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.
A vector system has been developed to express isoenzyme A1 of sweet potato peroxidase (POD) and was introduced into Saccharomyces cerevisiae. The system contains the signal sequence of Aspergillus oryzae -amylase to facilitate the extracellular secretion of peroxidase under the control of constitutive glyceraldehyde-3-phosphate dehydrogenase (GPD) promoter. In a batch culture using YNBDCA medium (yeast nitrogen base without amino acids 6.7 g l–1, Casamino acids 5 g l–1 and glucose 20 g l–1), the recombinant strain expressed the swpa1 gene giving a secretion yield of POD activity of ca. 90% of total expressed peroxidase. Supplementation with PMSF (0.05 mM) and Casamino acids (5 g/50 ml) increased extracellular POD activity to nearly 10 kU ml–1, equivalent to 1.5 kU g–1 cell dry wt. This is 9 fold higher than that obtained in medium without PMSF. From SDS-PAGE and native-PAGE analyses POD has an M r of 53 kDa.  相似文献   

2.
Glucose repressed xylose utilization inCandida tropicalis pre-grown on xylose until glucose reached approximately 0–5 g l–1. In fermentations consisting of xylose (93 g l–1) and glucose (47 g l–1), xylitol was produced with a yield of 0.65 g g–1 and a specific rate of 0.09 g g–1 h–1, and high concentrations of ethanol were also produced (25 g l–1). If the initial glucose was decreased to 8 g l–1, the xylitol yield (0.79 g g–1) and specific rate (0.24 g g–1 h–1) increased with little ethanol formation (<5 g l–1). To minimize glucose repression, batch fermentations were performed using an aerobic, glucose growth phase followed by xylitol production. Xylitol was produced under O2 limited and anaerobic conditions, but the specific production rate was higher under O2 limited conditions (0.1–0.4 vs. 0.03 g g–1 h–1). On-line analysis of the respiratory quotient defined the time of xylose reductase induction.  相似文献   

3.
Morphological and physiological measurements on individual leaves of Leucaena leucocephala seedlings were used to study acclimation to neutral shading. The light-saturated photosynthetic rate (Pn max) ranged from 19.6 to 6.5 mol CO2 m–2 s–1 as photosynthetic photon flux density (PPFD) during growth decreased from 27 to 1.6 mol m–2 s–1. Stomatal density varied from 144 mm–2 in plants grown in high PPFD to 84 mm–2 in plants grown in low PPFD. Average maximal stomatal conductance for H2O was 1.1 in plants grown in high PPFD and 0.3 for plants grown in low PPFD. Plants grown in low PPFD had a greater total chlorophyll content than plants grown in high PPFD (7.2 vs 2.9 mg g–1 on a unit fresh weight basis, and 4.3 vs 3.7 mg dm–2 on a unit leaf area basis). Leaf area was largest when plants were grown under the intermediate PPFDs. Leaf density thickness was largest when plants were grown under the largest PPFDs. It is concluded that L. leucocephala shows extensive ability to acclimate to neutral shade, and could be considered a facultative shade plant.Abbreviations the initial slope of the photosynthesis vs PPFD curve - Pn max the light-saturated photosynthetic rate - PPFD photosynthetic photon flux density  相似文献   

4.
PVA-cryogels entrapping about 109 cells of Acidithiobacillus ferrooxidans per ml of gel were prepared by freezing-thawing procedure, and the biooxidation of Fe2+ by immobilized cells was investigated in a 0.365 l packed-bed bioreactor. Fe2+ oxidation fits a plug-flow reaction model well. A maximum oxidation rate of 3.1 g Fe2+ l–1 h–1 was achieved at the dilution rate of 0.4 h–1 or higher, while no obvious precipitate was determined at this time. In addition, cell-immobilized PVA-cryogels packed in bioreactor maintained their oxidative ability for more than two months under non-sterile conditions. Nomenclature: C A0 – Concentration of Fe2+ in feed stream (g l–1) C A – Concentration of Fe2 + in outlet stream (g l– 1) D – Dilution rate of the packed-bed bioreactor (h–1) F – Volumetric flow rate of iron solution (l h–1) F A0 – Mass flow rate of Fe2+ in the feed stream (g h–1) K – Kinetic constant (l l–1 h–1) r A – Oxidation rate of Fe2+ (g l–1 h–1) V – Volume of packed-bed bioreactor (l) X A – Conversion ratio of Fe2+ (%)  相似文献   

