations using methanol (or methanol/glycerol mixtures) as inducer and carbon source. To overcome the observed limitations of methanol use such as high heat development, cell lysis, and explosion hazard, we here revisited the possibility to produce proteins with P. pastoris using glucose as sole carbon source. Using a recombinant P. pastoris strain in glucose limited fed‐batch cultivations, very high‐cell densities were reached (more than 200 gCDW L?1) resulting in a recombinant protein titer of about 6.5 g L?1. To investigate the impact of recombinant protein production and high‐cell density fermentation on the metabolism of P. pastoris, we used 13C‐tracer‐based metabolic flux analysis in batch and fed‐batch experiments. At a controlled growth rate of 0.12 h?1 in fed‐batch experiments an increased TCA cycle flux of 1.1 mmol g?1 h?1 compared to 0.7 mmol g?1 h?1 for the recombinant and reference strains, respectively, suggest a limited but significant flux rerouting of carbon and energy resources. This change in flux is most likely causal to protein synthesis. In summary, the results highlight the potential of glucose as carbon and energy source, enabling high biomass concentrations and protein titers. The insights into the operation of metabolism during recombinant protein production might guide strain design and fermentation development. Biotechnol. Bioeng. 2010;107: 357–368. © 2010 Wiley Periodicals, Inc.  相似文献   

7.
Klebsiella pneumoniae HR526, a new isolated 1,3‐propanediol (1,3‐PD) producer, exhibited great productivity. However, the accumulation of lactate in the late‐exponential phase remained an obstacle of 1,3‐PD industrial scale production. Hereby, mutants lacking D ‐lactate pathway were constructed by knocking out the ldhA gene encoding fermentative D ‐lactate dehydrogenase (LDH) of HR526. The mutant K. pneumoniae LDH526 with the lowest LDH activity was studied in aerobic fed‐batch fermentation. In experiments using pure glycerol as feedstock, the 1,3‐PD concentrations, conversion, and productivity increased from 95.39 g L?1, 0.48 and 1.98 g L?1 h?1 to 102. 06 g L?1, 0.52 mol mol?1 and 2.13 g L?1 h?1, respectively. The diol (1,3‐PD and 2,3‐butanediol) conversion increased from 0.55 mol mol?1 to a maximum of 0.65 mol mol?1. Lactate would not accumulate until 1,3‐PD exceeded 84 g L?1, and the final lactate concentration decreased dramatically from more than 40 g L?1 to <3 g L?1. Enzymic measurements showed LDH activity decreased by 89–98% during fed‐batch fermentation, and other related enzyme activities were not affected. NADH/NAD+ enhanced more than 50% in the late‐exponential phase as the D ‐lactate pathway was cut off, which might be the main reason for the change of final metabolites concentrations. The ability to utilize crude glycerol from biodiesel process and great genetic stability demonstrated that K. pnemoniae LDH526 was valuable for 1,3‐PD industrial production. Biotechnol. Bioeng. 2009; 104: 965–972. © 2009 Wiley Periodicals, Inc.  相似文献   

8.
Many β‐galactosidases show large differences in galacto‐oligosaccharide (GOS) production and lactose hydrolysis. In this study, a kinetic model is developed in which the effect of lactose, glucose, galactose, and oligosaccharides on the oNPG converting activity of various β‐galactosidases is quantified. The use of oNPG as a competing substrate to lactose yields more information than can be obtained by examining only the conversion of lactose itself. The reaction rate with lactose or oligosaccharides as substrate relative to that with water as acceptor is much higher for the β‐galactosidase of Bacillus circulans than the β‐galactosidases of Aspergillus oryzae and Kluyveromyces lactis. In addition, the β‐galactosidase of B.circulans has a high reaction rate with galactose as acceptor, in contrast to those of A. oryzae and K. lactis. The latter two are strongly inhibited by galactose. These differences explain why β‐galactosidase of B. circulans gives higher yields in GOS production than other β‐galactosidases. Many of the reaction rate constants for the β‐galactosidase isoforms of B. circulans increase with increasing molecular weight of the isoform. This indicates that the largest isoform β‐gal‐A is most active in GOS production. However, its hydrolysis rate is also much higher than that of the other isoforms, which results in a faster hydrolysis of oligosaccharides as well. © 2013 American Institute of Chemical Engineers Biotechnol. Prog., 30:38–47, 2014  相似文献   

