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.
Copper deficiency in rats produces a hypercholesterolemia with a marked increase in HDL fraction. This study investigated changes in the plasma distribution and composition of HDL subclasses as affected by copper deficiency. Plasma HDL were separated into the following three subclasses by heparin-affinity chromatography: HDL containing no apo E but high in apo A-I (HDL-E0); HDL with an intermediate level of apo E (HDL-E1); and HDL highly enriched in apo E but low in apo A-I (HDL-E2). The compositional analysis showed that the hypercholesterolemia observed in copper-deficient rats was due specifically to an increase in plasma cholesterol carried by HDL-E0. Copper deficiency did not alter the percent distribution of apo A-I in HDL-E0, but lowered the apo A-I content in HDL-E1 and HDL-E2, with an increase in apo E in these subclasses. The total plasma concentration of apo A-I was, however, significantly elevated in Cu-deficient rats, which was attributable to an increase in the total number of circulating HDL particles. No difference was noted between Cu-deficient and control groups in the distribution of free cholesterol or the ratio of free cholesterol to esterified cholesterol in any of the HDL subclasses. The present results and earlier observations suggest that copper deficiency may produce a defect in the plasma clearance or tissue uptake of the HDL subclass high in apo A-I but devoid of apo E (HDL-E0), which may be mediated by the specific apo A-I receptor or non-endocytotic transfer of HDL-E0 cholesterol to the liver. Such metabolic defects may partly explain the simultaneous increases in both plasma HDL cholesterol and apo A-I and altered cholesterol homeostasis observed in copper deficiency.  相似文献   

2.
Apolipoprotein E-(apoE-) rich high-density lipoprotein (HDL) of normal subjects showed marked inhibitory effects on platelet aggregation and ATP release as compared with apoE-poor HDL, suggesting that apoE has inhibitory effects on platelet function (Desai et al. J. Lipid Res. 30:831, 1989; Higashihara et al. FEBS Lett. 282:82, 1991). A patient with apoE deficiency showed evidence of decreased platelet aggregability in platelet-rich plasma, but normal aggregability in washed platelets. Both patient's plasma and HDL fraction inhibited platelet aggregation of normal subjects. Patient's HDL reconstituted with recombinant apoE showed further inhibitory effects on platelet function. These results suggest that apoE is a potent, but not unique, inhibitory factor for HDL.  相似文献   

3.
The aggregation of gel-filtered rabbit platelets by 50 microM ADP was inhibited by a labile factor produced by suspensions of cultured bovine pulmonary artery endothelial cells. Inhibition of aggregation occurred when indomethacin-treated endothelial cells (6.10(5) per ml) and rabbit platelets (3.2.10(8) per ml) were incubated together. This anti-aggregatory activity was characterized as similar to endothelium-derived relaxing factor (EDRF) in that it was unstable at neutral pH and by its inhibition by hemoglobin. The activity was unaffected by treatment of the platelets and endothelial cells with the cyclooxygenase inhibitor, indomethacin, and by the lipoxygenase inhibitor, BW755c. In association with the anti-aggregatory activity, the levels of cyclic GMP were elevated 4-fold. The effect of the EDRF-like product on the levels of cyclic nucleotides was mimicked by treatment of platelets with sodium nitroprusside, an activator of soluble guanylate cyclase; sodium nitroprusside had no measurable effect on the levels of cyclic nucleotides of endothelial cells. We conclude that a factor with the properties of EDRF inhibits platelet aggregation, and that this is associated with an activation of guanylate cyclase as in smooth muscle. Thus, EDRF may exert an inhibitory effect on platelets in a manner analogous to its actions on vascular smooth muscle.  相似文献   

