molecular perspectives on goethite dissolution in...

52
Molecular Perspectives on Goethite Dissolution in the Presence of Oxalate and Desferrioxamine-B Anna Simanova Department of Chemistry Umeå University Umeå 2011

Upload: others

Post on 23-Mar-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

Molecular Perspectives on Goethite Dissolution in the Presence of Oxalate and Desferrioxamine-B

Anna Simanova

Department of Chemistry

Umeå University

Umeå 2011

Page 2: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

Copyright © Anna Simanova, 2011

Cover designed by Roman Rosluk

ISBN: 978-91-7459-208-5

Electronic version available at http://umu.diva-portal.org/

Printed by Print&Media

Umeå, Sweden, 2011

Page 3: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

To my beloved husband…

You are in my heart…

Page 4: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution
Page 5: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

i

Table of Contents Abstract ..................................................................................................... ii 1. List of Papers ....................................................................................... 1 2. Introduction .......................................................................................... 3 3. The Aim of the Thesis .......................................................................... 4 4. Experimental Techniques and Data Analysis ...................................... 5

4.1. Attenuated Total Reflectance Fourier Transform Infrared (ATR-FTIR) Spectroscopy ........................................................ 5

4.2. Extended X-Ray Absorption Fine Structure (EXAFS) Spectroscopy .............................................................................. 8

4.3. Macroscopic Measurements ...................................................... 9

5. Speciation and IR Spectra of Organic Ligands in Aqueous Solution 10

5.1. Oxalate ..................................................................................... 10 5.2. Desferrioxamine-B (DFOB) .................................................... 13

6. Surface Characteristics of Goethite (α-FeOOH) ................................ 16 7. Types of Surface Complexes ............................................................. 17 8. Interaction of C2O4

2- and Me(C2O4)33- with Goethite Suspension

at pH 4 and 6 ...................................................................................... 20

8.1. Time-resolved IR Spectra at the Water-Goethite Interface in the Presence of C2O4

2- .................................................................... 20 8.2. Ternary Surface Complexes and Their Structures ................... 22

9. Interaction of DFOB and MeHDFOB+ with Goethite Suspension at pH 6 ................................................................................................ 25

9.1. Time-resolved IR Spectra at the Water-Goethite Interface in the Presence of DFOB ......................................................... 25

9.2. Macroscopic Data .................................................................... 27

9.2.1 At a Total DFOB Concentration of 1 µmol/m2 ........................... 27 9.2.2 At a Total DFOB Concentration of 0.25 µmol/m2 ...................... 30

9.3. Interaction of MeHDFOB+ with Goethite: Ternary Complexation ........................................................................... 32

10. Dissolution of Goethite in the Presence of Oxalate and DFOB at pH 6 ............................................................................................. 34

11. Conclusions ..................................................................................... 37 12. Acknowledgements ......................................................................... 39 13. References ....................................................................................... 41

Page 6: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

ii

Abstract Iron, an essential nutrient, is primarily present in soils in the form of iron-

bearing minerals characterized with low solubilities. Under iron deficient conditions, some plants and microorganisms exude a mixture of iron-complexing agents, including carboxylates and siderophores, that can cause minerals to dissolve and increase iron solubility. Siderophores are chelating agents with functional groups such as hydroxamate, catecholate, or α-hydroxycarboxylate, that have high selectivity and specificity for Fe(III). This thesis is focused on adsorption/dissolution processes at the surface of a common soil mineral, goethite (α-FeOOH), in the presence of oxalate and a trihydroxamate siderophore, desferrioxamine-B (DFOB) at pH 4 and/or 6 in the absence of visible light. In order to characterize these processes at a molecular level and to understand the reaction mechanisms, a combination of attenuated total reflectance Fourier transform infrared (ATR-FTIR) spectroscopy, extended X-ray absorption fine structure (EXAFS) spectroscopy and quantitative solution phase measurements were applied.

In the oxalate-goethite system, four surface species were detected: 1) an electrostatically attracted outer-sphere complex, 2) a hydrogen bonded outer-sphere complex, 3) an inner-sphere oxalate coordinated to surface iron and 4) a ternary type A complex formed during a dissolution-readsorption process. Addition of Al(C2O4)3

3- or Ga(C2O4)33- to a goethite suspension resulted in the formation of an

additional surface complex - oxalate coordinated to Al or Ga in a ternary type A complex.

In the DFOB-goethite system, DFOB is subjected to surface-mediated hydrolysis followed by the reduction of Fe(III) as evidenced by the release of acetate and a nitroso-DFOB fragment into the aqueous phase. It is postulated that Fe(II) is not detected in the solution phase due to its adsorption at the surface. At low surface coverage, a small fraction of dissolved FeHDFOB+ complex is also likely to form ternary surface complexes and hydrolyze. These observations suggest that DFOB-promoted dissolution of goethite may proceed not only via purely ligand-exchange reactions, but also through reductive pathways.

In the oxalate-DFOB-goethite system, the dissolution rates are greater than the sum of the dissolution rates in the single-ligand systems. Results presented demonstrate that this synergistic effect is due to the formation of the above mentioned ternary oxalate surface complex via dissolution and readsorption. Iron in this ternary complex is more labile than iron in the crystal lattice and thus more readily accessible for other complexing agents, e.g. siderophores.

The results presented in this thesis provide a molecular-level view of ligand-promoted mineral dissolution in the presence of small carboxylates and/or siderophores, which improves our fundamental understanding of the role of surface complexation in mineral dissolution and iron bioavailability.

Page 7: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

1

1. List of Papers

I. Formation of Ternary Metal-Oxalate Surface Complexes on α-FeOOH particles

Anna A. Simanova, John S. Loring, Per Persson

Submitted to Journal of Physical Chemistry C

II. Evidence for Ligand Hydrolysis and Fe(III) Reduction in the Dissolution of Goethite by Desferrioxamine-B

Anna A. Simanova, Per Persson, John S. Loring

Geochimica et Cosmochimica Acta 2010, 74, 6706-6720

Reprinted with permission from Elsevier

III. Surface Reactions Controlling the Decomposition of DFOB in the Presence of Goethite

Anna A. Simanova, John. S. Loring, Per Persson

Manuscript

IV. Highly Mobile Iron Pool from a Dissolution-Readsorption Process

John S. Loring, Anna A. Simanova, Per Persson

Langmuir 2008, 24, 7054-7057

Reprinted with permission from American Chemical Society

Page 8: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

2

Page 9: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

3

2. Introduction

The reactivity, mobility, bioavailability and ecological impact of a metal are controlled to a great extent by the metal speciation. In the environment, metals are present in numerous chemical forms such as free ions in different oxidation states and their hydrolysis products; inorganic or organic aqueous complexes; (nano)crystalline and amorphous solids; and different surface complexes. For example, iron, an essential nutrient, is found in soils mainly in the form of Fe(III)-(hydr)oxide minerals characterized with low solubilities. The low solubilities and slow dissolution rates of iron-bearing minerals often limit the bioavailability of iron.1-3 For example, at circumneutral pH and in the absence of any complexing agents, the concentration of soluble iron is at or below nanomolar level, which is far below the demands of plants and microorganisms.2,4-5 However, the solubility of minerals can be profoundly increased in the presence of anionic complexing agents, including carboxylic acids and siderophores.

Carboxylates are present in soils at variable concentrations as byproducts of organic matter decomposition and root exudates.6-8 Plants, bacteria, and fungi excrete carboxylic acids along with other chelating agents under moderate iron deficiency.9 Carboxylic acids are able to accelerate mineral dissolution, not only by lowering pH and complexing metals in solution but most importantly by complexing surface metals and promoting metal detachment from mineral surfaces either via non-reductive or reductive mechanisms.3,5,10-19

In response to severe iron deficiency, some plants and microorganisms can exude siderophores, low molecular weight organic compounds characterized by a high affinity and selectivity for Fe(III).20 Siderophores are more efficient than carboxylates at solubilizing iron at circumneutral pH and transporting it into living cells.21 Siderophores often contain such metal-binding groups as hydroxamate, catecholate, or α-hydroxy carboxylate, which form very stable complexes with Fe(III) (the stability constants are 1023-1052).22-23 Some siderophores (e.g. hydroxamate siderophores) are thought to dissolve iron via a non-reductive ligand-promoted mechanism, especially in the absence of light.22,24 In the presence of visible light, photoreductive dissolution mechanisms can be induced.25-28

Page 10: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

4

It has also been reported that carboxylates and siderophores act synergistically on the dissolution rates of iron-bearing minerals, i.e. the rates in the presence of both types of ligands are higher than a sum of the dissolution rates in the presence of only one ligand.29-33 Most of the previous investigations of this phenomenon have been based on macroscopic solution-phase experiments. Although a possible mechanism of the synergistic effect has been proposed,33 a fundamental molecular understanding has been lacking.

Mineral dissolution pathways and rates in the presence of organic ligands, such as carboxylates and siderophores, are partly controlled by the manner in which the ligands are coordinated to the mineral surfaces. Thus, in order to understand the role of organic ligands in mineral dissolution, as well as model environmental metal speciation and bioavailability, it is important to gain detailed structural information about surface complexes formed upon the interaction of ligands with minerals.

3. The Aim of the Thesis

The present study was aimed to improve our understanding of the interaction of oxalate, the simplest representative of the dicarboxylates, and desferrioxamine-B (DFOB), a commercially available trihydroxymate siderophore, with goethite (α-FeOOH) in aqueous suspension in the absence of visible light. Single-ligand systems were studied to characterize time-dependent surface speciation of oxalate C2O4

2- and H4DFOB+ adsorbed at the water-goethite interface at pH 4 and/or 6. In the DFOB-goethite system, solution phase composition was also determined as a function of time in order to shed light on the mechanisms of interfacial reactions. In addition, interaction of Me(C2O4)3

3- (Me=Fe(III), Co(III), Al(III), Ga(III)) and MeHDFOB+ (Me=Fe(III), Al(III), Ga(III)) complexes with goethite was explored to elucidate the role of ternary surface complexation in the dissolution processes. The two-ligand system was investigated at pH 6 in order to gain molecular-level insight into the synergistic role of these ligands in iron acquisition from goethite.

