Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Critical Review

Phenylalanine Hydroxylase: Function, Structure, Marte I. Flydal


Aurora Martinez*
and Regulation

Department of Biomedicine and K.G. Jebsen Centre for Research on


Neuropsychiatric Disorders, University of Bergen, Jonas Lies vei 91, 5009-
Bergen, Norway

Abstract
Mammalian phenylalanine hydroxylase (PAH) catalyzes the PAH by its substrate and cofactor, in addition to improved
rate-limiting step in the phenylalanine catabolism, consum- correlations between genotype and phenotype in PKU.
ing about 75% of the phenylalanine input from the diet and Importantly, there has also been an increased number of
protein catabolism under physiological conditions. In studies on the structure and function of PAH from bacteria
humans, mutations in the PAH gene lead to phenylketonuria and lower eukaryote organisms, revealing an additional ana-
(PKU), and most mutations are mainly associated with PAH bolic role of the enzyme in the synthesis of melanin-like pig-
misfolding and instability. The established treatment for PKU ments. In this review, we discuss these recent studies,
is a phenylalanine-restricted diet and, recently, supplementa- which contribute to define the evolutionary adaptation of the
tion with preparations of the natural tetrahydrobiopterin PAH structure and function leading to sophisticated regula-
cofactor also shows effectiveness for some patients. Since tion for effective catabolic processing of phenylalanine in
1997 there has been a significant increase in the understand- mammalian organisms. V C 2013 IUBMB Life, 65(4):341–349,

ing of the structure, catalytic mechanism, and regulation of 2013

Keywords: enzymology; evolution; protein function; protein structure

Introduction mitters and other L-Tyr derivatives. The function of PAH in


mammals has been studied since the end of the 1950s (see ref.
Phenylalanine hydroxylase (PAH, EC 1.14.16.1) catalyzes the 1 for a review of the earlier studies). PAH is primarily present
conversion of L-phenylalanine (L-Phe) to L-tyrosine (L-Tyr) by in the liver, where removal of excess L-Phe prevents the neuro-
para-hydroxylation of the aromatic side-chain. In mammals, toxic effect of hyperphenylalaninemia (HPA). However, L-Phe is
this tetrahydrobiopterin (BH4)-dependent reaction is the initial also an essential proteinogenic amino acid and it is important
and rate-limiting step in the degradation of excess L-Phe from to avoid that it is fully catabolized. To accomplish this dual
dietary proteins, where L-Tyr is further degraded to products role of preserving, yet removing excess L-Phe effectively, mam-
that feed into the citric acid cycle (Fig. 1). L-Tyr is thus a non- malian PAH has developed several regulatory mechanisms and
essential amino acid in PAH-containing organisms, where it is a specific structure—the framework through which the regula-
used for protein synthesis or as precursor for neurotrans- tory properties are exerted. In recent years, there have been
important advances in the elucidation of the catalytic mecha-
nism and structure–function relationships (2–4), but less is
C 2013 International Union of Biochemistry and Molecular Biology, Inc.
V known about the structure–regulation relationships (5). PAH is
Volume 65, Number 4, April 2013, Pages 341–349 also present in non-mammalian eukaryote organisms and
*Address for correspondence to: Aurora Martinez, Department of some bacteria, and notably in the last decade several studies
Biomedicine and K.G. Jebsen Centre for Research on Neuropsychiatric have investigated PAH from these sources. In addition to cata-
Disorders, University of Bergen, Jonas Lies vei 91, 5009-Bergen, Norway.
Tel: þ47-55586427. E-mail: aurora.martinez@biomed.uib.no.
bolic degradation (6,7), PAH from bacteria and non-mamma-
Received 4 December 2012; accepted 9 January 2013 lian eukaryote organisms has a prominent anabolic role con-
DOI: 10.1002/iub.1150 tributing to the synthesis of melanin-type pigments (8–11) (Fig.
Published online 4 March 2013 in Wiley Online Library 1). This review focuses on the present understanding of the
(wileyonlinelibrary.com) structure, function, and regulation of mammalian PAH, and

IUBMB Life 341


15216551, 2013, 4, Downloaded from https://iubmb.onlinelibrary.wiley.com/doi/10.1002/iub.1150 by National Health And Medical Research Council, Wiley Online Library on [19/05/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
IUBMB LIFE

The PAH reaction and complete L-Phe catabolism (black pathway). In blue (and squared gray background), the abnormal me-
FIG 1 tabolism of L-Phe in phenylketonuria, through transamination and decarboxylation of accumulated L-Phe. In red (and round
grey background), the production of pyomelanin from homogentisate occurring in some bacteria. [Color figure can be viewed
in the online issue, which is available at wileyonlinelibrary.com.]

includes a comparative discussion of the enzyme from different ria (PKU) (OMIM 261600) is the most severe form with plasma L-
organisms along an evolutionary perspective. Phe levels >1,200 lM (12). The accumulation of L-Phe and the
subsequent disturbance in brain neurotransmitters lead to neu-
rological symptoms including mental retardation, purposeless
movements, and depression (for a complete list of symptoms,
Phenylketonuria
see www.omim.org). The dietary intake of L-Phe must therefore
Dysfunctional PAH leads to increased concentration of L-Phe in be strictly controlled in PKU patients (12–14).
the blood and the appearance in urine of metabolites that arise PKU is inherited as an autosomal recessive disorder, with
from the transamination of L-Phe to phenylpyruvate (Fig. 1). more than 500 disease-causing mutations (www.pahdb.
This is the hallmark of the HPAs, of which classic phenylketonu- mcgill.ca and www.biopku.org). In vitro studies of mutant

