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.
Fluorescence photobleaching recovery methods reveal marked changes in lateral mobilities of rabbit lymphocyte membrane components during the course of stimulation with succinyl concanavalin A (S Con A). The diffusion constant of S Con A receptors on T lymphocytes falls from 1.6×10?10 cm2/sec to 6.5×10?11 cm2/sec within 4 hr after stimulation, remains constant for 14 hr, and returns to its former value. The mobility of B cell receptors similarly falls from 1.4×10?10 cm2/sec to 5.5×10?11 cm2/sec but regains its unstimulated value much more slowly. In contrast, a fluorescent phospholipid analog shows constant mobilities of 1.9×10?8 cm2/sec and 1.5×10?8 cm2/sec in T and B cells, respectively, throughout the experiment.  相似文献   

2.
K L Wun  W Prins 《Biopolymers》1975,14(1):111-117
Quasi-elastic light scattering as measured by intensity fluctuation (self-beat) spectroscopy in the time domain can be profitably used to follow both the translational diffusion D and the dominant internal flexing mode τint of DNA and its complexes with various histones in aqueous salt solutions. Without histones, DNA is found to have D = 1.6 × 10?8 cm2/sec and τint ? 5 × 10?4 sec in 0.8 M NaCl, 2 M urea at 20°C. Total histone as well as fraction F2A induce supercoiling (D = 2.6 × 10?8 cm2/sec, τint ? 2.8 × 10?4 sec) whereas fraction F1 induces uncoiling (D = 1.0 × 10?8 cm2/sec, τint ? 9.4 × 10?4 sec). Upon increasing the salt concentration to 1.5 M the DNA–histone complex dissociates (D = 1.8 × 10?8 cm2/sec). Upon decreasing the salt concentration to far below 0.8 M, the DNA–histone complex eventually precipitates as a chromatin gel.  相似文献   

3.
Concanavalin A (conA) modulates the lateral mobility of cell surface receptors differently on different cell types. This was demonstrated by using fluorescence photobleaching recovery (FPR) to measure the inhibition of the lateral mobility of conA receptors by localized binding of conA on lymphocytes, fibroblasts, and macrophages. On mouse spleen lymphocytes, binding of conA platelets above a threshold coverage (about 12% of the upper cell-surface area) reduced the diffusion coefficient of mobile TMR-SconA-receptor complexes from 3.0×10?10 cm2/sec to 0.6× 10?10 cm2/sec (a 5-fold decrease), and the fraction of mobile receptors was concomitantly reduced from 0.4 to 0.11. Below the threshold occupancy, no effect on either parameter was detected. On 3T3 cells, a qualitatively similar threshold phenomenon was observed: coverage of over 9% of the upper cell surface by conA platelets induced a 3-fold reduction in the diffusion coefficient of TMR-SconA-receptor complexes from 5×10?10 cm2/sec to 1.7× 10?10 cm2/sec. However, no effect on the mobile fraction (about 0.4) was observed. In contrast, neither the diffusion coefficient nor the mobile fraction of TMR-SconA-receptor complexes on mouse peritoneal macrophages (both resident and thioglycolate-stimulated) or on the mouse macrophage cell line P388D1 were affected by the binding of conA platelets in amounts covering over 50% of the upper cell surface (approx. 4.6× 10?10 cm2/sec and 0.5 for the diffusion coefficient and mobile fraction, respectively). These differences are correlated to the different cytoskeletal functions of the various cell types studied, and are discussed regarding the mechanism of the conA-induced modulation.  相似文献   

