Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 43
Filter
Add more filters










Publication year range
1.
Biochim Biophys Acta Proteins Proteom ; 1869(1): 140544, 2021 01.
Article in English | MEDLINE | ID: mdl-32971286

ABSTRACT

Murine serine racemase (SR), the enzyme responsible for the biosynthesis of the neuromodulator d-serine, was reported to form a complex with glyceraldehyde 3-phosphate dehydrogenase (GAPDH), resulting in SR inhibition. In this work, we investigated the interaction between the two human orthologues. We were not able to observe neither the inhibition nor the formation of the SR-GAPDH complex. Rather, hSR is inhibited by the hGAPDH substrate glyceraldehyde 3-phosphate (G3P) in a time- and concentration-dependent fashion, likely through a covalent reaction of the aldehyde functional group. The inhibition was similar for the two G3P enantiomers but it was not observed for structurally similar aldehydes. We ruled out a mechanism of inhibition based on the competition with either pyridoxal phosphate (PLP) - described for other PLP-dependent enzymes when incubated with small aldehydes - or ATP. Nevertheless, the inhibition time course was affected by the presence of hSR allosteric and orthosteric ligands, suggesting a conformation-dependence of the reaction.


Subject(s)
Adenosine Triphosphate/chemistry , Enzyme Inhibitors/chemistry , Glyceraldehyde 3-Phosphate/chemistry , Glyceraldehyde-3-Phosphate Dehydrogenases/chemistry , Pyridoxal Phosphate/chemistry , Racemases and Epimerases/chemistry , 2,3-Diphosphoglycerate/chemistry , 2,3-Diphosphoglycerate/metabolism , Adenosine Triphosphate/metabolism , Aldehydes/chemistry , Aldehydes/metabolism , Catalytic Domain , Cloning, Molecular , Enzyme Inhibitors/metabolism , Escherichia coli/genetics , Escherichia coli/metabolism , Gene Expression , Genetic Vectors/chemistry , Genetic Vectors/metabolism , Glyceraldehyde/chemistry , Glyceraldehyde/metabolism , Glyceraldehyde 3-Phosphate/metabolism , Glyceraldehyde-3-Phosphate Dehydrogenases/genetics , Glyceraldehyde-3-Phosphate Dehydrogenases/metabolism , Humans , Kinetics , Models, Molecular , Protein Binding , Pyridoxal Phosphate/metabolism , Racemases and Epimerases/antagonists & inhibitors , Racemases and Epimerases/genetics , Racemases and Epimerases/metabolism , Recombinant Proteins/chemistry , Recombinant Proteins/genetics , Recombinant Proteins/metabolism , Stereoisomerism , Substrate Specificity
2.
Hemoglobin ; 43(1): 42-49, 2019 Jan.
Article in English | MEDLINE | ID: mdl-31060394

ABSTRACT

Nonenzymatic glycation (NEG) of human hemoglobin (Hb A) consists of initial non covalent, reversible steps involving glucose and amino acid residues, which may also involve effector reagent(s) in the formation of labile Hb A1c (the conjugate acid of the Schiff base). Labile Hb A1c can then undergo slow, largely irreversible, formation of stable Hb A1c (the Amadori product). Stable Hb A1c is measured to assess diabetic progression after labile Hb A1c removal. This study aimed to increase the understanding of the distinctions between labile and stable Hb A1c from a mechanistic perspective in the presence of 2,3-bisphosphoglycerate (2,3-BPG). 2,3-Bisphosphoglycerate is an effector reagent that reversibly binds in the Hb A1c pocket and modestly enhances overall NEG rate. The deprotonation of C2 on labile Hb A1c in the formation of the Amadori product was previously proposed to be rate-limiting. Computational chemistry was used here to identify the mechanism(s) by which 2,3-BPG facilitates the deprotonation of C2 on labile Hb A1c. 2,3-Bisphosphoglycerate is capable of abstracting protons on C2 and the α-nitrogen of labile Hb A1c and can also deprotonate water and/or amino acid residues, therefore preparing these secondary reagents to deprotonate labile Hb A1c. Parallel reactions not leading to an Amadori product were found that include formation of the neutral Schiff base, dissociation of glucose from the protein, and cyclic glycosylamine formation. These heretofore under appreciated parallel reactions may help explain both the selective removal of labile from stable Hb A1c and the slow rate of NEG.


