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 20 毫秒
1.
Evidence is presented that xylose metabolism in the anaerobic cellulolytic fungus Piromyces sp. E2 proceeds via a xylose isomerase rather than via the xylose reductase/xylitol-dehydrogenase pathway found in xylose-metabolising yeasts. The XylA gene encoding the Piromyces xylose isomerase was functionally expressed in Saccharomyces cerevisiae. Heterologous isomerase activities in cell extracts, assayed at 30 degrees C, were 0.3-1.1 micromol min(-1) (mg protein)(-1), with a Km for xylose of 20 mM. The engineered S. cerevisiae strain grew very slowly on xylose. It co-consumed xylose in aerobic and anaerobic glucose-limited chemostat cultures at rates of 0.33 and 0.73 mmol (g biomass)(-1) h(-1), respectively.  相似文献   

2.
Regulation of β-xylosidase formation by xylose in Trichoderma reesei   总被引:1,自引:0,他引:1  
The soft-rot fungus Trichoderma reesei forms -xylosidase (EC 3.2.1.37) activity during cultivation on xylan and xylose, but not on glucose. When mycelia precultivated on glycerol were washed and transferred to fresh medium without a carbon and nitrogen source, -xylosidase formation was induced by xylan, xylobiose and xylose. A supply of 4 mm xylose and a pH of 2.5 provided optimal conditions for induction. -Xylosidase accounted for the major portion of total extracellular protein under these conditions, and could be purified to physical homogeneity by a single anion exchange chromatography step. A recombinant strain of T. reesei that carries multiple copies of the homologous xylanase II-encoding gene has a six-fold increased xylanase activity, but forms comparable -xylosidase activities. This shows that the rate of xylan hydrolysis has no effect on the induction of -xylosidase. Methyl--d-xyloside inhibited -xylosidase competitively and was a weak -xylosidase inducer. The induction by xylobiose and xylan was strongly enhanced by the simultaneous addition of methyl--d-xylosidese and xylan or xylobiose. The results suggest that a slow supply of xylose is a trigger for -xylosidase induction.  相似文献   

3.
Clostridium acetobutylicum TISTR 1462 and Clostridium beijerinckii TISTR 1461 were chosen to optimize acetone–butanol–ethanol (ABE) fermentation by using glucose as a carbon source. The enhancement in its productivity by adding various concentrations of ammonium acetate was studied. Then, the variation of glucose/xylose ratios in the pre-grown medium was investigated. The results showed that both increased ammonium acetate in the production medium and D–xylose in the pre-grown medium could produce more ABE. With these conditions, using corncob hydrolysate as a substrate, 20.58 g/L ABE was produced from C. beijerinckii TISTR 1461 with 0.44 g/L/h and 0.45 of ABE productivity and yield, respectively.  相似文献   

4.
Xylose oligomers are the intermediate products of xylan depolymerization into xylose monomers. An understanding of xylan depolymerization kinetics is important to improve the conversion of xylan into monomeric xylose and to minimize the formation of inhibitory products, thereby reducing ethanol production costs. The study of xylan depolymerization requires copious amount of xylose oligomers, which are expensive if acquired commercially. Our approach consisted of producing in-house oligomer material. To this end, birchwood xylan was used as the starting material and hydrolyzed in hot water at 200 °C for 60 min with a 4 % solids loading. The mixture of xylose oligomers was subsequently fractionated by a centrifugal partition chromatography (CPC) with a solvent system of butanol:methanol:water in a 5:1:4 volumetric ratio. Operating in an ascending mode, the butanol-rich upper phase (the mobile phase) eluted xylose oligomers from the water-rich stationary phase at a 4.89 mL/min flow rate for a total fractionation time of 300 min. The elution of xylose oligomers occurred between 110 and 280 min. The yields and purities of xylobiose (DP 2), xylotriose (DP 3), xylotetraose (DP 4), and xylopentaose (DP 5) were 21, 10, 14, and 15 mg/g xylan and 95, 90, 89, and 68 %, respectively. The purities of xylose oligomers from this solvent system were higher than those reported previously using tetrahydrofuran:dimethyl sulfoxide:water in a 6:1:3 volumetric ratio. Moreover, the butanol-based solvent system improved overall procedures by facilitating the evaporation of the solvents from the CPC fractions, rendering the purification process more efficient.  相似文献   