342 Evolutionary Adaptation of PAH Structure and Function


15216551, 2013, 4, Downloaded from https://iubmb.onlinelibrary.wiley.com/doi/10.1002/iub.1150 by National Health And Medical Research Council, Wiley Online Library on [19/05/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
proteins have revealed the kinetic and conformational defects ganization to enable sophisticated regulation, thereby adapting
caused by the mutations. PKU appears largely as a misfolding to the emerging metabolic and neurological needs of high-
disease where loss of enzymatic function is caused mainly by complexity organisms.
folding defects leading to decreased stability, and PKU is con-
sidered a paradigm for misfolding loss-of-function genetic dis-
orders (13,15,16). The molecular basis for the neurological The Phenylalanine Hydroxylase System
symptoms is not completely understood, but saturation by
As for the other AAAHs, PAH catalyzes the hydroxylation of its
L-Phe of the LAT-1 transporter at the blood–brain barrier
substrate by incorporation of one oxygen atom into the aro-
and defective myelination seem critical. Recently, a new
matic ring, and the final reaction also includes the reduction
amyloidosis-like etiology for PKU has been suggested, since
of the second oxygen atom to water using two electrons sup-
L-Phe has been shown to self-assemble into toxic fibrils (17).
plied by BH4. The cofactor BH4 functions as a co-substrate
While displaying highest affinity for L-Phe, LAT-1 is also
that is also hydroxylated at each turnover to pterin-4a-carbi-
the only route for eight other large neutral amino acids
nolamine (4a-OH-BH4), with consequent dissociation from the
(LNAAs) into the brain (18). LNAAs are important for cerebral
enzyme (23) (Fig. 1). The dehydration and reduction of 4a-OH-
protein synthesis, and L-Tyr and L-Tryptophan (L-Trp) are also
BH4 back to BH4 is catalyzed sequentially by pterin
precursors for neurotransmitters. Saturation of LAT-1 by ele-
carbinolamine dehydratase, which dehydrates 4a-OH-BH4 to
vated levels of L-Phe thus creates an imbalance in both neuro-
dihydrobiopterin quinonoid (q-BH2), and the NADH-dependent
transmitter- and protein synthesis, including hypomyelination.
dihydropteridine reductase, respectively (23). These two
Consequently, supplementation with LNAAs is one treatment
enzymes are therefore considered as part of the PAH system,
strategy currently investigated, in addition to enzymatic therapy
making the degradation of L-Phe sensitive to defects in several
with pegylated phenylalanine ammonia lyase, gene therapy,
genes. If q-BH2 rearranges to BH2, reduction to BH4 can be
pharmacological chaperones and the already approved supple-
catalyzed by dihydrofolate reductase (23).
mentation with the cofactor BH4 (for a review see 14). A number
of studies have contributed to reveal the most recurrent geno-
types associated with BH4-responsive PKU and the molecular Structure of Mammalian PAH
mechanisms behind the corrective effects of BH4 (12,19).
Mammalian PAH is a homo-tetrameric enzyme of 50 kDa sub-
units. The structure of a full-length mammalian PAH has not
yet been solved, but truncated PAH structures are available
Aromatic Amino Acid Hydroxylases
(Table 1). A composite model of full-length tetrameric PAH can
PAH is a member of the aromatic amino acid hydroxylase be prepared based on crystal structures from dimeric trun-
(AAAH) enzyme family. The AAAHs share a requirement for a cated rat PAH (PDB 2PHM) and tetrameric human PAH (hPAH)
catalytic non-heme ferrous iron, BH4 as cofactor, and molecu- (PDB 2PAH) (Fig. 2B). Although only the BH4-responsive PKU
lar oxygen as additional substrate. The mammalian genes mutant A313T-PAH has been amenable to crystal structure
encode PAH, tyrosine hydroxylase (TH), and two tryptophan determination (19), the available structures have been crucial
hydroxylases (TPH1 and TPH2), which are named after their for correlating genotypes and phenotype in PKU, defining hot-
specific amino acid substrates. The products of TH (L-3,4-dihy- spots for enzyme destabilization (4,15).
droxyphenylalanine (L-DOPA)) and the TPHs (5-hydroxytrypto- Each PAH subunit is composed of an N-terminal regula-
phan) are precursors for important neurotransmitters and tory domain (residues 1–110), a central catalytic domain
hormones in the brain and the neuroendocrine system (2). The (residues 111–410), and a C-terminal oligomerization domain
AAAHs show high sequence identity (Fig. 2A), similar struc- (residues 411–452) (24,25) (Fig. 2B). The N-terminal regula-
ture, and presumed similar catalytic mechanism (2,3). Metazo- tory domain is classified as an ACT-domain, a regulatory mod-
ans have three or four AAAHs-encoding genes, but only a sin- ule present in several proteins, which often dimerizes and
gle gene has so far been found in protozoans, such as binds amino acids (26). This domain is flexibly attached to the
Leishmania major (20), and in the slime mold Dictyostelium catalytic domain via a hinge region (Arg111–Thr117), but
discoideum (21). These single AAAHs have been identified as establishes extensive contacts with the catalytic domain of the
PAHs (20,21). Moreover, PAH is the only AAAH found in bacte- adjacent subunit within the dimer. The regulatory domain is
ria. In fact, based on the homology between the regulatory necessary for manifestation of the regulatory properties, such
AAAH domains and prephenate dehydratase, Gjetting et al. as activation by L-Phe, and it is still debated whether or not it
proposed that bacterial PAH and prephenate dehydratase includes an allosteric binding-site for L-Phe. In support of the
were the precursors of multi-domain AAAHs (22). A role of existence of a regulatory site, Gjetting et al. remarked that the
PAH as the ancestral function in the AAAH family appears rea- regulatory domain presents two motifs, GAL (residues 46–48
sonable since the TH and TPH functions are likely to be of in hPAH) and (I/L)ESRP (residues 65–69), which are involved
increasing importance in pluricellular organisms. As will be in L-Phe binding in a bacterial prephenate dehydratase, and
shown in the next sections, PAH has evolved its structural or- demonstrated that isolated regulatory domains of hPAH

