Journal articles on the topic 'Stomatocytosi'

To see the other types of publications on this topic, follow the link: Stomatocytosi.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Stomatocytosi.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Das, Reena, Manu Jamwal, Anu Aggarwal, Prashant Sharma, Arindam Maitra, Deepak Bansal, and Pankaj Malhotra. "Phenotype-Genotype Spectrum of Stomatocytic Disorders Encountered in India Using Next Generation Sequencing." Blood 132, Supplement 1 (November 29, 2018): 2326. http://dx.doi.org/10.1182/blood-2018-99-114554.

Full text
Abstract:
Abstract Introduction Stomatocytes in peripheral blood are pathognomonic findings in multiple conditions along with hemolysis and reticulocytosis, often suggestive of erythrocyte membrane transport defects. These uncommon disorders are usually difficult to diagnose due to a wide range of overlapping phenotypes and a perception of stomatocytes being artefacts. Genes involved in these disorders are multiple (RHAG,SLC4A1, ABCG5, ABCG8, PIEZO1,KCNN4, ABCB6, SLC2A1 etc) rendering Sanger sequencing costly and labor intensive. Phenotypes vary from transfusion-dependent anemia to compensated hemolysis. Methods Seventeen patients were encountered in 12 families and enrolled in this study. Majority of the cases showed the presence of significant numbers of red blood cells showing stomatocytes with or without thrombocytopenia. Various hematological, biochemical and molecular tests were used to exclude thalassemia syndromes and hemoglobinopathies, glucose-6-phosphate dehydrogenase (G6PD) deficiency, autoimmune hemolytic anemia, hereditary spherocytosis (HS) and pyruvate kinase deficiency. Genomic DNA was extracted by the QIAamp DNA Blood Midi Kit and quantified on NanoDrop 2000 spectrophotometer and QubitFluorometer. DNA libraries were prepared using Illumina's custom panels (TruSight One Sequencing Panel and TruSeq Custom Amplicon v1.5) and sequenced on a MiSeq Sequencing System. MiSeq Reporter and VariantStudio were used for analysis, classification, and reporting of variants. Variants which were predicted pathogenic by in silico analysis using PolyPhen-2, SIFT, PROVEAN (http://provean.jcvi.org/), Mutpred (http://mutpred.mutdb.org/) and Human Splicing Finder as indicated, were subjected to Sanger sequencing in the patient and family members (where available). Results Of these 17 patients, 10 patients in 6 families were diagnosed to have Mediterranean stomatocytosis/ macrothrombocytopenia. All had the presence of stomatocytes along with macrothrombocytopenia, short stature, continuous abdominal discomfort and marked pleiotropic effects in different cases. This is a syndromic form of stomatocytosis and defects in ABCG5/ABCG8 genes were found and showed an autosomal recessive inheritance pattern. One case did not show any mutation. This number of cases suggests that this disorder is not rare in India and is probably underdiagnosed as patients have mild or moderate anemia and are often misdiagnosed as cases with HS. One of the patients also had coinheritance of G6PD deficiency (G6PD Kerela Kalyan). Patients with Mediterranean stomatocytosis/macrothrombocytopenia are advised to take sterol-absorption inhibitor 'ezetimibe' to reduce sterol accumulation. We also found 2 unrelated patients with stomatocytosis, reticulocytosis and splenomegaly with overhydrated hereditary stomatocytosis (OHSt) and pathogenic mutation in RHAG gene was found in both of them. The pattern of inheritance is sporadic or autosomal dominant. Splenectomy was deferred in a patient with OHSt as postsplenectomy thrombotic complications are known to occur and is contraindicated. Four patients in 3 families were found to have mutations in PIEZO1 gene which translates to red cell membrane mechanosensitive cation channel protein, causing xerocytosis/dehydrated hereditary stomatocytosis and 3 patients had severe anemia and were transfusion dependent. One patient showed the presence of stomatocytes and macrothrombocytopenia was found to have probably disease causing variant in SLC2A1 gene. She was incidentally undergoing treatment for infertility when the stomatocytosis was noted. Conclusions Stomatocytic disorders appear to be underdiagnosed in India which is compounded by the protean clinical manifestations, milder phenotypes, low index of suspicion and non-availability of molecular confirmation. Astute phenotype characterization is critical as it will help in establishing the causality of the variants identified and appropriate genetic counseling. Recently NGS for hemolytic anemias has led to rapid molecular characterization and accurate phenotypic correlation. Disclosures No relevant conflicts of interest to declare.
APA, Harvard, Vancouver, ISO, and other styles
2

Truong, HT, JE Jr Ferrell, and WH Huestis. "Sulfhydryl reducing agents and shape regulation in human erythrocytes." Blood 67, no. 1 (January 1, 1986): 214–21. http://dx.doi.org/10.1182/blood.v67.1.214.214.

Full text
Abstract:
Abstract Metabolic crenation of red cells is reversible; on addition of nutrients, echinocytes recover the normal discoid shape. When the shape recovery takes place in the presence of reducing agents such as dithiothreitol (DTT), morphological change continues until the cells are stomatocytic. The degree of stomatocytosis varies, depending on the cell morphology when the nutrients and reducing agent are added. DTT has minimal effect on the shape of normal discocytes, but in its presence, mildly echinocytic cells become slightly cupped and advanced- stage echinocytes become severely stomatocytic. DTT must be present continuously for development and retention of stomatocytosis; echinocytes preincubated with or metabolically depleted in DTT do not become stomatocytic when supplemented in the absence of DTT, and DTT- induced stomatocytes revert to discocytes when the reducing agent is removed. DTT has no effect on adenosine triphosphate synthesis or equilibrium cell glutathione levels, and the induced stomatocytosis is not inhibited by excluding oxygen from cells during depletion. Spectrin phosphorylation and phosphate turnover are not affected by DTT. The echinocyte-to-discocyte transformation coincides with phosphorylation of membrane inner monolayer lipids (diacylglycerol to phosphatidic acid and phosphatidylinositol to phosphatidylinositol-4,5-bisphosphate). Overphosphorylation of these phospholipids is not responsible for the exaggerated shape recovery seen with reducing agents; phosphorylation of inner monolayer lipids proceeds identically in the presence and absence of DTT.
APA, Harvard, Vancouver, ISO, and other styles
3

Truong, HT, JE Jr Ferrell, and WH Huestis. "Sulfhydryl reducing agents and shape regulation in human erythrocytes." Blood 67, no. 1 (January 1, 1986): 214–21. http://dx.doi.org/10.1182/blood.v67.1.214.bloodjournal671214.

Full text
Abstract:
Metabolic crenation of red cells is reversible; on addition of nutrients, echinocytes recover the normal discoid shape. When the shape recovery takes place in the presence of reducing agents such as dithiothreitol (DTT), morphological change continues until the cells are stomatocytic. The degree of stomatocytosis varies, depending on the cell morphology when the nutrients and reducing agent are added. DTT has minimal effect on the shape of normal discocytes, but in its presence, mildly echinocytic cells become slightly cupped and advanced- stage echinocytes become severely stomatocytic. DTT must be present continuously for development and retention of stomatocytosis; echinocytes preincubated with or metabolically depleted in DTT do not become stomatocytic when supplemented in the absence of DTT, and DTT- induced stomatocytes revert to discocytes when the reducing agent is removed. DTT has no effect on adenosine triphosphate synthesis or equilibrium cell glutathione levels, and the induced stomatocytosis is not inhibited by excluding oxygen from cells during depletion. Spectrin phosphorylation and phosphate turnover are not affected by DTT. The echinocyte-to-discocyte transformation coincides with phosphorylation of membrane inner monolayer lipids (diacylglycerol to phosphatidic acid and phosphatidylinositol to phosphatidylinositol-4,5-bisphosphate). Overphosphorylation of these phospholipids is not responsible for the exaggerated shape recovery seen with reducing agents; phosphorylation of inner monolayer lipids proceeds identically in the presence and absence of DTT.
APA, Harvard, Vancouver, ISO, and other styles
4

Flatt, Joanna F., Hélène Guizouarn, Nicholas M. Burton, Franck Borgese, Richard J. Tomlinson, Robert J. Forsyth, Stephen A. Baldwin, et al. "Stomatin-deficient cryohydrocytosis results from mutations in SLC2A1: a novel form of GLUT1 deficiency syndrome." Blood 118, no. 19 (November 10, 2011): 5267–77. http://dx.doi.org/10.1182/blood-2010-12-326645.