5.
《Journal of biotechnology》1999,67(2-3):159-171
Xylitol production from xylose was studied using recombinant Saccharomyces cerevisiae 2805 containing xylose reductase genes (XYL1) of Pichia stipitis at chromosomal δ-sequences. S. cerevisiae 2805-39-40, which contains about 40 copies of the XYL1 gene on the chromosome, was obtained by a sequential transformation using a dominant selection marker neor and an auxotrophic marker URA3. The multiple XYL1 genes were stably maintained on the chromosome even after 21 and 10 days in the non-selective sequential batch and chemostat cultures, respectively, whereas S. cerevisiae 2805:pVTXR, which harbors the episomal plasmid pVTXR having the XYL1 gene, showed mitotic plasmid instability and more than 95% of the cells lost the plasmid under the same culture conditions. In the first batch (3 days) of the sequential batch culture, volumetric xylitol productivity was 0.18 g l−1 h−1 for S. cerevisiae 2805-39-40, as compared to 0.21 g l−1 h−1 for S. cerevisiae 2805:pVTXR. However, the xylitol productivity of the latter started to decrease rapidly in the third batch and dropped to 0.04 g l−1 h−1 in the seventh batch, whereas the former maintained the stable xylitol productivity at 0.18 g l−1 h−1 through the entire sequential batch culture. The xylitol production level in the chemostat culture was about 8 g l−1 for S. cerevisiae 2805-39-40, as compared to 2.0 g l−1 for S. cerevisiae 2805:pVTXR after 10 days of cultures even though the xylitol production level of the latter was higher than that of the former for the first 5 days. The results of this experiment indicate that S. cerevisiae containing the multiple XYL1 genes on the chromosome is much more efficient for the xylitol production in the long-term non-selective culture than S. cerevisiae harboring the episomal plasmid containing the XYL1 gene.  相似文献   

6.
2′-Fucosyllactose (2′-FL), a human milk oligosaccharide with confirmed benefits for infant health, is a promising infant formula ingredient. Although Escherichia coli, Saccharomyces cerevisiae, Corynebacterium glutamicum, and Bacillus subtilis have been engineered to produce 2′-FL, their titers and productivities need be improved for economic production. Glucose along with lactose have been used as substrates for producing 2′-FL, but accumulation of by-products due to overflow metabolism of glucose hampered efficient production of 2′-FL regardless of a host strain. To circumvent this problem, we used xylose, which is the second most abundant sugar in plant cell wall hydrolysates and is metabolized through oxidative metabolism, for the production of 2′-FL by engineered yeast. Specifically, we modified an engineered S. cerevisiae strain capable of assimilating xylose to produce 2′-FL from a mixture of xylose and lactose. First, a lactose transporter (Lac12) from Kluyveromyces lactis was introduced. Second, a heterologous 2′-FL biosynthetic pathway consisting of enzymes Gmd, WcaG, and WbgL from Escherichia coli was introduced. Third, we adjusted expression levels of the heterologous genes to maximize 2′-FL production. The resulting engineered yeast produced 25.5 g/L of 2′-FL with a volumetric productivity of 0.35 g/L∙h in a fed-batch fermentation with lactose and xylose feeding to mitigate the glucose repression. Interestingly, the major location of produced 2′-FL by the engineered yeast can be changed using different culture media. While 72% of the produced 2′-FL was secreted when a complex medium was used, 82% of the produced 2′-FL remained inside the cells when a minimal medium was used. As yeast extract is already used as food and animal feed ingredients, 2′-FL enriched yeast extract can be produced cost-effectively using the 2′-FL-accumulating yeast cells.  相似文献   

7.
Summary A two-stage culture method employing Lactococcus.lactis IO-1 and Alcaligenes eutrophus was developed for production of poly--hydroxybutyric acid (PHB) from xylose. In this method, xylose was converted to L-lactic acid and acetic acid by L.lactis IO-1, and then the organic acids were converted to PHB by A.eutrophus. When the supernatant of the IO-1 culture broth, containing 10 g·dm-3 L-lactate derived from xylose, was used as medium for A. eutrophus, the concentration of cells increased to 8.5 g·dm-3 in 24 h and 55% of the content in the cells by weight was PHB.  相似文献   

8.
Shi X  Wu Y  Li P  Yi H  Yang M  Wang G 《Carbohydrate research》2011,346(4):480-487
Al-promoted SO? 2?-/ZrO?SBA-15 catalysts were prepared and characterized by XRD, BET, ICP and NH?-TPD techniques. The influence of introducing aluminum on the structure and surface properties of the catalyst and the catalytic activity for dehydration of xylose to furfural has been investigated. The introduction of the Al stabilizes the tetragonal phase of the ZrO? and thus increases the number and intensity of acid sites. Based on the characterization of the deactivated catalyst, the accumulation of byproducts is the main reason for the deactivation of the catalyst. Regeneration with H?O? can completely recover the catalytic activity of the deactivated catalyst.  相似文献   

