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 26 毫秒
1.
Semicontinuous fermentation using pellets of Rhizopus oryzae has been recognized as a promising technology for l-lactic acid production. In this work, semicontinuous fermentation of R. oryzae AS 3.819 for l-lactic acid production has been developed with high l-lactic acid yield and volumetric productivity. The effects of factors such as inoculations, CaCO3 addition time, and temperature on l-lactic acid yield and R. oryzae morphology were researched in detail. The results showed that optimal fermentation conditions for the first cycle were: inoculation with 4% spore suspension, CaCO3 added to the culture medium at the beginning of culture, and culture temperature of 32–34°C. In orthogonal experiments, high l-lactic acid yield was achieved when the feeding medium was (g/l): glucose, 100; (NH4)2SO4, 2; KH2PO4, 0.1; ZnSO4·7H2O, 0.33; MgSO4·7H2O, 0.15; CaCO3, 50. Twenty cycles of semicontinuous fermentation were carried out in flask culture. l-lactic acid yield was 78.75% for the first cycle and 80–90% for the repeated cycles; the activities of lactate dehydrogenases (LDH) were 7.2–9.2 U/mg; fermentation was completed in 24 h for each repeated cycle. In a 7-l magnetically stirred fermentor, semicontinuous fermentation lasted for 25 cycles using pellets of R. oryzae AS 3.819 under the optimal conditions determined from flask cultures. The final l-lactic acid concentration (LLAC) reached 103.7 g/l, and the volumetric productivity was 2.16 g/(l·h) for the first cycle; in the following 19 repeated cycles, the final LLAC reached 81–95 g/l, and the volumetric productivities were 3.40–3.85 g/(l·h).  相似文献   

2.
In addition to the general amino acid transport system (GAP) ofS. cerevisiae l-tryptophan is transported by another system with approximately 25% capacity of GAP, with aK T of 0.41±0.08 mmol/L and with a similar specificity as GAP (lower inhibition by Met, Pro, Ser, Thr and 2-aminoisobutyric acid; greater inhibition by Glu and His). The pH optimum of this system is at 5.0–5.5, activation energy above the transition point (20°C) was 20 kJ/mol, below the transition point 55 kJ/mol. The transport by this system was virtually unidirectional, efflux amounting to at most 10% into a tryptophan-free medium. The transport itself was blocked by 2,4-dinitrophenol, antimycin A and uranyl nitrate. The system was synthesized de novo during preincubation with glucose=fructose>trehalose >ethanol within 30 min, and was degraded with a half-time of 15 min in the absence of further synthesis. The accumulation ratios ofl-tryptophan ingap1 mutants were concentration-dependent (200∶1 at 1 μmoll-Trp/L, 4∶1 at 2.5 mmoll-Trp/L) and decreased with increasing suspension density from 200∶1 to 5∶1 (for 10 μmoll-Trp/L). The involvement of hydrogen ions in the uptake was clearly demonstrated by the effect of D2O even if it could not be established by either shifts of pHout or membrane depolarization.  相似文献   

3.
The uptake of 3H-labeled choline by a suspension of isolated type II epithelial cells from rat lung has been studied in a Ringer medium. Uptake was linear for 4 min at both 0.1 μm and 5.0 μm medium choline; at 5 μm, only 10% of the label was recovered in a lipid fraction. Further experiments were conducted at the low concentration (0.1 μm), permitting characterization of the properties of high-affinity systems. Three fractions of choline uptake were detected: (i) a sodium-dependent system that was totally inhibited by hemicholinium-3 (HC-3); (ii) a sodium-independent uptake, when Na+ was replaced by Li+, K+ or Mg2+, inhibited by HC-3; (iii) a residual portion persisting in the absence of Na+ and unaffected by HC-3. Choline uptake was sigmoidally related to the medium Na+ concentration. Kinetic properties of the uptake of 0.1 μm 3H-choline in the presence and absence of medium Na+ were examined in two ways. (a) Inhibition by increasing concentrations of unlabeled choline (0.5–100 μm) was consistent with the presence of two Michaelis-Menten-type systems in the presence of Na+; a Na+-dependent portion (a mean of 0.52 of the total) had a K m for choline of 1.5 μm while K m in the absence of Na+ (Li+ substituting) was 18.6 μm. (b) Inhibition by HC-3 (0.3–300 μm) gave Ki values of 1.7 μm and 5.0 μm HC-3 for the Na+-dependent and -independent fractions. The apparent K m of the Na+-dependent uptake is lower than that reported previously for lung-derived cells and is in the range of the K m values reported for high-affinity, Na+-dependent choline uptake by neuronal cells. Received: 18 February 1997/Revised: 7 December 1997  相似文献   

