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.
A simple scaling (semigroup) property is manifest in the functional form of the effective logistic rate for the increase in the HIV seropositive fraction in the San Francisco (City Clinic) cohort. Witht i=4.5 years, this scaling property—r→λ-2r undert→[λt+(λ−1)t i] for all parameter values λ≧1—encapsulates the effects of relevant biological and sociological changes in the key epidemiological variables during the 8-year seropositive rise period, 1978–1985 inclusive.  相似文献   

2.
The rabbit Na+/glucose cotransporter (SGLT1) exhibits a presteady-state current after step changes in membrane voltage in the absence of sugar. These currents reflect voltage-dependent processes involved in cotransport, and provide insight on the partial reactions of the transport cycle. SGLT1 presteady-state currents were studied as a function of external Na+, membrane voltage V m , phlorizin and temperature. Step changes in membrane voltage—from the holding V h to test values, elicited transient currents that rose rapidly to a peak (at 3–4 msec), before decaying to the steady state, with time constants τ≈4–20 msec, and were blocked by phlorizin (K i ≈30 μm). The total charge Q was equal for the application of the voltage pulse and the subsequent removal, and was a function of V m . The Q-V curves obeyed the Boltzmann relation: the maximal charge Q max was 4–120 nC; V 0.5, the voltage for 50% Q max was −5 to +30 mV; and z, the apparent valence of the moveable charge, was 1. Q max and z were independent of V h (between 0 and −100 mV) and temperature (20–30°C), while increasing temperature shifted V 0.5 towards more negative values. Decreasing [Na+] o decreased Q max, and shifted V 0.5 to more negative voltages 9by −100 mV per 10-fold decrease in [Na+] o ). The time constant τ was voltage dependent: the τ-V relations were bell-shaped, with maximal τmax 8–20 msec. Decreasing [Na+] o decreased τmax, and shifted the τ-V curves towards more negative voltages. Increasing temperature also shifted the τ-V curves, but did not affect τmax. The maximum temperature coefficient Q 10 for τ was 3–4, and corresponds to an activation energy of 25 kcal/mole. Simulations of a 6-state ordered kinetic model for rabbit Na+/glucose cotransport indicate that charge-movements are due to Na+-binding/dissociation and a conformational change of the empty transporter. The model predicts that (i) transient currents rise to a peak before decay to steady-state; (ii) the τ-V relations are bell-shaped, and shift towards more negative voltages as [Na+] o is reduced; (iii) τmax is decreased with decreasing [Na+] o ; and (iv) the Q-V relations are shifted towards negative voltages as [Na+] o is reduced. In general, the kinetic properties of the presteady-state currents are qualitatively predicted by the model. Received: 12 August 1996/Revised: 30 September 1996  相似文献   

3.
A stretch-activated (SA) Cl channel in the plasma membrane of the human mast cell line HMC-1 was identified in outside-out patch-clamp experiments. SA currents, induced by pressure applied to the pipette, exhibited voltage dependence with strong outward rectification (55.1 pS at +100 mV and an about tenfold lower conductance at −100 mV). The probability of the SA channel being open (P o) also showed steep outward rectification and pressure dependence. The open-time distribution was fitted with three components with time constants of τ1o = 755.1 ms, τ2o = 166.4 ms, and τ3o = 16.5 ms at +60 mV. The closed-time distribution also required three components with time constants of τ1c = 661.6 ms, τ2c = 253.2 ms, and τ3c = 5.6 ms at +60 mV. Lowering extracellular Cl concentration reduced the conductance, shifted the reversal potential toward chloride reversal potential, and decreased the P o at positive potentials. The SA Cl currents were reversibly blocked by the chloride channel blocker 4,4′-diisothiocyanatostilbene-2,2′-disulfonic acid (DIDS) but not by (Z)-1-(p-dimethylaminoethoxyphenyl)-1,2-diphenyl-1-butene (tamoxifen). Furthermore, in HMC-1 cells swelling due to osmotic stress, DIDS could inhibit the increase in intracellular [Ca2+] and degranulation. We conclude that in the HMC-1 cell line, the SA outward currents are mediated by Cl influx. The SA Cl channel might contribute to mast cell degranulation caused by mechanical stimuli or accelerate membrane fusion during the degranulation process.  相似文献   

