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 83 毫秒
1.
Aims: The purification and biochemical properties of the 1,4‐β‐xylosidase of an oenological yeast were investigated. Methods and Results: An ethanol‐tolerant 1,4‐β‐xylosidase was purified from cultures of a strain of Pichia membranifaciens grown on xylan at 28°C. The enzyme was purified by sequential chromatography on DEAE cellulose and Sephadex G‐100. The relative molecular mass of the enzyme was determined to be 50 kDa by SDS‐PAGE. The activity of 1,4‐β‐xylosidase was optimum at pH 6·0 and at 35°C. The activity had a Km of 0·48 ± 0·06 mmol l?1 and a Vmax of 7·4 ± 0·1 μmol min?1 mg?1 protein for p‐nitrophenyl‐β‐d ‐xylopyranoside. Conclusions: The enzyme characteristics (pH and thermal stability, low inhibition rate by glucose and ethanol tolerance) make this enzyme a good candidate to be used in enzymatic production of xylose and improvement of hemicellulose saccharification for production of bioethanol. Significance and Impact of the Study: This study may be useful for assessing the ability of the 1,4‐β‐xylosidase from P. membranifaciens to be used in the bioethanol production process.  相似文献   

2.
The objective of this study was to determine the effect of freezing on the function in Atlantic salmon Salmo salar spermatozoa. The semen was frozen in Cortland's medium + 1.3M dimethyl sulphoxide + 0.3M glucose + 2% bovine serum albumin (final concentration) in a ratio of 1:3 (semen:cryoprotectant) as the treatment (T) and fresh semen as the control (F). Straws of 0·5 ml of sperm suspension were frozen in 4 cm of N2L. They were thawed in a thermoregulated bath (40° C). After thawing, the percentage of spermatozoa with fragmented DNA [transferase dUTP (deoxyuridine triphosphate) nick‐end labelling (TUNEL)], plasma membrane integrity (SYBR‐14/PI) and mitochondrial membrane potential (ΔΨMMit, JC‐1) were evaluated by flow cytometry and motility was evaluated by optical microscope under stroboscopic light. The fertilization rates of the control and treatment semen were tested at a sperm density of 1·5 × 107 spermatozoa oocyte?1, by observation of the first cleavages after 16 h incubation at 10° C. In the cryopreserved semen (T), the mean ± s.d . DNA fragmentation was 4·8 ± 2·5%; plasma membrane integrity 75·2 ± 6·3%; mitochondrial membrane potential 51·7 ± 3·6%; motility 58·5 ± 5·3%; curved line velocity (VCL) 61·2 ± 17·4 µm s?1; average‐path velocity (VAP) 50·1 ± 17·3 µm s?1; straight‐line velocity (VSL) 59·1 ± 18·4 µm s?1; fertilization rate 81·6 ± 1·9%. There were significant differences in the plasma membrane integrity, mitochondrial membrane potential, motility, fertilization rate, VCL, VAP and VSL compared with the controls (P < 0·05). Also the mitochondrial membrane potential correlated with motility, fertilization rate, VCL and VSL (r = 0·75; r = 0·59; r = 0·77 and r = 0·79, respectively; P < 0·05); and the fertilization rate correlated with VCL and VSL (r = 0·59 and r = 0·55, respectively).  相似文献   

3.
To obtain general rules of peptide design using α,β-dehydro-residues, a sequence with two consecutive ΔPhe-residues, Boc-L -Val-ΔPhe–ΔPhe- L -Ala-OCH3, was synthesized by azlactone method in solution phase. The peptide was crystallized from its solution in an acetone/water mixture (70:30) in space group P61 with a=b=14.912(3) Å, c= 25.548(5) Å, V=4912.0(6) Å3. The structure was determined by direct methods and refined by a full matrix least-squares procedure to an R value of 0.079 for 2891 observed [I?3σ(I)] reflections. The backbone torsion angles ?1=?54(1)°, ψ1= 129(1)°, ω1=?177(1)°, ?2 =57(1)°, ψ2=15(1)°, ω2 =?170(1)°, ?3=80(1)°, ψ3 =7(2)°, ω3=?177(1)°, ?4 =?108(1)° and ψT4=?34 (1)° suggest that the peptide adopts a folded conformation with two overlapping β-turns of types II and III′. These turns are stabilized by two intramolecular hydrogen bonds between the CO of the Boc group and the NH of ΔPhe3 and the CO of Val1 and the NH of Ala4. The torsion angles of ΔPhe2 and ΔPhe3 side chains are similar and indicate that the two ΔPhe residues are essentially planar. The folded molecules form head-to- tail intermolecular hydrogen bonds giving rise to continuous helical columns which run parallel to the c-axis. This structure established the formation of two β-turns of types II and III′ respectively for sequences containing two consecutive ΔPhe residues at (i+2) and (i+3) positions with a branched β-carbon residue at one end of the tetrapeptide.  相似文献   