4.
The biocatalytic ability of transgenic crown galls of Panax quinquefolium was evaluated by using eugenol (1) as a substrate and suspension cultures of Nicotiana tabacum as control system. Three biotransformed products, namely: 2-methoxy-4-(2-propenyl)phenyl-O-β-d-glucopyranoside (2, 67.11%), 2-methoxy-4-(2-propenyl)phenyl-O-β-d-glucopyranosyl (6′ → 1″)-β-d-xylopyranoside (3, 2.85%) and methyl eugenol (4, 14.30%) were obtained after 5 days of administration of eugenol to the suspension cultures of transgenic crown galls of P. quinquefolium. In contrast, only one product, compound 2 (15.41%), was obtained in suspension cultures of N. tabacum after 5 days of incubation. The results indicated that the glycosylation ability of transgenic crown galls of P. quinquefolium was much higher than that of the cultured cells of N. tabacum.  相似文献   

5.
Biosynthesis of six saponins (ginsenosides) in suspension culture of P. quinquefolium Z5 was investigated. Ginsenoside content in biomass reached the highest level, nearly 30 mg g−1 d.w., between 25 and 30 days of the culture. Saponins were synthesized simultaneously with cell growth but their synthesis rate was not proportional to the growth rate. During the phase of rapid biomass multiplication, after which biomass reached 90% of its maximum yield, only half examined ginsenosides was produced. The second half of the final saponins yield was produced during the slow growth phase, in which only 10% of biomass was grown. During the intensive growth phase the productivity of six saponins examined per biomass (dry weight) unit was 3.4 μg mg−1 d.w. day−1, however, this parameter calculated for slow growth phase reached nearly 30 μg mg−1 d.w. day−1. There were differences in increase of the contents of six saponins determined in biomass, and it was the highest for saponins Re (20(S)-protopanaxatriol-6-[O-α-l-rhamnopyranosyl(1 → 2)-β-d-glucopyranoside]-20-O-β-d-glucopyranoside) and Rg1 (20(S)-protopanaxatriol-6,20-di-O-β-d-glucoside).  相似文献   

6.
λ-Glutamylcysteine synthetase activity (EC 6.3.2.2) was analysed in Sephacryl S-200 eluents of extracts from cell suspension cultures ofNicotiana tabacum L. cv. Samsun by determination of λ-glutamylcysteine as its monobromobimane derivative. The enzyme has a relative molecular mass (Mr) of 60000 and exhibits maximal activity at pH 8 (50% at pH 7.0 and pH9.0) and an absolute requirement for Mg2+. With 0.2mM Cd2+ or Zn2+, enzyme activity was reduced by 35% and 19%, respectively. Treatment with 5 mM dithioerythritol led to a heavy loss of activity and to dissociation into subunits (Mr 34000). Buthionine sulfoximine andl-methionine-sulfoximine, known as potent inhibitors of λ-glutamylcysteine synthetase from mammalian cells, were found to be effective inhibitors of the plant enzyme too. The apparent Km values forl-glutamate,l-cysteine, and α-aminobutyrate were, respectively, 10.4mM, 0.19 mM, and 6.36 mM. The enzyme was completely inhibited by glutathione (Ki=0.42 mM). The data indicate that the rate of glutathione synthesis in vivo may be influenced substantially by the concentration of cysteine and glutamate and may be further regulated by feedback inhibition of λ-glutamylcysteine synthetase by glutathione itself. λ-Glutamylcysteine synthetase is, like glutathione synthetase, localized in chloroplasts as well as in the cytoplasm. Chloroplasts fromPisum sativum L. isolated on a Percoll gradient contained about 72% of the λ-glutamylcysteine synthetase activity in leaf cells and 48% of the total glutathione synthetase activity. In chloroplasts ofSpinacia oleracea L. about 61% of the total λ-glutamylcysteine synthetase activity of the cells were found and 58% of the total glutathione synthetase activity. These results indicate that glutathione synthesis can take place in at least two compartments of the plant cell. Dedicated to Professor A. Prison on the occasion of his 80th birthday  相似文献   

