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

https://doi.org/10.1016/S0022-2836(02)00560-0Get rights and content

Abstract

Phenylalanine hydroxylase catalyzes the stereospecific hydroxylation of l-phenylalanine, the committed step in the degradation of this amino acid. We have solved the crystal structure of the ternary complex (hPheOH–Fe(II)·BH4·THA) of the catalytically active Fe(II) form of a truncated form (ΔN1–102/ΔC428–452) of human phenylalanine hydroxylase (hPheOH), using the catalytically active reduced cofactor 6(R)-l-erythro-5,6,7,8-tetrahydrobiopterin (BH4) and 3-(2-thienyl)-l-alanine (THA) as a substrate analogue. The analogue is bound in the second coordination sphere of the catalytic iron atom with the thiophene ring stacking against the imidazole group of His285 (average interplanar distance 3.8 Å) and with a network of hydrogen bonds and hydrophobic contacts. Binding of the analogue to the binary complex hPheOH–Fe(II)·BH4 triggers structural changes throughout the entire molecule, which adopts a slightly more compact structure. The largest change occurs in the loop region comprising residues 131–155, where the maximum r.m.s. displacement (9.6 Å) is at Tyr138. This loop is refolded, bringing the hydroxyl oxygen atom of Tyr138 18.5 Å closer to the iron atom and into the active site. The iron geometry is highly distorted square pyramidal, and Glu330 adopts a conformation different from that observed in the hPheOH–Fe(II)·BH4 structure, with bidentate iron coordination. BH4 binds in the second coordination sphere of the catalytic iron atom, and is displaced 2.6 Å in the direction of Glu286 and the iron atom, relative to the hPheOH–Fe(II)·BH4 structure, thus changing its hydrogen bonding network. The active-site structure of the ternary complex gives new insight into the substrate specificity of the enzyme, notably the low affinity for l-tyrosine. Furthermore, the structure has implications both for the catalytic mechanism and the molecular basis for the activation of the full-length tetrameric enzyme by its substrate. The large conformational change, moving Tyr138 from a surface position into the active site, may reflect a possible functional role for this residue.

Introduction

The non-heme iron enzyme phenylalanine hydroxylase (PheOH, phenylalanine 4-monooxygenase, EC 1.14.16.1) catalyzes the hydroxylation of the essential aromatic amino acid l-phenylalanine (l-Phe) to l-tyrosine (l-Tyr) in the presence of the specific pterin cofactor 6(R)-l-erythro-5,6,7,8-tetrahydrobiopterin (BH4) and dioxygen. The reaction is the rate-limiting step in the degradation of l-Phe to carbon dioxide and water.1 Inborn errors that reduce or destroy the activity of the enzyme are responsible for the human autosomal recessive disease phenylketonuria (PKU)/hyperphenylalaninemia (HPA). The disease causes elevated concentrations of l-Phe in the blood, which can impair the normal development of the brain and cause severe mental retardation. In most of Europe, approximately 1 in 10,000 live births reportedly has the disorder2 and more than 400 different mutations are associated with PKU/HPA.3 Most of the mutations are found in the catalytic domain3 and they demonstrate different clinical, metabolic and enzymatic phenotypes.4., 5. Recent crystallographic studies on human phenylalanine hydroxylase (hPheOH)6., 7., 8., 9. and rat phenylalanine hydroxylase (rPheOH)10 have made it possible to define the structural phenotypes of the different genotypes.11 A limitation in the assignment of the structural phenotypes has been that they have been based on the structures of catalytically inactive Fe(III) forms of the enzyme, which also lack structural information on substrate binding. Following our recently solved crystal structure of the catalytically active Fe(II) form of the truncated form ΔN1–102/ΔC428–452-hPheOH and its binary complex with the reduced pterin cofactor (BH4),12 we now present the crystal structure of a ternary complex with the substrate analogue 3-(2-thienyl)-l-alanine (THA). THA is a substrate for rPheOH13 and hPheOH14 and binds competitively to l-Phe at the active site.15 Binding of the substrate analogue also triggers a conformational change similar to that observed upon binding of l-Phe.14., 16., 17.

The structure reveals the binding sites of the pterin cofactor and the substrate under near-turnover conditions, i.e. in the absence of dioxygen, and provides new insights into the substrate specificity and catalytic mechanism of the enzyme. It shows that substrate binding triggers a substantial structural change in the catalytic domain, particularly in the active-site region. This change may represent the “epicenter” of the global conformational transition and catalytic activation that occurs in the full-length tetrameric enzyme upon substrate binding.18

Section snippets

Results

Well defined crystals of the binary hPheOH–Fe(II)·BH4 complex were treated with the substrate analogue THA by adding solid THA to the crystallization drops. The whole procedure of crystallization, post-crystallization diffusion soaking in THA, flash-cooling in liquid nitrogen and mounting of crystals was carried out anaerobically as described.12 When observed in the microscope, the crystals appeared to be unaffected by the THA soaking, but the diffraction pattern revealed a mosaicity that was

Discussion

The present crystal structure of the ternary complex hPheOH–Fe(II)·BH4·THA has given valuable new information related to the question of the substrate-binding site, the substrate specificity and the conformational transition (hysteresis) that occurs in the enzyme upon substrate binding. Furthermore, the structure has important implications for the catalytic mechanism and defines clearly the amino acid residues of the active-site crevice structure that are involved in the binding of pterin

Crystallization and data collection

Expression and purification of the double truncated mutant (ΔN1–102/ΔC428–452) of hPheOH were carried out as described.50., 51. Anaerobic co-crystallization of hPheOH–Fe(II) in complex with BH4 was undertaken essentially as described12 but with some modifications. The drops contained initial concentrations of BH4 of 5 mM, and 15 mM sodium dithionite was used as the reducing agent. After four days of growth, solid THA was added in excess to the drop (anaerobic) and left for 24 hours to allow

Acknowledgements

This work has been supported by grants from the Norwegian Research Council (NFR), the Norwegian Council on Cardiovascular Diseases, Rebergs legat, L. Meltzer Høyskolefond, the Novo Nordisk Foundation and the European Commission. We thank Ali Sepulveda Muñoz for expert technical assistance in preparing the bacterial extracts and fusion protein, and the staff of the Swiss–Norwegian Beamlines in Grenoble (France).

