首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Electrical cell-substrate impedance sensing (ECIS) was used to measure the time-dependence and frequency-dependence of impedance for current flowing underneath and between cells. Osteosarcoma cells with a topology similar to a short cylinder (coin-like) surmounted by a dome were used in this study. Application of a small step increase in net vertical stress to the cells (4 and 7 dyn/cm2), via magnetic beads bound to the dorsal (upper) surface, causes an increase in cell body height and an increase in cell-cell separation, as well as stretching of the cell-substrate adhesion bonds. This results in a fast drop in measured resistance (less than 2 s), followed by a slower change with a time constant of 60–150 s. This time constant is about 1.5 times longer at 22 °C than that at 37 °C; it also increases with applied stress. Our frequency scan data, as well as our data for the time course of resistance and capacitance, show that the fast change is associated with both the under-the-cells and between-the-cells resistance. The slower change in resistance mainly reflects the between-the-cells resistance. To obtain viscoelastic parameters from our data we use a simple viscoelastic model comprising viscous and elastic elements (i.e., a dashpot and two springs) for the cell body, and an elastic element (a spring) for the cell-substrate adhesion system. Our results show that the spring constants and the viscosity of the cell body components of this viscoelastic model decrease as the temperature increases, whereas the elastic modulus of cell-substrate adhesion increases with temperature. At 37 °C, for the cell body we obtain a value of about 105 P for the viscous element of the viscoelastic model, and a spring constant expressed in units of an elastic modulus of about 104 dyn/cm2 for the spring in series with the viscous element, with another spring with a modulus of about 2×103 dyn/cm2 in parallel with these. In comparable units, we have a modulus for the cell-substrate adhesion system of about 3×103 dyn/cm2. Received: 23 March 1998 / Revised version: 23 June 1998 / Accepted: 1 July 1998  相似文献   

2.
The biophysical properties of the interaction between fibronectin and its membrane receptor were inferred from adhesion tests on living cells. Individual fibroblasts were maintained on fibronectin-coated glass for short time periods (1–16 s) using optical tweezers. After contact, the trap was removed quickly, leading to either adhesion or detachment of the fibroblast. Through a stochastic analysis of bond kinetics, we derived equations of adhesion probability versus time, which fit the experimental data well and were used to compute association and dissociation rates (k +=0.3–1.4 s−1 and k off=0.05–0.25 s−1, respectively). The bond distribution is binomial, with an average bond number ≤10 at these time scales. Increasing the fibronectin density (100–3000 molecules/μm2) raised k + in a diffusion-dependent manner, leaving k off relatively unchanged. Increasing the temperature (23–37 °C) raised both k + and k off, allowing calculation of the activation energy of the chemical reaction (around 20 k B T). Increasing the compressive force on the cell during contact (up to 60 pN) raised k + in a logarithmic manner, probably through an increase in the contact area, whereas k off was unaffected. Finally, by varying the pulling force to detach the cell, we could distinguish between two adhesive regimes, one corresponding to one bond, the other to at least two bonds. This transition occurred at a force around 20 pN, interpreted as the strength of a single bond. Received: 2 November 1999 / Revised version: 6 March 2000 / Accepted: 19 April 2000  相似文献   

