首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 265 毫秒
1.
The colloidal properties of transferrin receptor, isolated from human placenta, in detergent free solution has been investigated by light scattering techniques and analytical ultracentrifugation. In detergent free solution at 293.2 K, hTfR forms stable aggregates with an apparent hydrodynamic radius of 17 nm. The molecular mass was determined by ultracentrifugation to lie between (1722+/-87) kDa (sedimentation equilibrium) and (1675+/-46) kDa (sedimentation velocity). This implies that the aggregates are build up from nine hTfR dimers. Based on model calculations, which are in good agreement with the experimental data, we propose a torus-like structure for the aggregates. Upon pH shift from pH 7.5 to 5.0 or removal of the N-linked carbohydrate chains, formation of larger aggregates is induced. These aggregates can be described in terms of porous fractal structures. We propose a simple model, which accounts for that behaviour assuming that the aggregation is mainly due to the reduction of negative surface charge.  相似文献   

2.
The kinetics and thermodynamics of Bi(III) exchange between bismuth mononitrilotriacetate (BiL) and human serum transferrin as well as those of the interaction between bismuth-loaded transferrin and transferrin receptor 1 (TFR) were investigated at pH 7.4-8.9. Bismuth is rapidly exchanged between BiL and the C-site of human serum apotransferrin in interaction with bicarbonate to yield an intermediate complex with an effective equilibrium constant K(1) of 6 +/- 4, a direct second-order rate constant k(1) of (2.45 +/- 0.20) x 10(5) M(-1) s(-1), and a reverse second-order rate constant k(-1) of (1.5 +/- 0.5) x 10(6) M(-1) s(-1). The intermediate complex loses a single proton with a proton dissociation constant K(1a) of 2.4 +/- 1 nM to yield a first kinetic product. This product then undergoes a modification in its conformation followed by two proton losses with a first-order rate constant k(2) = 25 +/- 1.5 s(-1) to produce a second kinetic intermediate, which in turn undergoes a last modification in the conformation to yield the bismuth-saturated transferrin in its final state. This last process rate-controls Bi(III) uptake by the N-site of the protein and is independent of the experimental parameters with a constant reciprocal relaxation time tau(3)(-1) of (3 +/- 1) x 10(-2) s(-1). The mechanism of bismuth uptake differs from that of iron and probably does not involve the same transition in conformation from open to closed upon iron uptake. The interaction of bismuth-loaded transferrin with TFR occurs in a single very fast kinetic step with a dissociation constant K(d) of 4 +/- 0.4 microM, a second-order rate constant k(d) of (2.2 +/- 1.5) x 10(8) M(-1) s(-1), and a first-order rate constant k(-d) of 900 +/- 400 s(-1). This mechanism is different from that observed with the ferric holotransferrin and implies that the interaction between TFR and bismuth-loaded transferrin probably takes place on the helical domain of the receptor which is specific for the C-site of transferrin and HFE. The relevance of bismuth incorporation by the transferrin receptor-mediated iron acquisition pathway is discussed.  相似文献   

3.
The kinetics of cycling of the transferrin receptor in A431 human epidermoid-carcinoma cells was examined in the presence or absence of bound diferric transferrin. In order to investigate the properties of the receptor in the absence of transferrin, the cells were maintained in defined medium without transferrin. It was demonstrated that Fab fragments of a monoclonal anti-(transferrin receptor) antibody (OKT9) did not alter the binding of diferric 125I-transferrin to the receptor or change the accumulation of [59Fe]diferric transferrin by cells. OKT9 125I-Fab fragments were prepared and used as a probe for the function of the receptor. The first-order rate constants for endocytosis (0.16 +/- 0.02 min-1) and exocytosis (0.056 +/- 0.003 min-1) were found to be significantly lower for control cells than the corresponding rate constants for endocytosis (0.22 +/- 0.02 min-1) and exocytosis (0.065 +/- 0.004 min-1) measured for cells incubated with 1 microM-diferric transferrin (mean +/- S.D., n = 3). The cycling of the transferrin receptor is therefore regulated by diferric transferrin via an increase in both the rate of endocytosis and exocytosis. Examination of the accumulation of OKT9 125I-Fab fragments indicated that diferric transferrin caused a marked decrease in the amount of internalized 125I-Fab fragments associated with the cells after 60 min of incubation at 37 degrees C. Diferric transferrin therefore increases the efficiency of the release of internalized 125I-Fab fragments compared with cells incubated without diferric transferrin. These data indicate that transferrin regulates the sorting of the transferrin receptor at the cell surface and within endosomal membrane compartments.  相似文献   

