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 46 毫秒
1.
There is increasing interest in supported membranes as models of biological membranes and as a physiological matrix for studying the structure and function of membrane proteins and receptors. A common problem of protein-lipid bilayers that are directly supported on a hydrophilic substrate is nonphysiological interactions of integral membrane proteins with the solid support to the extent that they will not diffuse in the plane of the membrane. To alleviate some of these problems we have developed a new tethered polymer-supported planar lipid bilayer system, which permitted us to reconstitute integral membrane proteins in a laterally mobile form. We have supported lipid bilayers on a newly designed polyethyleneglycol cushion, which provided a soft support and, for increased stability, covalent linkage of the membranes to the supporting quartz or glass substrates. The formation and morphology of the bilayers were followed by total internal reflection and epifluorescence microscopy, and the lateral diffusion of the lipids and proteins in the bilayer was monitored by fluorescence recovery after photobleaching. Uniform bilayers with high lateral lipid diffusion coefficients (0.8-1.2 x 10(-8) cm(2)/s) were observed when the polymer concentration was kept slightly below the mushroom-to-brush transition. Cytochrome b(5) and annexin V were used as first test proteins in this system. When reconstituted in supported bilayers that were directly supported on quartz, both proteins were largely immobile with mobile fractions < 25%. However, two populations of laterally mobile proteins were observed in the polymer-supported bilayers. Approximately 25% of cytochrome b(5) diffused with a diffusion coefficient of approximately 1 x 10(-8) cm(2)/s, and 50-60% diffused with a diffusion coefficient of approximately 2 x 10(-10) cm(2)/s. Similarly, one-third of annexin V diffused with a diffusion coefficient of approximately 3 x 10(-9) cm(2)/s, and two-thirds diffused with a diffusion coefficient of approximately 4 x 10(-10) cm(2)/s. A model for the interaction of these proteins with the underlying polymer is discussed.  相似文献   

2.
Monoclonal antibodies (MoAbs) were developed against epidermal growth factor (EGF) receptor on the human epidermoid carcinoma cell line A431. The A431 antigen recognized by the MoAbs has an apparent molecular weight of approximately 170,000, with the same molecular weight as the CNE-2 cell line (poorly differentiated nasopharyngeal carcinoma). Administration of anti-EGF receptor MoAbs inhibited tumor formation, caused by the CNE-2 and A431 cell lines, in athymic mice. When the same MoAbs were used in therapy against Tca8113 (a human tongue carcinoma) and HeLa cells (a human cervical carcinoma), tumor growth was not affected. The number of EGF receptors and the apparent dissociation constants for 125I-EGF on CNE-2 and A431 were 1.3 x 10(5)/cell (Kd 7.7 x 10(-8) M) and 1.4 x 10(6)/cell (Kd 2.4 x 10(-9) M), respectively. Three anti-EGF receptor MoAbs were used in these studies. MoAbs 3 and 176, capable of competing with EGF for receptor binding, showed significant tumor growth inhibition. MoAb 101 was incapable of blocking the binding of EGF to its receptor and was not as effective as MoAbs 3 and 176 in tumor growth inhibition. Our observation is that in vitro, MoAb anti-EGF receptor is cytostatic, rather than cytocidal, against CNE-2 and A431.  相似文献   

