Glyoxylate cycle

Last updated
Overview of the Glyoxylate Cycle Glyoxylatepath.svg
Overview of the Glyoxylate Cycle

The glyoxylate cycle, a variation of the tricarboxylic acid cycle, is an anabolic pathway occurring in plants, bacteria, protists, and fungi. The glyoxylate cycle centers on the conversion of acetyl-CoA to succinate for the synthesis of carbohydrates. [1] In microorganisms, the glyoxylate cycle allows cells to use two carbons (C2 compounds), such as acetate, to satisfy cellular carbon requirements when simple sugars such as glucose or fructose are not available. [2] The cycle is generally assumed to be absent in animals, with the exception of nematodes at the early stages of embryogenesis. In recent years, however, the detection of malate synthase (MS) and isocitrate lyase (ICL), key enzymes involved in the glyoxylate cycle, in some animal tissue has raised questions regarding the evolutionary relationship of enzymes in bacteria and animals and suggests that animals encode alternative enzymes of the cycle that differ in function from known MS and ICL in non-metazoan species. [1] [3]

Contents

Plants as well as some algae and bacteria can use acetate as the carbon source for the production of carbon compounds. Plants and bacteria employ a modification of the TCA cycle called the glyoxylate cycle to produce four carbon dicarboxylic acid from two carbon acetate units. The glyoxylate cycle bypasses the two oxidative decarboxylation reactions of the TCA cycle and directly converts isocitrate through isocitrate lyase and malate synthase into malate and succinate.

The glyoxylate cycle was discovered in 1957 at the University of Oxford by Sir Hans Kornberg and his mentor Hans Krebs, resulting in a Nature paper Synthesis of Cell Constituents from C2-Units by a Modified Tricarboxylic Acid Cycle. [4]

Similarities with TCA cycle

The glyoxylate cycle uses five of the eight enzymes associated with the tricarboxylic acid cycle: citrate synthase, aconitase, succinate dehydrogenase, fumarase, and malate dehydrogenase. The two cycles differ in that in the glyoxylate cycle, isocitrate is converted into glyoxylate and succinate by isocitrate lyase (ICL) instead of into α-ketoglutarate. [1] This bypasses the decarboxylation steps that take place in the citric acid cycle (TCA cycle), allowing simple carbon compounds to be used in the later synthesis of macromolecules, including glucose. [2] Glyoxylate is subsequently combined with acetyl-CoA to produce malate, catalyzed by malate synthase. [1] Malate is also formed in parallel from succinate by the action of succinate dehydrogenase and fumarase.

Role in gluconeogenesis

Fatty acids from lipids are commonly used as an energy source by vertebrates as fatty acids are degraded through beta oxidation into acetate molecules. This acetate, bound to the active thiol group of coenzyme A, enters the citric acid cycle (TCA cycle) where it is fully oxidized to carbon dioxide. This pathway thus allows cells to obtain energy from fat. To use acetate from fat for biosynthesis of carbohydrates, the glyoxylate cycle, whose initial reactions are identical to the TCA cycle, is used.

Cell-wall containing organisms, such as plants, fungi, and bacteria, require very large amounts of carbohydrates during growth for the biosynthesis of complex structural polysaccharides, such as cellulose, glucans, and chitin. In these organisms, in the absence of available carbohydrates (for example, in certain microbial environments or during seed germination in plants), the glyoxylate cycle permits the synthesis of glucose from lipids via acetate generated in fatty acid β-oxidation.