7.
To assess the activation of the charybdotoxin-insensitive K+ channel responsible for Regulatory Volume Decrease (RVD) after substantial volume increases, we measured intracellular pH (pH i ), intracellular calcium ([Ca2+] i ) and inhibitors of kinases and phosphoprotein phosphatases in guinea pig jejunal villus enterocytes in response to volume changes. Fluorescence spectroscopy was used to measure pH i and [Ca2+] i of cells in suspension, loaded with 2,7,bis-carboxyethyl-5-6-carboxyfluorescein and Indo-1, respectively, and cell volume was assessed using electronic cell sizing. A modest 7% volume increase or substantial 15 to 20% volume increase caused [Ca2+] i to increase proportionately but the 7% increase caused alkalinization while the larger increases resulted in acidification of ≃0.14 pH units. Following a 15% volume increase, 1-N-0-bis (5-isoquinoline-sulfonyl)-N-methyl-l-4-phenyl-piperazine (KN-62, 50 μm), an inhibitor of Ca2+/calmodulin kinase II, blocked RVD. Gramicidin (0.5 μm) bypassed this inhibition suggesting that the K+ channel had been affected by the KN-62. RVD after a modest 7% volume increase was not influenced by KN-62 unless the cell was acidified. Okadaic acid, an inhibitor of phosphoprotein phosphatases 1 and 2A, accelerated RVD after a 20% volume increase; inhibition of RVD generated by increasing the K+ gradient was bypassed by okadaic acid. Tyrosine kinase inhibitor, genistein (100 μm) had no effect on RVD after 20% volume increases. We conclude that activation of charybdotoxin-insensitive K+ channels utilized for RVD after substantial (>7%) `nonphysiological' volume increases requires phosphorylation mediated by Ca2+/calmodulin kinase II and that increases in cytosolic acidification rather than larger increases in [Ca2+] i are a critical determinant of this activation. Received: 30 March 1999/Revised: 6 July 1999  相似文献   

8.
The physiologic effect of gibberellins (GA) in seed development is poorly understood. We examined the effect of gibberellic acid (GA3) on growth, protein secretion, and starch accumulation in cultured maize (Zea mays L.) endosperm suspension cells. GA3 (5 and 30 μm) increased the fresh weight, dry weight, and protein content of the cultured cells, but the effect of GA3 at 50 μm was not significantly different. However, the protein content in the culture medium was increased by these three concentrations of GA3. The effect of GA3 on the amount of cellular structural polysaccharides was not significant, but GA3 had a dramatic effect on the starch content. At 5 μm, GA3 caused an increase in the starch content, but at 50 μm the starch accumulation was reduced. Chlorocholine chloride (CCC), an inhibitor of GA biosynthesis, significantly increased the starch content and decreased the structural polysaccharide content of the cultured cells. The effects of CCC at 500 μm on the starch and polysaccharide content were partially reversed by 5 μm GA3 applied exogenously. Based on these results we suggest that GA does not favor starch accumulation in the cell cultures and that the addition of lower concentrations of GA3 in the medium may provide an improved balance among the endogenous GA in the cultured cells. Received October 31, 1995; accepted March 25, 1997  相似文献   

