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 453 毫秒
1.
Glutamate transport in blood platelets resembles that in brain nerve terminals because platelets contain neuronal Na+-dependent glutamate transporters, glutamate receptors in the plasma membrane, vesicular glutamate transporters in secretory granules, which use the proton gradient as a driving force, and can release glutamate during aggregation/activation. The acidification of secretory granules and glutamate transport were assessed during acute treatment of isolated platelets with cholesterol-depleting agent methyl-β-cyclodextrin (MβCD). Confocal imaging with the cholesterol-sensitive fluorescent dye filipin showed a quick reduction of cholesterol level in platelets. Using pH-sensitive fluorescent dye acridine orange, we demonstrated that the acidification of secretory granules of human and rabbit platelets was decreased by ∼15% and 51% after the addition of 5 and 15 mM MβCD, respectively. The enrichment of platelet plasma membrane with cholesterol by the application of complex MβCD-cholesterol (1:0.2) led to the additional accumulation of acridine orange in secretory granules indicating an increase in the proton pumping activity of vesicular H+-ATPase. MβCD did not evoke release of glutamate from platelets that was measured with glutamate dehydrogenase assay. Flow cytometric analysis did not reveal alterations in platelet size and granularity in the presence of MβCD. These data showed that the dissipation of the proton gradient of secretory granules rather than their exocytosis caused MβCD-evoked decrease in platelet acidification. Thus, the depletion of plasma membrane cholesterol in the presence of MβCD changed the functional state of platelets affecting storage capacity of secretory granules but did not evoke glutamate release from platelets.  相似文献   

2.
Prostacyclin prolongs viability of washed human platelets   总被引:5,自引:0,他引:5  
The functional viability of stored human platelets, washed in the presence and absence of prostacyclin, was examined over a 96 h period. Platelet counts, aggregation responses and cyclic AMP levels were monitored as well as the spontaneous generation of thromboxane B2 and the liberation of labelled oleate from cellular phosphatides. In suspensions prepared without prostacyclin in the medium, platelet counts declined rapidly as did the sensitivity to aggregating agents. In addition, substantial amounts of thromboxane B2 were generated during preparation and storage and oleate liberation occurred at a rapid rate. In contrast, in prostacyclin-washed platelets, aggregation was maintained throughout the study period and there was little generation of thromboxane B2. Moreover, only a gradual decrease in platelet count and a slow increase in the rate of oleate liberation was observed when compared with controls. However, cyclic AMP levels rapidly declined when platelets were resuspended in prostacyclin-free medium.  相似文献   

3.
Vitamin E (α-tocopherol) and tocopherol acetate produced a slightly increased amount of thromboxane in treated compared to untreated platelets. In tocopherol acetate-treated platelets significantly more lipoxygenase products were produced. α-tocopherol induced an increased, but not significant, production of thromboxane B2 during blood clotting. α-tocopherol was not found to affect platelet phospholipase activity as determined by its effect on the release of labelled arachidonic acid from platelet phospholipids by challenging the platelets with calcium ionophore A23,187. α-tocopherol potentiated the incorporation of labelled arachidonate in the platelet phospholipids. Inspite of having no effect on the arachidonic acid cascade in platelets, α-tocopherol inhibited aggregation induced by several aggregating agents including A23,187. Inhibition of aggregation may be explained by the ability of α-tocopherol to inhibit intracellular mobilization of sequestered calcium from the dense tubular system to the cytoplasm.  相似文献   