Page 11: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

5

4. Experimental Techniques and Data Analysis

A combination of attenuated total reflectance Fourier transform infrared spectroscopy (ATR-FTIR), extended X-ray absorption fine structure (EXAFS) spectroscopy, and an array of macroscopic analyses was used to provide complementary information on ligand-adsorption/mineral-dissolution processes at the water-goethite interface. Molecular-level insights into the interfacial processes were obtained using ATR-FTIR and EXAFS spectroscopies. The former technique is useful in elucidating structures of metal-ligand complexes from the perspective of the organic or inorganic ligand, while the latter provides this type of information from the point of view of the metal. The results from these spectroscopic techniques were combined with macroscopic solution-phase data obtained using ion chromatography (IC), inductively coupled plasma optical emission spectroscopy (ICP-OES), liquid chromatography mass spectrometry (LC-MS), ultraviolet-visible spectroscopy (UV-Vis), and potentiometry. All experiments were carried out in the absence of light to avoid possible photoinduced reactions such as photoreductive ligand decomposition and photoreductive mineral dissolution.

4.1. Attenuated Total Reflectance Fourier Transform Infrared (ATR-

FTIR) Spectroscopy

Attenuated total reflectance infrared (ATR-IR) spectroscopy is a technique widely used to collect the infrared spectra of dense and highly absorbing materials such as solids and liquids. In ATR-IR, a sample is placed on an inert infrared-transparent crystal called an internal reflection element (IRE), characterized with a high refractive index (Fig. 1). The infrared radiation is directed into the IRE so that it undergoes total internal reflection at the

Goethite

Z-Axis

Evanescent Wave

ZnSe

Goethite Suspension

SourceDetector

θ

Adsorbates

Figure 1. An internal reflection element covered with a thin goethite layer.

Page 12: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

6

crystal-sample interface. Total internal reflection occurs when the incidence angle θ is greater than the critical angle θc (θc = sin-1n21, n21 = n2/n1, where n1 and n2 are the refractive indices of the IRE and the sample, respectively). Under this condition, light penetrates into the sample in the form of an “evanescent wave” that exponentially decreases in intensity with distance. The evanescent wave travels into the sample, where it is absorbed or “attenuated” at vibrational frequencies characteristic of the substances of which the sample is comprised. The penetration depth depends on the wavelength of the radiation, the angle of incidence, and the refractive indices of the IRE and sample. Because the penetration depth is short (a few microns for infrared light), ATR-IR is especially convenient for investigating highly absorbing samples. In this study, a nine-reflection diamond crystal was used to measure the IR spectra of some aqueous solutions (Papers II, III), while a single-reflection ZnSe (45o) crystal was used for infrared titrations of other aqueous solutions and mineral suspensions (Papers I-IV).

In a FTIR spectrometer, polychromatic infrared radiation is directed to an interferometer where the beam is split by a beamsplitter into two parts: one part is directed to a fixed mirror, while another is sent to a moving mirror. The beams are then reflected from the mirrors back through the beamsplitter. The beams recombine and constructively or destructively interfere depending on the pathlength difference between the fixed mirror and the beamsplitter and the moving mirror and the beamsplitter. The resulting interference pattern is recorded by a detector and is called an interferogram. This interferogram contains information in the wavelength-domain but can be mathematically converted into a spectrum in the frequency (wavenumber) domain using a Fourier transform. Fourier transform instruments have the advantages of high light throughput, good resolution (dictated by the total length of travel by the moving mirror), and the possibility to average a large number of scans.

In this thesis, a majority of the IR spectroscopic data of aqueous solutions or mineral suspensions was obtained by coupling ATR-FTIR spectroscopy with the pH measurements using a method called simultaneous infrared and potentiometric titrations (SIPT). This method provides a tool for in situ monitoring of molecular-scale changes in solution or at the water-mineral interface as a function of pH, or as a function of time at a constant

Page 13: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

7

pH. A scheme of the experimental setup for SIPT experiments is shown in Fig. 2 and explained in detail in Papers I, II. Briefly, an aqueous solution or mineral suspension was pumped peristaltically from a titration vessel equipped with a propeller stirrer and a combination pH electrode to an ATR cell placed inside the spectrometer. The ATR cell consisted of a flow-through attachment mounted on a single-reflection, 45o, ZnSe IRE.

For titrations of mineral suspensions, the ZnSe IRE was covered with a thin film, or an overlayer, of the mineral (Fig. 1) by evaporating a dilute suspension of the mineral onto the crystal. This method concentrates and immobilizes mineral particles in the probing area of the infrared evanescent wave and, therefore, increases the signal and the sensitivity of the measurements.34 This procedure has been widely used to study adsorption and desorption of different ligands at water-mineral interfaces.34-40

Figure 2. A schematic representation of the SIPT experimental setup

Infrared data obtained in the SIPT experiments were processed and interpreted using different mathematical algorithms including second derivative and the two-dimensional (2D) correlation analyses. Second derivatization was used here to enhance the resolution of the infrared spectra. This procedure sharpens narrow but overlapping infrared bands, but suppresses broad bands and eliminates baseline changes. The second-order derivative spectrum has negative bands with minima at the same wavenumbers as maxima of corresponding bands in the original spectrum.

Time-resolved SIPT spectra obtained at a fixed pH and ligand concentration were also subjected to two-dimensional (2D) correlation analyses (Papers I-II). These analyses41 can be used to reveal perturbation-induced variations in positions and intensities, and they also provide information on correlations between spectral variations. Spectra obtained as

Page 14: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

8

a function of a perturbation (in this study time is the perturbation) are converted to a 2D correlation spectrum that is comprised of two orthogonal components: synchronous and asynchronous. The synchronous plot consists of auto-peaks, which appear at the diagonal due to the major changes in the spectra as a function of the perturbation, and cross-peaks, which appear at the off-diagonal positions due to simultaneous changes at two wavenumbers. Autopeaks are always positive while cross-peaks can be positive (when intensities at two separate wavenumbers increase or decrease together) or negative (when intensity at one wavenumber increases while intensity at another wavenumber decreases). The asynchronous plot contains only cross-peaks that are generated by the non-correlated variations at two wavenumbers. These changes occur at distinct rates and are assignable to different species. As a result, the asynchronous plot can be helpful in resolving overlapping peaks resulting from species with different structures.

4.2. Extended X-Ray Absorption Fine Structure (EXAFS)

Spectroscopy

X-ray Absorption Spectroscopy (XAS) is a technique to probe local atomic surroundings of X-ray absorbing atoms. This technique is selective for a particular element and informative only on a short Ångstrom-range distance, which is advantageous for non-crystalline samples.

In XAS, a sample is irradiated with a monochromatic x-ray beam, typically arising from a synchrotron radiation source. When the energy of the incident x-ray radiation equals the binding energy of an electron in an atom, the radiation is absorbed giving rise to a sudden increase in absorption (the absorption edge) in the x-ray absorption spectrum. At the energies higher than this threshold energy, a photoelectron is released and backscattered from the neighboring atoms. The backscattered wave interferes in-phase or out-of-phase with the original photoelectron wave, resulting in increase or decrease in the electron density of the absorbing atom. In the absorption spectrum, this is observed as intensity oscillations after the adsorption edge. The analysis of an absorption spectrum allows the extraction of structural information about the absorbing atom and its surroundings. The region near the absorption edge, X-ray Absorption Near Edge Structure (XANES), gives information on coordination geometry and

Page 15: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

9

oxidation states of the absorbing atom, while the region of oscillations (Extended X-ray Absorption Fine structure, EXAFS) provides information on local structure, i.e. interatomic distances and the number and types of neighboring atoms, up to 5-6 Å around the absorbing atom.

In this work, EXAFS spectroscopy was used to determine the structures of the ternary Ga(C2O4)3

3- complex at the water-goethite interface. The X-ray absorption spectrum was measured using the fluorescence method that is most suitable for samples with low concentrations of the X-ray absorber and/or high absorbing matrix. With this method, detection is based on the measurement of the fluorescent X-ray photons that are emitted when the excited electron of the absorbing atom relaxes to the ground state. More details on the experimental data collection and data analysis can be found in Paper I.

4.3. Macroscopic Measurements

The composition of the solution phase was studied to assess the extent of ligand adsorption, mineral dissolution, and DFOB hydrolysis. For example, the solution phase composition was determined as a function of time at pH 6 in the following systems: goethite-DFOB (Paper II), goethite-FeHDFOB+ (Paper III), goethite-oxalate-DFOB (Paper IV). In these studies, a volume of 10 g/L of goethite suspension was placed into a titration vessel and titrated to pH 6, and ligand(s) were added to reach the desired total concentration(s). During the experiments, the pH was held constant at pH 6 by adding standardized acid or base. Samples were withdrawn as a function of time, centrifuged, and the supernatant was filtered and then analyzed using different methods listed below. When a direct correlation between macroscopic and spectroscopic measurements was needed (Paper IV), samples were collected from the same titration vessel containing the mineral suspension that was circulated during the SIPT experiments.

Briefly, the concentrations of iron dissolved in supernatant were measured by inductively coupled plasma optical emission spectroscopy (ICP-OES) using two emission lines at 259.939 and 238.204 nm. The concentration of DFOB in supernatant was analyzed as a FeHDFOB+ complex using ultraviolet-visible (UV-Vis) spectroscopy at a wavelength of

Page 16: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

10

430 nm. The concentrations of oxalate, acetate and mesylate were measured by means of ion chromatography (IC). Other products of DFOB hydrolysis were identified by liquid chromatography-mass spectrometry (LC-MS). More detailed description of analytical methods can be found in Papers II-IV.