4.
We hypothesized that inhibition and activation of basolateral to luminal chloride transport mechanisms were associated with respective decreases and increases in basolateral to luminal water fluxes. The luminal to basolateral (J W L→B ) and basolateral to luminal (J W B→L ) water fluxes across ovine tracheal epithelia were measured simultaneously. The mean J W L→B (6.5 μl/min/cm2) was larger than J W B→L (6.1 μl/min/cm2). Furosemide reduced J W B→L from 6.0 to 5.6 μl/min/cm2. Diphenylamine-2-carboxylate (DPC) reduced J W B→L from 7.9 to 7.3 μl/min/cm2 and reduced the membrane potential difference by 38%. Furosemide together with DPC decreased J W L→B by 30% and J W B→L by 15%. Norepinephrine increased J W B→L from 4.9 to 6.0 μl/min/cm2. Neuropeptide Y in the presence of norepinephrine decreased J W L→B (6.4 to 5.2 μl/min/cm2) and returned J W B→L to its baseline value. Vasopressin increased J W B→L from 4.1 to 5.1 μl/min/cm2. Endothelin-1 induced a simultaneous increase in J W B→L (7.0 to 7.7 μl/min/cm2) and decrease in J W L→B (7.4 to 6.4 μl/min/cm2); and decreased the membrane resistance. These data indicate that in tracheal epithelia under homeostatic conditions J W B→L has a ∼15% actively coupled component. Consistent with our hypothesis, inhibition and receptor-induced stimulation of chloride effluxes were associated with decreases and increases in J W B→L , respectively. However, as inhibition of transcellular chloride transport always decreased J W L→B more than J W B→L , reducing transepithelial chloride transport did not result in less water being transported into the airway lumen. Received: 12 October 1999/Revised: 14 March 2000  相似文献   

5.
Voltage clamp experiments, which determine the kinetic parameters of calcium conductance of cardiac muscle, (d ,f , τ d and τ f ) are analyzed with a generally accepted expression for slow inward currentI s=g sdf (E-E R). Activation (d) and inactivation (f) reach the final valuesd andf with time constants τ d and τ f respectively. The analysis indicates that the measuredf agrees with the theoreticalf , but the measuredd differs from the theoreticald by a factor which depends on τ d . The peak tension can be made to correlate closely with the theoreticald after a correction factor is applied to the raw measurements of activation. It can be shown that experiments designed to measure τ f can also be used to determine τ d with greater accuracy.  相似文献   

6.
Depolarization-activated H+-selective currents were studied using whole-cell and excised-patch voltage clamp methods in human monocytic leukemia THP-1 cells, before and after being induced by phorbol ester to differentiate into macrophage-like cells. The H+ conductance, g H, activated slowly during depolarizing pulses, with a sigmoidal time course. Fitted by a single exponential following a delay, the activation time constant, τact was roughly 10 sec at threshold potentials, decreasing at more positive potentials. Tail currents upon repolarization decayed mono-exponentially at all potentials. The tail current time constant, τtail, was voltage dependent, decreasing with hyperpolarization from 2–3 sec at 0 mV to ∼200 msec at −100 mV. Surprisingly, although τact depended strongly on pH o , τtail was completely independent of pH o . H+ currents were inhibited by Zn2+. Increasing pH o or decreasing pH i shifted the voltage-activation relationship to more negative potentials, tending to activate the g H at any given voltage. Studied in excised, inside-out membrane patches, H+ currents were larger and activated much more rapidly at lower bath pH (i.e., pH i ). In THP-1 cells differentiated into macrophages, the H+ current density was reduced by one-half, and τact was slower by about twofold. The properties of H+ channels in THP-1 cells and in other macrophage-related cells are compared. Received: 19 September 1995/Revised: 14 March 1996  相似文献   