9.
10.
The N-glycans from soluble extracts of ten pollens were examined. The pyridylaminated oligosaccharides derived from these sources were subject to gel filtration and reverse-phase HPLC, in conjunction with exoglycosidase digests, and in some cases matrix-assisted laser desoprofessional-articletion-ionisation mass spectrometry. In comparison to known structures, it was possible to determine the major structures of the N-glycans derived from Kentucky blue grass (Poa pratensis), rye (Secale cerale), ryegrass (Lolium perenne), short ragweed (Ambrosia elatior), giant ragweed (Ambrosia trifida), birch (Betula alba), hornbeam (Caprofessional-articleinus betulus), horse chestnut (Aesculus hippocastanum), olive (Olea europaea) and snake-skin pine (Pinus leucodermis) pollen extracts. For grass pollens the major glycans detected were identical in properties to:$&1[-]Grass pollens also contained some minor structures with one or two non-reducing terminal N-acetylglucosamine residues. In the ragweed pollens, the major structures carried core 1,3-linked fucose with or without the presence of xylose. In tree pollen extracts, the major structures were either xylosylated, with or without fucose and terminal N-acetylglucosamine residues, with also significant amounts of oligomannose structures. These results are compatible with the hypothesis that the carbohydrate structures are another potential source of immunological cross-reaction between different plant allergens.  相似文献   

11.
Several alcohol dehydrogenase (ADH)-related genes have been identified as enzymes for reducing levels of toxic compounds, such as, furfural and/or 5-hydroxymethylfurfural (5-HMF), in hydrolysates of pretreated lignocelluloses. To date, overexpression of these ADH genes in yeast cells have aided ethanol production from glucose or glucose/xylose mixture in the presence of furfural or 5-HMF. However, the effects of these ADH isozymes on ethanol production from xylose as a sole carbon source remain uncertain. We showed that overexpression of mutant NADH-dependent ADH1 derived from TMB3000 strain in the recombinant Saccharomyces cerevisiae, into which xylose reductase (XR) and xylitol dehydrogenase (XDH) pathway of Pichia stipitis has been introduced, improved ethanol production from xylose as a sole carbon source in the presence of 5-HMF. Enhanced furan-reducing activity is able to regenerate NAD+ to relieve redox imbalance, resulting in increased ethanol yield arising from decreased xylitol accumulation. In addition, we found that overexpression of wild-type ADH1 prevented the more severe inhibitory effects of furfural in xylose fermentation as well as overexpression of TMB3000-derived mutant. After 120 h of fermentation, the recombinant strains overexpressing wild-type and mutant ADH1 completely consumed 50 g/L xylose in the presence of 40 mM furfural and most efficiently produced ethanol (15.70 g/L and 15.24 g/L) when compared with any other test conditions. This is the first report describing the improvement of ethanol production from xylose as the sole carbon source in the presence of furan derivatives with xylose-utilizing recombinant yeast strains via the overexpression of ADH-related genes.  相似文献   

12.
The hydrolytic activity of a recombinant β-glycosidase from Dictyoglomus turgidum that specifically hydrolyzed the xylose at the C-6 position and the glucose in protopanaxatriol (PPT)-type ginsenosides followed the order Rf > Rg1 > Re > R1 > Rh1 > R2. The production of aglycone protopanaxatriol (APPT) from ginsenoside Rf was optimal at pH 6.0, 80 °C, 1 mg ml?1 Rf, and 10.6 U ml?1 enzyme. Under these conditions, D. turgidum β-glycosidase converted ginsenoside R1 to APPT with a molar conversion yield of 75.6 % and a productivity of 15 mg l?1 h?1 after 24 h by the transformation pathway of R1 → R2 → Rh1 → APPT, whereas the complete conversion of ginsenosides Rf and Rg1 to APPT was achieved with a productivity of 1,515 mg l?1 h?1 after 6.6 h by the pathways of Rf → Rh1 → APPT and Rg1 → Rh1 → APPT, respectively. In addition, D. turgidum β-glycosidase produced 0.54 mg ml?1 APPT from 2.29 mg ml?1 PPT-type ginsenosides of Panax ginseng root extract after 24 h, with a molar conversion yield of 43.2 % and a productivity of 23 mg l?1 h?1, and 0.62 mg ml?1 APPT from 1.35 mg ml?1 PPT-type ginsenosides of Panax notoginseng root extract after 20 h, with a molar conversion yield of 81.2 % and a productivity of 31 mg l?1 h?1. This is the first report on the APPT production from ginseng root extract. Moreover, the concentrations, yields, and productivities of APPT achieved in the present study are the highest reported to date.  相似文献   