5.
Kurasová  I.  Čajánek  M.  Kalina  J.  Špunda  V. 《Photosynthetica》2000,38(4):513-519
The adaptation of barley (Hordeum vulgare L. cv. Akcent) plants to low (LI, 50 µmol m–2 s–1) and high (HI, 1000 µmol m–2 s–1) growth irradiances was studied using the simultaneous measurements of the photosynthetic oxygen evolution and chlorophyll a (Chl a) fluorescence at room temperature. If measured under ambient CO2 concentration, neither increase of the oxygen evolution rate (P) nor enhancement of non-radiative dissipation of the absorbed excitation energy within photosystem 2 (PS2) (determined as non-photochemical quenching of Chl a fluorescence, NPQ) were observed for HI plants compared with LI plants. Nevertheless, the HI plants exhibited a significantly higher proportion of QA in oxidised state (estimated from photochemical quenching of Chl a fluorescence, qP), by 49–102 % at irradiances above 200 µmol m–2 s–1 and an about 1.5 fold increase of irradiance-saturated PS2 electron transport rate (ETR) as compared to LI plants. At high CO2 concentration the degree of P stimulation was approximately three times higher for HI than for LI plants, and the irradiance-saturated P values at irradiances of 2 440 and 2 900 µmol m–2 s–1 were by 130 and 150 % higher for HI plants than for LI plants. We suggest that non-assimilatory electron transport dominates in the adaptation of the photosynthetic apparatus of barley grown at high irradiances under ambient CO2 rather than an increased NPQ or an enhancement of irradiance-saturated photosynthesis.  相似文献   

6.
AxenicTrentepohlia odorata was cultured at three different NH4Cl levels (3.5 × 10–2, 3.5 × 10–3, 3.5 × 10–4 M) and three different light intensities (48, 76, 122 µmol m–2 s–1). Chloride had no effect on growth over this range of concentration. High light intensity and high NH4Cl concentration enhanced the specific growth rate. The carotenoid content increased under a combination of high light intensity and low N concentration. WhenD. bardawil was exposed to the same combination of growth conditions, there was an increase in its carotenoid content. The light saturation and the light inhibition constants (K s andK i, respectively) for growth, and the saturation constant (K m) for NH4Cl were determined. TheK s andK i values were higher inT. odorata (66.7 and> 122 mol m–2 s–1, respectively) than inD. bardawil (5.1 and 14.7 µmol m–2 s–1, respectively). TheK m value determined at 122 µmol m–2 s–1, however, was lower inT. odorata (0.048 µM) than inD. bardawil (0.062 µM).Author for correspondence  相似文献   

7.
Kage  H.  Alt  C.  Stützel  H. 《Plant and Soil》2002,246(2):201-209
Data from field experiments carried out in three consecutive years under contrasting N supply and radiation environment altered by artificial shading were used to identify (a) the relationship between N concentration and organ size under conditions of unrestricted N supply and (b) critical levels of soil nitrate (Nmincrit), where nitrogen concentration of cauliflower organs begin to decline because of N limitations. The decline of N concentrations in cauliflower was analysed at different levels of morphological aggregation, i.e., the whole shoot level, the organ level (leaves, stem, and curd), and within different leaf groups within the canopy. Nmincrit values (0–60 cm soil depth) for total nitrogen concentration of cauliflower organs leaves, stem and curd were estimated at 85, 93 and 28 kg N ha–1, respectively. Within the canopy, Nmincrit values for total N of leaves increased from the top to the bottom from 44 to 188 kg N ha–1. Nmincrit values for protein N in leaves from different layers of the canopy were much lower at around 30 kg N ha–1, without a gradient within the canopy. It is discussed that these differences in Nmincrit values are most likely a consequence of N redistribution associated with nitrogen deficiency. The decline of average shoot nitrogen concentrations, [Nm] (%N DM), with shoot dry matter, W sh, (t ha–1) under conditions of optimal N supply was [Nm]= 4.84 (±0.071) W sh –0.089(± 0.011), r 2=0.67 (±S.E.). The reduction of radiation intensity by artificial shading (60% of control) had no significant influence on total nitrogen concentrations of leaves and only a small influence on protein nitrogen concentrations in lower layers of the canopy. The leaf nitrate nitrogen fraction of nitrogen, f nitr (–), within the canopy decreased linearly with increased average incident irradiance in different canopy layers (I av, W PAR m–2) (f Nitr. = 0.2456(±0.0188) – 0.0023(±0.0004)I av, r 2 = 0.67.  相似文献   

