The nature of Cu bonding to natural organic matter

https://doi.org/10.1016/j.gca.2010.01.027Get rights and content

Abstract

Copper biogeochemistry is largely controlled by its bonding to natural organic matter (NOM) for reasons not well understood. Using XANES and EXAFS spectroscopy, along with supporting thermodynamic equilibrium calculations and structural and steric considerations, we show evidence at pH 4.5 and 5.5 for a five-membered Cu(malate)2-like ring chelate at 100–300 ppm Cu concentration, and a six-membered Cu(malonate))1–2-like ring chelate at higher concentration. A “structure fingerprint” is defined for the 5.0–7.0 Å−1 EXAFS region which is indicative of the ring size and number (i.e., mono- vs. bis-chelate), and the distance and bonding of axial oxygens (Oax) perpendicular to the chelate plane formed by the four equatorial oxygens (Oeq) at 1.94 Å. The stronger malate-type chelate is a C4 dicarboxylate, and the weaker malonate-type chelate a C3 dicarboxylate. The malate-type chelate owes its superior binding strength to an –OH for –H substitution on the α carbon, thus offering additional binding possibilities. The two new model structures are consistent with the majority of carboxyl groups being clustered and α-OH substitutions common in NOM, as shown by recent infrared and NMR studies. The high affinity of NOM for Cu(II) is explained by the abundance and geometrical fit of the two types of structures to the size of the equatorial plane of Cu(II). The weaker binding abilities of functionalized aromatic rings also is explained, as malate-type and malonate-type structures are present only on aliphatic chains. For example, salicylate is a monocarboxylate which forms an unfavorable six-membered chelate, because the OH substitution is in the β position. Similarly, phthalate is a dicarboxylate forming a highly strained seven-membered chelate.

Five-membered Cu(II) chelates can be anchored by a thiol α-SH substituent instead of an alcohol α-OH, as in thio-carboxylic acids. This type of chelate is seldom present in NOM, but forms rapidly when Cu(II) is photoreduced to Cu(I) at room temperature under the X-ray beam. When the sample is wet, exposure to the beam can reduce Cu(II) to Cu(0). Chelates with an α-amino substituent were not detected, suggesting that malate-like α-OH dicarboxylates are stronger ligands than amino acids at acidic pH, in agreement with the strong electronegativity of the COOH clusters. However, aminocarboxylate Cu(II) chelates may form after saturation of the strongest sites or at circumneutral pH, and could be observed in NOM fractions enriched in proteinaceous material. Overall, our results support the following propositions:

  • (1)

    The most stable Cu–NOM chelates at acidic pH are formed with closely-spaced carboxyl groups and hydroxyl donors in the α-position; oxalate-type ring chelates are not observed.

  • (2)

    Cu(II) bonds the four equatorial oxygens to the heuristic distance of 1.94 ± 0.01 Å, compared to 1.97 Å in water. This shortening increases the ligand field strength, and hence the covalency of the Cu–Oeq bond and stability of the chelate.

  • (3)

    The chelate is further stabilized by the bonding of axial oxygens with intra- or inter-molecular carboxyl groups.

  • (4)

    Steric hindrances in NOM are the main reason for the absence of Cu–Cu interactions, which otherwise are common in carboxylate coordination complexes.

Introduction

The biogeochemistry of copper is largely controlled by its interactions with natural organic matter (NOM), not so much because of NOM’s abundance and polyfunctional character, as because of its remarkable affinity towards Cu(II) relative to other divalent cations (McLaren et al., 1983, Hering and Morel, 1988, Town and Powell, 1993, Ramos et al., 1994, Benedetti et al., 1996, McBride et al., 1997, Leenheer et al., 1998, Kinniburgh et al., 1999, Kogut and Voelker, 2001, Santos-Echeandia et al., 2008). The strong binding ability of NOM for Cu(II) likely results from the excellent match in size between the cupric ion and one or several ligands. The copper ligands must be well-defined structurally and chemically, and also numerous because NOM has a high sorption capacity and selectivity for Cu(II) over a large concentration range (Gao et al., 1997, Covelo et al., 2004). Since carboxylate moieties comprise the large majority of reactive sites below pH 7 in NOM, and can form a vast number of synthetic coordination complexes with Cu(II) (Melnik et al., 1998a, Melnik et al., 1998b, Melnik et al., 1999), the strongest bonds are expected to be with carboxyl ligands (Sposito et al., 1979, Boyd et al., 1981). The predominance of oxygen ligands does not exclude other electron donors from being involved in Cu bonding, such as nitrogen as suggested by electron spin resonance (ESR) spectroscopy (Boyd et al., 1983, Senesi and Sposito, 1984, Senesi et al., 1985, Luster et al., 1996). Based on solution chemistry the most likely ligands are dicarboxylate (malonate, log K = 5.04), mixed alcohol-carboxylate (citrate, malate, log K = 3.70 and 3.63), and aminocarboxylate (glutamate, log K = 8.32) groups attached to aliphatic chains (Fig. 1, Table 1; Gregor et al., 1989a, Gregor et al., 1989b, Town and Powell, 1993, Croué et al., 2003). Aromatic dicarboxylate (phthalate, log K = 3.22), alcohol-carboxylate (salicylate, log K = 2.22 for pH 5), and furan-carboxylate (furanate, log K = 1.10) ligands are less likely, because they have weaker binding strengths than aliphatic ligands. This interpretation is reinforced by infrared spectroscopy and two-dimensional NMR which showed that carboxylate moieties with –COOH, –OH, or –OR substituents on the α carbon from aliphatic chains or alicyclic structures constitute the majority of carboxyl structures, with the salicylate and furan-carboxylate aromatic structures being less reactive (Deshmukh et al., 2007, Hay and Myneni, 2007). Therefore, malonate, citrate/malate, and amino acid structures most likely play the important roles in Cu(II) chelation. Although stability constants can be used to predict the a priori nature of Cu(II)–NOM complexes, they do not provide unequivocal information on the true bonding environment of the metal in natural systems. They are, however, extremely valuable for making educated choices on the relevance of model compounds used in structural studies.

