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.

Unveiling the determinants for transferase and hydrolase activity in glycoside hydrolases would allow using their vast diversity for creating novel transglycosylases, thereby unlocking an extensive toolbox for carbohydrate chemists. Three different amino acid substitutions at position 220 of a GH1 β-glucosidase from Thermotoga neapolitana caused an increase of the ratio of transglycosylation to hydrolysis (r s/r h) from 0.33 to 1.45–2.71. Further increase in r s/r h was achieved by modulation of pH of the reaction medium. The wild-type enzyme had a pH optimum for both hydrolysis and transglycosylation around 6 and reduced activity at higher pH. Interestingly, the mutants had constant transglycosylation activity over a broad pH range (5–10), while the hydrolytic activity was largely eliminated at pH 10. The results demonstrate that a combination of protein engineering and medium engineering can be used to eliminate the hydrolytic activity without affecting the transglycosylation activity of a glycoside hydrolase. The underlying factors for this success are pursued, and perturbations of the catalytic acid/base in combination with flexibility are shown to be important factors.

  相似文献   

2.
We examined a 6‐year record of automated chamber‐based soil CO2 efflux (Fs) and the underlying processes in relation to climate and canopy gas exchange at an AmeriFlux site in a seasonally drought‐stressed pine forest. Interannual variability of Fs was large (CV=17%) with a range of 427 g C m?2 yr?1 around a mean annual Fs of 811 g C m?2 yr?1. On average, 76% of the variation of daily mean Fs could be quantified using an empirical model with year‐specific basal respiration rate that was a linear function of tree basal area increment (BAI) and modulated by a common response to soil temperature and moisture. Interannual variability in Fs could be attributed almost equally to interannual variability in BAI (a proxy for above‐ground productivity) and interannual variability in soil climate. Seasonal total Fs was twice as sensitive to soil moisture variability during the summer months compared with temperature variability during the same period and almost insensitive to the natural range of interannual variability in spring temperatures. A strong seasonality in both root respiration (Rr) and heterotrophic respiration (Rh) was observed with the fraction attributed to Rr steadily increasing from 18% in mid‐March to 50% in early June through early July before dropping rapidly to 10% of Fs by mid‐August. The seasonal pattern in Rr (10‐day averages) was strongly linearly correlated with tree transpiration (r2=0.90, P<0.01) as measured using sap flux techniques and gross ecosystem productivity (GEP, r2=0.83, P<0.01) measured by the eddy‐covariance approach. Rr increased by 0.43 g C m?2 day?1 for every 1 g C m?2 day?1 increase in GEP. The strong linear correlation of Rr to seasonal changes in GEP and transpiration combined with longer‐term interannual variability in the base rate of Fs, as a linear function of BAI (r2=0.64, P=0.06), provides compelling justification for including canopy processes in future models of Fs.  相似文献   

3.
Extracellular hydrolysis of 4-methylumbelliferyl β-N-acetylglucosaminide was measured in the oligomesotrophic Piburger See and the eutrophic Římov reservoir during spring and summer phytoplankton blooms, respectively. Total enzymatic activity (TEA) ranged between 0.2 and 19.1 nmol 1−1 h−1 in the reservoir and between 0.8 and 12.4 nmol 1−1 h−1 in the lake. High-affinity (Km < 1 μmol 1−1) and lowaffinity (Km > 100 μmol 1−1) enzymes were kinetically identifiable in most samples from both localities. The low-affinity enzyme activity (LEA) usually accounted for >60% (mean: 80%) of TEA. LEA and diatom biomass significantly correlated over time in the reservoir epilimnion (rs = 0.578) and in the lake metalimnion (rs = 0.862). As diatoms possess chitin and take up its monomer, N-acetylglucosamine, two explanations of the observed relationships are suggested: extracellular β-N-acetylglucosaminidase activity partly originates either from ectoenzymes of chitinolytic bacteria attached to diatom cells or from ectoenzymes of diatoms, enabling them to take up N-acetylglucosamine from ambient amino sugars instead of synthesizing it de novo. A significant positive correlation of LEA with crustacean abundance was found in the lake epilimnion (rs = 0.850), apparently reflecting the growing spring populations of frequently moulting juvenile crustaceans. A possible contribution of chitinolytic bacteria, accompanying the crustacean populations, to LEA is discussed.  相似文献   