4.
Regional differences in the lateral mobility properties of plasma membrane lipids have been studied in unfertilized and fertilized Xenopus eggs by fluorescence photobleaching recovery (FPR) measurements. Out of a variety of commonly used lipid probes only the aminofluorescein-labeled fatty acids HEDAF (5-(N-hexadecanoyl)-aminofluorescein) and TEDAF (5-(N-tetradecanoyl)-aminofluorescein) appear to partition into the plasma membrane. Under all experimental conditions used these molecules show partial recovery upon photobleaching indicating the existence of lipidic microdomains. In the unfertilized egg the mobile fraction of plasma membrane lipids (~50%) has a fivefold smaller lateral diffusion coefficient (D = 1.5 × 10?8 cm2/sec) in the animal than in the vegetal plasma membrane (D = 7.6 × 10?8 cm2/sec). This demonstrates the presence of an animal/vegetal polarity within the Xenopus egg plasma membrane. Upon fertilization this polarity is strongly (>100×) enhanced leading to the formation of two distinct macrodomains within the plasma membrane. At the animal side of the egg lipids are completely immobilized on the time scale of FPR measurements (D ? 10?10 cm2/sec), whereas at the vegetal side D is only slightly reduced (D = 4.4 × 10?8 cm2/sec). The immobilization of animal plasma membrane lipids, which could play a role in the polyspermy block, probably arises by the fusion of cortical granules which are more numerous here. The transition between the animal and the vegetal domain is sharp and coincides with the boundary between the presumptive ecto- and endoderm. The role of regional differences in the plasma membrane is discussed in relation to cell diversification in early development.  相似文献   

5.
We have constructed an apparatus for the simultaneous measurement of electrophoretic mobility, μ, and diffusion coefficient, D, of macromolecules and cells. It combines band electrophoresis in a vertical, sucrose-gradient stabilized column, with quasielastic laser light-scattering determination of the diffusion coefficient of the species within the band. The entire electrophoresis cell is scanned through the laser beam of the quasielastic laser light-scattering apparatus by a vertical translation stage. Total intensity light-scattering measurement at each point in the cell gives the macromolecular concentration at that point. Solvent viscosity and electrical potential are measured at each point in the cell. Application of this apparatus to resealed red blood cell ghosts and to bovine hemoglobin indicates that measurements of field, viscosity, and migration distance are reliable, and that electroosmosis is insignificant. Application to T4D bacteriophage gives μ20,w = (?1.05 ± 0.05) × 10?4 cm2/V sec and D20,w = (3.35 ± 0.10) × 10?8 cm2/sec for fiberless particles, and μ20,w = ?(0.59 ± 0.03) × 10?4 cm2/V sec and D20,w = (2.86 ± 0.09) × 10?8 cm2/sec for whole phage with 6 fibers. Approximate analysis of these results with the Henry electrophoresis theory for spheres in dicates that each fiber contributes about 193 positive charges to the phage particle, compared with 327 from amino-acid analysis. The advantages and disadvantages of this apparatus, relative to conventional electrophoresis and to electrophoretic light scattering, are discussed.  相似文献   

6.
Dermal exposure to volatile compounds (VC) in municipal water while showering is typically estimated using a steady-state condition between VC in water impacting on skin and skin exposed to water. The lag times to achieve steady-state between VC and skin can vary in the range of 7.5–218.3 min, while shower duration is often less than these values. Estimates of dermal exposure to VC using steady-state while showering may misinterpret exposure. This study developed models and estimated exposure to some disinfection byproducts (DBPs) through dermal pathway by considering lag times while showering. Dermal uptakes of VC were compared using different approaches. In the proposed approach, uptakes of trihalomethanes were estimated between 9.55 × 10?10–1.43 × 10?8 mg/cm2 of skin during the lag times from exposure to water with trihalomethanes of 50 μg/L. These values were higher than the steady-state estimates (1.37 × 10?10–4.34 × 10?9 mg/cm2), and lower than the average exposure analysis (4.12 × 10-8–1.93 × 10?6 mg/cm2). Using the Drinking Water Surveillance Program data in Ontario, chronic daily intakes of trihalomethanes were estimated to be 9.40 × 10?7 (1.85 × 10?7–1.65 × 10?6), 3.89 × 10?6 (7.11 × 10?7–2.33 × 10?5), and 1.40 × 10?6 (4.0 × 10?7–1.77 × 10?6) mg/kg/day in Toronto, Ottawa, and Hamilton, respectively. The findings can be useful in understanding THMs exposure and risk through dermal pathway.  相似文献   