9.
The objective of this study was to purify and characterize a mouse hepatic enzyme that directly generates CH3SeH from seleno-l-methionine (l-SeMet) by the α,γ-elimination reaction. The l-SeMet α,γ-elimination enzyme was ubiquitous in tissues from ICR mice and the activity was relatively high in the large intestine, brain, and muscle, as well as the liver. Aging and sex of the mice did not have any significant influence on the activity in the liver. The enzyme was purified from the mouse liver by ammonium sulfate precipitation and four kinds of column chromatography. These procedures yielded a homogeneous enzyme, which was purified approx 1000-fold relative to mouse liver extract. Overall recovery was approx 8%. The purified enzyme had a molecular mass of approx 160 kDa with four identical subunits. The K m value of the enzyme for the catalysis of l-SeMet was 15.5 m M, and the V max was 0.29 units/mg protein. Pyridoxal 5′-phosphate (pyridoxal-P) was required as a cofactor because the holoenzyme could be resolved to the apoenzyme by incubation with hydroxylamine and reconstituted by addition of pyridoxal-P. The enzyme showed the optimum activity at around pH 8.0 and the highest activity at 50°C; it catalyzed the α,γ-elimination reactions of several analogs such as d,l-homocysteine and l-homoserine in addition to l-SeMet. This enzyme also catalyzed the α,β-elimination reaction of Se-methylseleno-l-cysteine. However, l-methionine was inerts. Therefore, the purified enzyme was different from the bacterial l-methionine γ-lyase that metabolizes l-SeMet to CH3SeH, in terms of the substrate specificity. These results were the first identification of a mammalian enzyme that specifically catalyzes the α,γ-elimination reaction of l-SeMet and immediately converts it to CH3SeH, an important metabolite of Se.  相似文献   

10.
A cell suspension culture, prepared fromPerilla frutescens var.crispa callus induced by Murashige and Skoog (1962) medium containing 2,4-dichlorophenoxyacetic acid (2,4-D, 1.0 ml/l) and kinetin (0.1 mg/l), contained caffeic acid derivatives as the phenolic components. Fresh and dry weights of the cells increased exponentially for about 11 days after transfer to a fresh medium. The contents of caffeic acid and protein also reached a maximum on the 11th day, but α-amino nitrogen phenylalanine and tyrosine continued to increase in amount until the 20th to 23rd day. Caffeic acid formation in the cells was increased by lowering the concentration of 2,4-D. The administration ofl-2-aminooxy-3-phenylpropionic acid (l-AOPP), 2-aminooxyacetic acid (AOA) andN-(phosphonomethyl)glycine (glyphosate) to the cells inhibited caffeic acid formation to a large extent. An 80% inhibition of caffeic acid formation was caused by 10−4Ml-AOPP whereas phenylalanine and tyrosine contents of the cells became 7.5 and 2.3 times higher at thisl-AOPP concentration than those in the control. An 85% inhibition of caffeic acid formation was achieved at 10−3M glyphosate concentration, while 10−3M AOA inhibited caffeic acid formation by 95% and also growth rate by 80%. The influence of inhibitors on caffeic acid formation is discussed in relation to the level of α-amino nitrogen, particularly aromatic amino acids, in the cell suspension cultures.  相似文献   

11.
d-Aspartate (d-Asp) uptake by suspensions of cerebral rat brain astrocytes (RBA) maintained in long-term culture was studied as a means of characterizing function and regulation of Glutamate/Aspartate (Glu/Asp) transporter isoforms in the cells. d-Asp influx is Na+-dependent with K m = 5 μm and V max= 0.7 nmoles · min−1· mg protein−1. Influx is sigmoidal as f[Na+] with Na+ K m ∼ 12 μm and Hill coefficient of 1.9. The cells establish steady-state d-Asp gradients >3,000-fold. Phorbol ester (PMA) enhances uptake, and gradients near 6,000-fold are achieved due to a 2-fold increase in V max, with no change in K m . At initial [d-Asp] = 10 μm, RBA take up more than 90% of total d-Asp, and extracellular levels are reduced to levels below 1 μm. Ionophores that dissipate the ΔμNa+ inhibit gradient formation. Genistein (GEN, 100 μm), a PTK inhibitor, causes a 40% decrease in d-Asp. Inactive analogs of PMA (4α-PMA) and GEN (daidzein) have no detectable effect, although the stimulatory PMA response still occurs when GEN is present. Further specificity of action is indicated by the fact that PMA has no effect on Na+-coupled ALA uptake, but GEN is stimulatory. d-Asp uptake is strongly inhibited by serine-O-sulfate (S-O-S), threohydroxy-aspartate (THA), l-Asp, and l-Glu, but not by d-Glu, kainic acid (KA), or dihydrokainate (DHK), an inhibition pattern characteristic of GLAST and EAAC1 transporter isoforms. mRNA for both isoforms was detected by RT-PCR, and Western blotting with appropriate antibodies shows that both proteins are expressed in these cells. Received: 11 January 2001/Revised: 26 March 2001  相似文献   