The glyoxylate cycle bypasses the steps in the citric acid cycle where carbon is lost in the form of CO2. The two initial steps of the glyoxylate cycle are identical to those in the citric acid cycle: acetate → citrate → isocitrate. In the next step, catalyzed by the first glyoxylate cycle enzyme, isocitrate lyase, isocitrate undergoes cleavage into succinate and glyoxylate (the latter gives the cycle its name). Glyoxylate condenses with acetyl-CoA (a step catalyzed by malate synthase), yielding malate. Both malate and oxaloacetate can be converted into phosphoenolpyruvate, which is the product of phosphoenolpyruvate carboxykinase, the first enzyme in gluconeogenesis. The net result of the glyoxylate cycle is therefore the production of glucose from fatty acids. Succinate generated in the first step can enter into the citric acid cycle to eventually form oxaloacetate. [2]

Function in organisms

Plants

In plants the glyoxylate cycle occurs in special peroxisomes which are called glyoxysomes. This cycle allows seeds to use lipids as a source of energy to form the shoot during germination. The seed cannot produce biomass using photosynthesis because of lack of an organ to perform this function. The lipid stores of germinating seeds are used for the formation of the carbohydrates that fuel the growth and development of the organism.

The glyoxylate cycle can also provide plants with another aspect of metabolic diversity. This cycle allows plants to take in acetate both as a carbon source and as a source of energy. Acetate is converted to acetyl CoA (similar to the TCA cycle). This acetyl CoA can proceed through the glyoxylate cycle, and some succinate is released during the cycle. The four carbon succinate molecule can be transformed into a variety of carbohydrates through combinations of other metabolic processes; the plant can synthesize molecules using acetate as a source for carbon. The acetyl CoA can also react with glyoxylate to produce some NADPH from NADP+, which is used to drive energy synthesis in the form of ATP later in the electron transport chain. [5]

Pathogenic fungi

The glyoxylate cycle may serve an entirely different purpose in some species of pathogenic fungi. The levels of the main enzymes of the glyoxylate cycle, ICL and MS, are greatly increased upon contact with a human host. Mutants of a particular species of fungi that lacked ICL were also significantly less virulent in studies with mice compared to the wild type. The exact link between these two observations is still being explored, but it can be concluded that the glyoxylate cycle is a significant factor in the pathogenesis of these microbes. [6] [7]

Vertebrates

Vertebrates were once thought to be unable to perform this cycle because there was no evidence of its two key enzymes, isocitrate lyase and malate synthase. However, some research suggests that this pathway may exist in some, if not all, vertebrates. [8] [9] Specifically, some studies show evidence of components of the glyoxylate cycle existing in significant amounts in the liver tissue of chickens. Data such as these support the idea that the cycle could theoretically occur in even the most complex vertebrates. [10] Other experiments have also provided evidence that the cycle is present among certain insect and marine invertebrate species, as well as strong evidence of the cycle's presence in nematode species. However, other experiments refute this claim. [11] Some publications conflict on the presence of the cycle in mammals: for example, one paper has stated that the glyoxylate cycle is active in hibernating bears, [12] but this report was disputed in a later paper. [13] Evidence exists for malate synthase activity in humans due to a dual functional malate/B-methylmalate synthase of mitochondrial origin called CLYBL expressed in brown fat and kidney. [14] Vitamin D may regulate this pathway in vertebrates. [10] [15]

Inhibition of the glyoxylate cycle

Due to the central role of the glyoxylate cycle in the metabolism of pathogenic species including fungi and bacteria, enzymes of the glyoxylate cycle are current inhibition targets for the treatment of diseases. Most reported inhibitors of the glyoxylate cycle target the first enzyme of the cycle (ICL). Inhibitors were reported for Candida albicans for potential use as antifungal agents. [16] The mycobacterial glyoxylate cycle is also being targeted for potential treatments of tuberculosis. [17] [18]

Engineering concepts

The prospect of engineering various metabolic pathways into mammals which do not possess them is a topic of great interest for bio-engineers today. The glyoxylate cycle is one of the pathways which engineers have attempted to manipulate into mammalian cells. This is primarily of interest for engineers in order to increase the production of wool in sheep, which is limited by the access to stores of glucose. By introducing the pathway into sheep, the large stores of acetate in cells could be used in order to synthesize glucose through the cycle, allowing for increased production of wool. [19] Mammals are incapable of executing the pathway due to the lack of two enzymes, isocitrate lyase and malate synthase, which are needed in order for the cycle to take place. It is believed by some that the genes to produce these enzymes, however, are pseudogenic in mammals, meaning that the gene is not necessarily absent, rather, it is merely "turned off". [1]