Subject(s)
2,3-Diphosphoglycerate/metabolism , Glycated Hemoglobin/metabolism , 2,3-Diphosphoglycerate/chemistry , Amino Acids/chemistry , Glucose/chemistry , Glucose/metabolism , Glycated Hemoglobin/chemistry , Glycosylation , Humans , Protein Binding , Protein Multimerization
3.
Biochem Biophys Res Commun ; 511(4): 800-805, 2019 04 16.
Article in English | MEDLINE | ID: mdl-30837154

ABSTRACT

The antibiotic moenomycin A is a phosphoglycerate derivative with a C25-moenocinyl chain and a branched oligosaccharide. Formation of the C25-chain is catalyzed by the enzyme MoeN5 with geranyl pyrophosphate (GPP) and the sugar-linked 2-Z,E-farnesyl-3-phosphoglycerate (FPG) as its substrates. Previous complex crystal structures with GPP and long-chain alkyl glycosides suggested that GPP binds to the S1 site in a similar way as in most other α-helical prenyltransferases (PTs), and FPG is likely to assume a bent conformation in the S2 site. However, two FPG derivatives synthesized in the current study were found in the S1 site rather than S2 in their complex crystal structures with MoeN5. Apparently S1 is the preferred site for prenyl-containing ligand, and S2 binding may proceed only after S1 is occupied. Thus, like most trans-type PTs, MoeN5 may employ a sequential ionization-condensation-elimination mechanism that involves a carbocation intermediate.


Subject(s)
Bacterial Proteins/metabolism , Dimethylallyltranstransferase/metabolism , Streptomyces/metabolism , 2,3-Diphosphoglycerate/chemistry , 2,3-Diphosphoglycerate/metabolism , Amino Acid Sequence , Bacterial Proteins/chemistry , Bambermycins/metabolism , Crystallography, X-Ray , Dimethylallyltranstransferase/chemistry , Molecular Docking Simulation , Protein Conformation , Sequence Alignment , Streptomyces/chemistry , Substrate Specificity
4.
Sci Rep ; 7(1): 15281, 2017 11 10.
Article in English | MEDLINE | ID: mdl-29127281

ABSTRACT

Elevated sphingosine 1-phosphate (S1P) is detrimental in Sickle Cell Disease (SCD), but the mechanistic basis remains obscure. Here, we report that increased erythrocyte S1P binds to deoxygenated sickle Hb (deoxyHbS), facilitates deoxyHbS anchoring to the membrane, induces release of membrane-bound glycolytic enzymes and in turn switches glucose flux towards glycolysis relative to the pentose phosphate pathway (PPP). Suppressed PPP causes compromised glutathione homeostasis and increased oxidative stress, while enhanced glycolysis induces production of 2,3-bisphosphoglycerate (2,3-BPG) and thus increases deoxyHbS polymerization, sickling, hemolysis and disease progression. Functional studies revealed that S1P and 2,3-BPG work synergistically to decrease both HbA and HbS oxygen binding affinity. The crystal structure at 1.9 Å resolution deciphered that S1P binds to the surface of 2,3-BPG-deoxyHbA and causes additional conformation changes to the T-state Hb. Phosphate moiety of the surface bound S1P engages in a highly positive region close to α1-heme while its aliphatic chain snakes along a shallow cavity making hydrophobic interactions in the "switch region", as well as with α2-heme like a molecular "sticky tape" with the last 3-4 carbon atoms sticking out into bulk solvent. Altogether, our findings provide functional and structural bases underlying S1P-mediated pathogenic metabolic reprogramming in SCD and novel therapeutic avenues.


Subject(s)
Anemia, Sickle Cell/metabolism , Erythrocytes, Abnormal/metabolism , Hemoglobin A/metabolism , Hemoglobin, Sickle/metabolism , Lysophospholipids/metabolism , Sphingosine/analogs & derivatives , 2,3-Diphosphoglycerate/chemistry , 2,3-Diphosphoglycerate/metabolism , Anemia, Sickle Cell/pathology , Animals , Erythrocytes, Abnormal/pathology , Female , Hemoglobin A/chemistry , Hemoglobin, Sickle/chemistry , Hemolysis , Humans , Lysophospholipids/chemistry , Male , Mice , Mice, Transgenic , Oxidative Stress , Pentose Phosphate Pathway , Sphingosine/chemistry , Sphingosine/metabolism
5.
Biophys Chem ; 228: 87-97, 2017 09.
Article in English | MEDLINE | ID: mdl-28743047

ABSTRACT

As a prelude to separating tertiary from quaternary structure contributions to the Bohr effect, we employed the Wyman equation to analyze Bohr data for human hemoglobin to which 2,3-bisphosphoglycerate, 2,3-BPG, is bound. Changes in the pKas of the histidine Bohr groups result in a net reduction of their contributions to the Bohr effect at pH 7.4 compared to their contributions in stripped hemoglobin. The non-histidine 2,3-BPG binding groups - the ß-chain terminal amino group and Lys82ß - make negative and positive contributions, respectively, to the Bohr effect. The final result is that the Bohr effect at physiological pH is higher for 2,3-BPG bound compared to stripped hemoglobin. Contributions linked to His2ß, His77ß and His143ß enable us to separate tertiary from quaternary Bohr contributions in stripped and in 2,3-BPG bound hemoglobin. Both contributions serve to make the Bohr effect for 2,3-BPG bound hemoglobin higher than for stripped hemoglobin at physiological pH.