4.
The kinetics and thermodynamics of Al(III) exchange between aluminum citrate (AlL) and human serum transferrin were investigated in the 7.2-8.9 pH range. The C-site of human serum apotransferrin in interaction with bicarbonate removes Al(III) from Al citrate with an exchange equilibrium constant K1 = (2.0 +/- 0.6) x 10(-2); a direct second-order rate constant k1 = 45 +/- 3 M(-1) x s(-1); and a reverse second-order rate constant k(-1) = (2.3 +/- 0.5) x 10(3) M(-1) x s(-1). The newly formed aluminum-protein complex loses a single proton with proton dissociation constant K1a = (15 +/- 3) nM to yield a first kinetic intermediate. This intermediate then undergoes a modification in its conformation followed by two proton losses; first-order rate constant k2 = (4.20 +/- 0.02) x 10(-2) s(-1) to produce a second kinetic intermediate, which in turn undergoes a last slow modification in the conformation to yield the aluminum-loaded transferrin in its final state. This last process rate-controls Al(III) uptake by the N-site of the protein and is independent of the experimental parameters with a constant reciprocal relaxation time tau3(-1) = (6 +/- 1) x 10(-5) x s(-1). The affinities involved in aluminum uptake by serum transferrins are about 10 orders of magnitude lower than those involved in the uptake of iron. The interactions of iron-loaded transferrins with transferrin receptor 1 occur with average dissociation constants of 3 +/- 1 and 5 +/- 1 nM for the only C-site iron-loaded and of 6.0 +/- 0.6 and 7 +/- 0.5 nM for the iron-saturated ST in the absence or presence of CHAPS, respectively. No interaction is detected between receptor 1 and aluminum-saturated or mixed C-site iron-loaded/N-site aluminum-loaded transferrin under the same conditions. The fact that aluminum can be solubilized by serum transferrin in biological fluids does not necessarily imply that its transfer from the blood stream to cytoplasm follows the receptor-mediated pathway of iron transport by transferrins.  相似文献   

5.
The endocytosis and recycling of the human transferrin receptor were evaluated by several experimental modalities in K562 cells perturbed with 10(-5) M monensin. The work presented is an extension of a previous study demonstrating both complete inhibition of release of internalized human transferrin and a 50% reduction in the number of cell surface transferrin binding sites in K562 cells treated with monensin (Stein, B. S., Bensch, K. G., and Sussman, H. H. (1984) J. Biol. Chem. 259, 14762-14772). The data directly reveal the existence of two distinct transferrin receptor recycling pathways. One pathway is monensin-sensitive and is felt to represent recycling of transferrin receptors through the Golgi apparatus, and the other pathway is monensin-resistant and most likely represents non-Golgi-mediated transferrin receptor recycling. A transferrin-free K562 cell culture system was developed and used to demonstrate that cell surface transferrin receptors can be endocytosed without antecedent ligand binding, indicating that there are factors other than transferrin binding which regulate receptor internalization. Evidence is presented suggesting that two transferrin receptor recycling pathways are also operant in K562 cells under ligand-free conditions, signifying that trafficking of receptor into either recycling pathway is not highly ligand-dependent.  相似文献   

6.
Recombinant human transferrin receptor has been produced in a baculovirus expression system. Magnetic particles coated with an anti-transferrin receptor monoclonal antibody were used to immunoselect virus-infected Sf9 insect cells expressing the human transferrin receptor on their cell surface. Recombinant virus containing the human transferrin receptor cDNA was then plaque-purified from these cells. Biosynthetic labeling studies of infected cells showed that the human transferrin receptor is one of the major proteins made 2-3 days postinfection. The recombinant receptor made in insect cells is glycosylated and is also posttranslationally modified by the addition of a fatty acid moiety. However, studies with tunicamycin and endoglycosidases H and F showed that the oligosaccharides displayed on the recombinant receptor differ from those found on the naturally occurring receptor in human cells. As a consequence, the human receptor produced in the baculovirus system has an Mr of 82,000 and is smaller in size than the authentic receptor. About 30% of human transferrin receptors made in insect cells do not form intermolecular disulfide bonds, but are recognized by the anti-transferrin receptor antibody, B3/25, and bind specifically to a human transferrin-Sepharose column. Binding studies using 125I-labeled human transferrin showed that insect cells infected with the recombinant virus expressed an average of 5.8 +/- 0.9 X 10(5) transferrin receptors (Kd = 63 +/- 9 nM) on their cell surface. Thus, the human transferrin receptor produced in insect cells is biologically active and appears suitable for structural and functional studies.  相似文献   

