Molecular Determinants of Substrate Specificity in Plant 5′-Methylthioadenosine Nucleosidases

https://doi.org/10.1016/j.jmb.2008.01.088Get rights and content

Abstract

5′-Methylthioadenosine (MTA)/S-adenosylhomocysteine (SAH) nucleosidase (MTAN) is essential for cellular metabolism and development in many bacterial species. While the enzyme is found in plants, plant MTANs appear to select for MTA preferentially, with little or no affinity for SAH. To understand what determines substrate specificity in this enzyme, MTAN homologues from Arabidopsis thaliana (AtMTAN1 and AtMTAN2, which are referred to as AtMTN1 and AtMTN2 in the plant literature) have been characterized kinetically. While both homologues hydrolyze MTA with comparable kinetic parameters, only AtMTAN2 shows activity towards SAH. AtMTAN2 also has higher catalytic activity towards other substrate analogues with longer 5′-substituents. The structures of apo AtMTAN1 and its complexes with the substrate- and transition-state-analogues, 5′-methylthiotubercidin and formycin A, respectively, have been determined at 2.0–1.8 Å resolution. A homology model of AtMTAN2 was generated using the AtMTAN1 structures. Comparison of the AtMTAN1 and AtMTAN2 structures reveals that only three residues in the active site differ between the two enzymes. Our analysis suggests that two of these residues, Leu181/Met168 and Phe148/Leu135 in AtMTAN1/AtMTAN2, likely account for the divergence in specificity of the enzymes. Comparison of the AtMTAN1 and available Escherichia coli MTAN (EcMTAN) structures suggests that a combination of differences in the 5′-alkylthio binding region and reduced conformational flexibility in the AtMTAN1 active site likely contribute to its reduced efficiency in binding substrate analogues with longer 5′-substituents. In addition, in contrast to EcMTAN, the active site of AtMTAN1 remains solvated in its ligand-bound forms. As the apparent pKa of an amino acid depends on its local environment, the putative catalytic acid Asp225 in AtMTAN1 may not be protonated at physiological pH and this suggests the transition state of AtMTAN1, like human MTA phosphorylase and Streptococcus pneumoniae MTAN, may be different from that found in EcMTAN.

Introduction

Methionine-recycling pathways have evolved in eukaryotes and most bacterial species because the amino acid is often limited in bioavailability and is energetically costly to synthesize de novo.1, 2 Methionine is essential for cellular metabolism and development, and accretion of the nucleoside metabolites, 5′-methylthioadenosine (MTA) and S-adenosylhomocysteine (SAH) (Fig. 1), negatively regulates transmethylation reactions and quorum sensing, and polyamine biosynthesis, respectively.3, 4 In mammals, MTA is converted to 5-methylthioribose 1-phosphate (MTR 1-P) and adenine by MTA phosphorylase (MTAP),5 while SAH is metabolized to homocysteine and adenosine by SAH hydrolase. MTAP is not found in plants and many bacterial species. In its place, MTA nucleosidase (MTAN) is used to hydrolyze MTA, and in some cases SAH. MTAN catalyzes the irreversible cleavage of the N-glycosidic bond of MTA and SAH to form adenine, and methylthioribose (MTR) and S-ribosylhomocysteine (SRH), respectively.6, 7, 8, 9, 10 Adenine enters the nucleotide biosynthetic pathway, while MTR, depending on the species, is either excreted or phosphorylated at the C1-hydroxyl by MTR kinase to form MTR 1-P. MTR 1-P is subsequently isomerized to 5-methylthioribulose 1-phosphate by MTR 1-P epimerase.11 The ensuing downstream steps in the methionine recycling pathway appear to vary somewhat between species, depending largely on whether the environment is aerobic or anaerobic, and this has been reviewed in detail by Sekowska et al. 11 The last step is common to all species and involves an amino group transfer from a number of potential amino acid donors to 2-keto-4-methylthiobutyrate to form methionine.11

While bacterial MTANs show dual-substrate specificity for both MTA and SAH, plant MTANs (typically referred to in the plant literature as MTN) exhibit a distinct preference for MTA with little or no activity towards SAH.12, 13, 14 Like mammals, plants appear to utilize SAH hydrolase to metabolize SAH.15, 16 In addition, the methionine-recycling pathway in plants is conjugated to ethylene synthesis.17 In the first step of ethylene synthesis, 1-aminocyclopropane-1-carboxylate (ACC) is produced from S-adenosylmethionine (AdoMet) with MTA released as a side product.18 Accumulation of MTA inhibits ethylene production.17, 19 Regulation of the methionine recycling pathway by ethylene is observed in plants that synthesize high levels of ethylene for prolonged periods.17 These include plants with seasonal fruit ripening, such as tomatoes, and those that are submerged in water for extended periods, such as rice. In these species, MTAN is suggested to be important in seed germination and fruit ripening.16