In order to engineer the pathway into cells, the genes responsible for coding for the enzymes had to be isolated and sequenced, which was done using the bacteria E.coli, from which the AceA gene, responsible for encoding for isocitrate lyase, and the AceB gene, responsible for encoding for malate synthase were sequenced. [19] Engineers have been able to successfully incorporate the AceA and AceB genes into mammalian cells in culture, and the cells were successful in translating and transcribing the genes into the appropriate enzymes, proving that the genes could successfully be incorporated into the cell’s DNA without damaging the functionality or health of the cell. However, being able to engineer the pathway into transgenic mice has proven to be difficult for engineers. While the DNA has been expressed in some tissues, including the liver and small intestine in test animals, the level of expression is not high, and not found to be statistically significant. In order to successfully engineer the pathway, engineers would have to fuse the gene with promoters which could be regulated in order to increase the level of expression, and have the expression in the right cells, such as epithelial cells. [20]

Efforts to engineer the pathway into more complex animals, such as sheep, have not been effective. This illustrates that much more research needs to be done on the topic, and suggests it is possible that a high expression of the cycle in animals would not be tolerated by the chemistry of the cell. Incorporating the cycle into mammals will benefit from advances in nuclear transfer technology, which will enable engineers to examine and access the pathway for functional integration within the genome before its transfer to animals. [19]

There are possible benefits, however, to the cycle's absence in mammalian cells. The cycle is present in microorganisms that cause disease but is absent in mammals, for example humans. There is a strong plausibility of the development of antibiotics that would attack the glyoxylate cycle, which would kill the disease-causing microorganisms that depend on the cycle for their survival, yet would not harm humans where the cycle, and thus the enzymes that the antibiotic would target, are absent. [2]

See also

Related Research Articles

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

The citric acid cycle —also known as the Krebs cycle, Szent-Györgyi-Krebs cycle or the TCA cycle (tricarboxylic acid cycle)—is a series of chemical reactions to release stored energy through the oxidation of acetyl-CoA derived from carbohydrates, fats, and proteins. The Krebs cycle is used by organisms that respire (as opposed to organisms that ferment) to generate energy, either by anaerobic respiration or aerobic respiration. In addition, the cycle provides precursors of certain amino acids, as well as the reducing agent NADH, that are used in numerous other reactions. Its central importance to many biochemical pathways suggests that it was one of the earliest components of metabolism. Even though it is branded as a 'cycle', it is not necessary for metabolites to follow only one specific route; at least three alternative segments of the citric acid cycle have been recognized.

<span class="mw-page-title-main">Succinic acid</span> Dicarboxylic acid

Succinic acid is a dicarboxylic acid with the chemical formula (CH2)2(CO2H)2. In living organisms, succinic acid takes the form of an anion, succinate, which has multiple biological roles as a metabolic intermediate being converted into fumarate by the enzyme succinate dehydrogenase in complex 2 of the electron transport chain which is involved in making ATP, and as a signaling molecule reflecting the cellular metabolic state.

<span class="mw-page-title-main">Cellular respiration</span> Process to convert glucose to ATP in cells

Cellular respiration is the process by which biological fuels are oxidized in the presence of an inorganic electron acceptor, such as oxygen, to drive the bulk production of adenosine triphosphate (ATP), which contains energy. Cellular respiration may be described as a set of metabolic reactions and processes that take place in the cells of organisms to convert chemical energy from nutrients into ATP, and then release waste products.

<span class="mw-page-title-main">Acetyl-CoA</span> Chemical compound

Acetyl-CoA is a molecule that participates in many biochemical reactions in protein, carbohydrate and lipid metabolism. Its main function is to deliver the acetyl group to the citric acid cycle to be oxidized for energy production. Coenzyme A consists of a β-mercaptoethylamine group linked to the vitamin pantothenic acid (B5) through an amide linkage and 3'-phosphorylated ADP. The acetyl group of acetyl-CoA is linked to the sulfhydryl substituent of the β-mercaptoethylamine group. This thioester linkage is a "high energy" bond, which is particularly reactive. Hydrolysis of the thioester bond is exergonic (−31.5 kJ/mol).