4.
Following vessel wall injury, platelets adhere to the exposed subendothelium, become activated and release mediators such as TXA2 and nucleotides stored at very high concentration in the so-called dense granules. Released nucleotides and other soluble agents act in a positive feedback mechanism to cause further platelet activation and amplify platelet responses induced by agents such as thrombin or collagen. Adenine nucleotides act on platelets through three distinct P2 receptors: two are G protein-coupled ADP receptors, namely the P2Y1 and P2Y12 receptor subtypes, while the P2X1 receptor ligand-gated cation channel is activated by ATP. The P2Y1 receptor initiates platelet aggregation but is not sufficient for a full platelet aggregation in response to ADP, while the P2Y12 receptor is responsible for completion of the aggregation to ADP. The latter receptor, the molecular target of the antithrombotic drugs clopidogrel, prasugrel and ticagrelor, is responsible for most of the potentiating effects of ADP when platelets are stimulated by agents such as thrombin, collagen or immune complexes. The P2X1 receptor is involved in platelet shape change and in activation by collagen under shear conditions. Each of these receptors is coupled to specific signal transduction pathways in response to ADP or ATP and is differentially involved in all the sequential events involved in platelet function and haemostasis. As such, they represent potential targets for antithrombotic drugs.  相似文献   

5.
Platelets were harvested by a Hemonetics Model-30 discontinuous cell separator from 20 normal volunteers and were cryopreserved in the presence of 5% DMSO at a controlled rate of freezing of -1 degrees C/min and stored in liquid nitrogen for up to 3 months. A significant loss of platelets occurred at the platelet concentration step through adhesion of platelets to the bag walls. A small reduction in aggregation associated with this was also seen and may reflect some damage to the platelets during the pheresis procedure. A small, but significant loss of platelet aggregation was seen with all agents following cryopreservation. Mean percentage aggregation post-thaw for all the agents was 75.4% (range 74-78%) and platelet recovery was approximately 90%. No significant changes in aggregation or recovery were seen over the 3 months' storage period. The cryoprotectant DMSO was shown to have no deleterious effect on platelet function in vitro.  相似文献   

6.
The receptor for ADP on the platelet membrane, which triggers exposure of fibrinogen-binding sites and platelet aggregation, has not yet been identified. Two enzymes with which ADP interacts on the platelet surface, an ecto-ATPase and nucleosidediphosphate kinase, have been proposed as possible receptors for ADP in ADP-induced platelet aggregation. In the present study, experiments were conducted with washed human platelets to examine if a relationship existed between platelet aggregation, fibrinogen binding and the enzymatic degradation of ADP. With 12 different platelet suspensions, a good correlation (P less than 0.01) was found between the extent of platelet aggregation and the amount of 125I-fibrinogen bound to platelets after ADP stimulation. No correlation was found between these parameters and the rate or extent of transformation of [14C]ADP to [14C]ATP or [14C]AMP. The binding of fibrinogen to platelets was inhibited in parallel with aggregation when ADP stimulation was impaired by the enzymatic degradation of ADP by the system creatine phosphate/creatine phosphokinase, or by the use of specific antagonists, such as ATP and AMP. These antagonists also influenced the enzymatic degradation of ADP. This effect occurred at lower concentrations of ATP or AMP than those required to inhibit ADP-induced platelet aggregation and fibrinogen binding. Our results demonstrate that ATP and AMP may be used as specific antagonists of the ADP-induced fibrinogen binding to platelets. They do not provide evidence to suggest that enzymes which metabolize ADP on the platelet surface are involved in the mechanism of ADP-induced platelet aggregation.  相似文献   