8.
Mature human growth hormone (hGH) cDNA was cloned by homologous recombination into the yeast Pichia pastoris genome. The hGH gene expression was placed under the control of the methanol-inducible alcohol oxidase 1 (AOX1) gene promoter and the Saccharomyces cerevisiae -factor signal sequence to direct the secretion of recombinant human growth hormone (rhGH) into the growth medium. O2-limited induction of recombinant yeast strains in shake tubes with 3 ml of culture medium produced up to 11 mg rhGH l–1, while high cell density cultures using a 2-l bioreactor produced about 49 mg rhGH l–1 achieving 40% of total protein of the culture medium supernatant.  相似文献   

9.
M. M. Babiker 《Hydrobiologia》1984,110(1):351-363
The respiratory behaviour and partitioning of O2 uptake between air and water were investigated in Polypterus genegalus using continuous-flow and two-phase respirometers and lung gas replacement techniques P. senegalus rarely resorts to aerial respiration under normal conditions. Partitioning of O2 consumption depends on the activity and age of fish and the availability of aquatic oxygen. Immature fish (12–22 g) cannot utilize aerial O2 but older fish exhibit age-dependent reliance on aerial respiration in hypoxic and hypercarbic waters. Pulmonary respiration accounts for 50% of the total requirement at aquatic O2 concentrations of about 3.5 mg · l–1 (or CO2 of about 5%) and fish rely exclusively on aerial respiration at O2 concentrations of less than 2.5 mg · l–1. Branchial respiration is initially stimulated by hypercarbia (CO2: 0.5–0.8%) but increased hypercarbia (CO2 – 1%) greatly depresses (by over 90%) brancial respiration and initiates (CO2: 0.5%) and sustains pulmonary respiration.  相似文献   

10.
Sulphate uptake by Amphidinium carterae, Amphidinium klebsii and Gymnodinium microadriaticum grown on artificial seawater medium with sulphate, cysteine, methionine or taurine as sulphur source occurred via an active transport system which conformed to Michaelis-Menten type saturation kinetics. Values for K m ranged from 0.18–2.13 mM and V max ranged from 0.2–24.2 nmol · 105 cells–1 · h–1. K m for symbiotic G. microadriaticum was 0.48 mM and V max was 0.2 nmol · 105 cells–1 · h–1. Sulphate uptake was slightly inhibited by chromate and selenate, but not by tungstate, molybdate, sulphite or thiosulphate. Cysteine and methionine (0.1 mM), but not taurine, inhibited sulphate uptake by symbiotic G. microadriaticum, but not by the two species of Amphidinium. Uptake was inhibited 45–97% under both light and dark conditions by carbonylcyanide 3-chlorophenylhydrazone (CCCP); under dark conditions sulphate uptake was 40–60% of that observed under light conditions and was little affected by 3-(3,4-dichlorophenyl) 1,1-dimethylurea (DCMU).The uptake of taurine, cysteine and methionine by A. carterae, A. klebsii, cultured and symbiotic G. microadriaticum conformed to Michaelis-Menten type saturation kinetics. K m values of taurine uptake ranged from 1.9–10 mM; for cysteine uptake from 0.6–3.2 mM and methionine from 0.001–0.021 mM. Cysteine induced a taurine uptake system with a K m of 0.3–0.7 mM. Cysteine and methionine uptake by all organisms was largely unaffected by darkness or by DCMU in light or darkness. CCCP significantly inhibited uptake of these amino acids. Thus energy for cysteine and methionine uptake was supplied mainly by respiration. Taurine uptake by A. carterae was independent of light but was inhibited by CCCP, whereas uptake by A. klebsii and symbiotic G. microadriaticum was partially dependent on photosynthetic energy. Taurine uptake by cultured G. microadriaticum was more dependent on photosynthetic energy and was more sensitive to CCCP. Cysteine inhibited uptake of methionine and taurine by cultured and symbiotic G. microadriaticum to a greater extent than in the Amphidinium species. Methionine did not greatly affect taurine uptake, but did inhibit cysteine uptake. Taurine did not affect the uptake of cysteine or methionine.  相似文献   