Gluconeogenesis (GNG) is a metabolic pathway that results in the biosynthesis of glucose from certain non-carbohydrate carbon substrates. It is a ubiquitous process, present in plants, animals, fungi, bacteria, and other microorganisms. In vertebrates, gluconeogenesis occurs mainly in the liver and, to a lesser extent, in the cortex of the kidneys. It is one of two primary mechanisms – the other being degradation of glycogen (glycogenolysis) – used by humans and many other animals to maintain blood sugar levels, avoiding low levels (hypoglycemia). In ruminants, because dietary carbohydrates tend to be metabolized by rumen organisms, gluconeogenesis occurs regardless of fasting, low-carbohydrate diets, exercise, etc. In many other animals, the process occurs during periods of fasting, starvation, low-carbohydrate diets, or intense exercise.

The term amphibolic is used to describe a biochemical pathway that involves both catabolism and anabolism. Catabolism is a degradative phase of metabolism in which large molecules are converted into smaller and simpler molecules, which involves two types of reactions. First, hydrolysis reactions, in which catabolism is the breaking apart of molecules into smaller molecules to release energy. Examples of catabolic reactions are digestion and cellular respiration, where sugars and fats are broken down for energy. Breaking down a protein into amino acids, or a triglyceride into fatty acids, or a disaccharide into monosaccharides are all hydrolysis or catabolic reactions. Second, oxidation reactions involve the removal of hydrogens and electrons from an organic molecule. Anabolism is the biosynthesis phase of metabolism in which smaller simple precursors are converted to large and complex molecules of the cell. Anabolism has two classes of reactions. The first are dehydration synthesis reactions; these involve the joining of smaller molecules together to form larger, more complex molecules. These include the formation of carbohydrates, proteins, lipids and nucleic acids. The second are reduction reactions, in which hydrogens and electrons are added to a molecule. Whenever that is done, molecules gain energy.

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

Biological carbon fixation or сarbon assimilation is the process by which inorganic carbon is converted to organic compounds by living organisms. The compounds are then used to store energy and as structure for other biomolecules. Carbon is primarily fixed through photosynthesis, but some organisms use a process called chemosynthesis in the absence of sunlight.

<span class="mw-page-title-main">Oxaloacetic acid</span> Organic compound

Oxaloacetic acid (also known as oxalacetic acid or OAA) is a crystalline organic compound with the chemical formula HO2CC(O)CH2CO2H. Oxaloacetic acid, in the form of its conjugate base oxaloacetate, is a metabolic intermediate in many processes that occur in animals. It takes part in gluconeogenesis, the urea cycle, the glyoxylate cycle, amino acid synthesis, fatty acid synthesis and the citric acid cycle.

<span class="mw-page-title-main">Mitochondrial matrix</span> Space within the inner membrane of the mitochondrion

In the mitochondrion, the matrix is the space within the inner membrane. The word "matrix" stems from the fact that this space is viscous, compared to the relatively aqueous cytoplasm. The mitochondrial matrix contains the mitochondrial DNA, ribosomes, soluble enzymes, small organic molecules, nucleotide cofactors, and inorganic ions.[1] The enzymes in the matrix facilitate reactions responsible for the production of ATP, such as the citric acid cycle, oxidative phosphorylation, oxidation of pyruvate, and the beta oxidation of fatty acids.

Fatty acid metabolism consists of various metabolic processes involving or closely related to fatty acids, a family of molecules classified within the lipid macronutrient category. These processes can mainly be divided into (1) catabolic processes that generate energy and (2) anabolic processes where they serve as building blocks for other compounds.

<span class="mw-page-title-main">Mixed acid fermentation</span> Biochemical conversion of six-carbon sugars into acids in bacteria

In biochemistry, mixed acid fermentation is the metabolic process by which a six-carbon sugar is converted into a complex and variable mixture of acids. It is an anaerobic (non-oxygen-requiring) fermentation reaction that is common in bacteria. It is characteristic for members of the Enterobacteriaceae, a large family of Gram-negative bacteria that includes E. coli.