4.
The binding of low-density lipoproteins (LDL) as well as LDL modified by cyclohexanedione (CHD-LDL) to gel-filtered platelets (GFP) and its effect on platelet function were studied in normal and in homozygous familial hypercholesterolaemic (HFH) subjects. Only normal-derived LDL could significantly compete with normal 125I-labelled LDL for binding to normal platelets. When GFP from normal subjects were incubated with normal LDL at concentrations of 25-200 micrograms of protein/ml, platelet aggregation in the presence of thrombin (0.5 i.u./ml) was increased by 65-186%. CHD-LDL, at similar concentrations, caused the opposite effect and decreased platelet aggregation by 26-47%. Both LDL and CHD-LDL (100 micrograms/ml) from HFH patients, when incubated with normal GFP, caused a significant reduction in platelet aggregation (33 and 50% respectively). When HFH-derived platelets were used, both patient LDL and CHD-LDL (but not the normal lipoprotein) could markedly compete with the patient 125I-labelled LDL for binding to the platelets. LDL and CHD-LDL (100 micrograms/ml) from normal subjects decreased aggregation of HFH-platelets by 52 and 85% respectively, while corresponding concentrations of LDL derived from HFH subjects (HFH-LDL) and CHD-LDL derived from HFH subjects (CHD-HFH-LDL) increased platelet aggregation by 165 and 65% respectively. The present results support the following conclusions: platelet activation by LDL in normal subjects is through the arginine-rich apoprotein-binding site; more than one binding site for LDL exists on platelets; under certain circumstances, LDL binding can cause a reduction in platelet activity; specificity for LDL binding to the platelets resides in different regions of the lipoprotein in HFH and in normal subjects. We have thus suggested a model for LDL-platelet interaction in normal and in HFH subjects.  相似文献   

5.
A new apolipoprotein, termed apolipoprotein J (apoJ), was purified from human plasma by immunoaffinity chromatography. ApoJ is a glycoprotein consisting of disulfide-linked subunits of 34-36 and 36-39 kDa. Each subunit is glycosylated and has a pI range of 4.9-5.4. ApoJ exists in the plasma associated with high density lipoproteins (HDL) and specifically with subclasses of HDL which also contain apoAI and cholesteryl ester transfer protein activity. Immunoaffinity purified apoJ-HDL subclasses have apparent molecular masses of 80, 160, 240, 340, and 520 kDa, as determined by gradient gel electrophoresis. By negative staining electron microscopy, apoJ-HDL range in diameter from 5 to 16 nm. Fractionation of plasma by vertical gradient density centrifugation revealed apoJ-HDL in HDL2 (d 1.063-1.125 g/ml) with the majority overlapping HDL3 (d 1.125-1.21 g/ml) and very high density lipoprotein (d 1.21-1.25 g/ml). The bimodal density distribution of apoJ-HDL suggests that these subclasses have a unique metabolic relationship and may play a role in the transport of cholesterol from peripheral tissues to the liver.  相似文献   

6.
The concentration of cholesterol, apolipoproteins A-I, B, and E has been determined in lymphedema fluid from nine patients with chronic primary lymphedema. The concentrations were: 38.14 +/- 21.06 mg/dl for cholesterol, 15.6 +/- 6.17 mg/dl for apolipoprotein A-I, 7.5 +/- 2.8 mg/dl for apolipoprotein B, and 1.87 +/- 0.50 mg/dl for apolipoprotein E. These values represent 23%, 12%, 6%, and 38% of plasma concentrations, respectively. The ratio of esterified to unesterified cholesterol in lymphedema fluid was 1.46 +/- 0.45. Lipoproteins of lymphedema fluid were fractionated according to particle size by gradient gel electrophoresis and by exclusion chromatography. Gradient gel electrophoresis showed that a majority of high density lipoproteins (HDL) of lymphedema fluid were larger than ferritin (mol wt 440,000) and smaller than low density lipoproteins (LDL); several discrete subpopulations could be seen with the large HDL region. Fractionation by exclusion chromatography showed that more than 25% of apolipoprotein A-I and all of apolipoprotein E in lymphedema fluid was associated with particles larger than plasma HDL2. Apolipoprotein A-I also eluted in fractions that contained particles the size of or smaller than albumin. Isolation of lipoproteins by sequential ultracentrifugation showed that less than 25% of lymphedema fluid cholesterol was associated with apolipoprotein B. The majority of apolipoprotein A-containing lipoproteins of lymphedema fluid were less dense than those in plasma. Ultracentrifugally separated fractions of lipoproteins were examined by electron microscopy. The fraction d less than 1.019 g/ml contained little material, while fraction d 1.019-1.063 g/ml contained two types of particles: round particles 17-26 nm in diameter and square-packing particles 13-17 nm on a side. Fractions d 1.063-1.085 g/ml had extensive arrays of square-packing particles 13-14 nm in size. Fractions d 1.085-1.11 g/ml and fractions d 1.11-1.21 g/ml contained round HDL, 12-13 nm diameter and 10 nm diameter, respectively. Discoidal particles were observed infrequently.  相似文献   