3.
Previous studies have compared the oestrogenic properties of phytoestrogens in a wide variety of disparate assays. Since not all phytoestrogens have been tested in each assay, this makes inter-study comparisons and ranking oestrogenic potency difficult. In this report, we have compared the oestrogen agonist and antagonist activity of eight phytoestrogens (genistein, daidzein, equol, miroestrol, deoxymiroestrol, 8-prenylnaringenin, coumestrol and resveratrol) in a range of assays all based within the same receptor and cellular context of the MCF7 human breast cancer cell line. The relative binding of each phytoestrogen to oestrogen receptor (ER) of MCF7 cytosol was calculated from the molar excess needed for 50% inhibition of 3H]oestradiol binding (IC50), and was in the order coumestrol (35x)/8-prenylnaringenin (45x)/deoxymiroestrol (50x)>miroestrol (260x)>genistein (1000x)>equol (4000x)>daidzein (not achieved: 40% inhibition at 10(4)-fold molar excess)>resveratrol (not achieved: 10% inhibition at 10(5)-fold molar excess). For cell-based assays, the rank order of potency (estimated in terms of the concentration needed to achieve a response equivalent to 50% of that found with 17beta-oestradiol (IC50)) remained very similar for all the assays whether measuring ligand ability to induce a stably transfected oestrogen-responsive ERE-CAT reporter gene, cell growth in terms of proliferation rate after 7 days or cell growth in terms of saturation density after 14 days. The IC50 values for these three assays in order were for 17beta-oestradiol (1 x 10(-11)M, 1 x 10(-11)M, 2 x 10(-11)M), and in rank order of potency for the phytoestrogens, deoxymiroestrol (1 x 10(-10)M, 3 x 10(-11)M, 2 x 10(-11)M)>miroestrol (3 x 10(-10)M, 2 x 10(-10)M, 8 x 10(-11)M)>8-prenylnaringenin (1 x 10(-9)M, 3 x 10(-10)M, 3 x 10(-10)M)>coumestrol (3 x 10(-8)M, 2 x 10(-8)M, 3 x 10(-8)M)>genistein (4 x 10(-8)M, 2 x 10(-8)M, 1 x 10(-8)M)/equol (1 x 10(-7)M, 3 x 10(-8)M, 2 x 10(-8)M)>daidzein (3 x 10(-7)M, 2 x 10(-7)M, 4 x 10(-8)M)>resveratrol (4 x 10(-6)M, not achieved, not achieved). Despite using the same receptor context of the MCF7 cells, this rank order differed from that determined from receptor binding. The most marked difference was for coumestrol and 8-prenylnaringenin which both displayed a relatively potent ability to displace [3H]oestradiol from cytosolic ER compared with their much lower activity in the cell-based assays. Albeit at varying concentrations, seven of the eight phytoestrogens (all except resveratrol) gave similar maximal responses to that given by 17beta-oestradiol in cell-based assays which makes them full oestrogen agonists. We found no evidence for any oestrogen antagonist action of any of these phytoestrogens at concentrations of up to 10(-6)M on either reporter gene induction or on stimulation of cell growth.  相似文献   

4.
Short peptides sequences were selected that showed binding selectivity towards healthy or oxidised (unhealthy) low density lipoprotein (LDL), respectively. These were investigated for application in atherosclerosis risk monitoring. Comparison was also made with the LDL receptor ligand repeat peptide (LR5). The peptides were immobilised on a gold surface plasmon resonance surface and LDL binding detected as a shift in the resonance. 3.7x10(7) (+/-5.6x10(6)) LDL/mm(2)/microg/ml solution LDL were bound on GlySerAspGlu-OH and 6.8x10(7) (+/-9.2x10(6)) LDL/mm(2)/microg/ml on GlyCystineSerAspGlu, compared with approximately 10(8) LDL/mm(2)/microg/ml on LR5. In this first group, binding of LDL decreased with oxidation level and a good correlation was found between LDL binding and residual amino groups on the apoprotein of the LDL following oxidation, or the change in relative electrophoretic mobility (REM) of LDL. The decrease in binding was 1.1x10(7) LDL particles/mm(2) per% oxidation for GlySerAspGlu-OH, 1.8x10(7) LDL particles/mm(2) per% oxidation for GlyCystineSerAspGlu and 2.4x10(7) LDL particles/mm(2) per% oxidation for LR5. A second group of three peptides were also selected showing increased binding with LDL oxidation: GlyCystineCysCys (1.5x10(7) LDL/mm(2) per microg/ml), GlyLysLysCys-SH (10(7) LDL/mm(2) per microg/ml) and GlyLysLys-OH (5.6x10(7) LDL/mm(2) per microg/ml). The latter gave a linear increase in LDL binding with oxidation level (1.2x10(7) LDL particles/mm(2) per% oxidation). LDL concentration is around 2-3 mg/ml in plasma compared with the low detection levels with this method (1-10 microg/ml), allowing a strategy to be developed requiring the minimum sample volume and diluting with physiological buffer prior to assay. By using a comparative reading between LDL adsorption on surfaces from the first and second group of peptides (e.g. GlyCystineSerAspGlu and GlyLysLys-OH, respectively), LDL oxidation could be determined without knowledge of LDL concentration. Higher binding was seen on GlyCystineSerAspGlu than GlyLysLys-OH below 30% LDL oxidation, whereas above 30% oxidation the binding on the latter surface was greater. Simple correlation of this form could provide good tests for atherosclerosis risk.  相似文献   

