Acetyl-CoA carboxylase

Source: Wikipedia, the free encyclopedia.
acetyl-CoA carboxylase
ExPASy
NiceZyme view
KEGGKEGG entry
MetaCycmetabolic pathway
PRIAMprofile
PDB structuresRCSB PDB PDBe PDBsum
Gene OntologyAmiGO / QuickGO
Search
PMCarticles
PubMedarticles
NCBIproteins
Chr. 17 q21
Search for
StructuresSwiss-model
DomainsInterPro
Chr. 12 q24.1
Search for
StructuresSwiss-model
DomainsInterPro

Acetyl-CoA carboxylase (ACC) is a

covalent modification. The human genome contains the genes for two different ACCs[2]ACACA[3] and ACACB.[4]

Structure

eukaryotes have evolved an ACC with CT and BC catalytic domains and BCCP domains on a single polypeptide. Most plants also have this homomeric form in cytosol.[5] ACC functional regions, starting from the N-terminus to C-terminus are the biotin carboxylase (BC), biotin binding (BB), carboxyl transferase (CT), and ATP-binding (AB). AB lies within BC. Biotin is covalently attached through an amide bond to the long side chain
of a lysine reside in BB. As BB is between BC and CT regions, biotin can easily translocate to both of the active sites where it is required.

In mammals where two isoforms of ACC are expressed, the main structural difference between these isoforms is the extended ACC2 N-terminus containing a

Genes

The polypeptides composing the multi-subunit ACCs of

plants are encoded by distinct genes. In Escherichia coli, accA encodes the alpha subunit of the acetyl-CoA carboxylase,[6] and accD encodes its beta subunit.[7]

Mechanism

The overall reaction of ACAC(A,B) proceeds by a two-step mechanism.

malonyl CoA
in the second reaction, which is catalyzed by CT.

The reaction mechanism of ACAC(A,B).
The color scheme is as follows: enzyme, coenzymes, substrate names, metal ions, phosphate, and carbonate

In the

pKa of bicarbonate is 10.3. The enzyme apparently manipulates the pKa to facilitate the deprotonation of bicarbonate. The pKa of bicarbonate is decreased by its interaction with positively charged side chains of Arg338 and Arg292. Furthermore, Glu296 interacts with the side chain of Glu211, an interaction that has been shown to cause an increase in the apparent pKa. Following deprotonation of bicarbonate, the oxygen of the bicarbonate acts as a nucleophile and attacks the gamma phosphate on ATP. The carboxyphosphate intermediate quickly decomposes to CO2 and PO43−. The PO43− deprotonates biotin, creating an enolate, stabilized by Arg338, that subsequently attacks CO2 resulting in the production of carboxybiotin.[9] The carboxybiotin translocates to the carboxyl transferase (CT) active site, where the carboxyl group is transferred to acetyl-CoA. In contrast to the BC domain, little is known about the reaction mechanism of CT. A proposed mechanism is the release of CO2 from biotin, which subsequently abstracts a proton from the methyl group from acetyl CoA carboxylase. The resulting enolate attacks CO2 to form malonyl CoA. In a competing mechanism, proton
abstraction is concerted with the attack of acetyl CoA.

Function

The function of ACC is to regulate the metabolism of fatty acids. When the enzyme is active, the product, malonyl-CoA, is produced which is a building block for new fatty acids and can inhibit the transfer of the fatty acyl group from acyl CoA to

mitochondria
.

In

mammary glands, where fatty acid synthesis is important.[10] In oxidative tissues, such as the skeletal muscle and the heart, the ratio of ACC2 expressed is higher. ACC1 and ACC2 are both highly expressed in the liver where both fatty acid oxidation and synthesis are important.[11] The differences in tissue distribution indicate that ACC1 maintains regulation of fatty acid synthesis
whereas ACC2 mainly regulates fatty acid oxidation (beta oxidation).

A mitochondrial isoform of ACC1 (mACC1) plays a partially redundant role in lipoic acid synthesis and thus in protein lipoylation by providing malonyl-CoA for mitochondrial fatty acid synthesis (mtFASII) in tandem with ACSF3.[12][13]

Regulation

Control of Acetyl CoA Carboxylase. The AMP regulated kinase triggers the phosphorylation of the enzyme (thus inactivating it) and the phosphatase enzyme removes the phosphate group.

The regulation of mammalian ACC is complex, in order to control two distinct pools of malonyl CoA that direct either the inhibition of beta oxidation or the activation of lipid biosynthesis.[14]

Mammalian ACC1 and ACC2 are regulated transcriptionally by multiple

carbohydrates diets.[15][16]

Through a feed-forward loop,

dicarboxylic acids.[18] Long and short chain fatty acyl CoAs are negative feedback inhibitors of ACC.[19] One such negative allosteric modulator is palmitoyl-coA.[20]

Phosphorylation can result when the hormones

epinephrine[22] bind to cell surface receptors, but the main cause of phosphorylation is due to a rise in AMP levels when the energy status of the cell is low, leading to the activation of the AMP-activated protein kinase (AMPK). AMPK is the main kinase regulator of ACC, able to phosphorylate a number of serine residues on both isoforms of ACC.[23] On ACC1, AMPK phosphorylates Ser79, Ser1200, and Ser1215. Protein kinase A also has the ability to phosphorylate ACC, with a much greater ability to phosphorylate ACC2 than ACC1. Ser80 and Ser1263 on ACC1 may also serve as a site of phosphorylation as a regulatory mechanism.[24] However, the physiological significance of protein kinase A in the regulation of ACC is currently unknown. Researchers hypothesize there are other ACC kinases important to its regulation as there are many other possible phosphorylation sites on ACC.[25]

When

cellular membrane, it activates a phosphatase enzyme called protein phosphatase 2A (PP2A) to dephosphorylate the enzyme; thereby removing the inhibitory effect. Furthermore, insulin induces a phosphodiesterase that lowers the level of cAMP in the cell, thus inhibiting PKA, and also inhibits AMPK directly.[citation needed
]

This protein may use the morpheein model of allosteric regulation.[26]

Clinical implications

At the juncture of lipid synthesis and oxidation pathways, ACC presents many clinical possibilities for the production of novel

antibiotics and the development of new therapies for diabetes, obesity, and other manifestations of metabolic syndrome.[27] Researchers aim to take advantage of structural differences between bacterial and human ACCs to create antibiotics specific to the bacterial ACC, in efforts to minimize side effects to patients. Promising results for the usefulness of an ACC inhibitor include the finding that mice with no expression of ACC2 have continuous fatty acid oxidation, reduced body fat mass, and reduced body weight despite an increase in food consumption. These mice are also protected from diabetes.[14] A lack of ACC1 in mutant mice is lethal already at the embryonic stage. However, it is unknown whether drugs targeting ACCs in humans must be specific for ACC2.[28]

non-alcoholic steatohepatitis (NASH), believed to be an increasing cause of liver failure.[31]

In addition, plant-selective ACC inhibitors are in widespread use as herbicides,[32] which suggests clinical application against Apicomplexa parasites that rely on a plant-derived ACC isoform,[33] including malaria.

The heterogeneous clinical phenotypes of the metabolic disease combined malonic and methylmalonic aciduria (CMAMMA) due to ACSF3 deficiency are thought to result from partial compensation of a mitochondrial isoform of ACC1 (mACC1) for deficient ACSF3 in mitochondrial fatty acid synthesis (mtFASII).[34]

See also

References

Further reading