Full text
Abstract:
Abstract The hereditary stomatocytoses are a series of dominantly inherited hemolytic anemias in which the permeability of the erythrocyte membrane to monovalent cations is pathologically increased. The causative mutations for some forms of hereditary stomatocytosis have been found in the transporter protein genes, RHAG and SLC4A1. Glucose transporter 1 (glut1) deficiency syndromes (glut1DSs) result from mutations in SLC2A1, encoding glut1. Glut1 is the main glucose transporter in the mammalian blood-brain barrier, and glut1DSs are manifested by an array of neurologic symptoms. We have previously reported 2 cases of stomatin-deficient cryohydrocytosis (sdCHC), a rare form of stomatocytosis associated with a cold-induced cation leak, hemolytic anemia, and hepatosplenomegaly but also with cataracts, seizures, mental retardation, and movement disorder. We now show that sdCHC is associated with mutations in SLC2A1 that cause both loss of glucose transport and a cation leak, as shown by expression studies in Xenopus oocytes. On the basis of a 3-dimensional model of glut1, we propose potential mechanisms underlying the phenotypes of the 2 mutations found. We investigated the loss of stomatin during erythropoiesis and find this occurs during reticulocyte maturation and involves endocytosis. The molecular basis of the glut1DS, paroxysmal exercise-induced dyskinesia, and sdCHC phenotypes are compared and discussed.
APA, Harvard, Vancouver, ISO, and other styles
5

De Falco, Luigia, Lucia De Franceschi, Frank Borgese, Carmelo Piscopo, Maria Rosaria Esposito, Rosa Anna Avvisati, Pietro Izzo, Helene Guizouarn, Andrea Biondani, and Achille Iolascon. "BAND 3CEINGE (Gly796Arg) Mutation Causes Dehydrated Hereditary Stomatocytosis (DHS) with Dyserythropoietic Phenotype." Blood 112, no. 11 (November 16, 2008): 2874. http://dx.doi.org/10.1182/blood.v112.11.2874.2874.

Full text
Abstract:
Abstract Stomatocytosis is an inherited autosomal dominant hemolytic anemia and includes overhydrated hereditary stomatocytosis (OHS), dehydrated hereditary stomatocytosis (DHS), hereditary cryohydrocytosis (CHC) and familial pseudohyperkalemia (FP). Here, we report a novel variant of hereditary stomatocytosis due to a de-novo band 3 mutation due to G>A transition at nucleotide 2500 in exon 17 (p. G796R, band3CEINGE) associated with dyserythropoietic phenotype. This 43-years-old Caucasian female (II-2) with unrelated parents was admitted to our hospital for mild anemia evaluation. The patient was in good health until 7 years when she frequently experienced asthenia. Anemia was first recognized at the age of eighth years with presence of jaundice and hyperchromic urine, but she had never received blood transfusions. We observed a mild hypochromic macrocytic anemia with a hemoglobin level of 11.5 g/dL, a mean cell volume (MCV) of 110 fL, and a mean hemoglobin concentration (MCH) of 36.1 pg, the reticulocyte count was 64 × 103/μL. There was a typical hemolytic features: high levels of indirect bilirubin (3.48 mg/dL) and lactate dehydrogenase ( 567 U/l, v.n. 240– 480 U/l ) with negativity at direct and indirect Coomb’s test. Spleen was enlarged and ultrasonography detected 15 cm of longitudinal size. She was cholecystectomized at the age of 14 years because of numerous symptomatic small stones. Serum iron, soluble transferrin receptor, serum ferritin and transferrin saturation levels were all increased, while the transferrin was in the normal range.Other blood tests including osmotic fragility with incubated and fresh erythrocytes, serum electrolytes, B12 and folate levels, erythrocyte enzyme levels, EMA test and Pink test were normal. Peripheral blood smear showed anisopoikilocytosis with rare stomatocytes and no spherocytes. Bone marrow aspirate showed remarkable dyserythropoiesis with increased number of erythroblasts and binucleate erythroblasts, basophilic erythroblasts with alterations, irregular nuclei maturation, intererythroblastic bridges and erythroblasts with basophilic stippling. She received since the age of 14 yrs a diagnosis for congenital dyserythropoietic anemia type I. Patients red cells showed increase Na+ content and decrease K+ content; reduced Na-K pump activity and increased Na-H exchange, NKCC cotransport and KCC cotransport activities. We then functionally characterized band 3 CEINGE in Xenopus oocytes, showing that the mutated band 3 is converted from anion exchanger (Cl−, HCO3 −) function to unregulated cation pathway for Na+ and K+. The mutated band 3 was also associated with increased tyrosine phosphorylation pattern of some red cell membrane proteins. During erythropoiesis band 3 protein is the last cytoskeletal protein to appear, thus the dyserythropoietic phenotype may be related to a possible role of the mutated band 3 in perturbation of cytoskeleton assembly in the late stage of erythropoiesis, allowing us to conclude for a new variant of stomatocytosis with dyserythropoietic phenotype.
APA, Harvard, Vancouver, ISO, and other styles
6

Marcello, Anna P., Cristina Vercellati, Elisa Fermo, Paola Bianchi, Wilma Barcellini, and Alberto Zanella. "Coexistence of Congenital Red Cell Pyruvate Kinase Deficiency and Hereditary Stomatocytosis." Blood 110, no. 11 (November 16, 2007): 1738. http://dx.doi.org/10.1182/blood.v110.11.1738.1738.

Full text
Abstract:
Abstract We describe a case of chronic hemolytic anemia due to the co-presence of pyruvate kinase (PK) deficiency and Hereditary Stomatocytosis (HSto). The propositus was a 30 years old adopted male with no known family history; he had severe neonatal jaundice requiring exchange transfusion, followed by a life-long history of moderate to severe chronic hemolytic anemia (Hb 7–10 g/dL), with jaundice and splenomegaly. At the age of 6 months hemoglobin screening was made and a beta trait was found. At the age of 20 splenectomy and cholecystectomy were performed. Surgery resulted in an increase of 1.5 g/dL in haemoglobin, and in a conspicuous rise of reticulocytes (from 125×109/L to 562×109/L). Two thrombotic events occurred thereafter, the former 6 days after surgery, and the latter two years later, during a toxoplasmosis infection. At the time of the study Hb was 10.8 g/dL, MCV 82.2 fL, reticulocytes 562×109/L, unconjugated bilirubin 2.19 mg/dL, LDH 335 U/L, haptoglobin <20 mg/dL, serum ferritin 342 ng/mL and transferrin saturation 71%. The peripheral blood smear examination showed the presence of echinocytes (13%), stomatocytes (11%), acantocytes (10%), schistocytes (7%), elliptocytes (6%), spherocytes (4%), target cells (4%) and a few erythroblasts. Erythrocyte osmotic fragility was decreased; screening test for unstable hemoglobins and sodium dodecyl sulphate-polyacrylamide gel electrophoresis (SDS-PAGE) of red cell membrane gave normal results. The study of the most important red cell enzymes revealed reduced PK activity (6.0 UI/gHb, normal range 11.1–15.59 UI/gHb) and thermal stability (43%, normal range 57–100%). Direct sequencing of PK-LR gene showed a compound heterozygosity for mutation 1456T (Arg486Trp) and the new variant −73g>c. Mutation −73g>c occurs in the most proximal of the four GATA motifs in the R-type promoter region and possibly result in a decrease of mRNA synthesis, as already reported for the variant −72a>g (Manco et al, 2000). Molecular analysis of HFE gene showed heterozygosity for H63D mutation. The history of post splenectomy thrombosis and the presence of stomatocytes in peripheral blood smear prompted us to investigate for the coexistence of hereditary stomatocytosis. The determination of plasma potassium and sodium concentration revealed an increase in intracellular sodium (16.3 mmol/LRBC, reference range 5.0–12.0) and a decrease in intracellular potassium (74.73 mmol/LRBC, reference range 90–103), suggestive for a diagnosis of dehydrated HSto, or hereditary xerocytosis. This defect likely accounts for the thrombophilic state in this case, since HSto is known to be associated with hypercoagulability, particulary after splenectomy.
APA, Harvard, Vancouver, ISO, and other styles
7