4.
We have developed a new model describing the relationship between plasma and red cell tracers flowing through the lung. The model is the result of an analysis of the transport of radiolabeled plasma albumin between two flowing phases and shows that differences between red cell and plasma tracer curves are related to microvascular hematocrit. The model was tested in an isolated, blood-perfused dog lung preparation in which we injected51Cr-labeled red cells and125I-labeled plasma albumin into the pulmonary artery. From the tracer concentration-time curves at the venous outflow, we calculatedh r, the ratio of microvascular hematocrit to large-vessel hematocrit. In 18 baseline experiments,h r=0.92±0.01 (mn±sem) at a blood flow rate of 10.7±0.3 ml s−1. We determined the effects of (a) glass bead embolization, (b) alloxan, and (c) lobe ligation onh r. Embolization attenuated the separation between plasma and red cells (increasedh r), probably as a consequence of passive vasodilation. Alloxan enhanced separation of plasma and red cells (decreasedh r), possibly as a result of arteriolar vasoconstriction. Ligation of a fraction of the perfused tissue at constant flow did not cause significant change inh r in the remaining perfused tissue. The model assumes that large-vessel transit times are uniform and that all dispersion occurs in the microvasculature. A theoretical analysis apportioning dispersion between large and small vessels disclosed that the error associated with these assumptions is likely to be less than 15% of the measuredh r. We conclude from this study that the microvascular hematocrit model describes experimental plasma and red cell curves. The results imply thath r can be readily deduced from tagged red cells and plasma and can be accounted for in calculating permeability-surface area in diffusing tracer experiments.  相似文献   

5.
Summary To increase the solvent productivity of the acetone-butanol fermentation, a continuous culture of Clostridium acetobytylicum with cell recycling was used. At a dry cell mass concentration of 8 g l-1 and a dilution rate of D=0.64 h-1, a solvent productivity of 5.4 g l-1 h-1 was attained. To prevent degeneration of the culture, which occurs with high concentrations of solvents (acetone, butanol and ethanol), different reactor cascades were used. A two-stage cascade with cell recycling and turbidostatic cell concentration control turned out to be the best solution, the first stage of which was kept at relatively low cell and product concentrations. A solvent productivity of 3 and 2.3 g l-1 h-1, respectively, was achieved at solvent concentrations of 12 and 15 g l-1.Symbols D Dilution rate (h-1) - r p solvent productivity (g l-1 h-1) - s residual glucose concentration (g l-1) - V R reactor volume (l) - V O overall volume (l) - x (dry) cell mass concentration (g l-1) - Y P/S solvent yield (g g-1)  相似文献   

6.
Genetic parameters for stem diameter and wood density were compared at selection (4–5 years) and harvest (16–17 years) age in an open-pollinated progeny trial of Eucalyptus globulus in Tasmania (Australia). The study examined 514 families collected from 17 subraces of E. globulus. Wood density was assessed on a subsample of trees indirectly using pilodyn penetration at both ages and directly by core basic density at harvest age. Significant additive genetic variance and narrow-sense heritabilities ( h\textop2 h_{\text{op}}^2 ) were detected for all traits. Univariate and multivariate estimates of heritabilities were similar for each trait except harvest-age diameter. Comparable univariate estimates of selection- and harvest-age heritabilities for diameter masked changes in genetic architecture that occurred with stand development, whereby the loss of additive genetic variance through size-dependent mortality was countered by the accentuation of additive genetic differences among survivors with age. Regardless, the additive genetic (r a) and subrace (r s) correlations across ages were generally high for diameter (0.95 and 0.61, respectively) and pilodyn penetration (0.77 and 0.96), as were the correlations of harvest-age core basic density with selection- and harvest-age pilodyn (r a −0.83, −0.88; r s −0.96, −0.83). While r s between diameter and pilodyn were close to zero at both ages, there was a significant change in r a from adverse at selection age (0.25) to close to zero (−0.07) at harvest age. We argue that this change in the genetic correlation reflects a decoupling of the genetic association of growth and wood density with age. This result highlights the need to validate the use of selection-age genetic parameters for predicting harvest-age breeding values.  相似文献   

