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.
Connexons and cell adhesion: a romantic phase   总被引:3,自引:1,他引:2  
Recent evidence indicates, that gap junction forming proteins do not only contribute to intercellular communication (Kanno and Saffitz in Cardiovasc Pathol 10:169-177, 2001; Saez et al. in Physiol Rev 83:1359-1400, 2003), ion homeostasis and volume control (Goldberg et al. in J Biol Chem 277:36725-36730, 2002; Saez et al. in Physiol Rev 83:1359-1400, 2003). They also serve biological functions in a mechanical sense, supporting adherent connections between neighbouring cells of epithelial and non-epithelial tissues (Clair et al. in Exp Cell Res 314:1250-1265, 2008; Shaw et al. in Cell 128:547-560, 2007), where they stabilize migratory pathways in the developing central nervous system (Elias et al. in Nature 448:901-907, 2007; Malatesta et al. in Development 127:5253-5263, 2000; Noctor et al. in Nature 409:714-720, 2001; Rakic in Brain Res 33:471-476, 1971; J Comp Neurol 145:61-83 1972; Science 241:170-176, 1988), or mediate polarized movements and directionality of neural crest cells during organogenesis (Kirby and Waldo in Circ Res 77:211-215, 1995; Xu et al. in Development 133:3629-3639, 2006). Since, most data describing adhesive properties of gap junctions delt with connexin 43 (Cx43) (Beardslee et al. in Circ Res 83:629-635, 1998), we will focus our brief review on this isoform.  相似文献   

2.
In this study, filtration flows through the walls of the rat aorta, pulmonary artery (PA), and inferior vena cava (IVC), vessels with very different susceptibilities to atherosclerosis, were measured as a function of transmural pressure (DeltaP), with intact and denuded endothelium on the same vessel. Aortic hydraulic conductivity (L(p)) is high at 60 mmHg, drops approximately 40% by 100 mmHg, and is pressure independent to 140 mmHg. The trends are similar in the PA and IVC, dropping 42% from 10 to 40 mmHg and flat to 100 mmHg (PA) and dropping 33% from 10 to 20 mmHg and essentially flat to 60 mmHg (IVC). Removal of the endothelium renders L(p)(DeltaP) flat: it increases L(p) of the aorta by approximately 75%, doubles L(p) of the PA, and quadruples L(p) of the IVC. Specific resistance (1/L(p)) of the aortic endothelium is approximately 47% of total resistance; i.e., the endothelium accounts for approximately 47% of the DeltaP drop at 100 mmHg. The PA value is 55% at >40 mmHg, and the IVC value is 23% at 10 mmHg. L(p) of the intact aorta, PA, and IVC are order 10(-8), 10(-7), and 5 x 10(-7) cm.s(-1).mmHg(-1), and wall thicknesses are 145.8 microm (SD 9.3), 78.9 microm (SD 3.3), and 66.1 microm (SD 4.1), respectively. These data are consistent with the different wall structures of the three vessels. The rat aortic L(p) data are quantitatively consistent with rabbit L(p)(DeltaP) (Tedgui A and Lever MJ. Am J Physiol Heart Circ Physiol 247: H784-H791, 1984; Baldwin AL and Wilson LM. Am J Physiol Heart Circ Physiol 264: H26-H32, 1993), suggesting that intimal compression under pressure loading may also play a role in L(p)(DeltaP) in these other vessels. Despite very different driving DeltaP, nominal transmural water fluxes of these three vessels are very similar and, therefore, cannot alone account for their differences in disease susceptibility. The different fates of macromolecular tracers convected by these water fluxes into the walls of these vessels may account for this difference.  相似文献   

