Coordination abilities of the 1–16 and 1–28 fragments of β-amyloid peptide towards copper(II) ions: a combined potentiometric and spectroscopic study

https://doi.org/10.1016/S0162-0134(03)00128-4Get rights and content

Abstract

Stoichiometry, stability constants and solution structures of the copper(II) complexes of the (1–16H), (1–28H), (1–16M), (1–28M), (Ac-1–16H) and (Ac-1–16M) fragments of human (H) and mouse (M) β-amyloid peptide were determined in aqueous solution in the pH range 2.5–10.5. The potentiometric and spectroscopic data (UV–Vis, CD, EPR) show that acetylation of the amino terminal group induces significant changes in the coordination properties of the (Ac-1–16H) and (Ac-1–16M) peptides compared to the (1–16H) and (1–16M) fragments, respectively. The (Ac-1–16H) peptide forms the 3N {NIm6, NIm13, NIm14} complex in a wide pH range (5–8), while for the (Ac-1–16M) fragment the 2N {NIm6, NIm14} complex in the pH range 5–7 is suggested. At higher pH values sequential amide nitrogens are deprotonated and coordinated to copper(II) ions. The N-terminal amino group of the (1–16) and (1–28) fragments of human and mouse β-amyloid peptide takes part in the coordination of the metal ion, although, at pH above 9 the complexes with the 4N {NIm, 3N} coordination mode are formed. The phenolate –OH group of the Tyr10 residue of the human fragments does not coordinate to the metal ion.

Introduction

Patients with Alzheimer’s disease (AD) contain large quantities of insoluble amyloid plaques that are primarily found in brain tissue. A major component of amyloid plaque is the β-peptide (Aβ) [1], [2], a small (39–43 amino acids) polypeptide with heterogenous termini that is generated from the cleavage of a larger amyloid precursor protein (APP) [3], [4]. Amyloid deposition is likely a critical step in the neurodegenerative processes associated with AD [5]. Amyloid plaques are invariably associated with areas of nerve death, and the injection of synthetic β-peptides directly into rat brain produced cytotoxic effects [6]. It has also recently been established that soluble extracellular β-peptide is normally produced in cultured cells and human biological fluids [7], [8].

There is an emerging consensus in the literature to indicate that the homeostases of zinc, copper and iron are significantly altered in AD brain tissue [9]. Evidence for abnormal Cu homeostasis in AD includes a 2.2-fold increase in the concentration of cerebrospinal fluid (CSF) Cu [10], and an accompanying increase in ceruloplasmin in the brain and CSF of AD patients [11]. Aβ is resolubilized and extracted from postmortem AD and non-AD control brains using metal chelators. High-affinity Cu/Zn/Fe chelators markedly enhanced the resolubilization of AD deposits from postmortem AD and non-AD brain samples [12]. Aβ in vitro binds metal ions, including Zn2+, Cu2+, and Fe3+, inducing peptide aggregation that may be reversed by treatment with chelators such as ethylenediaminetetraacetic acid (EDTA) [13], [14]. Rats and mice do not develop amyloid [15], probably due to the three amino acid substitutions in their homologue of Aβ (Arg5→Gly, Tyr10→Phe, and His13→Arg) [16]. These changes have been shown to alter the structure and properties of the Aβ peptides [17], [18], [19] as well as the processing of APP. These sequence alternations might therefore be responsible for the virtual absence of Aβ deposits in normal or aged rodent brain [19]. In vitro it has been shown that, compared with human Aβ, rat Aβ binds Zn2+ and Cu2+ less avidly [13], that the coordination of Cu2+ or Fe3+ does not induce redox chemical reactions, and that limited reactive oxygen species are generated [20]. Through the use of synthetic peptide corresponding to the first 28 residues of human Aβ, rat Aβ, and single-residue variations, Liu et al. [21] reported a key role for His13 in the zinc interaction. Substitution His/Ala of the potential histidine ligands in Aβ suggests that residues His13 and His14 represent a critical domain for zinc interaction [22]. On the basis of nuclear magnetic resonance (NMR) and electron paramagnetic resonance (EPR) experiments, it has been proposed that Aβ binds Cu2+ via three His residues (His6, His13 and His14) and an oxygen ligand, probably Tyr10 [23].