7.
Specific binding of ferric bovine transferrin to the human transferrin receptor was investigated using K562 cells propagated in serum-free medium without transferrin supplemented with 10(-5) elemental iron. Affinity chromatography of solubilized extracts of K562 cells surface-labeled with 125I was performed using bovine transferrin- and human transferrin-Sepharose 4B resins. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis of resin eluates reveal that bovine transferrin specifically binds a Mr = 188,000 protein which dissociates into a Mr = 94,000 protein under reducing conditions, a finding identical to what is seen with human transferrin. The Mr = 94,000 reduced protein isolated by bovine transferrin resin shows an identical one-dimensional partial proteolytic digestion map with that of the human transferrin receptor. Unlabeled bovine transferrin was shown to specifically compete 125I-labeled human transferrin from the human transferrin receptor on the surface of K562 cells at 4 degrees C in a similar manner as unlabeled human transferrin; however, approximately a 2,000-fold higher concentration of bovine ligand was required to achieve comparable competition (50% inhibition of binding). Indirect immunofluorescence cytolocalization of bovine transferrin in K562 cells grown in serum-free medium supplemented with ferric bovine transferrin reveal patterns similar to those seen for human transferrin (both focal perinuclear and diffuse cytoplasmic fluorescence). Monensin treatment results in a dramatic accumulation of bovine ligand in perinuclear aggregates, suggesting that it is recycled through the Golgi, as is human transferrin. K562 cells grown in serum-free medium supplemented with either 300 micrograms/ml of ferric human or ferric bovine transferrin were found to demonstrate superimposable growth curves.  相似文献   

8.
The expression of transferrin receptors by blood monocytes, human alveolar macrophages, and in vitro matured macrophages was evaluated by immunofluorescence, radioligand binding, and Northern analysis, using the monoclonal anti-human transferrin receptor antibody OKT9, [125I]-labeled human transferrin and a [32P]-labeled human transferrin receptor cDNA probe, respectively. By immunofluorescence, the majority of alveolar macrophages expressed transferrin receptors (86 +/- 3%). The radioligand binding assay demonstrated the affinity constant (Ka) of the alveolar macrophage transferrin receptor was 4.4 +/- 0.7 X 10(8) M-1, and the number of receptors per cell was 4.4 +/- 1.2 X 10(4). In marked contrast, transferrin receptors were not present on the surface or in the cytoplasm of blood monocytes, the precursors of the alveolar macrophages. However, when monocytes were cultured in vitro and allowed to mature, greater than 80% expressed transferrin receptors by day 6, and the receptors could be detected by day 3. Consistent with these observations, a transferrin receptor mRNA with a molecular size of 4.9 kb was demonstrated in alveolar macrophages and in vitro matured macrophages but not in blood monocytes. Thus, although blood monocytes do not express the transferrin receptor gene, it is expressed by mature macrophages, an event that probably occurs relatively early in the process of monocyte differentiation to macrophages.  相似文献   

9.
The incorporation of iron into human cells involves the binding of diferric transferrin to a specific cell surface receptor. We studied the process of endocytosis in K562, a human erythroid cell line, by using tetramethylrhodamine isothiocyanate-labeled transferrin (TRITC- transferrin) and fluorescein isothiocyanate-labeled Fab fragments of goat antireceptor IgG preparation (FITC-Fab-antitransferrin receptor antibody). Because the antireceptor antibody and transferrin bind to different sites on the transferrin receptor molecule it was possible to simultaneously and independently follow ligand and receptor. At 4 degrees C, the binding of TRITC-transferrin or FITC-Fab antitransferrin receptor antibody exhibited diffuse membrane fluorescence. At 20 degrees C, the binding of TRITC-transferrin was followed by the rapid formation of aggregates. However, the FITC-Fab antitransferrin receptor did not show similar aggregation at 20 degrees C unless transferrin was present. In the presence of transferrin, the FITC-Fab antitransferrin receptor antibody formed aggregates at the same sites and within the same time period as TRITC transferrin, indicating co-migration. Although the diffuse surface staining of either label was removed by proteolysis, the larger aggregates were not susceptible to enzyme degradation, indicating that they were intracellular. The internal location of the aggregates was also demonstrated using permeabilized cells that had been preincubated with transferrin and fixed with 4% paraformaldehyde. These cells showed aggregated receptor in the interior of the cell when reacted with fluorescein-labeled antibody to the receptor. This indicated that the transferrin and the transferrin receptor co-internalize and migrate to the same structures within the cell.  相似文献   