5. Speciation and IR Spectra of Organic Ligands in Aqueous Solution

An understanding of the IR spectra of the different protonation states and metal complexes of a ligand in aqueous solution can facilitate the interpretation of spectra of the ligand adsorbed at a mineral surface. The molecular structures of ligands in aqueous solution are usually well understood. Furthermore, if the structures of ligands adsorbed at mineral surfaces are similar to the structures of aqueous species, then often their infrared spectra will also be similar. Hence, a comparison of the IR spectra of aqueous species with the spectra of adsorbed ligands has often been used to assess the coordination geometries of ligands in surface complexes.39,42-45

The IR spectra of the different protonation states and metal complexes of oxalate, as well as the metal complexes of DFOB, were collected for solutions made in 1 M NaCl and at appropriate total metal and ligand concentrations and pH values (Papers I-III). The IR spectra of the different protonation states of DFOB were obtained using a SIPT titration of a ligand solution over the pH range from 4 to 10 and a speciation modeling analysis of the recorded IR spectra (for details see Paper II). All spectra were examined in the range 1800-1100 cm-1 where carboxylic and hydroxamic groups have their characteristic bands.46

5.1. Oxalate

Oxalic acid has two carboxylic groups with pKa1=1.01 and pKa2=3.7923 (I=0.1 M) indicating that depending on the pH of a solution it can exist as deprotonated, singly protonated and doubly protonated species. Each of these species exhibits a distinctly different infrared spectrum (Table 1). The free oxalate anion has the D2h symmetry, which is lowered to the D2d

Page 17: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

11

symmetry in aqueous solution.47 In the region 1800-1100 cm-1, the IR spectrum of oxalate in aqueous solution has characteristic bands at 1569 and 1308 cm-1 that are assigned to the asymmetric and symmetric stretching vibrations of C-O groups, respectively.48-50 Protonation of the carboxylic groups results in a change in the

ligands symmetry affecting the number and position of infrared-active bands.

The spectrum of singly and doubly protonated oxalate was not possible to record within this study due to the chemical instability of ZnSe ATR crystal at pH values below 4 and, therefore, we reference previously collected spectra reported in the literature.44,48 Protonation of both carboxylic groups lowers the symmetry to C2v and gives rise to two absorptions at 1735 and 1233 cm-1 in the studied region due to the stretching vibrations of C=O and C-O-H groups, respectively. These bands have been previously used as spectroscopic evidence for the protonated species of oxalate.51 Protonation of one of the carboxylic group lowers the symmetry of the molecule to C and results in a more complex IR spectrum which exhibits at least four bands at 1728, 1609, 1307, 1242 cm-1 within the measured spectral range. Persson et al.44 have assigned the bands at 1728 and 1242 cm-1 to the C=O and C-OH stretches of a protonated group, and the bands at 1609 and 1307 cm-1 to the asymmetric and symmetric vibrations, respectively, of deprotonated carboxylate group. In addition, a weak feature at about 1410 cm-1 has been reported previously for singly-protonated oxalate by various authors,34-35,50,52 but it has not been assigned to a distinct vibration. However, it has been speculated that this weak band may be due to the coupling of the C-O stretching and the C-OH bending vibrations.53

Oxalate can bind to a metal ion, forming five-membered chelating rings.44,54-56 Table 2 presents the stability constants of the 1:1, 1:2, 1:3 metal-oxalate complexes with Fe(III), Al(III), and Ga(III). Coordination of oxalate to metal (Me) ions drastically changes its structural and electronic environment, thereby leading to the appearance of more complex IR spectra

Table 1. Infrared frequencies of characteristic bands of different protonation states of oxalate in the range 1800-1100 cm-1.44,48 C2O4

2- HC2O4- H2C2O4

O=Cν - 1728 1735 asC O−ν 1569 1609 - sC O−ν 1308 1307 -

OH−Cν - 1242 1233

Page 18: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

12

compared to the spectrum of free oxalate (Fig. 3, Table 3). Figure 3 shows the IR spectra of aqueous solutions of the trioxalatometal(III) complexes studied here. As can be seen from the figure, the positions of the IR bands are different for each metal-oxalate complex, since the metal-ligand bond strengths (therefore, the force constants) are different for each metal.

Figure 3. The IR spectra of Me(C2O4)3

3- complexes in aqueous solution. Table 2. Stability constants for oxalate coordinated to different Me3+ cations. The constants were taken from23 at the ionic strength of 1M and corrected to 0.1 M using the Davies equation.

log β1 log β2 log β3 Fe3+ 7.29 13.42 18.31 Al3+ 5.81 10.70 14.83 Ga3+ 6.16 11.99 17.57

Table 3. The IR frequencies of vibrational modes of Me(C2O4)3

3- complexes in aqueous solution at pH 4. Band assignments are according to the literature.44

Fe(C2O4)33- Co(C2O4)3

3- Al(C2O4)33- Ga(C2O4)3

3- nbC O−ν 1710, 1681 1704, 1677 1722, 1697 1714, 1690

CCbC −− +νν O 1391 1400 1406 1400

OCObC −−− + δν O 1269, 1252* 1255 1295, 1271 1288, 1266

*Resolved in the second derivative b denotes a vibration mode of a carboxylic group bound to a metal nb denotes a vibration mode of a group not bound to a metal

Page 19: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

13

5.2. Desferrioxamine-B (DFOB)

Desferrioxamine-B (DFOB) is a microbial siderophore containing three hydroxamic groups (Fig. 4). These three metal binding groups as well as a terminal primary amine group can be protonated or deprotonated leading to five different protonation states of DFOB in aqueous solution (Fig. 5, Table 4). A completely protonated and positively charged species H4DFOB+ dominates at pH values below 6. Increasing pH results in the stepwise deprotonation of the hydroxamic groups followed by deprotonation of the terminal amine, the most basic group.

Figure 4. Desferrioxamine-B

Table 4. Stability constants for different protonation states of DFOB and Me(III)-DFOB complexes. The constants were taken from the literature23 at the ionic strength of 1M and corrected to 0.1 using the Davies equation.

N

OH O

C(CH2)5 CH3

NHC

O

N

OH O

C(CH2)5 (CH2)2

NHC

O

N

OH O

C(CH2)5 (CH2)2

+H3N

log K DFOB3- + H+ ↔ HDFOB2- 10.79 DFOB3- + 2H+ ↔ H2DFOB- 20.34 DFOB3- + 3H+ ↔ H3DFOB 29.30 DFOB3- + 4H+ ↔ H4DFOB+ 37.62 Fe3++DFOB3- ↔ FeDFOB 30.99 Fe3++H++DFOB3-↔FeHDFOB+ 41.39 Fe3++2H++DFOB3-↔ FeH2DFOB2+ 42.33 Ga3+ + DFOB3- ↔GaDFOB 28.65 Ga3++H++DFOB3-↔ GaHDFOB+ 38.96 Al3++DFOB3-↔ AlDFOB 24.5 Al3++H++DFOB3-↔AlHDFOB+ 34.93 Al3++2H++DFOB3-↔ AlH2DFOB2+ 36.11

Page 20: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

14

The IR spectra of all five protonation states were determined by titrating the DFOB solution as a function of pH and using a speciation-modeling analysis (Fig. 6, Table 5, Paper III). Comparison of the IR spectra of totally protonated and deprotonated forms of DFOB shows that protonation leads to significant changes in the IR signature. First, the CO and CN stretching vibrations of hydroxamic groups shift from 1579 and 1472 cm-1 in DFOB3- to 1605 and 1468 cm-1 in H4DFOB+. Second, the NO stretching mode is present in the spectrum of DFOB3- at 1151 cm-1 and absent in the spectrum of H4DFOB+. The shoulder at about 1620 cm-1 in the spectra of DFOB3-, HDFO2-, and H2DFOB- (not resolved in H4DFOB+ and H3DFOB) has been assigned to the CO stretching vibration of the amide groups that should not be affected by protonation. The rest of the bands are related to the carbon

Figure 5. Distribution of aqueous DFOB species in the absence and presence of Fe(III)

Figure 6. Infrared spectra of the five protonation states of DFOB (left) and the 1:1 complexes of DFOB with Fe(III), Al(III) and Ga(III) (right) in aqueous solution.

Page 21: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

15

backbone vibrations and, therefore, are also not directly influenced by protonation of DFOB.

Complexation of DFOB with Fe(III) proceeds via the coordination of hydroxamic groups to the metal center forming five-membered rings. The coordination of all three hydroxamic groups results in the formation of the stable FeHDFOB+ complex that is predominant over a wide pH range (Fig. 5). Even though DFOB is an iron-specific ligand, it can also form stable complexes with a range of other metals including Al(III) and Ga(III),57 but with stability constants lower than those for Fe(III) complexes (Table 4). In aqueous solutions containing DFOB and Al(III) or Ga(III), the corresponding MeHDFOB+ species are also predominant in the pH range from 2 to 9 (the distribution diagrams not shown).

Table 5. Frequencies and assignments of the bands in the IR spectra of different species of DFOB (L) and MeHDFOB+ in aqueous solution in the range 1700-1150 cm-1. Assignments are according to the literature.43,45,58-60

L3- HL2- H4L+ FeHL+ AlHL+ GaHL+

)( O amideCNC νν +

1616 1616 nr 1622 1606 1632

)( O hydroxamicCNC νν +

1579 1579 1605 1571 1565 1594

)(amideCOCNNH ννδ ++

nr nr 1562 nr nr 1561

)(hydroxamicCCCOCN ννν ++

1472 1472 1468* 1461 1469 1456

NHCHCC δδν ++

1433-1198

1433-1198

1444-1201

1445-1190

1445-1190

1445-1190

)(hydroxamicCNNONOH ννδ ++

- - 1181 - - -

)(hydroxamicCNNO νν +

1151 1151 - - - -

* hydroxamicOHN−δ also contributes to this band when one or more hydroxamic groups are

protonated nr = not resolved

The IR spectra of MeHDFOB+ collected in aqueous solution at pH 6 (Fig. 6, Table 5, Paper III) display significant changes compared to the IR spectrum of uncomplexed singly protonated DFOB. As expected, most of the changes occur in the bands due to hydroxamic groups, which are involved in metal complexation. The CO and CN stretches of the hydroxamic groups shift from 1472 cm-1 in the spectrum of uncomplexed

Page 22: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

16

singly protonated DFOB to 1461, 1469, and 1456 cm-1 in the spectra of DFOB complexed to Fe(III), Al(III), and Ga(III), respectively. The band at 1579 cm-1 in the spectrum of uncomplexed HDFOB2-, that is also due to the hydroxamic groups, shifts to 1571, 1565, and 1594 cm-1 in the spectra of the complexes with Fe(III), Al(III), and Ga(III), respectively. The position and relative intensity of the amide band are also sensitive to the nature of a complexing metal, even though this group does not directly participate in complexation.