Peptide complexes with metal ions have been extensively studied in order to mimic specific metalloprotein structures and functions [24], [25]. A prominent role of some amino acid side chains and, in particular, the imidazole moieties of histidine in the binding of metal ion has been observed in proteins and in natural [26] or synthetic peptides [27], [28]. The studies of metal ion–peptide complex formation reactions indicate that solution equilibria are rather complicated especially in the presence of coordinating side chains. For the determination of formation constants of peptide complexes potentiometry is usually used, but for the justification of potentiometric results the combined application of various spectroscopic techniques: UV–visible (UV–Vis), EPR, NMR and/or circular dichroism (CD) spectroscopies is required.

The interaction of copper(II) ions with the (1–6), (1–9), (1–10) [29] (contain one His6 residue) and (11–16) [30], (11–20), (11–28) [31] human (contain two His13 and His14 residues) and mouse (one His14 residue) fragments of β-amyloid peptide was studied. The differences of the binding mode of human and mouse fragments to copper(II) ions and stability constants for the complexes formed were determined.

The present paper reports the results of combined spectroscopic and potentiometric studies on the copper(II) complexes of the (1–16) and (1–28) human (three His6, His13, His14 residues) and mouse (two His6, His14 residues) fragments of β-amyloid peptide. The β-amyloid fragments studied here are: human peptide (1–16H), H-Asp–Ala–Glu–Phe–Arg–His–Asp–Ser–Gly–Tyr–Glu–Val–His–His–Gln–Lys-NH2, DAEFRHDSGYEVHHQK-NH2 and mouse peptide (1–16M), H-Asp–Ala–Glu–Phe–Gly–His–Asp–Ser–Gly–Phe–Glu–Val–Arg–His–Gln–Lys-NH2, DAEFGHDSGFEVRHQK-NH2; the fragments (1–28), human peptide (1–28H), H-Asp–Ala–Glu–Phe–Arg–His–Asp–Ser–Gly–Tyr–Glu–Val–His–His–Gln–Lys–Leu–Val–Phe–Phe–Ala–Glu–Asp–Val–Gly–Ser–Asn–Lys-NH2, DAEFRHDSGYEVHHQKLVFFAEDVGSNK-NH2 and mouse peptide (1–28M) H-Asp–Ala–Glu–Phe–Gly–His–Asp–Ser–Gly–Phe–Glu–Val–Arg–His–Gln–Lys–Leu–Val–Phe–Phe–Ala–Glu–Asp–Val–Gly–Ser–Asn–Lys-NH2, DAEFGHDSGFEVRHQKLVFFAEDVGSNK-NH2. To determine the involvement of N-terminal amino group in the metal ion coordination, the analogues with blocked N-terminal amino group (by acetylation) of the human and mouse fragments (1–16): human (Ac-1–16H), Ac-Asp–Ala–Glu–Phe–Arg–His–Asp–Ser–Gly–Tyr–Glu–Val–His–His–Gln–Lys-NH2, Ac-DAEFRHDSGYEVHHQK-NH2; mouse (Ac-1–16M), Ac-Asp–Ala–Glu–Phe–Gly–His–Asp–Ser–Gly–Phe–Glu–Val–Arg–His–Gln–Lys-NH2, Ac-DAEFGHDSGFEVRHQK-NH2 were also studied. These fragments contain the complete bonding site of Aβ and, with the exception of the C-terminal carboxylate, are representative of the full-length peptide. Therefore, the 1–16 and 1–28 fragments are considered as valid models to examine the contribution of the key histidine residues (His6, His14 in mouse and His6, His13, His14 in human fragments) to the Aβ–Cu2+ interaction. This study was performed in order to examine the difference of the binding ability of the human and mouse fragments, especially the effect of the substitutions Arg with Gly, Tyr with Phe and His with Arg residues in positions 5, 10 and 13, respectively, on the formation of complexes with Cu2+ ions. The (17–28) fragment of β-amyloid peptide does not contain any additional bonding site to copper(II) ions, however, the influence of these residues on the stability of the complexes formed may be estimated.