10.
A soluble form of transferrin receptor has been detected in human serum and has been shown recently to be a truncated form of the intact membrane bound receptor. Mechanisms governing the release of transferrin receptor by cells are poorly understood and could be better defined by tissue culture. The present investigation was undertaken to characterize the transferrin receptor released by K562 erythroleukemic cells. In contrast with maturing sheep reticulocytes, which have been shown to release transferrin receptor in small vesicles termed exosomes, we demonstrated, with a monoclonal enzyme-linked immunoassay, that less than 30% of the transferrin receptor released by K562 cells in log phase growth was in a particulate form. The relative amounts of soluble and particulate receptor released to the supernatant did not change significantly during 48 hr of incubation. Soluble receptor was purified by immunoaffinity chromatography. On polyacrylamide gel electrophoresis, its mobility was the same (85 kDa) as that of the truncated monomeric form recently identified in human serum. Further evidence that serum and soluble receptors released by K562 cells are identical was provided by amino acid sequence analysis, which demonstrated that 16 of the first 19 residues of the N-terminal sequence of soluble K562 receptor are homologous with the serum receptor. The remaining three were not identifiable. K562 cells provide a useful in vitro model for studying the production of membrane-bound and soluble forms of released transferrin receptor.  相似文献   

11.
Sensitive Raman difference spectroscopy was used to monitor the protonation and deprotonation of histidine residues in apo-transferrin. We have shown previously that the behavior of small molecules and/or small molecular groups bound to proteins or other large macromolecules can be studied by Raman difference spectroscopy (Yue, K.T. et al. (1989) J. Raman Spectrosc. 20, 541-545). Using this method, we have measured the Raman difference spectra of human transferrin at different pH values with respect to pH 8.9, titrating its various histidine residues. About 12 +/- 2 of the 19 residues were titrated. The pH difference spectrum of transferrin obtained is very similar to that of histidine in solution, but with clear differences in the 1200-1400 cm-1 region. A titration curve with pKa of 6.08 +/- 0.01 fit the data of histidine in solution and a value of 6.56 +/- 0.02 was found for the average value of the 12 histidine residues inside transferrin. The technique has enough sensitivity at present to monitor a single histidine residue in a 130 kDa molecule and to determine the titration curve of one residue in a 40 kDa protein.  相似文献   

12.
Complete inhibition of transferrin recycling by monensin in K562 cells   总被引:17,自引:0,他引:17  
Monensin blocks human transferrin recycling in a dose-dependent and reversible manner in K562 cells, reaching 100% inhibition at a noncytocidal dose of 10(-5) M, whereas transferrin recycling is virtually unaffected by noncytocidal doses of chloroquine. The intracellular pathway of human transferrin in K562 cells, both in the presence and absence of 10(-5) M monensin, was localized by indirect immunofluorescence. Monensin blocks transferrin recycling by causing internalized ligand to accumulate in the perinuclear region of the cell. The effect of 10(-5) M monensin on human transferrin kinetics was quantitatively measured by radioimmunoassay and showed a positive correlation with immunofluorescent studies. Immunoelectron microscopic localization of human transferrin as it cycles through K562 cells reveals the appearance of perinuclear transferrin-positive multivesicular bodies within 3 min of internalization, with subsequent exocytic delivery of the ligand to the cell surface via transferrin-staining vesicles arising from these perinuclear structures within 5 min of internalization. Inhibition of ligand recycling with 10(-5) M monensin causes dilated transferrin-positive multivesicular bodies to accumulate within the cell with no evidence of recycling vesicles. A coordinated interaction between multivesicular bodies and the Golgi apparatus appears to be involved in the recycling of transferrin in K562 cells. Cell-surface-binding sites for transferrin were reduced by 50% with 10(-5) M monensin treatment; however, this effect was not attenuated by 80% protein synthesis inhibition with cycloheximide, supporting the idea that the transferrin receptor is also recycled through the Golgi.  相似文献   