11.
Ray  D.  Dey  S.K.  Das  G. 《Photosynthetica》2004,42(1):93-97
Adjustment in leaf area : mass ratio called leaf area ratio (LAR) is one of the strategies to optimize photon harvesting. LAR was recorded for 10 genotypes of Hevea brasiliensis under high irradiance and low temperature and the genotypes were categorized into two groups, i.e. high LAR and low LAR types. Simultaneously, the growth during summer as well as winter periods, photosynthetic characteristics, and in-vitro oxidative damage were studied. Low LAR (19.86±0.52 m2 kg–1) types, recorded an average of 18.0 % chlorophyll (Chl) degradation under high irradiance and 7.1 % Chl degradation under low temperature. These genotypes maintained significantly higher net photosynthetic rate (P N) of 10.4 mol(CO2) m–2 s–1 during winter season. On the contrary, the high LAR (24.33±0.27 m2 kg–1) types recorded significantly lower P N of 4 mol(CO2) m–2 s–1 and greater Chl degradation of 37.7 and 13.9 % under high irradiance and low temperature stress, respectively. Thus LAR may be one of the physiological traits, which are possibly involved in plant acclimation process under both stresses studied.  相似文献   

12.
P64k is a Neisseria meningitidis high molecular weight protein present in meningococcal vaccine preparations. The lpdA gene, codifying for this protein, was cloned in Escherichia coli and the P64k protein was expressed in Escherichia coli K12 W3110 under the control of the tryptophan promoter. The recombinant bacteria were grown in batch or fed-batch cultures. P64k was expressed as an intracellular soluble form at about 40% of the total cellular protein. A final productivity of 215 mg l–1 h–1 and 11 g cell dry wt l–1 were obtained when the fed-batch culture conditions were optimised, compared to 30% of total protein, and a productivity of 76 mg l–1 h–1 and 5.1 g cell dry wt l–1 in batch cultivation.  相似文献   

13.
W. E. Robe  H. Griffiths 《Oecologia》1994,100(4):368-378
The decline and disappearance of Littorella uniflora from oligotrophic waters which have become eutrophic has been associated with shading or reduced CO2 supply. However NO inf3 sup– concentrations can reach very high levels (100–2000 mmol m–3 compared with <1–3 in oligotrophic habitats). To investigate the impact of NO inf3 sup– loading alone, plants were grown under three NO inf3 sup– regimes (very low, near-natural and high). The interactive effects of NO inf3 sup– and photon flux density (low and high regimes) on N assimilation and accumulation, CO2 concentrating mechanisms, C3 photosynthesis and growth were also examined. The results were unexpected. Increased NO inf3 sup– supply had very little effect on photosynthetic capacity, crassulacean acid metabolism (CAM) or lacunal CO2 concentrations ([CO2]i), although there was considerable plasticity with respect to light regime. In contrast, increased NO inf3 sup– supply resulted in a marked accumulation of NO inf3 sup– , free amino acids and soluble protein in shoots and roots (up to 25 mol m–3, 30 mol m–3 and 9 mg g–1 fresh weight respectively in roots), while fresh weight and relative growth rate were reduced. Total N content even under the very low NO inf3 sup– regime (1.6–2.3%) was mid-range for aquatic and terrestrial species (and 3.1–4.3% under the high NO inf3 sup– regime). These findings, together with field data, suggest that L. uniflora is not growth limited by low NO inf3 sup– supply in natural oligotophic habitats, due not to an efficient photosynthetic nitrogen use but to a slow growth rate, a low N requirement and to the use of storage to avoid N stress. However the increased NO inf3 sup– concentrations in eutrophic environments seem likely have detrimental effects on the long-term survival of L. uniflora, possibly as a consequence of N accumulation.  相似文献   

