Maturation of the reticulocyte in vitro

1984 ◽  
Vol 71 (1) ◽  
pp. 177-197
Author(s):  
G. Gronowicz ◽  
H. Swift ◽  
T.L. Steck

The maturation of reticulocytes into erythrocytes was demonstrated in vitro. Reticulocytosis was induced in rats by repeated bleeding or by phenylhydrazine injections. Whole blood samples were then incubated for 2 days at 37 degrees C. Reticulocytes in culture changed from polylobulated, monoconcave or triconcave forms to biconcave disks. During the first 12 h in vitro, the average reticulocyte count decreased from 39% to 12%, and the membrane-bound organelles, ribosomes and exocytic figures in the remaining reticulocytes were markedly diminished. In contrast, the number of red cells containing inclusions of denatured haemoglobin (Heinz bodies) in phenylhydrazine-treated blood did not decline. The reduction in reticulocyte count was not the result of differential cell destruction, since little haemolysis occurred in vitro. During red cell maturation three modes of organelle removal were observed particularly well when mitochondria were followed by cytochrome oxidase cytochemistry. First, some mitochondria degenerated, presumably through autolysis, by swelling, losing cristae and forming small single membrane-bound vesicles. Second, individual mitochondria became enclosed in vacuoles that fused with the plasma membrane and expelled their mitochondria by exocytosis. Third, autophagic vacuoles containing mitochondria, cytosol and membrane fragments fused with existing lysosomes. We conclude that all aspects of normal reticulocyte maturation occur in vitro, independent of the spleen, including the removal of organelles and the assumption of the mature biconcave disk shape.

1960 ◽  
Vol XXXIV (II) ◽  
pp. 305-311 ◽  
Author(s):  
M. G. Woldring ◽  
A. Bakker ◽  
H. Doorenbos

ABSTRACT The red cell triiodothyronine uptake technique as used in our hospital is described. Incubation time is of almost no importance. The temperature during incubation should be 37° C. Further improvement of the technique is obtained when all blood samples are brought up to 40 % haematocrit prior to incubation. Clinical results are discussed. It is yet too early to give a definite assessment of its clinical value, but it is definitely superior to the measurement of the BMR.


Blood ◽  
2007 ◽  
Vol 110 (6) ◽  
pp. 2173-2181 ◽  
Author(s):  
Benjamin T. Spike ◽  
Benjamin C. Dibling ◽  
Kay F. Macleod

Abstract Definitive erythropoiesis occurs in islands composed of a central macrophage in contact with differentiating erythroblasts. Erythroid maturation including enucleation can also occur in the absence of macrophages both in vivo and in vitro. We reported previously that loss of Rb induces cell-autonomous defects in red cell maturation under stress conditions, while other reports have suggested that the failure of Rb-null erythroblasts to enucleate is due to defects in associated macrophages. Here we show that erythropoietic islands are disrupted by hypoxic stress, such as occurs in the Rb-null fetal liver, that Rb−/− macrophages are competent for erythropoietic island formation in the absence of exogenous stress and that enucleation defects persist in Rb-null erythroblasts irrespective of macrophage function.


1958 ◽  
Vol 193 (2) ◽  
pp. 244-248 ◽  
Author(s):  
Perry Ruth Stahl ◽  
Homer E. Dale

In a repeated study on 17 dairy calves, T-1824 dye plasma dilution showed significantly higher blood volumes than were found by any other technique or computation method using Cr51-tagged red blood cells. Five blood samples taken at 20-minute intervals after injection showed consistent decrease in radioactivity count from the first to the last sample, indicating greater accuracy in radioactivity dilution regressed to zero time figures than in average counts of several postinjection samples. In vitro studies suggest a loss of Cr51 from red blood cells to plasma after saline washings are Cr-free. Percentage blood volumes computed from whole blood samples of calves injected with Cr51-tagged red blood cells decreased in a straight line relationship with increase of body weight. Percentage plasma and whole blood volumes estimated with the T-1824 dye technique decreased regularly with body weight increase until a second determination was made when there was a rapid rise nearly to the level of the smallest calves, followed by another regular decrease with increase in weight. It is suggested that repeated dye injections do not always measure the same space. Regressed values of five whole blood samples taken at 20-minute intervals after injection of Cr51 tagged red blood cells gave more consistent blood volume determinations than either the weighed red cells or the plasma dye dilutions of the same samples.


