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 15 毫秒
1.
The tendency of sickle cells to adhere to the endothelium reflects the surface features not only of the red cells but also of the endothelial cells. Sickle cell disease is a prototype of a condition where the erythrocyte is under stress, ischemic, oxidative, or shear stress, that causes changes in the erythrocyte morphology. This change leads eventually to enhanced erythrocyte-endothelial cell adhesion.Reactive oxygen species generated by cytokine-activated inflammatory cells oxidize lipoproteins such as LDL and lipoprotein(a) within the vessel wall, facilitating uptake of these particles by activated macrophages and smooth muscle cells, with conversion into lipid-laden foam cells. Notably, the membranes of sickle RBCs have undergone excessive cytoskeletal protein thiol oxidation, and sickle RBCs are abnormally prone to vesiculation during mechanical stress in vitro and apparently in vivo.This abnormality was successfully reproduced in normal RBCs by causing stress conditions using PMS-induced stimulation of intracellular superoxide generation, a process similar to that occurring in sickle RBCs. It could be that the generation of reactive oxygen species in atherosclerosis activates red blood cells, and microvesicles of red blood cells are formed, enhancing the activation of the vascular endothelium and leading to vascular inflammation and atherogenesis.  相似文献   

2.
R P Hebbel  A Leung  N Mohandas 《Blood》1990,76(5):1015-1020
It has been hypothesized that some of the irreversible microrheologic abnormalities of sickle red blood cell (RBC) membranes could result from autoxidative perturbation. To model this possibility, we used micromechanical manipulation to examine the static extensional rigidity and inelastic or plastic behavior of normal RBCs exposed to phenazine methosulfate (PMS), an agent that generates superoxide from within the cell. In response to this stress, RBC membranes became stiff as evidenced by increasing extensional rigidity. At 50 mumol/L PMS they were as stiff as the membranes of most dense, dehydrated sickle RBCs; and at 25 mumols/L PMS the membranes were similar to somewhat less dense sickle RBCs. When examined for inelastic behavior, RBCs exposed to PMS even at 10 mumols/L showed hysteresis in loading and unloading phases of the curve relating aspiration length to suction pressure, and they developed membrane bumps that persisted after RBC release from the pipette. Examination of single cells in both isotonic and hypotonic buffers showed that the effect of PMS on RBC microheology is not mediated by cellular dehydration. Independent confirmation of the membrane stiffening effect of PMS was obtained by ektacytometric analysis of resealed RBC ghosts, with sickle-like increases in membrane rigidity observed between 50 and 100 mumol/L PMS. The rigidity of these ghosts was partially ameliorated by exposure to a thiol reductant. In terms of biochemical abnormalities, treated RBCs became significantly different from control RBCs at 25 mumol/L PMS, at which point they just began to enter the sickle range for amounts of membrane thiol oxidation and membrane-associated heme. The sickle average was achieved at 50 mumol/L PMS (for thiol oxidation) to 100 mumol/L PMS (for membrane heme). Thus, micromolar concentrations of PMS induce abnormalities of membrane microrheology that closely mimic those of unmanipulated sickle RBCs while reproducing similar degrees of oxidative biochemical change. We conclude that membrane protein oxidation could explain existence of an irreversible component to the abnormal rheology of the sickle membrane.  相似文献   