4.
Ventilation frequency (FV) in motionless common sole Solea solea was measured before and after a startling stimulus in normoxia and in hypoxia (15% air saturation). Startling reduced FV in normoxia (from mean ±s.e. 41 ± 3·3 beats min?1 to near zero, i.e. 2·0 ± 1·8 beats min?1) and in hypoxia (from mean ±s.e. 80 ± 4·4 to 58·8 ± 12·9 beats min?1). It is suggested that the maintenance of high FV in hypoxia may increase the probability of detection by predators compared to normoxia.  相似文献   

5.
Abstract: High-affinity μ-opioid receptors have been solubilized from rat brain membranes. In most experiments, rats were treated for 14 days with naltrexone to increase the density of opioid receptors in brain membranes. Occupancy of the membrane-associated receptors with morphine during solubilization in the detergent 3-[(3-cholamidopropyl)dimethyl]-1-propane sulfonate appeared to stabilize the μ-opioid receptor. After removal of free morphine by Sephadex G50 chromatography and adjustment of the 3-[(3-cholamidopropyl)dimethyl]-1-propane sulfonate concentration to 3 mM, the solubilized opioid receptor bound [3H][d -Ala2,N-Me-Phe4,Gly-ol5]-enkephalin ([3H]DAMGO), a μ-selective opioid agonist, with high affinity (KD = 1.90 ± 0.93 nM; Bmax = 629 ± 162 fmol/mg of protein). Of the membrane-associated [3H]-DAMGO binding sites, 29 ± 7% were recovered in the solubilized fraction. Specific [3H]DAMGO binding was completely abolished in the presence of 10 µM guanosine 5′-O-(3-thiotriphosphate). The solubilized receptor also bound [3H]diprenorphine, a nonselective opioid antagonist, with high affinity (KD = 1.4 ± 0.39 nM, Bmax = 920 ± 154 fmol/mg of protein). Guanosine 5′-O-(3-thiotriphosphate) did not diminish [3H]diprenorphine binding. DAMGO at concentrations between 1 nM and 1 µM competed with [3H]diprenorphine for the solubilized binding sites; in contrast, [d -Pen2,d -Pen5]-enkephalin, a δ-selective opioid agonist, and U50488H, a κ-selective opioid agonist, failed to compete with [3H]diprenorphine for the solubilized binding sites at concentrations of <1 µM. In the absence of guanine nucleotides, the DAMGO displacement curve for [3H]diprenorphine binding sites better fit a two-site than a one-site model with KDhigh = 2.17 ± 1.5 nM, Bmax = 648 ± 110 fmol/mg of protein and KDlow = 468 ± 63 nM, Bmax = 253 ± 84 fmol/mg of protein. In the presence of 10 µM guanosine 5′-O-(3-thiotriphosphate), the DAMGO displacement curve better fit a one- than a two-site model with KD = 815 ± 33 nM, Bmax = 965 ± 124 fmol/mg of protein.  相似文献   