6. Surface Characteristics of Goethite (α-FeOOH)

The reactivity of minerals towards ligand exchange and dissolution processes is controlled to a great extent by the crystal structure. The type of mineral structure determines the type of crystal planes present and, therefore, the number and types of reactive surface functional groups.

Goethite is a commonly occurring iron hydroxide that has a well-defined structure and is characterized with a high thermodynamic stability.61 Goethite consists of FeO3(OH)3 octahedra that are arranged in double chains by edge sharing; one double chain is connected to another through the corners of the octahedra (Fig. 7). These double-chain layers are elongated in the [010] direction forming needle-shaped particles. The dominant crystal plane that is found along the needles is the {110} plane. In the goethite used in this study, this crystal plane accounts for about 90% of the surface area and contains singly-, doubly-, and triply-coordinated oxygens, designated as ≡FeOH0.5-, ≡Fe2OH, and ≡Fe3OH0.5-, respectively. Singly- and triply-

coordinated oxygens are responsible for the surface charging and therefore acid-base properties of the goethite in the pH range 1-11,62 but primarily singly-coordinated oxygens are assumed to take part in ligand exchange reactions. Doubly-coordinated groups are generally thought to be inert, i.e. are not Figure 7. Schematic representation of the

structure of goethite.

Page 23: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

17

expected to contribute to the surface charge and, therefore, do not play a significant role in the adsorption-dissolution processes.63

The terminations of the needles (the remaining 10 % of the surface area of the goethite particles in this study) are represented by the {021} plane 63-64 but can be roughly approximated to the {001} plane.65 This crystal plane contains only singly- and doubly-coordinated oxygens. The high density of the most reactive singly-coordinated oxygens at the {021} plane 66 results in the high reactivity of this plane with respect to ligand exchange and dissolution processes15,61 as well as to the crystal growth.67 Singly-coordinated oxygens are also believed to be present at irregularities in mineral structure such as defects, dislocations, and fractures, which are therefore also preferential sites for ligand exchange and dissolution reactions.

In this study, goethite with high surface area was synthesized according to a method described previously.63 The BET surface area was determined to be 90.1 (Paper IV) and 104.9 m2/g (Paper I-III). The point of zero charge of such goethite has been determined to be 9.4 at zero ionic strength (9.3 at I=0.1 M, which was the ionic strength typically used in this thesis research),65 which means that the surface of goethite has a net negative charge at the pH values above 9.4 and a net positive charge below pH 9.4. It follows that the surface of goethite is positively charged in experiments performed in this thesis at pH 4 and 6 in the absence of absorbed ligands.

7. Types of Surface Complexes

Rates of ligand-promoted mineral dissolution are greatly influenced not only by the morphology and the crystal structure of a mineral, but also by the types of surface complexes formed upon interaction of a ligand with mineral surfaces. Depending on the specificity of this interaction, surface complexes can be classified as inner-sphere (specific) and outer-sphere (non-specific). In an inner-sphere surface complex, a direct chemical bond is formed between a ligand and a surface metal via a ligand exchange process (Fig. 8 A-C). In an outer-sphere complex there is no such direct bond, and the ligand is attracted to the surface by electrostatic forces or hydrogen bonding and can be separated from the surface by one, two, or more water

Page 24: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

18

molecules (Fig. 8 D-E). Outer-sphere complexes of some organic ligands (e.g. EDTA39) have been reported to indirectly affect the dissolution rates by bringing more negative charge to the surface and thus enhancing protolytic attack of surface sites.

Inner-sphere complexation is of much greater importance for mineral dissolution than outer-sphere coordination. Inner-sphere complexes can be divided into several types: 1) monodentate with one functional group bound to a surface metal (Fig. 8 A), 2) mononuclear chelating bidentate with two functional groups bound to one metal site (Fig. 8 B), or 3) bridging (binuclear) bidentate where each group is bound to a different metal site (Fig. 8 C). The latter, binuclear, surface complexes are thought to inhibit dissolution processes.3,68 In contrast, mononuclear chelating ligands may enhance mineral dissolution rates by polarizing and weakening a bond between surface metal and the underlying lattice oxygen and leading to the detachment of a metal ion. Among the bidentate chelating structures, five-membered rings are more effective in promoting dissolution than six- or seven-membered rings.3,19

In the presence of a dissolved metal ion, a negatively charged ligand can be complexed with this metal and interact with mineral surface forming one of the following possible structures: 1) intact Me-L complex outer-spherically coordinated to the surface by electrostatic forces or via hydrogen bonding, similarly to Fig. 8 D-E, 2) a type A ternary complex where the metal Me links the surface to the ligand (Fig. 8 G), and 3) a type B ternary complex where the ligand bridges the metal ion and the surface (Fig. 8 H).

Page 25: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

19

Figure 8. Schematic representations of the possible coordination modes of:

I) carboxylates (e.g. oxalate) at the water-goethite interface: inner-sphere monodentate mononuclear (A), side-on and end-on bidentate (chelating) mononuclear inner-sphere (B), bidentate binuclear inner-sphere (C); electrostatically attracted outer-sphere (D); hydrogen bonded outer-sphere (E);

II) metal-carboxylate surface complexes: type A ternary complex (G); type B ternary complex (H).

Page 26: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

20

8. Interaction of C2O42- and Me(C2O4)3

3- with Goethite Suspension at pH 4 and 6

8.1. Time-resolved IR Spectra at the Water-Goethite Interface in the

Presence of C2O42-

Infrared spectra of oxalate adsorbed at the water-goethite interface were collected as a function of time at pH 4 (Papers I and IV, goethite 90.1 m2/g) and pH 6 (Paper IV, goethite 104.9 m2/g) at a total ligand concentration of 1 µmol/m2 (Fig. 9).

At pH 4 these time-resolved spectra (Fig. 9, left) show a decrease in relative intensities at 1610, 1558, 1431, 1307, and 1290 cm-1 accompanied by an increase in intensities at 1713, 1693, 1674, 1413, 1394, 1268, and 1254 cm-1 with time. Closer inspection of the infrared spectra using the 2D correlation analysis suggests that the former group of bands is due to two outer-sphere surface species: electrostatically attracted oxalate characterized by the bands at 1558 and 1307 cm-1, and hydrogen bonded oxalate characterized by the bands at 1610, 1290 cm-1. The band at 1431 cm-1 was not resolved in the 2D analysis, but might arise from hydrogen bonded oxalate. The second group of the bands closely resembles the bands of

Figure 9. Infrared spectra collected as a function of time after addition of 1 µmol/m2 of oxalate to a goethite suspension at pH 4 (left) and 6 (right) in the dark and the corresponding second derivatives. The IR spectra of aqueous species are also shown for comparison.

Page 27: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

21

metal-coordinated oxalate (Fig. 3, Table 3), and they were therefore attributed to oxalate inner-spherically coordinated to iron. The presence of three bands in the asymmetric CO stretching region suggests that there are actually more than one inner-sphere oxalate species. This hypothesis is further corroborated by the 2D analysis (Fig. 10): the asynchronous slice spectra at 1694 and 1671 cm-1 display differences indicating that the bands at these two wavenumbers are originated from two structurally distinct surface species.

Figure 10. The asynchronous slice spectra at 1694 and 1671 cm-1 obtained using the 2D analysis of the normalized time-resolved infrared spectra of oxalate adsorbed at the water-goethite interface at pH 4.

Similar to the infrared spectra at pH 4, the spectra at pH 6 show the gradual conversion from predominantly outer-spherically to predominantly inner-spherically coordinated oxalate as a function of time (Fig. 9, right). The band positions resolved in the second derivatives and the 2D analysis closely resemble the band positions at pH 4, suggesting that the same set of the surface species were present at both pH values.

However, there are some significant differences between the spectra at these two pH values. First, the transition from outer- to inner-sphere oxalate was much slower at pH 6 than at pH 4; it took approximately five days to reach close-to-equilibrium state at pH 6 compared to two days at pH 4. Secondly, by the end of the experiments, the band at 1307 cm-1 originating from the electrostatically attracted outer-sphere oxalate was more intense at pH 6 suggesting a higher prevalence of this surface complex at pH 6 than at pH 4. Thirdly, hydrogen bonded outer-sphere oxalate seemed to be present

Page 28: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

22

at much lower concentrations at pH 6 as indicated by lower intensities of the band at 1290 cm-1. Lastly, the bands that appeared at 1694 and 1671 cm-1 due to two inner-sphere oxalate complexes at pH 4 are shifted to 1689 and 1674 cm-1 in the spectra at pH 6, respectively, leading to more severe overlap of these bands. The 2D analysis of the spectra at pH 6 allows us to resolve these two bands and also indicates that they are asynchronously correlated to each other as evidenced by the presence of the peak at 1674 cm-1 in the 1689 slice spectrum (Fig. 11).

In summary, the same surface complexes are proposed for both pH 4 and 6 but with different relative concentrations: electrostatically attracted outer-sphere oxalate complexes with the bands at 1558 and 1307 cm-1, hydrogen bonded outer-sphere oxalate complexes with the bands at 1610, 1290 cm-1, and two inner-sphere oxalate complexes characterized with the bands at about 1690 and 1670 cm-1.

8.2. Ternary Surface Complexes and Their Structures

In order to further investigate the structures of inner-sphere surface species of oxalate, we collected the IR spectra as a function of time after addition of Me(C2O4)3

3- complexes (Me = Fe(III), Al(III), Ga(III), Co(III)) to a goethite suspension at pH 4 (Fig.12, Paper I).

Figure 11. The asynchronous slice spectrum at 1689 cm-1 obtained using the 2D correlation analysis of the normalized time-resolved infrared spectra of oxalate adsorbed at the water-goethite interface at pH 6.

Page 29: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

23

Figure 12. Infrared spectra collected after addition of Me(C2O4)3

3- to a goethite suspension at pH 4 and the corresponding second derivatives. The IR spectra of aqueous species are also shown for comparison.

Co(III) is relatively inert to ligand exchange,69 and the Co(C2O4)33-

complex does not lose its oxalate coordination sphere during the time scale in which the surface spectrum was collected in Fig. 12 (5 minutes). Hence, Co(C2O4)3

3- adsorbs at the water-goethite interface outer-spherically, retaining the same structure as it has in aqueous solution and resulting in almost identical surface and solution spectra (Fig. 12). All other Me(C2O4)3

3-

Page 30: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

24

complexes studied here may also form intact outer-sphere surface complexes in their initial interaction with goethite. With time these intact outer-sphere complexes fall apart to form other surface species, leading to surface IR spectra that are significantly different from the IR spectra of corresponding aqueous Me(C2O4)3

3- species (Fig. 12).