Subject(s)
2,3-Diphosphoglycerate/chemistry , Hemoglobins/chemistry , 2,3-Diphosphoglycerate/metabolism , Hemoglobins/metabolism , Humans , Hydrogen-Ion Concentration , Kinetics , Protein Binding
6.
J Chromatogr B Analyt Technol Biomed Life Sci ; 1061-1062: 89-96, 2017 Sep 01.
Article in English | MEDLINE | ID: mdl-28711785

ABSTRACT

Many hemolytic anemias results in major metabolic abnormalities: two common metabolite abnormalities include increased levels of 2,3-diphosphoglycerate (2,3-DPG) and decreased levels of adenosine triphosphate (ATP). To better monitor the concentration changes of these metabolites, the development of a reliable LC-MS/MS method to quantitatively profile the concentrations of 2, 3-DPG and ATP in whole blood is essential to understand the effects of investigational therapeutics. Accurate quantification of both compounds imposes great challenges to bioanalytical scientists due to their polar, ionic and endogenous nature. Here we present an LC-MS/MS method for the reliable quantification of 2,3-DPG and ATP from K2EDTA human whole blood (WB) simultaneously. Whole blood samples were spiked with stable isotope labeled internal standards, processed by protein precipitation extraction, and analyzed using zwitterionic ion chromatography-hydrophilic interaction chromatography (ZIC-HILIC) coupled with tandem mass spectrometry. The linear analytical range of the assay was 50-3000µg/mL. The fit-for-purpose method demonstrated excellent accuracy and precision. The overall accuracy was within ±10.5% (%RE) for both analytes and the intra- and inter-assay precision (%CV) were less than 6.7% and 6.2% for both analytes, respectively. ATP and 2,3-DPG were found to be stable in human K2EDTA blood for at least 8h at 4°C, 96days when stored at -70°C and after three freeze/thaw cycles. The assay has been successfully applied to K2EDTA human whole blood samples to support clinical studies.


Subject(s)
2,3-Diphosphoglycerate/blood , Adenosine Triphosphate/blood , Chromatography, Liquid/methods , Tandem Mass Spectrometry/methods , 2,3-Diphosphoglycerate/chemistry , Adenosine Triphosphate/chemistry , Edetic Acid , Humans , Limit of Detection , Linear Models , Reproducibility of Results
7.
Proteins ; 85(8): 1507-1517, 2017 Aug.
Article in English | MEDLINE | ID: mdl-28459140

ABSTRACT

A new algorithm for comparison of protein dynamics is presented. Compared protein structures are superposed and their modes of motions are calculated using the anisotropic network model. The obtained modes are aligned using the dynamic programming algorithm of Needleman and Wunsch, commonly used for sequence alignment. Dynamical comparison of hemoglobin in the T and R2 states reveals that the dynamics of the allosteric effector 2,3-bisphosphoglycerate binding site is different in the two states. These differences can contribute to the selectivity of the effector to the T state. Similar comparison of the ionotropic glutamate receptor in the kainate+(R,R)-2b and ZK bound states reveals that the kainate+(R,R)-2b bound states slow modes describe upward motions of ligand binding domain and the transmembrane domain regions. Such motions may lead to the opening of the receptor. The upper lobes of the LBDs of the ZK bound state have a smaller interface with the amino terminal domains above them and have a better ability to move together. The present study exemplifies the use of dynamics comparison as a tool to study protein function. Proteins 2017; 85:1507-1517. © 2014 Wiley Periodicals, Inc.


Subject(s)
2,3-Diphosphoglycerate/chemistry , Alanine/analogs & derivatives , Excitatory Amino Acid Agonists/chemistry , Hemoglobins/chemistry , Kainic Acid/chemistry , Receptors, AMPA/chemistry , Sequence Alignment/methods , Uracil/chemistry , 2,3-Diphosphoglycerate/metabolism , Alanine/chemistry , Alanine/metabolism , Algorithms , Allosteric Site , Animals , Binding Sites , Excitatory Amino Acid Agonists/metabolism , Halogenation , Hemoglobins/metabolism , Humans , Kainic Acid/metabolism , Ligands , Molecular Dynamics Simulation , Protein Binding , Protein Conformation, alpha-Helical , Protein Conformation, beta-Strand , Protein Interaction Domains and Motifs , Rats , Receptors, AMPA/metabolism , Sequence Homology, Amino Acid , Thermodynamics , Uracil/metabolism
8.
Oncotarget ; 8(7): 12093-12107, 2017 Feb 14.
Article in English | MEDLINE | ID: mdl-28076845

ABSTRACT

Phosphoglycerate mutase 1 (PGAM1) catalyzes the eighth step of glycolysis and is often found upregulated in cancer cells. To test the hypothesis that the phosphorylation of tyrosine 26 residue of PGAM1 greatly enhances its activity, we performed both conventional and steered molecular dynamics simulations on the binding and unbinding of PGAM1 to its substrates, with tyrosine 26 either phosphorylated or not. We analyzed the simulated data in terms of structural stability, hydrogen bond formation, binding free energy, etc. We found that tyrosine 26 phosphorylation enhances the binding of PGAM1 to its substrates through generating electrostatic environment and structural features that are advantageous to the binding. Our results may provide valuable insights into computer-aided design of drugs that specifically target cancer cells with PGAM1 tyrosine 26 phosphorylated.