6.
N-acetylglucosaminyl transferases from pupae of Stomoxys calcitrans (L.) were studied in 10,000g pellet suspensions. Characterization of these enzymes was based on formation of glycolipids (ie, Dol·PP-GlcNAc and Dol·PP-(GlcNAc)2), oligosaccharide lipids, and giycoproteins. Studies on transferase activity during the pupal instar showed that there were two peaks of activity; the first peak was on day 0 (prepupae) and the second at 3 days after pupation. Subcellular fractionation indicated that 10,000g and 100,000g pellets contained most of the transferase activities. The transferases required divalent cations (either Mn2+ or Mg2+). The pH optimum, which varied for each of the products formed, was 7.5 for glycolipids, 7.0 for oligosaccharide lipids, and 6.5 for glycoprotein. Inclusion of dolichol monophosphate doubled the amount of Dol·PP-GlcNAc and Dol·PP·(GlcNAc)2 formed, but had little effect on oligosaccharide lipid and glycoprotein formation. Tunicamycin was a potent inhibitor of glycolipid formation with an I50 of 1.8–4.8 nM. It was confirmed that tunicamycin acts by preventing the transfer of GlcNAc-1-P from UDP-GlcNAc to Dol·P. UMP reverses glycolipid formation, yielding UDP-GlcNAc. Some characterization of the products was performed. Glycolipids were shown to be Dol·PP-GlcNAc and Dol·PP-(GlcNAc)2. Glycoprotein was rapidly solubilized by protease and detergent treatments, whereas oligosaccharide lipids appeared to be acid-labile, pyrophosphate-containing lipids. The apparent kinetic constants for the formation of glycolipids were as follows: UDP-GlcNAc Km = 1.55 ± 0.47 μM, Vmax = 0.66 ± 0.21 pmol·min?1·mg?1; Dol·P Km = 2.08 ± 0.85 μM, Vmax = 0.13 ± 0.06 pmol·min?1·mg?1 protein.  相似文献   

7.
Abstract: We have cloned and expressed a rat brain cDNA, TS11, that encodes a μ-opioid receptor based on pharmacological, physiological, and anatomical criteria. Membranes were prepared from COS-7 cells transiently expressing TS11 bound [3H]diprenorphine with high affinity (KD = 0.23 ± 0.04 nM). The rank order potency of drugs competing with [3H]diprenorphine was as follows: levorphanol (Ki = 0.6 ± 0.2 nM) ≈β-endorphin (Ki = 0.7 ± 0.5 nM) ≈ morphine (Ki = 0.8 ± 0.5 nM) ≈ [d -Ala2, N-Me-Phe4,Gly-ol5]-enkephalin (DAMGO; Ki = 1.6 ± 0.5 nM) ? U50,488 (Ki = 910 ± 0.78 nM) > [d -Pen2,5]-enkephalin (Ki = 3,170 ± 98 nM) > dextrorphan (Ki = 4,100 ± 68 nM). The rank order potencies of these ligands, the stereospecificity of levorphanol, and morphine's subnanomolar Ki are consistent with a μ-opioid binding site. Two additional experiments provided evidence that this opioid-binding site is functionally coupled to G proteins: (a) In COS-7 cells 50 µM 5′-guanylylimidodiphosphate shifted a fraction of receptors with high affinity for DAMGO (IC50 = 3.4 ± 0.5 nM) to a lower-affinity state (IC50 = 89.0 ± 19.0 nM), and (b) exposure of Chinese hamster ovary cells stably expressing the cloned μ-opioid receptor to DAMGO resulted in a dose-dependent, naloxone-sensitive inhibition of forskolin-stimulated cyclic AMP production. The distribution of mRNA corresponding to the μ-opioid receptor encoded by TS11 was determined by in situ hybridization to brain sections prepared from adult female rats. The highest levels of μ-receptor mRNA were detected in the thalamus, medial habenula, and the caudate putamen; however, significant hybridization was also observed in many other brain regions, including the hypothalamus.  相似文献   