Agrawal, Akanksha, Deepanshu Jain, and Mitchell Goldstein. "Reversible stomatocytosis." International Journal of Case Reports and Images 8, no. 5 (2017): 417. http://dx.doi.org/10.5348/ijcri-201714-cl-10124.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Chasis, JA, and SL Schrier. "Membrane deformability and the capacity for shape change in the erythrocyte." Blood 74, no. 7 (November 15, 1989): 2562–68. http://dx.doi.org/10.1182/blood.v74.7.2562.2562.

Full text
Abstract:
Abstract Erythrocytes must have the capacity to undergo marked membrane deformation and shape changes in order to circulate through capillaries and respond to a range of shear stresses. To study the interrelationships between membrane deformability and the capacity for shape transformation, we created rigid membranes using several agents and then examined the ability of these erythrocytes with rigid membranes to undergo amphipath-induced shape change. We have previously shown that wheat germ agglutinin (WGA) and a monoclonal antibody to glycophorin A (R-10) cause membrane rigidity as measured by ektacytometry. Experiments were therefore designed to produced comparably rigid membranes using WGA, R-10, and diamide, and then to test the ability of lysophosphatidylcholine to produce echinocytes, and primaquine to produce stomatocytes. We found that diamide treatment substantially blocked both types of shape change. In contrast, R-10 binding did not impair either primaquine- or lysophosphatidylcholine- induce shape change. Further, WGA blocked echinocyte transformation, as previously reported, but not stomatocytosis. Using reduced and unreduced gel electrophoresis and Triton extraction, we compared the biochemical changes associated with WGA-, diamide-, and R-10-induced rigidity, and found them to be different. We conclude that not all rigid cells are incapable of shape change, and therefore that decreased membrane deformability is not predictive of impaired capacity for shape change.
APA, Harvard, Vancouver, ISO, and other styles
9

Chasis, JA, and SL Schrier. "Membrane deformability and the capacity for shape change in the erythrocyte." Blood 74, no. 7 (November 15, 1989): 2562–68. http://dx.doi.org/10.1182/blood.v74.7.2562.bloodjournal7472562.

Full text
Abstract:
Erythrocytes must have the capacity to undergo marked membrane deformation and shape changes in order to circulate through capillaries and respond to a range of shear stresses. To study the interrelationships between membrane deformability and the capacity for shape transformation, we created rigid membranes using several agents and then examined the ability of these erythrocytes with rigid membranes to undergo amphipath-induced shape change. We have previously shown that wheat germ agglutinin (WGA) and a monoclonal antibody to glycophorin A (R-10) cause membrane rigidity as measured by ektacytometry. Experiments were therefore designed to produced comparably rigid membranes using WGA, R-10, and diamide, and then to test the ability of lysophosphatidylcholine to produce echinocytes, and primaquine to produce stomatocytes. We found that diamide treatment substantially blocked both types of shape change. In contrast, R-10 binding did not impair either primaquine- or lysophosphatidylcholine- induce shape change. Further, WGA blocked echinocyte transformation, as previously reported, but not stomatocytosis. Using reduced and unreduced gel electrophoresis and Triton extraction, we compared the biochemical changes associated with WGA-, diamide-, and R-10-induced rigidity, and found them to be different. We conclude that not all rigid cells are incapable of shape change, and therefore that decreased membrane deformability is not predictive of impaired capacity for shape change.
APA, Harvard, Vancouver, ISO, and other styles
10

Houston, Brett L., Teresa Zelinski, Donald S. Houston, Sara J. Israels, Gail Coghlan, Bernie N. Chodirker, Patrick G. Gallagher, and Ryan Zarychanski. "Genetic Linkage of the Dehydrated Hereditary Stomatocytosis Locus to Chromosome 16 in Two Kindreds and Refinement of the Candidate Gene Region to 16q24.2 - 16qter,." Blood 118, no. 21 (November 18, 2011): 3170. http://dx.doi.org/10.1182/blood.v118.21.3170.3170.

Full text
Abstract:
Abstract Abstract 3170 BACKGROUND: The hereditary stomatocytoses are a rare, diverse group of clinical conditions associated with chronic red blood cell hemolysis and increased erythrocyte membrane permeability to monovalent cations. The most common form of hereditary stomatocytosis is dehydrated stomatocytosis (DHSt, also called hereditary xerocytosis) first reported by Miller et al. (Blood 38:184, 1971). We studied two genetically unrelated kindreds with DHSt, one from Western Canada and the other, the original DHSt kindred reported by Miller et al., from upstate New York. Although the DHSt causative gene is unknown, previous studies have mapped the DHSt locus to 16q23 – 16qter. OBJECTIVES: To define the chromosomal region carrying the disease locus in these two families using DNA linkage analysis. METHODS: PCR-based genotyping was performed on genomic DNA according to standard methods. Fine mapping was performed using markers on the telomeric end of chromosome 16. Linkage data was generated from both paternal and maternal meioses. LOD scores (logarithm (base 10) of odds) were determined at assumed recombination fractions (theta) 0.00, 0.05, 0.10, 0.20, 0.30 and 0.40. Continuous allelic alignments were manually generated. RESULTS: Based on reticulocyte count, MCHC, and osmotic fragility, disease phenotypes were identified in 29 individuals of the Canadian kindred, and 27 individuals of the New York kindred. Linkage data from both families confirmed that the disease-causing locus mapped to chromosome 16. In the Canadian kindred, significant LOD scores (> +3.00) were established between the disease-causing locus (DHSt) and D16S3074, D16S2621 and D16S3026. Three recombinants were observed between DHSt:D16S3074, two between DHSt:D16S2621 and none between DHSt:D16S3026. Significant LOD scores were also established between DHSt and D16S3074, D16S476, and D16S413 in the New York kindred, and recombination between DHSt:D16S3074 was also observed. Definition of critical recombination events allowed us to define the centromeric boundary of the region containing the disease-causing locus as D16S2621, a 2.4 cM region containing 51 known or predicted genes. CONCLUSIONS: Genetic linkage analysis has confirmed DHSt linkage to chromosome 16 in two unrelated kindreds and has refined the assignment of DHSt to chromosome 16q24.2 – 16qter. Disclosures: No relevant conflicts of interest to declare.
APA, Harvard, Vancouver, ISO, and other styles
11

Layton, David Mark, and Barbara J. Bain. "Dehydrated hereditary stomatocytosis." American Journal of Hematology 91, no. 2 (November 17, 2015): 266. http://dx.doi.org/10.1002/ajh.24212.