7.
1. Human platelet-rich plasma prelabelled with [(3)H]adenine was incubated at 37 degrees C with antimycin A and 2-deoxy-d-glucose. Variations in the amounts of ATP, ADP and P(i), and in the radioactivity of ATP, ADP, AMP, IMP, hypoxanthine+inosine and adenine were determined during incubation. Adrenaline- and ADP-induced platelet aggregation and the ADP-induced shape change of the platelets were determined concurrently. 2. 2-Deoxyglucose caused conversion of [(3)H]ATP to [(3)H]hypoxanthine+inosine. The rate of this conversion increased with increasing 2-deoxyglucose concentration and was markedly stimulated by addition of antimycin, which had no effect alone. At maximal ATP-hypoxanthine conversion rates, the IMP radioactivity remained at values tenfold higher than control, whereas [(3)H]ADP and [(3)H]AMP radioactivity gave variations typical for product/substrates in consecutive reactions. The specific radioactivityof ethanol-soluble platelet ATP decreased during incubation to less than one-tenth of its original value. The amounts and radioactivity of ethanol-insoluble ADP did not vary during incubation with the metabolic inhibitors. 3. The rate of ADP- and adrenaline-induced primary aggregation decreased as the amount of radioactive ATP declined, and complete inhibition of aggregation was obtained at a certain ATP concentration (metabolic ATP threshold). This threshold decreased with increasing concentration of inducer ADP. 4. Secondary platelet aggregation (release reaction) had a metabolic ATP threshold markedly higher than that of primary aggregation. 5. Shape change was gradually inhibited as the ATP radioactivity decreased, and had a metabolic ATP threshold distinctly lower than that of primary aggregation, and which decreased with increasing concentration of ADP. 6. A small but distinct fraction of [(3)H]ATP disappeared rapidly during the combined shape change-aggregation process induced by ADP in platelets incubated with metabolic inhibitors, whereas no ATP disappearance occurred during aggregation in their absence.  相似文献   

8.
The platelet content of PPi is 1.90 +/- mumol/10(11) platelets (S.E.M., n = 19) or about 10.5 nmol/mg of protein, several hundred times that found for rat liver. Some 80% of this PPi is secreted by platelets treated with thrombin with a time course and dose-response relationship similar to secretion of ATP, ADP and 5-hydroxytryptamine (serotonin) from the platelet dense granules. During platelet aggregation induced by ADP and adrenaline, substantial amounts of PPi were secreted, but no release of acid hydrolases was observed. Subcellular-fractionation studies showed that the PPi is highly enriched in the same fraction that contains the storage organelles which store ATP, ADP, Ca2+ and 5-hydroxytryptamine. Inorganic pyrophosphatase was present mainly in the soluble fraction and in the mitochondria. Secretion studies done with platelets prelabelled with [32P]Pi showed that the sequestered PPi was relatively metabolically inactive, as is the ATP and ADP in the storage organelles. The possible participation of PPi in the formation of a bivalent-cation-nucleotide complex associated with amine storage is discussed.  相似文献   

9.
The equilibrium binding of 14C-labeled ADP to intact washed human blood platelets and to platelet membranes was investigated. With both intact platelets and platelet membranes a similar concentration dependence curve was found. It consisted of a curvilinear part below 20 microM and a rectilinear part above this concentration. At high ADP concentrations, the rectilinear part appeared to be saturable. Because of this, two classes of saturable ADP binding sites were proposed. ADP was partly converted to ATP and AMP with intact platelets while this conversion was virtually absent in isolated platelet membranes. ADP was bound to platelet membranes with the same type of curves found for intact platelets. The ADP binding to the high affinity system, which was stimulated by calcium ions, was nearly independent of temperature and had a pH optimum at 7.8. A number of agents were investigated for inhibiting properties. Of the sulfhydryl reagents only p-chloromercuribenzene sulfonate inhibited both high and low affinity binding systems while iodoacetamide and N-ethylmaleimide were without effect. Compounds acting via cyclic AMP on platelet aggregation, such as adenosine and cyclic AMP itself, had no influence on binding. Some nucleosidediphosphates and nucleotide analogs at a concentration of 100 microM had no, or only a slight, effect on high affinity ADP binding. For some other nucleotides inhibitor constants were determined for both platelet ADP aggregation and ADP binding. The inhibitor constants of ATP, adenyl-5'-yl-(beta,gamma-methylene)diphosphate, IDP, adenosine-5'(2-O-thio)diphosphate, for aggregation and high affinity binding were in good correlation with each other. Exceptions formed fluorosulfonylbenzoyl adenosine and AMP. The ATP formation found with intact platelets could be attributed to a nucleosidediphosphate kinase. It was investigated in some detail. The enzyme was magnesium dependent, had a Q10 value of 1.41, a pH optimum at 8.0, was competitively inhibited by AMP and reacted via a ping pong mechanism. All findings described in this paper indicate that platelets as well as platelet membranes bind ADP with the same characteristics and they suggest that the high affinity binding of ADP is involved in platelet aggregation induced by ADP. The results on nucleosidediphosphate kinase did not permit a firm conclusion about the role of the enzyme in induction of platelet aggregation by ADP.  相似文献   