14.
Miniature heat balance-sap flow gauges were used to measure water flows in small-diameter roots (3–4 mm) in the undisturbed soil of a mature beech–oak–spruce mixed stand. By relating sap flow to the surface area of all branch fine roots distal to the gauge, we were able to calculate real time water uptake rates per root surface area (Js) for individual fine root systems of 0.5–1.0 m in length. Study aims were (i) to quantify root water uptake of mature trees under field conditions with respect to average rates, and diurnal and seasonal changes of Js, and (ii) to investigate the relationship between uptake and soil moisture θ, atmospheric saturation deficit D, and radiation I. On most days, water uptake followed the diurnal course of D with a mid-day peak and low night flow. Neighbouring roots of the same species differed up to 10-fold in their daily totals of Js (<100–2000 g m−2 d−1) indicating a large spatial heterogeneity in uptake. Beech, oak and spruce roots revealed different seasonal patterns of water uptake although they were extracting water from the same soil volume. Multiple regression analyses on the influence of D, I and θ on root water uptake showed that D was the single most influential environmental factor in beech and oak (variable selection in 77% and 79% of the investigated roots), whereas D was less important in spruce roots (50% variable selection). A comparison of root water uptake with synchronous leaf transpiration (porometer data) indicated that average water fluxes per surface area in the beech and oak trees were about 2.5 and 5.5 times smaller on the uptake side (roots) than on the loss side (leaves) given that all branch roots <2 mm were equally participating in uptake. Beech fine roots showed maximal uptake rates on mid-summer days in the range of 48–205 g m−2 h−1 (i.e. 0.7–3.2 mmol m−2 s−1), oak of 12–160 g m−2 h−1 (0.2–2.5 mmol m−2 s−1). Maximal transpiration rates ranged from 3 to 5 and from 5 to 6 mmol m−2 s−1 for sun canopy leaves of beech and oak, respectively. We conclude that instantaneous rates of root water uptake in beech, oak and spruce trees are above all controlled by atmospheric factors. The effects of different root conductivities, soil moisture, and soil hydraulic properties become increasingly important if time spans longer than a week are considered.  相似文献   

15.
Braud  Jean-Paul  Amat  Mireille A. 《Hydrobiologia》1996,326(1):335-340
The injection of exogenous carbon into intensively cultivated algal tanks is necessary to insure a maximum growth rate by stabilizing the dissolved inorganic carbon (DIC) pool, but represents the major part of the cultivation cost (ca. 73%). This study was conducted in paddle-wheel tanks ranging in size from 260 m2 to 1000 m2. Additional carbon was provided by carbon dioxide mixed into the incoming sea water through a tubular reactor. Production vs pH was analysed on 120 growth measurements covering two years of continuous cultivation. Whereas production peaked at pH 8.0–8.2, the economic optimum for pH regulation was in the range 8.4–8.5, where CO2 injection was greatly reduced (–29%) for only a slight decrease in production (–4%). Expressed as a function of pH level, the specific carbon injection (g c gdw–1 of Chondrus produced) showed an inverse exponential relationship, whereas gross photoconversion ratio (gdw mol photons–1) varied according to a second degree equation with a low amplitude. The photoconversion ratio was not improved when the culture was maintained at a DIC concentration higher than the natural equilibrium (0.64 ± 0.11 gdw mol photons–1 at 2.35 mM and 0.65 ± 0.15 gdw mol photons–1 at 3.19 MM).A complementary source of carbon was found in underground salt water with a high and stable DIC concentration (10.15 ± 0.25 mmole Cl–1). The mixing of the well water with natural sea water allowed another economy of CO2 (–20% at pH 8.5) and nutrients (–12%), the total unitary cost of production being cut by about 17%.  相似文献   

