Phenylalanine hydroxylase

Last updated

PAH
Phenylalanine hydroxylase.jpg
Available structures
PDB Ortholog search: PDBe RCSB
Identifiers
Aliases PAH , PH, PKU, PKU1, phenylalanine hydroxylase
External IDs OMIM: 612349 MGI: 97473 HomoloGene: 234 GeneCards: PAH
Orthologs
SpeciesHumanMouse
Entrez
Ensembl
UniProt
RefSeq (mRNA)

NM_000277
NM_001354304

NM_008777

RefSeq (protein)

NP_000268
NP_001341233

NP_032803

Location (UCSC) Chr 12: 102.84 – 102.96 Mb Chr 10: 87.36 – 87.42 Mb
PubMed search [3] [4]
Wikidata
View/Edit Human View/Edit Mouse

Phenylalanine hydroxylase (PAH) (EC 1.14.16.1) is an enzyme that catalyzes the hydroxylation of the aromatic side-chain of phenylalanine to generate tyrosine. PAH is one of three members of the biopterin-dependent aromatic amino acid hydroxylases, a class of monooxygenase that uses tetrahydrobiopterin (BH4, a pteridine cofactor) and a non-heme iron for catalysis. During the reaction, molecular oxygen is heterolytically cleaved with sequential incorporation of one oxygen atom into BH4 and phenylalanine substrate. [5] [6] In humans, mutations in its encoding gene, PAH , can lead to the metabolic disorder phenylketonuria.

Reaction catalyzed by PAH Overallreaction.png
Reaction catalyzed by PAH

Enzyme mechanism

The reaction is thought to proceed through the following steps:

  1. formation of a Fe(II)-O-O-BH4 bridge.
  2. heterolytic cleavage of the O-O bond to yield the ferryl oxo hydroxylating intermediate Fe(IV)=O
  3. attack on Fe(IV)=O to hydroxylate phenylalanine substrate to tyrosine. [7]
PAH mechanism, part I Mechanismpart1.png
PAH mechanism, part I

Formation and cleavage of the iron-peroxypterin bridge. Although evidence strongly supports Fe(IV)=O as the hydroxylating intermediate, [8] the mechanistic details underlying the formation of the Fe(II)-O-O-BH4 bridge prior to heterolytic cleavage remain controversial. Two pathways have been proposed based on models that differ in the proximity of the iron to the pterin cofactor and the number of water molecules assumed to be iron-coordinated during catalysis. According to one model, an iron dioxygen complex is initially formed and stabilized as a resonance hybrid of Fe2+O2 and Fe3+O2. The activated O2 then attacks BH4, forming a transition state characterized by charge separation between the electron-deficient pterin ring and the electron-rich dioxygen species. [9] The Fe(II)-O-O-BH4 bridge is subsequently formed. On the other hand, formation of this bridge has been modeled assuming that BH4 is located in iron's first coordination shell and that the iron is not coordinated to any water molecules. This model predicts a different mechanism involving a pterin radical and superoxide as critical intermediates. [10] Once formed, the Fe(II)-O-O-BH4 bridge is broken through heterolytic cleavage of the O-O bond to Fe(IV)=O and 4a-hydroxytetrahydrobiopterin; thus, molecular oxygen is the source of both oxygen atoms used to hydroxylate the pterin ring and phenylalanine.

PAH mechanism, part II Mechanismpart2revised.png
PAH mechanism, part II

Hydroxylation of phenylalanine by ferryl oxo intermediate. Because the mechanism involves a Fe(IV)=O (as opposed to a peroxypterin) hydroxylating intermediate, oxidation of the BH4 cofactor and hydroxylation of phenylalanine can be decoupled, resulting in unproductive consumption of BH4 and formation of H2O2. [7] When productive, though, the Fe(IV)=O intermediate is added to phenylalanine in an electrophilic aromatic substitution reaction that reduces iron from the ferryl to the ferrous state. [7] Although initially an arene oxide or radical intermediate was proposed, analyses of the related tryptophan and tyrosine hydroxylases have suggested that the reaction instead proceeds through a cationic intermediate that requires Fe(IV)=O to be coordinated to a water ligand rather than a hydroxo group. [7] [11] This cationic intermediate subsequently undergoes a 1,2-hydride NIH shift, yielding a dienone intermediate that then tautomerizes to form the tyrosine product. [12] The pterin cofactor is regenerated by hydration of the carbinolamine product of PAH to quinonoid dihydrobiopterin (qBH2), which is then reduced to BH4. [13]