13.
Paenibacillus curdlanolyticus B-6 Xyn10C is a single module xylanase consisting of a glycoside hydrolase family-10 catalytic module. The recombinant enzyme, rXyn10C, was produced by Escherichia coli and characterized. rXyn10C was highly active toward soluble xylans derived from rye, birchwood, and oat spelt, and slightly active toward insoluble wheat arabinoxylan. It hydrolyzed xylooligosaccharides larger than xylotetraose to produce xylotriose, xylobiose, and xylose. When rye arabinoxylan and oat spelt xylan were treated with the enzyme and the hydrolysis products were analyzed by thin layer chromatography (TLC), two unknown hydrolysis products, U1 and U2, were detected in the upper position of xylose on a TLC plate. Electrospray ionization mass spectrometry and enzymatic analysis using Bacillus licheniformis α-l-arabinofuranosidase Axh43A indicated that U1 was α-l-Araf-(1  2)-[α-l-Araf-(1  3)]-d-Xylp and U2 was α-l-Araf-(1  2)-d-Xylp, suggesting that rXyn10C had strong activity toward a xylosidic linkage before and after a doubly arabinose-substituted xylose residue and was able to accommodate an α-1,2- and α-1,3-linked arabinose-substituted xylose unit in both the −1 and +1 subsites. A molecular docking study suggested that rXyn10C could accommodate a doubly arabinose-substituted xylose residue in its catalytic site, at subsite −1. This is the first report of a xylanase capable of producing α-l-Araf-(1  2)-[α-l-Araf-(1  3)]-d-Xylp from highly arabinosylated xylan.  相似文献   

14.
Using a set of methods (C-banding, DAPI-staining, fluorescence hybridization in situ (FISH) with probes of 26S and 5S rDNA, and analysis of meiosis), the first comparative cytogenetic study of three species of Macleaya, producers of complex isoquinoline alkaloids, cordate Macleaya cordata (Willd.) R. Br. (2n = 20), small-fruited Macleaya microcarpa (Maxim.) Fedde (2n = 20) and Macleaya kewensis Turrill (2n = 20), was first carried out. On the basis of morphometric analysis, formulas of karyotypes were made for each species. Species ideograms for M. cordata, M. microcarpa, and M. kewensis were constructed taking into account the polymorphic variants of the C-banding patterns and indicating the location of 26S and 5S rDNA sites. A comparative study revealed that the karyotypes of M. microcarpa and M. kewensis have more in common with each other than with M. cordata. Analysis of meiotic chromosomes suggests of genetic stability of Macleaya genomes. The results of chromosome analysis were used to confirm the close relationship of Macleaya and to clarify their phylogenetic relationships.  相似文献   

15.
A thermosensitive uracil requiring mutant of Bacillus subtilis Marburg 168 thy trp2 ts42 was examined as to the colony forming ability at the permissive and nonpermissive temperatures. The viability of the mutant cells decreased rapidly at the restrictive temperature in the modified Woese’s (MW) medium. However, the cells retained viability when sodium succinate or potassium chloride was added to the medium at that temperature although uracil deficiency was unchanged. A little but significant incorporation of adenine-8-14C into RNA still continued even after the incorporation of N-acetyl-3H-d-glucosamine into acid insoluble fraction of the cells terminated in the MW medium at 48°C. Both incorporations as well as increase of absorbance were slowed down in the presence of sodium succinate at 48°C. This mutant, ts42, was more sensitive to deoxycholate (DOC) than the parent strain. The restoration of colony forming ability after the temperature shift back from 48 to 37°C was suppressed by the addition of DOC to the medium. However, the cell became resistant to DOC when uracil was added to the medium prior to the temperature shift.  相似文献   