5.
We have produced recombinant proteins for a cytokine, L2G25BP (macrophage inflammatory protein-1 alpha) (MIP-1 alpha). By using the recombinant protein (rMIP-1 alpha), receptors for MIP-1 alpha were identified on Con A-stimulated and unstimulated CTLL-R8, a T cell line, and LPS-stimulated RAW 264.7, a macrophage cell line. The 125I-rMIP-1 alpha binds to the receptor in a specific and saturable manner. Scatchard analysis indicated a single class of high affinity receptor, with a Kd of approximately 1.5 x 10(-9) M and approximately 1200 binding sites/Con A-stimulated CTLL-R8 cell and a Kd of 0.9 x 10(-9) M and approximately 380 binding sites/RAW 264.7 cell. 125I-rMIP-1 alpha binding was inhibited by unlabeled rMIP-1 alpha in a dose-dependent manner, but not by IL-1 alpha or IL-2. rMIP-1 alpha inhibited the proliferation of unstimulated CTLL-R8 cells. Rabbit anti-rMIP-1 alpha antibodies blocked the growth-inhibitory effect of the rMIP-1 alpha on CTLL-R8 cells.  相似文献   

6.
This study presents a generic numerical model to simulate the coupled solute and solvent transport in human ovarian tissue sections during addition and removal of chemical additives or cryoprotective agents (CPA). The model accounts for the axial and radial diffusion of the solute (CPA) as well as axial convection of the CPA, and a variable vascular surface area (A) during the transport process. In addition, the model also accounts for the radial movement of the solvent (water) into and out of the vascular spaces. Osmotic responses of various cells within an human ovarian tissue section are predicted by the numerical model with three model parameters: permeability of the tissue cell membrane to water (L(p)), permeability of the tissue cell membrane to the solute or CPA (omega) and the diffusion coefficient of the solute or CPA in the vascular space (D). By fitting the model results with published experimental data on solute/water concentrations within an human ovarian tissue section, I was able to determine the permeability parameters of ovarian tissue cells in the presence of 1.5M solutions of each of the following: dimethyl sulphoxide (DMSO), propylene glycol (PROH), ethylene glycol (EG), and glycerol (GLY), at two temperatures (4 degrees C and 27 degrees C). Modeling Approach 1: Assuming a constant value of solute diffusivity (D = 1.0 x 10(-9) m(2)/sec), the best fit values of L(p) ranged from 0.35 x 10(-14) to 1.43 x 10(-14) m(3)/N-sec while omega ranged from 2.57 x 10(-14) to 70.5 x 10(-14) mol/N-sec. Based on these values of L(p) and omega, the solute reflection coefficient, sigma defined as sigma = 1-omega v(CPA)/L(P) ranged from 0.9961 to 0.9996. Modeling Approach 2: The relative values of omega and sigma from our initial modeling suggest that the embedded ovarian tissue cells are relatively impermeable to all the CPAs investigated (or omega approximately 0 and sigma approximately 1.0). Consequently the model was modified and used to predict the values of L(p) and D assuming omega = 0 and sigma = 1.0. The best fit values of L(p) ranged from 0.44 x 10(-14) to 1.2 x 10(-14) m(3)/N-sec while D ranged from 0.85 x 10(-9) to 2.08 x 10(-9) m(2)/sec. Modeling Approach 3: Finally, the best fit values of D from modeling approach 2 were incorporated into model 1 to re-predict the values of L(p) and omega. It is hoped that the ovarian tissue cell parameters reported here will help to optimize chemical loading and unloading procedures for whole ovarian tissue sections and consequently, tissue cryopreservation procedures.  相似文献   