12.
Summary RPMI 1788 lymphocytes growtn in semi-suspension culture proliferate as separate cells and in clumps. The addition of 10−3 m dibutyryl cyclic AMP (Bu2cAMP) to the culture resulted in the attachment of the cells to the substratum and a subsequent conversion of a portion of the adherent cells to a fibroblast-like morphology. Growth of the adherent cells proceeded at nearly the same rate as that of the control cells. When the cells cultured in the presence of Bu2cAMP were periodically disturbed, they remained in suspension and under this condition a distinct inhibition of growth by Bu2cAMP was observed. Cortisol at 10−5 m, a concentration having no effect on the proliferation of RPMI 1788 cells, when added to cells cultured in the presence of Bu2cAMP, prevented cell attachment, caused detachment of already adherent cells and thereby brought about the Bu2cAMP-mediated inhibition of growth in suspension. At a higher concentration (10−4 m), cortisol alone reduced the growth rate of RPMI 1788 lymphocytes. Under the combined effects of 10−4 m cortisol and 10−3 m Bu2cAMP, the proliferation and viability of cells in suspension were significantly lower than in the presence of either agent alone.  相似文献   

13.
Microorganisms capable of cleaving the urethane bond of t-butoxycarbonyl (Boc) amino acids in a whole-cell reaction were screened among stock cultures, and Corynebacterium aquaticum IFO12154 was the most promising. The conversion of Boc-Ala to Ala was stimulated by CoSO4 in the medium and reaction mixture. The optimum whole-cell concentration was 25 mg lyophilized cells/ml. Boc-l-Met was the best substrate for this reaction, and other Boc-L-amino acids, as well as benzyloxycarbonyl-l-amino acids with hydrophobic residues, were also good substrates. Boc-d- and Z-d-amino acids were inert. When the reactions had proceeded for 24 h with each substrate at 10 mM, the molar conversion rates from Boc-l-, dl- and d-Met were 100%, 50%, and 0% respectively. From 150 mM Boc-l-Met, 143 mM l-Met was formed at a molar yield of 95.3%. Received: 3 September 1996 / Received last revision: 7 April 1997 / Accepted: 19 April 1997  相似文献   

14.
To assess whether alterations in membrane fluidity of neonatal rat heart cells modulate gap junctional conductance (g j ), we compared the effects of 2mm 1-heptanol and 20 μm 2-(methoxyethoxy)ethyl 8-(cis-2-n-octylcyclopropyl)-octanoate (A2C) in a combined fluorescence anisotropy and electrophysiological study. Both substances decreased fluorescence steady-state anisotropy (rss), as assessed with the fluorescent probe 1-(4-trimethylammoniumphenyl)-6-phenyl-1,3,5-hexatriene (TMA-DPH) by 9.6±1.1% (mean ±sem,n=5) and 9.8±0.6% (n=5), respectively, i.e., both substances increased bulk membrane fluidity. Double whole-cell voltage-clamp experiments showed that 2mm heptanol uncoupled cell pairs completely (n=6), whereas 20 μm A2C, which increased bulk membrane fluidity to the same extent, did not affect coupling at all (n=5). Since gap junction channels are embedded in relatively cholesterol-rich domains of the membrane, we specifically assessed the fluidity of the cholesterol-rich domains with dehydroergosterol (DHE). Using DHE, heptanol increased rss by 14.9±3.0% (n=5), i.e., decreased cholesterol domain fluidity, whereas A2C had no effect on rss (−0.4±6.7%,n=5). Following an increase of cellular “cholesterol” content (by loading the cells with DHE), 2mm heptanol did not uncouple cell pairs completely:g j decreased by 80±20% (range 41–95%,n=5). The decrease ing j was most probably due to a decrease in the open probability of the gap junction channels, because the unitary conductances of the channels were not changed nor was the number of channels comprising the gap junction. The sensitivity of non-junctional membrane channels to heptanol was unaltered in cholesterol-enriched myocytes. These results indicate that the fluidity of cholesterol-rich domains is of importance to gap junctional coupling, and that heptanol decreasesg j by decreasing the fluidity of cholesterol-rich domains, rather than by increasing the bulk membrane fluidity.  相似文献   