Full text
APA, Harvard, Vancouver, ISO, and other styles
12

Fermo, Elisa, Anna Yu Bogdanova, Polina Petkova-Kirova, Anna Zaninoni, Anna Paola Marcello, Asya Makhro, Pascal Hänggi, et al. "Gardos Channel Mutation Is Associated with Hereditary Dehydrate Stomatocytosis: a Complex Channelopathy." Blood 126, no. 23 (December 3, 2015): 3333. http://dx.doi.org/10.1182/blood.v126.23.3333.3333.

Full text
Abstract:
Abstract The Gardos channel (KCNN4) is a Ca2+ sensitive, intermediate conductance, K+ selective channel abundant in several tissues including red blood cells (RBC) where it is involved in cell volume regulation (Hoffman et al. 2003, Cahalan et al. 2015). Hereditary dehydrate stomatocytosis (DHSt) is characterized by mild to moderate congenital hemolytic anemia with reticulocytosis and splenomegaly associated with RBC dehydration. Affected RBC are characterized by a nonspecific cation leak, reflected in elevated Na+ content, decreased K+, elevated MCHC and MCV, and decreased osmotic fragility. The definitive diagnosis of DHSt is made by osmotic gradient ektacytometry, which shows a leftward shift of the bell-shaped curve. Heterozygous mutations associated with delayed channel inactivation have been identified for PIEZO1 (Zarychanski et al. 2012), although there are patients lacking a mutation in this gene. Differential diagnosis in hereditary stomatocytosis is important because in DHSt patients, splenectomy is contraindicated due to increased risk of thromboembolic complications. We studied a 40 year-old Italian man, affected since infancy by chronic hemolyitc anemia (Hb 7-9 g/dL, reticulocytes 10%). The unrelated parents and three siblings were healthy with normal hematologic parameters. He was occasionally transfused until splenectomy, performed at the age of 11 years. After splenectomy Hb levels maintained around 10 g/dL, thrombotic events did not occur. Increased ferritin levels (>1000 ng/ml) and moderate iron overload required chelation therapy. Peripheral blood smear showed anysopoikilocytosis with stomatocytes (13%) echinocytes (7%) schistocytes (6%) ellyptocytes (5%) spherocytes (4%). Bone marrow evaluation revealed erythroid hyperplasia with some dyserytrhopoietc changes, in particular binucleated erythroblasts. Eosin 5 maleimide (EMA) binding test was negative; despite the high number of stomatocytes in peripheral blood, the osmotic gradient ektacytometry curve was not suggestive for hereditary xerocytosis. The proband's first daughter, born at term after an uneventful pregnancy, presented with severe anemia (Hb 6.1 g/dL) requiring RBC transfusion at birth and at 4 months; at the age of 1 year she displayed a clinical pattern similar to the one of the father, with median Hb levels of 9 g/dL and no need of transfusions. To uncover the underlying etiologies, all the family members underwent whole exome sequencing. Only one candidate gene, the KCNN4 could be identified: a heterozygous missense mutation (c.1055G>A, p.R352H) was detected as a de novo mutation in the proband, and dominantly transmitted to the daughter. The mutation falling in exon 7 involved in the calmodulin binding domain was predicted to be pathogenic by in silico analysis. Sanger sequencing confirmed the presence of the mutation. Very recently (July 2015), the same mutation in the KCNN4 gene was described by other groups (Rapetti-Mauss et al. 2015; Andolfo et al. 2015). In heterologous expression systems it could be shown that the Ca2+ sensitivity of the mutated Gardos channel was increased. We performed single cell experiments, including patch-clamp recordings in RBCs and precursor cells, Ca2+ fluorescence imaging as well as tracer flux experiments in RBCs of the above described patients and healthy controls to get a mechanistic link between the Gardos channel mutation and the cellular symptoms. Despite the presence of a single mutation in Gardos channel alone, pathological alterations in function of multiple ion transport systems were revealed during these studies. Our findings together with recent reports (Rapetti-Mauss et al. 2015; Andolfo et al. 2015; Glogowska et al. 2015) suggest that the prevalence of mutations in the Gardos channel is higher than previously thought. Therefore it is important to understand the mechanisms implicated by the mutation and to develop reliable diagnostic tools and treatment strategies. Disclosures No relevant conflicts of interest to declare.
APA, Harvard, Vancouver, ISO, and other styles
13

Schrier, SL, A. Zachowski, and PF Devaux. "Mechanisms of amphipath-induced stomatocytosis in human erythrocytes." Blood 79, no. 3 (February 1, 1992): 782–86. http://dx.doi.org/10.1182/blood.v79.3.782.782.

Full text
Abstract:
Abstract We studied stomatocytosis induced in human red blood cells (RBC) by vinblastine and chlorpromazine, monitoring the movements of spin- labeled phosphatidylcholine (PC*) and sphingomyelin (SM*) by electron spin resonance (ESR) spectroscopy. This technique allows determination of the fraction of labeled lipids, respectively, on the external leaflet, on the cytosol face, or trapped in endocytic vacuoles. Both vinblastine and chlorpromazine produce a time- and concentration- dependent stomatocytic shape change, which is paralleled by a shift of approximately 10% to 33% of outer leaflet SM* and PC* inward. Of this amount, 8% to 12% was trapped in endocytic vacuoles and 8% to 19% had flipped to the inner leaflet. Vanadate, while inhibiting the stomatocytosis, did not block the flip of either SM* or PC* to the inner leaflet. To explain the inhibiting effect of vanadate, as well as the adenosine triphosphate (ATP) requirement for drug-induced stomatocytosis, we propose the following model: (1) addition of amphipath partially scrambles the bilayer; and (2) the flop of phosphatidylserine (PS) and phosphatidylethanolamine (PE) to the outer leaflet provides substrate for the aminophospholipid translocase (APLT), which flips back PS and PE inward faster than PC or SM can diffuse outward--thereby producing inner layer expansion or stomatocytosis. This role of APLT accounts for the vanadate inhibition of amphipath stomatocytosis. However, the vanadate effect can be overcome by increasing the amphipath concentration, which at such levels probably passively expands the inner leaflet.
APA, Harvard, Vancouver, ISO, and other styles
14

Schrier, SL, A. Zachowski, and PF Devaux. "Mechanisms of amphipath-induced stomatocytosis in human erythrocytes." Blood 79, no. 3 (February 1, 1992): 782–86. http://dx.doi.org/10.1182/blood.v79.3.782.bloodjournal793782.

Full text
Abstract:
We studied stomatocytosis induced in human red blood cells (RBC) by vinblastine and chlorpromazine, monitoring the movements of spin- labeled phosphatidylcholine (PC*) and sphingomyelin (SM*) by electron spin resonance (ESR) spectroscopy. This technique allows determination of the fraction of labeled lipids, respectively, on the external leaflet, on the cytosol face, or trapped in endocytic vacuoles. Both vinblastine and chlorpromazine produce a time- and concentration- dependent stomatocytic shape change, which is paralleled by a shift of approximately 10% to 33% of outer leaflet SM* and PC* inward. Of this amount, 8% to 12% was trapped in endocytic vacuoles and 8% to 19% had flipped to the inner leaflet. Vanadate, while inhibiting the stomatocytosis, did not block the flip of either SM* or PC* to the inner leaflet. To explain the inhibiting effect of vanadate, as well as the adenosine triphosphate (ATP) requirement for drug-induced stomatocytosis, we propose the following model: (1) addition of amphipath partially scrambles the bilayer; and (2) the flop of phosphatidylserine (PS) and phosphatidylethanolamine (PE) to the outer leaflet provides substrate for the aminophospholipid translocase (APLT), which flips back PS and PE inward faster than PC or SM can diffuse outward--thereby producing inner layer expansion or stomatocytosis. This role of APLT accounts for the vanadate inhibition of amphipath stomatocytosis. However, the vanadate effect can be overcome by increasing the amphipath concentration, which at such levels probably passively expands the inner leaflet.
APA, Harvard, Vancouver, ISO, and other styles
15

Kapil, Menka, and Rateesh Sareen. "Case of Acquired stomatocytosis in an alcoholic patient." Annals of Advance Medical Sciecnes 1, no. 1 (December 9, 2017): C4—C6. http://dx.doi.org/10.21276/aams.1781.