Enzyme regulation

PAH is proposed to use the morpheein model of allosteric regulation. [14] [15]

Mammalian PAH exists in an equilibrium consisting of tetramers of two distinct architectures, with one or more dimeric forms as part of the equilibrium. This behavior is consistent with a dissociative allosteric mechanism. [15]

Many studies suggest that mammalian PAH shows behavior comparable to porphobilinogen synthase (PBGS), wherein a variety of factors such as pH and ligand binding are reported to affect enzyme activity and protein stability. [15]

Structure

The PAH monomer (51.9 kDa) consists of three distinct domains: a regulatory N-terminal domain (residues 1–117) that contains a Phe-binding ACT subdomain, the catalytic domain (residues 118–427), and a C-terminal domain (residues 428–453) responsible for oligomerization of identical monomers. Extensive crystallographic analysis has been performed, especially on the pterin- and iron-coordinated catalytic domain to examine the active site. The structure of the N-terminal regulatory domain has also been determined, and together with the solved structure of the homologous tyrosine hydroxylase C-terminal tetramerization domain, a structural model of tetrameric PAH was proposed. [13] Using X-ray crystallography, the structure of full-length rat PAH was determined experimentally and showed the auto-inhibited, or resting-state form of the enzyme. [16] The resting-state form (RS-PAH) is architecturally distinct from the activated form (A-PAH). [17] A full-length structure of A-PAH is currently lacking, but the Phe stabilized ACT-ACT interface that is characteristic of A-PAH has been determined and a structural model of A-PAH based on SAXS analysis has been proposed. [18] [19]

Model of the active site of PAH bound to BH4, ferrous, and a phenylalanine analogue. (from PDB 1KW0) Phenylalanine analogue, BH4, iron, Fe(II)-coordinated His and Glu residues PheOHactivesite.png
Model of the active site of PAH bound to BH4, ferrous, and a phenylalanine analogue. (from PDB 1KW0) Phenylalanine analogue, BH4, iron, Fe(II)-coordinated His and Glu residues

Catalytic domain

Solved crystal structures of the catalytic domain indicate that the active site consists of an open and spacious pocket lined primarily by hydrophobic residues, though three glutamic acid residues, two histidines, and a tyrosine are also present and iron-binding. [13] Contradictory evidence exists about the coordination state of the ferrous atom and its proximity to BH4 within the active site. According to crystallographic analysis, Fe(II) is coordinated by water, His285, His290, and Glu330 (a 2-his-1-carboxylate facial triad arrangement) with octahedral geometry. [20] Inclusion of a Phe analogue in the crystal structure changes both iron from a six- to a five-coordinated state involving a single water molecule and bidentate coordination to Glu330 and opening a site for oxygen to bind. BH4 is concomitantly shifted toward the iron atom, although the pterin cofactor remains in the second coordination sphere. [21] On the other hand, a competing model based on NMR and molecular modeling analyses suggests that all coordinated water molecules are forced out of the active site during the catalytic cycle while BH4 becomes directly coordinated to iron. [22] As discussed above, resolving this discrepancy will be important for determining the exact mechanism of PAH catalysis.

N-terminal regulatory domain

The regulatory nature of the N-terminal domain (residues 1–117) is conferred by its structural flexibility. [23] Hydrogen/deuterium exchanges analysis indicates that allosteric binding of Phe globally alters the conformation of PAH such that the active site is less occluded as the interface between the regulatory and catalytic domains is increasingly exposed to solvent. [23] [24] [25] This observation is consistent with kinetic studies, which show an initially low rate of tyrosine formation for full-length PAH. This lag time is not observed, however, for a truncated PAH lacking the N-terminal domain or if the full-length enzyme is pre-incubated with Phe. Deletion of the N-terminal domain also eliminates the lag time while increasing the affinity for Phe by nearly two-fold; no difference is observed in the Vmax or Km for the tetrahydrobiopterin cofactor. [26] Additional regulation is provided by Ser16; phosphorylation of this residue does not alter enzyme conformation but does reduce the concentration of Phe required for allosteric activation. [25] This N-terminal regulatory domain is not observed in bacterial PAHs but shows considerable structural homology to the regulatory domain of phosphogylcerate dehydrogenase, an enzyme in the serine biosynthetic pathway. [25]