Subject(s)
Glycolysis , Molecular Dynamics Simulation , Phosphoglycerate Mutase/metabolism , Tyrosine/metabolism , 2,3-Diphosphoglycerate/chemistry , 2,3-Diphosphoglycerate/metabolism , Algorithms , Amino Acid Sequence , Glyceric Acids/chemistry , Glyceric Acids/metabolism , Humans , Hydrogen Bonding , Neoplasms/genetics , Neoplasms/metabolism , Neoplasms/pathology , Phosphoglycerate Mutase/chemistry , Phosphoglycerate Mutase/genetics , Phosphorylation , Principal Component Analysis , Protein Binding , Sequence Homology, Amino Acid , Static Electricity , Substrate Specificity , Thermodynamics , Tyrosine/chemistry , Tyrosine/genetics
9.
J Zhejiang Univ Sci B ; 16(5): 395-403, 2015 May.
Article in English | MEDLINE | ID: mdl-25990057

ABSTRACT

The safe use of intraoperative blood salvage (IBS) in cancer surgery remains controversial. Here, we investigated the killing effect of cisplatin combined with hyperthermia on human hepatocarcinoma (HepG2) cells and erythrocytes from IBS in vitro. HepG2 cells were mixed with concentrated erythrocytes and pretreated with cisplatin (50, 100, and 200 µg/ml) alone at 37 °C for 60 min and cisplatin (25, 50, 100, and 200 µg/ml) combined with hyperthermia at 42 °C for 60 min. After pretreatment, the cell viability, colony formation and DNA metabolism in HepG2 and the Na(+)-K(+)-ATPase activity, 2,3-diphosphoglycerate (2,3-DPG) concentration, free hemoglobin (Hb) level, osmotic fragility, membrane phosphatidylserine externalization, and blood gas variables in erythrocytes were determined. Pretreatment with cisplatin (50, 100, and 200 µg/ml) combined with hyperthermia (42 °C) for 60 min significantly decreased HepG2 cell viability, and completely inhibited colony formation and DNA metabolism when the HepG2 cell concentration was 5×10(4) ml(-1) in the erythrocyte (P<0.01). Erythrocytic Na(+)-K(+)-ATPase activity, 2,3-DPG level, phosphatidylserine externalization, and extra-erythrocytic free Hb were significantly altered by hyperthermia plus high concentrations of cisplatin (100 and 200 µg/ml) (P<0.05), but not by hyperthermia plus 50 µg/ml cisplatin (P>0.05). In conclusion, pretreatment with cisplatin (50 µg/ml) combined with hyperthermia (42 °C) for 60 min effectively eliminated HepG2 cells from IBS but did not significantly affect erythrocytes in vitro.


Subject(s)
Antineoplastic Agents/therapeutic use , Cisplatin/therapeutic use , Erythrocytes/drug effects , Operative Blood Salvage , 2,3-Diphosphoglycerate/chemistry , Adult , Aged , Cell Survival , Combined Modality Therapy , Dose-Response Relationship, Drug , Drug Screening Assays, Antitumor , Female , Hemoglobins/chemistry , Hep G2 Cells , Humans , Hyperthermia, Induced , Male , Middle Aged , Osmosis , Phosphatidylserines/chemistry , Phospholipids/chemistry , Sodium-Potassium-Exchanging ATPase/chemistry
10.
Biochemistry ; 53(43): 6793-9, 2014 Nov 04.
Article in English | MEDLINE | ID: mdl-25325574