The two-hour reaction of Fe(C2O4)33- complex with a goethite

suspension resulted in an infrared spectrum that is identical to the spectrum collected after a two-day reaction of oxalate with goethite (Figs. 9 and 12). The 2D analysis detects the same four surface species as for the oxalate-goethite system discussed above: two outer- and two inner-sphere oxalate.

When Al(C2O4)33- or Ga(C2O4)3

3- are added to the goethite suspension, we clearly detect oxalate that is inner-spherically coordinated to Al(III) or Ga(III). The surface species are characterized by the bands at 1729, 1412, 1279 cm-1 for Al(III) and 1720, 1700, 1263 cm-1 for Ga(III), as resolved in a 2D correlation analysis. It should be noted that the wavenumbers identified in the 2D analysis slightly differ from the wavenumbers that are determined in the second derivatives and shown in Fig.12. The bands from the four surface species identified for the oxalate-goethite system are also resolved in these systems.

EXAFS spectra collected after the addition of Ga(C2O4)33- to a goethite

suspension, under the same conditions as in the infrared spectroscopic experiments, suggest the formation of a type A ternary complex, where the metal links the ligand and the surface site (Fig. 8G). The EXAFS results also reveal a shorter distance between Ga and C in this ternary complex compared to that in the aqueous Ga(C2O4)3

3- species, which suggests a stronger interaction between Ga and oxalate. This stronger interaction is also demostrated by the IR data displaying a blue shift of 2-13 cm-1 in the asymmetric C-O stretching band of the ternary surface complex compared to that of the aqueous species.

It is likely that a type A ternary complex of similar structure as in Fig. 8G is formed when oxalate or Fe(C2O4)3

3- reacts with the surface of goethite. In the case of oxalate, this ternary complex is most probably formed as a result of a dissolution-readsorption process: oxalate promotes the dissolution of iron from the high-energy sites and the dissolved Fe-oxalate complex readsorbs as a ternary type A complex at the low-energy surface sites.

Page 31: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

25

Oxalate-promoted dissolution is accelerated at lower pH values,2,11 which explains the faster formation of the ternary complex at pH 4 compared to that at pH 6. Therefore, we can conclude that the band at about 1690 cm-1 that is growing in with time is due to oxalate in this ternary complex (Fig. 8G), while the band at 1670 cm-1 is due to oxalate coordinated inner-spherically to surface iron (Fig. 8B, side-on).

9. Interaction of DFOB and MeHDFOB+ with Goethite Suspension at pH 6

9.1. Time-resolved IR Spectra at the Water-Goethite Interface in the

Presence of DFOB

IR spectra were collected as a function of time after addition of 0.25 or 1 µmol/m2 total concentrations of DFOB to goethite suspensions at pH 6 in the dark (Papers II, III). A plot of the area under the CH stretching band at 2867 cm-1 as a function of time (Fig. 13) plateaus after 4-h reaction for both DFOB concentrations suggesting that most of the DFOB adsorption occurs during first 4 hours. At both DFOB concentrations, the IR spectra collected after 4 hours of the reaction are very similar with respect to the band positions (Fig. 14).

Figure 13. Integrated absorbance in the CH stretching region between 2826 and 2895 cm-1 from IR spectra collected at the water-goethite interface as a function of time at 0.25 or 1 µmol/m2 total DFOB concentration at pH 6 in the dark.

Page 32: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

26

Figure 14. IR spectra collected at the water-goethite interface after about 4-h and 4-d reaction between goethite suspension and a 0.25 or 1 µmol/m2 total DFOB concentration at pH 6 in the dark. A difference spectrum obtained by subtracting a spectrum collected after 4-h reaction from a spectrum collected after 4-d reaction of a 1 µmol/m2 total DFOB concentration with goethite is also shown.

The IR spectrum of adsorbed DFOB (Fig. 14) exhibits significant differences compared to the IR spectra of aqueous DFOB species (Fig. 6). First, the NO stretching mode that appears at 1151 cm-1 in the spectra of DFOB species with deprotonated hydroxamic groups is absent in the spectrum of adsorbed DFOB. Second, changes in the region between 1490 and 1350 cm-1 are evidence of conformational changes of DFOB upon adsorption, since most of the bands in this region are due to vibrations of the ligand’s carbon backbone. The frequencies and intensities of the bands due to the CO and CN stretches of the hydroxamate and amide groups at 1630-1570 cm-1 in the spectrum of adsorbed DFOB seem to be similar to these bands in the spectrum of FeHDFOB+ (Fig. 6). This similarity has been previously used to suggest inner-sphere coordination of DFOB and to estimate the number of coordinated hydroxamic groups.45 However, we suggest that this region is unreliable due to the severe overlap of these bands with the OH bending mode of water and the sensitivity of the CO stretching bands of the amide groups to conformational changes. Thus, based on these IR spectra, we conservatively suggest that the hydroxamic groups of DFOB

Page 33: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

27

are coordinated to the surface either inner-spherically or outer-spherically but protonated.

Even though the adsorption of DFOB seems to be complete after 4 hours, significant changes in the IR spectra are observed during the remainder of the experiments at both 0.25 and 1 µmol/m2 total DFOB concentrations. In both cases, the growth of the bands at 1548 and 1404 cm-1 is revealed (Fig. 14). These bands can contain a contribution from monocarboxylate- or nitroso-DFOB fragments, the products of the hydrolysis of hydroxamic groups. At 0.25 µmol/m2 of total DFOB concentration, the growth of these bands is coupled with the decrease in intensities of other bands after about 10 hours of the reaction.

9.2. Macroscopic Data

9.2.1 At a Total DFOB Concentration of 1 µmol/m2

Figure 15 shows that the reaction of a 1 µmol/m2 total DFOB concentration with a goethite suspension at pH 6 results in the release of Fe(III) into the solution phase, an increase in the base consumption (base was added to keep pH at 6), the release of acetate, and an increase in a concentration of unaccounted DFOB (see also Paper II). The unaccounted DFOB, was calculated according to:

tSupernatanDFOB

TotalDFOB

dUnaccounteDFOB cc c −= (1)

where dUnaccounteDFOBc is the concentration of DFOB that cannot be detected in

the supernatant, TotalDFOBc is the known total DFOB concentration, and

tSupernatanDFOBc is the concentration of DFOB measured in the supernatant.

Page 34: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

28

Figure 15. The concentrations of Fe(III) dissolved, OH- consumed to keep pH of the suspension at 6, acetate released and DFOB unaccounted as a function of time at pH 6 and a 1 µmol/m2 total DFOB concentration.

The increase in dUnaccounteDFOBc during first four hours of the reaction is

most probably due to DFOB adsorption, as indicated by the IR data discussed above. The increase in dUnaccounte

DFOBc during the remainder of the

experiment, as well as the release of acetate and the base consumption, can be explained by the hydrolysis of DFOB at the terminal hydroxamic group, forming a large hydroxylamine-DFOB fragment (R-NHOH):

Independent experiments performed under the same conditions (pH 6, 1 µmol/m2 total DFOB concentration, dark) showed that H4DFOB+ and FeHDFOB+ in aqueous solutions are stable with respect to hydrolysis. This suggests that the hydrolysis observed in the goethite suspension is most probably induced by the surface. Inner-sphere complexation of DFOB to the

H3N

(H2C)2

N

OH O

C

NH

O

C

(H2C)5 (H2C)2

N

OH O

C

NH

O

C

(H2C)5 (H2C)5R =

Page 35: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

29

goethite surface could cause a shift of the electron density from a coordinated hydroxamic group to a surface metal, making it susceptible to nucleophilic attack by water molecules. The terminal hydroxamic group is most likely to be hydrolyzed since this group can more readily bind inner-spherically to the surface as it is least sterically hindered. The similarity in the rates of the DFOB removal, acetate formation, and the consumption of base supports the hypothesis of hydrolysis at the terminal hydroxamate group.

If we assume that the concentration of unaccounted DFOB is the sum of adsorbed and hydrolyzed DFOB and that the measured concentration of acetate is equivalent to the concentration of hydrolyzed DFOB, then we can calculate the concentration of adsorbed DFOB:

AcetatetSupernatan

DFOBTotalDFOB

AdsorbedDFOB cccc −−= (2)

where AdsorbedDFOBc is the concentration of DFOB adsorbed intact and Acetatec is

the concentration of released acetate.

The concentration AdsorbedDFOBc plateaus after 4 hours of the reaction (Fig.

16) for the experiment at 1 µmol/m2, which is consistent with the IR data in Fig. 13. It should be noted that Adsorbed

DFOBc may have a small contribution from

the decomposition products that could form as a result of the hydrolysis at other hydroxamic groups and remain at the surface.

Figure 16. The concentrations of DFOB unaccounted and adsorbed in a goethite suspension as a function of time at pH 6 and a 1 µmol/m2 total DFOB concentration.

Page 36: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

30

The hydroxylamine-DFOB fragment (R-NHOH, 519 m/z), which is highly reductive, was not found in the solution phase by LC-MS analysis. Instead, its oxidation product, a nitroso-DFOB fragment (517 m/z), was detected. Thus, we propose that the hydroxylamine-DFOB fragment is oxidized at the surface of goethite:

The presence of the nitroso-DFOB fragment in the suspension is indirect evidence of Fe(III) reduction. Fe(II) is not detected in solution, which can be explained by its adsorption at the goethite surface followed by electron transfer into the bulk crystal.70

9.2.2 At a Total DFOB Concentration of 0.25 µmol/m2

At the lower total concentration of DFOB, the concentration of dissolved Fe(III) plateaus after 60 hours of the reaction (Fig. 17, Paper III). At the same time, the concentration of uncomplexed DFOB calculated according to Eq. 3 reaches zero (Fig. 18).

ComplexedDFOB

tSupernatanDFOB

dUncomplexeDFOB cc c −= (3)

where dUncomplexeDFOBc is the concentration of free DFOB in the solution phase,

ComplexedDFOBc is the concentration of DFOB complexed to Fe(III) in the solution

phase and is equal to the concentration of dissolved Fe(III).