3.
R P Hebbel  W J Miller 《Blood》1984,64(3):733-741
Hemolytic anemia in sickle disease involves both intravascular and extravascular destruction of erythrocytes. Since the latter presumably involves the reticuloendothelial system, we have examined interactions between sickle erythrocytes and macrophages. In erythrophagocytosis assays, 18.9 +/- 7.2% of human marrow macrophages ingest sickle RBCs, while only 3.1 +/- 2.1% ingest normal RBCs. This abnormality is not explained by reticulocytosis, and it is strongly dependent upon RBC density. The interaction between sickle RBCs and macrophages appears to be partly immunologic, since it is partially blocked by Fc receptor blockade. Also, admixture of sickle RBCs (pretreated with rabbit anti- human-Ig) and Fc-receptor-bearing K562 cells results in 15.6 +/- 10.6% K562-RBC rosette formation compared with only 0.5 +/- 1.2% for normal RBCs. Regarding other factors that might promote erythrophagocytosis, sickle RBCs are found to spontaneously generate twice-normal amounts of dialdehyde byproducts of lipid peroxidation ("malondialdehyde" or MDA). Peroxide or reagent-MDA treatment of normal RBCs significantly enhances their phagocytosis, and MDA is at least 50 times more potent than other aldehydes studied here. Oxidative and immunologic effects may be related, since exposure of MDA-treated RBCs to immunoglobulin- containing human sera results in a further significant enhancement of erythrophagocytosis. For comparison of different sickle patients, an adherence assay demonstrates that sickle RBCs are 1.03 to 6.85 times more adherent to macrophages than are normal RBCs, and degree of adherence correlates significantly with irreversibly sickled cell (ISC) counts and hematologic variables reflecting hemolytic rate. We conclude that propensity for RBC interaction with macrophages is likely to be a determinant of hemolytic rate in sickle disease. Pertinent mechanisms appear to involve modification of RBC membranes by dialdehyde byproducts of excessive autoxidation and the abnormal acquisition of surface immunoglobulin on sickle RBCs, although participation of other membrane defects has not been excluded. Interestingly, the data further suggest the possibility that appearance of the "senescence antigen" in old normal RBCs represents modification of the membrane by "MDA."  相似文献   

4.
Kuross  SA; Hebbel  RP 《Blood》1988,72(4):1278-1285
Previous studies documented the abnormal association of heme and heme proteins with the sickle RBC membrane. We have now examined RBC ghosts and inside-out membranes (IOM) for the presence of nonheme iron as detected by its formation of a colored complex with ferrozine. Sickle ghosts have 33.8 +/- 18.2 nmol nonheme iron/mg membrane protein, and sickle IOM have 4.3 +/- 3.0 nmol/mg. In contrast, normal RBC ghosts and IOM have no detectable nonheme iron. The combination of heme and nonheme iron in sickle IOM averages nine times the amount of membrane- associated iron in normal IOM. Kinetics of the ferrozine reaction show that some of this nonheme iron on IOM reacts slowly and is probably in the form of ferritin, but most (72% +/- 18%) reacts rapidly and is in the form of some other biologic chelate. The latter iron compartment is removed by deferoxamine and by treatment of IOM with phospholipase D, which suggests that it represents an abnormal association of iron with polar head groups of aminophospholipids. The biologic feasibility of such a chelate was demonstrated by using an admixture of iron with model liposomes. Even in the presence of tenfold excess adenosine diphosphate, iron partitions readily into phosphatidylserine liposomes; there is no detectable association with phosphatidylcholine liposomes. To examine the bioavailability of membrane iron, we admixed membranes and t-butylhydroperoxide and found that sickle membranes show a tenfold greater peroxidation response than do normal membranes. This is not due simply to a deficiency of vitamin E, and this is profoundly inhibited by deferoxamine. Thus, while thiol oxidation in sickle membranes previously was shown to correlate with heme iron, the present data suggest that lipid peroxidation is related to nonheme iron. In control studies, we did not find this pathologic association of nonferritin, nonheme iron with IOM prepared from sickle trait, high-reticulocyte, postsplenectomy, or iron-overloaded individuals. These data provide additional support for the concept that iron decompartmentalization is a characteristic of sickle RBCs.  相似文献   

5.
Spectrin oxidation correlates with membrane vesiculation in stored RBCs   总被引:2,自引:0,他引:2  
Wagner  GM; Chiu  DT; Qju  JH; Heath  RH; Lubin  BH 《Blood》1987,69(6):1777-1781
An increase in spectrin oxidation in a variety of erythrocytes displaying a tendency to vesiculate has been previously described. To explore this relationship in more detail, we have studied blood stored in citrate-phosphate-dextrose-adenine under blood bank conditions because, in this system, vesiculation occurs slowly. Vesiculation was quantitated by measuring acetylcholinesterase release, and the extent of spectrin oxidation was detected by using thiol-disulfide exchange chromatography. A strong correlation (r = .92) was found between the extent of spectrin oxidation and vesiculation when blood from five donors was analyzed at weekly intervals during storage. This strongly suggests that spectrin oxidation plays a role in the formation of spectrin-free vesicles, thereby limiting the shelf life of stored blood.  相似文献   