ABSTRACT

Cross-linked human hemoglobins have been evaluated for clinical use as circulating oxygen carriers. However, their induction of vasoactivity was sufficiently problematic to lead to the cessation of clinical trials. The source of vasoactivity is likely to be endothelial extravasation causing the scavenging of endogenous nitric oxide. It was recently shown that species that consist of two coupled hemoglobin tetramers do not evoke vasoactivity in a sensitive murine model. Presumably these materials are too large to extravasate. In order to make this class of material more readily available, there is a need for improved methods that can form a cross-linked bis-tetramer without producing smaller species at the same time. A potentially efficient route to cross-linking and coupling two Hb tetramers is through phase-directed copper-catalyzed azide alkyne cycloaddition (PDCuAAC). However, introduction of the necessary azide-containing cross-link gives mixtures of tetrameric and bis-tetrameric proteins, as the PDCuAAC process appears to be limited to only those proteins where the cross-link containing the azide is exclusively within the ß-subunits. In order to block formation of the azide cross-link within the α-subunits, subunit-specific introduction of the azide is necessary. This is achieved by blocking reaction at the reactive amino groups of the ß-subunits in the site that binds the allosteric activator 2,3-diphosphoglycerate (DPG) with inositol hexaphosphate (IHP), permitting α-selective acetylation with acetyl 3,5-dibromosalicylate. After removal of IHP, reaction with an anionic cross-linker containing an azide group occurs within the ß-subunits. The resulting α-acetylated ß-ß'-cross-linked hemoglobin azide (acHb>-N3) undergoes efficient PDCuAAC with bis-alkynes to produce cross-linked bis-tetramers. Analysis of circular dichroism spectra of the modified species shows that there is little change in the structure of the globin chains as a result of the chemical modifications. The oxygenation properties are consistent with those needed for effective oxygenation in circulation, while the bis-tetrameric structure is sufficiently large to avoid extravasation and depletion of nitric oxide.


Subject(s)
2,3-Diphosphoglycerate/chemistry , Hemoglobin A/chemistry , Phytic Acid/chemistry , Protein Multimerization , Acetylation , Humans , Protein Structure, Quaternary
11.
Biochem J ; 458(3): 439-48, 2014 Mar 15.
Article in English | MEDLINE | ID: mdl-24423178

ABSTRACT

The p53-induced protein TIGAR [TP53 (tumour protein 53)-induced glycolysis and apoptosis regulator] is considered to be a F26BPase (fructose-2,6-bisphosphatase) with an important role in cancer cell metabolism. The reported catalytic efficiency of TIGAR as an F26BPase is several orders of magnitude lower than that of the F26BPase component of liver or muscle PFK2 (phosphofructokinase 2), suggesting that F26BP (fructose 2,6-bisphosphate) might not be the physiological substrate of TIGAR. We therefore set out to re-evaluate the biochemical function of TIGAR. Phosphatase activity of recombinant human TIGAR protein was tested on a series of physiological phosphate esters. The best substrate was 23BPG (2,3-bisphosphoglycerate), followed by 2PG (2-phosphoglycerate), 2-phosphoglycolate and PEP (phosphoenolpyruvate). In contrast the catalytic efficiency for F26BP was approximately 400-fold lower than that for 23BPG. Using genetic and shRNA-based cell culture models, we show that loss of TIGAR consistently leads to an up to 5-fold increase in the levels of 23BPG. Increases in F26BP levels were also observed, albeit in a more limited and cell-type dependent manner. The results of the present study challenge the concept that TIGAR acts primarily on F26BP. This has significant implications for our understanding of the metabolic changes downstream of p53 as well as for cancer cell metabolism in general. It also suggests that 23BPG might play an unrecognized function in metabolic control.


Subject(s)
Glycolates/chemistry , Intracellular Signaling Peptides and Proteins/chemistry , Phosphoric Monoester Hydrolases/chemistry , 2,3-Diphosphoglycerate/chemistry , Animals , Apoptosis Regulatory Proteins , Glycolates/metabolism , Humans , Intracellular Signaling Peptides and Proteins/genetics , Mice , Muscle, Skeletal/enzymology , Phosphoric Monoester Hydrolases/metabolism , Recombinant Proteins/chemistry , Substrate Specificity , Transcription, Genetic
12.
Phys Chem Chem Phys ; 16(9): 3946-54, 2014 Mar 07.
Article in English | MEDLINE | ID: mdl-24441588

ABSTRACT

Bisphosphoglycerate mutase (BPGM) is a multi-activity enzyme. Its main function is to synthesize the 2,3-bisphosphoglycerate, the allosteric effector of hemoglobin. This enzyme can also catalyze the 2,3-bisphosphoglycerate to the 3-phosphoglycerate. In this study, the reaction mechanisms of both the phosphatase and the synthase activities of human bisphosphoglycerate mutase were theoretically calculated by using the quantum mechanics/molecular mechanics method based on the metadynamics and umbrella sampling simulations. The simulation results not only show the free energy curve of the phosphatase and the synthase reactions, but also reveal the important role of some residues in the active site. Additionally, the energy barriers of the two reactions indicate that the activity of the synthase in human bisphosphoglycerate mutase is much higher than that of the phosphatase. The estimated reaction barriers are consistent with the experimental data. Therefore, our work can give important information to understand the catalytic mechanism of the bisphosphoglycerate mutase family.