8.
Marine phytoplankton and macroalgae acquire important resources, such as inorganic nitrogen, from the surrounding seawater by uptake across their entire surface area. Rates of ammonium and nitrate uptake per unit surface area were remarkably similar for both marine phytoplankton and macroalgae at low external concentrations. At an external concentration of 1 μM, the mean rate of nitrogen uptake was 10±2 nmol·cm?2·h?1 (n=36). There was a strong negative relationship between log surface area:volume (SA:V) quotient and log nitrogen content per cm2 of surface (slope=?0.77), but a positive relationship between log SA:V and log maximum specific growth rate (μmax; slope=0.46). There was a strong negative relationship between log SA:V and log measured rate of ammonium assimilation per cm2 of surface, but the slope (?0.49) was steeper than that required to sustain μmax (?0.31). Calculated rates of ammonium assimilation required to sustain growth rates measured in natural populations were similar for both marine phytoplankton and macroalgae with an overall mean of 6.2±1.4 nmol·cm?2·h?1 (n=15). These values were similar to maximum rates of ammonium assimilation in phytoplankton with high SA:V, but the values for algae with low SA:V were substantially less than the maximum rate of ammonium assimilation. This suggests that the growth rates of both marine phytoplankton and macroalgae in nature are often constrained by rates of uptake and assimilation of nutrients per cm2 surface area.  相似文献   

9.
Increase in 4‐hydroxy‐2‐nonenal (4HNE) due to oxidative stress has been observed in a variety of cardiac diseases such as diabetic cardiomyopathy. 4HNE exerts a damaging effect in the myocardium by interfering with subcellular organelles like mitochondria by forming adducts. Therefore, we hypothesized that increased 4HNE adduct formation in the heart results in proteasome inactivation in isoproterenol (ISO)‐infused type 1 diabetes mellitus (DM) rats. Eight‐week‐old male Sprague Dawley rats were injected with streptozotocin (STZ, 65 mg kg?1). The rats were infused with ISO (5 mg kg?1) for 2 weeks by mini pumps, after 8 weeks of STZ injection. We studied normal control (n = 8) and DM + ISO (n = 10) groups. Cardiac performance was assessed by echocardiography and Millar catheter at the end of the protocol at 20 weeks. Initially, we found an increase in 4HNE adducts in the hearts of the DM + ISO group. There was also a decrease in myocardial proteasomal peptidase (chymotrypsin and trypsin‐like) activity. Increases in cardiomyocyte area (446 ± 32·7 vs 221 ± 10·83) (µm2), per cent area of cardiac fibrosis (7·4 ± 0·7 vs 2·7 ± 0·5) and cardiac dysfunction were also found in DM + ISO (P < 0·05) relative to controls. We also found increased 4HNE adduct formation on proteasomal subunits. Furthermore, reduced aldehyde dehydrogenase 2 activity was observed in the myocardium of the DM + ISO group. Treatment with 4HNE (100 μM) for 4 h on cultured H9c2 cardiomyocytes attenuated proteasome activity. Therefore, we conclude that the 4HNE‐induced decrease in proteasome activity may be involved in the cardiac pathology in STZ‐injected rats infused with ISO. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

10.
The effects of CO2 enrichment on photosynthesis and ribulose‐1,5‐bisphosphate carboxylase/oxygenase (rubisco) were studied in current year and 1‐year‐old needles of the same branch of field‐grown Pinus radiata D. Don trees. All measurements were made in the fourth year of growth in large, open‐top chambers continuously maintained at ambient (36 Pa) or elevated (65 Pa) CO2 partial pressures. Photosynthetic rates of the 1‐year‐old needles made at the growth CO2 partial pressure averaged 10·5 ± 0·5 μmol m?2 s?1 in the 36 Pa grown trees and 11·8 ± 0·4 μmol m?2 s?1 in the 65 Pa grown trees, and were not significantly different from each other. The photosynthetic capacity of 1‐year‐old needles was reduced by 25% from 23·0 ± 1·8 μmol m?2 s?1 in the 36 Pa CO2 grown trees to 17·3 ± 0·7 μmol m?2 s?1 in the 65 Pa grown trees. Growth in elevated CO2 also resulted in a 25% reduction in Vcmax (maximum carboxylation rate), a 23% reduction in Jmax (RuBP regeneration capacity mediated by maximum electron transport rate) and a 30% reduction in Rubisco activity and content. Total non‐structural carbohydrates (TNC) as a fraction of total dry mass increased from 12·8 ± 0·4% in 1‐year‐old needles from the 36 Pa grown trees to 14·2 ± 0·7% in 1‐year‐old needles from the 65 Pa grown trees and leaf nitrogen content decreased from 1·30 ± 0·02 to 1·09 ± 0·10 g m?2. The current‐year needles were not of sufficient size for gas exchange measurements, but none of the biochemical parameters measured (Rubisco, leaf chlorophyll, TNC and N), were effected by growth in elevated CO2. These results demonstrate that photosynthetic acclimation, which was not found in the first 2 years of this experiment, can develop over time in field‐grown trees and may be regulated by source‐sink balance, sugar feedback mechanisms and nitrogen allocation.  相似文献   

