首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The pulmonary artery (PA) wall, which has much higher hydraulic conductivity and albumin void space and approximately one-sixth the normal transmural pressure of systemic arteries (e.g, aorta, carotid arteries), is rarely atherosclerotic, except under pulmonary hypertension. This study constructs a detailed, two-dimensional, wall-structure-based filtration and macromolecular transport model for the PA to investigate differences in prelesion transport processes between the disease-susceptible aorta and the relatively resistant PA. The PA and aorta models are similar in wall structure, but very different in parameter values, many of which have been measured (and therefore modified) since the original aorta model of Huang et al. (23). Both PA and aortic model simulations fit experimental data on transwall LDL concentration profiles and on the growth of isolated endothelial (horseradish peroxidase) tracer spots with circulation time very well. They reveal that lipid entering the aorta attains a much higher intima than media concentration but distributes better between these regions in the PA than aorta and that tracer in both regions contributes to observed tracer spots. Solutions show why both the overall transmural water flow and spot growth rates are similar in these vessels despite very different material transport parameters. Since early lipid accumulation occurs in the subendothelial intima and since (matrix binding) reaction kinetics depend on reactant concentrations, the lower intima lipid concentrations in the PA vs. aorta likely lead to slower accumulation of bound lipid in the PA. These findings may be relevant to understanding the different atherosusceptibilities of these vessels.  相似文献   

2.
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   

3.
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   

4.
The present study was designed to evaluate the expression of dopamine D1 and D2 receptor mRNAs in systemic and pulmonary vasculatures. Using specific antisense riboprobes for dopamine D1 and D2 receptor cDNAs, in situ hybridization histochemistry was performed in the aorta, common carotid artery, vertebral artery, pulmonary artery, and superior vena cava of the adult male Sprague Dawley rat. In the case of the aorta, common carotid artery, and vertebral artery, dopamine D1 receptor mRNAs localized mainly in the smooth muscle cells of the tunica media. However, the signals of dopamine D2 receptor mRNAs were found in the endothelium and subendothelial layer of tunica intima, and interstitial cells of tunica adventitia. In the case of the pulmonary artery, signals of dopamine D1 receptor mRNAs were detected within the tunica intima, media, and adventitia. Expression of D2 receptor mRNAs was detected in the walls of small blood vessels within the tunica adventitia of the pulmonary artery. There were no detectable signals of dopamine D1 and D2 receptor mRNAs in the vein. The uneven distribution of dopamine D1 and D2 receptor mRNAs in the rat systemic vasculatures and pulmonary artery suggests that dopamine differentially regulates the vasodilation of the systemic and pulmonary arteries through the differential stimulation of dopamine D1 and D2 receptor.  相似文献   

5.
The spatial distribution of sites of enhanced permeability to the macromolecule horseradish peroxidase (HRP) in the normal rabbit aorta after one min circulation was studied using image analysis. These sites, referred to as "HRP spots," exhibit a nonuniform distribution that is qualitatively similar in all rabbits studied. The density of HRP spots is highest in the aortic arch, decreases distally, reaches a minimum in the lower descending thoracic aorta, and then increases again in the abdominal aorta. The region of highest spot density follows a clockwise helical pattern in the aortic arch and outside the arch occurs in streaks largely oriented in the bulk flow direction. The streaks in the abdominal aorta localize along the anatomical right lateral wall and occasionally along the left lateral wall proximal to the celiac artery and along the ventral wall between the celiac and superior mesenteric arteries. The density of spots is high in the immediate vicinity of aortic ostia with the most elevated density being distal to ostia in most cases. At a short distance from the ostium edge of the celiac and superior mesenteric branches the proximal density is comparably high, and no preferred spot orientation is observed around the brachiocephalic vessel. These results are consistent with an influence of localizing factors such as detailed hemodynamic phenomena and/or arterial wall structural and/or functional variations.  相似文献   