NH

OH

R2 Fe(OH)3 N

O

R2 Fe(OH)2 2H2O+ + +

Page 37: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

31

Figure 17. The concentrations of Fe(III) dissolved, acetate released and DFOB unaccounted as a function of time at pH 6 and a 0.25 µmol/m2 total DFOB concentration.

Figure 18. The concentrations of DFOB unaccounted, uncomplexed and adsorbed in a goethite suspension as a function of time at pH 6 and a 0.25 µmol/m2 total DFOB concentration.

In spite of the fact that at this time point all free DFOB is consumed and present either as adsorbed or complexed to dissolved Fe(III), the concentration of acetate continues to increase. The hydrolysis of adsorbed DFOB seems to partially explain the formation of acetate after 60 hours of the reaction. The concentration of adsorbed DFOB, calculated according to Eq. 2, starts to decrease after about 20 hours of the reaction, which is in agreement with the above-mentioned IR data. Another contribution to the acetate formation may be from the FeHDFOB+ complex via either 1) its

Page 38: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

32

adsorption as a ternary complex and surface-mediated DFOB hydrolysis or 2) its dechelation, separate adsorption of Fe(III) and DFOB and surface-mediated DFOB hydrolysis. The decrease in the concentration of dissolved Fe(III) after the 60-hour reaction supports the idea of the removal of FeHDFOB+ from the solution phase. The possibility of adsorption of MeHDFOB+ at the goethite surface is further elaborated in the next section.

9.3. Interaction of MeHDFOB+ with Goethite: Ternary Complexation

The addition of 1.2 µmol/m2 total FeHDFOB+ concentration to a goethite suspension at pH 6 resulted in the continuous removal of Fe(III) from aqueous phase and the release of acetate, suggesting hydrolysis of DFOB at the surface (Fig. 19, Paper III). The concentration of adsorbed DFOB calculated according to Eq. 2 (see Section 9.2.1) is estimated to be 10-12% of the total FeHDFOB+ concentration.

Figure 19. The concentrations of Fe(III) dissolved and acetate released as a function of time in a goethite suspension containing 1.2 µmol/m2 total FeHDFOB+ concentration at pH 6.

IR spectra were collected as a function of time after addition of FeHDFOB+, AlHDFOB+, or GaHDFOB+ to the goethite suspension (Fig. 20). In all three cases, the spectra collected after a 15-minute reaction time closely resemble the IR spectra of the corresponding aqueous species, while the spectra after 6 hours of reaction more closely resemble the IR

Page 39: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

33

spectrum of adsorbed DFOB. This may be explained by a partial dissociation of the MeHDFOB+ complex at the surface and adsorption of a metal ion and DFOB at different surface sites.71 However, a subtraction of the IR spectrum of adsorbed DFOB from the IR spectra collected after 6-hour reaction of AlHDFOB+ or GaHDFOB+ with goethite revealed that a fraction of MeHDFOB+ remained adsorbed at the surface (see Fig. 7 in Paper III). We propose that MeHDFOB+ complexes are coordinated at the water-goethite interface as ternary complexes. Outer-sphere coordination of intact complexes would be unlikely due to electrostatic repulsion between positively charged complexes and the surface.

Figure 20. The IR spectra recorded 15 minutes and 6 hours after addition of 1 µmol/m2 MeHDFOB+ to a goethite suspension at pH 6. The IR spectra of the corresponding aqueous species are also shown.

Page 40: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

34

10. Dissolution of Goethite in the Presence of Oxalate and DFOB at pH 6

Figure 21 shows that the presence of both oxalate and DFOB in a goethite suspension has a synergistic effect on the rates of goethite dissolution, i.e. the dissolution rate in the two-ligand system is substantially higher than the sum of the dissolution rates in the single-ligand systems. The reaction of oxalate with a goethite suspension does not lead to a net mineral dissolution, but the addition of DFOB to goethite pre-reacted with oxalate triggers a fast release of Fe(III) into the solution phase. This synergistic effect of siderophores and carboxylates on dissolution of iron oxides has been under extensive investigation.29,31-33 Based on the macroscopic solution-phase studies, the following reaction mechanism has been proposed33: 1) adsorption of a carboxylate at mineral surface, 2) formation of a highly labile Fe(III)-carboxylate surface species, 3) detachment of this Fe(III)-carboxylate species and substitution of the carboxylate with a siderophore, 4) readsorption of the freed carboxylate to the mineral surface and the formation of a new portion of the labile species and so on. However, the understanding of this process at a molecular level has been missing and the nature of the highly labile surface species has been unknown.

Figure 21. The concentration of dissolved Fe(III) in a goethite suspension at pH 6, in the presence of a 1 µmol/m2 total DFOB concentration (triangles), in the presence of both oxalate (1 µmol/m2) and DFOB (1 µmol/m2) added simultaneously (squares), 3) during prereaction with 1 µmol/m2 of oxalate for 5 days and after addition of 1 µmol/m2 of DFOB (circles).

Page 41: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

35

In order to gain more insight on the nature of the labile Fe(III)-carboxylate surface species, we collected the IR spectra in a goethite suspension during the reaction with oxalate and after addition of DFOB (Fig. 22). As shown in Section 8, at pH 6 the interaction of oxalate with goethite results in the growth of the IR bands at 1713, 1689, 1413, 1270 cm-1 due to the ternary type A oxalate-Fe(III)-goethite complex formed during the dissolution-readsorption process. When DFOB is added to the oxalate-goethite system, we observe a dramatic decrease in these bands and an increase in the bands at 1610 and 1307 cm-1 due to outer-sphere oxalate complexes (Fig. 22). There is also an increase in the bands at 1630, 1577, 1470, 1443 cm-1 due to the adsorbed DFOB. A fraction of oxalate seems to remain coordinated inner-spherically to Fe(III) as indicated by the bands at 1713 and 1681 cm-1 resolved in the IR spectra after reaction with DFOB. These observations suggest that a key to the synergistic effect between oxalate and DFOB on the dissolution of goethite is the formation of the ternary type A oxalate-Fe(III)-goethite complexes at the water-goethite interface. Since the ternary complex is connected to the surface only via two bonds that need to be broken upon dissolution (Fig. 23), iron in this ternary

Figure 23. Schematic representations of two types of inner-sphere complexes of oxalate at the surface of goethite

Figure 22. The IR spectra collected as a function of time after addition of 1 µmol/m2 total DFOB concentration to the goethite suspension pre-equilibrated with 1 µmol/m2 total oxalate concentration for five days at pH 6.

Page 42: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

36

complex is likely more labile and therefore accessible to other complexing agents such as siderophores than the iron incorporated into the crystal lattice.

It follows that the concentration of the ternary complexes may be estimated by the measurement of the Fe(III) released after addition of DFOB to a goethite suspension pre-equilibrated with oxalate at different prereaction times. Figure 24 shows a plot of the concentration of Fe(III) released after addition of DFOB versus the integrated absorbance of the CC

bC −− +νν O band

between 1495 and 1345 cm-1 mostly originating from inner-sphere oxalate. Initially, the Fe(III) release is relatively constant up to the integrated absorbance of 0.19 and can be attributed to DFOB-promoted dissolution. The integrated absorbance in this region increases mostly due to oxalate coordinated to the relatively immobile surface iron. Once the ternary complexes are formed at the detectable amounts, the addition of DFOB to the oxalate-goethite system leads to a release of Fe(III) into the solution phase, which is linearly correlated to the integrated absorbance. The plot shows that by the end of the experiment (2 days) the concentration of iron in ternary complexes corresponds to 0.055 µmol/m2.

Figure 24. The concentration of Fe(III) dissolved after DFOB was added to a goethite suspension prereacted with oxalate (at a given prereaction time) versus the integrated absorbance of the CC

bC −− +νν O band measured before the DFOB

addition.

Page 43: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

37

11. Conclusions

The results obtained within this research project provide new valuable molecular-level information explaining processes occurring at the water-goethite interface in the presence of oxalate and DFOB.

In the oxalate-goethite system, IR data complemented with EXAFS results revealed the formation of four oxalate surface species at pH 4 and 6: 1) electrostatically attracted outer-sphere complex, 2) hydrogen bonded outer-sphere complex, 3) inner-sphere oxalate complex coordinated to surface iron and 4) a ternary type A complex. The time-dependency of the surface speciation is explained by a dissolution-readsorption process which is faster at pH 4 than 6. This process leads to a build-up of ternary surface complexes at the expense of the outer-sphere species.

In the DFOB-goethite system, the determination of the coordination mode was complicated due to interference from water bending mode and the sensitivity of the spectra to conformational changes. Nevertheless, we suggest that some of the hydroxamic groups are likely to coordinate inner-spherically, while the others might remain protonated. The number of inner-spherically coordinated hydroxamic groups is still an open question.

Macroscopic solution-phase measurements in the DFOB-goethite system showed that at pH 6 and in the absence of visible light, DFOB is subjected to surface-mediated hydrolysis accompanied by the reduction of surface Fe(III), as evidenced by the formation of acetate and nitroso-DFOB fragments. These observations might suggest that DFOB-promoted dissolution of goethite may proceed not only via pure ligand-exchange reactions, but may involve reductive pathways.

Similar experiments but at lower total DFOB concentration, as well as experiments in the FeHDFOB+-goethite system demonstrated that a small fraction of dissolved FeHDFOB+ can adsorb at the goethite surface, presumably as a ternary complex. The otherwise stable FeHDFOB+ complex undergoes surface-induced hydrolysis when coordinated to goethite surface, and this process might be a pathway of iron release from FeHDFOB+ in nature.

In the two-ligand system, the synergistic effect between oxalate and DFOB on goethite dissolution is associated with the formation of the ternary

Page 44: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

38

type A oxalate-Fe(III)-goethite complexes. Such a synergistic effect is not exclusive for oxalate, and it is most likely that similar ternary complexes are also formed in the presence of other carboxylates.

The findings of this work improve our understanding of the mechanisms of iron oxide dissolution in the presence of carboxylate and siderophores. However, some questions remain for further investigation, e.g. the actual coordination mode of DFOB, and the role of Fe(II) in the DFOB-promoted dissolution. It would also be interesting to investigate origins of synergistic effects of other carboxylates and siderophores on mineral dissolution under different conditions (pH, ligand concentrations). With more quantitative data, surface complexation and kinetic models could be created, that would substantially help to predict iron bioavailability in soils.

