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 19 毫秒
1.
HAEMOGLOBIN Hiroshima is a variant with interesting physiological properties1,2 discovered in a Japanese family. Its Bohr effect is halved, its oxygen affinity at physiological pH increased about three-fold and haem-haem interaction is somewhat reduced compared with normal haemoglobin. In 0.1 M NaCl solutions initially stripped of phosphate, 2,3-diphospho-glycerate (2,3-DPG) diminishes the oxygen affinity as in haemoglobin A (H. F. Bunn, unpublished results). The amino-acid substitution originally deduced for this abnormal haemoglobin was histidine 143 (H21)β ? aspartic acid1. It was possible to conceive of a mechanism which accounted for its diminished Bohr effect3, but the normal response of its oxygen affinity to 2,3-DPG was inconsistent with the proposed role of histidine 143 in 2,3-DPG binding by haemoglobin A4,5. An X-ray crystallographic study of deoxyhaemoglobin Hiroshima has now revealed that the replacement occurs not in position 143 but 146β. This was confirmed by chemical methods and the physiological properties of this haemoglobin are now satisfactorily accounted for. The results support the role of histidine 146β in the alkaline Bohr effect6.  相似文献   

2.
2,3-Diphosphoglycerate (2,3-DPG), an intracellular metabolite of glycolytic pathway is known to affect the oxygen binding capacity of haemoglobin and mechanical properties of the red blood cells. 2,3-DPG levels have been reported to be elevated during anaemic conditions including visceral leishmaniasis. 2,3-DPG activity in P. falciparum infected red blood cells, particularly in cells infected with different stages of the parasite and its relationship with structural integrity of the cells is not known. Chloroquine sensitive and resistant strains of P. falciparum were cultured in vitro and synchronized cultures of ring, trophozoite and schizont stage rich cells along with the uninfected control erythrocytes were assayed for 2,3-DPG activity and osmotic fragility. It was observed that in both the strains, in infected erythrocytes the 2,3-DPG activity gradually decreased and osmotic fragility gradually increased as the parasite matured from ring to schizont stage. The decrease in 2,3-DPG may probably be due to increased pyruvate kinase activity of parasite origin, which has been shown in erythrocytes infected with several species of Plasmodium. The absence of compensatory increase in 2,3-DPG in P. falciparum infected erythrocytes may aggravate hypoxia due to anaemia in malaria and probably may contribute to hypoxia in cerebral malaria. As 2,3-DPG was not found to be increased in erythrocytes parasitized with P. falciparum, the increased osmotic fragility observed in these cells is not due to increased 2,3-DPG as has been suggested in visceral leishmaniasis.  相似文献   

3.
The effects of 2,3 diphosphoglyceric acid (2,3-DPG), adenosine triphosphate (ATP), and inositol hexaphosphate (IHP) on the oxygen affinity of whole “stripped” hemoglobin (WSH), hemoglobin H (Hb-H), hemoglobin A (Hb-A) and hemoglobin D (Hb-D) isolated from 18-day chick embryo blood have been determined. The effect of the three organic phosphates upon the oxygen dissociation curves is similar and the following order of decreasing oxygen affinity of the organic phosphates was observed for each hemoglobin: 2,3-DPG < ATP < IHP. 2,3-DPG appears to have a slightly greater effect upon the P50 of Hb-H than upon that of either of the two adult-type hemoglobins. However, this effect seems insufficient to suggest a preferential interaction of 2,3-DPG with Hb-H which would account for either the large amounts of 2,3-DPG in the erythrocytes of embryos or the higher oxygen affinity of the whole blood. The effects of the organic phosphates upon the Hill constant of the purified hemoglobins are variable. It is concluded that since the distribution of hemoglobins H, A, and D in the erythrocytes during the developmental period from 18-day embryos to 6-day chicks remains fairly constant, the previously described progressive decrease in oxygen affinity of the whole blood during this period results from changes in the total amount and distribution of the intraerythrocytic organic phosphates.2  相似文献   

4.
THERE is a decrease in the O2 affinity of mammalian haemoglobin (Hb) as the levels of 2,3-DPG or ATP are increased, which is explained by an allosteric effect on the HbO2 binding1,2. Similar observations on amphibians3 and fish4, which have molar ratios of ATP to Hb similar to those of DPG to Hb in mammals, suggest that red cell organic phosphates modulate Hb function in all vertebrates. The adaptation of mammals to various hypoxic stresses involves reduced HbO2 affinity5–9, the attendant increase in O2 “unloading” capacity being mediated by an increase in the concentration of red cell 2,3-DPG. We have found the opposite response in hypoxic fish and suggest that an increased O2 affinity results in increased O2 transport for the fish.  相似文献   

