Amidophosphoribosyltransferase

From WikiProjectMed
Jump to navigation Jump to search
PPAT
Identifiers
AliasesPPAT, ATASE, GPAT, PRAT, phosphoribosyl pyrophosphate amidotransferase, Amidophosphoribosyltransferase
External IDsOMIM: 172450 MGI: 2387203 HomoloGene: 68272 GeneCards: PPAT
Orthologs
SpeciesHumanMouse
Entrez
Ensembl
UniProt
RefSeq (mRNA)

NM_002703

NM_172146

RefSeq (protein)

NP_002694

NP_742158

Location (UCSC)Chr 4: 56.39 – 56.44 MbChr 5: 77.06 – 77.1 Mb
PubMed search[3][4]
Wikidata
View/Edit HumanView/Edit Mouse

Amidophosphoribosyltransferase (ATase), also known as glutamine phosphoribosylpyrophosphate amidotransferase (GPAT), is an enzyme responsible for catalyzing the conversion of 5-phosphoribosyl-1-pyrophosphate (PRPP) into 5-phosphoribosyl-1-amine (PRA), using the amine group from a glutamine side-chain. This is the committing step in de novo purine synthesis. In humans it is encoded by the PPAT (phosphoribosyl pyrophosphate amidotransferase) gene.[5][6] ATase is a member of the purine/pyrimidine phosphoribosyltransferase family.

Structure and function

amidophosphoribosyltransferase
Identifiers
EC no.2.4.2.14
CAS no.9031-82-7
Databases
IntEnzIntEnz view
BRENDABRENDA entry
ExPASyNiceZyme view
KEGGKEGG entry
MetaCycmetabolic pathway
PRIAMprofile
PDB structuresRCSB PDB PDBe PDBsum
Gene OntologyAmiGO / QuickGO
Search
PMCarticles
PubMedarticles
NCBIproteins

The enzyme consists of two domains: a glutaminase domain that produces ammonia from glutamine by hydrolysis and a phosphoribosyltransferase domain that binds the ammonia to ribose-5-phosphate.[7] Coordination between the two active sites of enzyme give it special complexity.

The glutaminase domain is homologous to other N-terminal nucleophile (Ntn) hydrolases[7] such as carbamoyl phosphate synthetase (CPSase). Nine invariant residues among the sequences of all Ntn amidotransferases play key catalytic, substrate binding or structural roles. A terminal cysteine residue acts as the nucleophile in the first part of the reaction, analogous to the cysteine of a catalytic triad.[7][8] The free N terminus acts as a base to activate the nucleophile and protonate the leaving group in the hydrolytic reaction, in this case ammonia. Another key aspect of the catalytic site is an oxyanion hole which catalyzes the reaction intermediate, as shown in the mechanism below.[9]

The PRTase domain is homologous to many other PRTases involved in the purine nucleotide synthesis and salvage pathways. All PRTases involve the displacement of pyrophosphate in PRPP by a variety of nucleophiles.[10] ATase is the only PRTase that has ammonia as a nucleophile.[7] Pyrophosphate from PRPP is an excellent leaving group, so little chemical assistance is needed to promote catalysis. Rather, the primary function of the enzyme appears to be bringing the reactants together appropriately and preventing the wrong reaction, such as hydrolysis.[7]

Besides having their respective catalytic abilities, the two domains also coordinate with one another to ensure that all the ammonia produced from glutamine is transferred to PRPP and no other nucleophile than ammonia attacks PRPP. This is achieved mainly by blocking formation of ammonia until PRPP is bound and channelling the ammonia to the PRTase active site.[7]

Initial activation of the enzyme by PRPP is caused by a conformational change in a "glutamine loop", which repositions to be able to accept glutamine. This results in a 200-fold higher Km value for glutamine binding[11] Once glutamine has bound to the active site, further conformational changes bring the site into the enzyme, making it inaccessible.[7]

These conformational changes also result in the formation of a 20 Å long ammonia channel, one of the most striking features of this enzyme. This channel lacks any hydrogen bonding sites, to ensure easy diffusion of ammonia from one active site to the other. This channel ensures ammonia released from glutamine reaches the PRTase catalytic site, and it differs from the channel in CPSase[12] in that it is hydrophobic rather than polar, and transient rather than permanent.[7]

Reaction mechanism

Arrow pushing mechanism for the reaction catalyzed by ATase.
First half of the catalytic mechanism of ATase occurring in the glutaminase domain active site. The catalytic cysteine performs a nucleophilic attack on the substrate to form an acyl-enzyme intermediate, which is resolved by hydrolysis. Ammonia is produced in the third step, which is used in the second half of the mechanism.
Arrow pushing mechanism for the reaction catalyzed by ATase.
Second half of the catalytic mechanism of ATase occurring in the phosphoribosyltransferase domain active site. The ammonia liberated in the first half of the reaction replaces pyrophosphate in PRPP, yielding phosphoribosylamine. A tyrosine residue stabilizes the transition state and allows the reaction to occur.

The overall reaction catalyzed by ATase is the following:

PRPP + glutaminePRA + glutamate + PPi

Within the enzyme, the reaction is broken down into two half-reactions that occur at different active sites:

  1. glutamineNH
    3
    + glutamate
  2. PRPP + NH
    3
    PRA + PPi

The first part of the mechanism occurs in the active site of the glutaminase domain and releases an ammonia group from glutamine by hydrolysis. The ammonia released by the first reaction is then transferred to the active site of the phosphoribosyltransferase domain via a 20 Å channel, where it then binds to PRPP to form PRA.

