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.
Intraplatelet serotonin (5-HT), beta-thromboglobulin (beta-TG), and histamine content as well as platelet total thromboxane A2 (TXA2) synthesizing capacity were measured in 53 patients with chronic renal disease: nephrotic syndrome (n = 18); end-stage renal failure (ESRF; n = 13); continuous ambulatory peritoneal dialysis (CAPD; n = 9); hemodialysis (HD; n = 13). These indices of platelet function were correlated with plasma total cholesterol (TC), low-density lipoprotein cholesterol (LDL-C), high-density lipoprotein cholesterol (HDL-C), and triglyceride (TG) concentrations. When compared with controls, intraplatelet 5-HT was significantly reduced in all patient groups studied and beta-TG was diminished in all patient groups except CAPD. Total platelet TXA2 synthesizing capacity was increased in ESRF and HD groups. Intraplatelet histamine content was not altered in any of the patient groups studied. There was a significant inverse correlation between intraplatelet 5-HT content on the one hand and plasma TC, LDL-C, and TG on the other. The depletion of intraplatelet 5-HT and beta-TG and the increase in total TXA2 synthesizing capacity are consistent with platelet activation in chronic renal disease. The correlation between these indices of platelet activation and TC, LDL-C, HDL-C, and TG suggests that changes in the concentrations of these lipids may contribute to the activation of platelets in these conditions.  相似文献   

2.
The mechanisms involved in the hemostatic abnormality of uremic patients remain obscure. We have explored the response of normal and uremic platelets to surface activation at the ultrastructural level and analyzed changes in the composition of proteins associated with normal and uremic platelet cytoskeletons after stimulation with thrombin (0.01 and 0.1 U/ml). Cytoskeletons were obtained by extraction with Triton X-100, processed by sodium dodecylsulfate-polyacrylamide gel electrophoresis, and the presence of cytoskeletal proteins analyzed by densitometry. Under static conditions, uremic platelets spread with difficulty on formvar-coated grids. The percentage of platelets that spread fully on this polymer surface was statistically reduced compared with that of control platelets (11 +/- 1.4 vs. 21 +/- 1.6; P < 0.05). An impairment of cytoskeletal organization was observed in resting uremic platelets but abnormalities were more evident after thrombin activation. The incorporation of actin into the cytoskeletons of thrombin-stimulated uremic platelets was significantly reduced with respect to controls (6 +/- 3% vs. 29 +/- 5%; P < 0.01 after 0.01 U/ml and 28 +/- 9% vs. 59 +/- 10%; P < 0.05 after 0.1 U/ml). Decreased associations of actin-binding protein (P < 0.01), alpha-actinin (P < 0.05), and tropomyosin (P < 0.05) with the cytoskeletons of uremic platelets were also noted. No difference was observed for the incorporation of myosin into the cytoskeletons of activated uremic platelets. These results suggest functional and biochemical alterations of the platelet cytoskeleton in uremia, which may contribute to the impairment of platelet function observed in uremic patients.  相似文献   

3.
AIM: To investigate the morphology and function of platelets in nephropathic cystinosis (NC).METHODS: Seven patients (mean age, 6.5 years; SD, 20 months) with NC were investigated. Their platelets were examined by transmission electron microscopy (TEM) and the characteristics of the dense granules (DGs) were determined by mepacrine labelling and the uranaffin reaction. Bleeding time, turbidometric aggregation, and luminescence aggregation were studied and intraplatelet cystine was measured. RESULTS: Increased intraplatelet cystine, primary and secondary aggregation defects, and the absence of ATP release were demonstrated. TEM revealed DGs of various shapes and sizes and lamellary or amorphous cytoplasmic inclusions. Viscous material had been released into the vacuolar spaces and enlarged open canalicular system. Mepacrine labelling revealed that the numbers of DGs/platelet were comparable between the patients and the controls (mean, 2.9 (SD, 0.22) v 3.32 (0.18); p = 0.34). The uranaffin reaction revealed that the numbers of type 1, 3, and 4 DGs were comparable between the patients and the controls, but that there were fewer type 2 DGs in the patients (mean, 8.5 (SD, 1.95) v 17.22 (1.58); p = 0.01). TEM for platelet aggregation revealed a lack of induction and/or defective execution and/or delayed transmission. The patients' intraplatelet cystine concentrations were higher than the controls (mean, 1.56 (SD, 0.84) v 0.08 (0.01) nmol/mg protein; p = 0.009). CONCLUSIONS: This is the first report to demonstrate raised intraplatelet cystine, abnormal platelet ultrastructural findings, and defective aggregation in NC.  相似文献   