11.
The characteristics of the enzyme γ-glutamyltranspeptidase were determined in frog liver and compared to those of the rat. In Rana pipiens, tissue distribution studies indicated the order of activity to be: kidney >>> liver >> nerve > egg > lung > heart > skeletal muscle in homogenates. In the Rana pipiens relative to the Fischer 344 rat, the activity of the liver enzyme was somewhat greater (1·8-fold) and the kidney enzyme substantially less (25-fold). Frog liver γ-glutamyltranspeptidase displayed strain-dependent differences in activity with Rana pipiens and Rana sylvatica exhibiting comparable activities and Xenopus laevis exhibiting 20-fold lower activities. No influence of sex was apparent in Rana pipiens in contrast to the sex dependent differences observed in the Fischer 344 rat: ♀ : ♂ = 7:1. In homogenates and plasma membrane fractions of Rana pipiens, Xenopus laevis and the Fischer 344 rat, high, and comparable relative specific activities, were observed, 8–11, coupled with protein yields of 2·2–2·5 per cent indicating the enzyme to be plasma membrane bound and associated with the sinusoidal surface of the liver cell. Both the frog Rana pipiens and Xenopus laevis and Fischer 344 rat liver plasma membrane enzymes displayed comparable temperature-induced activation (1·51–1·74-fold) but with a peak for the frogs at 60°C and for the rat at 50°C. Both Acivicin and maleate inhibited the liver plasma membrane γ-glutamyltranspeptidase of both Rana pipiens and the Fischer 344 rat, but the frog enzyme was less sensitive (89 per cent decrease versus 97 per cent decrease) to 150 μM Acivicin and more sensitive (65 per cent decrease versus 35 per cent decrease at 150 mM maleate) to maleate. Kinetic studies indicated that the liver plasma membrane enzyme from Rana pipiens had a Km of 0·61 mM and Vmax of 55·6 nmol mg?1 min?1 and that from the Fischer 344 rat had a Km of 3·57 mM and Vmax of 71·4 nmol mg?1 min?1.  相似文献   

12.
Uptake rates of dissolved inorganic phosphorus and dissolved inorganic nitrogen under unsaturated and saturated conditions were studied in young sporophytes of the seaweeds Saccharina latissima and Laminaria digitata (Phaeophyceae) using a “pulse‐and‐chase” assay under fully controlled laboratory conditions. In a subsequent second “pulse‐and‐chase” assay, internal storage capacity (ISC) was calculated based on VM and the parameter for photosynthetic efficiency Fv/Fm. Sporophytes of S. latissima showed a VS of 0.80 ± 0.03 μmol · cm?2 · d?1 and a VM of 0.30 ± 0.09 μmol · cm?2 · d?1 for dissolved inorganic phosphate (DIP), whereas VS for DIN was 11.26 ± 0.56 μmol · cm?2 · d?1 and VM was 3.94 ± 0.67 μmol · cm?2 · d?1. In L. digitata, uptake kinetics for DIP and DIN were substantially lower: VS for DIP did not exceed 0.38 ± 0.03 μmol · cm?2 · d?1 while VM for DIP was 0.22 ± 0.01 μmol · cm?2 · d?1. VS for DIN was 3.92 ± 0.08 μmol · cm?2 · d?1 and the VM for DIN was 1.81 ± 0.38 μmol · cm?2 · d?1. Accordingly, S. latissima exhibited a larger ISC for DIP (27 μmol · cm?2) than L. digitata (10 μmol · cm?2), and was able to maintain high growth rates for a longer period under limiting DIP conditions. Our standardized data add to the physiological understanding of S. latissima and L. digitata, thus helping to identify potential locations for their cultivation. This could further contribute to the development and modification of applications in a bio‐based economy, for example, in evaluating the potential for bioremediation in integrated multitrophic aquacultures that produce biomass simultaneously for use in the food, feed, and energy industries.  相似文献   