5.
The changes in structure and function of 2,3-diphosphoglycerate-hemoglobin (2,3-DPG-Hb) induced by Ln3+ binding were studied by spectroscopic methods. The binding of lanthanide cations to 2,3-DPG is prior to that to Hb. Ln3+ binding causes the hydrolysis of either one from the two phosphomonoester bonds in 2,3-DPG non-specifically. The results using the ultrafiltration method indicate that Ln3+ binding sites for Hb can be classified into three categories: i.e. positive cooperative sites (NI), non-cooperative strong sites (NS) and non-cooperative weak sites (NW) with binding constants in decreasing order: KI>KS>KW. The total number of binding sites amounts to about 65 per Hb tetramer. Information on reaction kinetics was obtained from the change of intrinsic fluorescence in Hb monitored by stopped-flow fluorometry. Fluctuation of fluorescence dependent on Ln3+ concentration and temperature was observed and can be attributed to the successive conformational changes induced by Ln3+ binding. The results also reveal the bidirectional changes of the oxygen affinity of Hb in the dependence on Ln3+ concentration. At the range of [Ln3+]/[Hb]<2, the marked increase of oxygen affinity (P50 decrease) with the Ln3+ concentration can be attributed to the hydrolysis of 2,3-DPG, while the slight rebound of oxygen affinity in higher Ln3+ concentration can be interpreted by the transition to the T-state of the Hb tetramer induced by Ln3+ binding. This was indicated by the changes in secondary structure characterized by the decrease of α-helix content.  相似文献   

6.
The kinetics of 2,3-diphosphoglycerate (2,3-DPG) net breakdown was examined in intact human erythrocytes incubated at pH 7.00 and 37 °C. The concentrations of 2,3-DPG, 1,3-diphosphoglycerate (1,3-DPG), 3-phosphoglycerate, ATP, Pi, glucose, and lactate were determined during 10 to 12 h. Since the concentration of 1,3-DPG has been suggested to be the main regulating factor with respect to the rate of 2,3-DPG net breakdown the interdependence between the concentration of 1,3-DPG and pH was determined in the range of pH 6.9 to 7.4. It was found that the stationary level of 1,3-DPG decreased strongly with decreasing pH within this range. Qualitatively, the net breakdown of 2,3-DPG observed at pH 7.00 can be explained by the lowered level of 1,3-DPG. The influence of the concentration of Pi upon the rate of net degradation of 2,3-DPG at pH 7.00 was studied at low cell volume fraction (0.04), where given concentrations of Pi could be maintained for several hours. A marked increase in the rate of 2,3-DPG net breakdown by Pi was demonstrated. Computer simulations showed that activation of diphosphoglycerate phosphatase by the increasing concentration of Pi and decrease of degree of inhibition of the diphosphoglycerate mutase by the decreasing concentration of 2,3-DPG may well keep the rate of the degradation balanced at the time constant value observed. On the basis of the observed kinetics and a computer simulation, the flux through the phosphoglycerate bypass was estimated to be 10 to 15% of the total glycolytic flux at physiological conditions.  相似文献   

7.
  • 1.1. A cytoplasmic casein kinase II (CKII) has been purified more than 10,000-fold from Artemia sp.
  • 2.2. The reaction mechanism of the cytoplasmic CKII was determined to be random bi bi, using ATP and casein as substrates, which is in agreement with the results obtained for a DrosophilaCKII [Glover, Shelton and Brutlag (1983) J. biol. Chem.258, 3258–3265]. Km values for ATP and casein are 8 μM and 0.2 mg/ml respectively. The binding of either substrate lowers the enzyme-affinity for the other by a factor α = 1.65.
  • 3.3. In vitro, the enzyme is inhibited by poly(A)2-mRNA and 2,3-diphosphoglyceric acid (2,3-DPG). The inhibition by 2,3-DPG is due to competition with the protein substrate.
  • 4.4. The possible in vivo effects of these inhibitors in CKII-mediated translational regulation is discussed.
  相似文献   