6.
Adhesion of sickle neutrophils and erythrocytes to fibronectin   总被引:2,自引:0,他引:2  
Kasschau  MR; Barabino  GA; Bridges  KR; Golan  DE 《Blood》1996,87(2):771-780
The pathophysiology of vaso-occlusive crisis in sickle cell disease involves interactions among blood cells, plasma proteins, and vessel wall components. The initial goal of this work was to quantify the adhesion of sickle red blood cells (RBCs) to fibronectin immobilized on glass under both static and dynamic shear stress conditions. High-power microscopic inspection of static assay plates showed striking numbers of adherent neutrophils as well as RBCs. Sickle neutrophils and RBCs were significantly more adherent to fibronectin than the corresponding normal cells in static adhesion assays. Adhesion of both sickle neutrophils and sickle RBCs in dynamic adhesion assays was promoted by a period of static incubation preceding initiation of shear stress conditions. Adherent neutrophils remained attached at shear stresses up to 51 dyne/cm2; most adherent RBCs were attached at shear stresses up to 13 dyne/cm2, but detached at a shear stress of 20 dyne/cm2. Sickle neutrophil adhesion was enhanced significantly by autologous plasma. Elevated levels of plasma interleukin-6 (IL-6; but not IL-1 or IL-8) were found in 6 of 9 sickle cell disease samples examined, and elevated levels of tumor necrosis factor were found in 2 of 9 samples. Plasma IL- 6 levels correlated positively with both the number of sickle neutrophils adherent to fibronectin and the ability of sickle plasma to enhance adhesion of normal neutrophils to fibronectin. These data suggest possible roles for neutrophil activation and for fibronectin in mediating sickle neutrophil and RBC adhesion.  相似文献   

7.
Johnson  RM; Ravindranath  Y; el-Alfy  M; Goyette  G Jr 《Blood》1994,83(4):1117-1123
Chronic nonspherocytic hemolytic anemia has been observed in a recently described glucose-6-phosphate dehydrogenase (G6PD) variant, G6PDWayne. The mechanical properties of these erythrocytes and other G6PD variants were examined. The deformability of G6PD-deficient erythrocytes was normal, as determined by osmotic scan ektacytometry, and was not significantly affected by hemolytic crisis. In the common varieties of G6PD deficiency, the mechanical stability of the red blood cell (RBC) membrane was greater than normal, but G6PDWayne membranes were abnormally susceptible to shear-induced fragmentation. There was no evidence for a concurrent genetic defect in spectrin, because self- association constants and tryptic digests were normal. The fragility of G6PDWayne membranes appeared to be a consequence of oxidative damage to membrane thiol groups associated with a low glutathione (GSH) level in these RBCs. Associations among GSH level, thiol oxidation, and membrane instability were also found when a larger group of G6PD-deficient RBCs were examined. In normal erythrocytes, 1-chloro-2,4-dinitrobenzene was used to reduce GSH levels by 50%. Membrane thiol oxidation and membrane fragility both increased when these cells were kept at 4 degrees C for 3 to 5 days. Our findings suggest that chronic depletion of GSH leads to the destabilization of membrane skeleton through oxidation of membrane protein thiols.  相似文献   