10.
Platelet activation may explain the storage lesion in platelet concentrates   总被引:5,自引:0,他引:5  
A P Bode 《Blood cells》1990,16(1):109-25; discussion 125-6
While the exact nature of the dysfunction of stored platelets is not known, it is generally agreed that the platelet's metabolic activity with lactate accumulation presents a significant impediment to prolonged storage. There is an increasing body of evidence that stored platelets have become activated in the preparation and handling of platelet concentrates. Changes in platelet function and structure in concentrates can be explained in terms of sequelae of activation, especially heightened metabolic activity and activation-specific changes in surface glycoproteins on stored platelets. With the use of inhibitors of platelet activation in the preparation of platelet concentrates, the loss of platelet function and integrity is less rapid and platelet metabolic rate is decreased during an extended storage period. Surface levels of glycoprotein Ib, normally decreased during prolonged storage of platelets, are well-preserved in the presence of activation inhibitors. When the use of inhibitors is combined with replacement of plasma with an artificial medium, platelets stored for up to 20 days appear to be metabolically and structurally intact and responsive to stimuli. In summary, platelet activation appears to play a major role in the generation of the storage lesion in platelet concentrates.  相似文献   

11.
In continuation of our studies with the oil of cloves--a common kitchen spice and a crude drug for home medicine--we have isolated yet another active component identified as acetyl eugenol (AE); the earlier reported active component being eugenol. The isolated material (IM) was found to be a potent platelet inhibitor; IM abolished arachidonate (AA)-induced aggregation at ca. 12 microM, a concentration needed to abolish the second phase of adrenaline-induced aggregation. Chemically synthesized acetyl eugenol showed similar effects on AA- and adrenaline-induced aggregation. A dose-dependent inhibition of collagen-induced aggregation was also observed. AE did not inhibit either calcium ionophore A23187- or thrombin-induced aggregation. Studies on aggregation and ATP release were done using whole blood (WB). AA-induced aggregation in WB was abolished at 3 micrograms/ml (14.6 microM) which persisted even after doubling the concentration of AA. ATP release was inhibited. Inhibition of aggregation appeared to be mediated by a combination of two effects: reduced formation of thromboxane and increased generation of 12-lipoxygenase product (12-HPETE). These effects were observed by exposing washed platelets to (14C)AA or by stimulating AA-labelled platelets with ionophore A23187. Acetyl eugenol inhibited (14C)TxB2 formation in AA-labelled platelets on stimulation with thrombin. AE showed no effect on the incorporation of AA into platelet phospholipids.  相似文献   

12.
H M Rinder  E L Snyder 《Blood cells》1992,18(3):445-56; discussion 457-60
This review will discuss how stored platelets become activated and will examine their ability to function and survive in vivo, posttransfusion. Experimental methods which have been shown to alter platelets during storage will be detailed. Using beta-thromboglobulin (beta-TG) and surface adhesion receptors as markers, investigators have examined the activation changes in platelet concentrates during preparation and storage. Resuspension of the platelet pellet after isolation of platelet-rich plasma appears to play a major role in producing platelet activation and beta-TG release during preparation. However, there is a significant amount of interdonor variability in platelet activation even at this early stage of storage. Over 5 days of storage, platelets release approximately 50% of their beta-TG contents. Furthermore, between 40% and 60% of the platelets express the alpha-granule membrane protein, P-selectin (GMP-140), during storage, which is also indicative of platelet activation. These activation changes correlate to some degree with platelet recovery posttransfusion but clearly do not explain the full lesion of platelet storage. The surface density of two platelet membrane receptors, glycoproteins (GP) Ib and IIb/IIIa, also change with activation, although in opposite directions. Platelet surface GPIb decreases initially with storage and then recovers, perhaps due to its relocation to the platelet surface from an intracellular pool. In contrast to GPIb, mean platelet surface GPIIb/IIIa increases slightly during storage, probably as a consequence of platelet activation and release of alpha-granule GPIIb/IIIa to the surface. Some hypotheses are offered regarding how these activated platelets can continue to circulate after transfusion. Further exploration of the platelet storage lesion will hopefully provide needed answers and thus permit better treatment of hemostatic disorders in the future.  相似文献   