3.
The solvent drag reflection coefficient (sigma) for total proteins can be estimated by comparing the relative degrees of concentration of erythrocytes and plasma proteins that occur during fluid filtration in an isolated perfused organ. In this analysis, we evaluated the accuracy of equations proposed by Pilati and Maron [Am. J. Physiol. 247 (Heart Circ. Physiol. 16): H1-H7, 1984] and Wolf et al. [Am. J. Physiol. 253 (Heart Circ. Physiol. 22): H194-H204, 1987] to calculate sigma from these concentration changes. We calculated sigma with each equation using data generated from a mathematical model of fluid and solute flux in membranes with known sigma's. We found that the equation of Wolf et al. provided the closest approximation to the true sigma over the entire range of filtration fractions tested (0.1-0.6), with the differences between the two equations increasing with filtration fraction. At low filtration fractions, the difference in sigma obtained using either approach was found to be inconsequential. At larger filtration fractions, a closer approximation of the true sigma can be obtained using the equation of Wolf et al.  相似文献   

4.
In microvessels, acute inflammation is typified by an increase in leukocyte-endothelial cell interactions, culminating in leukocyte transmigration into the tissue, and increased permeability to water and solutes, resulting in tissue edema. The goal of this study was to establish a method to quantify solute permeability (P(s)) changes in microvessels in intact predominantly blood-perfused networks in which leukocyte transmigratory behavior could be precisely described using established paradigms. We used intravital confocal microscopy to measure solute (BSA) flux across microvessel walls, hence P(s). A quantitative fluorescence approach (Huxley VH, Curry FE, and Adamson RH. Am J Physiol Heart Circ Physiol 252: H188-H197, 1987) was adapted to the imaged confocal tissue slice in which the fluorescent source volume and source surface area of the microvessel were restricted to the region of vessel that was contained within the imaged confocal tissue section. P(s) measurements were made in intact cremaster muscle microvasculature of anesthetized mice and compared with measurements of P(s) made in isolated rat skeletal muscle microvessels. Mouse arteriolar P(s) was 9.9 +/- 1.1 x 10(-7) cm/s (n = 16), which was not different from 8.4 +/- 1.3 x 10(-7) cm/s (n = 6) in rat arterioles. Values in venules were significantly (P < 0.05) higher: 44.4 +/- 7.9 x 10(-7) cm/s (n = 14) in mice and 25.0 +/- 3.7 x 10(-7) cm/s in rats. Convective coupling was estimated to contribute <10% to the measured P(s) in both microvessel types and both animal models. We conclude that this approach provides an appropriate quantification of P(s) in the intact microvasculature and that arteriolar P(s), while lower than in venules, is nevertheless consistent with arterioles being a significant source of interstitial protein.  相似文献   

5.
The present study aims to experimentally elucidate subtle structural features of the rat valve leaflet and the related nature of macromolecular transport across its endothelium and in its subendothelial space, information necessary to construct a rational theoretical model that can explain observation. After intravenous injection of horseradish peroxidase (HRP), we perfusion-fixed the aortic valve of normal Sprague-Dawley rats and found under light microscopy that HRP leaked through the leaflet's endothelium at very few localized brown spots, rather than uniformly. These spots grew nearly as rapidly with HRP circulation time before euthanasia as aortic spots, particularly when the time axis only included the time the valve was closed. These results suggest that macromolecular transport in heart valves depends not only on the direction normal to, but also parallel to, the endothelial surface and that convection, as well as molecular diffusion, plays an important role in macromolecular transport in heart valves. Transmission electron microscopy of traverse leaflet sections after 4-min HRP circulation showed a very thin ( approximately 150 nm), sparse layer immediately beneath the endothelium where the HRP concentration was much higher than that in the matrix below it. Nievelstein-Post et al.'s (Nievelstein-Post P, Mottino G, Fogelman A, Frank J. Arterioscler Thromb 14: 1151-1161, 1994) ultrarapid freezing/rotary shadow etching of the normal rabbit valve's subendothelial space supports the existence of this very thin, very sparse "valvular subendothelial intima," in analogy to the vascular subendothelial intima.  相似文献   