7.
Lolium perenne selection lines with high calculated stomatal resistances to diffusion (rs) as a result of either few or short stomata, maintained leaf extension rates and photosynthetic rates longer than selection lines with low resistances when deprived of water. There were no significant differences between high and low rs plants in light saturated CO2 uptake of turgid attached leaves. When grown in soil drying to 21% moisture, plants with low calculated rs exhibited lower minimum leaf resistances (rl) than those with high, measured with a diffusion porometer, on all except the last day. The daily maximum rl (1.5 h after the start of the light period) became greater among low than high rs plants as the difference in rate of leaf extension between the two groups of plants increased with drying soil. Rate of leaf extension was negatively correlated with daily maximum rl and started to decline when relative leaf water content (RLWC), at 5 h after the start of the light period, fell below about 88%. Transpiration rate of plants grown in different soil moisture regimes was correlated (r=+ 0.83, P < 0.01) with mean maximum adaxial leaf conductance (reciprocal of resistance). There was a highly significant correlation (r=+ 0.62, P < 0.01) between calculated adaxial rs and mean minimum measured rl among plants growing at high or intermediate soil moisture, but not at low. Therefore, some random variation in minimum rl, even with adequate moisture, seemed to be unaccounted for by variation in stomatal numbers or size. Selection for increased numbers of adaxial stomata also resulted in more on the abaxial surface, but mean adaxial/abaxial ratio in the ‘frequent’ stomata plants was still only about 9:1.  相似文献   

8.
This is the first part of a survey of hierarchical clustering algorithms using joining methods: the Single-Linkage algorithm. Complete-Linkage and general algorithms defined by d(Ai, B) = = α,d(Ai, Ar)±αsd(Ai, As)±βd(Ar, As) will be discussed in two subsequent papers.  相似文献   

9.
A trenching method was used to determine the contribution of root respiration to soil respiration. Soil respiration rates in a trenched plot (R trench) and in a control plot (R control) were measured from May 2000 to September 2001 by using an open-flow gas exchange system with an infrared gas analyser. The decomposition rate of dead roots (R D) was estimated by using a root-bag method to correct the soil respiration measured from the trenched plots for the additional decaying root biomass. The soil respiration rates in the control plot increased from May (240–320 mg CO2 m–2 h–1) to August (840–1150 mg CO2 m–2 h–1) and then decreased during autumn (200–650 mg CO2 m–2 h–1). The soil respiration rates in the trenched plot showed a similar pattern of seasonal change, but the rates were lower than in the control plot except during the 2 months following the trenching. Root respiration rate (R r) and heterotrophic respiration rate (R h) were estimated from R control, R trench, and R D. We estimated that the contribution of R r to total soil respiration in the growing season ranged from 27 to 71%. There was a significant relationship between R h and soil temperature, whereas R r had no significant correlation with soil temperature. The results suggest that the factors controlling the seasonal change of respiration differ between the two components of soil respiration, R r and R h.  相似文献   

10.
Genetic variability within and among 10 geographically distinct populations of Greenfinches (Carduelis chloris) was assayed by directly sequencing a 637 BP part of the mtDNA control region from 194 individuals. Thirteen variable positions defined 18 haplotypes with a maximum sequence divergence of 0.8%. Haplotype (h = 0.28–0.77) and nucleotide (π = 0.058–0.17%) diversities within populations were low, and decreased with increasing latitude (h:rs = –0.81; π: rs = –0.89). The distribution of pairwise nucleotide differences fit better with expectations of a “sudden expansion” than of an “equilibrium” model, and the estimates of long term effective population sizes were considerably lower than current census estimates, especially in northern European samples. Selection is an unlikely cause of observed patterns because the distribution of variability conformed to expectations of neutral infinite alleles model and haplotype diversity across populations was positively correlated with heterozygosity (HE) in nuclear genes (rs = 0.74, P < 0.05). Hence, a recent bottleneck, followed by serial bottlenecking during the process of post-Pleistocene recolonization of northern Europe, together with recent population expansion provide a plausible explanation for the low genetic diversity in the north. Genetic distances among populations showed a clear pattern of isolation-by-distance, and 14% of the haplotypic variation was among populations, the rest being distributed among individuals within populations. In accordance with allozyme and morphological data, a hierarchical analysis of nucleotide diversity recognized southern European populations as distinct from northern European ones. However, the magnitude of divergence in mtDNA, allozymes and morphology were highly dissimilar (morphology > mtDNA > allozymes).  相似文献   