Subject(s)
Bisphosphoglycerate Mutase/metabolism , Molecular Dynamics Simulation , Quantum Theory , 2,3-Diphosphoglycerate/chemistry , 2,3-Diphosphoglycerate/metabolism , Binding Sites , Biocatalysis , Bisphosphoglycerate Mutase/chemistry , Catalytic Domain , Glyceric Acids/chemistry , Glyceric Acids/metabolism , Humans , Kinetics , Thermodynamics
13.
Vox Sang ; 106(1): 61-7, 2014 Jan.
Article in English | MEDLINE | ID: mdl-24117672

ABSTRACT

BACKGROUND AND OBJECTIVES: There are few studies on transient warming of red blood cells (RBCs). Occasional storage outside restricted temperature range often results in destroying of the RBC unit, even after a short period of time due to national guidelines. This study evaluates the in vitro effects associated with such accidental warming on RBCs stored in saline-adenine-glucose-mannitol (SAGM) and prepared within 8 h after blood collection. STUDY DESIGN AND METHODS: This study includes both repeated short-term exposure of RBCs to room temperature for 6 h as wells as warming for either 6, 12, 18 or 24 h after 1 week or after 3 weeks of storage in two separate studies. RBCs were stored for 42 days. We weekly measured pH, K(+) , glucose, lactate, haemolysis, red cell ATP and 2,3-diphosphoglycerate. RESULTS: The lowest individual ATP value observed in any of the groups of warmed units was 2·6 µmol/g haemoglobin. Increased haemolysis in warmed units was noted in two of the studies. None of the individual units exceeded the European maximum limit of 0·8% haemolysis. CONCLUSION: Our results suggest that quality of RBCs after transient warming will be maintained at acceptable levels specified in standards and in previous studies. However, increased haemolysis was observed when transient warming occurred during the second part of the storage period of 6 weeks suggesting that RBCs are more vulnerable to warming by the end of storage.


Subject(s)
Blood Preservation/methods , Erythrocytes , Hemolysis , Organ Preservation Solutions/chemistry , 2,3-Diphosphoglycerate/chemistry , Adenine/chemistry , Environmental Exposure , Glucose/chemistry , Hot Temperature , Humans , Mannitol/chemistry , Sodium Chloride/chemistry
14.
J Phys Chem B ; 117(42): 13058-68, 2013 Oct 24.
Article in English | MEDLINE | ID: mdl-23957820

ABSTRACT

Hemoglobin is a complex system that undergoes conformational changes in response to oxygen, allosteric effectors, mutations, and environmental changes. Here, we study allostery and polymerization of hemoglobin and its variants by application of two previously described methods: (i) AllosMod for simulating allostery dynamics given two allosterically related input structures and (ii) a machine-learning method for dynamics- and structure-based prediction of the mutation impact on allostery (Weinkam et al. J. Mol. Biol. 2013, 425, 647-661), now applicable to systems with multiple coupled binding sites, such as hemoglobin. First, we predict the relative stabilities of substates and microstates of hemoglobin, which are determined primarily by entropy within our model. Next, we predict the impact of 866 annotated mutations on hemoglobin's oxygen binding equilibrium. We then discuss a subset of 30 mutations that occur in the presence of the sickle cell mutation and whose effects on polymerization have been measured. Seven of these HbS mutations occur in three predicted druggable binding pockets that might be exploited to directly inhibit polymerization; one of these binding pockets is not apparent in the crystal structure, but only in structures generated by AllosMod. For the 30 mutations, we predict that mutation-induced conformational changes within a single tetramer tend not to significantly impact polymerization; instead, these mutations more likely impact polymerization by directly perturbing a polymerization interface. Finally, our analysis of allostery allows us to hypothesize why hemoglobin evolved to have multiple subunits and a persistent low frequency sickle cell mutation.


Subject(s)
Hemoglobin, Sickle/genetics , Hemoglobin, Sickle/metabolism , 2,3-Diphosphoglycerate/chemistry , 2,3-Diphosphoglycerate/metabolism , Allosteric Regulation , Binding Sites , Entropy , Hemoglobin, Sickle/chemistry , Humans , Oxygen/chemistry , Oxygen/metabolism , Point Mutation , Polymerization , Protein Binding , Protein Structure, Tertiary
15.
Artif Cells Nanomed Biotechnol ; 41(2): 109-15, 2013 Apr.
Article in English | MEDLINE | ID: mdl-23305389

ABSTRACT

Four phosphonate derivates of 2,3-diphosphoglycerate (2,3-DPG), in which the phosphate group is replaced by a methylene or difluoromethylene, were successfully synthesized for use as allosteric modulators of hemoglobin (Hb) O2 affinity. The syntheses were accomplished in four steps and the reagents were converted to their potassium salts to allow for effective binding with Hb in aqueous media. O2 equilibrium measurements of the chemically modified Hbs exhibited P50 values in the range 8.9-12.8 with Hill coefficients in the range of 1.5-2.4.