3.
Cell culture density is shown to alter the parameters characterizing phagocytic activity of cells in vitro. Phagocytosis index (PI, mean number of beads per cell in the bead-containing population) and phagocytosis percent (PP, percentage of bead-containing cells in cell population under study) for IC-21 macrophages incubated in the presence of non-opsonized 2-μm fluorescent latex beads were determined using fluorescent microscopy and ImageJ software specially adapted for the purpose. Under control conditions (DMEM without serum), increase in cell culture density was accompanied with a decrease of both parameters of the phagocytic activity. At a mean density of 4 cells/105 μm2 (9 cells per a viewfield) PI was 7.1 ± 0.2 beads/cell and at 20 cells/105 μm2 (40 cells per a viewfield) PI dropped to 4.6 ± 0.1 beads/cell. PP was less sensitive, varied in the range of 95–100% but also decreased as the cell density grew. At any density, PI was 1.5–2 times higher than the expected value (number of beads per μm2 × cell contour area); apparently this divergence can be accounted for by cell locomotion and capture of a larger number of beads than could drop onto a motionless cell with a constant contour area. Increase in cell density was also accompanied by a decrease of the cell contour area (S c), which amounted to 750 ± 16 μm2 at a density of 4 cells/105 μm2 and 346 ± 4 μm2 at a density of 20 cells/105 μm2. As the bead concentration was the same in all experiments, density-dependent decrease in PI and PP may be related with the observed decrease in cell contour area. Yet, the bead number per cell area unit (PI/S c) was bigger at higher density and PI/S c was higher in cells with smaller S c. Thus, individual (specific) activity of the cells did not lessen with an increase of the cell culture density in the range studied (4–20 cells/105 μm2). Reduction of the cell contour area may reflect alteration in cell adhesion to the substrate as well as competitive relations between adhesion and phagocytic processes. The data obtained imply that cell culture density has to be controlled as a factor notably altering the phagocytic activity parameters. The effects of serum, methyl-β-cyclodextrin, and carbenoxolon reported earlier [Golovkina et al. 2009. Biol membrany. 26 (5), 379–386] are re-evaluated and confirmed here.  相似文献   

4.
Integrins are transmembrane adhesion receptors that play important roles in the cardiovascular system by interacting with the extracellular matrix (ECM). However, direct quantitative measurements of the adhesion properties of the integrins on cardiomyocyte (CM) and their ECM ligands are lacking. In this study, we used atomic force microscopy (AFM) to quantify the adhesion force (peak force and mean force) and binding probability between CM integrins and three main heart tissue ECM proteins, ie, collagen (CN), fibronectin (FN), and laminin (LN). Functionalizing the AFM probes with ECM proteins, we found that the peak force (mean force) was 61.69 ± 5.5 pN (76.54 ± 4.0 pN), 39.26 ± 4.4 pN (59.84 ± 3.6 pN), and 108.31 ± 4.2 pN (129.63 ± 6.0 pN), respectively, for the bond of CN‐integrin, FN‐integrin, and LN‐integrin. The binding specificity between CM integrins and ECM proteins was verified by using monoclonal antibodies, where α10‐ and α11‐integrin bind to CN, α3‐ and α5‐integrin bind to FN, and α3‐ and α7‐integrin bind to LN. Furthermore, adhesion properties of CM integrins under physiologically high concentrations of extracellular Ca2+ and Mg2+ were tested. Additional Ca2+ reduced the adhesion mean force to 68.81 ± 4.0 pN, 49.84 ± 3.3 pN, and 119.21 ± 5.8 pN and binding probability to 0.31, 0.34, 0.40 for CN, FN, and LN, respectively, whereas Mg2+ caused very minor changes to adhesion properties of CM integrins. Thus, adhesion properties between adult murine CM integrins and its main ECM proteins were characterized, paving the way for an improved understanding of CM mechanobiology.  相似文献   

5.
Summary Atomic force microscopy (AFM) was used to measure the morphology and physicochemical properties of rhizobia and to probe cell-surface polymers with tips modified with soybean agglutinin (SBA). AFM measurements of the length, width, and height of Sinorhizobium fredii CCRC15769 were 1580±450, 870±70, and 270±50 nm, respectively (means±SD). Different AFM image modes revealed the morphology, adhesion, viscoelasticity, and surface roughness of rhizobia in air using the tapping operation. Force–distance relationships between SBA-terminated AFM probes and Bradyrhizobium japonicum USDA110 were recorded and the retraction curves showed an unbinding force of 106±48 pN with a loading rate of 1 nN/s in PBS containing 0.1 mM Mn2+ and 0.1 mM Ca2+ (pH 7.2). The technique of AFM provides information about the morphology and molecular interaction forces of rhizobia under physiological conditions.  相似文献   