Flydal and Martinez 343


15216551, 2013, 4, Downloaded from https://iubmb.onlinelibrary.wiley.com/doi/10.1002/iub.1150 by National Health And Medical Research Council, Wiley Online Library on [19/05/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
IUBMB LIFE

Domain organization and structure of PAH. A: Alignment of human AAAHs and PAHs from different organisms. Bars represent
FIG 2 gaps (white), non-identical residues (gray), residues identical in 60% of the sequences (dark gray), and residues identical in
all the sequences (black). RD, regulatory domain; CD, catalytic domain; OD, oligomerization domain. The highest homology
(80%) is encountered in the catalytic domains. B: Composite model of full-length tetrameric human PAH prepared with PDBs
2PHM and 2PAH. Inset, domain organization of each subunit. C: The structure of the ternary PAH  Fe(II)  BH4  L-Phe complex,
based on PDB 1MMK, with L-Phe modeled at the 3-(2-thienyl)-L-alanine binding site. D: The crystal structure of PAH from
C. violaceum PAH (PDB 1LTV). [Color figure can be viewed in the online issue, which is available at wileyonlinelibrary.com.]

actually form a dimer and bind L-Phe (22). However, whether nism (2,3,31). Formation of the high-spin Fe(IV) hydroxylating
this intersubunit regulatory binding site is functional in full- species—identified experimentally in the catalytic cycles of
length PAH, where (according to the available structural mod- both TH (32) and PAH from Chromobacterium violaceum
els) regulatory domains are physically separated (Fig. 2B), has (33)—is a crucial step in the reaction.
not been convincingly demonstrated (see also next section). The oligomerization domain starts by an antiparallel-sheet
Preceding the ACT-domain is an intrinsic autoregulatory (residues 411–414, 421–424), responsible for dimerization, fol-
sequence (residues 1–33 in hPAH) that extends over the active lowed by a 40 Å long-helix (428–452) that mediates tetrameri-
site (24) and limits L-Phe access, especially when its physiologi- zation through domain swapping and antiparallel coiled-coil
cal cofactor (6R)-BH4 is bound (24,27–29). The highly mobile formation with the other monomers (25) (Fig. 2B).
N-terminal (residues 1–18) of the intrinsic autoregulatory
sequence is not seen in the crystal structure, and conforma-
tional information has been obtained by molecular modeling Regulation of Mammalian PAH by L-Phe,
(30). BH4, and Phosphorylation
The catalytic domain contains the binding sites for iron,
cofactor, and substrate. At the active site iron binds to two his- Several mechanisms act together to tightly control the activity of
tidines (His285 and His290 in hPAH) and a glutamate (Glu330 mammalian PAH, and some short-term effects known to play a
in hPAH) (Fig. 2C). Structures with bound L-Phe analogues physiological role are caused by L-Phe, the natural cofactor (6R)-
3-(2-thienyl)-L-alanine or L-norleucine, and reduced or oxi- BH4, and phosphorylation at Ser16 by cAMP-dependent protein
dized cofactor at the active site (Table 1; Fig. 2C) have kinase. Liver PAH from different mammals displays the same
provided the molecular frames to elucidate the AAAH mecha- physical and regulatory properties as the human enzyme,

344 Evolutionary Adaptation of PAH Structure and Function


15216551, 2013, 4, Downloaded from https://iubmb.onlinelibrary.wiley.com/doi/10.1002/iub.1150 by National Health And Medical Research Council, Wiley Online Library on [19/05/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Representative PAH structures solved by crystallography
TABLE 1

PDB ID Domains (residues) State Ligands Organism

1PAH CD (117–427) D H. sapiens


2PAH CD, OD (118–452) T H. sapiens
3PAH CD (103–427) D L-adrenaline H. sapiens
4PAH CD (103–427) D L-noradrenaline H.sapiens
5PAH CD (103–427) D L-dopamine H. sapiens
2PHMa RD, CD(1–429) D R. norvegicus
a,b
1PHZ RD, CD(1–429) D R. norvegicus
1DMW CD (118–424) D BH2 H. sapiens
1J8U CD (103–427) D BH4 H. sapiens
1LTV CD (1–275) M C. violaceum
1LTZ CD (1–275) M BH2 C. violaceum
1MMK CD (103–427) D BH4, 3-(2-thienyl)-L-alanine H. sapiens
1MMT CD (103-427) D BH4, L-norleucine H. sapiens
1TG2 CD A313T (103-427) D BH2 H. sapiens
2V27 CD (1-275) M C. psychrerythraea