6.
Male Sprague-Dawley rats (n = 18) weighing 548 +/- 30 g were anesthetized with pentobarbital sodium (40-65 mg/kg body wt ip), intubated via tracheotomy, and mechanically ventilated. After exposure of the great vessels in the thorax, blood was withdrawn from the pulmonary artery (PA), right ventricle (RV), right atrium (RA), inferior vena cava (IVC), and ascending aorta. The O2 content of these blood samples was determined by direct measurements and/or was calculated from the measured hemoglobin concentration, percent of O2 saturation, and PO2. Ventilatory rates and the inspired fraction of O2 were manipulated to vary the mixed venous O2 content (CvO2) of blood withdrawn from the PA from 1.4 to 12.9 ml O2/dl blood (vol%). Our results demonstrate that O2 contents of blood withdrawn from the PA, RV, and RA are not significantly different from one another (CPAO2 - CrvO2 = -0.02 +/- 0.25 and CPAO2 - CRAO2 = -0.07 +/- 0.41 vol%, n = 28, P greater than 0.05); however, the O2 content of blood withdrawn from the IVC is significantly lower than that withdrawn from the PA (CPAO2 - CIVCO2 = 2.11 +/- 0.34 vol%, P less than 0.001). In addition, the directly measured O2 contents were equivalent to those that were calculated. These results suggest that the O2 content of blood found in the RA and RV of the rat are indicative of the O2 content of blood found in the PA. Thus blood sampled from these areas can be used to estimate mixed venous oxygenation.  相似文献   

7.
In this paper a new theoretical framework is presented for analyzing the filtration and macromolecular convective-diffusive transport processes in the intimal region of an artery wall with widely dispersed macromolecular cellular leakage sites, as proposed in the leaky junction-cell turnover hypothesis of Weinbaum et al. In contrast to existing convection-diffusive models, which assume that the transport is either 1-D, or convection is primarily in a direction normal to the endothelial surface, the present model considers for the first time the nonuniform subendothelial pressure field that arises from the different hydraulic resistances of normal and leaky endothelial clefts and the special role of the internal elastic lamina (IEL) in modulating the horizontal transport of macromolecules after they have passed through the leaky clefts of cells that are either in mitosis or demonstrate IgG labeling. The new theory is able to quantitatively explain the growing body of recent experiments in which an unexpectedly rapid early-time growth of the leakage spot has been observed and the longer time asymptotic behavior in which the leakage spot appears to approach an equilibrium diameter. The new theory also predicts the observed doubling in macromolecular permeability between EBA labeled blue and white areas when the frequency of leakage sites is doubled. This frequency for doubling of permeability, however, is an order of magnitude smaller than predicted by the author's previous model, Tzeghai et al., in which only convection normal to the endothelial surface was considered and the pressure was uniform in the intima. The longer time model predictions are used to explain the time scale for the formation of liposomes in subendothelial tissue matrix in animal feeding experiments where it has been observed that the extracellular lipid concentration rises sharply prior to the entry of monocytes into the intima.  相似文献   

8.
Enzymatic activity of cells, antigenic cellular markers and extracellular matrix of the hyperplastic intima of the aorta and carotid arteries was investigated in non-specific aorto-arteritis by immunomorphological and histochemical techniques. The cells of subendothelial layer of thickened arterial intima contained smooth muscle cell myosin, gave positive reactions to myosin ATP-ase and revealed high activity of thiamine pyrophosphatase. Fibronectin and type IV and V collagen were located in close proximity to these cells. The data obtained make it possible to consider these cells as modified smooth muscle cells. Type III collagen was the prevalent type of extracellular matrix of the thickened intima. A great number of blood vessels of the capillary and precapillary types have been found to penetrate into the intima from the adventitia. A possible role of pericytes surrounding newly formed capillaries as the precursors of subendothelial cell population in the hyperplastic intima is discussed.  相似文献   