6.
The adhesive properties of the mouse P388D1 macrophage-like line were explored. Cells were deposited in glass capillary tubes, and the kinetics of adhesion and spreading were studied. Binding involved the cell metabolism since it was decreased by cold, azide, or a divalent cation chelator. Glass-adherent cells were subjected to calibrated laminar shear flows with a highly viscous dextran solution. A tangential force of about 5×10−3 dyn/cell was required to achieve substantial detachment. The duration of application of the shearing force strongly influenced cell-substrate separation when this was varied from 1–10 s. Further, this treatment resulted in marked cell deformation, with the appearance of an elongated shape. Hence, cell-substrate separation is a progressive process, and binding strength is expected to be influenced by cell deformability. The minimum time required for adhesion was also investigated by making cells adhere under flow conditions. The maximum flow rate compatible with adhesion was about 1000-fold lower than that required to detach glass-bound cells. A simple model was devised to provide a quantitative interpretation for the experimental results of kinetic studies. It is concluded that cell-to-glass adhesion required a cell-substrate contact longer than a few seconds. This first step of adhesion was rapidly followed by a large (about 1000-fold) increase of adhesion strength. It is therefore emphasized that adhesion is heavily dependent on the duration of cell-to-cell encounter, as well as the force used to remove so-called unbound cells.  相似文献   

7.
The gold-rhenium (AuRe) composite nanoparticle was prepared by NaBH4 reduction procedure, and was modified by the aptamer to obtain an AuRe nanoprobe (AuRessDNA) for thrombin. In pH 7.0 Tris–HCl buffer solution and in the presence of salt, the nanoprobe specifically combined with thrombin to form AuRe-aptamer-thrombin cluster that resulted in the resonance scattering intensity (I 560 nm) increasing at 560 nm. The increased intensity ΔI 560 nm was linear to the thrombin concentration in the range of 0.115–6.93 nmol/L, with a regression equation of ΔI 560 nm = 53.0 C + 2.5, a correlation coefficient of 0.9989, and a detection limit of 13 pmol/L. This method was applied to detect thrombin in human plasma samples, with satisfactory results.  相似文献   

8.
Undifferentiated THP-1 cells from Cell Culture Collection of the Institute of Cytology, RAS (St. Petersburg), are characterized by weak expression of Toll-like receptor-4 (TLR4) on the cell surface (up to 2%) and by almost undetectable expression of CD14 and CD11b receptors. Differentiation agent phorbol-12-myristate-13-acetate independently of its concentration (2 × 10−7 M or 10−8 M) and incubation time (24 or 48 h) did not initiate CD11b surface expression and did not change the parameter Sapp (0.605 ± 0.005 at 37°C) reflecting the cell membrane viscosity. Differentiation of THP-1 cells induced by another differentiation agent, 1α,25-dihydroxyvitamin D3, caused expression of CD14 (up to 70–80%) and CD11b (up to 15–20%) receptors, again without changes in plasma membrane viscosity. The rate constants of the reduction of 5- and 16-doxyl-stearic acids by THP-1 cells were in the range of 6–8 × 10−3 s−1 at 37°C. During cell differentiation significant changes in cell electrophoretic mobility (EM, μm s−1 V−1 cm) were observed. Mean value of EM for undifferentiated THP-1 cells was −1.332 ± 0.011, whereas for phorbol-12-myristate-13-acetate- and 1α,25-dihydroxyvitamin D3-treated cells it was −1.432 ± 0.030 and −1.212 ± 0.016, respectively.  相似文献   