Tetramerization domain

Prokaryotic PAH is monomeric, whereas eukaryotic PAH exists in an equilibrium between homotetrameric and homodimeric forms. [7] [13] The dimerization interface is composed of symmetry-related loops that link identical monomers, while the overlapping C-terminal tetramerization domain mediates the association of conformationally distinct dimers that are characterized by a different relative orientation of the catalytic and tetramerization domains (Flatmark, Erlandsen). The resulting distortion of the tetramer symmetry is evident in the differential surface area of the dimerization interfaces and distinguishes PAH from the tetramerically symmetrical tyrosine hydroxylase. [13] A domain-swapping mechanism has been proposed to mediate formation of the tetramer from dimers, in which C-terminal alpha-helixes mutually alter their conformation around a flexible C-terminal five-residue hinge region to form a coiled-coil structure, shifting equilibrium toward the tetrameric form. [7] [13] [27] Although both the homodimeric and homotetrameric forms of PAH are catalytically active, the two exhibit differential kinetics and regulation. In addition to reduced catalytic efficiency, the dimer does not display positive cooperativity toward L-Phe (which at high concentrations activates the enzyme), suggesting that L-Phe allosterically regulates PAH by influencing dimer-dimer interaction. [27]

Biological function

PAH is a critical enzyme in phenylalanine metabolism and catalyzes the rate-limiting step in its complete catabolism to carbon dioxide and water. [13] [28] Regulation of flux through phenylalanine-associated pathways is critical in mammalian metabolism, as evidenced by the toxicity of high plasma levels of this amino acid observed in phenylketonuria (see below). The principal source of phenylalanine is ingested proteins, but relatively little of this pool is used for protein synthesis. [28] Instead, the majority of ingested phenylalanine is catabolized through PAH to form tyrosine; addition of the hydroxyl group allows for the benzene ring to be broken in subsequent catabolic steps. Transamination to phenylpyruvate, whose metabolites are excreted in the urine, represents another pathway of phenylalanine turnover, but catabolism through PAH predominates. [28]

In humans, this enzyme is expressed both in the liver and the kidney, and there is some indication that it may be differentially regulated in these tissues. [29] PAH is unusual among the aromatic amino acid hydroxylases for its involvement in catabolism; tyrosine and tryptophan hydroxylases, on the other hand, are primarily expressed in the central nervous system and catalyze rate-limiting steps in neurotransmitter/hormone biosynthesis. [13]

Disease relevance

Deficiency in PAH activity due to mutations in PAH causes hyperphenylalaninemia (HPA), and when blood phenylalanine levels increase above 20 times the normal concentration, the metabolic disease phenylketonuria (PKU) results. [28] PKU is both genotypically and phenotypically heterogeneous: Over 300 distinct pathogenic variants have been identified, the majority of which correspond to missense mutations that map to the catalytic domain. [13] [20] When a cohort of identified PAH mutants were expressed in recombinant systems, the enzymes displayed altered kinetic behavior and/or reduced stability, consistent with structural mapping of these mutations to both the catalytic and tetramerization domains of the enzyme. [13] BH44 has been administered as a pharmacological treatment and has been shown to reduce blood levels of phenylalanine for a segment of PKU patients whose genotypes lead to some residual PAH activity but have no defect in BH44 synthesis or regeneration. Follow-up studies suggest that in the case of certain PAH mutants, excess BH44 acts as a pharmacological chaperone to stabilize mutant enzymes with disrupted tetramer assembly and increased sensitivity to proteolytic cleavage and aggregation. [30] Mutations that have been identified in the PAH locus are documented at the Phenylalanine Hydroxylase Locus Knowledgbase (PAHdb, https://web.archive.org/web/20130718162051/http://www.pahdb.mcgill.ca/).

Since phenylketonuria can cause irreversible damage, it is imperative that deficiencies in the phenylalanine hydroxylase are determined early on in development. Originally, this was done using a bacterial inhibition assay known as the Guthrie Test. Now, PKU is part of newborn screening in many countries, and elevated phenylalanine levels are identified shortly after birth by measurement with tandem mass spectrometry. Placing the individual on a low phenylalanine, high tyrosine diet can help prevent any long-term damage to their development.

PAH Knock-out Model

The first attempt at creating a Pah-KO mouse model was reported in a research article published in 2021. [31] This knockout mouse was created to be homozygous through its development within the C57BL/6 J strain using CRISPR/Cas9. [32] Codon 7, GAG, in the Pah gene was altered to the stop codon TAG, depicting an intentional point mutation. Two to six-month-old, male, homozygous mice were studied by scientific methods such as behavioral and biochemical assays, MRI, and histopathology. The homozygous mice were routinely compared to control mice, age and sex-matched heterozygous Pah-KO mice that expressed sufficient PAH enzyme activity to supply phenylalanine and tyrosine levels similar to wild type mice.