6.
The structure and motion of phospholipids in human plasma lipoproteins have been studied by using 31P NMR. Lateral diffusion coefficients, DT, obtained from the viscosity dependence of the 31P NMR line widths, were obtained for very low density lipoprotein (VLDL), low-density lipoprotein (LDL), high-density lipoproteins (HDL2, HDL3), and egg PC/TO microemulsions at 25 degrees C, for VLDL at 40 degrees C, and for LDL at 45 degrees C. At 25 degrees C, the rate of lateral diffusion in LDL (DT = 1.4 x 10(-9) cm2/s) is an order of magnitude slower than in the HDLs (DT = 2 x 10(-8) cm2/s). At 45 degrees C, DT for LDL increases to 1.1 x 10(-8) cm2/s. In contrast, DT for VLDL increases only slightly going from 25 to 40 degrees C. The large increase in diffusion rate observed in LDL occurs over the same temperature range as the smectic to disordered phase transition of the core cholesteryl esters, and provides evidence for direct interactions between the monolayer and core. In order to prove the orientation and/or order of the phospholipid head-group, estimates of the residual chemical shift anistropy, delta sigma, have been obtained for all the lipoproteins and the microemulsions from the viscosity and field dependence of the 31P NMR line widths. For VLDL and LDL, the anisotropy is 47-50 ppm at 25 degrees C, in agreement with data from phospholipid bilayers. For the HDLs, however, significantly larger values of 69-75 ppm (HDL2) and greater than 120 ppm (HDL3) were obtained.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
Anionic Arabidopsis thaliana peroxidase ATP A2 was expressed in Escherichia coli and used as a model for the 95% identical commercially available horseradish peroxidase HRP A2. The crystal structure of ATP A2 at 1.45 A resolution at 100 K showed a water molecule only 2.1 A from heme iron [Ostergaard, L., et al. (2000) Plant Mol. Biol. 44, 231-243], whereas spectroscopic studies of HRP A2 in solution at room temperature [Feis, A., et al. (1998) J. Raman Spectrosc. 29, 933-938] showed five-coordinated heme iron, which is common in peroxidases. Presented here, the X-ray crystallographic, single-crystal, and solution resonance Raman studies at room temperature confirmed that the sixth coordination position of heme iron of ATP A2 is essentially vacant. Furthermore, electronic absorption and resonance Raman spectroscopy showed that the heme environments of recombinant ATP A2 and glycosylated plant HRP A2 are indistinguishable at neutral and alkaline pH, from room temperature to 12 K, and are highly flexible compared with other plant peroxidases. Ostergaard et al. (2000) also demonstrated that ATP A2 expression and lignin formation coincide in Arabidopsis tissues, and docking of lignin precursors into the substrate binding site of ATP A2 predicted that coniferyl and p-coumaryl alcohols were good substrates. In contrast, the additional methoxy group of the sinapyl moiety gave rise to steric hindrance, not only in A2 type peroxidases but also in all peroxidases. We confirm these predictions for ATP A2, HRP A2, and HRP C. The specific activity of ATP A2 was lower than that of HRP A2 (pH 4-8), although a steady-state study at pH 5 demonstrated very little difference in their rate constants for reaction with H2O2 (k1 = 1.0 microM(-1) x s(-1). The oxidation of coniferyl alcohol, ferulic, p-coumaric, and sinapic acids by HRP A2, and ATP A2, however, gave modest but significantly different k3 rate constants of 8.7 +/- 0.3, 4.0 +/- 0.2, 0.70 +/- 0.03, and 0.04 +/- 0.2 microM(-1) x s(-1) for HRP A2, respectively, and 4.6 +/- 0.2, 2.3 +/- 0.1, 0.25 +/- 0.01, and 0.01 +/- 0.004 microM(-1) x s(-1) for ATP A2, respectively. The structural origin of the differential reactivity is discussed in relation to glycosylation and amino acid substitutions. The results are of general importance to the use of homologous models and structure determination at low temperatures.  相似文献   

