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.
Using 133Xe, the vertical distribution of regional volume (Vr) was measured in three regions of excised canine lobes both in air and when completely submerged in saline at 40, 60, 70, and 80% lobar vital capacity (VC). The estimated pleural pressure gradient, derived from values of Vr, distance between regions, and the lobar pressure-volume (PV) curve, underestimated the true gradient by 45%. Conversely, the gradient of Vr was substantially less than predicted. From the mean depth of each region below the waterline, pleural, and hence transpulmonary, pressure (PL) was computed. The values of Vr-PL for each region at 40, 60, and 80% lung volume (VL) were related to the lobar PV curve. Slopes of lines joining initial VL-PL points on the lobar PV curve to corresponding Vr-PL points in submerged lobes represent an effective regional compliance of a lobe undergoing deformation. With one exception this was less than the corresponding homogeneous compliance, indicating a stiffening of the lobe during deformation. Slopes of lines joining Vr-PL points of each region at the three lobar volumes represent effective regional compliance of a deformed lobe undergoing volume change. This was not significantly different from the homogeneous compliance. However, effective compliance can only be an approximate indicator of the forces required for a given volume change due to the inadequacy of PL to represent the unequal stress components induced by lobe deformation.  相似文献   

2.
To determine the effects of alveolar hypoxia on pulmonary microvascular volume, X-ray microfocal angiographic images of isolated perfused dog lung lobes were obtained during passage of a bolus of radiopaque contrast medium during both normoxic (alveolar gas, 15% O(2), 6% CO(2), and 79% N(2)) and hypoxic (3% O(2), 6% CO(2), and 91% N(2)) conditions. Regions of interest (ROIs) over the lobar artery and vein at low magnification and a feeding artery ( approximately 500 microm diameter) and the nearby microvasculature (vessels smaller than approximately 50 microm) at high magnification were identified, and X-ray absorbance vs. time curves were acquired under both conditions from the same ROIs. The total pulmonary vascular volume was calculated from the flow and the mean transit time for the contrast medium passage from the lobar artery to lobar vein. The fractional changes in microvascular volume were determined from the areas under the high-magnification X-ray absorbance curves. Hypoxia decreased lobar volume by 13 +/- 3% (SE) and regional microvascular volume by 26 +/- 4% (SE). Given the morphometry of the lung vasculature, these results suggest that capillary volume was decreased by hypoxia.  相似文献   

3.
In adult canines following major lung resection, the remaining lobes expand asymmetrically, associated with alveolar tissue regrowth, remodeling, and progressive functional compensation over many months. To permit noninvasive longitudinal assessment of regional growth and function, we performed serial high-resolution computed tomography (HRCT) on six male dogs (~9 mo old, 25.0 ± 4.5 kg, ±SD) at 15 and 30 cmH(2)O transpulmonary pressure (Ptp) before resection (PRE) and 3 and 15 mo postresection (POST3 and POST15, respectively) of 65-70% of lung units. At POST3, lobar air volume increased 83-148% and tissue (including microvascular blood) volume 120-234% above PRE values without further changes at POST15. Lobar-specific compliance (Cs) increased 52-137% from PRE to POST3 and 28-79% from POST3 to POST15. Inflation-related parenchyma strain and shear were estimated by detailed registration of corresponding anatomical features at each Ptp. Within each lobe, regional displacement was most pronounced at the caudal region, whereas strain was pronounced in the periphery. Regional three-dimensional strain magnitudes increased heterogeneously from PRE to POST3, with further medial-lateral increases from POST3 to POST15. Lobar principal strains (PSs) were unchanged or modestly elevated postresection; changes in lobar maximum PS correlated inversely with changes in lobar air and tissue volumes. Lobar shear distortion increased in coronal and transverse planes at POST3 without further changes thereafter. These results establish a novel use of functional HRCT to map heterogeneous regional deformation during compensatory lung growth and illustrate a stimulus-response feedback loop whereby postresection mechanical stress initiates differential lobar regrowth and sustained remodeling, which in turn, relieves parenchyma stress and strain, resulting in progressive increases in lobar Cs and a delayed increase in whole lung Cs.  相似文献   

4.
Upright posture and lower body negative pressure (LBNP) both induce reductions in central blood volume. However, regional circulatory responses to postural changes and LBNP may differ. Therefore, we studied regional blood flow and blood volume changes in 10 healthy subjects undergoing graded lower-body negative pressure (-10 to -50 mmHg) and 8 subjects undergoing incremental head-up tilt (HUT; 20 degrees , 40 degrees , and 70 degrees ) on separate days. We continuously measured blood pressure (BP), heart rate, and regional blood volumes and blood flows in the thoracic, splanchnic, pelvic, and leg segments by impedance plethysmography and calculated regional arterial resistances. Neither LBNP nor HUT altered systolic BP, whereas pulse pressure decreased significantly. Blood flow decreased in all segments, whereas peripheral resistances uniformly and significantly increased with both HUT and LBNP. Thoracic volume decreased while pelvic and leg volumes increased with HUT and LBNP. However, splanchnic volume changes were directionally opposite with stepwise decreases in splanchnic volume with LBNP and stepwise increases in splanchnic volume during HUT. Splanchnic emptying in LBNP models regional vascular changes during hemorrhage. Splanchnic filling may limit the ability of the splanchnic bed to respond to thoracic hypovolemia during upright posture.  相似文献   

5.
Pulmonary vascular responses to sarafotoxins 6a and 6c (S6a and S6c) were investigated in the intact-chest cat under constant flow conditions. Injections of S6a and S6c into the perfused lobar artery caused dose-related increases in lobar arterial pressure, increased left atrial pressure, and produced biphasic changes in systemic arterial (aortic) pressure. When left atrial pressure was maintained constant, injections of S6a, S6c, and endothelin 1 (ET-1) caused dose-related increases in lobar arterial pressure. The increases in lobar arterial pressure in response to S6a and S6c were not altered by treatment with a cyclooxygenase inhibitor or a thromboxane receptor blocking agent. Increases in lobar arterial pressure in response to S6a and S6c were not altered when airflow to the left lower lung lobe was interrupted by bronchial occlusion, and pressor responses were not diminished when the left lower lobe was perfused with low-molecular-weight dextran. Under conditions of controlled blood flow and constant left atrial pressure, S6a, S6b, S6c, and ET-1 had similar pressor activity, whereas the thromboxane A2 mimic, U-46619, had far greater activity when compared on a nanomolar basis. The present studies demonstrate that S6a and S6c have significant vasoconstrictor activity in the feline pulmonary vascular bed. These data suggest that pulmonary vasoconstrictor responses to the endothelin peptides are not dependent on release of cyclooxygenase products and the activation of thromboxane A2 receptors, alterations in bronchomotor tone, or interaction with formed elements in blood.  相似文献   

6.
The effect of a bronchodilator in asthmatics is only partially described by changes in spirometric values since no information on regional differences can be obtained. Imaging techniques like high-resolution computed tomography (HRCT) provide further information but lack detailed information on specific airway responses. The aim of the present study was to improve the actual imaging techniques by subsequent analysis of the imaging data using computational fluid dynamics (CFD). We studied 14 mild to moderately severe asthmatics. Ten patients underwent HRCT before and 4h after inhalation of a novel long acting beta(2) agonist (LABA) that acts shortly after inhalation. Four patients were studied for chronic effects and underwent CT scans twice after adequate wash-out of bronchodilators. In the active group, a significant bronchodilator response was seen with a forced expiratory volume in 1s (FEV1) increase of 8.78 +/- -6.27% pred vs -3.38 +/- 6.87% pred in the control group. The changes in FEV1 correlated significantly with the changes in distal airway volume (r = 0.69, p = 0.007), total airway resistance (r = -0.73, p = 0.003) and distal airway resistance (r = -0.76, p = 0.002) as calculated with the CFD method. The changes in distal R(aw) were not fully homogeneous. In some patients with normal FEV1 at baseline, CFD-based changes in R(aw) were still detectable. We conclude that CFD calculations, based on airway geometries of asthmatic patients, provide additional information about changes in regional R(aw). All changes in the CFD-based calculated R(aw) significantly correlate with the observed changes in spirometric values therefore validating the CFD method for the studied application.  相似文献   

7.
When right atrial pressure (Pra) is greater than zero (atmospheric pressure), cardiac output is determined by the intersection of two functions, cardiac function and return function, which is used here to mean the determinants of venous return. When Pra < or = 0, flow is only determined by circuit function. The objective of this analysis was to determine the potential changes in return function that need to occur to allow the maximum cardiac output during exercise when Pra < or = 0 or is constant. The analysis expands on the model of Green and Jackman and includes the effects of changes in circuit parameters, including venous resistance, changes in capacitance, and muscle contractions. The analysis is based on the model of the circulation proposed by Permutt and co-workers, which assumes that the systemic circulation has two lumped compliant regions in parallel with independent inflow and outflow resistances. Changes in total flow in this model can come about by changes in the distribution of flow between the regions, recruitment of unstressed vascular volume, and changes in the regional venous resistances. The data for the analysis are from previous animal studies and are normalized to a 70-kg man. The major conclusions are that, to achieve the high cardiac output that occurs at peak exercise, there need to be marked changes in the distribution of blood flow, recruitment of unstressed volume, and the venous resistance draining vascular beds. A consequence of the increase in peripheral flow is a marked increase in pressure in the veins of the working muscle. Muscle contractions are potentially a very important mechanism for transiently decreasing this pressure and preventing excessive filtration of plasma during exercise.  相似文献   

8.
Since pulmonary blood flow to regions involved in adult respiratory disease syndrome (ARDS) is reduced by hypoxic vasoconstriction, compression by cuffs of edema, and local thromboses, we postulated that the bronchial circulation must enlarge to provide for the inflammatory response. We measured anastomotic bronchial systemic to pulmonary blood flow [QBr(s-p)] serially in a lung lobe in 31 open-chest dogs following a generalized lobar injury simulating ARDS. The pulmonary circulation of the weighed left lower lobe (LLL) was isolated and perfused (zone 2) with autologous blood in anesthetized dogs. QBr(s-p) was measured from the amount of blood which overflowed from this closed vascular circuit corrected by any changes in the lobe weight. The LLL was ventilated with 5% CO2 in air. The systemic blood pressure (volume infusion), gases, and acid-base status (right lung ventilation) were kept constant. We injured the LLL via the airway by instilling either 0.1 N HCl or a mixture of glucose and glucose oxidase or via the pulmonary vessels by injecting either alpha-naphthylthiourea or oleic acid into the LLL pulmonary artery. In both types of injury, there was a prompt rise in QBr(s-p) (mean rise = 247% compared with control), which was sustained for the 2 h of observation. The cause of this increase in flow was studied. Control instillation of normal saline into the airways or into the pulmonary vessels did not change QBr(s-p) nor did a similar increase in lobar fluid (weight) due to hydrostatic edema. Neither cardiac output nor systemic blood pressure increased.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
Pleural pressure is usually estimated with a balloon catheter (BC) positioned in the middle third of the esophagus. An alternate method, which avoids potential inaccuracies associated with changes in balloon volume, is a catheter-mounted transducer (CMT) system. To assess the accuracy of a CMT system in defining the elastic properties of the lungs, we compared the static pressure-volume (PV) properties of the lungs measured sequentially with CMT and BC systems in six healthy subjects each on two occasions, relating static transpulmonary pressure (Pst,L) to lung volume during interrupted exhalations from total lung capacity (TLC). PV data were fitted with an exponential function (least-squares method), and the exponent (k) was used to define the shape of the PV curve; position was defined by Pst,L at TLC and at 90 and 60% TLC. These data were examined for agreement (paired t test) and repeatability (coefficient of repeatability). No significant differences were demonstrated: k was 0.10 +/- 0.02 and 0.11 +/- 0.03 (SD) and Pst,L at 60% TLC was 8.27 +/- 2.09 and 8.37 +/- 1.63 cmH2O for the CMT and BC systems, respectively. The coefficient of repeatability for each parameter was not significantly different but was consistently less with the BC, suggesting better repeatability. We conclude that a CMT system is an acceptable alternative to a BC system for defining the elastic properties of lungs.  相似文献   

10.

Purpose

Flow diverters (FD) are increasingly being considered for treating large or giant wide-neck aneurysms. Clinical outcome is highly variable and depends on the type of aneurysm, the flow diverting device and treatment strategies. The objective of this study was to analyze the effect of different flow diverting strategies together with parent artery curvature variations on altering intra-aneurysmal hemodynamics.

Methods

Four ideal intracranial aneurysm models with different parent artery curvature were constructed. Computational fluid dynamics (CFD) simulations of the hemodynamics before and after applying five types of flow diverting strategies (single FD, single FD with 5% and 10% packing density of coils, two FDs with 25% and 50% overlapping rate) were performed. Changes in pressure, wall shear stress (WSS), relative residence time (RRT), inflow velocity and inflow volume rate were calculated and compared.

Results

Each flow diverting strategy resulted in enhancement of RRT and reduction of normalized mean WSS, inflow volume rate and inflow velocity in various levels. Among them, 50% overlapped FD induced most effective hemodynamic changes in RRT and inflow volume rate. The mean pressure only slightly decreased after treatment. Regardless of the kind of implantation of FD, the mean pressure, inflow volume rate and inflow velocity increased and the RRT decreased as the curvature of the parent artery increased.

Conclusions

Of all flow diverting strategies, overlapping FDs induced most favorable hemodynamic changes. Hemodynamics alterations post treatment were substantially influenced by parent artery curvature. Our results indicate the need of an individualized flow diverting strategy that is tailored for a specific aneurysm.  相似文献   

11.
Pulmonary vascular compliance and viscoelasticity   总被引:1,自引:0,他引:1  
When dog lung lobes were perfused at constant arterial inflow rate, occlusion of the venous outflow (VO) produced a rapid jump in venous pressure (Pv) followed by a slower rise in both arterial pressure (Pa) and Pv. During the slow rise Pa(t) and Pv(t) tended to converge and become concave upward as the volume of blood in the lungs increased. We compared the dynamic vascular volume vs. pressure curves obtained after VO with the static volume vs. pressure curves obtained by dye dilution. The slope of the static curve (the static compliance, Cst) was always larger than the slope of the dynamic curve (the dynamic compliance, Cdyn). In addition, the Cdyn decreased with increasing blood flow rate. When venous occlusion (VO) was followed after a short time interval by arterial occlusion (AO) such that the lobe was isovolumic, both Pa and Pv fell with time to a level that was below either pressure at the instant of AO. In an attempt to explain these observations a compartmental model was constructed in which the hemodynamic resistance and vascular compliance were volume dependent and the vessel walls were viscoelastic. These features of the model could account for the convergence and upward concavity of the Pa and Pv curves after VO and the pressure relaxation in the isovolumic state after AO, respectively. According to the model analysis, the difference between Cst and Cdyn and the flow dependence of Cdyn are due to wall viscosity and volume dependence of compliance, respectively. Model analysis also suggested ways of evaluating changes in the viscoelasticity of the lobar vascular bed. Hypoxic vasoconstriction that increased total vascular resistance also decreased Cst and Cdyn and appeared to increase the vessel wall viscosity.  相似文献   

12.
Intrahepatic pressure (9.4 +/- 0.3 mmHg; 1 mmHg = 133.32 Pa), measured proximal to a hepatic venous resistance site, was insignificantly different from portal venous pressure (9.6 +/- 0.4 mmHg). This lobar venous pressure is not wedged hepatic venous pressure as it is measured from side holes in a catheter with a sealed tip. Validation of the lobar venous pressure measurement was done in a variety of ways and using different sizes and configurations of catheters. The site of hepatic venous resistance in the dog is localized to a narrow sphincterlike region about 0.5 cm in length and within 1-2 cm (usually within 1 cm) of the junction of the vena cava and hepatic veins. Sinusoidal and portal venous resistance appears insignificant in the basal state and large increases in liver blood volume (histamine infusion or passive vena caval occlusion) or large decreases in liver blood volume (passive vascular occlusion) do not alter the insignificant pressure gradient between portal and lobar venous pressures. Norepinephrine infusion (1.25 microgram X kg-1 X min-1 intraportal) and hepatic sympathetic nerve stimulation (10 Hz) led to a significantly greater rise in portal venous pressure than in lobar venous pressure, indicating some presinusoidal (and (or) sinusoidal) constriction and this indicates that lobar venous pressure cannot be assumed under all conditions to accurately reflect portal pressure. However, most of the rise in portal venous pressure induced by intraportal infusion of norepinephrine or nerve stimulation and virtually all of the pressure rise induced by histamine could be attributed to the postsinusoidal resistance site.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
The low-viscosity bolus method was used to determine the longitudinal distributions of vascular resistance and intravascular pressure with respect to cumulative vascular volume from the lobar artery to the lobar vein in isolated dog lung lobes near functional residual capacity under zone 3 conditions. We found that the resistance distribution had two modes, a larger one upstream and a smaller one downstream from a local minimum. Over the range of vascular pressures studied the total vascular resistance decreased and the vascular volume increased with increasing vascular pressure. However, the shape of the normalized resistance distribution was independent of vascular pressure. Comparisons of the resistance distributions with the distributions of arterial, capillary, and venous volumes suggest that the modes represent regions of relatively high resistance proximal and distal to the capillary bed. These results are consistent with the concept that within the lobar vascular bed the highest resistance per unit blood volume is in the smallest arteries and veins, as suggested by morphometric data from other sources.  相似文献   