The Pah-KO mouse model presented high blood phenylalanine and low tyrosine levels, hypocholesterolemia, high phenylalanine and low levels of neurotransmitters and tyrosine in the brain, hypomyelination, low brain and body weight, high grade of ocular pathology, hypopigmentation, and a progressive behavioral deficit compared to heterozygous mice. [31] After termination of the homozygous mice, no hepatic PAH proteins were detected by western blot analysis. [33] The increased amount of phenylalanine in whole brain homogenates of homozygous mice was similar to PKU patients. The hair coat color of both mice was dependent on the amount of melanin pigment in hair shafts; thus, homozygous mice were prone to be a lighter brown based on less melanin present. Mice behavior was tested with a nesting building assay. [31]

This model sets up the potential for an extensive investigation of PKU biology resembling PKU patients and for the evaluation of modern therapeutic strategies for PKU treatment.

Phenylalanine hydroxylase is closely related to two other enzymes:

The three enzymes are homologous, that is, are thought to have evolved from the same ancient hydroxylase.

Related Research Articles

<span class="mw-page-title-main">Phenylketonuria</span> Amino acid metabolic disorder

Phenylketonuria (PKU) is an inborn error of metabolism that results in decreased metabolism of the amino acid phenylalanine. Untreated PKU can lead to intellectual disability, seizures, behavioral problems, and mental disorders. It may also result in a musty smell and lighter skin. A baby born to a mother who has poorly treated PKU may have heart problems, a small head, and low birth weight.

<span class="mw-page-title-main">Tyrosine</span> Amino acid

L-Tyrosine or tyrosine or 4-hydroxyphenylalanine is one of the 20 standard amino acids that are used by cells to synthesize proteins. It is a non-essential amino acid with a polar side group. The word "tyrosine" is from the Greek tyrós, meaning cheese, as it was first discovered in 1846 by German chemist Justus von Liebig in the protein casein from cheese. It is called tyrosyl when referred to as a functional group or side chain. While tyrosine is generally classified as a hydrophobic amino acid, it is more hydrophilic than phenylalanine. It is encoded by the codons UAC and UAU in messenger RNA.

<span class="mw-page-title-main">Phenylalanine</span> Type of α-amino acid

Phenylalanine is an essential α-amino acid with the formula C
9
H
11
NO
2
. It can be viewed as a benzyl group substituted for the methyl group of alanine, or a phenyl group in place of a terminal hydrogen of alanine. This essential amino acid is classified as neutral, and nonpolar because of the inert and hydrophobic nature of the benzyl side chain. The L-isomer is used to biochemically form proteins coded for by DNA. Phenylalanine is a precursor for tyrosine, the monoamine neurotransmitters dopamine, norepinephrine (noradrenaline), and epinephrine (adrenaline), and the biological pigment melanin. It is encoded by the codons UUU and UUC.

In chemistry, hydroxylation can refer to:

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

Tetrahydrobiopterin (BH4, THB), also known as sapropterin (INN), is a cofactor of the three aromatic amino acid hydroxylase enzymes, used in the degradation of amino acid phenylalanine and in the biosynthesis of the neurotransmitters serotonin (5-hydroxytryptamine, 5-HT), melatonin, dopamine, norepinephrine (noradrenaline), epinephrine (adrenaline), and is a cofactor for the production of nitric oxide (NO) by the nitric oxide synthases. Chemically, its structure is that of a (dihydropteridine reductase) reduced pteridine derivative (quinonoid dihydrobiopterin).

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

Pterin is a heterocyclic compound composed of a pteridine ring system, with a "keto group" and an amino group on positions 4 and 2 respectively. It is structurally related to the parent bicyclic heterocycle called pteridine. Pterins, as a group, are compounds related to pterin with additional substituents. Pterin itself is of no biological significance.

<span class="mw-page-title-main">Tetrahydrobiopterin deficiency</span> Medical condition

Tetrahydrobiopterin deficiency (THBD, BH4D) is a rare metabolic disorder that increases the blood levels of phenylalanine. Phenylalanine is an amino acid obtained normally through the diet, but can be harmful if excess levels build up, causing intellectual disability and other serious health problems. In healthy individuals, it is metabolised (hydroxylated) into tyrosine, another amino acid, by phenylalanine hydroxylase. However, this enzyme requires tetrahydrobiopterin as a cofactor and thus its deficiency slows phenylalanine metabolism.