In biochemistry, fatty acid synthesis is the creation of fatty acids from acetyl-CoA and NADPH through the action of enzymes called fatty acid synthases. This process takes place in the cytoplasm of the cell. Most of the acetyl-CoA which is converted into fatty acids is derived from carbohydrates via the glycolytic pathway. The glycolytic pathway also provides the glycerol with which three fatty acids can combine to form triglycerides, the final product of the lipogenic process. When only two fatty acids combine with glycerol and the third alcohol group is phosphorylated with a group such as phosphatidylcholine, a phospholipid is formed. Phospholipids form the bulk of the lipid bilayers that make up cell membranes and surrounds the organelles within the cells. In addition to cytosolic fatty acid synthesis, there is also mitochondrial fatty acid synthesis (mtFASII), in which malonyl-CoA is formed from malonic acid with the help of malonyl-CoA synthetase (ACSF3), which then becomes the final product octanoyl-ACP (C8) via further intermediate steps.

<span class="mw-page-title-main">Reverse Krebs cycle</span>

The reverse Krebs cycle is a sequence of chemical reactions that are used by some bacteria to produce carbon compounds from carbon dioxide and water by the use of energy-rich reducing agents as electron donors.

<span class="mw-page-title-main">Isocitrate lyase</span>

Isocitrate lyase, or ICL, is an enzyme in the glyoxylate cycle that catalyzes the cleavage of isocitrate to succinate and glyoxylate. Together with malate synthase, it bypasses the two decarboxylation steps of the tricarboxylic acid cycle and is used by bacteria, fungi, and plants.

<span class="mw-page-title-main">Methylisocitrate lyase</span>

The enzyme methylisocitrate lyase catalyzes the chemical reaction

<span class="mw-page-title-main">Malate synthase</span> Class of enzymes

In enzymology, a malate synthase (EC 2.3.3.9) is an enzyme that catalyzes the chemical reaction

Glutaminolysis (glutamine + -lysis) is a series of biochemical reactions by which the amino acid glutamine is lysed to glutamate, aspartate, CO2, pyruvate, lactate, alanine and citrate.

Glyoxylate and dicarboxylate metabolism describes a variety of reactions involving glyoxylate or dicarboxylates. Glyoxylate is the conjugate base of glyoxylic acid, and within a buffered environment of known pH such as the cell cytoplasm these terms can be used almost interchangeably, as the gain or loss of a hydrogen ion is all that distinguishes them, and this can occur in the aqueous environment at any time. Likewise dicarboxylates are the conjugate bases of dicarboxylic acids, a general class of organic compounds containing two carboxylic acid groups, such as oxalic acid or succinic acid.

<span class="mw-page-title-main">Acetate kinase</span> Class of enzymes

In molecular biology, acetate kinase (EC 2.7.2.1), which is predominantly found in micro-organisms, facilitates the production of acetyl-CoA by phosphorylating acetate in the presence of ATP and a divalent cation. Short-chain fatty acids (SCFAs) play a major role in carbon cycle and can be utilized as a source of carbon and energy by bacteria. Salmonella typhimurium propionate kinase (StTdcD) catalyzes reversible transfer of the γ-phosphate of ATP to propionate during l-threonine degradation to propionate. Kinetic analysis revealed that StTdcD possesses broad ligand specificity and could be activated by various SCFAs (propionate>acetate≈butyrate), nucleotides (ATP≈GTP>CTP≈TTP; dATP>dGTP>dCTP) and metal ions (Mg2+≈Mn2+>Co2+). Inhibition of StTdcD by tricarboxylic acid (TCA) cycle intermediates such as citrate, succinate, α-ketoglutarate and malate suggests that the enzyme could be under plausible feedback regulation. Crystal structures of StTdcD bound to PO4 (phosphate), AMP, ATP, Ap4 (adenosine tetraphosphate), GMP, GDP, GTP, CMP and CTP revealed that binding of nucleotide mainly involves hydrophobic interactions with the base moiety and could account for the broad biochemical specificity observed between the enzyme and nucleotides. Modelling and site-directed mutagenesis studies suggest Ala88 to be an important residue involved in determining the rate of catalysis with SCFA substrates. Molecular dynamics simulations on monomeric and dimeric forms of StTdcD revealed plausible open and closed states, and also suggested role for dimerization in stabilizing segment 235-290 involved in interfacial interactions and ligand binding. Observation of an ethylene glycol molecule bound sufficiently close to the γ-phosphate in StTdcD complexes with triphosphate nucleotides supports direct in-line phosphoryl transfer. The enzyme is important in the process of glycolysis, enzyme levels being increased in the presence of excess glucose. The growth of a bacterial mutant lacking acetate kinase has been shown to be inhibited by glucose, suggesting that the enzyme is involved in excretion of excess carbohydrate. A related enzyme, butyrate kinase, facilitates the formation of butyryl-CoA by phosphorylating butyrate in the presence of ATP to form butyryl phosphate.