7.
BACKGROUND: Tandem stem cell transplantation is an important treatment option for patients with myeloma and some additional tumors. In an attempt to reduce the contamination of the stem cell graft with tumor cells, patients with myeloma who entered complete remission after the first transplant underwent a second episode of mobilization to obtain progenitor cells for the second transplant. METHODS: Twenty-two patients with myeloma participated in the study. The first mobilization utilized CY, etoposide and filgrastim. The second mobilization used the same regimen, but seven patients received only filgrastim. The interval between the two collection periods was 6 months (median; range 4-9 months). The preparative regimen for the first transplant consisted of melphalan 200 mg/m(2). RESULTS: The number of total white cells collected during the two collection episodes was similar: 10.8+/-1.6 x 10(8)/kg white cells vs. 11.8+/-1.7 x 10(8)/kg white cells (P=0.63). The collected CD34(+) cell dose was much larger during the first collection: 45.2+/-8.4 x 10(6)/kg vs. 6.9+/-2.7 x 10(6)/kg (P<0.001). Similarly, the collected colony-forming unit (CFU)-GM dose was much larger during the first collection: 295.4+/-59.3 x 10(4)/kg vs. 67.3+/-21.6x10(4)/kg (P<0.001). While the CD34(+) cells collected during the two collection episodes correlated significantly (r=0.55, P<0.01); the first dose was a median of 14.9-fold larger. DISCUSSION: No laboratory parameter was able reliably to predict the results of the second collection. A second mobilization/collection episode as part of a tandem transplant approach carries a considerable risk of failing to obtain sufficient progenitor cells.  相似文献   

8.
The agonist-induced dynamic regulation of the beta(2)-adrenergic receptor (beta(2)-AR) on living cells was examined by means of fluorescence correlation spectroscopy (FCS) using a fluorescence-labeled arterenol derivative (Alexa-NA) in hippocampal neurons and in alveolar epithelial type II cell line A549. Alexa-NA specifically bound to the beta(2)-AR of neurons with a K(D) value of 1.29 +/- 0.31 nM and of A549 cells with a K(D) of 5.98 +/- 1.62 nM. The receptor density equaled 4.5 +/- 0.9 microm(-2) in neurons (rho(N)) and 19.9 +/- 2.0 microm(-2) in A549 cells (rho(A549)). Kinetic experiments revealed comparable on-rate constants in both cell types (k(on) = 0.49 +/- 0.03 s(-1) nM(-1) in neurons and k(on) = 0.12 +/- 0.02 s(-1) nM(-1) in A549 cells). In addition to the free ligand diffusing with a D(free) of (2.11 +/- 0.04) x 10(-6) cm(2)/s, in both cell types receptor-ligand complexes with two distinct diffusion coefficients, D(bound1) (fast lateral mobility) and D(bound2) (hindered mobility), were observed [D(bound1) = (5.23 +/- 0.64) x 10(-8) cm(2)/s and D(bound2) = (6.05 +/- 0.23) x 10(-10) cm(2)/s for neurons, and D(bound1) = (2.88 +/- 1.72) x 10(-8) cm(2)/s and D(bound2) = (1.01 +/- 0.46) x 10(-9) cm(2)/s for A549 cells]. Fast lateral mobility of the receptor-ligand complex was detected immediately after addition of the ligand, whereas hindered mobility (D(bound2)) was observed after a delay of 5 min in neurons (up to 38% of total binding) and of 15-20 min in A549 cells (up to 40% of total binding). Thus, the receptor-ligand complexes with low mobility were formed during receptor regulation. Consistently, stimulation of receptor internalization using the adenylate cyclase activator forskolin shifted the ratio of receptor-ligand complexes toward D(bound2). Intracellular FCS measurements and immunocytochemical studies confirmed the appearance of endocytosed receptor-ligand complexes in the cytoplasm subjacent to the plasma membrane after stimulation with the agonist terbutaline (1 microM). This regulatory receptor internalization was blocked after preincubation with propranolol and with a cholesterol-complexing saponin alpha-hederin.  相似文献   

