Sie sind auf Seite 1von 6

2,3-Bisphosphoglyceric acid

From Wikipedia, the free encyclopedia

2,3-Bisphosphoglyceric acid

IUPAC name 2,3-Bisphosphoglycerate Other names 2,3-Diphosphoglyceric acid; 2,3-Diphosphoglycerate; 2,3-Bisphosphoglycerate

Identifiers Abbreviations 2,3-BPG; 2,3-DPG; 23BPG CAS number 138-81-8 PubChem ChEBI Jmol-3D images 61 CHEBI:17720 Image 1 (http://chemapps.stolaf.edu/jmol/jmol.php? model=O%3DP%28O%29%28OC%28C%28%3DO%29O%29COP%28%3DO%29%28O%29O%29O) SMILES InChI Properties Molecular formula C3H8O10P2 ChemSpider 161681 , 60 (Racemic)

Molar mass

266.04 g mol1 (what is this?) (verify)

Except where noted otherwise, data are given for materials in their standard state (at 25 C, 100 kPa) Infobox references

2,3-Bisphosphoglyceric acid (2,3-Bisphosphoglycerate or 2,3-BPG, also known as 2,3-diphosphoglycerate or 2,3-DPG) is a three-carbon isomer of the glycolytic intermediate 1,3-bisphosphoglyceric acid (1,3-BPG). 2,3BPG is present in human red blood cells (RBC; erythrocyte) at approximately 5 mmol/L. It binds with greater affinity to deoxygenated hemoglobin (e.g. when the red cell is near respiring tissue) than it does to oxygenated hemoglobin (e.g., in the lungs) due to spatial changes: 2,3-BPG (whose size is estimated at about 9 angstroms) fits in the deoxygenated hemoglobin configuration (11 angstroms), but not as well in the oxygenated (5 angstroms). It interacts with deoxygenated hemoglobin beta subunits by decreasing their affinity for oxygen, so it allosterically promotes the release of the remaining oxygen molecules bound to the hemoglobin, thus enhancing the ability of RBCs to release oxygen near tissues that need it most. 2,3-BPG is thus an allosteric effector. Its function was discovered in 1967 by Reinhold Benesch and Ruth Benesch.[1]

Contents
1 Metabolism 2 Effects of binding 3 Fetal hemoglobin 4 Diseases related to 2,3-BPG 5 2,3 BPG during haemodialysis 6 References 7 External links

Metabolism
2,3-BPG is formed from 1,3-BPG by the enzyme 2,3-BPG mutase. It can then be broken down by 2,3-BPG phosphatase to form 3-phosphoglycerate. Its synthesis and breakdown are, therefore, a way around a step of glycolysis.

Erythrocytes synthesize and degrade the 2.3-BPG by a diversion of the glycolytic pathway. The first phase of glucose catabolism includes glucose phosphorylation, isomerization and another phosphorylation to bear fructose-1,6-bisphosphate (F-1,6-BP). Cleavage of fructose 1, 6-bisphosphate yields two molecules:

glyceraldehyde 3-phosphate (G3P and DHAP respectively)and dihydroxyacetone 3 phosphate. These two molecules are isomers and are readily converterd into one another by triose phosphate isomerase. The equilibrium of this conversion lies heavily on the side of DHAP for two reasons. Firstly so the glycolytic pathway does not get oversaturated and secondly so that the biochemistry of glycerol can be tied into glycolysis. DHAP can be converted into glycerol when the supply of DHAP is plentiful and glycerol can then be used as a substrate for phase 2 glycolysis when glycolytic intermediates are scarce. The second phase of glucose catabolism converts G3P to 3phophoglycerate (3-PG). During the first reaction step, G3P is phosphorylated with a high-energy phosphate and oxidized to 1,3-bisphosphoglycerate (1,3-BPG), through the action of glyceralgehyde-3-phosphate dehydrogenase (G3PD). 1,3-BPG may be dephosphorylated by phosphoglycerate kinase (PGK), generating ATP, or it may be shunted into the Luebering-Rapapport pathway, where bisphosphoglycerate mutase catalyzes the transfer of a phosphoryl group from C1 to C2 of 1,3-BPG, giving 2,3-BPG. 2,3-BPG, the most concentrated organophosphate in the erythrocyte, forms 3-PG by the action of diphosphoglycerate phosphatase. The concentration on 2,3-BPG varies inversely with the pH, which is inhibitory to catalytic action of bisphosphoglyceromutase. The third phase of anaerobic glucose catabolism involves conversion of 3-PG to pyruvate with the generation of ATP. There is a delicate balance between the need to generate ATP to support energy requirements for cell metabolism and the need to maintain appropriate oxygenation/deoxygenation status of hemoglobin. This balance is maintained by dephosphorilation of 1,3-BPG to 2,3-BPG, which enhances the deoxygenation of hemoglobin. Low pH inhibits the activity of biphosphoglyceromutase and activates bisphosphoglyerate phosphatase, which favors generation of ATP.[2]