16.
The -amylase of Micromonospora melanosporea was produced extracellularly during batch fermentation in a 5.0-1 fermentor. The absence of an organic nitrogen source in its growth medium facilitated subsequent purification of the enzyme by ammonium sulphate fractionation and two consecutive Superose-12 gel-filtration steps. The enzyme exhibited maxima for activity at pH 7.0 and 55° C and was 72% stable at pH 6.0–12.0 for 30 min at 40° C. It had a relative molecular mass of 45 000 and an isoelectric point at pH 7.6. The enzyme catalyses the conversion of starch to maltose (53%, w/w) as the predominant final end-product. Initial hydrolysis of this substrate, however, gave rise to the formation of maltooligosaccharides in the range maltotriose to maltohexaose. Maximum yields of these intermediate sugars accumulated to between 31 and 42% (w/w) as the reaction proceeded. The action of the M. melanosporea amylase on high concentrations of saccharides larger than maltotriose resulted in the formation of mainly maltose and maltotriose without concomitant glucose production. A combination of hydrolytic and transfer events is postulated to be responsible for this phenomenon and for the high maltose levels achieved. Correspondence to: C. T. Kelly  相似文献   

17.
The Gō-like models of proteins are constructed based on the knowledge of the native conformation. However, there are many possible choices of a Hamiltonian for which the ground state coincides with the native state. Here, we propose to use experimental data on protein stretching to determine what choices are most adequate physically. This criterion is motivated by the fact that stretching processes usually start with the native structure, in the vicinity of which the Gō-like models should work the best. Our selection procedure is applied to 62 different versions of the Gō model and is based on 28 proteins. We consider different potentials, contact maps, local stiffness energies, and energy scales—uniform and nonuniform. In the latter case, the strength of the nonuniformity was governed either by specificity or by properties related to positioning of the side groups. Among them is the simplest variant: uniform couplings with no i, i + 2 contacts. This choice also leads to good folding properties in most cases. We elucidate relationship between the local stiffness described by a potential which involves local chirality and the one which involves dihedral and bond angles. The latter stiffness improves folding but there is little difference between them when it comes to stretching.  相似文献   

18.
《Phytochemistry》1987,26(5):1299-1300
The effect ofpH on Km and Vmax values of coconut α-galactosidase indicates the involvement of two ionizing groups with pKa values of 3.5 and 6.5 in catalysis. Chemical modification has indicated the presence of two carboxyl groups, a tryptophan and a tyrosine, at or near the active site of α-galactosidase. Based on these facts a new mechanism of action for α-galactosidase is proposed in which the ionizing group with a pKa of 3.5 is a carboxyl group involved in stabilizing a carbonium ion intermediate and the ionizing group with a pKa of 6.5 is a carboxyl group perturbed due to the presence of a hydrophobic residues in its vicinity which donates a H+ ion in catalysis.  相似文献   

19.
20.
The complementary fragments of human Hb α, α1–30, and α31–141 are spliced together by V8 protease in the presence of 30%n-propanol to generate the full-length molecule (Hb α-semisynthetic reaction). Unlike the other protease-catalyzed protein/peptide splicing reactions of fragment complementing systems, the enzymic condensation of nonassociating segments of Hb α is facilitated by the organic cosolvent induced α-helical conformation of product acting as the “molecular trap” of the splicing reaction. The segments α24–30 and α31–40 are the shortest complementary segments that can be spliced by V8 protease. In the present study, the chemistry of the contiguous segment (product) α24–40 has been manipulated by engineering the amino acid replacements to the positions α27 and α31 to delineate the structural basis of the molecular trap. The location of Glu27 and Arg31 residues in the contiguous segment α24–40 (as well as in other larger segments) is ideal to generate (i, i+4) side-chain carboxylate-guanidino interaction in its α-helical conformation. The amino acid residue replacement studies have confirmed that the side chains at α27 and α31 facilitate the semisynthetic reaction. The relative influence of the substitute at these sites on the splicing reaction depends on the chemical nature of the side chain and the location. The γ-carboxylate guanidino side-chain interaction appears to contribute up to a maximum of 85% of the thermodynamic stability of the molecular trap. The studies also demonstrate that the thermodynamic stability of the molecular trap is determined by two interdependent conformational aspects of the peptide. One is an amino acid-sequence-specific event that facilitates the induction of an α-helical conformation to the contiguous segment in the presence of organic cosolvent that imparts some amount of protease resistance to Glu30-Arg31 peptide bond. The second structural aspect is a site-specific event, ani, i+4 side-chain interaction in the α-helical conformation of the peptide which imparts an additional thermodynamic stability to the molecular trap. The results suggest that conformationally driven “molecular traps” of protease-mediated ligation reactions of peptides could be designed into products to facilitate the modular assembly of peptides/proteins.  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