9.
The microstructure and permeability of rehydrated 20-100 microm thick partially coalesced (vinyl-actetate acrylic copolymer) SF091 latex coatings and a 118 microm thick model trilayer biocatalytic coating consisting of two sealant SF091 layers containing a middle layer of viable E. coli HB101 + latex were studied as delaminated films in a diffusion apparatus with KNO(3) as the diffussant. The permeability of the hydrated coatings is due to diffusive transport through the pore space between the partially coalesced SF091 latex particles. Coating microstructure was visualized by fast freeze cryogenic scanning electron microscopy (cryo-SEM). The effective diffusion coefficient of SF091 latex coatings (diffusive permeability/film thickness) was determined as the ratio of the effective diffusivity of KNO(3) to its diffusivity in water (D(eff)/D). Polymer particle coalescence was arrested by two methods to increase coating permeability. The first used glycerol with coating drying at 4 degrees C, near the glass transition temperature (T(g)). The second method used sucrose or trehalose as a filler to arrest coalescence; the filler was then dissolved away. D(eff)/D was measured as a function of film thickness; content of glycerol, sucrose, and trehalose; drying time; and rehydration time. D(eff)/D varied from 3 x 10(-4) for unmodified SF091 coatings to 6.8 x 10(-2) for coatings containing sucrose. D(eff)/D was reduced by the flattening of latex particles against the surface of the solid substrate, as well as by the presence of the colloid stabilizer hydroxyethylcellulose (HEC). When corrected for the flattened particle layer, D(eff)/D of HEC-free coatings was as high as 0.20, which agreed with the value predicted from analysis of cryo-SEM images of the coat surface. D(eff)/D decreased by one-half in approximately 5 days in rehydrated SF091 coatings, indicating that significant wet coalescence occurs after glycerol, sucrose, or trehalose are leached from the films. D(eff)/D of SF091 latex trilayer coatings containing viable E. coli HB101 cells decreased as cell loading was increased from 2.2 x 10(-2) for 64 g dry cell weight per liter of coat volume to 5 x 10(-3) for 151 g DCW/L of coat volume. The reduction in coating permeability with increasing cell loading is predicted by Maxwell's equation for D(eff)/D in periodic composites.  相似文献   

10.
Canada generates approximately 1.45 x 10(8)t of residual biomass per year, containing an estimated energy value of 2.28 x 10(9)GJ, which is equivalent to about 22% of Canada's current annual energy use. Anaerobic digestion of these biomass residues using conventional technologies could generate 1.14 x 10(10)m(3)/year of CH(4) with a heating value of 4.56 x 10(8)GJ. Conversion of these residues using emerging technologies that favor the synthesis of H(2) and represses the synthesis of CH(4) could generate 1.47 x 10(10)m(3)/year renewable H(2), with a heating value of 1.89 x 10(8)GJ. While CH(4)-production results in a larger amount of energy recovery, generating H(2) from waste biomass is a renewable alternative that could fuel the hydrogen economy. Additional research to further both the technical and commercial development of microbial bio-energy from biomass is warranted.  相似文献   

11.
A partially purified enzyme (120-fold) from Leuconostoc mesenteroides catalyzed the reversible N-acetylation of d-glucosamine-6-phosphate. Coenzyme A was not required and inhibited the reaction rate. Neither d-glucosamine nor N-acetyl-d-glucosamine served as a substrate for the reversible reaction. The enzyme preparation retained 50% of its original activity after 5 min at 100 C. The K(m) for acetate was 7.7 x 10(-2)m in the presence of 2 x 10(-2)md-glucosamine-6-phosphate. The K(m) for d-glucosamine-6-phosphate was 5.0 x 10(-3)m in the presence of 0.64 m acetate. The product of the reaction was characterized by comparison with N-acetyl-d-glucosamine-6-phosphate prepared by enzymatic phosphorylation of N-acetyl-d-glusamine. The characterization tests were: chromatographic migration, acid hydrolysis, enzymatic dephosphorylation, sodium borohydride reduction, and periodate oxidation. The equilibrium constant for the reaction was about 7.5 m for the expression K = (d-glucosamine-6-phosphate)(acetate)/N-acetyl-d-glucosamine-6-phosphate. The standard free energy of the reaction was approximately 1,200 cal per mole.  相似文献   