Section snippets

Peptide synthesis, purification and characterization

Syntheses of peptide amides: fragments of human (H) and mouse (M) β-amyloid peptide (Aβ): (1–16H), (1–16M), (Ac-1–16H), (Ac-1–16M), (1–28H) and (1–28M) were synthesized by solid phase methodology and purified according to the procedure described earlier [29].

The purity of the peptides was assessed by reversed-phase high-performance liquid chromatography (RP-HPLC) [29] and fast atom bombardment mass spectrometry (FAB-MS) or matrix-assisted laser desorption/ionization (MALDI). Purity of all the

Protonation constants

Protonation constants (log β, log K values) for the peptides studied and comparable ligands are given in Table 1, Table 2. The first protonation constant, log βHL, for the peptides studied is the ε-amino group of the Lys residue. For the (1–16M), (1–28H), (1–28M), (Ac-1–16H) and (Ac-1–16M) peptide fragments, the protonation values of the ε-amino nitrogen of the lysine residue (log K=10.46–9.64) agree well with literature data (Table 1, Table 2) [39], [40], [41]. The (1–28) fragments of human

Conclusions

The fragments of β-amyloid peptide studied here did not reduce copper(II) ions in accordance with the observation reported earlier [20]. The imidazole nitrogen of the histidine residue acts as an anchoring bonding site for the human and mouse fragments, (Ac-1–16H) and (Ac-1–16M). In a wide pH range, Cu2+ is bound to the (Ac-1–16M) peptide through imidazole nitrogens on both of its histidine residues (His6, His14), while, for the (Ac-1–16H) fragment the 3N complex with {NIm6, NIm13, NIm14}

Acknowledgements

This work was supported by the Polish State Committee for Scientific Research (KBN 3 T09A 06918).

References (59)

  • G.G. Glenner et al.

    Biochem. Biophys. Res.

    (1984)
  • D.J. Selkoe

    Neuron

    (1991)
  • D.A. Loeffler et al.

    Brain Res.

    (1996)
  • R.A. Cherny et al.

    J. Biol. Chem.

    (1999)
  • C.S. Atwood et al.

    J. Biol. Chem.

    (1998)
  • X. Huang et al.

    J. Biol. Chem.

    (1997)
  • T. Dyrks et al.

    FEBS Lett.

    (1993)
  • E.M. Johnstone et al.

    Mol. Brain Res.

    (1991)
  • X. Huang et al.

    J. Biol. Chem.

    (1999)
  • C.C. Curtain et al.

    J. Biol. Chem.

    (2001)
  • G. Xing et al.

    Curr. Opin. Chem. Biol. Rev.

    (2001)
  • P. Tsiveriotis et al.

    Co-ord. Chem. Rev.

    (1999)
  • T. Kowalik-Jankowska et al.

    J. Inorg. Biochem.

    (2001)
  • T. Kowalik-Jankowska et al.

    J. Inorg. Biochem.

    (2002)
  • H. Irving et al.

    Anal. Chim. Acta

    (1967)
  • C.J. Barrow et al.

    J. Mol. Biol.

    (1992)
  • C. Hilbich et al.

    J. Mol. Biol.

    (1991)
  • M.A. Zoroddu et al.

    Biochim. Biophys. Acta

    (2000)
  • T. Kowalik-Jankowska et al.

    J. Inorg. Biochem.

    (1999)
  • I. Sovago et al.

    Inorg. Chim. Acta

    (1984)
  • H. Kozłowski et al.

    Coord. Chem. Rev.

    (1999)
  • D. Sanna et al.

    Polyhedron

    (2001)
  • R. Prados et al.

    Bioinorg. Chem.

    (1975)
  • M.A. Zoroddu et al.

    J. Inorg. Biochem.

    (2001)
  • M. Casolaro et al.

    J. Inorg. Biochem.

    (2002)
  • W. Bal et al.

    J. Inorg. Biochem.

    (1993)
  • C.L. Masters et al.

    Proc. Natl. Acad. Sci. USA

    (1985)
  • T.E. Golde et al.

    Science

    (1992)
  • S.S. Sisodia

    Proc. Natl. Acad. Sci. USA

    (1992)
  • Cited by (0)

    View full text