2,3-Bisphosphoglyceric acid

Last updated
2,3-Bisphosphoglyceric acid
Structure of 2,3-bisphosphoglyceric acid.png
Names
Preferred IUPAC name
2,3-Bis(phosphonooxy)propanoic acid
Other names
2,3-Diphosphoglyceric acid; 2,3-Diphosphoglycerate; 2,3-Bisphosphoglycerate
Identifiers
3D model (JSmol)
Abbreviations2,3-BPG; 2,3-DPG; 23BPG
ChEBI
ChemSpider
PubChem CID
UNII
  • InChI=1S/C3H8O10P2/c4-3(5)2(13-15(9,10)11)1-12-14(6,7)8/h2H,1H2,(H,4,5)(H2,6,7,8)(H2,9,10,11)/t2-/m1/s1 Yes check.svgY
    Key: XOHUEYCVLUUEJJ-UWTATZPHSA-N Yes check.svgY
  • Key: XOHUEYCVLUUEJJ-UWTATZPHBL
  • O=P(O)(OC(C(=O)O)COP(=O)(O)O)O
Properties
C3H8O10P2
Molar mass 266.035 g·mol−1
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
X mark.svgN  verify  (what is  Yes check.svgYX mark.svgN ?)

2,3-Bisphosphoglyceric acid (conjugate base 2,3-bisphosphoglycerate) (2,3-BPG), also known as 2,3-diphosphoglyceric acid (conjugate base 2,3-diphosphoglycerate) (2,3-DPG), is a three-carbon isomer of the glycolytic intermediate 1,3-bisphosphoglyceric acid (1,3-BPG).

Contents

D-2,3-BPG 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 blood cell is near respiring tissue) than it does to oxygenated hemoglobin (e.g., in the lungs) due to conformational differences: 2,3-BPG (with an estimated size of about 9 Å) fits in the deoxygenated hemoglobin conformation (with an 11-Angstrom pocket), but not as well in the oxygenated conformation (5 Angstroms). It interacts with deoxygenated hemoglobin beta subunits and decreases the affinity for oxygen and allosterically promotes the release of the remaining oxygen molecules bound to the hemoglobin. Therefore, it enhances 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]

Metabolism

2,3-BPG is formed from 1,3-BPG by the enzyme 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, with the net expense of one ATP per molecule of 2,3-BPG generated as the high-energy carboxylic acid-phosphate mixed anhydride bond is cleaved by bisphosphoglycerate mutase.

Pathway of generation of 2,3-bisphosphoglycerate.png

The normal glycolytic pathway generates 1,3-BPG, which may be dephosphorylated by phosphoglycerate kinase (PGK), generating ATP, or it may be shunted into the Luebering-Rapoport 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 bisphosphoglycerate phosphatase. The concentration of 2,3-BPG varies proportionately to the [H+].

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 isomerisation of 1,3-BPG to 2,3-BPG, which enhances the deoxygenation of hemoglobin.

Effects of binding

Oxygen-haemoglobin dissociation curve Oxyhaemoglobin dissociation curve.png
Oxygen-haemoglobin dissociation curve

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 ß 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. In selectively binding to deoxyhemoglobin, 2,3-BPG stabilizes the T state conformation, making it harder for oxygen to bind hemoglobin and more likely to be released to adjacent tissues. 2,3-BPG is part of a feedback loop 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, because changes in pH and oxygen modulate the enzymes that make and degrade it. [2] The accumulation of 2,3-BPG decreases the affinity of hemoglobin for 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 which hemoglobin's binding affinity for oxygen is also reduced by a lower pH and high concentration of carbon dioxide. In tissues with high energetic demands, oxygen is rapidly consumed, which increases the concentration of H+ and carbon dioxide. Through the Bohr effect, hemoglobin is induced to release more oxygen to supply cells that need it. In contrast, 2,3-BPG has no effect on the related compound myoglobin.(reference required)

In pregnant women, there is a 30% increase in intracellular 2,3-BPG. This lowers the maternal hemoglobin affinity for oxygen, and therefore allows more oxygen to be offloaded to the fetus in the maternal uterine arteries. The fetus has a low sensitivity to 2,3-BPG, so its hemoglobin has a higher affinity for oxygen. Therefore, although the pO2 in the uterine arteries is low, the fetal umbilical artery (which carries deoxygenated blood) can still get oxygenated from them.

Fetal hemoglobin

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 α/γ dimers as opposed to the two α/β dimers of HbA. The positive histidine residues of HbA β-subunits that are essential for forming the 2,3-BPG binding pocket are replaced by serine residues in HbF γ-subunits. Like that, histidine nº143 gets lost, so 2,3-BPG has difficulties in linking to the fetal hemoglobin, and it looks like the pure hemoglobin. Increased binding affinity of fetal hemoglobin relative to HbA facilitates the passage of oxygen across the placental membrane from the mother to the fetus.