4.
We analyzed the RNA in platelets by fluorescence flow cytometry after staining with thiazole orange(TO) in whole blood samples from hematologically normal subjects and patients with thrombocytopenia. The percentage of TO-positive platelets and their mean fluorescence channel number in 32 control subjects were 6.2 +/- 2.5% (mean +/- SD) and 6.9 +/- 0.7, respectively. In 11 patients with idiopathic thrombocytopenic purpura, the percentage of fluorescently labeled platelets was significantly elevated (p less than 0.05) to 21.5 +/- 14.3%. By contrast, the proportion of positively stained platelets in 14 patients with thrombocytopenia due to impaired platelet production did not significantly differ from that of the controls, whereas the absolute counts of TO-positive platelets were significantly lowered (p less than 0.05). In both patient groups, the mean fluorescence channel numbers of TO-positive platelets were significantly elevated to 16.1 +/- 16.8 and 6.9 +/- 0.7, respectively (p less than 0.05, 0.005). We conclude that flow cytometric analysis of platelets after staining with TO provides information on the thrombopoietic activity in thrombocytopenic disorders. The main advantages of this method for clinical use are its simplicity and the rapidity.  相似文献   

5.
OBJECTIVE AND DESIGN: P-selectin, a membrane glycoprotein which is expressed on activated platelets and endothelial cells, plays a crucial role in the inflammatory response. The main action is adhesion of leukocytes, facilitation of diapedesis and induction of cytokine production from monocytes (MCP-1 and IL-8), mediated via RANTES released from activated platelets. An abnormal platelet activity has been reported in patients with ulcerative colitis (UC) and Crohn's disease (CD), jointly referred to as inflammatory bowel disease (IBD), which could have an aggravating influence on the inflammatory response. In addition, an up-regulation of platelet IL-8 receptors among patients with IBD has been reported. To reveal a presumptuous platelet dysfunction we analysed the expression of platelet surface P-selectin at resting state and after stimulation with thrombin, collagen, epinephrine and interleukin 8 (IL-8), and plasma levels of soluble P-selectin, neuropeptide Y (NPY) and RANTES in patients with IBD. SUBJECTS: Blood from twelve healthy subjects (control group) and twenty-one patients with IBD who had not taken any anti-platelet drugs or steroids were analysed. METHODS: Patients were sub-grouped according to disease entity, disease activity and 5ASA medication. Surface P-selectin expression on isolated human platelets and plasma P-selectin, NPY and RANTES were analysed with ELISA. All values are presented as mean +/- standard error of the mean (SEM). Mann-Whitney U test and Wilcoxon matched rank test were used for statistical analyses. RESULTS: Patients with IBD in remission (n = 9) had higher basal P-selectin expression, 0.38+/-0.04, compared to the control group (n = 12), 0.22+/-0.03,p < 0.01. UC patients (n = 16) showed down-regulation of P-selectin expression after stimulation with IL-8, 0.26+/-0.03 to 0.22+/-0.02, p < 0.05. No significant differences could be observed concerning soluble P-selectin and NPY in plasma. Patients with 5ASA (n = 12) had lower levels of plasma RANTES, 2.39+/-0.06 microg/l, compared to the control group (n = 12), 3.29+/-0.19 microg/l, p < 0.01, and patients without 5ASA (n = 9), 2.90+/-0.17 microg/l, p < 0.05. CONCLUSIONS: Patients with IBD in remission have higher basal platelet surface P-selectin expression. An exaggerated platelet activity with increased expression of platelet P-selectin and release of inflammatory mediators such as RANTES, which is chemotactic and induce chemokine production, could have a reinforcing and aggravating influence on the inflammatory response and increase the susceptibility to IBD. In addition IL-8 has a down-regulating effect on platelet surface P-selectin expression and 5ASA medication seems to lower plasma RANTES. If 5ASA is responsible for lowering the concentration of RANTES this could be one of the beneficial outcomes of 5ASA medication.  相似文献   