The metabolic variation between the bacterial and plant MTAN's suggests structural differences that affect substrate recognition and binding. The structures and mutational analysis of a representative bacterial MTAN, Escherichia coli MTAN (EcMTAN), have identified the catalytic residues,20, 21, 22 and kinetic isotope effect studies23 support the proposed SN1-type catalytic mechanism, which is thought to begin with the donation of a proton from the catalytic acid (Asp197) to N7 of the substrate. The protonated adenine base withdraws electrons from the ribose group, leading to elongation of the ribosidic bond and the development of a partial positive charge on the ribose group that is stabilized by the catalytic water molecule (WAT) and Glu174. Glu12 abstracts a proton from the catalytic water molecule, which then attacks C1′ of the oxocarbenium-like intermediate to form the products.

In contrast, structural and kinetic studies on the plant enzyme(s) are more limited. The first structure of a representative plant MTAN from Arabidopsis thaliana (AtMTAN1) was determined recently in complex with the product, adenine (PDB code 2H8G).24 To better understand the mechanisms of substrate-recognition and catalysis, we have examined the substrate specificities and activity of the two MTAN homologues present in A. thaliana, and have determined the structure of AtMTAN1 in its apo form, and in complex with the substrate-analogue, 5′-methylthiotubercidin (MTT), and the transition-state analogue, formycin A (FMA) (Fig. 1). A structural homology model of AtMTAN2 has been constructed, which highlights differences in the active sites of the two isozymes that are likely critical for the substrate specificity observed. Analysis of the ligand-bound structures of EcMTAN and AtMTAN1 further suggest that differences in the dynamics, size and specificity of the active sites may account for the divergence in substrate selectivity between the bacterial and plant enzymes.

Section snippets

Catalytic activity and substrate specificity of AtMTAN1 and AtMTAN2

In preparation for kinetic analysis of AtMTAN1 and AtMTAN2 enzymes, the optimal pH and buffer conditions for activity were examined. Distinct pH optima were apparent for the two homologues (Fig. 2a). The optimal pH for AtMTAN1 is 8, while AtMTAN2 functions best at pH 6. When the activity was assayed at pH 7, the choice of buffer also influenced the activity (Fig. 2b). Imidazole buffer is clearly the best buffer for both enzymes. Phosphate and Hepes buffers have been commonly used in MTAN

Expression and purification of AtMTAN1 and AtMTAN2

The A. thaliana MTAN1 (At4g38800) cDNA was cloned from A. thaliana genomic DNA into the pET28a+ vector (Novagen) at the NdeI and HindIII sites to allow over-expression of the enzyme with a cleavable N-terminal His6 tag. AtMTAN1 (molecular mass 28,451 Da) was over-expressed in BL21 Codon+ cells by growing in Luria-Bertani broth (LB) to exponential phase and then inducing with 1 mM isopropyl thiogalactopyranoside (IPTG) for 4 h at 37 °C. The cells were harvested by centrifugation in a JA-10 rotor

Acknowledgements

The authors thank Dr G. David Smith for his help with data analysis and Dr Li Zhang at the Advanced Protein Technology Centre at The Hospital for Sick Children for his help with the mass spectrometry. This work is supported by research grants from the Canadian Institutes of Health Research (CIHR) to P.L.H. (#43998), the Natural Sciences and Engineering Research Council of Canada (NSERC) to B.A.M., and the USDA (#02-0047) and NIH NCRR (#P20RR016454) to K.A.C. P.L.H. is the recipient of a Canada

References (58)

  • A.J. Ferro et al.

    Kinetic properties and the effect of substrate analogues on 5′-methylthioadenosine nucleosidase from Escherichia coli

    Biochim. Biophys. Acta

    (1976)
  • J.J. Dwyer et al.

    High apparent dielectric constants in the interior of a protein reflect water penetration

    Biophys. J.

    (2000)
  • D.D. Boehr et al.

    Analysis of the pi-pi stacking interactions between the aminoglycoside antibiotic kinase APH(3′)-IIIa and its nucleotide ligands

    Chem. Biol.

    (2002)
  • K.G. Wagner et al.

    Dynamics of nucleotides in plants studied on a cellular basis

    Int. Rev. Cytol.