7.
Concentrations in serum were determined for 18 fatty acids (FAs) and 21 lipoprotein main and subclasses by chromatographic analyses and the average size was calculated for very low density (VLDL), low density (LDL) and high density (HDL) particles. 283 ethnic Norwegian children and adults from the rural Fjord region of Western Norway were compared with the objectives to reveal patterns and gender differences during the development from prepuberty to adulthood and during aging in adults. Both genders showed a large increase in eicosapentaenoic acid (EPA) and docosahexaenoic acid (DHA) from child to adult. Males, but not females, show a significant increase in most C16–C18 FAs from prepuberty to adulthood. These changes in males correlate to a pattern of increased concentrations of triglycerides, VLDL and LDL particles, especially the atherogenic subclasses of small and very small LDL particles. Furthermore, concentrations of medium, large and very large HDL particles decrease, while concentration of very small HDL particles increase leading to reduced average size of HDL particles. Females only showed significant increase in concentrations of small and very small LDL particles, very small HDL particles and apolipoprotein B. While EPA and DHA continued to increase during aging in women, no validated model for connecting age to FA profile was obtained for men. Women showed significant increase in concentrations of all subclasses of LDL particles during aging, while men exhibited a more complex pattern with increase also in apolipoprotein A1 and HDL particles.  相似文献   

8.
Stimulated human blood platelets release thrombospondin, an alpha-granule glycoprotein of 450 kDa. The aim of this work was to characterize an anti-thrombospondin monoclonal antibody (P8) in order to study the role of thrombospondin in platelet functions. The presence of thrombospondin receptor sites on resting and thrombin-stimulated platelets of three Glanzmann's thrombasthenia patients and normal donors was investigated using the P8 monoclonal antibody. Monoclonal antibody P8 was extensively characterized using ELISA, immunoprecipitation, immunoadsorbent affinity chromatography combined with tryptic peptide map analysis and crossed immunoelectrophoretic techniques. Labelled P8 bound strongly to thrombin-stimulated normal platelets (n = 14917 +/- 420, mean +/- SD) (Kd = 9.2 +/- 3.0 nM) and poorly to resting platelets (n = 2697 +/- 1278) (Kd = 24.8 +/- 18.6 nM). Moreover, the number of binding sites for P8 on thrombin-stimulated platelets from three Glanzmann's thrombasthenia patients, lacking the IIb-IIIa glycoprotein complex, were found similar to normal samples. F(ab')2 fragments of P8 inhibited aggregation of, and reduced secretion from, washed platelets stimulated by low concentrations of thrombin (0.05-0.06 U/ml) and collagen (0.5-0.6 microgram/ml). F(ab')2 fragments of P8 inhibited thrombin-induced platelet aggregation, but did not reduce fibrinogen binding (n) nor affect its dissociation constant (Kd). Inhibition of platelet aggregation by P8 suggests that thrombospondin plays an active role in promoting platelet aggregation, at low concentrations of thrombin and collagen. Normal binding of P8 to thrombin-stimulated Glanzmann thrombasthenic platelets indicates the presence of a thrombospondin receptor on the platelet surface distinct from the GPIIb-IIIa complex.  相似文献   

