Dual catalytic role of the metal ion in nickel-assisted peptide bond hydrolysis

https://doi.org/10.1016/j.jinorgbio.2014.03.008Get rights and content

Abstract

In our previous research we demonstrated the sequence specific peptide bond hydrolysis of the R1-(Ser/Thr)-Xaa-His-Zaa-R2 in the presence of Ni(II) ions. The molecular mechanism of this reaction includes an N–O acyl shift of the R1 group from the Ser/Thr amine to the side chain hydroxyl group of this amino acid. The proposed role of the Ni(II) ion is to establish favorable geometry of the reacting groups. In this work we aimed to find out whether the crucial step of this reaction – the formation of the intermediate ester – is reversible. For this purpose we synthesized the test peptide Ac-QAASSHEQA-am, isolated and purified its intermediate ester under acidic conditions, and reacted it, alone, or in the presence of Ni(II) or Cu(II) ions at pH 8.2. We found that in the absence of either metal ion the ester was quickly and quantitatively (irreversibly) rearranged to the original peptide. Such reaction was prevented by either metal ion. Using Cu(II) ions as CD spectroscopic probe we showed that the metal binding structures of the ester and the final amine are practically identical. Molecular calculations of Ni(II) complexes indicated the presence of steric strain in the substrate, distorting the complex structure from planarity, and the absence of steric strain in the reaction products. These results demonstrated the dual catalytic role of the Ni(II) ion in this mechanism. Ni(II) facilitates the acyl shift by setting the peptide geometry, and prevents the reversal of the acyl shift, by stabilizing subsequent reaction products.

Graphical abstract

The dual catalytic role of Ni(II) ions consists of (i) arranging steric strain on the susceptible peptide bond, and (ii) stabilizing the intermediate and final reaction products.

  1. Download : Download full-size image

Introduction

Nickel and its compounds are toxic to humans, via inhalatory, dermal and alimentary routes. Contact allergy and respiratory cancers are two major health hazards related to nickel exposure, under environmental and occupational conditions, respectively. While it has become clear that Ni(II) ions are the ultimate toxic species, the exact molecular mechanisms of toxicity are the subject of dispute [1], [2], [3], [4], [5], [6], [7], [8], [9].

Nickel(II) dependent peptide bond hydrolysis (in brief, nickel-assisted hydrolysis) is a promising molecular concept for several aspects of nickel toxicity. This sequence specific reaction may impair function of important intracellular proteins, such as zinc finger transcription factors, at the same time yielding stable and reactive Ni(II) complexes, capable of damaging DNA and proteins [10], [11], [12], [13], [14], [15], [16]. Peptides susceptible to nickel hydrolysis have a general sequence R1-(Ser/Thr)-Xaa-His-Zaa-R2 (where Xaa and Zaa are any amino acids except of Pro and Cys as Xaa, and Cys as Zaa, and R1 and R2 are nonspecific N-terminal and C-terminal peptide sequences). The hydrolysis occurs, with an absolute selectivity, at the peptide bond preceding Ser/Thr. Scheme 1 illustrates key steps of the reaction [13], [14]. The reaction is enabled by the formation of a square-planar complex in which the Ni(II) ion is bonded by the imidazole nitrogen of the His residue and three preceding amide nitrogens. The first crucial step of the reaction is the N–O acyl shift of the carbonyl group of to hydrolysable R1-Ser/Thr peptide bond to the Ser/Thr hydroxyl group. This step follows an apparent 1st order kinetic regime (k1 in Scheme 1). The resulting ester intermediate product (IP) is unstable in water solution, hydrolyzing into two peptides, also according to the apparent 1st order kinetics (k2 in Scheme 1). The C-terminal peptide product of this reaction step remains bonded to the Ni(II) ion. Therefore, the overall reaction is stoichiometric (1:1), rather than catalytic, with respect to Ni(II). The reaction rate is strongly dependent on the spatial orientations of side chains of amino acids belonging to the Ni(II) coordination site [17], and is strongly (up to a factor of 200) enhanced in the presence of bulky residues C-terminal to the His residue [18]. These facts confirm the notion that crucial role of the Ni(II) ion is largely geometrical. The square-planar coordination mode induces strain on the hydrolysable peptide bond, modulated by sterical crowding of the above mentioned side chains.