RD, regulatory domain; CD, catalytic domain; OD, oligomerization domain; D, dimeric; T, tetrameric; M, monomeric.
a
No structural information for residues 1–18.
b
Phosphorylated at Ser16.

although there are some differences in the activation constants indication that binding to a regulatory site might be unnecessary
and notably the extent of the activation by L-Phe (5). to generate positive cooperative response arose when it was
The activity of tetrameric mammalian PAH—when assayed shown that noradrenaline, which binds directly to the ferric iron
with 6R-BH4—is activated by incubation with L-Phe and in the active site, also elicited positive cooperativity (37). Later,
responds with positive cooperativity toward increasing sub- differential scanning calorimetry investigations further sup-
strate concentrations (Hill coefficient 2) (1,34). The need for ported that L-Phe only binds to the catalytic domain of hPAH
fine-tuned conformational changes to elicit an allosteric (38). Moreover, surface plasmon resonance analyses and site-
response by L-Phe is manifested by the loss of positive coopera- directed mutagenesis on crystallographically defined hinges
tivity in many PKU-associated hPAH mutants that still maintain have also identified the active site as the epicenter of the global
the tetrameric structure (19). Current models to explain alloste- conformational changes resulting in activation and positive
ric regulation with positive cooperative ligand-binding describe cooperativity observed in the full-length tetrameric enzyme
the transition from a tense (T), less active state with low affinity (29). Nevertheless, evidences for L-Phe binding by the isolated
for the ligand, to a relaxed (R), more active, high-affinity state. A regulatory domains have been put forward by isothermal titra-
large conformational transition is indeed visualized for mam- tion calorimetry (22) and nuclear magnetic resonance spectros-
malian PAH upon activation by substrate as shown among other copy (36). Interestingly, recent analyses reveal that substrat
by the surfacing of a buried tryptophan residue and an increase binding sites in fact exist in the regulatory ACT domain in hPAH,
in hydrophobicity and size (1,5,29,34). The mechanism of this but binding appears obstructed by residues from the catalytic
activation is not yet understood and there is no consensus in the domain (39). It seems therefore not surprising, neither contra-
field as to whether the activation is caused by L-Phe binding to dictory, that L-Phe binds to isolated ACT domains (22,36).
an allosteric site in the regulatory domain (22,35,36) or to the The cofactor BH4 also exerts a regulatory inhibitory effect
active site itself (37–39). The numerous studies related to this on mammalian PAH, manifested as a stabilization of the
subject have been recently reviewed by Fitzpatrick (5). The first T-state, which is reverted by L-Phe activation (1,5,27). The

Flydal and Martinez 345


15216551, 2013, 4, Downloaded from https://iubmb.onlinelibrary.wiley.com/doi/10.1002/iub.1150 by National Health And Medical Research Council, Wiley Online Library on [19/05/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
IUBMB LIFE