13.
We applied fluorescence correlation spectroscopy (FCS) to characterize the interaction dynamics of fluorescence-labeled transferrin with transferrin receptor (hTfR) associates isolated from human placenta. The dissociation constant for the equilibrium binding of TMR-labeled ferri-transferrin to hTfR in detergent free solution was determined to be 7 +/- 3 nM. Binding curves were compatible with equal and independent binding sites present on the hTfR associates. Under pseudo-first-order conditions, with respect to transferrin, complex formation is monophasic. From these curves, association and dissociation rate constants for a reversible bimolecular binding reaction were determined, with (1.1 +/- 0.1) x 10(4) M-1 s-1 for the former and (6 +/- 4) x 10(-)4 s-1 for the latter. In dissociation exchange experiments, biphasic curves and concentration-independent reciprocal relaxation times were determined. From isothermal titration calorimetry experiments, we obtained an enthalpy change of -44.4 kJ/mol associated with the reaction. We thus conclude that the reaction is mainly enthalpy driven.  相似文献   

14.
The binding of heparin to human antithrombin III (ATIII) was investigated by titration calorimetry (TC) and differential scanning calorimetry (DSC). TC measurements of homogeneous high-affinity pentasaccharide and octasaccharide fragments of heparin in 0.02 M phosphate buffer and 0.15 M sodium chloride (pH 7.3) yielded binding constants of (7.1 +/- 1.3) x 10(5) M-1 and (6.7 +/- 1.2) x 10(6) M-1, respectively, and corresponding binding enthalpies of -48.3 +/- 0.7 and -54.4 +/- 5.4 kJ mol-1. The binding enthalpy of heparin in phosphate buffer (0.02 M, 0.15 M NaCl, pH 7.3) was estimated from TC measurements to be -55 +/- 10 kJ mol-1, while the enthalpy in Tris buffer (0.02 M, 0.15 M NaCl, pH 7.3) was -18 +/- 2 kJ mol-1. The heparin-binding affinity was shown by fluorescence measurements not to change under these conditions. The 3-fold lower binding enthalpy in Tris can be attributed to the transfer of a proton from the buffer to the heparin-ATIII complex. DSC measurements of the ATIII unfolding transition exhibited a sharp denaturation peak at 329 +/- 1 K with a van 't Hoff enthalpy of 951 +/- 89 kJ mol-1, based on a two-state transition model and a much broader transition from 333 to 366 K. The transition peak at 329 K accounted for 9-18% of the total ATIII. At sub-saturate heparin concentrations, the lower temperature peak became bimodal with the appearance of a second transition peak at 336 K. At saturate heparin concentration only the 336 K peak was observed. This supports a two domain model of ATIII folding in which the lower stability domain (329 K) binds and is stabilized by heparin.  相似文献   

15.
The thermodynamics of the binding of D-galactopyranoside (Gal), 2-acetamido-2-deoxygalactopyranoside (GalNAc), methyl-alpha-D-galactopyranoside, and methyl-beta-D-galactopyranoside to the basic agglutinin from winged bean (WBAI) in 0.02 M sodium phosphate and 0.15 M sodium chloride buffer have been investigated from 298.15 to 333.15 K by titration calorimetry and at the denaturation temperature by differential scanning calorimetry (DSC). WBAI is a dimer with two binding sites. The titration calorimetry yielded single-site binding constants ranging from 0.56 +/- 0.14 x 10(3) M-1 for Gal at 323.15 K to 7.2 +/- 0.5 x 10(3) M-1 for GalNAc at 298.15 K and binding enthalpies ranging from -28.0 +/- 2.0 kJ mol-1 for GalNAc at 298.15 K to -14.3 +/- 0.1 kJ mol-1 for methyl-beta-D-galactopyranoside at 322.65 K. The denaturation transition consisted of two overlapping peaks over the pH range 5.6-7.4. Fits of the differential scanning calorimetry data to a two-state transition model showed that the low temperature transition (341.6 +/- 0.4 K at pH 7.4) consisted of two domains unfolding as a single entity while the higher temperature transition (347.8 +/- 0.6 K at pH 7.4) is of the remaining WBAI dimer unfolding into two monomers. Both transitions shift to higher temperatures and higher calorimetric enthalpies with increase in added ligand concentration at pH 7.4. Analysis of the temperature increase as a function of added ligand concentration suggests that one ligand binds to the two domains unfolding at 341.6 +/- 0.6 K and one ligand binds to the domain unfolding at 347.8 +/- 0.6 K.  相似文献   