Using XANES and EXAFS spectroscopy, and Cu–glutamate as the best-fit structural analog, Karlsson and Skyllberg (2006) showed that Cu(II) forms a five-membered chelate ring with one amino nitrogen (α-NH2) or alcohol oxygen (α-OH) and one carboxylate oxygen from an α-substituted aliphatic carboxylic structure (Fig. 1). Copper is bridged to four O/N ligands (CuL2 complex) in a square-planar geometry at low metal concentration, and to two (CuL complex) at higher concentration. Today the glutamate-like model is the most detailed, but despite its own merit the topic is far from being exhausted. Here are some remaining open questions, which are addressed in this article:

  • Is it possible to distinguish oxygen from nitrogen ligands, at least at low Cu concentration when heterogeneity is minimized?

  • What are the limits of the glutamate model? Does it always provide an adequate fit to EXAFS data? Since NOM contains several well-defined types of carboxyl structures with specific protonation constants, the coordination mode of Cu(II) should vary with pH, metal to ligand ratio, and chemical composition of the NOM.

  • At medium and high Cu concentration, can the data be modeled by a mixture of discrete ligands, and if so, how many are needed?

Answering these questions rests upon three capabilities (i) to detect small variations of the local structure about Cu atoms up to approximately 4 Å (i.e., second C atomic shell), and over two orders of magnitude in concentration (i.e., Cu/C ratio); (ii) to determine from a large dataset of multicomponent spectra the number of independent patterns (or “principal components”, PCs) that represent the number of identifiable Cu binding environments (i.e., species) present in the set of spectra; (iii) to build a spectral database of references of known identity and structure, which includes all distinct binding environments seen by EXAFS and plausibly present in NOM.

We recorded our data at liquid helium (LHe) temperature to enhance the signal from higher atomic shells by reducing the lability of organics, and by using a high-flux spectrometer at the European Synchrotron Radiation Facility (ESRF, Grenoble) equipped with a 30-element Ge detector for the study of highly diluted samples (concentration range 100–6500 mg/kg). The significant number of independent patterns accounting for 98–99% of variance in the dataset was determined with principal component analysis (PCA, also known as abstract factor analysis, Malinowski, 1991, Wasserman et al., 1999: Ressler et al., 2000, Manceau et al., 2002). Here, linear algebra can be used for quantification because the EXAFS response from a multispecies sample is the weighted-superposition of the response of each species present in a sample (Manceau et al., 1996). The PCs obtained from the PCA are not real spectra (i.e., they are abstract components), but the single species spectra which make up the multicomponent spectra in the dataset are linear combinations of PCs. Thus, the spectra of all unknown species contained in a sample can be identified from a database by target transformation, provided the unknown is present in the library of reference spectra. Care has been taken to include in the database all plausible Cu(II) species discussed in the literature and to not miss any major species. With such an extensive and representative database, target testing goes beyond the usual fingerprinting approach between known and unknown spectra, because the entire dataset is analyzed at one time in a statistically meaningful way for similarity to a specific structural reference.