8.
Allen  DW; Manning  N 《Blood》1994,84(4):1283-1287
Spur cell anemia is a hemolytic anemia seen in severe alcoholic cirrhosis that is characterized by unusual morphology and a decreased ratio of phospholipids to cholesterol in the erythrocyte membrane. We hypothesized that defective phospholipid repair may contribute to the red blood cell (RBC) phospholipid abnormalities of spur cell anemia. Therefore, we compared RBCs from normal control subjects with RBCs from spur cell anemia patients. The incorporation of [14C] arachidonic acid into the phospholipids and acylcarnitine (acyl-Cn) of spur cells and normal RBCs was analyzed by a direct-phase high performance liquid chromatography column to separate both the phospholipids and acyl-Cn. There was less uptake of the [14C] arachidonate into phosphatidylethanolamine of spur cell RBCs (12.9% +/- 1.0%) compared with normal RBCs (20.5% +/- 2.8%; P = .0245). However, more arachidonate was incorporated into the acyl-Cn of spur cells (spur cell acyl-Cn [24.5% +/- 2.9%] v normal control acyl-Cn [10.1% +/- 1.9%]; P = .0018). We conclude that phospholipid biosynthesis is inhibited and that acyl-Cn formation is spared in spur cell anemia RBCs. These metabolic changes may help account for the lipid abnormalities seen in spur cell anemia RBCs and contribute to the hemolytic process.  相似文献   

9.
Red blood cell (RBC) membranes from patients with the thalassemic and sickle hemoglobinopathies carry abnormal deposits of iron presumed to mediate a variety of oxidative-induced membrane dysfunctions. We hypothesized that the oral iron chelator deferiprone (L1), which has an enhanced capacity to permeate cell membranes, might be useful in chelating these pathologic iron deposits from intact RBCs. We tested this hypothesis in vitro by incubating L1 with RBCs from 15 patients with thalassemia intermedia and 6 patients with sickle cell anemia. We found that removal of RBC membrane free iron by L1 increased both as a function of time of incubation and L1 concentration. Thus, increasing the time of incubation of thalassemic RBCs with 0.5 mmol/L L1 from 0.5 to 6 hours, enhanced removal of their membrane free iron from 18% +/- 9% to 96% +/- 4%. Dose-response studies showed that incubating thalassemic RBC for 2 hours with L1 concentrations ranging from 0.125 to 0.5 mmol/L resulted in removal of membrane free iron from 28% +/- 15% to 68% +/- 11%. Parallel studies with sickle RBCs showed a similar pattern in time and dose responses. Deferoxamine (DFO), on the other hand, was ineffective in chelating membrane free iron from either thalassemic or sickle RBCs regardless of dose (maximum, 0.333 mmol/L) or time of incubation (maximum, 24 hours). In vivo efficacy of L1 was shown in six thalassemic patients whose RBC membrane free iron decreased by 50% +/- 29% following a 2-week course of L1 at a daily dose of 25 mg/kg. As the dose of L1 was increased to 50 mg/kg/d (n = 5), and then to 75 mg/kg/d (n = 4), 67% +/- 14% and 79% +/- 11%, respectively, of their RBC membrane free iron was removed. L1 therapy-- both in vitro and in vivo--also significantly attenuated the malondialdehyde response of thalassemic RBC membranes to in vitro stimulation with peroxide. Remarkably, the heme content of RBC membranes from L1-treated thalassemic patients decreased by 28% +/- 10% during the 3-month study period. These results indicate that L1 can remove pathologic deposits of chelatable iron from thalassemic and sickle RBC membranes, a therapeutic potential not shared by DFO. Furthermore, membrane defects possibly mediated by catalytic iron, such as lipid peroxidation and hemichrome formation, may also be alleviated, at least in part, by L1.  相似文献   

10.
Test  ST; Woolworth  VS 《Blood》1994,83(3):842-852
A prominent clinical manifestation of sickle cell disease (SCD) is hemolytic anemia. Although complement activation can lead to intravascular hemolysis, its role in the hemolysis of SCD is not known. Because normal red blood cells induced to vesiculate by treatment with calcium and ionophore become sensitive to damage by activated complement and because sickle cells release microvesicles as they circulate, we postulated that sickle cells might also be unusually sensitive to complement-dependent hemolysis. Complement activation is tightly regulated on the membrane of the normal erythrocyte; therefore, defective complement regulation by the sickle cell would be necessary for complement-dependent hemolysis to occur. These studies show a defect in the regulation of membrane attack complex (C5b-9) formation in sickle erythrocytes, particularly in the most dense cells. The defect is characterized by increased binding of C5b-7 and of C9 to denser sickle cells and results in increased susceptibility of sickle cells to C5b-9-mediated (reactive) lysis initiated by either C5b6 or activated cobra venom factor. Among the densest sickle cells, irreversibly sickled cells are especially sensitive to reactive lysis. The similarity of this defect to that previously described in a patient with paroxysmal nocturnal hemoglobinuria suggests that complement- mediated hemolysis could play a role in the anemia of SCD.  相似文献   