12.
D. E. Fosket  J. G. Torrey 《Plant physiology》1969,44(6):871-873,875-880
The relationship between tracheary element differentiation, cell proliferation and growth hormones was examined in agar-grown soybean callus. The time course of cell division and tracheary element formation in tissues grown on a medium containing 5 x 10(-7)m kinetin and 10(-5)m NAA was determined by means of maceration technique. After a slight lag period, a logarithmic increase in cell number was observed through the twelfth day of the culture period. Cell numbers increased at a considerably slower rate after the twelfth day. The rate of tracheary element formation varied with the rate of cell proliferation. Tracheary elements increased logarithmically during the log phase of growth. As the rate of cell division decreased after the twelfth day of culture, the rate of tracheary element formation also decreased. In the presence of 10(-5)m NAA, cell number increased as the kinetin concentration was increased between 10(-9) and 10(-6)m. However, tracheary element formation was not initiated unless the kinetin concentration was 5 x 10(-8)m or above. When the Biloxi callus was subcultured repeatedly on media containing 10(-8)m kinetin, a tracheary element-free population of cells was obtained. This undifferentiated tissue produced tracheary elements upon transfer to a medium containing 5 x 10(-7)m kinetin. In the presence of 5 x 10(-7)m kinetin, NAA stimulated cell proliferation between 10(-7) and 10(-5)m, but no tracheary elements were formed without auxin, or with 10(-7)m NAA. Neither NAA nor kinetin at any concentration tested stimulated tracheary element formation in the absence of an effective level of the other hormone. However, 2,4-D at 10(-7) or 10(-6)m promoted both cell proliferation and tracheary element differentiation in the absence of an exogenous cytokinin.  相似文献   

13.
The frequency of sister-chromatid exchange (SCE) was studied in Chinese hamster ovary (CHO) cell lines with stable insertions of the vector pIII-14gpt which contains 2 truncated neomycin resistance (neo) gene fragments. Recombination between regions of homology in the 2 fragments can restore a functional neo gene and make the cell resistant to the antibiotic G418, a neomycin analogue. Unequal SCE would be one of several possible mechanisms for this event. The observed spontaneous rate of formation of G418-resistant subclones was approximately 6.4 x 10(-6) per cell per generation, as compared to the estimated spontaneous frequency of 3 SCE per cell per generation. Given this SCE frequency, the probability of an SCE occurring in a target site of about 1600 bp (the distance separating the homologous regions in the neo fragments) would be about 8 x 10(-7) per cell per generation, or approximately one tenth of the estimated rate of recombination. Treatment of the cells with methyl methanesulfonate (MMS, 50 x 10(-6) M) induced about 80-90 SCE per cell, corresponding to a probability of 2 x 10(-5) SCE per 1600-bp target per cell. In the same cell culture, MMS treatment induced 4-8 x 10(-4) recombination events per cell giving rise to G418 resistance. Cells treated with HN2 (up to 4 x 10(-6) M) showed a significant increase in SCEs, but no change in the frequency of G418-resistant revertants. These results suggest that the 2 pathways leading to SCE and recombination respectively are uncoupled, and only a small fraction of the recombination events, if any, are due to unequal SCE in this system.  相似文献   

14.
B. G. Hall 《Genetics》1988,120(4):887-897
Escherichia coli K12 strain chi 342LD requires two mutations in the bgl (beta-glucosidase) operon, bglR0----bglR+ and excision of IS103 from within bglF, in order to utilize salicin. In growing cells the two mutations occur at rates of 4 x 10(-8) per cell division and less than 2 x 10(-12) per cell division, respectively. In 2-3-week-old colonies on MacConkey salicin plates the double mutants occur at frequencies of 10(-8) per cell, yet the rate of an unselected mutation, resistance to valine, is unaffected. The two mutations occur sequentially. Colonies that are 8-12 days old contain from 1% to about 10% IS103 excision mutants, from which the Sal+ secondary bglR0----bglR+ mutants arise. It is shown that the excision mutants are not advantageous within colonies; thus, they must result from a burst of independent excisions late in the life of the colony. Excision of IS103 occurs only on medium containing salicin, despite the fact that the excision itself confers no detectable selective advantage and serves only to create the potential for a secondary selectively advantageous mutation.  相似文献   