9.
Microcarrier culture was investigated for the propagation of attenuated hepatitis A vaccine in the anchorage-dependent human fibroblast cell line, MRC-5. Cells were cultivated at 37°C for one to two weeks, while virus accumulation was performed at 32°C over 21 to 28 days. The major development focus for the microcarrier process was the difference between the cell and virus growth phases. Virus antigen yields, growth kinetics, and cell layer/bead morphology were each examined and compared for both the microcarrier and stationary T-flask cultures. Overall, cell densities of 4–5×106 cells/ml at 5–10 g/l beads were readily attained and could be maintained in the absence of infection at either 37°C or 32°C. Upon virus inoculation, however, substantial cell density decreases were observed as well as 2.5 to 10-fold lower per cell and per unit surface area antigen yields as compared to stationary cultures. The advantages as well as the problems presented by the microcarrier approach will be discussed.  相似文献   

10.
Summary. Heat shock proteins (HSPs) are synthesised by cells subsequent to a stress exposure and are known to confer protection to the cell in response to a second challenge. HSP induction and decay are correlated to thermotolerance and may therefore be used as a biomarker of thermal history. The current study tested the temperature-dependent nature of the heat shock response and characterised its time profile of induction. Whole blood from 6 healthy males (Age: 26 ± (SD) 2 yrs; Body mass 74.2 ± 3.8 kgs; VO2max: 49.1 ± 4.0 ml·kg−1·min−1) were isolated and exposed to in vitro heat shock (HS) at 37, 38, 39, 40, and 41 °C for a period of 90 min. After HS the temperature was returned to 37 °C and intracellular HSP70 was quantified from the leukocytes at 0, 2, 4, and 6 h after heat treatment. The concentration of HSP70 was not different between temperatures (P > 0.05), but the time-profile of HSP70 synthesis appeared temperature-dependent. At control (37 °C) and lower temperatures (38–39 °C) the mean HSP70 concentration increased up to 4 h post HS (P < 0.05) and then returned towards baseline values by 6 h post HS. With in vitro hyperthermic conditions (40–41 °C), the time-profile was characterised by a sharp rise in HSP70 levels immediately after treatment (P < 0.05 for 40 °C at 0 h), followed by a progressive decline over time. The results suggest a temperature-dependent time-profile of HSP70 synthesis. In addition, the temperature at which HSP70 is inducted might be lower than 37 °C.  相似文献   

11.
To characterize the dynamics of cell-substrate adhesive rupture, we used a novel micromanipulation technique, in which individual fibroblasts seized on a rigid microplate were placed into contact with a fibronectin-coated flexible microplate, then pulled away. The fibronectin density (0–3000 molecules/μm2) and the pulling rate (1–10 μm/s) were controlled. The extent of the contact zone decreased to zero at a time threshold corresponding to adhesive rupture. The uniaxial force at the interface, computed from the deflection of the microplate, increased linearly with time and reached a maximum before dropping to zero. A deterministic model, focusing on the mean number of bonds between fibronectin and its membrane receptor on the cell surface, shows rapid rupture when the force reaches a critical value, in agreement with experimental observations. Increasing the ligand density and the rate of load raises the maximal force (30–200 nN), in reasonably good agreement with the model predictions. Minimization of error between experimental and simulated forces allowed identification of two physicochemical properties of the bond, i.e. its association rate constant (k 2D on=3 × 10−4 μm2/s) and structural length (d=3 nm). These results may help understand better fibroblast locomotion and interaction with the extracellular matrix. Received: 2 November 1999 / Revised version: 23 March 2000 / Accepted: 19 April 2000  相似文献   

12.
Dey K  Roy P 《Biotechnology letters》2011,33(6):1101-1105
A Bacillus sp., capable of degrading chloroform, was immobilized in calcium alginate. The beads in 20 g alginate l−1 (about 2 × 108 cells/bead) could be re-used nine times for degradation of chloroform at 40 μM. The immobilized cells had a higher range of tolerance (pH 6.5–9 and 20–41°C) than free cells (pH 7–8.5 and 28–32°C). At 5 g alginate l−1, leakage of the cells from the beads was 0.51 mg dry wt ml−1. This species is the first reported Bacillus that can degrade chloroform as the sole carbon source.  相似文献   