Blood ◽  
1989 ◽  
Vol 74 (1) ◽  
pp. 475-481 ◽  
Author(s):  
NA Noble ◽  
QP Xu ◽  
JH Ward

Abstract Studies of reticulocyte maturation have been limited by the inability to obtain pure populations of age-synchronized reticulocytes and by the absence of well-defined methods for the maturation of reticulocytes in vitro. Many of these problems were overcome using temporary suppression of erythropoiesis with thiamphenicol and phlebotomy resulting in a highly reproducible reticulocyte response, Percoll density gradient separation of cells yielding essentially pure populations of age- synchronized reticulocytes, and liquid culture techniques where cell lysis is minimal. The system allows reproducible study of well-defined cohorts of reticulocytes as they mature into erythrocytes. During in vitro maturation we serially monitored changes in reticulocyte count, glucose consumption, 125I-transferrin binding, fluorescein (FITC)- labeled transferrin binding, the activities of four erythrocyte enzymes (glucose-6-phosphate dehydrogenase, pyruvate kinase, phosphofructokinase, and lactate dehydrogenase) and the appearance of cells on scanning electron microscopy. These variables changed at different rates suggesting that multiple mechanisms underlie these maturational events. Transferrin binding and reticulocyte count decreased most rapidly and reached values near zero after three to four days in culture. The four enzyme activities decreased much more slowly, and only two reached pretreatment values after seven days in culture. In contrast to the findings of others, scanning electron microscopy suggested that cells do not assume the normal biconcave shape in this system. The methods described make it feasible to study the process of reticulocyte maturation in vitro. The data presented represent a first step in the study of the coordination and interrelationships of various maturational processes.


Blood ◽  
2004 ◽  
Vol 104 (11) ◽  
pp. 577-577 ◽  
Author(s):  
Silverio Perrotta ◽  
Borriello Adriana ◽  
Lucia De Franceschi ◽  
Bruno Nobili ◽  
Achille Iolascon ◽  
...  

Abstract The 911 amino acid human erythroid AE1 (eAE1) Cl-/HCO3- exchanger SLC4A1 (usually called band 3) is the major intrinsic membrane protein of red cells. The N-terminal cytoplasmic domain of AE1 represents the anchoring site for membrane-associated proteins such as ankyrin, protein 4.2, protein 4.1, glycolytic enzymes (including aldolase and glyceraldeyde-3-phosphate dehydrogenase (GAPDH) and hemoglobin. We identified marked band 3 deficiency in the second son of a consanguineous marriage with a life-threatening nonimmune hemolytic anemia. The patient was transfusion-dependent prior to splenectomy. SDS-PAGE and immunoblotting analysis of the proband red cell membrane proteins showed approximately 12±4% of band 3 and protein 4.2 compared to controls. Direct nucleotide sequence of SLC4A1 gene showed a single base substitution (T->C) at position +2 in the donor splice site of intron 2 (Band 3 Neapolis). Functionally, the mutation causes an altered splicing with the consequent formation of two different mature mRNAs, one including intron 2 and one skipping exon 2. While intron 2 retention leads to premature translation termination, exon 2 skipping causes the loss of the normal start site of eAE1 protein translation. The purification of mutant band 3 and its characterization by MALDI mass spectrometry demonstrated the lack of the first 11 amino acids due to the usage of second in frame start site. Real-time RT-PCR analyses of reticulocyte mRNA showed a marked decrement in band 3 transcription accounting for protein deficiency. The lack of the 11 N-terminal amino acids resulted in complete absence of membrane bound aldolase while other glycolitic enzymes (for example GAPDH) were membrane bound. Syk tyrosine kinase recognized the truncated band 3 as a substrate in vitro. In spite of this ability to be phosphorylated by Syk and to recruit Lyn tyrosine kinase in vitro, we were unable to demonstrate Tyr-phosphorylation of mutant band 3 in intact erythrocytes following stimulation by oxidative stress. This finding implies a requirement for the 11 N-terminal amino acids for the sequential Tyr-phosphorylation of band 3 in intact red cell membranes. The mutant band 3 was largely present in the high molecular weight aggregate fraction (about 5.2 fold higher than control), indicating its increased tendency to cluster in the membrane. The spontaneous clustering of truncated band 3 strongly suggests that the negatively charged N-terminal domain may regulate oligomeric state of band 3 in the membrane. Biophysical characterization showed that band 3 deficiency resulted in decreased cohesion between lipid bilyer and spectrin based membrane skeleton accounting for membrane loss. The structural and functional characterization of the naturally occuring mutant band 3 has enabled us to identify a significant role for the 11 N-terminal amino acids in band 3 function and in red cell membrane physiology.