<span class="mw-page-title-main">6-Pyruvoyltetrahydropterin synthase deficiency</span> Medical condition

6-Pyruvoyltetrahydropterin synthase deficiency is an autosomal recessive disorder that causes malignant hyperphenylalaninemia due to tetrahydrobiopterin deficiency. It is a recessive disorder that is accompanied by hyperphenylalaninemia. Commonly reported symptoms are initial truncal hypotonia, subsequent appendicular hypertonia, bradykinesia, cogwheel rigidity, generalized dystonia, and marked diurnal fluctuation. Other reported clinical features include difficulty in swallowing, oculogyric crises, somnolence, irritability, hyperthermia, and seizures. Chorea, athetosis, hypersalivation, rash with eczema, and sudden death have also been reported. Patients with mild phenotypes may deteriorate if given folate antagonists such as methotrexate, which can interfere with a salvage pathway through which dihydrobiopterin is converted into tetrahydrobiopterin via dihydrofolate reductase. Treatment options include substitution with neurotransmitter precursors, monoamine oxidase inhibitors, and tetrahydrobiopterin. Response to treatment is variable and the long-term and functional outcome is unknown. To provide a basis for improving the understanding of the epidemiology, genotype–phenotype correlation and outcome of these diseases, their impact on the quality of life of patients, and for evaluating diagnostic and therapeutic strategies a patient registry was established by the noncommercial International Working Group on Neurotransmitter Related Disorders (iNTD).

<span class="mw-page-title-main">GTP cyclohydrolase I</span>

GTP cyclohydrolase I (GTPCH) (EC 3.5.4.16) is a member of the GTP cyclohydrolase family of enzymes. GTPCH is part of the folate and biopterin biosynthesis pathways. It is responsible for the hydrolysis of guanosine triphosphate (GTP) to form 7,8-dihydroneopterin triphosphate (7,8-DHNP-3'-TP, 7,8-NH2-3'-TP).

<span class="mw-page-title-main">Tyrosine hydroxylase</span> Enzyme found in Homo sapiens that converts l-tyrosine to l-dopa, the precursor of cathecolamines

Tyrosine hydroxylase or tyrosine 3-monooxygenase is the enzyme responsible for catalyzing the conversion of the amino acid L-tyrosine to L-3,4-dihydroxyphenylalanine (L-DOPA). It does so using molecular oxygen (O2), as well as iron (Fe2+) and tetrahydrobiopterin as cofactors. L-DOPA is a precursor for dopamine, which, in turn, is a precursor for the important neurotransmitters norepinephrine (noradrenaline) and epinephrine (adrenaline). Tyrosine hydroxylase catalyzes the rate limiting step in this synthesis of catecholamines. In humans, tyrosine hydroxylase is encoded by the TH gene, and the enzyme is present in the central nervous system (CNS), peripheral sympathetic neurons and the adrenal medulla. Tyrosine hydroxylase, phenylalanine hydroxylase and tryptophan hydroxylase together make up the family of aromatic amino acid hydroxylases (AAAHs).

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

Queuine (Q) is a hypermodified nucleobase found in the first position of the anticodon of tRNAs specific for Asn, Asp, His, and Tyr, in most eukaryotes and prokaryotes. Because it is utilized by all eukaryotes but produced exclusively by bacteria, it is a putative vitamin.

<span class="mw-page-title-main">Tryptophan hydroxylase</span> Class of enzymes

Tryptophan hydroxylase (TPH) is an enzyme (EC 1.14.16.4) involved in the synthesis of the monoamine neurotransmitter serotonin. Tyrosine hydroxylase, phenylalanine hydroxylase, and tryptophan hydroxylase together constitute the family of biopterin-dependent aromatic amino acid hydroxylases. TPH catalyzes the following chemical reaction

<span class="mw-page-title-main">Aromatic amino acid</span> Amino acid having an aromatic ring

An aromatic amino acid is an amino acid that includes an aromatic ring.