13.
 Direct cyclic voltammetry and 1H NMR spectroscopy have been combined to investigate the electrochemical and spectroscopic properties of cytochrome c 553 isolated from the alkaliphilic soil bacterium Bacillus pasteurii. A quasi-reversible diffusion-controlled redox process is exhibited by cytochrome c 553 at a pyrolitic graphite edge microelectrode. The temperature dependence of the reduction potential, measured using a non-isothermal electrochemical cell, revealed a discontinuity at 308 K. The thermodynamic parameters determined in the low-temperature range (275–308 K;ΔS°′=–162.7±1.2 J mol–1 K–1, ΔH°′=–53.0±0.5 kJ mol–1, ΔG°′=–4.5±0.1 kJ mol–1, E°′=+47.0±0.6 mV) indicate the presence of large enthalpic and entropic effects, leading, respectively, to stabilization and destabilization of the reduced form of cytochrome c 553. Both effects are more accentuated in the high-temperature range (308–323 K;ΔS°′=–294.1±8.4 J mol–1 K–1, ΔH°′=–93.4±3.1 kJ mol–1, ΔG°′=–5.8±0.6 kJ mol–1, E°′=+60.3±5.8 mV), with the net result being a slight increase of the standard reduction potential. These thermodynamic parameters are interpreted using the compensation theory of hydration of biopolymers as indicating the extrusion, upon reduction, of water molecules from the hydration sphere of the cytochrome. The low-T and high-T conformers differ by the number of water molecules in the solvation sphere: in the high-T conformer, the number of water molecules extruded upon reduction increases, as compared to the low-T conformer. The ionic strength dependence of the reduction potential at 298 K, treated within the frame of extended Debye-Hückel theory, yields values of E °′ (I=0) =–25.4±1.4 mV, z red=–11.3, and z ox=–10.3. The pH dependence of the reduction potential at 298 K shows a plateau in the pH range 7–10 and an increase at more acidic pH, allowing the calculation of pK O=5.5 and pK R=5.7, together with the estimate of the reduction potentials of completely protonated (+71 mV) and deprotonated (+58 mV) forms of cytochrome c 553. 1H NMR spectra of the oxidized paramagnetic cytochrome c 553 indicate the presence of a His-Met axial coordination of the low-spin (S=1/2) heme iron, which is maintained in the temperature interval 288–340 K at pH 7 and in the pH range 4.8–10.0 at 298 K. The temperature dependence of the hyperfine-shifted signals shows both Curie-type and anti-Curie-type behavior, with marked deviations from linearity, interpreted as indicating the presence of a fast equilibrium between the low-T and high-T conformers, having slightly different heme electronic structures resulting from the T-induced conformational change. Increasing the NaCl concentration in the range 0–0.2 M causes a slight change of the 1H NMR chemical shifts of the hyperfine-shifted signals, with no influence on their linewidth. The calculated lower limit value of the apparent affinity constant for specific ion binding is estimated as 5.2±1.1 M–1. The pH dependence of the isotropically shifted 1H NMR signals of the oxidized cytochrome displays at least one ionization step with pK O=5.7. The thermodynamic and spectroscopic data indicate a large solvent-derived entropic effect as the main cause for the observed low reduction potential of B. pasteurii cytochrome c 553. Received: 9 January 1998 / Accepted: 8 April 1998  相似文献   

14.
The increasing emergence of antibiotic-resistant bacteria has produced a growing interest among scientists in bacteriophages as alternative antimicrobial agents. This article reports a lytic phage against an antibiotic-resistant strain of Pseudomonas aeruginosa. Phage BVPaP-3 is a member of the Podoviridae family and morphologically similar to the T7-like phage gh-1. The phage has a hexagonal head of 58–59 nm in diameter and a short tail of 10 × 8 nm. It is stable at a wide range of pH (6–10) and temperatures (4–40°C). Its optimal growth temperature is 37°C and the adsorption rate constant is 1.19 × 10−9. Latent and eclipse periods are 20 and 15 min, respectively, and the burst size is 44 after 35 min at 37°C. The phage has a DNA size of 41.31 kb and a proteome of 11 proteins. The major protein is 33 kDa in size.  相似文献   