Page 45: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

39

12. Acknowledgements

I would like to express my sincere gratitude to those people who had helped and supported me during this work.

First of all, I want to say billions of thanks to my supervisors Dr. John Loring and Prof. Per Persson for invaluable discussions, help, encouragement, never-ending enthusiasm, and optimism… John, it has been an honor to be your first PhD student! Thank you for everything you had taught me during these years, for all advices and support you gave me not only in the “science” world, but also in the “real life” world . Per, thank you for encouraging me to think at both the microscopic and global scales and also thanks for your wise advice (not necessarily related to chemistry).

The Swedish Research Council and the Kempe Foundation are acknowledged for financial support.

I would like to thank Prof. Staffan Sjöberg for helping me with thermodynamic calculations, for fruitful discussions and valuable comments on the manuscripts and summary.

Dr. Madeleine Ramstedt, thank you so much for inspirational conversations, positive attitude and helpful comments on the summary.

Big hug and thank go to Prof. Lage Pettersson for the optimism and endless funny stories! I will never forget those parties that we had in our old fikarum on the 6th floor and that were often ending with debates about the history .

I wish to thank Andrei and Irina for their incredible help from our first days in Umeå, Caroline and Jörgen for their great company during the Goldschmidt conferences and proofreading, Janice for the linguistic revision of the summary, Ingegärd for helping with practical problems in the lab, Gun-Britt for solving all the administrative problems, for being kind and patient through all of my questions.

I am also grateful to Erik Björn, Knut Irgum, Nhat Bui Thi Hong, Rickard Lindberg and people from the organic chemistry for their help with ICP, HPLC and MS analyses.

All colleagues, former and present PhD students and postdocs at the 6th floor of the KBC building, thank you for creating a warm and friendly

Page 46: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

40

working environment. Special thanks to my officemates during these years, Laura, Peter, Sandra, Kristoffer, for all conversations, small breaks and good laughs!

I also want to send a huge thank to all my friends in Sweden and beyond for their support, great parties, for all the good times and fun we had together. Спасибо вам, Марина, Рома, Оля, Аня и Вадим, Олена и Патрик, Микола, Алёна и Вова, Надя и Миша, Настя, Наташа, Филипп…

Огромное спасибо моим родителям, брату, любимой свекрови за неоценимую поддержку и помощь во всем.

Спасибо моему солнышку, моей радости, Сонюшке!

And most of all I would love to thank a person who this thesis is dedicated to and who is not among us anymore - to my beloved husband Artem Liseenkov…. There are no words in the world to express everything I want… Thank you for everything, for being with me for ten wonderful years, for countless happy moments that we had, for your invaluable support… You are forever in my heart…...

Page 47: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

41

13. References

(1) Zinder, B.; Furrer, G.; Stumm, W., The Coordination Chemistry of Weathering. 2. Dissolution of Fe(III) Oxides. Geochim. Cosmochim. Acta 1986, 50 (9), 1861-1869.

(2) Schwertmann, U., Solubility And Dissolution of Iron-Oxides. Plant Soil 1991, 130 (1-2), 1-25.

(3) Stumm, W., Reactivity at the mineral-water interface: dissolution and inhibition. Colloids and Surfaces A: Physicochemical and Engineering Aspects 1997, 120 (1-3), 143-166.

(4) Kappler, A.; Straub, K. L., Geomicrobiological cycling of iron. In Molecular Geomicrobiology, 2005, 59, 85-108.

(5) Lindsay, W. L., Iron oxide solubilization by organic matter and its effect on iron availability. Plant Soil 1991, 130 (1), 27-34.

(6) Jones, D. L., Organic acids in the rhizosphere – a critical review. Plant Soil 1998, 205 (1), 25-44.

(7) Prapaipong, P.; Shock, E. L.; Koretsky, C. M., Metal-organic complexes in geochemical processes: temperature dependence of the standard thermodynamic properties of aqueous complexes between metal cations and dicarboxylate ligands. Geochim. Cosmochim. Acta 1999, 63 (17), 2547-2577.

(8) Vančura, V., Root exudates of plants. Plant Soil 1964, 21 (2), 231-248.

(9) Jones, D. L.; Darrah, P. R.; Kochian, L. V., Critical evaluation of organic acid mediated iron dissolution in the rhizosphere and its potential role in root iron uptake. Plant Soil 1996, 180 (1), 57-66.

(10) Zhang, Y.; Kallay, N.; Matijevic, E., Interaction of metal hydrous oxides with chelating agents. 7. Hematite-oxalic acid and -citric acid systems. Langmuir 1985, 1 (2), 201-206.

(11) Drever, J. I.; Stillings, L. L., The role of organic acids in mineral weathering. Colloids and Surfaces A: Physicochemical and Engineering Aspects 1997, 120 (1-3), 167-181.

(12) Bhandari, N.; Hausner, D. B.; Kubicki, J. D.; Strongin, D. R., Photodissolution of Ferrihydrite in the Presence of Oxalic Acid: An In Situ ATR-FTIR/DFT Study. Langmuir 2010, 26 (21), 16246-16253.

(13) Panias, D.; Taxiarchou, M.; Paspaliaris, I.; Kontopoulos, A., Mechanisms of dissolution of iron oxides in aqueous oxalic acid solutions. Hydrometallurgy 1996, 42 (2), 257-265.

Page 48: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

42

(14) Panias, D.; Taxiarchou, M.; Douni, I.; Paspaliaris, I.; Kontopoulos, A., Dissolution of hematite in acidic oxalate solutions: The effect of ferrous ions addition. Hydrometallurgy 1996, 43 (1-3), 219-230.

(15) Cwiertny, D. M.; Hunter, G. J.; Pettibone, J. M.; Scherer, M. M.; Grassian, V. H., Surface Chemistry and Dissolution of alpha-FeOOH Nanorods and Microrods: Environmental Implications of Size-Dependent Interactions with Oxalate. J. Phys. Chem. C 2009, 113 (6), 2175-2186.

(16) Siffert, C.; Sulzberger, B., Light-Induced Dissolution of Hematite in the Presence of Oxalate - a Case-Study. Langmuir 1991, 7 (8), 1627-1634.

(17) Duckworth, O. W.; Martin, S. T., Surface complexation and dissolution of hematite by C-1-C-6 dicarboxylic acids at pH=5.0. Geochim. Cosmochim. Acta 2001, 65 (23), 4289-4301.

(18) Wang, X.; Li, Q.; Hu, H.; Zhang, T.; Zhou, Y., Dissolution of kaolinite induced by citric, oxalic, and malic acids. J. Colloid Interf. Sci. 2005, 290 (2), 481-488.

(19) Li, J. Y.; Xu, R. K.; Tiwari, D.; Ji, G. L., Effect of low-molecular-weight organic acids on the distribution of mobilized Al between soil solution and solid phase. Appl. Geochem. 2006, 21 (10), 1750-1759.

(20) Marschner, H.; Romheld, V., Strategies of Plants for Acquisition of Iron. Plant Soil 1994, 165 (2), 261-274.

(21) Watteau, F.; Berthelin, J., Microbial Dissolution of Iron and Aluminum from Soil Minerals - Efficiency and Specificity of Hydroxamate Siderophores Compared to Aliphatic Acids. Eur. J. Soil Biol. 1994, 30 (1), 1-9.

(22) Kraemer, S. M., Iron oxide dissolution and solubility in the presence of siderophores. Aquat. Sci. 2004, 66 (1), 3-18.

(23) Martell, A. E.; Smith, R. M.; Motekaitis, R. J., Critically selected stability constants of Metal complexes Database version 5.0. Texas A & M University, College station: TX, USA, 1998.

(24) Hersman, L.; Lloyd, T.; Sposito, G., Siderophore-Promoted Dissolution of Hematite. Geochim. Cosmochim. Acta 1995, 59 (16), 3327-3330.

(25) Borer, P.; Sulzberger, B.; Hug, S. J.; Kraemer, S. M.; Kretzschmar, R., Wavelength-Dependence of Photoreductive Dissolution of Lepidocrocite (gamma-FeOOH) in the Absence and Presence of the Siderophore DFOB. Environ. Sci. Technol. 2009, 43 (6), 1871-1876.

(26) Borer, P.; Kraemer, S. M.; Sulzberger, B.; Hug, S. J.; Kretzschmar, R., Photodissolution of lepidocrocite (gamma-FeOOH) in the presence of desferrioxamine B and aerobactin. Geochim. Cosmochim. Acta 2009, 73 (16), 4673-4687.

Page 49: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

43

(27) Borer, P. M.; Sulzberger, B.; Reichard, P.; Kraemer, S. M., Effect of siderophores on the light-induced dissolution of colloidal iron(III) (hydr)oxides. Marine Chemistry 2005, 93 (2-4), 179-193.

(28) Borer, P.; Sulzberger, B.; Hug, S. J.; Kraemer, S. M.; Kretzschmar, R., Photoreductive Dissolution of Iron(III) (Hydr)oxides in the Absence and Presence of Organic ligands: Experimental Studies and Kinetic Modeling. Environ. Sci. Technol. 2009, 43 (6), 1864-1870.

(29) Reichard, P. U.; Kretzschmar, R.; Kraemer, S. M., Rate laws of steady-state and non-steady-state ligand-controlled dissolution of goethite. Colloids and Surfaces A-Physicochemical and Engineering Aspects 2007, 306 (1-3), 22-28.

(30) Cervini-Silva, J.; Sposito, G., Steady-state dissolution kinetics of aluminum-goethite in the presence of desferrioxamine-B and oxalate ligands. Environ. Sci. Technol. 2002, 36 (3), 337-342.

(31) Cheah, S. F.; Kraemer, S. M.; Cervini-Silva, J.; Sposito, G., Steady-state dissolution kinetics of goethite in the presence of desferrioxamine B and oxalate ligands: implications for the microbial acquisition of iron. Chem. Geol. 2003, 198 (1-2), 63-75.

(32) Reichard, P. U.; Kraemer, S. M.; Frazier, S. W.; Kretzschmar, R., Goethite dissolution in the presence of phytosiderophores: Rates, mechanisms, and the synergistic effect of oxalate. Plant Soil 2005, 276 (1-2), 115-132.

