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 31 毫秒
1.
Small-angle neutron scattering (SANS) experiments were performed on unilamellar 1,2-dimyristoylphosphatidylcholine (DMPC) vesicles prepared in heavy water by extrusion through polycarbonate filters with 500 Å pores. The data obtained at 30±0.1 °C were evaluated using a five-strip function model of the bilayer coherent neutron scattering length density, three different approximate form factors describing scattering from vesicles, and different methods of evaluation of the experimental data. It is shown that the results obtained from the SANS data in the range of scattering vector values 0.0316 Å–1<q<0.0775 Å–1 are not sensitive to the vesicle form factor, nor to the evaluation method. Using the hollow sphere model of vesicles convoluted with the Gaussian distribution of their sizes, a constrained bilayer polar region thickness of 9 Å and a DMPC headgroup volume of 325.5 Å3, it was possible to obtain from the experimental data the DMPC surface area as 58.9±0.8 Å2, the bilayer thickness as 44.5±0.3 Å and the number of water molecules as 6.8±0.2 per DMPC located in the bilayer polar region.  相似文献   

2.
The translational diffusion coefficient D 20,w 0 , of monomeric human immunoglobulin G (IgG) has been studied by photon-correlation spectroscopy as a function of pH and protein concentration. At pH 7.6, we find D 20,w 0 =3.89×10–7±0.02 cm2/sec, in good agreement with the value determined by classic mehods. This value corresponds to an effective hydrodynamic radius R, of 55.1±0.3 Å. As pH is increased to 8.9; with the same ionic strength, the molecule appears to expand slightly (3.5% increase in hydrodynamic radius). The concentration dependence of the IgG diffusion constant is interpreted in terms of solution electrostatic effects and shows that long-range repulsive interactions are negligible in the buffer used. The diffusion coefficient for dimeric IgG has also been determined to be D20,w=2.81×10–7±0.04 cm2/sec at 1.6 mg/ml, which corresponds to a hydrodynamic radius of 75 Å. For light-scattering studies of protein molecules in the dimension range of 5–10 nm (Mr=105–107) we find monomeric horse spleen ferritin well suited as a reference standard. Ferritin is a spherical molecule with a hydrodynamic radius R of 6.9±0.1 nm and is stable for years in our standard Tris-HCl-NaCl buffer even at room temperature.  相似文献   

3.
Summary A new purification procedure for endo-\-1,3-1,4-d-glucanase from Bacillus licheniformis is described. The secreted enzyme was purified both from B. licheniformis and from recombinant Escherichia coli harbouring the cloned gene by ion exchange chromatography on a CM-Sepharose matrix at pH 5.6. The mature enzyme was resistant to proteolysis by trypsin and chymotrypsin but it was slowly digested by protease V8. It showed a continuous trimming where no large-limit polypeptides were noticeable thus supporting a monodomain structure. Former appearing peptides have been assigned theoretically according to the protein sequence and predictive methods of accessible areas. Kinetic parameters for the hydrolysis of barley \-glucan and lichenan by measuring the net release of reducing sugars at the optimum pH (7.02) and temperature (55° C) are k cat=3500 ±800 s–1 (turnover number) and K m=1.45±0.21 mg/ml for barley \-glucan and k cat=3000±750 s–1 and K m=1.98±0.40 mg/ml for lichenan. Correspondence to: E. Querol  相似文献   

4.
Summary The effect of chloride on 4,4-dibenzamido-2,2-disulfonic stilbene (DBDS) binding to band 3 in unsealed red cell ghost membranes was studied in buffer [NaCl (0 to 500mm) + Na citrate] at constant ionic strength (160 or 600mm). pH 7.4, 25°C. In the presence of chloride, DBDS binds to a single class of sites on band 3. At 160mm ionic strength, the dissociation constant of DBDS increases linearly with chloride concentration in the range [Cl]=450mm. The observed rate of DBDS binding to ghost membranes, as measured by fluorescence stopped-flow kinetic experiments, increases with chloride concentration at both 160 and 600mm ionic strength. The equilibrium and kinetic results have been incorporated into the following model of the DBDS-band 3 interaction: The equilibrium and rate constants of the model at 600mm ionic strength areK 1=0.67±0.16 m,k 2=1.6±0.7 sec–1,k –2=0.17±0.09 sec–1,K 1=6.3±1.7 m,k 2=9±4 sec–1 andk –2=7±3 sec–1. The apparent dissociation constants of chloride from band 3,K Cl, are 40±4mm (160mm ionic strength) and 11±3mm (600mm ionic strength). Our results indicate that chloride and DBDS have distinct, interacting binding sites on band 3.  相似文献   