8.
The 2,3-diphosphoglycerate (2,3-DPG) concentration, oxygen half saturation pressure at pH 7.4 (P50), pH in plasma and red cells, and mean corpuscular hemoglobin concentration (MCHC) of venous blood were determined during unrestricted daily activity (series I) throughout 24 hrs as well as during prolonged bed rest until noon (series II). In series I almost synchronous dirunal behavior of P50 2,3-DPG, and plasma pH as well as red cell pH became significantly apparent with highest values in the afternoon. The [2,3-DPG] yielded most pronounced alterations, which made up to 13.5% of the average day value. During prolonged recumbency the [2,3-DPG] showed a nonsignificant tendency to decline; the P50 remained unchanged throughout that period. The possible reason for the missing [2,3-DPG] increase is a reduced change of red cell pH in series II. An influence of a posture dependent aldosterone secretion either directly on the 2,3-DPG metabloism of indirectly via mediating the red cell pH and thus ruling the formation of this organic PHOSPHORIS COMPOUND IS DISCUSSED.  相似文献   

9.
A systematic study of the pH dependent changes in the range 6.6--7.4 of 2,3 bisphosphoglycerate (2,3-DPG) was performed in the presence and absence of glucose during transitional and steady states. The results indicate that 2,3-DPGase breaks down 2,3-DPG nealy independent of pH at a rate of 480 mu moles 2,3-DPG/1 cells.h. The 2,3-DPG mutase is practically completely inhibited below pH 6.9. The 2,3-DPG level in the presence of glucose reaches a pH dependent steady state after about 18 h. The share of the 2,3-DPG bypass in the steady state decreases from 24% at pH 7.4 to 12% at pH 7.0. The formation of pyruvate corresponds to the beadkdown of 2,3-DPG after consumption of an unknown reducing substance.  相似文献   

10.
Red blood cell 2,3 diphosphoglycerate (2,3-DPG) levels increase after ascent to high altitude. Studies were undertaken to identify the biochemical mechanisms responsible for eliciting the 2,3-DPG response in several types of subjects. These included (1) short-term exposure to 3400 m in ten subjects; (2) exposure to 4300 m in an additional ten subjects; (3) studies in 28 high-altitude normal residents of 3100 m; and (4) studies in 28 high-altitude residents with chronic mountain polycythemia. Controls were 41 residents of 240 m. Regression analysis identified the glycolytic variables, termed “key variables,” on which variation in 2,3-DPG levels was dependent (P < .05). Key variables common to the short-term studies were glucose-6-phosphate, phosphoenolpyruvate, and the ratio of the levels of adenosine diphosphate to adenosine triphosphate. The positions of these key variables in the glycolytic pathway and their mean levels suggest erythrocyte hexokinase and pyruvate kinase activation as possible enzymatic mechanisms. Key variables unique to the 3400 m study suggested phosphofructokinase activation also acted to increase 2,3-DPG levels. 2,3-DPG levels in the normal 3100 m residents were not different from low-altitude values, and 2,3-DPG levels in these samples did not appear to be dependent on any of the glycolytic variables examined. Among the high-altitude residents with polycythemia, higher 2,3-DPG levels were dependent on glucose-6-phosphate, fructose diphosphate, dihydroxyacetone phosphate, and the ratio of adenosine diphosphate to adenosine triphosphate levels. The positions of these variables in the glycolytic pathway and their mean levels suggested activation of the hexokinase and phosphofructokinase enzymes.  相似文献   

11.
The early posttransfusion period after acute blood loss using erythromass without ozone was characterized by decreased electrophoretic mobility of erythrocytes (EPME), ATP, 2,3-diphosphoglycerate (2,3- DPG) concentration, Na+-K+-ATPase activity, and increased concentration of malonic dialdehyde (MDA) and catalase activity in erythrocytes. Most parameters gradually were restored over 5 days, but a reduced concentration of 2,3-DPG was observed throughout the entire experiment with transfusion of erythromass without ozone. Transfusion of the ozonized erythrocyte mass 1 h after its administration caused an increase in 2,3- DPG concentration in erythrocytes circulating in the vascular bed. Twenty-four hours later, increased catalase activity, EPPE, and ATP concentration were registered, and, after 5 days, enhanced Na+-K+-ATPase activity was registered. Their values were maintained until the end of the experiment at a higher level than in animals that had undergone transfusion of the nonozonized erythromass. Ozonized erythrocyte mass improves the functional metabolic state of erythrocytes, promotes an earlier recovery of the oxygen transport in blood, and limits the damage to the microcirculatory bed of the myocardium and cardiomyocytes during blood loss.  相似文献   