9.
The green microalga Chlorella sp. TISTR 8990 was grown heterotrophically in the dark using various concentrations of a basal glucose medium with a carbon‐to‐nitrogen mass ratio of 29:1. The final biomass concentration and the rate of growth were highest in the fivefold concentrated basal glucose medium (25 g L?1 glucose, 2.5 g L?1 KNO3) in batch operations. Improving oxygen transfer in the culture by increasing the agitation rate and decreasing the culture volume in 500‐mL shake flasks improved growth and glucose utilization. A maximum biomass concentration of nearly 12 g L?1 was obtained within 4 days at 300 rpm, 30°C, with a glucose utilization of nearly 76% in batch culture. The total fatty acid (TFA) content of the biomass and the TFA productivity were 102 mg g?1 and 305 mg L?1 day?1, respectively. A repeated fed‐batch culture with four cycles of feeding with the fivefold concentrated medium in a 3‐L bioreactor was evaluated for biomass production. The total culture period was 11 days. A maximum biomass concentration of nearly 26 g L?1 was obtained with a TFA productivity of 223 mg L?1 day?1. The final biomass contained (w/w) 13.5% lipids, 20.8% protein and 17.2% starch. Of the fatty acids produced, 52% (w/w) were saturated, 41% were monounsaturated and 7% were polyunsaturated (PUFA). A low content of PUFA in TFA feedstock is required for producing high quality biodiesel. © 2017 American Institute of Chemical Engineers Biotechnol. Prog., 33:1589–1600, 2017  相似文献   

10.
In these studies, liquid hot water (LHW) pretreated and enzymatically hydrolyzed Sweet Sorghum Bagasse (SSB) hydrolyzates were fermented in a fed‐batch reactor. As reported in the preceding paper, the culture was not able to ferment the hydrolyzate I in a batch process due to presence of high level of toxic chemicals, in particular acetic acid released from SSB during the hydrolytic process. To be able to ferment the hydrolyzate I obtained from 250 g L?1 SSB hydrolysis, a fed‐batch reactor with in situ butanol recovery was devised. The process was started with the hydrolyzate II and when good cell growth and vigorous fermentation were observed, the hydrolyzate I was slowly fed to the reactor. In this manner the culture was able to ferment all the sugars present in both the hydrolyzates to acetone butanol ethanol (ABE). In a control batch reactor in which ABE was produced from glucose, ABE productivity and yield of 0.42 g L?1 h?1 and 0.36 were obtained, respectively. In the fed‐batch reactor fed with SSB hydrolyzates, these productivity and yield values were 0.44 g L?1 h?1 and 0.45, respectively. ABE yield in the integrated system was high due to utilization of acetic acid to convert to ABE. In summary we were able to utilize both the hydrolyzates obtained from LHW pretreated and enzymatically hydrolyzed SSB (250 g L?1) and convert them to ABE. Complete fermentation was possible due to simultaneous recovery of ABE by vacuum. © 2018 American Institute of Chemical Engineers Biotechnol. Prog., 34:967–972, 2018  相似文献   

11.
Schizochytrium mangrovei strain PQ 6 was investigated for coproduction of docosahexaenoic acid (C22: 6ω‐3, DHA ) and squalene using a 30‐L bioreactor with a working volume of 15 L under various batch and fed‐batch fermentation process regimes. The fed‐batch process was a more efficient cultivation strategy for achieving higher biomass production rich in DHA and squalene. The final biomass, total lipid, unsaponifiable lipid content, and DHA productivity were 105.25 g · L?1, 43.40% of dry cell weight, 8.58% total lipid, and 61.66 mg · g?1 · L?1, respectively, after a 96 h fed‐batch fermentation. The squalene content was highest at 48 h after feeding glucose (98.07 mg · g?1 of lipid). Differences in lipid accumulation during fermentation were correlated with changes in ultrastructure using transmission electron microscopy and Nile Red staining of cells. The results may be of relevance to industrial‐scale coproduction of DHA and squalene in heterotrophic marine microalgae such as Schizochytrium .  相似文献   

