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
KEGG
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.

Structural binding to hemoglobin

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 ꞵ 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.

Physiological effects

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

An increase in 2,3-BPG essentially facilitates the delivery of oxygen from hemoglobin in target tissues, at a cost of also making it somewhat more difficult for hemoglobin to take up oxygen in the lungs. This mechanisms makes maternal-fetal oxygenation more efficient, as fetal 2,3-BPG is lower than maternal levels, resulting in a higher uptake of oxygen by the fetal blood in the placenta.

2,3-BPG may also serve to physiologically counteract certain metabolic disturbances to the oxygen-hemoglobin dissociation curve. For example, at high altitudes, low atmospheric oxygen content of oxygen can cause hyperventilation and resultant metabolic alkalosis which causes an abnormal left-shift of the oxygen-hemoglobin dissociation curve, and this can be counteracted by an increase in 2,3-BPG. [2] Traditional teaching has claimed that the physiologic increased 2,3-BPG seen at high altitudes is simply to make it easier for oxygen to be delivered in target tissues, but this mechanism by itself is refuted by the reasoning that the decreased oxygen affinity would also inhibit oxygen uptake in the lungs, and arguably result in a net decrease in total oxygen delivery to target tissues. [2]

Maternal-fetal oxygenation

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.

The increased maternal 2,3-BPG also causes a decreased affinity for oxygen takeup in the lungs, but this is usually compensated by a physiologic increased respiratory rate in pregnancy. [3]

Fetal hemoglobin (HbF), on the other hand, 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. [4]

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. [5]

See also

Related Research Articles

<span class="mw-page-title-main">Glycolysis</span> Series of interconnected biochemical reactions

Glycolysis is the metabolic pathway that converts glucose into pyruvate and, in most organisms, occurs in the liquid part of cells. 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.

<span class="mw-page-title-main">Hypoxia (medicine)</span> Medical condition of lack of oxygen in the tissues

Hypoxia is a condition in which the body or a region of the body is deprived of adequate oxygen supply at the tissue level. Hypoxia may be classified as either generalized, affecting the whole body, or local, affecting a region of the body. Although hypoxia is often a pathological condition, variations in arterial oxygen concentrations can be part of the normal physiology, for example, during strenuous physical exercise.

<span class="mw-page-title-main">Hemoglobin</span> Metalloprotein that binds with oxygen

Hemoglobin is a protein containing iron that facilitates the transport of oxygen in red blood cells. Almost all vertebrates contain hemoglobin, with the exception of the fish family Channichthyidae. Hemoglobin in the blood carries oxygen from the respiratory organs to the other tissues of the body, where it releases the oxygen to enable aerobic respiration which powers the animal's metabolism. A healthy human has 12 to 20 grams of hemoglobin in every 100 mL of blood. Hemoglobin is a metalloprotein, a chromoprotein, and globulin.

<span class="mw-page-title-main">Tumor hypoxia</span> Situation where tumor cells have been deprived of oxygen

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 microenvironments in solid tumors are a result of available oxygen being consumed within 70 to 150 μm of tumor 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.

<span class="mw-page-title-main">Fetal hemoglobin</span> Oxygen carrier protein in the human fetus

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 than adult forms of hemoglobin, allowing it to bind oxygen more strongly; this in turn enables the developing fetus to retrieve oxygen from the mother's bloodstream, which occurs through the placenta found in the mother's uterus.

<span class="mw-page-title-main">Bohr effect</span> Concept in physiology

The Bohr effect is a phenomenon first described in 1904 by the Danish physiologist Christian Bohr. Hemoglobin's oxygen binding affinity (see oxygen–haemoglobin dissociation curve) is inversely related both to acidity and to the concentration of carbon dioxide. That is, the Bohr effect refers to the shift in the oxygen dissociation curve caused by changes in the concentration of carbon dioxide or the pH of the environment. Since carbon dioxide reacts with water to form carbonic acid, an increase in CO2 results in a decrease in blood pH, resulting in hemoglobin proteins releasing their load of oxygen. Conversely, a decrease in carbon dioxide provokes an increase in pH, which results in hemoglobin picking up more oxygen.