13.
S P Colgan  M A Hull Thrall  P W Gasper 《Blood cells》1989,15(3):585-95; discussion 596-600
A rapid and simple technique using the Whole Blood Lumi-Aggregometer was used to study storage pool disease in Chediak-Higashi homozygote and heterozygote cats. Feline Chedlak-Higashi platelets aggregated after the addition of both ADP and collagen. During platelet aggregation, ATP secretion was assayed; the whole blood aggregometer is effective in detecting decreased levels of secretable ATP in homozygote cats. No storage pool deficiency was found in heterozygote cats. However, upon analysis of impedance tracings, a decreased platelet aggregation response was seen in both homozygote and heterozygote cats. These results suggest that prolonged bleeding times in Chediak-Higashi cats may involve a mechanism in addition to a dense granule deficiency.  相似文献   

14.
Platelet abnormalities of Tester Moriyama (TM) rats, which have prolonged bleeding time with normal platelet count, were characterized by comparison with those of fawn-hooded (FH) rats with platelet storage pool deficiency (SPD). Morphologically, the dense granules were virtually lacking in platelets from TM and FH rats. Platelets from TM and FH rats aggregated in response to adenosine diphosphate (ADP), but failed to have secondary aggregation. In contrast, platelet aggregation was completely absent in response to 1 to 20 micrograms of collagen/ml, although partial aggregation was observed at the higher dosage of 50 micrograms/ml. Normal amounts of platelet membrane glycoproteins IIb/IIIa were expressed in TM and FH rats, but platelet adenosine triphosphate (ATP) and ADP contents were lower than those in platelets from control Wistar rats. Platelet ATP-to-ADP ratio of TM and FH rats was significantly higher than that of Wistar rats. Serotonin content in platelets from TM and FH rats was 20 to 25% that of Wistar rat platelets. These results suggested that platelet abnormalities of TM rats are a typical characteristic of platelet SPD and are similar to those of FH rats, which are genetically different from TM rats. Therefore, TM rats may serve as a useful animal model for the study of platelet SPD.  相似文献   

15.
Shear stress triggers von Willebrand factor (VWF) binding to platelet glycoprotein Ibalpha and subsequent integrin alpha(IIb)beta(3)-dependent platelet aggregation. Concomitantly, nucleotides are released from plateletdense granules, and ADP is known to contribute to shear-induced platelet aggregation (SIPA). We found that the impaired SIPA of platelets from a Hermansky-Pudlak patient lacking dense granules was restored by exogenous l-beta,gamma-methylene ATP, a stable P2X(1) agonist, as well as by ADP, confirming that in addition to ADP (via P2Y(1) and P2Y(12)), ATP (via P2X(1)) also contributes to SIPA. Likewise, SIPA of apyrase-treated platelets was restored upon P2X(1) activation with l-beta,gamma-methylene ATP, which promoted granule centralization within platelets and stimulated P-selectin expression, which is a marker of alpha-granule release. In addition, during SIPA, platelet degranulation required both extracellular Ca(2+) and VWF-glycoprotein Ibalpha interactions without involving alpha(IIb)beta(3). Neither platelet release nor SIPA was affected by protein kinase C inactivation, even though protein kinase C blockade inhibits platelet responses to collagen and thrombin in stirring conditions. In contrast, inhibiting myosin light chain (MLC) kinase with ML-7 reduced platelet release and SIPA by 30%. Accordingly, the potentiating effect of P2X(1) stimulation on the aggregation of apyrase-treated platelets coincided with intensified phosphorylation of MLC and was abrogated by ML-7. SIPA-induced MLC phosphorylation occurred exclusively through released nucleotides and selective antagonism of P2X(1) with MRS2159-reduced SIPA, ATP release, and potently inhibited MLC phosphorylation. We conclude that the P2X(1) ion channel induces MLC-mediated cytoskeletal rearrangements, thus contributing to SIPA and degranulation during VWF-triggered platelet activation.  相似文献   