16.
George  T.S.  Gregory  P.J.  Robinson  J.S.  Buresh  R.J.  Jama  B. 《Plant and Soil》2002,246(1):53-63
A field experiment in western Kenya assessed whether the agroforestry species Tithonia diversifolia (Hemsley) A. Gray, Tephrosia vogelii Hook f., Crotalaria grahamiana Wight & Arn. and Sesbania sesban (L) Merill. had access to forms of soil P unavailable to maize, and the consequences of this for sustainable management of biomass transfer. The species were grown in rows at high planting density to ensure the soil under rows was thoroughly permeated by roots. Soil samples taken from beneath rows were compared to controls, which included a bulk soil monolith enclosed by iron sheets within the tithonia plot, continuous maize, and bare fallow plots. Three separate plant biomass samples and soil samples were taken at 6-month intervals, over a period of 18 months. The agroforestry species produced mainly leaf biomass in the first 6 months but stem growth dominated thereafter. Consequently, litterfall was greatest early in the experiment (0–6 months) and declined with continued growth. Soil pH increased by up to 1 unit (from pH 4.85) and available P increased by up to 38% (1 g P g–1) in agroforestry plots where biomass was conserved on the field. In contrast, in plots where biomass was removed, P availability decreased by up to 15%. Coincident with the declines in litterfall, pH decreased by up to 0.26 pH units, plant available P decreased by between 0.27 and 0.72 g g–1 and Po concentration decreased by between 8 and 35 g g–1 in the agroforestry plots. Declines in Po were related to phosphatase activity (R2=0.65, P<0.05), which was greater under agroforestry species (0.40–0.50 nmol MUB s–1 g–1) than maize (0.28 nmol MUB s–1 g–1) or the bare fallow (0.25 nmol MUB s–1 g–1). Management of tithonia for biomass transfer, decreased available soil P by 0.70 g g–1 and Po by 22.82 g g–1. In this study, tithonia acquired Po that was unavailable to maize. However, it is apparent that continuous cutting and removal of biomass would lead to rapid depletion of P stored in organic forms.  相似文献   

17.
A fermentation medium based on millet (Pennisetum typhoides) flour hydrolysate and a four-phase feeding strategy for fed-batch production of baker's yeast,Saccharomyces cerevisiae, are presented. Millet flour was prepared by dry-milling and sieving of whole grain. A 25% (w/v) flour mash was liquefied with a thermostable 1,4--d-glucanohydrolase (EC 3.2.1.1) in the presence of 100 ppm Ca2+, at 80°C, pH 6.1–6.3, for 1 h. The liquefied mash was saccharified with 1,4--d-glucan glucohydrolase (EC 3.2.1.3) at 55°C, pH 5.5, for 2 h. An average of 75% of the flour was hydrolysed and about 82% of the hydrolysate was glucose. The feeding profile, which was based on a model with desired specific growth rate range of 0.18–0.23 h–1, biomass yield coefficient of 0.5 g g–1 and feed substrate concentration of 200 g L–1, was implemented manually using the millet flour hydrolysate in test experiments and glucose feed in control experiments. The fermentation off-gas was analyzed on-line by mass spectrometry for the calculation of carbon dioxide production rate, oxygen up-take rate and the respiratory quotient. Off-line determination of biomass, ethanol and glucose were done, respectively, by dry weight, gas chromatography and spectrophotometry. Cell mass concentrations of 49.9–51.9 g L–1 were achieved in all experiments within 27 h of which the last 15 h were in the fedbatch mode. The average biomass yields for the millet flour and glucose media were 0.48 and 0.49 g g–1, respectively. No significant differences were observed between the dough-leavening activities of the products of the test and the control media and a commercial preparation of instant active dry yeast. Millet flour hydrolysate was established to be a satisfactory low cost replacement for glucose in the production of baking quality yeast.Nomenclature C ox Dissolved oxygen concentration (mg L–1) - CPR Carbon dioxide production rate (mmol h–1) - C s0 Glucose concentration in the feed (g L–1) - C s Substrate concentration in the fermenter (g L–1) - C s.crit Critical substrate concentration (g L–1) - E Ethanol concentration (g L–1) - F s Substrate flow rate (g h–1) - i Sample number (–) - K e Constant in Equation 6 (g L–1) - K o Constant in Equation 7 (mg L–1) - K s Constant in Equation 5 (g L–1) - m Specific maintenance term (h–1) - OUR Oxygen up-take rate (mmol h–1) - q ox Specific oxygen up-take rate (h–1) - q ox.max Maximum specific oxygen up-take rate (h–1) - q p Specific product formation rate (h–1) - q s Specific substrate up-take rate (g g–1 h–1) - q s.max Maximum specific substrate up-take rate (g g–1 h–1) - RQ Respiratory quotient (–) - S Total substrate in the fermenter at timet (g) - S 0 Substrate mass fraction in the feed (g g–1) - t Fermentation time (h) - V Instantaneous volume of the broth in the fermenter (L) - V 0 Starting volume in the fermenter (L) - V si Volume of samplei (L) - x Biomass concentration in the fermenter (g L–1) - X 0 Total amount of initial biomass (g) - X t Total amount of biomass at timet (g) - Y p/s Product yield coefficient on substrate (–) - Y x/e Biomass yield coefficient on ethanol (–) - Y x/s Biomass yield coefficient on substrate (–) Greek letters Moles of carbon per mole of yeast (–) - Moles of hydrogen atom per mole of yeast (–) - Moles of oxygen atom per mole of yeast (–) - Moles of nitrogen atom per mole of yeast (–) - Specific growth rate (h–1) - crit Critical specific growth rate (h–1) - E Specific ethanol up-take rate (h–1) - max.E Maximum specific ethanol up-take rate (h–1)  相似文献   