11.
Oxygen uptake of fully deoxygenated sickle (SS) erythrocytes is slower than that of normal (AA) erythrocytes, as demonstrated by the half-times of the overall oxygenation reactions: at 25 degrees in an isotonic phosphate buffer the normal red cells have a t1/2 = 82 +/- 4.7 msec, as compared to sickle red cells where t1/2 = 135 +/- 17.6 msec. The effects of temperature, extracellular osmolality, and the presence of an antisickling agent (n-butylurea) on the rate of red cell oxygenation strongly suggest that the differences in oxygenation rates encountered with sickle red cells is directly related to the intracellular polymerization of deoxyhemoglobin S.  相似文献   

12.
Kuross  SA; Rank  BH; Hebbel  RP 《Blood》1988,71(4):876-882
It has been suggested that the development of sickle RBC membrane defects might be related to abnormal amounts of membrane-associated heme (a term we use in its generic sense to include hemoglobins, hemichromes, and free heme). Techniques previously used to measure membrane heme, however, would not distinguish between what is truly membrane-associated and what is merely trapped in RBC ghost preparations. Consequently, we have examined extensively washed inside- out membranes (IOM) prepared from normal and sickle RBC. Approximately 25% of the sickle ghost heme is lost upon conversion to IOM, but sickle IOM still have a significant excess (1.6 +/- 0.3 nmol heme/mg membrane protein compared with 0.7 +/- 0.2 nmol/mg for normal IOM, P less than .001). Amounts of ghost heme are only poorly predictive of amounts of IOM heme (r = .664). Preparation of IOM by using isotonic lysis with saponin yields virtually identical amounts of IOM heme. Small amounts of heme (less than 15%) can be displaced from IOM by using manipulations that elute spectrin, displace electrostatically bound proteins, or cleave the cytoplasmic portion of band 3. Treatment of IOM with dithiothreitol (DTT), however, displaces the most heme (35%), and this is almost reproduced (25% displacement) by the treatment of intact RBC with DTT before IOM preparation. Sequential treatment with all manipulations still leaves about 40% of the heme in sickle IOM, which indicates a compartment more intimately associated with the membrane. At least part of this is free heme without globin, as evidenced by abnormal binding of radiochloroquine to sickle IOM. Conversely, some IOM-associated globin is globin without heme because the measurement of globin per se markedly overpredicts amount of IOM heme. There is a strong correlation between RBC density and amounts of either ghost or IOM heme. Finally, the amount of membrane thiol oxidation (as measured by thiol-disulfide-exchange chromatography) does not correlate at all with ghost heme (r = .105), but it correlates well with IOM heme (r = .877, P less than .001). These data demonstrate that there are abnormal amounts of heme truly associated with sickle RBC membranes, and they are consistent with the hypothesis that this membrane-associated heme participates in the pathobiology of the sickle RBC membrane, particularly those aspects perhaps related to thiol oxidation.  相似文献   

13.
The effect of hypothermic cardiopulmonary bypass techniques on the sickling process was evaluated in patients with sickle cell hemoglobin. It was presumed that intraoperative hemolysis, as identified by hemoglobinuria, reflected increased sickling. Data of 43 patients with sickle cell traits and 2 with sickle cell disease, who were operated on under cardiopulmonary bypass and cold cardioplegic arrest in a tertiary center from the beginning of 1995 to the end of 2004, were retrospectively analyzed. A mean nasal temperature of 30.8 degrees C +/- 2.1 degrees C was achieved. Three patients with sickle cell trait developed intraoperative hemoglobinuria, albeit with normal surrogate values for hemolysis. However, they had significantly lower mean hemoglobin levels during cardiopulmonary bypass compared to those sickle cell patients who did not exhibit hemoglobinuria (hemoglobin, 6.0 +/- 0.2 vs. 7.4 +/- 0.9 g x dL(-1), p < 0.01). Total drainage and blood transfusion requirements in patients with normal and sickle cell hemoglobin were similar. It was concluded that hypothermic cardiopulmonary bypass with cold cardioplegia is safe in sickle cell patients. Maintenance of adequate hemoglobin levels during cardiopulmonary bypass may be important to avoid triggering a sickling process.  相似文献   