Effects of binding
When 2,3-BPG binds to deoxyhemoglobin, it acts to stabilize the low oxygen affinity state (T state) of the oxygen carrier. It fits neatly into the cavity of the deoxy- conformation, exploiting the molecular symmetry and positive polarity by forming salt bridges with lysine and histidine residues in the four subunits of hemoglobin. The R state, with oxygen bound to a heme group, has a different conformation and does not allow this interaction. By itself, hemoglobin has sigmoid-like kinetics, which makes easier another subunits binding (the first molecule of oxygen helps the following to link). By selectively binding to deoxyhemoglobin, 2,3-BPG stabilizes the T state conformation, making it harder for oxygen to bind hemoglobin and more Oxygen-haemoglobin dissociation likely to be released to adjacent tissues. 2,3-BPG is part of a feedback loop curve that can help prevent tissue hypoxia in conditions where it is most likely to occur. Conditions of low tissue oxygen concentration such as high altitude (2,3-BPG levels are higher in those acclimated to high altitudes), airway obstruction, or congestive heart failure will tend to cause RBCs to generate more 2,3-BPG in their effort to generate energy by allowing more oxygen to be released in tissues deprived of oxygen. Ultimately, this mechanism increases oxygen release from RBCs under circumstances where it is needed most. This release is potentiated by the Bohr effect in tissues with high energetic demands. Bohr effect is another useful way to solve the affinity problem of the hemoglobin, and its related to the pH and the CO2. Its important to highlight that the behaviour of myoglobin doesnt work in the same way, as 2,3BPG has no effect on it.

Fetal hemoglobin
It is interesting to note that fetal hemoglobin (HbF) exhibits a low affinity for 2,3-BPG, resulting in a higher binding

affinity for oxygen. This increased oxygen-binding affinity relative to that of adult hemoglobin (HbA) is due to HbF's having two / dimers as opposed to the two / dimers of HbA. The positive histidine residues of HbA -subunits that are essential for forming the 2,3-BPG binding pocket are replaced by serine residues in HbF -subunits. Like that, histidine n143 gets lost, so 2,3-BPG has difficulties in linking to the fetal hemoglobin, and it looks like the pure hemoglobin. Thats the way O2 flows from the mother to the fetus. As we can see in the following image, fetal hemoglobin has more affinity to oxygen than adult hemoglobin. Moreover, myoglobin has the highest affinity to oxygen.

Differences between myoglobin (Mb), fetal hemoglobin (Hb F), adult hemoglobin (Hb A)

Diseases related to 2,3-BPG


Hyperthyroidism A 2004 study checked the effects of thyroid hormone on 2,3-BPG levels. The result was that the hyperthyroidism modulates in vivo 2,3-BPG content in erythrocytes by changes in the expression of phosphoglycerate mutase (PGM) and 2,3-BPG synthase. This result shows that the increase in the 2,3-BPG content of erythrocytes observed in hyperthyroidism doesnt depend on any variation in the rate of circulating hemoglobin, but seems to be a direct consequence of the stimulating effect of thyroid hormones on erythrocyte glycolytic activity.[3]

Iron deficiency anaemia This illness causes a lack of iron, and as 2,3-BPG needs this chemical element to be synthesized, BPG concentration decreases and hemoglobin binds tightly to oxygen. That makes difficult to release the oxygen in tissues. Chronic respiratory disease with hypoxia Recently, scientists have found similarities between low amounts of 2,3-BPG with the occurrence of high altitude pulmonary edema at high altitudes.