<span class="mw-page-title-main">Methylcitrate cycle</span>

The methylcitrate cycle, or the MCC, is the mechanism by which propionyl-CoA is formed, generated by β-oxidation of odd-chain fatty acids, and broken down to its final products, succinate and pyruvate. The methylcitrate cycle is closely related to both the citric acid cycle and the glyoxylate cycle, in that they share substrates, enzymes and products. The methylcitrate cycle functions overall to detoxify bacteria of toxic propionyl-CoA, and plays an essential role in propionate metabolism in bacteria. Incomplete propionyl-CoA metabolism may lead to the buildup of toxic metabolites in bacteria, and thus the function of the methylcitrate cycle is an important biological process.

References

  1. 1 2 3 4 5 Kondrashov FA, Koonin EV, Morgunov IG, Finogenova TV, Kondrashova MN (October 2006). "Evolution of glyoxylate cycle enzymes in Metazoa: evidence of multiple horizontal transfer events and pseudogene formation". Biology Direct. 1: 31. doi: 10.1186/1745-6150-1-31 . PMC   1630690 . PMID   17059607.
  2. 1 2 3 4 Lorenz MC, Fink GR (October 2002). "Life and death in a macrophage: role of the glyoxylate cycle in virulence". Eukaryotic Cell. 1 (5): 657–62. doi:10.1128/EC.1.5.657-662.2002. PMC   126751 . PMID   12455685.
  3. Popov, EA; Moskalev, EA; Shevchenko, MU; Eprintsev, AT (November 2005). "Comparative analysis of glyoxylate cycle key enzyme isocitrate lyase from organisms of different systematic groups". Journal of Evolutionary Biochemistry and Physiology. 41 (6): 631–639. doi:10.1007/s10893-006-0004-3. S2CID   30856607.
  4. Kornberg, H. L.; Krebs, H. A. (1957). "Synthesis of Cell Constituents from C 2 -Units by a Modified Tricarboxylic Acid Cycle". Nature. 179 (4568): 988–991. Bibcode:1957Natur.179..988K. doi:10.1038/179988a0. ISSN   1476-4687. PMID   13430766. S2CID   40858130.
  5. Berg JM, Tymoczko JL, Stryer L (2002). Biochemistry. New York: W. H. Freeman.
  6. Lorenz MC, Fink GR (July 2001). "The glyoxylate cycle is required for fungal virulence". Nature. 412 (6842): 83–6. Bibcode:2001Natur.412...83L. doi:10.1038/35083594. PMID   11452311. S2CID   4330954.
  7. Dunn MF, Ramírez-Trujillo JA, Hernández-Lucas I (October 2009). "Major roles of isocitrate lyase and malate synthase in bacterial and fungal pathogenesis". Microbiology. 155 (Pt 10): 3166–75. doi: 10.1099/mic.0.030858-0 . PMID   19684068.
  8. V. N. Popov; E. A. Moskalev; M. U. Shevchenko; A. T. Eprintsev (December 2005). "Comparative Analysis of Glyoxylate Cycle Key Enzyme Isocitrate Lyase from Organisms of Different Systematic Groups". Journal of Evolutionary Biochemistry and Physiology. 41 (6): 631–639. doi:10.1007/s10893-006-0004-3. S2CID   30856607.
  9. Davis WL, Goodman DB (December 1992). "Evidence for the glyoxylate cycle in human liver". The Anatomical Record. 234 (4): 461–8. doi:10.1002/ar.1092340402. PMID   1456449. S2CID   13181926.