Blood ◽  
2012 ◽  
Vol 120 (21) ◽  
pp. 85-85 ◽  
Author(s):  
Laurel G. Mendelsohn ◽  
Leah Pedoeim ◽  
Eduard J. van Beers ◽  
Rehan Saiyed ◽  
James S Nichols ◽  
...  

Abstract Abstract 85 Background: Aes-103 (5-hydroxymethylfurfural, 5-HMF) is a putative anti-sickling agent that has undergone pre-clinical testing for potential treatment for sickle cell anemia (SCA). It is an organic compound derived from dehydration of certain sugars, found commonly in small amounts in foods such as coffee and prunes. It binds to alpha subunits of hemoglobin and increases its oxygen affinity. At millimolar levels, it inhibits hypoxia-induced sickling in vitro and when dosed orally it protects sickle cell mice against hypoxia-induced death. We investigate the in vitro effects of a range of concentrations of Aes-103 on oxygen affinity and red cell stability in blood from healthy volunteers, and from patients with SCA with or without hydroxyurea treatment. Methods: Blood specimens from healthy control adults and adults with SCA were incubated in vitro with a range of concentrations of Aes-103 between 0.1 and 5 mM for one hour at 37 degrees C. Oxygen equilibrium curves were determined for each sample using the HemOx Analyzer. Samples were diluted in HemOx buffer and then loaded into the Analyzer, which exposed the samples to increasing partial pressure and then deoxygenated with nitrogen gas to produce the oxygen equilibrium curve. The P50 value for each curve was determined at the oxygen tension that produced 50% oxygen saturation. In additional experiments, samples of human control blood and SCA blood were treated with Aes-103 and incubated at 37°C for 1 hr, and then the samples in tubes were subjected to shear stress by rotation on a vertical rotator at 21 revolutions per minute for 3 hrs. The samples were centrifuged for 2 minutes and plasma was collected and free hemoglobin levels as an indicator of red cell membrane disruption were measured by ELISA. Results: Blood samples from SCA patients off hydroxyurea (n=6) without Aes-103 tended to have higher baseline p50 values than healthy controls (n=6)(30.3 ± 1.1 vs. 28.3 ± 0.8 torr, p=0.15), consistent with previous reports of high intracellular 2,3-DPG, known to increase P50. The P50 remained right shifted in SCA compared to controls at Aes-103 concentration below 1mM, converging with controls at higher concentrations (p=0.035). At baseline, P50 of SCA patients on chronic hydroxyurea (n=9) was significantly lower than SCA patients not on hydroxyurea (26.3 ± 0.8 vs. 30.3 ± 1.1 torr, p=0.008), compatible with the lower P50 contributed by fetal hemoglobin induced by hydroxyurea. At every concentration of Aes-103, P50 was lower for specimens SCA on hydroxyurea compared to those off hydroxyurea (p<0.001) (Figure 1A). Overall, the delta decrease in P50 from baseline in all subjects at all concentrations of Aes-103 was comparable, on regression analysis showing −2.16 torr for each mM increase in Aes-103 (r2=0.64, p<0.001). In vitro shear stress under normoxia promoted hemolysis in blood samples from patients with SCA compared to baseline (n=10, free hemoglobin 29.4 ± 3.4 vs. 8.4 ± 0.9 uM, p<0.001). Addition of Aes-103 at increasing concentrations reduced the extent of shear-stress induced hemolysis, by 15% at 1mM Aes-103; by 28% at 2mM Aes-103; and by 37% at 5mM Aes-103 (p<0.001, Figure 1B). Interestingly, although shear stress promoted less hemolysis in blood samples from healthy controls, Aes-103 at these concentrations also reduced this hemolysis to a comparable extent, suggesting a red cell stabilizing mechanism distinct from anti-sickling effect. Shear stress experiments under hypoxic conditions are underway. In pilot experiments using an imaging flow cytometry assay described in detail in a separate abstract, Aes-103 showed preliminary ability to repress sickling induced by hypoxia in vitro. Conclusions: Red cells from SCA adults treated with hydroxyurea have significantly higher affinity for oxygen than those from patients not treated with hydroxyurea, presumably related in part to the high affinity of fetal hemoglobin induced by the drug. Aes-103 increases oxygen affinity in sickle erythrocytes in a concentration-dependent fashion, and this effect is even more prominent when combined with that of hydroxyurea. Aes-103 at high concentrations stabilizes red cells against shear stress in vitro. With our collaborators at AesRx, LLC, a phase 1 safety and pharmacokinetics study of Aes-103 in healthy volunteers has been completed and we now are conducting a similar study at the NIH Clinical Center in adults with sickle cell anemia. Disclosures: No relevant conflicts of interest to declare.