In the course of our studies we noted that the stability of IP varies significantly among various peptide sequences studied. In all cases the overall reaction was irreversible, resulting in the full conversion of the substrate into final products. This was assured by the irreversible character of the 2nd step of the reaction, the well-known acid- or base-catalyzed ester hydrolysis [19]. Nevertheless, the data we collected so far did not provide information on the reversibility of the 1st step, the IP formation. This issue can be quite important for many purposes, including the practical application of the reaction for purification of recombinant proteins [20].

We decided to study this issue using a peptide of the sequence Ac-QAASSHEQA-am (Ac- denotes the N-terminal acetylation, and -am the C-terminal amidation of the peptide chain), which yields a long-lived and easy to extract IP. This peptide was chosen on the basis of our current research on hydrolysis of filaggrin, a protein involved in the process of outer skin keratinization [21].

Section snippets

Reagents

N-α-9-Fluorenylmethyloxycarbonyl (F-moc) amino acids were purchased from Sigma-Aldrich, and Fluka Co. Trifluoroacetic acid (TFA), piperidine, O-(Benzotriazol-1-yl)-N,N,N′,N′-tetramethyluronium hexafluorophosphate (HBTU), triisopropylsilane (TIS), N,N-diisopropylethylamine (DIEA) and nickel(II) nitrate hexahydrate were obtained from Sigma-Aldrich. TentaGel® S RAM resin was obtained from Rapp Polymer Inc. Acetonitrile (HPLC grade) was obtained from Rathburn Chemicals Ltd. Pure sodium hydroxide

Ni(II) dependent Ac-QAASSHEQA-am peptide hydrolysis

The first step of our research was the characterization of Ni(II) binding by the Ac-QAASSHEQA-am peptide and its susceptibility to nickel-assisted hydrolysis. We started with the UV–vis spectroscopic pH titration of the equimolar Ac-QAASSHEQA-am/Ni(II) system. The spectra and the resulting titration curve are presented in Fig. 1. The pKa value for the formation of the square-planar Ni(II) complex, determined from the pH dependence of the absorption band at 461 nm was 8.31 ± 0.01, with a high

Reversibility of hydrolysis

Our experiments started with the confirmation that Ac-QAASSHEQA-am is a suitable hydrolysis substrate to study the properties of the IP. As shown in Fig. 1, This peptide forms a square planar Ni(II) complex in the alkaline pH range (pKa value of 8.31), as confirmed by parameters of the d–d band (λmax = 461 nm, ε = 92 M 1 cm 1) [14], [35], [36]. The pKa value, somewhat lower than typical for similar peptides, containing a His residue in the middle of the peptide chain (ca. 9), is responsible for a high

Conclusion

In the experiments described above we demonstrated that the key steps of nickel-assisted peptide bond hydrolysis are irreversible The process of reverse (O–N) acyl shift from the metal free IP, in the amine-deprotonated form to the peptidic substrate of hydrolysis is much faster than the IP formation in the presence of Ni(II) ions. The direction of the process from the peptide through IP to the final hydrolyzed products (Scheme 2) is ensured by the engagement of the amine lone pair in the

Abbreviations

    DIPEA

    N,N-diisopropylethylamine, or Hünig's base

    DMF

    dimethylformamide

    ESI-MS

    electrospray ionization mass spectrometry or electrospray mass spectrometry

    HEPES

    4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid or 2-[4-(2-hydroxyethyl)piperazin-1-yl]ethanesulfonic acid

    HBTU

    O-(Benzotriazol-1-yl)-N,N,N′,N′-tetramethyluronium hexafluorophosphate

    IP

    intermediate product of hydrolysis reaction

    P

    final C-terminal product of the hydrolysis reaction

    S

    substrate of the hydrolysis reaction

    TFA

    trifluoroacetic acid

    TIS

Acknowledgments

This research was supported by the project “Metal-dependent peptide hydrolysis. Tools and mechanisms for biotechnology, toxicology and supramolecular chemistry” carried out as part of the Foundation for Polish Science TEAM program, co-financed from the European Regional Development Fund resources within the framework of Operational Program Innovative Economy. The equipment used was sponsored in part by the Centre for Preclinical Research and Technology (CePT), a project co-sponsored by European

References (48)

  • K.S. Kasprzak et al.