13.
Aims: To analyse the production of different metabolites by dark‐grown Euglena gracilis under conditions found to render high cell growth. Methods and Results: The combination of glutamate (5 g l?1), malate (2 g l?1) and ethanol (10 ml l?1) (GM + EtOH); glutamate (7·15 g l?1) and ethanol (10 ml l?1); or malate (8·16 g l?1), glucose (10·6 g l?1) and NH4Cl (1·8 g l?1) as carbon and nitrogen sources, promoted an increase of 5·6, 3·7 and 2·6‐fold, respectively, in biomass concentration in comparison with glutamate and malate (GM). In turn, the production of α‐tocopherol after 120 h identified by LC‐MS was 3·7 ± 0·2, 2·4 ± 0·1 and 2 ± 0·1 mg [g dry weight (DW)]?1, respectively, while in the control medium (GM) it was 0·72 ± 0·1 mg (g DW)?1. For paramylon synthesis, the addition of EtOH or glucose induced a higher production. Amino acids were assayed by RP‐HPLC; Tyr a tocopherol precursor and Ala an amino acid with antioxidant activity were the amino acids synthesized at higher concentration. Conclusions: Dark‐grown E. gracilis Z is a suitable source for the generation of the biotechnologically relevant metabolites tyrosine, α‐tocopherol and paramylon. Significance and Impact of the Study: By combining different carbon and nitrogen sources and inducing a tolerable stress to the cell by adding ethanol, it was possible to increase the production of biomass, paramylon, α‐tocopherol and some amino acids. The concentrations of α‐tocopherol achieved in this study are higher than others reported previously for Euglena, plant and algal systems. This work helps to understand the effect of different carbon sources on the synthesis of bio‐molecules by E. gracilis and can be used as a basis for future works to improve the production of different metabolites of biotechnological importance by this organism.  相似文献   

14.
Synechococcus R-2 (PCC 7942) actively accumulated Cl? in the light and dark, under control conditions (BG-11 media: pHo, 7·5; [Na+]o, 18 mol m?3; [Cl?]o, 0·508 molm?3). In BG-11 medium [Cl?], was 17·2±0·848 mol m?3 (light), electrochemical potential of Cl? (ΔμCl?i,o) =+211±2mV; [Cl?]i= 1·24±0·11 mol m?3(dark), ΔμCl?i,o=+133±4mV. Cl? fluxes, but not permeabilities, were much higher in the light: ?Cl?i,o= 4·01±5·4 nmol m?2 s?1, PCl?i,o= 47±5pm s?1 (light); ?Cl?i,o= 0·395±0·071 nmol m?2 s?1, PCl?i,o= 69±14 pm s?1 (dark). Chloride fluxes are inhibited by acid pHo (pHo 5; ?Cl?i,o= 0·14±0·04 nmol m?2 s?1); optimal at pHo 7·5 and not strongly inhibited by alkaline pHo (pHo 10; ?Cl?1i,o= 1·7±0·14 nmol m?2 s?1). A Cl?in/2H+in coporter could not account for the accumulation of Cl? alkaline pHo. Permeability of Cl? is very low, below 100pm s?1 under all conditions used, and appears to be maximal at pHo 7·5 (50–70 pm s?1) and minimal in acid pHo (20pm s?1). DCCD (dicyclohexyl-carbodiimide) inhibited ?Cl?i,o in the light about 75% and [Cl?]i fell to 2·2±0·26 (4) mol m?3. Valinomycin had no effect but monensin severely inhibited Cl? uptake ([Cl?]i= 1·02±0·32 mol m?3; ?Cl?i,o= 0·20±0·1 nmol m?2 s?1). Vanadate (200 mmol m?3) accelerated the Cl? flux (?Cl?i,o= 5·28±0·64 nmol m?2 s?1) but slightly decreased accumulation of Cl? ([Cl?], = 13·9±1·3 mol m?3) in BG-11 medium but had no significant effect in Na+-free media. DCMU (dichlorophenyldimethylurea) did not reduce [Cl?], or ?Cl?i,o to that found in the dark ([Cl?]i= 8·41±0·76 mol m?3; ?Cl?i,o= 2·06±0·36 nmol m?2 s?1). Synechococcus also actively accumulated Cl? in Na+-free media, [Cl?]i was lower but ΔΨi,o hyperpolarized in Na+-free media and so the ΔμCl?i,o was little changed ([Cl?]i= 7·98±0·698 mol m?3; ΔμCl?i,o=+203±3 mV). Net Cl? uptake was stimulated by Na+; Li+ acted as a partial analogue for Na+. Synechococcus has a Na+ activated Cl? transporter which is probably a primary 2Cl?/ATP pump. The Cl? pump is voltage sensitive. ΔμCl?i,o is directly proportional to ΔΨi,o(P»0·01%): ΔμCl?i,o= -1·487 (±0·102) ×ΔΨi,o, r= -0·983, n= 31. The ΔμCl?i,o increased (more positive) as the Δμi,o became more negative. The ΔμCl?i,o has no known function, but might provide a driving force for the uptake of micronutrients.  相似文献   