negative effect on activity is attributed to the intrinsic autoreg- nolamine dehydratase, in the phh operon, which also includes
ulatory sequence (28), and molecular dynamics (MD) simula- the gene for an aromatic aminotransferase (45). The similarity
tions complemented with thermodynamic characterizations of of the regulatory domain of mammalian PAH with pterin car-
BH4-binding also support a restructuring of the intrinsic auto- binolamine dehydratase has been remarked, and the bacterial
regulatory sequence, which possibly occupies the L-Phe binding operon-structure has been proposed as a possible basis for
site (27). In the liver, PAH exists largely in this BH4-bound, evolution of the modular gene encoding eukaryotic PAH (24).
low-activity, stable state (40). The stabilizing effect of BH4 is Several bacterial PAHs have been studied experimentally
one of the most important molecular mechanisms behind the to different extents, all showing an absolute requirement for
stimulation of PKU-mutant activity by BH4 supplementation ferrous iron and tetrahydropterin (11,45–47). It has been
(i.e., the chaperone effect) (19). speculated that most bacterial PAHs use alternative pterins to
Phosphorylation of mammalian PAH at Ser16 by glucagon- BH4, for example, tetrahydromonapterin (44). For the Colwel-
stimulated cAMP-dependent protein kinase increases the affin- lia psychrerythraea enzyme, highest activity was however
ity for L-Phe and thus activates the enzyme in synergy with measured with BH4 (47). Except for the recently studied PAH
L-Phe activation (1,5). Crystal structures of unphosphorylated from Legionella pneumophila, which appears to be dimeric
and phosphorylated forms of dimeric rat PAH (24) (Table 1) (11), presently characterized bacterial PAHs are monomeric
lack structural information on the region around Ser16. Site- and show a similar fold to the catalytic domain of mammalian
directed mutagenesis and molecular modeling indicate that PAH (46,47) (Table 1; Fig. 2D).
phosphorylation elicits local conformational changes, mostly Several pathogens encode a PAH and produce a melanin-
driven by electrostatics, that increase the accessibility of the pigment, notably a pyomelanin derived from homogentisate
substrate binding site (30). autooxidation and polymerization (Fig. 1). The high mutational
rate of Pseudomonas aeruginosa during the course of lung
infection leads to adaptive mutations in the gene encoding
PAH in Non-Mammalian Eukaryote homogentisate oxidase (Fig. 1) (48). This induces increased
homogentisate accumulation and hyperproduction of pyomela-
Organisms nin that render the infection very difficult to treat (48). The
The investigation of PAH in non-mammalian organisms includ- pyomelanin of the opportunistic pathogen Burkholderia ceno-
ing bacteria readily followed the earliest studies of mammalian cepacia was recently shown to have antioxidant properties,
PAH (41). Recently, PAH has also been identified in some pro- and mutants not producing pyomelanin were more sensitive to
tozoans and slime molds, and even in nonflowering plants oxidative stress (49). Furthermore, the crucial role of L. pneu-
(20,21,42). While animal PAH uses BH4 as cofactor, other nat- mophila PAH in both the growth of the bacterium in tyrosine-
ural tetrahydropterins, such as tetrahydrodictyopterin, are limited media and in the synthesis of pyomelanin has been
present in D. discoideum together with BH4 and support the recently demonstrated (11) (Fig. 3). The secreted pyomelanin
PAH function at least in vitro (21,43). However, BH4 seems to has intrinsic ferric reductase activity, contributing to the
be the PAH cofactor in vivo, while tetrahydrodictyopterin func- capacity of Legionella to acquire iron and possibly to its viru-
tions mainly as an antioxidant (43). Plant PAH seems to use lence (50).
tetrahydrofolate as cofactor (42).
In addition to their role in the catabolic L-Phe/L-Tyr degra-
dation pathway (7), PAH in lower eukaryotes has been impli-
cated in the synthesis of (pyo)melanin, a well-known pigment Adaptive Strategies: Regulation Of PAH
that confers advantageous properties (8–10,20). Melanin is Activity By L-Phe
used by insects to encapsulate parasites (9) and by the sponge
Geodia cydonium to target cells for apoptosis (8). In the nema- PAH shows an evolutionary adaptation of its structure to
tode Caenorhabditis elegans, PAH is expressed in the hypoder- accommodate for changes in the regulation of its function. The
mal cells, and knock-out pah mutants lack a melanin in the characterization of PAHs in organisms at different levels of
cuticle which appears to protect from oxygen-radicals (10). cellular complexity indicates that the tetrameric structure
appears early—in an evolutionary scale—in the animal king-
dom (21). Clear manifestations of allosteric regulation of PAH
Bacterial PAH activity by L-Phe have only been proven for mammals, but the
need for PAH activity in anabolic/catabolic pathways with si-
In bacteria, the gene coding for PAH (phhA) occurs within sev- multaneous preservation of threshold levels of L-Phe for pro-
eral phyla, but appears to be relatively scattered within these. tein synthesis would require a strict enzyme regulation in all
The phylogenetic distribution of phhA in bacteria has not been organisms. Several mechanisms appear to contribute to main-
investigated per se, but Pribat et al. report that 184 out of 724 tain L-Phe homeostasis in different organisms, and, in addition
bacterial genomes encode a PAH (44). In some PAH-containing to the regulation by positive cooperativity in mammals, other
bacteria, phhA is located next to phhB, encoding pterin carbi- mechanisms such as non-catalytic binding of L-Phe, and low

346 Evolutionary Adaptation of PAH Structure and Function


15216551, 2013, 4, Downloaded from https://iubmb.onlinelibrary.wiley.com/doi/10.1002/iub.1150 by National Health And Medical Research Council, Wiley Online Library on [19/05/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Role of PAH in growth and pigment-synthesis in Legionella pneumophila 130b. Wild type 130b and mutant L. pneumophila
FIG 3 lacking PAH (NU406) with empty vector (pMMB2002) or vector containing the pah gene (pPhhA) were grown in chemically
defined medium (CDM) lacking tyrosine (A) and in medium containing standard (CDM) (B,C) and twice the standard (CDM^ 2
Tyr) amount of tyrosine (B). The importance of PAH both for growth in media limited in tyrosine (A), and for the production of
pyomelanin pigment (B,C) was established. Figure adapted from ref. 11. [Color figure can be viewed in the online issue, which
is available at wileyonlinelibrary.com.]

