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 62 毫秒
1.
The effect of intravenous dipyridamole (0.7 mg/kg) on cerebral blood flow (CBF), mean arterial blood pressure (MABP), heart rate, respiration rate, cerebral electrical activity, arterial blood gases, pH, and glucose was investigated in 14 normotensive and 14 stroke-prone spontaneously hypertensive anesthetized rabbits. CBF was measured by hydrogen and heat clearance. In both groups, MABP decreased (normotensive: -24 mm Hg, hypertensive: -47 mm Hg; ANOVA: P < 0.0001) and CBF increased (normotensive: +59 ml/100 g/min, hypertensive: +72 ml/100 g/min; ANOVA: P < 0.0002). CBF returned to the initial level 21 min later in hypertensive than in normotensive rabbits. Changes in other parameters were insignificant. In additional experiments, 30 mg/kg theophylline entirely prevented the cerebral vasodilator and systemic hypotensive effects of dipyridamole in both normotensive and hypertensive rabbits. We conclude that, in stroke-prone spontaneously hypertensive rabbits, the longer-lasting and larger CBF increase in response to dipyridamole may be attributed to reversible functional changes in the cerebral vasculature resulting from hypertension.  相似文献   

2.
Cerebral autoregulation, the physiological regulatory mechanism that maintains a constant cerebral blood flow (CBF) over wide ranges of arterial blood pressure, was investigated in normotensive and spontaneously hypertensive rats by means of laser-Doppler flowmetry. Systemic arterial hypertension was produced at rates ranging from 0.02 mm Hg/second to 11 mm Hg/second by constant infusion of epinephrine and norepinephrine. Systemic arterial hypotension was produced at rates ranging from -0.03 mm Hg/second to -12 mm Hg/second, either by bleeding the animals into a reservoir or by compressing the abdomen. In those cases with a low rate of change in systemic arterial blood pressure (SABP), the measurements lasted for 5 +/- 2 minutes, and in those with a high rate of change in SABP, measurements lasted for 40 +/- 30 seconds. The purpose was to record the time of onset and course of autoregulation in the basal ganglia in response to slow or rapid changes in SABP. CBF in the basal gray matter remained at baseline values (i.e., autoregulation was functioning) if the rate of increase of SABP did not exceed a critical value (0.10 mm Hg/second in the normotensive rats; 0.35 mm Hg/second in the spontaneously hypertensive rats). When hypertension was produced at faster rates, CBF followed arterial blood pressure passively, and no autoregulatory response was observed for 2 +/- 1 minutes. Hypotension did not change the baseline CBF when it was not produced at a rate faster than -0.4 mm Hg/second in normotensive rats and -0.15 mm Hg/second in spontaneously hypertensive rats.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
The cardiac hypertrophy observed in hypertension is thought to be responsible for the accompanying deficiency in the baroreflex control of heart rate. In this study, we assessed the baroreflex relationship between heart rate and arterial pressure on a group of seven rabbits during a normotensive period, during the early phase of angiotensin II (Ang II)-induced hypertension II week) (50 ng/kg per minute i.v. via osmotic minipumps), after 7 weeks of continuous hypertension, then 2 days after Ang II was stopped, and finally 7 days after Ang II. Left ventricles were weighed for measurement of left ventricular weight-body weight ratio. One week of intravenous Ang II infusion produced hypertension (mean arterial pressure from 80 +/- 2 up to 115 +/- 8 mm Hg), with significantly increased heart rate and hematocrit. The heart rate-arterial pressure baroreflex curve was shifted to the right, with a significant 45% reduction in the gain of the reflex (-6.4 +/- 1.5 to -3.5 +/- 0.2 beats per minute/mm Hg). After 7 weeks of Ang II, arterial pressure was still elevated (112 +/- 4 mm Hg) and the gain of the baroreflex curve still somewhat attenuated, although it was no longer markedly different from normotensive levels (gain, -5.09 +/- 0.95, 20% reduction from normotensive level). Two days after the Ang II infusion was stopped, arterial pressure had returned to normotensive levels, although hematocrit and heart rate remained elevated. At this time, the baroreflex curve was similar to prehypertensive control levels, with no further changes when measured again 7 days after Ang II. Cardiac hypertrophy was present when measured at 7 days after angiotensin (left ventricular weight-body weight ratio: 1.78 +/- 0.05 versus 1.35 +/- 0.04 g/kg, hypertensive versus normotensive, P < .05). Thus, although Ang II infusion produced an initial deficit in the baroreflex control of heart rate, this effect became less as the hypertension continued. Furthermore, although cardiac hypertrophy developed, its presence did not appear to be sufficient to produce a decrease in barosensitivity independent of raised arterial pressure.  相似文献   