5.
Sequence analysis of the metabolically rich 8.7-Mbp genome of the model actinomycete Streptomyces coelicolor A3(2) revealed three genes encoding predicted type III polyketide synthases (PKSs). We report the inactivation, expression, and characterization of the type III PKS homologous SCO1206 gene product as 1,3,6,8-tetrahydroxynaphthalene synthase (THNS). Incubation of recombinant THNS with malonyl-CoA showed THN production, as demonstrated by UV and HPLC analyses. The Km value for malonyl-CoA and the kcat value for THN synthesis were determined spectrophotometrically to be 3.58±0.85 µM and 0.48±0.03 min–1, respectively. The C-terminal region of S. coelicolor THNS, which is longer than most other bacterial and plant type III PKSs, was shortened by 25 amino acid residues and the resulting mutant was shown to be slightly more active (Km=1.97±0.19 µM, kcat=0.75±0.04 min–1) than the wild-type enzyme.  相似文献   

6.
Summary Cell K activity,a k, was measured in the short-circuited frog skin by simultaneous cell punctures from the apical surface with open-tip and K-selective microelectrodes. Strict criteria for acceptance of impalements included constancy of the open-tip microelectrode resistance, agreement within 3% of the fractional apical voltage measured with open-tip and K-selective microelectrodes, and constancy of the differential voltage recorded between the open-tip and the K microelectrodes 30–60 sec after application of amiloride or substitution of apical Na. Skins were bathed on the serosal surface with NaCl Ringer and, to reduce paracellular Cl conductance and effects of amiloride on paracellular conductance, with NaNO3 Ringer on the apical surface.Under control conditionsa k r was nearly constant among skins (mean±SD=92±8mM, 14 skins) in spite of a wide range of cellular currents (5 to 70 A/cm2). Cell current (and transcellular Na transport) was inhibited by either apical addition of amiloride or substitution of Na by other cations. Although in some experiments the expected small increase ina k r after inhibition of cell current was observed, on the average the change was not significant (98±11mM after amiloride, 101±12mM after Na substitution), even 30 min after the inhibition of cell current. The membrane potential, which in the control state ranged from –42 to –77 mV, hyperpolarized after inhibition of cell current, initially to –109±5mV, then depolarizing to a stable value (–88±5mV) after 15–25 min. At this time K was above equilibrium (E k=98±2mV), indicating that the active pump mechanism is still operating after inhibition of transcellular Na transport.The measurement ofa k r permitted the calculation of the passive K current and pump current under control conditions. assuming a constant current source with almost all of the basolateral conductance attributable to K. We found a significant correlation between pump current and cell current with a slope of 0.31, indicating that about one-third of the cell current is carried by the pump, i.e., a pump stoichiometry of 3Na/2K.  相似文献   

7.
This study calculated the compositional nutrient diagnosis (CND) norms of cowpea (Vigna unguiculata (L.) Walp.), as well as identified significant nutrient interactions of this crop growing in an irrigated calcareous desert soil. Three genotypes were distributed in rows in a 2-ha field. The soil showed high heterogeneity in its chemical properties. For statistical analysis, 86 foliar composite samples from healthy plants were used. Preliminary CND norms were developed using a cumulative variance ratio function and the 2 distribution function. Means and standard deviations of row-centered log ratios VX of five nutrients (N, P, K, Ca, and Mg) and a filling value R, which included all nutrients not chemically analyzed. Preliminary CND norms are: VN*=0.174±0.095, VP*=–2.172±0.234, VK*=–0.007±0.267, VCa*=–0.022±0.146, VMg*=–1.710±0.132, and VR5*=3.728±0.084. These CND norms are associated with dry bean yields higher than 1.88 t ha–1, and are associated with the following foliar concentrations: 26.2 g N kg–1, 2.5 g P kg–1, 22.9 g K kg–1, 21.6 g Ca kg–1, and 4 g Mg kg–1. Cowpea plants growing in desert calcareous soils took up lower amounts of N, P, and K than those considered as optimum in a previous report. Six interactions were strongly indicated for cowpea through principal component analyses: positive for Ca–Mg, and negative for N–Ca, N–Mg, Ca–P, Mg–P, and K–P. Furthermore, two interactions were identified using simple correlations, negative N–P and positive K–Ca.  相似文献   