9.
Addition of prostacyclin (PGI2) temporarily inhibits platelet aggregation and permits the isolation of platelets free from plasma proteins, which have the same sensitivity as those in plasma [Moncada, Radomski & Vargas (1982) Br. J. Pharmacol. 75, 165P]. By using a modification of this technique we have established that platelets isolated from normal subjects aggregate more readily in response to ADP and adrenaline when physiological concentrations of low-density lipoproteins (LDL) are present. At high LDL concentrations spontaneous aggregation occurs. High-density lipoproteins (HDL) and very-low-density lipoproteins (VLDL) had no effect on agonist-induced platelet aggregation at normal concentrations, but HDL sensitized at higher concentrations. These effects by lipoproteins are not accompanied by changes in platelet lipid content. Cyclohexanedione treatment of LDL to modify apolipoproteins appeared to abolish the sensitization effect, indicating that binding to receptors was essential for the effects of LDL. LDL, but not HDL, overcame the inhibitory effect of PGI2 on platelet aggregation, except at very high concentrations of PGI2. PGI2 raised the cyclic AMP content of isolated platelets, but LDL only partially prevented this rise. These results suggest that LDL may have a greater role in platelet aggregation than previously recognized and may also regulate effects of PGI2. These findings may be of relevance to an understanding of cardiovascular diseases.  相似文献   

10.
By determining the sum of the supernatant concentrations of nitrite and nitrate the stimulated generation of nitric oxide (NO) by human washed platelets induced by a range of fibrillar collagen concentrations (0.0156-25 microg ml(-1)) was investigated. Platelet serotonin (5-hydroxytryptamine, 5-HT) efflux and platelet aggregation were also measured. Under resting conditions (0 microg ml(-1) collagen) platelet NO release was equivalent to 1.06+/-0.17 nmol per 10(8) platelets. Maximal NO release, equivalent to 2.1+/-0. 37 nmol per 10(8) platelets, was observed with only 0.0625 microg ml(-1) collagen (P<0.02, stimulated vs. resting release), higher collagen concentrations producing no further increases in platelet NO output. By contrast, maximal platelet aggregation and 5-HT efflux did not occur until collagen concentrations of 2.5 microg ml(-1) and 10-25 microg ml-1), respectively, had been achieved. L-NAME (1 mmol l(-1)) and L-NMMA (1 mmol l(-1)) inhibited stimulated platelet NO generation by 78+/-6% and 72%, respectively. Contrasting with fibrillar collagen, fibrillar beta-amyloid protein had no effect on platelet NO generation, or on 5-HT efflux or aggregation. These data perhaps indicate that NO generation by human platelets is stimulated by concentrations of fibrillar collagen insufficient to elicit an aggregatory response. Such a mechanism could operate in vivo to inhibit platelet aggregation which might otherwise be induced by low concentrations of circulating agonists.  相似文献   

11.
Antiplatelet effect of butylidenephthalide   总被引:1,自引:0,他引:1  
Butylidenephthalide inhibited, in a dose-dependent manner, the aggregation and release reaction of washed rabbit platelets induced by collagen and arachidonic acid. Butylidenephthalide also inhibited slightly the platelet aggregation induced by PAF and ADP, but not that by thrombin or ionophore A23187. Thromboxane B2 formation caused by collagen, arachidonic acid, thrombin and ionophore A23187 was in each case markedly inhibited by butylidenephthalide. Butylidenephthalide inhibited the aggregation of ADP-refractory platelets, thrombin-degranulated platelets, chymotrypsin-treated platelets and platelets in the presence of creatine phosphate/creatine phosphokinase. Its inhibition of collagen-induced aggregation was more marked at lower Ca2+ concentrations in the medium. The aggregability of platelets inhibited by butylidenephthalide could be recovered after the washing of platelets. In human platelet-rich plasma, butylidenephthalide and indomethacin prevented the secondary aggregation and blocked ATP release from platelets induced by epinephrine. Prostaglandin E2 formed by the incubation of guinea-pig lung homogenate with arachidonic acid could be inhibited by butylidenephthalide, indomethacin and aspirin. It is concluded that the antiplatelet effect of butylidenephthalide is mainly due to an inhibitory effect on cyclo-oxygenase and may be due partly to interference with calcium mobilization.  相似文献   