7.
By using the static correlations of fluctuations in the dihedral angles of the α-helices of polyglycine and poly(L -alanine) calculated previously, geometrical fluctuations of a section (consisting of up to 18 peptide units) of the α-helices of infinite length are calculated. These fluctuations are found to differ in some respects (e.g., the dependence of amplitudes on the length of section) from those of a circular rod made of homogeneous continuous material. However, the moduli of the mechanical strengths (tensile Young's modulus, bending Young's modulus, and the shear modulus) of a circular rod are calculated, whose geometrical fluctuations are approximately equal to the fluctuations of a section consisting of 18 peptide units. They are of the order of 1011 dyn/cm2. The tensile rigidity, flexural rigidity, and torsional rigidity are calculated to be 1.20 × 10?3 dyn, 2.46 × 10?19 dyn·cm2 and 1.79 × 10?19 dyn·cm2 for polyglycine, and 1.96 × 10?3 dyn, 4.05 × 10?19 dyn·cm2 and 3.28 × 10?19 dyn·cm2 for poly(L -alanine), respectively.  相似文献   

8.
The structure of thermally denatured Type I collagen has been studied using laser light scattering. The results indicate that the diffusion coefficients of α-chains and β- and γ-components are 1.550 ± 0.08 × 10?7, 1.000 ± 0.05 × 10?7, and 0.835 ± 0.04 × 10?7 cm2/sec, respectively, at temperatures between 20 and 40°C. It is concluded from diffusion data that these species have hydrodynamic radii of about 13.8 nm (α-chain), 21.5 nm (β-component), and 25.7 nm (γ-component), consistent with previous studies of thermal denaturation by light scattering. It is also concluded, based on volume calculations, that a large volume increase occurs when the triple helix unfolds. Homodyne correlation functions for two component mixtures of α-chains and β-and γ-components appeared to decay exponentially. In all but one case discussed the correlation function could be fitted with a single component having a translational diffusion coefficient which was an intensity weighted average of the diffusion coefficient of each component present.  相似文献   

9.
T. Raj  W. H. Flygare 《Biopolymers》1977,16(3):545-549
The translational diffusion coefficient of a pure sample of α-chymotrypsinogen A is measured by laser light scattering to give a value of D20,w0 = (8.40 ± 0.15) × 10?7 cm2/sec.  相似文献   

10.
Fluorescence photobleaching recovery techniques have allowed us to measure the lateral mobility of T-independent antigens bound to antigen-specific mouse B cells. The in vitro immunogenicity or tolerogenicity of antigens we have examined, DNP-polymerized flagellin (DNP-POL), and DNP-linear dextran (DNP-DEX), depend upon the antigen dose and epitope density. These factors also determine the mobility of antigen bound to B cell surfaces. For DNP-POL bound to DNP-specific cells, the observed diffusion constants D decrease monotonically with increasing antigen dose and epitope density. Values of D range from 10.4 × 10?11 cm2 sec?1 for DNP0.4-POL at 0.15 μg/ml to 0.8 × 10?11cm2 sec?1for DNP3.5-POL at 30 μg/ml. For receptor-bound DNP-DEX, D depends strongly on antigen epitope density but not observably on antigen concentration. For epitope densities of 1.2 or less, D is close to the value of 21 × 10?11cm2sec?1 observed for single slg receptors. By an epitope density of 4.8, D has fallen to 2.1 × 10?11cm2sec?1. Peak immunogenicities for DNP-POL and DNP-DEX arc observed when antigen- receptor aggregates have mobilities 14-fold and 3-fold lower, respectively, than a single slg molecule.  相似文献   