References (64)

  • S. Olafsdottir et al.

    The accessibility of iron at the active site of recombinant human phenylalanine hydroxylase to water as studied by 1H NMR paramagnetic relaxation. Effect of l-Phe and comparison with the rat enzyme

    J. Biol. Chem.

    (1999)
  • D. Chen et al.

    Phenylalanine hydroxylase from Chromobacterium violaceum. Uncoupled oxidation of tetrahydropterin and the role of iron in hydroxylation

    J. Biol. Chem.

    (1998)
  • B. Dworniczak et al.

    Phenylalanine hydroxylase gene: Novel missense mutation in exon 7 causing severe phenylketonuria

    Genomics

    (1991)
  • T. Senda et al.

    Three-dimensional structures of free form and two substrate complexes of an extradiol ring-cleavage type dioxygenase, the BphC enzyme from Pseudomonas sp. strain KKS102

    J. Mol. Biol.

    (1996)
  • I.G. Jennings et al.

    Essential role of the N-terminal autoreguatory sequence in the regulation of phenylalanine hydroxylase

    FEBS Letters

    (2001)
  • J. Li et al.

    Identification of three novel missense PKU mutations among Chinese

    Genomics

    (1992)
  • Y. Okano et al.

    Phenylketonuria missense mutations in the Mediterranean

    Genomics

    (1991)
  • Z. Otwinowski et al.

    Processing of X-ray diffraction data collected in oscillation mode

    Methods Enzymol.

    (1997)
  • G.J. Kleywegt et al.

    Checking your imagination: applications of the free R value

    Structure

    (1996)
  • R.M. Esnouf

    An extensively modified version of MolScript that includes greatly enhanced coloring capabilities

    J. Mol. Graph. Model

    (1997)
  • S. Kaufman

    The phenylalanine hydroxylating system

    Advan. Enzymol. Relat. Areas. Mol. Biol.

    (1993)
  • H. Bickel et al.

    Neonatal mass screening for metabolic disorders

    Eur. J. Pediatr.

    (1981)
  • C.R. Scriver et al.

    PAHdb: A locus-specific knowledgebase

    Hum. Mutat.

    (2000)
  • T. Flatmark et al.

    Structural insight into the aromatic amino acid hydroxylases and their disease-related mutant forms

    Chem. Rev.

    (1999)
  • P.J. Waters et al.

    In vitro expression analysis of mutations in phenylalanine hydroxylase: linking genotype to phenotype and structure to function

    Hum. Mutat.

    (1998)
  • H. Erlandsen et al.

    Crystal structure of the catalytic domain of human phenylalanine hydroxylase reveals the structural basis for phenylketonuria

    Nature Struct. Biol.

    (1997)
  • H. Erlandsen et al.

    Crystal structure and site-specific mutagenesis of pterin-bound human phenylalanine hydroxylase

    Biochemisty

    (2000)
  • H. Erlandsen et al.

    Crystallographic analysis of the human phenylalanine hydroxylase catalytic domain with bound catechol inhibitors at 2.0 Å resolution

    Biochemistry

    (1998)
  • B. Kobe et al.

    Structural basis of autoregulation of phenylalanine hydroxylase

    Nature Struct. Biol.

    (1999)
  • A.J. Stokka et al.

    3-(2-Thienyl)-l-alanine as a competitive substrate analogue and activator of human phenylalanine hydroxylase

  • R.J. Read

    Improved Fourier coefficients for maps using phases from partial structures with errors

    Acta Crystallog. sect. A

    (1986)
  • K.E. Goodwill et al.

    Crystal structure of tyrosine hydroxylase with bound cofactor analogue and iron at 2.3 Å resolution: self-hydroxylation of Phe300 and the pterin-binding site

    Biochemistry

    (1998)
  • Cited by (111)

    • The effects of ligand deprotonation on the binding selectivity of the phenylalanine hydroxylase active site

      2019, Computational and Theoretical Chemistry
      Citation Excerpt :

      The related Minnesota Functional, M06-L, was compared against many methods when studying gas-phase interaction energies of chemical fragments [14]. The crystal structure of the PheOH active site with bound thienylalanine, a known inhibitor, was isolated from the Protein Data Bank (PDB ID: 1KW0)[15]. Another crystal structure was published a year later with slightly higher resolution (PDB ID: 1MMK), but the main features of the structure are confirmed between both structures [16].

    • In silico analyses of the effects of a point mutation and a pharmacological chaperone on the thermal fluctuation of phenylalanine hydroxylase

      2017, Biophysical Chemistry
      Citation Excerpt :

      The three-dimensional structure of PAH is important for studying the relationship between protein structure and enzymatic function in atomic detail. The structure of wild type PAH (WT(PAH)) has been extensively investigated using X-ray crystallography [6–11] and the catalytic domain (PDB ID: 1J8T) is shown in Fig. 1a [9]. The active site of the enzyme is located around the iron ion highlighted as a yellow sphere in Fig. 1a.

    View all citing articles on Scopus
    View full text