9.
The aim of this study was to investigate the protective effects of vanadyl sulfate on aorta tissue of normal and streptozotocin (STZ)-induced diabetic rats, morphologically and biochemically. The animals were made diabetic by an intraperitoneal injection of streptozotocin (65 mg/kg) and vanadyl sulfate (100 mg/kg) that was given every day for 60 days by gavage technique to rats. Under the light and transmission electron microscopes, hypertrophy of the vessel wall, focal disruption in the elastic lamellae, an increase in thickness of total aortic wall, tunica intima, subendothelial space and adventitial layer, and a disorganization in smooth muscular cells of the tunica media were observed in diabetic animals. The aorta lipid peroxidation (LPO) levels were significantly increased and the aorta glutathione (GSH) levels were significantly reduced in STZ diabetic rats. In diabetic rats administered vanadyl sulfate for 60 days, aorta LPO levels significantly decreased and the aorta GSH level significantly increased. In conclusion, in vivo treatment with vanadyl sulfate of diabetic rats prevented the morphological and biochemical changes observed in thoracic aorta of diabetic animals.  相似文献   

10.
Summary Aortic endothelium presents a continuous barrier to diffusion of macromolecules. The cell margins overlap for long distances and there are multiple points of contact between the cell membranes at which the intercellular cleft is reduced to 30–40 Å or less, and free diffusion of lanthanum is impeded at some points of apposition. Macromolecular transport through the endothelium of mouse aorta was studied with the help of horseradish peroxidase (HRP) and bovine milk lactoperoxidase. Following injection of 0.25–0.5 mg of HRP no tracer was detected in the intercellular clefts even though it was seen in plasmalemmal vesicles and subendothelial space. However, when 5 mg of HRP was injected in either 0.05 or 0.5 ml of saline, transport of the enzyme occurred through both the intercellular clefts and via the plasmalemmal vesicles. On the other hand, lactoperoxidase of m.w. 82000 was transported through the plasmalemmal vesicles only. The findings were discussed with reference to the transport of serum lipoproteins and it was suggested that low and high density lipoproteins would be transported via the plasmalemmal vesicles.The excellent technical help of Miss R. Ben-Moshe and Mrs. A. Mandeles is gratefully acknowledged. This study was supported in part by a grant from the Myra Kurland Heart Fund, Chicago, Ill., and by a grant 06-101-1 of the National Institute of Health, United States Public Health Service.  相似文献   

11.
Microheterogeneity of different vinculin and meta-vinculin isoforms in adult human tissues and cultured cells was studied by two-dimensional gel electrophoresis and immunoblotting technique. Four isoforms of vinculin (alpha, alpha', beta, and gamma) and two isoforms of meta-vinculin (alpha and beta) were resolved. alpha-, alpha'-, and beta-isoforms of vinculin were found in all cell types and tissue samples analyzed in the present study. gamma-Isoform of vinculin and both alpha- and beta-isoforms of meta-vinculin were found in smooth (aorta wall and myometrium) and cardiac muscle, rather than in skeletal muscle, liver, foreskin fibroblasts, and macrophages. In the primary culture of human aorta smooth muscle cells, the fractional content of gamma-isoform of vinculin and meta-vinculin was dramatically reduced, and, by the onset of intensive cell division, the proteins could hardly be detected. Subcultured human aorta smooth muscle cells did not contain gamma-vinculin and meta-vinculin. We analyzed the microheterogeneity of vinculin and meta-vinculin in three smooth muscle layers of human aorta wall--media, muscular-elastic (adjacent to media) intima, and subendothelial (juxtaluminal) intima. It was shown that in media the fractional content of gamma-isoform of vinculin was 45% and meta-vinculin, 42%; in muscular-elastic intima the fractional content of gamma-vinculin was 42% and meta-vinculin, 36%. However, in subendothelial intima, the share of these proteins was significantly lower than in adjacent muscular-elastic intima and media. Isoactin pattern that is characteristic of smooth muscle was identical in all aortic layers, thus proving the smooth muscle origin of subendothelial intima cells. These findings demonstrate that human aortic smooth muscle cells in vivo and in vitro undergo coordinated differential expression of smooth muscle specific variants of vinculin, i.e. gamma-vinculin and meta-vinculin.  相似文献   