11.
Dictyostelium discoideum cells were allowed to differentiate on agar for 600 min at room temperature. All of the cells were then competent to relay or amplify a cAMP signal, but none to produce a cAMP signal autonomously. The cells were stimulated with cAMP concentrations ranging from 10?9 to 3.5 × 10?7M. Populations of 106 cells could amplify an initial cAMP concentration of 2.5 × 10?9M with a low probability, while an initial cAMP concentration of 5 × 10?8M always induced a response. An initial cAMP concentration of 1.2 × 10?7M induced the maximum cellular release of cAMP observed; this corresponded to 3 × 107 molecules per cell. No cellular release of cAMP was detected for initial cAMP concentrations of 3 × 10?7M or more. The amplification of a 10?7M cAMP stimulus was complete within 8 sec, indicating the pulsatile nature of the cellular release of cAMP. The phosphodiesterase (PDE) activities of D. discoideum cells were measured over a wide range of cell densities. At densities above 7.5 × 104 cells/cm2, both cell-bound and extracellular (ePDE) activities declined, per cell, as cell density increased. These results are compared to ePDE activities derived from critical density measurements. We found that PDE activities were in the range of 10?13–10?14 moles of cAMP converted/cell/min under culture conditions consistent with normal aggregation.  相似文献   

12.

A conspicuous bioluminescence during nighttime was reported in an aquaculture farm in the Cochin estuary due to Gonyaulax spinifera bloom on March 20, 2020. In situ measurements on bioluminescence was carried out during nighttime to quantify the response of G. spinifera to various mechanical stimuli. The bioluminescence intensity (BI) was measured using Glowtracka, an advanced single channel sensor, attached to a Conductivity–Temperature–Depth Profiler. In steady environment, without any external stimuli, the bioluminescence generated due to the movement of fishes and shrimps in the water column was not detected by the sensor. However, stimuli such as a hand splash, oar and swimming movements, and a mixer could generate measurable bioluminescence responses. An abundance of?~?2.7?×?106 cells L?1 of G. spinifera with exceptionally high chlorophyll a of 25 mg m?3 was recorded. The BI in response to hand splash was recorded as high as 1.6?×?1011 photons cm?2 s?1. Similarly, BI of?~?1–6?×?1010 photons cm?2 s?1 with a cumulative bioluminescence of?~?2.51?×?1012 photons cm?2 (for 35 s) was recorded when there is a mixer with a constant force of 494 N/800 rpm min?1. The response of G. spinifera was spontaneous with no time lapse between application of stimuli and the bioluminescence response. Interestingly, in natural environment, application of stimulus for longer time periods (10 min) does not lower the bioluminescence intensity due to the replenishment of water thrusted in by the mixer from surrounding areas. We also demonstrated that the bioluminescence intensity decreases with increase in distance from the source of stimuli (mixer) (av. 1.84?×?1010 photons cm?2 s?1 at 0.2 m to av. 0.05?×?1010 photons cm?2 s?1 at 1 m). The BI was highest in the periphery of the turbulent wake generated by the stimuli (av. 3.1?×?1010 photons cm?2 s?1) compared to the center (av. 1.8?×?1010 photons cm?2 s?1). When the stimuli was applied vertically down, the BI decreased from 0.2 m (0.3?×?1010 photons cm?2 s?1) to 0.5 m (0.10?×?1010 photons cm?2 s?1). Our study demonstrates that the BI of G. spinifera increases with increase in mechanical stimuli and decreases with increase in distance from the stimuli.

  相似文献   

13.
An LBO (Li2B4O7) walled ionization chamber was designed to monitor the epithermal neutron fluence in boron neutron capture therapy clinical irradiation. The thermal and epithermal neutron sensitivities of the device were evaluated using accelerator neutrons from the 9Be(d, n) reaction at a deuteron energy of 4 MeV (4 MeV d-Be neutrons). The response of the chamber in terms of the electric charge induced in the LBO chamber was compared with the thermal and epithermal neutron fluences measured using the gold-foil activation method. The thermal and epithermal neutron sensitivities obtained were expressed in units of pC cm2, i.e., from the chamber response divided by neutron fluence (cm?2). The measured LBO chamber sensitivities were 2.23 × 10?7 ± 0.34 × 10?7 (pC cm2) for thermal neutrons and 2.00 × 10?5 ± 0.12 × 10?5 (pC cm2) for epithermal neutrons. This shows that the LBO chamber is sufficiently sensitive to epithermal neutrons to be useful for epithermal neutron monitoring in BNCT irradiation.  相似文献   