16.
During iron acquisition by the cell, complete homodimeric transferrin receptor 1 in an unknown state (R1) binds iron-loaded human serum apotransferrin in an unknown state (T) and allows its internalization in the cytoplasm. T also forms complexes with metals other than iron. Are these metals incorporated by the iron acquisition pathway and how can other proteins interact with R1? We report here a four-step mechanism for cobalt(III) transfer from CoNtaCO(3)(2-) to T and analyze the interaction of cobalt-loaded transferrin with R1. The first step in cobalt uptake by T is a fast transfer of Co(3+) and CO(3)(2-) from CoNtaCO(3)(2-) to the metal-binding site in the C-lobe of T: direct rate constant, k(1)=(1.1+/-0.1) x 10(6) M(-1) s(-1); reverse rate constant, k(-1)=(1.9+/-0.6) x 10(6) M(-1) s(-1); and equilibrium constant, K=1.7+/-0.7. This step is followed by a proton-assisted conformational change of the C-lobe: direct rate constant, k(2)=(3+/-0.3) x 10(6) M(-1) s(-1); reverse rate constant, k(-2)=(1.6+/-0.3) x 10(-2) s(-1); and equilibrium constant, K(2a)=5.3+/-1.5 nM. The two final steps are slow changes in the conformation of the protein (0.5 h and 72 h), which allow it to achieve its final thermodynamic state and also to acquire second cobalt. The cobalt-saturated transferrin in an unknown state (TCo(2)) interacts with R1 in two different steps. The first is an ultra-fast interaction of the C-lobe of TCo(2) with the helical domain of R1: direct rate constant, k(3)=(4.4+/-0.6)x10(10) M(-1) s(-1); reverse rate constant, k(-3)=(3.6+/-0.6) x 10(4) s(-1); and dissociation constant, K(1d)=0.82+/-0.25 muM. The second is a very slow interaction of the N-lobe of TCo(2) with the protease-like domain of R1. This increases the stability of the protein-protein adduct by 30-fold with an average overall dissociation constant K(d)=25+/-10 nM. The main trigger in the R1-mediated iron acquisition is the ultra-fast interaction of the metal-loaded C-lobe of T with R1. This step is much faster than endocytosis, which in turn is much faster than the interaction of the N-lobe of T with the protease-like domain. This can explain why other metal-loaded transferrins or a protein such as HFE-with a lower affinity for R1 than iron-saturated transferrin but with, however, similar or higher affinities for the helical domain than the C-lobe-competes with iron-saturated transferrin in an unknown state towards interaction with R1.  相似文献   

17.
The endocytosis of diferric transferrin and accumulation of its iron by freshly isolated rabbit reticulocytes was studied using 59Fe-125I-transferrin. Internalized transferrin was distinguished from surface-bound transferrin by its resistance to release during treatment with Pronase at 4 degrees C. Endocytosis of diferric transferrin occurs at the same rate as exocytosis of apotransferrin, the rate constants being 0.08 min-1 at 22 degrees C, 0.19 min-1 at 30 degrees C, and 0.45 min-1 at 37 degrees C. At 37 degrees C, the maximum rate of transferrin endocytosis by reticulocytes is approximately 500 molecules/cell/s. The recycling time for transferrin bound to its receptor is about 3 min at this temperature. Neither transferrin nor its receptor is degraded during the intracellular passage. When a steady state has been reached between endocytosis and exocytosis of the ligand, about 90% of the total cell-bound transferrin is internal. Endocytosis of transferrin was found to be negligible below 10 degrees C. From 10 to 39 degrees C, the effect of temperature on the rate of endocytosis is biphasic, the rate increasing sharply above 26 degrees C. Over the temperature range 12-26 degrees C, the apparent activation energy for transferrin endocytosis is 33.0 +/- 2.7 kcal/mol, whereas from 26-39 degrees C the activation energy is considerably lower, at 12.3 +/- 1.6 kcal/mol. Reticulocytes accumulate iron atoms from diferric transferrin at twice the rate at which transferrin molecules are internalized, implying that iron enters the cell while still bound to transferrin. The activation energies for iron accumulation from transferrin are similar to those of endocytosis of transferrin. This study provides further evidence that transferrin-iron enters the cell by receptor-mediated endocytosis and that iron release occurs within the cell.  相似文献   