Carboxyhemoglobin 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 historic name, "carbonic oxide", and evolved through Germanic and British English etymological influences; the preferred IUPAC nomenclature is carbonylhemoglobin.

<span class="mw-page-title-main">Fetal circulation</span> Circulatory system of fetuses

In humans, the circulatory system is different before and after birth. The fetal circulation is composed of the placenta, umbilical blood vessels encapsulated by the umbilical cord, heart and systemic blood vessels. A major difference between the fetal circulation and postnatal circulation is that the lungs are not used during the fetal stage resulting in the presence of shunts to move oxygenated blood and nutrients from the placenta to the fetal tissue. At birth, the start of breathing and the severance of the umbilical cord prompt various changes that quickly transform fetal circulation into postnatal circulation.

<span class="mw-page-title-main">1,3-Bisphosphoglyceric acid</span> 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.

<span class="mw-page-title-main">Oxygen–hemoglobin dissociation curve</span> Visual tool used to understand how human blood carries and releases oxygen

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. Vice versa, it is true that a high concentration of CO2 facilitates dissociation of oxyhemoglobin, though this is the result of two distinct processes (Bohr effect and Margaria-Green effect) and should be distinguished from Haldane effect.

<span class="mw-page-title-main">Phosphoglycerate kinase</span> 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 :

<span class="mw-page-title-main">Bisphosphoglycerate mutase</span> Enzyme

Bisphosphoglycerate mutase is an enzyme expressed in 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 glycolytic cousin, phosphoglycerate mutase (PGM), which favors these two functions, but can also catalyze the synthesis of 2,3-BPG to a lesser extent.

<span class="mw-page-title-main">Phosphoglycerate mutase</span> Class of enzymes

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.

Carbamino refers to an adduct generated by the addition of carbon dioxide to the free amino group of an amino acid or a protein, such as hemoglobin forming carbaminohemoglobin.

<span class="mw-page-title-main">Phosphoglycolate phosphatase</span>

Phosphoglycolate phosphatase(EC 3.1.3.18; systematic name 2-phosphoglycolate phosphohydrolase), 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:

<span class="mw-page-title-main">Oxygen saturation (medicine)</span> Medical measurement

Oxygen saturation is the fraction of oxygen-saturated haemoglobin relative to total haemoglobin 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 96–100 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.

The Root effect is a physiological phenomenon that occurs in fish hemoglobin, named after its discoverer R. W. Root. It is the phenomenon where an increased proton or carbon dioxide concentration (lower pH) lowers hemoglobin's affinity and carrying capacity for oxygen. The Root effect is to be distinguished from the Bohr effect where only the affinity to oxygen is reduced. Hemoglobins showing the Root effect show a loss of cooperativity at low pH. This results in the Hb-O2 dissociation curve being shifted downward and not just to the right. At low pH, hemoglobins showing the Root effect don't become fully oxygenated even at oxygen tensions up to 20kPa. This effect allows hemoglobin in fish with swim bladders to unload oxygen into the swim bladder against a high oxygen gradient. The effect is also noted in the choroid rete, the network of blood vessels which carries oxygen to the retina. In the absence of the Root effect, retia will result in the diffusion of some oxygen directly from the arterial blood to the venous blood, making such systems less effective for the concentration of oxygen. It has also been hypothesized that the loss of affinity is used to provide more oxygen to red muscle during acidotic stress.

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. Fish respond to hypoxia with varied behavioral, physiological, and cellular responses 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 nutrients 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.

<span class="mw-page-title-main">Luebering–Rapoport pathway</span>

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. 1 2 Webb KL, Dominelli PB, Baker SE, Klassen SA, Joyner MJ, Senefeld JW; et al. (2021). "Influence of High Hemoglobin-Oxygen Affinity on Humans During Hypoxia". Front Physiol. 12: 763933. doi: 10.3389/fphys.2021.763933 . PMC   8795792 . PMID   35095551.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  3. LoMauro A, Aliverti A (2015). "Respiratory physiology of pregnancy: Physiology masterclass". Breathe (Sheff). 11 (4): 297–301. doi:10.1183/20734735.008615. PMC   4818213 . PMID   27066123.
  4. 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.
  5. 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.