12.
β‐poly(l ‐malic acid) (PMLA) is a biopolyester, which has attracted growing attention due to its potential applications in medicine and other industries. In this study, the biosynthetic pathway of PMLA and the fermentation strategies with mixed sugars were both investigated to enhance PMLA production by Aureobasidium pullulans ipe‐1. Metabolic intermediates and inhibitors were used to study the biosynthetic pathway of PMLA. It showed that exogenous addition of l ‐malic acid, succinic acid, TFA, and avidin had negligible effect on PMLA production, while pyruvic acid and biotin were the inhibitors, indicating that PMLA biosynthesis was probably related to phosphoenolpyruvate via oxaloacetate catalyzed by phosphoenolpyruvate carboxylase. Sucrose was suitable for achieving the highest PMLA concentration, while fructose generated a higher yield of PMLA (PMLA produced per biomass). Furthermore, the fed‐batch culture using fed solution with different sugar mixture for PMLA production was implemented. During the fed‐batch culture with mixed solution, fructose could increase PMLA production. Compared with the batch culture, the feeding with mixed sugar (sucrose and glucose) increased PMLA concentration by 23.9%, up to 63.2 g/L, and the final volume of the broth was increased by 25%. These results provide a good reference for process development and optimization of PMLA production.  相似文献   

13.
Pandoraea sp. MA03 wild type strain was subjected to UV mutation to obtain mutants unable to grow on propionic acid (PA) but still able to produce poly(3‐hydroxybutyrate‐co‐3‐hydroxyvalerate) [P(3HB‐co‐3HV)] from glycerol and PA at high 3HV yields. In shake flask experiments, mutant prp25 was selected from 52 mutants affected in the propionate metabolism exhibiting a conversion rate of PA into 3HV units of 0.78 g g?1. The use of crude glycerol (CG) plus PA or valeric acid resulted in a copolymer with 3HV contents varying from 21.9 to 30 mol% and 22.2 to 36.7 mol%, respectively. Fed‐batch fermentations were performed using CG and PA and reached a 3HV yield of 1.16 g g?1, which is 86% of the maximum theoretical yield. Nitrogen limitation was a key parameter for polymer accumulation reaching up to 63.7% content and 18.1 mol% of 3HV. Henceforth, mutant prp25 is revealed as an additional alternative to minimize costs and support the P(3HB‐co‐3HV) production from biodiesel by‐products. © 2017 American Institute of Chemical Engineers Biotechnol. Prog., 33:1077–1084, 2017  相似文献   

14.
Chemical synthesis of lactones from cycloalkanes is a multi‐step process challenged by limitations in reaction efficiency (conversion and yield), atom economy (by‐products) and environmental performance. A heterologous pathway comprising novel enzymes with compatible kinetics was designed in Pseudomonas taiwanensis VLB120 enabling in‐vivo cascade for synthesizing lactones from cycloalkanes. The respective pathway included cytochrome P450 monooxygenase (CHX), cyclohexanol dehydrogenase (CDH), and cyclohexanone monooxygenase (CHXON) from Acidovorax sp. CHX100. Resting (non‐growing) cells of the recombinant host P. taiwanensis VLB120 converted cyclohexane, cyclohexanol, and cyclohexanone to ?‐caprolactone at 22, 80–100, and 170 U gCDW?1, respectively. Cyclohexane (5 mM) was completely converted with a selectivity of 65% for ?‐caprolactone formation in 2 hr without accumulation of intermediate products. Promiscuity of the whole‐cell biocatalyst gave access to analogous lactones from cyclooctane and cyclodecane. A total product concentration of 2.3 g L?1 and a total turnover number of 36,720 was achieved over 5 hr with a biocatalyst concentration of 6.8 gCDW L?1.
  相似文献   