14.
T Sugihara  R P Hebbel 《Blood》1992,80(9):2374-2378
An abnormal susceptibility of the sickle red blood cell (RBC) membrane to deformation could compromise its permeability barrier function and contribute to the exuberant cation leakiness occurring during the sickling phenomenon. We examined this hypothesis by subjecting RBCs at ambient oxygen tension to elliptical deformation, applying shear stress in a viscous medium under physiologic conditions. Compared with normal and high-reticulocyte control RBCs, sickle RBCs manifest an exaggerated K leak response to deformation. This leak is fully reversible, is both Cl and Ca independent, and at pHe 7.4 is fully balanced so that Kefflux equals Nainflux. This abnormal susceptibility is also evident in that the K leak in response to deformation occurs at an applied shear stress of only 141 dyne/cm2 for sickle RBCs, as compared to 204 dyne/cm2 for normal RBCs. Fresh sickle RBC membranes contain elevated amounts of lipid hydroperoxide, the presence of which is believed to provide the biochemical basis for enhanced deformation susceptibility. When examined at pHe 6.8, oxygenated sickle RBCs acquire an additional, unbalanced (Kefflux > Nainflux) component to the K leak increment specifically ascribable to deformation. Studies with inhibitors suggest that this additional component is not caused by a known leak pathway (eg, either K:Cl cotransport or the Gardos channel). This abnormal susceptibility of the sickle membrane to development of cation leakiness during deformation probably contributes to the exuberant cation leak taking place during RBC sickling.  相似文献   

15.
Liu  SC; Derick  LH; Palek  J 《Blood》1993,81(2):522-528
Red blood cells (RBCs) in sickle cell anemia, transformed into a sickled shape by prolonged deoxygenation, or normal RBCs deformed by a prolonged micropipette aspiration become permanently stabilized in their abnormal shape. This semisolid plastic behavior is thought to involve an irreversible reorganization of the membrane skeleton, but the exact nature of this skeletal rearrangement is not known. In this study, we first asked whether the irreversible deformation is associated with a permanent stretching of the skeletal network, and then whether it is due to a rearrangement of skeletal components involving a disruption of pre-existing protein associations and the subsequent reassociation of new protein contacts. Having found no ultrastructural evidence of stretching of the skeletal lattice in membranes derived from permanently deformed RBCs, we addressed the possibility of reorganization of the proteins of the membrane skeleton. We examined the temperature dependence of irreversible cell deformation to see if it correlated with the known temperature dependence of spectrin tetramers to dimer dissociation and reassociation. Testing the shape irreversibility of both deoxygenated reversibly sickled cells and Nucleopore-aspirated normal cells, we found that both types of cells became permanently deformed when the prolonged incubation of applied force or deoxygenation was performed at 37 degrees C, the temperature at which spectrin tetramers were free to dissociate and reassociate. In contrast, both types of cells were able to regain their original discocytic shape if the prolonged incubation was performed at the lower temperature: at less than 13 degrees C instead of 37 degrees C. Furthermore, normal RBCs were incubated with inosine and pyruvate to elevate intracellular 2,3-diphosphoglycerate, the polyanion shown to destabilize spectrin-actin-protein 4.1 association. This did not result in a promotion of irreversible deformation of these cells. We conclude that the irreversible cell deformation observed at physiologic temperature is associated with a skeletal rearrangement through dissociation of spectrin tetramers to dimers and a subsequent reassociation of dimers to tetramers in the new (deformed) configuration. These findings may explain a permanent stabilization of irreversibly sickled cells in their abnormal shape in vivo.  相似文献   