Full text
APA, Harvard, Vancouver, ISO, and other styles
16

Neff, Anne T. "Sitosterolemia's stomatocytosis and macrothrombocytopenia." Blood 120, no. 22 (November 22, 2012): 4283. http://dx.doi.org/10.1182/blood-2012-06-429449.

Full text
APA, Harvard, Vancouver, ISO, and other styles
17

Flatt, J. F., and L. J. Bruce. "The hereditary stomatocytoses." Haematologica 94, no. 8 (July 30, 2009): 1039–41. http://dx.doi.org/10.3324/haematol.2009.010041.

Full text
APA, Harvard, Vancouver, ISO, and other styles
18

Diallo, Safiatou, Anne Demulder, Kathleen Freson, Alina Ferster, and Laurence Rozen. "Macrothrombocytopenia and stomatocytosis in sitosterolaemia." British Journal of Haematology 194, no. 5 (April 26, 2021): 804. http://dx.doi.org/10.1111/bjh.17518.

Full text
APA, Harvard, Vancouver, ISO, and other styles
19

Göttgens, Eva‐Leonne, Peter C. Ligthart, Barbera Veldhuisen, Martijn Veldthuis, Masja Haas, and Adriaan J. Gammeren. "Rh‐null phenotype and stomatocytosis." British Journal of Haematology 194, no. 5 (April 26, 2021): 803. http://dx.doi.org/10.1111/bjh.17486.

Full text
APA, Harvard, Vancouver, ISO, and other styles
20

Frederiksen, Henrik. "Dehydrated hereditary stomatocytosis: clinical perspectives." Journal of Blood Medicine Volume 10 (July 2019): 183–91. http://dx.doi.org/10.2147/jbm.s179764.

Full text
APA, Harvard, Vancouver, ISO, and other styles
21

Bruce, L. J., J. F. Flatt, H. Guizouarn, N. M. Burton, F. Borgese, P. Quittet, P. Aguilar-Martinez, G. W. Stewart, and J. Delaunay. "Membrane alterations in hereditary stomatocytosis." Transfusion Clinique et Biologique 17, no. 3 (September 2010): 201. http://dx.doi.org/10.1016/j.tracli.2010.06.016.

Full text
APA, Harvard, Vancouver, ISO, and other styles
22

MORITA, Hiroshi, Hideo WADA, Hirohisa TSUJINOUE, Takaaki KURIOKA, Yukio UENO, Kanji SHICHIKAWA, Hiroyuki AMANO, et al. "Variant Type of Congenital Stomatocytosis." Internal Medicine 33, no. 2 (1994): 103–6. http://dx.doi.org/10.2169/internalmedicine.33.103.

Full text
APA, Harvard, Vancouver, ISO, and other styles
23

Andolfo, Immacolata, Roberta Russo, Antonella Gambale, and Achille Iolascon. "Hereditary stomatocytosis: An underdiagnosed condition." American Journal of Hematology 93, no. 1 (October 23, 2017): 107–21. http://dx.doi.org/10.1002/ajh.24929.

Full text
APA, Harvard, Vancouver, ISO, and other styles
24

Shao, Jingxin, Shoupeng Cao, Hanglong Wu, Loai K. E. A. Abdelmohsen, and Jan C. M. van Hest. "Therapeutic Stomatocytes with Aggregation Induced Emission for Intracellular Delivery." Pharmaceutics 13, no. 11 (November 2, 2021): 1833. http://dx.doi.org/10.3390/pharmaceutics13111833.

Full text
Abstract:
Bowl-shaped biodegradable polymersomes, or stomatocytes, have much potential as drug delivery systems, due to their intriguing properties, such as controllable size, programmable morphology, and versatile cargo encapsulation capability. In this contribution, we developed well-defined therapeutically active stomatocytes with aggregation-induced emission (AIE) features by self-assembly of biodegradable amphiphilic block copolymers, comprising poly(ethylene glycol) (PEG) and AIEgenic poly(trimethylene carbonate) (PTMC) moieties. The presence of the AIEgens endowed the as-prepared stomatocytes with intrinsic fluorescence, which was employed for imaging of cellular uptake of the particles. It simultaneously enabled the photo-mediated generation of reactive oxygen species (ROS) for photodynamic therapy. The potential of the therapeutic stomatocytes as cargo carriers was demonstrated by loading enzymes (catalase and glucose oxidase) in the nanocavity, followed by a cross-linking reaction to achieve stable encapsulation. This provided the particles with a robust motile function, which further strengthened their therapeutic effect. With these unique features, enzyme-loaded AIEgenic stomatocytes are an attractive platform to be exploited in the field of nanomedicine.
APA, Harvard, Vancouver, ISO, and other styles
25

Barneaud-Rocca, Damien, Bernard Pellissier, Franck Borgese, and Hélène Guizouarn. "Band 3 Missense Mutations and Stomatocytosis: Insight into the Molecular Mechanism Responsible for Monovalent Cation Leak." International Journal of Cell Biology 2011 (2011): 1–8. http://dx.doi.org/10.1155/2011/136802.

Full text
Abstract:
Missense mutations in the erythroid band 3 protein (Anion Exchanger 1) have been associated with hereditary stomatocytosis. Features of cation leaky red cells combined with functional expression of the mutated protein led to the conclusion that the AE1 point mutations were responsible for and leak through a conductive mechanism. A molecular mechanism explaining mutated AE1-linked stomatocytosis involves changes in AE1 transport properties that become leaky to and . However, another explanation suggests that point-mutated AE1 could regulate a cation leak through other transporters. This short paper intends to discuss these two alternatives.
APA, Harvard, Vancouver, ISO, and other styles
26

Smith, Brian D., and George B. Segel. "Abnormal Erythrocyte Endothelial Adherence in Hereditary Stomatocytosis." Blood 89, no. 9 (May 1, 1997): 3451–56. http://dx.doi.org/10.1182/blood.v89.9.3451.

Full text
Abstract:
Abstract Hereditary stomatocytosis is a red cell membrane protein disorder, which results in hemolytic anemia. Some patients with hereditary stomatocytosis experience dyspnea, chest pain, and abdominal pain, particularly after splenectomy. These symptoms may represent vaso-occlusion secondary to adherence of an abnormal erythrocyte membrane to vascular endothelium. We studied three members of a family with varying clinical expression of hereditary stomatocytosis. Adherence of red cells to endothelium was quantified by measuring the shear force required to separate individual cells from endothelial monolayers using a micropipette technique. Two patients with symptoms of in situ thromboses had a higher percentage of adherent cells compared with their asymptomatic sibling and normal controls. Correlation between this in vitro phenomenon and the clinical course suggests that flow abnormalities in the microcirculation attributable to erythrocyte endothelial adherence may play an important pathogenetic role in the illness. When the proportion of adherent red cells was reduced by a chronic transfusion program in one patient and pentoxifyllin therapy in another, the vaso-occlusive complications were eliminated.
APA, Harvard, Vancouver, ISO, and other styles
27

Slappendel, RJ, W. Renooij, and JJ de Bruijne. "Normal cations and abnormal membrane lipids in the red blood cells of dogs with familial stomatocytosis-hypertrophic gastritis." Blood 84, no. 3 (August 1, 1994): 904–9. http://dx.doi.org/10.1182/blood.v84.3.904.bloodjournal843904.