16.
A23187 induced shape change, aggregation and secretion of platelets in plasma. When rapid cooling was used to stop secretion and centrifugation to separate the cells from the medium, maximal amounts of storage ATP plus ADP and preadsorbed [14C]serotonin were found in the supernatant immediately (less than 5 s) after A23187 addition. These results suggested that A23187 could cause shape change and aggregation through secreted ADP and not directly. When secretion was stopped with chilling and formaldehyde treatment before centrifugation, the secreted substances appeared after a lag of 60-120 s, i.e. after shape change was terminated and aggregation was well on its way. These two platelet responses thus seemed to be independent of secretion and induced directly by A23187. The absence of a lag period when secretion was stopped by chilling alone was thought to be due to centrifugation-induced secretion of platelets conditioned by A23187. Formaldehyde completely inhibited centrifugation-induced secretion. At 37 degrees C, formaldehyde caused rapid breakdown of metabolic ATP in platelets with a pattern dependent on the formaldehyde concentration: Below 50 mM, ATP was converted to inosine plus hypoxanthine via ADP, AMP and IMP and the adenylate energy charge was preserved. Above 100 mM, AMP was the end product with a drastic reduction in the adenylate energy charge. These changes were not due to lysis of the platelets, but were apparently caused by an formaldehyde-induced increase in cellular ATP consumption. Platelet secretion is usually associated with a conversion of metabolic ATP to hypoxanthine. Formaldehyde had to be used to stop secretion and since it caused breakdown of ATP, additional smaples were taken out for nucleotide determination during stirring of platelet-rich plasma with A23187. It was found that metabolic ATP was converted to inosine plus hypoxanthine only during the secretion step.  相似文献   

17.
The correlation between energy consumption and platelet responses induced by collagen, A23187 and ADP was investigated and compared with the energetics of thrombin-stimulated platelets established in earlier work. Aggregation, measured as single-platelet disappearance, and secretion correlated quantitatively with the increment but not with the total consumption of energy, suggesting that the former reflects the energy cost of these responses. The cost of complete aggregation was 2-3 mumol of ATP equivalents/10(11) platelets with collagen, ADP and thrombin as the stimulus. The cost of complete dense-granule secretion was 0.5-0.8 mumol of ATP equivalents/10(11) platelets with all agonists tested. The cost of combined secretion of alpha-granule and acid hydrolase granule contents was 5-7 mumol of ATP equivalents/10(11) platelets with thrombin and collagen. However, in the presence of A23187 much more energy was consumed during aggregation and secretion. Also ADP triggered more energy consumption during secretion than was seen with the other inducers. The effect of inhibitors of aggregation and secretion was investigated in thrombin-stimulated platelets. Raising the cellular cyclic AMP content sharply decreased the increment in energy consumption as well as aggregation and secretion. The cytoskeleton-disrupting agents cytochalasin B and colchicine left the increment in energy consumption intact, but decreased the basal consumption seen in unstimulated platelets. This was accompanied by normal (cytochalasin B) or diminished (colchicine) aggregation and secretion. Apart from the latter exception, all inhibitors decreased secretion and incremental energy consumption in parallel, thereby preserving the energy-versus-secretion relationship established in earlier work. In contrast, aggregation and energy consumption varied independently, suggesting that the coupling with energy consumption is much weaker for this response.  相似文献   