8.
We have studied gas flow and particle deposition in a realistic three-dimensional (3D) model of the bronchial tree, extending from the trachea to the segmental bronchi (7th airway generation for the most distal ones) using computational fluid dynamics. The model is based on the morphometrical data of Horsfield et al. (Horsfield K, Dart G, Olson DE, Filley GF, and Cumming G. J Appl Physiol 31: 207-217, 1971) and on bronchoscopic and computerized tomography images, which give the spatial 3D orientation of the curved ducts. It incorporates realistic angles of successive branching planes. Steady inspiratory flow varying between 50 and 500 cm(3)/s was simulated, as well as deposition of spherical aerosol particles (1-7 microm diameter, 1 g/cm(3) density). Flow simulations indicated nonfully developed flows in the branches due to their relative short lengths. Velocity flow profiles in the segmental bronchi, taken one diameter downstream of the bifurcation, were distorted compared with the flow in a simple curved tube, and wide patterns of secondary flow fields were observed. Both were due to the asymmetrical 3D configuration of the bifurcating network. Viscous pressure drop in the model was compared with results obtained by Pedley et al. (Pedley TJ, Schroter RC, and Sudlow MF. Respir Physiol 9: 387-405, 1970), which are shown to be a good first approximation. Particle deposition increased with particle size and was minimal for approximately 200 cm(3)/s inspiratory flow, but it was highly heterogeneous for branches of the same generation.  相似文献   

9.
A simple capillary zone electrophoresis (CZE) method was used to determine native, in vitro Cu(2+) and glucose modified low-density lipoprotein (LDL) particles for four healthy subjects. The LDL electropherograms are highly reproducible with good precisions of effective mobility and peak area. The native LDL capillary electrophoresis (CE) profile shows a major peak with lower mobility and two minor peaks with higher mobilities. For three-hour Cu(2+) oxidation, one major peak with mobility close to that of the native major peak, and one minor peak with mobility extending to -47 x 10(-5)cm(2)V(-1)s(-1) appear. For eighteen-hour Cu(2+) oxidation, one major peak with mobility much higher than that of the native major peak appears. As the reaction time for LDL and Cu(2+) increases from 0 to 24h, effective mobility of the LDL major peak increases, suggesting that LDL particles become more negatively charged and oxidized as the time increases. The in vitro glycated LDL particles are characterized by a major peak and two minor peaks. Mobility of the major peak is close to that of native major peak, but the second minor peak is much more negatively charged with mobility extending to -53 x 10(-5)cm(2)V(-1)s(-1). Native, oxidized and glycated LDL particles show distinctive differences in their CZE profiles. Agarose electrophoresis shows that the charge to mass ratios of native, three-hour Cu(2+) and glucose modified LDL particles are similar, but that of eighteen-hour Cu(2+) oxidized LDL particles is higher.  相似文献   

10.
The recent experiments in Hu et al. (Am J Physiol Heart Circ Physiol 279: H1724-H1736, 2000) and Adamson et al. (J Physiol 557: 889-907, 2004) in frog and rat mesentery microvessels have provided strong evidence supporting the Michel-Weinbaum hypothesis for a revised asymmetric Starling principle in which the Starling force balance is applied locally across the endothelial glycocalyx layer rather than between lumen and tissue. These experiments were interpreted by a three-dimensional (3-D) mathematical model (Hu et al.; Microvasc Res 58: 281-304, 1999) to describe the coupled water and albumin fluxes in the glycocalyx layer, the cleft with its tight junction strand, and the surrounding tissue. This numerical 3-D model converges if the tissue is at uniform concentration or has significant tissue gradients due to tissue loading. However, for most physiological conditions, tissue gradients are two to three orders of magnitude smaller than the albumin gradients in the cleft, and the numerical model does not converge. A simpler multilayer one-dimensional (1-D) analytical model has been developed to describe these conditions. This model is used to extend Michel and Phillips's original 1-D analysis of the matrix layer (J Physiol 388: 421-435, 1987) to include a cleft with a tight junction strand, to explain the observation of Levick (Exp Physiol 76: 825-857, 1991) that most tissues have an equilibrium tissue concentration that is close to 0.4 lumen concentration, and to explore the role of vesicular transport in achieving this equilibrium. The model predicts the surprising finding that one can have steady-state reabsorption at low pressures, in contrast to the experiments in Michel and Phillips, if a backward-standing gradient is established in the cleft that prevents the concentration from rising behind the glycocalyx.  相似文献   