8.
Summary We have measured the intracellular potassium activity, [K+]i and the mechanisms of transcellular K+ transport in reabsorptive sweat duct (RSD) using intracellular ion-sensitive microelectrodes (ISMEs). The mean value of [K+]i in RSD is 79.8±4.1mm (n=39). Under conditions of microperfusion, the [K+]i is above equilibrium across both the basolateral membrane, BLM (5.5 times) and the apical membrane, APM (7.8 times). The Na+/K+ pump inhibitor ouabain reduced [K+]i towards passive distribution across the BLM. However, the [K+]i is insensitive to the Na+/K+/2 Cl cotransport inhibitor bumetanide in the bath. Cl substitution in the lumen had no effect on [K+]i. In contrast, Cl substitution in the bath (basolateral side) depolarized BLM from –26.0±2.6 mV to –4.7*±2.4 mV (n=3;* indicates significant difference) and decreased [K+]i from 76.0±15.2mm to 57.7* ±12.7mm (n=3). Removal of K+ in the bath decreased [K+]i from 76.3±15.0mm to 32.3*±7.6mm (n=4) while depolarizing the BLM from –32.5±4.1 mV to –28.3*±3.0 mV (n=4). Raising the [K+] in the bath by 10-fold increased [K+]i from 81.7±9.0mm to 95.0*±13.5mm and depolarized the BLM from –25.7±2.4 mV to –21.3*±2.9 mV (n=4). The K+ conductance inhibitor, Ba2+, in the bath also increased [K+]i from 85.8±6.7mm to 107.0*±11.5mm (n=4) and depolarized BLM from –25.8±2.2 mV to –17.0*±3.1 mV (n=4). Amiloride at 10–6 m increased [K+]i from 77.5±18.8mm to 98.8*±21.6mm (n=4) and hyperpolarized both the BLM (from –35.5±2.6 mV to –47.8*±4.3 mV) and the APM (from –27.5±1.4 mV to –46.0* ±3.5 mV,n=4). However, amiloride at 10–4 m decreased [K+]i from 64.5±0.9mm to 36.0*±9.9mm and hyperpolarized both the BLM (from –24.7±1.4 mV to –43.5*±4.2 mV) and APM (from –18.3±0.9 mV to –43.5*±4.2 mV,n=6). In contrast to the observations at the BLM, substitution of K+ or application of Ba2+ in the lumen had no effect on the [K+]i or the electrical properties of RSD, indicating the absence of a K+ conductance in the APM. Our results indicate that (i) [K+]i is above equilibrium due to the Na+/K+ pump; (ii) only the BLM has a K+ conductance; (iii) [K+]i is subject to modulation by transport status; (iv) K+ is probably not involved in carrier-mediated ion transport across the cell membranes; and (v) the RSD does not secrete K+ into the lumen.  相似文献   

9.
Cobo  J. G.  Barrios  E.  Kass  D. C. L.  Thomas  R. J. 《Plant and Soil》2002,240(2):331-342
The decomposition and nutrient release of 12 plant materials were assessed in a 20-week litterbag field study in hillsides from Cauca, Colombia. Leaves of Tithonia diversifolia (TTH) and Indigofera constricta (IND) decomposed quickly (k=0.035±0.002 d–1), while those of Cratylia argentea (CRA) and the stems evaluated decomposed slowly (k=0.007±0.002 d–1). Potassium presented the highest release rates (k>0.085 d–1). Rates of N and P release were high for all leaf materials evaluated (k>0.028 d–1) with the exception of CRA (N and P), TTH and IND (P). While Mg release rates ranged from 0.013 to 0.122 d–1, Ca release was generally slower (k=0.008–0.041 d–1). Initial quality parameters that best correlated with decomposition (P>0.001) were neutral detergent fibre, NDF (r=–0.96) and in vitro dry matter digestibility, IVDMD (r=0.87). It is argued that NDF or IVDMD could be useful lab-based tests during screening of plant materials as green manures. Significant correlations (P>0.05) were also found for initial quality parameters and nutrient release, being most important the lignin/N ratio (r=–0.71) and (lignin+polyphenol)/N ratios (r=–0.70) for N release, the C/N (r=0.70) and N/P ratios (r=–0.66) for P release, the hemicellulose content (r=–0.75) for K release, the Ca content (r=0.82) for Ca release, and the C/P ratio (r=0.65) for Mg release. After 20 weeks, the leaves of Mucuna deerengianum released the highest amounts of N and P (144.5 and 11.4 kg ha–1, respectively), while TTH released the highest amounts of K, Ca and Mg (129.3, 112.6 and 25.9 kg ha–1, respectively). These results show the potential of some plant materials studied as sources of nutrients in tropical hillside agroecosystems.  相似文献   