Fetal hemoglobin chart.jpg

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

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 doesn’t 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]

Chronic anemia

Red cells increase their intracellular 2,3-BPG concentration as much as five times within one to two hours in patients with chronic anemia, when the oxygen carrying capacity of the blood is diminished. This results in a rightward shift of the oxygen dissociation curve and more oxygen being released to the 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
nHb (g/dl)2,3-BPG (mM)
1Normality12014.2 ± 1.64.54 ± 0.57
2Hyperthyroidism3513.7 ± 1.45.66 ± 0.69
3Iron deficiency anaemia4010.0 ± 1.75.79 ± 1.02
4Chronic respiratory disease with hypoxia4716.4 ± 2.25.29 ± 1.13

Hemodialysis

In a 1998 study, erythrocyte 2,3-BPG concentration was analyzed during the hemodialysis process. The 2,3-BPG concentration was expressed relative to the hemoglobin 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 hemodialysis. The concentrations of calcium, phosphate, creatinine, urea and albumin did not 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 erythrocyte formation) given to the patients. [4]

See also

Related Research Articles

Blood Organic fluid which transports nutrients throughout the organism

Blood is a body fluid in the circulatory system of humans and other vertebrates that delivers necessary substances such as nutrients and oxygen to the cells and transports metabolic waste products away from those same cells.

Glycolysis Metabolic pathway

Glycolysis is the metabolic pathway that converts glucose (C6H12O6), into pyruvic acid (CH3COCO2H). The free energy released in this process is used to form the high-energy molecules adenosine triphosphate (ATP) and reduced nicotinamide adenine dinucleotide (NADH). Glycolysis is a sequence of ten reactions catalyzed by enzymes.

Hemoglobin Oxygen-transport metalloprotein in red blood cells of most vertebrates

Hemoglobin,, abbreviated Hb or Hgb, is the iron-containing oxygen-transport metalloprotein in red blood cells (erythrocytes) of almost all vertebrates as well as the tissues of some invertebrates. Hemoglobin in blood carries oxygen from the respiratory organs to the rest of the body. There it releases the oxygen to permit aerobic respiration to provide energy to power functions of an organism in the process called metabolism. A healthy individual human has 12 to 20 grams of hemoglobin in every 100 mL of blood.

Hemoglobinopathy Medical condition

Hemoglobinopathy is the medical term for a group of inherited blood disorders and diseases that primarily affect red blood cells. They are single-gene disorders and, in most cases, they are inherited as autosomal co-dominant traits.

Tumor hypoxia

Tumor hypoxia is the situation where tumor cells have been deprived of oxygen. As a tumor grows, it rapidly outgrows its blood supply, leaving portions of the tumor with regions where the oxygen concentration is significantly lower than in healthy tissues. Hypoxic microenvironements in solid tumors are a result of available oxygen being consumed within 70 to 150 μm of tumour vasculature by rapidly proliferating tumor cells thus limiting the amount of oxygen available to diffuse further into the tumor tissue. In order to support continuous growth and proliferation in challenging hypoxic environments, cancer cells are found to alter their metabolism. Furthermore, hypoxia is known to change cell behavior and is associated with extracellular matrix remodeling and increased migratory and metastatic behavior.

Pyruvate kinase Class of enzymes

Pyruvate kinase is the enzyme involved in the last step of glycolysis. It catalyzes the transfer of a phosphate group from phosphoenolpyruvate (PEP) to adenosine diphosphate (ADP), yielding one molecule of pyruvate and one molecule of ATP. Pyruvate kinase was inappropriately named before it was recognized that it did not directly catalyze phosphorylation of pyruvate, which does not occur under physiological conditions. Pyruvate kinase is present in four distinct, tissue-specific isozymes in animals, each consisting of particular kinetic properties necessary to accommodate the variations in metabolic requirements of diverse tissues.

Fetal hemoglobin

Fetal hemoglobin, or foetal haemoglobin is the main oxygen carrier protein in the human fetus. Hemoglobin F is found in fetal red blood cells, and is involved in transporting oxygen from the mother's bloodstream to organs and tissues in the fetus. It is produced at around 6 weeks of pregnancy and the levels remain high after birth until the baby is roughly 2–4 months old. Hemoglobin F has a different composition from the adult forms of hemoglobin, which allows it to bind oxygen more strongly. This way, the developing fetus is able to retrieve oxygen from the mother's bloodstream, which occurs through the placenta found in the mother's uterus.