1987 ◽  
Vol 65 (12) ◽  
pp. 1080-1090 ◽  
Author(s):  
Linda Orr ◽  
M. Adam ◽  
R. M. Johnstone

During the maturation of sheep reticulocytes in vitro, there is release of material that can be pelleted from the cell-free incubation medium by centrifugation at 100 000 × g. This pellet contains activities that are derived from both the plasma membrane and lysosomes. No evidence was obtained for the presence of mitochondrial activities or cytosolic enzyme activities. The release of these activities is ATP and temperature dependent, since reduction of either results in a greater retention of the activities by the cells and a lesser amount in the 100 000 × g pellet. The pelleted material is vesicular in nature, and the production and (or) release of the material are reduced upon ATP depletion or lowering of the temperature. It is concluded that the externalization of specific membrane components is a normal metabolic process that occurs during reticulocyte maturation and represents a means by which reticulocytes shed specific types of membrane-associated functions that are known to decrease during reticulocytes maturation.


Blood ◽  
1989 ◽  
Vol 74 (1) ◽  
pp. 475-481 ◽  
Author(s):  
NA Noble ◽  
QP Xu ◽  
JH Ward

Studies of reticulocyte maturation have been limited by the inability to obtain pure populations of age-synchronized reticulocytes and by the absence of well-defined methods for the maturation of reticulocytes in vitro. Many of these problems were overcome using temporary suppression of erythropoiesis with thiamphenicol and phlebotomy resulting in a highly reproducible reticulocyte response, Percoll density gradient separation of cells yielding essentially pure populations of age- synchronized reticulocytes, and liquid culture techniques where cell lysis is minimal. The system allows reproducible study of well-defined cohorts of reticulocytes as they mature into erythrocytes. During in vitro maturation we serially monitored changes in reticulocyte count, glucose consumption, 125I-transferrin binding, fluorescein (FITC)- labeled transferrin binding, the activities of four erythrocyte enzymes (glucose-6-phosphate dehydrogenase, pyruvate kinase, phosphofructokinase, and lactate dehydrogenase) and the appearance of cells on scanning electron microscopy. These variables changed at different rates suggesting that multiple mechanisms underlie these maturational events. Transferrin binding and reticulocyte count decreased most rapidly and reached values near zero after three to four days in culture. The four enzyme activities decreased much more slowly, and only two reached pretreatment values after seven days in culture. In contrast to the findings of others, scanning electron microscopy suggested that cells do not assume the normal biconcave shape in this system. The methods described make it feasible to study the process of reticulocyte maturation in vitro. The data presented represent a first step in the study of the coordination and interrelationships of various maturational processes.