<span class="mw-page-title-main">Phenylalanine racemase (ATP-hydrolysing)</span>

The enzyme phenylalanine racemase is the enzyme that acts on amino acids and derivatives. It activates both the L & D stereo isomers of phenylalanine to form L-phenylalanyl adenylate and D-phenylalanyl adenylate, which are bound to the enzyme. These bound compounds are then transferred to the thiol group of the enzyme followed by conversion of its configuration, the D-isomer being the more favorable configuration of the two, with a 7 to 3 ratio between the two isomers. The racemisation reaction of phenylalanine is coupled with the highly favorable hydrolysis of adenosine triphosphate (ATP) to adenosine monophosphate (AMP) and pyrophosphate (PP), thermodynamically allowing it to proceed. This reaction is then drawn forward by further hydrolyzing PP to inorganic phosphate (Pi), via Le Chatelier's principle.

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

Biopterins are pterin derivatives which function as endogenous enzyme cofactors in many species of animals and in some bacteria and fungi. The prototypical compound of the class is biopterin, as shown in the infobox. Biopterins act as cofactors for aromatic amino acid hydroxylases (AAAH), which are involved in synthesizing a number of neurotransmitters including dopamine, norepinephrine, epinepherine, and serotonin, along with several trace amines. Nitric oxide synthesis also uses biopterin derivatives as cofactors. In humans, tetrahydrobiopterin (BH4) is the endogenous cofactor for AAAH enzymes.

<span class="mw-page-title-main">Maleylacetoacetate isomerase</span> Class of enzymes

In enzymology, maleylacetoacetate isomerase is an enzyme that catalyzes the chemical reaction

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

The enzyme phenylalanine ammonia lyase (EC 4.3.1.24) catalyzes the conversion of L-phenylalanine to ammonia and trans-cinnamic acid.:

<span class="mw-page-title-main">PCBD1</span> Protein-coding gene in the species Homo sapiens

Pterin-4-alpha-carbinolamine dehydratase is an enzyme that in humans is encoded by the PCBD1 gene.

<span class="mw-page-title-main">Hyperphenylalaninemia</span> Medical condition

Hyperphenylalaninemia is a medical condition characterized by mildly or strongly elevated concentrations of the amino acid phenylalanine in the blood. Phenylketonuria (PKU) can result in severe hyperphenylalaninemia. Phenylalanine concentrations are routinely screened in newborns by the neonatal heel prick, which takes a few drops of blood from the heel of the infant. Standard phenylalanine concentrations in unaffected persons are about 2-6mg/dl phenylalanine concentrations in those with untreated hyperphenylalaninemia can be up to 20 mg/dL. Measurable IQ deficits are often detected as phenylalanine levels approach 10 mg/dL. Phenylketonuria (PKU)-like symptoms, including more pronounced developmental defects, skin irritation, and vomiting, may appear when phenylalanine levels are near 20 mg/dL .Hyperphenylalaninemia is a recessive hereditary metabolic disorder that is caused by the body's failure to convert phenylalanine to tyrosine as a result of the entire or partial absence of the enzyme phenylalanine hydroxylase.

<span class="mw-page-title-main">Biopterin-dependent aromatic amino acid hydroxylase</span>

Biopterin-dependent aromatic amino acid hydroxylases (AAAH) are a family of aromatic amino acid hydroxylase enzymes which includes phenylalanine 4-hydroxylase, tyrosine 3-hydroxylase, and tryptophan 5-hydroxylase. These enzymes primarily hydroxylate the amino acids L-phenylalanine, L-tyrosine, and L-tryptophan, respectively.