11.
A key early event in the development of atherosclerosis is the oxidation of low density lipoprotein (LDL) via different mechanisms including free radical reactions with both protein and lipid components. Nitric oxide (( small middle dot)NO) is capable of inhibiting LDL oxidation by scavenging radical species involved in oxidative chain propagation reactions. Herein, the diffusion of ( small middle dot)NO into LDL is studied by fluorescence quenching of pyrene derivatives. Selected probes 1-(pyrenyl)methyltrimethylammonium (PMTMA) and 1-(pyrenyl)-methyl-3-(9-octadecenoyloxy)-22,23-bisnor-5-cholenate (PMChO) were chosen so that they could be incorporated at different depths of the LDL particle. Indeed, PMTMA and PMChO were located in the surface and core of LDL, respectively, as indicated by changes in fluorescence spectra, fluorescence quenching studies with water-soluble quenchers and the lifetime values (tau(o)) of the excited probes. The apparent second order rate quenching constants of ( small middle dot)NO (k(NO)) for both probes were 2.6-3.8 x 10(10) m(-1) s(-1) and 1.2 x 10(10) m(-1) s(-1) in solution and native LDL, respectively, indicating that there is no significant barrier to the diffusion of ( small middle dot)NO to the surface and core of LDL. Nitric oxide was also capable of diffusing through oxidized LDL. Considering the preferential partitioning of ( small middle dot)NO in apolar milieu (6-8 for n-octanol:water) and therefore a larger ( small middle dot)NO concentration in LDL with respect to the aqueous phase, a corrected k(NO) value of approximately 0.2 x 10(10) m(-1) s(-1) can be determined, which still is sufficiently large and consistent with a facile diffusion of ( small middle dot)NO through LDL. Applying the Einstein-Smoluchowsky treatment, the apparent diffusion coefficient (D(')NO) of ( small middle dot)NO in native LDL is on average 2 x 10(-5) cm(2) s(-1), six times larger than that previously reported for erythrocyte plasma membrane. Thus, our observations support that ( small middle dot)NO readily traverses the LDL surface accessing the hydrophobic lipid core of the particle and affirm a role for ( small middle dot)NO as a major lipophilic antioxidant in LDL.  相似文献   

12.
It was recently shown that the structure of the fluorophore attached to the acyl chain of phosphatidylcholine analogs determines their mechanism of transport across the plasma membrane of yeast cells (Elvington et al., J. Biol Chem. 280:40957, 2005). In order to gain further insight into the physical properties of these fluorescent phosphatidylcholine (PC) analogs, the rate and mechanism of their intervesicular transport was determined. The rate of spontaneous exchange was measured for PC analogs containing either NBD (7-nitrobenz-2-oxa-1,3-diazol-4-yl), Bodipy FL (4,4-difluoro-5,7-dimethyl-4-bora-3a,4a-diaza-s-indacene), Bodipy 530 (4,4-difluoro-5,7-diphenyl-4-bora-3a,4a-diaza-s-indacene), or Bodipy 581 (4,4-difluoro-5-(4-phenyl-1,3-butadienyl)-4-bora-3a,4a-diaza-s-indacene) attached to a five or six carbon acyl chain in the sn-2 position. The rate of transfer between phospholipid vesicles was measured by monitoring the increase in fluorescence as the analogs transferred from donor vesicles containing self-quenching concentrations to unlabeled acceptor vesicles. Kinetic analysis indicated that the transfer of each analog occurred by diffusion through the water phase as opposed to transfer during vesicle collisions. The vesicle-to-monomer dissociation rate constants differed by over four orders of magnitude: NBD-PC (k(dis)=0.115 s(-1); t(1/2)=6.03 s); Bodipy FL-PC (k(dis)=5.2x10(-4); t(1/2)=22.2 min); Bodipy 530-PC (k(dis)=1.52x10(-5); t(1/2)=12.6 h); and Bodipy 581-PC (k(dis)=5.9x10(-6); t(1/2)=32.6 h). The large differences in spontaneous rates of transfer through the water measured for these four fluorescent PC analogs reflect their hydrophobicity and may account for their recognition by different mechanisms of transport across the plasma membrane of yeast.  相似文献   