10.
Photosynthetic complexes in bacteria absorb light and undergo photochemistry with high quantum efficiency. We describe the isolation of a highly purified, active, reaction center-light-harvesting 1–PufX complex (RC–LH1–PufX core complex) from a strain of the photosynthetic bacterium, Rhodobacter sphaeroides, which lacks the light-harvesting 2 (LH2) and contains a 6 histidine tag on the H subunit of the RC. The complex was solubilized with diheptanoyl-sn-glycero-3-phosphocholine (DHPC), and purified by Ni-affinity, size-exclusion and ion-exchange chromatography in dodecyl maltoside. SDS-PAGE analysis shows the complex to be highly purified. The quantum efficiency was determined by measuring the charge separation (DQA → D+QA-) in the RC as a function of light intensity. The RC–LH1–PufX complex had a quantum efficiency of 0.95 ± 0.05, indicating full activity. The stoichiometry of LH1 subunits per RC was determined by two independent methods: (i) solvent extraction and absorbance spectroscopy of bacteriochlorophyll, and (ii) density scanning of the SDS-PAGE bands. The average stoichiometry from the two measurements was 13.3 ± 0.9 LH1/RC. The presence of PufX was observed in SDS-PAGE gels at a stoichiometry of 1.1 ± 0.1/RC. Crystals of the core complex have been obtained which diffract X-rays to 12 Å. A preliminary analysis of the space group and unit cell analysis indicated a P1 space group with unit cell dimensions of a = 76.3 Å, b = 137.2 Å, c = 137.5 Å; α = 60.0°, β = 89.95°, γ =90.02°.  相似文献   

11.
Patel  R.  Yago  M.D.  Mañas  M.  Victoria  E.M.  Shervington  A.  Singh  J. 《Molecular and cellular biochemistry》2004,261(1):83-89
This study investigated the effects of cholecystokinin-octapeptide (CCK-8) on pancreatic juice flow and its contents, and on cytosolic calcium (Ca2+) and magnesium (Mg2+) levels in streptozotocin (STZ)-induced diabetic rats compared to healthy age-matched controls. Animals were rendered diabetic by a single injection of STZ (60 mg kg–1, I.P.). Age-matched control rats obtained an equivalent volume of citrate buffer. Seven weeks later, animals were either anaesthetised (1 g kg–1 urethane; IP) for the measurement of pancreatic juice flow or humanely killed and the pancreas isolated for the measurements of cytosolic Ca2+ and Mg2+ levels. Non-fasting blood glucose levels in control and diabetic rats were 92.40 ± 2.42 mg dl–1 (n= 44) and >500 mg dl–1 (n= 27), respectively. Resting (basal) pancreatic juice flow in control and diabetic anaesthetised rats was 0.56 ± 0.05 ul min–1 (n= 10) and 1.28 ± 0.16 ul min–1 (n= 8). CCK-8 infusion resulted in a significant (p < 0.05) increase in pancreatic juice flow in control animals compared to a much larger increase in diabetic rats. In contrast, CCK-8 evoked significant (p < 0.05) increases in protein output and amylase secretion in control rats compared to much reduced responses in diabetic animals. Basal [Ca2+]i in control and diabetic fura-2-loaded acinar cells was 109.40 ± 15.41 nM (n= 15) and 130.62 ± 17.66 nM (n= 8), respectively. CCK-8 (10–8M) induced a peak response of 436.55 ± 36.54 nM (n= 15) and 409.31 ± 34.64 nM (n= 8) in control and diabetic cells, respectively. Basal [Mg2+]i in control and diabetic magfura-2-loaded acinar cells was 0.96 ± 0.06 nM (n= 18) and 0.86 ± 0.04 nM (n= 10). In the presence of CCK-8 (10–8) [Mg2+]i in control and diabetic cells was 0.80 ± 0.05 nM (n= 18) and 0.60 ± 0.02 nM (n= 10), respectively. The results indicate that diabetes-induced pancreatic insufficiency may be associated with derangements in cellular Ca2+ and Mg2+ homeostasis. (Mol Cell Biochem 261: 83–89, 2004)  相似文献   