Carboxyhemoglobin, or carboxyhaemoglobin, is a stable complex of carbon monoxide and hemoglobin (Hb) that forms in red blood cells upon contact with carbon monoxide. Carboxyhemoglobin is often mistaken for the compound formed by the combination of carbon dioxide (carboxyl) and hemoglobin, which is actually carbaminohemoglobin. Carboxyhemoglobin terminology emerged when carbon monoxide was known by its ancient name carbonic oxide and evolved through Germanic and British English etymological influences; the preferred IUPAC nomenclature is carbonylhemoglobin.

Hemoglobin A

Hemoglobin A (HbA), also known as adult hemoglobin, hemoglobin A1 or α2β2, is the most common human hemoglobin tetramer, accounting for over 97% of the total red blood cell hemoglobin. Hemoglobin is an oxygen-binding protein, found in erythrocytes, which transports oxygen from the lungs to the tissues. Hemoglobin A is the most common adult form of hemoglobin and exists as a tetramer containing two alpha subunits and two beta subunits (α2β2). Hemoglobin A2 (HbA2) is a less common adult form of hemoglobin and is composed of two alpha and two delta-globin subunits. This hemoglobin makes up 1-3% of hemoglobin in adults.

Blood doping is a form of doping in which the number of red blood cells in the bloodstream is boosted in order to enhance athletic performance. Because such blood cells carry oxygen from the lungs to the muscles, a higher concentration in the blood can improve an athlete's aerobic capacity (VO2 max) and endurance. Blood doping can be achieved by making the body produce more red blood cells itself using drugs, giving blood transfusions either from another person or back to the same individual, or by using blood substitutes.

1,3-Bisphosphoglyceric acid Chemical compound

1,3-Bisphosphoglyceric acid (1,3-Bisphosphoglycerate or 1,3BPG) is a 3-carbon organic molecule present in most, if not all, living organisms. It primarily exists as a metabolic intermediate in both glycolysis during respiration and the Calvin cycle during photosynthesis. 1,3BPG is a transitional stage between glycerate 3-phosphate and glyceraldehyde 3-phosphate during the fixation/reduction of CO2. 1,3BPG is also a precursor to 2,3-bisphosphoglycerate which in turn is a reaction intermediate in the glycolytic pathway.

Oxygen–hemoglobin dissociation curve

The oxygen–hemoglobin dissociation curve, also called the oxyhemoglobin dissociation curve or oxygen dissociation curve (ODC), is a curve that plots the proportion of hemoglobin in its saturated (oxygen-laden) form on the vertical axis against the prevailing oxygen tension on the horizontal axis. This curve is an important tool for understanding how our blood carries and releases oxygen. Specifically, the oxyhemoglobin dissociation curve relates oxygen saturation (SO2) and partial pressure of oxygen in the blood (PO2), and is determined by what is called "hemoglobin affinity for oxygen"; that is, how readily hemoglobin acquires and releases oxygen molecules into the fluid that surrounds it.

The Haldane effect is a property of hemoglobin first described by John Scott Haldane, within which oxygenation of blood in the lungs displaces carbon dioxide from hemoglobin, increasing the removal of carbon dioxide. Consequently, oxygenated blood has a reduced affinity for carbon dioxide. Thus, the Haldane effect describes the ability of hemoglobin to carry increased amounts of carbon dioxide (CO2) in the deoxygenated state as opposed to the oxygenated state. A high concentration of CO2 facilitates dissociation of oxyhemoglobin.

Phosphoglycerate kinase Enzyme

Phosphoglycerate kinase is an enzyme that catalyzes the reversible transfer of a phosphate group from 1,3-bisphosphoglycerate (1,3-BPG) to ADP producing 3-phosphoglycerate (3-PG) and ATP :

Bisphosphoglycerate mutase

Bisphosphoglycerate mutase (BPGM) is an enzyme unique to erythrocytes and placental cells. It is responsible for the catalytic synthesis of 2,3-Bisphosphoglycerate (2,3-BPG) from 1,3-bisphosphoglycerate. BPGM also has a mutase and a phosphatase function, but these are much less active, in contrast to its glycolitic cousin, phosphoglycerate mutase (PGM), which favors these two functions, but can also catalyze the synthesis of 2,3-BPG to a lesser extent.

Phosphoglycerate mutase

Phosphoglycerate mutase (PGM) is any enzyme that catalyzes step 8 of glycolysis - the internal transfer of a phosphate group from C-3 to C-2 which results in the conversion of 3-phosphoglycerate (3PG) to 2-phosphoglycerate (2PG) through a 2,3-bisphosphoglycerate intermediate. These enzymes are categorized into the two distinct classes of either cofactor-dependent (dPGM) or cofactor-independent (iPGM). The dPGM enzyme is composed of approximately 250 amino acids and is found in all vertebrates as well as in some invertebrates, fungi, and bacteria. The iPGM class is found in all plants and algae as well as in some invertebrate, fungi, and Gram-positive bacteria. This class of PGM enzyme shares the same superfamily as alkaline phosphatase.