13.
We previously demonstrated that physiologic levels of shear stress enhance endothelial repair. Cell spreading and migration, but not proliferation, were the major mechanisms accounting for the increases in wound closure rate (Albuquerque et al., 2000, Am. J. Physiol. Heart Circ. Physiol. 279, H293-H302). However, the patterns and movements of beta-actin filaments responsible for cell motility and translocation in human coronary artery endothelial cells (HCAECs) have not been previously investigated under physiologic flow. HCAECs transfected with beta-actin-GFP were cultured on type I collagen-coated coverslips. Confluent cell monolayers were subjected to laminar shear stress of 12 dynes/cm(2) for 18 h in a parallel-plate flow chamber to attain cellular alignment and then wounded by scraping with a metal spatula and subsequently exposed to a laminar shear stress of 20 dynes/cm(2) (S-W-sH) or static (S-W-sT) conditions. Time-lapse imaging and deconvolution microscopy was performed during the first 3 h after imposition of S-W-sH or S-W-sT conditions. The spatial and temporal dynamics of beta-actin-GFP motility and translocation during wound closure in HCAEC monolayers were analyzed under both conditions. Compared with HCAEC under S-W-sT conditions, our data show that HCAEC under S-W-sH conditions demonstrated greater beta-actin-GFP motility, filament and clumping patterns, and filament arcs used during cellular attachment and detachment. These findings demonstrate intriguing patterns of beta-actin organization and movement during wound closure in HCAEC exposed to physiological flow.  相似文献   

14.
We examined the factors that influence the permeability characteristics of isolated perfused rat lungs and compared the ex vivo permeability-surface area product (PS) with that obtained in vivo. In lungs perfused for 20 min with homologous blood or a physiological salt solution (PSS) containing 4 g/100 ml albumin, mean PS values, obtained by the single-sample method of Kern et al. [Am. J. Physiol. 245 (Heart Circ. Physiol. 14): H229-H236, 1983], were 9.9 +/- 0.6 (SE) and 6.8 +/- 0.3 cm3.min-1.g wet lung-1.10(-2), respectively. These values were similar to lung PS obtained in intact rats (7.7 +/- 0.4 cm3.min-1.g wet lung-1.10(-2). In perfused lungs, PS values were influenced by the perfusate albumin concentration, the length of perfusion time, and the degree of vascular recruitment. Twenty minutes after lung isolation, PS was 126% higher in lungs perfused with albumin-free PSS containing Ficoll than in lungs perfused with albumin-PSS. Moreover, PS in Ficoll-PSS-perfused lungs increased even higher after 2 h of perfusion, and this time-dependent increase in PS was attenuated by addition of 0.1 g/100 ml albumin to the perfusate. Two hours of ex vivo ventilation with hypoxic (0 or 3% 0(2)) or hyperoxic (95% 0(2)) gas mixture did not affect PS values in perfused lungs. However, PS was elevated in lungs perfused ex vivo with protamine, which causes endothelial cell injury, or in lungs from rats exposed in vivo to human recombinant tumor necrosis factor.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
A novel inexpensive and simple amperometric biosensor, based on the immobilization of HRP into redox active [Zn-Cr-ABTS] layered double hydroxide, is applied to the determination of cyanide. The electrochemical transduction step corresponds to the reduction at 0.0 V of ABTS+* enzymatically formed in the presence of H2O2. The biosensor has a fast response to H2O2 (8s) with a linear range of 1.7 x 10(-9) to 2.1 x 10(-6) M and a sensitivity of 875 mA M(-1) cm(-2). The apparent Michaelis-Menten constant (KMapp) is 12 microM. The detection of cyanide is performed via its non competitive inhibiting action on the HRP/[Zn-Cr-ABTS] electrode. The concentration range of the linear response and the apparent inhibition constant (ki) are 5 x 10(-9) to 4 x 10(-8) and 1.4 x 10 (-7) M, respectively.  相似文献   