15.
Microtuberization in potato (Solanum tuberosum L.)   总被引:7,自引:0,他引:7  
Twenty-two genotypes of potato (Solanum tuberosum L.) were induced to form microtubers under six in vitro culture conditions. Cultures maintained under a short photoperiod (10 h of 6–12 μmol m–2 s–1) and low temperatures (day 20°±2°C and night 18°±2°C) had both a higher yield (255 mg/plantlet) and a greater number (2/plantlet) of microtubers than those maintained under long days (16 h of 38–50 μmol m–2 s–1) combined with high temperatures (day 28°±2°C and night 25°±2°C) (yield 207 mg/plantlet; microtuber number, 0.9/plantlet), over a wide range of genotypes. After the plantlets had been cultured under long days for an initial period of 60 days, continuous darkness advanced microtuberization by 2–3 months in various genotypes. Under short-day and low-temperature conditions the addition of 6-benzylaminopurine increased microtuber yield from 255 mg/plantlet to 645 mg/plantlet and average microtuber weight from 115 mg to 364 mg. A similar pattern was observed under conditions of long days and high temperature, and continuous darkness and low-temperature. Microtubers produced under light had a greater number of eyes (maximum average: 5.96/microtuber) than those produced in the dark (maximum average: 3.50/plantlet). The genotype × cultural conditions interactions were significant indicating the importance of developing genotype-specific protocols to maximize microtuberization. Received: 17 September 1997 / Revision received: 12 December 1997 / Accepted: 1 January 1998  相似文献   

16.
Atomic force microscopy (AFM) indentation has become an important technique for quantifying the mechanical properties of live cells at nanoscale. However, determination of cell elasticity modulus from the force–displacement curves measured in the AFM indentations is not a trivial task. The present work shows that these force–displacement curves are affected by indenter-cell adhesion force, while the use of an appropriate indentation model may provide information on the cell elasticity and the work of adhesion of the cell membrane to the surface of the AFM probes. A recently proposed indentation model (Sirghi, Rossi in Appl Phys Lett 89:243118, 2006), which accounts for the effect of the adhesion force in nanoscale indentation, is applied to the AFM indentation experiments performed on live cells with pyramidal indenters. The model considers that the indentation force equilibrates the elastic force of the cell cytoskeleton and the adhesion force of the cell membrane. It is assumed that the indenter-cell contact area and the adhesion force decrease continuously during the unloading part of the indentation (peeling model). Force–displacement curves measured in indentation experiments performed with silicon nitride AFM probes with pyramidal tips on live cells (mouse fibroblast Balb/c3T3 clone A31-1-1) in physiological medium at 37°C agree well with the theoretical prediction and are used to determine the cell elasticity modulus and indenter-cell work of adhesion. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

17.
In this study, we report the cloning, recombinant expression, and biochemical characterization of a heat-stable CMP-N-acylneuraminic acid (NeuAc) synthetase from Clostridium thermocellum ATCC 27405. A high throughput electrospray ionization mass spectrometry (ESI-MS)-based assay demonstrates that the enzyme has an absolute requirement for a divalent cation for activity and reaches maximum activity in the presence of 10 mM Mn2+. The enzyme is active at pH 8–13 in Tris–HCl buffer and at 37–60 °C, and maximum activity is observed at pH 9.5 and 50 °C in the presence of 0.2 mM dithiothreitol. In addition to NeuAc, the enzyme also accepts the analog N-glycolylneuraminic acid (NeuGc) as a substrate. The apparent Michaelis constants for cytidine triphosphate and NeuAc or NeuGc are 240 ± 20, 130 ± 10, and 160 ± 10 μM, respectively, with corresponding turnover numbers of 3.33, 2.25, and 1.66 s−1, respectively. An initial velocity study of the enzymatic reaction indicates an ordered bi–bi catalytic mechanism. In addition to demonstration of a thermostable and substrate-tolerant enzyme, confirmation of the biochemical function of a gene for CMP-NeuAc synthetase in C. thermocellum also opens the question of the biological function of CMP-NeuAc in such nonpathogenic microorganisms.  相似文献   