15.
A new ligand L, 1-[3-(2-pyridyl)pyrazol-1-ylmethyl]naphthalene, and its two metal complexes, [Cu(L)3](ClO4)2 (1) and [Zn(L)3](ClO4)2(H2O)2 (2), have been synthesized and characterized. The crystal structure of complex 1 was determined by single crystal X-ray diffraction, which crystallized in monoclinic, space group P2(1)/n with unit cell parameters, a = 12.710(4) angstroms, b = 12.135(3) angstroms, c = 33.450(9) angstroms, beta = 93.281(5) degrees and Z = 4. The Cu atom was six-coordinated to N(1), N(2), N(4), N(5), N(7) and N(8) from three L ligands and formed a slightly distorted octahedral geometry. Complexes 1 and 2, and ligand L were subjected to biological tests in vitro using three different cancer cell lines (HL-60, BGC-823 and MDA-MB-435). Complex 1 showed significant cytotoxic activity against three cancer cell lines. The interactions of complexes 1 and 2, and ligand L with calf thymus DNA were then investigated by thermal denaturation, viscosity measurements and spectrophotometric methods. The experimental results indicated that complexes 1 and 2 bound to DNA by intercalative mode via the ligand L. The intrinsic binding constants of complexes 1 and 2, and ligand L with DNA were 1.8 x 10(4), 5.4 x 10(3) and 2.76 x 10(3) M(-1), respectively.  相似文献   

16.
Synchronization and synchronous growth of a cell wall-less mutant of Chlamydomonas reinhardii have been described. The following growth conditions were used: A modified Sueokas' high salt minimal medium, 1410 h light-dark cycle, growth temperature 30°C, light intensity 12–18 Klux and dilution of the culture at the end of the dark to a constant cell density of 1.0·106 cells/ml. The time course of increase and distribution of cell volume, cytoplasmic and nuclear division, release of motile cells after the division period and accumulation of DNA, RNA and protein are reported. These mutant cells did not make any sporangium in which the dividing cells were kept as a unit inside a mother cell wall. However, they usually adhered during the period of division, thus making clumps containing 2, 4 and 8 cells. Several of these cell clumps dissolved releasing either single or couples of 2 and 4 cells. After the end of division the cells became flagellated and motile and thereby releasing themselves from the aggregate.Non-Standard Abbreviations AWV average weighed cell volume - MM minimal medium - HSM high salt medium - TCA trichloroacetic acid  相似文献   

17.
Mutations in the aquaporin-2 (AQP2) water channel cause the hereditary renal disease nephrogenic diabetes insipidus (NDI). The missense mutation AQP2-T126M causes human recessive NDI by retention at the endoplasmic reticulum (ER) of renal epithelial cells. To determine whether the ER retention of AQP2-T126M is due to relative immobilization in the ER, we measured by fluorescence recovery after photobleaching the intramembrane mobility of green fluorescent protein (GFP) chimeras containing human wild-type and mutant AQP2. In transfected LLC-PK1 renal epithelial cells, GFP-labeled AQP2-T126M was localized to the ER, and wild-type AQP2 to endosomes and the plasma membrane; both were localized to the ER after brefeldin A treatment. Photobleaching with image detection indicated that the GFP-AQP2 chimeras were freely mobile throughout the ER. Quantitative spot photobleaching revealed a diffusion-dependent irreversible process whose recovery depended on spot size and was abolished by paraformaldehyde fixation. In addition, a novel slow reversible fluorescence recovery (t(12) approximately 2 s) was characterized whose recovery was independent of spot size and not affected by fixation. AQP2 translational diffusion in the ER was not slowed by the T126M mutation; diffusion coefficients were (in cm(2)/s x 10(-)10) 2.6 +/- 0.5 (wild-type) and 3.0 +/- 0.4 (T126M). Much faster diffusion was found for a lipid probe (diOC(4)(3), 2.7 x 10(-)8 cm(2)/s) in the ER membrane and for unconjugated GFP in the aqueous ER lumen (6 x 10(-)8 cm(2)/s). ER diffusion of GFP-T126M was not significantly affected by up-regulation of molecular chaperones, cAMP activation, or actin filament disruption. ATP depletion by 2-deoxyglucose and azide resulted in comparable slowing/immobilization of wild-type and T126M AQP2. These results indicate that the ER retention of AQP2-T126M does not result from restricted or slowed mobility and suggest that the majority of AQP2-T126M is not aggregated or bound to slowly moving membrane proteins.  相似文献   