18.
Platelets are uniquely stored at room temperature, during which they gradually loss their quality owing to deteriorating functions of mitochondria over time. Given the well‐documented beneficial effect of near infrared low‐level light (LLL) on mitochondrial functions, we explored a potential for LLL to protect mitochondrial function and extend the shelf‐life of platelets beyond the current 5 days. We found that exposure of a platelet‐containing storage bag to 830 nm light‐emitting diode (LED) light at 0.5 J/cm2 prior to storage could significantly retain a pH value and viability of the platelets stored for 8 days with improved quality compared to those stored similarly for 5 days in controls. The LLL inhibited reactive oxygen species (ROS) and lactate production, while sustaining ATP synthesis and mitochondrial membrane potential and morphology in the stored platelets. It also sustained aggregation capacity and in vivo survival of stored platelets, concomitant with no significant activation, as suggested by similar CD62p expression and enhanced agonist‐induced aggregation and recovery following infusion in the presence compared to absence of LLL treatment. This simple, additive‐free, cost‐effective, noninvasive approach can be readily incorporated into the current platelet storage system to potentially improve quality of stored platelets.   相似文献   

19.
BACKGROUND: Nitrite is a nitric oxide (NO) metabolite in tissues and blood, which can be converted to NO under hypoxia to facilitate tissue perfusion. Although nitrite is known to cause vasodilation following its reduction to NO, the effect of nitrite on platelet activity remains unclear. In this study, the effect of nitrite and nitrite+erythrocytes, with and without deoxygenation, on platelet activity was investigated. METHODOLOGY/FINDING: Platelet aggregation was studied in platelet-rich plasma (PRP) and PRP+erythrocytes by turbidimetric and impedance aggregometry, respectively. In PRP, DEANONOate inhibited platelet aggregation induced by ADP while nitrite had no effect on platelets. In PRP+erythrocytes, the inhibitory effect of DEANONOate on platelets decreased whereas nitrite at physiologic concentration (0.1 μM) inhibited platelet aggregation and ATP release. The effect of nitrite+erythrocytes on platelets was abrogated by C-PTIO (a membrane-impermeable NO scavenger), suggesting an NO-mediated action. Furthermore, deoxygenation enhanced the effect of nitrite as observed from a decrease of P-selectin expression and increase of the cGMP levels in platelets. The ADP-induced platelet aggregation in whole blood showed inverse correlations with the nitrite levels in whole blood and erythrocytes. CONCLUSION: Nitrite alone at physiological levels has no effect on platelets in plasma. Nitrite in the presence of erythrocytes inhibits platelets through its reduction to NO, which is promoted by deoxygenation. Nitrite may have role in modulating platelet activity in the circulation, especially during hypoxia.  相似文献   

20.
A protocol for the biochemical study of platelet stored for transfusional use at 22 degrees C and under continuous shaking in a plastic bag highly permeable to gases and with a suitable area/volume ratio, is described. Plasmatic dextrose, lactic acid, lactic dehydrogenase activity, cellular ATP and malonyldialdehyde were monitored during the storage, as well as some acid-base indexes namely: pH, pCO2, HCO3-, pO2. The platelet functional status was checked as aggregating power induced by ADP and collagen and by beta-thromboglobulin release. The results obtained are indicative of a discrete maintenance of aerobic metabolism by platelets which are able to give up CO2 and take up O2 so that the plasmatic pH is constant during the storage. However, the malonyldialdehyde increase suggests that platelets become increasingly susceptible to peroxidative attacks. The aggregating response was dramatically reduced even on the third day of storage. The data obtained point out that, under the conditions reported, platelets can be transfused up to the third day of storage.  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