16.
BACKGROUND: Most mammals experience cardiac arrest during hypothermia. In contrast, hibernators remain in sinus rhythm even at body temperatures of 0 degrees C. OBJECTIVES: The purpose of this study was to quantify electrical activity and connexin expression in the heart of hibernating Siberian ground squirrel Citellus undulatus. METHODS: Optical imaging and microelectrode recordings were conducted in Langendorff-perfused hearts and isolated papillary muscles of summer active (SA, n = 19), winter hibernating (WH, n = 21), interbout arousal (IBA, n = 12), and winter active (WA, n = 3) ground squirrels and rabbits (n = 14) at temperatures from +37 degrees C to +3 degrees C. RESULTS: All studied SA and WH hearts maintained spontaneous sinus rhythm, safe propagation through the entire conduction system, and normal pattern of ventricular excitation at all temperatures. However, three of the seven IBA and all rabbit hearts lost excitability at 10 degrees C +/- 1 degrees C and 12 degrees C +/- 1 degrees C, respectively. In WH, SA, and IBA ground squirrels, temperature reduction from 37 degrees C to 3 degrees C resulted in a 10-fold slowing of ventricular conduction velocity and increased excitation threshold. At any temperature, WH ventricles had faster conduction velocity and lower excitation threshold compared with SA and IBA. Immunolabeling demonstrated that connexin43 (Cx43) was significantly up-regulated in WH and WA compared with SA myocardium: Cx43 area density was 12.4 +/- 1.3, 15.0 +/- 3.0 and 8.6 +/- 1.1 microm(2)/1,000 microm(2), respectively. Moreover, Cx45 was expressed in the WH but not in the SA or WA ventricles. CONCLUSION: Hibernator Citellus undulatus has evolved to maintain safe conduction at extreme hypothermia via up-regulation of Cx43 and Cx45 in order to protect the heart against arrhythmia associated with hypothermia.  相似文献   

17.
Due to their role in oxygen transport and the presence of redox active haemoglobin molecules, red blood cells (RBC) generate relatively high levels of reactive oxygen species (ROS). To counteract the potential deleterious effects of ROS, RBCs have a well-integrated network of anti-oxidant mechanisms to combat this oxidative stress. ROS formation is increased in sickle-cell disease (SCD) and our studies in a murine SCD model showed a significant increase in the generation of ROS when compared with normal mice. Our data also indicated that murine sickle RBCs exhibit a significantly increased ATP catabolism, partly due to the increased activity of glucose-6-phosphate dehydrogenase and glutathione reductase to regenerate intracellular glutathione (GSH) levels to neutralize the adverse milieu of oxidative stress. Higher ATP consumption by the murine sickle RBCs, together with the increased ROS formation and impairment of the aminophospholipid translocase or flipase may underlie the exposure of phosphatidylserine on the surface of these cells.  相似文献   

18.
The decrease in Na/K adenosine triphosphatase (ATPase) activity observed in several tissues of type 1 diabetic patients is thought to play a role in the development of long-term complications. Infusion of insulin may restore this enzyme activity in red blood cells (RBCs), and recent arguments have been developed for a similar role of C-peptide. The aims of this study were to determine whether insulin acts directly on the RBC enzyme and to evaluate the effect of C-peptide on Na/K ATPase activity. Thirty-nine C-peptide-negative type 1 diabetic patients were studied (blood glucose, 11.2 +/- 1.49 mmol/L; hemoglobin A1c [HbA1c], 8.9% +/- 0.1%, mean +/- SEM). Blood samples were obtained in the morning, before breakfast and insulin injection. Intact and living RBCs were resuspended in their own plasma and incubated with or without insulin (50 microU/mL) or C-peptide (6 nmol/L). Ex vivo by microcalorimetry, the heat produced after 1 hour by the enzyme-induced hydrolysis of adenosine triphosphate (ATP), was measured in a thermostated microcalorimeter at 37 degrees C. The results showed that Na/K ATPase activity was significantly increased by insulin (12.4 +/- 0.5 v 15.4 +/- 0.9 mW/L RBCs, P < .05, n = 23) but not by C-peptide (11.9 +/- 0.7 v 12.9 +/- 0.9 mW/L RBCs, NS, P = .26, n = 12). In another experiment, RBC suspensions were incubated at 37 degrees C in a water bath with or without insulin (50 microU/mL) or C-peptide (6 nmol/L) for 10 minutes. RBC membranes were isolated and Na/K ATPase activity was assessed by measuring inorganic phosphate release at saturating concentrations of all substrates. The results showed that insulin and C-peptide significantly increased RBC Na/K ATPase activity (342 +/- 25, P < .005 and 363 +/- 30, P < .005, respectively v255 +/- 22 nmol Pi x mg protein(-1) x h(-1), n = 14). We conclude that insulin and C-peptide act directly on RBC Na/K ATPase, thus restoring this activity in type 1 diabetic patients. The stimulatory effect of C-peptide observed in vitro on RBC Na/K ATPase activity confirms that C-peptide plays a physiological role.  相似文献   