11.
We report here on configurational and conformational studies undertaken on the bromofluoro-1,4-benzodiazepinooxazole derivative I, which has previously been found capable of undergoing total spontaneous resolution under racemizing conditions. Due to its bridgehead N-atom I may exist in any of four diastereomeric forms,C r N r (orC s N s ) andC r N s (orC s N r ). Molecular mechanics calculations revealed that in their lowest energy conformations theC r N r (orC s N s ) diastereomers were favored over theC r N s (orC s N r ) diastereomers by some 3.42 kcal/mol, while semi-empirical quantum mechanical calculations indicated heats of formation also favoring theC r N (orC s N s ) diastereomers by 3.83 kcal/mol. The1H NMR spectra of the three crystalline modifications (, and) of I were examined in acetone, and found to be identical. These data, combined with the results of the above calculations, suggest that I exists in methanol or methanol-acetone solutions as theC r N (orC s N s ) enantiomer prior to its racemization.  相似文献   

12.
Mirit Eynan  Razi Dmi'el 《Oecologia》1993,95(2):290-294
Many desert lizards show reduced rates of cutaneous water loss (CWL) compared to their counterparts from more humid environments. It is not clear yet whether reduced CWL is connected to the taxonomic position of the lizard studied, or is affected more by environmental or experimental conditions. To investigate this, we measured the skin resistance to water transfer, R s, in five closely related lizard taxa of the genus Agama. These diurnal lizards are distributed in Israel from mesic-Mediterranean to extreme desert biotopes. The highest R s (738 s cm-1) and the lowest CWL (0.160 mg cm-2 h-1) were found in Agama sinaita, which lives in the most arid habitat. The lowest R s (234 s cm-1) and the highest CWL (0.548 mg cm-2 h-1) were found in A. stellio ssp., which occupies mesic habitats. In addition, only the desert species were able to change their R s in accordance with the changing experimental conditions. These R s changes, which probably reflect vasomotor responses, were more pronounced in A. sinaita and presumably enable the desert species to control their CWL in a hot and dry environment.  相似文献   

13.
Reduced stomatal conductance (gs) during soil drought in angiosperms may result from effects of leaf turgor on stomata and/or factors that do not directly depend on leaf turgor, including root‐derived abscisic acid (ABA) signals. To quantify the roles of leaf turgor‐mediated and leaf turgor‐independent mechanisms in gs decline during drought, we measured drought responses of gs and water relations in three woody species (almond, grapevine and olive) under a range of conditions designed to generate independent variation in leaf and root turgor, including diurnal variation in evaporative demand and changes in plant hydraulic conductance and leaf osmotic pressure. We then applied these data to a process‐based gs model and used a novel method to partition observed declines in gs during drought into contributions from each parameter in the model. Soil drought reduced gs by 63–84% across species, and the model reproduced these changes well (r2 = 0.91, P < 0.0001, n = 44) despite having only a single fitted parameter. Our analysis concluded that responses mediated by leaf turgor could explain over 87% of the observed decline in gs across species, adding to a growing body of evidence that challenges the root ABA‐centric model of stomatal responses to drought.  相似文献   

14.
Oxygen consumption rates of adult spring chinook salmon Oncorhynchus tshawytscha increased with swim speed and, depending on temperature and fish mass, ranged from 609 mg O2 h?1 at 30 cm s?1 (c. 0·5 BL s?1) to 3347 mg O2 h?1 at 170 cm s?1 (c. 2·3 BL s?1). Corrected for fish mass, these values ranged from 122 to 670 mg O2 kg?1 h?1, and were similar to other Oncorhynchus species. At all temperatures (8, 12·5 and 17° C), maximum oxygen consumption values levelled off and slightly declined with increasing swim speed >170 cm s?1, and a third‐order polynomial regression model fitted the data best. The upper critical swim speed (Ucrit) of fish tested at two laboratories averaged 155 cm s?1 (2·1 BL s?1), but Ucrit of fish tested at the Pacific Northwest National Laboratory were significantly higher (mean 165 cm s?1) than those from fish tested at the Columbia River Research Laboratory (mean 140 cm s?1). Swim trials using fish that had electromyogram (EMG) transmitters implanted in them suggested that at a swim speed of c. 135 cm s?1, red muscle EMG pulse rates slowed and white muscle EMG pulse rates increased. Although there was significant variation between individual fish, this swim speed was c. 80% of the Ucrit for the fish used in the EMG trials (mean Ucrit 168·2 cm s?1). Bioenergetic modelling of the upstream migration of adult chinook salmon should consider incorporating an anaerobic fraction of the energy budget when swim speeds are ≥80% of the Ucrit.  相似文献   