7.
8.
Current research into the dynamics of iterative ecological and biological models has lead to a number of theorems concerning the existence of various types of iterative dynamical behavior. In particular, much study has been done on the dynamical behavior of the “simplest dynamical system”f b(x)=bx(1−x), which is just the canonical discrete form of logistic growth equations found in ecology, sociobiology, and population biology. In this paper, we make use of some of the techniques and concepts of topological dynamics to construct a number of generalized conjugacy theorems. These theorems are then used to demonstrate that the mappingf b has a number of conjugacy classes in which the dynamics of the iterates is equivalent to within a change of variables. The concepts of fitness and survival in logistic equations are then shown to be independent, if we follow certain intuitive definitions for these concepts. This conclusion follows from a comparison of the conjugacy classes of the functionf b and the extinction sets off b.  相似文献   

9.
The alkali extractable and water-soluble cell wall polysaccharides F1SS from Aspergillus wentii and Chaetosartorya chrysella have been studied by methylation analysis, 1D- and 2D-NMR, and MALDI-TOF analysis. Their structures are almost identical, corresponding to the following repeating unit: [→ 3)-β-D-Galf-(1 → 5)-β-D-Galf-(1 →] n → mannan core. The structure of this galactofuranose side chain differs from that found in the pathogenic fungus Aspergillus fumigatus, in other Aspergillii and members of Trichocomaceae: [→ 5)-β-D-Galf-(1 →] n → mannan core. The mannan cores have also been investigated, and are constituted by a (1 → 6)-α-mannan backbone, substituted at positions 2 by chains from 1 to 7 residues of (1 → 2) linked α-mannopyranoses. Published in 2004. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

10.
The growth and production of the Baltic clam Macoma balthica in the southeastern part of the Baltic Sea were studied. The shell length of M. balthica reached 23.5 mm, the maximum age was 13+ years. The linear growth was described by the von Bertalanffy equation for shallow-water area (depths 9–40 m): L τ = 23.99(1 − e −0.1293(τ − (−0.9578))), and for the deep-water area (41–81 m): L τ = 20.61(1 − e −0.1813(τ − (−0.5608))). The annual production was lower (25.35 ± 1.72 kJ/m2) in the shallow-water area than in the deep-water area (71.23 ± 4.48 kJ/m2), with values of P s /B ratio 0.44 and 0.38, respectively.  相似文献   

11.
The effect of root growth temperature on maximal photosynthetic CO2 assimilation (P max), carbohydrate content, 14C-photoassimilate partitioning, growth, and root morphology of lettuce was studied after transfer of the root system from cool root-zone temperature (C-RZT) of 20 °C to hot ambient-RZT (A-RZT) and vice versa. Four days after RZT transfer, P max and leaf total soluble sugar content were highest and lowest, respectively, in C-RZT and A-RZT plants. P max and total leaf soluble sugar content were much lower in plants transferred from C-to A-RZT (C→A-RZT) than in C-RZT plants. However, these two parameters were much higher in plants transferred from A-to C-RZT (A→C-RZT) than in A-RZT plants. A-RZT and C→A-RZT plants had higher root total soluble sugar content than A→C-RZT and C-RZT plants. Leaf total insoluble sugar content was similar in leaves of all plants while it was the highest in the roots of C-RZT plants. Developing leaves of C-RZT plants had higher 14C-photoassimilate content than A-RZT plants. The A→C-RZT plants also had higher 14C-photoassimilate content in their developing leaves than A-RZT plants. However, more 14C-photoassimilates were translocated to the roots of A-RZT and C→A-RZT plants, but they were mainly used for root thickening than for its elongation. Increases in leaf area, shoot and root fresh mass were slower in C→A-RZT than in C-RZT plants. Conversely, A→C-RZT plants had higher increases in these parameters than A-RZT plants. Lower root/shoot ratio (R/S) in C-RZT than in A-RZT plants confirmed that more photoassimilates were channelled to the shoots than to the roots of C-RZT plants. Roots of C-RZT plants had greater total length with a greater number of tips and surface area, and smaller average diameter as compared to A-RZT plants. In C→A-RZT plants, there was root thickening but the increases in its length, tip number and surface area decreased. The reverse was observed for A→C-RZT plants. These results further supported the idea that newly fixed photoassimilates contributed more to root thickening than to root elongation in A-RZT and C→A-RZT plants.  相似文献   