18.
The growing impact of protein aggregation pathologies, together with the current high need for extensive information on protein structures are focusing much interest on the physics underlying the nucleation and growth of protein aggregates and crystals. Sickle Cell Hemoglobin (HbS), a point-mutant form of normal human Hemoglobin (HbA), is the first recognized and best-studied case of pathologically aggregating protein. Here we reanalyze kinetic data on nucleation of deoxy-HbS aggregates by referring them to the (concentration-dependent) temperature T(s) characterizing the occurrence of the phase transition of liquid-liquid demixing (LLD) of the solution. In this way, and by appropriate scaling of kinetic data at different concentrations, so as to normalize their spans, the apparently disparate sets of data are seen to fall on a master curve. Expressing the master curve vs. the parameter epsilon = (T - T(s)) / T(s), familiar from phase transition theory, allows eliciting the role of anomalously large concentration fluctuations associated with the LLD phase transition and also allows decoupling quantitatively the role of such fluctuations from that of microscopic, inter-protein interactions leading to nucleation. Referring to epsilon shows how in a narrow temperature span, that is at T - T(s), nucleation kinetics can undergo orders-of-magnitude changes, unexpected in terms of ordinary chemical kinetics. The same is true for similarly small changes of other parameters (pH, salts, precipitants), capable of altering T(s) and consequently epsilon. This offers the rationale for understanding how apparently minor changes of parameters can dramatically affect protein aggregation and related diseases.  相似文献   

19.
J Bentz  N Düzgüne?  S Nir 《Biochemistry》1985,24(4):1064-1072
The effect of temperature and divalent cation binding (Ca2+, Sr2+, Ba2+) on the kinetic rate constants of aggregation and fusion of large phosphatidylserine liposomes is measured for the first time. Fusion is monitored by the Tb3+/dipicolinate assay. Fusion rate constants increase with temperature (15-35 degrees C) in a roughly linear fashion. These rate constants are not otherwise sensitive to whether the temperature is above or below the phase transition temperature of the Ba2+ or Sr2+ complex of phosphatidylserine, as measured by differential scanning calorimetry. Hence, the isothermal transition of the acyl chains from liquid-crystalline to gel phase induced by the cations is not the driving force of the initial fusion event. The aggregation rate constants increase with temperature, and it is the temperature dependence of the energetics of close approach of the liposomes which underlies this increase. On the other hand, the aggregation becomes more reversible at higher temperatures, which has also been observed with monovalent cation induced liposome aggregation where there is no fusion. Calculations on several cases show that the potential energy minimum holding the liposome dimer aggregates together is approximately 5-6 kT deep. This result implies that the aggregation step is highly reversible; i.e., if fusion were not occurring, no stable aggregates would form.  相似文献   

20.
The binding of cadmium(II) to human serum transferrin in 0.01 M N-(2-hydroxyethyl)-piperazine-N'-2-ethanesulfonic acid with 5 mM bicarbonate at 25 degrees C has been evaluated by difference ultraviolet spectroscopy. Equilibrium constants were determined by competition versus three different low molecular weight chelating agents: nitrilotriacetic acid, ethylenediamine-N,N'-diacetic acid, and triethylenetetramine. Conditional equilibrium constants for the sequential binding of two cadmium ions to transferrin under the stated experimental conditions are log K1 = 5.95 +/- 0.10 and log K2 = 4.86 +/- 0.13. A linear free energy relationship for the complexation of cadmium and zinc has been prepared by using equilibrium data on 243 complexes of these metal ions with low molecular weight ligands. The transferrin binding constants for cadmium and zinc are in good agreement with this linear free energy relationship. This indicates that the larger size of the cadmium(II) ion does not significantly hinder its binding to the protein.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号