4.
The length-active tension relation has been previously reported to be decreased or unchanged in hypertensive vessels whereas resting distensibility was unchanged or increased. We found the maximum active stress and the internal ring circumference, in millimeters, at which it occurs (Lmax) to be lower in arterial rings from perinephritic hypertensive dogs than in rings from normotensive dogs. The internal circumferences (length) at which resting force and active force became zero (L0 and Lmin, respectively) were unchanged. Lmax, L0 and Lmin were used to normalize length-tension diagrams. Active stress was significantly lower in hypertensive vessels at most of the lengths tested with the diagram normalized to Lmax. When the length-tension diagram was normalized to Lmin there was no difference in the active stress at any of the lengths tested. The length-resting stress curves were identical when rhe diagram was normalized to Lmax but the curve for hypertensive vessels was higher when the diagram was normalized to L0. An important characteristic of these length-tension curves is that normalized lengths correspond to the same absolute length in each group of vessels when the reference length has the same absolute value (L0 and Lmin in this study). This separates differences due to absolute length from differences associated with hypertension. We conclude that perinephritic hypertension in the dog is accompanied by a decrease in resting distensibility of the arterial wall. The results indicate that the choice of reference length may affect the values of stress and tension that are obtained for comparison of length-tension relationships in hypertensive and normotensive blood vessels.  相似文献   

5.
BACKGROUND: Endothelial dysfunction with a loss of endothelium-dependent vasodilation has been reported in patients with arterial hypertension. The purpose of the present study was to evaluate coronary vasomotor response to dynamic exercise in patients with coronary artery disease with and without arterial hypertension and to determine the effect of calcium antagonists on coronary vasomotion. METHODS AND RESULTS: Cross-sectional areas of a normal and a stenotic coronary vessel segment were examined in 79 patients with coronary artery disease at rest and during supine bicycle exercise (Ex). Change in luminal area after acute administration of a calcium antagonist (diltiazem or nicardipine), during exercise, and after sublingual nitroglycerin (percent change compared with rest = 100%) was assessed by biplane quantitative coronary arteriography. Patients were divided into two groups: Group 1 (control) consisted of 48 patients without (normotensive subjects, n = 30; hypertensive subjects, n = 18) and group 2 of 31 patients with (normotensive subjects, n = 15; hypertensive subjects, n = 16) pretreatment with a calcium antagonist immediately before exercise. The groups did not differ with regard to clinical characteristics or hemodynamic data measured during exercise. Mean aortic pressure at rest, however, was significantly increased in hypertensive patients compared with normotensive subjects in group 1 (103 mm Hg versus 92 mm Hg, P < .01) and group 2 (110 mm Hg versus 98 mm Hg, P < .025). In group 1, exercise-induced vasomotor response was significantly different between normotensive and hypertensive patients in normal (+20% versus +1%, P < .003) and stenotic vessels (-5% versus -20%, P < .025). However, in group 2 there was coronary vasodilation in normotensive and hypertensive patients for both normal (delta Ex +23% versus +21%, P = NS) and stenotic vessel segments (+24% versus +26%, P = NS). CONCLUSIONS: Abnormal coronary vasomotion during exercise can be observed in hypertensive patients with reduced vasodilator response in normal arteries and enhanced vasoconstrictor response in stenotic arteries. Calcium antagonists prevent the abnormal response of normal and stenotic coronary arteries to exercise in hypertensive patients and thus may compensate for endothelial dysfunction with reduced vasodilator response to exercise.  相似文献   

6.
This study was performed to compare metabolic and endocrine characteristics of untreated hypertensive patients and normal controls. Measurements were made in age-matched, body mass index (BMI) matched, normotensive patients with (n = 40; age = 53; BMI = 28) and without (n = 39; age = 54; BMI = 27) a family history of hypertension and hypertensive patients with (n = 38; age = 53; BMI = 28) and without (n = 25; age = 54; BMI = 29) a family history of hypertension. Norepinephrine, renin activity, and total cholesterol blood concentrations were similar in normotensive patients with a positive family history of hypertension and in hypertensive patients with or without a family history. Similarly, there were no differences in plasma insulin concentrations or insulin/glucose ratios between the normotensive patients with a family history of hypertension and hypertensive patients with or without a family history. But in all three groups the values were significantly greater (at least p < 0.05 for each) than in the normotensive patients without a family history. Increases in systolic blood pressure during treadmill testing were 51 +/- 4 mm Hg in the normotensive patients with a family history, 50 +/- 3 mm Hg in hypertensives with a family history, and 45 +/- 5 mm Hg in hypertensives without a family history; these changes were all less (p < 0.05 for each) than in normotensives without a family history (65 +/- 3 mm Hg).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
To assess whether patients with mild essential hypertension have excessive activities of the sympathoneuronal and adrenomedullary systems, we examined total body and forearm spillovers and norepinephrine and epinephrine clearances in 47 subjects with mild essential hypertension (25 men, 22 women, aged 38.1 +/- 6.7 years) and 43 normotensive subjects (19 men, 24 women, aged 36.5 +/- 5.9 years). The isotope dilution method with infusions of tritiated norepinephrine and epinephrine was used at rest and during sympathetic stimulation by lower body negative pressure at -15 and -40 mm Hg. Hypertensive subjects had a higher arterial plasma epinephrine concentration (0.20 +/- 0.01 nmol.L-1: mean +/- SE) than normotensive subjects (0.15 +/- 0.01) (P < .01). The increased arterial plasma epinephrine levels appeared to be due to a higher total body epinephrine spillover rate in the hypertensive subjects (0.23 +/- 0.02 nmol.min-1.m-2) than the normotensive subjects (0.18 +/- 0.01) (P < .05) and not to a decreased plasma clearance of epinephrine. The arterial plasma norepinephrine level, total body and forearm norepinephrine spillover rates, and plasma norepinephrine clearance were not altered in the hypertensive subjects. The responses of the catecholamine kinetic variables to lower body negative pressure were not consistently different between normotensive and hypertensive individuals. These data indicate that individuals with mild essential hypertension (1) have elevated arterial plasma epinephrine concentrations that are due to an increased total body epinephrine spillover rate, indicating an increased adrenomedullary secretion of epinephrine; (2) have no increased generalized sympathoneuronal activity and no increased forearm norepinephrine spillover; and (3) have similar responses of both the sympathoneuronal and adrenomedullary systems to sympathetic stimulation by lower body negative pressure.  相似文献   

8.
Neurological and vascular complications of Arnold-Chiari malformation treated with ventriculoatrial shunting may result in sudden or unexpected death. Two patients with Arnold-Chiari malformation and ventriculoatrial shunting had variable clinical manifestations and diagnostic difficulties. A 3-year-old girl with a 1-day history of right-sided heart failure died unexpectedly soon after cardiac catheterization. At autopsy examination an adherent thrombus around the ventriculoatrial catheter tip, pulmonary infarction, and embolic pulmonary arterial hypertensive changes were found. In the second case, a 21-year-old man died suddenly after a brief episode of dyspnea. He had a 1-year history of "asthma" before death. Autopsy examination confirmed pulmonary infarction and embolic pulmonary arterial hypertensive changes. There was no histological evidence of asthma. Deaths in both cases were due to pulmonary infarction stemming from thromboemboli derived from ventriculoatrial catheterization. Both patients had evidence of long-standing clinically unsuspected vascular disease, which may have contributed to death. Cardiac catheterization may also have precipitated death in the first patient. Other possible problems leading or contributing to sudden death in such patients include pulmonary hypertension with chronic cor pulmonale, airway obstruction from recurrent laryngeal nerve paralysis, and shunt blockage with acute hydrocephalus. Lethal brainstem compression may also accompany relatively minor trauma associated with chronic cerebellar tonsillar herniation in these patients.  相似文献   

9.
Changes in aortic lipolytic enzyme activities (cholesterol esterase and lipoprotein lipase) and acid phosphatase activity during aging were investigated in three strains of rats with different blood pressures; stroke prone spontaneously hypertensive rats (SHRSP), spontaneously hypertensive rats (SHR) and normotensive Wistar Kyoto (WKR). The blood pressures of male, 7 month old animals, was 234 (SHRSP), 173 (SHR) and 128 (WKR) mmHg. The cholesterol esterase activity markedly decreased with age in the aortas of SHRSP, SHR and normotensive WKR rats, while acid phosphatase activity decreased only slightly, if at all, and lipoprotein lipase activity remained unchanged. This effect was enhanced by increasing blood pressure in SHRSP, SHR and WKR. The total aortic cholesterol content increased significantly with hypertension in a inverse relation with cholesterol esterase activity. These results suggest that cholesterol deposition in aged arteries is, at least partialy, ascribable to an age-related decrease in cholesterol esterase, and that hypertension aggravates the deposition of arterial cholesterol by accelerating the age-related decrease in aortic cholesterol esterase activity.  相似文献   

10.
OBJECTIVES: This study sought to determine the site of increased pulmonary vascular resistance (PVR) in primary pulmonary hypertension by standard bedside hemodynamic evaluation. BACKGROUND: The measurement of pulmonary vascular pressures at several levels of flow (Q) allows the discrimination between active and passive, flow-dependent changes in mean pulmonary artery pressure (Ppa), and may detect the presence of an increased pulmonary vascular closing pressure. The determination of a capillary pressure (Pc') from the analysis of a Ppa decay curve after balloon occlusion allows the partitioning of PVR in an arterial and a (capillary + venous) segment. These approaches have not been reported in primary pulmonary hypertension. METHODS: Ppa and Pc' were measured at baseline and after an increase in Q induced either by exercise or by an infusion of dobutamine, at a dosage up to 8 microg/kg body weight per min, in 11 patients with primary pulmonary hypertension. Reversibility of pulmonary hypertension was assessed by the inhalation of 20 ppm nitric oxide (NO), and, in 6 patients, by an infusion of prostacyclin. RESULTS: At baseline, Ppa was 52+/-3 mm Hg (mean value+/-SE), Q 2.2+/-0.2 liters/min per m2, and Pc' 29+/-3 mm Hg. Dobutamine did not affect Pc' and allowed the calculation of an averaged extrapolated pressure intercept of Ppa/Q plots of 34 mm Hg. Inhaled NO had no effect. Prostacyclin decreased Pc' and PVR. Exercise increased Pc' to 40+/-3 mm Hg but did not affect PVR. CONCLUSIONS:ns. These findings are compatible with a major increase of resistance and reactivity at the periphery of the pulmonary arterial tree.  相似文献   

11.
We hypothesized that left atrial hypertension results in pulmonary vasoconstriction, which is obscured by the expected passive decrease in pulmonary vascular resistance. The objectives of this study were to demonstrate and quantify the vasoconstrictive changes that occur in the pulmonary circulation during experimental left atrial hypertension, to determine the site of vasoconstriction, and to explore its mechanism. Sheep were instrumented for measurement of pulmonary arterial (Ppa), left atrial (Pla), and systemic arterial pressures (Psa) with a Foley balloon catheter to variably obstruct the mitral valve. Distal pulmonary arterial wedge pressure (Ppaw) was determined by using a 5-Fr Swan-Ganz catheter that was advanced until it wedged with the balloon deflated. Cardiac output (CO) was estimated by thermodilution; pulmonary vascular resistances (PVR) were calculated as mean (Ppa - Pla)/CO = total PVR, (Ppa - Ppaw)/CO = upstream PVR, and (Ppaw - Pla)/CO = downstream PVR. We studied 15 awake sheep at baseline and during increases in Pla of 10 and 20 cmH2O, with and without inhalation of approximately 36 parts per million of nitric oxide. Left atrial hypertension resulted in elevation of Ppa. CO decreased only slightly at both levels of Pla elevation. Nitric oxide inhalation caused a significant decrease in PVR, which was greater as Pla increased. This vasodilator effect was most striking in downstream vessels. Experiments with phentolamine, atropine, and ibuprofen failed to reveal the mechanism of the reactive pulmonary vasoconstriction.  相似文献   

12.
Twelve patients with predominantly obstructive type sleep apnea underwent cardiac catheterization, hemodynamic monitoring, and arterial blood gas analysis during wakefulness and sleep. Abnormalities during wakefulness included systemic hypertension in four of 12, exercise-induced mild pulmonary hypertension in five of 12, and alveolar hypoventilation in one. During sleep nine patients had cyclic elevations of arterial pressure with each apneic episode, exceeding 200 mm Hg systolic in three of 12. Pulmonary artery pressures increased in 10 of 12, exceeding 60 mm Hg systolic in five. Marked degrees of hypoxemia (arterial P02, less than 50 mm Hg in eight of 12) and moderate hypercapnia with respiratory acidosis were associated with these hemodynamic changes. Cyclic upper airway obstruction during sleep may result in hypercapnia, acidosis, and pronounced hypoxemia, which can lead to hemodynamic abnormalities during sleep. Sustained pulmonary hypertension and possibly systemic hypertension may follow. Tracheostomy is an effective therapy and is recommended to symptomatic patients who have predominantly obstructive apnea but no relievable anatomic cause of upper airway obstruction.  相似文献   

13.
BACKGROUND: Inhibition of an endothelium-derived relaxing factor (EDRF) may contribute to the pathogenesis of thrombotic arterial occlusions. EXPERIMENTAL DESIGN: We measured the blood pressure and urinary excretion of protein, sodium, and potassium and histologically examined the brains, hearts, and kidneys in normotensive Wistar Kyoto rats (WKY) and stroke-prone spontaneously hypertensive rats (SHRSP) fed on a diet containing: (a) EDRF inhibitor (L-N-nitroarginine:L-NNA); (b) L-arginine, which reverses the effect of L-NNA; or (c) both L-NNA and L-arginine for 1 to 8 weeks. In addition, we examined L-NNA-treated SHRSP, the blood pressures of which were lowered using hydralazine. Furthermore, we produced and examined Goldblatt's renal hypertensive rats, which are of a different type from those resulting from the L-NNA treatment. RESULTS: Both WKY and SHRSP rats fed on a diet containing L-NNA suffered from hypertension and cerebral infarctions in a dose-dependent manner. Cerebral infarctions occurred whether or not SHRSP rats were treated with an antihypertensive agent when they were fed a high dosage of L-NNA. In contrast, SHRSP rats, treated simultaneously with both L-NNA and L-arginine, suffered few cerebral infarctions, although they were severely hypertensive. In addition, there were no cerebral infarctions in Goldblatt's renal hypertensive rats, although they suffered from advanced hypertension. CONCLUSIONS: The data indicate that the inhibition of EDRF injures the vessel walls and encourages platelet adhesion to the damaged areas. The adhering platelets narrow the lumen with resultant thrombotic arterial occlusions. Pathophysiologic conditions that decrease EDRF synthesis appear to play an important role in cerebral, renal, and myocardial infarctions.  相似文献   

14.
Radiation-induced genomic instability   总被引:1,自引:0,他引:1  
OBJECTIVES: To evaluate the effects of hypertension on heart rate and left ventricular responses to beta-agonist in young and older subjects, as well as the modulating effect of the arterial baroreflex on these responses. METHODS: Isoproterenol (INN, isoprenaline) alone was infused in 14 young normotensive subjects (mean age, 30 +/- 2 years), 18 older normotensive subjects (mean age, 60 +/- 2 years), 11 young hypertensive subjects (mean age, 36 +/- 1 years), and 17 older hypertensive subjects (mean age, 59 +/- 1 years); isoproterenol combined with ganglionic blockade (trimethaphan [INN, trimetaphan]) was administered to eight young normotensive subjects and eight young hypertensive subjects. Isoproterenol was infused at three to four incremental rates, each rate for 8 minutes. Left ventricular responses were assessed by echocardiography. RESULTS: Isoproterenol caused similar increases in heart rate in all four groups. With ganglionic blockade, heart rate responses were enhanced but were similar in the young normotensive and hypertensive subjects. In young subjects, hypertension did not affect left ventricular responses to isoproterenol alone, whereas older hypertensive subjects showed some blunting of left ventricular responses compared with older normotensive subjects. With ganglionic blockade, young hypertensive subjects also showed mild blunting of left ventricular responses. CONCLUSION: These results show that, in humans, hypertension does not lead to a decrease in chronotropic responses to infusion of the beta-agonist isoproterenol and causes only a modest decrease in left ventricular responses.  相似文献   

15.
OBJECTIVES: To evaluate whether the extent of autonomic activation following brain infarction differs between hypertensive and normotensive humans, and to investigate the role of the insular cortex for this sympathetic activation. DESIGN: Prospective, hospital-based study. SETTING: Department of Neurology of a university medical center. SUBJECTS: Forty-two patients with essential hypertension and 45 patients who were normotensive. MAIN OUTCOME MEASURES: Extent of autonomic activation following stroke as indicated by circadian blood pressure patterns, serum norepinephrine levels, and cardiovascular variables. RESULTS: Normotensive patients with insular infarction showed a significantly reduced circadian blood pressure variation and a higher frequency of nocturnal blood pressure increase compared with patients suffering from essential hypertension and insular stroke. These findings were also associated with higher serum norepinephrine concentrations and more frequent electrocardiographic abnormalities. No significant changes in these variables were seen between normotensive and hypertensive patients without insular involvement. CONCLUSIONS: Our findings suggest a difference in cortical control of autonomic function between hypertensive and normotensive patients after stroke and point to a possible role of the insular cortex in the pathogenesis of essential hypertension.  相似文献   

16.
OBJECTIVES: To assess the relation between white coat hypertension and alterations of left ventricular structure and function. DESIGN: Cross sectional survey. SETTING: Augsburg, Germany. SUBJECTS: 1677 subjects, aged 25 to 74 years, who participated in an echocardiographic substudy of the monitoring of trends and determinants in cardiovascular disease Augsburg study during 1994-5. OUTCOME MEASURES: Blood pressure measurements and M mode, two dimensional, and Doppler echocardiography. After at least 30 minutes' rest blood pressure was measured three times by a technician, and once by a physician after echocardiography. Subjects were classified as normotensive (technician <140/90 mm Hg, physician <160/95 mm Hg; n=849), white coat hypertensive (technician <140/90 mm Hg, physician >=160/95 mm Hg; n=160), mildly hypertensive (technician >=140/90 mm Hg, physician <160/95 mm Hg; n=129), and sustained hypertensive (taking antihypertensive drugs or blood pressure measured by a technician >=140/90 mm Hg, and physician >=160/95 mm Hg; n=538). RESULTS: White coat hypertension was more common in men than women (10.9% versus 8.2% respectively) and positively related to age and body mass index. After adjustment for these variables, white coat hypertension was associated with an increase in left ventricular mass and an increased prevalence of left ventricular hypertrophy (odds ratio 1.9, 95% confidence interval 1.2 to 3.2; P=0.009) compared with normotensive patients. The increase in left ventricular mass was secondary to significantly increased septal and posterior wall thicknesses whereas end diastolic diameters were similar in both groups with white coat hypertension or normotension. Additionally, the systolic white coat effect (difference between blood pressures recorded by a technician and physician) was associated with increased left ventricular mass and increased prevalence of left ventricular hypertrophy (P<0.05 each). Values for systolic left ventricular function (M mode fractional shortening) were above normal in subjects with white coat hypertension whereas diastolic filling and left atrial size were similar to those in normotension. CONCLUSION: About 10% of the general population show exaggerated inotropic and blood pressure responses when mildly stressed. This is associated with an increased risk of left ventricular hypertrophy.  相似文献   

17.
BACKGROUND: Endothelial dysfunction of coronary arteries with impaired vasodilation has been reported in patients with arterial hypertension. However, the effect of dynamic exercise on coronary vasomotion of a stenotic vessel segment before and after PTCA has not yet been evaluated in these patients. METHODS AND RESULTS:Coronary vasomotion of a normal and a stenotic vessel segment was studied in 39 patients with coronary artery disease during supine bicycle exercise before and 9+/-3 months after PTCA. Luminal area changes were determined by biplane quantitative coronary arteriography. There were 21 normotensive and 18 hypertensive patients who did not differ with regard to clinical characteristics. Percent area stenosis decreased after PTCA from 90% to 39% (P<0.001) in normotensive and from 86% to 33% (P<0.001) in hypertensive patients. Exercise-induced vasomotion of the normal vessel segment was significantly different between normotensives and hypertensives before (+19% versus +1%, P<0.01) and after (+16% versus +3%, P<0.01) PTCA. In contrast, stenotic vessel segments showed vasoconstriction in both normotensive and hypertensive patients (Deltaexercise, -11% versus - 20%, P=NS), which was reversed after PTCA (+3% versus +2%, P=NS). CONCLUSIONS: Normal coronary arteries show reduced vasodilation during exercise in hypertensive patients that may be explained by the presence of endothelial dysfunction. Stenotic vessels demonstrate paradoxical vasoconstriction during exercise in both normotensive and hypertensive patients. PTCA reverses vasoconstriction by elimination of the flow-limiting stenosis and prevention of coronary stenosis narrowing during exercise in normotensive and hypertensive patients.  相似文献   

18.
The mechanisms responsible for reduced arterial distensibility in renal transplant recipients remain to be evaluated. The present longitudinal study was aimed to evaluate the effect of hypertension on the evolution of vessel wall properties in renal transplant recipients. The mechanical properties of the common carotid artery were determined in 24 normotensive and 24 treated hypertensive renal transplant recipients 6-12 weeks after transplantation. The measurements were repeated after 2 years. Arterial distension was determined by using a multigate pulsed Doppler system, blood pressure (BP) was measured by a mercury sphygmomanometer. BP was 127 +/- 3/80 +/- 2 mm Hg at entry and 133 +/- 3/82 +/- 2 mm Hg after 2 years in the normotensive group, 146 +/- 4/90 +/- 3 mm Hg at entry and 145 +/- 3/87 +/- 2 mm Hg after 2 years in the hypertensive group (P < 0.01, normotensives vs hypertensives). The distensibility coefficient (DC) decreased significantly after 2 years in the hypertensive group (DC 18.3 +/- 1.3 10(-3)/kPa before, 15.1 +/- 1.2 10(-3)/kPa after 2 years, P < 0.05) whereas no significant change was observed in the normotensive group (DC 19.0 +/- 1.4 10(-3)/kPa before, DC 17.8 +/- 1.3 10(-3)/kPa after 2 years, NS). There was a significant correlation between the change of the distensibility coefficient after 2 years and mean arterial pressure (n = 48, r = 0.42, P < 0.01). The results show that the decrease of arterial distensibility after 2 years is accelerated in hypertensive renal transplant recipients despite effective anti-hypertensive treatment. Since BP levels were not different at entry into the study and after 2 years, differences in distending pressure along cannot explain the more pronounced decrease of arterial distensibility over time in hypertensive renal transplant recipients.  相似文献   

19.
The objective of our study was: (1) to compare the influence of moderate exercise on circulatory after-response in mildly hypertensive (n = 8) and normotensive male subjects (n = 9); (2) to examine the circulatory response to 3-min hyperoxic inactivation of arterial chemoreceptors at rest and during postexercise period in both groups. Hypertensive men (HTS) with a systolic blood pressure (SBP) 148 +/- 5 mm Hg, diastolic blood pressure (DBP) 92.4 +/- 4 mm Hg; and normotensive men (NTS), with a SBP 126 +/- 3 mm Hg, DBP 75.6 +/- 1.3 mm Hg, were submitted to 20-min of moderate exercise on a cycloergometer (up to the level of 55% of each subject's resting heart rate reserve). Finger arterial BP was recorded continuously with Finapres, impedance reography was used for recording stroke volume, cardiac output and arm blood flow. In HTS a significant decrease in SBP by 14.5 +/- 3.4 mm Hg, DBP by 8.9 +/- 1.9 mm Hg, total peripheral resistance (TPR) by 0.45 +/- 0.05 TPR u. (33.7 +/- 2.7%), and in arm vascular resistance (AVR) by 11.0 +/- 2.7 PRU u. (35.6 +/- 7%), was observed over a 60-min postexercise period. NTS exhibited insignificant changes in SBP, DBP, AVR except a significant decrease in TPR limited only to 20-min postexercise period. Hyperoxia decreased SBP, DBP and TPR in HTS. This effect was significantly attenuated during the postexercise period. Long-lasting antihypertensive effect of a single dynamic exercise in HTS suggests that moderate exercise may be applied as an effective physiological procedure to reduce elevated arterial BP in mild hypertension. We suggest also that the attenuation of the sympathoexcitatory arterial chemoreceptor reflex may contribute to a postexercise decrease in arterial BP and in TPR in mildly hypertensive subjects.  相似文献   

20.
OBJECTIVE: To investigate the roles of brain angiotensin II and C-type natriuretic peptide (CNP) in the hypertensive mechanism of deoxycorticosterone acetate (DOCA)-salt hypertension. METHODS: We injected 50 microg/kg CV-11 974, an angiotensin II type-1 receptors antagonist, 30 nmol/kg CNP-22, or the vehicle (artificial cerebrospinal fluid) into the cerebral ventricle or intravenously 5 min before the intracerebroventricular infusion of 1.5 mol/I NaCl solution for 30 min into either male normotensive Wistar rats or DOCA-salt hypertensive rats anesthetized with urethane, and their arterial pressures and heart rates were continuously recorded. Blood (2 ml) was collected at the end of the infusion for the measurement of plasma concentration of arginine vasopressin. We infused 10 or 50 microg/kg per day CV-11 974, 10 or 50 nmol/kg per day CNP-22, or the vehicle (1 microl/h) into the cerebral ventricles of DOCA-salt hypertensive rats for 7 days by using osmotic minipumps, and measured their systolic arterial pressures, pulse rates, and urinary excretions of vasopressin. RESULTS: Intracerebroventricular pre-administrations of CV-11 974 and of CNP-22 inhibited increases in mean arterial pressure, heart rate, and plasma vasopressin concentration induced by intracerebroventricular infusion of 1.5 mol/l NaCl into normotensive rats; increases in hemodynamics and plasma level of vasopressin induced by intracerebroventricular infusion of 1.5 mol/l NaCl were suppressed by intracerebroventricular pre-injections of CV-11 974, but not of CNP-22, into DOCA-salt hypertensive rats. Continuous intracerebroventricular infusions of 50 microg/kg per day CV-11 974 attenuated hypertension in DOCA-salt treated rats, accompanied by a reduction in urinary excretion of vasopressin. Continuous intracerebroventricular infusions of 50 nmol/kg per day CNP-22, however, affected neither hypertension nor urinary excretion of vasopressin in DOCA-salt hypertensive rats. CONCLUSION: Brain angiotensin II could play a role in the pressor mechanism of DOCA-salt hypertension by increasing release of vasopressin via type 1 receptors. That brain CNP has an inhibitory effect on release of vasopressin in acute experiments indicates that the impairment of this inhibitory effect of brain CNP on secretion of vasopressin could be involved in the pathogenesis of DOCA-salt hypertension in rats.  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