substrate affinity have been revealed in recent in vitro studies Evolutionary Change in the Function of
with purified PAHs from C. elegans and bacteria.
PAH from C. elegans, which is tetrameric and includes the
the PAH-Catalyzed Reaction
canonical three-domain PAH organization and 57% sequence It is broadly accepted that the main role of mammalian PAH is
identity with hPAH, does not show positive cooperativity (10). the catabolic degradation of dietary L-Phe and protection from
This sophisticated regulatory mechanism that allows mamma- HPA. On the other hand, combining the studies from Calvo
lian PAH to avoid toxic accumulation of L-Phe is not present in et al. and Fisher et al. shows that the role of C. elegans PAH is
the worm most probably because it is simply not required at catabolic, but also anabolic, providing L-Phe for melanin syn-
its level of nervous system complexity. In fact, knock-out pah thesis (7,10). In fact, functional divergence analysis acknowl-
worms appear healthy even when grown in media supple- edge that an important functional switch has occurred
mented with excess L-Phe (10), suggesting that accumulation between the nematode PAH and mammalian PAH on the evo-
of L-Phe is not detrimental to the C. elegans nervous system. lutionary time-scale (39). These results further indicate that
Nevertheless, it is reasonable to infer that the worm would the specific residue substitutions are associated to fundamen-
face the need of maintaining threshold values of L-Phe for protein tal changes in regulation of the activity (10). The more sophis-
synthesis. In this context, our discovery of a second binding site for ticated regulation in the mammalian organisms seems to tune
L-Phe in the regulatory domain of worm PAH (39) suggests a sim- the enzyme toward the effective catabolic processing of L-Phe.
pler regulatory mechanism where the non-catalytic binding of L-
Phe, which does not seem to induce activating conformational
changes, rather ensures the preservation of certain L-Phe levels. Acknowledgements
In the case of bacteria, the cold-adapted PAH from C. psy- Authors are very thankful to all members of the Biorecognition
chrerythraea shows a low affinity for L-Phe, and the concen- research group for discussions, especially to Jarl Underhaug who
tration of substrate providing half maximal activity ([S]0.5, a also contributed to the figures. This research is supported by
parameter used in non-Michaelis–Menten kinetics to provide a grants from the Research Council of Norway, the Western Nor-
constant comparable to Km) is 1300 6 300 lM, much higher way Health Authorities, K.G. Jebsen Centre for Research on Neu-
than [S]0.5 < 200 lM for eukaryote PAH (1,10). This could be a ropsychiatric Disorders and Novo Seeds (Novo Nordisk Fonden).
result of the increased flexibility and accessibility of the active
site due to cold adaptation, but might also indicate a primitive
regulatory mechanism to preserve threshold amounts of L-Phe. References
Indeed, also the thermostable PAH from L. pneumophila dis- [1] Kaufman, S. (1993) The phenylalanine hydroxylating system. Adv. Enzymol.
Relat. Areas Mol. Biol. 67, 77–264.
played low affinity for L-Phe (Km ¼ 735 6 50 lM). But not all
[2] Fitzpatrick, P. F. (2000) The aromatic amino acid hydroxylases. Adv. Enzymol.
bacterial PAHs seem to have low affinity for its substrate and Relat. Areas Mol. Biol. 74, 235–294.
PAH from C. violaceum has Km (L-Phe) comparable to eukar- [3] Olsson, E., Teigen, K., Martinez, A., and Jensen, V. R. (2010) The aromatic
yote PAH (51). Additional studies are necessary to reveal the amino acid hydroxylase mechanism: a perspective from computational
extent of low affinity for L-Phe among bacterial PAH. In fact chemistry. Adv. Inorg. Chem. 62, 437–500.
[4] Flatmark, T. and Stevens, R. C. (1999) Structural insight into the aromatic
other mechanisms at the transcriptional level, such as the reg-
amino acid hydroxylases and their disease-related mutant forms. Chem. Rev.
ulation of transcription from phhA by L-Phe shown to take 99, 2137–2160.
place in Pseudomonas putida (52), might also be important to [5] Fitzpatrick, P. F. (2012) Allosteric regulation of phenylalanine hydroxylase.
safeguard L-Phe (and aromatic amino acid) homeostasis. Arch. Biochem. Biophys. 519, 194–201.

Flydal and Martinez 347


15216551, 2013, 4, Downloaded from https://iubmb.onlinelibrary.wiley.com/doi/10.1002/iub.1150 by National Health And Medical Research Council, Wiley Online Library on [19/05/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
IUBMB LIFE