    (1992)
  • V. Singh et al.

    Femtomolar transition state analogue inhibitors of 5′-methylthioadenosine/S-adenosylhomocysteine nucleosidase from Escherichia coli

    J. Biol. Chem.

    (2005)
  • S.F. Altschul et al.

    Basic local alignment search tool

    J. Mol. Biol.

    (1990)
  • J.E. Lee et al.

    Structure of E. coli 5′-methylthioadenosine/S-adenosylhomocysteine nucleosidase reveals similarity to the purine nucleoside phosphorylases

    Structure

    (2001)
  • A.J. Cooper

    Biochemistry of sulfur-containing amino acids

    Annu. Rev. Biochem.

    (1983)
  • J.R. Sufrin et al.

    Methionine recycling pathways and antimalarial drug design

    Antimicrob. Agents Chemother.

    (1995)
  • F. Della Ragione et al.

    Escherichia coli S-adenosylhomocysteine/5′-methylthioadenosine nucleosidase. Purification, substrate specificity and mechanism of action

    Biochem. J.

    (1985)
  • V. Zappia et al.

    Biological Methylation and Drug Design: Experimental and Clinical Roles of S-Adenosylmethionine

    (1985)
  • S. Shimizu et al.

    Enzymatic process for the production of pharmacologically useful nucleosides

    Ann. NY Acad. Sci.

    (1988)
  • A. Sekowska et al.

    Bacterial variations on the methionine salvage pathway

    BMC Microbiol.

    (2004)
  • M.M. Kushad et al.

    5′-Methylthioadenosine nucleosidase and 5-methylthioribose kinase activities and ethylene production during tomato fruit development and ripening

    Plant Physiol.

    (1985)
  • G. Rzewuski et al.

    OsMTN encodes a 5′-methylthioadenosine nucleosidase that is up-regulated during submergence-induced ethylene synthesis in rice (Oryza sativa L.)

    J. Expt. Bot.

    (2007)
  • A. Guranowski et al.

    Adenosylhomocysteinase from yellow lupin seeds. Purification and properties

    Eur. J. Biochem.

    (1977)
  • A.B. Guranowski et al.

    5′-Methylthioadenosine nucleosidase. Purification and characterization of the enzyme from Lupinus luteus seeds

    Eur. J. Biochem.

    (1981)
  • K. Burstenbinder et al.

    The role of methionine recycling for ethylene synthesis in Arabidopsis

    Plant J.

    (2007)
  • M. Sauter et al.

    The immediate-early ethylene response gene OsARD1 encodes an acireductone dioxygenase involved in recycling of the ethylene precursor S-adenosylmethionine

    Plant J.

    (2005)
  • Cited by (15)

    • Purine salvage in plants

      2018, Phytochemistry
    • Mechanism of substrate specificity in 5'-methylthioadenosine/S-adenosylhomocysteine nucleosidases

      2011, Journal of Structural Biology
      Citation Excerpt :

      Titration of SAH or MTA into EcMTAN-Glu174Gln leads to favourable enthalpic and unfavourable entropic changes (Table 1, Fig. 6e and f). As previous modeling had suggested that SAH could not be accommodated in the confined active site of the plant enzyme (Siu et al., 2008a) but our ITC results demonstrate unequivocally that AtMTAN1 can bind SAH, the crystal structure of AtMTAN1 in complex with SAH was determined (Table 2). Analysis by Molprobity shows that 98% of residues are in the favored regions and no residues fall in the disallowed regions (Lovell et al., 2003).

    • Biochemical and structural characterization of 5′-methylthioadenosine nucleosidases from Arabidopsis thaliana

      2009, Biochemical and Biophysical Research Communications
      Citation Excerpt :

      This is usually interpreted as AtMTAN1 being able to hydrolyze smaller 5′-methyl, 5′-ethyl, 5′-n-butyl, and 5′-isobutylthioadenosine, but not SAH [11]. Furthermore, this collision involving Phe148′ is predictable when compared with adenine-free AtMTAN1, even though it has more opened conformation [21]. Therefore, the SAH model demonstrates the substitution of this position in AtMTAN2 results in the dual activity of MTA and SAH (Figs. 3E, F and 4A).

    View all citing articles on Scopus
    1

    Present address: J. E. Lee, Department of Immunology, The Scripps Research Institute, 10550 North Torrey Pines Road, La Jolla, CA, 92037, USA.

    2

    Present address: Global Pharmaceutical Services, Inc., 14A Florence Street, East Patchogue, NY 11772-6186, USA.

    View full text