19.
Diaz  C; Morkowski  J; Schroit  AJ 《Blood》1996,87(7):2956-2961
In vitro stored red blood cells (RBC) and RBC artificially induced to vesiculate by incubation with dilauroylphosphatidyl-choline were monitored for age- and vesiculation-dependent changes in cell density, membrane lipid asymmetry, and their ability to be recognized and cleared by reticuloendothelial cells. RBC demonstrated a progressive increase in density on self-forming Percoll gradients upon vesiculation and in vitro "aging." Uptake of vesiculated RBC by in vitro cultivated macrophages was increased threefold over non-vesiculated control RBC. The clearance rate of dense vesiculated RBC was biphasic and contained a rapid component and a slower second component consistent with the clearance rates of normal control populations. Determination of phosphatidylserine (PS) in the outer leaflet of RBC by the PS-dependent prothrombinase assay revealed that PS redistributed to the cell's outer leaflet upon in vitro storage and vesiculation. Inhibition of PS movement by oxidation of membrane sulfhydryls with pyridyldithioethylamine resulted in higher prothrombinase levels and enhanced clearance of vesiculated RBC. These experiments suggest that vesiculation contributes to alterations in membrane lipid asymmetry and cell density characteristic of the aged RBC phenotype.  相似文献   

20.
Multicomponent apheresis procedures offer the possibility to collect standardized blood components as compared to whole blood donations. A new program for the concurrent collection of platelets (PLTs) and red blood cells (RBCs) was evaluated in a prospective study. Apheresis donors ( n=18) underwent concurrent collection of PLTs and RBCs using the Haemonetics MCS+ blood cell separator. Aliquots of PLTs and RBCs were collected during five to six passes of the discontinuous flow procedure. The platelet product was in-line filtered during the last pass of the separation procedure. After collection, saline-adenine-glucose-mannitol (SAGM) preservative solution was automatically added to the RBCs. Thereafter, the RBCs were in-line leukodepleted by gravity filtration at room temperature. The PLTs and RBCs were subsequently stored at 22+/-2 degrees C for 5 days and 4+/-2 degrees C for 35 days, respectively. The following in vitro parameters were evaluated over the storage periods: blood cell counts, glucose, lactate, lactate dehydrogenase, pH, plasma hemoglobin, and potassium. Two ready-to-use blood components from one donor were collected in an average procedure time of 86+/-10 min; 2.47+/-0.74 x 10(11) PLTs were collected in a product volume of 232+/-43 ml. The RBC volume averaged 280+/-20 ml and the hemoglobin content was 56.8+/-2.4 g per unit. The leukocyte contamination of the platelet product was 0.44+/-0.56 x 10(5) and the residual leukocyte content of the RBC product was 0.28+/-0.02 x 10(5). Storage data showed no relevant drop in pH. Day 35 results of the RBC products showed that all of the units had less than 0.8% hemolysis. Standardized PLT and RBC products of good quality can be concurrently collected with the MCS+ blood cell separator. In vitro testing of the products collected and stored for 5 and 35 days, respectively, met the Council of Europe criteria for leukodepleted blood products.  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