Regulation

In an example of feedback inhibition, ATase is inhibited mainly by the end-products of the purine synthesis pathway, AMP, GMP, ADP, and GDP.[7] Each enzyme subunit from the homotetramer has two binding sites for these inhibitors. The allosteric (A) site overlaps with the site for the ribose-5-phosphate of PRPP, while the catalytic (C) site overlaps with the site for the pyrophosphate of PRPP.[7] The binding of specific nucleotide pairs to the two sites results in synergistic inhibition stronger than additive inhibition.[7][13][14] Inhibition occurs via a structural change in the enzyme where the flexible glutamine loop gets locked in an open position, preventing the binding of PRPP.[7]

Due to the chemical lability of PRA, which has a half-life of 38 seconds at pH 7.5 and 37 °C, researchers have suggested that the compound is channeled from Amidophosphoribosyltransferase to GAR synthetase in vivo.[15]

Interactive pathway map

Click on genes, proteins and metabolites below to link to respective articles.[§ 1]

[[File:
FluoropyrimidineActivity_WP1601go to articlego to articlego to articlego to pathway articlego to pathway articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to PubChem Compoundgo to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to pathway articlego to pathway articlego to articlego to articlego to articlego to articlego to articlego to WikiPathwaysgo to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to article
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
FluoropyrimidineActivity_WP1601go to articlego to articlego to articlego to pathway articlego to pathway articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to PubChem Compoundgo to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to pathway articlego to pathway articlego to articlego to articlego to articlego to articlego to articlego to WikiPathwaysgo to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to article
|alt=Fluorouracil (5-FU) Activity edit]]
Fluorouracil (5-FU) Activity edit
  1. ^ The interactive pathway map can be edited at WikiPathways: "FluoropyrimidineActivity_WP1601".

Gallery

References

  1. ^ a b c GRCh38: Ensembl release 89: ENSG00000128059 - Ensembl, May 2017
  2. ^ a b c GRCm38: Ensembl release 89: ENSMUSG00000029246 - Ensembl, May 2017
  3. ^ "Human PubMed Reference:". National Center for Biotechnology Information, U.S. National Library of Medicine.
  4. ^ "Mouse PubMed Reference:". National Center for Biotechnology Information, U.S. National Library of Medicine.
  5. ^ "Entrez Gene: phosphoribosyl pyrophosphate amidotransferase".
  6. ^ Brayton KA, Chen Z, Zhou G, Nagy PL, Gavalas A, Trent JM, Deaven LL, Dixon JE, Zalkin H (Feb 1994). "Two genes for de novo purine nucleotide synthesis on human chromosome 4 are closely linked and divergently transcribed". The Journal of Biological Chemistry. 269 (7): 5313–21. doi:10.1016/S0021-9258(17)37689-5. PMID 8106516.
  7. ^ a b c d e f g h i j k l Smith JL (Dec 1998). "Glutamine PRPP amidotransferase: snapshots of an enzyme in action". Current Opinion in Structural Biology. 8 (6): 686–94. doi:10.1016/s0959-440x(98)80087-0. PMID 9914248.
  8. ^ Smith JL, Zaluzec EJ, Wery JP, Niu L, Switzer RL, Zalkin H, Satow Y (Jun 1994). "Structure of the allosteric regulatory enzyme of purine biosynthesis". Science. 264 (5164): 1427–1433. Bibcode:1994Sci...264.1427S. doi:10.1126/science.8197456. PMID 8197456.
  9. ^ "Overview for MACiE Entry M0214". EMBL-EBI. Archived from the original on 2015-04-02. Retrieved 2015-03-10.
  10. ^ Musick WD (1981). "Structural features of the phosphoribosyltransferases and their relationship to the human deficiency disorders of purine and pyrimidine metabolism". CRC Critical Reviews in Biochemistry. 11 (1): 1–34. doi:10.3109/10409238109108698. PMID 7030616.
  11. ^ Kim JH, Krahn JM, Tomchick DR, Smith JL, Zalkin H (Jun 1996). "Structure and function of the glutamine phosphoribosylpyrophosphate amidotransferase glutamine site and communication with the phosphoribosylpyrophosphate site". The Journal of Biological Chemistry. 271 (26): 15549–15557. doi:10.1074/jbc.271.26.15549. PMID 8663035.
  12. ^ Thoden JB, Holden HM, Wesenberg G, Raushel FM, Rayment I (May 1997). "Structure of carbamoyl phosphate synthetase: a journey of 96 A from substrate to product". Biochemistry. 36 (21): 6305–6316. CiteSeerX 10.1.1.512.5333. doi:10.1021/bi970503q. PMID 9174345.
  13. ^ Chen S, Tomchick DR, Wolle D, Hu P, Smith JL, Switzer RL, Zalkin H (Sep 1997). "Mechanism of the synergistic end-product regulation of Bacillus subtilis glutamine phosphoribosylpyrophosphate amidotransferase by nucleotides". Biochemistry. 36 (35): 10718–10726. doi:10.1021/bi9711893. PMID 9271502.
  14. ^ Zhou G, Smith JL, Zalkin H (Mar 1994). "Binding of purine nucleotides to two regulatory sites results in synergistic feedback inhibition of glutamine 5-phosphoribosylpyrophosphate amidotransferase". The Journal of Biological Chemistry. 269 (9): 6784–6789. doi:10.1016/S0021-9258(17)37444-6. PMID 8120039.
  15. ^ Antle VD, Liu D, McKellar BR, Caperelli CA, Hua M, Vince R (1996). "Substrate specificity of glycinamide ribonucleotide synthetase from chicken liver". The Journal of Biological Chemistry. 271 (14): 8192–5. doi:10.1074/jbc.271.14.8192. PMID 8626510.

Further reading

External links

This article incorporates text from the United States National Library of Medicine, which is in the public domain.