15.
Zeng X  Sun Y  Ye H  Liu J  Uzawa H 《Biotechnology letters》2007,29(7):1105-1110
When α-d-GlcNAc-OC6H4NO2 -p and β-d-(6-sulfo)-GlcNAc-OC6H4NO2-p (2) were used as substrates, β-N-acetylhexosaminidase from Aspergillus oryzae transferred the β-d-(6-sulfo)-GlcNAc(unit from 2 to α-d-GlcNAc-OC6H4NO2 -p to afford β-d-(6-sulfo)-GlcNAc-(1→4)-α-d-GlcNAc-OC6H4NO2-p (3) in a yield of 94% based on the amount of donor, 2, added. β-d-(6-sulfo)-GlcNAc-(1→4)-α-d-Glc-OC6H4NO2-p (4) was obtained with α-d-Glc-OC6H4NO2 -p as acceptor in a similar manner. With a reaction mixture of 2 and β-d-GlcNAc-OC6H4NO2-p (1) in a molar ratio of 6:1, the enzyme mediated the transfer of β-d-GlcNAc from 1 to 2, affording disaccharide β-d-GlcNAc-(1→4)-β-(6-sulfo)-d-GlcNAc-OC6H4NO2-p (5) in a yield of 13% based on the amount of 1 added.  相似文献   

16.
α-l-Rhamnosidase was extracted and purified from the cells of Pseudomonas paucimobilis FP2001 with a 19.5% yield. The purified enzyme, which was homogeneous as shown by SDS-PAGE and isoelectric focusing, had a molecular weight of 112,000 and an isoelectric point of 7.1. The enzyme activity was accelerated by Ca2+ and remained stable for several months when stored at –20 °C. The optimum pH was 7.8; the optimum temperature was 45 °C. The K m, V max and k cat for p-nitrophenyl α-l-rhamnopyranoside were 1.18 mM, 92.4 μM · min–1 and 117,000 · min–1, respectively. Examination of the substrate specificity using various synthetic and natural l-rhamnosyl glycosides showed that this enzyme had a relatively broader substrate specificity than those reported so far. Received: 24 May 1999 / Accepted: 7 October 1999  相似文献   

17.
The thermophilic fungus Humicola sp constitutively produces intracellular α-galactosidase (1.33 U mg−1 protein) within 48 h at 45°C in shaken flasks, when grown in a medium containing 7% wheat bran extract as a carbon source and 0.5% yeast extract as a nitrogen source. The enzyme has been purified to homogeneity by ultrafiltration, ethanol precipitation, DEAE cellulose and Sephacryl S-300 chromatography with a 124-fold increase in specific activity and 29.5% recovery. The molecular weight of the enzyme is 371.5 kDa by gel filtration on Sephacryl S-300 and 87.1 kDa by SDS-polyacrylamide gel electrophoresis. The enzyme has an optimum temperature of 65°C and an optimum pH of 5.0. Humicola α-galactosidase is a glycoprotein with 8.3% carbohydrate content and is acidic in nature with a pI of 4.0. The K m S for p-nitrophenyl-α-D-galactopyranoside, O-nitrophenyl-α-D-galactopyranoside, raffinose and stachyose are 0.279, 0.40, 1.45 and 1.42 mM respectively. The enzyme activity was strongly inhibited by Ag+ and Hg2+. D-Galactose inhibited α-galactosidase competitively and the inhibition constant (K i) for galactose was 11 mM. Received 28 January 1999/ Accepted in revised form 07 April 1999  相似文献   