12.
In vitro, high concentrations of ethanol (EtOH) reduce platelet aggregation. Less is known about the effect of low EtOH doses on platelet function in a selected human population of long-life abstainers and low moderate-wine drinkers to avoid rebound effect of EtOH on platelet aggregation. Results of our experiments suggest that moderate-wine drinkers have higher levels of high density lipoprotein (HDL) than long-life abstainers while fibrinogen levels are unchanged. Furthermore, platelets obtained from these individuals do not differ in their response when stimulated by agonists such as AA and collagen. The effect of in vitro exposure of low doses of EtOH has been studied in PRP and in washed platelets. EtOH (0.1-10 mM) inhibits platelet aggregation induced by collagen at its ED50 while is ineffective when aggregation was triggered by U-46619 and by 1 microM adenosine diphosphate (ADP). 5-10 mM EtOH partially reduces the second wave of aggregation induced by 3 microM ADP. 0.1-10 mM EtOH dose-dependently lowers the aggregation induced by AA at its ED50 but it is less effective at ED75 of AA. The antiaggregating effect of EtOH on aggregation induced by AA is unchanged by inhibitor of nitric oxide synthase. In addition, 10 mM EtOH reduces thromboxane (Tx) formation. In washed platelets, 1-10 mM EtOH partially inhibits platelet aggregation induced by thrombin. In washed resting platelets, 10 mM EtOH does not change the resting [Ca++]i while significantly reduces the increase in [Ca++]i triggered by AA. The results of ex vivo experiments have demonstrated that wine increases the HDL. However, this observation may or may not influence the response of platelets to agonists. Results of our studies demonstrate that low doses of alcohol reduces platelet function.  相似文献   

13.
We previously found that sodium 2-propenyl thiosulfate (2PTS) has an anti-aggregatory effect in vitro on both canine and human platelets at relatively low concentrations, but the extent of aggregation tends to return to the control level at high concentrations. To clarify the mechanism of this modulatory influence of 2PTS on the aggregation of platelets, we investigated the effects of 2PTS on cyclooxygenase (COX) activity and the reduced glutathione (GSH) concentration in canine platelets. Platelet COX activity was inhibited by 2PTS in a dose-dependent manner up to 0.1 mM, but tended to return to the control level at 1 mM. In contrast, the platelet GSH concentration decreased in a dose-dependent manner after treatment with 2PTS and a significant decrease was observed at 0.1 mM (P<0.05) and 1 mM (P<0.001). Furthermore, the activity of purified COX-1 was directly inhibited by addition of GSH in a dose-dependent manner. From these results, we conclude that the 2PTS-induced inhibition of platelet aggregation occurs as a result of inhibition of COX activity. Additionally, 2PTS may have a modulatory effect on platelet aggregation by affecting the platelet GSH concentration.  相似文献   

14.
As widely assumed, platelets and coagulation system heavily influence the pathogenesis and progression of cardiovascular diseases. Some 1,4-naphthoquinone derivatives, such as vitamin K3, have been reported to increase the synthesis of coagulation proteins. In this study, we examine how 2-p-mercaptophenyl -1,4-naphthoquinone (NTP), a newly synthesized 1,4-naphthoquinone derivative, affects the platelet function in humans. A tapered parallel plate chamber which provided a range of shear stress covering the entire physiological range in human circulation is used to assess platelet adhesiveness on fibrinogen coated-surface. In addition, platelet aggregation and thromboxane B2 (TXB2) production by inducers are evaluated by the turbidimetric method and enzyme immunoassay kit, respectively. Moreover, platelets [Ca2+]i are measured using a dual-wavelength fluorescence spectrophotometer. Analysis results indicate that 1) NTP decreases the percentages of attached platelets at the locations in various shear stresses and the levels of platelet adhesiveness, denoted as the slope; 2) NTP can inhibit the platelet aggregation by ADP (2 microM) and collagen (25 microg/ml), and the IC50 are: 0.32 and 26.83 microg/ml, respectively; and 3) NTP markedly inhibits TXB2 formation and platelet [Ca2+]i elevation caused by ADP and collagen. Therefore, we conclude that NTP may inhibit platelet adhesiveness on fibrinogen coated-surface, aggregability, [Ca2+]i, and thromboxane production, and that it may be used as an antiplatelet agent.  相似文献   