18.
Thirty-three enterococcal strains and 10 Streptococcus bovis strains were investigated for their protein-binding cell surface components. Seven extracellular matrix (ECM) proteins were immobilized on Difco latex beads to detect these components on the surface of all enterococcal strains and eight non-autoaggregating S. bovis strains by a particle agglutination assay (PAA). Twenty-three selected strains were also examined in microtiter plate assays. According to the absorbance readings (A570nm), 11 strains were classified as nonadherent (A570nm < 0.1), 10 strains as weakly adherent (0.1 < A570nm > 0.3), and 2 strains as strongly adherent (A570nm > 0.3) in these assays. A direct correlation was found between the values obtained in PAA and A570nm readings of microtiter plate assays. Binding of 125I-labeled bovine lactoferrin to enterococci and streptococci was in the range of 6%–30% and of 125I-labeled human vitronectin in the range of 9%–33% to streptococci. The binding of 125I-labeled ECM proteins to selected strains was much more effectively inhibited by sulfated carbohydrates than by non-sulfated hyaluronic acid, indicating the importance of the sulfate groups of these inhibitors. An inhibition effect of heparin on bLf binding to four selected strains was higher in comparison with fucoidan in the microtiter plates. Thirty-five out of 44 strains had agglutinated rabbit erythrocytes. However, these strains showed no ability to agglutinate bovine or sheep erythrocytes. Received: 28 April 1999 / Accepted: 26 July 1999  相似文献   

19.
Lipase-catalyzed synthesis of isoamyl acetate in hexane at 10–250 MPa at 80°C and 1–100 MPa at 40°C resulted in activation volumes of −12.9 ± 1.7 and −21.6 ± 2.9 cm3 mol−1, respectively. Increasing pressure from 10 to 200 MPa resulted in approximately 10-fold increase in V max at both 40 and 80°C. Pressure increased the K m from 2.4 ± 0.004 to 38 ± 0.78 mM at 40°C. In contrast, at 80°C the pressure did not affect the K m.  相似文献   

20.
Single-cell force spectroscopy was used to investigate the initial adhesion of L929 fibroblasts onto periodically grooved titanium microstructures (height ~6 μm, groove width 20 μm). The position-dependent local adhesion strength of the cells was correlated with their rheological behavior. Spherical cells exhibited a significantly lower Young’s modulus (<1 kPa) than that reported for spread cells, and their elastic properties can roughly be explained by the Hertz model for an elastic sphere. While in contact with the planar regions of the substrate, the cells started to adapt their shape through slight ventral flattening. The process was found to be independent of the applied contact force for values between 100 and 1,000 pN. The degree of flattening correlated with the adhesion strength during the first 60 s. Adhesion strength can be described by fast exponential kinetics as C1 [ 1 - exp( - C2 ·t ) ] C_{1} \left[ {1 - \exp \left( { - C_{2} \cdot t} \right)} \right] with C 1 = 2.34 ± 0.19 nN and C 2 = 0.09 ± 0.02 s−1. A significant drop in the adhesion strength of up to 50% was found near the groove edges. The effect can be interpreted by the geometric decrease of the contact area, which indicates the inability of the fibroblasts to adapt to the shape of the substrate. Our results explain the role of the substrate’s topography in contact guidance and suggest that rheological cell properties must be considered in cell adhesion modeling.  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