18.
Human angiotensin-converting enzyme-related carboxypeptidase (ACE2) is a zinc metalloprotease whose closest homolog is angiotensin I-converting enzyme. To begin to elucidate the physiological role of ACE2, ACE2 was purified, and its catalytic activity was characterized. ACE2 proteolytic activity has a pH optimum of 6.5 and is enhanced by monovalent anions, which is consistent with the activity of ACE. ACE2 activity is increased approximately 10-fold by Cl(-) and F(-) but is unaffected by Br(-). ACE2 was screened for hydrolytic activity against a panel of 126 biological peptides, using liquid chromatography-mass spectrometry detection. Eleven of the peptides were hydrolyzed by ACE2, and in each case, the proteolytic activity resulted in removal of the C-terminal residue only. ACE2 hydrolyzes three of the peptides with high catalytic efficiency: angiotensin II () (k(cat)/K(m) = 1.9 x 10(6) m(-1) s(-1)), apelin-13 (k(cat)/K(m) = 2.1 x 10(6) m(-1) s(-1)), and dynorphin A 1-13 (k(cat)/K(m) = 3.1 x 10(6) m(-1) s(-1)). The ACE2 catalytic efficiency is 400-fold higher with angiotensin II () as a substrate than with angiotensin I (). ACE2 also efficiently hydrolyzes des-Arg(9)-bradykinin (k(cat)/K(m) = 1.3 x 10(5) m(-1) s(-1)), but it does not hydrolyze bradykinin. An alignment of the ACE2 peptide substrates reveals a consensus sequence of: Pro-X((1-3 residues))-Pro-Hydrophobic, where hydrolysis occurs between proline and the hydrophobic amino acid.  相似文献   

19.
The nitric oxide synthase inhibitor N(G)-nitro-L-arginine methyl ester (l-NAME) increased vascular resistance (VR) 10% above baseline of 3.08+/-0.08 (n=11) mmHg/mL/min at 10 mg/kg and 20% above 3.05+/-0.08 (n=9) at 50 mg/kg in anesthetized toads (Bufo marinus). Blood pressure was unaffected by either dose of L-NAME. Blood flow decreased at the higher dose of L-NAME. L-arginine (300 mg/kg) reversed the effects of L-NAME on VR and blood flow in toads treated with 10 mg/kg but not with 50 mg/kg. Injection of 50 mg/kg L-NAME into empty-bladder toads produced a 10% decrease in water uptake, J(v), resulting in a J(v) of 1,267+/-11 cm(3)/cm(2)/s x 10(-7) (n=9) compared to 1,385+/-12 (n=8) for controls. Injection of 10 microg/kg angiotensin II (ANG II) increased J(v) 15% across the pelvic patch (J(v), cm(3)/cm(2)/s x 10(-7)), resulting in a J(v) of 1,723+/-12 cm(3)/cm(2)/s x 10(-7) (n=8) compared to 1,471+/-12 (n=8) for controls. It is hypothesized that during cutaneous drinking blood flow into the capillary bed of the pelvic patch is regulated by nitric oxide and ANG II.  相似文献   

20.
Anchorage-dependent HeLa cells were successfully cultured on two fibrous materials (A07 and R100) with porosities of 75-125 and 40 mum, void fractions of 92% and 81%, and fiber diameters of 7.6 and 10.2 mum, respectively, in 100-mL spinner flasks and 2-L stirred tank bioreactors. The matrix was formed into a fixed vertical spiral configuration. All cultures displayed rapid (/=95%) to the matrix, uniform coverage of the immobilizing area with viable cells, and no significant amount of cell debris in the medium. Spinner flask cultures indicated that the denser material R100 showed better results in terms of final cell density. The growth of HeLa cells on material R100 in both culture systems was similar to that observed in tissue culture dishes (specific growth rate approximately 0.03-0.04 h(-1), maximum cell density of 8 x 10(6)-9 x 10(6) cells . mL(-1), and yields of 0.4 x 10(8) cells . mM(-1) on glucose and 2 x 10(8)-3 x 10(8) cells . mM(-1) on glutamine). Scale-up of this culture technique in a 2-L bioreactor under perfusion with pH and dissolved oxygen (DO) control yielded cell densities of up to 1.6 x 10(6) cells . mL(-1). Two other anchorage-dependent mammalian cells (ADC) known to be cultured with difficulty in roller bottles or with micro carriers were easily grown on material R100 in spinner flasks. The performance of this culture technique was compared to other ADC culture systems.  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