6.
A factor(s) from human platelets enhances IgE-mediated histamine release from human basophils and mast cells. This effect is directly related to the platelet number; at physiological platelet/leukocyte ratios (40:1), the enhancement was 66 +/- 11%. Platelet stimulation by thrombin more than doubled the enhancement, to 172 +/- 10% at 40:1. Mast cell release was also enhanced by platelets although the magnitude was more limited (86 +/- 13% at 40:1 with thrombin). Direct basophil/platelet contact was unnecessary in that platelet supernatants were fully active; a direct platelet factor/basophil interaction is suggested, however, by the fact that basophils purified 100-fold with respect to other leukocytes were enhanced by the platelet factors. The appearance of platelet-enhancing activity is associated with the release of an alpha-granule marker (PF4) rather than with products of arachidonic acid metabolism (thromboxane B2). The platelet factor(s) responsible for these effects are not dialyzable, are heat stable and do not appear to be identical to PF4 or platelet-derived growth factor (PDGF). Since anti-IgE-stimulated basophils cause PF4 release and this correlates with the release of enhancing factor, we suggest that a pro-inflammatory feed forward relationship exists. Together with our previous data showing that platelets are activated in vivo during antigen challenge of allergic asthmatic subjects, these results suggest that platelets may be important in modulating IgE-mediated allergic reactions in man.  相似文献   

7.
In order to study the ionic efflux or granule release from human platelets following pulse exposure to various stimuli, a method for continuous perfusion of platelets was developed. The method was applied to compare the effects of membrane depolarization and thrombin stimulation on the release of 86Rb and [3H]5-HT. Washed and preloaded human platelets were placed on a membrane filter in a temperature controlled polypropylene chamber, and subsequently perfused with buffer. After an initial washout period the efflux of 86Rb or [3H]5-HT reached steady, low levels. K+ induced concentration dependent increases in 86Rb efflux, corresponding to a depolarization of the membrane potential, whereas the efflux of [3H]5-HT was unaltered. Thrombin induced concentration dependent increases in the efflux of both 86Rb and [3H]5-HT. Pretreatment with K+ 12 or 30 mM did not alter the [3H]5-HT efflux induced by thrombin 0.1 U ml-1. Scanning electron micrographs of platelets on the filter showed that the unstimulated platelets had regular shape, whereas after addition of thrombin there was formation of pseudopods and minor aggregates. The effect of potassium-induced membrane depolarization on platelet aggregation was also studied. High concentration of K+ did not induce aggregation or shape change during 2 or 10 minutes of incubation. K+ had little or no effect on aggregation induced by ADP 2 microM or thrombin 0.4 U ml-1. The results from release experiments and aggregation tests argue against an immediate coupling between membrane potential and platelet reactivity.  相似文献   