16.
Palmitic acid (Pal) is known to promote apoptosis (Sparagna G et al (2000) Am J Physiol Heart Circ Physiol 279: H2124–H2132) and its amount in blood and mitochondria increases under some pathological conditions. Yet, the mechanism of the proapoptotic action of Pal has not been elucidated. We present evidence for the involvement of the mitochondrial cyclosporin A-insensitive pore induced by Pal/Ca2+ complexes in the apoptotic process. Opening of this pore led to a fall of the mitochondrial membrane potential and the release of the proapoptotic signal cytochrome c. The addition of cytochrome c prevented these effects and recovered membrane potential, which is in contrast to the cyclosporin A-sensitive mitochondrial permeability transition pore. Oleic and linoleic acids prevented the Pal/Ca2+-induced pore opening in the intact mitochondria, this directly and significantly correlating with the effect of these fatty acids on Pal-induced apoptosis in cells (Hardy S et al (2003) J Biol Chem 278: 31861–31870). The specific probe for cardiolipin, 10-N-nonyl acridine orange, inhibited formation of this pore.  相似文献   

17.
Hypoxic pulmonary vasoconstriction (HPV) is encountered during ascent to high altitude. Atrial natriuretic peptide (ANP) could be an option to treat HPV because of its natriuretic, diuretic, and vasodilatory properties. Data on effects of ANP on pulmonary and systemic circulation during HVP are conflicting, partly owing to anesthesia, surgical stress or uncontrolled dietary conditions. Therefore, ten conscious, chronically tracheotomized dogs were studied under standardized dietary conditions. The dogs were trained to breathe spontaneously at a ventilator circuit. Protocol: 30min of normoxia [inspiratory oxygen fraction (F(i)O(2))=0.21] were followed by 30min of hypoxia without ANP infusion (Hypoxia I, F(i)O(2)=0.1). While maintaining hypoxia an intravenous infusion of atrial natriuretic peptide was started with 50ng x kg body wt(-1) x min(-1) for 30min (Hypoxia+ANP1=low dose), followed by 1000ng x kg body wt(-1) x min(-1) for 30min (Hypoxia+ANP2=high dose). Thereafter, ANP infusion was stopped and hypoxia maintained for a final 30min (Hypoxia II). Compared to normoxia, mean pulmonary arterial pressure (MPAP) (16+/-0.7 vs. 26+/-1.3mmHg) and pulmonary vascular resistance (PVR) (448+/-28 vs. 764+/-89dyn x s(-1) x cm(-5)) increased during Hypoxia I and decreased during Hypoxia+ANP 1 (MPAP 20+/-1mmHg, PVR 542+/-55dyn x s(-1) x cm(-5)) (P<0.05). The higher dose of ANP did not further decrease MPAP or PVR, but started to have a tendency to decrease mean arterial pressure and cardiac output. We conclude that low dose ANP is able to reduce HPV without affecting systemic circulation during acute hypoxia.  相似文献   

18.
In Part 1 of this study (Weinbaum et al., 1988) a short time model has been proposed to describe the initial time dependent leakage of macromolecules at short distances (5 microns or less) from the exit of a transient open junction which the authors have hypothesized as a characteristic feature of endothelial cells in the process of turnover (Weinbaum et al., 1985). This open junction pathway has also been proposed (Weinbaum et al., 1988) to be the primary ultrastructural correlate of the 20 nm diameter large pore suggested by Renkin et al. (1977) using the predictions of cylindrical pore theory. The short time model in (Weinbaum et al., 1988), however, has major limitations in that it neglects the interaction between leakage sites, macromolecular entry through other pathways, the finite thickness of the vessel wall and the curvature of the cell perimeter. The longer time model developed herein will attempt to describe each of these features and also present an improved model and analytic solution for the steady state flux and uptake. In the previous steady state model developed by Weinbaum et al. (1985) the effect of the resistance of the transient open junctions and the non-isotropic diffusion in the underlying tissue due to the internal elastic lamina (IEL) were both neglected. New solutions are first presented which describe the effect of these important model refinements on the steady state macromolecular permeability of the major arteries. Time dependent solutions are then presented to predict the transient longer time labeling following the introduction of tracer macromolecules of varying size. These solutions and the corresponding short time solutions in Weinbaum et al. (1988) are the first solutions to our knowledge to describe the difficult time-dependent boundary value problem to determine how the channel exit concentration and flux at a leaky junction vary with time. This is accomplished by casting the boundary value problem in the form of an integral equation for the unknown flux at the cleft exit and then solving this problem using a specially designed numerical technique. The theoretical predictions are used to interpret the behavior of the localized leaks to HRP and albumin that have been reported in Stemerman et al. (1986) and our own recent experiments (Lin et al., 1988).  相似文献   