[6] Arias-Barrau, E., Olivera, E. R., Luengo, J. M., Fernandez, C., Galan, B., et al. [26] Liberles, J. S., Tho  ro
 lfsson, M., and Martinez, A. (2005) Allosteric mecha-
(2004) The homogentisate pathway: a central catabolic pathway involved in nisms in ACT domain containing enzymes involved in amino acid metabo-
the degradation of L-phenylalanine, L-tyrosine, and 3-hydroxyphenylacetate lism. Amino Acids 28, 1–12.
in Pseudomonas putida. J. Bacteriol. 186, 5062–5077. [27] Pey, A. L., Tho  ro
 lfsson, M., Teigen, K., Ugarte, M., and Martı́nez, A. (2004)
[7] Fisher, A. L., Page, K. E., Lithgow, G. J., and Nash, L. (2008) The Caenorhab- Thermodynamic characterization of the binding of tetrahydropterins to phe-
ditis elegans K10C2.4 gene encodes a member of the fumarylacetoacetate nylalanine hydroxylase. J. Am. Chem. Soc. 126, 13670–13678.
hydrolase family: a Caenorhabditis elegans model of type I tyrosinemia. J. [28] Jennings, I. G., Teh, T., and Kobe, B. (2001) Essential role of the N-terminal
Biol. Chem. 283, 9127–9135. autoregulatory sequence in the regulation of phenylalanine hydroxylase.
[8] Wiens, M., Koziol, C., Batel, R., and Muller, W. E. (1998) Phenylalanine FEBS Lett. 488, 196–200.
hydroxylase from the sponge Geodia cydonium: implication for allorecogni- [29] Stokka, A. J., Carvalho, R. N., Barroso, J. F., and Flatmark, T. (2004) Probing
tion and evolution of aromatic amino acid hydroxylases. Dev. Comp. Immu- the role of crystallographically defined/predicted hinge-bending regions in
nol. 22, 469–478. the substrate-induced global conformational transition and catalytic activa-
[9] Infanger, L. C., Rocheleau, T. A., Bartholomay, L. C., Johnson, J. K., Fuchs, tion of human phenylalanine hydroxylase by single-site mutagenesis. J.
J., et al. (2004) The role of phenylalanine hydroxylase in melanotic encapsu- Biol. Chem. 279, 26571–26580.
lation of filarial worms in two species of mosquitoes. Insect. Biochem. Mol. [30] Miranda, F. F., Teigen, K., Thorolfsson, M., Svebak, R. M., Knappskog, P. M.,
Biol. 34, 1329–1338. et al. (2002) Phosphorylation and mutations of Ser(16) in human phenylalanine
[10] Calvo, A. C., Pey, A. L., Ying, M., Loer, C. M., and Martinez, A. (2008) Ana- hydroxylase. Kinetic and structural effects. J. Biol. Chem. 277, 40937–40943.
bolic function of phenylalanine hydroxylase in Caenorhabditis elegans. [31] Andersen, O. A., Flatmark, T., and Hough, E. (2002) Crystal structure of the ter-
FASEB J. 22, 3046–3058. nary complex of the catalytic domain of human phenylalanine hydroxylase
[11] Flydal, M. I., Chatfield, C. H., Zheng, H., Gunderson, F. F., Aubi, O., et al. with tetrahydrobiopterin and 3-(2-thienyl)-L-alanine, and its implications for the
(2012) Phenylalanine hydroxylase from Legionella pneumophila is a ther- mechanism of catalysis and substrate activation. J. Mol. Biol. 320, 1095–1108.
mostable enzyme with a major functional role in pyomelanin synthesis. [32] Eser, B. E., Barr, E. W., Frantom, P. A., Saleh, L., Bollinger, J. M., Jr., et al.
PLoS One 7, e46209. (2007) Direct spectroscopic evidence for a high-spin Fe(IV) intermediate in
[12] Blau, N., Hennermann, J. B., Langenbeck, U., and Lichter-Konecki, U. (2011) tyrosine hydroxylase. J. Am. Chem. Soc. 129, 11334–11335.
Diagnosis, classification, and genetics of phenylketonuria and tetrahydro- [33] Panay, A. J., Lee, M., Krebs, C., Bollinger, J. M., and Fitzpatrick, P. F. (2011)
biopterin (BH4) deficiencies. Mol. Genet. Metab 104, Suppl:S2–9. Evidence for a high-spin Fe(IV) species in the catalytic cycle of a bacterial
[13] Scriver, C. R. (2007) The PAH gene, phenylketonuria, and a paradigm shift. phenylalanine hydroxylase. Biochemistry 50, 1928–1933.
Hum. Mutat. 28, 831–845. [34] Tho ro
 lfsson, M., Teigen, K., and Martı́nez, A. (2003) Activation of phenylala-
[14] van Spronsen, F. J. (2010) Phenylketonuria: a 21st century perspective. Nat. nine hydroxylase: effect of substitutions at Arg68 and Cys237. Biochemistry
Rev. Endocrinol. 6, 509–514. 42, 3419–3428.
[15] Pey, A. L., Stricher, F., Serrano, L., and Martinez, A. (2007) Predicted effects [35] Shiman, R., Xia, T., Hill, M. A., and Gray, D. W. (1994) Regulation of rat liver
of missense mutations on native-state stability account for phenotypic out- phenylalanine hydroxylase. II. Substrate binding and the role of activation
come in phenylketonuria, a paradigm of misfolding diseases. Am. J. Hum. in the control of enzymatic activity. J. Biol. Chem. 269, 24647–24656.
Genet. 81, 1006–1024. [36] Li, J., Ilangovan, U., Daubner, S. C., Hinck, A. P., and Fitzpatrick, P. F. (2011)
[16] Gersting, S. W., Kemter, K. F., Staudigl, M., Messing, D. D., Danecka, M. K., Direct evidence for a phenylalanine site in the regulatory domain of phenyl-
et al. (2008) Loss of function in phenylketonuria is caused by impaired mo- alanine hydroxylase. Arch. Biochem. Biophys. 505, 250–255.
lecular motions and conformational instability. Am. J. Hum. Genet. 83, 5–17. [37] Martı́nez, A., Haavik, J., and Flatmark, T. (1990) Cooperative homotropic
[17] Adler-Abramovich, L., Vaks, L., Carny, O., Trudler, D., Magno, A., et al. interaction of L-noradrenaline with the catalytic site of phenylalanine 4-
(2012) Phenylalanine assembly into toxic fibrils suggests amyloid etiology monooxygenase. Eur. J. Biochem. 193, 211–219.
in phenylketonuria. Nat. Chem. Biol. 8, 701–706. [38] Tho ro
 lfsson, M., Ibarra-Molero, B., Fojan, P., Petersen, S. B., Sanchez-Ruiz,
[18] Pratt, O. E. (1980) A new approach to the treatment of phenylketonuria. J. J. M., et al. (2002) L-Phenylalanine binding and domain organization in
Ment. Defic. Res. 24, 203–217. human phenylalanine hydroxylase: a differential scanning calorimetry
[19] Erlandsen, H., Pey, A. L., Gamez, A., Perez, B., Desviat, L. R., et al. (2004) study. Biochemistry 41, 7573–7585.
Correction of kinetic and stability defects by tetrahydrobiopterin in phenyl- [39] Flydal, M. I., Mohn, T. C., Pey, A. L., Siltberg-Liberles, J., Teigen, K., et al.
ketonuria patients with certain phenylalanine hydroxylase mutations. Proc. (2010) Superstoichiometric binding of L-Phe to phenylalanine hydroxylase
Natl. Acad. Sci. USA 101, 16903–16908. from Caenorhabditis elegans: evolutionary implications. Amino Acids 39,
[20] Lye, L. F., Kang, S. O., Nosanchuk, J. D., Casadevall, A., and Beverley, S. M. 1463–1475.
(2011) Phenylalanine hydroxylase (PAH) from the lower eukaryote Leishma- [40] Mitnaul, L. J. and Shiman, R. (1995) Coordinate regulation of tetrahydro-
nia major. Mol. Biochem. Parasitol. 175, 58–67. biopterin turnover and phenylalanine hydroxylase activity in rat liver cells.
[21] Siltberg-Liberles, J., Steen, I. H., Svebak, R. M., and Martinez, A. (2008) The Proc. Natl. Acad. Sci. USA 92, 885–889.
phylogeny of the aromatic amino acid hydroxylases revisited by character- [41] Guroff, G. and Ito, T. (1963) Induced, soluble phenylalanine hydroxylase
izing phenylalanine hydroxylase from Dictyostelium discoideum. Gene 427, from Pseudomonas Sp. grown on phenylalanine or tyrosine. Biochim. Bio-
86–92. phys. Acta 77, 159–161.
[22] Gjetting, T., Petersen, M., Guldberg, P., and Guttler, F. (2001) Missense [42] Pribat, A., Noiriel, A., Morse, A. M., Davis, J. M., Fouquet, R., et al. (2010)
mutations in the N-terminal domain of human phenylalanine hydroxylase Nonflowering plants possess a unique folate-dependent phenylalanine
interfere with binding of regulatory phenylalanine. Am. J. Hum. Genet. 68, hydroxylase that is localized in chloroplasts. Plant Cell 22, 3410–3422.
1353–1360. [43] Kim, H. L., Park, M. B., and Park, Y. S. (2011) Tetrahydrobiopterin is func-
[23] Werner, E. R., Blau, N., and Thony, B. (2011) Tetrahydrobiopterin: biochem- tionally distinguishable from tetrahydrodictyopterin in Dictyostelium discoi-
istry and pathophysiology. Biochem. J. 438, 397–414. deum Ax2. FEBS Lett. 585, 3047–3051.
[24] Kobe, B., Jennings, I. G., House, C. M., Michell, B. J., Goodwill, K. E., et al. [44] Pribat, A., Blaby, I. K., Lara-Nunez, A., Gregory, J. F.,III,de Crecy-Lagard, V.,
(1999) Structural basis of autoregulation of phenylalanine hydroxylase. Nat. et al. (2010) FolX and FolM are essential for tetrahydromonapterin synthesis
Struct. Biol. 6, 442–448. in Escherichia coli and Pseudomonas aeruginosa. J. Bacteriol. 192,
[25] Fusetti, F., Erlandsen, H., Flatmark, T., and Stevens, R. C. (1998) Structure of 475–482.
tetrameric human phenylalanine hydroxylase and its implications for phe- [45] Zhao, G., Xia, T., Song, J., and Jensen, R. A. (1994) Pseudomonas aerugi-
nylketonuria. J. Biol. Chem. 273, 16962–16967. nosa possesses homologues of mammalian phenylalanine hydroxylase and