12.
The scattering spectrum properties of highly turbid and eutrophic inland case 2 water from Taihu Lake were studied during three cruises from 2006 to 2007. The scattering [b p(λ)] and backscattering [b bp(λ)] coefficients and the backscattering probability (B) for Taihu Lake were found to show a clear spectral dependence, and this dependence was well simulated by a power-law function. This dependence, however, became weak when algae dominated the sample points. The mean values of the power-law index for b p(λ), v, in Oct 2006, Mar 2007 and Nov 2007 were −0.6712, −0.8129 and −0.7600, respectively. To interpret the spectral characteristics and mechanisms of b p(λ) and b bp(λ), water samples were collected simultaneously for the biogeochemical characterization of suspended particles. The average values of the specific scattering coefficients for total suspended matter, inorganic suspended matter (ISPM) and organic suspended matter (OSPM) were 0.634 (550 nm), 1.057 (532 nm), and 0.396 g m−2 (532 nm), respectively. The power-law index of b bp(λ) (Y) was significantly related to ISPM/OSPM and b bp(532 nm), but only weakly related to the particle size distribution index. The mean (spatial and wavelength) values of B in Oct 2006, Mar 2007, and Nov 2007 were 0.0108, 0.0138, and 0.0125, respectively. B decreases with increasing ISPM concentration because of the large contribution of ISPM to b b(λ) and the strong restraint on b bp(λ) caused by the multi-scattering effect under high-turbidity conditions.  相似文献   

13.
Summary Using a direct Monte Carlo simulation, population growth of helper T-cells (N H) and viral cells (N v) is studied for an immune response model with an enhanced spatial inter-cellular interaction relevant to HIV as a function of viral mutation. In the absence of cellular mobility (P mob=0), the helper T-cells grow nonmonotonically before reaching saturation and the viral population grows monotonically before reaching a constant equilibrium. Cellular mobility (P mob=1) enhances the viral growth and reduces the stimulative T-cell growth. Below a mutation threshold (P c), the steady-state density of helper T-cell (p H) is larger than that of the Virus (p v); the density difference Δp o(=pV−pH) remains a constant at P mob=1 while −Δp o→0 as P mutP c at P mob=0. Above the mutation threshold, the difference Δp o in cell density, grows with ΔP=P mutP c monotonically: ΔP o ∞ (ΔP)β ≃ with β≈0.574±0.016 in absence of mobility, while Δp o≈6(ΔP) with P mob=1.  相似文献   

14.
We have attempted to investigate the correlation between the detergent-perturbed structural integrity of the Cyt b 6 f complex from the marine green alga Bryopsis corticulans and its photo-protective properties, for which the nonionic detergents n-octyl-β-d-glucopyranoside (β-OG) and n-dodecyl-β-d-maltoside (β-DM), respectively, were used for the preparation of Cyt b 6 f, and the singlet oxygen (1O2*) production as well as the triplet excited-state chlorophyll a (3Chl a*) formation and deactivation were examined by spectroscopic means. Near-infrared luminescence of 1O2 * (~1,270 nm) on photo-irradiation was detected for the β-OG preparation where the complex is mainly in oligomeric state, but not for the β-DM one in which the complex exists in dimeric form. Under anaerobic condition, photo-excitation of Chl a in the β-DM preparation generated 3Chl a* with a lower quantum yield of ΦT ~ 0.02 and a longer lifetime of ~600 μs with respect to those as in the case of β-OG preparation, ΦT ~ 0.12 and 200–300 μs. These results prove that the enzymatically active and intact Cyt b 6 f complex on photo-excitation tends to produce little 3Chl a* or 1O2 *, which implies that the pigment–protein assembly of Cyt b 6 f complex per se is crucial for photo-protection. F. Ma and X.-B. Chen contributed equally to this work.  相似文献   

15.
Summary Degradation of aflatoxin B1 byCorynebacterium rubrum and byAspergillus niger was analysed by adding14C-labeled aflatoxin B1 to cultures of these microorganisms. Two blue fluorescent compounds, formed byA. niger from aflatoxin B1 with Rf-values 0.42 and 0.48 (Rf of aflatoxin B1=0.54) were accumulated and characterized by UV-, fluorescence and mass spectrometry. Based on their properties both products were identified to be aflatoxin Ro. Under the same conditionsMucor ambiguus andTrichoderma viride also produced aflatoxin Ro.  相似文献   