12.
The thoracic aorta and basilar artery, in which the incidence of atherosclerosis is known to be different, were examined to elucidate the correlation between the structure of the intercellular cleft junction between adjacent endothelial cells and its permeability to HRP. Tannic acid or HRP in the vessel lumen passed through the intercellular clefts of the thoracic aorta into the subendothelial space, whereas in the basilar artery they were unable to penetrate beyond the tight junction of the intercellular clefts. Freeze-fracture replicas revealed that the tight junctions of the thoracic aorta consisted of one to two junctional strands in most areas of the cleaved planes, with discontinuities in some places, whereas those of the basilar artery consisted of a continuous belt-like meshwork of six anastomosing junctional strands on average. These observations confirm that the structure of endothelial junctions in arteries has a close correlation with the permeability of the intercellular clefts to HRP.  相似文献   

13.
Pharmacomechanical coupling of vascular smooth muscle is believed to be mediated by inositol trisphosphate (IP3). Numerous studies have demonstrated an increase in inositol phosphates following tissue stimulation using either intact aortic strips or cultured cells from aorta. However, little information is available concerning inositol phosphates in vascular tissue other than in the large conduit vessel, the aorta. This present study was designed to examine the role of inositol phosphate metabolism following adrenergic stimulation of the muscular rat tail artery as compared to the aorta. Segments of thoracic aorta and tail artery from male Sprague Dawley rats were labeled with [3H]inositol and stimulated with norepinephrine. The norepinephrine concentration that resulted in a half-maximal stimulation of inositol phosphates was approximately 10(-6) M in both the aorta and tail artery. Although the sensitivity of the two vessels to norepinephrine stimulation were similar, the stimulated levels of IP, IP2, and IP3 were from 1 to 2 orders of magnitude greater in the tail artery than in aorta. IP production in aorta and tail artery was a linear function of time (from 0 to 30 min). Significant levels of IP3 (the 1,4,5-IP3 isomer as determined by HPLC) could only be detected in the tail artery and appeared to be produced optimally after 5 min of stimulation. The several order of magnitude increase in adrenergic stimulated inositol phosphate production in the tail artery was not due to either an increased magnitude of [3H]inositol incorporated into PI, PIP, and PIP2 or to a greater percentage of smooth muscle cells per unit tissue of the rat tail artery. We believe the results of this study demonstrate that the increased inositol phosphate metabolism in the vascular smooth muscle cells of the tail artery is an intrinsic property of the cell. Moreover, due to the significant levels of all inositol phosphates produced in the tail artery, this muscular artery may be a better model, as compared to the aorta, for future studies investigating pharmacomechanical coupling of vascular smooth muscle.  相似文献   

14.
L-Arginine is a common substrate for the enzymes arginase and nitric oxide synthase (NOS). Acute inhibition of arginase enzyme activity improves endothelium-dependent vasorelaxation, presumably by increasing availability of substrate for NOS. Arginase is activated by manganese (Mn), and the consumption of a Mn-deficient (Mn-) diet can result in low arginase activity. We hypothesize that endothelium-dependent vasorelaxation is greater in rats fed Mn- versus Mn sufficient (Mn+) diets. Newly weaned rats fed Mn+ diets (0.5 microg Mn/g; n = 12) versus Mn+ diets (45 microg Mn/g; n = 12) for 44 +/- 3 days had (i) lower liver and kidney Mn and arginase activity (P < or = 0.05), (ii) higher plasma L-arginine (P < or = 0.05), (iii) similar plasma and urine nitrate + nitrite, and (iv) similar staining for endothelial nitric oxide synthase in thoracic aorta. Vascular reactivity of thoracic aorta (approximately 720 microm i.d.) and small coronary arteries (approximately 110 microm i.d.) was evaluated using wire myographs. Acetylcholine (ACh; 10(-8)-10(-4) M) produced greater (P < or = 0.05) vasorelaxation in thoracic aorta from Mn- rats (e.g., maximal percent relaxation, 79 +/- 7%) versus Mn + rats (e.g., maximal percent relaxation, 54 +/- 9%) at 5 of 7 evaluated doses. Tension produced by NOS inhibition using N(G) monomethyl-L-arginine (L-NMMA; 10(-3) M) and vasorelaxation evoked by (i) arginase inhibition using difluoromethylornithine (DFMO; 10(-7) M), (ii) ACh (10(-8)-10(-4) M) in the presence of DFMO, and (iii) sodium nitroprusside (10(-9)-10(-4) M) were unaffected by diet. No differences existed between groups concerning these responses in small coronary arteries. These findings support our hypothesis that endothelium-dependent vasorelaxation is greater in aortic segments from rats that consume Mn- versus Mn+ diets; however, responses from small coronary arteries were unaffected.  相似文献   