Blood ◽  
2008 ◽  
Vol 112 (11) ◽  
pp. 1418-1418
Author(s):  
Vanessa Bourgeaux ◽  
Olivier Hecquet ◽  
Dominique Rigal ◽  
Alain Francina ◽  
Yann Godfrin

Abstract Sickle cell disease (SCD) is characterised by abnormal haemoglobin S (HbS). Under hypoxic conditions, HbS crystallizes, inducing sickling of red blood cells. Consequently, patients have a high risk of vaso-occlusive painful crisis. Red cell exchange transfusions remain an effective therapy in the acute and chronic treatment of SCD: the patient’s red blood cells (RBC) are removed and replaced by homologous normal red cells. Red cell exchange can provide needed oxygen carrying capacity while reducing the overall viscosity of blood (P.S. Swerdlow, 2006). We propose a novel preventive and therapeutic approach for SCD based on red blood cell transfusion. We hypothesise that loading RBC with an allosteric effector of hemoglobin can reduce RBC sickling. Indeed, the entrapment of Inositol Hexaphosphate (IHP) inside RBC reduces the oxygen-hemoglobin affinity, which is measured by a right shift of the oxygen dissociation curve. Thus, RBC-IHP have an increased capacity to deliver oxygen to tissues. It is also expected that the deoxygenation of SS RBC is reduced and sickling is avoided. IHP was entrapped into human RBC by hypotonic reversible lysis followed by a resealing step. RBC-IHP were characterised by the amount of IHP entrapped into RBCs and the P50 measurement. Unprocessed human RBC were used as control. The potential anti-sickling effect of RBC-IHP was investigated using an in vitro model. Firstly, an experimental model to observe the relationship between sickling and oxygen concentration was set up : patients cells were submitted to deoxygenation by nitrogen bubbling for 30 min, and then re-oxygenated with different concentrations of oxygen (2, 5, 8, 15, 22%) for 30 min. The percentage of sickled cells was assessed by microscopy (about 500 cells checked). We observed that sickled cells recovered a normal shape upon reoxygenation (&gt;15%O2), and a steady state between 5 and 8 % of oxygen, allowing the development of a reliable experimental model. Next, patient blood samples (n=6), harvested just prior to red cell exchange, were studied. RBC were washed 3 times with phoshate buffer before use. Different proportions of RBC-IHP (10%, 30% or 50%) were mixed with patients red cells and submitted to deoxygenation (0% O2) for 30 min and reoxygenation (5% O2) for 30 min. The final hematocrit of the suspensions was approximately 15%. The percentage of sickled cells in the suspensions was evaluated by microscopy and corrected according to the appropriate dilution factor. After full deoxygenation, 10% to 50% of cells were sickled, which appeared to be dependent on the HbS level in the blood samples. For all patients, RBC-IHP exhibited an enhanced anti-sickling effect: sickling was reduced by 19, 34, and 67% according to the RBC-IHP proportions 10%, 30% and 50%, respectively. Indeed, for equivalent RBC proportions RBC-IHP (50%) was 1.4 to 9 times more efficient compared to the unprocessed control RBC. Thus, RBC-IHP has the capacity to prevent sickling in a dose-dependent manner and is efficient at low proportions (10%). Consequently, RBC-IHP can improve classical transfusion therapy in terms of transfused volume, frequency and preventive sickling effect.


Sign in / Sign up

Export Citation Format

Share Document