Subject(s)
2,3-Diphosphoglycerate/chemistry , Hemoglobins/chemistry , Organophosphonates/chemical synthesis , Oxygen/chemistry , Allosteric Regulation , Allosteric Site , Humans , Hydrogen-Ion Concentration , Organophosphonates/chemistry , Protein Binding , Solutions , Structure-Activity Relationship
16.
Am J Physiol Regul Integr Comp Physiol ; 304(1): R43-9, 2013 Jan 01.
Article in English | MEDLINE | ID: mdl-23174858

ABSTRACT

During winter hibernation, brown bears (Ursus arctos) reduce basal O(2) consumption rate to ∼25% compared with the active state, while body temperature decreases moderately (to ∼30°C), suggesting a temperature-independent component in their metabolic depression. To establish whether changes in O(2) consumption during hibernation correlate with changes in blood O(2) affinity, we took blood samples from the same six individuals of hibernating and nonhibernating free-ranging brown bears during winter and summer, respectively. A single hemoglobin (Hb) component was detected in all samples, indicating no switch in Hb synthesis. O(2) binding curves measured on red blood cell lysates at 30°C and 37°C showed a less temperature-sensitive O(2) affinity than in other vertebrates. Furthermore, hemolysates from hibernating bears consistently showed lower cooperativity and higher O(2) affinity than their summer counterparts, regardless of the temperature. We found that this increase in O(2) affinity was associated with a significant decrease in the red cell Hb-cofactor 2,3-diphosphoglycerate (DPG) during hibernation to approximately half of the summer value. Experiments performed on purified Hb, to which DPG had been added to match summer and winter levels, confirmed that the low DPG content was the cause of the left shift in the Hb-O(2) equilibrium curve during hibernation. Levels of plasma lactate indicated that glycolysis is not upregulated during hibernation and that metabolism is essentially aerobic. Calculations show that the increase in Hb-O(2) affinity and decrease in cooperativity resulting from decreased red cell DPG may be crucial in maintaining a fairly constant tissue oxygen tension during hibernation in vivo.


Subject(s)
2,3-Diphosphoglycerate/metabolism , Hemoglobins/physiology , Hibernation/physiology , Oxygen/physiology , Ursidae/physiology , 2,3-Diphosphoglycerate/analysis , 2,3-Diphosphoglycerate/chemistry , Animals , Erythrocytes/metabolism , Female , Hemoglobins/chemistry , Lactic Acid/blood , Male , Oxygen/blood , Oxygen/chemistry , Seasons , Ursidae/blood
17.
Transfusion ; 52(1): 195-200, 2012 Jan.
Article in English | MEDLINE | ID: mdl-21790622

ABSTRACT

The focus of this study was to determine if there is significant data to prohibit short-term storage of red blood cells (RBCs; i.e., <24 hr) at 1 to 10°C rather than 1 to 6°C, which occurs not uncommonly when RBCs are stored in a cooler for a patient during surgery. This document will describe the evidence in the literature to date regarding the potential impact of having RBCs temporarily in the 1 to 10°C range versus in the 1 to 6°C range, if any, on key measures of the quality of RBC storage: potassium, adenosine triphosphate, 2,3-diphosphoglycerate, posttransfusion survival, and bacterial contamination.


Subject(s)
Blood Preservation/methods , Erythrocytes/chemistry , 2,3-Diphosphoglycerate/chemistry , Adenosine Triphosphate/chemistry , Humans , Potassium/chemistry , Temperature
18.
Ann Biomed Eng ; 38(4): 1683-701, 2010 Apr.
Article in English | MEDLINE | ID: mdl-20162361

ABSTRACT

New mathematical model equations for O(2) and CO(2) saturations of hemoglobin (S(HbO)(2) and S(HbCO)(2) are developed here from the equilibrium binding of O(2) and CO(2) with hemoglobin inside RBCs. They are in the form of an invertible Hill-type equation with the apparent Hill coefficients KHbO(2) and KHbCO(2) in the expressions for SHbO(2) and SHbCO(2) dependent on the levels of O(2) and CO(2) partial pressures (P(O)(2) and P(CO)(2)), pH, 2,3-DPG concentration, and temperature in blood. The invertibility of these new equations allows PO(2) and PCO(2) to be computed efficiently from S(HbO)(2) and S(HbCO)(2) and vice versa. The oxyhemoglobin (HbO(2)) and carbamino-hemoglobin (HbCO(2)) dissociation curves computed from these equations are in good agreement with the published experimental and theoretical curves in the literature. The model solutions describe that, at standard physiological conditions, the hemoglobin is about 97.2% saturated by O(2) and the amino group of hemoglobin is about 13.1% saturated by CO(2). The O(2) and CO(2) content in whole blood are also calculated here from the gas solubilities, hematocrits, and the new formulas for S(HbO)(2) and S(HbCO)(2). Because of the mathematical simplicity and invertibility, these new formulas can be conveniently used in the modeling of simultaneous transport and exchange of O(2) and CO(2) in the alveoli-blood and blood-tissue exchange systems.