CONCENTRATION OF 2.3-BPG ERYTHROCYTE FOUND IN DIFFERENT CLINICAL SITUATIONS STUDIED


n 1 2 3 4 Normality Hyperthyroidism Chronic respiratory disease with hypoxia 120 35 Hb (g/dl) 14.2 1.6 13.7 1.4 10.0 1.7 16.4 2.2 2,3-BPG (mM) 4.54 0.57 5.66 0.69 5.79 1.02 5.29 1.13

Iron deficiency anaemia 40 47

2,3 BPG during haemodialysis


In a 1998 study, erythrocyte 2,3-BPG concentration was analysed during the haemodialysis process. The 2,3-BPG concentration was expressed relative to the haemoglobin tetramer (Hb4) concentration as the 2,3-BPG/Hb4 ratio. Physiologically, an increase in 2.3-BPG levels would be expected to counteract the hypoxia that is frequently observed in this process. Nevertheless, the results show a 2,3-BPG/Hb4 ratio decreased. This is due to the procedure itself: mechanical stress on the erythrocytes is believed to cause the 2,3-BPG escape, which is then removed by haemodialysis. The concentrations of calcium, phosphate, creatinine, urea and albumin didnt correlate significantly with the total change in 2,3-BPG/Hb4 ratio. However, the ratio sampled just before dialysis correlated significantly and positively with the total weekly dosage of erythropoietin (main hormone in the erythrocytes formation) given to the patients.[4]

References
1. ^ Benesch, R.; Benesch, R.E. (1967). "The effect of organic phosphates from the human erythrocyte on the allosteric properties of hemoglobin.". Biochem Biophys Res Commun 26 (2): 1627. doi:10.1016/0006291X(67)90228-8 (http://dx.doi.org/10.1016%2F0006-291X%2867%2990228-8) . PMID 6030262 (http://www.ncbi.nlm.nih.gov/pubmed/6030262) . 2. ^ Mller-Sterl, W.; Mller-Sterl, W. (2008). "Biochemistry: Fundamentals Of Medicine And The Science Of Life.". Editorial Revert: 660. 3. ^ Gonzlez-Cinca N; Prez de la Ossa P, Carreras J, Climent F (2004). ."Effects of thyroid hormone and hypoxia on 2,3-bisphosphoglycerate, bisphosphoglycerate synthase and phosphoglycerate mutase in rabbit erythroblasts and reticulocytes in vivo".. 4. ^ Nielsen AL; Andersen EM, Jrgensen LG, Jensen HA. (1998). "Oxygen and 2,3 biphosphoglycerate (2,3-BPG) during haemodialysis"..

Berg, J.M., Tymockzko, J.L. and Stryer L. Biochemistry (5th ed). W.H. Freeman and Co, New York, 1995. ISBN 0-7167-4684-0. -684392408 (http://www.gpnotebook.co.uk/simplepage.cfm?ID=-684392408) at GPnotebook Online medical dictionary (http://www.online-medical-dictionary.org/2,3-DPG.asp?q=2%2C3-DPG) Nelson, David L.; Cox, Michael M.; Lehninger, Albert L. "Principles of Biochemistry (4th ed)". W.H. Freeman, 2005. ISBN 978071674339. Muller-esterl,W. "Biochemistry: "Fundamentals Of Medicine And The Science Of Life (2nd ed.)". Revert, 2008. ISBN 8429173935. Rodak. "Hematology. Clinical principles and applications (2nd ed.)" Elsevier Science, Philadelphia, 2003. ISBN 950-06-1876-1. Gonzlez-Cinca N, Prez de la Ossa P, Carreras J, Climent F."Effects of thyroid hormone and hypoxia on 2,3-bisphosphoglycerate, bisphosphoglycerate synthase and phosphoglycerate mutase in rabbit erythroblasts and reticulocytes in vivo". Unitat de Bioqumica, Departament de Cincies Fisiolgiques I, Institut d'Investigacions Biomdiques August Pi i Sunyer, Universitat de Barcelona, Barcelona, Spain, 2004. Nielsen AL, Andersen EM, Jrgensen LG, Jensen HA. "Oxygen and 2,3 biphosphoglycerate (2,3BPG) during haemodialysis". Department of Nephrology, Hvidovre University Hospital, Denmark, 1998. "Anales de la Real Academia Nacional de Medicina (cuaderno cuarto)". ISSN: 0034-0634

External links
A live model of the effect of changing 2,3 BPG on the oxyhaemoglobin saturation curve (http://www.altitude.org/hemoglobin_saturation.php) Retrieved from "http://en.wikipedia.org/wiki/2,3-Bisphosphoglyceric_acid" Categories: Organophosphates | Respiratory physiology This page was last modified on 8 September 2011 at 12:06. Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. See Terms of use for details. Wikipedia is a registered trademark of the Wikimedia Foundation, Inc., a non-profit organization.

Das könnte Ihnen auch gefallen