12.
2,3-Diphosphoglycerate (2,3-DPG) has been found to be the product of the enzymatic degradation of cyclic 2,3-diphosphoglycerate (cDPG) in the archaebacterium Methanobacterium thermoautotrophicum delta H. Although 2,3-DPG has not previously been detected as a major soluble component of M. thermoautotrophicum, large pools accumulated at an incubation temperature of 50 degrees C (below the optimum growth temperature of 62 degrees C). Under these conditions, cellular activity was significantly decreased; a return of the culture to the optimum growth temperature restored the 2,3-DPG pool back to original low levels and caused steady-state cDPG levels to increase again. While 13CO2-pulse/12CO2-chase experiments at 50 degrees C showed that the cDPG turned over, the appearance of 2,3-DPG at NMR-visible concentrations required at least 10 h. Production of 2,3-DPG in vivo was prevented by exposure of the cells to O2. The enzyme responsible for this hydrolysis of cDPG was purified by affinity chromatography and appears to be a 33-kDa protein. Activity was detected in the presence of oxygen and was enhanced by a solution of 1 M KCl, 25 mM MgCl2, and dithiothreitol. Both Km and Vmax have been determined at 37 degrees C; kinetics also indicate that in vitro the product, 2,3-DPG, is an inhibitor of cDPG hydrolysis. These findings are discussed in view of a proposed role for cDPG in methanogens.  相似文献   

13.
The physiological adaptation to anemia and other hypoxic states includes an increase in the level of 2,3-bisphosphoglycerate (2,3-DPG) in the red cell. We suggest that the high level of 2,3-DPG may have adverse effects in vivo. It has been found that red cells incubated with glycolate lose 2,3-DPG at a rapid rate relative to controls. ATP is stable. Net 2,3-DPG synthesis is observed after the glycolate is removed from the cells suggesting that they are not harmed. The effect appears to be specific for glycolate since lactate, glyoxylate, glycerate, acetate, and citrate were without effect. This procedure could be used to assess the effects of decreasing the 2,3-DPG level to normal in the erythrocytes of sickle cell and other anemias.  相似文献   

14.
A statistically significant 10% increase (p less than 0.005) in mean red cell 2,3-diphosphoglycerate (2,3-DPG) concentration, concomitantly with a mean 16% increase (p less than 0.001) in the predicted maximal oxygen uptake (VO2max) was observed in 29 recruits, who were studied during 6 months of physical training in military service. The increase in 2,3-DPG was higher, the lower the initial 2,3-DPG and VO2max levels. The mean initial 2,3-DPG level was higher in the subjects with a higher initial VO2max. A strenuous but highly aerobic 21-km marching exercise elicited a mean 9% increase (p less than 0.005) in red cell 2,3-DPG concentration. A significantly greater response of 2,3-DPG to marching exercise was observed in subjects with a lower pre-test VO2max than in those with a higher pre-test VO2max. During another more competitive march 2,3-DPG remained almost unchanged and was associated with a tendency towards a negative correlation with the acccompanying lactate response (r = -0.60, p less than 0.05). Red cell 2,3-DPG response to a standardized exercise is considered to be a suitable indicator for evaluating the effect of training on an individual.  相似文献   

15.
We have previously shown that 2,3-diphosphoglycerate (2,3-DPG) inhibits the phosphorylation of erythrocyte membrane cytoskeletal proteins by endogenous casein kinases. Here, we report that 2,3-DPG stimulates the phosphorylation of protein 4.1 by protein kinase C. Studies with red cell membrane preparations showed that while the phosphorylation of most of the membrane proteins by endogenous membrane-bound kinases and purified kinase C was inhibited by 2,3-DPG, the phosphorylation of protein 4.1 was slightly enhanced by the metabolite. The effect of 2,3-DPG was further examined using purified protein 4.1 preparations. Our results indicate that 2,3-DPG stimulates both the rate and the extent of phosphorylation of purified protein 4.1 by kinase C. The amount of phosphate incorporated was found to double to 2 mol of phosphate per mole of protein 4.1 in the presence of 10 mM 2,3-DPG. The increase in phosphorylation was distributed over all phosphorylation sites as revealed by an analysis of the labeling patterns of phosphopeptides resolved by high performance liquid chromatography, but a significantly higher incorporation was detected in two of the phosphopeptides. The stimulatory effect of 2,3-DPG on the phosphorylation of protein 4.1 was observed only with kinase C. Phosphorylation by the cytosolic erythrocyte casein kinase and the cyclic AMP-dependent protein kinase was inhibited by 2,3-DPG. Moreover, the stimulatory effect of 2,3-DPG seemed to be unique to the phosphorylation of protein 4.1 since a similar effect had not been observed with other protein kinase C substrates. Our results suggest that 2,3-DPG may play an important role in the regulation of cytoskeletal interactions.  相似文献   