15.
Enteric neurons controlling various gut functions are prone to oxidative insults that might damage mitochondria (e.g., intestinal inflammation). To resume local energy supply, mitochondria need to be transported. We used MitoTracker dyes and confocal microscopy to investigate basic characteristics of mitochondrial transport in guinea pig myenteric neurites. During a 10-s observation of 1 mm nerve fiber, on average, three mitochondria were transported at an average speed of 0.41 +/- 0.02 microm/s. Movement patterns were clearly erratic, and velocities were independent of mitochondrial size. The velocity oscillated periodically ( approximately 6 s) but was not consistently affected by structures such as en route boutons, bifurcations, or stationary mitochondria. Also, mitochondria transported in opposite directions did not necessarily affect each others' mobility. Transport was blocked by microtubule disruption (100 microM colchicine), and destabilization (1 microM cytochalasin-D) or stabilization (10 microM phalloidin) of actin filaments, respectively, decreased (0.22 +/- 0.02 microm/s, P < 0.05) or increased (0.53 +/- 0.02 microm/s, P < 0.05) transport speed. Transport was inhibited by TTX (1 microM), and removal of extracellular Ca(2+) (plus 2 mM EGTA) had no effect. However, depletion of intracellular stores (thapsigargin) reduced (to 33%) and slowed the transport significantly (0.18 +/- 0.02 microm/s, P < 0.05), suggesting an important role for stored Ca(2+) in mitochondrial transport. Transport was also reduced (to 21%) by the mitochondrial uncoupler FCCP (1 microM) in a time-dependent fashion and slowed by oligomycin (10 microM). We conclude that mitochondrial transport is remarkably independent of structural nerve fiber properties. We also show that mitochondrial transport is TTX sensitive and speeds up by stabilizing actin and that functional Ca(2+) stores are required for efficient transport.  相似文献   

16.
Within the artery intima, endothelial cells respond to mechanical cues and changes in subendothelial matrix stiffness. Recently, we found that the aging subendothelial matrix stiffens heterogeneously and that stiffness heterogeneities are present on the scale of one cell length. However, the impacts of these complex mechanical micro-heterogeneities on endothelial cells have not been fully understood. Here, we simulate the effects of matrices that mimic young and aged vessels on single- and multi-cell endothelial cell models and examine the resulting cell basal strain profiles. Although there are limitations to the model which prohibit the prediction of intracellular strain distributions in alive cells, this model does introduce mechanical complexities to the subendothelial matrix material. More heterogeneous basal strain distributions are present in the single- and multi-cell models on the matrix mimicking an aged artery over those exhibited on the young artery. Overall, our data indicate that increased heterogeneous strain profiles in endothelial cells are displayed in silico when there is an increased presence of microscale arterial mechanical heterogeneities in the matrix.  相似文献   

17.
Y Lanir 《Biorheology》1984,21(4):643-647
The possibility of fluid flux within the thickened subendothelial intima is considered. Both the media and the endothelium were already shown to be major hydraulic barriers. It is hypothesized that if the hydraulic conductivity of the inbetween layer of the subendothelial intima is considerably higher, then fluid flux in the downstream (axial) direction is likely to occur within the intima as a result of the luminal blood pressure wave. Macromolecular species (as lipoproteins) would then be transported axially by the fluid. This convective transport may give rise to the formation of early atheromas. The proposed mechanism is in accord with several clinical and experimental observations.  相似文献   