16.
A physical model that incorporates all the experimental information on the formation of the visual pigment rhodopsin is presented. The visual pigments consist of a chromophore bound to an appropriate protein. Thus rhodopsin (λm 505 mμ) is formed by a Schiff’s base linkage C19H27CH=NH+-opsin (λm 440 mμ) between 11-cis retinal (λm 380 mμ) and the protein opsin (λm 280 mμ). It is found that there exists a red shift in the spectrum of rhodopsin from the Schiff’s base. The model brings an explanation for this red shift. It is shown that such a shift may be due to a charge transfer process (R. S. Mulliken,J. Am. Chem. Soc.,74, 811–824, 1952) between an electron at the double bond of carbons C11−C12 and an atomic orbital of the sulphur present in cysteine. This provides an explanation of the presence of SH-groups in the protein after the absorption of light. A one-electron approximation is used and the dipole momentμ NV ; hence, the oscillator strengthf of the transitionNV is estimated and compared with the experimentally determined extinction coefficient ∈m by mixing 3.5×10−3 M of 11-cis retinal with 8.3×10−5 M of cysteine at pH ranges 6 through 8. Reasonable agreement is found. Solvent, concentration and temperature dependence are shown also.  相似文献   

17.
As was done by Sinclair and Ross (1969(, we consider a cellular population that consists initially (at time zero) ofN 0 newborn cells, all with the same volumev o. It is assumed that the occurrence of cell division is determined only by a cell’s age, and not by its volume. The frequency function of interdivision times, τ, is denoted byf(τ). If cell death is negligible, the expected number of cells,N(t), will increase according to the laws of a simple age-dependent branching process. The expression forN(t) is obtained as a sum over all generations; thevth term of this sum, in turn, is a multiple convolution integral, reflecting the life history ofvth generation cells (i.e., the lengths of thev successive interdivision periods plus the age of the cell at timet). Assuming that cell volume is a given function of cell age, e.g., linear or exponential, and that cellular volume is exactly halved at each division, it is possible to calculate the volume of a cell with a given life history, and thus the average cellular volume of the whole population as a function of time. If at time zero the volumes differ from cell to cell, the final equation must be modified by averaging over initial volumes. In the case of linear volume increase with age, a very simple asymptotic expression is found for the average cellular volume ast→∞. The case of exponential volume increase with age also leads to a simple asymptotic formula, but the resulting volume distribution is unstable. The mean cellular volume at birth and the second moment of the volume distribution can be calculated in a similar manner. Work supported by the U.S. Atomic Energy Commission.  相似文献   

18.
Experimental study of a glow discharge with electrostatic confinement of electrons is carried out in the vacuum chamber volume V ≈ 0.12 m3 of a technological system “Bulat-6” in argon pressure range 0.005–5 Pa. The chamber is used as a hollow cathode of the discharge with the inner surface area S ≈ 1.5 m2. It is equipped with two feedthroughs, which make it possible to immerse in the discharge plasma interchangeable anodes with surface area S a ranging from ∼0.001 to ∼0.1 m2, as well as floating electrodes isolated from both the chamber and the anode. Dependences of the cathode fall U c = 0.4−3 kV on the pressure p at a constant discharge current in the range I = 0.2−2 A proved that aperture of the electron escape out of the electrostatic trap is equal to the sum S o = S a + S f of the anode surface S a and the floating electrode surface S f . The sum S o defines the lower limit p o of the pressure range, in which U c is independent of p. At p < p o the cathode fall U c grows up dramatically, when the pressure decreases, and the pressure p tends to the limit p ex, which is in fact the discharge extinction pressure. At pp ex electrons emitted by the cathode and the first generation of fast electrons produced in the cathode sheath spend almost all their energy up to 3 keV on heating the anode and the floating electrode up to 600–800°C and higher. In this case the gas in the chamber is being ionized by the next generations of electrons produced in the cathode sheath, their energy being one order of magnitude lower. When S a < (2m/M)1/2 S, where m is the electron mass and M is the ion mass, the anode may be additionally heated by plasma electrons accelerated by the anode fall of potential U a up to 0.5 kV.  相似文献   