(33) Reichard, P. U.; Kretzschmar, R.; Kraemer, S. M., Dissolution mechanisms of goethite in the presence of siderophores and organic acids. Geochim. Cosmochim. Acta 2007, 71 (23), 5635-5650.

(34) Hug, S. J.; Sulzberger, B., In situ Fourier Transform Infrared Spectroscopic Evidence for the Formation of Several Different Surface Complexes of Oxalate on TiO2 in the Aqueous Phase. Langmuir 1994, 10 (10), 3587-3597.

(35) Hug, S. J.; Bahnemann, D., Infrared spectra of oxalate, malonate and succinate adsorbed on the aqueous surface of rutile, anatase and lepidocrocite measured with in situ ATR-FTIR. Journal of Electron Spectroscopy and Related Phenomena 2006, 150 (2-3), 208-219.

(36) Young, A. G.; McQuillan, A. J., Adsorption/Desorption Kinetics from ATR-IR Spectroscopy. Aqueous Oxalic Acid on Anatase TiO2. Langmuir 2009, 25 (6), 3538-3548.

(37) Mendive, C. B.; Bahnemann, D. W.; Blesa, M. A., Microscopic characterization of the photocatalytic oxidation of oxalic acid adsorbed onto TiO2 by FTIR-ATR. Catalysis Today 2005, 101 (3-4), 237-244.

(38) Dobson, K. D.; McQuillan, A. J., In situ infrared spectroscopic analysis of the adsorption of aliphatic carboxylic acids to TiO2, ZrO2, Al2O3, and

Page 50: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

44

Ta2O5 from aqueous solutions. Spectrochimica Acta Part A-Molecular and Biomolecular Spectroscopy 1999, 55 (7-8), 1395-1405.

(39) Noren, K.; Loring, J. S.; Bargar, J. R.; Persson, P., Adsorption Mechanisms of EDTA at the Water-Iron Oxide Interface: Implications for Dissolution. J. Phys. Chem. C 2009, 113 (18), 7762-7771.

(40) Lindegren, M.; Loring, J. S.; Persson, P., Molecular Structures of Citrate and Tricarballylate Adsorbed on alpha-FeOOH Particles in Aqueous Suspensions. Langmuir 2009, 25 (18), 10639-10647.

(41) Noda, I.; Ozaki, Y., Two-dimensional correlation spectroscopy: applications in vibrational and optical spectroscopy. John Wiley & Sons: 2004.

(42) Yost, E. C.; Tejedortejedor, M. I.; Anderson, M. A., Insitu CIR-FTIR Characterization of Salicilate Complexes at the Goethite Aqueous-Solution Interface. Environ. Sci. Technol. 1990, 24 (6), 822-828.

(43) Yang, J.; Bremer, P. J.; Lamont, I. L.; McQuillan, A. J., Infrared spectroscopic studies of siderophore-related hydroxamic acid ligands adsorbed on titanium dioxide. Langmuir 2006, 22 (24), 10109-10117.

(44) Persson, P.; Axe, K., Adsorption of oxalate and malonate at the water-goethite interface: molecular surface speciation from IR spectroscopy. Geochim. Cosmochim. Acta 2005, 69 (3), 541-552.

(45) Borer, P.; Hug, S. J.; Sulzberger, B.; Kraemer, S. M.; Kretzschmar, R., ATR-FTIR spectroscopic study of the adsorption of desferrioxamine B and aerobactin to the surface of lepidocrocite (gamma-FeOOH). Geochim. Cosmochim. Acta 2009, 73 (16), 4661-4672.

(46) Socrates, G., Infrared characteristic group frequencies. Wiley: 1980.

(47) Begun, G. M.; Fletcher, W. H., Vibrational Spectra of Aqueous Oxalate Ion. Spectrochimica Acta 1963, 19 (8), 1343-1349.

(48) Axe, K.; Persson, P., Time-dependent surface speciation of oxalate at the water-boehmite (gamma-AlOOH) interface: Implications for dissolution. Geochim. Cosmochim. Acta 2001, 65 (24), 4481-4492.

(49) Cabaniss, S. E.; Leenheer, J. A.; McVey, I. F., Aqueous infrared carboxylate absorbances: aliphatic di-acids. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 1998, 54 (3), 449-458.

(50) Yoon, T. H.; Johnson, S. B.; Musgrave, C. B.; Brown, G. E., Adsorption of organic matter at mineral/water interfaces: I. ATR-FTIR spectroscopic and quantum chemical study of oxalate adsorbed at boehmite/water and corundum/water interfaces. Geochim. Cosmochim. Acta 2004, 68 (22), 4505-4518.

Page 51: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

45

(51) Degenhardt, J.; McQuillan, A. J., Mechanism of oxalate ion adsorption on chromium oxide-hydroxide from pH dependence and time evolution of ATR-IR spectra. Chemical Physics Letters 1999, 311 (3-4), 179-184.

(52) Johnson, S. B.; Yoon, T. H.; Slowey, A. J.; Brown, G. E., Adsorption of organic matter at mineral/water interfaces: 3. Implications of surface dissolution for adsorption of oxalate. Langmuir 2004, 20 (26), 11480-11492.

(53) Specht, C. H.; Frimmel, F. H., An in situ ATR-FTIR study on the adsorption of dicarboxylic acids onto kaolinite in aqueous suspensions. Phys. Chem. Chem. Phys. 2001, 3 (24), 5444-5449.

(54) Fujita, J.; Nakamoto, K.; Martell, A. E., Infrared Spectra of Metal Chelate Compounds.6. Normal Coordinate Treatment of Oxalato Metal Complexes. J. Chem. Phys. 1962, 36 (2), 324-&.

(55) Fujita, J.; Nakamoto, K.; Martell, A. E., Infrared Spectra of Metal Chelate Compounds.7. Normal Coordinate Treatments on 1-2 and 1-3 Oxalato Complexes. J. Chem. Phys. 1962, 36 (2), 331-&.

(56) Clausen, M.; Ohman, L. O.; Axe, K.; Persson, P., Spectroscopic studies of aluminum and gallium complexes with oxalate and malonate in aqueous solution. J. Mol. Struct. 2003, 648 (3), 225-235.

(57) Duckworth, O. W.; Bargar, J. R.; Sposito, G., Quantitative Structure-Activity Relationships for Aqueous Metal-Siderophore Complexes. Environ. Sci. Technol. 2009, 43 (2), 343-349.

(58) Edwards, D. C.; Nielsen, S. B.; Jarzecki, A. A.; Spiro, T. G.; Myneni, S. C. B., Experimental and theoretical vibrational spectroscopy studies of acetohydroxamic acid and desferrioxamine B in aqueous solution: Effects of pH and iron complexation. Geochim. Cosmochim. Acta 2005, 69 (13), 3237-3248.

(59) Holmen, B. A.; TejedorTejedor, M. I.; Casey, W. H., Hydroxamate complexes in solution and at the goethite-water interface: A cylindrical internal reflection Fourier transform infrared spectroscopy study. Langmuir 1997, 13 (8), 2197-2206.

(60) Domagal-Goldman, S. D.; Paul, K. W.; Sparks, D. L.; Kubicki, J. D., Quantum chemical study of the Fe(III)-desferrioxamine B siderophore complex-Electronic structure, vibrational frequencies, and equilibrium Fe-isotope fractionation. Geochim. Cosmochim. Acta 2009, 73 (1), 1-12.

(61) Cornell, R. M. S. U., The iron oxides: structure, properties, reactions, occurences and uses. Weinheim ; Wiley-VCH: 2003.

(62) Venema, P.; Hiemstra, T.; Weidler, P. G.; van Riemsdijk, W. H., Intrinsic proton affinity of reactive surface groups of metal (hydr)oxides: Application to iron (hydr)oxides. J. Colloid Interf. Sci. 1998, 198 (2), 282-295.

Page 52: Molecular Perspectives on Goethite Dissolution in …umu.diva-portal.org/smash/get/diva2:439370/FULLTEXT01.pdfa fundamental molecular understanding has been lacking. Mineral dissolution

46

(63) Hiemstra, T.; VanRiemsdijk, W. H., A surface structural approach to ion adsorption: The charge distribution (CD) model. J. Colloid Interf. Sci. 1996, 179 (2), 488-508.

(64) Filius, J. D.; Hiemstra, T.; Van Riemsdijk, W. H., Adsorption of Small Weak Organic Acids on Goethite: Modeling of Mechanisms. J. Colloid Interf. Sci. 1997, 195 (2), 368-380.

(65) Boily, J. F.; Lutzenkirchen, J.; Balmes, O.; Beattie, J.; Sjoberg, S., Modeling proton binding at the goethite (alpha-FeOOH)-water interface. Colloids and Surfaces A-Physicochemical and Engineering Aspects 2001, 179 (1), 11-27.

(66) Barrón, V.; Torrent, J., Surface Hydroxyl Configuration of Various Crystal Faces of Hematite and Goethite. J. Colloid Interf. Sci. 1996, 177 (2), 407-410.

(67) Weidler, P. G.; Hug, S. J.; Wetche, T. P.; Hiemstra, T., Determination of growth rates of (100) and (110) faces of synthetic goethite by scanning force microscopy. Geochim. Cosmochim. Acta 1998, 62 (21-22), 3407-3412.

(68) Biber, M. V.; Afonso, M. D.; Stumm, W., The Coordination Chemistry of Weathering .4. Inhibition of the Dissolution of Oxide Minerals. Geochim. Cosmochim. Acta 1994, 58 (9), 1999-2010.

(69) Maurice, P. A.; Haack, E. A.; Mishra, B., Siderophore sorption to clays. Biometals 2009, 22 (4), 649-658.

(70) Williams, A. G. B.; Scherer, M. M., Spectroscopic evidence for Fe(II)-Fe(III) electron transfer at the iron oxide-water interface. Environ. Sci. Technol. 2004, 38 (18), 4782-4790.

(71) Duckworth, O. W.; Bargar, J. R.; Sposito, G., Sorption of ferric iron from ferrioxamine B to synthetic and biogenic layer type manganese oxides. Geochim. Cosmochim. Acta 2008, 72 (14), 3371-3380.