Full text
Abstract:
Examination of the red blood cells (RBCs) of eight dogs with familial stomatocytosis-hypertrophic gastritis (FS-HG), a multiorgan disease associated with hemolytic anemia, hereditary stomatocytosis (HSt), and hypertrophic gastritis resembling Menetrier's disease in man, showed abnormal osmotic fragility, normal mean corpuscular volume, slightly increased cell water, and normal cation content and cation fluxes. Cholesterol was decreased in RBC and increased in plasma. In both RBCs and plasma, total phospholipid (PL) was normal, phosphatidylcholine (PC) decreased, and sphingomyelin increased. The palmitic acid content of PC was increased, and the stearic acid content of PC was decreased. Sodium dodecyl sulfate electrophoresis of RBC membrane proteins was normal. These findings have not been described previously in HSt. They suggest that in FS-HG, abnormal composition of the PL in RBCs secondary to abnormal PL in plasma causes defective membrane function and stomatocytic shape-change. This conclusion was supported by a shortened half-life of 51Cr-labeled RBCs from normal dogs after transfusion in dogs with FS-HG. It was concluded (1) that not all hereditary forms of stomatocytosis are necessarily associated with an intrinsic structural defect of the RBC membrane, but that the change in shape of RBC may also be induced by abnormal composition of the plasma; (2) that stomatocytosis may be caused by loss of membrane surface area rather than by the increased cation uptake such as has been shown in some human kindreds with HSt, (3) that FS-HG is a disorder of lipid metabolism, and by consequence, (4) that abnormal lipid metabolism might be involved in the pathogenesis of Menetrier's disease.
APA, Harvard, Vancouver, ISO, and other styles
28

Slappendel, RJ, W. Renooij, and JJ de Bruijne. "Normal cations and abnormal membrane lipids in the red blood cells of dogs with familial stomatocytosis-hypertrophic gastritis." Blood 84, no. 3 (August 1, 1994): 904–9. http://dx.doi.org/10.1182/blood.v84.3.904.904.

Full text
Abstract:
Abstract Examination of the red blood cells (RBCs) of eight dogs with familial stomatocytosis-hypertrophic gastritis (FS-HG), a multiorgan disease associated with hemolytic anemia, hereditary stomatocytosis (HSt), and hypertrophic gastritis resembling Menetrier's disease in man, showed abnormal osmotic fragility, normal mean corpuscular volume, slightly increased cell water, and normal cation content and cation fluxes. Cholesterol was decreased in RBC and increased in plasma. In both RBCs and plasma, total phospholipid (PL) was normal, phosphatidylcholine (PC) decreased, and sphingomyelin increased. The palmitic acid content of PC was increased, and the stearic acid content of PC was decreased. Sodium dodecyl sulfate electrophoresis of RBC membrane proteins was normal. These findings have not been described previously in HSt. They suggest that in FS-HG, abnormal composition of the PL in RBCs secondary to abnormal PL in plasma causes defective membrane function and stomatocytic shape-change. This conclusion was supported by a shortened half-life of 51Cr-labeled RBCs from normal dogs after transfusion in dogs with FS-HG. It was concluded (1) that not all hereditary forms of stomatocytosis are necessarily associated with an intrinsic structural defect of the RBC membrane, but that the change in shape of RBC may also be induced by abnormal composition of the plasma; (2) that stomatocytosis may be caused by loss of membrane surface area rather than by the increased cation uptake such as has been shown in some human kindreds with HSt, (3) that FS-HG is a disorder of lipid metabolism, and by consequence, (4) that abnormal lipid metabolism might be involved in the pathogenesis of Menetrier's disease.
APA, Harvard, Vancouver, ISO, and other styles
29

CARRAGHER, PETER, HELEN HEGARTY, IAN TEMPERLEY, and MICHAEL McKILLEN. "The characterization of an hereditary stomatocytosis." Biochemical Society Transactions 15, no. 2 (April 1, 1987): 283–84. http://dx.doi.org/10.1042/bst0150283.

Full text
APA, Harvard, Vancouver, ISO, and other styles
30

Bienzle, U., D. Niethammer, U. Kleeberg, K. Ungefehr, E. Kohne, and E. Kleihauer. "Congenital Stomatocytosis and Chronic Haemolytic Anaemia." Scandinavian Journal of Haematology 15, no. 5 (April 24, 2009): 339–46. http://dx.doi.org/10.1111/j.1600-0609.1975.tb01089.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
31

Wislöff, F., and D. Boman. "Acquired Stomatocytosis in Alcoholic Liver Disease." Scandinavian Journal of Haematology 23, no. 1 (April 24, 2009): 43–50. http://dx.doi.org/10.1111/j.1600-0609.1979.tb02852.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
32

Paessler, Michele, and Helge Hartung. "Dehydrated hereditary stomatocytosis masquerading as MDS." Blood 125, no. 11 (March 12, 2015): 1841. http://dx.doi.org/10.1182/blood-2014-11-612184.

Full text
APA, Harvard, Vancouver, ISO, and other styles
33

Bonfanti, Ugo, Stefano Comazzi, Saverio Paltrinieri, and Walter Bertazzolo. "Stomatocytosis in 7 related Standard Schnauzers." Veterinary Clinical Pathology 33, no. 4 (December 2004): 234–39. http://dx.doi.org/10.1111/j.1939-165x.2004.tb00379.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
34

Wang, Gaifeng, Lijuan Cao, Zhaoyue Wang, Minghua Jiang, Xionghua Sun, Xia Bai, and Changgeng Ruan. "Macrothrombocytopenia/Stomatocytosis Specially Associated With Phytosterolemia." Clinical and Applied Thrombosis/Hemostasis 18, no. 6 (January 31, 2012): 582–87. http://dx.doi.org/10.1177/1076029611435090.

Full text
APA, Harvard, Vancouver, ISO, and other styles
35

Zhu, Yiwen, Chris Paszty, Tikva Turetsky, Susan Tsai, Frans A. Kuypers, Gloria Lee, Philip Cooper, et al. "Stomatocytosis Is Absent in “Stomatin”-Deficient Murine Red Blood Cells." Blood 93, no. 7 (April 1, 1999): 2404–10. http://dx.doi.org/10.1182/blood.v93.7.2404.

Full text
Abstract:
Abstract To examine the relationship between erythrocyte membrane protein 7.2b deficiency and the hemolytic anemia of human hereditary stomatocytosis, we created 7.2b knock-out mice by standard gene targeting approaches. Immunoblots showed that homozygous knock-out mice completely lacked erythrocyte protein 7.2b. Despite the absence of protein 7.2b, there was no hemolytic anemia and mouse red blood cells (RBCs) were normal in morphology, cell indices, hydration status, monovalent cation content, and ability to translocate lipids. The absence of the phenotype of hereditary stomatocytosis implies that protein 7.2b deficiency plays no direct role in the etiology of this disorder and casts doubt on the previously proposed role of this protein as a mediator of cation transport in RBC.
APA, Harvard, Vancouver, ISO, and other styles
36

Zhu, Yiwen, Chris Paszty, Tikva Turetsky, Susan Tsai, Frans A. Kuypers, Gloria Lee, Philip Cooper, et al. "Stomatocytosis Is Absent in “Stomatin”-Deficient Murine Red Blood Cells." Blood 93, no. 7 (April 1, 1999): 2404–10. http://dx.doi.org/10.1182/blood.v93.7.2404.407k13_2404_2410.