8.
We assessed the role of platelet activation markers (PMPs, Annexin V and CD62P on activated platelets), cytokines (IL-1 beta, IL-4, IL-6, IFN- gamma, GM-CSF, and TNF alpha ), and soluble factors (sIL-2R, TM, sHLA-1, beta(2) -m, sVCAM-1, sPECAM-1, sP-selectin and sE-selectin) in vascular damage related to SLE. There were differences in the levels of PMPs and platelet activation markers between the SLE patients and controls (PMPs: 493+/-82 vs. 328+/-36, p<0.05; plt-CD62P; 8.5%+/-1.2 % vs. 4.6%+/-0.7 %, p<0.05; plt-Annexin V: 11.3%+/-2.1 % vs. 4.9%+/-0.6 %, p<0.01). There were no differences in the levels of IFN- gamma between the groups. However, the levels of IL-1 beta, IL-4, IL-6, GM-CSF, TNF alpha, and soluble factors were higher in the SLE patients than in the controls. The levels of IL-4, IL-6, beta2 -m, sIL-2R, sVCAM-1, sP-selectin, and sE-selectin in SLE patients with elevated sTM levels were higher than those in the SLE patients without elevated sTM levels. On the other hand, elevations of sIL-2R, sVCAM-1, and sP-selectin were not found in patients with Beh?et disease or rheumatoid arthritis. The levels of platelet CD62P, platelet annexin V, and PMP were significantly elevated in high-sTM patients. These findings suggest the possibility that activated platelets and cytokines participate in the pathogenesis of SLE in patients with elevated sTM levels.  相似文献   

9.
Data from animal studies indicate that platelets play a key role in tumor dissemination and metastasis. We therefore hypothesized that metastastic cancer patients may display a specific platelet phenotype. Percentage of activated, p-selectin positive platelets as well as platelet contents (i.e., plasma and platelet count-corrected serum levels of VEGF-A, CXCL12, CXCL4, and thrombospondin-1) were analyzed in 43 patients with newly diagnosed metastatic disease prior to treatment. Tumor patients had increased platelet counts and significantly elevated percentages of activated platelets. Moreover, the platelet content of VEGF-A in cancer patients was significantly increased compared to healthy controls, while thrombospondin-1, CXCL12 and CXCL4 were significantly decreased. Our data contain several unexpected results: firstly, CXCL12 was found in minute quantities in the serum as compared with murine studies. Secondly, CXCL4, which was found by mass spectrometry to be the single massively upregulated intraplatelet chemokine in mice after tumor xenotransplantation, was decreased in tumor patient platelets. While increased contents of VEGF-A have been attributed to platelet scavenger activity, the differential decrease of specific platelet contents may be due to differential secretion or altered megakaryopoiesis in metastatic cancer patients.  相似文献   

10.
The characteristics of spontaneous platelet aggregation (SPA) in a hereditary giant platelet syndrome (Montreal platelet syndrome, MPS) are examined. SPA was quantitated by microscopy from the decrease in single platelets in platelet-rich plasma (PRP). In contrast to normal donors, a significant proportion (20-50%) of platelets in MPS whole blood and PRP occurred in microaggregates typically containing 2-6 disk-shaped platelets. Stirring MPS-PRP at 1000 rpm for 10 minutes further increased the fraction of platelets in aggregates by 10-170%, the percentage increase not being correlated to the donor's platelet count (5000-220,000 microliters-1). Normal platelets resuspended in MPS platelet-poor plasma (PPP) did not undergo SPA, whereas MPS platelets resuspended in normal PPP or Ca2+-free, fibrinogen-free Tyrode's continued to show SPA. The increase in SPA could be inhibited by 10 microM prostaglandin (PG) E1, 150 mM ASA or glutaraldehyde or formaldehyde fixation; however, it was not inhibited by 10 nM PGI2 and was only partially inhibited by 1 microM 2-chloroadenosine and 1-10 units/ml apyrase. SPA in Acid-citrate-dextrose-PRP was much less than in PRP; however, SPA reoccurred on returning the platelets to platelet-free plasma or Tyrode's. Platelet aggregation (PA) could be increased over that due to SPA alone by the addition of adenosine diphosphate, adrenaline, collagen, ionophore A-23187, arachidonic acid and ristocetin, with results suggesting that the response to these agents is normal. The ristocetin-induced increase in PA was completely blocked by an IgG specific for Bernard-Soulier syndrome. In contrast, MPS platelets had a reduced sensitivity to thrombin, which appeared to be more pronounced at low platelet counts. There was no correlation between the thrombin insensitivity and the extent of SPA. Total adenosine triphosphate (ATP) and thrombin-induced release of ATP and platelet factor 4 appeared normal for MPS platelets. The ultrastructural features of MPS platelets were within normal limits except for an increased frequency of giant granules. SPA was observed for 5/5 MPS donors, but only one of three MPS donors' platelets evaluated for glycoprotein I and sialic acid content showed any measurable reduction as compared with normal controls. The above observations point to the existence of an as yet undetermined anomaly of MPS plasma membrane related to a fibrinogen and Ca2+ independent form of platelet aggregation.  相似文献   