12.
Metabolic rate and evaporative water loss (EWL) were measured for a small, arid-zone marsupial, the stripe-faced dunnart (Sminthopsis macroura), when normothermic and torpid. Metabolic rate increased linearly with decreasing ambient temperature (Ta) for normothermic dunnarts, and calculated metabolic water production (MWP) ranged from 0.85±0.05 (Ta=30°C) to 3.13±0.22 mg H2O g–1 h–1 (Ta=11°C). Torpor at Ta=11 and 16°C reduced MWP to 24–36% of normothermic values. EWL increased with decreasing Ta, and ranged from 1.81±0.37 (Ta=30°C) to 5.26±0.86 mg H2O g–1 h–1 (Ta=11°C). Torpor significantly reduced absolute EWL to 23.5–42.3% of normothermic values, resulting in absolute water savings of 50–55 mg H2O h–1. The relative water economy (EWL/MWP) of the dunnarts was unfavourable, remaining >1 at all Ta investigated, and did not improve with torpor. Thus torpor in stripe-faced dunnarts results in absolute, but not relative, water savings.  相似文献   

13.
A biotinylated mannotriose (Man3-bio) was dispersively immobilized in the matrix of biotinylated lactose (Gal-Glc-bio) on a streptavidin-covered, 27-MHz quartz crystal microbalance (QCM), and binding kinetics of concanavalin A (Con A) to Man3-bio in the Gal-Glc-bio matrix could be obtained from frequency decreases (mass increases) of the QCM. Association constants (Ka) and binding and dissociation rate constants (kon and koff) could be determined separately as the 1:1 and 1:2 bindings of Con A to Man3-bio on the surface. When Man3-bio was immobilized with content of 1 to 5 mol% in the matrix, the 1:1 binding of Con A to Man3-bio was obtained as Ka = (4 ± 1) × 106 M−1, kon = (4 ± 1) × 104 M−1 s−1, and koff = (12 ± 2) × 10–3 s−1. On the contrary, when Man3-bio was immobilized with content of 20 to 100 mol% in the matrix, the 1:2 binding of Con A to Man3-bio was obtained as Ka = (14 ± 2) × 106 M−1, kon = (14 ± 2) × 104 M−1 s−1, and koff = (7 ± 2) × 10–3 s−1. Thus, Ka for the 1:2 binding was 10 times larger than that for the 1:1 binding, with a three times larger binding rate constant (kon) and a three times smaller dissociation rate constant (koff). This is the first example to obtain separate kinetic parameters for the 1:1 and 1:2 bindings of lectins to carbohydrates on the surface.  相似文献   

14.
The chain-melting transition temperature of dipalmitoyl phosphatidylcholine (DPPC) bilayer membranes containing poly(ethylene glycol)-grafted dipalmitoyl phosphatidylethanolamine (PEG-DPPE) was determined by optical turbidity measurements. The dependence on content, Xp, of PEG-DPPE lipid was studied for different polar headgroup sizes, np, of the polymer lipid, throughout the lamellar phase of the mixtures with DPPC. Mean-field theory for the polymer brush regime predicts that the downward shift in transition temperature should vary with polymer size and content as npXp5/3 (∼npXp11/6 for scaling theory). Any shift induced by the charge on PEG-lipids is independent of polymer size. These predictions are reasonably borne out for the longer polymer lipids (PEG molecular masses 750, 2000 and 5000 Da). Transition temperature shifts in the lamellar phase, before the onset of micellisation, are in the region of −1 to −2 °C (±0.1-0.2 °C) in reasonable accord with theoretical estimates of the lateral pressure exerted by the polymer brush. Shifts of this size are significant to the design of liposomes for controlled release of contents by mild hyperthermia.  相似文献   