14.
Dog left upper lobes (LUL) were perfused in situ via the left lower lobe artery. Lobe weight was continuously monitored. Increasing lobar flow from normal to 10 times normal had little effect on left atrial pressure, which ranged from 1 to 5 mmHg. There was a flow threshold (Qth) below which lobar weight was stable. Qth ranged from 1.1 to 1.55 l/min (mean 1.27) corresponding to four times normal LUL blood flow. Above Qth, step increases in lobar flow resulted in progressive weight gain at a constant rate that was proportional to flow. The effective pressure at the filtration site (EFP) at different flow rates was estimated from the static vascular pressure that resulted in the same rate of weight gain. From this value and from mean pulmonary arterial (PA) and left atrial (LA) pressures, we calculated resistance upstream (Rus) and downstream (Rds) from filtration site. At Qth, Rds accounted for 60% of total resistance. This fraction increased progressively with flow, reaching 83% at Q of 10 times normal. We conclude that during high pulmonary blood flow EFP is closer to PA pressure than it is to LA pressure, and that this becomes progressively more so as a function of flow. As a result, the lung accumulates water at flow rates in excess of four times normal despite a normal left atrial pressure.  相似文献   

15.
Five dogs underwent left pneumonectomy at 10 wk of age, whereas four littermates underwent a sham operation. At 26 wk of age the postpneumonectomy dogs had total lung vital capacity (VC) and lung weight similar to controls, but maximum expiratory flow was reduced. Pressure capsules were glued to right lower (RLL) and right cardiac (RCL) lobes, and alveolar pressures (PA) were measured during forced expiration. In postpneumonectomy dogs RLL and RCL both emptied more slowly than in control dogs, and emptying was especially delayed in RCL, which underwent the most growth. When both lobes deflated together, PA in RCL and RLL were similar in control dogs, but in postpneumonectomy dogs PA in RCL exceeded that in RLL by approximately 3 cmH2O from 80 to 20% VC. Because the higher driving pressure in RCL compensated for the relatively high resistance of RCL, the pattern of lobar emptying was relatively uniform over these lung volumes. This result was compatible with interdependence of lobar maximum expiratory flows. In addition, at PA of 6-10 cmH2O in postpneumonectomy dogs, maximum emptying rates of RCL were less when RCL deflated alone than when RCL and RLL emptied together, again demonstrating interdependence of lobar maximum expiratory flow.  相似文献   

16.
The purpose of this preliminary study was to test the hypothesis that subsyndromal depression is associated with the volume of medial prefrontal regional gray matter and that of white matter lesions (WMLs) in the brains of cognitively normal older people. We also explored the relationships between subsyndromal depression and medial prefrontal regional gray matter volume, limbic regional gray matter volume, and lobar WMLs in the brains of patients with mild cognitive impairment (MCI) and Alzheimer''s disease (AD). We performed a cross-sectional study comparing patients with subsyndromal depression and nondepressed controls with normal cognition (n = 59), MCI (n = 27), and AD (n = 27), adjusting for sex, age, years of education, and results of the Mini-Mental State Examination. Frontal WML volume was greater, and right medial orbitofrontal cortical volume was smaller in cognitively normal participants with subsyndromal depression than in those without subsyndromal depression. No volume differences were observed in medial prefrontal, limbic, or WML volumes according to the presence of subsyndromal depression in cognitively impaired patients. The absence of these changes in patients with MCI and AD suggests that brain changes associated with AD pathology may override the changes associated with subsyndromal depression.  相似文献   

17.
Although several factors are known to influence nonuniformity of ventilation, including lung mechanical properties (regional structure and compliance), external factors (chest wall, pleural pressure, heart), and ventilatory parameters (tidal and preinspiratory volume, flow rate), their relative contributions are poorly understood. We studied five excised, unperfused, canine right-middle lobes under varied levels of tidal volume (VT), thus eliminating many factors affecting heterogeneity. Multiple-breath washouts of N(2) were analyzed for anatomic dead space volume (VD(anat)), nonuniformity of N(2) washout, and nonuniformity between joined acinar regions vs. that occurring between larger joined regions. Approximately 80% of ventilation heterogeneity was found among joined acinar regions at resting levels of VT, but increasing VT reduced intra-acinar heterogeneity to about 25% of that found at resting levels. Increasing VT had essentially no effect on VD(anat) and heterogeneity among larger joined regions. The results indicate that the magnitude of VT is a major influence on the dominant intra-acinar component of ventilation heterogeneity and that VT effects on VD(anat) are likely due to perfusion and/or influences normally external to the lobar structure.  相似文献   