18.
Transformation of the monocot Alstroemeria by Agrobacterium rhizogenes   总被引:1,自引:0,他引:1  
An efficient procedure is described for transformation of calli of the monocotyledonous plant Alstroemeria by Agrobacterium rhizogenes. Calli were co-cultivated with A. rhizogenes strain A13 that harbored both a wild-type Ri-plasmid and the binary vector plasmid pIG121Hm, which included a gene for neomycin phosphotransferase II (NPTII) under the control of the nopaline synthase (NOS) promoter, a gene for hygromycin phosphotransferase (HPT) under the control of the cauliflower mosaic virus (CaMV) 35S promoter, and a gene for -glucuronidase (GUS) with an intron fused to the CaMV 35S promoter. Inoculated calli were plated on medium that contained cefotaxime to eliminate bacteria. Four weeks later, transformed cells were selected on medium that contained 20 mg L–1 hygromycin. A histochemical assay for GUS activity revealed that selection by hygromycin was complete after eight weeks. The integration of the T-DNA of the Ri-plasmid and pIG121Hm into the plant genome was confirmed by PCR. Plants derived from transformed calli were produced on half-strength MS medium supplemented with 0.1 mg L–1 GA3 after about 5 months of culture. The presence of the gusA, nptII, and rol genes in the genomic DNA of regenerated plants was detected by PCR and Southern hybridization, and the expression of these transgenes was verified by RT-PCR.  相似文献   

19.
A method for isolation of d-amino acid oxidase (DAAO) from disrupted Trigonopsis variabilis cells has been developed. In an aqueous two-phase system consisting of PEG6000 (220 g l–1), potassium phosphate (110 g l–1, K2HPO4 + KH2PO4 = 10.1:1, mol mol–1) and dl-methionine (11 g l–1), the major portion of cellular proteins (87%) was partitioned into the salt phase. By sequential extraction, 48% of DAAO was recovered in PEG phase, giving a yield of 211 U mg protein–1.  相似文献   

20.
A thermophilic bacterium, which we designated as Geobacillus thermoleovorans 47b was isolated from a hot spring in Beppu, Oita Prefecture, Japan, on the basis of its ability to grow on bitter peptides as a sole carbon and nitrogen source. The cell-free extract from G. thermoleovorans 47b contained leucine aminopeptidase (LAP; EC 3.4.11.10), which was purified 164-fold to homogeneity in seven steps, using ammonium sulfate fractionation followed by the column chromatography using DEAE-Toyopearl, hydroxyapatite, MonoQ and Superdex 200 PC gel filtration, followed again by MonoQ and hydroxyapatite. The enzyme was a single polypeptide with a molecular mass of 42,977.2 Da, as determined by matrix-assisted laser desorption ionization and time-of-flight mass spectrometry, and was found to be thermostable at 90°C for up to 1 h. Its optimal pH and temperature were observed to be 7.6–7.8 and 60°C, respectively, and it had high activity towards the substrates Leu-p-nitroanilide (p-NA)(100%), Arg-p-NA (56.3%) and LeuGlyGly (486%). The Km and Vmax values for Leu-p-NA and LeuGlyGly were 0.658 mM and 25.0 mM and 236.2 mol min–1 mg–1 protein and 1,149 mol min–1 mg–1 protein, respectively. The turnover rate (kcat) and catalytic efficiency (kcat/ Km) for Leu-p-NA and LeuGlyGly were 10,179 s–1 and 49,543 s–1 and 15,470 mM–1 s–1 and 1981.7 mM–1 s–1, respectively. The enzyme was strongly inhibited by EDTA, 1,10-phenanthroline, dithiothreitol, -mercaptoethanol, iodoacetate and bestatin; and its apoenzyme was found to be reactivated by Co2+ .  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