  10. 1 2 Davis WL, Jones RG, Farmer GR, Dickerson T, Cortinas E, Cooper OJ, Crawford L, Goodman DB (July 1990). "Identification of glyoxylate cycle enzymes in chick liver--the effect of vitamin D3: cytochemistry and biochemistry". The Anatomical Record. 227 (3): 271–84. doi:10.1002/ar.1092270302. PMID   2164796. S2CID   39607621.
  11. Storrey, Kenneth, ed. (2004). Functional Metabolism: Regulation and Adaption. Hobocken, New Jersey: John Wiley and Sons, Inc. pp. 221–223. ISBN   978-0-471-41090-4.
  12. Davis WL, Goodman DB, Crawford LA, Cooper OJ, Matthews JL (March 1990). "Hibernation activates glyoxylate cycle and gluconeogenesis in black bear brown adipose tissue". Biochimica et Biophysica Acta (BBA) - Molecular Cell Research. 1051 (3): 276–8. doi:10.1016/0167-4889(90)90133-X. PMID   2310778.
  13. Jones JD, Burnett P, Zollman P (October 1999). "The glyoxylate cycle: does it function in the dormant or active bear?". Comparative Biochemistry and Physiology. Part B, Biochemistry & Molecular Biology. 124 (2): 177–9. doi:10.1016/S0305-0491(99)00109-1. PMID   10584301.
  14. Strittmatter L, Li Y, Nakatsuka NJ, Calvo SE, Grabarek Z, Mootha VK (May 2014). "CLYBL is a polymorphic human enzyme with malate synthase and β-methylmalate synthase activity". Human Molecular Genetics. 23 (9): 2313–23. doi:10.1093/hmg/ddt624. PMC   3976331 . PMID   24334609.
  15. Davis WL, Jones RG, Farmer GR, Cortinas E, Matthews JL, Goodman DB (1989). "The glyoxylate cycle in rat epiphyseal cartilage: the effect of vitamin-D3 on the activity of the enzymes isocitrate lyase and malate synthase". Bone. 10 (3): 201–6. doi:10.1016/8756-3282(89)90054-9. PMID   2553083.
  16. Cheah HL, Lim V, Sandai D (Apr 2014). "Inhibitors of the glyoxylate cycle enzyme ICL1 in Candida albicans for potential use as antifungal agents". PLOS ONE. 9 (4): e95951. Bibcode:2014PLoSO...995951C. doi: 10.1371/journal.pone.0095951 . PMC   4004578 . PMID   24781056.
  17. Bhusal RP, Bashiri G, Kwai BX, Sperry J, Leung IK (July 2017). "Targeting isocitrate lyase for the treatment of latent tuberculosis". Drug Discovery Today. 22 (7): 1008–1016. doi:10.1016/j.drudis.2017.04.012. PMID   28458043.
  18. Lee YV, Wahab HA, Choong YS (2015). "Potential inhibitors for isocitrate lyase of Mycobacterium tuberculosis and non-M. tuberculosis: a summary". BioMed Research International. 2015: 895453. doi: 10.1155/2015/895453 . PMC   4306415 . PMID   25649791.
  19. 1 2 3 Ward KA (March 2000). "Transgene-mediated modifications to animal biochemistry". Trends in Biotechnology. 18 (3): 99–102. doi:10.1016/S0167-7799(99)01417-1. PMID   10675896.
  20. Ward, Kevin; C. D. Nancarrow (1 Sep 1991). "The genetic engineering of production traits in domestic animals". Cellular and Molecular Life Sciences. 47 (9): 913–922. doi:10.1007/BF01929882. PMID   1915775. S2CID   41676957.