15.
The relative affinities of various muscarinic drugs in the antagonist ([3H]N-methyl scopolamine ([3H]NMS)) and agonist ([3H]Oxotremorine-m ([3H]OXO-M)) binding assays using a mixture of tissues containing M1–M4 receptor subtypes have been determined. [3H]NMS bound with high affinity (Kd=25±5.9 pM; n=3) and to a high density (Bmax=11.8±0.025 nmol/g wet weight) of muscarinic receptors. [3H]OXO-M appeared to bind to two binding sites with differing affinities (Kd1=2.5±0.1 nM; Kd2=9.0±4.9 M; n=4) and to a different population of binding sites (Bmax1=5.0±0.26 nmol/g wet weight; Bmax2=130±60 nmol/g wet weight). Well known antagonists exhibited high affinity for [3H]NMS binding but a lower affinity for [3H]OXO-M binding. The opposite was true for acetylcholine and other known agonists. However, pilocarpine and McN-A-343 had similar affinities for sites labeled by both radioligands. Using the ratios of antagonist-to-agonist binding affinities, it was possible to group compounds into apparently distinct full agonist (ratios of 180–665; e.g. carbachol, muscarine, OXO-M, OXO-S and arecoline), partial agonist (ratios of 14–132; e.g. McN-A-343, pilocarpine, aceclidine, bethanechol, OXA-22 and acetylcholine) and antagonist (ratios of 0.22–1.9; e.g. atropine, NMS, pirenzepine, methoctramine, 4-DAMP and p-fluorohexahydrosialo-difenidol) classes. These data suggest that the NMS/OXO-M affinity ratios using a mixture of M1–M4 muscarinic receptors may be a useful way to screen and group a large number of compounds into apparent agonist, partial agonist, and antagonist classes of cholinergic agents.  相似文献   

16.
Summary 1. The purpose of this study was (a) to identify if astrocytes show a similar non-Nernstian depolarization in low K+ or low Ca2+ solutions as previously found in human glial and glioma cells, and (b) to analyze the influence of the K+ conductance on the membrane potential of astrocytes.2. The membrane potential (Em) and the ionic conductance were studied with whole-cell patch-clamp technique in neonatal rat astrocytes (5–9 days in culture) and in human glioma cells (U-251MG).3. In 3.0 mM K+, Em was –75 ± 1.0 mV (mean ± SEM,n=39) in rat astrocytes and –79 ± 0.7 mV (n=5) in U-251MG cells. In both cell types Em changed linearly to the logarithm of [K+]0 between 3.0 and 160 mM K+. K+ free medium caused astrocytes to hyperpolarize to –93 ± 2.7 mV (n=21) and U-251MG cells to depolarize to –27 ± 2.1 mV (n=3).4. The I-E curve did not show inward rectification in astrocytes at this developmental stage. The slope conductance (g) exhibited only a small decrease (–19%) in K+ free solution and no significant change in 160 mM K+.5. Ba2+ (1.0 mM) depolarized astrocytes to –45 ± 2.9 mV (n=11), decreasing the slope conductance (g) by 42.4 ± 8.3% (n=11). Ca2+ free solution depolarized astrocytes to –53 ± 3.4 mV (n=12) and resulted in a positive shift of the I-E curve, increasing g by 15.3 ± 8.2% (n=8).6. Calculations indicated that a block of K+ channels explains the depolarizing effect of Ba2+. The effects of K+ free or Ca2+ free solutions on Em can be explained by a transformation of K+ channels to non-specific leakage channels. That astrocytes show a different reaction to low K+ than glioma cells can be related to the lack of inwardly rectifying K+ channels in astrocytes at this developmental stage.  相似文献   

17.
We investigated the effects of various phospholipids on the presynaptic levels of newly synthesized [3H]acetylcholine (ACh) in rat cerebral cortical synaptosomes. When administered as small unilamellar vesicles (200–500 Å diameters) dipalmitoylphosphatidylcholine (DPPC) reduced [3H]ACh levels in concentration and time-related manners, while increasing the efflux of labelled choline to a similar extent. The reductions in synaptosomal [3H]-ACh levels induced by DPPC (3 mg/ml) were found in the cytosolic S3 but not microsomal P3 fraction, arguing for a cytoplasmic, nonvesicular site of action. DPPC-induced reductions in [3H]ACh levels were blocked by 100 M eserine, a tertiary amine cholinesterase inhibitor, but not with 100 M neostigmine, a quaternary ammonium inhibitor. Large unilamellar vesicles (2000–5000 Å diameters) consisting of soybean-phosphatidylcholine reduced [3H]ACh levels to the same extent that small vesicles did at the same concentration (3 mg/ml). Taken together, these results suggest that DPPC can fuse with membranes to increase the hydrolysis of cytoplasmic ACh via a small intra-terminal subpopulation of cholinesterases.  相似文献   