19.
Kubitscheck U  Homann U  Thiel G 《Planta》2000,210(3):423-431
The dye FM1-43 was used alone or in combination with measurements of the membrane capacitance (Cm) to monitor membrane changes in protoplasts from Viciafaba L. guard cells. Confocal images of protoplasts incubated with FM1-43 (10 μM) at constant ambient osmotic pressure (πo) revealed in confocal images a slow internalisation of FM1-43-labelled membrane into the cytoplasm. As a result of this process the relative fluorescence intensity of the cell interior (fFM,i) increased with reference to the total fluorescence (fFM,t) by 7.4 × 10−4 min−1. This steady internalisation of dye suggests the occurrence of constitutive endocytosis under constant osmotic pressure. Steady internalisation of FM1-43 labelled membrane caused a prominent staining of a ring-like structure located beneath the plasma membrane. Abrupt elevation of πo by 200 mosmol kg−1 caused, over the first minutes of incubation, a rapid internalisation of FM1-43 fluorescence into the cytoplasm concomitant with a decrease in cell perimeter. Within the first 5 min the cell perimeter decreased by 7.9%. Over the same time fFM,i/fFM,t increased by 0.13, reflecting internalisation of fluorescent label into the cytoplasm. Combined measurements of Cm and total fluorescence of a protoplast (fFM,p) showed that an increase in πo evoked a decrease in Cm but no change in fFM,p. This means that surface contraction of the protoplast is due to retrieval of excess membrane from the plasma membrane and internalisation into the cytoplasm. Further inspection of confocal images revealed that protoplast shrinking was only occasionally associated with internalisation of giant vesicles (median diameter 2.7 μm) with FM1-43-labelled membrane. But, in all cases, osmotic contraction was correlated with a diffuse distribution of FM1-43 label throughout the cytoplasm. From this, we conclude that endocytosis of small vesicles into the cytoplasm is the obligatory process by which cells accommodate an osmotically driven decrease in membrane surface area. Received: 4 May 1999 / Accepted: 19 August 1999  相似文献   

20.
The patch clamp K+-conductance G of the nicotinic acetylcholine receptor (AcChoR) dimer (Mr≈ 590 000) of Torpedo californica, reconstituted in lipid vesicles, which decreases with increasing Ca2+-concentration in the range 0.1≤[Ca2+]/mM≤2, can be quantitatively rationalized by Ca2+-binding to negatively charged sites, causing charge reversal reducing the normal K+-accumulation in the channel vestibules. Cleavage of the sialic acid residues (up to 20±2 per dimer) reduces the K+-accumulation factor α = G0/G from α = 3±0.8 of the normal AcChoR to α = 2±0.7 for the desialyated AcChoR. Desialysation also decreases the Ca2+-sensitivity of the conductance from G0 = 96.6±6 pS at [Ca2+]→0 of the normal AcChoR to G0 = 84.2±6 pS. Endogenous hyperphosphorylation (to up to 28±4 phosphates per dimer) enhances the vestibular K+-accumulation to α = 3.6±0.7, without affecting the Ca2+-dissociation equilibrium constant KCa = 0.34± 0.05 mM at 295 K (22 °C). Most interestingly, even in the absence of AcCho, the hyperphosphorylated AcChoR dimer exhibits spontaneously long-lasting open channel events (τ = 200±50 ms). At [AcCho] = 2 μM there are two open states (τ 1 = 20±10 ms, τ 2 = 140±60 ms) whereas the normal AcChoR dimer has only one open state (τ = 6±4 ms). – Physiologically important is that (i) the sialic acid and phosphate residues render the AcChoR conductance sensitive to control by divalent ions and (ii) the channel behavior of the hyperphosphorylated AcChoR without AcCho appears to indicate pathophysiologically high phosphorylation activity of the cell leading, among others, to myasthenic syndromes. Received: 10 November 1997 / Revised version: 12 January 1998 / Accepted: 7 March 1998  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