15.
Incubation of human high density lipoprotein (HDL) particles (density = 1.063-1.21 g/ml) with catalytic amounts of Manduca sexta lipid transfer particle (LTP) resulted in alteration of the density distribution of HDL protein such that the original HDL particles were transformed into new particles with an equilibrium density = 1.05 g/ml. Concomitantly, substantial amounts of protein were recovered in the bottom fraction of the density gradient. The LTP-induced alteration in HDL protein density distribution was dependent on the LTP concentration and incubation time. Electrophoretic analysis revealed that the lower density fraction contained apolipoprotein A-II (apoA-II) as the major apoprotein component while nearly all of the apoA-I was recovered in the bottom fraction. Lipid analysis of the HDL substrate and product fractions revealed that the apoA-I-rich fraction was nearly devoid of lipid (less than 1%, w/w). The lipid originally associated with HDL was recovered in the low density, apoA-II-rich, lipoprotein fraction, and the ratios of individual lipid classes were the same as in control HDL. Electron microscopy and gel permeation chromatography experiments revealed that the LTP-induced product lipoprotein population comprised particles of larger size (19.7 +/- 1.4-nm diameter) than control HDL (10.6 +/- 1.4-nm diameter). The results suggest that facilitated net lipid transfer between HDL particles altered the distribution of lipid such that apoprotein migration occurred and donor particles disintegrated. Similar results were obtained when human HDL3 or HDL2 density subclasses were employed as substrates for LTP. The lower surface area to core volume ratio of the larger, product lipoprotein particles compared with the substrate HDL requires that there be a decrease in the total exposed lipid/water interface which requires stabilization by apolipoprotein. Selective displacement of apoA-I by apoA-II or apoC, due to their greater surface binding affinity, dictates that apoA-I is preferentially lost from the lipoprotein surface and is therefore recovered as lipid-free apoprotein. Thus, it is conceivable that the structural arrangement of HDL particle lipid and apoprotein components isolated from human plasma may not represent the most thermodynamically stable arrangement of lipid and protein.  相似文献   

16.
Plasma lipoproteins from 5-week old male chickens were separated over the density range 1.006-1.172 g/ml into 22 subfractions by isopycnic density gradient ultracentrifugation, in order to establish the distribution of these particles and their constituent apolipoproteins as a function of density. Lipoprotein subfractions were characterized by electrophorectic, chemical and morphological analyses, and their protein moieties were defined according to net charge at alkaline pH, molecular weight and isoelectric point. These analyses have permitted us to reevaluate the density limits of the major chicken lipoprotein classes and to determine their main characteristics, which are as follows: (1) very-low-density lipoproteins (VLDL), isolated at d less than 1.016 g/ml, were present at low concentrations (less than 0.1 mg/ml) in fasted birds; their mean diameter determined by gradient gel electrophoresis and by electron microscopy was 20.5 and 31.4 nm respectively; (2) as the the density increased from VLDL to intermediate density lipoproteins (IDL), d 1.016-l.020 g/ml) and low-density lipoproteins (LDL, d 1.020-1.046 g/ml), the lipoprotein particles contained progressively less triacylglycerol and more protein, and their Stokes diameter decreased to 20.0 nm; (3) apolipoprotein B-100 was the major apolipoprotein in lipoproteins of d less than 1.046 g/ml, with an Mr of 350000; small amounts of apolipoprotein B-100 were detectable in HDL subfractions of d less than 1.076 g/ml; urea-soluble apolipoproteins were present in this density range as minor components of Mr 38000-39000, 27000-28000 (corresponding to apolipoprotein A-1) and Mr 11000-12000; (4) high density lipoprotein (HDL, d 1.052-1.130 g/ml) was isolated as a single band, whose protein content increased progressively with increase in density; the chemical composition of HDL resembled that of human HDL2, with apolipoprotein A-1 (M 27000-28000) as the major protein component, and a protein of Mr 11000-12000 as a minor component; (5) heterogeneity was observed in the particle size and apolipoprotein distribution of HDL subfractions: two lipoprotein bands which additional apolipoproteins of Mr 13000 and 15000 were detected. These studies illustrate the inadequacy in the chicken of the density limits applied to fractionate the lipoprotein spectrum, and particularly the inappropriateness of the 1.063 g/ml density limit as the cutoff for LDL and HDL particle populations in the species.  相似文献   