18.
Summary Intracellular pH (pHi) regulation was studied in crayfish neurons with pH-, and Na+-sensitive microelectrodes. It was confirmed to involve both a HCO 3 -dependent and a HCO 3 -independent mechanism. The latter was identified as the amiloride-sensitive Na+/H+ exchange described in vertebrate cells. Its dependence on extracellular pH (pHe) and Na+ concentration ([Na+]e) was studied in CO2-free external solutions at 20°C. The steady state pHi and the rate constant (k) of the exponential pHi recovery following an acid load were determined. At pHe=7.5 and [Na+]e=200 mM, the average steady state pHi was 7.09±0.12 (as compared to 7.30±0.10 in the presence of 5 mM bicarbonate). The dependence of the rate constant of recovery on [Na+]e could be described by Michaelis-Menten kinetics; at pHe=7.5 the apparentK m andK max were 39 mM and 1.4 mmol·l–1·min–1, respectively. Decreasing pHe reduced the rate of recovery, the variations ofk with pHe conforming to a simple titration curve with an apparent pK of 7.05±0.21. These kinetic properties of the Na+/H+ exchange in crayfish neurons are similar to those described in vertebrate cells.Preliminary results were presented at the First International Congress of Comparative Physiology and Biochemistry (Liège, Belgium, 1984)  相似文献   

19.
A fourth order Runge–Kutta approximation was used to determine the Monod kinetics of Candida rugopelliculosa by using unsteady state data from only one continuous unsteady state operation at a fixed dilution rate. The maximum microbial growth rates, max, and half saturation coefficient, K s, were 0.82 ± 0.22 h–1 and 690 ± 220 mg soluble chemical oxygen demand (SCOD) l–1, respectively. The microbial yield coefficient, Y, and microbial decay rate coefficient, k d, were 1.39 ± 0.22 × 104 cells mg–1 SCOD and 0.06 ± 0.01 h–1, respectively.  相似文献   

20.
Temperate forests of North America are thought to besignificant sinks of atmospheric CO2. Wedeveloped a below-ground carbon (C) budget forwell-drained soils in Harvard Forest Massachusetts, anecosystem that is storing C. Measurements of carbonand radiocarbon (14C) inventory were used todetermine the turnover time and maximum rate ofCO2 production from heterotrophic respiration ofthree fractions of soil organic matter (SOM):recognizable litter fragments (L), humified lowdensity material (H), and high density ormineral-associated organic matter (M). Turnover timesin all fractions increased with soil depth and were2–5 years for recognizable leaf litter, 5–10 years forroot litter, 40–100+ years for low density humifiedmaterial and >100 years for carbon associated withminerals. These turnover times represent the timecarbon resides in the plant + soil system, and mayunderestimate actual decomposition rates if carbonresides for several years in living root, plant orwoody material.Soil respiration was partitioned into two componentsusing 14C: recent photosynthate which ismetabolized by roots and microorganisms within a yearof initial fixation (Recent-C), and C that is respiredduring microbial decomposition of SOM that resides inthe soil for several years or longer (Reservoir-C).For the whole soil, we calculate that decomposition ofReservoir-C contributes approximately 41% of thetotal annual soil respiration. Of this 41%,recognizable leaf or root detritus accounts for 80%of the flux, and 20% is from the more humifiedfractions that dominate the soil carbon stocks.Measurements of CO2 and 14CO2 in thesoil atmosphere and in total soil respiration werecombined with surface CO2 fluxes and a soil gasdiffusion model to determine the flux and isotopicsignature of C produced as a function of soil depth. 63% of soil respiration takes place in the top 15 cmof the soil (O + A + Ap horizons). The average residencetime of Reservoir-C in the plant + soil system is8±1 years and the average age of carbon in totalsoil respiration (Recent-C + Reservoir-C) is 4±1years.The O and A horizons have accumulated 4.4 kgC m–2above the plow layer since abandonment by settlers inthe late-1800's. C pools contributing the most to soilrespiration have short enough turnover times that theyare likely in steady state. However, most C is storedas humified organic matter within both the O and Ahorizons and has turnover times from 40 to 100+ yearsrespectively. These reservoirs continue to accumulatecarbon at a combined rate of 10–30 gC mminus 2yr–1. This rate of accumulation is only 5–15% of the total ecosystem C sink measured in this stand using eddy covariance methods.  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