Phosphoglycolate phosphatase

Phosphoglycolate phosphatase(PGP), also commonly referred to as phosphoglycolate hydrolase, 2-phosphoglycolate phosphatase, P-glycolate phosphatase, and phosphoglycollate phosphatase, is an enzyme responsible for catalyzing the conversion of 2-phosphoglycolate into glycolate and phosphate. First studied and purified within plants, phosphoglycolate phosphatase plays a major role in photorespiratory 2-phosphoglycolate metabolism, an essential pathway for photosynthesis in plants. The occurrence of photorespiration in plants, due to the lack of substrate specificity of rubisco, leads to the formation of 2-phosphoglycolate and 3-phosphogylcerate(PGA). PGA is the normal product of carboxylation and will enter the Calvin cycle. Phosphoglycolate, which is a potent inhibitor of phosphofructokinase and triosephosphate isomerase, must be quickly metabolized and transformed into a useful substrate, and phosphoglycolate phosphatase catalyzes the first step in the regeneration of 3-phosphoglycerate from 2-phosphoglycolate at the expense of energy in the form of ATP.

Oxygen saturation (medicine) Medical measurement

Oxygen saturation is the fraction of oxygen-saturated hemoglobin relative to total hemoglobin in the blood. The human body requires and regulates a very precise and specific balance of oxygen in the blood. Normal arterial blood oxygen saturation levels in humans are 95–100 percent. If the level is below 90 percent, it is considered low and called hypoxemia. Arterial blood oxygen levels below 80 percent may compromise organ function, such as the brain and heart, and should be promptly addressed. Continued low oxygen levels may lead to respiratory or cardiac arrest. Oxygen therapy may be used to assist in raising blood oxygen levels. Oxygenation occurs when oxygen molecules enter the tissues of the body. For example, blood is oxygenated in the lungs, where oxygen molecules travel from the air and into the blood. Oxygenation is commonly used to refer to medical oxygen saturation.

Fish are exposed to large oxygen fluctuations in their aquatic environment since the inherent properties of water can result in marked spatial and temporal differences in the concentration of oxygen (see oxygenation and underwater). Fish respond to hypoxia with varied behavioral, physiological, and cellular responses in order to maintain homeostasis and organism function in an oxygen-depleted environment. The biggest challenge fish face when exposed to low oxygen conditions is maintaining metabolic energy balance, as 95% of the oxygen consumed by fish is used for ATP production releasing the chemical energy of O2 through the mitochondrial electron transport chain. Therefore, hypoxia survival requires a coordinated response to secure more oxygen from the depleted environment and counteract the metabolic consequences of decreased ATP production at the mitochondria. This article is a review of the effects of hypoxia on all aspects of fish, ranging from behavior down to genes.

Luebering–Rapoport pathway

In biochemistry, the Luebering–Rapoport pathway is a metabolic pathway in mature erythrocytes involving the formation of 2,3-bisphosphoglycerate (2,3-BPG), which regulates oxygen release from hemoglobin and delivery to tissues. 2,3-BPG, the reaction product of the Luebering-Rapoport pathway was first described and isolated in 1925 by the Austrian biochemist Samuel Mitja Rapoport and his technical assistant Jane Luebering.

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): 162–7. doi:10.1016/0006-291X(67)90228-8. PMID   6030262.
  2. Mulquiney, PJ; Bubb, WA; Kuchel, PW (1999). "Model of 2,3-bisphosphoglycerate metabolism in the human erythrocyte based on detailed enzyme kinetic equations: in vivo kinetic characterization of 2,3-bisphosphoglycerate synthase/phosphatase using 13C and 31P NMR". Biochem J. 342 (3): 567–80. doi:10.1042/0264-6021:3420567. PMC   1220498 . PMID   10477268.
  3. González-Cinca N, Pérez de la Ossa P, Carreras J, Climent F (September 2004). "Effects of thyroid hormone and hypoxia on 2,3-bisphosphoglycerate, bisphosphoglycerate synthase and phosphoglycerate mutase in rabbit erythroblasts and reticulocytes in vivo". Hormone Research in Paediatrics. 62 (4): 191–196. doi:10.1159/000080897. PMID   15375329. S2CID   34271262.
  4. Nielsen AL, Andersen EM, Jørgensen LG, Jensen HA (Oct 1998). "Oxygen and 2,3 biphosphoglycerate (2,3-BPG) during haemodialysis". Scandinavian Journal of Clinical and Laboratory Investigation. 58 (6): 459–67. doi:10.1080/00365519850186256. PMID   9832337.