19.
Nitric oxide (*NO) has been proposed to play a relevant role in modulating oxidative reactions in lipophilic media like biomembranes and lipoproteins. Two factors that will regulate *NO reactivity in the lipid milieu are its diffusion and solubility, but there is no data concerning the actual diffusion (D) and partition coefficients (KP) of *NO in biologically relevant hydrophobic phases. Herein, a "equilibrium-shift" method was designed to directly determine the *NO and O2 partition coefficients in liposomes and low density lipoprotein (LDL) relative to water. It was found that *NO partitions 4.4- and 3.4-fold in liposomes and LDL, respectively, whereas O2 behaves similarly with values of 3.9 and 2.9, respectively. In addition, actual diffusion coefficients in these hydrophobic phases were determined using fluorescence quenching and found that *NO diffuses approximately 2 times slower than O2 in the core of LDL and 12 times slower than in buffer (DNOLDL=3.9 x 10(-6) cm2 s(-1),DO2LDL=7.0 x 10(-6) cm2 s(-1),DNObuffer=DO2buffer=4.5 x 10(-5) cm2 s(-1)). The influence of *NO and O2 partitioning and diffusion in membranes and lipoproteins on *NO reaction with lipid radicals and auto-oxidation is discussed. Particularly, the 3-4-fold increase in O2 and *NO concentration within biological hydrophobic phases provides quantitative support for the idea of an accelerated auto-oxidation of *NO in lipid-containing structures, turning them into sites of enhanced local production of oxidant and nitrosating species.  相似文献   

20.
Histamine dehydrogenase (NSHADH) can be isolated from cultures of Nocardioides simplex grown with histamine as the sole nitrogen source. A previous report suggested that NSHADH might contain the quinone cofactor tryptophan tryptophyl quinone (TTQ). Here, the hdh gene encoding NSHADH is cloned from the genomic DNA of N. simplex, and the isolated enzyme is subjected to a full spectroscopic characterization. Protein sequence alignment shows NSHADH to be related to trimethylamine dehydrogenase (TMADH: EC 1.5.99.7), where the latter contains a bacterial ferredoxin-type [4Fe-4S] cluster and 6-S-cysteinyl FMN cofactor. NSHADH has no sequence similarity to any TTQ containing amine dehydrogenases. NSHADH contains 3.6+/-0.3 mol Fe and 3.7+/-0.2 mol acid labile S per subunit. A comparison of the UV/vis spectra of NSHADH and TMADH shows significant similarity. The EPR spectrum of histamine reduced NSHADH also supports the presence of the flavin and [4Fe-4S] cofactors. Importantly, we show that NSHADH has a narrow substrate specificity, oxidizing only histamine (K(m)=31+/-11 microM, k(cat)/K(m)=2.1 (+/-0.4)x10(5)M(-1)s(-1)), agmatine (K(m)=37+/-6 microM, k(cat)/K(m)=6.0 (+/-0.6)x10(4)M(-1)s(-1)), and putrescine (K(m)=1280+/-240 microM, k(cat)/K(m)=1500+/-200 M(-1)s(-1)). A kinetic characterization of the oxidative deamination of histamine by NSHADH is presented that includes the pH dependence of k(cat)/K(m) (histamine) and the measurement of a substrate deuterium isotope effect, (D)(k(cat)/K(m) (histamine))=7.0+/-1.8 at pH 8.5. k(cat) is also pH dependent and has a reduced substrate deuterium isotope of (D)(k(cat))=1.3+/-0.2.  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