    Mutat. Res.

    (2003)
  • E. Kurowska et al.

    Adv. Mol. Toxicol.

    (2010)
  • A. Hartwig

    Free Radic. Biol. Med.

    (2013)
  • A.M. Protas et al.

    J. Inorg. Biochem.

    (2013)
  • L. Acerenza et al.

    Biochem. Biophys. Acta

    (1997)
  • H. Kozlowski et al.

    Coord. Chem. Rev.

    (1999)
  • International Agency for Research on Cancer

    IARC monographs on the evaluation of carcinogenic risk to humans

    (1990)
  • K.S. Kasprzak et al.

    J. Environ. Monit.

    (2003)
  • J.P. Thyssen et al.

    Chem. Res. Toxicol.

    (2010)
  • A. Arita et al.

    Metallomics

    (2009)
  • K.S. Kasprzak et al.
  • W. Bal et al.

    Met. Ions Life Sci.

    (2011)
  • W. Bal et al.

    Chem. Res. Toxicol.

    (1998)
  • W. Bal et al.

    Chem. Res. Toxicol.

    (2000)
  • A.A. Karaczyn et al.

    Chem. Res. Toxicol.

    (2003)
  • A. Krężel et al.

    J. Am. Chem. Soc.

    (2010)
  • E. Kopera et al.

    Inorg. Chem.

    (2010)
  • E. Kurowska et al.

    Metallomics

    (2011)
  • R. Liang et al.

    Carcinogenesis

    (1999)
  • H.H. Ariani et al.

    Inorg. Chem.

    (2013)
  • R.A. Cox

    Int. J. Mol. Sci.

    (2011)
  • E. Kopera et al.

    PLoS One

    (2012)
  • E.I. Podobas et al.
  • W. Chan et al.

    Fmoc Solid Phase Peptide Synthesis: A Practical Approach

    (2000)
  • Cited by (10)

    • Modulation of the catalytic activity of a metallonuclease by tagging with oligohistidine

      2020, Journal of Inorganic Biochemistry
      Citation Excerpt :

      The self-eliminating BsmBI restriction enzyme sites allowed for the insertion of the gene of the target protein between the atg “start” codon and the nucleotide sequence encoding for the (Ser/Thr)-X-His-X ((S/T)XHX) segment instead of the stop codon [1]. The latter sequence can be hydrolyzed by Ni(II) ions at the N-terminus of the Ser/Thr amino acid [2–7], and in this way, any kind of C-terminal affinity tags can be easily removed from the expressed protein without leaving any additional amino acids at the termini of the native protein sequence. Such peptide bond cleavage initiated by Ni(II) ions has already found application in the affinity tag removal from fusion proteins [8–10].

    • Metal assisted peptide bond hydrolysis: Chemistry, biotechnology and toxicological implications

      2016, Coordination Chemistry Reviews
      Citation Excerpt :

      In addition to its indirect structural role (imposing the strain on the hydrolysable peptide bond via the square-planar complex structure and by placing the Ser/Thr side chain in a position enabling its role as the acyl group acceptor), it has a direct function, destabilizing the R1–Ser peptide bond directly through its peptide nitrogen coordination. The engagement of the lone pair of the Ser/Thr amine by Ni(II) in the complex with the intermediate ester prevents its reversal to the thermodynamically more stable amide [110]. Comparative studies of model oligopeptides derived from proteins, and of their larger domains, or even whole proteins revealed that the hydrolysis rate may be increased or decreased, compared to a prediction based simply on Xaa and Zaa substitutions, as a result of a specific conformational context of the reactive complex provided by secondary protein structures (loops, helices and β-sheets [107,111,112].

    • Reactions of Carboxylic, Phosphoric, and Sulfonic Acids and their Derivatives

      2017, Organic Reaction Mechanisms 2014: An Annual Survey Covering the Literature Dated January to December 2014
    View all citing articles on Scopus
    View full text