16.
The behaviour of 2,3-diphosphoglycerate concentration (2,3-DPG) of red blood cells of babies and children with cyanotic heart diseases is studied before and after shunt operations. In babies with cyanotic heart diseases at the age of up to 6 months an increase of 2,3-DPG-level and haematocrit (HCT) is seen. Later, the compensation of chronic hypoxia is effected by further increase of HCT at unchanged high 2,3-DPG-level. The 2,3-DPG concentration which is still increased after successfull shunt operations as compared with the normal value is considered as an effective adaptation mechanism to the residual hypoxia presenting only a small load on the circulatory system. The 2,3-DPG concentration alone does not represent a criterion for the assessment of chronic hypoxia.  相似文献   

17.
The influence of 1-éburnamonine (1-E) and vincamine (Vi) on 2,3-disphosphoglycerate (2,3-DPG) blood level was investigated in awake rats when cyanide (KCN) induced hypoxia was present or not. Used alone, KCN, 1-E and Vi (i.p. route) increased 2,3-DPG blood level. Used with KCN, 1-E or Vi produced a very more important increase of 2,3-DPG than that observed when both drugs were used alone. In all cases, the observed increase was attributed to red cells 2,3-DPG since hematocrite, red-cells count and hemoglobin level were unmodified. The results suggest that the KCN induced increase of 2,3-DPG constitutes a response to hypoxia. On the contrary, that of 1-E or Vi seems to be the result of a metabolic stimulation and could explain in part their antihypoxic properties previously described at cerebral level.  相似文献   

18.
The relationship between hemorheology, erythrocyte ATP and 2,3-diphosphoglycerate (2,3-DPG) concentrations, and von Willebrand factor antigen was studied in athletes and peripheral arterial disease patients. Lower blood viscosity, mainly due to a higher erythrocyte deformability, was found in athletes compared to control subjects. Higher 2,3-DPG/Ht levels in athletes were correlated with blood viscosity, erythrocyte deformability, the rigidity index, and erythrocyte suspension viscosity at low shear stress. It is suggested that these relationships might be determined by the predominance of immature erythrocytes in the blood circulation of the athletes. In the group of patients, a decrease in ATP/Ht was related to increased erythrocyte aggregation and a higher erythrocyte suspension viscosity. Moreover, the concentration of von Willebrand factor was positively correlated with the erythrocyte aggregation index, erythrocyte suspension viscosity, and plasma viscosity. The results show that alterations in erythrocyte and plasma rheology may be involved in the modification of the functional state of the vascular endothelium and the development of atherosclerosis.  相似文献   

19.
Mehta M  Sonawat HM  Sharma S 《FEBS letters》2005,579(27):6151-6158
The erythrocytic stages of the malaria parasite depend on anaerobic glycolysis for energy. Using [2-13C]glucose and nuclear magnetic resonance, the glucose utilization rate and 2,3-diphosphoglycerate (2,3-DPG) level produced in normal RBCs and Plasmodium falciparum infected red blood cell populations (IRBCs, with <4% parasite infected red cells), were measured. The glucose flux in IRBCs was several-folds greater, was proportional to parasitemia, and maximal at trophozoite stage. The 2,3-DPG levels were disproportionately lower in IRBCs, indicating a downregulation of 2,3-DPG flux in non-parasitized RBCs. This may be due to lowered pH leading to selective differential inhibition of the regulatory glycolytic enzyme phosphofructokinase. This downregulation of the glucose utilization rate in the majority (>96%) of uninfected RBCs in an IRBC population may have physiological implications in malaria patients.  相似文献   

20.
1. Erythrocytes in whole blood samples from dogs with phosphofructokinase (PFK) deficiency had lower 2,3-diphosphoglycerate (2,3-DPG) concentrations, higher ATP concentrations, and were more alkaline fragile than normal canine erythrocytes. 2. Reticulocytes from a PFK-deficient dog contained nearly three times the ATP concentration of normal canine erythrocytes, and had 2,3-DPG concentrations similar to normal canine erythrocytes. 3. PFK-deficient reticulocytes are not alkaline fragile. 4. The erythrocyte 2,3-DPG concentration in whole blood samples from PFK-deficient dogs was increased to normal by in vitro incubation with dihydroxyacetone, pyruvate and phosphate. This incubation resulted in only a slight increase in ATP concentration. 5. The alkaline fragility of these 2,3-DPG replenished PFK-deficient erythrocytes was normal. 6. Findings in this study indicate that the increased alkaline fragility of canine PFK-deficient erythrocytes is the result of decreased intracellular 2,3-DPG concentration.  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