18.
Transport Pathways for Therapeutic Concentrations of Lithium in Rat Liver   总被引:1,自引:0,他引:1  
Although both amiloride- and phloretin-sensitive Na+/Li+ exchange activities have been reported in mammalian red blood cells, it is still unclear whether or not the two are mediated by the same pathway. Also, little is known about the relative contribution of these transport mechanisms to the entry of therapeutic concentrations of Li+ (0.2–2 mm) into cells other than erythrocytes. Here, we describe characteristics of these transport systems in rat isolated hepatocytes in suspension. Uptake of Li+ by hepatocytes, preloaded with Na+ and incubated in the presence of ouabain and bumetanide, comprised three components. (a) An amiloride-sensitive component, with apparent K m 1.2 mm Li+, V max 40 μmol · (kg dry wt · min)−1, showed increased activity at low intracellular pH. The relationship of this component to the concentration of intracellular H+ was curvilinear suggesting a modifier role of [H+] i . This system persisted in Na+-depleted cells, although with apparent K m 3.8 mm. (b) A phloretin-sensitive component, with K m 1.2 mm, V max 21 μmol · (kg · min)−1, was unaffected by pH but was inactive in Na+-depleted cells. Phloretin inhibited Li+ uptake and Na+ efflux in parallel. (c) A residual uptake increased linearly with the external Li+ concentration and represented an increasing proportion of the total uptake. The results strongly suggest that the amiloride-sensitive and the phloretin-sensitive Li+ uptake in rat liver are mediated by two separate pathways which can be distinguished by their sensitivity to inhibitors and intracellular [H+]. Received: 8 April 1999/Revised: 19 July 1999  相似文献   

19.
An extracellular enzyme with glucose dehydrogenase activity was purified from liquid cultures of the basidiomycete Agaricus bisporus after growth with d-cellobiose or d-glucose as carbon source. The molecular mass was measured as 57 kDa by gel filtration and 55 kDa by sodiumdodecyl sulphate/polyacrylamide gel electrophoresis, while the isoelectric point was at pH 3.6. By analysis of 1H-NMR spectra in D2O, the product of d-glucose oxidation was identified as 3-ketoglucose. The substrates oxidized included d-cellobiose, l-arabinose, d-xylose and sucrose, but the specificity parameter (k cat/K m) was highest for d-glucose. Two electron acceptors were identified, namely 2,6-dichloroindophenol and p-benzoquinone, but reduction of dioxygen, ferricyanide or cytochrome c was not detectable. The selective C-3 oxidation of d-glucose is well-characterized for Agrobacterium and Flavobacterium, but this is the first report for a fungus. Received: 19 June 1998 / Received revision: 15 September 1998 / Accepted: 17 September 1998  相似文献   

20.
Surwase SN  Jadhav JP 《Amino acids》2011,41(2):495-506
l-DOPA is an amino acid derivative and most potent drug used against Parkinson’s disease, generally obtained from Mucuna pruriens seeds. In present communication, we have studied the in vitro production of l-DOPA from l-tyrosine by novel bacterium Bacillus sp. JPJ. This bacterium produced 99.4% of l-DOPA from l-tyrosine in buffer (pH 8) containing 1 mg ml−1 cell mass incubated at 40°C for 60 min. The combination of CuSO4 and l-ascorbic acid showed the inducing effect at concentrations of 0.06 and 0.04 mg ml−1, respectively. The activated charcoal 2 mg ml−1 was essential for maximum bioconversion of l-tyrosine to l-DOPA and the crude tyrosinase activity was 2.7 U mg−1 of tyrosinase. Kinetic studies showed significant values of Y p/s (0.994), Q s (0.500) and q s (0.994) after optimization of the process. The production of l-DOPA was confirmed by analytical techniques such as HPTLC, HPLC and GC–MS. This is the first report on rapid and efficient production of l-DOPA from l-tyrosine by bacterial source which is more effective than the plant, fungal and yeast systems.  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