15.
We quantified metabolic power consumption as a function of wind speed in the presence and absence of simulated solar radiation in rock squirrels, Spermophilus variegatus, a diurnal rodent inhabiting arid regions of Mexico and the western United States. In the absence of solar radiation, metabolic rate increased 2.2-fold as wind speed increased from 0.25 to 4.0 m·s-1. Whole-body thermal resistance declined 56% as wind speed increased over this range, indicating that body insulation in this species is much more sensitive to wind disruption than in other mammals. In the presence of 950 W·m-2 simulated solar radiation, metabolic rate increased 2.3-fold as wind speed was elevated from 0.25 to 4.0 m·s-1. Solar heat gain, calculated as the reduction in metabolic heat production associated with the addition of solar radiation, increased with wind speed from 1.26 mW·g-1 at 0.25 m·s-1 to 2.92 mW·g-1 at 4.0 m·s-1. This increase is opposite to theoretical expectations. Both the unexpected increase in solar heat gain at elevated wind speeds and the large-scale reduction of coat insulation suggests that assumptions often used in heat-transfer analyses of animals can produce important errors.Abbreviations absorptivity of coat to solar radiation - kinematic viscosity of air (mm2·s-1) - reflectivity of coat to solar radiation - a r B expected at zero wind speed (s·m-1) - A P projected surface area of animal on plane perpendicular to solar beam (cm2) - A SKIN skin surface area (cm2) - b Coefficient describing change in r B with change in square-root of wind speed (s1.5·m1.5) - d hair diameter (m) - d characteristic dimension of animal (m) - D H thermal diffusivity of air (m2·s-1) - E evaporative heat loss (W·m-2) - I probability per unit coat depth that photon will strike hair - k constant equalling 1200 J·m-3·°C-1 - l C coat depth m) - l H hair length (m) - M metabolic rate (W·m-2) - n density of hairs of skin (m-2) - Q A solar heat gain to animal (W·m-2) - Q I solar irradiance intercepted by animal (W·m-2) - RQ respiratory quotient - r A thermal resistance of boundary layer (s·m-1) - r B whole-body thermal resistance (s·m-1) - r E thermal resistance between animal surface and environment s·m-1) - r R radiative resistance (s·m-1) - r S sum of r B and r E at 0.25 m·s-1 (s·m-1) - r T tissue thermal resistance s·m-1) - T AIR air temperature (°C) - T B body temperature (°C) - T E operative temperature of environment (°C) - T ES standard operative temperature of environment (°C) - u wind speed (m·s-1)  相似文献   