15.
Agmatine is a kind of important biogenic amine. The chemical synthesis route is not a desirable choice for industrial production of agmatine. To date, there are no reports on the fermentative production of agmatine by microorganism. In this study, the base Escherichia coli strain AUX4 (JM109 ?speC ?speF ?speB ?argR) capable of excreting agmatine into the culture medium was first constructed by sequential deletions of the speC and speF genes encoding the ornithine decarboxylase isoenzymes, the speB gene encoding agmatine ureohydrolase and the regulation gene argR responsible for the negative control of the arg regulon. The speA gene encoding arginine decarboxylase harboured by the pKK223‐3 plasmid was overexpressed in AUX4, resulting in the engineered strain AUX5. The batch and fed‐batch fermentations of the AUX5 strain were conducted in a 3‐L bioreactor, and the results showed that the AUX5 strain was able to produce 1.13 g agmatine L?1 with the yield of 0.11 g agmatine g?1 glucose in the batch fermentation and the fed‐batch fermentation of AUX5 allowed the production of 15.32 g agmatine L?1 with the productivity of 0.48 g agmatine L?1 h?1, demonstrating the potential of E. coli as an industrial producer of agmatine.  相似文献   

16.
Aims: To characterize the kinetics of growth, sugar uptake and xylitol production in batch and fed‐batch cultures for a xylitol assimilation‐deficient strain of Candida tropicalis isolated via chemical mutagenesis. Methods and Results: Chemical mutagenesis using nitrosoguanidine led to the isolation of the xylitol‐assimilation deficient strain C. tropicalis SS2. Shake‐flask fermentations with this mutant showed a sixfold higher xylitol yield than the parent strain in medium containing 25 g l?1 glucose and 25 g l?1 xylose. With 20 g l?1 glycerol, replacing glucose for cell growth, and various concentrations of xylose, the studies indicated that the mutant strain resulted in xylitol yields from xylose close to theoretical. Under fully aerobic conditions, fed‐batch fermentation with repeated addition of glycerol and xylose resulted in 3·3 g l?1 h?1 xylitol volumetric productivity with the final concentration of 220 g l?1 and overall yield of 0·93 g g?1 xylitol. Conclusions: The xylitol assimilation‐deficient mutant isolated in this study showed the potential for high xylitol yield and volumetric productivity under aerobic conditions. In the evaluation of glycerol as an alternative low‐cost nonfermentable carbon source, high biomass and xylitol yields under aerobic conditions were achieved; however, the increase in initial xylose concentrations resulted in a reduction in biomass yield based on glycerol consumption. This may be a consequence of the role of an active transport system in the yeast requiring increasing energy for xylose uptake and possible xylitol secretion, with little or no energy available from xylose metabolism. Significance and Impact of the Study: The study confirms the advantage of using a xylitol assimilation‐deficient yeast under aerobic conditions for xylitol production with glycerol as a primary carbon source. It illustrates the potential of using the xylose stream in a biomass‐based bio‐refinery for the production of xylitol with further cost reductions resulting from using glycerol for yeast growth and energy production.  相似文献   

17.
The design of an optimal process is particularly crucial when the reactants deactivate the biocatalyst. The reaction cascades of the chemo‐enzymatic epoxidation where the intermediate peroxy acid is produced by an enzyme are still limited by enzyme inhibition and deactivation by hydrogen peroxide. To avoid additional effects caused by interfaces (aq/org) and to reduce the process limiting deactivation by the substrate hydrogen peroxide, a single‐phase concept was applied in a fed‐batch and a continuous process (stirred tank), without the commonly applied addition of a carrier solvent. The synthesis of peroxyoctanoic acid catalyzed by Candida antarctica lipase B was chosen as the model reaction. Here, the feasibility of this biocatalytic reaction in a single‐phase system was shown for the first time. The work shows the economic superiority of the continuous process compared to the fed‐batch process. Employing the fed‐batch process reaction rates up to 36 mmol h?1 per gramcat, and a maximum yield of 96 % was achieved, but activity dropped quickly. In contrast, continuous operation can maintain long‐term enzyme activity. For the first time, the continuous enzymatic reaction could be performed for 55 h without any loss of activity and with a space‐time yield of 154 mmol L?1 h?1, which is three times higher than in the fed‐batch process. The higher catalytic productivity compared to the fed‐batch process (34 vs. 18 gProd g?1cat) shows the increased enzyme stability in the continuous process.  相似文献   