15.
N‐alkylated trans‐diaziridines are an intriguing class of compounds with two stereogenic nitrogen atoms which easily interconvert. In the course of our investigations of the nature of the interconversion process via nitrogen inversion or electrocyclic ring opening ring closure, we synthesized and characterized the three constitutionally isomeric diaziridines 1,2‐di‐n‐propyldiaziridine 1 , 1‐isopropyl‐2‐n‐propyldiaziridine 2 , and 1,2‐diisopropyldiaziridine 3 to study the influence of the substituents on the interconversion barriers. Enantiomer separation was achieved by enantioselective gas chromatography on the chiral stationary phase Chirasil‐β‐Dex with high separation factors α (1‐isopropyl‐2‐n‐propyldiaziridine: 1.18; 1, 2‐diisopropyldiaziridine: 1.24; 100°C 50 kPa He) for the isopropyl substituted diaziridines. These compounds showed pronounced plateau formation between 100 and 150°C, and peak coalescence at elevated temperatures. The enantiomerization barriers ΔG? and activation parameters ΔH? and ΔS? were determined by enantioselective dynamic gas chromatography (DGC) and direct evaluation of the elution profiles using the unified equation implemented in the software DCXplorer. Interestingly, 1‐isopropyl‐2‐n‐propyldiaziridine and 1,2‐diisopropyldiaziridine exhibit similar high interconversion barriers ΔG? (100°C) of 128.3 ± 0.4 kJ mol?1 and 129.8 ± 0.4 kJ mol?1, respectively, which indicates that two sterically demanding substituents do not substantially increase the barrier as expected for a distinct nitrogen inversion process. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

16.
Particulate fractions (10,000g) from pupae of Stomoxys calcitrans transfer [14C]-mannose from GDP-[14C]-mannose to dolichol monophosphate and proteins. Production of the mannosyl lipid was inhibited by Mn2+, UDP, GMP, GDP, and EDTA. The insect growth regulator diflubenzuron had no effect on mannosyl transferase activity. Dolichol monophosphate and Mg2+ stimulated mannosyl transferase activity. The mannosyl lipid product was identified as mannosyl-phosphoryl-dolichol (Man-P-Dol). The apparent Km and Vmax values for the formation of Man-P-Dol using GDP-[14C]-Man while holding dolichol phosphate constant were 2.4 ± 0.9 μM and 9.4 ± 2.3 pmol Man-P-Dol·min?1·mg?1 protein, respectively. The apparent Km and Vmax values using dólichol phosphate while holding GDP-Man constant were 2.2 ± 1.2 μM and 18.5 ± 1.7 pmol Man-P-Dol·min?1·mg?1 protein.  相似文献   