18.
Past work in the isolated perfused cat lung has shown that acute hypoxia (H) changes the response to norepinephrine (NE) from vasoconstriction to vasodilation but has no effect on the response to serotonin (S). These results could be related to the increase in pulmonary arterial pressure or vascular resistance during the hypoxic pressor response or a direct effect of H. We addressed this question, in the same preparation, by comparing responses to NE under four conditions in each experimental animal (n = 12): 1) NE infused during normoxia; 2) NE infused after vascular resistance (Rpv) was increased with serotonin; 3) NE infused after Rpv was increased by H; 4) NE infused after lobar pressure was raised by an increase in flow (P/F). PO2 values during H were varied (27-56 Torr). S and H produced a 137 +/- 35 and 43 +/- 8% delta Rpv increase in lobar vascular resistance, respectively. P/F increased lobar pressure 91 +/- 10%. Only NE infusion during H demonstrated significant differences in lobar pressure and Rpv compared with control normoxic periods. There was no correlation between responses to NE during S, H, and P/F and degree to which each stimulus increased Rpv or lobar pressure (r = 0.003, 0.28, 0.24). A significant relationship between response to NE during H vs. PO2 during H was observed (r = 0.78; P less than 0.001). In a subset of animals, we repeated the infusion of NE during H and P/F post-beta-blockade. The decrease in vascular response to NE during H and the correlation of PO2 with NE response were abolished (n = 7).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
We examined maximum expiratory flow (Vmax) in two canine preparations in which regional changes in lung mechanical properties were produced. In one experiment serial bronchial obstructions were made to determine whether flow-limiting sites (choke points, CP) would occur in series. With the right lung tied off, constrictions were placed at the left lower lobar bronchus (LLL) and left main-stem bronchus. On deflation from total lung capacity, the obstructed LLL and nonobstructed left upper lobe (LUL) emptied into the obstructed left main-stem bronchus. Although a CP common to both lobes was identified at the main-stem obstruction, which limited total Vmax, we questioned whether there was also a CP at the lobar obstruction that fixed LLL flow. In that case the rate of LLL emptying would not be dependent on the presence of the common (i.e., central) CP and thus the flow contribution of the LUL. We found that when the LUL was removed, the LLL increased its rate of emptying. Thus a lobar CP did not fix LLL flow and CP did not occur in series. In a second experiment emphysema was produced in the left lung to reduce lung recoil, whereas the right lung was normal. CP were identified at approximately lobar bronchi of each lung, and the lungs were emptied at different rates. A CP common to both lungs was not identified. Our results indicate that in localized lung disease, if flows from the different regions are high enough, then wave speed is reached in proximal airways, and a CP occurs centrally rather than peripherally. On the other hand, if flows are low, then wave speed is reached peripherally and a CP common to all lung regions does not occur.  相似文献   

20.
The equilibrium pressure obtained during simultaneous occlusion of hepatic vascular inflow and outflow was taken as the reference estimate of hepatic vascular distending pressure (P(hd)). P(hd) at baseline was 1.1 +/- 0.2 (mean +/- SE) mmHg higher than hepatic vein pressure (P(hv)) and 0.7 +/- 0.3 mmHg lower than portal vein pressure (P(pv)). Norepinephrine (NE) infusion increased P(hd) by 1. 5 +/- 0.5 mmHg and P(pv) by 3.7 +/- 0.6 mmHg but did not significantly increase P(hv). Hepatic lobar vein pressure (P(hlv)) measured by a micromanometer tipped 2-Fr catheter closely resembled P(hd) both at baseline and during NE-infusion. Dynamic pressure-volume (PV) curves were constructed from continuous measurements of P(hv) and hepatic blood volume increases (estimated by sonomicrometry) during brief occlusions of hepatic vascular outflow and compared with static PV curves constructed from P(hd) determinations at five different hepatic volumes. Estimates of hepatic vascular compliance and changes in unstressed blood volume from the two methods were in close agreement with hepatic compliance averaging 32 +/- 2 ml. mmHg(-1). kg liver(-1). NE infusion reduced unstressed blood volume by 110 +/- 38 ml/kg liver but did not alter compliance. In conclusion, P(hlv) reflects hepatic distending pressure, and the construction of dynamic PV curves is a fast and valid method for assessing hepatic compliance and changes in unstressed blood volume.  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