18.
The production of norovirus virus‐like particles (NoV VLPs) displaying NY‐ESO‐1 cancer testis antigen in Pichia pastoris BG11 Mut+ has been enhanced through feed‐strategy optimization using a near‐infrared bioprocess monitor (RTBio® Bioprocess Monitor, ASL Analytical, Inc.), capable of monitoring and controlling the concentrations of glycerol and methanol in real‐time. The production of NoV VLPs displaying NY‐ESO‐1 in P. pastoris has potential as a novel cancer vaccine platform. Optimization of the growth conditions resulted in an almost two‐fold increase in the expression levels in the fermentation supernatant of P. pastoris as compared to the starting conditions. We investigated the effect of methanol concentration, batch phase time, and batch to induction transition on NoV VLP‐NY‐ESO‐1 production. The optimized process included a glycerol transition phase during the first 2 h of induction and a methanol concentration set point of 4 g L?1 during induction. Utilizing the bioprocess monitor to control the glycerol and methanol concentrations during induction resulted in a maximum NoV VP1‐NY‐ESO‐1 yield of 0.85 g L?1. © 2016 American Institute of Chemical Engineers Biotechnol. Prog., 32:518–526, 2016  相似文献   

19.
Aims: To characterize of a thermostable recombinant α‐l ‐arabinofuranosidase from Caldicellulosiruptor saccharolyticus for the hydrolysis of arabino‐oligosaccharides to l ‐arabinose. Methods and Results: A recombinant α‐l ‐arabinofuranosidase from C. saccharolyticus was purified by heat treatment and Hi‐Trap anion exchange chromatography with a specific activity of 28·2 U mg?1. The native enzyme was a 58‐kDa octamer with a molecular mass of 460 kDa, as measured by gel filtration. The catalytic residues and consensus sequences of the glycoside hydrolase 51 family of α‐l ‐arabinofuranosidases were completely conserved in α‐l ‐arabinofuranosidase from C. saccharolyticus. The maximum enzyme activity was observed at pH 5·5 and 80°C with a half‐life of 49 h at 75°C. Among aryl‐glycoside substrates, the enzyme displayed activity only for p‐nitrophenyl‐α‐l ‐arabinofuranoside [maximum kcat/Km of 220 m(mol l?1)?1 s?1] and p‐nitrophenyl‐α‐l ‐arabinopyranoside. This substrate specificity differs from those of other α‐l ‐arabinofuranosidases. In a 1 mmol l?1 solution of each sugar, arabino‐oligosaccharides with 2–5 monomer units were completely hydrolysed to l ‐arabinose within 13 h in the presence of 30 U ml?1 of enzyme at 75°C. Conclusions: The novel substrate specificity and hydrolytic properties for arabino‐oligosaccharides of α‐l ‐arabinofuranosidase from C. saccharolyticus demonstrate the potential in the commercial production of l ‐arabinose in concert with endoarabinanase and/or xylanase. Significance and Impact of the Study: The findings of this work contribute to the knowledge of hydrolytic properties for arabino‐oligosaccharides performed by thermostable α‐l ‐arabinofuranosidase.  相似文献   

20.
Co‐cultures for simultaneous production of ethanol and xylitol were studied under different operation bioreactor modes using Candida tropicalis IEC5‐ITV and Saccharomyces cerevisiae ITV01‐RD in a simulated medium of sugarcane bagasse hydrolyzates. Xylitol and ethanol tolerance by S. cerevisiae and C. tropicalis, respectively, was evaluated. The results showed that C. tropicalis was sensitive to ethanol concentrations up to 30 g/L, while xylitol had no effect on S. cerevisiae viability and metabolism. The best condition found for simultaneous culture was S. cerevisiae co‐culture and C. tropicalis sequential cultivation at 24 h. Under these conditions, productivity and yield for ethanol were QEtOH = 0.72 g L?1 h?1 and YEtOH/s = 0.37 g/g, and for xylitol, QXylOH = 0.10 g L?1 h?1 and YXylOH/S = 0.31 g/g, respectively; using fed‐batch culture, the results were QEtOH = 0.87 g L?1 h?1 and YEtOH/s = 0.44 g L?1 h?1, and QEtOH = 0.27 g L?1 h?1 and YEtOH/s = 0.57 g/g, respectively. Maximum volumetric productivity in continuous multistep cultures of ethanol and xylitol was at dilution rates of 0.131 and 0.074 h?1, respectively. Continuous multistep production, QEtOH increased up to 50% more than in fed‐batch culture, even though xylitol yield remained unchanged.  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