16.
The growth of Hansenula polymorpha and Kloeckera sp. 2201 with a mixture of glucose and methanol (38.8%/61.2%, w/w) and the regulation of the methanol dissimilating enzymes alcohol oxidase, catalase, formaldehyde dehydrogenase and formate dehydrogenase were studied in chemostat culture, as a function of the dilution rate. Both organisms utilized and assimilated glucose and methanol simultaneously up to dilution rates of 0.30 h-1 (H. polymorpha) and 0.26h-1, respectively (Kloeckera sp. 2201) which significantly exceeded max found for the two yeasts with methanol as the only source of carbon. At higher dilution rates methanol utilisation ceased and only glucose was assimilated. Over the whole range of mixed-substrate growth both carbon sources were assimilated with the same efficiency as during growth with glucose or methanol alone.In cultures of H. polymorpha, however, the growth yield for glucose was lowered by the unmetabolized methanol at high dilution rates. During growth on both carbon sources the repression of the synthesis of all catabolic methanol enzymes which is normally caused by glucose was overcome by the inductive effect of the simultaneously fed methanol. In both organisms the synthesis of alcohol oxidase was found to be regulated differently as compared to catalase, formaldehyde and formate dehydrogenase. Whereas increasing repression of the synthesis of alcohol oxidase was found with increasing dilution rates as indicated by gradually decreasing specific activities of this enzyme in cell-free extracts, the specific activities of this enzyme in cell-free extracts, the specific activities of catalase and the dehydrogenases increased with increasing growth rates until repression started. The results indicate similar patterns of the regulation of the synthesis of methanol dissimilating enzymes in different methylotrophic yeasts.Abbreviations and Terms C1 Methanol - C6 glucose; D dilution rate (h-1) - D c critical dilution rate (h-1) - q s specific, rate of substrate consumption (g substrate [g cell dry weight]-1 h-1) - q CO2 and q O2 are the specific rates of carbon dioxide release and oxygen consumption (mmol [g cell dry weight]-1 h-1) - RQ respiration quotient (q CO2 q O2 1 ) - s 0(C1) and s 0(C6) are the concentrations of methanol and glucose in the inflowing medium (g l-1) - s residual substrate concentration in the culture liquid (g l-1) - Sp. A. enzyme specific activity - x cell dry weight concentration (gl-1) - Y X/C6 growth yield on glucose (g cell dry weight [g substrate]-1  相似文献   

17.
A general model of a large 2m-ploid breeding population, withr loci ands h alleles at the h th locus is considered. It is assumed that the population is bisexual, non-overlapping and breeds by random mating. The genotypic structure of the population is presented as a bilinear form in the gametic output vectors where the genotype distribution is in the matrix form. Using the concept of the segregation distribution, the genotype proportions in the (n+1)st generation are given. An equilibrium condition for random chromosome segregation is obtained in terms of gene frequencies.  相似文献   

18.
Effects of undecanoic acid (UDA) on germination of microconidia and elongation of germ tubes in UDA sensitive (uda s) wild type Trichophyton rubrum and UDA resistant (uda r) mutant derived from it, were studied. UDA inhibited conidial germination of uda s and uda r strains at 30 g/ml and 120 g/ml respectively which were minimum inhibitory concentrations of UDA for these two strains. When spores from both uda s and uda r were germinated in presence of subinhibitory concentration of UDA, germ tube growth was short. The elongation of germ tubes of spores pregerminated in absence of UDA was also inhibited by dose of UDA not sufficient to inhibit germination.  相似文献   

19.
Summary The influence of variations in the boundary air layer thickness on transpirtion due to changes in leaf dimension or wind speed was evaluated at a given stomatal resistance (r s) for various combinations of air temperature (T a) and total absorbed solar energy expressed as a fraction of full sunlight (S ffs). Predicted transpiration was found to either increase or decrease for increases in leaf size depending on specific combinations of T a, S ffs, and r s. Major reductions in simulated transpiration with increasing leaf size occurred for shaded, highly reflective, or specially oriented leaves (S ffs=0.1) at relatively high T a when r s was below a critical value of near 500 s m-1. Increases in S ffs and decreases in T a lowered this critical resistance to below 50 s m-1 for S ffs=0.7 and T a=20°C. In contrast, when r s was above this critical value, an increase in leaf dimension (or less wind) resulted in increases in transpiration, especially at high T a and S ffs. For several combinations of T a, S ffs, and r s, transpiration was minimal for a specific leaf size. These theoretical results were compared to field measurements on common desert, alpine, and subalpine plants to evaluate the possible interactions of leaf and environmental parameters that may serve to reduce transpiration in xeric habitats.  相似文献   

20.
Candida utilis was grown on a pineapple cannery effluent in a chemostat at dilution rates ranging between 0.05 and 0.65 h–1 to establish optimal conditions for biomass production and chemical oxygen demand (COD) reduction. Sucrose, fructose and glucose were the main sugars in the effluent. Maximum value for cell yield coefficient and productivity were (0.686, gx/gs) and (2.96, gx/l/h) at a dilution rate of 0.425 and 0.475 h–1, respectively, while maximum COD reduction (98%) was attained at a dilution rate of 0.1 h–1. The maintenance coefficient attained a value of (0.093, gs/gx/h). An increase in dilution rate produced a higher protein content of the biomass.  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