17.
Platelet aggregability was studied in 154 apparently healthy subjects aged 18–60 years. Both spontaneous platelet aggregation and aggregation induced by inducers with different mechanisms of action were studied. In 25-to 27-year-old subjects, the parameters of both the spontaneous aggregation, which characterized the platelet activation, and ADP-induced platelet aggregation, which characterized the total platelet aggregability, were lower than in older subjects (45–50 and 60–70 years of age). The parameters of the platelet activity and aggregation indicated the absence of age-related changes in the structure and functions of platelets in response to inducers, including epinephrine, collagen, and ristocetin.  相似文献   

18.
Lipoprotein (a) (Lp (a)) may be pro-thrombotic in humans due to its apolipoprotein (a) (apo(a))-mediated decreases in fibrinolysis. Such decreased fibrinolysis arises putatively from interference with plasminogen conversion to plasmin due to the considerable homology between apolipoprotein (a) and plasminogen. However, in vitro, most studies have shown that human Lp (a) decreases agonist-stimulated platelet aggregation while in vivo it appears to decrease aggregation as implied by increased bleeding times with higher blood serum concentrations of Lp(a). Lp (a) binding to platelets mediated by apo (a) increases platelet intracellular c-AMP levels in resting platelets, and decreases platelet production of thromboxane A2 and fibrinogen binding to platelets all of which reduce platelet aggregation. One, though not the only, explanation of these conflicting data may be that Lp(a) self-regulates its interference with fibrinolysis by reducing platelet aggregation and platelet binding of fibrinogen and hence the degree of requirement for fibrinolysis. However, it is concluded more in vivo work needs to be done to fully understand whether, if at all, Lp(a) in varying concentrations and isoforms, favours reduced platelet aggregation or fibrinolysis.  相似文献   

19.
Platelets were activated with freezing/thawing and thrombin stimulation, and platelet microparticles generated following platelet activation were isolated with ultracentrifugation. The effects of platelet microparticles on platelet activation were studied with annexin V assay, protein tyrosine phosphorylation, and platelet aggregation. Freezing-induced platelet microparticles decreased but thrombin-induced platelet microparticles increased platelet annexin V binding and aggregation. Freshly washed platelets were cryopreserved using epinephrine and dimethyl sulfoxide (Me(2)SO) as combined cryoprotectants, and stimulated with thrombin-induced platelet microparticles. Following incubation of thrombin-induced platelet microparticles, the reaction time of platelets to agonists decreased but the percentages of aggregation increased, such as washed platelets from 44% +/- 30 to 92% +/- 7, p < 0.001, and cryopreserved platelets from 66% +/- 10 to 77% +/- 7, p < 0.02. By increasing platelet aggregability, platelet microparticles recovered after thrombin stimulation improved platelet function for transfusion. A 53-kDa platelet microparticle protein showed little phosphorylation if it was released from resting platelets or platelets stimulated with ADP, epinephrine, propyl gallate or dephosphorylation if it was derived from ionophore A 23187-stimulated platelets. However, the same protein released from frozen platelets showed significant tyrosine phosphorylation. Since a microparticle protein with 53 kDa was compatible with protein tyrosine phosphatase-1B (PTP-1B), its phosphorylation suggests the inhibition of enzyme activity. The microparticle proteins derived from thrombin-stimulated platelets were significantly phosphorylated at 64 kDa and pp60c-src, suggesting that the activation of tyrosine kinases represents a possible mechanism of thrombin-induced platelet microparticles to improve platelet aggregation.  相似文献   

20.
I Fuse  S Ito  A Takagi  A Shibata 《Life sciences》1985,36(21):2047-2052
We studied the different effects of three kinds of somatostatin (somatostatin 1-14, somatostatin 15-28, somatostatin 1-28) on the aggregation of rabbit's platelets. It was clarified that somatostatin 15-28 had inhibitory effects on rabbit's platelet aggregation stronger than somatostatin 1-14 did, and that somatostatin 1-28 did not have any such effects. These anti-aggregatory effects of somatostatin were stronger when induced by collagen than induced by ADP.  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