348 Evolutionary Adaptation of PAH Structure and Function


15216551, 2013, 4, Downloaded from https://iubmb.onlinelibrary.wiley.com/doi/10.1002/iub.1150 by National Health And Medical Research Council, Wiley Online Library on [19/05/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
4 alpha-carbinolamine dehydratase/DCoH as part of a three-component [49] Keith, K. E., Killip, L., He, P., Moran, G. R., and Valvano, M. A. (2007) Bur-
gene cluster. Proc. Natl. Acad. Sci. USA 91, 1366–1370. kholderia cenocepacia C5424 produces a pigment with antioxidant proper-
[46] Erlandsen, H., Kim, J. Y., Patch, M. G., Han, A., Volner, A., et al. (2002) ties using a homogentisate intermediate. J. Bacteriol. 189, 9057–9065.
Structural comparison of bacterial and human iron-dependent phenylala- [50] Chatfield, C. H. and Cianciotto, N. P. (2007) The secreted pyomelanin pig-
nine hydroxylases: similar fold, different stability and reaction rates. J. Mol. ment of Legionella pneumophila confers ferric reductase activity. Infect.
Biol. 320, 645–661. Immun. 75, 4062–4070.
[47] Leiros, H. K., Pey, A. L., Innselset, M., Moe, E., Leiros, I., et al. (2007) Struc- [51] Volner, A., Zoidakis, J., and Abu-Omar, M. M. (2003) Order of substrate
ture of phenylalanine hydroxylase from Colwellia psychrerythraea 34H, a binding in bacterial phenylalanine hydroxylase and its mechanistic implica-
monomeric cold active enzyme with local flexibility around the active site tion for pterin-dependent oxygenases. J. Biol. Inorg. Chem. 8, 121–128.
and high overall stability. J. Biol. Chem. 282, 21973–21986. [52] Herrera, M. C., Daddaoua, A., Fernandez-Escamilla, A., and Ramos, J. L.
[48] Rodriguez-Rojas, A., Oliver, A., and Blazquez, J. (2012) Intrinsic and environ- (2012) Involvement of the global Crp regulator in cyclic AMP-dependent uti-
mental mutagenesis drive diversification and persistence of Pseudomonas lization of aromatic amino acids by Pseudomonas putida. J. Bacteriol. 194,
aeruginosa in chronic lung infections. J. Infect. Dis. 205, 121–127. 406–412.

Flydal and Martinez 349

You might also like