Section snippets

Materials

Copper was sorbed on four types of organic materials considered good representatives of the diversity of NOM: two humic acid standards from the International Humic Substance Society (IHSS #1S102H from the Elliott Soil, and #1S103H from Pahokee Peat in Florida), one peat of heath vegetation (mostly Carex sp.) from Mazerolles, northeast of Nantes, France, and one moss peat (mostly Sphagnum sp.) from Estonia. The humic acid (HA) and fulvic acid (FA) fractions were extracted from the Carex (CP) and

Cu(II) reference compounds

When Cu is monovalent and coordinated to O (e.g., Cu2O), a well-defined absorption feature from the 1s to 4px,y transition occurs at 8982 eV (Kau et al., 1987) (Fig. 2a). This feature is less salient when Cu(I) is coordinated to S (e.g., Cu2S), because of the hybridization of Cu-d and S-p states. When S is bonded to Cu(II) (e.g., CuS) instead of Cu(I), it occurs at higher energy (8986 eV), partly because of the chemical shift of the 1s core level of Cu(II) relative to Cu(I). The high degree of

Nature of the five- and six-O-ring chelates

Despite the polyfunctional character of NOM and the many possible binding sites and coordination geometries, all samples can be described with variable proportions of only three key model ligands, all single metal ring chelates: malate, malonate, and thiolactate. There was no hint under our experimental conditions for dimerization of Cu, as observed for some phthalate, benzoate, acetate, and succinate molecular complexes (Koizumi et al., 1963, Cingi et al., 1970, Cingi et al., 1977, Cingi et

Acknowledgments

Olivier Proux and Jean Louis Hazemann from the FAME beamline at ESRF, and Matthew Marcus and Sirine Fakra from beamline 10.3.2 at the ALS, are thanked for their assistance during X-ray measurements. We thank Tourbières de France (http://www.tourbieres.com/) for providing the peat samples. This manuscript benefited from comments and suggestions by two anonymous reviewers. This research was funded by the EC2CO program from the CNRS, and supported by the Centre National de la Recherche

References (133)

  • M.P. Isaure et al.

    Quantitative Zn speciation in a contaminated dredged sediment by μPIXE, μSXRF, EXAFS spectroscopy and principal component analysis

    Geochim. Cosmochim. Acta

    (2002)
  • D.G. Kinniburgh et al.

    Ion binding to natural organic matter: competition, heterogeneity, stoichiometry and thermodynamic consistency

    Colloid Surf. A Phys. Eng. Aspects

    (1999)
  • T. Kirpichtchikova et al.

    Speciation and solubility of heavy metals in contaminated soil using X-ray microfluorescence, EXAFS spectroscopy, chemical extraction, and thermodynamic modelling

    Geochim. Cosmochim. Acta

    (2006)
  • N. Kosugi et al.

    Polarized Cu K-edge XANES of square planar CuCl42− ion. Experimental and theoretical evidence for shake-down phenomena

    Chem. Phys.

    (1984)
  • Y.J. Lee et al.

    Cu(II) adsorption at the calcite–water interface in the presence of natural organic matter: kinetic studies and molecular-scale characterization

    Geochim. Cosmochim. Acta

    (2005)
  • M. Mizutani et al.

    An infinite chiral single-helical structure formed in Cu(II)-l-/d-glutamic acid system

    Inorg. Chim. Acta

    (1998)
  • K. Ozutsumi et al.

    EXAFS and spectrophotometric studies on the structure of mono- and bis(aminocarboxylato)copper (II) complexes in aqueous solution

    J. Inorg. Biochem.

    (1991)
  • F. Panfili et al.

    The effect of phytostabilization on Zn speciation in a dredged contaminated sediment using scanning electron microscopy, X-ray fluorescence, EXAFS spectroscopy and principal components analysis

    Geochim. Cosmochim. Acta

    (2005)
  • L. Powers

    X-ray absorption spectroscopy. Application to biological molecules

    Biochim. Biophys. Acta

    (1982)
  • J. Prietzel et al.

    Sulfur K-edge XANES spectroscopy reveals differences in sulfur speciation of bulk soils, humic acid, fulvic acid, and particle size separates

    Soil Biol. Biochem.

    (2007)
  • A.L. Ankudinov et al.

    Relativistic calculations of spin-dependent X-ray-absorption spectra

    Phys. Rev. B

    (1997)
  • R.A. Bair et al.

    Ab initio studies of the X-ray absorption edge in copper complexes. I. Atomic Cu2+ and Cu(II)Cl2

    Phys. Rev. B

    (1980)
  • H. Bartl et al.

    Neutronenbeugungsuntersuchung der extrem kurzen wasserstoffbrücke in kupfer-dihydrogen-diphthalat-dihydrat

    Z. Kristallogr.

    (1980)
  • C.F. Bell

    Principles and Applications of Metal Chelation

    (1977)
  • E. Bohlin et al.

    Botanical and chemical characterization of peat using multivariate methods

    Soil Sci.