References

  1. 1 2 3 GRCh38: Ensembl release 89: ENSG00000171759 - Ensembl, May 2017
  2. 1 2 3 GRCm38: Ensembl release 89: ENSMUSG00000020051 - 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. Fitzpatrick PF (1999). "Tetrahydropterin-dependent amino acid hydroxylases". Annual Review of Biochemistry. 68: 355–81. doi:10.1146/annurev.biochem.68.1.355. PMID   10872454.
  6. Kaufman S (February 1958). "A new cofactor required for the enzymatic conversion of phenylalanine to tyrosine". The Journal of Biological Chemistry. 230 (2): 931–9. doi: 10.1016/S0021-9258(18)70516-4 . PMID   13525410.
  7. 1 2 3 4 5 6 Fitzpatrick PF (December 2003). "Mechanism of aromatic amino acid hydroxylation". Biochemistry. 42 (48): 14083–91. doi:10.1021/bi035656u. PMC   1635487 . PMID   14640675.
  8. Panay AJ, Lee M, Krebs C, Bollinger JM, Fitzpatrick PF (March 2011). "Evidence for a high-spin Fe(IV) species in the catalytic cycle of a bacterial phenylalanine hydroxylase". Biochemistry. 50 (11): 1928–33. doi:10.1021/bi1019868. PMC   3059337 . PMID   21261288.
  9. Bassan A, Blomberg MR, Siegbahn PE (January 2003). "Mechanism of dioxygen cleavage in tetrahydrobiopterin-dependent amino acid hydroxylases". Chemistry: A European Journal. 9 (1): 106–15. doi: 10.1002/chem.200390006 . PMID   12506369.
  10. Olsson E, Martinez A, Teigen K, Jensen VR (March 2011). "Formation of the iron-oxo hydroxylating species in the catalytic cycle of aromatic amino acid hydroxylases". Chemistry: A European Journal. 17 (13): 3746–58. doi:10.1002/chem.201002910. PMID   21351297.
  11. Bassan A, Blomberg MR, Siegbahn PE (September 2003). "Mechanism of aromatic hydroxylation by an activated FeIV=O core in tetrahydrobiopterin-dependent hydroxylases". Chemistry: A European Journal. 9 (17): 4055–67. doi:10.1002/chem.200304768. PMID   12953191.
  12. Pavon JA, Fitzpatrick PF (September 2006). "Insights into the catalytic mechanisms of phenylalanine and tryptophan hydroxylase from kinetic isotope effects on aromatic hydroxylation". Biochemistry. 45 (36): 11030–7. doi:10.1021/bi0607554. PMC   1945167 . PMID   16953590.
  13. 1 2 3 4 5 6 7 8 9 10 Flatmark T, Stevens RC (August 1999). "Structural Insight into the Aromatic Amino Acid Hydroxylases and Their Disease-Related Mutant Forms". Chemical Reviews. 99 (8): 2137–2160. doi:10.1021/cr980450y. PMID   11849022.
  14. Selwood T, Jaffe EK (March 2012). "Dynamic dissociating homo-oligomers and the control of protein function". Archives of Biochemistry and Biophysics. 519 (2): 131–43. doi:10.1016/j.abb.2011.11.020. PMC   3298769 . PMID   22182754.
  15. 1 2 3 Jaffe EK, Stith L, Lawrence SH, Andrake M, Dunbrack RL (February 2013). "A new model for allosteric regulation of phenylalanine hydroxylase: implications for disease and therapeutics". Archives of Biochemistry and Biophysics. 530 (2): 73–82. doi:10.1016/j.abb.2012.12.017. PMC   3580015 . PMID   23296088.
  16. Arturo EC, Gupta K, Héroux A, Stith L, Cross PJ, Parker EJ, Loll PJ, Jaffe EK (March 2016). "First structure of full-length mammalian phenylalanine hydroxylase reveals the architecture of an autoinhibited tetramer". Proceedings of the National Academy of Sciences of the United States of America. 113 (9): 2394–9. Bibcode:2016PNAS..113.2394A. doi: 10.1073/pnas.1516967113 . PMC   4780608 . PMID   26884182.
  17. Jaffe EK (August 2017). "New protein structures provide an updated understanding of phenylketonuria". Molecular Genetics and Metabolism. 121 (4): 289–296. doi:10.1016/j.ymgme.2017.06.005. PMC   5549558 . PMID   28645531.
  18. Patel D, Kopec J, Fitzpatrick F, McCorvie TJ, Yue WW (April 2016). "Structural basis for ligand-dependent dimerization of phenylalanine hydroxylase regulatory domain". Scientific Reports. 6 (1): 23748. Bibcode:2016NatSR...623748P. doi:10.1038/srep23748. PMC   4822156 . PMID   27049649.
  19. Meisburger SP, Taylor AB, Khan CA, Zhang S, Fitzpatrick PF, Ando N (May 2016). "Domain Movements upon Activation of Phenylalanine Hydroxylase Characterized by Crystallography and Chromatography-Coupled Small-Angle X-ray Scattering". Journal of the American Chemical Society. 138 (20): 6506–16. doi:10.1021/jacs.6b01563. PMC   4896396 . PMID   27145334.