17.
The recovery of oxygen uptake to the standard metabolic rate (SMR) following exhaustive chasing exercise in Atlantic salmon Salmo salar parr occurred in three phases (rapid, plateau and slow). The initial recovery phase lasted 0·7 h and contributed 16% to the total excess post‐exercise oxygen consumption (EPOC). It was followed by a longer plateau phase that contributed 53% to the total EPOC. The slow recovery phase that completed recovery of SMR, which has not been reported previously, made a 31% contribution to the total EPOC. The plasticity of EPOC was demonstrated in exercise‐trained fish. Exercise training increased EPOC by 39% when compared with control fish (mean ± S.E., 877·7 ± 73·1 v . 629·2 ± 53·4 mg O2 kg?1, d.f. = 9, P <  0·05), with the duration of the plateau phase increasing by 38% (4·7 ± 0·58 v . 3·4 ± 0·16 h, d.f. = 9, P <  0·05) and the contribution of the slow phase to the total EPOC increasing by 80% (173·9 ± 23·9 v . 312·5 ± 50·4 mg O2 kg?1, d.f. = 9, P  < 0·05). As a result, the combination of the plateau and slow phases of exercise‐trained fish increased by 47% compared with control fish (756·6 ± 71·4 v . 513·6 ± 43·1 mg O2 kg?1; d.f. = 9, P  = 0·01). To substantiate the hypothesis that the plateau and slow recovery phase of EPOC was related to general metabolic recovery following exhaustive exercise, the time‐course for recovery of SMR was compared with previously published metabolite recovery profiles. The final phase of metabolic recovery was temporally associated with the final phases of gluconeogenesis, lactate oxidation and muscle intracellular pH regulation. Therefore, the plasticity of the latter phase of EPOC agreed with the known effects of exercise training in fishes.  相似文献   

18.
Aims: A microbiological bioassay using Geoacillus stearothermophilus was optimized to detect betalactams at concentrations near to the Maximum Residue Limits (MRLs), with low cross‐specificity for tetracycline. Methods and Results: A factorial design (3 × 4) was used to evaluate the effects of concentration of spores (2·0 × 106, 4·0 × 106 and 8·0 × 106 spores ml?1) and incubation time (3·0, 3·5, 4·0 and 4·5 h) on the response of the bioassay. Then, desirability function to raise the detection capabilities (CCβ) of tetracyclines and increase sensitivity to betalactams was implemented. Significant effects of Log[S] and incubation time [It] on the CCβ of betalactams and tetracyclines were observed. Finally, high value of global desirability (D = 0·853), adequate betalactams CCβ (3·8 μg l?1 of penicillin ‘G’, 27 μg l?1 of oxacillin, 8·1 μg l?1 of ampicillin, 48 μg l?1 of cloxacillin) and high tetracyclines CCβ (5260 μg l?1 chlortetracycline, 1550 μg l?1 of oxytetracycline, 1070 μg l?1 of tetracycline) were calculated. Conclusions: The application of chemometric tools allows the optimization of a bioassay that detects betalactam residues in milk. The more robust conditions have been achieved in Log[S] = 6·30 and [It] = 4·20 h. Significance and Impact of the Study: The logistic regression model and the desirability function are adequate chemometric techniques to improve the properties of the methods, because it is possible to increase sensitivity and decrease cross‐specificity simultaneously.  相似文献   

19.
The stress response of Oncorhynchus mykiss in high‐altitude farms in central Mexico was investigated over two seasons: the cool (9·1–13·7° C) dry winter season, and the warmer (14·7–15·9° C), wetter summer season. Fish were subjected to an acute stress test followed by sampling of six physiological variables: blood cortisol, glucose, lactate, total antioxidant capacity, haemoglobin concentration and per cent packed cell volume (VPC%). Multivariate analyses revealed that lactate and total antioxidant capacity were significantly higher in the summer, when water temperatures were warmer and moderate hypoxia (4·9–5·3 mg l?1) prevailed. In contrast, plasma cortisol was significantly higher in the winter (mean ± s.e .: 76·7 ± 4·0 ng ml?1) when temperatures were cooler and dissolved oxygen levels higher (6·05–7·9 mg l?1), than in the summer (22·7 ± 3·8 ng ml?1). Haemoglobin concentrations (mg dl?1) were not significantly different between seasons, but VPC% was significantly higher in the summer (50%) than in the winter (35%). These results suggest that in summer, effects of high altitude on farmed fish are exacerbated by stresses of high temperatures and hypoxia, resulting in higher blood lactate, increased total antioxidant capacity and elevated VPC% levels.  相似文献   

20.
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