    (1989)
  • R.C. Bott et al.

    The preparation and crystal structure of ammonium bis[citrato(3-)]cuprate(II)

    Aust. J. Chem.

    (1991)
  • S.A. Boyd et al.

    Copper(II) and iron(III) complexation by the carboxylate group of humic acid

    Soil Sci. Soc. Am. J.

    (1981)
  • S.A. Boyd et al.

    Copper(II) binding by humic acid extracted from sewage sludge: an electron spin resonance study

    Soil Sci. Soc. Am. J.

    (1983)
  • G.G. Briand et al.

    Identification, isolation, and characterization of cysteinate and thiolactate complexes of bismuth

    Inorg. Chem.

    (2004)
  • G.M. Brown et al.

    Dinuclear copper(II) acetate monohydrate: a redetermination of the structure by neutron-diffraction analysis

    Acta Crystallogr. B

    (1973)
  • I.D. Brown et al.

    Bond-valence parameters obtained from a systematic analysis of the inorganic crystal structure database

    Acta Cryst.

    (1985)
  • J. Burgess

    Ions in Solution. Basic Principles of Chemical Interactions

    (1988)
  • R. Calvo et al.

    Crystal structure and magnetic properties of diaqua(l-aspartato)copper(II)

    Inorg. Chem.

    (1993)
  • F. Carrera et al.

    Nature of metal binding sites in Cu(II) complexes with histidine and related N-coordinating ligands, as studied by EXAFS

    Inorg. Chem.

    (2004)
  • X.Y. Chen et al.

    Multi-dimensional copper(II) coordination polymers via self-assembly induced by sodium ions

    Z. Anorg. Allg. Chem.

    (2003)
  • M.B. Cingi et al.

    The crystal and molecular structure of bis(hydrogen o-phthalato)diaquocopper(II)

    Acta Crystallogr. B

    (1969)
  • M.B. Cingi et al.

    The crystal and molecular structure of diammine-(o-phthalato)copper (II)

    Acta Crystallogr. B

    (1970)
  • M.B. Cingi et al.

    Influence of the alkaline cation on the structures of polymeric o-phthalatoeuprate(ll). 1. The crystal structures of dilithium catena-di-μ-(o-phthalato)-cuprate(ll) tetrahydrate and dirubidium catena-di-μ-(o-phthalato)-cuprate(ll) dihydrate

    Acta Crystallogr. B

    (1977)
  • M.B. Cingi et al.

    The crystal and molecular structure of o-phthalatocopper(II) dihydrate

    Acta Crystallogr. B

    (1978)
  • M.B. Cingi et al.

    The crystal and molecular structures of magnesium di-o-phthalatocuprate(II) dihydrate and strontium di-o-phthalatocuprate(II) trihydrate

    Acta Crystallogr. B

    (1978)
  • M.B. Cingi et al.

    The crystal and molecular structures of barium diaquadi(o-phthalato)cuprate(II) dihydrate. An example of the chelating behavior of the o-phthalate anion

    Acta Crystallogr. B

    (1978)
  • M.B. Cingi et al.

    Polymeric chains of tetranuclear hydroxo-o-phthalatocuprate(II) units and silver(I)-aromatic interactions in the crystal structure of Ag[Cu2(C8H4O4)2(OH)]·5H2O

    Acta Crystallogr. B

    (1979)
  • F.A. Cotton et al.

    Advanced Inorganic Chemistry

    (1999)
  • E.F. Covelo et al.

    Competitive adsorption and desorption of cadmium, chromium, copper, nickel, lead, and zinc by humic umbrisols

    Commun. Soil Sci. Plant Anal.

    (2004)
  • J.P. Croué et al.

    Characterization and copper binding of humic and nonhumic organic matter isolated from the South Platte River: evidence for the presence of nitrogenous binding site

    Environ. Sci. Technol.

    (2003)
  • P. D’Angelo et al.

    X-ray absorption study of copper(II)–glycinate complexes in aqueous solution

    J. Phys. Chem. B

    (1998)
  • G. Davies et al.

    Tight metal binding by humic acids and its role in biomineralization

    J. Chem. Soc. Dalton Trans.

    (1997)
  • P. de Meester et al.

    Refined crystal structure of tetra-μ-acetato-bisaquodiccoper(II)

    J. Chem. Soc. Dalton Trans.

    (1973)
  • L. Dupont et al.

    EXAFS and XANES studies of retention of copper and lead by a lignocellulosic biomaterial

    Environ. Sci. Technol.

    (2002)
  • B. Evertsson

    The crystal structure of bis-l-histidinecopper(II) dinitrate dihydrate

    Acta Crystallogr. B

    (1969)
  • Cited by (166)

    View all citing articles on Scopus
    View full text