18.
Hindlimb unloading (HU) of rats induces a cephalic shift in body fluids. We hypothesized that the putative increase in cranial fluid pressure and decrease in peripheral fluid pressure would alter the morphology of resistance arteries from 2-wk HU male Sprague-Dawley rats. To test this hypothesis, the cerebral basilar, mesenteric, and splenic arteries were removed from control (C) and HU animals. The vessels were cannulated, and luminal pressure was set to 60 cmH(2)O. The resistance arteries were then relaxed with 10(-4) M nitroprusside, fixed, and cut into transverse cross sections (5 microm thick). Media cross-sectional area (CSA), intraluminal CSA, media layer thickness, vessel outer perimeter, and media nuclei number were determined. In the basilar artery, both media CSA (HU 17, 893 +/- 2,539 microm(2); C 12,904 +/- 1,433 microm(2)) and thickness (HU 33.9 +/- 4.1 microm; C 22.3 +/- 3.2 microm) were increased with hindlimb unloading (P < 0.05), intraluminal CSA decreased (HU 7,816 +/- 3,045 microm(2); C 13,469 +/- 5,500 microm(2)) (P < 0.05), and vessel outer perimeter and media nuclei number were unaltered. There were no differences in mesenteric or splenic resistance artery morphology between HU and C rats. These findings suggest that hindlimb unloading-induced increases in cephalic arterial pressure and, correspondingly, increases in circumferential wall stress result in the hypertrophy of basilar artery smooth muscle cells.  相似文献   

19.
Summary Collagen types I and III were located by immunofluorescence procedures in the aorta and coronary arteries of the rat. Type I collagen was most prevalent in the adventitia of the aorta with only small amounts present in the intima and media. Type III collagen appeared to be a significant component in the media of the aorta and also in the adventitia of both blood vessels. The intima and media of the coronary arteries did not stain strongly for either type I or III collagen. Neither staining procedure was altered with preincubation of the sections with hyaluronidase or chondroitinase ABC. These studies indicate that type III collagen is a major component of the adventitia which has previously not been recognized by immunohistochemical techniques, possibly due to masking of collagen staining with glycosaminoglycans.  相似文献   

20.
Inducible nitric oxide synthase (iNOS) is associated with vascular hypocontractility in systemic vessels after endotoxin lipopolysaccharide (LPS) administration. Although lung iNOS is increased after LPS, its role in the pulmonary circulation is unclear. We hypothesized that whereas iNOS upregulation is responsible for LPS-induced vascular dysfunction in systemic vessels, iNOS does not play a significant role in the pulmonary artery (PA). Using isolated aorta (AO) and PA rings, we examined the effect of nonselective NOS inhibition [N(G)-monomethyl-L-arginine (L-NMMA); 100 micromol/l] and selective iNOS inhibition (aminoguanidine, AG; 100 micromol/l) on alpha(1)-adrenergic-mediated vasoconstriction (phenylephrine; 10(-9) to 10(-3) M) after LPS (Salmonella typhimurium, 20 mg/kg ip). We also determined the presence of iNOS using Western blot and immunohistochemistry. LPS markedly impaired AO contractility (maximal control tension 1,076 +/- 33 mg vs. LPS 412 +/- 39 mg, P < 0.05), but PA contractility was unchanged (control 466 +/- 29 mg vs. LPS 455 +/- 27 mg, P > 0.05). Selective iNOS inhibition restored the AO's response to vasoconstriction (LPS + AG 1,135 +/- 54 mg, P > 0.05 vs. control and P < 0.05 vs. LPS), but had no effect on the PA (LPS + AG 422 +/- 38 mg, P > 0.05 vs. control and LPS). Western blot and immunohistochemistry revealed increased iNOS expression in the AO after LPS, but iNOS was not detected in the PA. Our results suggest that differential iNOS expression after LPS in systemic and pulmonary vessels contributes to the phenomenon of sepsis/endotoxemia-induced systemic hypotension and pulmonary hypertension.  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