14.
The use of fluorescence recovery after photobleaching (FRAP) techniques to monitor the lateral mobility of plant lectin-receptor complexes on the surface of single, living mammalian cells is described in detail. FRAP measurements indicate that over 75% of the wheat germ agglutinin receptor (WGA-receptor) complexes on the surface of human embryo fibroblasts are mobile. These WGA-receptor complexes diffuse laterally (as opposed to flow) on the cell surface with a diffusion coefficient in the range of 2 × 10?11 to 2 × 10?10 cm2/sec. Both the percentage of mobile WGA-receptor complexes and the mean diffusion coefficient of these complexes are higher than that obtained from earlier FRAP measurements of the mobility of concanavalin A-receptor (Con A-receptor) complexes in a variety of cell types. The possible reasons for the differing mobilities of WGA and Con A receptors are discussed.  相似文献   

15.
The exposure of red beet root tissue to ultraviolet (254 nm) at 2 × 106 erg × cm?2× min?1 (0.2 J × cm?2× min?1) causes release of betacyanin after a 20 minute induction period. Ultraviolet-photolysis is temperature-sensitive having a thermal threshold at about 10°C. Reduction in pigment release was effected by chlorides of Mg, Ca and Sr, but not by Li, Na or K. This effect was marked but not complete, even at 40 mM concentration. It is concluded that photolysis is indirect, and involves a lytic factor, possibly an oxidant, derived from an original photochemical product.  相似文献   

16.
A new aminopeptidase was isolated from Agave americana by fractionation, chromatography and gel filtration. The mw of the enzyme, determined by different procedures was 86000 ± 1500; the enzyme had a sedimentation coefficient of 4.96 S, a diffusion coefficient of 5.2 × 10?7 cm2/sec, a Stokes radius of 3.8 nm, a partial specific volume of 0.733 cm3/g, a frictional ratio of 1.40, a molecular absorbancy index at 280 nm of 8.36 × 104, an isoelectric point of 4.53 and contained 1.25% carbohydrate. The amino acid composition of the enzyme was determined and the aminoterminal residue was identified as lysine whilst the carboxylterminal residue was either leucine or isoleucine. No subunit structure was observed for the enzyme.  相似文献   

17.
Colony formation by variant Chinese hamster cells highly resistant to adenine analogs and deficient in adenine phosphoribosyltransferase (APRT) activity was measured after co-cultivation with APRT+, CHO-K1 cells in medium containing one of three different adenine analogs. Depending upon the density of APRT+ cells and the specific adenine analog, large differences in the recovery of APRT? colonies were observed. The particular adenine analog and APRT+ cell density were more significant factors in the recovery of APRT? colonies than the concentration of the analog or the level of APRT activity. The number of wild-type cells (CHO-K1) required to inhibit formation of APRT? colonies by 50% (mean lethal density; MLD50) with 65 μg/ml 8-aza-adenine (AzA) as the selective drug was 8.0 × 105 cells/100 mm dish (1.5 × 104/cm2). With 100 μg/ml 2,6-diaminopurine (DAP) the MLD50 for CHO-K1 was 4.0 × 105 cells/100 mm dish (7.3 × 103/cm2). The MLD50 for CHO-K1 when the DAP concentration was decreased to 50 μg/ml was only slightly higher, 5 × 105 cells/100 mm dish (9.1 × 103/cm2). The most toxic effect was observed with 2-fluoroadenine (FA). The MLD50 for CHO-K1 in 2 μg/ml FA was 4.5 × 104 cells/100 mm dish (8.2 × 102/cm2), a cell density which permits minimal direct contact between APRT+ and APRT? cells. The toxic effects of FA on individually resistant, APRT? cells were found to be mediated by metabolites released into the medium by dying APRT+ cells. This metabolite toxicity to APRT? cells was also demonstrated in mixtures with cells having only 8% of wild-type APRT activity. The MLD50 for these APRT+ (8%) cells in 2 μg/ml FA was 7.5 × 104 cells/100 dish (1.4 × 103/cm2), a small difference from the MLD50 for cells with wild-type levels of APRT activity. The differences in the recovery of APRT? colonies from mixtures with APRT+ cells in these three adenine analogs are critical to the design of procedures for the selection of APRT? cells from populations of APRT+ cells and emphasize the importance of establishing the parameters of metabolic cooperation, not only in terms of cell density but also with regard to the particular selective agent, in any experiment designed to determine precise mutation rates or to test putative mutagens upon mammalian cells in culture.  相似文献   