  20. 1 2 Erlandsen H, Fusetti F, Martinez A, Hough E, Flatmark T, Stevens RC (December 1997). "Crystal structure of the catalytic domain of human phenylalanine hydroxylase reveals the structural basis for phenylketonuria". Nature Structural Biology. 4 (12): 995–1000. doi:10.1038/nsb1297-995. PMID   9406548. S2CID   6293946.
  21. Andersen OA, Flatmark T, Hough E (July 2002). "Crystal structure of the ternary complex of the catalytic domain of human phenylalanine hydroxylase with tetrahydrobiopterin and 3-(2-thienyl)-L-alanine, and its implications for the mechanism of catalysis and substrate activation". Journal of Molecular Biology. 320 (5): 1095–108. doi:10.1016/S0022-2836(02)00560-0. PMID   12126628.
  22. Teigen K, Frøystein NA, Martínez A (December 1999). "The structural basis of the recognition of phenylalanine and pterin cofactors by phenylalanine hydroxylase: implications for the catalytic mechanism". Journal of Molecular Biology. 294 (3): 807–23. doi:10.1006/jmbi.1999.3288. PMID   10610798.
  23. 1 2 Li J, Dangott LJ, Fitzpatrick PF (April 2010). "Regulation of phenylalanine hydroxylase: conformational changes upon phenylalanine binding detected by hydrogen/deuterium exchange and mass spectrometry". Biochemistry. 49 (15): 3327–35. doi:10.1021/bi1001294. PMC   2855537 . PMID   20307070.
  24. Li J, Ilangovan U, Daubner SC, Hinck AP, Fitzpatrick PF (January 2011). "Direct evidence for a phenylalanine site in the regulatory domain of phenylalanine hydroxylase". Archives of Biochemistry and Biophysics. 505 (2): 250–5. doi:10.1016/j.abb.2010.10.009. PMC   3019263 . PMID   20951114.
  25. 1 2 3 Kobe B, Jennings IG, House CM, Michell BJ, Goodwill KE, Santarsiero BD, Stevens RC, Cotton RG, Kemp BE (May 1999). "Structural basis of autoregulation of phenylalanine hydroxylase". Nature Structural Biology. 6 (5): 442–8. doi:10.1038/8247. PMID   10331871. S2CID   11709986.
  26. Daubner SC, Hillas PJ, Fitzpatrick PF (December 1997). "Expression and characterization of the catalytic domain of human phenylalanine hydroxylase". Archives of Biochemistry and Biophysics. 348 (2): 295–302. doi:10.1006/abbi.1997.0435. PMID   9434741.
  27. 1 2 Bjørgo E, de Carvalho RM, Flatmark T (February 2001). "A comparison of kinetic and regulatory properties of the tetrameric and dimeric forms of wild-type and Thr427-->Pro mutant human phenylalanine hydroxylase: contribution of the flexible hinge region Asp425-Gln429 to the tetramerization and cooperative substrate binding". European Journal of Biochemistry. 268 (4): 997–1005. doi:10.1046/j.1432-1327.2001.01958.x. PMID   11179966.
  28. 1 2 3 4 Kaufman S (March 1999). "A model of human phenylalanine metabolism in normal subjects and in phenylketonuric patients". Proceedings of the National Academy of Sciences of the United States of America. 96 (6): 3160–4. Bibcode:1999PNAS...96.3160K. doi: 10.1073/pnas.96.6.3160 . PMC   15912 . PMID   10077654.
  29. Lichter-Konecki U, Hipke CM, Konecki DS (August 1999). "Human phenylalanine hydroxylase gene expression in kidney and other nonhepatic tissues". Molecular Genetics and Metabolism. 67 (4): 308–16. doi:10.1006/mgme.1999.2880. PMID   10444341.
  30. Muntau AC, Gersting SW (December 2010). "Phenylketonuria as a model for protein misfolding diseases and for the development of next generation orphan drugs for patients with inborn errors of metabolism". Journal of Inherited Metabolic Disease. 33 (6): 649–58. doi:10.1007/s10545-010-9185-4. PMID   20824346. S2CID   20843095.
  31. 1 2 3 Singh K, Cornell CS, Jackson R, Kabiri M, Phipps M, Desai M, Fogle R, Ying X, Anarat-Cappillino G, Geller S, Johnson J, Roberts E, Malley K, Devlin T, DeRiso M, Berthelette P, Zhang YV, Ryan S, Rao S, Thurberg BL, Bangari DS, Kyostio-Moore S (31 March 2021). "CRISPR/Cas9 generated knockout mice lacking phenylalanine hydroxylase protein as a novel preclinical model for human phenylketonuria". Scientific Reports. 11 (1): 7254. doi:10.1038/s41598-021-86663-8. PMC   8012645 . PMID   33790381.
  32. "Knockout Mice Fact Sheet". Genome.gov.
  33. "PAH phenylalanine hydroxylase [Homo sapiens (human)] - Gene - NCBI". www.ncbi.nlm.nih.gov.

Further reading