Full text
Abstract:
To examine the relationship between erythrocyte membrane protein 7.2b deficiency and the hemolytic anemia of human hereditary stomatocytosis, we created 7.2b knock-out mice by standard gene targeting approaches. Immunoblots showed that homozygous knock-out mice completely lacked erythrocyte protein 7.2b. Despite the absence of protein 7.2b, there was no hemolytic anemia and mouse red blood cells (RBCs) were normal in morphology, cell indices, hydration status, monovalent cation content, and ability to translocate lipids. The absence of the phenotype of hereditary stomatocytosis implies that protein 7.2b deficiency plays no direct role in the etiology of this disorder and casts doubt on the previously proposed role of this protein as a mediator of cation transport in RBC.
APA, Harvard, Vancouver, ISO, and other styles
37

Jamwal, Manu, Anu Aggarwal, Prashant Sharma, Deepak Bansal, Pankaj Malhotra, Arindam Maitra, and Reena Das. "The Spectrum of Genetic Defects in Indian Patients with Rare Congenital Anemias: Next Generation Sequencing Based Approach." Blood 132, Supplement 1 (November 29, 2018): 2328. http://dx.doi.org/10.1182/blood-2018-99-116092.

Full text
Abstract:
Abstract Introduction Diagnosis of inherited anemias is based on automated red cell indices, morphology and reticulocyte count in an appropriate clinical presentation and a stepwise diagnostic algorithm needs to be followed. Major inherited causes of hemolysis include the hemoglobinopathies, membranopathies and enzymopathies. In contrast, ineffective erythropoiesis, a term that refers to a disturbance in proliferation, differentiation and maturation of erythroblasts, includes megaloblastic anemia, myelodysplastic syndromes, thalassemias, and congenital dyserythropoietic anemias (CDA). Phenotypes vary from severe transfusion-dependent hemolytic anemia (HA) to fully compensated hemolysis. Patients are encountered where the entire spectrum of diagnostic tests cannot identify the etiology. Causal genes implicated are numerous, making a gene-by-gene approach time consuming, expensive and laborious. Hence, the use of targeted resequencing can expedite molecular diagnosis, genetic counseling and prenatal diagnosis if indicated. Objectives This study aimed at determining the spectrum of mutations in uncommon HA and CDA. Methods Sixty-one patients (HA=44, CDA=17) with clinical and laboratory evidence suggestive of HA/CDA were enrolled after excluding common causes of anemias. Various biochemical and molecular tests were used to exclude glucose-6-phosphate dehydrogenase (G6PD) deficiency, hemoglobinopathies, autoimmune HA, hereditary spherocytosis and pyruvate kinase (PK) deficiency. Peripheral blood genomic DNA was extracted using QIAamp DNA Blood Midi Kit, quantified on NanoDrop 2000 spectrophotometer and Qubit® 2.0 Fluorometer. Common G6PD, PKLR variants, SEC23B mutation were excluded by molecular tests. Since the genetic defects in Indian patients with unexplained HA/CDA are not available, we first tested 16 patients with a comprehensive commercially available sequencing panel that covers 4,813 clinically-relevant genes. We next sought to streamline testing and designed a customized panel of 55 genes (Illumina). Variants were annotated and classified and the most likely disease-causing variants were validated by Sanger sequencing in the patient and available affected and unaffected family members. Results Out of 61 patient's sequenced, unexpected pyruvate kinase (PK) deficiency were found in 11 patients. PK enzyme activity assay was within normal limits in all these cases. Three patients with G6PD deficiency, 2 patients with glucose-6-phosphate isomerase deficiency and 1 case with hexokinase deficiency were found. Of 5 patients with stomatocytes on peripheral blood film, 3 had Mediterranean stomatocytosis/ macrothrombocytopenia (ABCG5/ABCG8) and 2 were found to have overhydrated hereditary stomatocytosis (RHAG). We also found 6 cases of HA to have a mutation in erythrocyte membrane protein-coding genes. Heterozygous nonsense mutation in the ANK1 gene was found in 2 cases and compound heterozygous missense mutations in ANK1 were seen in 1. Missense mutations were also found in SPTB gene in other 3 cases. Xerocytosis or dehydrated hereditary stomatocytosis (PEIZO1) was found in 3 patients. Seventeen cases for suspected CDA were studied and SEC23B mutations were present in 7. In one patient only heterozygous mutation was found. There could be the possibility of intronic mutation or long deletion/insertion in her. Notably, 6 patients with CDA showed mutations in PKLR (n=3), 1 each had mutations in MTRR, SPTB and PIEZO1 genes. We also found a mutation in GATA1 gene in a patient of CDA with thrombocytopenia. Conclusion(s) HA/CDA showed highly varied etiology. Our experience demonstrates the high diagnostic yield (>75%) of targeted resequencing for molecular diagnosis of unexplained anemias. It is cost effective and can expedite the diagnosis where conventional testing is not helpful. Timely detection of the etiology was helpful in genetic counseling. It not only offered therapeutic benefits to our patients but may help us in future antenatal diagnosis if required. Therapeutic implications include performing splenectomy which can ameliorate the anemia in selected subgroups of patients (HS, PK deficiency, CDA type II). Disclosures No relevant conflicts of interest to declare.
APA, Harvard, Vancouver, ISO, and other styles
38

Westhoff, Connie M., and Seth Alper. "Hereditary Stomatocytosis Associated with a Loss of Function Mutation In Rh-Associated Glycoprotein (RhAG)." Blood 116, no. 21 (November 19, 2010): 2040. http://dx.doi.org/10.1182/blood.v116.21.2040.2040.

Full text
Abstract:
Abstract Abstract 2040 The erythroid Rh family of proteins includes RhCE and RhD which carry the common Rh antigens, and the related Rh-associated glycoprotein, RhAG. RhAG is required for trafficking of the blood group proteins to the membrane and forms the core of a macro-complex in the membrane which includes glycophorin B, Band 3, CD47, and LW. The Rh proteins are structurally and functionally related to the Amt superfamily of NH3/NH4+ transport proteins, and RhAG and its nonerythroid paralogs, RhCG and RhBG, have been shown to mediate NH3/NH4+ transport. RhCG is responsible for part of renal collecting duct epithelial cell NH3/NH4+ secretion, and Rhcg-/- mice exhibit incomplete distal renal tubular acidosis due to impaired urinary NH4+ excretion. The Rhag-/- mouse is grossly normal, and the significance of RhAG-mediated NH3/NH4+ transport in human erythrocytes remains unclear. Over-hydrated hereditary stomatocytosis (OHSt) is a rare dominant disorder characterized by moderate hemolytic anemia, increased mean red cell volumes, stomatocytes and echinocytes, and increased red cell permeability to the monovalent cations, Na+ and K+. Six of the seven OHSt kindred studied by Bruce et al. (Blood. 2009;113:1350) displayed a heterozygous Phe65Ser mutation in RhAG. Expression studies of the mutant 65Ser-RhAG in Xenopus oocytes induced a monovalent cation flux compatible with the cation leak seen in RBCs. The increased Na+ and decreased K+ contents of mutant RhAG-expressing oocytes suggested that F65S is a gain-of-function mutation that opens a cation leak, likely within the RhAG polypeptide. In this study the ammonia transport properties of the OHSt mutant 65Ser-RhAG were investigated. Xenopus oocytes were injected with cRNA encoding wild-type RhAG, the OHSt mutant 65Ser-RhAG, and 65Val-RhAG, an engineered mutation with a smaller hydrophobic side chain at position 65. Wild-type and mutant RhAG polypeptides were well-expressed in the oocyte membrane as measured by quantitative immunoblotting. Uptake of the NH3/NH4+ substrate analog 14C-methylammonium (MA), was assayed in oocytes previously injected with water (control) or with cRNA. Expression of wild-type RhAG mediated MA uptake at rates 6-fold greater than that of water-injected controls. Uptake of MA by oocytes expressing 65Val-RhAG was equivalent to that of wild type RhAG. However, MA uptake by oocytes expressing OHSt mutant 65Ser-RhAG was greatly reduced to less than 20% that of oocytes expressing wild-type RHAG or 65Val-RhAG, and was only 1.5-fold greater than that of water-injected control oocytes. Co-expression with other, individual Rh complex members glycophorin B, RhD, RhCE, or Band 3 did not alter MA-mediated uptake by RhAG-expressing oocytes. Importantly, this study reveals that the RhAG mutation Phe65Ser found in patients with type 1 over-hydrated stomatocytosis is a loss of function mutation. Further study is required to define the relationship between loss of NH3/NH4+ transport and erythrocyte Na+ and K+ cation content. Disclosures: Westhoff: Immucor: Scientific Advisor.
APA, Harvard, Vancouver, ISO, and other styles
39