18.
Zoeae of Paralithodes camtschatica were positively phototactic to white light intensities above 1 × 1013 q cm?2 s?1. Negative phototaxis occurred at low (1 × 1012 q cm?2 s?1), but not high intensities (2.2 × 1016q cm?2 s?1). Phototactic response was directly related to light intensity. Zoeae also responded to red, green and blue light. Zoeae were negatively geotactic, but geotaxis was dominated by phototaxis. Horizontal swimming speed of stage 1 zoeae <4 d old was 2.4 ± 0.1 (SE) cms?1 and decreased to 1.7 ± 0.1 cm s?1 in older zoeae (P <0.01). Horizontal swimming speed of stage 2 zoeae was not significantly different from ≥4 d old stage 1 zoeae. Vertical swimming speed, 1.6 ± 0.1 cm s?1, and sinking rate, 0.7 ± 0.1 cm s?1, did not change with ontogeny. King crab zoeae were positively rheotactic and maintained position in horizontal currents less than 1.4 cm s?1. Starvation reduced swimming and sinking rates and phototactic response.  相似文献   

19.
A stearamide spin probe was used to study the light-induced structural changes in Rod Outer Segment Membranes in the presence of sodium and calcium ions. The correlation time (τC) for the reorientation of the probe was calculated in the dark and light. In the presence of sodium ions, τC = 3.3 × 10?9 sec in the dark, and 2.7 × 10?9sec in the light while the opposite was noticed in the presence of calcium ions, τC = 2.9 × 10?9 sec in the dark and 3.6 × 10?9 sec in the light. The correlation times for reorientation of the probe were also calculated in aqueous glycerol solutions of varying viscosities at 20°C. Comparison of the values of τC (dark and light) suggests a change in local mobility in the ROS corresponding to a macroscopic viscosity difference of approximately 150 cp. The significance of calcium ion interaction with negatively charged groups and the formation of a Schiff base is emphasized.  相似文献   

20.
Morphology and recordings of electrical activity of Kuruma shrimp (Penaeus japonicus) giant medullated nerve fibers were carried out. A pair of giant fibers with external diameter of about 120 μ and 10 μ in myelin thickness were found in the ventral nerve cord. The diameter of the axon is about 10 μ. Thus there is a wide gap between the axon and the external myelin sheath. Each axon is doubly coated directly by Schwann cells and indirectly by the myelin sheath layer which is produced by those Schwann cells. Impulse conduction velocities of these giant fibers showed a range between 90–210 m/sec at about 22°C. Large action potentials (up to 113 mV, rise time of 0.16–0.3 msec, maximum rate of rise of 650–1250 V/sec, half decay time of 0.2–0.3 msec, maximum rate of fall of 250–450 V/sec and total duration of less than 1.5 msec) could be obtained by inserting microelectrodes or by longitudinal insertion of 25 μ diameter capillary electrodes into the gap but no DC-potential difference was observed across the myelin sheath. Transmyelin electrical parameters were very favorable for fast impulse conduction: myelin resistance of 3 × 104 Ω cm2; time constant of 0.38 msec; myelin capacitance of 1.35 × 10?8 F/cm2; gap fluid resistivity of 23 Ω cm. The existence of nodes of Ranvier could not be demonstrated morphologically, but electrophysiological evidence suggests that a type of saltatory conduction occurs in these giant fibers.  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