Subject(s)
2,3-Diphosphoglycerate/chemistry , Carbon Dioxide/chemistry , Carboxyhemoglobin/chemistry , Oxygen/chemistry , Oxyhemoglobins/chemistry , Hydrogen-Ion Concentration , Kinetics , Models, Chemical , Temperature
19.
Proc Natl Acad Sci U S A ; 105(16): 5998-6003, 2008 Apr 22.
Article in English | MEDLINE | ID: mdl-18413611

ABSTRACT

The Rapoport-Luebering glycolytic bypass comprises evolutionarily conserved reactions that generate and dephosphorylate 2,3-bisphosphoglycerate (2,3-BPG). For >30 years, these reactions have been considered the responsibility of a single enzyme, the 2,3-BPG synthase/2-phosphatase (BPGM). Here, we show that Dictyostelium, birds, and mammals contain an additional 2,3-BPG phosphatase that, unlike BPGM, removes the 3-phosphate. This discovery reveals that the glycolytic pathway can bypass the formation of 3-phosphoglycerate, which is a precursor for serine biosynthesis and an activator of AMP-activated protein kinase. Our 2,3-BPG phosphatase activity is encoded by the previously identified gene for multiple inositol polyphosphate phosphatase (MIPP1), which we now show to have dual substrate specificity. By genetically manipulating Mipp1 expression in Dictyostelium, we demonstrated that this enzyme provides physiologically relevant regulation of cellular 2,3-BPG content. Mammalian erythrocytes possess the highest content of 2,3-BPG, which controls oxygen binding to hemoglobin. We determined that total MIPP1 activity in erythrocytes at 37 degrees C is 0.6 mmol 2,3-BPG hydrolyzed per liter of cells per h, matching previously published estimates of the phosphatase activity of BPGM. MIPP1 is active at 4 degrees C, revealing a clinically significant contribution to 2,3-BPG loss during the storage of erythrocytes for transfusion. Hydrolysis of 2,3-BPG by human MIPP1 is sensitive to physiologic alkalosis; activity decreases 50% when pH rises from 7.0 to 7.4. This phenomenon provides a homeostatic mechanism for elevating 2,3-BPG levels, thereby enhancing oxygen release to tissues. Our data indicate greater biological significance of the Rapoport-Luebering shunt than previously considered.


Subject(s)
2,3-Diphosphoglycerate/metabolism , Evolution, Molecular , Glycolysis , Phosphoric Monoester Hydrolases/metabolism , 2,3-Diphosphoglycerate/analysis , 2,3-Diphosphoglycerate/chemistry , Alkalosis , Amino Acid Sequence , Animals , Birds , Dictyostelium/enzymology , Erythrocytes/chemistry , Erythrocytes/enzymology , Erythrocytes/metabolism , Hemoglobins/metabolism , Humans , Hydrogen-Ion Concentration , Hydrolysis , Molecular Sequence Data , Oxygen/metabolism , Phosphoric Monoester Hydrolases/chemistry , Phosphorylation , Protozoan Proteins/chemistry , Rats
20.
Biophys J ; 94(7): 2737-51, 2008 Apr 01.
Article in English | MEDLINE | ID: mdl-18096633

ABSTRACT

Recent functional studies reported on human adult hemoglobin (HbA) show that heterotropic effector-linked tertiary structural changes are primarily responsible for modulating the oxygen affinity of hemoglobin. We present the results of 6-ns molecular dynamics simulations performed to gain insights into the dynamical and structural details of these effector-linked tertiary changes. All-atom simulations were carried out on a series of models generated for T- and R-state HbA, and for 2,3-diphosphoglycerate-bound models. Cross-correlation analyses identify both intra- and intersubunit correlated motions that are perturbed by the presence of the effector. Principal components analysis was used to decompose the covariance matrix extracted from the simulations and reconstruct the trajectories along the principal coordinates representative of functionally important collective motions. It is found that HbA in both quaternary states exists as ensembles of tertiary conformations that introduce dynamic heterogeneity in the protein. 2,3-Diphosphoglycerate induces significant perturbations in the fluctuations of both HbA states that translate into the protein visiting different tertiary conformations within each quaternary state. The analysis reveals that the presence of the effector affects the most important components of HbA motions and that heterotropic effectors modify the overall dynamics of the quaternary equilibrium via tertiary changes occurring in regions where conserved functionally significant residues are located, namely in the loop regions between helices C and E, E and F, and F and G, and in concerted helix motions. The changes are not apparent when comparing the available x-ray crystal structures in the presence and absence of effector, but are striking when comparing the respective dynamic tertiary conformations of the R and T tetramers.


Subject(s)
2,3-Diphosphoglycerate/chemistry , Hemoglobin A/chemistry , Hemoglobin A/ultrastructure , Models, Chemical , Models, Molecular , Binding Sites , Computer Simulation , Kinetics , Protein Binding , Protein Conformation
SELECTION OF CITATIONS
SEARCH DETAIL
...