Andolfo, Immacolata, Seth L. Alper, Lucia De Franceschi, Carla Auriemma, Roberta Russo, Luigia De Falco, Fara Vallefuoco, et al. "Multiple clinical forms of dehydrated hereditary stomatocytosis arise from mutations in PIEZO1." Blood 121, no. 19 (May 9, 2013): 3925–35. http://dx.doi.org/10.1182/blood-2013-02-482489.

Full text
Abstract:
Key Points Dehydrated hereditary stomatocytosis is characterized by abnormal RBC morphology but may involve pseudohyperkalemia and perinatal edema. This syndrome is associated with germline mutations in PIEZO1, encoding a transmembrane protein that induces mechanosensitive currents.
APA, Harvard, Vancouver, ISO, and other styles
40

Seyed, Z. B., S. S. Tabrizi, and G. R. Abdollahpo. "Stomatocytosis in the Ruminant: A First Report." Asian Journal of Animal and Veterinary Advances 6, no. 11 (October 15, 2011): 1089–93. http://dx.doi.org/10.3923/ajava.2011.1089.1093.

Full text
APA, Harvard, Vancouver, ISO, and other styles
41

Konstantopoulos, K., G. Vassilopoulos, S. Adamides, M. Alexandrakis, and J. Zervas. "Stomatocytosis as a presenting symptom of myelodysplasia." Medical Oncology and Tumor Pharmacotherapy 9, no. 4 (December 1992): 213–14. http://dx.doi.org/10.1007/bf02987759.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Poret, H., E. G. Simon, P. Hervé, and F. Perrotin. "Dehydrated hereditary stomatocytosis and recurrent prenatal ascites." Journal of Obstetrics and Gynaecology 33, no. 5 (July 2013): 527. http://dx.doi.org/10.3109/01443615.2013.781142.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Basu, A. P. "Dehydrated hereditary stomatocytosis with transient perinatal ascites." Archives of Disease in Childhood - Fetal and Neonatal Edition 88, no. 5 (September 1, 2003): 438F—439. http://dx.doi.org/10.1136/fn.88.5.f438.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Stewart, Gordon W., John Lloyd, and Karl Pegel. "Mediterranean stomatocytosis/macrothrombocytopenia: update from Adelaide, Australia." British Journal of Haematology 132, no. 5 (March 2006): 660–61. http://dx.doi.org/10.1111/j.1365-2141.2006.05973.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Stewart, G. W., H. O'Brien, S. A. Morris, J. S. Owen, J. K. Lloyd, and J. A. L. Ames. "Stomatocytosis, abnormal platelets and pseudo-homozygous hypercholesterolaemia." European Journal of Haematology 38, no. 4 (April 24, 2009): 376–80. http://dx.doi.org/10.1111/j.1600-0609.1987.tb00014.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Stewart, Gordon W. "3 The membrane defect in hereditary stomatocytosis." Baillière's Clinical Haematology 6, no. 2 (June 1993): 371–99. http://dx.doi.org/10.1016/s0950-3536(05)80151-9.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

Delaunay, Jean, Gordon Stewart, and Achille Iolascon. "Hereditary dehydrated and overhydrated stomatocytosis: recent advances." Current Opinion in Hematology 6, no. 2 (March 1999): 110. http://dx.doi.org/10.1097/00062752-199903000-00009.

Full text
APA, Harvard, Vancouver, ISO, and other styles
48

Yoshimoto, Akihiro, Masaki Fujimura, and Shinji Nakao. "Pulmonary Hypertension after Splenectomy in Hereditary Stomatocytosis." American Journal of the Medical Sciences 330, no. 4 (October 2005): 195–97. http://dx.doi.org/10.1097/00000441-200510000-00008.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Frumence, Etienne, Sandrine Genetet, Pierre Ripoche, Achille Iolascon, Immacolata Andolfo, Caroline Le Van Kim, Yves Colin, Isabelle Mouro-Chanteloup, and Claude Lopez. "Rapid Cl−/HCO3−exchange kinetics of AE1 in HEK293 cells and hereditary stomatocytosis red blood cells." American Journal of Physiology-Cell Physiology 305, no. 6 (September 15, 2013): C654—C662. http://dx.doi.org/10.1152/ajpcell.00142.2013.

Full text
Abstract:
Anion exchanger 1 (AE1) or band 3 is a membrane protein responsible for the rapid exchange of chloride for bicarbonate across the red blood cell membrane. Nine mutations leading to single amino-acid substitutions in the transmembrane domain of AE1 are associated with dominant hereditary stomatocytosis, monovalent cation leaks, and reduced anion exchange activity. We set up a stopped-flow spectrofluorometry assay coupled with flow cytometry to investigate the anion transport and membrane expression characteristics of wild-type recombinant AE1 in HEK293 cells, using an inducible expression system. Likewise, study of three stomatocytosis-associated mutations (R730C, E758K, and G796R), allowed the validation of our method. Measurement of the rapid and specific chloride/bicarbonate exchange by surface expressed AE1 showed that E758K mutant was fully active compared with wild-type (WT) AE1, whereas R730C and G796R mutants were inactive, reinforcing previously reported data on other experimental models. Stopped-flow analysis of AE1 transport activity in red blood cell ghost preparations revealed a 50% reduction of G796R compared with WT AE1 corresponding to a loss of function of the G796R mutated protein, in accordance with the heterozygous status of the AE1 variant patients. In conclusion, stopped-flow led to measurement of rapid transport kinetics using the natural substrate for AE1 and, conjugated with flow cytometry, allowed a reliable correlation of chloride/bicarbonate exchange to surface expression of AE1, both in recombinant cells and ghosts and therefore a fine comparison of function between different stomatocytosis samples. This technical approach thus provides significant improvements in anion exchange analysis in red blood cells.
APA, Harvard, Vancouver, ISO, and other styles
50

Mohandas, Narla. "To shrink or not to shrink." Blood 121, no. 19 (May 9, 2013): 3783–84. http://dx.doi.org/10.1182/blood-2013-04-491753.

Full text
Abstract:
In this issue of Blood, Andolfo and colleagues show that dehydrated hereditary stomatocytosis (DHSt), an inherited red cell disorder, is associated with a number of distinct germline mutations in PIEZO1, a stretch activated cation channel, in 26 affected individuals from 7 families.1
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography