Academic literature on the topic 'Human ferritin'

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

Select a source type:

Consult the lists of relevant articles, books, theses, conference reports, and other scholarly sources on the topic 'Human ferritin.'

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.

Journal articles on the topic "Human ferritin"

1

Ghosh, Sharmistha, Sarah Hevi, and Steven L. Chuck. "Regulated secretion of glycosylated human ferritin from hepatocytes." Blood 103, no. 6 (March 15, 2004): 2369–76. http://dx.doi.org/10.1182/blood-2003-09-3050.

Full text
Abstract:
Abstract Serum ferritin has been used widely in clinical medicine chiefly as an indicator of iron stores and inflammation. Circulating ferritin also can have paracrine effects. Despite the clinical significance of serum ferritin, its secretion remains an enigma. The consensus view is that serum ferritin arises from tissue ferritins— principally ferritin light—which can be glycosylated. Ferritin heavy and light chains are cytosolic proteins that form cages of 24 subunits to store intracellular iron. We show that ferritin light is secreted when its expression is increased in stable, transfected HepG2 cells or adenovirus-infected HepG2 cells. Export occurs through the classical secretory pathway and some chains are N-glycosylated. Ferritins do not need to form cages prior to secretion. Secretion is blocked specifically, effectively, and rapidly by a factor in serum. The timing of this inhibition of ferritin secretion suggests that normally cytosolic ferritin L is targeted to the secretory pathway during translation despite the absence of a conventional signal sequence. Thus, secretion of glycosylated and unglycosylated ferritin is a regulated and not a stochastic process.
APA, Harvard, Vancouver, ISO, and other styles
2

CORSI, Barbara, Federica PERRONE, Monique BOURGEOIS, Carole BEAUMONT, C. Maria PANZERI, Anna COZZI, Romina SANGREGORIO, et al. "Transient overexpression of human H- and L-ferritin chains in COS cells." Biochemical Journal 330, no. 1 (February 15, 1998): 315–20. http://dx.doi.org/10.1042/bj3300315.

Full text
Abstract:
The understanding of the in vitro mechanisms of ferritin iron incorporation has greatly increased in recent years with the studies of recombinant and mutant ferritins. However, little is known about how this protein functions in vivo, mainly because of the lack of cellular models in which ferritin expression can be modulated independently from iron. To this aim, primate fibroblastoid COS-7 cells were transiently transfected with cDNAs for human ferritin H- and L-chains under simian virus 40 promoter and analysed within 66 h. Ferritin accumulation reached levels 300-500-fold higher than background, with about 40% of the cells being transfected. Thus ferritin concentration in individual cells was increased up to 1000-fold over controls with no evident signs of toxicity. The exogenous ferritin subunits were correctly assembled into homopolymers, but did not affect either the size or the subunit composition of the endogenous heteropolymeric fraction of ferritin, which remained essentially unchanged in the transfected and non-transfected cells. After 18 h of incubation with [59Fe]ferric-nitrilotriacetate, cellular iron incorporation was similar in the transfected and non-transfected cells and most of the protein-bound radioactivity was associated with ferritin heteropolymers, while H- and L-homopolymers remained iron-free. Cell co-transfection with cDNAs for H- and L-chains produced ferritin heteropolymers that also did not increase cellular iron incorporation. It is concluded that transient transfection of COS cells induces a high level of expression of ferritin subunits that do not co-assemble with the endogenous ferritins and have no evident activity in iron incorporation/metabolism.
APA, Harvard, Vancouver, ISO, and other styles
3

Lobreaux, S., S. J. Yewdall, J. F. Briat, and P. M. Harrison. "Amino-acid sequence and predicted three-dimensional structure of pea seed (Pisum sativum) ferritin." Biochemical Journal 288, no. 3 (December 15, 1992): 931–39. http://dx.doi.org/10.1042/bj2880931.

Full text
Abstract:
The iron storage protein, ferritin, is widely distributed in the living kingdom. Here the complete cDNA and derived amino-acid sequence of pea seed ferritin are described, together with its predicted secondary structure, namely a four-helix-bundle fold similar to those of mammalian ferritins, with a fifth short helix at the C-terminus. An N-terminal extension of 71 residues contains a transit peptide (first 47 residues) responsible for plastid targetting as in other plant ferritins, and this is cleaved before assembly. The second part of the extension (24 residues) belongs to the mature subunit; it is cleaved during germination. The amino-acid sequence of pea seed ferritin is aligned with those of other ferritins (49% amino-acid identity with H-chains and 40% with L-chains of human liver ferritin in the aligned region). A three-dimensional model has been constructed by fitting the aligned sequence to the coordinates of human H-chains, with appropriate modifications. A folded conformation with an 11-residue helix is predicted for the N-terminal extension. As in mammalian ferritins, 24 subunits assemble into a hollow shell. In pea seed ferritin, its N-terminal extension is exposed on the outside surface of the shell. Within each pea subunit is a ferroxidase centre resembling those of human ferritin H-chains except for a replacement of Glu-62 by His. The channel at the 4-fold-symmetry axes defined by E-helices, is predicted to be hydrophilic in plant ferritins, whereas it is hydrophobic in mammalian ferritins.
APA, Harvard, Vancouver, ISO, and other styles
4

Ebrahimi, Kourosh Honarmand, Eckhard Bill, Peter-Leon Hagedoorn, and Wilfred R. Hagen. "Spectroscopic evidence for the role of a site of the di-iron catalytic center of ferritins in tuning the kinetics of Fe(ii) oxidation." Molecular BioSystems 12, no. 12 (2016): 3576–88. http://dx.doi.org/10.1039/c6mb00235h.

Full text
Abstract:
Spectroscopic studies of human H-type ferritin in comparison with an archaeal ferritin from Pyrococcus furiosus reveal how kinetics of a common mechanism of Fe(ii) oxidation is tuned differently in these two ferritins.
APA, Harvard, Vancouver, ISO, and other styles
5

Pozzi, Cecilia, Flavio Di Pisa, Caterina Bernacchioni, Silvia Ciambellotti, Paola Turano, and Stefano Mangani. "Iron binding to human heavy-chain ferritin." Acta Crystallographica Section D Biological Crystallography 71, no. 9 (August 25, 2015): 1909–20. http://dx.doi.org/10.1107/s1399004715013073.

Full text
Abstract:
Maxi-ferritins are ubiquitous iron-storage proteins with a common cage architecture made up of 24 identical subunits of five α-helices that drive iron biomineralization through catalytic iron(II) oxidation occurring at oxidoreductase sites (OS). Structures of iron-bound human H ferritin were solved at high resolution by freezing ferritin crystals at different time intervals after exposure to a ferrous salt. Multiple binding sites were identified that define the iron path from the entry ion channels to the oxidoreductase sites. Similar data are available for another vertebrate ferritin: the M protein fromRana catesbeiana. A comparative analysis of the iron sites in the two proteins identifies new reaction intermediates and underlines clear differences in the pattern of ligands that define the additional iron sites that precede the oxidoreductase binding sites along this path. Stopped-flow kinetics assays revealed that human H ferritin has different levels of activity compared with itsR. catesbeianacounterpart. The role of the different pattern of transient iron-binding sites in the OS is discussed with respect to the observed differences in activity across the species.
APA, Harvard, Vancouver, ISO, and other styles
6

Fargion, S., AL Fracanzani, B. Brando, P. Arosio, S. Levi, and G. Fiorelli. "Specific binding sites for H-ferritin on human lymphocytes: modulation during cellular proliferation and potential implication in cell growth control." Blood 78, no. 4 (August 15, 1991): 1056–61. http://dx.doi.org/10.1182/blood.v78.4.1056.1056.

Full text
Abstract:
Abstract Interactions between human recombinant H- and L-ferritins and human lymphocytes were studied in vitro by direct binding assays and by flow cytometry. L-ferritin did not cause detectable specific binding, whereas H-ferritin showed a specific and saturable binding that increased markedly in phytohemagglutinin (PHA)-stimulated cells. This ferritin bound up to 30% of CD4+ and CD8+ T-lymphocytes and most B cells, indicating that expression of ferritin binding sites is not related to cell lineage or function. Dual-color flow cytometry experiments showed that ferritin binding sites were present on cells expressing the proliferation markers HLA-DR, MLR3, interleukin 2 (IL- 2), and transferrin receptors (Tf-R). In addition, after PHA induction, the time course of the expression of H-ferritin binding sites was similar to those of the above proliferation markers. Ferritin binding sites were observed in lymphocytes at all cell cycle phases, including the early S-phase. H-Ferritin at nanomolar and picomolar concentrations had an inhibitory effect on PHA-induced blastogenesis. We propose that H-ferritin binding sites behave like proliferation markers, with the unusual function of downregulating proliferation.
APA, Harvard, Vancouver, ISO, and other styles
7

Fargion, S., AL Fracanzani, B. Brando, P. Arosio, S. Levi, and G. Fiorelli. "Specific binding sites for H-ferritin on human lymphocytes: modulation during cellular proliferation and potential implication in cell growth control." Blood 78, no. 4 (August 15, 1991): 1056–61. http://dx.doi.org/10.1182/blood.v78.4.1056.bloodjournal7841056.

Full text
Abstract:
Interactions between human recombinant H- and L-ferritins and human lymphocytes were studied in vitro by direct binding assays and by flow cytometry. L-ferritin did not cause detectable specific binding, whereas H-ferritin showed a specific and saturable binding that increased markedly in phytohemagglutinin (PHA)-stimulated cells. This ferritin bound up to 30% of CD4+ and CD8+ T-lymphocytes and most B cells, indicating that expression of ferritin binding sites is not related to cell lineage or function. Dual-color flow cytometry experiments showed that ferritin binding sites were present on cells expressing the proliferation markers HLA-DR, MLR3, interleukin 2 (IL- 2), and transferrin receptors (Tf-R). In addition, after PHA induction, the time course of the expression of H-ferritin binding sites was similar to those of the above proliferation markers. Ferritin binding sites were observed in lymphocytes at all cell cycle phases, including the early S-phase. H-Ferritin at nanomolar and picomolar concentrations had an inhibitory effect on PHA-induced blastogenesis. We propose that H-ferritin binding sites behave like proliferation markers, with the unusual function of downregulating proliferation.
APA, Harvard, Vancouver, ISO, and other styles
8

Morikawa, K., F. Oseko, and S. Morikawa. "H- and L-rich ferritins suppress antibody production, but not proliferation, of human B lymphocytes in vitro." Blood 83, no. 3 (February 1, 1994): 737–43. http://dx.doi.org/10.1182/blood.v83.3.737.737.

Full text
Abstract:
Abstract The effect of human spleen(L-rich) and heart(H-rich) ferritins on the proliferation and differentiation of human B lymphocytes was studied in comparison with that of holo- and apo-transferrins. Ferritins rich in H and L chain, as well as the transferrins, did not inhibit the proliferative response of resting and activated B cells stimulated with polyclonal B-cell mitogen, Staphylococcus aureus Cowan strain I. In contrast, the ferritins, but not the transferrins, clearly suppressed the antibody production by B blasts in T-cell-independent as well as T- cell-dependent system. Kinetic study showed that inhibitory action of ferritins on immunoglobulin (Ig) production was caused at an early stage of B-cell differentiation. The cytoplasmic Ig-containing cells decreased in proportion to the reduction of Ig secretion. The evidence that ferritin inhibited Ig synthesis of Epstein-Barr virus-transformed human B-lymphoblastoid cell line also supported the idea that the effect of ferritin was directed toward the antibody-producing B lymphocytes. The molecular analysis showed that the inhibitory effect of ferritin was regulated at the transcriptional level of the Ig generation signal. Our results suggest that H- and L-rich ferritins exert their inhibitory action on the differentiation of B cells maturing into Ig-producing cells.
APA, Harvard, Vancouver, ISO, and other styles
9

Morikawa, K., F. Oseko, and S. Morikawa. "H- and L-rich ferritins suppress antibody production, but not proliferation, of human B lymphocytes in vitro." Blood 83, no. 3 (February 1, 1994): 737–43. http://dx.doi.org/10.1182/blood.v83.3.737.bloodjournal833737.

Full text
Abstract:
The effect of human spleen(L-rich) and heart(H-rich) ferritins on the proliferation and differentiation of human B lymphocytes was studied in comparison with that of holo- and apo-transferrins. Ferritins rich in H and L chain, as well as the transferrins, did not inhibit the proliferative response of resting and activated B cells stimulated with polyclonal B-cell mitogen, Staphylococcus aureus Cowan strain I. In contrast, the ferritins, but not the transferrins, clearly suppressed the antibody production by B blasts in T-cell-independent as well as T- cell-dependent system. Kinetic study showed that inhibitory action of ferritins on immunoglobulin (Ig) production was caused at an early stage of B-cell differentiation. The cytoplasmic Ig-containing cells decreased in proportion to the reduction of Ig secretion. The evidence that ferritin inhibited Ig synthesis of Epstein-Barr virus-transformed human B-lymphoblastoid cell line also supported the idea that the effect of ferritin was directed toward the antibody-producing B lymphocytes. The molecular analysis showed that the inhibitory effect of ferritin was regulated at the transcriptional level of the Ig generation signal. Our results suggest that H- and L-rich ferritins exert their inhibitory action on the differentiation of B cells maturing into Ig-producing cells.
APA, Harvard, Vancouver, ISO, and other styles
10

Bauminger, E. R., A. Treffry, A. J. Hudson, D. Hechel, N. W. Hodson, S. C. Andrews, S. Levi, et al. "Iron incorporation into ferritins: evidence for the transfer of monomeric Fe(III) between ferritin molecules and for the formation of an unusual mineral in the ferritin of Escherichia coli." Biochemical Journal 302, no. 3 (September 15, 1994): 813–20. http://dx.doi.org/10.1042/bj3020813.

Full text
Abstract:
Iron that has been oxidized by H-chain ferritin can be transferred into other ferritin molecules before it is incorporated into mature ferrihydrite iron cores. Iron(III) dimers are formed at the ferroxidase centres of ferritin H chains at an early stage of Fe(II) oxidation. Mössbauer spectroscopic data now show that the iron is transferred as monomeric species arising from dimer dissociation and that it binds to the iron core of the acceptor ferritin. Human H-chain ferritin variants containing altered threefold channels can act as acceptors, as can the ferritin of Escherichia coli (Ec-FTN). A human H-chain ferritin variant with a substituted tyrosine (rHuHF-Y34F) can act as a donor of Fe(III). Since an Fe(III)-tyrosinate (first identified in bullfrog H-chain ferritin) is absent from variant rHuHF-Y34F, the Fe(III) transferred is not derived from this tyrosinate complex. Mössbauer parameters of the small iron cores formed within Ec-FTN are significantly different from those of mammalian ferritins. Analysis of the spectra suggests that they are derived from both ferrihydrite and non-ferrihydrite components. This provides further evidence that the ferritin protein shell can influence the structure of its iron core.
APA, Harvard, Vancouver, ISO, and other styles

Dissertations / Theses on the topic "Human ferritin"

1

Flink, Håkan. "Unstimulated human whole saliva flow rate in relation to hyposalivation and dental caries /." Stockholm, 2005. http://diss.kib.ki.se/2005/91-7140-265-9/.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Sakamoto, Souichiro. "H-Ferritin Is Preferentially Incorporated by Human Erythroid Cells through Transferrin Receptor 1 in a Threshold-Dependent Manner." Kyoto University, 2016. http://hdl.handle.net/2433/215433.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Wilkins, Jennie P. "Relationship between maternal prenatal vitamin use and infant iron status." Morgantown, W. Va. : [West Virginia University Libraries], 2002. http://etd.wvu.edu/templates/showETD.cfm?recnum=2381.

Full text
Abstract:
Thesis (M.S.)--West Virginia University, 2002.
Title from document title page. Document formatted into pages; contains vi, 43 p. Vita. Includes abstract. Includes bibliographical references (p. 34-36).
APA, Harvard, Vancouver, ISO, and other styles
4

McVie, Alison. "Immunological characterisation of Echinococcus granulosus recombinant antigen B and ferritin for the serodiagnosis of human cystic echinococcosis." Thesis, University of Salford, 1997. http://ethos.bl.uk/OrderDetails.do?uin=uk.bl.ethos.244927.

Full text
APA, Harvard, Vancouver, ISO, and other styles
5

Govus, Andrew. "The regulation of human iron metabolism in hypoxia." Thesis, Edith Cowan University, Research Online, Perth, Western Australia, 2015. https://ro.ecu.edu.au/theses/1719.

Full text
Abstract:
Athletes commonly use altitude exposure in an attempt to improve their aerobic performance at sea level. Altitude exposure enhances erythropoiesis and iron-dependent oxidative and glycolytic enzyme production, for this reason, athletes must maintain a healthy iron balance at altitude. A negative iron balance at altitude may limit such physiological adaptations, potentially reducing the performance benefits of altitude exposure. This thesis examined the regulation of iron metabolism during acute (~31 min, Study One) and prolonged altitude exposure (14 days, Study Two). Finally, Study Three examined how daily oral iron supplementation influenced haemoglobin mass (Hbmass) and iron parameter responses to prolonged, moderate altitude exposure in a large cohort of elite athletes. Specifically, Study One found acute (~31 min) interval exercise [5 × 4 min at 90% of the maximal aerobic running velocity (vVO2max)] increased post-exercise interleukin-6 (IL-6) production and elevated hepcidin production 3 h thereafter in both normoxia (fraction of inspired oxygen (FIO2) = 0.2093) and normobaric hypoxia (i.e. 3,000 m simulated altitude; FIO2 = 0.1450). These results suggest exercise performed in acute hypoxia does not alter the post-exercise hepcidin response, relative to exercise in normoxia, possibly owing to the short duration of the hypoxic stimulus. Prolonged altitude exposure suppresses resting hepcidin levels in sojourning mountaineers, but its influence on the post-exercise hepcidin response exercise has not yet been investigated. Therefore, Study Two investigated how 14 days of live high: train low (LHTL) (exposure to 3,000 m simulated altitude for 14 h.d-1) influenced resting levels of hepcidin, erythropoietin (EPO) and blood iron parameters. Study Two also examined the post-exercise hepcidin and iron parameter responses to interval exercise (5 × 1,000 m at 90% of the maximal aerobic running velocity) performed in normoxia (600 m natural altitude) and normobaric hypoxia (i.e. ~3,000 m simulated altitude), following 11 and 14 days of LHTL. The post-exercise hepcidin response was compared with interval exercise performed at a matched exercise intensity in normoxia or hypoxia before LHTL. Here, LHTL suppressed resting hepcidin levels after two days of exposure, but the post-exercise hepcidin response to interval exercise was similar in normoxia and hypoxia, both before and after LHTL. Additionally, Hbmass increased by 2.2% and plasma ferritin levels decreased following LHTL. In conclusion, prolonged, moderate altitude exposure suppresses resting hepcidin levels, which likely ensures more iron can be transported to the erythron to support accelerated erythropoiesis. Prolonged altitude exposure places a large burden on body iron stores because additional iron is required to support accelerated erythropoiesis. Accordingly, athletes often ingest oral iron supplements during altitude exposure to ensure they maintain a healthy iron balance. By analysing ten years of haematological data collected from welltrained athletes who undertook two-to-four weeks of LHTL at simulated (3,000 m) or natural (1,350-2,700 m) altitudes, Study Three established how oral iron supplement dose moderates the Hbmass, serum ferritin and transferrin saturation response to prolonged moderate altitude exposure. In general, athletes supplemented with 105 mg.d- 1 or 210 mg.d-1 of oral iron supplement increased their Hbmass from pre-altitude levels by 3.3% and 4.0% respectively. Serum ferritin levels decreased by 33.2% in non-iron supplemented athletes and by 13.8% in athletes supplemented with 105 mg.d-1 of oral iron, however, those athletes who ingested 210 mg.d-1 markedly increased their iron storage compartment by 36.8% after moderate altitude exposure. Thus, daily oral iron supplementation at altitude assists athletes to maintain a healthy iron balance, providing them with sufficient iron to sustain accelerated erythropoiesis. In conclusion, this thesis suggests exercise in acute hypoxia does not seem to alter the post-exercise hepcidin response relative to exercise in normoxia, but prolonged altitude exposure suppresses resting hepcidin levels and may attenuate the magnitude of postexercise hepcidin response after 14 days of LHTL. Finally, daily oral iron supplementation may support iron balance and Hbmass production in athletes undertaking prolonged moderate altitude exposure.
APA, Harvard, Vancouver, ISO, and other styles
6

Englehardt, Kimberly G. "The Effects of a Vegetarian Diet on Iron Status in Female Students." DigitalCommons@CalPoly, 2008. https://digitalcommons.calpoly.edu/theses/65.

Full text
Abstract:
Iron deficiency anemia is the most common nutritional deficiency disease worldwide (Mahan & Escott-Stump, 2004). Iron deficiency anemia is of major concern especially in women of child bearing age and those who follow a vegetarian diet. The objective of this study was to compare the nutrient and hematological values related to iron status in female university students following a vegetarian versus following a nonvegetarian diet. This study took a cross sectional analysis of 39 female students at California Polytechnic State University (Cal Poly State University) in San Luis Obispo, CA. Of the participants 19 were following a vegetarian diet and 20 were following a nonvegetarian diet. To participate, individuals had to be female, current Cal Poly students, and between the ages of 18 and 22 years old. Those taking vitamin or mineral supplements, medications (including oral contraceptives), smokers, and pregnant women were excluded. Characteristic, demographic, and anthropometric data were collected through interview, nutrient intake was accessed by averaging three day food records, and hematological parameters were measured. Statistical analysis used nonparametric techniques including the Mann-Whitney Wilcoxon statistical test for demographics and baseline characteristics, the Spearman Rank Correlation analysis and Fisher’s Exact statistical test for associations between vegetarians and nonvegetarians. Results found no significant difference in iron intake between vegetarians and nonvegetarians, however nonvegetarians had higher mean intakes of iron at 16.82 (SD 6.36) mg/day compared to vegetarians at 14.84 (SD 7.10) mg/day (p=0.482). A similar percentage of vegetarians at 66.7% (n=8) compared to nonvegetarians at 65% (n=13) were under the Recommended Daily Allowance (18 mg per day for females 19 to 30 years of age) for mean iron consumption. There were slightly more nonvegetarians at 10% (n=2) compared to vegetarians at 8.3% (n=1) under the Estimated Average Requirement (8.1 mg/day for females 19 to 30 years of age) for mean iron intake. No significant difference was found for serum iron, serum ferritin, transferrin saturation, and total iron binding capacity between vegetarians and nonvegetarians. Finding revealed serum ferritin, the most common iron status indicator, was lower for vegetarians at 23.16 (SD15.54) ng/mL compared to nonvegetarians at 27.75 (SD 18.01) ng/mL (p=0.47). When looking at the stages of iron balance, there was greater percentage of vegetarians with hematological results (serum iron <40 µg/dL, total iron binging capacity of >410 µg/dL, transferrin saturation <15%, and serum ferritin <10 ng/mL) indicating iron deficiency anemia or stage IV negative iron balance compared to nonvegetarians. There was no significant correlation between iron intake and serum ferritin, however results showed a positive association (r=0.28, p=0.09). In conclusion, vegetarian participants are believed to be at higher risk of developing negative iron balance compared to nonvegetarians due to lower iron consumption and lower serum ferritin concentrations. Female university students following a vegetarian diet should be educated on iron deficiency anemia and prevention of iron depletion.
APA, Harvard, Vancouver, ISO, and other styles
7

Jühlen, Ramona, Jan Idkowiak, Angela E. Taylor, Barbara Kind, Wiebke Arlt, Angela Huebner, and Katrin Koehler. "Role of ALADIN in Human Adrenocortical Cells for Oxidative Stress Response and Steroidogenesis." Saechsische Landesbibliothek- Staats- und Universitaetsbibliothek Dresden, 2015. http://nbn-resolving.de/urn:nbn:de:bsz:14-qucosa-173705.

Full text
Abstract:
Triple A syndrome is caused by mutations in AAAS encoding the protein ALADIN. We investigated the role of ALADIN in the human adrenocortical cell line NCI-H295R1 by either over-expression or down-regulation of ALADIN. Our findings indicate that AAAS knock-down induces a down-regulation of genes coding for type II microsomal cytochrome P450 hydroxylases CYP17A1 and CYP21A2 and their electron donor enzyme cytochrome P450 oxidoreductase, thereby decreasing biosynthesis of precursor metabolites required for glucocorticoid and androgen production. Furthermore we demonstrate that ALADIN deficiency leads to increased susceptibility to oxidative stress and alteration in redox homeostasis after paraquat treatment. Finally, we show significantly impaired nuclear import of DNA ligase 1, aprataxin and ferritin heavy chain 1 in ALADIN knock-down cells. We conclude that down-regulating ALADIN results in decreased oxidative stress response leading to alteration in steroidogenesis, highlighting our knock-down cell model as an important in-vitro tool for studying the adrenal phenotype in triple A syndrome.
APA, Harvard, Vancouver, ISO, and other styles
8

Andersson, Ola. "Effects of Delayed versus Early Cord Clamping on Healthy Term Infants." Doctoral thesis, Uppsala universitet, Pediatrik, 2013. http://urn.kb.se/resolve?urn=urn:nbn:se:uu:diva-198167.

Full text
Abstract:
The aim of this thesis was to study maternal and infant effects of delayed cord clamping (≥180 seconds, DCC) compared to early (≤10 seconds, ECC) in a randomised controlled trial. Practice and guidelines regarding when to clamp the cord vary globally, and different meta-analyses have shown contradictory conclusions on benefits and disadvantages of DCC and ECC. The study population consisted of 382 term infants born after normal pregnancies and randomised to DCC or ECC after birth. The primary objective was iron stores and iron deficiency at 4 months of age, but the thesis was designed to investigate a wide range of suggested effects associated with cord clamping. Paper I showed that DCC was associated with improved iron stores at 4 months (45% higher ferritin) and that the incidence of iron deficiency was reduced from 5.7% to 0.6%. Neonatal anaemia at 2-3 days was less frequent in the DCC group, 1.2% vs. 6.3%. There were no differences between the groups in respiratory symptoms, polycythaemia, or hyperbilirubinaemia. In paper II we demonstrated that DCC versus ECC was not associated with higher risk for maternal post partum haemorrhage and rendered a comparable ratio of valid umbilical artery blood gas samples. In paper III, the Ages and Stages Questionnaire was used to assess neurodevelopment at 4 months. The total scores did not differ, but the DCC group had a higher score in the problem-solving domain and a lower score in the personal-social domain. Immunoglobulin G level was 0.7 g/L higher in the DCC group at 2–3 days, but did not differ at 4 months. Symptoms of infection up to 4 months were comparable between groups. Finally, in paper IV, iron stores and neurodevelopment were similar between groups at 12 months. Gender specific outcome on neurodevelopment at 12 months was discovered, implying positive effects from DCC on boys and negative on girls. We conclude that delaying umbilical cord clamping for 180 seconds is safe and associated with a significantly reduced risk for iron deficiency at 4 months, which may have neurodevelopmental effects at a later age.
APA, Harvard, Vancouver, ISO, and other styles
9

Berglund, Staffan. "Effects of iron supplementation on iron status, health and neurological development in marginally low birth weight infants." Doctoral thesis, Umeå universitet, Pediatrik, 2012. http://urn.kb.se/resolve?urn=urn:nbn:se:umu:diva-52079.

Full text
Abstract:
Background Due to small iron stores and rapid growth during the first months of life, infants with low birth weight (LBW) are at risk of iron deficiency (ID). ID in infancy is associated with irreversible impaired neurodevelopment. Preventive iron supplementation may reduce the risk of ID and benefit neurodevelopment, but there is also a possible risk of adverse effects. More than 50% of all LBW infants are born with marginally LBW (MLBW, 2000-2500g), and it is not known if they benefit from iron supplementation. Methods We randomized 285 healthy, Swedish, MLBW infants to receive 3 different doses of oral iron supplements; 0 (Placebo), 1, and 2 mg/kg/day from six weeks to six months of age. Iron status, during and after the intervention was assessed and so was the prevalence of ID and ID anemia (IDA), growth, morbidity and the interplay with iron and the erythropoetic hormones hepcidin and erythropoietin (EPO). As a proxy for conduction speed in the developing brain, auditory brainstem response (ABR) was analyzed at six months. In a follow up at 3.5 years of age, the children were assessed with a cognitive test (WPPSI-III) and a validated parental checklist of behavioral problems (CBCL), and compared to a matched reference group of 95 children born with normal birth weight. Results At six months of age, the prevalence of ID and IDA was significantly higher in the placebo group compared to the iron supplemented infants. 36% had ID in the placebo group, compared to 8% and 4 % in the 1 and 2mg/kg/day-groups, respectively. The prevalence of IDA was 10%, 3% and 0%, respectively. ABR-latencies did not correlate with the iron intake and was not increased in infants with ID or IDA. ABR wave V latencies were similar in all three groups. Hepcidin correlated to ferritin and increased in supplemented infants while EPO, which was negatively correlated to iron status indicators, decreased. At follow up there were no differences in cognitive scores between the groups but the prevalence of behavioral problems was significantly higher in the placebo group compared to those supplemented and to controls. The relative risk increase of CBCL-scores above a validated cutoff was 4.5 (1.4 – 14.2) in the placebo-group compared to supplemented children. There was no detected difference in growth or morbidity at any age. Conclusion MLBW infants are at risk of ID in infancy and behavioral problems at 3 years of age. Iron supplementation at a dose of 1-2 mg/kg/day from six weeks to six months of age reduces the risks with no adverse effects, suggesting both short and long term benefit. MLBW infants should be included in general iron supplementation programs during their first six months of life.
APA, Harvard, Vancouver, ISO, and other styles
10

Augustine, Robin. "Electromagnetic modelling of human tissues and its application on the interaction between antenna and human body in the BAN context." Phd thesis, Université Paris-Est, 2009. http://tel.archives-ouvertes.fr/tel-00499255.

Full text
Abstract:
Les réseaux BAN (Body Area Network) révolutionnent le concept de la surveillance et de la prise en charge à distance de la santé du patient. Le BAN fournit des informations sur l'état de santé du patient en temps réel quelque soit l'endroit où il se trouve. Dans le « télé monitoring », des capteurs de mouvement, de respiration ou du rythme cardiaque placés à l'intérieur ou sur le corps humain transmettent des données via le réseau sans fil constituant le BAN, une antenne étant associée à chaque nœud du réseau. La communication peut être in/on, on/on ou on/off selon que les antennes sont placées à l'intérieur, sur ou à l'extérieur du corps. Le développement des BAN nécessite la réalisation de modèles (ou fantômes) simulant au mieux les propriétés électromagnétiques du corps humain. Des antennes portables, miniaturisées doivent être réalisées avec des contraintes d'intégration d'une part (aux vêtements, à des objets type montre ou badge), des contraintes de résistance ou de prise en compte de l'influence du corps d'autre part. La réduction de l'impact des antennes sur les tissus en terme de SAR (Specific Absorption Rate) doit également être considérée. Dans ce travail, l'objectif est de développer des fantômes valables pour les communications dans et sur le corps. Les matériaux de base sélectionnés sont d'origine biologique (biocéramiques et biopolymères) avec des propriétés proches de celles des tissus humains. Ces fantômes étant biocompatibles, ils sont essentiellement non toxiques alors que les fantômes usuels le sont en général. D'autre part, différents types d'antennes conformables, fonctionnant dans la bande ISM 2.4 GHz ont été développées et étudiées dans la perspective du BAN. Les antennes voient leur adaptation et leur efficacité chuter au contact ou à proximité du corps, ce qui constitue un écueil majeur pour établir une bonne communication. Différentes méthodes permettant de réduire l'influence du corps (plan de masse à l'arrière, surface haute impédance, feuille de ferrite polymère) sont testés et leurs avantages et inconvénients développés. Des mesures de SAR permettent aussi de démontrer l'efficacité de ces méthodes sur la réduction de la puissance absorbée par les tissus. Au final, ce travail apporte une contribution à l'étude théorique et expérimentale de l'interaction entre corps humain et antenne dans le cadre des réseaux BAN appliqués à la télésurveillance de la santé
APA, Harvard, Vancouver, ISO, and other styles

Book chapters on the topic "Human ferritin"

1

Levi, Sonia, Alessandra Luzzago, Paolo Santambrogio, Anna Cozzi, Gianni Cesareni, and Paolo Arosio. "Mechanisms of Ferritin Iron Incorporation: A Study with Recombinant and Mutant Human Ferritins." In Iron Biominerals, 339–48. Boston, MA: Springer US, 1991. http://dx.doi.org/10.1007/978-1-4615-3810-3_24.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Arosio, Paolo, Alberto Albertini, Sonia Levi, Paolo Santambrogio, Anna Cozzi, Barbara Corsi, Elena Tamborini, Stefania Spada, and Ermanna Rovida. "Chemico-Physical and Functional Differences Between H and L Chains of Human Ferritin." In Advances in Experimental Medicine and Biology, 13–21. Boston, MA: Springer US, 1994. http://dx.doi.org/10.1007/978-1-4615-2554-7_2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Oshtrakh, M. I., O. B. Milder, V. A. Semionkin, P. G. Prokopenko, and L. I. Malakheeva. "Comparative Study of Human Liver Ferritin and Chicken Liver by Mössbauer Spectroscopy. Preliminary Results." In ICAME 2003, 279–84. Dordrecht: Springer Netherlands, 2004. http://dx.doi.org/10.1007/978-1-4020-2852-6_42.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Konijn, A. M., E. G. Meyron-Holtz, D. Gelvan, and E. Fibach. "Cellular Ferritin Uptake: A Highly Regulated Pathway for Iron Assimilation in Human Erythroid Precursor Cells." In Advances in Experimental Medicine and Biology, 189–97. Boston, MA: Springer US, 1994. http://dx.doi.org/10.1007/978-1-4615-2554-7_21.

Full text
APA, Harvard, Vancouver, ISO, and other styles
5

de Oliveira, Leandro, and Maria de Sousa. "Expression of Ferritin Messenger RNA in an in vitro Model of Human CD3-Mediated T-Lymphocyte Activation." In Cellular Regulation by Protein Phosphorylation, 453–57. Berlin, Heidelberg: Springer Berlin Heidelberg, 1991. http://dx.doi.org/10.1007/978-3-642-75142-4_56.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Oshtrakh, M. I., O. B. Milder, and V. A. Semionkin. "A study of human liver ferritin and chicken liver and spleen using Mössbauer spectroscopy with high velocity resolution." In HFI/NQI 2007, 565–72. Berlin, Heidelberg: Springer Berlin Heidelberg, 2008. http://dx.doi.org/10.1007/978-3-540-85320-6_88.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Theil, Elizabeth C. "Concentrating, Storing, and Detoxifying Iron: The Ferritins and Hemosiderin." In Iron Physiology and Pathophysiology in Humans, 63–78. Totowa, NJ: Humana Press, 2011. http://dx.doi.org/10.1007/978-1-60327-485-2_4.

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

Drysdale, James W. "Human Ferritin Gene Expression." In Progress in Nucleic Acid Research and Molecular Biology, 127–55. Elsevier, 1988. http://dx.doi.org/10.1016/s0079-6603(08)60612-1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Boumaiza, Mohamed, Samia Rourou, Paolo Arosio, and Mohamed Nejib Marzouki. "The Use of Ferritin as a Carrier of Peptides and Its Application for Hepcidin." In BioMechanics and Functional Tissue Engineering [Working Title]. IntechOpen, 2020. http://dx.doi.org/10.5772/intechopen.94408.

Full text
Abstract:
Hepcidin a 25-amino-acid and highly disulfide bonded hormone, is the central regulator of iron homeostasis. In this chapter we propose ferritin as a peptide carrier to promote the association of the hybrid hepcidin/ferritin nanoparticle with a particular cell or tissue for therapeutic or diagnostic use. Indeed, human ferritin H-chain fused directly (on its 5’end) with camel mature hepcidin was cloned into the pASK-43 plus vector and expressed using BL21 (DE3) pLys E. coli strain. The transformed E.coli produced efficiently hepcidin-ferritin construct (hepcH), consisting of 213 amino acids with a molecular weight of 24 KDa. The recovered product is a ferritin exposing hepcidin on outer surface. The hepcH monomer was characterized by immunoblotting using a monoclonal antibody specific for human ferritin and a polyclonal antibody specific for hepcidin-25. The results were also confirmed by MALDI-TOF mass spectrometry. The recombinant native human ferritin and the commercial human hepcidin-25 were used as controls in this experiment. The assembly of hepcH, as an heteropolymer molecule, was performed in presence of denatured human ferritin-H and -L chains. After cysteine oxidation of the recombinant nanoparticles, cellular binding assays were performed on mammalian cells such as mouse monocyte–macrophage cell line J774, HepG2 and COS7.
APA, Harvard, Vancouver, ISO, and other styles
10

Leggett*, Barbara A., and June W. Halliday**. "Iron Balance in Western Societies as Measured by Serum Ferritin." In Iron and Human Disease, 107–30. CRC Press, 2018. http://dx.doi.org/10.1201/9781351073899-4.

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

Conference papers on the topic "Human ferritin"

1

Shpyleva, Svitlana I., Volodymyr P. Tryndyak, Olga Kovalchuk, Athena Starlard-Davenport, Vasyl’ F. Chekhun, Frederick A. Beland, and Igor P. Pogribny. "Abstract 2313: Role of ferritin dysregulation in human breast cancer cells." In Proceedings: AACR 101st Annual Meeting 2010‐‐ Apr 17‐21, 2010; Washington, DC. American Association for Cancer Research, 2010. http://dx.doi.org/10.1158/1538-7445.am10-2313.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Pang, Min, Xiaoli Liu, A. B. Madhankumar, and James Connor. "Abstract 1471: Role of ferritin heavy chain in therapeutic sensitivity of human glioblastoma cells." In Proceedings: AACR 103rd Annual Meeting 2012‐‐ Mar 31‐Apr 4, 2012; Chicago, IL. American Association for Cancer Research, 2012. http://dx.doi.org/10.1158/1538-7445.am2012-1471.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Lihua, Xu, Guo Huatao, and Lv Shujie. "Serum ferritin and folate, but not uric acid, associate with coronary heart disease in a northern Chinese population." In 2011 International Conference on Human Health and Biomedical Engineering (HHBE). IEEE, 2011. http://dx.doi.org/10.1109/hhbe.2011.6027931.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Pang, Min, A. B. Madhankumar, Becky Webb, Xiaoli Liu, and James R. Connor. "Abstract 4436: The down-regulation of ferritin heavy chain as an adjuvant therapy in human glioma." In Proceedings: AACR 104th Annual Meeting 2013; Apr 6-10, 2013; Washington, DC. American Association for Cancer Research, 2013. http://dx.doi.org/10.1158/1538-7445.am2013-4436.

Full text
APA, Harvard, Vancouver, ISO, and other styles
5

Alenkina, Irina V., Michael I. Oshtrakh, Zoltán Klencsár, Ernő Kuzmann, and Vladimir A. Semionkin. "Mössbauer spectroscopy of human liver ferritin and its analogue, Ferrum Lek, in the temperature range of 295-90 K: Comparison within the homogeneous iron core model." In MOSSBAUER SPECTROSCOPY IN MATERIALS SCIENCE - 2014. AIP Publishing LLC, 2014. http://dx.doi.org/10.1063/1.4898623.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Greenwood, Michael, Rawan Eid, Nagla Arab, Chamel Khoury, and Paul Young. "Expression of human H ferritin prompts the identification of a hitherto elusive yeast orthologue and enables parsing of distinct iron-induced cell death pathways in Saccharomyces cerevisiae." In 1st Electronic Conference on Molecular Science. Basel, Switzerland: MDPI, 2015. http://dx.doi.org/10.3390/ecms-1-b002.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Andreas, Ruf, Eberhard Morgenstern, Heinrich Patscheke, Sentot Santoso, Christian Müller-Eckhardt, and Norbert Heimburger. "THE ROLE OF THE GP IIb/IIIa COMPLEX AND vWF IN PLATELET -COLLAGEN INTERACTION." In XIth International Congress on Thrombosis and Haemostasis. Schattauer GmbH, 1987. http://dx.doi.org/10.1055/s-0038-1643518.

Full text
Abstract:
Our recent studies showed that collagen fibrils (CF) are internalized by platelets in citrated plasm. This phenomenon was not observed in EDTA-PRP. In order to investigate whether collagen internalization is mediated by a receptor we studied the effect of monoclonal antibodies against the receptor molecules GP IIb/IIIa (Gi5) and GP Ib (AN51, Da-kopatts GmbH, Hamburg). Washed human platelets from healthy donors were incubated with lOμg of Gi5 or 4μg of AN51 per ml of platelet suspension (2 × 108 platelets per ml). After 10 min. at 37° C we added AOpg collagen (Hormonchemie, München) per ml of platelet suspension. After an additional incubation period of 10 min. the reaction was stopped by glutaraldehyde fixation (40μl of a 3% glutaraldehyde and of a 0.2 % tannin solution in phosphate buffer). To investigate the role of vWF in the internalization process we added 1Oμg/ml ferritin-labeled anti-vWF Fab-fragments to the platelet suspension two min. in advance of collagen stimulation (40μg/ml; 10 min). The values in percent of the inhibition of the internalization phenomena were obtained by statistical evaluation of these phenomena observed on ultrathin serial sections. Washed platelets interacted with CF in the same manner as platelets in citrated plasm, but aggregates were only formed after stirring. Washed platelets which had internalized CF were found to be completely degranulated in contrast to platelets which had only contacted CF without internalization. These platelets were hardly degranulated. The membrane system containing the internalized CF surrounded the contractile sphere of the platelets and always displayed an opening to the platelet surface.We found that Gi5 inhibited collagen internalization up to 95% and anti-vWF to approximately 50%, whereas AN51 showed no effect on this process. Futhermore, Gi5 inhibited the platelet function aggregation and the release reaction which was enhanced by CF internalization. These findings show that the GP Ilb/IIIa complex is strongly involved in the internalization of CF by platelets and suggest an involvement of vWF released from the stimulated platelets in this process by binding to the GP Ilb/IIIa complex.Supported by DFG, Grant MO 124/2-4
APA, Harvard, Vancouver, ISO, and other styles
8

Sugimoto, Seiichi, Kazuo Yagi, Yujiro Harada, and Masataka Tokuda. "Synthesis and Magnetic Properties of New Multi-components Spinel Ferrite Nanoparticles." In 2007 International Symposium on Micro-NanoMechatronics and Human Science. IEEE, 2007. http://dx.doi.org/10.1109/mhs.2007.4420915.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Maruo, S., T. Saeki, Y. Kanazawa, and Y. Ichiyanagi. "Magnetically driven micropump produced by microstereolithography with ferrite nanoparticle composite photopolymer." In 2008 International Symposium on Micro-NanoMechatronics and Human Science. IEEE, 2008. http://dx.doi.org/10.1109/mhs.2008.4752465.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Ozdemir, Mehmet Semih, Cemil Ocak, and Adem Dalcali. "Permanent Magnet Wind Generators: Neodymium vs. Ferrite Magnets." In 2021 3rd International Congress on Human-Computer Interaction, Optimization and Robotic Applications (HORA). IEEE, 2021. http://dx.doi.org/10.1109/hora52670.2021.9461291.

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

Reports on the topic "Human ferritin"

1

Spiegel, Yitzhak, Michael McClure, Itzhak Kahane, and B. M. Zuckerman. Characterization of the Phytophagous Nematode Surface Coat to Provide New Strategies for Biocontrol. United States Department of Agriculture, November 1995. http://dx.doi.org/10.32747/1995.7613015.bard.

Full text
Abstract:
Chemical composition and biological role of the surface coat (SC) of the root-knot nematodes, Meloidogyne spp. are described. SC proteins of M. incognita race 3 infective juveniles (J2) were characterized by electrophoresis and western blotting of extracts from radioiodine and biotin-labelled nematodes. J2 labelled with radioiodine and biotin released 125I and biotin-labelled molecules into water after 20 hours incubation, indicating that SC proteins may be loosely attached to the nematode. Antiserum to the principal protein reacted with the surface of live J2 and with surface proteins previously separated by electrophoresis. Human red blood cells (HRBC) adhered to J2 of several tylenchid nematodes over the entire nematode body. HRBC adhered also to nylon fibers coated with SC extracted from M. javanica J2; binding was Ca++/Mg++ dependent, and decreased when the nylon fibers were coated with bovine serum albumin, or pre-incubated with fucose and mannose. These experiments support a working hypothesis that RBC adhesion involves carbohydrate moieties of HRBC and carbohydrate-recognition domain(s) (CRD) distributed on the nematode surface. To our knowledge, this is the first report of a surface CRD i the phylum Nematoda. Gold-conjugated lectins and neoglycoproteins combined with silver enhancement have been used for the detection of carbohydrates and CRD, respectively, on the SC of M. javanica J2. Biotin reagents were used to trace surface proteins, specifically, on live J2. The labile and transitory nature of the SC was demonstrated by the dynamics of HRBC adherence to detergent-treated J2, J2 at different ages or fresh-hatched J2 held at various temperatures. SC recovery was demonstrated also by a SDS-PAGE profile. Monoclonal antibodies developed to a cuticular protein of M. incognita J2 gave a slight, but significant reduction in attachment of Pasteuria penetrans spores. Spore attachment as affected by several enzymes was inconsistent: alcian blue, which specifically blocks sulfyl groups, had no afffect on spore attachment. Treatment with cationized ferritin alone or catonized ferritin following monoclonal antibody caused significant decreases in spore attachment. Those results suggest a role in attachment by negatively charged groups.
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