11.
BACKGROUND. Isolated thrombocytopenia accompanied by increased amounts of platelet-associated antibody is a common manifestation of human immunodeficiency virus (HIV) infection, and the thrombocytopenia often improves with zidovudine. It is not clear whether the mechanism of HIV-related thrombocytopenia primarily involves autoimmune destruction of platelets or reduced platelet production by megakaryocytes. METHODS. We studied the survival of 111In-labeled autologous platelets and performed platelet imaging in 24 men with isolated HIV-related thrombocytopenia (16 who received no treatment and 8 who received zidovudine). We also studied 20 HIV-infected men with normal platelet counts (10 who received no treatment and 10 who received zidovudine) and studied 12 healthy seronegative men as controls. RESULTS. Mean (+/- SD) platelet survival was significantly decreased in both the untreated and the zidovudine-treated patients with HIV-related thrombocytopenia (to 92 +/- 33 and 129 +/- 44 hours, respectively; both P < 0.001), as compared with the normal controls (198 +/- 15 hours). Mean platelet survival was also significantly decreased in the HIV-infected patients with normal platelet counts (untreated, 162 +/- 23 hours, P < 0.01; zidovudine-treated, 166 +/- 35 hours, P < 0.05). Imaging studies, however, revealed no evidence of increased clearance of autologous platelets in the liver or spleen in any of these groups. Mean platelet production was significantly depressed in the untreated patients with thrombocytopenia (23,000 +/- 11,000 platelets per cubic millimeter per day, P < 0.001) as compared with the healthy controls (45,000 +/- 6,000 per cubic millimeter per day). Mean platelet production was significantly increased, however, in the men treated with zidovudine, both in those with thrombocytopenia (60,000 +/- 31,000 platelets per cubic millimeter per day, P < 0.01 vs. controls) and in those without thrombocytopenia (68,000 +/- 22,000 per cubic millimeter per day, P < 0.01). CONCLUSIONS. Although there was a moderate reduction in platelet survival in HIV-infected persons, these patients, regardless of platelet counts, also had decreased production of platelets, possibly due to viral infection of the megakaryocytes. Zidovudine appears to improve platelet production.  相似文献   

12.
Platelet serotonin uptake was measured in 45 patients (12 males, 33 females) with panic attacks and 21 controls (9 males, 12 females). Higher Vmax values were obtained in the patient group than in the controls (100.2 +/- 11.5 vs. 34.9 +/- 3.8 pmol/10(8) platelets/min; P less than 0.0005) while the affinity constant Km was not significantly different (2.28 +/- 0.3 vs. 1.71 +/- 0.18 microM). A value of Vmax in excess of 60 pmol/10(8) platelets/min was observed in 60% of patients and in only 5% of controls. The results point to a specific abnormality of platelet serotonin uptake in patients with panic attacks.  相似文献   

13.
Metabolic and functional studies of the amyloid precursor protein (APP) in platelets have advanced our understanding of Alzheimer's disease (AD). Here we report that human platelets contain Abeta peptides, process and secrete them constitutively. Platelets generate formerly unkown Abeta-species by differential processing of APP. Release of Abeta peptides were also increased by platelet activation with thrombin, indicating the existence of a regulated exocytotic pathway. We showed that Abeta-levels, Abeta-processing patterns and Abeta-release kinetics were regulated by thrombin. In controls, release of Abeta peptide species (Abeta 1-40/42 and 1-37/38/39/) continued for more than 4 h, while thrombin activated cells ceased secretion after 1 h at large. Treatment of platelets with prostaglandine 2 slowed this process down. Intracellular Abeta peptide concentrations decreased steadily until no peptides could be detected after 20 h (control) or after 4 h (thrombin) in cultured platelets.  相似文献   

14.
BACKGROUND. Constriction of small pulmonary arteries and arterioles and focal vascular injury are features of pulmonary hypertension. Because thromboxane A2 is both a vasoconstrictor and a potent stimulus for platelet aggregation, it may be an important mediator of pulmonary hypertension. Its effects are antagonized by prostacyclin, which is released by vascular endothelial cells. We tested the hypothesis that there may be an imbalance between the release of thromboxane A2 and prostacyclin in pulmonary hypertension, reflecting platelet activation and an abnormal response of the pulmonary vascular endothelium. METHODS. We used radioimmunoassays to measure the 24-hour urinary excretion of two stable metabolites of thromboxane A2 and a metabolite of prostacyclin in 20 patients with primary pulmonary hypertension, 14 with secondary pulmonary hypertension, 9 with severe chronic obstructive pulmonary disease (COPD) but no clinical evidence of pulmonary hypertension, and 23 normal controls. RESULTS. The 24-hour excretion of 11-dehydro-thromboxane B2 (a stable metabolite of thromboxane A2) was increased in patients with primary pulmonary hypertension and patients with secondary pulmonary hypertension, as compared with normal controls (3224 +/- 482, 5392 +/- 1640, and 1145 +/- 221 pg per milligram of creatinine, respectively; P less than 0.05), whereas the 24-hour excretion of 2,3-dinor-6-keto-prostaglandin F1 alpha (a stable metabolite of prostacyclin) was decreased (369 +/- 106, 304 +/- 76, and 644 +/- 124 pg per milligram of creatinine, respectively; P less than 0.05). The rate of excretion of all metabolites in the patients with COPD but no clinical evidence of pulmonary hypertension was similar to that in the normal controls. CONCLUSIONS. An increase in the release of the vasoconstrictor thromboxane A2, suggesting the activation of platelets, occurs in both the primary and secondary forms of pulmonary hypertension. By contrast, the release of prostacyclin is depressed in these patients. Whether the imbalance in the release of these mediators is a cause or a result of pulmonary hypertension is unknown, but it may play a part in the development and maintenance of both forms of the disorder.  相似文献   

15.
The serotonin content of platelets, serum and plasma from rats of various ages was examined. In male rats, platelet serotonin content, which was about 0.65 nmol/10(8) platelets at young age (6-7 months), increased slightly at middle age (12-14 months) but decreased markedly at old age (25-26 months). Significant difference (P less than 0.05) was observed between young and old rats, and between middle-aged and old rats. In female rats, on the other hand, no age-related change in the platelet serotonin content was found. In both sexes, the serotonin content of rat sera changed with age in the same pattern as that of the platelets. No plasma serotonin was detected in rats of either sex and at any ages examined. Serotonin release from rat platelets was also studied using collagen and thrombin as stimulants. In males, the responsiveness of platelets to these two stimulants showed almost the same age-dependent changes. It was lower in middle-aged rats than in young rats but increased greatly in old rats. Significant difference (P less than 0.05) was observed between middle-aged and old rats. In females, collagen and thrombin had the opposite effect on the sensitivity of the platelets as age increased. The amount of serotonin released in response to collagen was low until middle age but increased markedly at old age, while the content of serotonin released by thrombin remained high until middle age and decreased greatly at old age. These results imply that age-related changes in the serotonin release reaction in rat platelets differed according to the stimulants used.  相似文献   

16.
Lithium carbonate treatment for 2-3 weeks produced a significant decrease in the maximum velocity (Vmax) of serotonin (5-HT) uptake, a measure of the number of 5-HT uptake sites in blood platelets from bipolar patients. The decrease was more pronounced in bipolar manic patients than bipolar depressed patients. There was no significant affect on the affinity for 5-HT (Km) of the uptake sites in the platelets of manic or depressed bipolar patients although Km did decrease (indicating increased affinity) in the majority of subjects from both groups of patients. However, lithium treatment of at least 1 year duration was associated with significant increases in Vmax without affecting Km. Lithium in vitro, at concentrations up to 1 mM, had no effect on the Km or Vmax of 5-HT uptake in blood platelets of normal controls. The possible mechanisms of the inhibitory and stimulatory effect of lithium carbonate treatment on platelet 5-HT uptake are discussed.  相似文献   

17.
Platelet hyperactivity is likely to contribute to the progression of atherogenesis and organized thrombus formation on vascular surfaces. The purpose of this study was to examine the effect of hypercholesterolemia on the cholesterol content of platelets, on platelet responsiveness and other platelet indices using platelets from 5 groups of age-matched subjects (n = 30 each), which includes healthy controls. All groups except controls had a high plasma lipid profile. While subjects in group I had only hyperlipidemia, those in groups II and III had hyperlipidemia in conjunction with diabetes mellitus and hypertension, respectively. The fourth group consisted of patients with confirmed coronary artery disease (CAD). The parameters studied include packed cell volume of platelets (platelet crit), platelet distribution width (PDW), platelet cholesterol and platelet aggregation in response to adenosine diphosphate and collagen. All the patient groups showed increased platelet aggregation (p < 0.05) and low platelet crit compared with controls (p < 0.05). In addition, platelet cholesterol was increased in patients with coronary disease, hyperlipidemia and diabetes mellitus (p < 0.05) but not in patients with hypertension (p > 0.05); PDW was high only in CAD (p < 0.05). A higher PDW indicated a prothrombotic tendency in CAD patients. Our data suggest that hyperlipidemia increases the lipid content in platelets and enhances their reactivity. Hyperactive platelets with increased platelet cholesterol may contribute to accelerated atherogenesis associated with CAD.  相似文献   

18.
Platelet characteristics were assessed in 15 patients with essential thrombocytosis (ET), 89 patients with reactive thrombocytosis (RT), and 23 normal controls. A platelet volume distribution width (PDW) greater than or equal to 10.5 was found in 50%, 21%, and 14% of the three groups, respectively (P = 0.01 between patients with ET and patients with RT; P = 0.02 between patients with RT and controls), reflecting an excess of extreme values at both ends of the distribution. Compared with controls, the increase in platelet number in patients with RT was about twofold throughout the platelet volume range, whereas ET was characterized by a fivefold increase in small platelets less than 7.5 fL and threefold increase in larger size platelets. Mean platelet volume (MPV) was significantly lower in patients with ET versus patients with RT and in patients with RT versus controls (mean +/- SD 7.5 +/- 1.2 vs. 8.8 +/- 0.1 and 10.2 +/- 1.8 fL, respectively, P less than 0.01). Rate of in vitro platelet aggregation greater than or equal to 50% was significantly lower in patients with ET versus patients with RT and in patients with RT versus controls (0%, 23%, and 45%, respectively, P less than 0.01). Aggregation rate was positively correlated with MPV (r = 0.54; P less than 0.0001). Aggregation rate in patients with ET was significantly lower (P = 0.01) than expected from their reduced MPV alone. Despite these group differences, the overlap of individual platelet characteristics between the three groups precludes their usefulness for diagnostic purposes.  相似文献   

19.
We performed this study to investigate whether alpha- and beta-adrenergic agonists are able to regulate intracellular free magnesium concentrations [Mg2+]i in platelets from healthy and obese individuals. Twenty-six informed-consent men (14 healthy and 12 obese) were enrolled in the study. We measured fasting plasma glucose, insulin, epinephrine and norepinephrine. Platelet [Mg2+]i at the baseline and after stimulation with clonidine or isoproterenol was measured by fluorescent probe mag-fura-2. In platelets from healthy subjects, alpha-adrenergic stimulation by clonidine led to a dose-dependent decrease in [Mg2+]i (basal: 245 +/- 39 microM; clonidine 5 pg/mL: 109 +/- 27 microM, p < 0.05; clonidine 10 pg/mL: 77 +/- 26 microM, p < 0.01), while no significant change in platelet [Mg 2+]i was detected in obese men. Furthermore, the co-incubation with clonidine (10 pg/mL) and yohimbine (50-100 pg/mL) completely abated the effect of clonidine on [Mg2+]i in platelets from healthy individuals. Analysis of the time course for platelet magnesium showed that the intracellular magnesium loss induced by clonidine (10 pg/mL) was time-dependent. Conversely, the beta-adrenergic agonist isoproterenol was able to produce a significant rise in [Mg2+]i in platelets from healthy individuals (basal: 234 +/- 40 microM; isoproterenol 2.5 pg/mL: 594 +/- 44 microM, p < 0.05: isoproterenol 5 pg/mL: 681 +/- 56 microM, p < 0.01), while no such finding was detectable in platelets from obese patients. When platelets from healthy subjects were co-stimulated with isoproterenol (5 pg/mL) and propranolol (10-20 pg/mL), the ionophoric effect of the beta-adrenergic agonist was completely reverted. The time course of isoproterenol (5 pg/mL) effect on platelet [Mg2+]i showed that the ionophoric effect of isoproterenol was time-dependent. In conclusion, (1) the stimulation of alpha-adrenergic receptor by clonidine is able to induce a significant dose- and time-dependent fall in platelet [Mg2+]i; (2) the stimulation of beta-adrenoceptors by isoproterenol lead to a signifcant time- and dose-dependent rise in platelet [Mg2+]; (3) the ionic effect of alpha- and beta-adrenergic stimulation is not detectable in obese subjects, in whom is probably present a reduced sensitivity to the ionic effect of adrenergic agonists.  相似文献   

20.
A quantitative study of various aspects of platelet function was carried out in eight patients with typical hairy-cell leukaemia (HCL). In at least two patients platelet aggregation was convincingly reduced to more than one aggregating agent (ADP, adrenaline, collagen, thrombin, and ristocetin). Granular storage capacity for {(14)C} 5-HT was reduced in five of the six patients tested. The two patients with definitely abnormal aggregation had the greatest reduction in granular storage pool and the longest bleeding times of those tested but, like the other patients, they did not have a clinical haemostatic defect. It was concluded that a granular storage pool defect (SPD) was at least partly responsible for aggregation abnormalities in HCL since the platelet release reaction in response to thrombin appeared to be normal. All our patients ran a chronic course uncomplicated by any of the factors known to predispose to a platelet SPD acquired in the circulation. Although in the one patient tested before and after splenectomy there was some improvement in platelet aggregation after operation, there was no clear general relationship between defective platelet function and either previous splenectomy or platelet count. Since a direct involvement of the megakaryocytic series in the underlying cell proliferation of HCL seems unlikely, it is concluded that the platelet defect can most reasonably be attributed to the production of abnormal platelets as a result of marrow fibrosis and/or infiltration by hairy cells.  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